100% found this document useful (6 votes)
1K views

Fundamentals of Modern Physics

Fundamentals of Modern Physics En Teoría y relatividad encontraremos lo básico: Transformadas de Galileo, experimento de Michelson y Morley. Abarcando también los temas de radiación, electrones, descubrimiento del núcleo del átomo, Teoría de Bohr, ondas y partículas, soluciones a la ecuación de Schrödinger, Momento magnético.

Uploaded by

Israel Tuxpan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (6 votes)
1K views

Fundamentals of Modern Physics

Fundamentals of Modern Physics En Teoría y relatividad encontraremos lo básico: Transformadas de Galileo, experimento de Michelson y Morley. Abarcando también los temas de radiación, electrones, descubrimiento del núcleo del átomo, Teoría de Bohr, ondas y partículas, soluciones a la ecuación de Schrödinger, Momento magnético.

Uploaded by

Israel Tuxpan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 746

MuyjjjjJHpwi*

FUNDAMENTALS OF

MODERN

PHYSICS

FUNDAMENTALS OF

ROBERT MARTIN EISBERG


Associate Professor of Physics
University of California

Santa Barbara

John Wiley

&

Sons, Inc., New York

London

Sydney

MODERN PHYSICS

SoS'Q

13 14 15 16 17 18 19 20

Copyright

1961 by John Wiley & Sons,

All rights reserved.

Inc.

This book or any part

thereof must not be reproduced in any form

without the written permission of the publisher.


Library of Congress Catalog Card

Number: 61-6770

Printed in the United States of America

ISBN

471 23463

Dedicated to

JACOB LOUIS EISBERG

Prefao

This book has evolved from mimeographed notes which I have been
using at the University of Minnesota as a text for a one-year advanced
undergraduate course in modern physics. Its selection of material re-

some

definite ideas that I have concerning such a course.


convinced that a modern course in modern physics should have
two purposes: In addition to teaching many of the standard topics of
flects
I

the

am

field,

the course also should take full advantage of

tunity to introduce the student to

quantum mechanics

its

unique oppor-

an early stage
and in a way which integrates the theory into its historical development and its applications.
The introduction must be
elementary, but it must not be cursory because such treatment would
only amplify the problems that so often arise in a first contact with
quantum mechanics. This program takes time, but it pays great dividends by laying the foundation for a really mature discussion of the
atom and the nucleus. A student can surely learn much more about
these subjects in one year if half of it is invested in developing their
at

in his education,

theoretical bases than


tially

if all

of

it is

spent in discussing

them

at

an essen-

empirical level.

The
in this

inclusion of

book makes

more than
it

the usual

amount of quantum mechanics

necessary that there be certain omissions.

Because
an undergraduate course in solid state physics, I have chosen most of the omissions from
material that would be
covered in such a course.

most

institutions

now

offer

Although there are items in the table of contents of this book which
usually are found only in textbooks of a strictly postgraduate level,
this
should not be taken as a false indication of its prerequisites.
I have
tried to start near the beginning of each topic, and initially
to employ

PREFACE
only arguments that can be appreciated by a student well trained in
elementary physics, and in mathematics through intermediate calculus.

Almost

all the material of the book has been presented on two or


three
separate occasions to a very heterogeneous group of physics, chemistry,
mathematics, and engineering students, ranging from juniors to post-

graduates. Their performance has shown that students at all levels,


and
from all specialties, are quite capable of following the material.
However, in its final form the book contains about twenty percent
more material than can be covered in 90 lecture-hours. This, plus the
fact that there

is a gradual build-up of level in proceeding


through the
book, should make it adaptable to a variety of needs. A course
for
undergraduate students, who have had no prior contact with modern

physics, might thoroughly cover the material through


Chapter 8 and
leave for outside reading certain topics in the remaining
chapters.
course for postgraduate students, who have already studied
modem

physics at an elementary level, might begin with a very rapid


review of
the material through Chapter 5, leaving the details for
outside reading,
and then thoroughly cover the remaining chapters. There is also
enough
material for a one-semester course in introductory

quantum mechanics.

would

like to

particularly Drs.

thank

W.

my

colleagues at the University of Minnesota


B. Cheston, C. E. Porter, and D. R. Yennie,
for

reading parts of the mimeographed notes and


making

many

useful sug-

would also like to thank my wife Lila for the constant


aid
and encouragement which made the completion of this
book possible.

gestions.

R. M. ElSBERG
February, 1961

Contents

Introduction
1.

What

is

modern physics?

2. Historical sketch, 2.

1.

The theory of

1.

relativity

The Galilean transformation and

classical

mechanics, 3.

Galilean transformation and electromagnetic theory, 7.

Michelson-Morley experiment,
Simultaneity, 16.

9.

2.
3.

4. Einstein's postulates,

The
The
15.

Kinematical effects of relativity, 19.


The Lorentz transformation, 24. 8. Transformation of ve-

5.
7.

locity,

27.

9. Relativistic

momentum and
theory, 37.

6.

mechanics, 30.

energy, 36.

11.

Bibliography, 39.

10. Transformation of
Experimental verification of the

Exercises, 39.

Thermal radiation and the origin of quantum

theory
1.

41

Introduction, 41.

charges, 42.
46.

4.

2. Electromagnetic radiation by accelerated


Emission and absorption of radiation by surfaces,

3.

Black body radiation, 47.

Rayleigh- Jeans theory, 57.


tribution, 57.

theory, 63.

8.

10.

liography, 68.

7.

5.

Wien's law, 49.

The Boltzmann

Comparison with experiment, 62.

Some comments on
Exercises, 68.

6.

The

probability dis9.

Planck's

Planck's postulate, 66.

Bib-

CONTENTS

3
1.

Cathode

Electrons

rays, 70.

and quanta

The e/m

2.

ratio of electrons, 72.

charge and mass of electrons, 73.


5.

The

photoelectric effect, 76.

electric effect, 78.

79.

Quantum

7.

The Compton

8.

effect,

magnetic radiation, 85.

A
1.

3.

70

6. Classical

81.

9.

The dual nature

Bibliography, 86.

son's model, 92.

106.

5.

9.

8.

The

Exercises, 86.

2.

particles, 89.

ThomRuth-

6.

7. Predictions of Rutherford's

Experimental verification and determination of Z,

size of the nucleus,

Bibliography, 709.

Alpha

87

4. Predictions of

Comparison with experiment, 98.

model of the atom, 99.

model, 100.

effect,

of electro-

The discovery of the atomic nucleus

scattering of alpha particles, 91.

erford's

The

theory of the photo-

theory of the photoelectric

Thomson's model of the atom, 87.

The

3.

Bucherer's experiment, 75.

4.

707.

10.

problem, 108.

Exercises, 709.

Bohr's theory of atomic structure

10

1. Atomic spectra, 770.


2. Bohr's postulates, 113.
3. Bohr's
theory of the one-electron atom, 775.
4. Correction for finite
nuclear mass, 722.
5. Atomic energy states, 124.
6. The

Wilson-Sommerfeld quantization
relativistic theory,

9.

A critique of the old

rules,

725.

7.

Sommerfeld's

The correspondence principle, 135.


quantum theory, 756. Bibliography, 137.

131.

8.

Exercises, 138.

6
1.

De

5.

746.

The

3.

2.

4. Interpretation of the

uncertainty principle, 153.

139

Some

properties of the pilot

Experimental confirmation of de Broglie's postu-

uncertainty principle, 759.

1.

and waves

Broglie's postulate, 759.

waves, 747.
late,

Particles

Bohr quantization
6.

rule,

Some consequences

Bibliography, 765.

757.

of the

Exercises, 765.

Schroedinger's theory of quantum mechanics

Introduction, 164.

terpretation of the

2.

wave

The Schroedinger
function, 770.

4.

equation, 765.

3. In-

The time independent

164

CONTENTS
xi

Schroedinger equation, 776.


5. Energy quantization in the
Schroedinger theory, 178. 6. Mathematical properties
of wave
functions and eigenfunctions, 184.
7. The classical theory of
transverse waves in a stretched string, 192.

and

differential operators,

mechanics, 206.

1.

9.

The

Bibliography, 210.

8.

Expectation values

classical limit of

quantum

Exercises, 210.

Solutions of Schroedinger's equation

The

201.

212

free particle, 212.

2. Step potentials, 221.


3. Barrier
Square well potentials, 239. 5. The infinite
square well, 251. 6. The simple harmonic oscillator,
256. Bibliography, 266. Exercises, 266.

potentials, 231.

4.

Perturbation theory

1.

Introduction, 268.

3.

An

268

2. Time independent perturbations, 268.


example, 274. 4. The treatment of degeneracies,
278.
5. Time dependent perturbation theory, 285.
Bibliography, 291.
Exercises, 291.

10
1.

One-electron atoms

Quantum mechanics

cles,

293.

293

for several dimensions

The one-electron atom, 295.

2.

and several

parti-

Separation and
solution of the equation of relative motion,
298.
4. Quantum
numbers, eigenvalues, and degeneracies, 302. 5.
Eigenfunctions
3.

and probability

densities, 304.
6. Angular momentum operators,
Eigenvalue equations, 318. 8. Angular momentum
of
one-electron-atom eigenfunctions, 323.
Bibliography, 325.
Exercises, 325.

313.

7.

Magnetic moments,

spin,

and

relativistic

effects
1
.

397

Orbital magnetic moments, 327.

netic field, 330.


spin, 334.

4.

The

momentum, 346.
atoms, 353.

3.

2. Effects of

an external mag-

The Stern-Gerlach experiment and

spin-orbit interaction, 338.


6.

Relativistic

Bibliography, 358.

corrections

5.

electron

Total angular

for one-electron

Exercises, 358.

CONTENTS.

12
1.

2.

Quantum mechanical description of identical particles, 360.


Symmetric and antisymmetric eigenfunctions, 363. 3. The ex-

clusion principle, 365.


eigenfunctions, 369.
gas, 381.

4.

Additional properties of antisymmetric

The helium atom, 373.

5.

Bibliography, 390.

13
1.

360

Identical particles

Multi-electron

Introduction, 391.

Hartree theory, 396.


states of atoms, 414.

structure, 449.

10.

The Fermi

atoms

391

The Thomas-Fermi theory, 393. 3. The


The periodic table, 407. 5. Excited
6. Alkali atoms, 418.
1. Atoms with sev-

2.

4.

eral optically active electrons, 425.

coupling, 441.

6.

Exercises, 390.

The Zeeman

12. Transition rates

Lifetimes and line widths, 468.

8.

LS

and

coupling, 428.

443.

effect,

11.

9.

selection rules, 453.

Bibliography, 472.

JJ

Hyperfine
13.

Exercises,

473.

14
1.

X-rays

The discovery

spectra, 476.
spectra, 489.
electric effect

475

of X-rays, 475.

3.

X-ray

5.

The

Measurement of X-ray
4. X-ray continuum
of X-rays, 493.
6. The photo-

scattering

483.

and pair production, 505.

and attenuation coefficients, 509.


particles, 57 J. Bibliography, 577.

15
1.

2.

line spectra,

7.

Total cross sections

8. Positrons

and other

anti-

Exercises, 577.

Collision theory

Introduction, 575.

2.

518

Laboratory-center of mass transforma-

The Born approximation, 524. 4. Some applications of the Born approximation, 530.
5. Partial wave analysis,
534. 6. Some applications of partial wave analysis, 546. 7. Abtions,

579.

3.

sorption, 554.

16
1.

Bibliography, 558.

The nucleus

Introduction, 559.

clear sizes

Exercises, 558.

and the

2.

The composition

optical model, 565.

559
of nuclei, 567.
4.

3.

Nu-

Nuclear masses and

CONTENTS
abundances, 580.
cal

5.

The

mass formula, 589.

drop model and the semi-empiriMagic numbers, 592. 7. The Fermi

liquid

6.

gas model, 594.

8. The shell model, 599.


9. The collective
Alpha decay and fission, 609. 11. Beta decay,
Gamma decay and internal conversion, 640. 13. Prop-

model, 604.
620.

12.

10.

erties of excited states,

655.

Nuclear forces, 674.

16.

14. Nuclear reactions, 662.


15.
Mesons, 694.
Bibliography, 703.

Exercises, 704.

Index

707

Introduction

I.

What

Is

Modern

Physics?

Webster's dictionary gives as a primary definition of the word modem:


"pertaining to the present time, or time not long past." However, when a
physicist speaks of modern physics he is using the word in a somewhat
different way. Modern physics is used to designate certain specific fields
of physics. All these fields have two things in common: first, their
development has taken place since, roughly speaking, the year 1900; and,
second, the theories used to explain the phenomena with which these
fields are concerned are startlingly different from the theories
that were

in existence before 1900.

The term

that contrasts with

modern

physics

is

classical physics.

Classical physics comprises fields in

which the subjects studied are:


Newton's theory of mechanics and the varied phenomena which can be
explained in terms of that theory, Maxwell's theory of electromagnetic

phenomena and its applications, thermodynamics, and the kinetic theory


of gases. In modern physics the subjects studied are: the theory of
relativity and associated phenomena, the quantum theories and quantum
phenomena, and, in particular, the application of the relativity and
quantum theories to the atom and the nucleus.

Many physicists are working today in the fields of classical physics. On


the other hand, the theory of relativity, which dates from 1905, is certainly
not modern in the sense of Webster's definition. We see that the words
modern and classical, as used in physics, do not have quite their normal
temporal significance.
In this book we shall be concerned with modern physics. However,
as a point of departure, we shall first give a very brief description of the
status of classical physics

toward the end of the nineteenth century.


I

INTRODUCTION

Historical Sketch

2.

The astronomical observations of Tycho Brahe and their interpretation


by Kepler, plus the mechanical experiments of Galileo (all of which
happened within a few decades of the year 1600), were welded into an
by Newton (1687). By the end of
was very highly developed. It provided
mechanical phenomena known at that

elegant yet simple theory of mechanics

the nineteenth century this theory

a successful interpretation of all


time and served as a basis for the kinetic theory of gases which, in turn,
removed many mysteries from the theory of thermodynamics.

A great variety of facts concerning electric and magnetic fields and their
interaction were discovered during the nineteenth century.

phenomena were brought

All these

together within one framework by the beautiful

theory of Maxwell (1864). This theory provided an explanation for the


wave propagation of light which was in agreement with what was known
at that time of both geometrical and physical optics.
all the experimental data known at that time could be
Newton's theory of mechanics or the electromagnetic
theory of Maxwell. Physicists were beginning to feel quite self-satisfied,
and it appears that the majority were of the opinion that the work of their
successors would be 'merely to "make measurements to the next decimal

In fact, almost

fitted into either

place."

The turn of

saw the shattering of this tranquil


of quite revolutionary experimental and theoretical
developments, such as the theory of relativity, which demands that we
reject deeply seated intuitive ideas concerning space and time, and the
situation

by a

the present century

series

quantum theories, which make similar demands on our intuitive ideas


about the continuity of nature.
We shall begin our study of modern physics with the theory of relativity,
not because of chronological considerations since the relativity and early
quantum theories were developed concurrently, but because this order of
presentation will

make

the discussion

much more

convenient.

fairly

brief treatment of relativity will be followed

early

we

quantum

theories.

shall study in detail

atom and

the nucleus.

This will

by a similar treatment of the


lead us to quantum mechanics, which

and then apply to a thorough discussion of the

CHAPTER

The Theory
of Relativity

I.

The Galilean Transformation and

Classical Mechanics

In classical physics, the state of any mechanical system at some


time
can be specified by constructing a set of coordinate axes and giving the
coordinates and momenta of the various parts of the system
at that
t

time.

laws

If we know the forces acting between the various


parts, Newton's
make it possible to calculate the state of the system at any future time

t in terms of
a calculation

state at

its

we

t
It is often desirable that during or after such
specify the state of the system in terms of a new set of
.

coordinate axes, which is moving relative to the first set. The


two-fold
question then arises: How do we transform our description of the
system
from the old to the new coordinates, and what happens to the equations

which govern the behavior of the system when we make the transformation ?
The same question often arises when treating electromagnetic systems
with Maxwell's equations.

This is the question with which the theory


of relativity concerns itself.
Let us see what answer classical mechanics would give to this question.
Consider the simplest possible mechanical system one particle of mass
m acted on by a force F.f Using rectangular coordinates to describe the
system, we have the situation shown in figure (1-1).
t Boldface type will be used in this

The same symbol

in

normal type

book for symbols

will

that represent a vector quantity.


then represent the magnitude of the vector

quantity.

THE THEORY OF RELATIVITY

The

set

of four numbers

coordinates

x, y, z.

that

m
m

Fx Fy
,

and

Fz

we know

the particle has

Newton's laws,

F,

d 2x

F
7? = *
d2y

7?
dt

where

us that at time

(x, y, z, t) tells

Assuming

[Ch.

= F
,,

(1-1)

are the x, y,

and

components of the vector F,


motion of the

give us a set of differential equations which govern the


y

axis

y
Figure l-l.

system.

set of rectangular coordinates.

These equations are valid only

if

the frame of reference defined

by the coordinates x, y, z is an inertial frame, i.e., a frame of reference in


which a body not under the influence of forces, and initially at rest, will
remain at rest. If we know the three coordinates and three components
of momenta of the particle at an initial time t we can evaluate the two
arbitrary constants which appear in the general solution of each of the
differential equations (1-1) and can make specific predictions of the location
(and momentum) of the particle at any time t.
Now consider the problem of describing the system from the point of
view of a new frame of reference x', y', and z', which is moving to the right
with constant velocity u with respect to the old frame of reference, and
which maintains the relative orientation between the two frames shown in
figure (1-2). These two frames of reference are said to be in uniform
translation with respect to each other. There are now two sets of four
numbers (x, y, z, t) and (x', y', z', t') that can equally well be used to
specify the location of the particle at any instant of time t and t', which
are the times measured in the two frames of reference and defined such
that t t' =
when the two frames of reference coincide. What is the
,

Sec.

GALILEAN TRANSFORMATIONS

I]

AND MECHANICS

relationship between the sets of

numbers (x,y,z,t) and (x', y', z', /') ?


Newton, almost all physicists before 1900, and most probably the reader,
would certainly not hesitate to assert that

=X
y' = y
z' = z
=t
X

ut

(1-2)

t'

These equations are known as the Galilean transformation equations.


constitute the answer which classical physics gives to the first part
of the question that we have posed for the case of uniform translational
motion between the two frames of reference.

They

axis

axis

y
x

z' axis

z axis

Two

Figure 1-2.

supposed to be

To answer

dV/A"

by

axis

frames of reference

in

uniform translation.

The x and

x' axes are

collinear.

the second part of the question, let us evaluate the


quantity

differentiating the first

We have
dx'jdt

of equations (1-2) with respect to

= dxjdt

t.

and
Since

It is

t'

= d*x/dt*

dVfdt*

= d*xldt*

obvious that
2

d*y'ldt'
It is also true that

F in

<Px'ldt2
t,

d*yldt*

and

Fx = Fx Fy = Fy
.

the direction of the x or x' axis

is

F,

dV/dt"

=F

d*z/dt*

z because the component of


the same as seen from either frame

THE THEORY OF RELATIVITY

of reference

and

similarly for its other

two components.

for the unprimed quantities in equations (1-1),


motion in the primed system:

= Fx

we

[Ch.

Substituting

find the equations of

dt'

d 2 y'

dt'

Note that equations (1-3) have exactly the same mathematical form as
equations (1-1). Thus the answer to the second part of the question is the
very interesting one that Newton's laws, which govern the behavior of
the system, do not change when we make a Galilean transformation. We
have demonstrated

this for the case in

reference with respect to the other

is

which the velocity of one frame of

in the direction of the x or x' axis.

We leave it as an exercise for the reader to

show that these results are true


regardless of the direction of the relative velocity vector u. The x,
y, z
frame was an inertial frame because d2x/dt 2
d 2y\dt 2
d 2z/dt 2
if

From

we see that x', y', z' is also an inertial frame


because d 2x'\dt* = d^'jdt* = dV/dt" =
if F = 0.
The physical content of the result that we have just derived is quite easy
0.

equations (1-3)

to understand and, in fact, is something with which the reader is certainly


already familiar. Since Newton's laws are identical in any two inertial
frames, and since the behavior of a mechanical system is governed by

Newton's laws, it follows that the behavior of all mechanical systems will
be identical in all inertial frames even though they are in uniform translation with respect to each other.

Consider a laboratory which

is

located

With the car initially at rest with respect to the ground


be an inertial frame), let us perform a series of mechanical

in a railroad car.

(assumed to

experiments such as measuring the period of a swinging pendulum or


investigating the subject of collision mechanics by means of a serious game
of billiards. If we were to perform exactly the same experiments while the
car was traveling at constant speed over a very smooth and very straight
stretch of track, the fact that equations (1-1) and (1-3) are identical in
form requires that the experiments yield exactly the same results as we

obtain when the car is at rest with respect to the ground. This
in agreement with our everyday experience.

is

certainly

Thus we see that Newtonian mechanics predicts that all inertial frames
in uniform translation with respect to each other are equivalent in so far

GALILEAN TRANSFORMATIONS

Sec. 2]

AND ELECTROMAGNETICS

as mechanical phenomena are concerned.


corollary of this is: No
mechanical phenomena can be used to differentiate between inertial frames,
so it is impossible by mechanical experiments to prove that one of the
frames is in a state of absolute rest while the others move with respect to
it.

If the

windows were covered in the railroad car moving smoothly with


we would not be able to detect its motion.

respect to the ground,

2.

The

Galilean Transformation and Electromagnetic Theory

Let us next inquire into the behavior of electromagnetic phenomena


Galilean transformation. Electromagnetic phenomena
are discussed in classical physics in terms of a set of differential
equations
which govern their behavior (called Maxwell's equations), just as mechanics
is discussed in terms of the set of differential
equations which govern
mechanical phenomena (called Newton's laws). We shall not actually

when we perform a

carry through the transformation of Maxwell's equations because


the
calculation is quite complicated owing to the nature of these equations.

However, when a Galilean transformation is performed, it is found that


Maxwell's equations do change their mathematical formin sharp contrast to the behavior of Newton's equations. The point
is a very important
one in the development of the theory of relativity, and we must discuss
its
physical significance.

As mentioned in the Introduction, Maxwell's equations predict the


existence of electromagnetic disturbances which propagate through
space
in the characteristic

manner of wave motion, with a propagation velocity


on the frequency of the wave motion (at least in a

that does not depend

medium with constant index of refraction such


many different names have been given to

as the vacuum). Although


these disturbances (radio

~ 10
~ 10

waves for frequency


lO^sec, infrared radiation for frequency
light waves for frequency
10 15 /sec, X-rays for frequency
they are
physicists,

all

basically the

same phenomena.

13

19

/sec,

/sec)^

The nineteenth century

who were very mechanistic in their outlook,

felt quite sure that


the propagation of waves predicted by Maxwell's equations
requires the
existence of a propagation medium. Just as water waves
propagate
through water, so, according to their argument, electromagnetic

waves
must propagate through a medium. This medium was given the name
ether. The ether was obliged to have somewhat strange
properties in order
not to disagree with certain known facts. For instance, the ether must
be
massless since it was observed that electromagnetic waves such
as light
can travel through a vacuum. However, it must have elastic properties
in order to be able to sustain the vibrations which are
inherent in the idea

THE THEORY OF RELATIVITY

of wave motion. Despite these

[Ch.

the concept of the ether

was
which was to assume
that electromagnetic disturbances could be propagated without the aid of
a propagation medium. Existence of the ether being assumed, the
disturbances would certainly be expected to propagate with a fixed velocity
with respect to the ether, just as sound waves propagate with a fixed
considerably

more

difficulties,

attractive than the alternative,

velocity with respect to the air.


It was assumed that the electromagnetic equations, in the form
presented by Maxwell, were valid for the frame of reference which was at

rest

with respect to the ether

these equations leads to

electromagnetic

waves.

the so-called ether frame.

solution of

an evaluation of the propagation velocity of


The result was 3.00 x 10" cm/sec = c, in

agreement within experimental accuracy with the value for the velocity
of light measured by Fizeau (1849). However, in a new frame of reference
in uniform translational motion with respect to the ether, Maxwell's
equations would have a mathematical form different from their mathematical form in the ether frame since, as mentioned above, a Galilean
transformation changes the form of the equations. When the propagation
velocity of electromagnetic waves as seen from the new frame of reference
was evaluated by using Maxwell's equations in the form which they assume
in this frame, it was found to have a value different from c. The reason is
that the propagation velocity predicted by the equations depends on the
form of the equations, as certainly must be the case.
Thus the application of a Galilean transformation to Maxwell's equations leads to the prediction that a measurement of the velocity of fight
made in the new frame of reference would yield a result different from the
velocity c measured in the ether frame. It is not necessary to describe
in detail the predicted relation between the velocity of light in the ether
frame, the velocity of light in the new frame, and the velocity of the new
frame with respect to the ether frame, since it is exactly what would be
obtained from a simple vector addition calculation of the type done near
the beginning of any elementary physics text. That is,
* light wrt.

where

Flight wrt

new frame

" light wrt. ether

* new frame wrt. ether

\^~

The somewhat comether


plicated calculation of the velocity of light as measured in the new frame
of reference, performed by transforming Maxwell's equations and then
solving them in the new frame, is in complete agreement with the very
simple physical idea that light propagates in such a way that its velocity
with respect to the ether frame is always c, and that its velocity with respect
to any other frame can be found by doing a simple vector addition of
velocities. Note that the intuitive arguments made to justify the validity
c,

wrt.

with respect

to.

Sec. 3]

THE MICHELSON-MORLEY EXPERIMENT

of the vector addition of velocities are basically the same arguments which
behind the Galilean transformation.
In recapitulation, the theory of physics near the end of the nineteenth

lie

century was based

upon three fundamental assumptions: (a) the validity


of Newton's laws, (b) the validity of Maxwell's equations, and (c) the
validity of the Galilean transformation. Almost everything which
could
be derived from these assumptions was in good agreement with experiment
in every case in which the appropriate experiment had been performed.
With respect to the questions we have been discussing, these assumptions

predicted that

all frames of reference in uniform translation with


respect
to each other were equivalent as far as mechanical phenomena
were

concerned,

they were inertial frames, but in regard to electromagnetic


equivalent; there was only one frame of reference, the ether frame, in which the velocity of light was equal to c.
if

phenomena they were not

3.

The Michelson-Morley Experiment

In 1887 Michelson and Morley carried out an experiment which proved


to be of extreme importance. This experiment was designed to
demonstrate
the existence of the special frame of reference called the ether
frame, and
to determine the motion of the earth with respect to that frame.

The earth travels in a roughly circular orbit about the sun with a period
(by definition) of one year, and with an orbital velocity of the order of
6
3 x 10 cm/sec. It would seem unrealistic to make the a priori assumption
that the ether frame travels with the earth and, as we shall see later,
strong
arguments against this assumption were known at that time.
more

reasonable working hypothesis would be to assume that the ether frame


was at rest with respect to the center of mass of our solar system, or to
assume that it was at rest with respect to the center of mass of the universe.
In the first case the velocity of the earth with respect to the ether frame

would have a magnitude of the order of 106 cm/sec, and furthermore the
orientation of the velocity vector would change during the course of
a
year. (The velocity of a point on the surface of the earth, owing to its
daily rotation, is negligible compared to the orbital velocity.) In the
second case the velocity of the earth with respect to the ether frame would
be even greater.

The

basic idea of the Michelson-Morley experiment

was to measure the


from a frame of reference fixed with respect to
the earth, in two perpendicular directions. A moment's consideration of
the classical theory, as summarized by equation (1-4), will show that
the
apparent velocity of light with respect to the earth would be different for
velocity of light, as seen

THE THEORY OF RELATIVITY

10

light traveling in the different directions.

[Ch.

At first glance it might be expected


would be observed in such

that the order of magnitude of the effect which

a measurement would be
3

10 6

10

|velocity of earth wrt. ether|


r+tj

|velocity

of light wrt. ether|

10

1(T

where we have assumed that the ether frame is fixed with respect to the
solar system. But, in fact, there exists a theorem due to Fresnel and
Lorentz which says that it is impossible to devise an optical experiment

Figure 1-3.

An

interferometer.

where the velocity of the apparatus with respect to the ether can produce
a first order effect. The biggest possible effect is a second order effect.
This means that an apparatus designed to observe the difference in the
velocity of light traveling in two perpendicular directions must be able to
-4

We shall prove this


theorem for a specific case in the course of our discussion of the MichelsonMorley apparatus.
The numerical estimate which we have just made indicates that an
extremely sensitive apparatus would be required to observe effects due to
the motion of the earth with respect to the ether, and an even more
sensitive apparatus if quantitative measurements were desired. In order
to do the experiment, Michelson invented an interesting device called an
detect differences of the order of (10

interferometer,

shown schematically

10~ 8

in figure (1-3).

THE MICHELSON-MORLEY EXPERIMENT

Sec. 3]

||

Let us briefly consider the operation of this device.


(For more details
the reader should consult any text in physical
optics.)

Assume

that the apparatus

is

Light emitted by
Consider the central ray shown in the

source

at first

at rest with respect to the ether.

is collimated by lens C.
This ray is incident at 45 upon the half-silvered mirror
M. Such
a mirror has the property of reflecting half the light incident
upon

figure.

it

and

transmitting the remainder. Thus the incident ray is split


into ray 1 and
ray 2, which travel to mirrors
x and
2 and are then reflected back on
themselves. Half of each of the two rays returning to
is directed into
the telescope T. The intensity, due to the central ray,
of the light seen by

An

Figure 1-4.

the observer

interferometer moving through the ether.

be a function of the phase relationship of the two


Since the two rays are in phase when originally split

by M,

their relative

lengths

lx

I2

will

recombining rays.

and

phase upon recombining

will

depend only on the

as follows

rays in phase
2*

2.

where X

2.

rays out of phase

the wavelength of the light. Such a device can be used in many


ways; the standard meter is now defined in terms of the wavelength of a certain spectral line by using an interferometer.
Next, let us consider an interferometer moving, as indicated in
figure
(1^), with velocity v with respect to the ether, where v is parallel to the
is

different

direction of ray 2.

beam of

Assume

light is split

by

lx

l2

reflection

/.

In the present case the incident

from a mirror which

is

moving with

THE THEORY OF RELATIVITY

12

[Ch.

Assuming that light travels with a constant velocity


easy to show by a Huygens wavelet construction
that the reflected ray of light will be thrown forward at exactly the right
angle such that, after being reflected from
l5 it will be back at the axis
defined by S and C just in time to meet M. This is not a crucial point
because it would be possible to assume that
is a "rough" mirror which
reflects light diffusely. Figure (1-4) shows the situation as seen by an

respect to the ether.


in the ether frame,

it is

M
M

M' indicates the position of M at the instant


it;
M'2 indicates the position of 2

observer in the ether frame;


the ray reflected from

at the instant

ray

To

it

reflects

meets
2.

evaluate the phase relation of the two combining rays,

we

calculate

the time required for the two rays to travel their respective paths.

seen from the ether frame, ray


ct in
vt.

a time

From

t.

During

the figure

this

we

As

an oblique path of length


time the apparatus moves to the right a distance
travels along

see that

cV =

vh 2

2
I

so
c
2

Since v \c

<

1,

we can

use the binomial theorem and obtain

The

total time for ray

1 is

-"(> +

It

to calculate the time required by ray 2 from the point of view


of an observer fixed with respect to the apparatus. According to the last
of equations (1-2), no difficulties will arise from evaluating time intervals
It is easiest

in different frames of reference.

Equation (1-4) predicts

observer, the velocity of ray 2

is

v while

moving to the left. In each part of the journey


The time for the round trip is then

v while

distance

/.

that, for this

moving to the

=
C

+
V

Using the binomial theorem,

c
this

c\

becomes

-?('*3

2
c /

right,
it

and
a

travels

THE MICHELSON-MORLEY EXPERIMENT

Sec. 3]

13

The two combining rays are out of phase by an amount dt/r, where
T = Xjc = the period of the vibration for light waves of wavelength X.
This phase difference

is

dt_

(t 2

-t )c _2llv*c _lr?
1

c2c 2 X

as

Xc 2

In the actual experiment a measurement was made with the apparatus


shown in the figure, and then a second measurement was made after

the apparatus was rotated through 90.


moment's consideration will
show that the phase difference for the apparatus in the second position is
v2

~~_ ~x7
St'

The effect of rotating the apparatus is consequently expected to be a change


in the phase difference by an amount

_
AA =

dt

St'

[The experiment was done in this


that /x be precisely equal to

/2 ,

as

21 v

=T1
X

(1-6)y

<r

way in order to avoid the requirement


we have assumed in our calculation. It

show that equation (1-6) is true if lx is only approximately equal


even though (1-5) would then be incorrect.]

easy to

is

to

/2

In their experiment, Michelson and Morley used light of


wavelength
6 x 10- 5 cm and a length /
10* cm. If we use our estimate of
2
8
If)" , we find that equation (1-6) gives a value
v*lc
for A of about
0.5. This means that, if the combining rays were just in
phase in

the

first

measurement, they would be just out of phase in the second measurement.


An observer looking through the telescope while the apparatus was being
rotated

would

see brightness replaced

by darkness, or

vice versa.

To

the

extreme surprise of Michelson and Morley, no such effect was


observed.
To make sure that the null effect was not due to a fortuitous combination
of velocity vectors such that their laboratory happened to be stationary
with respect to the ether frame during the time their measurements
were
being made, Michelson and Morley made observations during

various
times of the day and various seasons of the year. An effect
was never
observed. Despite the predictions of the classical theory, the

Michelson-

Morley experiment showed that the velocity of light is the same


when
measured along two perpendicular axes in a reference frame which,
presumably, is moving relative to the ether frame in different directions
at
different times of the year.

THE THEORY OF RELATIVITY

14

[Ch.

These remarkable results captured the attention of the entire community


of physicists. The experiment was repeated with increased accuracy in a

number of laboratories, and it was confirmed that the observed effect was
zero to within the experimental accuracy of about 0.3 percent of the
theoretically predicted value. Many people tried to devise explanations for
the negative result.

Three separate ideas were put forth in an attempt to explain the results
of the Michelson-Morley experiment and yet retain as much as possible of
the physical theories then in existence. These were: the "ether drag"
the "Lorentz contraction" hypothesis, and the so-called
"emission theories."
hypothesis,

The
all

ether drag hypothesis

bodies of

finite

assumed that the ether frame was attached to


mass. The hypothesis was attractive because it would

Michelson-Morley experiment and yet


did not involve modification of the theories of mechanics or electro-

certainly lead to a null result in the


it

magnetism. But there were two reasons why it could not be accepted. In
1853 Fizeau had measured the velocity of light in a column of rapidly
flowing water. If the water were to drag the ether frame with it, the
observed velocity would just be the sum of the velocity of light in stationary
water and the velocity, of the water. Fizeau actually did observe a change in
the velocity of light, but the change

was only a

fraction of the velocity of

the .water and, furthermore, was fully accounted for by the electromagnetic

theory of Maxwell, without introducing the ether drag hypothesis, as due


to the motion of charges in water molecules polarized by the electro-

magnetic field. The second argument against this hypothesis comes from
a phenomenon known in astronomy as stellar aberration. The apparent
positions of stars are observed to

angular diameter.

This

of the earth about the sun; in


causes a vertical
at

move

in circular orbits of very small

a purely kinematical

effect due to the motion


same effect as that which
shower of rain to appear to a moving observer as falling
is

an angle. From

this

would not be present

fact, it is the

analogy

if

it is apparent that the astronomical effect


a light ray were to travel with constant velocity

with respect to the ether frame and

frame were fixed with respect

if that

to the earth.

The Lorentz contraction assumed


to the ether,

all

that,

owing to

their

material bodies contract by a factor of

motion
(1

relative

v 2 jc 2 ) 1A in

the direction of motion. This, of course, is a serious modification of the


concept of a rigid body that is basic to classical mechanics. The reader
would find it worth while to demonstrate for himself that, if an interferometer moving through the ether were distorted in this way, a null
result

would be obtained

in the Michelson-Morley experiment.

for an interferometer in which

lx

2,

However,

a calculation similar to the one

Sec. 4]

EINSTEIN'S

POSTULATES

15

above, but including the Lorentz contraction, predicts a


change in phase
difference

by an amount

when

the velocity of the interferometer with respect to the ether


changes
v'. Such a change in velocity could be
provided by the motion of
the earth in its orbit. The appropriate experiment has
been

from

v to

Kennedy, who observed no

effect.

done by
Thus the Lorentz contraction hypoth-

esis is ruled out.

In an emission theory Maxwell's equations are modified in such


a way
wave remains associated with the velocity of its
source. It is easy to see that this too would lead to a null
result in the
that the velocity of a light

Michelson-Morley experiment, but it must be rejected because it


conflicts
with certain astronomical evidence concerning binary stars. Binary
stars
are pairs of stars which are rotating about their common
center of gravity.
Consider such a pair at a time when one is moving toward the earth

and the
an emission theory were valid, the velocity
of light with respect to the earth from one star would be larger
than that
of light from the other star. This would cause them to appear
to have
very unusual orbits about their center of gravity. However,
the observed
motion of binary stars is completely accounted for by Newtonian
other

is

moving away. Then,

if

mechanics.

None of these

attempts to "patch

A more fundamental change was


4.

up"

the classical theories

was

tenable.

needed.

Einstein's Postulates

All the evidence we have presented seems to be consistent


only with the
conclusion that there is no special frame of reference, the ether
frame, with
the unique property that the velocity of light in the
frame
is

equal' to

c.

Just as for inertial frames

and mechanical phenomena, all frames in


uniform translation with respect to each other seem to be
equivalent in
that the propagation velocity of light, measured in any
direction, is equal
to the same value, c. Einstein (1905) was willing not only
to accept this
evidence but also to generalize on it. In fact, he stated as
a postulate:
The laws of electromagnetic phenomena (including the
fact that the
propagation velocity of light is equal to the constant value c), as
well as the
laws of mechanics, are the same in all inertial frames
of reference, despite
the fact that these frames may be in uniform translation
with respect to
each other. Consequently, all inertial frames are completely
equivalent.

THE THEORY OF RELATIVITY

16

[Ch.

This postulate required that Einstein change either Maxwell's equations


or the Galilean transformation, since the two together imply the contrary
of the postulate. Guided by the failure of the emission theories, he chose

not to modify Maxwell's equations.

This he indicated by stating as a

second postulate:

The

velocity

of light

is

independent of the motion of its source.

Einstein was then forced to modify the Galilean transformations. This


was a very bold move. The intuitive belief in the validity of the Galilean
transformation was so strong that his contemporaries had never seriously

questioned

it.

And

yet, as

we

shall see, the very different transformation

which Einstein proposed in lieu of the Galilean one is based on realistic


physical considerations whereas the Galilean transformation is grossly
unrealistic. Another indication of the boldness of this move is that the
arguments of section 1 show that any modification of the Galilean transformation will require some compensating modification of Newton's
equations in order that the first postulate will continue to be satisfied for
mechanics.
first

5.

We

shall see later

we must develop

the

what

interesting results this leads to, but

new transformation

equations.

Simultaneity

Consider the fourth of the Galilean transformation equations


which is
,

(1-2),

is the same time scale for any two inertial


frames of reference in uniform translation with respect to each other, and

This equation says that there

this is equivalent to saying that there exists

such frames.

Is this

true?

investigate the question of time

Let us for the

a universal time scale for

In order to find out

we must

all

realistically

measurement.

moment concern

ourselves with the problem of defining


a time scale in a single frame of reference. Now the basic process involved
in any time measurement is a measurement of simultaneity. For instance,

common

statement of the type "The train arrived at 7 o'clock,"


an abbreviated notation for what really occurred "The train
arrived and the little hand of a nearby clock pointed to 7 simultaneously."
No further discussion of the question of the simultaneity of events which
occur at essentially the same physical location is necessary. Not so
obvious is the problem of determining the simultaneity of events which
occur at different locations, but this, in fact, is the key problem involved
in the setting up of a time scale for a frame of reference. In order to have
the very
is

actually

SIMULTANEITY

Sec. 5]

17

a time scale valid for a whole frame of reference we must have a number
of clocks distributed throughout the frame so that there will everywhere
be a nearby clock which can be used to measure time in its vicinity. These
clocks must be synchronized

of the clocks

"The

that

is,

we must be

hand of clock

little

two
hand of clock B

able to say of any

and the

little

pointed to 7 simultaneously."

number of methods for


They all involve

the determination of simultaneity suggest

themselves.

two
means of transwould be no problem.

the transmission of signals between

physically separated locations. If we

had

at our disposal a

mitting information with infinite velocity, there


y

s Light

-*J

signals >.

-*-

-L>.

^-Midpoint

Illustrating Einstein's definition of simultaneity.

Figure 1-5.

is the point where the Galilean transformation goes wrong by tacitly


assuming the existence of such a method of synchronization. In fact, no
such method exists. Since we have agreed to be realistic in developing a

This

use a real process to transmit the information.


Electromagnetic waves, such as light signals, seem to be most appropriate

time scale,
for

we must

two reasons

first,

propagation velocity e is the highest known,


second, and more fundamental, since we have

their

a clearly desirable feature

postulated that the propagation velocity of light in any frame of reference


will

be

c, it

our time

Two

seems expedient to use the velocity of

scale.

Thus we

instants of time

light itself in setting

up

are lead to Einstein's definition of simultaneity:


x

and

2,

observed at two points x x and x z

in

particular frame, are simultaneous if light signals simultaneously emitted


from the geometrically measured midpoint between xx and x2 arrive at

x x at

This

is

and

at x % at

t2 .

illustrated in figure (1-5).

An equivalent definition is that

and t2 are simultaneous if light signals


x
emitted at tx from x 1 and at t 2 from 2 arrive at the midpoint simultaneously.
These definitions intimately mix the times tlt t2 and the spatial coordinates
xx x2 In Einstein's theory simultaneity does not have an absolute meaning,
,

?x

independent of the spatial coordinates, as

it

does in the classical theory.

THE THEORY OF RELATIVITY

18

[Ch.

consequence of these definitions is that two events which are simulwhen viewed from one frame of reference are in general not
simultaneous when viewed from a second frame which is moving relative
to the first. This can be seen by considering a simple example. Figure
taneous

from the point of view

(1-6) illustrates the following sequence of events

O who is at rest relative to the ground. This observer has so


charges of dynamite at Q and C2 that the distances OC[ and

of an observer
placed two

OC2 are equal. He can cause them to explode simultaneously by sending


simultaneous light signals to Cx and C2 which actuate detonators. Assume
he does this so that, in his frame of reference, the explosions are simultaneous with the instant that he

n3

is

O'

abreast of O', an observer stationed

Co

c
Figure 1-6.

Two

successive views of a train moving with uniform velocity.

arrows indicate the

on a

c.

The

small

light signals.

which is moving by with uniform velocity v. The explosions


marks C[ and % on the side of the train. After the experiment O'

train

leave

can measure distances 0'C[ and 0'C%. He will find they are equal.f The
explosions also produce flashes of light. Both O and O' record the times
at which they receive the flashes. Observer O will, of course, receive the
flashes simultaneously, confirming that in his frame of reference the
explosions occurred simultaneously. However, O' will receive the flash
which originated at C% before he receives the flash from C[ because the

moves during the finite time required for the light to reach him.
Since the explosions occurred at points equidistant from O' but the light
train

were not received simultaneously, he concludes that the explosions


were not simultaneous.
signals

Such disagreements concerning simultaneity lead to interesting

results.

From the point of view of O, C^C2 = C\C^. However, from the point of
view o0\ c; passed C2 before C[ passed Q. Therefore O' must conclude
that
t

QC <

We

C^C^.

shall see

he must find 0'C{

The

simultaneity problem will also bring the two

n t he ne xt paragraph

that, according to O',

0'C'% due to the homogeneity of space.

0'C[

OCl. However

Sec. 6]

KINEMATICAL EFFECTS OF RELATIVITY

19

observers into disagreement concerning the rates of


clocks in their reAs we shall see, the nature of their disagreements about the measurement of distances and time intervals
is such as to
allow both O and O' to find a value of c for the velocity
of the light pulse
which came from x or C2
spective frames of reference.

6.

Kinematical Effects of Relativity

Einstein's two postulates, plus his definition, uniquely


determine the
equations which must be used to transform the coordinates
and time from
B'-r

O'

Figure 1-7.

Two

measuring rods

in

uniform translation.

one frame of reference to another.

There are several ways of actually


carrying out the derivation required to find these equations.
We shall
use here a procedure which is particularly instructive
because of

its

emphasis on physical concepts.

Comparison of Lengths Oriented Perpendicular to the Direction of


Motion

(I)

O and O' so stationed in reference frames that


O with velocity v. As shown in figure (1-7), O
at the center of parallel measuring rods AB and

Consider two observers


O'

is

moving

relative to

and O' are each located


A'B' which, by superposition, are
rest with respect to each other.

known

to be of identical length when at


pose the question: What are the
relative lengths of the two measuring rods, from the point
of view of either
observer, when they are moving with respect to each other?

As

indicated in the figure,

rods coincide

when

we assume

passing. Let

near the ends of AB,


defined by AB, and

We

that the centers of the measuring

O have two helpers Oa and Ob

stationed

who send light signals when A' and B' cross the line
who also mark the location of the points at which

THE THEORY OF RELATIVITY

20

A' and B' cross the


with the length AB.

line.

It is

Afterwards

[Ch.

can compare the length A'B'

apparent from the symmetry of the arrangement

from Oa and Ob simultaneously. Furthermore, so will O'. Therefore O' must agree that Oa and B made their
measurements simultaneously and, consequently, must accept the results
that
of the measurement. Let us tentatively assume that O and O' conclude

that

will receive the signals

TW < AB.

Next repeat the experiment, but now requiring that the length
comparisons be carried out in the primed frame of reference. Since
th e two
Einstein's first postulate requires complete symmetry between

<

A'B'.
apparent that O and O' must then conclude that AB
consistent
possible
only
The
statement.
previous
This contradicts the
conclusion is that AB = A'B'. Thus we find that the lengths of identical
measuring rods, as seen by any observer, are the same independent of

frames,

it is

whether they are at

rest relative to the observer or

moving

in a direction

perpendicular to the direction of their orientation.

(II)

Comparison of Time
Relative Motion

An observer
compare

O',

Intervals

Measured by Clocks Which Are

in

moving with velocity v relative to observer O, wishes to


measured by his clock with a measurement of the

a time interval

Mirror

Mirror

K~
^<g)
Figure 1-8.

sets of clocks in uniform translation.

made with clocks belonging to O.


have established that, when at rest with respect

same time
they

Two

interval

We

assume that

to each other,

all

is
the clocks involved run at the same rate and are synchronized. It
reading
with
the
compared
be
apparent that the reading of an O' clock can
of an O clock that happens to be momentarily coincident with the former

without any complication, even though they are moving relative to each
clocks in the two
other. Thus measurements of a time interval made with
in figure (1-8).
illustrated
frames can be compared by the procedure

KINEMATICAL EFFECTS OF RELATIVITY

Sec. 6]

21

Observer O' sends a light signal to a mirror which reflects


it back to him.
Both O and 0' record the time of transmission of the signal with
clocks
and C", respectively. The time of reception of the signal is recorded
by
the two observers with clocks C and C". The diagram
on the left shows
2

the sequence of events

from the point of view of O', and the diagram on the


from the point of view of O.
The elapsed time between the two events measured by O' is 2At', where
At' = l'/c, and where /' is the distance to the mirror
as measured in his
frame. The elapsed time measured by O is 2At. From the diagram
it is
right

apparent that
c2

where

/ is

Af2

v 2 At 2

the distance to the mirror as measured

by O. Solving for

At,

we have
At2

_
=

c*-v 2

shows that

At

We

Vl

/'.

=L
c

v /c

At

But argument

v jc

Therefore
1

Vl

v /c

= _^1_
Vl - v 2jc

(1 _7)

see that a time interval

between two events occurring at the same


some frame of reference is longer by a factor l/Vl v 2/c 2
when viewed from a frame moving relative to the first frame and, consequently, in which the two events are spatially separated. The time
place in

interval

measured

place

called a proper time.

(Ill)

is

in the

frame in which the events occurred in the same

The

effect

Comparison of Lengths Oriented


Motion

involved

Parallel to

is

called time dilation.

the Direction of

Consider the experiment we have just described and imagine a measuring


rod placed in the O frame with one end at clock C and the other end at
x

clock

C2

Designate by

which it is at rest.
from the O' frame.
to

L the length of the rod in the O frame, with respect

We want to evaluate L', the length of the rod as seen

In this frame the rod is moving in a direction parallel to its own length.
Since the velocity of O' with respect to O is v, the velocity of O (and also
of the rod) with respect to O' must be precisely
v; otherwise there would

THE THEORY OF RELATIVITY

22

[Ch.

be an inherent asymmetry between the two frames which is not allowed by


Einstein's first postulate. Let t[ be the time at which observer O' sees the
front end of the rod pass, and 4 be the time of passage of the rear end.
Then the elapsed time 2At' = t' - t[ is related to the length of the rod as
2

measured in the O' frame, and to

velocity

its

measured

in that frame,

by

the equation!

L'

2Af

(1-8)

We must also establish an equation connecting the corresponding quantities


as seen

from the

v, travels

O frame. In this frame, C, which is moving with velocity


L in a time 2At. Thus

the distance

L =
From

the last two equations

(1-9)

2At

we obtain
L'

LAt'

At

Argument

II

shows that
At' I At

= Vl -

c/c

Therefore
L'

= LVl -

v*lc

(1-10)

We see that a measuring rod is shorter by a factor Vl


which

v 2 /c 2

compared

observed from a
length. The length of

when

its length in a frame


frame in which it is moving parallel to its own
the rod measured in the frame in which it is at rest is called the proper
The effect is called the Lorentz contraction because equation
length.
like the equation proposed by Lorentz (cf. section 3). In
looks
(1-10)

to

in

it is

at rest,

it is

a great distinction between these equations. In Lorentz's


proposals represented the velocity of a material body with respect to the

fact, there is

ether frame.

In the present equation v represents the relative velocity

between the two arbitrary inertial frames O and O


It is worth while to point out that the technique used by the observer to
measure the length of a moving rod in the experiment just described is
different from that used by the observer in the experiment of argument I.
The latter observer simultaneously marked the locations of the ends of the
rod and then measured the distance between the two marks. However,
'.

very easy to prove that the two techniques are equivalent in that, if
both can be applied to the same situation, they will give the same answer.

it is

must, of course, be
t Velocities measured in any given frame
distances measured in the frame and times measured in the frame.

denned

in terms of

Sec. 6]

(IV)

KINEMATICAL EFFECTS OF RELATIVITY

Synchronization of Clocks

Which Are

in

23

Relative Motion

In section 5 we saw that observers in motion


with respect to each other
disagree concerning the simultaneity of two
events. This will lead to a
disagreement between the two observers concerning
the synchronization
of clocks. Our task here is to determine the
quantitative nature of this
disagreement.

Consider again the experiment discussed in arguments


II and III with
between the two events which were
timed by the clocks in both frames. We called the time
interval measured in
the O frame 2&t, and the time interval measured
in the O' frame 2Af ',
and we already know one relation between these two
quantities. From
equation (1-7) we have
special reference to the time lapse

Vi -

Since this equation expresses the estimate which


would be made in the O
frame of the time interval which was measured in the
O' frame, and since
the measurement was made in the O' frame with
a single clock, no problem
of synchronization arose in its derivation. However,
if we wish to write
an equation giving the estimate which would be made in
the O' frame of the
time interval measured in the O frame, we must
account for the fact that
observer O' would contend that the two clocks in
the
frame were
unsynchronized. But, after taking this into account, there
must be the
same dilation of a time interval as it is transformed from
O to O' as there

was when

it

was transformed from O' to O. Thus we put

2At

0A ,,

VI

tr/c

where 8

is a quantity to be determined which


corrects for synchronization.
This quantity can be determined by requiring that
equations (1-8),
(1-9), and (1-10) be satisfied. Substitute (1-8) and (1-9) into
(1-11). We
then have

From

L'

Ljv

Vl -

3
2

/c

(1-10) this becomes

LVl -

ytjc 2

L/v

(.-4)-i +
V\

CI

+
i

THE THEORY OF RELATIVITY

24

So the value of 8

is

[Ch.

given by

Lv^_ _Lv
V c

The minus

sign

ahead of the

means that

first

clock

(1-12)

to observer O' the second clock

C2

appears

Cv

We see again that two clocks fixed in a certain frame with a separation
L, and synchronized in that frame, appear out of synchronization when
viewed from a frame moving relative to the clocks.

It is interesting

to note

from synchronization is proportional to the separation


of the two clocks. Even more interesting is that the sign of the effect
changes as the sign of v changes, since v enters linearly. Thus two observers
that the departure

in different frames investigating the

synchronization of each other's

Not only would each observer


contend that the other observer's clocks were unsynchronized, but the
two observers would even disagree on the sign of the effect. This should
be compared with the effects described by equations (1-7) and (1-10),
in which v enters as a squared quantity. Although in these cases there is

clocks

would be

in total disagreement.

also a disagreement

between the two observers, there

Each observer believes


measuring rod contracts and time scale dilates.

in their disagreement.

7.

is

a certain symmetry

that the other observer's

The Lorentz Transformation

We

are

now

only a short step from our goal of deriving the coordinate

transformation of the theory of relativity.

Consider two observers

and O' who view the same event while moving with

velocity v with respect

to each other. Relative to their respective coordinate systems, the location

and time of the event are specified by the set of numbers (x, y, z, t) or
(x', y', z', t').
The desired coordinate transformation consists of the
equations which allow us to calculate the primed set of numbers in terms
of the unprimed set, or vice versa.
Assume that the orientation of the two coordinate systems, and of the
vector v specifying the velocity of O' with respect to O, is as shown in
figure (1-9). We define the zeros of the t and the t' scales such that clocks
at the origins of the two coordinate systems read / = t' =
when the
origins are coincident.

The separation of the two origins at some later time is vt, as seen by O
by O'). However, the distance x' measured in the O'

(or vt' as seen

frame appears contracted by the factor

Thus

would say that the

Vl

tP/c 2

distance, parallel to v,

when viewed by O.

from

his yz plane to the

Sec. 7]

THE LORENTZ TRANSFORMATION

position of the event was vt + x'Vl - v 2 \c 2


x coordinate of the event. Consequently

x
Solving for

x',

we

vt

x'Vl

25

This

is,

by

definition, the

v 2/c 2

find

Vl - v 2lc2

x
vt
^=^^=^=^

Now O believes that the time


at the point x', y',

z',

t' read on the clock fixed in the O' frame,


must be corrected by an amount 6' = +x'v/c 2

in order to account for the lack of synchronization

axis

between that clock and

y' axis

(x,y,z,t)
(x',y',z',t')

vt (as seen by

O)

->~x axis

Figure 1-9.

Illustrating the

*-*' axis

argument leading to the Lorentz transformation.

the clock fixed at the origin of the O' frame. (It would
be instructive for
the reader to convince himself that in the present
case the quantity d'
positive.)

Furthermore, time intervals measured by the clock at the


O' origin appear to O to be dilated by the factor 1/Vl
v*/c 2
Thus,
according to O, the time t of the occurrence of the event is
is

+ x'v[c 2
Vl - v 2jc
t'

Substituting for x

from the equation above and solving for

xv/c
Vl - v jc

of section 6

tells

we

find

Argument

t',

us that y'

y and

z'

z,

since these

THE THEORY OF RELATIVITY

26
distances

are measured in directions perpendicular

together our results,

to

v.

[Ch.

Gathering

we have

vt
Vi - v /c
X

y'
z'

=
=

(1-13)

z
t

xv/c

Vl -

v jc

These equations constitute the so-called Lorentz transformation which,


according to Einstein, must be used for the transformation of the coordinates and time of an event from the inertial frame O into the inertial
frame O'.f

An algebraic solution of equations (1-13) for x, y, z, and t produces


another form of the Lorentz transformation which is useful in transforming from the primed frame to the unprimed frame. These are

+ vt'
Vi- v /c
x'

y'

(1-13')
z'

+ x'v/c
Vl - v /c
f

Equations (1-13') are identical with (1-13) except that v has been replaced
v. The minus sign is merely a consequence of the fact that, for (1-13'),
still
v
means the velocity of O' with respect to O. If we were to redefine

by

v to

mean

the velocity of

with respect to O', the resulting equations

would be completely identical with (1-13). We could have guessed the


form of (1-13') from the form of (1-13) without actually doing the algebra,
since the observed symmetry between the two is required by Einstein's
first postulate. The postulate demands that all inertial reference frames in
uniform translation be equivalent. Consequently the transformation
which we have derived from this postulate must be the same in going from
O to O' as it is in going from O' to O.
It is' immediately apparent that the Lorentz transformation reduces to
the Galilean transformation

when

vie

<

This makes good physical

t Equations of exactly this form, but with v representing a velocity relative to the
ether,

some

had been proposed in connection with a


work of Einstein.

years before the

classical theory of electrons

by Lorentz

TRANSFORMATION OF VELOCITY

Sec. 8]

27

The fundamental difference between the two transformations is


we realize that the signals, which are needed to synchro-

sense.

that in the former

nize the clocks of a reference frame, propagate


with the finite velocity c,
while in the latter we naively assume the velocity
of these signals to be
infinite. However, if we are required to
make a coordinate transformation
in a situation in which the relative velocity
v is very small compared to the
velocity of the synchronizing signals, it should
not make any difference
whether we take the latter velocity to be c or oo.
Consequently in this
limit the two transformations should merge.

On

the other hand, significant discrepancies arise


between the preof the Galilean transformation and the rigorously
correct

dictions

Lorentz transformation when v

had not been observed

is

in classical

comparable to c. These discrepancies


mechanics because experiments had not

been performed at such high

velocities. It is interesting to note that for


equations (1-13) are meaningless, in that real coordinates
and
times are transformed into imaginary ones. Thus
c appears to play the
role of a limiting velocity for all physical
phenomena.

v/c>

We

shall attain

a better understanding of this as we continue our


discussion.
To close this section, let us check the internal consistency of the Lorentz
transformation with the postulates from which it was
derived. Consider
again figure (1-9), and imagine that at the instant =
=
t

a spark gap located at the

As viewed by

common

origin of the

we

t'

discharge

two frames of reference.

observer O, the resulting light pulse will travel out


in

directions with the velocity

c.

Thus, at any instant

form a spherical wave front of radius r

t,

The wave

ct.

all

the light pulse will


front will

obey the

equation
**

z2

or

x*

/2

a2

c2 ? 2

(1

_ 14)

From the view point of O', according to Einstein's postulates, the


light
pulse must also travel in all directions with the
velocity c. Thus O' must
also see a spherical wave front
*'

y*

+ z' - cY =
2

(1_15)

despite the fact that he


it

as

is moving with velocity v relative


to O. We leave
an exercise for the reader to apply equations (1-13')
to (1-14) and

obtain (1-15), thus giving the desired demonstration


of consistency.

8.

Transformation of Velocity

Consider the particle shown in figure (1-10), moving


with velocity
as seen in a frame of reference O.
would like to evaluate the velocity

We

THE THEORY OF RELATIVITY

28

[Ch.

of the particle, as seen in the frame of reference O' which

moving relative to O with velocity v.


Measured in the O frame, the velocity vector of the

is itself

particle has

com-

ponents

Vx =
The same

Vy =

dxjdt,

dzjdt

measured in the O' frame, has components

velocity vector, as

V'x

V =

dy/dt,

V'y

dx'jdt',

dy'jdt',

V'

dz'/dt'

/
Particle

Figure 1-10.

moving

observed from two frames of reference

particle

in

uniform

translation.

Now from equations

we know that the

(1-13)

and the unprimed coordinates and times

relation between the

primed

is

(x

vt)

V1-/3 2

=
=

y'
z'

y
z
1

vx\

(^

where we have introduced the convenient notation ^ = v/c. Take the


differential of these equations, remembering that v is a constant. This gives
dx'

(dx

Vl-
dy'
dz'
i

dt'

=
=

v dt)

dy
dz
1

=
Vi

j
A/ dt

2V
i

dx\
1

TRANSFORMATION OF VELOCITY

Sec. 8]

29

Then
1

/j

(ax

V __= Vl -

j.\

v at)

dx
v

K.-

dr

ft

J=(dt - fl\
Vi-^
1

-M i_^ i-v

dy

n=
Vl

dt

<fy/rff

/?\

V!

flt

K,Vl

rfr/

ft

"

(1-16)
and, similarly,
dz

V',=

dz/dt

Vl-/S 2 ^

Vl-/32 V

__

vji -

1--.V.
c* dtl

These equations tell us how to transform the observed velocity from one
frame of reference to another frame of reference.

0.9

<

0.9 c
*-

Figure l-l

Illustrating an

example of the addition of

velocities.

First we note that, as Vjc and v\c approach zero,


equations (1-16)
approach those which would be derived from the Galilean transformation.
Another very interesting property of these equations is that it is impossible
to choose V and v such that V, the magnitude of the
velocity which
is

seen in the

new

in figure (1-1

of positive

To

x,

1).

frame,

As

is

greater than

c.

seen by O, particle

Consider the example illustrated


has velocity 0.9c in the direction

and

particle 2 has velocity 0.9c in the negative x direction.


evaluate the velocity of particle 1 with respect to particle 2
we transform

from the
velocity v

O frame

to the O' frame moving in the negative x direction


with
-0.9c, using the first of equations (1-16). We obtain

V{

0.9c

(-0.9c)

1.80c

<

THE THEORY OF RELATIVITY

30

not

It is

[Ch.

a rigorous proof of this property by considering


form of equations (1-16). The velocity transformation

difficult to give

the mathematical

equations demonstrate another aspect of the fact that in the theory of


a limiting velocity for all physical phenomena.

relativity c acts as

9.

Relativistic

Mechanics

It has been emphasized that Einstein's modification of the transformation


equations would, in turn, necessitate a compensating modification in the
laws of mechanics, so that these laws continue to satisfy the requirement

of not changing form under the influence of a coordinate transformation^


In this section we develop the laws of the new mechanics, which is called
relativistic

mechanics.

obviously desirable to carry over into relativistic mechanics as


of classical mechanics as possible. We shall see that it is possible to

It is

much

preserve the classical definition of

where p

and

is

the

momentum,

momentum,

= mu

(1-17)

and u the velocity of a particle,


law of conservation of momentum,

the mass,

also to preserve the classical

2*

Ip, _ initial

(1-18)
final

we are willing to modify the classical concept of mass. It will


be necessary to allow the mass of a particle to be a function of the magnitude of its velocity, i.e.,
providing

m=

m(u)

(1-19)'

The form of this function is to be determined. However, we know a priori


that we must have m(u) -> m as u -- 0, where m is a constant equal to
,

the classically measured mass of the particle. This is true since, as u -> 0,
the Lorentz transformation approaches the Galilean transformation and

no modification

in

mechanics

is

necessary.

In order to evaluate the function m(u),

As seen from the xyz frame


and 2 are moving in directions

ment.
(?!

we

consider the following experi-

indicated in figure (1-12), observers

parallel to the x axis with equal and


These observers have identical balls Br and B2 each
as measured when they are at rest. While passing, each observer

opposite velocities.

of mass

t Just the opposite is true for the laws governing electromagnetic


well's equations), since Einstein's

phenomena (Max-

second postulate prohibits modification of these laws.

RELATIVISTIC

Sec. 9]

MECHANICS

throws his ball toward the other with a velocity which, from
his own point
of view, is directed perpendicular to the x axis and is of
magnitude V.
In the xyz frame, Bx and B will be seen to approach with equal
2
velocity
along parallel lines making angles 6 =
with
x
the
axis,
and
rebound
X
2<
.

at angles 6

and

Assuming conservation of momentum and that the


collision is elastic, it is easy to show that 8
= 2/ and that the magnitude
lf
of the velocity of the balls is the same after the collision
as before. The
actual value of
or 2/ depends on the impact parameter d, which we
lf
assume to be such that 6 =
as shown in the figure.
X
Xf
lf

2/

(Out of paper)

Figure 1-12.

collision

between two

balls of identical rest

mass.

Now look at the process from the point of view of Ox as shown in figure
Observer Ox throws Bx along a line parallel to his y axis with
,

(1-13).

velocity

V.

It

returns along the

same

with velocity of the same


maintain a constant x velocity
relative velocity of
2 with respect to O x
The component of velocity of B2 along his y axis is observed by O to
x
change sign during the collision but to maintain a constant magnitude.

magnitude but opposite sign.


which is just equal to v, the

To

He

sees

line

B%

we realize that the y component of the velocity


Then we transform this to the Ox frame with the

evaluate this magnitude

of B2 as seen by
,

2 , is V.

aid of the second of equations (1-16) and obtain vVl


component of the velocity of B2 as seen by O x
The y momenta of both Bx and B2 as measured in the

v 2jc 2 for the

Ox

frame, simply

change sign during the collision. Consequently the total y momentum of


the system consisting of the two balls changes sign. But the law
of
conservation of momentum says that the total y momentum of the system
before the collision must equal the total y momentum of the system after
the collision. This can only be true if the total
y momentum, as seen by

THE THEORY OF RELATIVITY

32

x , is

zero both before and after the collision.

Evaluating the

[Ch.

momenta

according to the definition (1-17), and equating the total y momentum to


zero in order to satisfy the conservation of momentum (1-18), we obtain

an equation which

is

constant equal to

obviously self-contradictory
.

However,

if

we

allow

if

we

insist that

be a

to be a function of the

velocity as in equation (1-19), the equation reads

Vm(V)= VVl -

v*lc

m(V V\\ -

v 2 jc 2 )

v2)

or

m{V)

= Vl -

iP/c 2

m(V V\l -

where the argument of the function

Figure 1-13.

Acollison between two

balls

on the

v^/c

2
)

v2)

right side of the equations

of identical rest mass, as seen by observer

t.

simply the magnitude of the velocity of B2 as seen by O v This equation


perfectly satisfactory and, in fact, enables us to evaluate the function m
immediately if we take V-*-0. We then have
is
is

m(0)

= m = Vl -

m(v)

=w

2
t>

/c

m(v)

so

/Vl

v 2 jc 2

(1-20)

[Evaluation of the functional form (1-20) from the equation for m(V),
by taking V-*-0, constitutes a short-cut but not an approximation. It
is easy to show that (1-20) satisfies the equation for m(V) exactly.] Thus
a consistent theory of relativistic mechanics demands that the mass of
a particle moving with velocity v be larger than its mass when at rest by a
factor

v /c The mass m(v) is called the relativistic mass of the


mQ is called its rest mass.
quite high velocity v = 0.1c, the relativistic mass is only 0.5

1/Vl
and

particle,

For the

percent greater than the rest mass.

However, with increasing

v the rest

Sec. 9]

MECHANICS

RELATIVISTIC

33

mass rapidly becomes larger since m(v) ->ooasu^e. It is apparent


that the velocity of a particle cannot exceed c if (1-20) is valid.
Continuing our discussion of relativistic mechanics, let us consider a
particle of rest

magnitude

mass

but accelerated by a force of

initially at rest,

F which is applied in

the particle in the distance x


f

the x direction.

Jo

In order to evaluate this

We

assume that

total

work done on

tf

T = )'f dx = P'f dt =

law.

The

is

Jo

dt

we must know

Fv

dt

Jo

the relativistic

form of Newton's

it is

rF

dP
_
=
= d (mv)J

(1-21)

dt

dt

Note that, if m = constant, this is just F = m dvjdt = ma. Actually,


Newton's statement of the law was that of equation (1-21) and
not
F - ma. Inserting (1-21) into the equation for T, we have

T=

Jo

Integrating

by

parts,

we

dt

dt

vdp
Jo

obtain

T=

H? - ['p

dv

Jo

Substituting for
v dv as d(v*)j2,

p from

equations (1-17) and (1-20), and also expressing

we have

T =

Vf
mv
_
2 z
Lvi
v /c J

mo |V
2 Jo

The second term

is

a standard form.

T= m

v /c

U/l

d(v

2
)

Vl -

Integrating

it,

v /c

we have

+ Vl -

v /c

v /c

This reduces to

T= m

c''

-Vl

v /c

Evaluating at the limits and then dropping the subscript /to simplify
our
notation,

we have

T=
Vl

^2
c"
2

v /c 2

m nc'

(1-22)

THE THEORY OF RELATIVITY

34

[Ch.

Now the theorem of conservation of energy implies that the total work
done on the particle should equal its kinetic energy. Thus we would like
T

to call

To check

the kinetic energy of the particle.

this, let v/c

->

0.

Then

T = m

>+-\

V*

= -- m v*
m oCc -
2 v

*
T-^ z
,

x
1

This agrees with the classical expression for kinetic energy and confirms
our identification of T.

Continuing with our interpretation of equation (1-22), we observe that


a function of v which can be written as the difference between a term
depending on v and a constant term, as follows

T is

T(v)

where E(v)
E(v) for v

must

=w
= 0,

i.e.,

also be energies

when
when

its

its

velocity

v jc 2

E(0)

=m

/V

E(v)

= mc

2
,

(0)

and

also

Since

is

where E(0) is the value of


an energy, E(v) and E(0)

E(v) being some energy associated with the particle


and (0) some energy associated with the particle
zero. To identify the meaning of these energies,

is v,

velocity

is

rewrite the equation as

E(v)

The conclusion

is

inescapable.

T(v)

E(0)

We must interpret

of the particle moving with velocity

E(v) as the total energy

the sum of the kinetic


energy T(v) of the particle and an intrinsic energy (0) associated with the
particle when it is at rest. The energy E(v) is called the total relativistic
energy,

and E(0)

is

called the rest

v,

since

it is

mass energy.

We

have established Einstein's well-known relation between mass and


energy: The rest mass energy of a particle is equal to c 2 times its rest mass.
That is,
(0)

and the
That is,

=m

c2

(1-23)

total relativistic energy is equal to c 2 times its relativistic mass.

E(v)

Equation (1-22)
kinetic energy,

tells

and

= mc

(1-24)

us that the relation between total relativistic energy,

rest

mass energy

mc 2

is

= T+ m

c2

(1-22')

RELATIVISTIC

Sec. 9]

MECHANICS

35

For situations in which the particle also has a potential energy V,


possible to extend the argument to show that

mc 2

= T + V+ m

c2

This emphasizes the fact that the relativistic mass

(1-22")
is

a direct measure of the

total energy content of the particle.

content will increase


It is

its relativistic

it is

Anything that increases


mass.

its

energy

sometimes convenient to have a relation similar to (1-22') which


momentum p. It can be obtained by evaluating

explicitly involves the

the quantity

m 2c4 where

/5

vjc.

This

mfc*

is

m v - m y = my

= my

^
1

/r

1
.1

1
2

^1
mW = cY

/3

Thus

=w

2 4

= c/+ m
2

2 4

(1-25)

The relativistic properties of mass and energy, and the relation between
the two, were derived from the Lorentz transformation and the additional
assumptions we made concerning the new form of the laws of mechanics.
These additional assumptions seemed reasonable since they were obviously
designed to preserve as

much

of classical mechanics as possible. However,


can be found only by comparing the predictions
of relativistic mechanics with experiment. We reserve this
comparison for
section 1 1, but perhaps it is appropriate to point out here that
the existence
of a rest mass energy m c 2 is not in conflict with classical physics.
Since the
experiments of classical physics are all such that the total'rest mass involved
their ultimate justification

constant, the appropriate rest

mass energies could be added to both sides


energy balance equations without destroying their validity.
The relativistic theory of energy is, however, of more than academic
interest. One important reason is that there are
processes in nature in
is

of

all

which the total rest mass of an isolated system does not remain constant.
For such processes experimental evidence shows that the change in rest
mass energy is exactly compensated by a change in kinetic, or potential,
energy in a
This

is

way

that conserves the total relativistic energy of the system.


some very practical processes for the conversion of rest

the basis of

mass energy into other forms of energy, as in a nuclear reactor. It


also
shows that in our relativistic theory we must replace the separate classical
laws of conservation of mass and conservation of energy
by a single
comprehensive law of conservation of total relativistic energy:

36

THE THEORY OF RELATIVITY

As observed from a given frame of reference,^


of an isolated system remains constant.

10.

Transformation of

Momentum

[Ch.

the total relativistic energy

and Energy

In certain problems, particularly those involving collision mechanics,


convenient to transform the momentum and total relativistic energy of

it is

a particle from one frame of reference to another. In this section


the appropriate transformation.

Consider a particle of

rest

mass

positive x axis of reference frame O.

of

moving with

In that frame

it

we derive

V along the
have components

velocity
will

momentum
V,

and

total relativistic

Pv

o,

Pz

energy

Wgc 2

E=

Vl - V 2/c 2
Next consider an observer O' who is also moving along the positive x axis
of O, but with velocity v. According to this observer, the particle would
have components of momentum

p'

and

m v

'

p'y

Vi -

F'V

p'z

0,

total relativistic energy

'

where

To

is

we

m Qc 2
Vl - F'V

the velocity of the particle with respect to O'.

establish the relation

frames,

between

momentum and

energy in the two

evaluate the quantities

V'/Vl

Vic*

and

c 2 /Vl

V'

jc

The same restriction applies, of course, to the energy conservation law of classical
For example, the total energy content of a particle will increase if we transform
classically from a system in which the particle is at rest into a system in which it is
moving and consequently has kinetic energy. See section 10.
t

physics.

Sec. II]

EXPERIMENTAL VERIFICATION OF THE THEORY

37

in terms of V. This is
(1-16),

done by using the velocity transformation equation


which in the present case reads

V~

V'=
1

Vvjc 2

The

calculation will not be reproduced here.


forward but lengthy. The result is

'

vE/c
Vi - v

Px

Pz

= Py
= Pz

E'

- E~

Py

It is perfectly straight-

Vl

(1-26)

PxV
2

v /c

Although in the derivation we assumed the velocity of the particle to be


parallel to the x axis, it can be shown that equations (1-26) are valid for
any orientation of its velocity vector.

II.

Experimental Verification of the Theory

The theory of

was designed to agree with the experimental


same in frames of
reference which are in uniform translation with respect to each other.
However, in addition to achieving this, the theory predicts a number of
new phenomena, such as length contraction, time dilation, relativistic
increase in mass, and a relation between mass and energy. This is typical
of a scientific theory. Also typical is the fact that the initial acceptance of
the theory was only tentative even though it appeared to be based upon
relativity

fact that the velocity of light is observed to be the

correct logic.

The theory did not achieve full


new phenomena had been put

concerning the

status until

its

predictions

to the test of experiment.

To

close our discussion of the theory of relativity we shall briefly mention


some of the experiments which confirm its predictions.
The first of these was performed in 1909 by Bucherer. It consisted of a
measurement of the masses of high velocity electrons, using a fairly direct

technique that will be discussed in Chapter

3. Bucherer's results are shown


some more extensive results obtained in
recent years are shown by dots, and the predictions of equation (1-20)
are shown by the solid line. Note that these results prove not only that

by the

crosses in figure (1-14),

equation (1-20) has the correct functional form, but also that the velocity

THE THEORY OF RELATIVITY

38
c,

which

essentially enters the theory

of

3.00

is

equal to the velocity of

10 10 cm/sec.

length contraction

relativity as the limiting velocity

for the transmission of information, actually


light,

[Ch.

number of experimental verifications of


and time dilation are now known. One of the clearest

of these involves measuring the lifetime of unstable particles called mesons,


mesons of various velocities. It is observed that the lifetime of a

for

moving meson

rapidly

dilated in excellent agreement with equation

is

(1-7).

These experiments

chapter

we

"
.,

1.8

1.7

1.6
1.5

be mentioned in Chapter 16. In that


an overwhelming amount of evidence proving

will

shall also present

with c

3.00

x 10

cm/sec

-g-1.4
1.3

1.2
1.1

1.0
0.3

0.4

0.5

0.6
v/c

Figure 1-14.
I.

An experimental

verification of the

0.7

0.8

0.9

dependence of mass on

velocity.

From

Kaplan, Nuclear Physics, 1955, Addison Wesley, Reading, Mass.

the validity of equations (1-23) and (1-24), Einstein's relations between


mass and energy. The predictions of the theory of relativity have been
confirmed on every point, and there is now universal agreement on its
validity.

subject that might logically be discussed at this point

is

the so-called

general theory of relativity, which considers frames of reference that are


accelerating with respect to each other. It is found that there are very
relations between acceleration and gravitational effects.
However, the subject is such that any pretense at a serious treatment
must be given in terms of tensor calculus, and therefore we shall pass it
over. This will not be a handicap because the general theory has had very
little interaction with the other branches of modern physics, and also
because it is possible, by using various subterfuges, to employ the special
theory that we have developed in certain problems involving accelerated
motion. For instance, in Chapter 1 1 we shall make such a calculation for
a case of uniform circular motion.
interesting

Ch.

EXERCISES

I]

39

BIBLIOGRAPHY
Bergmann,

wood

P. G., Introduction to the

Theory of Relativity, Prentice-Hall Enele'

Cliffs, N.J.,

1942.

Leighton, R. B., Principles of Modern Physics,


McGraw-Hill

Book

Co.,

New

York, 1959.
Lorentz, H. A., A. Einstein, H. Minkowski, and
H. Weyl, The Principle of
Relativity, Dover Publications, New York, 1951.

EXERCISES
Write equations analogous to (1-2) for the case in
which the relative
is not parallel to any of the axes,
and use these to show that
equations (1-3) still obtain.
1.

velocity vector u

2.

Derive equation (1-6').

3.

The

years.

distance to the farthest star in our galaxy

Explain

why

it is

this star within his lifetime,


4.

Apply equations

5.

In frame O, particle

velocity

v.

is

possible, in principle, for a

and estimate the required

(1-13') to (1-14)
1

is

at rest

and obtain

and

of the order of 10 5 light


being to travel to

human

velocity.

(1-15).

moving to the right with


moving to the right
the two particles appear in O'

particle 2 is

Now consider a frame O' which,

relative to O, is

with velocity u. Find the value of u such that


to be approaching each other with equal and
opposite velocities.
6.

Verify that the form (1-20) for m{v) satisfies exactly


the equation for miV).

7.

Carry through the calculation leading to (1-26).

In the laboratory (LAB) frame, particle 1 is at rest


with total relativistic
energy Ex and particle 2 is moving to the right with
total relativistic energy
8.

and momentum p 2

Show

masses of the system

is

that the frame in which the center of the


relativistic
at rest is moving to the right with velocity

c-

+E

LAB frame, and show that the total momentum of the system is
zero in this center of mass (CM) frame. Now let
the two particles have the same
rest mass m , and let the total kinetic energy
of the system in the LAB frame be
TLAB Evaluate Tcu the total kinetic energy of the system
relative to the

in the

and show that

rCM

CM

frame

V2m n c* V Tj

in the extreme relativistic limit where T


m ^. What might be the practical
1AB
consequences of this in the study of high energy nuclear

>

reactions ?

EXERCISES

40
9.

[Ch.

Consider a collision in the LAB frame between the moving particle 2 and
Show that
1, both of which have the same rest mass m

the stationary particle

the angle between the velocity vectors of the two particles after the collision is
always 90 in classical mechanics, and show how this result is modified by
relativistic

mechanics.

Hint: Transform to the

then transform back to the

An

LAB

CM

frame, treat the collision,

frame.

-4.8 x 10~ 10

x 10~ 28 gm), initially


held at rest at one plate of a plane parallel condenser, is released and allowed
to fall through vacuum to the other plate under the influence of the electric field
in the condenser. The separation between the plates is 102 cm, and the voltage
10.

electron (charge

across the plates

is

esu,

mass

10 7 volts (10 7 /300 stat-volts).

9.1

Calculate the time required,

in the frame of reference of the condenser, for the electron to travel between
its plates.

Discuss the justification of using the theory developed in this chapter

Hint: The magnitude of


and the magnitude of the component of the electric field
to the direction of motion, are both unchanged by a Lorentz trans-

for the present system, which involves accelerations.

the electric charge,


parallel

formation.

CHAPTER

Thermal Radiation
and the Origin of

Quantum Theory
I.

Introduction

In the course of a successful attempt at resolving certain discrepancies


between the observed energy spectrum of thermal radiation and the
predictions of the classical theory, Planck was led to the idea that a system
executing simple harmonic oscillations only can have energies which are
integral multiples of a certain finite

amount of energy

(1901).

closely

was later applied by Einstein in explaining the photoelectric


effect (1905), and by Bohr in a theory which predicted with great
accuracy
many of the complex features of atomic spectra (1913). The work of these
related idea

three physicists, plus subsequent developments

and Heisenberg

(ca.

1925), constitutes

by de Broglie, Schroedinger,
what is known as the quantum

This theory and the theory of relativity together comprise the two
of modern physics. In this chapter we begin our
study of quantum theory with a discussion of thermal radiation.
The nature of the subject is such that any discussion of thermal radiation
must be based upon the results of advanced treatments of several different
fields of classical physics: thermodynamics, statistical mechanics,
and
theory.

most

significant features

electromagnetic theory. In

and

justify

them with

many

cases

we

shall simply quote these results

arguments since their proofs are most


appropriately left to the texts which treat these subjects. However, we
shall reproduce in detail that step in which Planck introduced the
quantization of energy, and we shall be able to see exactly how its introduction
resolved a serious conflict between experiment and the classical theory.
qualitative

41

THERMAL RADIATION AND QUANTUM THEORY

42

2.

[Ch. 2

Electromagnetic Radiation by Accelerated Charges

The surface of every body at a temperature greater than absolute zero


emits energy, in the form of thermal radiation, due to the motion of electric
charges near the surface. This radiation consists of electromagnetic
waves (often called infrared waves) of exactly the same nature as visible
but of longer wavelength. As we shall be concerned on several
occasions in this book with the emission of electromagnetic radiation by a

light,

moving charge, a description of the classical theory of the process is


appropriate at this point. The description will be qualitative, but some of
the quantitative results of the theory will be presented.

In elementary electrostatics

an electric
Coulomb's law
charge q

is

shown that surrounding every

it is

field specified

by the vector E. These are

stationary
related

^(-:)

by

<->

where r is the vector from the charge to the point at which E is evaluated,
and where (r//-) is a unit vector in that direction.! It is often convenient to
represent the electric field in terms of lines offorce which are constructed
to be everywhere parallel to the local direction of E and of density (number

of

lines crossing

local value of E.

cm2

area normal to the direction of E) equal to the


two dimensional representation of the lines of force

surrounding a stationary charge is shown in figure (2-1). This


time independent) electric field contains stored energy. In

static

energy stored per unit volume,


the equation

E by

(i.e.,

p, the

is

related to the local value of

fact,

=E

/Stt

(2-2)

the field is in a medium of unity permeability such as the vacuum. This


can be verified with little difficulty by considering the case of a parallel
plate condenser in which E is everywhere constant, and equating the
energy required to charge the condenser to the energy stored in the electric
if

field.

The energy stored in the field of a stationary charge is static and is not
away in the form of electromagnetic radiation; if this were not
so, the conservation of energy would be clearly violated. It is also true
that the energy stored in the field of a charge moving with uniform velocity
is not radiated away but moves along with the charge, even though there
radiated

t With a few exceptions that are explicitly indicated, in this book we shall write all
electromagnetic equations in the cgs-Gaussian system of units. In this system charges
are expressed in esu in all equations.

RADIATION BY ACCELERATED CHARGES

Sec. 2]

43

are associated with the charge a non-static electric


field E and also,
according to Ampere's law, a non-static magnetic field H. It is
easy to
see that this must be the case by mentally transforming
to a frame of

reference in which the charge

is stationary, realizing that it cannot radiate


and then applying the relativistic requirement that the
behavior of the charge including whether or not it radiates
cannot
depend upon the frame of reference from which it is viewed. The total
energy stored in the field of a uniformly moving charge is larger
than if it
were stationary because, when there is also a magnetic field H, electromagnetic theory shows that the energy density is

in that frame,

(2-2')
o7T

for unity permeability.

by the

The

excess energy

forces initially producing the

Figure 2-1.

The

lines of force

For a charge undergoing

is supplied from the work done


motion of the charge.

surrounding a stationary charge.

acceleration,

magnetic

fields

instant

suffered a constant acceleration

the non-static electric

and

cannot adjust themselves in such a way that none of the


stored energy is radiated, as they can in the case
of a charge moving with
uniform velocity. We can see this qualitatively by considering
the behavior
of the electric field. In figure (2-2) we describe this
field by drawing some
of the lines of force surrounding a charge which was
at rest at the initial

/,

a to the

right during the interval

t to t', and then continued moving


with constant final velocity. The figure
shows the lines of force at some later instant t", as viewed
from the frame
of reference moving with the final velocity, and
with the restriction that
all times are measured in that frame.
At small distances the lines of force
are directed radially outward from the present
position of the charge.

At

large distances they are directed radially


outward from the initial
position of the charge. The reason is simply that
information concerning
the position of the charge cannot be transmitted
to distant locations with
infinite velocity, but only with the velocity
c. As a consequence, there are

THERMAL RADIATION AND QUANTUM THEORY

44

[Ch. 2

breaks in the lines of force contained between a sphere centered on the


position and of radius c(t"
t), which is the minimum distance

initial

at

which the

field

can "know" the acceleration

started,

centered on the present position and of radius c{t"

t'),

and a sphere
which is the

minimum

distance at which the field can "know" that the acceleration


As t" increases, the region containing the breaks expands outward
with velocity c. The electric field in this region has components which are

stopped.

both longitudinal and transverse to the direction of expansion, but by


constructing diagrams for several values of t" it is easy to see that the
longitudinal component dies out very rapidly, and can soon be ignored,

Figure 2-2.

The

lines

offeree surrounding an accelerated charge. Only

some of the

lines

are shown.

whereas the transverse component dies out slowly. In fact, electromagnetic


theory shows, by calculations based upon the same idea as our qualitative
discussion, that at large distances
(large t") the transverse electric field

bx

from the region of the acceleration


obeys the equation

_ qa sin 6
=
i

(2-3)

c r

= c(t" t) is the magnitude of the vector


from the region at which the acceleration a took place to the point at
which the transverse field is evaluated, and where is the angle between
r and a. The dependence of E on 6 and r can be seen in our diagrams,
x
and it is clear from our discussion that E must be proportional to q and a.
L
The electromagnetic theory also shows that because of the acceleration
for unity permeability, where r
r

RADIATION BY ACCELERATED CHARGES

Sec. 2]

there will be a transverse magnetic field

that at large distances

45

HL moving along with E

and

from the region of acceleration


TI

Hi.

^sin 8
= qa
i
2

(2-3')

c r

for unity permeability, where r


c(t"
t).
These two transverse fields
propagating outward with velocity c form the electromagnetic radiation
emitted by the accelerated charge. For this, or any other, electromagnetic

radiation propagating in a

medium of unity

EX =

HL

and the energy density of the radiation

Figure 2-3.

Illustrating

permeability,

= {El +

(2-2') is

E\)

the energy transport

in

(2-4)

a parallel stream of radiation.

However, equations (2-3) and (2-3') are valid only if the dimensions
of
the region in which the acceleration took place are small
compared to c
times the period of its duration, as we have assumed.

The

intensity / of the emitted radiation

carries per second across a

cm2

is defined as the energy which it


area normal to the direction of pro-

pagation.
in

Consider figure (2-3), which shows the radiation propagating


from the region of acceleration
forms an essentially parallel stream. The radiation is moving with

some

that

it

direction at such a large distance

velocity c cm/sec.

In one second it moves a distance c cm. Consequently


the energy contained in the parallelepiped with a 1 cm 2
base normal to
the direction of propagation, and with length /
c cm, will be carried
across the base in 1 sec. The energy contained in the
parallelepiped is
equal to the energy density p times the volume c cm3 Thus
the intensity
all

pc

(2-5)

THERMAL RADIATION AND QUANTUM THEORY

46

Using equations (2-3) and

(2-4),

I(6)

we have

E!c

[Ch. 2

= tEl^l

(2 _ 6)

where we indicate explicitly that the intensity is a function of the direction


of emission. Note that 1(6) obeys the familiar inverse square law. The
total

energy radiated in

all

directions per second, R, can be evaluated

integrating 1(6) over the area of a sphere of arbitrary radius

R=

\l(0)

= \(6)2^ sin

dA

r.

That

by
is,

d6

which

yields

*=

lV

(2-7)

for the rate of radiation of energy.

Only some of the qualitative ideas of


be used in the next few chapters; the equations will be
used toward the end of the book.
this section will

3.

Emission and Absorption of Radiation by Surfaces

Let us return now to the phenomenon of thermal radiation. As we


have said, every body at a temperature greater than absolute zero emits
In the classical theory, this can be pictured as the result of
the accelerations of electric charges near the surface due to thermal
this radiation.

agitation.

Now,

in a single acceleration lasting for a certain period

of

most of the emitted radiation has a frequency approximately equal


to the reciprocal of the period, and a wavelength equal to c times the
period. But, in the many different acceleration processes which give rise
to the thermal radiation, a whole spectrum of wavelengths is emitted.
On the basis of this picture it would be expected that the rate of emission
of energy, integrated over the entire wavelength spectrum, would increase
as the temperature of the surface increases because of the enhanced
thermal agitation, and it would also be expected that the rate of emission of
energy would be proportional to the area of the surface. This is found to
be the case; an empirical equation due to Stefan (1879) states
time,

IT

aeT*

(2-8)

The quantity IT is the total energy emitted in all frequencies per second
cm2 from a surface at absolute temperature T; e is a constant called
the emissivity which ranges from
to 1 depending on the nature of the
per

Sec. 4]

BLACK BODY RADIATION

emitting

surface;

and

47

Stefan-Boltzmann constant.

0.567 x 10" 4 erg-cm^-deg-^-sec- 1 is the


The energy emitted as thermal radiation is

supplied by the thermal agitation energy.


We shall also be concerned with the absorption of thermal radiation by
a surface. In this process, energy is removed from incident
thermal radiation,

through

its

agitation energy.

action

There

on electric charges, and ends up as thermal


an interesting relation between the efficiency

is

of a surface as an emitter of thermal radiation, as measured


by e, and its
an absorber. The latter property is measured by a constant
called the absorptivity a that is denned as the ratio
of the total thermal
energy absorbed by the surface to the total thermal energy
incident
efficiency as

upon

it.

The

(1895),

relation between e

is

and

a,

which

is

a theorem due to Kirchhoff

simply

e=

(2_9)

This was proved in a thermodynamic argument,


independent of any
detailed assumptions about the emission and
absorption processes,

by

considering the thermal equilibrium of surfaces of different


nature which
are exchanging thermal energy by emission and
absorption. It has also
been verified experimentally.

4.

Black Body Radiation

Consider a body with a surface that has the property of


absorbing all
upon it, i.e., a = 1. Such a body is called a black
body. A number of objects found in nature are quite
good approximations
to a black body. One example would be any object
coated with a diffuse
layer of black pigment; another example will be
given shortly. Since the
absorptivity of a black body is equal to unity, it follows
from Kirchhoff
the radiation incident

's

law that
absorber,

its
is

emissivity
also the

to equation (2-8),

is also.

most

we

black body, which

efficient radiator.

is

the

most

efficient

Applying the condition e

see that the radiant

power emitted per cm2 is the


the same temperature. This strongly

same for all black bodies which are at


suggests that the other properties of the thermal
radiation emitted by
black bodies, such as the spectral distribution of the
radiation, also depend
only on the temperature of the black body but not
on its detailed nature.
It

can be proved by thermodynamic arguments, involving


two black bodies

in equilibrium, that this is indeed the case. It is


evident that the universal
properties of the thermal radiation emitted by black
bodies makes them of
particular theoretical interest.

The

spectral distribution of black

body

radiation

is

specified

by the

THERMAL RADIATION AND QUANTUM THEORY

48

quantity IT {X), which

is

so denned that IT (X) dX

[Ch. 2

equal to the energy

is

emitted per second, in radiation of wavelength in the interval X to X

from

cm2

of a surface at temperature T.

Pringsheim are associated with the


quantity (1899).

dX,

The names Lummer and

earliest accurate

measurements of

this

The measurements were made with an instrument


prism spectrometers used in measuring optical
had to be used in order to have the
transparent to the long wavelength thermal radiation.

essentially similar to the

spectra, except that special materials


lenses, prisms, etc.,

10

1750"K

r=1500K

r=1250K
T=1000K

12

X
Figure 2-4.

The

(in units of

spectral distribution of black

4
-4
10
cm)

body radiation

at several temperatures.

The observed dependence of IT (X) on X and Tis indicated in figure (2-4).


The arrow on the abscissa indicates the wavelength at which the eye has its

maximum

response (green

light).

IT (X) increases with increasing T. The


integral of IT (X) over all X is, of course, just equal to the quantity IT
previously defined. This integral, which is equal to the area under the
curves, does increase with the fourth power of T, in agreement with

We

see that, for

equation (2-8).

any fixed

X,

Figure (2-4) also shows that the spectrum

shifts

toward

Sec. 5]

WIEN'S

LAW

49

shorter wavelengths as
increases.
quantitative inspection of the
figure will demonstrate the validity of
the equation

^max c 1/r

where Xmax

is

the X at which IT {X) has

its

(2-10)

maximum value for a particular

T.

All these results are in agreement with the


everyday experience that
bodies emit more heat as their temperature
increases, and that with
increasing temperature their "color" shifts from
dull red to blue white
(i.e., increasingly more radiant energy
is emitted in the region of short
wavelengths).

Now consider an object containing a cavity which is connected to the


outside by a small hole. Radiation incident upon the
hole from the outside
enters the cavity and is reflected back and forth by the
walls of the cavity,
eventually being absorbed on these walls. If the area of
the hole is very
small compared with the area of the inner surface of the cavity,

a negligible

amount of the incident radiation will be reflected back to the hole. Then all
the radiation incident upon the hole is absorbed. For the hole,
a = 1 and
must have the properties of the surface of a black body.
Next assume that the walls of the cavity are uniformly heated to
a

therefore the hole

The walls will emit thermal radiation which will fill the
small fraction of this radiation incident from the inside upon

temperature T.
cavity.

The

the hole will pass through the hole.

Thus the hole will act as an emitter of


thermal radiation. Since the hole must have the properties of the
surface
of a black body, the radiation emitted by the hole must have
a black body
spectrum. But, as the hole is merely sampling the thermal radiation present
inside the cavity,

it is

clear that the radiation in the cavity

a black body spectrum.

In fact,

it

will

must also have


have a black body spectrum

characteristic of the temperature T of the walls since this is


the only
temperature defined for the system. The spectrum emitted by the hole in
the cavity is specified in terms of an energy flux I (X), but it is
convenient
T
to specify the spectrum of the radiation inside the cavity in
terms of an
energy density p T (X) which is defined such that
p T (X) dX is the energy contained in 1 cm3 of the cavity in the wavelength interval X to X
dX. It is
apparent from the discussion above that
Pt (X) is proportional to IT (X),
with a proportionality constant that does not depend on X or T. This statement, as well as all the other statements of the last two paragraphs, can,
of course, be proved rigorously.

5.

Wien's Law

Many of the properties of black body radiation can be understood from


the classical theory of thermodynamics. In 1884 Boltzmann produced a

THERMAL RADIATION AND QUANTUM THEORY

50

theoretical derivation of the equation giving IT

per

the total energy emitted

cm2

per second, for a black body radiator (equation 2-8 for the case

1).

In this derivation he considered a cavity with reflecting walls in

the form of a cylinder with a movable piston and


radiation at temperature T. Thermal radiation, in

electromagnetic radiation, can be


to

[Ch. 2

its

energy density.f Taking

this

shown

filled

common

with thermal
with

all

other

to exert a pressure proportional

system through a cycle of expansion and

,x-*<>x

X\

/
/

\\

/
/

N*

xrFigure 2-5.

I449K;

An

experimental verification of Wien's

0,T =

Modern

Introduction to

compression,

From

I259K.

K.

F.

Physics, 5th ed.,

called

in

Richtmyer,

E.

law.

#, T

I646K;

H. Kennard, and T.

McGraw-Hill Book Co.,

New

x,T =

Lauritsen,

York, 1955.

thermodynamics a Carnot cycle, Boltzmann


work done by the pressure of the radiation,

obtained a relation between the

and its temperature. This relation leads to the desired result since the
pressure can be expressed in terms of the energy density, which can in turn
be expressed in terms of IT
.

In the process of expansion or compression of such a cavity filled with


radiation, the wavelength of any spectral component of the radiation will

be changed as the result of a Doppler


piston.

From

shift

upon reflection from the moving


fact, Wien (1893) was able

a detailed consideration of this

to derive a general functional

body radiation known

form

for the spectral distribution of black

as Wien's law:

MA)=^P
where /(A T)

is

(2-11)

some function of the product of the wavelength and the


is not specified by Wien's derivation.
This

temperature, whose form

equation

is

in excellent

seen from figure (2-5).


t In this respect

it

agreement with the experimental data, as can be


This figure plots A 5 times p T (X) as a function of

behaves just

like a gas.

Sec. 6]

THE RAYLEIGH-JEANS THEORY

51

XT, using 1259, 1449, and 1646 data of


Lummer and Pringsheim for
IT (X) [which is proportional to
p T (X)]. We see that all sets of data fall
on the same smooth curve, confirming the prediction
of equation (2-11)
that X h p T (X) is equal to a universal function
of the variable XT.
It is

apparent that Wien's law agrees with the empirical


fact expressed by
J
equation (2-10).

6.

The Rayleigh-Jeans Theory

Wien's thermodynamical derivation performed the useful


role of allowall black body spectra measured at
different temperatures to be discussed theoretically in terms of the single function/pr).
But the derivation
did not evaluate the form of this function. It is
typical of a thermodynamical argument to show that certain relations must
obtain between the
variables which describe some physical system,
but not to provide a
complete theory of the behavior of the system. The
reason is that such
arguments are based on general principles that apply to all
physical systems,
but do not involve the details of the composition of
the particular system
in question. A theory which would be able
to evaluate the function
f{XT) must take into account some of the detailed properties
of a black
body.
ing

At

the turn of the present century several attempts


were made to formusuch a theory. In one of them the behavior of accelerating
electric
charges in the walls of a black body cavity was
considered. These
charges were assumed to be the source of the black
body radiation. It
was also assumed that each charge undergoes simple
harmonic oscillation,
of some particular frequency, about its equilibrium
late

position. According
to electromagnetic theory, each charge then
radiates at a frequency
precisely equal to its particular oscillation frequency.
In thermal equilibrium the energy of a particular frequency, or wavelength,
component of
the black body radiation must be proportional
to the average energy of the
corresponding charged oscillator because the radiation and
the oscillator
are constantly exchanging energy. From
detailed considerations of the
absorption and emission properties of a charged oscillator,
the proportion-

constant was evaluated. It contains the square of


the frequency.
Next, certain results of the theory of statistical
mechanics (to be described
later) were used to evaluate the average
energy of the charged oscillators
in terms of the temperature of the walls in which
they are located. From
ality

these results the spectral distribution of the black


body radiation,
the function f(XT), is immediately obtained.

therefore

pursue

this theory further.

We

Instead,

we

shall turn to the theory

and

not
of Rayleigh
shall

THERMAL RADIATION AND QUANTUM THEORY

52

[Ch. 2

and Jeans which


because

leads to the same results, and which has a sounder basis


does not require such a detailed picture of the origin of the black

it

body radiation. In fact, it has been shown that the form off(XT) obtained
by Rayleigh and Jeans is a necessary consequence of the theories of
classical physics.

Consider a cavity with metallic walls at temperature T. The walls emit


thermal electromagnetic radiation. In thermal equilibrium this radiation

has a black body spectrum characteristic of the temperature T. Furthermore, as we shall show, in the steady state attained at equilibrium the
electromagnetic radiation inside the cavity must exist in the form of
standing waves with nodes at the metallic surfaces.

Rayleigh and Jeans

number of standing waves with nodes at the surfaces of the


cavity in the wavelength interval A to 1 + dX. They then calculated the
average energy of these waves. The energy depends on the temperature T
and is evaluated from the theory of statistical mechanics. The number of
calculated the

standing waves in the wavelength interval times the average energy of the
waves divided by the volume of the cavity is equal to the average energy
per cm3 in the wavelength interval X to X + dX, which is just p T (X). From
this f(XT) is evaluated by using equation (2-11).
Let us assume that the metallic walled cavity filled with electromagnetic
radiation is in the form of a perfect cube of edge length a, as shown in
figure (2-6). Then the radiation reflecting back and forth between the
walls can be analyzed into three components along the three mutually
perpendicular directions defined by the edges of the cavity. As the opposing walls are exactly parallel to each other, the three components of the
radiation do not mix, and we may treat them separately. Consider the
x component and the metallic wall at x = 0. All the radiation of this
component which is incident upon the wall is reflected by it, and the
incident and reflected waves combine to form a standing wave. Now,
since electromagnetic radiation is a transverse vibration with the electric
field

vector

perpendicular to the propagation direction, and since the

propagation direction for


question,

an

its

is

this

component

parallel to the wall.

electric field parallel to its surface,

is

perpendicular to the wall in

But a metallic wall cannot support


as currents can always flow in such

as to neutralize the electric field. Therefore E for this component


must always vanish at the wall that is, the standing wave associated with
the x component of the radiation must have a node at x = 0. The standing
wave must also have a node at x = a because there can be no parallel

way

corresponding wall. Furthermore, similar conditions


apply to the other two components the standing wave associated with the
and y
a, and the standing wave
y component must have nodes at y
associated with the z component must have nodes at z
and z a.
electric field in the

THE RAYLEIGH-JEANS THEORY

Sec. 6]

53

These conditions put a limitation on the


possible wavelengths of the
electromagnetic radiation contained in the
cavity.

To determine this limitation, consider radiation of


wavelength X and
frequency v
c\X, propagating in the direction

defined by the three


The radiation must be a standing
wave since all three of its components are standing
waves. We have
indicated the locations of some of the fixed
nodes of this standing wave
by a set of planes perpendicular to the propagation
direction a, y. The
angles

a,

/?,

y, as

Figure 2-6.

shown

in figure (2-7).

cubical cavity containing electromagnetic


radiation.

distance between these nodal planes of the


radiation is just 1/2, where X
wavelength.
have also indicated the locations at the three axes
of the nodes of the three components. The distances
between these nodes
are
is its

We

X^ __

2 cos a

2 cos p

2 cos y

(2-12)

THERMAL RADIATION AND QUANTUM THEORY

54

[Ch. 2

Let us write expressions for the magnitudes at the three axes of the
of the three components. They are

electric

fields

E(x,

i)

E(y,

t)

E(z,

t)

= A sin (Inx/X^ sin {lirvi)


= B sin (2irylXy) sin (2nvt)
= C sin (2ttz/A ) sin (2m>t)

(2-13)

for the x component represents a wave with a maximum


amplitude A, with a spatial variation sin (2Trx/Xx), and which is oscillating

The expression

z
,

T
C\J

CM

Mil

,<

feft

Hji

^X
c)

*-\ xl2-+ ^A*/2-*


Figure 2-7.

The nodal

planes of a standing

wave propagating

in a certain direction in

a cubical cavity.

the
0, 1, 2, 3,
with frequency v. As sin (2irxlXx) vanishes for 2x\X x
separfixed
nodes
has
because
it
X
wavelength
wave
of
standing
wave is a
x

ated by the distance Ax

XJ2. The expressions for the y and

components

represent standing waves of maximum amplitudes B and C and wavelengths X y and X but all three component standing waves oscillate with
the frequency v of the radiation.

Note that these expressions automatically

requirement that the x component have a node at x = 0, the


0, and the z component have a node at
y component have a node at y
z
0. To make them also satisfy the requirement that the x component
satisfy the

THE RAYLEIGH-JEANS THEORY

Sec. 6]

55

have a node at x = a, the y component have


a node at y
component have a node at z = a, set

= nx
= nv
=

2*/4
1y\K
2/ l z
.

ycosa =

and the

=a
for y = a
for z = a

2,

W!C ,

a,

(2-14)

C os

,8

f"

^,

0, 1, 2, 3

Squaring both sides of these equations and adding,


we obtain
2

(2a/A) (cos 2

But the angles

a,

/S,

cos 2 p

cos 2 y)

nx

y have the property


cos 2 a

cos 2

,5

cos 2

Thus
2a/A

= Vn 2 +

where n., , n z take on all possible integral values.


This equation describes
the limitation on the possible wavelengths
of the electromagnetic radiation
contained in the cavity.
It is convenient to continue
the discussion in terms of the allowed
frequencies instead of the allowed wavelengths.
They are

= =
A

la

^ nl +

"**

"*

Now we shall count the number of allowed frequencies in a given


interval

by constructing a uniform cubic

(2

"15

frequency

one octant of a
rectangular coordinate system in such a way
that the three coordinates of
each point of the lattice are equal to a possible
set of the three integers
nx

lattice in

n y n z See figure (2-8). By construction,


each lattice point corresponds to an allowed frequency. Furthermore,
N(v) dv, the number of
allowed frequencies between v and v + dv, is
equal to N(r) dr, the number
of points contained between concentric
shells of radii r and r + dr
where
,

'

From

equation (2-15) this

= Vn 2 +

n2

is

for

where nx = 0, 1, 2, 3,
= 0, 1, 2, 3,
Using equations (2-12) these conditions become
.

2a

(2-16)

THERMAL RADIATION AND QUANTUM THEORY

56

Since N(r) dr

is

equal to the volume enclosed by the shells times the

density of lattice points, and since,

N(r) dr

is

Figure 2-8.

by construction, the density

is

unity,

simply

N(r) dr

frequencies

[Ch. 2

4jrr

dr

The rectangular coordinate system used in counting the number


Only a few of the points are shown.

of allowed

in a cubical cavity.

Setting this equal to N(v) dv,

and evaluating

r2

dr from equation (2-16),

we have

v)dv=^y dv

N{

This completes the calculation except that

by a factor of 2 because,

we must

multiply these results

we have
enumerated, there are actually two independent waves corresponding to
the two possible states of polarization of electromagnetic radiation.! Thus
for each of the allowed frequencies

t The fact that electromagnetic radiation in the visible range (light) can have either
of two independent circular polarizations is discussed in almost any elementary physics
text.

THE BOLTZMANN PROBABILITY DISTRIBUTION

Sec. 7]

the

number of allowed

57

frequencies in the frequency interval v to


v

N(v) dv

dv

is

(2 -17)

can be shown that N(v) dv is independent


of the assumed shape of the
and depends only on its volume, a3

It

cavity

The Boltzmann Probability Distribution

7.

The next step in the Rayleigh- Jeans calculation


the average energy contained in each
standing

is

the evaluation of

wave of frequency v
the particular energy of some wave
can have

According to classical physics,


any value from zero to infinity; the actual
value is proportional to the
square of its average amplitude (cf.
equation 2-4). But, if we have a
system containing a large number of
physical entities of the same kind
which are
thermal equilibrium with each other at a
temperature T such
as the system of standing waves in
equilibrium in

the black body c'avity


the classical theory of statistical
mechanics demands that the energies of
these entities be distributed according
to a definite probability distribution

whose form

is

specified

by

T.

As

the average energy

determined by the
depends on T
It is not difficult to see the
essential ideas underlying these results
of
statistical mechanics, if a sufficiently
simple system is considered. Imagine
a system consisting of entities of the
same kind which contain energy An
example would be a set of identical coil
springs which are vibrating
Assume that the system is isolated so that
its total energy
probability distribution,

it

must have a

is

definite value that

content

is

and assume also that the entities can exchange


energy with each
other through some mechanism so that
the system is in thermal equilibrium
Finally, assume that there are only
four of these
constant,

entities in the

that the total energy of

system

any entity is restricted to the values s =


Ae
2A 3A 4Ae, 5A
and that the total energy of the system
(which
now must be some integral multiple of As) has
the value 3Ae The last
assumptions are made strictly for the purpose
of simplifying the calculation
we are about to perform. After the calculation
we shall imagine letting
As approach zero so that e may have any
value, but at the same time
,

keeping the

energy of the system constant. We


shall also imagine
entities in the system become large.
Since the entities can exchange energy
with each other, all possible
divisions of the total energy 3As
between the four entities can occur The
only possible divisions are indicated
in the diagrams of figure (2-9)
Consider the diagram labeled i
1.
This corresponds to divisions of
lettmg the

total

number of

THERMAL RADIATION AND QUANTUM THEORY

58

total energy in

which

for three entities

But
can be the one in the energy

and e = 3Ae for the


any one of the four

there are actually four such divisions, since

state e

This

3Ae.

is

[Ch. 2

fourth.
entities

indicated in the

column marked "number of duplicate divisions." The same is true for the
divisions of the type = 3. The reader should verify for himself that there
/'

are twelve duplicate divisions of the type

any rearrangement of

entities

In evaluating this number,

2.

different energy states is to

co

CO

CO

CO

co

o c
o

<
to

to

40/20

24/20

P'(S)

Figure 2-9.

between

Illustrating

to

be

a>

^a;

ro
8.2

%'S.

4/20

12/20

12

4/20

12/20

a simple

4/20

calculation

0/20

leading to the

Boltzmann probability

distribution.

counted as a new division since the

entities

could presumably be

dis-

tinguished experimentally when in different energy states even though they

same kind, but any rearrangement of entities in the same energy


not counted because entities in the same energy state could not be

are of the
state is

distinguished.!

We now make the final assumption:

All possible divisions

of the total energy occur with the same probability. Then the probability
that the divisions of a given type will occur is proportional to the number
of duplicate divisions of that type, and the relative probability Pt is just
equal to that number divided by the total number of duplicate divisions.

The

are evaluated in the

'.

The

total

= 4x3x2x1.

state

let

us calculate

number of rearrangements (permutations) of the four entities is


But rearrangements of the two entities within the same energy
do not count, so this number must be divided by 2

t Hint:

column marked "Pt ." Next

THE BOLTZMANN PROBABILITY DISTRIBUTION

Sec. 7]

P'(e), the probability

0.

and the

For

of finding an entity in the energy state

divisions of the type

relative probability

59
s.

Consider

there are three entities in this state,


that these divisions occur is 4/20; for /
2
/

two entities, and Pt is 12/20; for i = 3 there is one entity, and


4/20. Thus P'(0), the probability of finding an entity in this
state,

there are

P
3

is

4/20

lated

12/20

m the same way

= 40/20.

4/20

2A

Figure 2-10.

is

The

values of P'(a) calcufor the other values of s are given in the row marked
1

3A

4A

comparison of the results of a simple calculation and the Boltzmann

probability distribution.

They are also plotted as points in figure (2-10).


the exponential function

"P'(e)."
is

P(e)

where

and

are constants.

adjusted to give the best

fit

Ae- t/

The

solid line

(2-18)

(In the figure the constants have been

to the points representing the results of our

calculation.)

Now we must imagine letting Ae go to zeroand at the same time


changing the condition on the total energy to read that it is equal to
Ae, with
increasing as rapidly as Ae decreases, so that the total energy
remains constant. The result of this step is that the calculated function

becomes defined for values of e which are closer and closer together.
In the limit, the energy e becomes a continuous variable as classical
physics

P'(e)

demands

must, and the probability distribution P'(e) becomes a conFinally, when the number of entities in the system is
allowed to become large, this function is found to be identical with the
P(e) of equation (2-18).
it

tinuous function.

The quantity P(e)


of

its

specifies the probability

energy states between e and s

of finding an entity in one


de for a system containing a large

THERMAL RADIATION AND QUANTUM THEORY

60

[Ch. 2

number of entities of the same kind in thermal equilibrium with each


other. The probability is simply P(e) de. It also specifies the average
energy of one of the

which

entities,

is

eP(e) de
*

Jo

(2-19)
P(e) de

I
The integrand

in the

numerator

the energy e weighted by the probability

is

The denominator

that an entity will have this energy .f

is

the probability

of finding an entity with any energy; it is just the total number of


in the system. Evaluating P(e) from equation (2-18),
P(e)

we

Ae- Zl

Ae-* e

entities

l/e

find

Aee'^de
-

ee~

Jo

de

Jo

Ae~ m de

e-" de

Jo

Jo

Now
X
'de
da.
da,

("e-de
da.

Jo

Jo
e

_a

f
Jq_

de
Jo

Jo

Thus

= -

In

da,

Evaluating the integral,

e~" de

Jo

we have
d

da

-1

da.

da.

a.

_2

1
a.

Therefore
(2-20)

By

treating a system containing

two

different kinds of entities in equili-

can be proved that for a given temperature the constant e


must be independent of the nature of the entities of which the system is
comprised. Consider then the system containing a set of identical coil
brium,

it

t We assume that the number of energy states in the interval e to e + de is independent


of e. The assumption is valid for a system of simple harmonic oscillators to which we
shall apply the equation.

THE BOLTZMANN PROBABILITY DISTRIBUTION

Sec. 7]

springs which can exchange energy

by some process or

are in equilibrium at temperature T.

Assume

other,

61

and which

that the lengths of the

springs can vibrate about their quiescent lengths, but that they cannot

otherwise move.

Then one "coordinate" per

spring

(its

length)

is sufficient

to describe the system at any instant,

and the system is said to have one


degree of freedom per entity. Now it is shown in almost any elementary
physics text that, for a system of gas molecules moving in thermal
equilibrium at temperature T, the average kinetic energy per degree of
-1
Boltztnanris
freedom is k T/2, where k
1.38 x 10~ 16 erg-deg
is

the law of equipartition of energy, and it is true for any


classical system. Thus, for our classical system of springs, the average

This

constant.

is

kinetic energy of vibration of

one of the springs

is

k T/2.

But, for a vibrat-

ing spring, or any entity in which the single "coordinate" executes simple

harmonic
energy.

oscillations, the average total

The reader may

easily

prove

energy

this for

one of the springs of our system, the average


e

With equations (2-20) and

(2-21),

twice the average kinetic

total

energy

e is

= kT
we

(2-21)

find

= kT

eQ

This determines the value of

is

a pendulum. Therefore, for

(2-22)

and equation (2-18) becomes

P(e)

Ae- elkT

(2-23)

is the famous Boltzmann probability distribution.


These equations should apply directly to a system of standing waves in
equilibrium in a black body cavity because each of the entities of this
system also has a single "coordinate" (its amplitude) which executes

This

simple harmonic oscillations.f In the Rayleigh- Jeans calculation for this

system

we

are interested only in equation (2-21).

It

could have been

written directly from the law of equipartition of energy, but later

we

shall

need (2-23) and the calculation preceding (2-20).


t Equations (2-22) and (2-23) also apply to a system of moving gas molecules
however, equation (2-21) is replaced by e = f&Tas the average total energy equals the
average kinetic energy, but there are 3 degrees of freedom. There is no inconsistency
here since equation (2-20) is also changed. It is replaced by s = ie because a factor
proportional to e^ must be included in the integrands of equation (2-19) to account
for the fact that the number of energy states in the interval e to s + fife is proportional to
e^ (cf equation 12-37). The product of this factor and P(e) gives the Maxwellian energy
.

distribution of gas molecules,

Ae ^e~ slkT
,

THERMAL RADIATION AND QUANTUM THEORY

62

8.

[Ch. 2

Comparison with Experiment

The energy per cm3 in the frequency interval v to r + dv of the black


body spectrum of a cavity at temperature T is just the product of the
average energy per standing wave (equation 2-21) and the number of
standing waves in the frequency interval (equation 2-17) divided by the
volume of the

cavity.

That

is,

p T {v) dv

Snv 2 kT

dv

(2-24)

cr

To make

a comparison with our earlier discussion,

the energy density in terms of the wavelength

X.

we must

express

This can be done by

using the relation


v

c/X

Thus

dv=-j- dX

(2-25)

Consider a certain interval of the spectrum, which can be defined in terms


of frequency as v to v + dv, or equally well in terms of wavelength as
X to X + dX. For this interval the following relation obtains

PT (X) dX

= -p T (v) dv

(2-26)

because both sides are equal to the energy content per cm3 in that interval
of the spectrum. [The minus sign compensates for the fact that, according
to equation (2-25), dv and dX are of the opposite sign.] Combining
equations (2-25) and (2-26),

we have

PT(X)dX
Using

this to

We

is

p T (v)j-2 dX

transform equation (2-24) to the variable

Pt {X) dX
This

&nk
X

(2-27)

X,

it

find

XT dX

(2-28)

the Rayleigh- Jeans spectral distribution of black

note that

we

body

radiation.

has the form required by Wien's law (equation 2-11)

it
can be written p T (X) dX =f(XT)/X5 where f(XT) = ZnkXT.
However, as is seen in figure (2-11), the Rayleigh- Jeans spectrum is in

since

PLANCK'S THEORY

Sec. 9]

63

complete disagreement with the experimental data;

form

it

predicts the

wrong

for the function /(AT).

In the limit of long wavelengths, the Rayleigh- Jeans spectrum approaches


the experimental

p T(X) -*

and

is

But, as

results.

A->0,

theory predicts that

their

whereas experiment shows that p T {X) always remains finite


equal to zero at zero wavelength. The unrealistic behavior of the
oo,

Rayleigh- Jeans distribution at short wavelengths


"ultraviolet catastrophe." This terminology

is

is

known in physics as the

suggestive of the importance

Rayleigh-Jeans

spectrum

T=1646K

X(in units of 10
Figure 2-1

1.

A comparison

-4

cm)

of the Rayleigh-Jeans spectrum and experiment.

of the failure of the Rayleigh-Jeans theory. Let us evaluate the total energy
cm3 for the Rayleigh-Jeans spectrum:

per

0. We have here an even more impressive


This goes to infinity unless T
complete
failure
of the Rayleigh-Jeans spectrum. Yet
indication of the
earlier, a necessary consequence
we
have
mentioned
spectrum
is,
as
this

of the theories of classical physics.

9.

Planck's

Theory

The discrepancy between experiment and theory was resolved in 1901


by Planck but only at the expense of introducing a postulate which was

not only new, but also drastically at variance with certain concepts of

THERMAL RADIATION AND QUANTUM THEORY

64

Planck's postulate

classical physics.

Any physical
oscillations

may

[Ch. 2

be stated as follows

whose single "coordinate" executes simple harmonic


a sinusoidal function of time)^ can possess only total

entity

(i.e.,

is

energies s which satisfy the relation


e

where v

An

nhv,

0, 1, 2, 3,

the frequency of the oscillation

is

and h

a universal constant.

is

way of

energy level diagram provides a convenient

behavior of an entity governed by


contrasting this behavior with

(2-29)

and

this postulate,

illustrating the

it is

also useful in

what would be predicted by

classical

= bhv

Classical

Figure 2-12.

= 4hv
= 3kv
= 2hv
= hv
=

Planck

Energy

level

diagrams for a

classical

simple harmonic oscillator and a

simple harmonic oscillator obeying Planck's postulate.

physics.

See figure (2-12).

we

In such diagrams

indicate each of the

possible energy states of the entity with a horizontal line.

The

distance

from the line to the zero energy line is proportional to the total energy to
which it corresponds. Since the entity may have any energy from zero to
infinity according to classical physics, the classical energy level diagram
consists of a continuum of lines extending from zero up. However, the
entity executing simple harmonic oscillations can have only one of the
discrete total energies s

This

postulate.

is

nhv, n

0, 1, 2, 3,

obeys Planck's

if it

indicated by the discrete set of lines in

its

diagram. The energy of the entity obeying Planck's postulate

energy level
is

said to be

and the
quantum number.
Let us recalculate p T (X) under the assumption that Planck's postulate
is obeyed by the electromagnetic standing waves whose amplitudes are,
according to equation (2-13), executing simple harmonic oscillation with

quantized, the allowed energy states are called


integer n

is

quantum

states,

called the

t The word "coordinate" is used in its general sense to mean


describes the instantaneous condition of the entity. Examples are

spring, the angular position of a

any quantity which


:

the length of a coil

pendulum bob, the amplitude of a wave. All

examples happen also to be sinusoidal functions of time.

these

PLANCK'S THEORY

Sec. 9]

frequency

65

No modification will be required in the calculation of N(v) dv

v.

no point

was it necessary to specify the energy


However, the evaluation of s involved integrals in
which the energy of the standing waves, e, was the variable of integration.
This was proper in the original calculation since classically the energy can
assume any value. But, under Planck's postulate, e can only have the
discrete values nhv where n = 0, 1, 2, 3, 4,
Thus we must recalculate
Then
e, replacing all integrals over e by summations over this variable.
since at

in that calculation

of the standing waves.

00

-Q

__

00

00

n=

__ n =

where we have assumed, as did Planck, that the Boltzmann probability


distribution

employed

In a manner completely analogous to that

applies.

still

in the calculation preceding equation (2-20),

we can show

this

yields
1

= -

oo

da.

2 e~"'
n=o

<*7iV

In

hv

Now
tznhv

Ip

p 2<xhv

_ 3aftv

_i_

_i_

or
00

-v

Je
n=
where x

e~

xhv

=1 + X+X +X +
Z

...

But

(1

x)' 1

an " v

(1

x%

x3

Therefore
GC

2=

e~

_0* v )~ 1

71

and

- ln(l

-e-"**)" 1

den

_(1

e -^)(_l)(l
1 '"

- xh

hv

hve""'
1

-"v

^v

*y%-{-hv)e- ahv
hv

e^l^T

(2-30)
i

Evaluating p T (v) dv as before, using equation (2-17), we have


Pt(v) dv

~\

= SJMdv =

hv

THERMAL RADIATION AND QUANTUM THEORY

66

[Ch. 2_

Transforming to the variable X with equation (2-25), we obtain

Pj{X) dX

Znhc
J5

This

is

the black

body

JiclkXT

dl

(2-32)

spectral distribution derived

by Planck.

It

Wien's law with/(Ar) = %-nhc\(e ullc, T 1). This function involves


the constant h whose numerical value is to be determined by fitting the
-

satisfies

\
Figure 2-13.

(in units of

10 ~* cm)

comparison of Planck's spectrum and experiment.


T = I646K.
is theoretical.

The dots are

experimental and the curve

theoretical spectrum to the experimental data.


figure (2-13).

The value providing the


h

6.63

best

X 10"27

fit

Such a

to the data

fit

is

shown

in

is

erg-sec

This constant is known as Planck's constant. The figure shows that Planck's
theoretical spectrum is in extremely good agreement with the experimental
spectrum for T 1646 K. Since it satisfies Wien's law, we know that it
will agree equally well with experiment at any other temperature.

10.

Some Comments on

Planck's Postulate

worth while to give a qualitative explanation of how Planck's


which in terms of the mathematics only has the effect of replacing
integrals by sums, is able to produce such a remarkable change in the
It is

postulate,

results of the calculation.

essentially

The

failure of the Rayleigh- Jeans theory is

due to the rapid increase in N(y) dv as

becomes

large.

Since

SOME COMMENTS ON PLANCK'S POSTULATE

Sec. 10]

67

the classical law of equipartition of energy requires that e be independent


of v, the spectrum itself will also increase rapidly with increasing v (or

decreasing

Planck's postulate has the effect of putting a cut-off into


does not obey the classical law. This keeps the spectrum
bounded and, in fact, brings the spectrum to zero as v -* ao (or as k -* 0).
This behavior of s can be seen directly from equation (2-30). The decrease
s so that

A).

it

comes about in the following way. Under Planck's postuthe energy of the standing waves is either e
0, or s
hv, or e
2hv,
etc. If we now consider the Boltzmann probability distribution
(equation
in e as v ->- oo

late,

we see that for v large enough that hv


kT the probability that a
standing wave will have an energy other than e
becomes negligible.
Consequently, when we evaluate s for standing waves whose frequency

>

2-23),

we find e
0. On the other hand, for the very low frequency
standing waves in which hv
kT, the quantum states are so close together
that they behave as if they were continuously distributed, as in the
classical
theory. Thus in the low frequency limit Planck's value of e
approaches the
v is large,

<

classical value

of i

kT, as

from equation (2-30).


was not so far reaching as it is
in the form we have given. Planck's initial work was done
with the theory,
described in section 6, in which the energy in a particular
frequency
component of the black body radiation is related to the energy of a
In

its

is

also apparent

original form, Planck's postulate

(charged) particle in the wall of the black body cavity oscillating


sinusoidsame frequency, and he postulated only that the energy of the

ally at the

oscillating particle is quantized.

After the development of the Rayleighit is equivalent to quantize directly


the oscillating electromagnetic waves, and the postulate
was broadened
to include any entity whose single "coordinate" oscillates
Jeans theory

it

became apparent that

sinusoidally.

Planck's

first

step consisted in finding,

by

and

an empirical
equation which agreed with the experimental spectra. Then he
set about
finding what modifications of the classical theory were
required to yield
this equation. In considering the derivation of the
Boltzmann probability
distribution P(e), which we described in terms of a simple
example in
section 7, he was led to try holding constant the quantity
that we called
trial

error,

Ae in our example, instead of taking it to zero as required by classical


theory. Although P(e) then has values only for discrete e,
these values are
not changed. However s is changed, as we have seen, and its new
value
depends on Ae. Planck soon found that the desired spectral distribution
is

obtained

Ae

proportional to the oscillation frequency v. Then the


possible energies e
n Ae become e
nhv, as stated in his postulate.
The success of Planck's postulate in leading to a theoretical black body
if

is

spectrum which agrees with experiment requires that we tentatively accept


at least until such time as it may be proved to lead to conclusions

its validity,

THERMAL RADIATION AND QUANTUM THEORY

68

[Ch. 2

which disagree with experiment. It may have occurred to the reader that
there are physical systems whose behavior seems to be obviously in disagreement with the postulate. For instance, an ordinary pendulum executes
simple harmonic oscillations, and yet this system certainly appears to be
capable of possessing a continuous range of energies. Before we accept
this argument, it would be prudent to make some simple numerical
calculations concerning such a system. Consider a pendulum consisting
of a 1-gm weight suspended from a 10-cm string. The oscillation frequency
of this pendulum is about 1 .6/sec. The energy of the pendulum depends on
the amplitude of the oscillations.

Assume the amplitude to be such that


extreme position makes an angle of 0. 1 radian with the
vertical; then the energy E is approximately 50 ergs. If the energy of the
the string in

its

pendulum is quantized, any decrease in the energy of the pendulum, due


for instance to frictional effects, will take place in discontinuous jumps of

AE =

magnitude

not observed. But, if we consider that


10" 26 ergs, whereas
50 ergs,
it is apparent that even the most sensitive experimental equipment would
be totally incapable of resolving the discontinuous nature of the energy

AE =

hv

6.63

hv.

10

This

-27

is

1.6

E~

decrease.

No

evidence, either positive or negative, concerning the validity of

Planck's postulate

The same

we
is

is

is

to be found

consider systems in which v

of the order of

example

is,

radiation.

from experiments involving a pendulum.


Only when

true of all other macroscopic mechanical systems.

E are

we

is

so large and/or

is

so small that hv

One

in a position to test Planck's postulate.

of course, the high frequency standing waves in black body


Other examples will be considered in the following chapters.

BIBLIOGRAPHY
Mayer,

New

J. E.,

and M. G. Mayer,

Statistical Mechanics,

John Wiley and Sons,

York, 1940.

Peck, E. R., Electricity and Magnetism, McGraw-Hill

Book

Co.,

New

York,

1953.

Richtmyer, F. K., E. H. Kennard, and T. Lauritsen, Introduction to Modern


Physics,

McGraw-Hill Book Co.,

New

York, 1955.

EXERCISES
1.

Verify (2-2) by using the procedure suggested immediately below that

equation.
2.

When

the sun

is

directly overhead, the thermal energy incident

upon

the

Ch.

EXERCISES

2]

69

about 1.4 x 10 6 ergs-crrf^-secr 1 The diameter of the sun is


about
1.6 x 10 u cm, and the distance from the sun to the earth is
about 1.3 x 10 13 cm.
Assuming that the sun radiates like a black body, use equation (2-8) to
estimate
earth

its

is

surface temperature.

3.

Repeat the calculation of section 7 for a system with

energy 6 Ac.

Compare

six entities

and

total

the results with equation (2-18), as in figure


(2-10).

4. Prove the fact, used in obtaining


equation (2-21), that the total energy
of a simple harmonic oscillator is twice its average
kinetic energy. Do this for
the case of a pendulum.
5. Integrate equation (2-32) over all
the results agrees with equation (2-8).
6.

To

A,

and show that the

T dependence

of

a certain approximation, an atom of a crystal executes


simple harmonic
about its equilibrium position because it possesses thermal

oscillations

agitation

energy and experiences an approximately linear


restoring force. Use the bulk
compressibility of some crystal to estimate the strength
of the restoring force
and, from this, the oscillation frequency. Use
equation (2-21) to estimate the
oscillation energy at room temperature.
From these results determine the
approximate value of the quantum number that describes
the motion of this
microscopic simple harmonic oscillator, and compare
it with the approximate
value of the quantum number that describes
the motion of the macroscopic
simple harmonic oscillator discussed in section
10.

CHAPTER

Electrons

and Quanta

Cathode Rays

I.

In this chapter

we

the photoelectric effect

shall primarily

be concerned with two phenomena,

and the Compton

effect, the interpretation of which


played a crucial role in the development of the quantum theory. These

phenomena

involve the interaction of electromagnetic radiation with

To pump

Glass tube

Cathode

A
matter

we

specifically

cathode ray tube.

with the electrons contained in matter. Consequently

shall first briefly describe the discovery of electrons

and the quantitative

measurement of their charge and mass.

The conduction of

electricity

through rarefied gases was a popular

subject of investigation in the second half of the nineteenth century.

70

Sec.

I]

CATHODE RAYS

71

typical experimental arrangement is shown in figure (3-1).


high voltage
applied between the negative electrode, or cathode, and the positive
electrode, or anode, causes current to flow through the
tube which is
filled with some gas (such as air) at a low pressure.
The current produces
a striking pattern of layers of glowing gas which alternate with dark
layers.

The exact nature of the pattern depends on the pressure of the gas as well
as on its composition and the magnitude of the voltage.
A complicated
set of phenomena, involving ionization of the gas, is
responsible for the
conduction of current and for the light emitted by the gas. We shall not
discuss these processes here.
pressures, in the region of 0.01

menon

develops.

The gas

When

the tube

is

pumped

to very low

mm of Hg or less, a quite different pheno-

glow (for one reason, because there


ammeter indicates that a current
is still flowing in the tube.
Another indication that something is still
happening in the tube is the presence of a luminous spot S
on the inner
surface of the glass wall at the end of the tube. The
size of the spot and
its position on the end wall of the tube
are observed to depend on the size
and location of the holes in diaphragms D and D' in such
a way as to
imply that the luminous spot is caused by rays of
particles which leave
from the cathode and travel in straight lines, finally impinging
upon the
is

very

little

gas

left).

ceases to

Nevertheless, the

wall.

Additional supporting evidence is obtained from the


observation
that a solid object placed to the right of D' will
cast a sharp shadow on the
luminous spot at the end of the tube. The position of the
luminous spot
can be influenced by the presence of a strong electric
or magnetic field in
the region to the right of D', showing that the
particles traveling from the
cathode to the wall of the tube are electrically charged.
The sense of the
deflection of the spot is such as to indicate that the
charge of the particles
is negative.f This was confirmed
by an experiment in which an insulated
metallic chamber was placed inside the tube
at S; the chamber was
observed to collect a negative charge. These particles,
which travel in
ray-like paths from the cathode, were called
cathode ray particles, or

sometimes simply cathode

rays. The apparatus shown in figure


(3-1)
called a cathode ray tube. It is now known that
the cathode ray particles
are emitted from the cathode as a result of its
bombardment by positive
is

ions of the gas contained in the tube.


t The reader may wonder why the negatively charged particles travel
in straight lines
in the region to the right of D\ instead of being
attracted to the positively charged anode
The reason is that most of the voltage drop across the

tube occurs near the cathode

Thus the charged particles are accelerated in the strong


electric field near the cathode
and then they continue moving in straight lines. The
electric field near the anode is too*
weak to produce an appreciable deflection.

ELECTRONS AND QUANTA

72

2.

[Ch. 3

The e/m Ratio of Electrons

In 1897 Thomson made accurate measurements or e/m, the ratio of


the charge of the cathode ray particles to their mass. This provided the

key to their identity. Thomson's apparatus is shown in figure (3-2).


Cathode ray particles of (negative) charge e and mass m travel from
the cathode C toward the anode A, which is pierced with a small hole and

H.v.

Glass tube-

An

Figure 3-2.

apparatus used to measure the e/m ratio for electrons.

first of two diaphragms, D and D', that define the particles


narrow beam. The particles then pass between two parallel metal
plates of length / and separation d. Finally they strike the end of the tube,
producing a luminous spot at Sv The application of a voltage V across
the two plates creates a force F e Vj'd which acts on the particles in a

also acts as the


to a

direction perpendicular to their velocity

verse acceleration a
t

l/v

acceleration lasts for the time

during which the particles are passing between the plates.

emerging,

The

This force produces a trans-

v.

The

(e/m)(V/d).

transverse

their

deflection

St S2

deflection

is

of the luminous spot

is

amplified

almost exactly 2L/1, where L


the center of the plates to the end of the tube. Thus
for small deflections,

is

SiS 2

VL

by a factor which,
the distance from

is

(3-1)

S^

known, equation

(3-1) allows the evaluation of e/m.

To

measured and the velocity v of the

Thomson made

Upon

\(e/m){V/d){f/vf.

If the deflection

is

%at 2

particles is

a separate measurement employing an


due to a voltage V across the two plates and also a
magnetic force resulting from the interaction of the moving charged
particles with a magnetic field
applied over the region between the plates

determine

v,

electrostatic force

THE CHARGE AND MASS OF ELECTRONS

Sec. 3]

and

73

With the
and magnetic

in a direction perpendicular to the plane of the figure (3-2).

correct orientation of the magnetic field, the electrostatic


forces will be opposed;

in fact, for a certain value of the magnetic field

strength the forces can be

made

to cancel out exactly.

For

this value of

the following relation obtains

eVjd

and the luminous spot


is

Hev/c

(3-2)

no deflection. The left side of this equation


and the right side the magnetic force. From

suffers

the electrostatic force,

equations (3-1) and (3-2), ejm can be evaluated in terms of the quantities
measured in the experiment. The currently accepted value is

ejm

=
=

5.27

10 17 esu/gm

1.76

10 8 coulombs/gm

The results of the experiment were observed to be independent of both the


composition of the cathode and the nature of the residual gas in the tube.
This showed that the value of ejm stated above is characteristic of the
cathode ray particles themselves and that all cathode ray particles have
same value of ejm.
The ratio of the charge to the mass of cathode ray

the

particles is very

larger than the value of that ratio for ionized atoms.


is

measured in experiments involving the

ionized atoms; an electrolytic

cell is

electrolysis

The

much

latter quantity

of solutions containing

operated for a certain period of time

and the total charge carried by the ionized atoms, as well as the total mass
of the atoms deposited at the electrode, is measured. The largest atomic
value of ejm is that for an ionized hydrogen atom, but ejm for a cathode
ray particle

is

1836 times greater than ejm for an ionized hydrogen atom.

Either the charge of a cathode ray particle

is

very

much

larger than the

charge of an ionized atom, or the mass of a cathode ray particle

much smaller than an atomic mass.


Thomson made the assumption that

the second alternative

was

In fact, he assumed that the charge of the cathode ray particles

equal in magnitude to the charge of a singly ionized atom.


charge

is

called the electronic charge.

modern terminology,

3.

The cathode ray

is

is

very

correct.

exactly

This unit of
particles,

in

are called electrons.

The Charge and Mass

of Electrons

The magnitude of the electronic charge was determined in a series of


experiments performed by Townsend and Thomson (ca. 1900) and by
Millikan (1909).

These experiments were based on the observation that

ELECTRONS AND QUANTA

74

[Ch. 3

very small drops of liquid often pick up electric charges. If a voltage V


is applied between two horizontal parallel plates located above and below
the region containing the charged droplets, a droplet can be suspended
against the force of gravity

when

q-=Mg

(3-3)

where d

is

the separation of the parallel plates, q

is

the charge

on the

980 cm-sec-2 In order to determine M,


g
the voltage is removed, and the droplet falls under the influence of gravity
until it reaches terminal velocity because the frictional drag F becomes
equal to the gravitational force Mg. According to Stokes' law, Fv =
6nrjav, where a is the radius of the droplet, v is its terminal velocity, and r\
droplet,

is

mass, and

is its

the coefficient of viscosity of the

Thus

is falling.

medium (air) through which


we have the relation

the droplet

at terminal velocity

Mg =

(3-4)

6irr]av

The radius of the droplet can be expressed in terms of its mass and density
= ^-n&p. Putting this into equation (3-4), we can
from the equation
solve for
in terms of the terminal velocity v. The latter quantity is
measured by observing with the aid of a microscope the distance through
which the droplet falls in a given time. Knowing M, we find q from equation

(3-3).

The results of a large number of measurements on the charge of different


droplets could

all

be expressed as
q

where n
That is,

is
it

4.80

lO

-10

esu

(3-5)

an integer, and where we quote currently accepted values.


was observed that the charge of a droplet was always some

integral multiple of 4.80

meaning that there

exists

10 -10 esu.

It is

reasonable to interpret this as

a basic unit of charge of magnitude

= 4.80
= 1.60

10" 10 esu

10- 19 coulombs

no droplet was ever found to have a charge less than this amount.
his colleagues assumed that the electronic charge was
equal to this basic unit of charge. Then, knowing the value of ejm for
an electron, they were able to evaluate the mass of an electron. The value
they obtained was not far from the currently accepted value,
since

Thomson and

m=

9.11

x 10-28 gm

BUCHERER'S EXPERIMENT

Sec. 4]

75

is now an overwhelming body of evidence proving the correctness


of the assumption that the electronic charge is equal to 4.80 X 10 -10 esu.

There

For

instance,

possible to evaluate the electron

it is

mass from the equation

(e/m) H +

e/m

1836

where (e/w) H+ represents the charge to mass ratio for an ionized hydrogen
atom and m n+ represents its mass. The latter quantity is equal to the
chemical atomic weight of hydrogen divided by Avogadro's number.f
Thus

L2JL_

6.02

L=

10 23 1836

9 ai
llxl0
x 1U
*

-28

This provides an independent check on the value of the electron mass


determined from e/m and the value of e and thereby confirms the correctness of the value of e given above.
In recapitulation, the electron is a particle having a negative charge
the magnitude of which is equal to the magnitude of the charge of a
singly ionized atom, and a mass which is smaller than the mass of an
ionized hydrogen

4.

atom by a

factor of 1836.

Bucherer's Experiment

This is an appropriate place to describe the work of Bucherer


(1909)
which provided the first experimental verification of the relativistic depend-

m=

ence of mass on velocity:


mjVl - ^/c 2 (equation 1-20). This
experiment consisted of a series of measurements of the e/m ratio of high
velocity electrons. The reader will recall that the theory of relativity does
not alter the laws of electromagnetic phenomena; the charge of an
electron,

just as

e, is

it is

a constant independent of its velocity in the theory of relativity

in the classical theory.

Consequently the

relativistic

theory

predicts

cV

W,

/e

cV

where (e/w),,/c_ is the value of e/m for electrons of velocities small compared to c. The quantity (e//n) e/( ^ is what was measured in Thomson's
experiments, since the velocity of the electrons in the cathode ray tubes
which he used is very small compared to c. In order to obtain electrons
t

We are ignoring the 0.05 percent difference between the mass of an ionized hydrogen

atom and

that of a neutral hydrogen atom.

76

ELECTRONS AND QUANTA

[Ch. 3

with velocities comparable to c, Bucherer used the electrons which


are
spontaneously emitted from the naturally radioactive elements (cf.
section 2, Chapter

These electrons are emitted with a continuous


4).
spectrum of velocities which extend, in some cases, to velocities as high as
Bucherer's apparatus was essentially similar to that used by
Electrons of a certain velocity were first selected from the
continuous spectrum by a combination of electric and magnetic fields.
These electrons were then deflected in a magnetic field, and the e\m ratio
0.99c.

Thomson.

was determined from the magnitude of the deflection. The results of the
experiment, which are shown in figure (1-14), are in excellent agreement
with equation (3-6) over the entire range of velocities investigated.

5.

The Photoelectric

Effect

In a cathode ray tube, electrons are emitted from a metallic electrode


bombardment of the electrode by positive
ions of the gas contained in the tube. Another process in which electrons
are emitted from the surface of a metal was discovered by Hertz in 1887.
(the cathode) as a result of the

later

form of

To pump

shown

his apparatus is

in figure (3-3).

-s-

The

glass tube

Ultraviolet light

&

III

Figure 3-3.

photoelectric

cell.

contains a polished metal electrode called a photocathode, and a second


electrode in the form of a perforated metal plate. The two electrodes are

maintained at a potential difference of a few volts (normally with the


second electrode positive with respect to the photocathode). When
ultraviolet light (frequency v

~ 10

16

sec" 1) passes through the perforated

second electrode and is incident upon the inner surface of the photocathode,
a current is observed to flow through the tube. This phenomenon is called
the photoelectric effect. The effect persists even when the tube is evacuated
to a very low pressure, implying that gaseous ions are not the carriers of

77

THE PHOTOELECTRIC EFFECT

Sec. 5]

Experiments in which a magnetic

the current.

field

was applied

to the

region between the photocathode and the second electrode indicated that
the current consisted of the flow of negatively charged particles. Thomson's
discovery of the existence of electrons suggested that the negatively charged
particles of the photoelectric effect could also be electrons which were
liberated

from the photocathode as a result of


This hypothesis was confirmed

violet light.

measured the ejm ratio of the photoelectric


was the same as that for electrons.

-2

v'max

Potential of the

Figure 3-4.

second electrode

The voltage dependence

in

bombardment by
1900 by Lenard,

particles

ultra-

who

and showed that

it

+4

+2

its

wrt.

photocathode

(volts)

of the current liberated in a photoelectric cell.

cleared up the question of the identity of the


but they also demonstrated some properties of the
which, as we shall see, were very difficult to understand

The experiments of Lenard


photoelectric particles,
photoelectric effect

in terms of the theories of classical physics.

Lenard measured the current

reaching the second electrode as a function of the potential between that


electrode and the photocathode, keeping all other parameters fixed.
His data are represented in figure (3-4). An interesting feature of these

data

is

that there

the potential

is still

some current reaching the second

of that electrode

is

electrode

when

negative with respect to the photoof course, tend to repel the negatively

A potential of this sign will,


charged photoelectrons. This must mean that the photoelectrons are
ejected from the photocathode with non-negligible kinetic energy. The
shape of the curve of current versus potential difference indicates that not

cathode.

photoelectrons have the same kinetic energy; but the well-defined


Fmax shows that there is a well-defined maximum
endpoint at V

all

kinetic energy

charge.

It

max = eFmax

was

where e

is

the magnitude of the electronic

soon suggested that the photoelectrons of

maximum

the surface of the photocathode, whereas

energy were emitted from


lower energy photoelectrons originated inside the surface and lost kinetic
energy in the process of reaching the surface. This suggestion implies that
Emax is a good measure of the energy given to an electron in the photoLenard made a series of measurements, of the type
electric process.

ELECTRONS AND QUANTA

78

[Ch. 3

indicated in figure (3-4), as a function of the intensity of the incident


ultraviolet light. He found that the photoelectric current for positive
values of V (for which photoelectrons of all kinetic energies will reach the

second electrode) was directly proportional to the light intensity. However


Vmax was observed to be independent of the light
intensity. Thus the energy E
max acquired by the photoelectrons does not
depend on the intensity of the incident light.
the cut-off potential,

6.

Classical

Theory of the Photoelectric

Effect

Let us consider what the classical theory would predict about the photoeffect.
Thomson's experiments showed that metals contain
electrons, although it was not really known whether the electrons were
electric

free to

move throughout

the metal or were bound to atoms. In either


mechanism for the ejection of these electrons by the absorption of
incident light would involve an interaction between the electric field of the
electromagnetic light waves and the electric charge of the electrons. The
oscillating electric field acting on the charged electrons would set them
into vibration with an amplitude that would be proportional to the amplitude of the electric field. Now, it is easy to show that the average kinetic
energy of any vibrating body is proportional to the square of the amplitude
of its vibration. Furthermore, the electromagnetic theory shows that the
case, the

amplitude of the oscillating

electric field is proportional to the

root of the light intensity

equation 2-6). Thus

(av. kin. energy) cc

(cf.

(amp. of vib.) 2 oc (amp. of

square

we have

elect, fid.) 2 oc

(Vlight

int.)

That is, the average kinetic energy of the vibrating electrons would be
proportional to the intensity of the incident light. This makes it difficult
to understand why the energy acquired by the photoelectrons was
observed experimentally to be independent of the light intensity.
An even more serious problem, which arises in any attempt to explain the
photoelectric effect in terms of the classical wave theory of light, is that
according to such theories a very long time is required for an electron to
absorb enough energy. The problem can be illustrated by a simple calculation.

This calculation requires an estimate of the radius of an atom. The


estimate can be obtained from a knowledge of the macroscopic density of
an element, p, its atomic weight A, and Avogadro's number N. Consider

a typical element,
gm.- 1

Now

of an atom

silver;

~ 10 gm-cm- 3 A ~ 100, also JV~ 6 x 1023


atom is AjN~ 100/6 x 1023 gm. The volume
where r is its radius. Assuming that the atoms
p

the mass of an
is

V=

fn-r

3
,


QUANTUM THEORY OF THE PHOTOELECTRIC

Sec. 7]

in the solid are closely packed, the density of an


mately equal to the density of the solid. So

mass/atom
'

or

1n
10

___"
gm-cm

atom

100/6

c=l

EFFECT

will

79

be approxi-

X 10-23 gm'1

sT

volume/atom

IQr8 cm. This result can be confirmed in several


from which we obtain r
analysis of the measured diffusion rate of a
an
from
ways. For instance,
terms
of the kinetic theory of gases, it is also
in
atoms,
gas composed of
8
10~
cm.
r
found that

us calculate the time required for photoelectrons to absorb


eFmax which is observed experimentally to be of the
the energy .Em ax
lO" 19 joules = 10~ 12 ergs. Let
order of 1.6 x 10~ 19 coulombs x 1 volt

Now

let

us assume that a source which emits energy in the form of ultraviolet


1
10 7 erg-sec- 1 is located 1
1 joule-seclight, at the rate of 1 watt
meter from the photoelectric tube. If the light is emitted with spherical
symmetry, the energy flux per cm2 at the tube is equal to the energy

cm2 to the area of a sphere of radius


X 1/4tt(102)2 cm2 ~ 102 erg-cm^-sec"1

emitted per second times the ratio of


1

that is, 10 erg-sec1 meter;


Let us assume that the electrons emitted in the photoelectric effect are
bound to atoms and that they are somehow able to absorb all the energy
incident upon the atom to which they are bound. The radius of an atom is
-8
of the order of 10 cm. Thus its cross sectional area is of the order of

10

-16

cm2 and
,

the rate at which energy

is

incident

cm =

upon

this area is

10" 14 erg-sec -1

If we
approximately 10 erg-cnr^-sec"" x
their
rate
bound
to
atoms,
were to assume that the electrons were not
can
estimate
Finally
we
be
less.
of absorption of energy would presumably
-12
that the time required for a photoelectron to absorb the observed 10
ergs is about 10 2 sec. This calculation, which is based on the assumption
that the light energy is uniformly
characteristic of the wave theory
2

10" 16

from the source


was turned on
and the emission of the first photoelectrons of about a minute. No
such time delay was observed by Lenard. In fact, experiments performed
in 1928 by Lawrence and Beams, using a light source many orders of
magnitude weaker than we assumed in the above calculation, set an
upper limit on the time delay of about 10~9 sec!
distributed over spherical

wave

fronts spreading out

predicts a delay between the time at

7.

Quantum Theory

light source

of the Photoelectric Effect

In 1905 Einstein enunciated a


effect

which the

quantum theory of

the photoelectric

which was closely related to the quantum theory of black body

80

ELECTRONS AND QUANTA

[Ch. 3

He reasoned that Planck's requirement that the energy content


of the electromagnetic waves of frequency v in a radiant source (e.g.,
the ultraviolet light source) can only be 0, or hv, or 2hv,
or nhv,
radiation.

implies that in the process of going

from energy

nhv to energy state


(n \)hv the source would emit a burst of electromagnetic energy of
energy content hv.
state

assumed that such a burst of emitted energy is initially localized


a small volume of space; and furthermore that it remains localized as it
moves away from the source with velocity c, instead of spreading out in the
Einstein

in

manner

characteristic

of moving waves. He assumed that the energy


quantum of energy, is related to itsfrequency v

content s of such a bundle, or

by the equation

He

also

assumed

hy

that in the photoelectric process one

completely absorbed by an electron

in the

of the quanta

is

photocathode.

This would impart to the electron an energy equal to hv. If this energy is
greater than the quantity &E, the electron can escape from the photocathode, where Aisis the energy which the electron must expend in reaching
the surface of the photocathode plus the energy

W required to overcome

we know must be present at the surface since


do not normally escape. The kinetic energy of the electron after

the attractive forces which


electrons

escaping from the photocathode will be equal to


electron originating at the surface,
will

have

its

maximum

is

will

E=

hv

A".

For an

be just equal to W, and

value

^max
where

AE

hv

(3-8)

a constant which depends on the composition of the photo-

cathode.

According to

this theory, the rate

of emission of photoelectrons will be

proportional to the flux of quanta incident upon the photocathode, which


in turn will be proportional to the intensity of the incident electromagnetic
radiation.

This

is

in agreement with Lenard's observation that the photo-

proportional to the intensity. As can be seen from


equation (3-8), the theory also agrees with the observation that is m a X
does not depend on the intensity. Furthermore, the theory removes the
electric current is

difficulty

concerning the time required for the photoelectrons to receive

enough energy. This difficulty arose in the wave theory because the energy
of the source was spread uniformly over the entire wave front. In the

quantum theory

the energy

concentrated in the quanta. Within a very


turned on, a quantum will strike the photowill be absorbed by one of the electrons, and the ejection

short time after the source

cathode where it
of that electron will

result.

is

is

THE COMPTON EFFECT

Sec. 8]

Einstein's theory, as

diction that the

81

summarized by equation

(3-8),

makes the

maximum kinetic energy of the photoelectrons

pre-

should be a

linear function of the frequency of the incident electromagnetic radiation.

This prediction was tested in 1916 by Millikan, who measured is m ax for


-1
14
The nature of his
radiation in the frequency range 6 to 12 x 10 sec
.

data
is

is

indicated in figure (3-5).

The experiment shows

a linear function of the frequency

v.

that .Emax actually

According to equation

(3-8), the

fci

Figure 3-5.

The frequency dependence

of the

maximum

kinetic energy of photo-

electrons.

slope of the straight line fitting the data should be equal to Planck's

constant

h,

and

its

intercept

on the

v axis should

be equal to Wjh. Thus

the photoelectric experiments, with the aid of Einstein's theory, could be

used to determine the numerical value of h. The value of h which was


obtained agreed to within better than 0.5 percent with the value of that
constant which had been determined by the completely independent method
of fitting Planck's theory to the experimental black body spectrum. This
finding represented a great triumph for the

8.

The Compton
In 1923

quantum

theories.

Effect

Compton

discovered that,

when a beam of X-rays of

well-

by sending the
radiation through a metallic foil, the scattered radiation contains a comThis
ponent of well-defined wavelength ^ which is longer than A
phenomenon is called the Compton effect. As we shall see, it provides
extremely convincing evidence for the existence of quanta. Although a
description of the production of X-rays and the measurement of their
wavelength will not be presented until Chapter 14, it is desirable to discuss

defined wavelength X

is

scattered through

an angle

the

Compton

effect at this time.

ELECTRONS AND QUANTA

82

[Ch. 3

X-rays consist of electromagnetic radiation of wavelength which is


by a factor of the order of 104 This is
proved by experiments which show that X-rays can be diffracted and
shorter than that of visible light

polarized just like light.

In Compton's experiment the wavelength A


of the incident radiation was 0.708 x 10~ 8 cm. The wavelength A was
x
observed to depend on the angle of scattering 6, but not on the material
comprising the foil. A typical set of data is shown in figure (3-6) for

45

:90

TT

\> ^1

Wavelength spectra
From A. H. Compton, Phys.

Figure 3-6.
foil.

135

of quanta scattered at various angles


Rev., 22,

from a carbon

409 (1923).

and 135. These curves are wavelength spectra of the


The abscissas are a measure of the wavelength, and
the ordinates are a measure of the intensity per unit wavelength interval.
At all angles the scattered radiation is observed to contain a component of
wavelength equal to the wavelength X of the incident radiation. The
reason for the presence of this component will be discussed in Chapter 14.
Here we shall be concerned only with the component in the scattered radiation of wavelength X v The wavelength of this component is observed
always to be greater than the wavelength of the incident radiation, and to
6

0, 45, 90,

scattered radiation.

of scattering. At 6 = 90, X x X =
x 10~8 cm. In terms of the frequency of the scattered radiation,
c\l x we see that v x is always less than the frequency of the incident

increase with increasing angle

0.024
v1

radiation v

and decreases with increasing angle of

This suggested to

Compton

scattering.

that, since v t is proportional to the

energy
of the quanta, which according to Einstein are associated with electromagnetic radiation of this frequency, the dependence of E1 on 6 is qualita-

angular dependence of the energy of a particle which


has been scattered from another particle. Pursuing this, he developed a
simple but very successful theory.
tively similar to the

Consider a quantum in the incident

For radiation of frequency

v,

beam of electromagnetic
quantum is

radiation.

the energy of the

E=

hv

(3-9)

Taking the idea of a quantum as a localized bundle of energy quite literally,

THE COMPTON EFFECT

Sec. 8]

we

shall consider

it

83

E and momentum p.

to be a particle of energy

Such a

particle must, however, have certain quite specialized properties. Consider


equations (1-20) and (1-24), which give the total relativistic energy of a
particle in terms of its rest

mass

E=w
Since the velocity of a

hv

is finite, it is

Thus

and

its

velocity v

-r

c 2 /Vl

quantum equals

apparent that the

/c

and since its energy content


mass of a quantum must be

c,

rest

quantum can be considered

to be a particle of zero rest


which is entirely kinetic. The
momentum of a quantum can be evaluated from the general relation
(1-25) between the total relativistic energy E, momentum p, and rest mass
zero.

the

mass, and of total

This

relativistic

energy

is

y+

For a quantum the second term on the

Ejc

(WoC 2)2
right

hv/c

is

zero,

h\l

and we have
(3-10)

where X is the wavelength of the electromagnetic radiation associated with


the quantum. It is quite interesting to note that Maxwell's classical wave
theory of electromagnetic radiation leads to an equation p = Ejc, where
p represents the momentum content per unit volume of the radiation and
E represents the energy content per unit volume.
Now the frequency v of the scattered radiation was observed to be
independent of the material in the
does not involve entire atoms.

due

He

to collisions

foil.

This implies that the scattering

Compton assumed

that the scattering

between the quanta and individual electrons in the

was
foil.

assumed that the electrons participating in this scattering process


and initially stationary. Some a priori justification of this
assumption can be found from considering the fact that the energy of an
X-ray quantum is several orders of magnitude greater than the energy of
an ultraviolet quantum, and from our discussion of the photoelectric
effect it is apparent that the energy of an ultraviolet quantum is comparable to the energy with which an electron is bound in a metal. Consider,
then, a collision between a quantum and a free stationary electron,
as in figure (3-7). In the diagram on the left, a quantum of total relativistic
energy E and momentum p is incident on a stationary electron of rest
mass energy m c 2 In the diagram on the right, the quantum is scattered
at an angle 6 and moves off with total relativistic energy Ex and momentum
with kinetic energy T and
p lt while the electron recoils at an angle
of momentum and total
conservation
applied
the
momentum p. Compton
also

are free

</>

AND QUANTA

ELECTRONS

84
relativistic

Momentum

energy to this collision problem.

Ch. 3

conservation

requires

= Pi cos + p cos

Po

(3-1 1)

<f>

and
sin 6

px

= p sin

(3-12)

<f>

Squaring these equations, we obtain

(p

=p

cos 8) 2

cos 2

<f>

and
p\ sin

2
2
p sin

<

o^o
Before scattering

After scattering

Figure 3-7.

collision

between a quantum and a free stationary electron.

Adding, we find

p
Conservation of total

p\

2poPl cos 6

p*

relativistic

energy requires

E +m

= E + T+ m

(3-13)

c2

(3-14)

Thus

- Ex = T
According to equation (3-10),

this is

c(Po~Pi)

=T

(3-15)

According to equations (1-22') and (1-25),

(T

+m

c 2) 2

= cy +

(w

c 2) 2

So

T2 + 27m

c2

c 2/? 2

or

T2/^ + 2Tm = p2
Evaluating/? 2 from equation (3-13), and

Oo

Pi)

+ 2

c(Po

Pi)

(3-16)

Tfrom
Pt

p\

(3-15),

we have

2PoPi cos

Sec. 9]

DUAL NATURE OF ELECTROMAGNETIC RADIATION

85

which reduces to
<fi(Po

~ Pi) = PoPitt -

or

(1 - -) =
(1 moc
pj
\pi

cos e )

cos 0)

Multiplying through by h and applying equation (3-10),

Ax

Ao

A c (l

we

obtain
(3-17)

cos 0)

where
la

=
m

0.02426

x 10

_8

cm

is

the so-called

equation.

Compton wavelength.

It predicts that

Equation (3-17)

is

the increase in wavelength (A t

scattered electromagnetic radiation depends only

on

the

Compton

A)

of the

the scattering angle

and the universal constant h/m c and is independent of the wavelength of


The equation was verified quantitatively by the
experiments of Compton and others. The recoil electrons were observed
experimentally by Bothe and by Wilson (1923). Bothe and Geiger (1925)
showed that, in the scattering of a single quantum by an electron, the
scattered quantum and the recoil electron appear simultaneously. The
energy of the recoil electrons were measured by Bless (1927) and found to
be in agreement with the predictions of the theory. These experiments, and

the incident radiation.

their interpretation in terms of the theory, provide

even stronger evidence

for the existence of quanta than does the photoelectric effect.

9.

The Dual Nature


Einstein's

of Electromagnetic Radiation

quantum theory of the photoelectric effect was very successful


phenomenon which simply could not be understood on the

in explaining a

basis of the theories of classical physics. Nevertheless, many physicists


were very reluctant to accept it because it was completely contrary to the
well-established wave nature of electromagnetic radiation. After the
discovery and interpretation of the Compton effect, the existence of quanta
could no longer be questioned. As a result physics was in what seemed at
the time to be a very uncomfortable situation. On the one hand, electromagnetic radiation manifests certain properties, such as diffraction, which
could only be explained in terms of wave motion. On the other hand,
some of its properties, such as those evident in the photoelectric and
Compton effects, could only be explained in terms of quanta which, being
localized,

have essentially the characteristics of particles. Electromagnetic

ELECTRONS AND QUANTA

86

[Ch. 3

radiation appeared to possess a split personality, sometimes behaving as

were composed of waves and sometimes behaving as


posed of particles.
if it

The experimental

if it

were com-

facts concerning the properties of electromagnetic

radiation, as well as the interpretations of these properties in terms of the


existence of both wave aspects and particle aspects, remain essentially

unchanged today. However, as a consequence of the broader point of


view provided by the development of the theory of quantum mechanics,
the attitude of physicists concerning this situation is now very different

from

their initial attitude.

The

of electromagnetic radiation
is

duality evident in the wave-particle aspects

is

no longer considered unusual because

now known to be a general characteristic of all physical entities.

more,

this duality is

it

Further-

no longer considered to represent a problem,

since

possible to reconcile the existence of both aspects with the aid of the

it is

theory of quantum mechanics.

This theory will be discussed in detail

later.

BIBLIOGRAPHY
Born, M., Atomic Physics, Blackie and Son, London, 1957.
Richtmyer, F. K., E. H. Kennard, and T. Lauritsen, Introduction
Physics,

McGraw-Hill Book Co.,

New York,

to

Modern

1955.

EXERCISES
1.

Electrons are emitted at thermal velocities from a heated cathode, are

accelerated through a voltage drop V,

and then move

through a magnetic field of strength


motion. Derive an expression relating
based upon this expression, with which

H perpendicular
V,
it

in a circle of radius

to their direction of

H, and R to e/m. Design equipment,


would be possible to make a 1 percent

measurement of e/m.
2. Prove the statement, made in section 6, that the average kinetic energy of
a vibrating system is proportional to the square of the amplitude of its vibration.
Do this for the case of a pendulum.
3.

Derive an expression for the kinetic energy of the recoiling electron in the

Compton
4.

the

scattering process.

Derive the

CM

Compton equation

for the case 8

180 by transforming to

frame, treating the collision, and then transforming back to the

LAB

frame.
5. List the experimental evidence, mentioned in section 9, for the properties
of electromagnetic radiation that can be explained only by wave motion.

CHAPTER

The Discovery
of the Atomic
Nucleus

I.

Thomson's Model of the Atom

That electrons are emitted from the metallic cathode of a cathode ray
or photoelectric tube implies that the atoms of which the cathode is
comprised contain electrons.

atoms contain

electrons.

If so,

it is

This assumption

the simple picture of a positively ionized

or

more

electrons have been removed.

reasonable to assume that


is

attractive because

atom
It

as an

it

all

leads to

atom from which one

agrees with the experimental

observation that the charge of a singly ionized

atom

is

equal in magnitude

to the charge of a single electron, or that the charge of a doubly ionized

atom

is

equal to the magnitude of the charge of two electrons,

etc.

Additional evidence for the existence of electrons in atoms was soon

obtained from the experiments of Barkla and others (1909) concerning the
scattering of X-rays by atoms. These experiments will be discussed in

Chapter 14, but it is appropriate to mention at this point that the experiments provided an estimate of Z, the number of electrons in an atom.
It was found that Z is roughly equal to A/2, where A is the chemical atomic
weight of the atom in question. Another set of experiments which provided
a measure of the number of electrons in an atom will be described in this
chapter.

Since atoms are normally neutral, they must also contain positive charge

equal in magnitude to the negative charge carried by their normal comple-

ment of electrons. Thus a neutral atom has a negative charge of magnitude


87

THE DISCOVERY OF THE ATOMIC NUCLEUS

88

[Ch. 4

Ze, where e

is the electronic charge, and also a positive charge of the same


magnitude. That the mass of an electron is very small compared to the
mass of even the lightest atom implies that most of the mass of the atom

must be associated with the

positive charge.

All these considerations naturally led to the question of the distribution


of the positive and negative charges within the atom. Thomson proposed

a tentative description, or model, of the composition of an atom, according


to which the negatively charged electrons were located within a continuous

Electron

Sphere of
positive charge

Figure 4-1.

Thomson's model of the atom.

The positive charge distribution was


assumed to be spherical in shape with a radius of the order of 10 8 cm,
which was known to be the order of magnitude of the radius of an atom.
Owing to their mutual repulsion, the electrons would be uniformly distridistribution of positive charge.

buted through the sphere of positive charge. This is illustrated in figure


(4-1). In an atom in its lowest possible energy state, the electrons would be
fixed at their equilibrium positions.

In excited atoms

material at high temperature), the electrons

equilibrium positions.

would

(e.g.,

atoms

in a

vibrate about their

Since the electromagnetic theory predicts that an

accelerated charged body, such as a vibrating electron, emits electro-

magnetic radiation, it was possible to understand qualitatively the emission


of such radiation by excited atoms on the basis of Thomson's model.

However, a calculation of the spectrum of radiation which would be


emitted showed that the model did not appear to be able to lead to quan-

agreement with the experimentally observed spectra.

titative

Really conclusive proof of the inadequacy of Thomson's model was

obtained in 1911 by Rutherford from the analysis of certain experiments


involving the scattering of alpha particles by atoms. Rutherford's analysis

showed
charge

that, instead of being spread

is

throughout the atom, the positive

concentrated in a very small region, or nucleus, at the center of the

atom. Rutherford's work comprised one of the most important developments in the subject of atomic physics, as well as the foundation of the
subject of nuclear physics.

Sec. 2]

2.

Alpha

ALPHA PARTICLES

89

Particles

Alpha particles are doubly ionized helium atoms which are spontaneously
emitted at very high velocities from several of the heaviest elements, such

U and Ra. This phenomenon will be discussed in detail later in this


book; here we shall present only a few pertinent points which are required
as background for a description of the experiments which led Rutherford
as

to the discovery of the nucleus.

Photographic

tr
\

To

Collimator

Figure 4-2.

An

pump

"Radioactive

element

apparatus used to study the types of radiation emitted by radioactive

elements.

In 1896-98 Becquerel and

Mme.

Curie discovered that several of the

heaviest elements spontaneously emit penetrating radiation capable of

Soon the radiation emitted by these


was being investigated in many laboratories.
With apparatus of the type indicated in figure (4-2), it was shown that the

blackening a photographic plate.


so-called radioactive elements

radiation consists of three separate components.

Radiation

is

emitted

from a radioactive element at the bottom of a channel in a lead block which


collimates the radiation into a well-defined beam. With no magnetic field,
a photographic plate exposed to the radiation shows, upon development,
a darkened spot at a point in line with the collimating channel.

magnetic

field is

in a direction perpendicular to the plane of the figure, three

spots are seen

When

applied to the region between the collimator and the plate,

on the

plate after

it is

developed.

The two

darkened

spots which are

deviated from the zero field position must be due to charged particles

emitted by the radioactive element, and which are deflected, by the application of the magnetic field, in a direction depending on the sign of the

The undeviated spot is presumably due to uncharged radiation


upon which the magnetic field would have no influence. The particles of
which the positively charged component of the radiation are comprised
charge.

THE DISCOVERY OF THE ATOMIC NUCLEUS

90

The

are called alpha particles.

[Ch. 4

negatively charged particles are called

beta particles, and the uncharged radiation

is

given the

name gamma

radiation.

In the experiments just described the chamber containing the collimator

and

plate is evacuated. If the experiment is repeated with the chamber at


atmospheric pressure, the spot due to the alpha particles does not appear.
Apparently a few centimeters of air is sufficient to stop the alpha particles.

Radioactive

Very

substance

thin-

walled tube

* To pump

Figure 4-3.

An

apparatus used to prove that alpha particles are doubly ionized helium

atoms.

Similar experiments establish that a foil of

some dense substance several


from reaching the plate, but

millimeters thick will stop the beta particles


that the intensity of the spot due to the

decreased only

when

collimator and the plate.

We know now

energy electrons, whereas the


radiation of the

gamma

radiation

is

significantly

several centimeters of lead are placed between the

same nature

that the beta particles are high

gamma radiation

consists of electromagnetic

as light, but of very

much

higher frequency.

The nature of the alpha particles was established in a series of measurements performed by Rutherford. Using apparatus basically similar to
that used by Thomson, he measured the charge to mass ratio of the alpha
particles as well as their velocity. These measurements showed that the
alpha particles were emitted with a well-defined velocity in the range
1.4 x 109 to 2.2 x 10 9 cm/sec, the exact value depending upon the radioactive element from which they were emitted. Note that these velocities

The charge to mass ratio


of the alpha particles was found to be one-half the value of that ratio for

are roughly one-twentieth the velocity of light.

singly ionized hydrogen atoms. Thus, if their charge was equal in


magnitude to a single electronic charge, the alpha particles would have a
mass twice the mass of hydrogen atoms. Alternatively, if the alpha
particles were assumed to have a charge twice the magnitude of an electronic
charge, their mass would be four times the mass of hydrogen atoms (i.e.,
equal to the mass of helium atoms). Rutherford was inclined toward the

THE SCATTERING OF ALPHA PARTICLES

Sec. 3]

latter interpretation since

it

would mean

91

that alpha particles were simply

doubly ionized helium atoms.

By a simple experiment, using the apparatus shown in figure (4-3),


Rutherford proved that this is indeed the case. A very thin-walled glass
tube containing a radioactive substance was surrounded by a glass
enclosure which was initially evacuated.

Some of

the alpha particles

emitted by the radioactive substance were able to penetrate the thin-

walled tube and enter the outer enclosure.

After a time, a sensitive test

of the contents of the outer enclosure showed that

amount of ordinary helium

gas.

it contained a detectable
This finding confirmed the argument that

alpha particles are doubly ionized helium atoms


electrons

which can pick up two


and become neutral helium atoms during or after their passage

through the thin-walled tubing.

3.

The Scattering

of Alpha Particles

Rutherford and his collaborators performed

many

experiments in the

course of a program designed to determine the properties of alpha particles

and of their interaction with matter. By far the most interesting of these
were the experiments concerning the scattering of alpha particles upon

-Microscope

Diaphragm

Alpha particle
source

Figure 4-4.

Thin

An

alpha particles

is

foil

alpha particle scattering experiment.

The region traversed by the

evacuated.

foils of various substances. A typical experimental


arrangement is shown in figure (4-4). The radioactive source emits alpha
particles which are collimated into a narrow parallel beam by the two
diaphragms. The beam is incident upon a foil of some substance, usually
a metal. The foil is so thin that the alpha particles pass completely through
with only a small decrease in their velocity. However, in the process of

passing through thin

traversing the

foil,

each alpha particle experiences

many

small deflections

THE DISCOVERY OF THE ATOMIC NUCLEUS

92

[Ch. 4

due to the Coulomb force acting between its charge and the positive and
negative charges of the atoms of the foil. Since the deflection of an alpha
particle in passing through a single atom depends on the details of its
trajectory through the atom, it is apparent that the net deflection in passing
through the entire foil is different for different alpha particles in the beam.
As a result, the beam emerging from the far side of the foil is divergent.
A quantitative measure of its divergent character is obtained by measuring
the

number of alpha

particles scattered into

each angular range

to

accomplished with the aid of an alpha particle detector


+
consisting of a layer of the crystalline compound ZnS and a microscope.
ZnS has the useful property of producing a small flash of light when struck

d. This

<D

by an alpha

is

particle.

If observed

by a microscope, the

flash

incidence of a single alpha particle can be distinguished.!

ment an observer counts

the

number of

light flashes

due to the

In the experi-

produced per unit

time as a function of the angular position of the detector.

4.

Predictions of Thomson's Model

The deflection of an alpha particle in following a given trajectory through


an atom can be calculated from Coulomb's law and the laws of mechanics,
if the distribution

distribution

of charges in the atom

is

given.

Thus, for the charge

assumed in Thomson's model of the atom,

it is

possible to

calculate the expected results of the experiment just described.

It

is

Alpha particle
trajectory

Figure 4-5.

*\

An

alpha particle passing through a Thomson's model atom.

apparent from the discussion above that the calculation will not be able
to predict the behavior of a single alpha particle, but only the average
behavior of a large number of alpha particles. The average behavior,
however, is precisely what is measured in the experiment.

Consider the traversal of an alpha particle through a single atom of the


type discussed in Thomson's model;
figure (4-5).

mum

We

this is indicated schematically in

shall estimate the order

of magnitude of

possible deflection angle of the alpha particle.

< max ,

the maxi-

calculation will

is used in the radium dials commonly found on wrist watches. Individual


can be seen by a dark-adapted eye with the aid of a magnifying glass.

t This effect
light flashes

The

PREDICTIONS OF THOMSON'S MODEL

Sec. 4]

use classical mechanics, which

is justifiable

93

since vjc c^ 1/20.

The

de-

assumed to be due to the Coulomb forces acting between the


charged alpha particle and the charges of the various parts of the atom.
According to Coulomb's law, the force due to some region of the atom at
a distance r from the instantaneous position of the alpha particle, and
flection is

containing charge dq,

where

(r/r) is

is

a vector of unit length directed from the location of the

charge dq to the position of the alpha particle, and where 2e


of the alpha particle.

Let us begin by evaluating the

maximum deflection

the forces due to the electrons in the atom.

would seem

At

first

is

the charge

to be expected

from

glance, equation (4-1)

to imply that the electrons could give very large deflections to

the alpha particle since

dV -*

oo as r

* 0,

so that the alpha particle

would

experience a very large force and consequently suffer a large deflection


if it

happened to pass very

close to

case because the mass of an electron

an

electron.

This

is

actually not the

compared to the mass


of the alpha particle. We leave it as an exercise for the reader to show that,
in the collision of a very massive body initially moving at velocity v,
with a stationary body of small mass, the velocity of the light body after
the collision cannot be greater than 2u.| This result follows strictly from
the conservation of momentum and energy and does not depend on the
nature of the force between the two bodies. Consequently the maximum
momentum which can be transferred to an electron is p = m(2v), where
m is the mass of an electron. This must be equal to the momentum lost
by the alpha particle in the collision. Thus Apx the maximum change in
the momentum of the alpha particle due to a collision with an electron,
so very small

is

is

To

equal to 2mv.

deflection angle

evaluate the order of magnitude of the

assume that the

max
perpendicular to the initial
in figure

mass of

<

is

change

maximum
is

directed

~ 10 -4 radians

=
4(1836)m

4(1836)

the order of magnitude of the

maximum

collision with a single electron in the

We

momentum

momentum p of the alpha particle as illustrated


2mv/Mv, where
is the
Apjpa
(4-6). Then we have < max
an alpha particle which is equal to 4 X 1836m. Consequently
<max
r

This

atom;

deflection angle

but

it

is

due to a

clear that the

assume the binding energy of the electron to the atom to be small compared
t
to the energy of the alpha particle. Then in the collision the electron will be essentially
free.

THE DISCOVERY OF THE ATOMIC NUCLEUS

94

maximum

value of the total deflection angle (due to

electrons in the atom) will be of the

same order

all

[Ch.

the collisions with

since the chance of having

a very close collision with more than one or two of the electrons

is

negligible.

To

estimate

msLX

<f>

due to the forces arising from the positive charge, we

integrate equation (4-1) over the positive charge distribution to obtain

"^max

Figure 4-6.

Illustrating

momentum

of

p a looses

the evaluation of the

momentum Apa

the total force acting on


inside the atom.

the.

maximum

deflection angle

when

a particle

alpha particle at some instant when

=m)
where

it is

This gives
(4-2)

from the position of the alpha particle at that


volume element containing the positive charge dq, and (r/r)
a vector of unit length directed from the volume element to the position
r is the distance

instant to a
is

of the alpha particle.

It is clear

that in evaluating the integral there will

be cancellation from volume elements symmetrically disposed with respect


to the position of the alpha particle. However, to estimate the order of
magnitude of the maximum total force, we ignore such cancellation and
take
Tmax

The order of magnitude of

~ a#
I

the integral can be evaluated

by writing

it

as

FmaK ~2e^jdq=2eq(^)
where

(1/r 2 ) is the

dimensions, and q
(1/r 2 )

value of 1/r 2 averaged over a region of space of atomic

= Ze

is

the total positive charge of the atom.

must be of the order of l/R2 where


,

is

Now

the radius of an atom.

Consequently

Fmax ~2Ze*/R*
To

estimate the

maximum

deflection angle

(4-3)
<

max

we

evaluate A/?a the

PREDICTIONS OF THOMSON'S MODEL

Sec. 4]

95

given to the alpha particle by the force Fmax acting for a time
of the order of At, the transit time through the atom. The time At is

momentum

approximately equal to Rjv, where v

is

the velocity of the alpha particle, so

Apx ~ .Fm3x At

~ 2Ze IRv
2

and we have

~ApJpx ~2Ze*lRMv 2

m!lx

<j>

(4-4)

yth

atom

Figure 4-7.

The

velocity vectors before and after an alpha particle traverses the

j'th

atom.

For atoms containing the

largest positive charge,

Zci

100.

The other

quantities entering into equation (4-4) have the values

e^5

10- 1C esu

<=

X 10-24 gm

C:i

R~

10- 8

10- 9 cm/sec

cm

Conseqilently
<Pmax

2
10

We

x 10 a x 2 x 10"-19
x 8 x KT 21 x 4 x 1018

10

_ 4

radians

see that the positive charge can produce only a very small deflection

+Ze is distributed uniformly over a sphere of


cm, the very high velocity alpha particles never get close
enough to a large enough quantity of charge that they experience a
Coulomb force of the intensity required to produce a large deflection.
Now that we know something about the angle through which an alpha
because, since the charge
radius

~10~8

particle is deflected in the traversal of

a single atom,

learn about the angle through which the particle

is

let

us see what

we can

deflected in the process

of traversing the large number of atoms which lie along its path through the
Clearly, summing the results of the individual deflections is involved,

foil.

sum we must

recognize the three dimensional character of

the individual deflections.

Consider figure (4-7), which indicates typical

and

in this

THE DISCOVERY OF THE ATOMIC NUCLEUS

96

[Ch. 4

v,- and vj, the velocity vectors before and after traversing
and a coordinate system useful in describing these vectors.

orientations of

the yth atom,

The coordinate system is chosen such that the x axis is in the direction of
the beam of alpha particles incident upon the foil. Because the change in
the

momentum

vector of the alpha particle in any single deflection

extremely small, the angles between

all

the velocity vectors

velocity vector remain small and, for any/, both y,

Figure 4-8.

A two

dimensional diagram representing the

y'th

and

is

and the

initial

v' are

nearly

deflection of an alpha

particle.

perpendicular to the yz plane.

Consequently, the nature of the deflection


can be specified by the angle 3 measured between the two vectors in the
plane which they define, and the angle y,- between that plane and the
z axis. This allows us to represent the deflection in terms of a two dimen</>

diagram consisting of the y and z axes and a vector 8,., as shown in


The length of the vector is equal to the magnitude of the
angle <,, and the angle between the vector and the z axis is equal to the
angle ipj. A moment's consideration will show that the relative orientation
between the velocity vector after the
deflection and the initial velocity
vector can be described in the same representation by a vector 8 where
sional

figure (4-8).

Mh
N

3=1

Thus the

entire deflection process can be represented by a two dimensional


diagram, a typical example of which is shown in figure (4-9). The net
angle of deflection resulting from
individual deflections is equal to the

4
et 10~
radians,
<f> s
the magnitude of the individual deflection angles <f> as well as the oriens
tation of the deflection planes specified by \p jt depends in an essentially

length of the vector 8.

Except for the constraint,


,

manner on both the exact trajectory of the alpha particle


through the atom and on the exact location of the charges within the atom
fortuitous

at the time the particle passes through.

random

directions

Thus the vectors 8 will have


and a distribution of lengths, as indicated in the figure.
3

PREDICTIONS OF THOMSON'S MODEL

Sec. 4]

97

The random nature of the &,-, and the fact that there are a large number
we apply the mathematical theory of statistics to the

of them, suggest that

problem of determining the length of the vector 8. Now the diagram in


figure (4-9) represents the deflections suffered by one alpha particle in its
traversal of the foil. If we were to observe the traversals of a number of
alpha particles and make similar diagrams for each of them, we could
evaluate two characteristic properties of these diagrams: the average

Figure 4-9. A two dimensional diagram representing the entire deflection process
experienced by an alpha particle in passing through a foil.

length of the vector 8, and the distribution of the lengths of this vector.

The theory of

statistics makes predictions concerning both of these


These predictions are particularly easy to obtain because we
have formulated the problem in such a way that it is identical with a well-

properties.

known

statistical

problem called the "random walk."t According to the

theory (d 2 YA the root-mean-square average length of the net deflection


vector 8, is related to the total number of individual deflections
and to
,

(dj)

lA
,

the root-mean-square average length of the individual deflection

vectors 8 3 by the equation


,

(J*f
It is interesting to

sjN{8f)

note that the average length of the vector 8 increases

because of the random directions of the 8


on the other hand, the 8 were all pointing in the same direction, which
just the opposite of being randomly directed, the average length of 8

only as the square root of


(If,

is

t See

any textbook on the theory of

statistics.

THE DISCOVERY OF THE ATOMIC NUCLEUS

98

would

be proportional to the

clearly

first

power of N.) The

[Ch. 4

theoretical

predictions of the distribution of the lengths of the vector 8 can be stated as

JT(6) dd

number of

e-**'

dd

The quantity ^V{d) dd

is

net deflection vector

within the range of values d to d

is

the

cases in which the length of the

dd;

d2

average value of the square of the length of the net deflection vectors

Jf

is

is
;

the

and

number of cases considered.


and 8, are equal, respectively, to the
angle O and the individual deflection angle p the equations

the total

Since the lengths of the vectors 8


net deflection

<f>

above can be immediately written in terms of the angles

jf(O)

rfO

(O a

The quantity

N(<&)

the angular range


is

JV

the

is

the

JO

to

is

the

= ^^ -''* dD

f = y/Nitfj*

(4-6)

number of alpha

O + JO O2 is
;

the

particles scattered within

mean

square scattering angle

mean square scattering angle in a deflection from a single atom;


number of alpha particles passing through the foil; and N\% the

number of atoms
foil.

(4-5)

traversed by an alpha particle in


These equations, plus the restriction
4>.j

<

its

passage through the

10~ 4 radians

(4-7)

that the angle of deflection in passing through a single

atom

is

very small,

comprise the predictions of Thomson's model of the atom for the scattering
of alpha particles traversing a thin foil.

5.

Comparison with Experiment


These predictions were tested by Rutherford and his collaborators. In

a typical experiment (Geiger and Marsden, 1909) alpha particles were


scattered by an Au foil 10~ 4 cm thick. The average scattering angle

was found to be about 1


2 x 10 -2 radians, and more than
99 percent of the alpha particles were scattered at angles less than about 3.
Now, the number of atoms traversed by the alpha particle is approximately
equal to the thickness of the foil divided by the diameter of an atom.
2
(4> )'^

Thus

#~ 10-

cm/10"8

cm

104

Knowing

N and

(<P) M ,

we can use

RUTHERFORD'S MODEL OF THE ATOM

Sec. 6]

99

equation (4-6) to evaluate the average deflection angle in traversing a


single atom. We obtain

,-HM
(<i?p

~2

2
= (W*
,__ ~
VN

* 10

->
2

10 *
x m-4

aradians

10

in reasonable agreement with the restriction stated in equation


Accurate measurements of ^T(O) d were difficult to obtain
because of the small angles involved. However, the available data were

This

is

(4-7).

in

agreement with equation (4-5), using

1 for

( 2) M

for angles less than

2 or 3.

On

the other hand, the angular distribution of the small

particles scattered at angles large

compared to

marked disagreement with equation

in

This equation predicts that

(4-5).

^V(<b) rfO will decrease very rapidly with increasing

we

number of

was observed to be

<J>.

For

instance, if

evaluate the predicted fraction of alpha particles scattered at angles

greater than 90,

we

find
fiso
1

JTifb

>

90)

Using the experimental value of


small

number jV{Q>

>

90)/^

>

J 90
J90"

1 for

e"*

(O 2) ^, we obtain the impressively


10" 3500 But experimentally it

90 * 2

In general, the number of


scattered alpha particles was observed to be very much larger than the
predicted number for all scattering angles greater than a few degrees.

was found that ^T(0

The

90)/^T

10-*.

existence of a small, but non-zero, probability for scattering at large

was totally inexplicable in terms of Thomson's model of the atom,


which could only lead to the multiple small angle scattering process
angles

described in section

6.

4.

Rutherford's Model of the

Atom

Experiments utilizing foils of various thicknesses showed that the


number of large angle scatterings was proportional to N, the number of
atoms traversed by the alpha particle. This is just the dependence on
that would arise if there was a small probability that an alpha particle
could be scattered through a large angle in traversing a single atom. Since
this cannot happen in Thomson's model of the atom, Rutherford proposed

a new model (1911).


In Rutherford's model of the structure of the atom, all the positive charge

THE DISCOVERY OF THE ATOMIC NUCLEUS

100

of the atom, and consequently essentially all its mass,


in a small region called the nucleus.

is

[Ch. 4

assumed

to

be

concentrated

From symmetry

considerations the nucleus was assumed to be located at

the center of thelitom, but

its

exact location played

no

role in Rutherford's

work.

The arguments of section 4

indicate that a large angle scattering

would

indeed be possible in the traversal of a single atom of such structure if the


alpha particle happened to pass very near the nucleus, provided that the

dimensions of the nucleus are small enough.

We

can get a preliminary

idea of the dimensions required from equation (4-4) since, as a review of


its

derivation will demonstrate,

it

can be used to estimate the scattering

angle of an alpha particle passing near (or through) the nucleus


interpret

to be the nuclear radius.

if

we

In the computation directly below

10~ 8 cm and obtained <


10~* radians.
equation (4-4), we used R
Consequently, to obtain a large angle scattering, </> <~ 1 radian, the value of
R must be ~104 times smaller; that is, R 10~ 12 cm. As will be indi-

cated later in this chapter, and discussed in detail in Chapter 16, 10~ 12
is, in fact, a good estimate of the radius of the atomic nucleus.

7.

cm

Predictions of Rutherford's Model

Rutherford made a detailed calculation of the angular distribution to be


expected for the scattering of alpha particles from atoms of the type
proposed in his model. The calculation was concerned only with scattering
than several degrees. Consequently, as we have seen
above, the scattering due to the atomic electrons can be ignored. The
at angles greater

scattering

is

then due to the repulsive

positively charged alpha particle

Coulomb

and the

force acting between the

positively charged nucleus.

Furthermore, the calculation considered only the scattering from heavy


atoms and thus permitted the assumption that the mass of the nucleus
is so large compared to the mass of the alpha particle that the nucleus does
not recoil (remains fixed in space) during the scattering process. It was
also assumed that the alpha particle does not actually penetrate the nuclear
region, so that the alpha particle and the nucleus act like point charges as

Coulomb force is concerned. We shall see later that all these


assumptions are quite valid except for the scattering of alpha particles
from the lighter nuclei. The calculation employs classical mechanics since
far as the

y/c~l/20.

+ze
+Ze. The nucleus is

Figure (4-10) illustrates the scattering of a particle, of charge

and mass M,

in passing near a nucleus of charge

PREDICTIONS OF RUTHERFORD'S MODEL

Sec. 7]

When

fixed at the origin of the coordinate system.

from the nucleus, the Coulomb force on

approaches the nucleus along a straight

move

After the scattering, the particle will

with constant velocity


radial coordinate r

drawn

v'.

The

line

very far
particle

with constant velocity

v.

off finally along a straight line


is specified by the
measured from an axis

position of the particle

and the polar angle

6,

which

is

parallel to the initial direction of motion.

+ ze

is

and so the

the particle

will vanish,

it

101

The perpendicular

"

1
Figure 4-10.

distance

The

+Ze
scattering of a positively charged particle in passing near a nucleus.

from that

axis to the line of initial

motion

is

specified

by the

impact parameter b. The scattering angle


is just the angle between the
axis and a line drawn through the origin parallel to the line of final motion,
and the perpendicular distance between these two lines is b'. We resolve
the velocity of the particle at any point along its trajectory into a radial
component of magnitude drjdt and a tangential component of magnitude
<f>

r (dd[dt).

Now,

the force acting on the particle

Consequently
That is,

its

angular

is

always in the radial direction.


has the constant value L.

momentum Mr 2 (ddjdt)

Mr 2 = L

(4-8)

dt
Specifically, the initial angular

momentum,

Mvb
Of

momentum

is

equal to the final angular

=L

(4-8')

or

Mv'b'

course, the kinetic energy of the particle does not remain constant

during the scattering.

However, the

energy must be equal


assumed to remain stationary.

initial kinetic

to the final kinetic energy since the nucleus

is

Thus
\Mv*>

\Mv' 2

THE DISCOVERY OF THE ATOMIC NUCLEUS

102

From

the last two equations

we

see that v

and b

v'

b',

[Ch. 4

as indicated

in figure (4-10).

Next, applying Newton's law to the radial component of motion,

we

have
,\2~|

The

left-hand term

Idt

Coulomb

the

is

(4-9)

\dtl.

force.

The

term in the bracket

first

and the second term

the radial acceleration due to the change in

r,

(radially directed) centrifugal acceleration.

To

is

the

effect the fastest solution

of this equation we transform from the coordinates


u, 6,

is

r,

6 to the coordinates

where
r

\\u

(4-10)

Then

_drdd

dr

dr du dd

dd dt'

du dd dt

_
'

dt

dr_
dt~
d

dd

dt

dd

'dr'

dd 2

we have

-dJ

The constant D,

l(Lu*\2

u\M/

dd 2

d u

_Jh?dSL _
2

d 2 u Lu 2

M dd M

d 2u

du

M dd

dd

L2 u 2

Substituting into equation (4-9),

~~

\-dt\ dt

d*r
dt

du Lu 2

M
-~ir
zZe

zZe 2u 2

M
M

zZe2
'

Mvb
2 2

(4-11)

defined by

D=

zZe 2
(4-12)

\Mv2

a convenient parameter equal to the distance of closest approach to the


nucleus in a head-on collision (b
is the distance at which the
0), since
is

potential energy zZe2 /D

is

point the particle would

equal to the

come

initial kinetic

energy

to a stop and then reverse

%Mv2

its

at this

direction of

motion. In terms of this parameter, equation (4-11) reads

d2 u\dd 2

= -

D\lb2

(4-13)

Sec. 7]

This

is

PREDICTIONS OF RUTHERFORD'S MODEL


a second order differential equation

vative) for u as

a function of

=A

6.

cos 6

is,

involving a second deri-

The general

solution to this equation is

+ B sin 6 -

D/2b*

A and

which contains two arbitrary constants,


equation (4-14)

(i.e.,

103

B.

(4-14)

We may

prove that

in fact, the solution to (4-13) by evaluating

= -A

du/dd

+ B cos 6

sin 6

and

= -A cos 6 - B sin

dhi/dd 2

and

substituting into (4-13). This gives

-A cos 6 - B sin +
To determine A and
initial conditions:

we

5,

+ B sin 6 -

cos 6

.4

-*

=i.

= - D/262

require that equation (4-14)

as r -> oo,

2>/2Z>

and

/lcosO

</r/</r

->

BsinO

Q.E.D.

conform to the
oo. Thus

v as r

>-

2b 2

or

A=
2b*

and
*!

= _ k iH = _ = _

k.

(-a

sin

+ B cos 0)

or

_ Mv _ Mv _
~ L ~ Mvb ~

Therefore the particular solution of the differential equation pertinent to


the problem at

hand

is

D=
cos a J*
+ - sin

2b

D
2b 2

or

This

is

sin 6

+ . (cos 0-1)

(4-15)

2b

the equation of a hyperbola in polar coordinates.

We

see that the

trajectory of the particle moving under the influence of a repulsive inverse


square Coulomb force is one of the conic sections. The trajectory of a
particle moving under an attractive inverse square force (e.g., a satellite

in the gravitational field of the earth) could be obtained

from the derivation


THE DISCOVERY OF THE ATOMIC NUCLEUS

104

[Ch. 4

above by simply reversing the sign of the term on the left side of equation
(4-9). The trajectory would then turn out to be one of the conic sections
exactly which one depending on the values of the parameters.

From

by

<f>

Irinfl'
sin 6'

letting r

-=--

solution to this equation

r before scattering.

The

is 0'

cos

is

figure (4-10) that for this solution

1)

the polar angle for large

is

6'

tan

sin 6'

which gives the polar angle for large

r after scattering.

<f>

6'.

It is

apparent from

Consequently

h=) = 2*

tan
is

6'

Then

oo.

This

0.

interesting solution

2b
=

which

(cos 6'

2b 2

One

we may

the equation (4-15) of the trajectory of the particle,

evaluate the scattering angle

equivalent to

=^
2D

cot^
It will

(4-16)

also be useful to evaluate R, the distance of closest approach of

the particle to the center of the nucleus (the origin).


it is

apparent that the radial coordinate

polar angle

is

for this angle,

equal to d'/2

(tt

will

</>)/2.

assume

From

figure (4-10)

this value

when

the

Evaluating equation (4-15)

and using (4-16), we obtain


1

(4-17)
sin (</2).

-> n, and that R -> b as


-> 0, as
easy to verify that R - D as
would be expected.
We see from equation (4-16) that, in the scattering of an alpha particle
by a single nucleus, if the impact parameter is in the range b to b + db
the scattering angle is in the range
to
+ dj>, where the relation between
b and
is given by the equation.
Thus the problem of calculating the
It is

(f>

</>

<f>

<f>

</>

number of alpha

particles scattered in the angular range

<f>

to

<f>

d<f>

in traversing the entire foil is equivalent to the

problem of calculating the


with impact parameter from b to b + db,

number which are incident,


upon the nuclei in the foil. Let

the number of nuclei (i.e., the number of


atoms) per cm 3 in the foil be p, and let the foil be t cm thick. Consider a
segment of the foil with a cross-sectional area of 1 cm2 as shown in figure
,

Sec. 7]

PREDICTIONS OF RUTHERFORD'S MODEL

105

This diagram indicates a ring, of inner radius b and outer radius


around an axis passing through each nucleus. The area of
drawn
b + db,
2-nb
db. The number of such rings in the segment is pt, where
is
each ring
equal to the volume of the segment. The probability
numerically
t is
P(b) db that an alpha particle will pass through one of these rings is equal
(4-11).

Incident

alpha
particles

Figure 4-11.

are very

much

A beam

upon a

of alpha particles incident

smaller and there are very

to the total area obscured

many more

cm 2

rings, as seen

by the

particles, divided by the total area of the segment.

P(b) db

pt2irb

foil.

In actuality

the rings

of them.

by the incident alpha


That is,

db

Now
u
b

D
=
cot 2

so

D
db= - ~2

\d$
sin

(#2)

and
>

bdb

cos

-~Y

(<ft/2)

S in

d<j>

= _

-g

sin

16sin

(#2)

</>

d<j>

(</2)

Thus

will

particles
But -P(b) db is equal to the probability that the incident
sign compenminus
(The
to
+
range
d<f>.
angular
the
in
scattered
be
opposite sign.) Using the
sates for the fact that db and d<f> are of the
notation employed in equation (4-5), this is
tf>

<f>

^m^
= _
jV

P{b)db

- ptD

sin 3)

sin (0/2)

106

THE DISCOVERY OF THE ATOMIC NUCLEUS

Evaluating D,

we have

8^
This

is

\1mW

[Ch. 4

sin (<t>/2)

the angular distribution derived by Rutherford for the scattering

composed of atoms with the


The quantity Jf is the number
of alpha particles incident on the foil of thickness t and p nuclei per cm3
^(O) d<& is the number scattered at angles from $ to O + c/O ze is the
charge of the alpha particle;
is its mass; v is its velocity; and Ze is
of alpha particles in passing through a

foil

nuclear structure proposed by his model.

the charge of the nuclei.

We note

that, although the angular factor in this


equation decreases rapidly with increasing angle, nevertheless the decrease
is

very

much

less rapid

than the decrease predicted by the multiple small

angle scattering angular distribution (equation 4-5).

8.

Experimental Verification and the Determination of

Detailed experimental tests of equation (4-18) were performed by Geiger

and Marsden within a few months

after its derivation (1911).

They investi-

gated the following points.


tested, using foils of Ag and Au, over
Although JV{<&) d<& varies by a factor of
about 10s over this range, the experimental data remained proportional
to the theoretical angular distribution to within a few percent.
2. The quantity jV{<&) afO was found indeed to be proportional to t
for a range of thickness of about 10 for all the elements investigated.
3. Equation (4-18) predicts that the number of scattered alpha particles
1.

The angular dependence was

the angular range 5 to 150.

will

be inversely proportional to the square of their kinetic energy,


This was tested by using alpha particles from several different

\m&.

radioactive elements.

The predicted energy dependence was observed

over an energy variation of about a factor of


4.

Finally,

the equation predicts JV($>)

(Ze) 2 the square of the nuclear charge.


,

3.

At the

to

be proportional to

time,

Z was not known for

the various atoms, except that, as mentioned in section

1,

X-ray scattering

was roughly equal to A/2, where A is the


chemical atomic weight of the atom. The validity of equation (4-18)
experiments had shown that

being assumed, the experimental data could be used to determine

was known. The resulting values agreed with the law


Z A/2. Furthermore it was observed that, to within the accuracy of
the experiment, the measured values of Z were equal to the chemical
atomic number of the atoms in question, i.e., equal to the ordering number
since everything else

THE

Sec. 9]

OF THE NUCLEUS

SIZE

of the atoms in the periodic

This implies that the

table.

in the periodic table contains

107

2 electrons, the third atom, Li, contains 3 electrons,

number of

first

atom, H,

electron, the second atom, He, contains

etc.,'

since

is

the

This was shortly confirmed by very


using X-ray techniques, which will be

electrons in the atom.

accurate determinations of
described in Chapter 14.

The Size

9.

By

of the Nucleus

evaluating R, the distance of closest approach, Rutherford was able

on the

to place limits

Ra

of
it

of the nucleus. As an example, in the scattering

size

alpha particles

(v

10 9 cm/sec)

1.6

was observed that the Rutherford

largest scattering angle investigated,

we

see that

takes

R^an

on

from Cu (A = 63, Z = 29)


law was obeyed to the

scattering
<1>

From

180.

smallest value at that angle.

its

= D=

equation (4-17)

So

zZe2
2 1V1U

*Mftn
180

V=

x 29 x

x 1CT

|(4

1-7

X 10" 12 cm

1.67

(4.8
24
)

x IP" 10)2
x

(1.6

x 109)2

Rutherford's derivation depends on the assumption that the force acting


particle is always strictly a Coulomb force between two point

on the alpha
charges.

This would not be true

region at

its

if

the particle penetrated the nuclear

point of closest approach.

the radius of the

Cu

nucleus

is

<

.7

The obvious implication

is

that

10 -12 cm.f

The equation above shows that R decreases as Z decreases. The question


arises

How much can R decrease before R <

nuclear radius ? Departures

light (low Z)
were actually observed in the scattering
nuclei, of
very
light
violation,
for
the
was
due
to
nuclei. Part of this
a
the assumption that the nuclear mass is large compared to the alpha
particle mass. However, deviations remained even after the finite nuclear
mass was taken into account in the theory. Figure (4-12) shows some data
for the scattering of alpha particles, of various velocities, at a fixed large

from the very

We

may also conclude that the use of classical mechanics to calculate the angular
t
distribution for scattering from a Coulomb field is valid, even for collisions in which

is

as small as JO

for the case of a


give only

-12

cm. In Chapter 15 we

Coulomb

shall find that

force, but that for

an approximation

quantum theory

any other force

in such collisions.

classical

verifies this

mechanics would

THE DISCOVERY OF THE ATOMIC NUCLEUS

108

[Ch. 4

angle from an Al

foil. The ordinate is the ratio of the observed number


of scattered particles to the number predicted by the Rutherford theory
(corrected for the finite nuclear mass). The abscissa is the distance of

approach calculated from equation (4-17).


12
is about 10~
cm.

closest

These data imply

that the radius of the Al nucleus

(J

1
*'-'

2?

>"x

CD

sz
ye,/

g,

0.5

X
1

R (in
Figure 4-12.

10.

Some

10
units of

12

14

16

10" 13

18

cm)

data obtained in the scattering of alpha particles from aluminum.

Problem

The detailed experimental verification of the predictions of the nuclear


model of the atom left little room for doubt concerning the validity of the
model. However, the question of the stability of such an atom presented a
serious problem.

we assume

atom are stationary, it is easy to


no stable arrangement of the electrons surrounding the
nucleus which would prevent the electrons from falling into the nucleus
under the influence of its Coulomb attraction. We cannot allow the atom
to collapse, because then its radius would be of the order of a nuclear
radius, which is four orders of magnitude smaller than what we know the
radius of an atom to be.
At first glance it appears that we can simply allow the electrons to
If

that the electrons in the

see that there exists

about the nucleus in orbits similar to the orbits of the planets


Such a system can be stable mechanically
because the centrifugal force can be made to balance the Coulomb
circulate

circulating about the sun.

attraction just as

system.

But a

it

balances the gravitational attraction in the planetary

difficulty arises in trying to carry

planetary system to the atomic system.


electrons

over this idea from the

The problem

would be constantly accelerating

in their

is

that the charged

motion around the

Ch.

EXERCISES

4]

109

nucleus and, according to the classical electromagnetic theory,


ating charged bodies radiate energy in the

all acceler-

form of electromagnetic

radia-

The energy would be emitted at the expense of the mechanical


energy of the electron, and the electron would spiral into the nucleus.
Again we have an atom which would rapidly collapse to nuclear dimention.

sions.

Furthermore, the continuous spectrum of the radiation which


is not in agreement with the discrete spectrum which

would be emitted

known

is

to be emitted

by atoms.

BIBLIOGRAPHY
Evans, R. D., The Atomic Nucleus, McGraw-Hill Book Co., New York, 1955.
Rutherford, E., J. Chadwick, and C. D. Ellis, Radiations from Radioactive
Substances, Cambridge University Press, Cambridge, England, 1930.

EXERCISES
1. Use classical mechanics to prove the statement, made in section 4, that in
a collision of a very massive body, initially moving at velocity v, with a free
stationary body of small mass, the velocity of the light body after the collision

can be no greater than

2v.

Hint:

Make

a Galilean transformation to the

system, treat the collision, then transform back to the


2.

Derive equation (4-8) from Newton's law.

3.

Adapt the

an

LAB

calculation of section 7 to the case of a particle

attractive inverse square force.

Use the

CM

system.

moving under
motion of a

results to discuss the

satellite.

4.
-<<f>

Verify from equation (4-17) that


0.

-*

as ^ -> n,

Give physical arguments explaining these

and that

-> b as

results.

5. Adapt the calculation of section 7 to the case of an electron moving in a


hydrogen atom, under the assumption that it radiates energy at the rate given
by equation (2-7) because it is accelerated. If the radius of its orbit is initially
10~ 8 cm, how long would it take for it to spiral into the nucleus of radius 1 -12 cm ?

CHAPTER

Bohr's Theory of
Atomic Structure

I.

Atomic Spectra
The

difficult

problem of the

theory of atomic structure.

of atoms actually led to a simple


key feature of this very successful theory,

stability

proposed in 1913 by Bohr, was the prediction of the spectrum of electromagnetic radiation emitted by certain atoms. Because of this prediction,
and because we have mentioned atomic spectra on two occasions in the
previous chapter,

it is

necessary at this point to describe

complex features of such


is,

spectra.

A detailed

however, best deferred until after

some of the

less

discussion of atomic spectra

we have developed

the theory of

quantum mechanics.

typical apparatus used in the

measurement of atomic spectra

is

The source consists of an electric discharge


passing through a region containing a monatomic gas. Owing to collisions
with electrons, and with each other, some of the atoms in the discharge are
indicated in figure (5-1).

put into a state in which their total energy is greater than it is in a normal
atom. In returning to their normal energy state, the atoms give up their

by emitting electromagnetic radiation. The radiation is


slit, and then it passes through a prism (or diffraction
consequently breaking up into its spectrum, which is recorded

excess energy

collimated by the
grating),

on the photographic plate.


The nature of the observed

spectra is indicated by the front view of the


photographic plate. In contrast to the continuous spectrum of electromagnetic radiation emitted, for instance, from the surface of solids at
high temperature, the electromagnetic radiation emitted by a free atom is
110

Sec.

ATOMIC SPECTRA

I]

III

concentrated at a number of discrete wavelengths. Each of these wavelength

components

is

photographic

kinds of atoms,
istic

spectrum,

which

called a line because of the line

Upon investigating the

plate.
it is
i.e.,

it

produces on the

spectra emitted

from

different

own character-

observed that each kind of atom has

its

characteristic set of wavelengths at

which the

the spectrum are found.

This feature

is

lines

of

of great practical importance

Source
Slit

Front view of
photographic plate

An

Figure 5-1.

because

it

apparatus used to measure atomic spectra.

makes spectroscopy a very

niques of chemical analysis.

useful addition to the usual tech-

Primarily as a result of this motivation,

considerable effort was devoted in the nineteenth century to the accurate

measurement of atomic

spectra.

And,

was
and in

in fact, considerable effort

required because the spectra consist of

many hundreds of

lines

general are very complicated.


Designation
of line

HR

Hy

in

X(A)

5fi
CO

"Color"

Red

Blue

H$ HgHg

Hoe

o do
* en
m
m 00

3
to

7-i

Violet

Near
ultraviolet

Figure 5-2.

The

visible part of

the hydrogen spectrum.

However, the spectrum of hydrogen is relatively simple. This is perhaps


not surprising since hydrogen, which contains just one electron, is itself
the simplest atom. The
spectrum is of considerable interest, not only
because of its simplicity, but also because of historical and theoretical
reasons which will become apparent later. Figure (5-2) represents that part

BOHR'S THEORY OF ATOMIC STRUCTURE

112

of the

H spectrum which falls approximately within

[Ch. 5

the wavelength range

The wavelength A of the lines is measured in units of 1


Angstrom = 1 A = 10-8 cm; the unit is named after a physicist who was
responsible for some of the first precise spectroscopic measurements.
of visible

light.

We

see that the spacing, in wavelengths, between adjacent lines of the


spectrum continuously decreases with decreasing wavelength of the lines,
so that the series of lines converges to the so-called series limit at 3645.6 A.
The short wavelength lines, including the series limit, are hard to observe
experimentally because of their close spacing and because they are in the
ultraviolet.

The obvious regularity of the


spectrum tempted several people to
look for an empirical formula which would represent the wavelength of
the lines. Such a formula was discovered in 1885 by Balmer. He found
that the simple equation
A

3646

where n

3 for

Ha

the wavelength of the

known

4 for

p,

(in

units)

(5-1)

5 for

etc.,

was able to predict

nine lines of the series, which were

first

at the time, to better than

all

that were

part in 1000. This discovery initiated

a search for similar empirical formulas that would apply to series of lines

which can sometimes be

identified in the complicated distribution of lines

Most of this work was done


convenient to deal with the wave

that constitute the spectra of other elements.

by Rydberg

mmber

(ca.

1890),

who found

'

it

B MX

of the lines, instead of their wavelength.

(5-2)

In terms of this quantity, the

Balmer formula can be rewritten

k
where

RK

is

2? H (l/2 2

l/ 2 ),

3, 4, 5,

(5-3)

the so-called Rydberg constant for hydrogen.

spectroscopic data,

its

value

Ru =

is

known

109677.576

From

recent

to be

0.012 cm-

(5-4)

This

is indicative of the accuracy possible in spectroscopic measurements.


Formulas of this type were found for a number of series. For instance,

we now know of the

existence of five series of lines in the

shown in table (5-1). For


same general structure:

where

is

the

alkali

H spectrum,

as

elements the series formulas are of the

Rydberg constant for the

particular element, a

and b are

Sec. 2]

BOHR'S POSTULATES

||3

TABLE

(5-1)

Names, Wavelength Ranges, and Formulas for the Hydrogen Series

Lyman

Ultraviolet

= i?H /p - -\,

n =2,3,4,..

Balmer

Near ultraviolet
and visible

= RK (

Paschen

Infrared

=RK (^--A,

n=4,5,6,..

Brackett

Infrared

= Ra (

n= 5,6,1,..

= Rs ( ^ - -% \

Pfund

Infrared

constants for the particular


particular series,

and n

is

series,

is

2%

-^

3, 4, 5,

'

- -^ \

6, 7, 8,

an integer which is fixed for the


To within about 0.05 percent

a variable integer.

the Rydberg constant has the same value for

all elements, although it does


with increasing atomic weight.
We have been discussing the emission spectrum of an atom.
closely
related property is the absorption spectrum. This may be measured with
apparatus similar to that shown in figure (5-1) except that a source

show a very

slight systematic increase

emitting a continuous spectrum

is used and a glass-walled cell, containing


monatomic gas to be investigated, is inserted somewhere between the
source and the prism. After exposure and development, the photographic
plate is found to be darkened everywhere except for a number of unexposed lines. These lines represent a set of discrete wavelength components which were missing from the otherwise continuous spectrum
incident upon the prism, and which must have been absorbed by the atoms

the

in the gas

cell. It is observed that for every line in the absorption spectrum


of an element there is a corresponding (same wavelength) line in its emission
spectrum. However, the reverse is not true. Only certain emission lines

show up

in the absorption spectrum. For hydrogen gas, normally only


corresponding to the Lyman series appear in the absorption spectrum;
but, when the gas is at very high temperatures, e.g., at the surface of a star,
lines corresponding to the Balmer series are found.
lines

2.

Bohr's Postulates
All these features of atomic spectra,

discussed,

must be explained by any

and many more which we have not

successful theory of atomic structure.

Furthermore, the very great precision which characterizes spectroscopic

BOHR'S THEORY OF ATOMIC STRUCTURE

114

[Ch. 5

measurements imposes particularly severe requirements on the accuracy


with which such a theory must be able to predict the quantitative features
of the spectra.
Nevertheless, in 1913

Bohr developed a theory which was

in accurate

quantitative agreement with certain of the spectroscopic data (e.g., the

hydrogen spectrum). This theory had the additional attraction that the
mathematics involved was very easy to understand. However, the postulates upon which it was based were not so transparent. These postulates
were:
1.

An

electron in an

atom moves

in

a circular orbit about the nucleus

under the influence of the Coulomb attraction between the electron and the
nucleus, and obeying the laws of classical mechanics.
2.

But, instead of the infinity of orbits which would be possible in classical

mechanics,
its

h, divided

only possible for an electron to

momentum L

an

is

move

an orbit for which


of Planck's constant

in

integral multiple

by 2n.

Despite the fact that

3.

in

it is

orbital angular

it is

constantly accelerating, an electron moving

such an allowed orbit does not radiate electromagnetic energy.

total energy

E remains

Thus

its

constant.

Electromagnetic radiation is emitted if an electron, initially moving


an orbit of total energy E t discontinuously changes its motion so that
it moves in an orbit of total energy E . The frequency of the emitted radiation
f
4.

in

is

equal to the quantity (Et

The

Ef

divided by Planck's constant h.

on the existence of the atomic


embodies some of the ideas concerning the stability of a
nuclear atom discussed at the end of the preceding chapter. The second
postulate introduces quantization. The reader should note the difference
between Bohr's quantization of the orbital angular momentum of an atomic
electron moving under the influence of an inverse square (Coulomb)
first

nucleus.

postulate bases Bohr's theory

It

force,

L=

nh\2iT,

1, 2, 3,

(5-6)

and Planck's quantization of the total energy of a particle, such as an


executing simple harmonic motion under the influence of a

electron,

harmonic restoring force: E


nhv, n
We shall see in the
0,1,2, ...
next section that the quantization of the orbital angular momentum of the
atomic electron does lead to the quantization of its total energy, but with
.

an energy quantization equation which is different from Planck's equation.


The third postulate removes the problem of the stability of an electron
moving in a circular orbit, due to of the emission of the electromagnetic
radiation required of the electron by classical theory, by simply postulating

BOHR'S THEORY OF THE ONE-ELECTRON

Sec. 3]

that this particular feature of the classical theory

is

ATOM

115

not valid for the case

of an atomic electron. The postulate was based on the fact that atoms are
observed by experiment to be stable

by the

classical theory.

The fourth
v

even though

this is

not predicted

postulate,

= Ei~ Et

(5 _7)

h
is

obviously closely related to Einstein's postulate (equation 3-7) that

the frequency of a

quantum of electromagnetic

radiation

is

equal to the

energy carried by the quantum divided by Planck's constant.

These postulates do a thorough job of mixing

classical and non-classical


moving in a circular orbit is assumed to obey
classical mechanics, and yet the non-classical idea of quantization of
orbital angular momentum is included. The electron is assumed to obey
one feature of classical electromagnetic theory (Coulomb's law), and yet
not to obey another feature (emission of radiation by an accelerated

physics.

The

electron

we should not be surprised if the laws of classical


which are based on our experience with macroscopic systems, are

charged body). However,


physics,

not completely valid when dealing with microscopic systems such as the

atom.

3.

Bohr's Theory of the One-Electron

Atom

The justification of Bohr's postulates, or of any set of postulates, can


be found only by comparing the predictions that can be derived from the
postulates with the results of experiment. In this section we derive some of
and compare them with the data of section 1
Consider an atom consisting of a nucleus of charge +Ze and mass M,
and a single electron of charge e and mass m. For a neutral hydrogen
atom Z = 1, for a singly ionized helium atom Z = 2, for a doubly
ionized lithium atom Z = 3, etc. We assume that the electron rotates in a
circular orbit about the nucleus. Initially we suppose the mass of the
these predictions

compared to the mass of the nucleus,


and consequently assume that the nucleus remains fixed in space. Thus
we consider at first the situation depicted in figure (5-3). The condition of
electron to be completely negligible

mechanical

stability

of the electron

is

= m2

where

(5-8)

v is the velocity of the electron in its orbit,

and

r is

the radius of the

BOHR'S THEORY OF ATOMIC STRUCTURE

116

[Ch. 5

The left side of equation (5-8) is the Coulomb force acting on the
and the right side is the centrifugal force required to keep it in

orbit.

electron,
its

Now,

circular orbit.

Figure 5-3.

An

the orbital angular

momentum

of the electron

electron moving in a circular orbit about a fixed nucleus.

must be a constant because the force acting on the electron


the radial direction.

It is

is

entirely in

equal to

L = mvr

(5-9)

Applying the quantization condition (5-6) to L, we have

mvr
or, in

nhjln,

1, 2, 3,

terms of the convenient symbol


h

mvr

h/2n
n

nh,

(5-10)
1,

2, 3,

Solving for v and substituting into equation (5-8),

Ze2

mrv2

mrn 2h2\m 2r 2

(5-11)

we obtain

= rpfpjmr

So
r

nWjmZe*,

1, 2, 3,

(5-12)

and
v

nhmZe2 lmn2 FP

= Ze \nh,
2

n=

1, 2, 3,

(5-13)

The application of the angular momentum quantization condition has


restricted the possible circular orbits to those

Note that these

of radii given by equation

of the quantum
Let us evaluate the radius of the smallest orbit (n
1) for a
hydrogen atom (Z
Inserting the known values of H, m, and e, we
1).
obtain r = 5.3 x lO" 9 cm. It is shown below that the electron has its
minimum total energy when in the orbit corresponding to n
1.
Consequently we may interpret the radius of this orbit as a measure of the
(5-12).

number

radii are proportional to the square

n.

radius of a hydrogen atom in its normal state. It is in good agreement


with the estimate, mentioned previously, that the order of magnitude of
an atomic radius is 10 -8 cm. Evaluating the orbital velocity of an electron

BOHR'S THEORY OF THE ONE-ELECTRON

Sec. 3]

ATOM

117

hydrogen atom from equation (5-13), we find


apparent from the equation that this is the
largest velocity possible for an electron in a hydrogen atom. The fact
that this velocity is less than 1 percent of the velocity of light is the justification for using classical mechanics instead of relativistic mechanics in
the Bohr theory. On the other hand, equation (5-13) shows that for large
values of Z the electron velocity becomes relativistic. The theory could not
be applied in such cases.
Next we calculate the total energy of an atomic electron moving in
one of the allowed orbits. Let us define the potential energy to be zero
when the electron is infinitely distant from the nucleus. Then the potential
energy V at any finite distance r can be obtained by integrating the energy
imparted to the electron by the Coulomb force acting from oo to r. Thus

in the smallest orbit of a

2.2

108 cm/sec.f

It is

_
-j: r

Ze2

dr

Ze2

>

The
The

potential energy

is

negative because the

Coulomb

force

attractive.

is

kinetic energy of the electron, T, can be evaluated, with the aid of

equation (5-8), to be

T=
The

= Ze*llr

total energy of the electron, E, is then

E= T+

V=

-Zc*l2r

= -T

Using equation (5-12), we have

We

-Z(*mZe*l2n*?i 2

-/wZ*e*/22 ft2

1, 2, 3, 4,

see that the quantization of the orbital angular

electron leads to a quantization of

contained in equation (5-14)


figure (5-4).
is

is

The energy of each

shown on the

left in

its

total energy.

momentum

(5-14)

of the

The information

presented as an energy level diagram in


level, as

evaluated from equation (5-14),

terms of ergs and also in terms of electron volts

5, and the quantum


shown on the right. The diagram is so constructed
that the distance from any level to the zero energy level is proportional
to the energy of that level. Note that the lowest (most negative) allowed
value of total energy occurs for the smallest quantum number n = 1.
As n increases, the total energy of the quantum state becomes less negative,

(ev),

a unit of energy which will be explained in section

number of the

with

level is

E approaching

est total

energy

is,

zero as n approaches infinity. Since the state of lowof course, the most stable state for the electron, we see

t Equation (5-13) explains why Bohr could not allow the quantum number n ever to
assume the value n = 0, as it may in Planck's quantization equation.

BOHR'S THEORY OF ATOMIC STRUCTURE

118

[Ch. 5

that the normal state of the electron in a one-electron atom is the state for
which n = 1.
Next we calculate the frequency v of the electromagnetic radiation
emitted when the electron makes a transition from the quantum state
n i to the quantum state nf that is, when an electron initially moving in an
orbit characterized by the quantum number n t discontinuously changes its
,

1.36

x 10

-12
-12

erg

= - 0.85

erg

ev

4
3

2.41

x 10

= -

1.51 ev

5.42

x 10 -12 erg = -

3.39 ev

- 21.7 x 10" 12 erg = -

13.6 ev

An energy

Figure 5-4.

diagram for the hydrogen atom.

level

motion so that it moves in an orbit characterized by quantum number nf


Using Bohr's fourth postulate (equation 5-7), and equation (5-14), we
have
.

E -E,
t

4ttH

h
In terms of the wave

number k
k

mZ (1.1)
3

1/A

vie, this is

= JL3 z *(l - 1)
\n 2f

4nch

ni

or

\n 2f

where n f and nf are

integers.

ny

4vcfi

ATOM

BOHR'S THEORY OF THE ONE-ELECTRON

Sec. 3]

119

Bohr theory are contained in equations


and (5-15). Let us first discuss the emission of electromagnetic
radiation by a one-electron Bohr atom in terms of these equations.
1. The normal state of the atom will be the state in which the electron

The

essential predictions of the

(5-14)

has the lowest energy,

the state n

i.e.,

This

1.

called the

is

ground

state.
2.

etc.

In an electric discharge the atom receives energy due to collisions,


make a transition to a state of higher

This means that the electron must

>

energy, or excited state, in which n

common

1.

all physical systems, the atom will


and return to the ground state. This is accomplished
by a series of transitions in which the electron drops to states of successively
lower energy, finally reaching the ground state. In each transition electromagnetic radiation is emitted with a wave number which depends on the
energy lost by the electron, i.e., upon the initial and final quantum
3.

emit

Obeying the

its

tendency of

excess energy

numbers. In a typical case, the electron might be excited into state n 1


and drop successively through the states n = 4 and n = 3 to the ground
state n \. Three lines of the atomic spectrum are emitted with wave
numbers given by equation (5-15) for n i = 1 and nf 4, n t = 4 and

3,

4.

In the very large

nf

and n i

3>

and nf

1.

number of

excitation

and de-excitation processes

which take place during a measurement of an atomic spectrum, all possible


transitions occur and the complete spectrum is emitted. The wave numbers
of the set of lines which constitute the spectrum are given by equation
(5-15), where we allow n t and n t to take on all possible integral values
subject only to the restriction that n i

>

nf

For hydrogen (Z = 1) let us consider the subset of


arises from transitions in which n t = 2. According
the wave numbers for these lines are
k

= RMIn 2, -

1/nf),

n,

2,n t

spectral lines

which

to equation (5-15)

>

nt

or
2

RJ1/2

1/n

2
),

3, 4, 5, 6,

This is identical with the series formula for the Balmer series of the hydrogen
spectrum (equation 5-3), if R x is equal to i? H According to the Bohr
tn^lAnct?. Although the numerical values of some of the
theory, R
.

quantities entering into this equation

were not very accurately known at the

Bohr evaluated R x in terms of these quantities and found that the


resulting number was in quite good agreement with the experimental
time,

In the next section we shall make a detailed comparison,


using recent data, between the experimental value of i? H and the prediction
of Bohr's theory, and show that the two agree almost perfectly.

value of

Ru

BOHR'S THEORY OF ATOMIC STRUCTURE

120

[Ch. S

According to the Bohr theory, each of the five known series of the
hydrogen spectrum arises from a subset of transitions in which the electron

For the Lyman series n t = 1


Balmer n t = 2; for the Paschen n, = 3; for the Brackett n f = 4;
and for the Pfund nf = 5. These series are conveniently illustrated in
terms of the energy level diagram of figure (5-5). The transition giving
rise to a particular line of a series is indicated in this diagram by an arrow
goes to a certain final quantum state n t

for the

Balmer

-n =

Lyman
Figure 5-5.

Illustrating the

production of the

five

known

series of the

hydrogen

spectrum.

going from the initial quantum state n t to the final quantum state n
t
Only the arrows corresponding to the first few lines of each series are
shown. Since the distance between any two energy levels in such a diagram
.

proportional to the difference between the energy of the two

levels, and
wave number k is
proportional to the energy difference, the length of any arrow is proportional to the frequency or wave number of the corresponding spectral
is

since equation (5-7) states that the frequency v or

line.

The wave numbers of the lines of all these series are fitted very accurately
by equation (5-15) by using the appropriate value of n f This was a great
triumph for Bohr's theory. The success of the theory was particularly
impressive because the Lyman, Brackett, and Pfund series had not been
discovered at the time the theory was developed by Bohr. The existence of
these series was predicted by the theory, and the series were soon observed
experimentally by the gentlemen they are named after.
The theory worked equally well when applied to the case of oneelectron atoms with Z = 2, i.e., singly ionized helium atoms He+. Such
atoms can be produced by passing a particularly violent electric discharge
.

BOHR'S THEORY OF THE ONE-ELECTRON

Sec. 3]

ATOM

121

through normal helium gas. They make their presence apparent


by emitting a simpler spectrum than that emitted by normal helium atoms.
In fact, the atomic spectrum of He+ is exactly the same as the hydrogen
spectrum except that the wave numbers of all the lines are almost exactly
four times as great. This is explained very easily, in terms of the Bohr

(a spark)

theory,

The

by

setting

Z2 =

4 in equation (5-15).

properties of the absorption spectrum of one-electron atoms are

Bohr theory. Since the atomic


of one of the
the atom can only absorb discrete amounts of

also easy to understand in terms of the

electron

must have a

allowed energy

total energy exactly equal to the energy

states,

energy from the incident electromagnetic radiation. This fact leads to the
idea that we consider the incident radiation to be a beam of quanta, and
that only those quanta can be absorbed whose frequency is given by

E=

where E is one of the discrete amounts of energy which can be


absorbed by the atom. The process of absorbing electromagnetic radiation
is then just the inverse of the normal emission process, and the lines of the
absorption spectrum will have exactly the same wave numbers as the lines
of the emission spectrum. Normally the atom is always initially in the
ground state n = 1, so that only absorption processes from n = 1 to
n > 1 can occur. Thus only the absorption lines which correspond (for
hydrogen) to the Lyman series will normally be observed. However, if the
hv,

gas containing the absorbing atoms

owing

to collisions,

at a very high temperature, then,

is

some of the atoms

will initially

be in the

first

excited

corresponding to the Balmer series will


2,
be observed. The required temperature can be estimated from the Boltzmann probability distribution (equation 2-23), which shows that the ratio
state

and absorption

lines

of the probability of finding an atom in the state n


of finding it in the ground state n = 1 is

2 to the probability

p -En = zlkT

En = dkT

Boltzmann's constant = 1.38 x 10" 16 ergs/K. For hydrogen


the energies of these two states are given in the energy level diagram (figure

where k
5_4):

is

En=1

21.7

pw~2 =

x 10- 12

ergs,

En=2 =

-( -5.42 + 21.7) xlo~

(1.38 x

la

-5.42 X 1(H2

Thus

-1.18 x 10 "K

ergs

NT 18 ergs/'-Kjr

ergs.

Pn=l
apparent that a significant fraction of the hydrogen atoms will
initially be in the n = 2 state only when T is of the order of, or greater
than, 10 5 K. As mentioned before, Balmer absorption lines are actually
observed in the hydrogen gas of some stellar atmospheres. Thus we have
It is

one way of estimating the temperature of the surface of a

star.

BOHR'S THEORY OF ATOMIC STRUCTURE

122

4.

[Ch. 5

Correction for Finite Nuclear Mass

We have assumed the mass of the atomic nucleus to be infinitely large


compared to the mass of the atomic electron, so that the nucleus remains
fixed in space. This is a good approximation even for hydrogen, which
contains the lightest nucleus, since the mass of that nucleus is about 2000
times larger than the electron mass. However, the spectroscopic data are

CM

The center of mass of an electron-nucleus system.

Figure 5-6.

so very accurate that before

we make a detailed numerical comparison


we must take into account the fact that

of these data with the Bohr theory


the nuclear mass

is

actually finite.

Consider the electron-nucleus system shown in figure (5-6). The electron


of mass m and the nucleus of mass
must both move so that the center
of mass (CM) remains fixed in space. Let r be the distance between the

nucleus and the electron.

CM

is

Then the

m+
If the electron rotates

must also

about the

rotate about the

and

from the nucleus to the

distance x

determined by the equation m{r

x)

Mx;

so

m+

(5-16)

CM with angular velocity

co,

the nucleus

CM

with the same angular velocity. Let us


calculate the total orbital angular momentum L of such a system. We
have

L = m(r

2
= mM r co
(m + Mf

2
Marco

Mm r

2 2

xfco

o)

(-

+ Mf

(m

So

mM(m
,
+ M)
i
L ro^(o =
L=
(m

As

Mjm -

+ Mf

mM
m+

L -> {mMjM)r 2 m and

._

._.

(5-17)

x -- 0. Thus, for an infinitely


heavy nucleus the electron rotates about the stationary nucleus with orbital
angular momentum L = mr 2 co. This is just what we had before in
oo,

equation (5-9), since

rco is

equal to the orbital velocity

v.

CORRECTION FOR

Sec. 4]

To handle

NUCLEAR MASS

FINITE

Bohr modified

the case at hand,

his

123

second postulate to read

Instead of the infinity of orbits which would be possible in classical


it is only possible for an electron to move in an orbit for which

mechanics,

the total orbital angular

Planck's constant

h,

momentum of the atom,

divided by

L,

is

an integral multiple of

2tt.

This obviously reduces to the original statement of the postulate for


an infinitely heavy nucleus discussed in the previous section. Applying

we have

this postulate to equation (5-17),

mM
m+
Comparing

mM

this

m+

vr

with equation (5-1

fi

we

1),

replace the mass of the electron, m,

nh,

1, 2, 3,

see that the net effect has been to

by the

so-called reduced mass,

-f

(5-18)

m by the factor 1/(1 + m/M).


go through the rest of Bohr's derivation for the case
of finite nuclear mass. It is found that all the equations are identical with
those derived before except that the electron mass m is replaced by the
reduced electron mass (x. In particular, the formula for the wave numbers
of the spectral lines becomes
which

It is

is

than

slightly less

not

difficult to

lc

mM
^s
~m --j^
+ MAirc&

>(L-L)
-= R MZ*[9
9
?/'

RM -

),

\nf

(5-19)

R M is the Rydberg constant for a nucleus of mass M. As M/m * co,


apparent that R M > R^, the Rydberg constant for an infinitely heavy

where
it is

nucleus which appears in equation (5-15).

constant

RM

than

is less

R^ by

extreme case of hydrogen,

than

R x by
we

about

In general, the Rydberg

the factor 1/(1

M/m =

1836

(see

mjM). For the most


3) and R H is less

Chapter

part in 2000.

Rn

from equation (5-19), using the currently accepted


-1
values of the quantities m, M, e, c, and h, we find R H = 109681 cm
R
given
in
equation
this
with
the
experimental
value
of
(5-4),
Comparing
H
we see that the Bohr theory, corrected for finite nuclear mass, agrees with
If

evaluate

the spectroscopic data to within 3 parts in 100,000!

worth while to present another demonboth the spectroscopic data and the
Bohr theory. This involves a comparison of the spectrum of hydrogen,
H, and the spectrum of singly ionized helium, He+. As mentioned above,
the lines of these two spectra are observed to differ in wave number by a
Before closing this section,

it is

stration of the numerical accuracy of

BOHR'S THEORY OF ATOMIC STRUCTURE

124

factor

F which has

a value very close to

4.

[Ch. 5

According to equation (5-19),

the factor should be

mMse+

MHe+ _ 4 MHe + (m + MH
mAfH
MH (m + MHe+)
m + MH

m
F _ ^2 +

The mass of the helium nucleus is just the mass of the helium atom less two
electron masses.f That

is,

MHe+ = MHe
and the mass of the hydrogen nucleus
less one electron mass.f So

atom

is

- 2W

equal to the mass of the hydrogen

atom

MB MH
From

the study of chemistry

atom

we have an

accurate value of the ratio

Mm atom _

4.004

Mh atom

1.0081

The four equations above, plus the experimental value of the factor F,
allow an evaluation of
n /m. The result is
n jm = 1837, in very good

agreement with the value 1836 obtained in the experiments discussed in


Chapter 3.

Atomic Energy States

5.

The Bohr theory predicts that the total energy of an atomic electron is
For example, equation (5-14) gives the allowed energy values

quantized.

for the electron in a one-electron atom.

Although we have not attempted

to derive similar expressions for the electrons in a multi-electron atom,

clear that the total energy of each of the electrons will also be quantized
and, consequently, that the total energy content of the atom must be
it is

restricted to certain discrete values.

Thus the theory

predicts that the

atomic energy states are quantized. The correctness of this prediction was
confirmed in a simple experiment performed by Franck and Hertz in 1914,
less than one year after the announcement of Bohr's theory.
Apparatus of the type used by these investigators is indicated in figure
(5-7).
t

We

Electrons are emitted thermally at low energy from the heated


ignore the relativistic contribution to the atomic mass, which arises from the
have a negative total energy of the order of 10- 11 ergs, since this
10- u /10"
10" 32 gm is small compared to an electron mass.
/c 2

fact that the electrons

mass

M=

ATOMIC ENERGY STATES

Sec. 5]

125

cathode C.f They are accelerated to the anode A by a potential V applied


between the two electrodes. Some of the electrons pass through holes in
A and travel to plate P, providing their kinetic energy upon leaving A is
enough to overcome a small retarding potential Vr applied between P

and A
atoms

The

entire tube

current reaching

at a low pressure with a gas or vapor of the


The experiment involves measuring the electron

is filled

to be investigated.

P (indicated

by the current / flowing through the meter)

as a function of the accelerating voltage

-V

Heater

An

Figure 5-7.

V.

apparatus used to prove that atomic energy states are quantized.

The first experiment was performed with the tube containing Hg vapor.
The results are indicated in figure (5-8). At low accelerating voltage, the

When V
is observed to increase with increasing voltage V.
reaches 4.9 volts, the current abruptly drops. This was interpreted as
indicating that some interaction between the electrons and the Hg atoms
suddenly commences when the electrons attain a kinetic energy of 4.9

current /

10~ 12 ergs, as a result of falling through a potential difference


of 4.9 volts.+ Apparently a significant fraction of the electrons of this
energy excite the Hg atoms and in so doing entirely lose their kinetic
energy. If V is only slightly more than 4.9 volts, the excitation process

ev

7.85

must occur just

in front of the

anode A, and

after the process the electrons

point at which the thermal agitation


t By raising the temperature of the cathode to the
required to overcome the
kinetic energy of its electrons is comparable to the energy
attractive forces acting on the electrons at the surface of the cathode, some of the

electrons will be spontaneously emitted.


%

An

electron falling

kinetic energy

unimpeded through a potential difference of 1 volt attains a


19
joules =
1.60 X 10" 19 coulomb x 1 volt = 1.60 x 10~

E = eV =

10~ 12 ergs
ergs. In the present case the electron energy is 4.9 x 1.60 x
12
volt (ev)
electron
define
the
energy,
we
10~
unit
of
convenient
For
a
very
ergs.
7.85 x
1.60 x 10~ 12 ergs. This unit is used extensively in atomic and nuclear
such that 1 ev
physics. The kinetic energy of the electron discussed above can be specified as 4.9 ev or
12
as 7.85 x 10~ ergs.
1.60

x 10" 12

BOHR'S THEORY OF ATOMIC STRUCTURE

126

[Ch. 5

cannot gain enough kinetic energy in falling toward A to overcome the


retarding potential Vr and reach plate P. At somewhat larger V, the
electrons can gain enough kinetic energy after the excitation process to
overcome Vr and reach P. The sharpness of the break in the curve indicates that electrons of energy less than 4.9 ev are not able to transfer their
energy to an Hg atom. This interpretation is consistent with the existence
of discrete energy states for the Hg atom, as predicted by the Bohr theory.

10

V (volts)
The voltage dependence

Figure 5-8.

of the current measured

the Franck-Hertz

in

experiment.

Assuming the first excited state of Hg to be 4.9 ev higher in energy than


the ground state, an Hg atom would simply not be able to accept energy
from the bombarding electrons unless these electrons had at least 4.9 ev.
Now, if the separation between the ground state and the first excited state

actually 4.9 ev
7.85 X 10~ 12
spectrum of wave numberf

is

ch

7.85

3.00

10

10

ergs, there

should be a line in the

x 1(T 12

6.62

KT 27

3.95

10*

Hg

cm -1

corresponding to the transition from the

first excited state to the ground


Franck and Hertz found a line in the spectrum of very nearly this
wave number. Furthermore, they found that when the energy of the
bombarding electrons is less than 4.9 ev no spectral lines at all are emitted
from the Hg vapor in the tube, and when the energy is not more than a
few electron volts greater than this value only the single line k = 3.95

state.

104

cm-1

is

seen in the spectrum.

what would be expected

if

All these observations are precisely

the energy of the

Hg atom

is

quantized.

The

Franck-Hertz experiment provided a striking confirmation of the predictions of the Bohr theory. It also provided a method for the direct
measurement of the energy differences between the quantum states of an
t In evaluating this

we use

quantities in the cgs system

E=

7.85

x 10-12

ergs since

and want the answer

we have

in that system.

expressed

all

other

ATOMIC ENERGY STATES

Sec. 5]

127

the answers appear on the dial of a voltmeter

atom

/ versus

Some

V is

When

the curve of

extended to higher voltages, additional breaks are found.

are due to electrons exciting the

first

excited state of the

C to A

atoms on

but some are due to


excitation of the higher excited states and, from the position of these
breaks, the energy differences between the higher excited states and the

several separate occasions in their trip

ground

state

from

can be directly measured.

EHgjnna

2nd excited
)

state

1st excited state


.

>

>

a)

a*

r**

^J-

to

'

'

'

Ground
Figure 5-9.

state

considerably simplified energy level diagram for mercury.

Another experimental method of determining the separations between


atom is to measure its atomic spectrum and then
empirically construct a set of energy states which would lead to such a
spectrum. In practice this is often quite difficult to do since the set of
the energy states of an

lines constituting the spectrum, as well as the set

very complicated. However, in


it is

of energy

states, is often

common with all spectroscopic techniques,

a very accurate method. In

all

cases in which determinations of the

separations between the energy states of a certain

atom have been made

using both this technique and the Franck-Hertz technique, the results
have been found to be in excellent agreement.

In order to illustrate the preceding discussion, we show in figure (5-9)


a considerably simplified representation of the energy states of Hg in
terms of an energy level diagram. The separations between the ground

and the first and second excited states are known, from the FranckHertz experiment, to be 4.9 ev and 6.7 ev. These numbers can be confirmed,
and in fact determined with much higher accuracy, by measuring the wave
numbers of the two spectral lines corresponding to transitions of an
state

electron in the

Hg atom from

these

two

states to the

ground

state.

The

BOHR'S THEORY OF ATOMIC STRUCTURE

128

[Ch. 5

= 10.4 ev,

of the ground state relative to a state of zero total


not determined by the Franck-Hertz experiment. However, it
can be found by measuring the wave number of the line corresponding to a
transition of an atomic electron from a state of zero total energy to the
ground state. This is the series limit of the series terminating on the ground
state. The energy e can also be measured by measuring the energy which
energy

energy,

is

must be supplied to an Hg atom in order to send one of its electrons from


the ground state to a state of zero total energy. Since an electron of zero
total energy is no longer bound to the atom, s is the energy required to
ionize the atom and is therefore called the ionization energy.
Lying above the highest discrete state at E =
are the energy states
of the system consisting of an unbound electron plus an ionized Hg atom.
The total energy of an unbound electron (a free electron with E > 0) is not
quantized. Thus any energy E >
is possible for the electron, and the
energy states form a continuum. The electron can be excited from its
ground state to a continuum state if the Hg atom receives an energy greater
than 10.4 ev. Conversely, it is possible for an ionized Hg atom to capture
a free electron into one of the quantized energy states of the neutral atom.
In this process, radiation of wave number greater than the series limit

corresponding to that state will be emitted. The exact value of the wave
number depends on the initial energy E of the free electron. Since E
can have any value, the spectrum of Hg should have a continuum extending

beyond every series limit in the direction of increasing wave number. This
can actually be seen experimentally, although with some difficulty. These
comments concerning the continuum of energy states for E > 0, and its
consequences, have been
equally true for

6.

all

made

in reference to the

Hg

atom, but they are

atoms.

The Wilson-Sommerfeld Quantization Rules


The

success of the

Bohr

theory, as measured

by its agreement with


But it only accentuated the
mysterious nature of the postulates on which the theory was based. One
of the biggest of these mysteries was the question of the relation between
Bohr's quantization of the angular momentum of an electron moving in a
circular orbit and Planck's quantization of the total energy of an entity,
such as an electron, executing simple harmonic motion. In 1916 some light
was shed upon this by Wilson and Sommerfeld, who enunciated a set of
rules for the quantization of any physical system for which the coordinates
are periodic functions of time. These rules included both the Planck and
the Bohr quantization as special cases. They were also of considerable
experiment, was certainly very striking.

THE WILSON-SOMMERFELD QUANTIZATION RULES

Sec. 6]

use in broadening the range of applicability of the


rules

quantum

129

theory. These

can be stated as follows

For any physical system in which the coordinates are periodic functions
of time, there exists a quantum condition for each coordinate. These

quantum conditions are

j>pt dq

= nji

(5-20)

one of the coordinates, pa is the momentum associated with that


coordinate, n q is a quantum number which takes on integral values, and
means that the integration is taken over one period of the coordinate q.

where q

is

Figure 5-10.

The time dependence of the coordinates of an electron

The meaning of these

rules

can best be

in

illustrated in terms of

Bohr

some

orbit.

specific

examples.

Consider a particle of mass

m moving with constant angular velocity a>


Bohr

in a circular orbit of radius r (an atomic electron in a

orbit).

The

position of the particle can be specified by the polar coordinates r and


0.
The behavior of these two coordinates is shown in figure (5-10) as

functions of time.
r

They

are both periodic functions of time, if we consider

to be a limiting case of this behavior.

The momentum

associated

mr2 dd/dt.
with the angular coordinate 6 is the angular momentum L
The momentum associated with the radial coordinate r is the radial

momentum pT
dd/dt

= m dr/dt.

w, a constant.

In the present case r

Thus L

mrgco and

pr

=
0.

= 0, and
do not need to

dr/dt

We

apply equation (5-20) to the radial coordinate r in the limiting case in


which the coordinate is constant. The application of the equation to the
is easy to carry through for the present example.
angular coordinate
n; then
andp
L, a constant. Write n a
We have q d
Q

P, dq

= i Ld6 =

L$>dd

So the condition

Pa

dq

n ah

L\

"dd

2ttL

BOHR'S THEORY OF ATOMIC STRUCTURE

130

[Ch. 5

becomes

2ttL

nh

or

L=

= nh

nhjltr

which is identical with Bohr's quantization law.


Next let us apply the Wilson-Sommerfeld quantization rules to a particle
of mass m executing simple harmonic motion with frequency v. The

-*o

Figure 5-1

The time dependence

1.

of the coordinate of a simple harmonic oscillator.

position of the particle can be specified by the single linear coordinate x.

The behavior of this coordinate with time


and can be expressed as

=x

where x

is

sin

2-TTVt

sin

is

illustrated in figure (5-11)

m=

cot,

2ttv

The momentum of the particle

the amplitude of the oscillation.

is

dx
= m
= mx a> cos cot
dt

In this case

pa dq

<p

To

evaluate this integral

it is

<p

p dx

(p

mx

a)

cos

cot

dx

convenient to express cos

x^sin

cot

(l

cos

cot

in terms of x:

cot)

Then
cos^

cot

cos

cot

2
= V* -

xa

and

p dx

tnx

v*o

co

dx

into

PV*

Jo

The

last step

depends upon the fact that

'O

'ito

Jo

J -x

Jo

x?

dx

SOMMERFELD'S RELATIVISTIC THEORY

Sec. 7]

because the integrand

<J>

an even function of x. This gives

is

p dx

131

mco 4

sin
2

(p

p dx

mco 2{x% sin

p dx

mx\ am

-1

We would like to write this in terms


it is

is

easy to do

equal to

its

if

we

E = -m\
since

when

when x

dth=o
0.

x\ sin

a-o-J o

-1
0)

of the total energy

recall that the total

kinetic energy

m(x

0.

co

E of the particle;

energy of a harmonic oscillator

So

cos

ci>0?=

- mx\co %

Thus we have

p dx

2tt
= E
= ^1 mxlcoi2
,

and
Ejv

which

7.

is

nQh

nh,

E=

nhv

identical with Planck's quantization law.

Sommerfeld's Reiativistic Theory

One of

the important applications of the Wilson-Sommerfeld quanti-

zation rules

was to the case of a hydrogen atom

in

which

it

was assumed

This was done by


Sommerfeld in an attempt to explain the fine structure of the hydrogen
spectrum. The fine structure is a splitting of the spectral lines, into several
distinct components, which is found in all atomic spectra. It can be
that the electron could

move

in elliptical orbits.

observed only by using equipment of very high resolution since the sepanumber, between adjacent components of a single
4
spectral line is of the order of 10~ times the separation between adjacent

ration, in terms of wave

lines. According to the Bohr theory, this must mean that what we had
thought was a single energy state of the hydrogen atom actually consists
of several states which are very close together in energy.
Sommerfeld first evaluated the size and shape of the allowed elliptical

BOHR'S THEORY OF ATOMIC STRUCTURE

132

orbits, as well as the total

[Ch. S

energy of an electron moving in such an orbit,

using the formulas of classical mechanics. Describing the motion in terms

of the polar coordinates r and

0,

The first condition


momentum,

yields the

L=

he applied the two quantum conditions

LdB =

nh

pr dr

nr h

same

restriction

n e h,

ng

1, 2, 3,

on the

orbital angular

that it does for the circular orbit theory. The second condition (which was
not applicable in the limiting case of purely circular orbits) leads to the

following relation between

and

ajb, the ratio

of the semimajor axis to

the semiminor axis of the ellipse

L{a\b

By applying

1)

nr

nft,

0, 1, 2, 3,

the condition of mechanical stability analogous to equation

From

(5-8), a third equation is obtained.

these equations

Sommerfeld

evaluated a and b, which give the size and shape of the elliptical orbits,

and

also the total energy

E of an electron in such an orbit.

n A

The results

2
<.

fiZe

= a^

are

(5-21)

2t?h%

where

/j,

the reduced mass of the electron, and where the

is

number n

n
Since n e

quantum

defined by

is

1, 2, 3,

and n r
n

=
=

nB

nr

0, 1, 2, 3,

1, 2, 3, 4,

n can take on the values

For a given value of n, n e can assume only the values


n

1, 2, 3,

SOMMERFELD'S RELATIVISTIC THEORY

Sec. 7]

The

integer n

is

called the principal

133

quantum number, and ng

is

called the

azimuthal quantum number.


(5-21) shows that the shape of the orbit (the
semimajor to the semiminor axes) is determined by the ratio of

The second of equations


ratio of the

n, the orbits are circles of radius a. Note that the


n e to n. For n e
equation giving a in terms of n is identical with equation (5-12), the equation giving the radius of the circular Bohr orbits.f Figure (5-12) shows,

,n

ne =

{2Zis- n

=l

re

n=2

Figure 512.

re

Some

elliptical

Bohr-Sommerfeld

to scale, the possible orbits corresponding to the

first

orbits.

three values of the

quantum number. Corresponding to each value of the principal


quantum number n there are n different allowed orbits. One of these, the
circular orbit, is just the orbit described by the original Bohr theory. The
others are elliptical. But, despite the very different paths followed by an
electron moving in the different possible orbits for a given n, the third of
principal

equations (5-21)

us that the total energy of the electron

tells

is

the same.

of the electron depends only on n. The several orbits


characterized by a common value of n are said to be degenerate.
This degeneracy in the total energy of an electron, following the orbits
of very different shape but common n, is the result of a very delicate balance
between potential and kinetic energy which is characteristic of treating an
inverse square law of force by the methods of classical mechanics.

The

total energy

Sommerfeld removed the degeneracy by treating the problem with the


theory of

In the discussion following equation


-2
an electron in a hydrogen atom, vjc ca 10 or

mechanics.

relativistic

we showed
Thus we would expect the
that, for

(5-13)
less.

due to the

relativistic variation

relativistic corrections to the total energy,


of the electron mass, which will be of the

order of (v/c) 2 to be only of the order of IQr*. However, this is just the
order of magnitude of the splitting in the energy states of hydrogen that
would be needed to explain the fine structure of the hydrogen spectrum.
,

The

actual size of the correction depends

calculation which
t

is

much

on the average

velocity of the

After a
too tedious to reproduce here, Sommerfeld

electron which, in turn, depends

on

the ellipticity of the orbit.

Remember that equation (5-12) will have m replaced by p if proper account is taken

of the

finite

nuclear mass.

BOHR'S THEORY OF ATOMIC STRUCTURE

134

showed that the total energy of an electron in an


quantum numbers n and ne is equal to

-&.

+ ^(i

Irftf

The quantity a = e%\hc is a pure number


The numerical value of this constant is
a

*\hc

\n B

orbit characterized

(5 _22)

1)1
4n/J

(5-23)

1/137

^n =

1
1

-n =
-n =

"*

Figure 5-13.

The

splitting

The
is

<

=4

n=
n=

r-l

n=

fine structure splitting of

greatly exaggerated.

The

by the

called the fine structure constant.

X 10"s ~

7.297

[Ch. 5

some energy

3, rig

3,

ne = 2

3, tig

2,

ne = 2
n = 1

1,

n =

2,

ft

fl

hydrogen atom.
between the levels are

levels of the

possible transitions

indicated by solid arrows.

In figure (5-13)

atom

in terms of

we

several levels with a

the sake of clarity.


states

represent the

an energy

level

common

Arrows

which produce the

first

few energy

diagram.

states

of the hydrogen

The separation between

the
value of n has been greatly exaggerated for

indicate transitions between the various energy

lines

of the atomic spectrum. Lines correspond-

ing to the transitions represented by the solid arrows are observed in the
hydrogen spectrum. The wave numbers of these lines are in very good

agreement with the predictions derived from equation (5-22). However,


the lines corresponding to the transitions represented by dashed arrows
are not found in the spectrum. The transitions concerned do not take
place. Inspection of the figure will demonstrate that transitions only
occur if
n 6i
This

is

called a selection rule.

-,,= 1

(5-24)

THE CORRESPONDENCE PRINCIPLE

Sec. 8]

8.

The Correspondence
The reason

Principle

for the existence of selection rules

we have described so far. However, with


postulate known as the correspondence principle, a
theory

of these rules could sometimes be found.

Bohr

in 1923, consists of

135

is

not contained in the

the aid of an auxiliary


theoretical justification

This principle, enunciated by

two parts:

1. The predictions of the quantum theory for the behavior of any physical
system must correspond to the predictions of classical physics in the limit
in which the quantum numbers specifying the state of the system become very

large.

A selection rule holds true over the entire range of the quantum number

2.

Thus any selection rules which are necessary

concerned.

to obtain the

required correspondence in the classical limit (large n) also apply in the

quantum

limit (small n).

Concerning the first part, it is obvious that the quantum theory must
correspond to the classical theory in the limit in which the system behaves
classically. The only question is: Where is the classical limit? Bohr's
assumption is that the classical limit is always to be found in the limit of
large quantum numbers. In making this assumption he was guided by
certain evidence available at the time.

For instance, the

classical Rayleigh-

Jeans theory of the black body spectrum agrees with experiment in the
limit of small v. Since Planck's quantum theory agrees with experiment

between the quantum and classical


v. But it is easy to see
from equation (2-30) that as v becomes small the average value n, of the
quantum number specifying the energy state of black body electromagnetic
waves of frequency v, will become large.f The second part of the corre-

everywhere,
theories

is

we

see that correspondence

found, in this case, in the limit of small

spondence principle was purely an assumption, but certainly a reasonable


one.

correspondence principle by applying it to a simple


such
harmonic oscillator,
as a pendulum oscillating at frequency v. One
of the predictions of the quantum theory for this system is that the allowed
nhv. In the discussion of
energy states are given by the equation E
Let us

illustrate the

section 10, Chapter 2,

we have

seen that, in the limit of large

n, this

not in disagreement with what we actually know about the


energy states of a classical pendulum. In this sense, the quantum and
classical theories do correspond for n * co in so far as the energy states of
a pendulum are concerned. Next assume that the pendulum bob carries
prediction

f e

nhv

is

nhv, so e

kT, which

=
is

nhv. Equation (2-30) shows that e ->


a constant. Thus n -* co as v -* 0.

kTas

->- 0,

so in this limit

BOHR'S THEORY OF ATOMIC STRUCTURE

136

[Ch. 5

we can compare the predictions of the two


concerning the emission and absorption of electromagnetic
radiation by such a system. Classically the system would emit radiation
an

electric charge, so that

theories

due to the accelerated motion of the charge, and the frequency of the
v. According to the quantum theory,
radiation is emitted as a result of the system making a transition from
quantum state n { to quantum state nf The energy emitted in such a transiemitted radiation would be exactly

tion

is

E Ef =

equal to

(n t

a quantum of frequency (Et

nf)hv. This energy


Ef)/h = (n nt)v.
4

obtain correspondence between the classical and


the frequency of the emitted radiation,
rule n t

nf

be valid in the

is

we must

carried

away by

Thus, in order to

quantum

predictions of

require that the selection

classical limit.

similar

argument

concerning the absorption of radiation by the charged pendulum shows


that in the classical limit there is also the possibility of a transition in which
n { nf
1
The validity of these selection rules in the quantum limit
can be tested by investigating the spectrum of radiation emitted by a

The

vibrating diatomic molecule.

vibrational energy states for such a

system are just those of a simple harmonic


leads to the equilibrium separation of the

oscillator, since the force which


two atoms has the same form as

a harmonic restoring force. From the vibrational spectrum it can be


determined that the selection rule n t n f 1 actually is in operation
in the limit of small quantum numbers.

number of selection rules were discovered empirically in the analysis


of atomic and molecular spectra. It was often possible to understand these
selection rules in terms of a correspondence principle argument, but

some-

times ambiguities arose.

9.

A Critique of the
In Chapters

2, 3,

ment of what

is

and

now

Old Quantum Theory


5,

we have described the high points in the develop-

referred to as the old

quantum

theory.

In

even

many

more so than may be


apparent to the reader since we have not mentioned a number of successful
applications of the theory to phenomena, such as the heat capacity of solids
respects this theory

was very

successful

low temperatures, which were inexplicable in terms of the classical


However, the old quantum theory was certainly not free of
criticism. To complete our discussion of this theory we must indicate some
at

theories.

of

its

1.

undesirable aspects.

The theory only

there are

many

tells

us

how

to treat systems

which are

periodic, but

systems of physical interest which are not periodic.

Ch.

5]

BIBLIOGRAPHY

137

Although it does tell us how to calculate the energies of the allowed


of a system, and the frequency of the quanta emitted or absorbed
when the system makes a transition between allowed states, the theory does
not tell us how to calculate the rate at which such transitions take place.
3. The theory is only really applicable to a one-electron atom. The
alkali elements (Li, Na, K, Rb, Cs) can be treated approximately, but only
because they are in many respects similar to a one-electron atom. The
theory fails badly when applied to the neutral He atom, which contains
two electrons.
4. Finally we might mention the subjective criticism that the entire
theory seems somehow to lack coherence to be intellectually unsatisfying.
2.

states

That some of these objections are really of a very fundamental nature


was realized by everyone concerned, and much effort was expended in
attempts to develop a quantum theory which would be free of these and
other objections. The effort was well rewarded. In 1925, following an
idea proposed in the preceding year by de Broglie, Schroedinger developed
his theory of quantum mechanics. This theory is now almost universally
considered completely acceptable.

As we shall see,

the Schroedinger theory

is

in

many respects very different

from the old quantum theory. For instance, the picture of atomic structure
provided by the theory of quantum mechanics is the antithesis of the
picture, used in the old quantum theory, of electrons moving in welldefined orbits. Nevertheless, the old quantum theory is still often employed
as a first approximation to the more accurate description of quantum
phenomena provided by quantum mechanics. The reasons are that the
old quantum theory is often capable of giving numerically correct results

with mathematical procedures which are considerably less complicated

than those used in quantum mechanics, and that the old quantum theory
is

often helpful in visualizing processes which are difficult to visualize in

terms of the somewhat abstract language of the theory of quantum


mechanics. Therefore the reader would be well advised to become thor-

oughly familiar with the content of the old quantum theory, particularly
the

Bohr

theory.

BIBLIOGRAPHY
Born, M., Atomic Physics, Blackie and Son, London, 1957.
Ruark, A. E., and H. C. Urey, Atoms, Molecules, and Quanta, McGraw-Hill
Book Co., New York, 1930.
Richtmyer, F. K., E. H. Kennard, and T. Lauritsen, Introduction to Modern
Physics, McGraw-Hill Book Co., New York, 1955.

BOHR'S THEORY OF ATOMIC STRUCTURE

138

[Ch. 5

EXERCISES
1.

Carry through the details of the Bohr calculation for a nucleus of

mass and obtain equation

finite

(5-19).

2. Consider a particle of mass m moving back and forth along the x axis
between two points x = and x = a from which it rebounds elastically. Apply
the Wilson-Sommerfeld quantization rules to find the predicted possible values
of the total energy of the particle.
3.

Consider a body rotating about an

quantization rules, and

show

axis.

Apply the Wilson-Sommerfeld

that the possible values of

its total

energy are

predicted to be

E=
where / is
4.

its

moment

AW/2/,

0, 1, 2, 3,

of inertia about the axis of rotation.

Apply the Wilson-Sommerfeld quantization rules to the results of exercise


4, and obtain equations (5-21).

3,

Chapter
5.

Prove the statement, made in section

classical expression

of

(v/c)

E = p2/2m + V is

7, that for

smaller than

small v/c the error in the

E by

a factor of the order

2
.

6. Derive a relation connecting the frequency of the electromagnetic radiation,


emitted in a transition between two states of a Bohr atom, and the orbital fre-

quencies of the electron in these states. Study this relation in the limit of large

quantum numbers, and comment on


classical physics.

its

correspondence with the predictions of

CHAPTER

Particles

and

I.

De

Waves

Broglie's Postulate

In his doctoral dissertation (1924) de Broglie put forth a simple but


extremely important idea which initiated the development of the theory of

quantum mechanics. De
following:

Broglie's line of thought ran something like the

In classical physics electromagnetic radiation had been con-

wave propagation phenomenon. However the


and Compton has demonstrated that under certain

sidered to be purely a

work of

Einstein

circumstances

displays properties characteristic of particles (quanta).

it

it also be true that physical entities which we


normally think of as particles (electrons, alpha particles, billiard balls,
etc.) will under certain circumstances display properties characteristic of
waves ?
The particle-like aspects of electromagnetic radiation become apparent
when the interaction of the radiation with matter is investigated, while the

This being the case, could

wave-like aspects

propagates

is

become apparent when the manner in which the radiation

investigated.

It

is

immaterial whether the situation

is

described by saying that electromagnetic radiation actually consists of

waves which upon interacting with matter are able to manifest a particleby saying that it actually consists of particles whose
motion is governed by the wave propagation properties of certain
associated waves. In fact, it is somewhat naive even to imply that there
is a choice to be made.
However, by temporarily adopting the latter
statement, de Broglie was guided by analogy to look for the proposed
wave-like aspects of particles in terms of some wave propagation
like behavior, or

139

PARTICLES

140

AND WAVES

[Ch. 6

Thus he was led to investigate the idea that the


governed by the wave propagation properties of
certain pilot waves (to use his terminology) which are associated with the
character in their motion.

motion of a

particle is

particle.

At

the time of de Broglie's proposal, wave-like behavior in the motion

of a particle had never been observed although the motion of particles had

been extensively investigated. But

this

does not prove anything because

Source

Photographic

Slit

plate

Figure 6-1.

An

apparatus used to demonstrate the wave-like nature of electromagnetic

radiation.

is so small, compared to
had been used to investigate
particle motion, that the wave-like properties would not have been observable. The meaning of this statement can be illustrated by considering a
single-slit diffraction experiment typical of those which can be used to
demonstrate the wave-like nature of electromagnetic radiation. The

it

might be that the wavelength of the pilot waves

the dimensions of the physical systems which

experiment

is

indicated in figure (6-1). It

an experiment depend
of the radiation,

X,

in a drastic

to the width of the

However, for 1

observed.
its

Under

<d

known

slit,

that the results of such

the ratio of the wavelength

d.

For A

d,

a diffraction

phenomena is recorded on the photographic

pattern characteristic of wave


plate.

is

manner on

a quite sharply defined image of the

these circumstances the radiation betrays

wave-like character, and

little

slit is

sign of

propagation can be very adequately

its

described in terms of the laws of ray optics, a particle-like theory.


It is clear that

de Broglie's

first

task was to specify

determining the wavelength of the pilot waves.

He

some procedure

for

did this by making the

reasonable assumption that the wavelength of pilot waves, and also their
frequency, could be evaluated from the same equations which are used to
evaluate these parameters for electromagnetic radiation. Consider a

beam

of electromagnetic radiation containing quanta of total relativistic energy E.

According to Einstein's equation

(3-7), the frequency v of the radiation is


v

Ejh

(6-1)

SOME PROPERTIES OF THE PILOT WAVES

Sec. 2]

141

The wavelength A of the radiation can be evaluated from the usual relation
between

A, v,

and w, the velocity of propagation of the wave.

= w/v
For electromagnetic radiation w = c, so
X = cjv = hcjE
According to equation (3-10), E\c = p, the momentum

This

Thus A can

also be evaluated

(6-3)

of the quantum.

h\p

(6-4)

Broglie postulated that the wavelength A and the frequency v of the

a particle of momentum
by the equations

p and

pilot waves associated with

energy

E are given

l
v

and

(6-2)

from the equation


A

De

is

that the motion

=
=

of the particle

total relativistic

hlp

E\h
is

governed by the wave propagation

properties of the pilot waves.

His reason for choosing equations analogous to (6-4) and (6-1), instead
of to (6-3) and (6-1), will become apparent shortly.

2.

Some

Properties of the Pilot

Waves

Let us calculate the propagation velocity w of the pilot waves associated


with a particle, using equation (6-2) and evaluating A and v from equations
(6-5).

This gives

vk

hp

(6-6)

Using equation (1-25) to evaluate the total relativistic energy of the particle,
we have

2 2
VcV+^mpC
)

cVp 2

(m c?

p
or

w
Note

that

w is greater than

c.

the velocity v of the particle

cVl

+ (m

c//>)

This seems, at

must

"7

C6

first,

quite disturbing because

necessarily be less than

appear that the particle could not keep up with

its

own

c,

and

it

would

pilot waves.

PARTICLES

142

Actually there

is

no

difficulty, as

that a free particle (experiencing

are

moving

coordinate

AND WAVES

[Ch. 6

the following argument shows. Imagine

no

and

forces)

its

associated pilot waves

way that they can be described by the single spatial


Assume that we have distributed at a large number of

in such a
x.

points along the x axis a set of hypothetical instruments which are capable

of measuring the intensity of the pilot waves. At a certain time t we


record the readings of these instruments. The results of the experiment
,

ii

Figure 6-2.

group of

pilot

waves.

could be presented as a plot of the instantaneous value of the pilot waves,


which we designate by the symbol *(x, i)t=t as a function of a;. Although
,

we know

about the pilot waves at present, it is evident that the plot


must look qualitatively like the one shown in figure (6-2). The amplitude
of the pilot waves must be modulated in such a way that their value is
non-zero only over a finite region of space in the vicinity of the particle.
This is necessary because the pilot waves must somehow be associated
spatially with the particle whose motion they control. The pilot waves
form a group of waves and, as a function of time, the group must move
along the x axis with the same velocity as the particle. The reader may
recall,

little

from

his study of

wave motion

a moving group of waves

g of

it is

the group and the velocity

waves.

Furthermore,

is

in elementary physics, that for such

necessary to distinguish between the velocity

of the individual oscillations of the

in general less than w.

This

is

encouraging,

but of course we must prove that g is equal to the velocity of the particle.
To do this, we develop a relation between g and the quantities v and
X comparable to the relation (6-2) between w and these two quantities.

We

start by considering the simplest type of wave motion, a sinusoidal


wave of frequency v and wavelength X, which is of constant amplitude
from -co to +oo, but which is moving with uniform velocity in the
direction of increasing x. Such a wave can be represented by the function

T(x,

i)

sin 2t7

(M

or

Y(x,

t)

sin

2n(kx

vi),

where k

1/A

(6-8)

Sec. 2]

SOME PROPERTIES OF THE PILOT WAVES

That

does represent the wave can be seen from the following con-

this

siderations
1.

Holding x fixed, we see that the function

with frequency
2.

on

x,

3.
it

143

Holding

oscillates in

time sinusoidally

v.

fixed,

we

see that the function has a sinusoidal

dependence

with wavelength A or wave number k.

The

zeros of the function, which correspond to the nodes of the

represents, are

wave

found at positions x n for which

2n(kx n

vi)

vn,

0,

1, 2,

or

xn

Thus these nodes, and

2k

in fact all points

on the wave, are moving

in the

direction of increasing x with velocity

w
which

is

dxjdt

equal to

w =

vjk

Note that this is identical with equation (6-2) since k = 1/A.


Next we discuss the case in which the amplitude of the waves is modulated to form a group. We can obtain mathematically one group of
waves moving in the direction of increasing x [something similar to the
group of pilot waves pictured in figure (6-2)] by adding together an
infinitely large number of waves of the form (6-8), each with infinitesimally
differing frequencies v and wave numbers k. However, the mathematical
techniques become a little involved and, for the purposes of the present
argument, it will suffice to consider what happens when we add together
only two such waves. Thus we take
W(x,

t)

= Y (x, t) + Y2 (z, t)

(6-9)

where
i)

sin 2ir\kx

2 (x, i)

sin 2n[{k

Yj(x,

vi\

(y

B)

sin

\(A

{2k

dk)

dk)x

dv)i\

Now
sin

Applying

T(x,

A +

sin

B=

this to the case at

2 cos

2n\

2 cos \(A

E)

hand, we have

1 sin 2n

{2v

dv)

/]

PARTICLES

144

Since dv

< 2v and dk < 2k, this


T(a;,

()

= 2 cos

2tt\

AND WAVES

[Ch. 6

(6-10)

is

sin 2ir{kx

vi)

A plot otW(x, t) as a function of x for a fixed value of t =

/ is shown in
The second term of Y(a;, /) is a wave of the same form as
equation (6-8), but this wave is modulated by the first term so that the

figure (6-3).

-ir*"

two

waves of

&

The sum

Figure 6-3.

of

sinusoidal

slightly different frequencies

and wave

numbers.

oscillations of

Y(x,

t)

fall

Two waves of

within an envelope of periodically varying

frequency and wave number


and reinforce in such a way as to produce an infinite
succession of groups. These groups, and the individual waves which they
contain, are both moving in the direction of increasing x. The velocity
w of the individual waves can be evaluated by considering the second
term of Y(a;, t), and the velocity g of the groups can be evaluated from the
amplitude.

slightly different

alternately interfere

first

Proceeding as in 3 above,

term.

w
8

_
~

we

find

vjk

dv/2

dkj2

_
~

(6-11)

dv

Tk

It can be shown that for an infinitely large number of waves combining to


form one moving group, the dependence of the wave velocity w, and
the group velocity g, on v, k, and dvjdk is exactly the same as for the

we have considered. Equations (6-1 1) have a general validity.


we are in a position to calculate the group velocity g of the

simple case
Finally

group of pilot waves associated with the moving particle. From equations
(6-5), we have
v = Ejh and k = \\'k=p\h

So
dv

dEjh

and

dk

dp/h

SOME PROPERTIES OF THE PILOT WAVES

Sec. 2]

Thus the group

velocity

145

is

dvjdk

dE\dp

ing to equation (1-25),

Kc

E 2 == c2p2 +
2EdE =
dE

2p dp

?
E

dp
>re

2 2

(6-12)

According to equations (1-24) and (1-17),

E = mc 2
where

m m /Vl

of the

particle.

mv

v 2 /c 2 , the total relativistic mass,

Thus we have the

g
The

and p

and

is

the velocity

satisfying result that

=
mc

(6-13)

group of pilot waves is just equal to the velocity of the


whose motion they govern, and de Broglie's postulate is internally

velocity of the

particle

consistent.
It is

easy to see

why de

Broglie did not take a pair of equations

k=\jX =
v

E\hc

E\h

analogous to equations (6-3) and (6-1) to specify the frequency and wave
the pilot waves associated with a particle. Applying (6-11),

number of

these equations lead to a

wave

velocity

w
and to a group

vjk

velocity

dv

dk

dEjh
'
dEjhc

=c

quantum which moves with a velocity which is always


but certainly not correct for a particle which moves with a
velocity v which is necessarily different from c.
This

is

correct for a

equal to

c,

From
the

AND WAVES

PARTICLES

146

equations (6-6) and (6-12)

wave

velocity

w and

we

find the following relation

[Ch. 6

between

the group velocity g:

w
According to equation (6-13),

this

c*/g

can be written in terms of the velocity

of the particle v:

c % \v

(6-14)

Since w is larger than g, the individual waves are constantly moving


through the group from the rear to the front. The same situation occurs
in a group of water waves.

Experimental Confirmation of de Broglie's Postulate

3.

As mentioned

in section 1, comparison with the propagation characterof electromagnetic radiation predicts that wave-like behavior in the
motion of a particle will only be apparent when the de Broglie wavelength
istics

hjp is of the order of, or larger than, the characteristic dimensions


of the system used to investigate the motion. Since Planck's constant h
has the very small value 6.62 x 10~ 27 erg-sec, it is immediately apparent
that, unless the

momentum p

also very small, X will be so small as to

is

preclude any hope of observing the postulated


p and X for some specific cases.

effects.

Let us evaluate

Consider a very small dust particle of radius r = 10~ 4 cm, density


-1
10, moving with the low velocity v = 1 cm-sec
Since v
c, we
p
may evaluate p from the classical formula

p = mv =
The de

is

~4

10 -11 gm-cm-sec -1

Broglie wavelength of this particle

This

7rr 3 pv

6.6

x 10~ 27

10~ u

<

is

1.6

__ 16
18
cm

X 10

extremely small compared to the dimensions of any physical system,


we have taken the momentum of the particle to be very much

although
less

than that of the particles typically discussed in classical mechanics.

Thus we would not expect to be able to either prove or disprove the validity
of de Broglie's postulate by investigating the motion of macroscopic
particles.

Next consider an electron with the


10 ev

1.6

10 _11 ergs.

This

is,

relatively

low

kinetic energy

T=

for example, approximately equal to

the kinetic energy of an electron in a hydrogen atom.

The

velocity of the

EXPERIMENTAL CONFIRMATION OF THE POSTULATE

Sec. 3]

electron

is

small compared to

c,

so

we may

evaluate

its

147

momentum from

the classical equation

p=

VlmT=V2 x 9.1 x 10- X 1.6


~ 1.7 x 10~ 19 gm-cm-sec-1
28

KH

/?

The corresponding de

Broglie wavelength

is

= 6.6xl0-^
39xl0_8cm
10- 19

1.7

Although this is still a very small wavelength, it is of the order of the


size of an atom and, consequently, of the interatomic spacing of the

o+

Surface of crystal

-J U
Figure 6-4.

An

by the electrons

atoms in a

apparatus used to verify de Eroglie's postulate.


is

The region traversed

evacuated.

crystal.

This suggests the possibility of looking for diffraction


or transmission of a beam of electrons by a crystal.

effects in the reflection

Such

effects

were

first

found by Davisson and Germer.

In the Davisson-Germer experiment (1927), illustrated schematically


electrons
in figure (6-4), a parallel beam of monoenergetic low energy
a
was produced by accelerating electrons, thermally emitted from cathode
C, through a voltage drop V.
surface of a

Ni

crystal.

Some

the surface of the crystal.

This

beam was

incident

normal to the
back from

of the electrons were scattered

The number

scattered at angle

(defined in

behind
the figure) was measured with a detector consisting of a plate P
allowed
V,
than
less
slightly
V
voltage
retarding
a diaphragm D.
r,

PARTICLES

148

only electrons scattered from the crystal with

AND WAVES

little

P and produce a current indicated by meter M. Low

[Ch. 6

energy loss to reach


energy electrons lose

energy rapidly in traversing a solid, so the detected electrons must have


essentially from the surface of the crystal. The periodic
distribution of the atoms of that surface is indicated in the figure. For a

been scattered

Ni

d of the periodic structure is 2.15 x 10~8


comparable to the de Broglie wavelength of the incident
Since the essential feature of a diffraction grating is its periodi-

crystal the repetition length

cm,f which
electrons.
city,

the pattern of electrons scattered from the crystal surface should

Figure 6-5.

is

The angular

distribution of electrons scattered from a nickel

show

crystal.

54.0 volts.

diffraction effects if the

de Broglie postulate

is

correct.

Plotted in figure

form of Davisson and Germer's data for the number of


scattered electrons N(6) as a function of 0, measured for an accelerating
voltage V of 54.0 volts. The angular distribution JV(0) becomes very large
for small 0; that is, most of the scattered electrons are specularly reflected
(6-5)

is

the

(angle of reflection

angle of incidence). This

is not particularly informof both particle motion and


wave motion. However, at an angle of 50 a peak is observed in N(6).
The existence of this peak proves the qualitative validity of the de Broglie
postulate because such a peak can only be explained as a constructive
interference of waves scattered by the periodically placed atoms
of the
crystal surface. This is not an interference between waves associated
with one electron and waves associated with another. Instead, it is an

ative because such behavior is characteristic

interference between waves, associated with a single electron, that have


been scattered from various parts of the crystal. This conclusion can be
demonstrated experimentally by making measurements with an electron
t This

is

known from measurements of the diffraction pattern produced in


known wavelength by a Ni crystal. Such measurements

scattering of X-rays of

be described in Chapter

14.

the
will

EXPERIMENTAL CONFIRMATION OF THE POSTULATE

Sec. 3]

beam of such low

intensity that the electrons

149

go through the apparatus

one at a time, and by showing that the pattern of the scattered electrons
remains the same.
The data of Davisson and Germer can also be used to show that de
Broglie's postulate is quantitatively correct. If we evaluate the de Broglie
wavelength of a 54 ev electron from the equation X = hjp, as above, we
find

1.67

10- 8

cm =

1.67

The wavelength of the waves which are interfering constructively can


be evaluated from the well-known grating equation,^
nX

Assuming

that the

d sin

6,

peak corresponds to a

1, 2, 3,

first

also

(6-15)

order diffraction (n

1),

we have

The two

2.15

10-8

cm x

sin 50

1.65

values of X agree to within the accuracy of the experiment.

parable agreement was found in the data obtained at a


accelerating voltages.

(n

2)

was also

At somewhat higher

Com-

number of different

voltages a second order peak

seen.

Diffraction effects in transmission through crystals were observed

by

Thomson. His experiment (1928) involved sending a collimated


beam of electrons through a polycrystalline foil. The foil thickness was
only about 10~ s cm, but even so it was necessary to use an electron
energy of about 10* ev so that the energy loss in traversing the foil would
G.

P.

not be too large. In passing through the randomly oriented crystals of the
foil, the electrons will find certain ones for which the set of scattering angles
wu^ satisfy the equation analogous to (6-15) which applies
0i 6 2 %>
,

to this case.

The

from these crystals


The angles are all small since Xjd is small
however, by placing a photographic plate some

electrons will be strongly scattered

at the angles indicated.

(A~5

x 10~2 A);

distance behind the scattering foil a series of concentric rings corresponding

shown
The photograph on the left shows a ring pattern resulting
from the scattering of electrons by a thin layer of Au crystals. The photograph on the right shows the completely similar pattern observed in the
scattering of electromagnetic radiation (X-rays) by a thin layer of ZrO a
to the allowed scattering angles can be detected. Typical results are
in figure (6-6).

crystals. This comparison gives another convincing demonstration of


wave-like behavior in the motion of electrons. Furthermore the values of
the scattering angles measured from the electron ring pattern agree with
t See any textbook

on elementary

physics.

PARTICLES

150

AND WAVES

[Ch. 6

those of the scattering angles calculated from the wavelength predicted

by the de Broglie

postulate.

Several years after the initial experiments with electrons, Estermann,


Frisch, and Stern demonstrated the existence of diffraction effects in the
scattering of a beam of He atoms from the surface of a LiF crystal.
The beam was obtained from an enclosure heated to a temperature of
400K. Emerging from a channel communicating with the interior of the

oxide crystals.

The ring pattern produced in the scattering of electrons by gold


The ring pattern produced in the scattering of X-rays by zirconium
From U. Fano and L. Fano, Basic Physics of Atoms and Molecules, John

Wiley and Sons,

New

Figure 6-6.
crystals.

Left:

Right:

York, 1959.

beam of atoms of average energy E = \kT, where k is


Boltzmann's constant. The de Broglie wavelength of an atom of average
energy was
enclosure was a

J2ME

j3MkT
6.6

V3 x

6.7

x 1(T 27

X 1(T 24 x

1.4

~ 0.62 A

X 1(T 16 x 4 x 102

The scattered atoms were detected by a very sensitive pressure gauge.


The pattern of scattered He atoms was found to be similar to that observed
by Davisson and Germer for the scattering of electrons. The angular
location of the diffraction peak was in good agreement with what would be
expected for the wavelength predicted by de Broglie's postulate.

We have described three of the original experiments verifying that there


are

wave propagation

effects in the

motion of particles, as well as verifying

INTERPRETATION OF THE BOHR QUANTIZATION RULE

Sec. 4]

Since then very

the wavelength predicted by de Broglie's postulate.

more examples of

151

many

these effects have been observed experimentally,

and

the validity of de Broglie's postulate has been undeniably confirmed.

4.

Interpretation of the Bohr Quantization Rule

We

have seen above that the de Broglie wavelength of an electron of


comparable to that of an electron in an H atom is about
10~ 8 cm. According to the Bohr theory, the radius of the ground state

kinetic energy

4x

atom is about 0.5 x 10-8 cm. Since X ;> r, we would


wave propagation properties of the pilot waves to play an

orbit of the

expect the

important part in determining the motion of the electron in the atom.


However, there is a significant difference between the type of particle
motion we have previously discussed in this chapter and the motion of an
electron in

an atom.

trajectories just once

In the examples above the particles travel their


they are unbound particles, and their associated pilot

waves are traveling waves (moving nodes). An electron in an atom is


bound to the atomic nucleus. It travels repeatedly through the same orbit
Consequently its associated pilot waves would be expected to be standing
waves (fixed nodes).
Standing waves are one of the most striking phenomena of wave propagation. In 1924 de Broglie showed how a simple consideration of their
properties could lead to an interpretation of the mysterious Bohr angular
momentum quantization rule. According to equation (5-11), this rule
can be written

mvr
where

radius

r.

is

the linear

If

we

= pr =

n=

nh\2-n,

momentum

of an electron in an allowed orbit of

substitute into this equation the expression

p=
for

1, 2, 3,

hjX

in terms of the corresponding de Broglie wavelength, the equation

becomes
hr/X

This

nhjln

is

Itrr

nX,

1,

2, 3,

(6-16)

orbits are those in which the circumference of the orbit


can contain exactly an integral number of de Broglie wavelengths. If
we think of an electron traveling repeatedly around the same orbit, it
becomes apparent that equation (6-16) is just the necessary condition that

Thus the allowed

PARTICLES

152

AND WAVES

[Ch. 6

the pilot waves associated with the particle in a particular traversal of the
orbit will

combine coherently with the

pilot

waves associated with the

up a standing wave. However, if this equation is


violated, then in a large number of traversals the pilot waves will interfere
with each other in such a way that their average intensity will vanish.
Since the intensity of the pilot waves must be some sort of a measure of
where the particle is located, we interpret this as meaning that an electron
previous traversals to set

could not be found in such an orbit. Figure (6-7)

Figure 6-7.

the

first

is

an attempt to illustrate

schematic representation of the patterns of standing waves set up

in

three allowed Bohr orbits.

the

Tj =(

the

first

some

waves set up in
the magnitude
and sign of the amplitudes of these oscillatory patterns might be different,
but the locations of the nodes would be the same since the nodes of any
standing wave are fixed.
It is not difficult to show that the requirement that the pilot waves
associated with a particle undergoing any sort of periodic motion be a set
,

at

particular value of

three allowed

of standing waves

is

Bohr

orbits.

for the standing

At some other value of t

equivalent to the requirement that the motion of the

particle satisfy the Wilson-Sommerfeld quantization rules (5-20).

see in the following chapters that the properties of standing

We

shall

waves also

play an important role in the Schroedinger theory of quantum mechanics.

For

instance, the time independent features of the standing

ated with an electron in one of


possible to understand

why

its

waves associ-

allowed states in an atom will

the motion described

make

it

by the standing wave

does not cause the electron to emit electromagnetic radiation.

THE UNCERTAINTY PRINCIPLE

Sec. 5]

5.

The Uncertainty

An

153

Principle

inescapable result of de Broglie's standing

electron in a

Bohr

orbit

is

wave description of an
and the orienta-

that the position of the electron,

its linear momentum, cannot be exactly


a given instant of time since the azimuthal symmetry of the

tion of the vector indicating


specified at

pattern of standing waves, pictured in figure (6-7) at the instant

indicates that at that instant the electron in a particular orbit could be

located essentially anywhere in the orbit.


specification of position

and

linear

This

is

in contrast to the exact

momentum which is

possible in Bohr's

motion of an atomic electron, just as it is


possible in Newton's description of the motion of a planet. This uncertainty
in the position and momentum of a particle at any instant is, as we shall
see, a general feature of de Broglie's description of the motion of a particle

particle-like description of the

its associated pilot waves.


Let us consider a freely moving particle and its associated group, or
groups, of pilot waves. We take first the mathematically simplified case

in terms of the propagation of

of the pilot waves, described by equation (6-9), obtained by combining


two unmodulated waves of slightly different k and v. This is

Y = sin 2n[kx where we use

Ak and Av

at the instant

vt]

sin 2-n-p

instead of

dk and

in figure (6-3).

It

Ak)x
dv.

(v

Av)f]

This function

represents

an

is

(6-17)
plotted,

infinite succession

of groups of waves which are traveling in the direction of increasing

Now,

in consideration of the connection that

we have made on

x.

several

previous occasions between the intensity of the pilot waves and the location

of the particle,

it is

apparent that we should interpret

this particular pilot

which it is associated has equal


probability of being located within any one of the groups at the time
/ = t
Considering momentarily a single group, our interpretation of
indicates that even the location of the particle within that group is undetermined to within a distanee comparable to the length A* of the group.
From equation (6-10), or from figure (6-3), we obtain the following
relation between Aa; and the difference AA: between the wave numbers

wave as indicating

that the particle with

of the two constituent waves


Aa; AA:

A
2, in

similar relation can be obtained for the case,

which an

infinitely large

infinitesimally different

k and

(6-18)

mentioned in section

number of unmodulated waves, each with

v,

combine to form a

single traveling group.

PARTICLES

154

AND WAVES

[Ch. 6

Such a group is pictured at t = t in figure (6-8). A single group is formed


by adjusting the phases of all the unmodulated waves so that at one value
of x (the center of the group) they are all in phase and combine to give
large net values of <==i
Proceeding away from this value of x, in either
direction, the unmodulated waves begin to get out of phase with each other
because their wave numbers differ. Beyond certain points they are completely out of phase, and never again get back into phase. Everywhere in
these regions there are, on the average, as many unmodulated waves with
positive values as with negative values, and the waves combine to give a
^= whose value is zero at all x. It is clear that, the larger the range

Figure 6-8.

group of

pilot waves.

of the wave numbers of the unmodulated waves, the smaller is the distance to the points at which they become completely out of phase.
In fact, the distance Ax between these two points is just inversely pro-

Ak of the wave numbers, as in equation (6-18).


However, the proportionality constant is different. From the theory of
Fourier analysis,! it can be shown that a combination of unmodulated
waves with wave numbers covering the range Ak will form a group of
length Ax, where
portional to the range

Ax Ak ~

1/2tt

(6-19)

The exact value of the constant depends on the relative amplitudes of the
different unmodulated waves. The relative amplitudes also determine the
exact shape of the group.

Equation (6-19) describes a property which is common to all wave


motion. Let us now consider the pilot waves which govern the motion of
a particle moving along the x axis. For these waves de Broglie postulates
Jc

1/2.

=pjh

t The "Fourier integral" is used to sum the contributions from the infinitely
number of unmodulated waves with infinitesimally differing wave numbers.

large

THE UNCERTAINTY PRINCIPLE

Sec. 5]

where px

tum

is

is

the

in the

momentum

x direction).

155

of the particle (the x

We

signifies that the

momen-

then have

Ak

Apjh

(6-20)

where Ak is the range of wave numbers involved in the composition of a


group of pilot waves having, at some instant, a length Ax. Npw, we may
say that Ak represents the uncertainty in the wave number which should
be associated with the group at that instant. Therefore we may also say
that Apx represents the uncertainty in the momentum which should be
associated with the particle at the same instant. Furthermore, the discussion above indicates that, for pilot waves, the length Ax of the group
represents the uncertainty in the instantaneous position of the particle.

Thus from equations (6-19) and (6-20) we have the following

relation

between the uncertainties in the instantaneous values of the position and


of the momentum of the particle

Ax APx ~H,

(6-21)

hjlir

is one statement of the uncertainty principle, which was first enunciated


by Heisenberg in the year 1927.
This relation sets a fundamental limit on the ultimate precision with
which we can simultaneously know both the x coordinate of a particle and
We may arbitrarily decrease the
its momentum in the x direction.
uncertainty in one of these quantities, but only at the expense of increasing
the uncertainty in the other. For example, consider a particle whose
associated pilot waves form a group of length ten times less than the group

This

The instantaneous uncertainty in the position


of the particle is then less by a factor of 10. However, equation (6-19)
says that the range of wave numbers present in the new group will be

pictured in figure (6-8).

increased by a factor of 1 0. Consequently the uncertainty in the

momentum

of the particle at that instant will also be increased by a factor of 10, and
the product Aa; Ap x remains approximately equal to h. If we demand that
the uncertainty in our knowledge of the position of the particle at

some

instant approach zero, then equation (6-21) requires that the uncertainty

knowledge of the momentum of the particle at that instant approach


Going to the other limit, we see that if at any time the momentum
of the particle is precisely known (Apx = 0), then at that time the position
must be completely unknown (Aa; oo). A pilot wave associated with a

in our

infinity.

particle in this situation is given in equation (6-8).

Y=

sin

2n(kx

Here px always has the precise value px


of the particle

is

vt)

hk, but at

This

is

(6-22)

any time the position

entirely unknown since Y _^ extends from oo to + oo.


t

PARTICLES

56
It

AND WAVES

[Ch. 6

should be pointed out that the uncertainty principle gives a lower


product Ax Apx That is, it is possible to find a group of waves

limit to the

which the relation corresponding to equation (6-19) is Ax Ak = y,


where y > \\2tt. For a particle whose pilot waves form such a group, the
product of the instantaneous uncertainties in the position and momentum
of the particle would be larger than h. As a limiting example of this,
consider a particle whose associated pilot wave is represented by the
for

function

T = sin 2Tr[kx -

vt]

sin 2n[(k

Ak)x

(r

Av)t]

(6-23)

Ax -- oo,
which Ax Apx is

given in equation (6-9), with a large value of AA:. For this particle

and Apx

= h Ak

situation in
be large.f
than h means simply that the experimentalist has not
achieved the ultimate limit of precision set by the uncertainty principle.
In fact, it is very difficult to obtain Ax Ap x comparable to h in an actual
will also

significantly larger

measurement because of the very small size of this constant. This is why
the uncertainty principle was never noticed in the experiments of classical

10~27 erg-sec, the unBecause of the very small size of h


but it is
is not important in classical mechanics;
extremely important when we are dealing with the very small distances and
momenta involved in atomic and nuclear systems.
If we are considering a three dimensional problem then, according to

mechanics.

certainty principle

Heisenberg:

There are three independently operating uncertainty relations, one for


For rectangular coordinates x, y, z and corresponding

each coordinate.

momentum components px py p z
,

these relations are

Ax Apx > h
Ay Apy > h
Az Ap z % h

(6-24)

If some of the coordinates are angular coordinates, then for each coordinate
6 and corresponding angular momentum L e the uncertainty relation is

A0AL >
fl

(6-25)

f It may seem that the use of functions (6-22) and (6-23) in these examples weakens
the arguments of the second paragraph of section 2, but this is not actually so. Consider

a 1 cm long beam of alpha particles which are being scattered by an atom. From the
point of view of the atom, and in terms of distances of the order of the atomic radius R,
the x position of some particular particle in the beam may be entirely unknown; Aa; ^> R.

on the accuracy with which the momentum is


known, could provide a. good approximation in the vicinity of the atom to the pilot waves
associated with a particle of the beam. However, \x is really 1 cm and not infinity,
and the arguments of section 2 remain valid.
If so, (6-22) or (6-23), depending

Sec. 5]

THE UNCERTAINTY PRINCIPLE

The symbol > has been used

157

to indicate specifically that the uncertainty

Note that there is no restriction at


on products such as Ax &py
The "physical reason" for the existence of a limit in the maximum
precision with which we can simultaneously know the position and the
momentum of a particle is quite easy to understand. What happens is
that a measurement performed to find the value of one of these quantities

principle specifies only a lower limit.


all

Observer

Microscope

Particle

Bohr's microscope experiment.

Figure 6-9.

will

always disturb the particle in such a way as to leave the value of the
good example of this was pointed out by

other quantity uncertain.

Consider a measurement, illustrated in figure (6-9), made


to determine the instantaneous location of a particle by means of a
microscope. In such a measurement the particle must be illuminated,
because it is actually the light quanta scattered by the particle that the
microscopist sees. The resolving power of the microscope determines the

Bohr

in 1928.

ultimate accuracy with which the particle can be located. The resolving
power is known to be approximately A/sin a, where X is the wavelength of

and a is the half-angle subtended by the


of the microscope. So the accuracy of the measurement is

the scattered light

Acc~

objective lens

A
sin

Let us assume that the microscopist needs to see only one scattered light
quantum in order to complete his measurement. The magnitude of the

PARTICLES

158

AND WAVES

[Ch. 6

hjk. But the quantum may have been


anywhere within the angle a, since there is no way of measuring
the exact scattering angle. Consequently the x component of the momentum of the quantum after scattering is uncertain by an amount

momentum

of that quantum

is

scattered

2p

sin

2h

sin

component of the momentum of the quantum can be known


exactly before the scattering (there being no need to know its x coordinate),
conservation of momentum requires that the particle receive an amount of
recoil momentum in the x direction which is uncertain by the same amount.
Thus
,

As

the x

2h sin a
&Px
.

and the product of the uncertainties in the knowledge of the x position


and the x component of momentum of the particle at the instant of
measurement is
Ax Ap^ c=L2h> h

By

using light of shorter wavelength, the microscopist can increase the

accuracy of the position measurement, but


the uncertainty of the

We
is

momentum

this will lead to

an increase in

particle.

see that the state of affairs described

by the uncertainty principle

The quantization of electromagnetic


least one "unit" of light (a quantum of

a direct result of quantization.

radiation requires that either at

momentum p
no

of the

light at all.

hjX)
If

reason in principle

it

must be scattered from the


were not for

why

particle, or else absolutely

would be no
would not be able to see the

this requirement, there

the microscopist

good resolution using an illumination of arbitrarily small


scattered light quantum provides the interaction
which must exist between the measuring instrument and the particle.
This interaction disturbs the particle in a way which is uncontrollable and
unpredictable. As a result, the coordinates and momentum of the particle
cannot be completely known after the measurement. The relation
Ax Apx > h shows that Planck's constant is a measure of the (minimum)

particle with

momentum content. The

magnitude of this uncontrollable disturbance.


There is another form of the uncertainty principle which again follows
from the mathematical properties of waves plus the de Broglie postulate,
in just the same way that equation (6-21) follows from (6-19) and (6-20).
From the theory of Fourier analysis it can be shown that the frequency
range Av, which must be involved in the composition of a single group
of waves of length Ax and group velocity g, is given by the relation

At Av >

1/277-

(6-26)

SOME CONSEQUENCES OF THE UNCERTAINTY PRINCIPLE

Sec. 6]

where At
point;

i.e.,

Ax/g

At

is

is

59

the time required for the group to pass any given

the duration of the pulse of waves.f

Now consider a group of pilot waves of duration At. This quantity


would be a measure of the uncertainty in the time at which the particle,
whose motion is governed by the waves, would pass by a given point.
Furthermore, the frequency of the pilot waves is related to the energy of
the particle by de Broglie's postulate
v

E/h

Av

AE/h

Consequently

The uncertainty Av

AE

in the frequency of the

(6-27)

waves leads to an uncertainty

in the energy associated with the particle,

and

that uncertainty

related to the uncertainty in the time associated with the particle

is

by the

relation

ArAEyh

(6-28)

Heisenberg's interpretation of this uncertainty relation was, however,


than that implied by the discussion above. He stated

much broader

measurement of the energy of a particle (or any system) performed


At must be uncertain by an amount AE, where the

during a time interval

relation between these two quantities

We

shall discuss several

is

At AE y

examples of

h.

this uncertainty relation in sub-

sequent chapters.

6.

Some Consequences

of the Uncertainty Principle

Consider a particle moving freely along the x


instant

the position of the particle

is

axis.

Assume that at the


amount Ax
particle at some later

uncertain by the

Let us calculate the uncertainty in the position of the


time t. According to equations (6-24), the momentum of the particle at
h/Ax Consequently the velocity
is uncertain by the amount Apx
t

t The property of wave motion described by equation (6-26) makes it necessary to


use very high frequencies in the transmission of television programs. The signal from a
10 -6 sec. Thus the range of
television station consists of pulses of duration At
-1
6
10 sec , and the entire broadcast band
frequencies which must be present is Av
-1
-1
6
1.5 x 10" sec ) would be able to accommodate only a
to v
(y ~0.5 x 10 sec
single television "channel." At the frequencies used in television transmission (y

108 sec -1 )

many channels can

fit

into a reasonable portion of the spectrum.

PARTICLES

160

AND WAVES

[Ch. 6

is uncertain by the amount Av = Apjm


and the distance traveled by the particle in the time t
uncertain by the amount

of the particle at that instant

hjm Ax

Ax

tAv

~
is

(6-29)

m Ax

If

by a measurement

range

Ax

at

we have

then in a measurement of

localized the particle within the

position at time / the particle


could be found anywhere within the range A*. Since Ax is inversely
proportional to Ax we see that, the more carefully we localize the particle
,

its

at the initial instant, the less

Furthermore,

we

shall

know about

this uncertainty increases linearly

with

its final

f.f

position.

The behavior of

a system, involving distances and momenta small enough that the uncertainty principle is important, is in complete contrast to the behavior of a
system described by classical mechanics.
In classical mechanics, if the position and momentum of each particle
in an isolated system are known exactly at any instant, the exact behavior
of the particles of the system can be predicted for all future time. The
uncertainty principle shows us that

involving small distances and

it is

impossible to do this for systems

momenta because

it is impossible to know,
with the required accuracy, the instantaneous positions and momenta of

the particles of such a system.

As a

result,

we

be able to

shall

dictions only of the probable behavior of these particles.

make pre-

This idea has

already crept into our discussion through the interpretation of the intensity

of pilot waves, at a particular position, as some sort of measure of the


probability that the associated particle

is

located at that position.

interpretation will be put into a quantitative

form

This

in the next chapter.

In the same chapter we shall develop the differential equation (Schroedinger 's
equation) to which the pilot waves are a solution.

equation

we

In discussing this

behavior of the pilot waves is exactly predictable. But the pilot waves only give information about the probable
behavior of the associated particle.
shall see that the

The uncertainty principle provides us with a deep insight into our ability
to describe the behavior of small scale physical entities.

It also

helps us

to reconcile conflicts which appear to arise

nature of these

entities.

from the dual (particle-wave)


Let us consider, as an example, a double-slit

t This statement implies that the group of pilot waves associated with the particle
spreads in time, so that its length at any instant is essentially the Aa; of equation (6-29).
This can be understood as a consequence of the fact that the unmodulated waves which

form the group have a range of wave velocities Ah> because they have a range of wave
numbers A and a range of frequencies Av. As these waves move with different velocities,
they cannot maintain the phase relations required to form the group and, in time, the
group spreads.

Sec. 6]

SOME CONSEQUENCES OF THE UNCERTAINTY PRINCIPLE

diffraction experiment, using the photoelectric effect as a detector.

161

This

The source emits electromagnetic radiation.


From a study of the energy and time distribution of the photoelectrons
ejected from the photocathode, we conclude that the radiation must
possess particle-like aspects. However, when we measure the number of
is

shown

in figure (6-10).

photoelectrons ejected as a function of the y coordinate, the periodic


pattern which we observe (characteristic double-slit intensity pattern)
forces us to conclude that the radiation
aspects.

If

we

must also possess wave-like

think of the radiation as quanta whose motion

is

governed

Photoelectrons

Source

Slits

Figure 6-10.

Photocathode

An experiment

illustrating the particle-wave duality.

by the wave propagation properties of certain associated waves, then we


are faced with the following problem: Each quantum must pass through
either one slit or the other; if so, how can its motion in the region
to the right of the slits be influenced by the interaction of its associated
waves with a slit through which it did not pass ?
Let us analyze this problem in the spirit of Einstein's analysis of

The statement that a particular


quantum must pass through either one slit or the other is really meaningless unless some provision is made to determine through which slit the
quantum actually passed. One way of doing this would be to place a
number of very small particles in the region immediately to the right of the
slits. After its passage through a slit, the quantum would strike one of the
particles, causing it to recoil. By observing the recoiling particle, we
could tell through which slit the quantum passed, provided the uncertainty
in the y coordinate of the particle after the collision is small compared
the Galilean transformation equations.

to the distance between the

slits;

that

by

is,

<d

provided that

AND WAVES

PARTICLES

162

[Ch. 6

quantum exchanges some y momentum with the


The exact amount exchanged depends on details of the collision
which we have no way of determining. Thus the y momentum exchange
is uncertain by an amount which we call Ap
In order not to destroy the
y
double-slit diffraction pattern produced in the passage of a number of
quanta through the system, A/> must always obey the restriction

Now,

in colliding the

particle.

4ft/P.

<

where 6 is the angle between adjacent minima and maxima of the pattern,
and px is the x component of the momentum of the quantum. If this
condition is not satisfied, a quantum could receive so much y momentum
in the collision that it would end up in what had been a minimum of the
diffraction pattern. Since we can assume that the y momentum of the
particle was precisely known before the collision because there was no
need to know its y position at that time, and since momentum must be
conserved in the collision, b.py

y momentum of the
angle

is

is

also equal to the uncertainty in the

particle after the collision.

Now it is known that the

related to the wavelength of the radiation

between the

slits

Expressing this in terms of the


equation (6-5),

this

Xjld

momentum

of the quantum by means of

we have
6

Combining

and to the distance

by the equation

h/2dp x

with the second inequality above gives


4P.//.

< 6 = h 2dPx
l

so
A/>v

Multiplying this into the

first

<^h/2d

inequality,

Ay Apv

we have

< h/2

is a condition on the uncertainty in the y position and y


of the particle which must always be satisfied in order that

This

able to determine through which

slit

momentum
we may be

each quantum passed without

destroying the double-slit diffraction pattern.

But the uncertainty princannot be satisfied. If we do demonstrate


that each quantum actually did pass through one slit or the other, we shall

ciple tells us that this condition

no longer observe

the double-slit diffraction pattern.

The uncertainty

shows us that the problem we have drawn attention to is actually


illusory. In Chapter 8 we shall see additional examples of the way in which
the uncertainty principle resolves apparent conflicts between the wavelike and particle-like aspects of the same entity.
principle

163

EXERCISES

Ch. 6]

BIBLIOGRAPHY
Bohm,

D.,

Quantum Theory, Prentice-Hall, Englewood Cliffs, N.J., 1951.


and Human Knowledge, John Wiley and Sons,

Bohr, N., Atomic Physics

New

York, 1958.

EXERCISES
1

Add graphically six unmodulated waves of 1 cm amplitudes and wavelengths


1.5, 1.6, 1.7, 1.8, 1.9,

and 2.0 cm, adjusting the phases so

that, at

0,

Carry out the addition to x = 10 cm,


and compare the length Aa; of the group formed, and the range Mc of the wave
numbers used, with equation (6-19). Note that, for a finite number of components with finite differences in wave numbers, a single group is not obtained
because there are always many values of x at which all the components will be

^(x, 0t=t

= + 1 cm for aU tne waves.

in phase again.

Find the possible values of the momentum of the particle, treated in


from the possible values of energy predicted by the WilsonSommerfeld quantization rules. Use these values to evaluate the predicted
2.

exercise 2, Chapter 5,

Show

possible wavelengths of the pilot waves.

that,

with the exception of the

lowest energy state where the momentum is zero, it is just for these wavelengths
that the waves can combine to form standing waves, as stated in section 4.
3.

Discuss the exception mentioned in exercise 2 in terms of the uncertainty

principle.
4.

Consider a

single-slit diffraction

experiment for

particles, similar to that

illustrated in figure (6-1) for electromagnetic radiation. First treat the diffraction

of the pilot waves, and show by a Huygens wavelet construction that the angular
X/d. Next treat the deflection of the
width of the diffraction pattern is

by using the uncertainty principle to estimate the transverse momentum


introduced in passing through the slit by the measurement of transverse position
X/d. Comment
that the slit provides, and show that the deflection angle is d
on why both treatments lead to the same predictions.
is dropping marbles of mass m to the
5. A boy on top of a ladder of height
floor and is trying to hit a crack in the floor. To aim, he is using equipment of
particles

the highest possible precision.

Show

that, despite his great care, the

marbles

will miss the crack by an average distance of the order of

where g

is

the acceleration due to gravity. Using reasonable values of

evaluate this distance.

H and m,

CHAPTER

Schroedinger's

Theory of

Quantum
Mechanics

I.

Introduction

In this chapter we begin our development of Schroedinger's theory of


quantum mechanics. This theory, and its applications to a number of

occupy our attention for much of the remainder of


Furthermore, the discussion of atomic and nuclear physics

typical problems, will

the book.

which follows the discussion of quantum mechanics will be largely based


on the Schroedinger theory. For these reasons our treatment of the
theory, particularly of the mathematical details, will be considerably
more thorough than the treatments we have given previously. This is
necessary because it is often easier to think in terms of the mathematics of
quantum mechanics than it is to think in terms of the somewhat elusive
physical pictures which the mathematics represents and the mathematics
is sufficiently involved that a cursory treatment could not help being
confusing. However, we shall, as usual, be perfectly willing to sacrifice
rigor for clarity when it seems advisable.
Before embarking on Schroedinger's theory, we should mention that an
alternative approach was developed at about the same time by Heisenberg.
His theory did not discuss pilot waves. Instead he dealt only with dynamical
quantities such as x, px E, which were represented by matrices. The

164

Sec. 2]

THE SCHROEDINGER EQUATION

quantum

aspects were introduced into the theory via the uncertainty

principle,

and applied

uncertainty principle

as
is

165

commutation conditions on the matrices. The


Con-

equivalent to the de Broglie postulate.

sequently the Heisenberg and Schroedinger theories are actually identical


in content, although very different in form. Since the Schroedinger theory
is

better adapted to

an introductory treatment of the

subject,

we

consider

here only that theory.

2.

The Schroedinger Equation

should be apparent to the reader that de Broglie's postulate is essenticorrect.


It should also be apparent that the postulate does not
provide us with a complete theory for the behavior of a particle, but only
It

ally

first step toward such a theory. The postulate says that the motion of
a particle is governed by the propagation of its associated pilot waves,
but it does not tell us the way in which these waves propagate. Only for
the simple case of a free particle have we been able to learn some of the
features of the propagation of pilot waves.f To handle the case of a
particle moving under the influence of forces, we must have an equation

the

which

will tell

circumstances.

how

the pilot waves propagate under these more general


we must also have a quantitative connection

In addition,

waves and the associated particle; that is, we must know


exactly how these waves "govern" the motion of the particle. The required
propagation equation for pilot waves, called the Schroedinger equation,
was developed in 1925, and the quantitative connection between these
between the

pilot

waves and the associated

With

particle

was developed

in the following year.

several exceptions, Schroedinger directly followed de Broglie's

He did not use the picturesque term "pilot waves" but instead
denoted both the waves, and the mathematical function T(x, t) which
represents them, by the term wave function. We shall adopt this terminology ourselves. A more serious change was that Schroedinger attempted
only to develop a theory which would be valid in the non-relativistic
range of velocities, although, as we have seen in section 2, Chapter 6,
the de Broglie postulate was consistent with the theory of relativity.

ideas.

Schroedinger adopted de Broglie's two equations


A

and

= h/p
= E\h
,,

(7-1)
._

_.

(7-2)

associated with a free particle


t We know that the velocity of a group of pilot waves
equal to the velocity of the particle (equation 6-13); and also that a group spreads
with increasing time, its length Ax being proportional to / (equation 6-29).
is

SCHROEDINGER'S THEORY OF

166

QUANTUM MECHANICS

but he did not adopt de Broglie's definition of


energy.

E as

Now, from the theory presented in Chapter

[Ch. 7

the total relativistic

equation 1-22"),
apparent that, in the case in which there is a potential energy V, the
total relativistic energy approaches the expression
1 (cf.

it is

+ V+ m

2
^total relativistic ->-/> /2m

C2

in

the non-relativistic limit v/c-+0.


In his non-relativistic theory
Schroedinger did not bother to carry along the constant m c 2 , and he
took for E the classical definition of total energy

E = film + V
m=m

where

(7-3)

We see from equation (7-2) that a change


E will change the value of v. But recall that the experi-

the rest mass.

in the definition of

ments described in the previous chapter were


validity of equation (7-1), not equation (7-2).

all

such as to check the

We

shall see later that the

is of no significance in Schroedinger's theory. However,


one point should be investigated immediately. Assuming equation (7-3),
can we still obtain the necessary result that the group velocity g of the

actual value of v

wave function
Take

for a free particle

equal to the velocity v of the particle?

is

= E\h= pt/lmh +

Vjh

and

k
For a

\\X=p\h

free particle the potential energy

dv

V is

a constant

(cf.

3 below), so

= 2pdp
2mh

Also

we have

From

dk

dp/h

dvjdk

equation (6-11),

Thus we obtain the acceptable

result

g-P d P

mh dp

We

begin our development of the Schroedinger equation by


we know a priori it must satisfy.

listing

three requirements which


1.

It

2. It

must be consistent with equations


must be linear in Y(ar, /). That is,

and (7-3).
and 2(ar, if) are

(7-1), (7-2),
if "x (x, t)

tions to the equation, a linear combination of these functions T(x,

ajTjix,

t)

a^Y^x,

t)

must

also be a solution for

solui)

any values of the

THE SCHROEDINGER EQUATION

Sec. 2]

constants ax and a 2

167

This requirement ensures that we shall be able to add

interference phenomena which


Davisson-Germer experiments. In the
previous chapter we have already assumed that the wave functions can be
added (cf. equation 6-9).
3. The potential energy V is, in general, a function of x and t. In the

together

wave functions to produce the

are observed, for example, in the

special case V(x,

t)

= V a constant, the momentum


= F, and the force
dV _
dV(x,t) _
Q
F=
,

of the particle will

be constant since dpjdt

dx

dx

Also the total energy E is constant in this case. This is the situation of
a free particle with constant values of k = pjh and v = Ejh discussed in
the previous chapter. For this case we require the Schroedinger equation
to have oscillatory traveling wave solutions of constant wave number
and frequency, similar to the wave functions used in that chapter.

Rewriting equations (7-1) and (7-2) in terms of the convenient parameters,

K=

2rrk,

co

(7-4)

2ttv

we have

p =
Combining equations

To

satisfy

(7-3)

requirement

E=hco

HK,

and

(7-5),

we

(7-5)

obtain

H2

K l2m +

the Schroedinger equation must be consistent

V(x,

t)

(7-6)

hco

with equation (7-6).


Requirement 2 demands that every term in the equation be linear in
itself, or a
Y(x, t). Consequently every term must contain either T(a;,
derivative of *(x,

3W(x,

dx

t)

such asf

d 2x(x,

Q
'

dx*

BVjx,

t)

or

""

dt

d^jx, Q

Q
'

at

The equation cannot contain terms independent of Y(x,

i),

'""

or terms such

as \T(x, OFWe shall use requirement 3 as an aid in finding the form of the Schroetake one of the free particle wave functions 'F/s, t),
dinger equation.

We

derivative of

T, such

as dY/dx,

is

linear

inT

T = <rfF + a^
x

then

because

if

SCHROEDINGER'S THEORY OF

168

and then

QUANTUM MECHANICS

[Ch. 7

differential

an equation involving only terms in this function and its


which is consistent with (7-6). This equation will give us a
equation which has the free particle wave function as a solution

when

t)

find

derivatives

V(x,

= V

We

a constant.

shall then postulate that,

general case in which the potential V(x,

t) is

even for the


not a constant, the solutions

of this differential equation are the correct wave functions to be associated


with a particle moving under the influence of such a potential.
Let us take the simplest free particle wave function discussed in the

Y = sin 2n(kx -

previous chapter,

Y,

sin

In terms of

vt).

(Kx

we have assumed

-'

V(x,

= K cos {Kx -

t)

cos

(Kx

=
2-y
2

cot),

(7-7)

must

-K

- = -co
f

cot),

dt

We

this is

K and

to

are constants

dx

co

co,

= V We find

ox

and

cot)

Evaluate some of its derivatives, remembering that


because

sin

(Kx

sin

(Kx

cot)

(7-8)
2

cot)

dt*

find

an equation consistent with

K
2m

+ V =

hco

(7-9)

The fact that differentiating twice with respect to x brings in a factor of


and differentiating once with respect to t brings in a" factor of co
suggests that we try the equation

K2

d 2xV

lit
where a and

/?

?W

**,->%

are constants to be determined.

into this equation,

-a sin (Kx -

we

cot)K 2

o-m

Inserting (7-7)

and (7-8)

find

sin

(Kx

cot)V

-/S cos (Kx

coi)co

(7-11)

Even though the constants a and /9 are at our disposal, we obviously


cannot make this equation agree with (7-9) for all possible values of x
and t, and so this attempt has failed.

The observation
into cos (Kx

cot),

that the differentiation process changes sin

and

vice versa, suggests that

we

(Kx

cot)

try again to construct

the differential equation by starting with a combination of two functions


of the form (7-7), one being a sine and the other a cosine. That is, we try

>

cos (Kx

cot)

sin

(Kx

cot)

(7-12)

Sec. 2]

THE SCHROEDINGER EQUATION

169

where y is a constant to be determined that gives us some freedom which,


might be guessed, we shall need. Evaluating some derivatives, we find

it

^ = -A

sin

(Kx

+ Ky cos (Kx -

cot)

cot)

dx

^-' = -K 2 cos {Kx 3T

*-

co sin

(Kx

- K2y sin (Kx -

cot)

cot)

coy cos (Aa;

cot)

(7-13)

wf)

dt

We

again assume the form of the differential equation to be given by


equation (7-10) and substitute these derivatives into that equation. This
gives

xK y sin (Kx - cot) + V cos (Kx - cot)


+ V y sin (Kx cot) = fico sin (Kx cot) /3coy cos (Kx cot)
[-a.K2 +V + ficoy] cos (Kx - cot)
sin (Ax - cot) =
+ [-<x.K y + V y olK2 cos (Kx

cot)

fico]

If this is to

be true for

all

values of x and

cosine and the sine must vanish. That

(7-15),

and three

the coefficients of both the

is,

-K + V =

-Py<o

(7-14)

-aA2 + V =

&
V

(7-15)

This looks encouraging;

/,

we have

co

three equations to satisfy, (7-9), (7-14),

free constants, a,

/S,

Subtracting (7-15) from (7-14)

y.

gives

jiyco

= -1/y
/ = -1
_
y = V-1 =

co

(7-16)

Substituting this into equation (7-14) yields

-ocA2
Comparing

this

+ V =

equation with (7-9),

we

Ttfco
see that

= -h2j2m
= n, p = ^hji =
a

^tp

ih

(7-17)

There are two possible

QUANTUM MECHANICS

SCHROEDINGER'S THEORY OF

170

+i

then

ft

or y

= + ih,

/.

and the

sets

of solutions, depending on whether

Following

common

differential

equation (7-10)

fc2

o2uy

we

we

take

+i;

take y

is

W""?
3 vp

-r7t
2m

car

This equation

practice,

[Ch. 7

(7-18)

3t

consistent with the three a priori requirements.

is

Equation (7-18) has been obtained for the special case of a constant
potential V(x,

= V

t)

from a

starting

free particle

wave function t
We now assume

similar to those used in the previous chapter for this case.

which V(x, t) is not a constant, the differential


equation controlling the propagation of the wave function is of the same
that, for the general case in

form

That

as (7-18).

is

It is

equation to be

-we postulate the

oxr2

2m
This

is,

T(x,

V(x,

t)

ih

(7-19)

ct

the famous Schroedinger equation.


a second order partial differential equation

partial derivatives) for

"^*'

T as a function of x and

similar to other partial differential equations

However, there

(i.e.,

In

t.

which

involving second

many

respects

it is

arise in the theories of

one feature of the Schroedinger


way from the equations of
classical physics.
The Schroedinger equation contains the imaginary
number i. As a consequence, its solutions are necessarily complex
functions of x and t. We have already seen one example of this; we were
forced to take a complex function
classical physics.f

equation which

sets it off in a

is

very striking

Y = cos (Kx
r

cot)

sin

(Kx

cot)

(7-20)

for the case of the free particle.

3.

Interpretation of the

Wave

Function

f) is an inherently complex function.


We did
Chapter 6 when we developed some of the qualitative
properties of the free particle wave functions. However, each of the arguments of that chapter could be stated again, always replacing the phrase
"wave function" by, for example, the phrase "real part of the wave

The wave function Y(x,

not

know

this in

t See the equation (7-58) for the propagation of transverse

waves in a stretched

string.

Sec. 3]

INTERPRETATION OF THE

function."!

WAVE FUNCTION

171

In every case the argument would remain valid and the


Recall also that in the discussion of

conclusion would remain the same.


the diiference between group

and wave

velocities, in section 2,

Chapter

6,

described the results of a hypothetical measurement of the value of


This
the wave function for various coordinates a; at a certain instant t
was done simply to provide something concrete to think about. If we

we

were to repeat that discussion, we would now have sufficient acquaintance


with the idea of a wave function to render such an artifice unnecessary.
We would also realize that the hypothetical instruments for measuring the
value of a wave function cannot possibly

exist, since the

value of an in-

herently complex quantity cannot be measured with an actual physical

instrument.

wave functions are complex functions should not be


considered a weak point of the quantum mechanical theory. Actually,
it is a desirable feature because it makes it immediately obvious that we
should not attempt to attribute to wave functions a physical existence in
the same sense that water waves have a physical existence. That is, we
should not try to answer, or even pose, the questions: Exactly what is
waving, and what is it waving in ? The reader will recall that consideration

The

fact that

of just such questions concerning the nature of electromagnetic waves


led nineteenth century physicists to the fallacious concept of the ether.
the wave functions are complex, there is no temptation to make the
same mistake again. Instead, it is apparent from the outset that the wave
functions are computational devices which have an existence or at least
a significance only in the context of the Schroedinger theory of which
they are a part. This point is emphasized by the fact that wave functions
never enter in the Heisenberg theory. Yet that theory is completely
equivalent, in terms of its end results, to the Schroedinger theory. These
comments should not be construed as implying that the wave functions
are lacking in physical interest. We shall see in this section, and in section
8, that a wave function actually contains all the information which the
uncertainty principle allows us to know about the associated particle.
In order to obtain this information we must have a quantitative connection between T(x, i) and the dynamical quantities describing the
associated particle. Now, it has been indicated in the previous chapter
that there must be a relation between some measure of the intensity of
Wipe, t), in the region of the coordinate x at the time t, and the probability

As

of finding the associated


apparent that T(x, t) cannot
because the probability density is a real

per unit length, or probability density P(x,


particle in that region at that instant.

simply be set equal to P{x,


t

The real

(7-22).

part of the

t)

wave function Y(x,

t),

It is

t) is

the function R(x,

t)

denned by equation

QUANTUM MECHANICS

SCHROEDINGER'S THEORY OF

172

[Ch. 7

wave function is complex. However, a way to make


was proposed in 1926 by Born in the form of the following

quantity whereas the


this association

postulate:
If,

at the instant

a measurement

t,

with the wave function Y(a;,


particle will be found at

made

is

to locate the particle associated

then the probability P(x,

r),

a coordinate between x and x

P(x,t)dx

dx

i)

dx that the

is

= xF*(x,t)Y(x,t)dx

(7-21)

The symbol W*(x, t) represents the complex conjugate of T(a;,


meaning of this term, consider the complex function

i).

To

illustrate the

*{x,

t)

R(x,

t)

(7-22)

I(x, i)

where R(x, f) and I(x, t) are real functions. (A complex function can
always be decomposed in this manner.) Then the complex conjugate of
this function is defined as

T*(*,

An

example we

i)

come

shall often

R(x,

across

Y(z,
This provides a convenient
(7-20) because e iz
(7-23),

it is

easy to

I(x, t)

(7-23)

is

i(-

Kx -<ot)

(7-24)

way of writing the free particle wave function


From this relation and the definition
i sin z.f

cos z

show

t)

that

Y*(x,

t)

<.-*<**-<><>

Now let us evaluate T*T from equations

T*T = (R Y*Y = R*-

(7-22)

(7-25)

and

(7-23).

This gives

+ il)
I =R? + P

U)(R
i

= I. Thus we see that Y* ?" is always a real function, and Born


i
not inconsistent in equating T*(a;, O^fo
to P{x, i).
Of course, this does not prove that these two quantities must be equated
because there are other functions of T(a;, t), such as the absolute value
|T(a;, t)\, which are also always real.
Some justification for Born's

since

is

choice of "*(#,

t)

T(a;,

f)

to give the probability density can be found

t This equality can easily be proved


these functions:
e( '

from a consideration of the

(izy
(fc)*
(fe)
+
+ (te)
- 4+
+
+ 2!
1!
3!
4!

1 4-

z2

-I-

z4

=l--+_-- +
sinz=z __ + _-_ +
z

co S2

z6

z>

--.

...

series

expansions of

Sec. 3]

WAVE FUNCTION

INTERPRETATION OF THE

from the following argument. Consider

173

T and Y* which are, respectively,

solutions to the Schroedinger equation,

aw
- + V =ih
t2 22ur

2m

(7-26)

da?

dt

and the complex conjugate of the Schroedinger equation,!

t2

2m dx

avp*

+ VY*

ih

3*2

3x2

(7-26')

dt

Multiply equation (7-26) by T* and (7-26')

'2m \

g2w*

JL _!_

byT and subtract.


\

This gives

dt J

dt

which reduces to
o
fc2

32W

\p* " T
dx*
2m\

_ _2_
and further

T
_ ip v3 *F*\
=
)

dx2

j^

_3 vp*ip
dt

to

2mdx\

dx

dx

dt

Integrating both sides over x between the limits x x to x2

2m

Jx!

dx \

dx

dx

we

find

J Xl dt

which gives
ih T
dW
dW*!**
+ Jl Y* 1 _ Y 2i-

2m L

3x

3a;

J Xl

=-3

C x*

Y*Wdx

(7-27)

dt J Xl

Take the complex conjugate of every term of equation

(7-26).

This gives

3Y\

easy to show from the definition (7-23) that the complex conjugate of a product
equal to the product of the complex conjugates. Thus we have
It is

(-)W)

is

="0*

According to equation (7-23), the complex conjugate of a purely real quantity, such as
(h 2 l2m) or V, is equal to that quantity. The complex conjugate of a purely imaginary
quantity, such as (ih), is equal to the negative of the quantity. Furthermore, it is easy
to show from (7-23) and the mathematical definition of a derivative that the complex
conjugate of the derivative of a function is equal to the derivative of the complex
conjugate of the function. From these properties we obtain immediately equation
(7-260.

QUANTUM MECHANICS

SCHROEDINGER'S THEORY OF

174

Now let us

take for
vp

Y the free particle wave function (7-24).

HKx-iot)

\p*

an(j

[Ch. 7

Then

-i(Kx-af)

Differentiation yields

7)W

dx

and

? XV*

-iKe~ Kx at)
i<-

= -iKY*

dx

Substituting into equation (7-27),

we have
>2

KV*^

is

v,

where v

the velocity of the

is

Therefore

(Y*Y)
This

= pjm =

equation (7-5), hK/rn

particle.

w*Ydx

ut Jxi

x1

From

r*2

=-

really

(vW**F)x=Xi

= |-

"*

Y*Y r

(7-28)

not a very interesting equation for the case of a free particle,

where V(x, t) = constant, because Y*Y = e -w-<t)S Kx m = 1, r =


= 0. However, if we consider a
constant, and the equation amounts to
case in which V(x, t) is a very slowly varying function of x and / (almost a
constant), then, at least for a restricted range of the x axis, the wave
function can be written
*p

_ A eHKx-wt)

where A, K, and co are very slowly varying functions of x and t. In this


= A 2 and also t> is not constant, but we shall still obtain equation
case Y*
as
we
(7-28)
can ignore derivatives of A, K, and co in evaluating dWjdx
and 3Y*/dx. Under these circumstances we can use equation (7-28) to
confirm the identification ofY*Y with the probability density by comparing
this equation with the one dimensional conservation equation of flowing

fluids,

{vp) x = Xl

(vp) x=Xt

= d

C Xi

p dx

(7-29)

Ot Jx,

(S) x=Xl

This equation

p
is

is

is

% dx

- (S)._ = f
at Jxt

mass

flux

The quantity
The quantity vp = S

illustrated schematically in figure (7-1).

the mass density of a fluid, and v

the

(7-29')

of the

fluid,

which

is its

is

the

velocity.

mass crossing a given point x

WAVE FUNCTION

INTERPRETATION OF THE

Sec. 3]

per unit time, and the

left side

of the equation

is

175

equal to the mass of fluid

flowing into a region per unit time at x1; less the mass flowing out per unit

The integral is the total mass contained in the region xx to


x 2 so the right side is just the change in mass in that region per unit time.
Thus the equation simply states that in the region xx to x2 the fluid is

time at x %

neither created nor destroyed

We

conserved.

is

it

show

in the next

paragraph that the same kind of conservation equation must apply to


the quantum mechanical probability. By comparing equations (7-28)
and (7-29), we see that this will be true if we accept Born's postulate.

S * = *2

*1

Figure 7-1.

Illustrating

*2
the one dimensional conservation equation.

The reason why we must have conservation of probability is easy


" T*(x, t)W{x,

Consider the integral

oo

and

r
and that

**(x,

must be true

this

conserved. Unless

we

+ oo

at the time

t)

for all

This

(x,t)dx

It is clear

t.

Thus the

t.

are willing to

is

to see.

just the total prob-

with the wave function T(x,

ability that the particle associated

somewhere between
must have

i)

dx.

make some

t)

be

then that

we

(7-30)

total probability

must be

very special assumptions,

can only be achieved by conserving probability in every region x L to x 2


we must have a conservation equation of the form (7-29).
Before leaving this argument, we take the opportunity to obtain from it
the formula for the probability flux S(x, t). This quantity is the probability per unit time that the particle associated with the wave function
this

In other words,

T(a;,

i)

will cross the point x, in the direction

By comparing

coordinate.

in the case where V(x,


v**(x,

(x,

i)

t).

An

i)

of increasing values of that

equations (7-28) and (7-29')


is

we

a slowly varying function, S(x,

expression for S(x,

t)

see that,
i)

is

just

of general validity can be

obtained by considering equation (7-27), which was derived without any


approximations. If

we postulate

the probability flux to be


"

S(x

t)

= -

J* \v*(x,

?*&<> - T(x,

2mL

3W * (X

dx

'

f)

(7-31)

dx

then equation (7-27) 'becomes a quantum mechanical probability conservation equation of general validity.
postulate.

It is

easy to

show

that S(x,

We take this as justification for the


i) is

always

real.

SCHROEDINGER'S THEORY OF

176

4.

QUANTUM MECHANICS

[Ch. 7

The Time Independent Schroedinger Equation


The Schroedinger equation

Y as a function of x and

(7-19)

is

a partial differential equation for

A standard technique for the solution of such

t.

an equation is to look for solutions T(, t) which are products of a function


of x times a function of t, that is, solutions of the form

Y(*,

We

xp(x)

such solutions exist

(7-32)

<f>(t)

the potential energy is a function


procedure reduces the problem of
solving the partial differential equation to the problem of solving two
ordinary differential equations.
shall see that

of x alone.

We

if

shall also see that this

Let us substitute (7-32) into (7-19), assuming V(x,

y (x)
-T-f-%
2m oxr

#*>

F(a:

>

*<*> #*>

ih

t)

V(x).

This gives

**) *
f
at

Now
ax2

dx2

dx2

the notation di ip(x)ldx2 being redundant with

function of x only.

d2y(x)\dx2

since

xp

is

we have

Similarly

dt

dt

and the equation becomes

2m

+ n*) **)

dx*

Dividing both sides by y(x)

i**) 4
dt

<f>(i),

tf.x)

<Kt)

2m dx2

we have

V(X) W(X)}
r

ill

4>{i)

dt

The

left side of the equation depends only on the variable x; the right
depends only on the variable t. Since x and t are independent variables, both sides must be equal to a quantity which depends on neither
x nor t in order that the equality of the left and right sides of the equation
can be true for arbitrary values of the variables. Thus both sides must be
equal to the same separation constant C. That is,

side

-M-f^ +
xp(x)

2m

dxr

**>*>}
1

(7 - 33)

THE TIME INDEPENDENT SCHROEDINGER EQUATION

Sec. 4]

177

and

f^ =

(7-33')
'

dt

<Kt)

Equation (7-33') is a simple first order ordinary


^ as a function of t. It has the solution

#t)

differential

equation for

- iCt/H

(7-34)

This solution can be verified by differentiating and substituting into


equation (7-33')

dt

which shows that

-^L

in-

#0 a

The function

<f>(t)

in

Q.E.D.

C;

#r)L

a complex oscillatory function of time,

is

tfi)

e~

iCtin

cos (Ctjh)

with frequency v given by 2ttv

Thus

Cjh.

sin (Cf///)

Cjlirh

C//i.

But,

according to equations (7-2) and (7-3), we must have v


E\h, where E
is the total energy of the particle, because <(f) gives W(x, i) its time
dependence. Thus C must be equal to the total energy E. Knowing this,

we may

write equation (7-33) as

^ +
f
2m

V(x)

y>(x)

Ey>(x)

(7-35)

ax*

C=E

Setting

we can

in equation (7-34),

write the solution to the

Schroedinger equation as

Y(x,

where

%p(x) is

y{x)e- iEtin

(7-36)

a solution to equation (7-35), which

is

called the time

independent Schroedinger equation.

an ordinary second order differential equation for y> as a function


contains no imaginary numbers and, consequently, its solutions
are not necessarily complex functions.! In section 7 we shall show

It is

of

x.

y>(x)

Of

It

course, they can be complex

linear in y>(x) so

we can always

if

we

wish, because the differential equation

take a solution which

is

the

sum of a

is

real solution plus

times a second real solution, provided the two solutions correspond to the same value
of E. This is proved in the footnote on page 189. An example is the wave function
(7-24), which is a solution for the potential V(x, t)
0. This can be written as
"{x, t)

y>(x) <(f),

and
y>(x)

where

eiKx

cos

=
#<) =

Kx

e~ imt
sin

Kx

-*/

SCHROEDINGER'S THEORY OF

178

that this equation

is

QUANTUM MECHANICS

[Ch. 7

very closely related to the time independent differ-

wave motion. The functions f(x) are called


The reader is cautioned to keep clearly in mind the
between the eigenfunctions ip{x) and the wave functions Y(a;, t),

ential equation for classical

the eigenfunctions.
difference

and

also

the

difference

between the time independent Schroedinger

equation and the Schroedinger equation.

5.

Energy Quantization

in

the Schroedinger Theory

Energy quantization appears in a very natural way in the Schroedinger


theory. In this section

of

how

we

comes about.

this

sented in Chapter

present, in

some

detail,

a qualitative description

A number of quantitative examples will be pre-

8.

Consider a particle of mass m moving under the influence of a force


F(x, t), Corresponding to this force is a potential energy V(x, t). The
two are related by the equation
ey
F(x,

a
t)=

dV{x,

t)

(7-37)

ox

To determine

the motion of the particle under the influence of this


from the Schroedinger theory, we must solve the Schroedinger
equation with the potential V{x, t). Now consider the class of problems
for which the potential is a function of x only. (The more complicated

potential

case of a time dependent potential will be discussed in a subsequent

For a time independent potential, we know that there are


form *(x, t) = f(x) <j>(t). Since <f>(i) has already been
evaluated (equation 7-34), the problem of finding these solutions reduces
to the problem of finding the solutions ip(x) to the time independent

chapter.)

solutions of the

Schroedinger equation for the potential V(x) pertinent to the problem.


The behavior of the particle is then specified by the functions W{x, t)

through the probability postulates (7-21) and (7-31). Energy quantiarises because, in solving the time independent Schroedinger
equation, we shall find that acceptable solutions exist only for certain values
zation

of the total energy E.

To be an
Iy)(x)/dx

acceptable solution, an eigenfunction

must obey the following

dwi iC)
la. y>(x)

must be

finite.

y>(x)

and

its

restrictions for all values of

lb.

derivative

a;:

must be finite.

dx
(7-38)

dwi
r CC must be continuous.
2b.
I

2a. w(x)

must be continuous.

dx

ENERGY QUANTIZATION

Sec. 5]

If f(x) or drp(x)jdx

Y(x,

t)

179

were to violate requirement

e-

lEt/n

y>(x)

dY(x,

and

la or lb,

t)

then

-iEt/n dip(x)

dx

dx

would

also.

This would

mean

that the probability density

^,0 = ^,0^,0
and/or the probability flux

2m
were not well defined
all

dx

that

is,

values of the variable s.f

dx

did not have a finite and definite value for


i)

and

between potential and

total

This cannot be allowed since P(x,

*o*
Figure 7-2.

energies

S(x,

t)

relation

time independent Schroedinger equation.

describe the behavior of a real physical particle

defined.
2a.

An energy diagram showing the

in a typical

In order to have requirement

The need

for requirement

lb,

we must

2b can be demonstrated by considering the

time independent Schroedinger equation (7-35), which

dx2

and must be well

also have requirement

we

write

J1P

(7-39)

tf

For finite V(x), E, and ip(x), we see that tPy^/dx2 must be finite. This
demands that we impose requirement 2b.
Keeping in mind these restrictions on the acceptable form of y(x), let
us mentally solve the time independent Schroedinger equation for a
potential energy V(x) of the
solve the equation"

in figure (7-2). By "mentally


we imagine going through exactly the

form indicated

we mean

that

same type of integration procedure which would be used

in finding

t All the eigenfunctions we shall deal with in the next few chapters are single valued,
so the requirement that P(pc, t) and S(x, t) have a definite value for all values of x is
automatically satisfied. However, for the
atom eigenfunctions treated in Chapter 10
it will be necessary to consider this requirement explicitly.

180

SCHROEDINGER'S THEORY OF

QUANTUM MECHANICS

[Ch. 7

solutions to the differential equation with the aid of a high speed electronic

computing machine. This technique


of the function V(x)

is

used in situations in which the form

such that the standard mathematical methods for


obtaining an explicit integration of the differential equation (which will
is

be presented in the following chapters) are not applicable.


In order to integrate the differential equation (7-39), we must know the
value of the constant total energy E of the particle. But E is not known

initially

and

see if

choose
Energy

we know is the potential V(x). Consequently we must


choose a value for this constant, put this value in the equation,
we can find an acceptable solution. The value of E which we

all

arbitrarily

on the

indicated

is

We

E.

plot in figure (7-2)

note that, for the value of

=E

divide the x axis into three regions


first

and

third

second region

line:

chosen, there are two

and the line Energy = V(x), which


x < x', x'
x ^ x", x > x". In
regions the quantity [V(x) E] is positive; in the
Energy

intersections of the line

the

by the horizontal

this quantity is negative.

In integrating the differential equation on a computing machine we must


start at some point x
which we shall take such that x'
x
x", and
initial

equation

linear in ip(x), the value

is

value for the function

Let us choose [v(x)L

<

<

assume an

We

xp{x) at

we choose
must

for

Since the

that point.
[fix)],,

is

immaterial.

assume an initial value of


Then, knowing the values of
the function and its derivative at the point x we can calculate their values
at the nearby point x as follows. For (x x x ) very small, [ip(x)] x is
given simply by

+l-t

the derivative [df(x)/dx] x

also

at that point.

ix*)]*,

[**)].

A similar formula for [dii>(x)ldx]x


equation (7-39), which

r^

1
Oi
l ax ^ Xo

*o)

can be found by using the Schroedinger

is

dx2

Lv(x)

\V{x)

VV{X)

~E

tf

*)

or

J tfy(x) = 2m
"|

E\y{x) dx

dx J

So
dy(x)}

dy>(x)
-

dx

provided (x 1

) is

dx*]

A?

^ ^^ ~
v

*o }

very small.

t As y>(x) is not necessarily a


that it is real.

complex function, we simplify

this

argument by assuming

ENERGY QUANTIZATION

Sec. 5]

181

Repeating the procedure from the starting point x x , we can find y>(x)
dy)(x)/dx at x2
Continuing in this manner, we can trace out the
function and its derivative over the entire x axis. Now note that in
x
x", where [V(x) E] < 0, the sign of the change in the
the region x'

and

^ ^

derivative for increasing values of

if

is

This means that the function

itself.

the value of the function

is

is positive,

opposite to the sign of the function

concave downwards in that region


and concave upwards if its value is

+1

A'
'
i

t.

Figure 7-3.
Illustrating three unsuccessful attempts at numerical integration of a typical
time independent Schroedinger equation.

In the regions x

negative.

function

concave downwards

On curve
x

< x'

concave upwards

is

if

and x"

< x,

where [V(x)

the value of the function

the value

of figure (7-3)

which traced

if

is

E]

>

the

0,

positive,

and

negative.

is

we record

the results of a calculation, starting

and dxpjdx in the direction of increasing x. For


x < x" the function was concave downwards, but on passing x" it became
concave upwards. Although the derivative was negative at x = x", it
soon became zero, and then positive. The function ip then started increasing, and matters rapidly went from bad to worse because, according to
at

ip

Schroedinger's equation, the rate of change of the derivative,


is

proportional to

ip.

we found

In this calculation

ip

i.e.

cPipjdx2 ,

oo as x

oo.

not acceptable behavior for an eigenfunction, we made a


second attempt in which we tried to avoid this divergent behavior by
Since this

is

changing our choice of the

y>

value of the derivative

initial

[dip/dx] x

The

We

were not quite successful because


became negative in the region where the function is concave downwards

results are indicated in curve 2.

if its

value

is

negative.

function should

now be

The

difficulty in obtaining

apparent.

It

an acceptable eigen-

should also be apparent that, by

making exactly the right choice of [dxpjdx]x we shall be able to find a function whose acceptable behavior with increasing x is as shown in curve 3.
,

SCHROEDINGER'S THEORY OF

182

This

QUANTUM MECHANICS

asymptotically approaches the x axis.

ip

concave upwards
In figure (7-3)

it

becomes, so

it

The

closer

it

[Ch. 7

gets the less

never turns up.

we also indicate with a dashed curve the results of extend-

ing the function of curve 3 in the direction of decreasing x.

From the dis-

cussion above we must expect that the function will, in general, diverge when

we extendit to small x. We cannot adjust \dy\dx\ as that would disturb the


situation at large x. Nor can we attempt to prevent divergence at both large
,

7-4.

Figure

The

first

three

acceptable

solutions

of a

time independent

typical

Schroedinger equation.

and small x by joining two functions of different derivative at x = x This


ruled out by the requirement that the derivative of an eigenfunction be
.

is

everywhere continuous.
value of the total energy

We are forced to conclude that, for the particular


E which we initially chose, there is no acceptable

solution to the time independent Schroedinger equation. The relation between a function y) and its second derivative d^ipjch?, imposed by the equation for that value of E, is such that the function must diverge at either
large x or small x (or both).
However, by repeating this procedure for many different choices of the
energy E, we shall eventually find a value Ex for which the time independent
Schroedinger equation has an acceptable solution, y>x(x). In fact, there
will, in general, be a number of allowed values of total energy, Ex E2
E3
for which the time independent Schroedinger equation has
,

acceptable solutions
the form of the
for both small

shown
%()

y) x (x), y> 2 (x), y>3 (x),

first

....

and large x

is

the

same

For x

essentially similar to the behavior of yt^x).

derivative

is relatively

we

indicate

The behavior of

>

the behavior of

But, since

its

second

larger in magnitude, y^fx) crosses the axis at

value of x less than x but greater than

tp^x)

as the behavior of the function

in curve 3 of figure (7-3) for large x.

is

In figure (7-4)

three acceptable solutions.

x'.

When this happens,

some

the sign of

the second derivative reverses and the function becomes concave upwards.

At x

x'

the second derivative reverses again and, for x

asymptotically approaches the axis.

<

x',

the function

ENERGY QUANTIZATION

Sec. 5]

183

From figure (7-4), and from the time independent Schroedinger equation,

&? = ^ \y

- em*)

{x)

k2

dx2

see that the allowed energy E2 is greater than the allowed energy
Consider the point x where both ip x (x) and y2 (x) have the same value.
is apparent from figure (7-4) that at this point

we can

Ev
It

>

dx2

dx2

Thus

- E2 >

- E]\,

E% > E
From a similar argument we can show that E3 > E It is also apparent
(E3 EJ, etc., are not infinitesimals since,
that the quantities {E2 E
|

V(x)

V(x)

and

x),

for example, the quantity

d\x)
dx

d2y>i(x)
dx2

not an infinitesimal. Thus the allowed values of energy are well


separated and form a discrete set of energies. For a particle moving under
the influence of a time independent potential V(x), acceptable solutions to
the Schroedinger equation exist only if the total energy of the particle is
quantized to be precisely equal to one of the discrete set of energies

is

E\,

E2 E3
,

....

This statement

is

true as long as the relation between the potential

E is essentially as

shown in figure (7-2);


and an *" such that [V(x) E] >
for all x < x' and all x > x". For a potential of the type shown in that
figure, in which V(x) has a finite limiting value V as x becomes very large,
there is generally room only for a finite number of discrete allowed energy
values which satisfy that condition^ This is illustrated in figure (7-5).
For E > Fj, the situation changes. There are now only two regions of the
x axis: x < x' and x > x'. In the second region [V(x) E] will be
negative for all values of x out to infinity. But, when [V(x) E] is
negative, y>(x) is concave downwards if its value is positive, and concave
upwards if its value is negative. It always tends to return to the axis and
energy V(x) and the total energy
that

is,

as long as there exists an *'

t Potentials which approach the limiting value Vx very slowly can be an exception,
because the separation between the allowed energies En and E +1 becomes very small
as V t En becomes very small. In such a case it is possible for there to be an infinite
number of discrete values of E which satisfy the condition Vi > En An important
_1
which very slowly approaches the
example is the Coulomb potential V(f) oc
.

limiting value V%

= 0.

184

SCHROEDINGER'S THEORY OF

QUANTUM MECHANICS

[Ch. 7

is, therefore, an oscillatory function.


Consequently there will be no
problem of %p(x) diverging for large values of a;. Since we can always make
ip(x) asymptotically approach the axis for small values of a; by a proper
choice of [dy)(x)/dx] x we shall be able to find an acceptable eigenfunction
for any value of E > V
Thus the allowed energy values for
are
continuously distributed. They are said to form a continuum. It is evident
,

E>V

Continuum

Figure 7-5. The discrete and continuum values of allowed energy for a typical time
independent Schroedinger equation.

that, if the potential V(x) is limited for small values

of x, or for both large

and small values of a;, then the allowed energy values will form a continuum
for all energies greater than the smallest limit.

We

note that

when

the relation between the potential energy V(x) and


such that classically the particle would be bound to a
finite region of the x axis, in order to satisfy the classical requirement that
the total energy E can never be less than the potential energy V(x), then
according to the Schroedinger theory the total energy of the particle can

the total energy

E is

only be one of a discrete set of energies. When the relation between


V(x) and E is such that classically the particle is unbound and could move
in an infinite region of the x axis, then according to the Schroedinger
theory the total energy can have any value and the allowed energies form a
continuum. All the conclusions we have obtained from "mentally solving"
the time independent Schroedinger equation will be verified in the following chapters by explicit integrations of this equation.

6.

Mathematical Properties of the Wave Functions and


Eigenfunctions

We

(a)
have found that for a particular time independent potential
V(x) acceptable solutions to the Schroedinger equation exist only for

MATHEMATICAL PROPERTIES

Sec. 6]

which we

certain values of the energy,

185
in order of increasing energy as

list

E* E2 ,...,En ,...

(7-40)

These energies are called the eigenvalues of the potential. The eigenvalues
early in the list are discrete. However, unless the potential V(x) increases
without limit for both very large and very small values of a;, the eigenvalues
become continuous beyond a certain point. Corresponding to each eigenvalue

is

an eigenfunction
V>i(x)>

%(*)>

C7 41 )

Vn(x),

>

a solution to the time independent Schroedinger equation for the


potential V(x). For each eigenvalue there is also a corresponding wave

which

is

function

^(x,

t),

Y (z, 0,

Yn

(7-42)

(*, /),...

which is a solution to the Schroedinger equation for the potential.


According to equation (7-36), we know that these wave functions are
iE ^ /n

e-

The index

n,

Vl (x), e

- iE * t/n

y> 2 (x),

..., e-

iE " tfn

which takes on integral values from

used to designate a particular eigenvalue and


function and wave function,
(b)

called the

is

Each of the wave functions

Tn

wave

to oo,

(7-43)

and which

quantum number.
is a particular solution to the
Since that equation

functions, any linear combination of these functions

Let us check this for the simple case

also a solution.

W(x,

t)

ajYjfx,

t)

ajV^x,

Substituting this function into the Schroedinger equation,

3Y
F

+ VW - =
2m
h 2 d 2x

ih

dx*

dt

we have
h2 (

T\
2

+ V[a&x +
r

is

corresponding eigen-

(a;, t)

Schroedinger equation for the potential V{x).


linear in the

its

f n(x),

T - in{a^ + a,^ ) =

a2

2)

is
is

SCHROEDINGER'S THEORY OF

186

Since

and

2 satisfy

QUANTUM MECHANICS

[Ch. 7

the Schroedinger equation, both brackets vanish.

Thus
1

2m dx2

dt

+
It is

obvious that

this

fli

dx2

2m

+ FY2 -i/^l =0;

Q.E.D.

dt J

argument may be extended to show that the function

= 2 aY K (z,
=

Y(z,

(7-44)

where the a are arbitrary constants, is a solution to the Schroedinger


equation. In fact, it is the most general form of the solution for the
potential V(x).
(c)

Let us calculate the probability density ~V*(x, t^ix, t) for the


wave function is of the form of

general case of a particle whose associated


the

wave function

(7-44).

We

write

= 1 a/-*"VW

Y(s,

Then

the complex conjugate

Y*(z,

We

it

(7-45)

is

0=1 a?e

+OBrt

(7-45')

"tyi*(*)

have used a different symbol to designate the index of summation


two series to remind ourselves that these two series are independent

for the

and that

summation

their

series together,

Multiplying the two

indices are independent.

we havef

Y*(*, t)Y(x,

= | a*H a nV Z(x) n (x)


+ IIa?"nV>X*)Vn(x)e- i(En El)t/n
y>

t For the benefit of readers not proficient in handling series,


writing out a few terms.

[afe^i'/V

Iaie- i

ale iE i t l>hp*
i'"tyi

(7-46)

a$eiE Jl h y*

a&- iBttl'hp

we

illustrate this

a 3 e- iEa t l' y! 3
i

-\

[fl*aiV*Vi + a*aaV>*V>2 + a? "aVJVs +


+ [(.alche-Wz-EiMy)*^ + a*a ae-^8-iWty*Y>, +
+ K^e-^i-^WfysVi + aJflse-^^s-^X/VJVs +
+ (a?a e- B i--B8)/'V* Vl + *a,e-* -*>/>?% +
+(
) +
) +
i <-

fl

by

MATHEMATICAL PROPERTIES

Sec. 6]

We

187

see that in general the probability density

depends on time. Thus the


measurement made

relative probabilities of the various possible results of a

to locate the particle depend

on the time

at

which the measurement

is

made.

Next consider the special case of a particle whose associated wave


is one of the wave functions Y(a;, i) corresponding to a single

(d)

function

eigenvalue

En

Then

V*n(x, t)Y %{z,

= JW'e-WriLx) Vn(x) =

y>* (x) y> (x)


n
n

(7-47)

and we see that the probability density is independent of time. Even


though the wave function is time dependent, the relative probabilities of
the possible results of a measurement of the location of the particle do not
depend on the time at which the measurement is made. In such a case the
particle is said to

be in a stationary

state,

or eigenstate, of the potential

K(*).t

According to equation (7-30), the probability interpretation of


t) demands that any wave function *F(a;, i) must be normalized

(e)
x

*(x, t)W(x,

so that

it satisfies

the relation

E
If

we

*(*,

t)dx=l

(7-48)

consider the general form of the wave function

Y(*.
it is

HT{x,

= t a nW n (x,
n=l

t)

apparent that normalization of the wave function can be achieved,


any given instant of time, by a proper choice of the arbitrary

at least at

constants a n .% Of course, equation (7-48) must be satisfied at all times.


Later we shall derive a time independent relation between the constants

a n which, when

satisfied, will effect the

normalization of

*F(a:, i)

at all

times.
(f)

Consider the special case of a particle whose wave function ^(x, t)


i).
Then we have

simply one of the eigenstate wave functions ^Jx,


from equation (7-47)
is

W*(x,t)V(x,t)=y>*n(x) Wn(x)
t In section 12, Chapter 1 3, it will be seen that an atom does not emit electromagnetic
radiation when the atomic electrons are in eigenstates because its charge distribution,

which is proportional to yif * xF, is time independent.


J There are certain difficulties connected with normalizing the wave functions for an
unbound particle, for instance the free particle wave function ^(x, t) = ex( Kx ~'t \ However, these difficulties are not fundamental; they will be discussed in the first section of
-

the next chapter.

SCHROEDINGER'S THEORY OF

188

QUANTUM MECHANICS

[Ch. 7

In this case equation (7-48) requires that

V*(*)

V0) dx

(7-49)

Thus the eigenfunctions must also be normalized. This can always be


accomplished as follows. Assume that we have obtained an eigenfunction
V( x) by solving the time independent Schroedinger equation. We evaluate
the integral

r
Then,

if

we

construct a

"
J

oo

a;)

V'*(

(g)

The

V>J.X) dx

set

y>'

new

A,

a constant

function y>Jx)

and that

\,

dx

n (x)

y>'*(x)

y> n {x) is

A~^tp^(x),

we

see that

a normalized eigenfunction.

of eigenfunctions of the time independent Schroedinger

equation for a particular potential V(x) have an interesting and very useful
property expressed by the equation

xp*{x)

/:
The

integral, over all x,

dx

y) n (x)

l^n

0,

(7-50)

of the product of one eigenfunction of the set

times the complex conjugate of a different eigenfunction of the set always


vanishes. Because of this property eigenfunctions are said to be orthogonal.

The proof of equation


eigenfunctions yjz) and

is not difficult.
Consider two different
which are, respectively, solutions to the time

(7-50)
y> t (x)

independent Schroedinger equations


h*

A". + VWn = E
nWn

2m dx2

(7-51)

and

+ ^.=^.
~f
2m 7T

(7-51')

dx'

Taking the complex conjugate of the second equation, we obtain the


equation for ff(x),
fc2

J2

+ Wt = E ^*
f
2m TT
dx*

since everything in that equation

Then we multiply (7-51) by

y>*

is

< 7 -52)

real except possibly the eigenfunction.

and (7-52) by

ip

and

subtract. This gives

MATHEMATICAL PROPERTIES

Sec. 6]

Next we

integrate with respect to

2m

-E

( E,

189

oo to + oo, and find

x from

jjy r * _
.

vt

&. - ,. M) dI

which reduces to

Integrating the right side gives

- * jf%f

(*,

En

If the eigenvalues

and y

and

V> n

dx

=
[

Vf

^-

M] "_

(7-53)

are discrete, thus corresponding to eigen-

a bound particle, then we have seen


not only remain finite for very
large and very small values of x, but actually go to zero at these limits.
In this case the right side of equation (7-53) vanishes at both limits and,
since (E E)
0, we obtain the desired proof. For eigenvalues in the
continuum, the corresponding eigenfunctions remain finite for very large or
functions

fn(x)

(x) representing
t

in the previous section that tpjx)

and

y>j(#)

very small values of x (or both).

way

things in such a

Sometimes

happens that

it

However,

possible to arrange

it is still

that the right side of equation (7-53) vanishes.f

En E

for

^ n.

That

the eigenvalues

is,

corresponding to two different eigenfunctions are numerically equal. In


this case the two different eigenstates of the particle are said to be
degenerate. (Recall the degeneracy discussed in the Sommerfeld treatment

of the hydrogen atom.)

It is,

eigenfunctions can always be


t This

unbound

is

however, easy to show that degenerate

made

orthogonal. %

associated with the question of normalizing the

particle;

it is

discussed in the

wave functions for an

section of the next chapter.

first

% Consider y> and y> t which are solutions to the time independent Schroedinger
equation for the potential V(x) and for the same eigenvalue E = E t = E. Then the
linear combination
V>

is

also a solution.

To

y>,

We have

+ " lV " + K ^"Vn + "W^ ~ E(a<W + <Wi) =

a "Vn

check, substitute into the equation.

~2~dx* ^
But

this is

h* dhpn

X^n^

^Incidentally this

**

makes

h* dhp,

\
" **) + a \~ 2^ I* +

it

Vy"

~ Em '
)

QED

obvious that a linear combination of eigenfunctions

is

"

not a

SCHROEDINGER'S THEORY OF

190

QUANTUM MECHANICS

The normality and orthogonality of the

(h)

particular potential V(x) can be

of eigenfunctions for a
single equation

set

summarized in the

V>*( x) V>ri x)

[Ch. 7

dx

1,

n
(7-54)

l=n

0,

These properties often allow a considerable simplification of quantum


mechanical calculations.
(i) We now derive the relation between the arbitrary constants a
n of
the general form for a wave function

= I a nW n(x,

Y(x,

which must be satisfied in order that the wave function

satisfies

the normali-

zation requirement

"*(*,
JI

t)

(x,t)dx

CO

Integrating both sides of equation (7-46) over

we have

all x,

*(*, t)W(x, t)dx


JJ

oo

= X a*n a n f
v
n

Wn{x) dx

y>* (x)

oo

"
afaj-***-'* f v *(x)
11
J
In

y> n (x)

dx

oo

solution to the time independent Schroedinger equation in the normal case in which

Ei.)

Now choose

JftVidx
-co

Oi

y>*yi n

dx

J CO

Then

ipa is

orthogonal to

yt n

since

/*O0

/*0O

W*Wa dx

an

/*0O

y>*y>

dx

a,\

J 00

J CO

y*y>

dx

J CO
VnVl
a

Of

J 00

By applying the condition


pletely determined.

The

J_ 0O
VJV dx
J ~ oo

that y be normalized, the values of a

and a can be comy> n and y> are

pair of orthogonal degenerate eigenfunctions

used instead of the original pair,

y>

and

y> t .

MATHEMATICAL PROPERTIES

Sec. 6]

Since the double

sum

191

contains only terms for which

^ n, we may apply

n) to each integral appearing in that sum. We may


equation (7-54) (/
also apply equation (7-54) (/
n) to each integral in the single sum. Then

we

obtain

J:

V*{x,t)W{x,i)dx=^a*na n
n=\

Thus normalization of the wave function

I a*na n =

achieved

is

we have

if

(7-55)

n=l

Since this condition does not involve time,

we

always remain normalized.


(j) Consider a general wave function Y(a;,

see that the

wave function

will

t)

which

Schroedinger equation for the potential V(x). If we

wave function

for

is

a solution to the
the form of this

know

any particular time, say / = 0, then it is easy (at least


wave function at any other time t. At

formally) to evaluate explicitly the


the time

0,

we

write the function Y(a;,

as a

t)

sum of

wave functions ^Jx, /) of the Schroedinger equation


V(x). That is, we take equation (7-44), as evaluated
instant

0.

the eigenstate

for the potential


in (7-45), at the

This gives'

= fa nfn(x)

V(x,0)

(7-56)

n=l

(This can be thought of as a generalized type of Fourier analysis.) Multi-

plying both sides by the complex conjugate of one of the eigenfunctions,

and integrating over


f"
J qo

From

all x,

we

V ?(x)Y(x,

equation (7-54)

0)

obtain

dx

= f an f"

n=l
**

we

will vanish, except for the

v *(x)

tp n {x)

dx

oo

term of the sum the integral


which the integral equals unity.

see that in every

term n

in

Thus we have
J

This

tells

oo

us the value of the constant a v The constant a n can be evaluated


if we replace / by n throughout. Furthermore, the

from the same equation

value of a definite integral does not depend on what


integration. Let us then replace x by *', and write

an

J oo

W n(x')W(x',0)dx'

we

call

the variable of

192

To

SCHROEDINGER'S THEORY OF
wave function T(,

evaluate the

these values of

a,n

=2
K=

Y(x,

7.

The

Classical

In this section

t)

in equation (7-45)

f"
l

V-oo

QUANTUM MECHANICS

for an arbitrary value of t, we


and obtain

rp*
n {x')W{x',

-iBnt/H,

0)dx'

Theory of Transverse Waves

we

[Ch. 7

Wn (x)

in

insert

(7-57)

a Stretched String

shall develop the classical theory

of transverse waves

At each step we shall compare and contrast the


wave theory with the quantum mechanical wave theory that we

in a stretched string.
classical

Figure 7-6.

A small

element of a string at an instant when a transverse wave

have developed in the previous sections.

As we

is

shall see, the

matical properties of these two theories are often very similar.

we

passing.

mathe-

Con-

be able to illustrate the significance of much of the


quantum mechanical wave theory in terms of the familiar phenomena

sequently

shall

described by the classical wave theory.

Consider a string which is under tension. If the string is given an up


and down displacement at one end, the displacement will travel down the
string in the manner characteristic of a moving wave. The propagation of
this wave can be completely described by the wave function *{x, t), the
value of which is numerically equal to the transverse displacement at time
t of a point on the string at longitudinal coordinate x. We note that in the
classical theory the wave function is a real, measurable function with an
immediate physical interpretation.
The quantum mechanical wave
function is a complex function and is, therefore, not measurable. Furthermore, a physical interpretation exists only for the product of the quantum
mechanical wave function times its complex conjugate.
To discuss the propagation of either the classical or the quantum
mechanical waves, we must know the differential equation which the wave
function obeys. The quantum mechanical wave equation cannot be

TRANSVERSE WAVES

Sec. 7]

A STRETCHED STRING

IN

derived from classical physics and

The

classical

193

by postulate.
from Newton's laws

obtained, essentially,

is

wave equation can be derived

directly

as follows.

Consider a small element of the string located between the coordinates


+ dx. In figure (7-6) we show that element at an instant t when
the displacement is passing the point x. The figure can also be thought of

x and x

as part of a plot of the function Y(x, t) versus x, at the instant t. Now let
us impose the restriction
1
Our classical wave equation will apply

<

<

only to waves on the string of small amplitude. If


sin

cos

0=1

tan

length of element of string


the tension

The

and the net

force in the

Tsin

Now
t)

(0

tan

(0

dd)

Tcos

y direction

dd)

Tsin

9Y(a;, t)jdx,

with respect to

x,

is

is

then

= T- T=

is

which

T(6
is

dd)

T6

Tdd

the derivative of the function

evaluated at the point x and the time

dx
which

dx

a constant

net force on the element in the x direction

Tcos

Y(x,

T is

1,

t.

Thus

dx

the change in the derivative, per unit change in x, times the change

in x. Therefore

d
TdO=T ^^dx

dx 2

According to Newton's laws, this net force in the y direction must equal the
mass of the element of string times its acceleration in the y direction. The
mass is p dx, where p is the mass per unit length of the string. Since there
is no net x force, the element moves only in the y direction and its acceleration is just <PF(x, t)jdt 2 Consequently we have
.

dx2
or

~dx^~ ~ T

H
dt

~~

(?_58)

SCHROEDINGER'S THEORY OF

194

This

is

the classical

wave equation.

(of small amplitude) in a string.

QUANTUM MECHANICS

It

[Ch. 7

governs the propagation of waves


same equation governs

Essentially the

the propagation of electromagnetic waves (of any amplitude) through


space.

The

Y(a;,

to

wave equation relates the second space derivative of


second
time
derivative. All the constants in the equation are
t)
The quantum mechanical wave equation

real.

classical

its

2m
relates the

oar

ot

second space derivative of T(*,

t)

to

its first

time derivative,

The constants in the quantum mechanical wave


equation are partly real and partly imaginary. Nevertheless, the procedures
involved in solving these two partial differential equations are very similar.
Let us consider an example of the solution of the classical wave equation.
A string of length a is fixed at both ends and under tension T. The
string is given a small but arbitrary initial displacement, and we are
and

to the function

itself.

its subsequent motion. We define an x axis extending


along the length of the string with an origin at its center. Then we must

required to evaluate
solve the

wave equation,
T(z,

which

state that the

subject to the conditions

for

0,

a/2, for

(7-59)

all t

ends of the string are fixed. Similar conditions obtain

quantum mechanical wave function when it describes the motion


of a particle which is bound to a certain region of space. Such quantum
mechanical wave functions always go to zero at both large and small x.

for the

It is

obviously appropriate in this problem to look for solutions of the

form
W(x,

t)

y(z) 4(f)

we

Substituting this into equation (7-58),

d>(^)

^
Dividing both sides by
1

y(z)
Since x and

ip(x)

dx2
<j>(t),

rfXaQ
dx2

obtain

dW

dt

we have
p

T<f>(t)

rfV(t)

dt

= -K 2
must be
Thus we have

are independent variables, both sides of this equation

equal to the same separation constant, which

we

call

K2

the two second order ordinary differential equations

^- + *>(*) =

(7-60)

Sec. 7]

TRANSVERSE WAVES

A STRETCHED

IN

STRING

195

and

2 + tfItfO-0
at

(7-61)

Consider equation (7-60). It is identical in form with the time independent Schroedinger equation for the case V(x) O.t The general
solution of the equation is
ip(x)

where

A and B are

= A sin Kx + B cos Kx

arbitrary constants.
dip(z)

To prove

(7-62)

this,

evaluate

= AK cos Kx BK sin Kx

dx
and
dPipix)

= -AK* sin Kx - BK2 cos Kx =

->(:*)

dxr

and

+ Khj>(x) =

-Khp(x)

To

This gives

substitute into equation (7-60).

Q.E.D.

Q-,

determine the allowed values of the separation constant,

we apply

the

conditions (7-59), which read

T(a/2,

/)

y(a/2)

<Hf)

0,

for all

So
V(a/2)

(7-63)

Then we have
.

A sin

Ka

B cos

Ka =

and

A sm Ka
.

B cos

Substracting and adding,

we

Ka

2
obtain

=
Ka
Bcos =
A sm

Both of these equations must be

satisfied.

t The analogy is even closer than this. If we consider a string of mass per unit length
p(x) which is a function of x, we shall obtain an equation identical in form with the time

independent Schroedinger equation for a general potential V(x).

196

SCHROEDINGER'S THEORY OF

QUANTUM MECHANICS

[Ch. 7

no value of K for which both sin (Ka/2) and cos (Ka/2)


However, we can satisfy these equations either
by choosing K such that cos (Ka/2) vanishes and setting A equal to zero,
or by choosing K such that sin (Ka/2) vanishes and setting B equal to zero.
That is, we take

Now

there

is

vanish simultaneously.

0; cos

0,

which implies

with n

1, 3, 5,

or

B=

0; sin

(7-64)
0,

which implies

-n
Figure 7-7.

The

Thus there are two

first

For the second

=3

For the first

= B cos YITTX

1, 3, 5,

class,

a
(7-65)

class,
.

2, 4, 6,

three eigenfunctions for a string fixed at both ends.

classes of solutions.

y> n (x)

with n

xp n (x)

= A sin
.

JITTX

2, 4, 6,

These are the eigenfunctions for the string of length a fixed at botn ends.
In figure (7-7) we plot the first three eigenfunctions. Note the essential
similarity of these functions to the three quantum mechanical eigenfunctions for a bound particle plotted in figure (7-4). Also note the difference in their physical significance. In the classical theory, y>Jx) gives the
spatial dependence of the amplitude of the vibration. In the quantum
mechanical theory, y>*(x) ip n (x) gives the spatial dependence of the probable
location of the associated particle.
Next consider equation (7-61). By comparison with equations (7-60)
and (7-62), it is apparent that the general solution to this equation is
rf(0

sin

(KVTfr

t)

p cos (kVt/^ t)

(7-66)

..

TRANSVERSE WAVES

Sec. 7]

where a and
time,

/?

we may

A STRETCHED

IN

are arbitrary constants.

set

0,

By proper

STRING

choice of the zero of

and simplify equation (7-66)

<f>(t)

fi

cos

197

to

(kVTTp t)

(7-66')

no complex constants, it has a real solution


The corresponding equation in the quantum mechanical case
contains the imaginary number i, and its solution (7-34) is

Since equation (7-61) involves


(7-66').

(7-33')

necessarily complex.

Combining equations (7-59), (7-64), (7-65), and (7-66'), we find the


wave functions for a string of length a, density per unit length p, and tension
T, which is fixed at both ends:

Y(a>,

a n cos (27rr0

(7-67)

y, n (x)

where

mrx

1, 3, 5,

2, 4, 6,

= K [r =

cos

mtx

sin

and where
"n

The a n

2n

are arbitrary constants,

2a

/I

and the r are the vibration frequencies

This discrete set of frequencies constitutes the familiar


"fundamental," "second harmonic," "third harmonic," etc., produced by
any stringed musical instrument. In the classical theory the frequency is
of the string.

quantized. In the
is

quantum theory the energy

related to the frequency by the equation

is

E=

quantized, but the energy


hv.

An interesting difference between the two theories

is

that in the classical

theory the wave functions are not required to satisfy the normalization
condition (7-48); the integral P

T*T dx can assume any value because

the value of the constants a n which determine the amplitude of the vibrations, can assume any value that is not too large to violate the restriction
,

<

1.

However, the eigenfunctions of the

string are

orthogonal. The orthogonality


a/2

/:

This

is

Vi(*) ?(*)

dx

classical theory

of a vibrating

relation reads

= 0,

l^n

(7-68)

a/2

completely equivalent to equation (7-50) because, in the present

SCHROEDINGER'S THEORY OF

198

case, the range of


y> t (x).

is

QUANTUM MECHANICS

only from a/2 to +a/2, and because

The reader should

verify equation (7-68)

some of the integrals.


The general form of the

classical

W(x,

t)

=J
n=

tf

^(x)

by actually evaluating

wave function

combination of the functions (7-67). That

[Ch. 7

is

an arbitrary linear

y n(x)

(7-69)

is,

cos (IttvJ)

Normally, the form of the x dependence of the function W(x, i), at some
t, depends on the value of /.
This is analogous to the
situation described by equation (7-46). However, if we set the string into

particular time

^-* *,o) = (f-H)^


(

-/2

o/2
Figure 7-8.

string plucked at

center.

its

by giving each segment an initial displacement which is in the


form of one of the eigenfunctions, the form of the x dependence of Y(#, t)
will be constant in time. For instance, let us give the string a displacement
at t =
of the form Y(a;, 0) = C cos (ttx/o). Then, in equation (7-69),
we have

vibration

Ccos
In this case
the

all

an

wave function

=
at

0,

a = n=l
%a

except a x

n ip n (x)

C. Since the a n are time independent,


is given by

any instant of time

Y(x,

= C cos

t)

/.
\2tt
\

nx
IT

t) cos
la's
\

is analogous to a quantum mechanical eigenstate.


Consider a string which, at the instant t = 0, is given a displacement
of a form different from that of an eigenfunction; say we pluck the string
at its center, so that T(x, 0) = (a/2 \x\) d/(a/2). This is shown in
figure (7-8). To determine the subsequent motion of the string, we use
its initial displacement to evaluate the coefficients a
n At / = 0, equation

This situation

(7-69) reads

Y(*,0)

= f ?(*)

(7-70)

..

TRANSVERSE WAVES

Sec. 7]

To

A STRETCHED

IN

evaluate a particular constant, say a

equation by

ipi(x)

and

This gives

|*o/2

oo

ViO) ^O* 0)

199

multiply both sides of this

integrate over the range of x.

fa/2

STRING

= 2 a
V>l x) VnO) dx
J -a/2
n=l

da;

J-a/Z

Because of the orthogonality condition (7-68), every term of the summation


I.
We consequently have
vanishes except the term for n

f"% (ar)'F(a:,0)te
I

2
l> (*)] dx
f

/-a/2

which

is
a/2

f
a,

mWj

J
\2

J-/i

'

7 a/2

/a/a
a/2

J -a/2

Evaluating the eigenfunctions from equation (7-67), and performing the


integrations,

we

find

MlifirP,

0,

where we have replaced

by

n.

= 1, 3, 5,
= 2, 4, 6,

The

coefficients

of fjx) in the

(7-70) are non-zero only for odd n because of the symmetry of Y(x,

series

0).

The

series is

Y(*,0)=

-^sv

n = l,3,5,...7r'n

From

equation (7-67) this

is

Y(, 0)

2
=i,3,5,...

cos
Tl
7rn*

The coefficients of the cosine functions decrease rapidly with increasing


Thus a good approximation to
n, and the series rapidly converges.
W(x, 0) can be obtained by taking only the first few terms of the series.
This is illustrated in figure (7-9) by plotting each of the first three terms of
the series, their sum, and also the function

T(*,o)=(fto which the series converges.

W )A
/2

SCHROEDINGER'S THEORY OF

200

What we have done


of a

series

and

cosines, this

is

QUANTUM MECHANICS

[Ch. 7

to write the initial displacement Y(x, 0) in terms

of the eigenfunctions

amounts

to

y> n (x).

[As the fjx) happen to be sines


analysis' of Yfo 0).] This

making a Fourier

allows us to evaluate the displacement of the string at any instant of time


by simply inserting, into the series for the general form of the wave function
(7-69), the values of

^0*.

a n which we have determined. That

0=2

cos 2^0 c s
-n
n

=i,3,5,...77

is,

?22 +M-^
+ -mcos
2 2
2
ir

Figure 7-9.

Illustrating

^5

"

the Fourier anal/sis of a string plucked at

its

cos See

center.

A good approximation to T(a;, t) can be obtained from the first few terms
of the

The physical

series.

significance of this series can be interpreted as

we simultaneously set into vibration a number of


different eigenstates. As time progresses, these eigenstates vibrate independently, each at its own characteristic frequency v. Then at any
instant the net displacement at x can be obtained by summing the
follows.

At

0,

instantaneous displacements of the eigenstates at that value of x.

Using the general expression (7-71) for the a n we can write a general
,

solution to the problem of a string with fixed ends that


result

of an arbitrary given

initial

displacement

is

*F(a;, 0).

vibrating as a

The

solution

is

fa/2

Y(*.

0=2

^%

y, n

(x')W(x',0)dx'
cos (2nv n t)

y> n (x)

J-aji

This equation

is

completely analogous to equation (7-57), which gives the

quantum mechanical wave function for a particle moving in a


potential in terms of a given initial form of the wave function.

certain

8.

AND

EXPECTATION VALUES

Sec. 8]

DIFFERENTIAL OPERATORS

201

Expectation Values and Differential Operators

we compared some of

In the previous section


perties

the mathematical pro-

of the classical wave theory with certain similar properties which we

have developed for the quantum mechanical wave theory. We return now
to the task of continuing the development of the latter theory.
Consider a particle and its associated wave function Y(cb, t). In a
measurement of the position of the particle, there would be a finite
probability of finding it at any x coordinate in the interval x to x + dx,
as long as the

wave function

wave function

is

is non-zero in that interval. In general, the


non-zero over a certain range of the x axis. Thus we are
generally not able to state that the x coordinate of the particle has a
certain definite value. However, it is possible to specify some sort of an

average position of the particle in the following way. Let us imagine


making a measurement of the position of the particle at the instant t.

The

probability of finding

between x and x

it

dx

is,

according to the

postulate (7-21),

P(x,

t)

dx

= Y*(x, t)Y(x, i) dx

(7-72)

this measurement a number of times, always at the


same value of /, and recording the observed values of a;.f We can then use

Imagine repeating

the average of the observed values to characterize the position of the particle
at the instant

t.

This

we

call x, the expectation value

of the particle at the instant

It is

t.

-r xP(x,
J

of the x coordinate

evident that x will be given by


t)

dx

is just the value of the x coordinate weighted by the probaof observing that value and, upon integrating, we obtain the
average of the observed values. From equation (7-72) this can be written!

The integrand
bility

Y*(x,
v

xY(x,

t)

dx

(7-73)

(.

easy to see that an expression of the same form as (7-73) would be


appropriate for the evaluation of the expectation value of any function of
It is

x.

That

is,

~~

<

We

Y*(x,

x ix(x,

t)

dx

can visualize doing this either by making a number of measurements on identical


t
particles in identical environments, each particle being in the state associated with the
t), or by making a number of repeated measurements on the same
always with the particle in the state associated with the wave function ^(x, t).
t The terms of the integrand are written in this order to preserve symmetry with a
notation which will be used later.

wave function 'Vix,


particle,

SCHROEDINGER'S THEORY OF

202

QUANTUM MECHANICS

[Ch. 7

and
X

f(xj=! W*(x,t)f(x)Y(x,t)dx
J oo

where f(x) is any function of x. Even for a function which


depends on the time, such as the potential energy V(x, t), we

explicitly

may

still

write

P(iT0

because

all

value of

t,

V(x,

T*(*.

W(x,

t)

t)

dx

(7-74)

CO

measurements made <o evaluate V(x, i) are made at the same


and so the arguments of the previous paragraph would still

hold.

The coordinate x and the potential energy V(x, i) are two examples of
the dynamical quantities which can be used to characterize the state of a
Examples of other dynamical quantities are the momentum p
and the total energy E. The expectation value of these quantities is always
given by the same type of expression. For instance, for the momentum
particle.

p the

expression

is

W*(x,

pY{x,

dx

(7-75)

J 00

In order to evaluate the integral in equation (7-75), the integrand

Y*(x, t)pY(x, i) must be expressed as a function of the variables x and


In a problem in classical mechanics, p can always be written as a
/.
function of the variables x and/or t after the problem has been solved. For
instance, for a particle moving in a time independent potential, p can be
written as a function of a; since, after the classical problem has been solved,
the momentum is precisely known at every point of the trajectory. But

quantum mechanics this cannot


be done because p and x cannot be simultaneously known with complete
precision. We must find some other way of expressing the integrand in
terms of x and t.

the uncertainty principle shows us that in

clue can be

(7-24),

which

found by considering the


W(x,

Differentiate with respect to x.

^j*'

t)

iiKx -'ot)

iKe*'-~t

>

K = pjh,
ox

This gives

ox
Since

free particle

is

iK{x,

t)

wave function

AND

EXPECTATION VALUES

Sec. 8]

DIFFERENTIAL OPERATORS

203

which can be written


*[(*. 0]

= -tt!- [TOM)]
ox

an association between the dynamical quantity


ih djdx. A similar association between
the dynamical quantity E and the differential operator ih djdt can be
found by differentiating the free particle wave function with respect to t:

This indicates that there

p and

is

the differential operator

8T(a;

'

f)

-itue*^--"*)

-icoW(x,
V

t)
'

dt

Since

co

E/h, this can be written

at
It

may seem

particle

wave

of showing

that these associations apply only to the case of a free

function, but this

this;

actually not so.

is

There are several ways

perhaps the most convincing one

Consider the equation (7-3) defining the

film +

V(x,

classical total
t)

Let us replace the dynamical quantities


differential operators.

is

the following.

energy E:

=E
p and E by

their associated

Then we have

-L(-it\v(x,t) =
2m \

ih?-

dx!

dt

or

V ^^ =
-ff-2+
2m dx

ih

This is an operator equation.

To

(7

l
dt

obtain a differential equation,

"76 )

we multiply

the equation into the equality

TO,

= W(x,

and obtain

2m
This

is

dx2

dt

just the Schroedinger equation (7-19).

We

see that postulating the

associations

p-^-ih
dX
(7-77)
<> ih

dt

is

QUANTUM MECHANICS

SCHROEDINGER'S THEORY OF

204

equivalent to postulating the Schroedinger equation.

This

is

[Ch. 7
essentially

procedure originally followed by Schroedinger. The procedure


provides us with a powerful method for obtaining Schroedinger equations
the

cases than we have discussed in this chapter. We


Chapter 10 to obtain Schroedinger equations for three
dimensional problems, and for problems involving more than one

for

more complicated

shall use

in

it

particle.

Let us use (7-77) in the equation (7-75) for the expectation value of the

momentum. Then

r W*(x,

p*(x,

t)

dx

00

becomes

Y*(x, t)(-ih

J - oo

dx!

Y(x,

t)

dx

or

p=
We

-ih\ Y*(a;,
J-x

J^hJl dx

t)

(7-78)

dx

thus obtain an expression which can be immediately evaluated

know W(x,

f).f

we can

Similarly,

if

we

evaluate the expectation value of the

energy as follows

E=\

E=

T*(a;, i)[ih

*{x,i)E x{x,f)dx

E=ih)

-)

^fo

Y*(x,

J -oo

From

W(x,

f)

t)

dx

dx

(7-79)

dt

equation (7-76), this can also be written

f " **{x,
./-oo

i)(-

%- +
^
2m
ox'

V(x, o) T(z,

dx

t We see now the reason for the ordering of the terms in the integrand of p.
not be possible to have

-ih

T*^,

f *F*(*, /)(,
|

J 00
since this

is

meaningless.

-'*

I
I

J 00

Nor
for would
4ax

it

to have
be possible
pos

^ *(*, /)Y(iB, t)dx=

because this always vanishes.

yx dx

-ih [(*,

O^x, t)f

It

would

EXPECTATION VALUES

Sec. 8]

In general,
that

is,

AND

DIFFERENTIAL OPERATORS

205

we can obtain the expectation value of any dynamical quantity

a quantity which

a function ofx,p, and possibly

is

by evaluating

the integral

fix, P.

J"

^*(*. 0/op (*,

- it

t)

dx

(*,

(7- 80)

where the operator fQV (x, ih 3/ dx, i) is obtained from the function
t) specifying the dynamical quantity, by everywhere replacing p
by ih d/dz.f
We see that the wave function W(x, t) contains more information than

f(x,p,

just the probability density

P{x,

and the probability

i)=Y*(x, t)Y(x,t)

flux

2mL
The wave function

dx

ox

also contains, through equation (7-80), the expectation

values of the coordinate x, the potential energy V(x,

t),

the

momentum p,

the total energy E, and, in general, of any dynamical quantity f(x, p,

In fact, the wave function Y(a;,

t)

contains

all

t).

the information which, in

we can ever expect to learn


about the particle associated with that wave function.
Let us calculate the expectation value E of the total energy of a particle
consideration of the uncertainty principle,

whose associated wave function


Y(x,

is

of the general form

= | a n e- iE^ K Wn (x)

= | a nW n {x,

(7-81)

n=l

n-l

According to equation (7-79),

J -oo

E=\

dt

fateiE t/n y>?(x)ih


>

dx
2 a n (- ^Ae^^f^x)
n

=i

J-ooi=i

E = 2 2 Erfaj*-* Jf"
n li =
l

From

y> n (x)

dx

oo

equation (7-54)

t Cf. exercise 8 at

tf(x)

we

end of

see that the integral vanishes except

this chapter.

when

SCHROEDINGER'S THEORY OF

206
I

n, for

QUANTUM MECHANICS

which the integral equals

contributes only the single term

n,

[Ch. 7

So the summation over


and we have

unity.

n=l

or

E = I EaU n

(7-82)

is quantized and can only be equal to


one of the eigenvalues En any single measurement of this quantity can
only yield one of these values. However, as the wave function (7-81) is
a sum of a number of different wave functions ^(a;, t), each corresponding
to a different eigenvalue, any one of these eigenvalues could be found in
the measurement. In light of this fact, it is apparent that we should
interpret equation (7-82) as meaning that the quantity a*a n is equal to the
probability that a single measurement of the total energy will yield the value
En ; i.e., it is equal to the probability that the particle would be found in the

Since the total energy of the particle


,

eigenstate n.

Then

that equation will just give the average of the values

found in a large number of measurements of the energy of a particle in a


state described by the wave function (7-81), which is the definition of the
expectation value of the total energy. In general, E will not happen to
be equal to one of the eigenvalues. However, if the particle is known to be
in single eigenstate, sayY (a;, /), then in equation (7-81) a n = 1 if n = k,
otherwise. In this case equation (7-82) gives E = Ek as
and an =
would be expected.
fc

9.

The

Classical Limit of

Quantum Mechanics

Consider a particle moving under the influence of a potential V(x,


In the limit in which the characteristic distances and

momenta

t).

involved in

describing the motion of the particle are so large that the uncertainty

be ignored, we shall prove that the predictions of quantum


mechanics coincide with Newton's laws:f
principle can

dp

__

dVixJ)

(7g3)

dt

where

dx
= m
dt

The

first

equation

is

simply dpjdt

= F, where F

is

the force. Cf. equation (7-37).

THE CLASSICAL

Sec. 9]

LIMIT

OF QUANTUM MECHANICS

207

Such a proof was first given by Ehrenfest in 1927. It is considerably more


involved than showing that the laws of relativistic mechanics approach the
laws of Newtonian mechanics in the limit in which the characteristic
velocities are small

compared to the

velocity of light.

obviously of the greatest importance to be sure that

However,

it

is

quantum mechanics

agrees with classical mechanics in the appropriate limit.

In Chapter 6 we saw that, for large characteristic distances and momenta,


possible to associate with the particle a wave function W(x, t) in the

it is

form of a group involving uncertainties in these quantities which will


always be relatively small. Now for any wave function, and in particular
for ^(x,

i),

the expectation value of the coordinate x

is

T*xxF dx

J 00

we obtain

Evaluating the time derivative of this quantity,

m
^-\
Wis -[*
fa, + **!#?)**
J-,\
dtJ-oc
dt

dt

Substituting for

(7-26)

dt

and

dt

dY/dt and dW*jdt from the Schroedinger equations

(7-26') gives

HJ-S

The terms involving

2m

dx2

V cancel

out,

2m dx2

and we have

-*[ fa,** -**!&) i*


dx
dx
2m J- A
2

dt

Separate out the second term of the integrand, and integrate

it

by parts

as follows

dx =
xy ^.
2

dx

J-oo

Since

_ p w*& dx
[WMdx
dx
dx
L

J -oo

dv

J-oo

du

Y must be in the form of a group of limited spatial extent in order

that the uncertainty in the x coordinate be relatively small, both the

function and
grated term

x >
equal to zero, and we have

its

is

derivatives vanish as

Consequently the

oo.

xdvn
rx^dx=-(
-^dx
dx
dx
dx
d

J-oo

J-oo

wave
inte-

208

QUANTUM MECHANICS

SCHROEDINGER'S THEORY OF
by parts

Integrating

Again

this

again,

dx2

J-oo

[Ch. 7

L dx

J-oo

dx2

J-oo

reduces to

dx2

J-oo

dx2

J-oo

Putting this back into the equation for dxfdt,

2m

dt

Consider the bracket in the integrand.

d 2xF

dx

2
d (xW)y

dx

dx2 J

dx2

J-oo

can be written

It

dx2

dx

_ <PY_
dx

we have

dx

\
2*?

dx

3T

9Y

dx

dx

_2^
dx

Consequently

dx

_ih^^dX dx
m

dt

J-oo

3a;

According to equation (7-78), the expectation value of the


the particle

of

dx

gvp

Too

p=-ih\

W*

(7-84)

dx

J-oo

Thus we

momentum

is

see that

dx

dt

or

= m^-

(7-85)

at

Next
(7-84),

let

us evaluate the time derivative of p.

Differentiating equation

we have
dt

Again we

substitute for

J-A

dt

dx

dxdtl

dYjdt and dW*jdt from the Schroedinger equation.

THE CLASSICAL

Sec. 9]

OF QUANTUM MECHANICS

LIMIT

209

Then

2m

dt

J-<x>\

dx2

dx

J-ao\

This

dx9 /

dx

dx

is

2m

dt

Then we

term on the

first

= _Jf_

dt

2m

dx2

dx dx

'

W*
Ydx
dx

-|
/:
Integrating the

J- oo dx

dw* av
dx

right,

we have
2

_ TN! 3 T'

dx

dx

dV
r
vp* v

J-ao

xrf

dx

dx

see that this term vanishes because of the behavior of the

function as x -*

wave

Thus

oo.

&

T* (

J -oo

dt

\dxf

Y dx

(7-86)

dV/dx = dV(x, t)jdx is simply some function of x and t. Consequently, on comparing equation (7-86) with the general equation for the
expectation value of a dynamical quantity (7-80), we see that

Now

dp

= _ dvpt^
dx

dt

where dV(x,

t)jdx

is

the expectation value of the spatial derivative of the

potential energy.

Equations (7-85) and (7-87) are identical with Newton's laws (7-83),
In the limit in which the

except that they involve expectation values.

and momenta involved in describing the motion of the particle


compared to the uncertainties in these quantities, the
dynamical quantities all have precise values and it is no longer necessary

distances

are very large

dV/dx and dV/dx. Then the


become
completely identical with the laws of Newtonian mechanics. Quantum
mechanics does indeed agree with Newtonian mechanics in the limit in
which the uncertainty principle is not important. However, we shall see
to distinguish between x

equations

we have

x,

p and

when we
number of extremely

in the next chapter that,

predicts a

and

p, or

derived from the theory of quantum mechanics

quantum mechanics
phenomena which have no

are not at this limit,


interesting

counterpart in Newtonian mechanics.

QUANTUM MECHANICS

SCHROEDINGER'S THEORY OF

210

[Ch. 7

BIBLIOGRAPHY
Bohm, D., Quantum
Pauling, L.,

Book

Hill

Theory, Prentice-Hall, Englewood Cliffs, N.J., 1951.


and E. B. Wilson, Introduction to Quantum Mechanics, McGraw-

Co.,

New

York, 1935.

Quantum Mechanics, McGraw-Hill Book Co., New York, 1955.


Sherwin, C. W., Introduction to Quantum Mechanics, Henry Holt and Co.,
Schiff, L. I.,

New

York, 1959.

EXERCISES
1.

Show

quantity
2.

is

that the probability flux (equation 7-31)


real if

it

equals

its

own complex

always

real.

Hint:

Consider a particle of mass m in the state of lowest energy E1 of the potential


v

'

* < -all, x >


-all <x <a/l

Pi).

,,, v

0,

V , and a are such that E,


eigenfunction y^x) for this particle.
where m,
of

is

conjugate.

all

= 0.5 V Make a sketch of the form of the


Now assume that V is increased by a factor
.

What

effect does this have on the shape of y,(a:) in the regions where
,? What effect does this have on the shape of y x {x) in the region where
V(x)
,? What effect does this have on ,? These questions are to be
answered qualitatively by a direct consideration of the behavior of the time
3.

V(x)

>
<

independent Schroedinger equations for the potentials concerned.


3.

4,

Take a potential of the form indicated in exercise 2, except


and use the formulas for [v>(x)\
- Mx)~\Xj and
Xj+l
+

that

^mVtfP/U?

[dy(x)/dx]x

developed in section 5, to make a numerical integration of the time


independent Schroedinger equation and find the first eigenfunction y^x) and
\dy>(x)ldx\Xj ,

first eigenvalue E
v Compare these with the results of the analytical treatment
of this potential presented in section 4, Chapter 8. Hint: Start the integration
at x = 0, where it is apparent from symmetry that dyx)/dx = 0, and take

the

(xi+1
4.

Xj )

0.1a/2.

Use numerical

integration to find the


oo,

V(x)

<

a/1,

= 0,

-a/1

a/4

<x <
<x <

first

>

eigenvalue

Ex for the potential

a/2

-a/4,

a/4

<

<

a/2

a/4

where
v

Tfiffl/Sma 2

Employ the conditions y>j(x) = 0, x < -a/2, x > a/2, and yi^x) need not be
continuous at x = a/2, which are justified in section 5, Chapter 8, for a
potential of this type. One of the exercises of that chapter consists of an
analytical treatment of this potential.

Ch.

211

EXERCISES

7]

Derive the classical wave equation for a string of variable density per unit
Separate it into two ordinary differential equations, and compare
the equation in x with the time independent Schroedinger equation for the
5.

length p(x).

general potential V(x).


6. Verify

equation (7-68) by evaluating some of the integrals.

solve the problem of the motion of a string fixed at both ends,


which is plucked at a point at distance b from its center. Check the solution
by seeing that when b -* it reduces to the solution of the problem treated in
7. Set

up and

section 7.
8.

In calculating the expectation value of the product of position times

momentum, an ambiguity

arises

because

it is

not apparent which of the two

expressions

should be used.

Show

that neither of these

the obvious requirement that xp be real.


/*00

xp

**

{- m Tx)

is

acceptable because they violate

Then show

that the expression

+ {~ ih
Tx} Vdx

J 00
is

acceptable because

integrate

by

parts.

it

does satisfy this requirement.

Hint:

Feel free to

CHAPTER

Solutions
of Schroedinger's

Equation

I.

The Free

Particle

In this chapter we shall demonstrate some of the features of the quantum


mechanical theory and, in particular, some of the interesting phenomena
predicted by the theory,
specific

by solving Schroedinger's equation for several


t) and finding the eigenfunctions

forms of the potential energy V(x,

and eigenvalues.
Let us begin by considering the simplest case

V(x,

t)

For

constant.

a particle under the influence of such a potential the force F(x, i) =


dV(x, t)/dx will vanish; it is a free particle. Since this is true regardless
of the value of the constant, we do not lose generality by defining the
potential energy such that V(x, t) = 0. We know that in classical mechanics
a free particle may be either at rest or moving with constant momentum
p. In either case its total energy E is a constant.
To find the behavior predicted by quantum mechanics for a free particle,
we solve the Schroedinger equation for V(x, t) = 0. Since this potential
energy is not a function of t, the problem reduces to the problem of
solving the

time independent Schroedinger equation.

that equation reads

_*<*V*)_
2m dx2
The

.,

(8-1)

f( *)

solutions to the equation are the eigenfunctions

functions are

For V(x)

y>(x).

The wave

._,,

W(x,t)=e-* Et" w(x)


l

212

(8-2)

Sec.

THE FREE PARTICLE

I]

where

213

E is the total energy of the particle. From the discussion of section


we know

5 of the last chapter,

any value of E

(8-1) exists for

that an acceptable solution to equation

^ 0.

A solution to the differential equation (8-1)


tfx)

where

and

AeiKx

can be written

Be~ iKx

(8-3)

K=

and where
demonstrated by evaluating

are arbitrary constants,

validity of this solution is

V2mE/h. The

^> = i*K Ae iKx + K Be~ iKx = -K y>(z) = - ^- f(z)


2

dx2

and

ti>

substituting into equation (8-1)

-(-^)^)s ^

);

QED

Since there are two arbitrary constants, equation (8-3)

is

the general form

of the solution to the second order ordinary differential equation (8-1).

Thus the wave function

for a free particle

T(z,

may be

e- iEtin f(x)

AeiKx

written
(S-A)

where
y>(x)

Be~ iKx

(8-4')

and

K=

^2mE\h

We consider first the case in which one


to zero.

Then

the

wave function

W(x,

of the constants, say B,

is

equal

is

Ae~ imn e iKx

Aei(Kx

mm

(8-5)

which can also be written


Yix,

t)

= A cos (Kx -

Et/h)

iA sin {Kx

Etjh)

Its
is a traveling wave, oscillating at angular frequency co = Ejh.
nodes move to the right with velocity w = vjk = wjK = EjhK. Let us

This

evaluate

its

probability density P(x,

t)

and

its

probability flux S(x,

t).

These are

P(x,

A*

(8-6)

SOLUTIONS OF SCHROEDINGER'S EQUATION

214

[Ch. 8

and

2m\

=-

S(x,

S(x,

t)

= -

2m

dx

dx

[A*e- iiKx m/ AiKe Kx Etm


~ Et/

AeiiKx

A*( iK)e~ i( Kx
-

~ miK} ~\

A*A2iK = A*
2m
m

(8-7)

According to the de Broglie postulate (equation 7-5),

HK = p
where

v is the velocity

of the

= mv
Thus the probability

particle.

S(x,

vA*A

vP(x,

flux is

(8-8)

t)

which is the same as the familiar classical relation between flux, velocity,
and density. For the wave function (8-5), both the probability density
and the probability flux are constant in time and independent of position.
Furthermore, the wave function involves only the single

= HK and the

function

is

momentum

Consequently we realize that this wave


associated with a particle which is traveling along the x axis

with perfectly

single energy E.

known and

momentum p, but with completely


the physical situation of a particle of

constant

unknown x

coordinate.

momentum

that

an

long beam of particles. For the wave function obtained

infinitely

the constant

is

is set

This

known

to be exactly p,

equal to zero, which

Y(x,

we

is

t)

and moving somewhere

in

when

is

Be- iiKx+Et/n)

(8-9)

find

P(x,

= B*B

(8-10)

-
B*B
m

(8-11)

t)

and
htr

S(x,

The

t)=

sign of the probability flux S(x,

motion of the

particle.

vector quantity.)

moving

in a

beam

The

indicates, of course, the direction of

in the direction
is

is

wave function (8-5) is


of increasing x, and the particle associmoving in a beam in the direction of

particle associated with

ated with wave function (8-9)

decreasing x.

t)

(In three dimensions the probability flux

Sec.

THE FREE PARTICLE

I]

215

Let us evaluate the integral of the probability density for the wave
function (8-5). This

is

foo

/"oo

foo

V* x'dx=\

A* A dx

J 00

A* A

00

We see that it diverges unless A =


the wave function of an

unbound

0.

This

is

particle,

dx

00

the difficulty of normalizing

which we have already come

across in the previous chapter. But the difficulty

is

not really fundamental.

This wave function represents the highly idealized situation of a particle


moving in a beam of infinite length and whose x coordinate is, consequently,
completely unknown.
instead

case;

it is

But

known

in

any

real physical situation this

that the particle

would

is

not the

certainly not be

found

some region of the* axis of finite length. In an experiment the beam


A realistic wave function describing
is actually always of finite length.
the actual situation would be of essentially constant amplitude over the
length of the beam but would vanish at very large and very small values of
a;.
It would be in the form of a group of length A* which is long, but
Furthermore, we know from the uncertainty principle that the
finite.
realistic wave function must involve not only the single momentum
outside

p = hK

but also a distribution of momenta spread over the range


h/ Ax which is centered about the momentum p = hK.
h A.K
The wave function ^8 5) can be thought of as representing the behavior of
the realistic wave function in the limit Aa; -> oo. There is never any diffiA/?

wave function as long as Ax is finite. In the limit


the
sense that the wave function can be normalized
does
arise
in
a
only if the multiplicative constant A vanishes. Even so, the limiting form of
the wave function can be used to calculate quantities which do not depend

culty in normalizing the


difficulty

on the value of the multiplicative constant. In subsequent sections we shall


perform a number of calculations using wave functions of the limiting
form but they will always be used in such a way that a physical significance
is attached only to the ratio of the multiplicative constant of one wave
function to the multiplicative constant of another wave function. This
;

procedure bypasses the normalization


In certain situations

it is

difficulty.

necessary to use a wave function in the form

of a group of finite length. Such a wave function can be constructed by


adding together a number of wave functions of the form (8-5) if the group
is moving in the direction of increasing x, or of the form (8-9) otherwise,
and E. That is, we take
each with a different value of

W(x,

t)

- f A n Wn (x,
n=l

t)

where

Yn(x,t) = e- iE

"t

" eiK"x
t

(8-12)

SOLUTIONS OF SCHROEDINGER'S EQUATION

216

[Ch. 8

and where

Kn =

Tn

Since each of the

(a;, t)

y/2mEJh

a solution to the Schroedinger equation,

is

equation (7-44) shows that (8-12) will also be a solution. The values of the
constants A n can be calculated as in equation (7-57) in terms of the form

of the wave function at some particular instant of time, say / = 0. As an


example, consider a free particle moving in the potential V{x) 0, in
the direction of increasing x, where at / =
the particle was known to have
an x coordinate within the range Ax centered about the expectation value

x and momentum
Furthermore,

let

in the range

Ap centered about the expectation value p.


Ax and Ap be equal to
the minimum

the product of

ft,

allowed by the uncertainty principle. Then the normalized wave function


for this particle at any time

YOr,

0=2=
m

can be shown to bef

!
J

iE " tlh

Wn (x')V(x',0)dx'

(8-13)

Vn (x)

CO

where
(z'

) 2

ipx'-i

[
and
Vn (x)

iK *

[For a particle moving with the same

initial

potential V{x), the expression for Y(x,

;) is essentially

the eigenfunctions

equation (8-13)

for that potential

y> n (x)

strictly for the

more general
same except that

conditions in a
the

must be

used.]

Fortunately

sake of interest.

We
it is

present

almost

always possible to formulate quantum mechanical problems in such a

way

wave function for a free particle moving in the potential V(x) =


can be represented by the wave function (8-5) or (8-9), or by a simple
that the

combination of the two, instead of something as complicated as equation


(8-13). [The same is true for a particle moving in a more general potential
V(x).] Only such problems will be considered in this book.
Consider the two wave functions, (8-5)

W(x,

and

Ae~ iEtin e iKx

Be- iEtin e~ iKx

(8-9)

W{x,

t)

which are solutions to the Schroedinger equation for a free particle.


Choose the ratio of the arbitrary constants such that A = B, and add
the two wave functions. The result is a new wave function,
W(x,
To be

t)

Ae- iEt/n (eiKx

e~

iKx
)

must be written as

integrals since
n is a
continuous variable for the case of a free particle. Equation (8-13) has been written
here as a sum to bring out its analogy with equation (7-57).
t

correct, (8-13),

and also

(8-12),

Sec.

THE FREE PARTICLE

I]

217

which can be written

Y(x,

Ae~ iEt,n [cos

Kx +

Y(x,
Y(x,

t)

Kx -

sin

(-Kx)

cos

sin

(-Kx)]

= lAie-'^sin Kx
= A'e- iEt/K sm Kx

(8-14)

an arbitrary constant which is equal to 2i times the arbitrary


Next choose the ratio of the constants such that A = B,
and add the two wave functions. The result is a second new wave

where A'

is

constant A.
function,

Be~ im \eiKx

Y(x,

e~

iKx
)

which can be written


*(x,

t)

Be- iml \co%

Kx +

Y(x,t)

TO,

sin

Kx +

cos

(-Kx)

sin

(-Kx)]

= 2Be- iEt/n cosKx


= B'e- iEtlH cos Kx

(8-15)

where B' is an arbitrary constant equal to 2 times B, and which is independent of A'. Since the Schroedinger equation is a linear differential
equation, the wave functions (8-14) and (8-15) are also solutions, and so
is their sum,
T(x,

- iEtl

\A'

sin

Kx +

B' cos

Kx)

which can be written


Y(x,

t)

e-

iEt/K

(8-16)

y>(x)

where
y>(x)

= A' sin Kx +

B' cos

Kx

(8-16')

and

K = VlmE/h
We recognize from the form of equation (8-16) that (8-16') must also be a
solution to the time independent Schroedinger equation (8-1) for a free
particle.

Since

it

contains two arbitrary constants,

the solutionjust as general as (8-4').

The

it is

the general

form of

solution (8-16') could have

been written down directly as was (8^4'). However, the procedure we have
followed has the advantage of emphasizing the relation between the two
general forms of the solution.
Furthermore, it helps us to interpret the meaning of the wave function
First of all, we note that the wave function was obtained by
adding and subtracting two traveling waves of equal amplitudes and must,
therefore, be a standing wave. This, of course, can be seen directly from
the form of (8-16), since the nodes of T(a;, t) are fixed in space at values

(8-16).

SOLUTIONS OF SCHROEDINGER'S EQUATION

218

[Ch. 8

of x for which A' sin Kx + B' cos Kx = 0. Next we recall that each of the
two traveling waves have been associated with a particle of precisely known
momentum moving in an infinitely long beam, with the momentum being
equal in magnitude but oppositely directed in the two cases. Consequently
we must associate the standing wave (8-16) with a particle which is moving
in an infinitely long beam in such a way that the magnitude of its momentum
p is precisely known but the direction of the momentum is not known.
That is, the wave function is to be associated with a particle which could
be moving in the direction of either increasing or decreasing x, with
momentum of magnitude precisely equal to p, and with a completely
unknown * coordinate.
Let us check our interpretation of equation (8-16) by evaluating the
probability density P(x, i) and the probability flux S(x, i). For simplicity,
set one of the constants, say B', equal to zero. Then

Y(x,

i)

A'e- iEt ' n sin

Kx

(8-17)

and
P(x,

= Y*Y = A'*eiEtlK sin KxA'e- imi1i sin Kx


= A'* A' sin 2 Kx
P(x,
t)

(8-18)

and also
S(*,

=-^F*^_T^)
2m
dx
dx
\

S(x,

t)

= -

1m

iEtin

[A'*e

sm Kx A'e-

A'e- imin
S(x,

The

sin

mtlK

Kcos Kx

KxA*eimin K cos Kx~]

(8-19)

agreement with
According to that interpretation, the probability per unit time that the particle associated with the wave function
will cross the point x in the direction of increasing x will be equal to the
probability per unit time that it will cross in the direction of decreasing x.
These will balance each other out, and the total probability flux S(x, i)
will be equal to zero. Next consider the probability density P(x, t), which
is plotted in figure (8-1). Also plotted in the same figure is the constant
probability density which would be expected classically for a particle in
the situation which we have described as being associated with the wave
function (8-17). The fact that the quantum mechanical P(x, t) extends to
infinitely large and small values of x agrees with our interpretation of the
wave function and with the classical behavior of the particle, but the
presence of points along the x axis at which the particle would never be
fact that the probability flux vanishes is certainly in

our interpretation of ^(x,

t).

Sec.

THE FREE PARTICLE

I]

found

in contrast to

is

219

what would be expected

This

classically.

is

manifestation of the wave-like aspects of a particle in the situation


described by this

wave

because the distance d

Such behavior

function.

length of the particle.

is

not observed classically

= hjlp is only one-half the de Broglie wave-

tt/K

In the classical limit this distance becomes ext) pattern is so compressed in the x direction

tremely small, and the P(x,

that only the average (constant) behavior of P(x,

t)

can be resolved

experimentally.

P(x,t)

oc sin jRj:

P(x,t)

Figure 8-1.

The

probability density for a free particle standing

For the standing wave function (8-16) there


concerning normalization as there

is

will

wave

constant

function.

be the same problem

for the traveling

wave

For

function.

a particularly simple way to remove the difficulty. The normalization problem arises because of the unrealistic way
the wave function extends to infinitely large and small values of the cothe standing wave there

ordinate x.

is

Consequently the problem can be removed by completely


but finite, region

suppressing the wave function outside the large,

L/2

We

<; x <: L/2.

function which
that region.

is

That

do

by constructing a more

this

realistic

wave

the function (8-16) inside the region, and zero outside


is,

we

take

(*,

_
=

-iEtlK
a
ip(x)
e

(8-20)

where

rw =
, ,

A' sin

Kx

0,

w(x)

It is clear

(8-20).

<

that there will be

B' cos Kx,

-L/2,

no

difficulty in

Furthermore, the eigenfunctions

-L/2

< x < L/2

> L/2

,D

^,

(8-20

normalizing the wave function

now

vanish as x

-*

oo,

and

immediately apparent from equation (7-53) that there will be no


difficulty in proving that such eigenfunctions are orthogonal. We shall
see in section 5 that the procedure indicated in equation (8-20') is
mathematically equivalent to saying that the potential energy V(x)
becomes infinitely large for x
L/2 and x
L/2. This is as if the
it is

<

SOLUTIONS OF SCHROEDINGER'S EQUATION

220

region of space

Z./2 < x < L/2

[Ch. 8

were enclosed in a (one dimensional)

box with completely impenetrable walls. The procedure is known as


box normalization. As long as the walls of the box are very far
away from the region of interest, their presence will have no physical
effect of importance, but they will remove the mathematical difficulties
associated with normality and orthogonality for a free particle The values
of the constants A' and B' required to normalize the wave function will
depend on the length of the box, L; however, these constants will not
enter into the results of a calculation of any quantity of physical interest.
Box normalization will also change the eigenvalues E from continuous to
discrete. But we shall see that the separation between adjacent eigenvalues
2
is proportional to Lr
With a very large value of L the eigenfunctions
will, for all practical purposes, remain continuous.
Normalization of the traveling wave function (8-4) in a box with
impenetrable walls is not possible because, as will be shown in section 5,
the wave function must always vanish at the walls of the box. This
boundary condition can be satisfied by a standing wave with fixed nodes,
but it cannot be satisfied by a traveling wave with moving nodes. However,
we shall see in Chapter 15 that a form of box normalization can be used
for the traveling wave function, if we assume that the walls of the box are
penetrable so that the wave function need not vanish at the walls. In
this form of box normalization we impose the so-called periodic boundary
conditions by demanding that at any instant the wave function and its
spatial derivative have the same values at both walls. It is immediately
apparent from equation (7-53) that with these boundary conditions the
eigenfunctions are orthogonal. These boundary conditions also lead to the
convenient result that both forms of box normalization produce the same
discreteness of the eigenvalues. With normalization in a box of penetrable
walls, we are interested only in what happens in the box, and we restrict
the domain of the variable x to the region between its walls. The wave
.

The value of the normalization constant


depends on the length of the box, but this always drops out of all calculations of physically measurable quantities.

function can then be normalized.

The

problem of normality and orthogonality


wave functions and eigenfunctions, both traveling
wave and standing wave, may be summarized briefly by saying that it is
usually possible simply to ignore the problem. The mathematical consistency of the theory is ensured by the fact that a wave function in the
form of a group may be used, or that one of the forms of box normalizasituation relative to the

for the free particle

tion

may be employed.

to do either, and the


is."

In practice,

it is

wave functions

not normally necessary actually

(8-4)

and (8-16) can be used "as

STEP POTENTIALS

Sec. 2]

2.

221

Step Potentials

we shall calculate the quantum mechanical


whose potential energy can be represented by a

In the next few sections

motion of a

particle

function V(x) which has a different constant value in each of several


adjacent ranges of the x axis, and which changes discontinuously in going

V(x) =

V(x)

=
*
Figure 8-2.

step potential.

from one range to the next. Of course, the situations represented by


potentials which are discontinuous functions of x do not really exist in
nature. However, such potentials are often used in quantum mechanics
to approximate real situations because, being constant in each range, they
are easy to treat mathematically. With such potentials we shall be able to
illustrate

a number of characteristic quantum mechanical phenomena.

Figure 8-3.

Let us consider
figure (8-2).

physical system approximating a step potential.

the simplest example, a step potential, as shown in


choose the origin of the x axis to be at the step. Then

first

We

V(x) can be written


V(x)

= V
0,

x>0
x<0

(8-21)

V is a constant We may think of V(x) as a limiting representation


of the potential energy function for a charged particle moving along the

where

axis of a system of

two electrodes which are held

illustrated in figure (8-3).

at different voltages, as

SOLUTIONS OF SCHROEDINGER'S EQUATION

222

mass m and total energy E is in the region


moving toward the point at which V(z) changes. According

Assume

that a particle of

<

is

and

[Ch. 8

move

to classical mechanics the particle will

freely in that region until it

an impulsive force F(x) =


dV(x)/dx acting in the direction of decreasing x. The subsequent
motion of the particle depends, classically, on the relation between E
and V
this is also true in quantum mechanics. We take first the case:

reaches x

where

0,

it

subjected to

is

E<V
This

is

Since the total energy

illustrated in figure (8-4).

is
-

EV(x)

The

V(x)

=V

-E

x
Figure 8-4.

a constant,

case in which classical mechanics predicts that the particle

is

reflected

at the discontinuity of the potential.

classical

mechanics says that the particle cannot enter the region x

as in that region

^<^

>

0,

f\lm <
change the momentum

The impulsive force will


way that it will exactly reverse

its

of the particle in such a


motion, traveling off in the direction of

momentum equal in magnitude (so that the total energy


remains constant) but opposite in direction to its initial momentum.
To determine the motion of the particle quantum mechanically, we must

decreasing x with

find the

wave function which

is

our discussion of section

5,

E< V
We know from

a solution, for the total energy

to the Schroedinger equation for the potential (8-21).

Chapter

7,

that acceptable solutions to the

E^

Schroedinger equation for the potential V(x) exist for any value of
0.
Since we are dealing with a time independent potential^ the actual problem
is

to solve the time independent Schroedinger equation

eigenfunction.

With

and

find the

x axis breaks up into two regions.


a solution to the simple time inde-

this potential the

In one region the eigenfunction


pendent Schroedinger equation

is

_^_d>^) =
2m dx2

In the other region

it

is

YK

a solution to the simple time independent

Schroedinger equation
h*

dV*) + V y>(x) =

2m dx2

Ey>(x),

x>0

(8-23)

223

STEP POTENTIALS

Sec. 2]

Then an eigenfunction valid


x is constructed by joining the two solutions
in such a way as to satisfy the conditions (7-38), which
and dy>(x)ldx be everywhere finite and continuous.

The two equations

are solved separately.

for the entire range of

together at x
require that

y>(x)

Consider equation (8-22). This is precisely the time independent


Schroedinger equation (8-1) for a free particle. The general solution to
equation can be written either in the traveling wave form (8-4')
or in the standing wave form (8-16'). Since the two are mathematically
equivalent, either may be used. However, by using the form most approthis

priate to the

problem

at hand,

it is

often possible to simplify the calculation

When

or the interpretation of the results.

unbound

particle

it is

usually

most

convenient to use the traveling

bound, the standing wave form


convenient. For the problem at hand we take

form. If the particle

is

f(x)

Ae iKlX

an
wave
usually the most

discussing the motion of

Be~ iKlX

is

x<0

(8-24)

where

K
x

Next consider equation

(8-23).

VlmEjh

The

general solution to this equation

= Ce K x + De~ K x

x>0

yix)

is

(8-25)

where

K = V2m(V - E)\h
2

To

E< V

and

verify this, calculate

^^ = CK\eK

*x

dx2

and

+ D(-K2fe- K x
*

substitute into equation (8-23). This gives

_Jt2m
2m
The

^s

fr

arbitrary constants A, B, C,

eigenfunction satisfies

and

the conditions (7-38).

yKx)

QED

must be so chosen that the


Consider

first

the behavior

+ co.

In this region of the x axis


the solution to the time independent Schroedinger equation is given by
equation (8-25). It will, in general, diverge as x -* + oo because of the
of the eigenfunction in the limit

first

term. In order to prevent

this,

->-

we must

set the coefficient

of the

first

term equal to zero. Thus

C=

(8-26)

SOLUTIONS OF SCHROEDINGER'S EQUATION

224

Now

consider the eigenfunction at the point x

solutions (8-24)

are continuous.

Continuity of

D(e- K >*) x=0


is satisfied.

At

0.

and (8-25) must join in such a way that


ip(x) is

obtained

A(eiK*) x=0

if

[Ch. 8

this point the

y>(x)

and

dy>(x)jdx

the relation

B{ e iK *) x=0

This yields

D=A+B

(8-27)

Continuity of the derivative of the solutions


dy>(x)

= -K2 De- K x

x>0

dx

and
dip(x)

iK^Ae*^*

dx
is

obtained

if

-K
is satisfied.

iKx Be- iKlX

<

the relation

2 D(e~

K*% =0 =

iKxA(e iK n*=o

~ iK^***'),x =

This yields

i^ D =
Adding equations (8-27) and (8-28)

A-B

(8-28)

gives

(8-29)

Subtracting gives
JB

_ ^^1
iEi)
R(i
=
ll
2\

Thus the eigenfunction

(1 +
w(x)

De~ KiX

for this potential,

iK2 IK^e iK * x

(8-30)

Kj

+ - (1 -

and for the energy E,

is

<

iKJKJe-***',

x^O

(8-31)

The arbitrary constant D has not been determined because we have not
normalized the eigenfunction. This could be done by the techniques
described in section 1, but it is not necessary.

The wave function

,
f)
^ {X A
'

is

Ae -iBtih e KlX +
~ De-We-x*,
i

Be -iEtm e -iK ^
1

^
x^O

*~ 5Z)

STEP POTENTIALS

Sec. 2]

In the region x

<

it

increasing x, plus a

the region x
function,

<

>

it is

225

consists of a

wave

traveling in the direction of

traveling in the direction of decreasing x;

wave

in

In interpreting this wave

consists of a standing wave.

it

useful to calculate the probability flux S(x,

t)

in the region

equations (8-7) and (8-11) and the wave


functions from which these expressions were derived, it is apparent that

By comparison with

0.

in the present case


S(x,

where

hKjm,

t)

= vA*A-

vB*B,

<

(8-33)

the magnitude of the velocity of the particle.

The

first

term in the wave function, is a probability flux flowing in the direction of increasing x. The second term,
which comes from the second term in the wave function, is a probability
term, which comes from the

first

flux flowing in the opposite direction.

Consequently,

we

associate the

first

term in Y(a;, i) or S(x, t), in the region x < 0, with the incidence of the
particle upon the point at which the potential energy changes, and the
second term with the reflection of the particle from the change in potential.
Let us calculate the ratio of the intensity of the reflected probability flux
to the intensity of the incident probability flux. We obtain

vB*B

vA*A

(1
(1

+
-

B*B
A* A

(1

(1

- jKJKt)*(1 - iKjKJ
+ iK,IKJ*(\ + iKjKJ

Kx)(l - iKJKJ _
iKJKtfL + iKjKJ

tK/ 2

The quantity B*BjA*A is also the ratio of the intensity of the reflected
traveling wave to the intensity of the incident traveling wave in the region
x < 0. That these ratios are unity means that a particle incident upon the
change in potential, with total energy E < VQ has unity probability of
being reflected. This is in complete agreement with the classical pre,

dictions.

Kxx,

<
x^O

(8-34)

iK
Consider the eigenfunction (8-31). Writing e ^ x = cos xx + / sin
etc., it is easy to show that the eigenfunction can be expressed as

y(x)=

D - sin K,x,
D cos K,x
*
Kj.
De~ K *x
,

The wave function corresponding to this eigenfunction is actually a


standing wave for all x. In this problem the incident and reflected traveling
waves combine to form a standing wave because they are of equal intensity.
Let us plot (8-34), which is a real function of a; if we take D real. The

SOLUTIONS OF SCHROEDINGER'S EQUATION

226
plot

is

shown

[Ch. 8

Here we find a feature which is in sharp


Although in the region x >
the

in figure (8-5).

contrast to the classical predictions.

probability density

P(x,

= Y*(a;, t)W(sc,

t)

t)

D*De~ iK*x

(8-35)

decreases rapidly with increasing x, there is a finite probability of finding


the particle in this region. In classical mechanics it would be absolutely

impossible to find the particle in the region x

>

*
Figure 8-5.

An

because there the total

=o

eigenfunction for a step potential in the case in which classical mechanics

predicts that the particle

is

reflected.

energy

is less than the potential energy.


Penetration of the classically
excluded region is one of the more striking predictions of quantum
mechanics.

We shall discuss later certain experiments which confirm this prediction.


Here we would like to point out that equation (8-35) is not in disagreement
with the experiments of classical mechanics. It is apparent from that
equation that probability of finding the particle with a coordinate x >
is only appreciable in a region starting at x =
and of length Ax of the
order of 1/A"2 Thus
.

Ax

~ 1/K

hlV2m(V

E)

E)

In the classical limit the product of


and
to h2 , that A* is immeasurably small.

(V

is

so large, compared

Furthermore, the uncertainty principle demonstrates that the wave-like


by an entity in penetrating the classically excluded
region are not really in conflict with its particle-like properties. Consider
an experiment capable of showing that the particle was located somewhere
properties manifested

>

>

x
is appreciable only in a region
0. Since P(x, t) for x
of length Ax, the experiment would amount to localizing the particle within

in the region

STEP POTENTIALS

Sec. 2]

that region.

Ap

in the

In doing

momentum

this,

227

the experiment

would lead to an uncertainty

of the particle, where


A/>

~ HIAx ~ V2m(V

- E)

Consequently the energy of the particle would be uncertain by an amount

A =

(Ap?/2m

~V

-E

would no longer be possible to say that the total energy of the


E was definitely less than the potential energy V
Next let us consider the motion of a particle in a step potential when:

and

it

particle

E>V
This

in figure (8-6).

is illustrated

traveling in the region

< 0,

Classically,

a particle of total energy

E
V(x)

V(x)

= 0x

Figure 8-6.

The case

in

which

in the direction of increasing x, will suffer

= V

classical

mechanics predicts that the particle

is

not reflected

at the discontinuity of the potential.

an impulsive retarding force F{x)

dV(x)/dx at the' point x = 0. But


and it will enter the region x > 0,

the impulse will only slow the particle

motion in the direction of increasing x. Its total energy E


its momentum in the region x <
is px where E =
pWlm; its momentum in the region x > is p2 where E = p\J2m + V
Quantum mechanically, the motion of the particle is described by the
wave function Y(a;, i) = e-iEtlHy(x), where the eigenfunction y(x) is a
continuing

its

remains constant;

solution to

-?-E,).
2m dar

x<0

(8-36)

x>o

(8 - 37)

and

-^

=(E - v

xi

and where the eigenfunction satisfies the conditions (7-38) at the point
* = 0. The first equation describes the motion of a free particle of momentum/^. The traveling wave form of its general solution is
ip{x)

Ae iKlX

Be~ iKlX

where

K = JlmE/H = pJH
t

x<0

(8-38)

SOLUTIONS OF SCHROEDINGER'S EQUATION

228

[Ch. 8

The second equation describes the motion of a free particle of momentum


p2 The traveling wave form of its general solution is
.

f(x)

Ce iK **

De~ iK

x>0

*x
.

(8-39)

where

K 2 = 72m( - V )/h =

and

p2 /h

E> V

The wave function

specified by equations (8-38) and (8-39) consists of


waves of de Broglie wavelength Ax = hjp1 = 2tt\Kx in the region
x < 0, and of longer de Broglie wavelength A2 = hjp2 = 2tt/K2 in the
region x > 0.

traveling

Classically,

the particle has unity probability of passing the point

and entering the region x

cally.

>

0. This is not true quantum mechaniBecause of the wave-like properties of the particle, there is a

certain probability that the particle will be reflected at the point x = 0,


where there is a discontinuous change in the de Broglie wavelength.
Thus we need to take both terms of the general solution (8-38) to describe
the incident and reflected traveling waves in the region x < 0. However,
we do not need to take the second term of the general solution (8-39).
This term describes a wave traveling in the direction of decreasing x in
the region x > 0. Since the particle is incident in the direction of increasing x, such a wave could only arise from a reflection at some point with
a large positive x coordinate. As there is nothing out there to cause a
reflection, we know that there is only a transmitted traveling wave in the
region x > 0, and so we set the arbitrary constant

The

arbitrary constants A, B,

dfix)\dx continuous at x

Ae

iKlX

)*=o

D=
and C must
The

0.

Bie-

be chosen to make fix) and

requirement

first

iK

(8-40)

n*=o

Cie

is satisfied if

iK **)

x=0

or

A+B=C
The second requirement

ilM(e*a=o

(8-41)

is satisfied if

iKx Bie- iK

x =o

iK2 CieiK**) x=0

or

K iA - B) = K C
x

From

we

equations (8-41) and (8-42)

b =

(8-42)

find

-k

and

(8-43)

C=

2Kl

K +K
1

A
2

STEP POTENTIALS

Sec. 2]

Thus the eigenfunction

is

K K
+ A i~ * <T**i",
K + K2

AeiK i*
w(x)

arbitrary constant

<

x^O

iK",
e

(8-44)

IF

A
The

229

can be chosen to

satisfy the

normalization

condition, but

we do not bother to do

satisfying the

two continuity conditions and the normalization condition

could not be found if we had


wave equal to zero, since then

down

wave

the

initially set the coefficient

we would have

an eigenfunction

B of the

reflected

only two arbitrary constants

We shall not plot the complex function (8-44).

to satisfy three conditions.

Writing

It is clear that

this.

function,

and evaluating the probability

flux at

< 0, we find

some point x

S(x,

t)

= v A*A - v B*B
x

where

= hKjm = px/m

vx

The first term is the incident flux, and the second term is the reflected flux.
The transmitted probability flux can be evaluated by calculating S(x, t)
at some point x > 0. This gives
S(x,

i)

= v 2 C*C

where
v2

The

ratio

HK^jm

= pjm

of the intensity of the reflected flux to the intensity of the


the probability that the particle will be reflected back into

incident flux

is

the region x

< 0.

This

is

R= viE B = (K -K f
2

vx

The

ratio of the intensity of the transmitted flux to the intensity of the

incident flux

region x

>

is the. probability

0.

This

v 2 C*C
Vl

show

_ K2

A*A

T
easy to

that the particle will be transmitted into the

is

T=

It is

(Kx+KJ2

A*A

{IK,)2

(Kx

+ K2 f

4KX K2
(K t

+Kf

(8-46)

that

R+

T=

(8^7)

SOLUTIONS OF SCHROEDINGER'S EQUATION

230

The

we

upon

probability flux incident

and a

into a transmitted flux

[Ch. 8

the potential discontinuity

reflected flux, but

is split

from equation (8-47)

see that the total probability flux is conserved.

that either the particle will be reflected or

will

it

This statement says


be transmitted, but it will

not vanish. Of course, the particle is never split at the discontinuity.


In any particular trial the particle will go one way or the other. For a large
number of trials, the average probability of going in the direction of
decreasing x is R, and the average probability of going in the direction of
increasing x is T. Note that R and T are both unchanged if x and
2

are exchanged.

A moment's

consideration will

show

that this

means the

same values of R and T would be obtained if the particle were incident


upon the potential discontinuity in the direction of decreasing x from the
region x > 0. The wave function describing the motion of the particle,
and consequently the probability flux, is partially reflected simply because
there is a discontinuity in V(x), and not because V(x) becomes larger in the
direction in which the particle is going.
It is interesting to investigate the value of RforE~ V and also for
,

E > V In the rst case, K2 = V2m(E - V )/h - 0,


V2mElH-*V2mV IH. Thus
\Kt
and the probability
second case,
the particle

while

fi

+K

reflected

becomes

Thus

XkJ

that the particle is reflected approaches unity.

K2 -> VlmEjh = Kx
is

K =

In the

R ->

negligible.

and the probability that


However, the case E > V
at which we know there will

not necessarily the case of the classical limit


be no reflection. This becomes apparent if we evaluate equation (8-45)
in terms of E and V , giving
is

R-

(l-J^m
Vi

+ vi -

(M8)

VJEl

Planck's constant does not enter into the equation; the value of R depends

only on the ratio of

to E.

Keeping

this ratio constant,

(8-48) predicts that there will be reflection even

when

and

that the predictions of classical mechanics should be valid.

paradoxical until

we

equation

E are so large
This seems

realize that, in the limit of large energies,

the

assumption that the change in V(x) is perfectly sharp can no longer be


even an approximation to a real physical situation. The point is that the
reflection is the result of an abrupt change in the de Broglie wavelength;
"abrupt" means that the wavelength changes by an appreciable amount in a
distance of 1 wavelength. If the change in wavelength in 1 wavelength
becomes very small, because the change in the potential in that distance is

BARRIER POTENTIALS

Sec. 3]

231

very small, the reflection becomes negligible.

This can be demonstrated

by solving the Schroedinger equation for a potential of the form shown in


figure (8-7). These statements, which follow from the characteristic
properties of wave motion, should be familiar to the reader from his study
of the behavior of water waves or light waves. In the classical limit the

V(x)

Figure 8-7.

A smoothed

step potential.

de Broglie wavelength is so very short that any physically realistic potential


V(x) changes only by a negligible amount in 1 wavelength, and no reflection

However, for particles in atomic or nuclear systems, the de Broglie


wavelength can be long relative to the distance in which the potential
changes significantly, and reflection phenomena can become very
results.

important.

3.

Barrier Potentials

We

consider next a barrier potential, illustrated in figure (8-8).

potential

This

is

V(x)

=F

'

0,

<x<a
< 0, x >

(8-49)

vo~
(

x=Q

x=a

-x

Figure 8-8.

Classically,

incident

barrier potential.

a particle of total energy

upon the

transmitted if

in the region

< 0,

which

is

barrier in the direction of increasing x, will have unity

E> V

<

E V , and unity probability of being


Neither of these conclusions obtains in quantum

probability of being reflected if

mechanics.

SOLUTIONS OF SCHROEDINGER'S EQUATION

232

For

this potential acceptable solutions to the time independent Schroe-

dinger equation exist for all

E^

Also the equation breaks into three


x < 0,
< x < a, and x > a.

0.

separate equations for the three regions

and
energy E. The
In the

[Ch. 8

first

third regions the equation is that of a free particle of total

traveling

w(x)
f(x)

wave

=
=

solutions are

AeiK x
Ce

iK x

Be~ iK x

'

De~

4-

'

iK

<
>a

'

(8-50)

where

K\

^jlmElh

In the second region the form of the equation and of the solutions depends

E< F

on whether

or

E> V

the previous section. In the


y>(x)

Both of these cases have been discussed in

first

Fe~ K x

case,

Ge K x

0<x<a

(8-51)

where

x =

V 2m (K> - E)l h

and

E<

In the second case,

v(x)

Fe iK "i x

Ge~ iK i" x

0<x<a

(8-52)

where

K m = ^2m{E - V )lh

and

E> V

In evaluating the motion of a particle incident in the direction of increasing


x,

we may

set

D=
we know

that there can only be a transmitted wave in the region


However, in the present situation we cannot set G in equation
< x < a. Nor can we set
(8-51) since x remains finite in the region
G = in equation (8-52) since there will be a reflection from the discontinuity at x = a.
since

>

a.

Consider

first

the case

<V
In matching

ip(x)

and df(x)/dx

at the points

equations in the constants A, B, C, F, and


it

as an exercise for the reader to set

evaluate B, C, F, and

up

G will

and x

be obtained.

these equations

a,

four

We

leave

and use them

to

G in terms of A. The value of A would be determined

were applied. The form of the eigenThis figure is meant to be


< x < a,
only an indication, since the exact form of f(x) in the region
as well as the ratio of its amplitude in the region x > a to its amplitude

by the normalization condition,


function obtained

is

if it

indicated in figure (8-9).

BARRIER POTENTIALS

Sec. 3]

<

in the region x
y>(x)

>

0,

depends on the values of E, V a, and m. Furthermore,


complex function. [For example, ip(x)
CelKlX for
,

actually a

is

In the figure

a.]

The most

233

we have

plotted the real part of

ip(x).

interesting result of the calculation is the ratio

intensity of the probability flux transmitted into the region

the intensity of the incident probability flux.

T=

A*A

which

sinh

(4/F )(l

The value of

K^a

T
x

of the

>

a to

T is

E<V

E/V )

is

sinh

T=

(i

n2

Figure &-9.

-l

2mV a 2

(4/F )(l

The

- to)

/F )

E<V

(8-53)

real part of a barrier penetration eigenfunction.

If the argument of the hyperbolic sine is large compared to


a very small value which is given approximately by

T~ 16
V

(1

EIV )e- 2K " a

1,

Twill have

(8-54)

These equations make the remarkable (from the point of view of classical
mechanics) prediction that a particle of mass m and total energy E, which
is

incident

on a potential

on the other
course,

T is

the quantity
is

side.

This

V > E and finite thickness a,


T of penetrating the barrier and appearing

barrier of height

actually has a certain probablity

phenomenon

is

called barrier penetration.

Of

vanishingly small in the classical limit because in that limit

2mV a2/H2

which

is

a measure of the "opacity" of the barrier,

extremely large.

Next consider the case

>v
In this case the wave function

is a traveling wave in all three regions,


but of longer wavelength in the region
x
a. Evaluation of the

< <

SOLUTIONS OF SCHROEDINGER'S EQUATION

234

constants B, C, F, and

and x

G by the application of the continuity conditions at

a, leads to the

value
2

T=

vx

which

[Ch. 8

A*A

sin JCjija

(4/F )(/F

E>V
1)

is

T=

The forms of

2wF a2

2
//2

E>V

(4/F )(/F

(8-55)

1)

and (8-55) depend on the value of

the functions (8-53)

Figure (8-10) shows a plot of

as a function of

/F

for

1.0

0.9

0.8

2mV.a %
1

0.7
0.6

~ q9

ft*

0.5
0.4

0.3
0.2
0.1

/III
1

Figure 8-10.

The transmission

10

coefficient for a barrier potential.

an electron incident upon a barrier of height V = 10 eV and thickness


a = 1.85 X 10- 8 cm; (2mV a/fP = 9). This is typical of the values for
the potentials encountered, for instance, by an electron moving through
the atoms of a crystal. For EjV < 1, T is very small. But, when E/V
Q
is only somewhat smaller than unity, Tis not at all negligible. It is apparent
that barrier penetration can be very important in atomic systems. For
EjV
1, T is in general less than 1, owing to reflection at the discon-

>

tinuities in the potential.

that

T=

whenever

However, from equation (8-55)

Km a =

n, 2n, 3n,

dition that the length a of the barrier region


integral

number of wavelengths A |N

particular barrier, electrons of energy

This
is

is

it can be seen
simply the con-

equal to an integral or half-

2irlKm in that region.

E~ 21 ev,

53 ev,

For

etc., satisfy

this

the

Sec. 3]

BARRIER POTENTIALS

condition

Kw a =

any

reflection.

reflections at

-rr,

and so pass

2tt, etc.,

This effect

into the region x

>

a without

a result of destructive interference between

is

and x

235

a.

It is closely related

to the

Ramsauer

effect

observed in the scattering of electrons by atoms, which will be discussed in

Chapter

15.

Barrier penetration

a manifestation of wave-like behavior in the

is

motion of a particle. The same phenomenon is also observed in other types


of wave motion. We have seen that the time independent differential
equation governing the propagation of waves in a string is essentially the
same as the time independent Schroedinger equation. This is also true of
the time independent differential equation for electromagnetic radiation.

For radiation of frequency


refraction

\i

this

equation

d?y>(x)
z

dx

When we compare

this

propagating through a

^V)
(

y(*)

with equation (7-35),

independent Schroedinger equation


gation of light through a

medium of index of

is

medium

is

in

(8-56)

it is

apparent that the time

identical with that for the propa-

which the index of refraction

is

the following function of position

M *)= 2i;v^ C In

fact, there are optical

phenomena which

F(*)]

~
(8 57)

are exactly analogous to each

of the quantum mechanical phenomena that arise in considering the

motion of an unbound

particle.

For

instance,

an optical phenomenon,

completely analogous to the quantum mechanical

phenomenon of

unity

transmission coefficient for barriers of length equal to an integral or halfintegral

number of wavelengths,

obtain very high

is

light transmission.

used in the "coating" of lenses to


In particular, the situation

E<

V(x)

encountered in quantum mechanical barrier penetration is equivalent to


the optical situation for an imaginary index of refraction. This occurs in
the case of total internal reflection.

Consider a ray of light incident upon a glass to air interface at an angle


This situation, which is called total
internal reflection, is illustrated in figure (8-1 1). A detailed treatment of

greater than the critical angle 6 C


this process in

refraction,

terms of electromagnetic theory shows that the index of


line ABC, is real in the region AB but

measured along the

imaginary in the region BC. Furthermore, it shows that there are electromagnetic vibrations in the region BC of exactly the same form as the
exponentially decreasing standing wave of equation (8-34) for the region
E < V(x). The flux of energy is zero in this electromagnetic standing wave,

SOLUTIONS OF SCHROEDINGER'S EQUATION

236

just as the flux of probability

wave.
first

However,

if

is

zero in the

a second block of glass

[Ch. 8

quantum mechanical standing


is placed near enough to the

block to be in the region in which the intensity of the electromagnetic

vibrations

is

still

appreciable, these vibrations are picked

up and they

propagate through the second block, and the electromagnetic vibrations


in the air gap

now

carry a flux of energy through to the second block.

Glass

Figure 8-1

This situation, which

Illustrating total internal reflection.

1.

is

called frustrated

total

internal reflection,

is

same thing happens in the


quantum mechanical case when the region in which E < V(x) is reduced
from infinite thickness (step potential) to finite thickness (barrier potential).
The transmission of light through an air gap, at an angle of incidence
greater than 6 C was first observed by Newton (ca. 1700). The equation
illustrated in figure (8-12).

Essentially the

Glass

Glass

Figure 8-12.

Illustrating frustrated total internal reflection.

relating the intensity of the transmitted

gap, and the other parameters,

and

it

is

beam

identical in

to the thickness of the air

form with equation

(8-53),

has been verified experimentally.

Barrier penetration of particles was first discussed in connection with


a long-standing problem concerning the emission of alpha particles by
radioactive elements.

As a

typical example, consider a nucleus of the

BARRIER POTENTIALS

Sec. 3]

U238

237

The potential energy V(r) of an alpha particle


from the nucleus had been investigated by means of Rutherford scattering experiments. Using as a probe the 8.8 X 10 6 ev = 8.8
Mevf alpha particles emitted from the radioactive element Po 212 it was
radioactive element

at a distance r

observed that the probability of scattering agreed with the predictions of


Rutherford's formula over the entire angular range. From this it was

concluded

that, for the

= zZe^jr at
V(r') = 8.8 Mev.
V(r)

U238 nucleus,

least in to
It

was

also

scattering of alpha particles

V{r) exactly follows the

a distance

r'

known from

from the

3.0

10~ 12

Coulomb law

cm

at

which

the experiments involving the

nuclei of light

atoms that

V(r)

must

u*
N
3 8.8

UJ

V(r)

ss-E

4.2

-Kinetic energy at large r

Figure 8-13.

the relation between potential and total

An energy diagram showing

energies in alpha particle emission.

eventually depart from a 1/r law


at that time the exact value of
nuclei.

by

U 238

when

R was

<

not

R, the nuclear radius, although

known

for the

heavy radioactive

Furthermore, since alpha particles are (very infrequently) emitted


nuclei, it was assumed that they must exist inside such nuclei

to which they are normally bound by the potential V(r). From these
arguments it was concluded that the form of V(r) in the region r < r'
must be qualitatively as depicted in figure (8-13). (This assumption has
been verified by recent experiments involving the scattering of alpha particles of energies high enough to investigate the potential over the entire
range of r.) It was also known that the kinetic energy of alpha particles
emitted by U 238 was 4.2 Mev. The kinetic energy was, of course, measured
and the kinetic
at a very large distance from the nucleus, at which V(r) =

energy equals the total energy.


also

This value of the constant total energy

this situation is

completely paradoxical.

E is initially in the region r <

An alpha particle

R. This region
space by a potential barrier of height which is
t

is

shown in figure (8-1 3). From the point of view of classical mechanics,

Mev =

10 6 ev

nuclear physics.

1.602

x 10 -6

ergs.

is

of total energy

separated from the rest of

known to be at least twice E.

This unit of energy

is

used extensively in

SOLUTIONS OF SCHROEDINGER'S EQUATION

238

[Ch. 8

Yet it is observed that on occasion the alpha particle penetrates the barrier
and moves off to large values of r.
In 1928 Gamow, Condon, and Gurney treated alpha particle emission as
a quantum mechanical barrier penetration problem. They assumed that
V(r) = zZe*\r for r > R, and V(r) < E for r < R, as shown in figure
(8-13). Equation (8-54) was used to evaluate the transmission factor T
since K^a > 1. In fact, this quantity is so large that the exponential
completely dominates the behavior of T, and it was sufficient to take

This expression was derived for a rectangular barrier of height

width

and

but in the limit that the expression is valid it can be applied to the
barrier V(f) by considering it to consist of a set of adjacent rectangular
barriers of height V(r?) and width Ar{ The reason is that in the limit rea,

flections

from the

discontinuities in the rectangular barriers

important contributions to the transmission factor, and


pletely

do not make
is com-

value

dominated by the exponential decrease in the eigenfunction in the

The transmission factor for

regions within the barriers.


barrier

its

the

z'th

rectangular

is
J*

If reflections

-2V'(2m/S 2 )[F(r < )-S]Ar i

can be ignored so that there

is

no interaction between the

individual barriers, the total transmission factor will be the product of

the individual transmission factors because independent probabilities are

Then

multiplicative.

_ e - V (2m/S )[F(r - El An g - V(2m/B )[F(r -J2] Ar2g -2 V(2m/B )F[(r - E] Ar


_ e -2(\ (2m/S )tF(r )-S] Ar + V(2m/B )[F(r )- E] Ar + V (2m/B )[F(r - E] Ar +
_ e -2^V (2m/ )[F(r,)-]Ar,/

j-

Ar become very
s

J_

s)

Letting the

2)

1)

small, this

becomes

-2j\/ (2m/S 2 )[F(r)- E]dr


J

(8-58)

rigorous treatment of this problem shows that this procedure

justified, in that it gives the correct

answer,

if

T is

very small, as

it is

is

for

the case of alpha particle emission.

The quantity

T gives

the probability that in one trial an alpha particle

will penetrate the barrier.

The number of trials per second can be estimated

as

N~

v/2R

(8-59)

Sec. 4]
if it is

SQUARE WELL POTENTIALS


assumed that an alpha

particle

239

bouncing back and forth with

is

Then

velocity v inside the nucleus of diameter 2R.

second that the nucleus will emit an alpha


designated by the symbol

particle,

the probability per

which

is

conventionally

X, is

~ e -2^V(2W )[(,zeV)-]*
2

2R

(8
v

_6Q)'

In applying this expression to a particular radioactive nucleus, all the


quantities in this expression except v and R are known. Assuming v to
be comparable to the velocity of the alpha particle after emission (i.e.,

mv*

~9

E), then X

is

a function only of the nuclear radius R.

x 10-13 cm, which was

Using

certainly in line with the values obtained

from Rutherford's analysis of the alpha particle scattering experiments


from light elements, Gamow, Condon, and Gurney obtained values of X
which were in good agreement with those experimentally measured,
although they vary over a tremendously large range. For U 238 X =
5 X 10~ 18 sec" 1
For Po212 X = 2 X 106 sec" 1 This variation is due
primarily to the variation, from one radioactive element to the next, of
the energy E of the emitted alpha particles. The height and breadth of the
barrier (Z and R) do not change significantly for nuclei in the limited
range of the periodic table in which the radioactive elements are found.
,

The successful application of the theory of quantum mechanics to the


process of alpha particle emission provides one of the most convincing
verifications of the theory.

Furthermore,

illustration of the wave-particle duality.

this

On

process gives a striking

either side of the barrier the

alpha "particle" could be localized and shown to behave like a particle.


In penetrating the barrier it behaves purely like a wave. Any attempt to
localize it in the barrier would, according to the uncertainty principle,
necessarily introduce such a large uncertainty in the energy that it would

no longer be

possible to say definitely that

it

was not going over the top of

the barrier.

4.

Square Well Potentials

In this section we shall discuss the simplest potential which is capable


of binding a particle to a limited region of space, a square well potential,
indicated in figure (8-14).

V(x)

This potential

VQ
0,

<

is

a/2,

a/2

<x<

>

+aj2
(8-61)

+a/2

SOLUTIONS OF SCHROEDINGER'S EQUATION

240

Consider

>

[Ch. 8

the case

first

(traveling waves)

under the influence of this potential with total


energy E > V would be able to move over the entire range of the x axis.
Upon entering the region aj2<x< +a/2, it would receive an accelerating impulse and travel through that region with momentum higher than
Upon leaving, it would receive a compensating
its initial momentum.
Classically, a particle

decelerating impulse. Its final

We

tum.

shall

momentum would

equal

its initial

momen-

not develop the traveling wave solutions to the quantum

= -

a/2

A square

Figure 8-14.

mechanical problem, which exist for

x = a/2

well potential.

E > V since
E> V

all

they are essentially

the same as those for a barrier potential with

In fact, the trans-

can be obtained directly from equation (8-55) if the


terms are appropriately redefined. This factor is, as before, less than unity
except for energies at which a is just equal to an integral or half-integral
number of de Broglie wavelengths in the region a/2 < x < +a/2.
mission factor

We

return later to discuss the standing

potential with

We

<

E> V

wave

solutions for the square well

consider instead the case:

V (standing waves)

If the particle has total energy

in the region

a/2 < x <

E< V

+a/2. The

then classically
particle

it

could only be

would be bound

to that

region and would bounce back and forth between the ends of the region

momentum

of constant magnitude but alternating direction. Classiany value of the total energy E is possible for a bound state. In
section 5, Chapter 7, we have seen qualitatively that in quantum mechanics

with

cally,

only certain values of the total energy


are possible. In this section we shall
evaluate quantitatively the discrete eigenvalues, and the corresponding
eigenfunctions, for a square well potential.

Such a potential

is

very often used in

quantum mechanics

to represent

SQUARE WELL POTENTIALS

Sec. 4]

a situation in which a particle

moves

the influence offerees which hold

some

potential loses

it

details of the

241

in a restricted region of space under

Although this simplified

in that region.

motion,

it

retains the essential feature

of binding the particle by forces of a certain strength to a region of a


As an example, many aspects of the motion of a nuclear

certain size.

can be explained by assuming that the particle is


about 50
ev and linear
dimensions of about 10 -12 cm.
The description of the classical motion of particle bound by the square
well suggests that it would be most appropriate to look for solutions to the
Schroedinger equation in the form of standing waves. Thus we take, as a
solution to the time independent Schroedinger equation in the region
a/2 < x
+a/2, in which V(x) = 0, the free particle standing wave

particle in a nucleus

bound

in a square well potential with a depth of

<

This

eigenfunction (8-16').
y>(x)

=A

KyC

sin

is

+ B cos K^x,

a/2

<

<

+a/2

(8-62)

where

= VlmElh
> +a/2 the time independent Schroedinger

K,

In the regions x

<

equation

same

is

the

a/2 and x

According to equation (8-25),

as equation (8-23).

the solutions are

y(x)

Ce K " x

De~ K " x

f{x)

Fe K x

Ge~ K " x

<

> + aj2

-a/2

(8-63)

and
,

where

(8-64)
<

X M = yj2m(V To determine

E)/n

the arbitrary constants

the eigenfunctions remain finite for


the limit x ->

oo.

It is

and
first

all x.

apparent that

this

<V

impose the requirement that


Consider equation (8-63) in
requirement demands

D=
Similarly,

it is

(8-65)

necessary to set

F=

(8-66)

remain finite in the limit


+ oo. Next
impose the requirement that the eigenfunctions and their first derivatives
be continuous at x = a/2 and x = +a/2. Four equations are obtained.
x -*

in order that equation (8-64)

They

are

-A sin (K a/2) + B cos (X,a/2) = Ce~ K a


AK\ cos CK,a/2) + BK sin (X,a/2) = CK^e-*""
A sin (K,a/2) + B cos (X,a/2) = Ge _JC
AK\ cos (X,a/2) - BK\ sin (K,a/2) = -GK^e~ K m
'<

'2

(8-67)

12

(8-68)

a/2
"

(8-69)

(8-70)

SOLUTIONS OF SCHROEDINGER'S EQUATION

242

[Ch. 8

Subtracting equation (8-67) from (8-69) yields

2A

sin (K|a/2)

Adding equation (8-67)

IB

(G

- Qe'^"' 2

(8-71)

Qe-^ii^

(8-72)

to (8-69) yields

cos (K,/2)

(G

Subtracting equation (8-70) from (8-68) yields

2BK

sin (.K,a/2)

+ QKnc-^'^2

(G

Adding equation (8-70)

2AK

(8-73)

to (8-68) yields

cos (X,o/2)

= -(G -

C)Ke- x o/2

(8-74)

Provided 5

and (G

C)

# 0, we may divide equation (8-73) by (8-72)

and obtain

= Ku
(G + C) #

K, tan (K,a/2)

# 0and(G -

C)

if

Provided A

and

(8-75)

# 0, we may divide equation (8-74) by (8-71)

and obtain

X, cot (X,a/2)
if

It is

and

(G

-K|,

C)

(8-76)

easy to see that both equations (8-75) and (8-76) cannot be satisfied

simultaneously. If they could, the equation obtained by adding these two,

X, tan (K,a/2)

would be
becomes

valid.

+K

cot (K,a/2)

Multiply through by tan(AT|o/2).

2C|

tan 2 (X,a/2)

K,

Then

the equation

or
tan 2 (K,a/2)

But

this

= -1

cannot be valid as both K\ and a/2 are

real.

Thus

it is

only

possible either to satisfy (8-75) but not (8-76) or to satisfy (8-76) but not
(8-75).

Just as for the eigenfunctions of the vibrating string, the eigen-

functions of the square well potential form

two

classes.

For the

first

class,

K|tan(lS:|/2>=K||

G-

A=
C=

(8-77)

SQUARE WELL POTENTIALS

Sec. 4]

Then equation

243

(8-69) reads

cos (2^/2)

Ge~ K " al2

G = 5cos(X|a/2)e K

a/2
i'

=C

and the eigenfunctions are

K a/2 K x
[ cos (K\al2)e ~\e

y,(x)

[B] cos

(K

<

-a/2

-a/2

x),

<x<

a/2 (8-77')

\B cos

^>

(A' a/2)e^M/ ] e -^n^


2

a \2

For the second

class,

X| cot

Then equation

= K"n
B=
G+ C =

(A'|a/2)

(8-78)

(8-69) reads

= Ge^"
G = A sin (X|a/2)e K n 0/2 = -C
yl sin

(K

'2

a/2)

and the eigenfunctions are

[-4

sin

(K

al2)e

K " al2 ~\e K " x

<

-a/2

y,()

[A] sin

[A
Consider the

sin

first

-a/2

(A", a),

(X|a/2> K a/2 Je- K M^

a;

>

< x < a/2

(8-78')

a/2

Evaluating K\ and K^, and

of equations (8-77).

multiplying through by a/2, the equation becomes


(8-79)
VmEa /2h2 tan (VmEa2 l2h 2) = Vm(K - E)a2 \2h
For a given particle of mass m and a given potential well of depth V and
2

width a, this is an equation in the single unknown E. Its solutions are the
allowed values of the total energy of the particle the eigenvalues for
eigenfunctions of the first class. Solutions of this transcendental equation
can be obtained only by numerical or graphical methods. We present a
simple graphical method which will illustrate the important features of the

equation.

Let us

make

the change of variable


e

= VmEa l2H
2

j2h2

(8-80)

so the equation becomes

e tan
If

we

= VmV a

plot the function


p(e)

tan e

and the function


q(e)

= VmV a

/2h 2

(8-81)

SOLUTIONS OF SCHROEDINGEFVS EQUATION

244

[Ch. 8

the intersections specify values of s which are solutions to equation


(8-81).
at e

Such a plot
0,

it, 277,

function q(e)

is

shown in figure (8-15). The function p(e) has zeros


and has asymptotes at e = irj2, ~h-n\2, 5ttJ2, .... The

is

VmV a l2H
2

a quarter-circle of radius

number of

figure that the

It is clear

from the

solutions which exist for equation (8-81)

depends on the radius of the quarter-circle. Each solution gives an


E < V corresponding to an eigenfunction of the first

eigenvalue for

Figure 8-15.

graphical solution of the equation for eigenvalues of the first class of a

particular square well potential.

cl ass.

There

one such eigenvalue

exists

VfflVffl < 2n;


VrnV^fih 2
solutions: e

[Solution of e tan e

three

is

if

if

= VmV^jlh* e

VmV a j2h <


2

< VmV^fth < 3tt;


2

2tt

illustrated in the figure.

~ 1.25 and e c 3.60.

For

tt;

etc.

or

p(s)

two

= q{e).]
if

The

this case there are

From equation (8-80),

<

case

two

the eigenvalues

are

mV a 2

ma 2

4 /

and

2tf

-,v'

mVaa

$f)v. = <mv.

The eigenvalues corresponding

to eigenfunctions of the second class are

found from the solutions of an analogous equation obtained from


(8-78), which is

-e cot e =

VmV a l2h 2

(8-82)

SQUARE WELL POTENTIALS

Sec. 4]

245

Figure (8-16) illustrates the solution of this equation.


there will be

no eigenvalues

of the second class

Vm

if

E< V

for

< n/2; there will be one if 7r/2 ^


lit\2 < V mF q /2A < 5ir/2;
etc.
The

two

if

the

case

VmV cPI2fP = 4.

figure

illustrates

equation (8-82)

E=

Figure 8-16.

e*

mV a

graphical solution of the

class of a particular
r(e)

The

ci 2.45, and the eigenvalue

is e

apparent that

a2l2fi 2

VmV a l2h < 3tt/2;


2

It is

corresponding to eigenfunctions

single

solution

to

is

equation for eigenvalues of the second

square well potential. [Solution of

cot s

= VmV

o 2 /2A a

e2

or

q(e).]

We see that for a given potential well there are only a restricted number
E for E < F These are the discrete

of allowed values of total energy


eigenvalues for the

know

bound

that any value of

E is

On

of the particle.

states

allowed for

E> F

the other hand,

we

the eigenvalues for the

form a continuum. For a potential well which is very


shallow or very narrow or both, only a single eigenstate of the first class

unbound

states

VwF

a2/2/z 2 an eigenstate of the


be bound. With increasing values of
second class will be bound. For even larger values of this parameter an
will

additional eigenstate of the

first

consider the case

VmF^/W =

continuum eigenvalues are

bound. Next, an additional


bound, etc. As an example

class will be

eigenstate of the second class will be


4.

The

potential

and the

discrete

illustrated to scale in figure (8-17).

We

and
have

SOLUTIONS OF SCHROEDINGER'S EQUATION

246

used the quantum numbers n

1, 2, 3, 4, 5,

For

in order of increasing energy.

[Ch. 8

to label the eigenvalues

this potential

only the

first

three

eigenstates are bound.

Continuum

E2 = Q.37V

(2nd class)

(1st class)

Figure 8-17.

The eigenvalues

of a particular square well potential.

-Jatyx)

-^aip (x)
3

The eigenfunctions

Figure 8-18.

for the

bound eigenstates of

a particular square well

potential.

From the

VmV a

may

l2fi

solutions

2
,

of equations (8-81) and (8-82) for a given value of


the explicit forms of the eigenfunctions (8-77') and (8-78')
e,

be evaluated. The required relations are

*l| =

and

K | = s/mV a 2J2h2 -

(8-83)

SQUARE WELL POTENTIALS

Sec. 4]

The value of
function

4,

the eigenvalues

Et E2

must be adjusted so that each eigenFor the case

the three normalized eigenfunctions corresponding to


,

and

4= e

17.9

w,(x)

or

normalization condition (7-49).

the

Vm V a l2h =
2

the constant

satisfies

247

E3

are

3 80

x^-a/2

^,

'

1.26 -4= cos (1.25

Va

'

17.9

4= e

-380

-18.6-|=e

-a/2

),

< x < a/2

a/2/

^,

316

a;>a/2

*<-a/2

^,

va
Was)

.23

4=

sin

w3(x)

1.13

3 16
"

4= cos
,/a

-5.80

The

'

- a/2 < x < a/2

J3.60
-1

eigenfunctions, multiplied

a;^a/2

x^-a\2

^,

4= e

^,

1 74

4= e

a/2/

18.6 -j= e"

-5.80

2.45

,ya

74

J,
a/2/

^,

-a/2

< x < a/2

a;>a/2
figure (8-18) as a

by \/a, are plotted in

function of /(a/2).

This figure makes quite apparent the essential difference between the
The eigenfunctions of the first class,
classes of eigenfunctions.

two

ip x (x)

and

y> s (x),

are even functions of

rp{-x)

a;;

that

is,

(8-84)

+y>(x)

In quantum mechanics these functions are said to be of even parity.


The eigenfunction of the second class, tp2 (x), is an oddfunction of a;; that is
ip(x)

and

is

said to be of

odd

parity.

(8-78') will demonstrate that, for

(8-85)

xp{x)

Inspection of equations (8-77') and

any value of

VwK a /2S
2

2
,

the functions

SOLUTIONS OF SCHROEDINGER'S EQUATION

248

of the

first class

class are all of

are

odd

[Ch. 8

of even parity and the functions of the second

all

Furthermore, the continuum standing wave

parity.

eigenfunctions also form two classes, one of even parity and the other of

odd

That all the standing wave eigenfunctions have a definite parity,


is a direct result of the fact that we have so chosen the
origin of the x axis that the square well potential V(x) is an even function
of ie.f For such a choice it is intuitively obvious that any measurable
quantity describing the motion of the particle in a bound eigenstate (or
in a standing wave continuum eigenstate) will be an even function of x.
parity.

either even or odd,

We

note that

P(-x,

t)

density function P(x,

this is true for the probability

for both even* parity

and odd parity

y*{-x) y,(-x)

t)

eigenstates since

[yj*(x)][f(x)]

y>*(x) y(x)

= P(x, t)
(8-86)

This

is

not true for the wave function

eigenstate ; such a

wave function

is

itself in

the case of an

odd parity

an odd function of x. Of course,

this

not a contradiction as the wave function is not measurable. The traveling wave eigenfunctions do not have a definite parity since they are linear
combinations of standing wave eigenfunctions of opposite parities. For
is

a one dimensional problem the fact that the standing wave eigenfunctions
have a definite parity, if V(x) = V(x), is of importance largely because
it leads to a simplification in the mathematics of certain quantum mechanical calculations. But in a three dimensional problem this property takes
on a deeper significance that will become apparent later.

The probability densities multiplied by a,


of

VmV a l2h = 4,
2

for the three

bound

eigenstates

are plotted in figure (8-19) as a function of x/(a/2).

Also

illustrated in the figure is the normalized probability density which


would be predicted classically for a bound particle bouncing back and forth
between a/2 and +a/2. Since classically the particle would spend an
equal amount of time in any element of the x axis in that region, it would
be equally likely to be found in any such element. The quantum mechanical

probability densities

show

that, in

each eigenstate, there

is

finite

prob-

of finding the particle in the classically excluded region. This


effect is largest for n
3, which corresponds to the smallest value of
ability

En

(V
). Also there are points
quantum mechanical probability
ability

of finding the

in the classically allowed region at


densities

show

that there

is

which

zero prob-

particle.

ymV

For a potential in which


a2 l2tP is large enough that there are
many bound eigenstates, the quantum mechanical probability density
t In figure (8-18), if we redefine the origin of the x axis to be, for instance, at the
point x
a/2, it becomes apparent that the eigenfunctions will no longer have a

definite parity.

SQUARE WELL POTENTIALS

Sec. 4]

more and more

oscillates

potential, consider

249

rapidly with increasing values of n.

ip%(%) y(),

where n

is

For such a

large but not so large as to

violate the condition

y/mV a 2l2h2

-2.0

-1.6

-1.2

-0.8

- mE n a 2j2h2 =

-0.4

0.4

jLa/2

0.8

1.2

>1

(8-87)

2.0

1.6

x
aj2

Figure 8-19.

The

probability densities for the

bound eigenstates of a

particular square

well potential.

The

probability density has the

form indicated in

figure (8-20).

The

function a f*{x) y>J%) oscillates rapidly, and because of condition (8-87)


there is a negligible chance of finding the particle in the classically excluded

(I

(I

#;w*w-

All

aP(x),classical-

-o/2

11 111

a/2

Figure 8-20. Illustrating the approach to the classical limit of the probability density for

a square well potential.

In the limit n -* oo, the oscillations in ay)*(x) y) n (x) are so rapid


that only the average behavior can be observed experimentally and the
region.

results

of the quantum mechanical theory approach those of classical

mechanics.

SOLUTIONS OF SCHROEDINGER'S EQUATION

250

It is interesting

[Ch. 8

quantum
quantum numbers, as
of the old quantum theory.

to note that in this case the classical limit of the

mechanical theory

is'

found in the

limit of large

predicted by the correspondence principle

However, it should also be noted that it is necessary to apply the restriction


(8-87), which was not anticipated by the correspondence principle. For

Figure 8-21.

An even

a square well potential and a

unbound eigenfunction for

parity

typical value of total energy.

we have

seen that the approach to the classical limit

unbound

eigenstates,

has

relation to the technique of letting n -> oo.

little

Finally, let us consider the case:

>

V (standing waves)

E>

V
value of total energy
wave eigenfunctions are not the

Any

allowed, but the features of the standing

is

same

for all values of E.

We

shall

not

bother actually to solve the time independent Schroedinger equation for


this case

(it is

qualitative

easier to

do than for

arguments and

E< V

but shall only present certain

results.

In each of the three regions of the x axis the eigenfunctions are sinuIn the region
+ a/2, the momentum

a/2<x<

soidal standing waves.

of the particle

is

p = V2mE,

so the wavelength of the eigenfunctions

is

A,
is

hlVlmE. In

the regions x

p u = Vlm(E V ), and

<

a/2 and x

the wavelength

Figure (8-21) depicts the qualitative


for a typical value of E.

It is

>
is

momentum
= hfV2m{E V ).

+o/2, th e
kn

form of an even parity eigenfunction

easy to see that, because A N

>

\,

it

is

possible to join smoothly the sinusoidal solutions of the three regions only

< <

x
+fl/2 is less than it is
such that exactly an integral number
of wavelengths fit in the region a/2 to -fa/2, the slope of the interior
solution is zero at the boundaries and the interior and exterior solutions
have the same amplitude. This is shown in figure (8-22), which depicts
if

the amplitude inside the region

outside.

However,

if

the value of E

a/2

is

THE

Sec. 5]

INFINITE

SQUARE WELL

the eigenfunction for a value of

For odd

figure.
is less

251

slightly larger

than for the previous

parity eigenfunctions, the amplitude in the interior region

than the amplitude in the exterior regions except for values of E such
number of wavelengths fit in the region.

that exactly a half-integral

we

Considering both classes of eigenfunctions,

Figure 8-22.

An even

the total energy

is

at

parity

unbound eigenfunction for

see that the probability

a square well potential

when

resonance.

of finding the particle in the interior region, which

is

equal to

f/2

ip*(x) ip(x)
/:-a/2
is

dx

generally less than the probability of finding

it

in an exterior region of

the same length. This is analogous to the classical statement that the
particle moves faster in the interior region and so it would be less likely
to find the particle in that region. However, when the length of the region
is equal to an integral or half-integral number of wavelengths in the region,

the probability of finding

it

in the interior region equals that of finding

in a comparable exterior region.

This

is

it

precisely the condition (which

arose in the discussion of the traveling wave eigenfunctions) that the

region x
lead to

5.

The

An
limit

T will

be unity for the transmission of the particle from the


a/2 to the region x > +a/2. These so-called resonances
some interesting situations in atomic and nuclear physics.

probability

<

Infinite

Square Well

important special case of a square well potential is found in the


oo, which is illustrated in figure (8-23). This potential is

V(x)

0,

a/2

oo,

<

<

<

a/2

(8-88)

a/2,

>

a/2

SOLUTIONS OF SCHROEDINGER'S EQUATION

252

We

shall see that

[Ch. 8

easy to find simple closed-form expressions for the


For small values of

it is

eigenvalues and eigenfunctions of such a potential.


the

quantum number

n, these

eigenvalues and eigenfunctions can often

be used to approximate the corresponding (same ri) eigenvalues and eigenAn infinite
functions of a square well potential with large but finite V
.

-a/2
Figure 8-23.

square well potential

a/2

An

infinite

also used in discussing the

is

=*-

square well potential.

quantum mechanical

properties of a system of gas molecules (or other particles) which are


strictly

confined within a box of certain dimensions. In addition, such a

potential

is

used in the technique for the impenetrable walled box normali-

zation of free particle eigenfunctions.

Let us consider the

on the eigenfunctions of

effect

Shown in figure (8-24) is


potential. As V -> oo,

the eigenfunction for n

Ex

will

change; but

it

letting

->

oo.

in a finite square well

will

change very slowly

compared to the change in V Thus the coefficient of x in the exponential


will become infinitely large and the value of y>y(x) will decrease infinitely
.

- a/2
Figure 8-24.

The

a/2
first

eigenfunction for a finite square well potential.

x for x > a/2. In the limit, y>i(#) must be zero for


Then the requirement that an eigenfunction be everywhere
continuous demands that ip^ix) =
at x = a/2. For an infinite square
well potential, y>i(x) must have the form shown in figure (8-25). It is
rapidly with increasing
all

>

a/2.

apparent that

this

argument holds for

tpjx)

= 0,

all

x^-a/2,

the eigenfunctions.

x^ a/2

That

is,

(8-89)

THE

Sec. 5]

SQUARE WELL

INFINITE

253

for all n. The condition (8-89) can only be satisfied by violating the requirement of (7-38) that the derivative d\p n {x)jdx of an eigenfunction be con-

tinuous everywhere. However, if the reader will review the argument


which was presented to justify this requirement, he will find that dy)Jx)/dx
must be continuous only when V is finite.
For the infinite square well potential, we must solve the time independent

Schroedinger equation

d 2 f n(x)

subject to the condition (8-89).

2m _

(8-90)

Comparing with the

0C

-a/2
Figure 8-25.

The

first

COS

differential

equation

(^S *)

a/2

eigenfunction for an infinite square well potential.

(7-60) and the condition (7-63),

we

problem

see that the present

is

mathe-

matically identical with the problem of a vibrating string fixed at both ends.

The

solution of the present problem could be obtained directly from


equation (7-65) by making the appropriate substitutions. It could also be
obtained from the equations of the preceding section by letting V ~* oo.

However, we shall solve the problem again in order to illustrate how the
knowledge that the eigenfunctions must have a definite parity, because
V{ x) = V(x), can be used to simplify the calculation.
From equation (8-16') the standing wave form of the general solution
to (8-90)

is

y) n (x)

sin

(Knx)

B cos

-a/2

(Kx),

< x < a/2

where

Kn = V2mEjh
Since cos

of

x,

(Knx)

is

an even function of x and

sin

(Knx)

is

an odd function

the even parity eigenfunctions are


y, n (x)

= Bn cos (Knx),

-a/2

< x < a/2

(8-91)

-a/2

< x < a/2

(8-92)

and the odd parity eigenfunctions are


y> n (x)

= A n sin (Knx),

SOLUTIONS OF SCHROEDINGER'S EQUATION

254

[Ch. 8

In both cases the values of n must be such as to satisfy condition (8-89).


Because of the evenness or oddness of the eigenfunctions, it is only
necessary to apply condition (8-89) at x = +a/2. From equation (8-91)

we have

= Bn cos (Kn a/2)


This demands

Kna\2 =
which

tt/2, 377/2, 5tt/2,

Kn = mr/a,
Thus the even
n

1, 3, 5,

is

n=

(8-93)

parity eigenfunctions are those for

Next apply condition (8-89)

1, 3, 5,

at

quantum numbers
= +a/2 to (8-92)

and obtain

= A n sin (Kna/2)
This demands

Knaj2 =
which

tt, 2tt, 3tt,

is

Kn = nirja,
The odd
4, 6, ...

quantum numbers n

parity eigenfunctions are those for

To

2,

normalize the eigenfunctions, evaluate the integral

r
For even parity eigenfunctions

r0) Wn(X) dx
this is

Bz\ '\o{K n x)dx =


'-a/2

Since this

(8-94)

2,4,6,...f

must equal

unity,

B^2

we have

B n = V2fa
Similarly

it is

found that

A n = V2ja
Thus the normalized eigenfunctions are

V2/a
V>n(x)

cos (mrx/a),

= V 2/a sin {mrxja),

=
=

1, 3, 5,

2, 4, 6,

* < a/2,

0,

...

< x < a/2


a/2 < x < a/2
x ^ a/2
a/2

(8-95)

t The value n
ing solution Vote)

satisfies

condition (8-89), but

for all x.

it

leads to the physically uninterest-

THE INFINITE SQUARE WELL

Sec. 5]
It is

hardly necessary to plot

that there

is

or f*(x) y n (x). Except for the fact


excluded region for an infinite

y> n (x)

no penetration of the

255

classically

square well potential, the eigenfunctions and probability densities are


similar to those for a finite square well potential.

n -> oo

The approach

to the

same as for a finite square well except that


there is no auxiliary condition on n comparable to (8-87).
From equations (8-93) and (8-94), the eigenvalues are

classical limit as

is

the

En = K^jlm =
The

and the

potential V(x)

(8-26).

Of course,

TT

all

k=

2 2 2

h n l2ma 2

first

(8-96)

1,2,3,4,5,

few eigenvalues are

illustrated in figure

the eigenvalues are discrete for

an

infinite

square

^ = 16 i^

E, = 9

E, = A

Et =
-a/2
Figure 8-26.

The

first

is

the

first

is

a/2

to the region

< <
a;

is

bound

infinite

for

square well potential.

any

finite eigenvalue.

Of

7T

h 2 l2ma 2
It is the

(8-97)

lowest possible total energy

bound by the infinite square well potential


+a/2. The particle cannot have zero total

a manifestation of the uncertainty principle. If the particle


bound by the potential, then we know its x coordinate to within an

energy. This
is

if it is

eigenvalue,

called the zero point energy.

the particle can have

a/2

Ex =
This

2roa 2

2ma2

few eigenvalues for an

well potential since the particle


particular interest

7T

2ma 2

uncertainty

is

Ax

must be Apx

>

a.

hjt\x

Consequently the uncertainty in

hja.

The uncertainty

its

momentum

principle cannot allow the

particle to be bound by the potential with zero total energy since that
would mean the uncertainty in its x momentum would be zero. For the

eigenstate

the magnitude of the

momentum

is

px =

\2mEx =

-nhja.

SOLUTIONS OF SCHROEDINGER'S EQUATION

256

Since the particle whose motion

is

[Ch. 8

described by the standing wave eigen-

function can be moving in either direction, the actual value of the x

momentum

is

uncertain by an

agreement with the lower limit


point energy

is

amount &px = 2px = 2-nhja. This is in


by the uncertainty principle. The zero

set

responsible for several interesting phenomena, particularly

in connection with the behavior of matter at very

6.

low temperatures.

The Simple Harmonic Oscillator

We have

discussed

most of the

interesting cases of potentials

which are

discontinuous functions of position with constant values in adjacent


regions. Let us

now consider potentials which are

position. It turns out that there are only a limited

for which

it is

closed form.

continuous functions of

number of such potentials

possible to obtain a solution to the Schroedinger equation in

But, fortunately, these potentials include

some of

the

most

important cases: for example, the Coulomb potential V(f) oc r~ x which


will be discussed in Chapter 10, and the simple harmonic oscillator potential
V(x) oc x2 , which is discussed in this section. Techniques for handling the
,

more

recalcitrant continuous potentials will be presented in the next

chapter.

The simple harmonic oscillator potential is of great importance because


can be used to describe the broad class of problems in which a particle
is executing small vibrations about a position of stable equilibrium. At
a position of stable equilibrium, V(x) must have a minimum. For any
V(x) which is continuous, the shape of the curve in the region near the
it

minimum, which is all that is of interest for small vibrations, can be


approximated by a parabola. Thus, if we choose the zeros of the x
and the energy axis to be at the minimum, we can write
V(x)

C*\l

well
axis

(8-98)

a constant. This potential is illustrated in figure (8-27). A


influence of such a potential experiences a
dVjdx = Cx. The theory of
simple harmonic restoring force
classical mechanics predicts that, if the particle is displaced by an amount
x from the equilibrium position, and then let go, it will oscillate sinu-

where

is

moving under the

particle

F=

soidally

about the equilibrium position with frequency


v

fiL

2ir 'V

where

m is

energy

the mass of the particle.

of the particle

is

(8-99)

According to that theory, the total


2
and can have any value,

proportional to x

THE SIMPLE HARMONIC OSCILLATOR

Sec. 6]

since

is

arbitrary.

predicts that

257

However, the old quantum theory (Planck's postulate)


only one of the values:

E can have

En =

nhv,

0, 1, 2, 3,

(8-100)

What

are the predictions of quantum mechanics?


Let us specify the simple harmonic oscillator potential (8-98) in terms
of the classical oscillation frequency (8-99). From these two equations

we have

= l^mvV

V(x)

Figure 8-27.

A simple

harmonic

(8-101)

oscillator potential.

The corresponding time independent Schroedinger equation

2m

dx*

or

r?5

is.

is

(^yi

(8-102)

Define

Then

(8-103)

the equation can be written


cPipjdx2

It is

ImE/H2

iTrmvjh,

(fi

- oAr> =

convenient to express this in terms of a


|

Vac

new

_dy> d

dx~~dldx~
dx2

/-dip
yJa

d\dx)dx

variable,

(8-105)

a;

Then
djp

(8-104)

'

d$

d?

SOLUTIONS OF SCHROEDINGER'S EQUATION

258

[Ch. 8

and the time independent Schroedinger equation becomes

or

We

must find solutions


from oo to +oo. The
the solutions

which are continuous, and finite, for all I


condition will be automatically satisfied by

ip(g)
first

we shall obtain. However, it

consideration of the requirement that


this

purpose

values of

it is

useful

ip()

be necessary to take explicit


remain finite as ||[ > oo. For

will

to consider the

first

form of y(|) for very

large

|||

For any finite value of the total energy E, the quantity /8/oc becomes
compared to I2 for very large values of || Thus we may write

negligible

dhp/d?

The

= IV

general solution to this equation


xp

To check

Ae?'

|||

(8-107)

is

-* oo

Be-?1*

(8-108)

the solution, calculate

= Mefl2 + B(-)e-s
tPyld? = A?e?l + Ae^ 2 + BlV^ - Be~^ 2
= A( 2 + l)e?l* + B(2 - l)e~^ 2
2/2

dy>ldS

But for

|||

-> oo

this is essentially
2

A?e?'2

d 2 y\de

e(Ae^ 12

y>[d

B2e~&2

or

Now
III

->

Be'^' 2 )

|>;

Q.E.D.

apply the condition that the eigenfunctions must remain finite as


oo. It is apparent from equation (8-108) that this demands we set
>

.4

Thus the form of the eigenfunctions for very


y<|)

= Be-& 2

III

large

^ oo

|||

must be
(8-109)

This suggests looking for solutions to equation (8-106), which will be


valid for all |, in the

form
V<)

e"

|2/2

H(|)

(8-110)

Sec. 6]

THE SIMPLE HARMONIC OSCILLATOR

The H(i)

are functions which must, at

pared to e

_f2/2
.

To

^=

-e-t 2/2H

d\ =

e~ n2

d?

Then

substitute

-172

>

?e-!' <*H

-l /2

dH _

f /8

dH_

d
2

,dH

f ff

dH
2| =il

d2

+
.

df

(8-106).

dH

This gives

H + ^ e -l H -| e-^ H =
2

/2

/2

1~

<PH
d$ 2

d|

Dividing by e~^ 12 and canceling the terms involving

d 2H

/2

df2 J

'

e _f

<?H

and d2 xpjd2 into equation

xp

be slowly varying com-

oo,

evaluate these functions, calculate

-H + H

-H +

|f|

259

2
,

H, we have

g-l)*-0

(8-111)

This equation determines the functions H().

Let us recapitulate.

As

equation (8-106).

We started with the time independent Schroedinger

stands, this equation cannot be solved in closed


However, by writing the solutions to the equation as products of
_|2/2
the function e
which is the form of the solution for ||| -> oo, times the
functions H(g), we transform the problem to one of solving equation
it

form.

This equation is soluable in closed form. In fact, it is identical


with an equation called the Hermite differential equation whose solutions
(8-1 1 1).

are functions which are well

known

in the mathematical literature.

However, we shall solve it ourselves.


To do this we must make use of the most general procedure available
for the purpose of obtaining an explicit solution to a differential equation.
This consists in assuming a solution in the form of a power series in the
independent variable. That is, we assume

H (0 = 1 a** =
fc

The

+ aJ +

a.f

"s?

(8-112)

(8-1 12) into (8-1 1 1)

are then determined by substituting equation


and demanding that the resulting equation be satisfied

for any value of f

Calculate

coefficients

ax a 2
,

= 2 ka^' 1 =
^T
a?
*=i

lfl!

2a 2

3a 3 | 2

SOLUTIONS OF SCHROEDINGER'S EQUATION

260

[Ch. 8

and

^=
Then

|(fc

l)fca t f*-

2a 2

2a2

-2

3a3 |

4a4 ^

5a5 3

- 2 3c3 |
+ 08/a - IK + (|9/a - lKf + (/S/a - IK** + (/J/a

l^f

2a 2 f

4a 4 | 2

This gives

substitute into the differential equation.


1

3a 3 |

l)a3 3

Since this is to be true for all values of , the coefficients of each power of
I must vanish individually so that the validity of the equation will not
depend on the value of . Gathering the coefficients together and equating

them to

zero,

we have

+ (PI* - IK =
-2- 1K =
2-3fl3 + (iS/a- 2 2)c2 =
3 4a4 + (0/a 4-5c5 + (^/a- -2-3)a3 =

?:
I

2o 2

For the kth power of the

+ 2K+.+ (--l-2fc)a t =

(fe+lXfc

?:

relation is

so

(fe

This

is

called the recursion relation.

+
It

l)(fc

2)

allows us to calculate successively

and the coefficients a3 a5


ay The coefficients a and a1 are not specified by
equation (8-113). But this is as it should be; the second order differential
equation for #(!) should have a general solution containing two arbitrary
constants. Thus the general solution splits up into two independent series,
which we write as
the coefficients a 2 a4 , a6
,

a~,

JJ(|)

in terms of a

in terms of

= ajl + -2 e + -4 -

+ 4* + Tj3 +

+ -6 -4 -

%%? + %%%?+)

The ratios a]c+ Jak are given by the recursion


an even function of I, and the second series
variable.

relation.
is

The

(8-lW)

first series is

an odd function of that

THE SIMPLE HARMONIC OSCILLATOR

Sec. 6]

For an arbitrary value of


contain an infinite

both the even and the odd

/?/a,

number of terms. As we
Consider either

acceptable eigenfunctions.

261

shall see, this will

and evaluate the

series

of the coefficients of successive powers of f for large k. This

_ Q3/q-l-2fc) _
(fc + l)(fc + 2)

a
iSsa
ak
Let us compare
e

|2
,

which

e?

with the ratio for the

this

2fc

2
fc

fc

expansion of the function

=i+

t
+L+L+

fc+2

tie

+ __ + _

3!

(fc/2)!

(fc/2

large k, the ratio of the coefficients of successive

l/(fc/2

ratio

is

is

2!

For

series

series will

not lead to

1)!

(k/2)!

l/(k/2)!

(fc/2

(fc/2

k/2

l)(fc/2)!

1)!

powers of

(k/2)!

1)!

+1

is

^J_ = 2
fc/2

means that the terms of high power in f


e^ can only differ from the corresponding terms in the
even series of i/(f) by a constant K. They can only differ from the terms
in the odd series of //() by I times another constant K'. But, for ||| <x>,
the terms of low power in I are not important in determining the value of
any of these series. Consequently we conclude that

The two ratios

are the same. This

in the series for

tf(f)

Ke?

Now, according to equation

ax

K'^\

|||

-+ oo

(8-1 10), the solutions to the time independent

Schroedinger equation are

V()
Thus,

if the series

of H(J-) contain an infinite number of terms, the behavior

of these solutions for


e-*

But

2/2

= e-^H{)

||

H(!)

this diverges as

- oo

is

= a Ke& +
2

||

-*

oo,

aJCSt?*,

which

is

|||

-> oo

not acceptable behavior for an

eigenfunction.

However, acceptable eigenfunctions can be obtained for certain values


Set either the arbitrary constant a or the arbitrary constant ax
/?/<x.
equal to zero. Force the remaining series of if() to terminate by setting
of

/?/

= 2n+

(8-115)

where
n
n

= 3, 5,
= 0, 2, 4, 6,
1,

iffl

==

if ax ==

SOLUTIONS OF SCHROEDINGER'S EQUATION

262
It is clear

[Ch. 8

from equation (8-113) that such a choice of /?/oc will cause the
nth term since we shall have for k = n

series to terminate at the

(/?/(n

The

coefficients

l-2n)

lXn

a+4 ,

portional to a n+i

(2n

2)

a+8

fl +6 ,

The

+ 1 - 1 - 2n)'a =
(n + lXn + 2)

_
.

will also

resulting solutions

0-a

be zero since they are pro-

Hn () are polynomials of order

Hermite polynomials. For polynomial solutions to the Hermite

n, called

equation, the corresponding eigenfunctions

r(f)

= e-^H H(S)

(8-116)

always have the acceptable behavior of going to zero as ||| - oo,


|2/2
varies so much more rapidly than
||[, the function e~
the polynomial //() that it completely dominates the behavior of the

will

because, for large

eigenfunctions.

In

fact,

y({) oc e-?'\

oo

||

in agreement with equation (8-109).

Acceptable solutions to the Schroedinger equation exist only for the


/J/a specified by equation (8-115). According to (8-103), this

values of

quantity

is

2mE

/?

Thus equation (8-115)

2,

_2E

hv

Ittttiv

specifies the

allowed values of the total energy E,

which are
2Ejhv

In

or

En =

(n

\)hv,

0, 1, 2, 3,

(8-117)

where v is the classical oscillation frequency of the particle, and where


we have used the quantum number n to label the eigenvalues. The potential
and the eigenvalues are illustrated in figure (8-28). All eigenvalues are
discrete for the simple harmonic oscillator potential, since the particle is
bound for any finite eigenvalue. Comparing equation (8-117) with the
energy quantization equation postulated by Planck (8-100), we see that in
quantum mechanics all the eigenvalues are shifted up by an amount
\hv. As a consequence,
bound to the potential
this

the
is

minimum

EQ =

potential, the existence of

principle. It

is

\hv.

possible total energy for a particle

This

which

is

is

the zero point energy for

required by the uncertainty

interesting to note that Planck's postulated quantization of

Sec. 6]

THE SIMPLE HARMONIC OSCILLATOR

the simple harmonic oscillator


\hv.

263

was actually in error by the additive constant

This constant canceled out in most applications of the Planck

postulate because they involved only calculating the difference between two
(cf. Einstein's arguments leading to the idea of quanta).
However, discrepancies were noticed when attempts were made to explain
certain low temperature phenomena in terms of the old quantum theory.
Each Hermite polynomial HJ) can be evaluated from equation (8-114)

eigenvalues

_
4 -

3 _
_
E2 _
Ei -

9
2
7

2
5

by

The

first

hv

3
2

hv

The

first

HJ =
#itf) =
Htf) =
H3 () =
Ht(g) =
H (i) =
H (S) =
5
6

few eigenvalues for a simple harmonic

calculating the coefficients

for that value of n.

from (8-113) with

few are

listed

hv

Eq ~
Figure 8-28.

hv

hv
oscillator potential.

/?/a

given by (8-115)

below.

21
2

4f

- 8f3
12 - 48| 2 + 16| 4
1201 - 160|3 + 32f*
120 - 720I2 + 480f* 12

(8-118)

64f

In each case the arbitrary constant a or ax has been so chosen that the

normalized eigenfunctions are given by

*c-(jkrJ e ~'"HM
The

eigenfunctions can be written in terms of x

= Va x,

by making the

(8 " 119)

substitution

more convenient to leave them in terms of


an even function, the parity of the eigenfunctions is the
same as that of the Hermite polynomials even for even n, and odd for
odd n. The eigenfunction corresponding to the eigenvalue of lowest
energy is of even parity, just as in the case of a square potential well.
I

Since e

but

_|2/2

it is

usually

is

SOLUTIONS OF SCHROEDINGER'S EQUATION

264

[Ch. 8

few eigenfunctions are plotted in figure (8-29). The vertical lines


of. the classically allowed region in each case. In figure
(8-30) we show the quantum mechanical probability density for n = 10,
and also the probability density which would be predicted classically for a

The

first

indicate the limits

f
i

-5

ra

An=

\li
+5
1

Vi

+5

ra

iii

-5

if

Jol

ill

/ +5
'

i-

s.

Figure 8-29.

+5

The

first

-6

-5

+5

-5

few eigenfunctions for a simple harmonic

-4

oscillator potential.

-2

Figure 8-30. Illustrating the approach to the classical limit of the probability density for
a simple harmonic oscillator potential.

From

Quantum Mechanics, McGraw-Hill Book Co.,


particle of total energy equal to

Ew

L.

Pauling and E. B. Wilson, Introduction to

New

York, 1935.

The

velocity of a particle classically

executing simple harmonic oscillations goes to zero at the extremes of

its

motion. Consequently the particle spends most of the time near the limits
of the classically allowed region, and the classical probability of finding the
particle in

an element of the axis behaves as shown. The figure makes

Sec. 6]

THE SIMPLE HARMONIC OSCILLATOR

265

quite apparent the nature of the correspondence between the

mechanical and

classical probability densities in the limit

n -

quantum

oo.

To close our discussion of the simple harmonic oscillator, let us calculate

T and V of the

the expectation values


particle in the state n

0.

kinetic and
The wave function is

= w^y- iE

fed,

" tin

potential energies of a

4i e_|2/2

"**"*

77

Now
2m
Thus, according to equation (7-80),

Change

to a derivative with respect to

,-d
r d

d = Jo.
v =- = J
v
a.

djxx

dx

3f

Then

The

derivative

is

SX =
ai

The complex conjugate

1
tt

is

\J-*

_ J_ e -^l2 e +iE

t/A

IT

So
Im^JTT J -oo

The

definite integrals

can be found in any

h2x

"V^
[

4 J

we have

T=

They

_ ViH _ a^

m^TT L 2
Evaluating a,

table.

fplirmv

4mh

hv

4m

give

SOLUTIONS OF SCHROEDINGER'S EQUATION

266

According to equation (8-117)

[Ch. 8

this is

f= /2
From

(8-101), (8-103),

and (8-105) we may write the potential energy as

hnv?

Then, according to (7-80),

00

yjTT

F=

>

-oo

fV*

Ivh^TT

2vhir\\

hvjA

/2

Jo

can be shown that the relation

It

T=P=EJ2
true for all values of the

is

relation

is

quantum number. Furthermore,

true classically if T and

T and V averaged

of

(8-120)
the

same

V are understood to represent the values

over one period of the oscillation.

BIBLIOGRAPHY
Bohm,

D.,

Quantum Theory, Prentice-Hall, Englewood Cliffs, N.J., 1951.


and E. B. Wilson, Introduction to Quantum Mechanics, McGraw-

Pauling, L.,
Hill

Book

Schiff, L. I.,

Co.,

New

York, 1935.

Quantum Mechanics, McGraw-Hill Book Co.,

New

York, 1955.

EXERCISES
1

Derive equation (8-53) and verify (8-54).

2.

Derive equation (8-55).

3.

Consider the potential


0,

<0

<
>

Vi,
y,

<

where

<
and a

particle of total energy

increasing x.

Show

that

its

Vi

E > V2

< v2

approaching x

in the direction of

probability of continuing into the region x

>

is

Ch.
a

8]

EXERCISES

maximum

if

in the region
4.

is

<

267

an integral or
x < a.

half-integral

number of de

Broglie wavelengths

Obtain the transmission factor for a square well potential by redefining

the terms of equation (8-55), as suggested near the beginning of section 4.


5.

Evaluate quantitatively the standing wave solutions for the square well

potential of section 4 in the case

E> V

qualitative conclusions reached at the

end of

Use

these solutions to verify the

that section.

Obtain an analytical solution to the time independent Schroedinger equation


and determine the first eigenvalue Ex
Compare with the results of the numerical integration performed in that exercise.
6.

for the potential of exercise 4, Chapter 7,

7.

n ($) by using equations (8-113), (8-114), and


Check these against the forms presented in (8-118), and also verify

Evaluate several of the

(8-115).

that several of the eigenfunctions (8-119) are properly normalized.


8. Verify

equation (8-120) forn

1.

CHAPTER

Perturbation

Theory

I.

Introduction

The technique which we employed

in solving the Schroedinger equation

for the simple harmonic oscillator potential will not, in general, be of use

form V(x). What happens is that


found to involve more than two coefficients,

in the case of a potential of arbitrary

the recursion relation

is

making it impossible to find

solutions to the differential equation in closed

In such cases the equation can always be solved by numerical


integration in the manner described in section 5, Chapter 7. This method
form.

can be tedious if the calculation must be done "manually." But, with the
aid of modern high speed computing machines, numerical integration is
completely feasible and is widely used. In addition, there are approximation
techniques that are very useful for treating certain potentials for which
the Schroedinger equation cannot be solved in closed form.

The study

of one of these techniques forms the subject of perturbation theory, to

which

2.

this chapter is devoted.

Time Independent Perturbations


Consider a potential

V'(x), for

which

it is

either difficult or impossible

to solve the Schroedinger equation in closed form, but which can be

decomposed as follows
V'(x)

where V(x)

is

V(x)

v(x)

(9-1)

a potential for which the Schroedinger equation has been


268

TIME INDEPENDENT PERTURBATIONS

Sec. 2]

269

a potential that is small compared to V(x). We


from which it will be easy to obtain good approximations to the eigenvalues and eigenfunctions of the perturbed potential
V'(x), in terms of the perturbation v(x) and the known eigenvalues and
eigenfunctions of the unperturbed potential V(x). An example of equation
(9-1) is illustrated in figure (9-1). The potential V'(x) has been decomposed into a square well potential V(x), plus a perturbation v(x) which is
small compared to V(x).
solved,

and where

v(x) is

shall develop expressions

Co ntinu um

Co ntinu um

~4~

-3
-2

3
2

V'(x)

V(x)

v(x)

Figure 9-1. Illustrating the decomposition of a perturbed potential into an unperturbed


potential plus a perturbation.

On several occasions in Chapter 7 we wrote a function of


certain set of eigenfunctions (cf equations 7-56
.

for

some

and

That

particular perturbed eigenfunction Y^(a;).

as a sum of a
Let us do this

a;

7-70).
is,

we

writef

(9-2)

The

fi(x) are the

specify

how much

tion runs over

unperturbed eigenfunctions, and the coefficients a nl


is contained in y'n(x)- The summa-

of each of the ^(z)

the values of the quantum number /, including those in


The unperturbed eigenfunctions are solutions to the time

all

the continuum.

independent Schroedinger equation for the potential V, which

^dSpi
2m dx2
The perturbed eigenfunctions
potential V, which is

Vxpi

=E

l y> l

are solutions to the

2m dx2

V'rp'n

is

(9-3)

same equation for the

E'n y>'n

t It can be shown that every set of eigenfunctions y>i(x) forms a "complete set" in
terms of which almost any function of x can be expanded (i.e., written as in equation
9-2).

PERTURBATION THEORY

270

Using equation (9-1)

can be written

this

d2 y>'n

+ VW H +

values.

'

2m dx 2
where

vrp'n

>'

E and E'n are, respectively, the unperturbed


Now substitute equation (9-2) into (9-4)
l

2a

h 2 d 2 fl
i

[Ch. 9

+ 2 a nl v

2m dx

According to equation (9-3) the bracket

is

(9-4)

and perturbed

= J a nl E'n

y> t

equal to

eigen-

xp x

Thus we have

i?,y>,.

2 anlElVl + 2 a nl V Vl = 2 a nlE'n Wl
or

2 am(E'n ~ EdVi = 2 "mm


i

Multiplying through by the complex conjugate of a certain unperturbed


eigenfunction

y>

m and
,

we have

integrating over all x,

Too

/"oo

2 a m(E'n -

E)

ip*y> t

dx

00

= J a nl
J
I

The unperturbed eigenfunctions

ip*py dx

00

are, of course, orthogonal.

Assuming

that they have also been normalized,! equation (7-54) shows that the

on the left will be equal to zero if /


m, and equal to unity if
m. Thus there will be only one non-vanishing term in the summation

integral
1

on the

left side

of the equation, and


a nm (E'n

- EJ = 2 ai
wlpfi dx
'-OO

(9-5)

Let us define the symbol


/*00

vmi

Then equation

=J

"*(*)
V>t

(*) 1P00

dx

(9-6)

00

(9-5) can be written

a n JE'n

- EJ = 2 a nl v ml

(9-7)

This equation
that
is

is

small

useful,

is

we

exact, but
shall

it is

not very useful. In order to obtain one

employ the condition that the perturbation

compared to the unperturbed potential

t This involves

box normalization for the continuum eigenfunctions. In

the continuum eigenvalues actually

become

doing; this,

discrete (although very closely spaced).

This removes any difficulty of interpreting the summation


of/.

v(x)

V(x), so that the perturbed

^ fr the continuum values


I

TIME INDEPENDENT PERTURBATIONS

Sec. 2]

from

potential V'(x) differs only slightly

271

For such a

V(x).

situation

reasonable to assume

from the unperturbed eigenfunctions. In terms of equation


means that we assume

slightly
this

< 1,

a nni

'c=: 1,

If

we

it is

it is

that the perturbed eigenfunctions will differ only

^n

(9-2),

,
.
Q
y."-)

compared to the eigenvalues of V(x),


be small compared to the unperturbed

also require that v(x) be small

clear that the v ml

must then

all

eigenvalues because, according to (9-6), these quantities are just certain

Now

averages of v (x).

unperturbed eigenvalue

us divide both sides of equation (9-7) by the

let

Em We have
.

~ h m)

(E n

a nm

v mt

-rp

Z. "nl

Em

hm

Every term in the summation, except the term / = n, is the product of


two small quantities a nl and v ml \Em We shall neglect such terms, keeping
only the term for I = n. Then we have
.

Em

Em

or
a nm\Pn

Now take m =

We

n.

~ Em
)

(~9)

a nn v mn

obtain

SO
If

we

m # n,

take

E'n

we

~a

a
Setting a nn

-E ~ v nn

(9-10)

obtain

^ _

because of (9-8), this becomes


a
"mm

r^

Ehz>

"n

(9-11)
c 'm

Using (9-10) to evaluate ^, we find that


_

"mil.

En

Em +

V<m.<n.

(E n

- Ejll +
\

E n Em \

%n

Ej

Vnn

En ~

En E m \

EJ)
En

Ej

PERTURBATION THEORY

272

We

have taken the

first

Em

quantity v n J(En

).

[Ch. 9

term in the binomial expansion of 1 plus the


shall drop the term involving the product

Next we

- EJ and v n J(En - EJ. The validity of


two steps depends on the additional requirement that v(x) be small
even compared to the difference between En and any other eigenvalue
Em which enters into our calculations. We have finally
of the two quantities v m J(En
these

(9-12)

^n

t'n

Equations (9-10) and (9-12) are the expressions which provide good

approximations to the eigenvalues and eigenfunctions of the perturbed


potential V'(x). Consider (9-10), and evaluate v nn from (9-6). This yields

E'n

-E n ~ v nn = Jr

vC(*) v(x) Wn(x) dx

(9-13)

oo

This gives an approximation to the nth perturbed eigenvalue in terms of the


nth unperturbed eigenvalue and a certain integral involving the corre-

sponding unperturbed eigenfunction and the perturbation

v(x).

This

integral is the expectation value of v(x) for the nth unperturbed eigenstate.

To

see this, consider (7-74), with V(x,

e -iEn tm y,jx).

t)

v(x)

and T(a;,

t)

= WJx, t) =

The equation reads

v(x)

=J

iEt/n

v *^

-iE n t/n

vj^ dx

or
iff)

=
J

V>Z( X)

< X) VnW ^

oo

Thus perturbation theory

gives the very reasonable result that the shift

in the energy of the nth eigenvalue, due to the presence of the perturbing
potential v(x),

is

approximately equal to the value of v{x) averaged over

the nth unperturbed eigenstate with a weighting factor equal to the probability density y>*{x) y>Jx) for that eigenstate.

Next consider (9-1 1), and evaluate the symbol v mn


a nm

1
==:

yt(x) v{x)

xp

n {x) dx,

This yields

m^n

(9-14)

This equation gives the approximate value of the coefficients a nm which


specify how much of each of the unperturbed eigenfunctions y> m (x) is
mixed in with the dominant unperturbed eigenfunction y>Jx) to form the

TIME INDEPENDENT PERTURBATIONS

Sec. 2]

273

perturbed eigenfunction y^O**). Then in the series (9-2), with /replaced by m,

= 2 anmV>m(x)

W'( x)

(9-15)

we may use (9-14) to evaluate all the coefficients except ann From
(9-8) we know o ~ 1. Its exact value can be determined by requiring
.

that fn(x) be normalized. Note that a nm is proportional to \f(E


m).
Thus the perturbation v(x) will mix in with the unperturbed eigenfunction

amount of any unperturbed eigenfunction

ipjx) only a negligibly small

Em is very different from the

eigenvalue En This
has the important consequence that a good approximation to the series
(9-15) may be obtained by taking only the term for
n, plus a few
terms for
not very different from n. The coefficient a nm is also pro-

ip

whose eigenvalue

m(x)

m=

portional to the quantity

v mn

This

is

ft lx) v{x)

a certain average of

v(x),

y> n {x)

dx

with a weighting factor ip%(x)

y>(x)

which depends on the eigenfunction for the mth unperturbed eigenstate as


well as the eigenfunction for the nth unperturbed eigenstate.

The

quantities v mn , for

m~

n as well as

m / n,

are called the matrix

elements of the perturbation v taken between the state n and the state m.
This terminology is used because in advanced treatments of quantum

mechanics it is convenient to consider a matrix in which each element


one of the quantities v mn Such a matrix,

is

^12

"13

"22

"23

"m2

contains

all

possible information concerning the application of a perturba-

tion v(x) to a system

whose unperturbed eigenfunctions are

ipxix), yi^x),

PERTURBATION THEORY

274

3.

[Ch. 9

An Example
Let us

illustrate the

tion calculation.

We

use of (9-13) and (9-14) by doing a simple perturbashall evaluate the first eigenvalue

for the potential indicated in figure (9-2)

a
V'(x)

-a/2

We

which

is

an

2
(9-16)

a/2

V-bottom

^=

0,
oo,

<

*-

potential.

<x<

-a/2

>

sum of an unperturbed

consider this as the

<-a

x
a

<

oo,

and eigenfunction
by the equation

specified

a/2'

Figure 9-2.

<

and

potential,

a/2

>

a/2, x

a/2

square well, plus a perturbation

infinite

v(x)

lasl

a/2

According to (8-95), the normalized unperturbed eigenfunctions are

Wm( x)

yjlja cos (rrnrx/a),

m=

1, 3, 5,

yj2ja sin (m-n-x/a),

m=

2, 4, 6,

According to (8-96), the unperturbed eigenvalues are

Em =
where we use
the

first

TT

h 2 m 2 l2Ma2

M for the mass of the

m=

1, 2, 3, 4,

particle.

If d is small

eigenvalue,

E1 = TrW/lMa2
the perturbation technique should be applicable.

compared to

AN EXAMPLE

Sec. 3]

To

evaluate

which

k2

7T

m) J -a/2

(1

-m
mr)) J -a/2

7T /i

m=

2, 4, 6,

C\m

w=

3, 5, 7,

16M6
o^o2
_2fc2

ir

Z=

sin

m/1

dx,

)dx,

l|x|cosl

the integrand

Thus we have

Let

ViO)

ax,

77x/a;

a-im
1

m=

0,

the integral

(1

-m

a/2

3, 5, 7,

m=

2, 4, 6,

(ttx\

^ cos

x.

Since the

the integral will

2,4,6,...

(mirx\
cos

^7
2

0,

an even function of x;

is

m=

an odd function of

is

taken over a range symmetrical about x

vanish.

im

v ( x)

\\x\cos\
\

is

fl

wt( x)

integral

For

cos

(1

SM6

For

This gives

in (9-14).

is

a lm

take n

rp'^x),

im

275

it

gives
_

, _
m = 3, 5, 1,

dx,

...

then this becomes

-^
2
7T

C*

-m

(1

,z

cos

;r
2
) Jo

(mZ) Z cos

Z dZ

where we have introduced the convenient dimensionless ratio 6jEx


2Ma2 6JTrzh2 The integral can be evaluated easily by writing cos (mZ)
.

l( e +imZ

=
~

8_

Ulm

The

77

first

e-imZ^ etc
_d_

Ex (1 - m

The
/

result j s

+ 1M2] 2(m + l)

cos [(m

) I

few non-vanishing

cos [(m

have the values

coefficients

13

32
15

77

77

1728

a 19

864
17

n E1
2

(5

arw\ ~2 F

1)tt/2]

- l)
m = 3,

2(m

5, 7,

PERTURBATION THEORY

276

[Ch. 9

not surprising that alm =


for m = 2, 4, 6,
The perturbed
is symmetrical about the origin, and so its first eigenstate
must be of even parity. Consequently there can be no odd parity unperturbed eigenfunctions mixed into the first perturbed eigenfunction,
It is

potential V'(x)

and the odd parity unperturbed eigenfunctions are


The perturbed eigenfunction ^[(x)
2, 4, 6,

m=

precisely those for


is

obtained by sub-

alm in the series (9-15). Since the alm decrease rapidly with
increasing m [owing partly to the l/CE^ Em ) term and partly to the
v ml term], it is apparent that we can get a very good approximation to the
stituting the

series

by taking only the terms

m=

for

m=

and

3.

Thus

8 d
%0)
^ aiiViC^) +
~l
32
E
1

Vi(x)

tt

must be adjusted so that yfa)

Finally the coefficient an

but we leave

an

this as

(9-17)

is

normalized,

exercise for the reader.

Figure (9-3) illustrates equation (9-17). The relative amount of y>3 (z)
has been exaggerated for the sake of clarity. Fixing our attention on
ipfo) and fifr), we see that the second derivative of the perturbed eigen-

ends of the region a/2 to +a/2,


compared to the second derivative of
the unperturbed eigenfunction. Consideration of the form of the time
independent Schroedinger equation for the perturbed and unperturbed
function

and

is

relatively small near the

relatively large near the center,

potentials will

Next

let

make

it

clear

us evaluate E{

why

happens.

this

Taking n

E{.

in equation (9-13),

inserting the appropriate unperturbed eigenfunction,

E'1

-E = -^
1

m
\x\

cos

a J-a/2
E'x

E =
x

x cos 2

a Jo

E1

-E =
1

2
TT

which

dx

a \dx
)

Zi

JO

77

( )

and

we have

\16

4/

is

E'1

-E = 0.297(5
1

Figure (9-4) shows the perturbed eigenvalue E{ in terms of the dimensionless ratio (E[ E^jE^, plotted as a function of the dimensionless
ratio d/E^ Perturbation theory predicts the straight line of slope 0.297.

The

points are the correct answer.

They were

calculated

from the

eigen-

values E[ obtained by an exact (numerical integration) solution of the

AN EXAMPLE

Sec. 3]

277

Schroedinger equation for the four potentials V'(x) corresponding to the


The shift in the energy of the first eigenvalue, as

values of 6jEx indicated.

oc

Wx)
oc

Figure 9-3.

Illustrating

the composition of the

first

4/Jx)

eigenfunction for a V-bottom

potential.

predicted by perturbation theory,


for

{E[

is

seen to be in error by about 10 percent

~ 0.9, which corresponds to (E[ E-^jE^ ~ 0.25. For


E )IE ~ 0.05, the error about 0.5 percent. Now
apparent
d/E1
1

it is

is

that the error in the perturbation theory

we have developed

is

of the order

of the square of a small quantity since, throughout the development, the

Figure 9-4. A comparison between the first eigenvalues for several V-bottom potentials
obtained from perturbation theory and from exact solutions of the Schroedinger equation.

squares of small quantities were always neglected.

The numbers just quoted


an approximate measure of the size of
this small quantity is the ratio (E{ E )/E
Note also that the eigenvalue
1
1
E{ calculated by perturbation theory is always too large. It can be shown

indicate that, in the present case,

PERTURBATION THEORY

278
that this

why

it

is

to

tp-jix)

true for

happens.

any form of the perturbation

and

E[

E =

of ip[(x),

we

y>*(x) v(x)

y^x)

is largest,

eigenvalue E[

it is

easy to see

Comparing

dx.

see that this procedure gives too

to the values of v(x) near the ends of the region.

region v(x)

and

Perturbation theory uses the unperturbed eigenvalue

evaluate

plots of ipi(x)

v{x),

[Ch. 9

the

much weight

But near the ends of the

and therefore the contribution of v(x)

to the perturbed

overestimated.

is

comparison of the exact form of the eigenfunction y>[(x) of the


form (9-17) predicted by perturbation theory shows

potential V'(x) with the

that the error in the coefficient a13

quantity {E[
If

is

also of the order of the square of the

E^)jEv

more accurate

estimates of E'n

and

y'Jx) are needed,

it is

possible to

extend the perturbation theory to obtain expressions in which the error

is

of the order of the cube, or even of a higher power, of the appropriate


small quantity. However, in practice equations (9-13) and (9-14) are
normally adequate.

The Treatment

4.

of Degeneracies

Consider the case of two different unperturbed eigenfunctions, which


label ip^x) and f 2 (x), whose corresponding unperturbed eigenvalues

we

Ex

and

E2

happen

to be exactly equal.

These eigenfunctions are said to

We shall see later that a number of important examples

be degenerate.

this situation actually arise in the

of

study of atomic and nuclear physics^

For instance, many of the eigenfunctions are degenerate for an electron


bound in the \jr Coulomb potential of a hydrogen atom. When eigenfunctions are degenerate, we shall often be interested in studying the effect
of a small perturbation which changes the potential in such a way as to
remove the degeneracy.
However, to apply perturbation theory in a case involving degenerate
we must make certain modifications in the theory. This
need is clearly indicated by equation (9-11), which makes the prediction
1 and a
1 for the case Ex = 2 .t This really tells us only that
a 12
21
the theory we have developed breaks down in this case. But it also provides
some clue to the nature of the difficulty by showing that, when E1 = E2
/ (equation 9-8) is not consistent with the
the assumption a nl < 1 for n
results obtained from that assumption in the cases n = 1, 2 and / = 1,2.
eigenfunctions,

~
~

Equation (9-11) states a 12


v 21 l(E{ E2 ). Taking E1 = E2 and using (9-10),
becomes a 12
v^KE^ JS,)
1, in general. Note that this result does
%/
not depend on the "additional requirement" that v(x) be small compared to the
difference between two unperturbed eigenvalues.
t

this

THE TREATMENT OF DEGENERACIES

Sec. 4]

The

when we

difficulty is resolved

279

realize that there is certainly

<

priori basis for the assumption that a 12

and a21

<

1,

no a

when fi(x)
Under such

correspond to eigenvalues which are exactly equal.


circumstances it might very well be that, in contrast to the assumption, a
small perturbation could have a big effect and thoroughly mix up the two

and

y> 2 (x)

degenerate eigenfunctions.

To
we

account for

we modify

this situation

the procedure as follows. First

consider only the mixing, due to the presence of the perturbation

of the two degenerate unperturbed eigenfunctions y^x) and


we ignore the mixing with ip^x) and

v(x),

with each other. In doing this

of any of the non-degenerate unperturbed eigenfunctions


y> b (%),

....

From

we

this calculation

y?(aO
(x)
ip
2

=
=

y> 3 (x),

y> 2 (x)
ip

2 (x)

%(x),

shall find the eigenfunctions

aiiViO*)

a 2 np x {x)

+
+

OiaVaOe)

a 22 y> 2 (x)

and the eigenvalues

In

but exceptional cases it will turn out that El = E\, i.e., the perturbaremoves the degeneracy. The eigenfunctions y\ and yfy are in error by

all

tion

first powers of small quantities (the a nl for n = 1 2 and / = 3, 4, 5, ...


which have been neglected), but the eigenvalues E\ and E% are in error only
by the squares of small quantities. Since it is usually necessary only to
know the eigenvalues, the calculation is often completed when E\ and E%
are obtained. However, if it is also necessary to know the eigenfunctions,
we may take the set of non-degenerate eigenfunctions and eigenvalues

the

yRx), y&r), rp%x),

rp\(x),

po pO pO pO
^"2' ^3' ^4)

^1'

where
Vh( x)
y/fcx)

=
=

x)

and

E%

f^x)

and

E\

V>3(

=E
= 4

and apply the normal non-degenerate perturbation theory to obtain the


set of eigenfunctions and eigenvalues
ViO0>

V2O), Vs(aD, viC*).


t'Xi

^2 "3> ^4

El

and

E'2

>

where
E[

El

PERTURBATION THEORY

280
All these eigenfunctions

and eigenvalues are

in error only

[Ch. 9

by the square of

a small quantity.

Now

us study the calculation which evaluates

let

Assume

that y>i(aO

the other

and

ip3 (x), y> (#), y> (x)


4
5

y>, ip 2 ,

E\, and E%.

are orthogonal to each other, as well as to

y> 2 (x)
.

Assume

.f

all

also that all the eigenfunctions

have been normalized. Then it is apparent that we may write an equation


identical with (9-7), but E'n replaced by E%. That is,
a nm\E n

Em =
)

2, a nl v ml
I

Now

set all

a nl

obtained, one for

except for n

m=

a n i(K

= 2
=

Ei)

a^iK - E2) =

2 and

1,

m=

and one for

Two

1,2.

equations are

2:

a nl v u

1,

1,2

n=

a n iVn,

1,2

1=1,2

Evaluating the summations explicitly, these equations give

-E) =

n=l,2
(9-21)
+ a n2 v12
(9-22)
n = 1, 2
a n2 (En - E) = a nl v 21 + a n2 v 22
where we have set E = E2 = E. Write these equations in the form
n = 1, 2
(El - E - v n )a nl + (-v 12 )a n2 = 0,
n = 1, 2
(-%)m + (E- E - v22)a n2 = 0,
"ni(En

a nl uu

Then

it

becomes apparent that they constitute a pair of homogeneous


unknowns a nl and a n2 In order that a solution

linear equations in the

such a set of equations, it is necessary that the determinant of the


coefficients of the unknowns be equal to zero. That is, we must have

exist for

- Vl2

l(K-E- Vll)
\

-v21

(En

Evaluating the determinant,


(

- E - vn)(En

we

= h2

--

(9 _23)

v 22 )J

find

-E- v

22 )

v 21 v 12

0,

= 1, 2 (9-24)
v n = v 22 and that
n

In many cases of physical interest it turns out that


v 12 and t>2i are real. To simplify the discussion, we shall assume that both
hold true. Since the definition (9-6) shows that in all cases v 21 = yf2
we shall then have v21 = v12 as well as vn = v22 and equation (9-24)
,

becomes
(En
t

-E- vnf -v\

The footnote on page 189 proves

that this

= 0,
is

possible.

1,

THE TREATMENT OF DEGENERACIES

Sec. 4]

This

a quadratic in

is

2s.

one for each value of


the

sign for n

has two solutions,

It

- E - vn = v12

En

n.

281

We

1,

arbitrarily take the

2
sign for

and

Then

2.

= E + v11 + va
E% = E + vn vXi
Now consider equation (9-21) with n = 1. It is
E1

ii(-E?

Substitute

-E) = a n vu +

E\ from (9-25) and


fliiO>u

transpose.

iz

u ii)

au

a 12

o 12 i;12

The

(9-25)

result is

iai2

Therefore

Thus the perturbed eigenfunction


eigenvalue E\,

is

= a u \%px{x) +

xp\(x)

The

coefficient

ip 2 {x)~\

an must be adjusted to normalize

that this requires an

ip%x).

It is

easy to see

1/V2. Then

>?(*)

and we

(9-18), corresponding to perturbed

equal to

= -4 [>i(*) +

%(*)]

(9-26)

also have

= E+v11 + vu
Consider equation (9-21) with n = 2. It is
a 2i( 2 E) = a 21 u + a 22
E1

t>

Evaluating

fijj,

(9-260

t;

12

we have

a 2l(>ll

^12

V ll)

"22^12

So

a 21 =

a 22

and
Normalization gives

#*)
and we

= -4 Oi(*> -

V2 (*)]

(9-27)

also have

Ez

= E + vn -

i>

12

(9-27')

PERTURBATION THEORY

282

Except for the unusual case in which the matrix element

i>

12 is

[Ch. 9

equal to

and (9-27') show that the perturbation v{x) has


removed the degeneracy. The two perturbed eigenvalues E\ and E\ are
separated by the energy

zero, equations (9-26')

E\-El =
Then

y>i(x), tp%(x),

tions

y3 (x),

2v 12

(9-28)

E\ and E\, plus the original non-degenerate eigenfuncand eigenvalues E3 Et


provide the completely
non-degenerate set of eigenfunctions and eigenvalues (9-19) to which we
can apply the normal perturbation theory and obtain (9-20). Note that
equations (9-26) and (9-27) certainly confirm our suspicion that even an
y>i(pc),

arbitrarily small perturbation will

produce a large mixing of the degenerate

eigenfunctions fi{x) and y> 2 (x), because both y>%x) and y>$(x) contain
equal amounts of these eigenfunctions, independent of the magnitude

of

v(x).

For a case in which v12


v 21 and/or v n ^ v 22 the procedure is
the same in essence but the details are more complicated. For a case in
which there are a. degenerate eigenfunctions, the procedure is the same
except that equation (9-23) becomes a determinant of order <x. This leads
to an equation analogous to (9-24) of order a in E%- For large values of a,
such an equation can be very difficult to solve, but it is often possible to
perform certain manipulations on the determinant so that the equation for
E comes to a form for which its roots are easy to find.
Furthermore,

it is

often possible to eliminate the entire procedure of

degenerate perturbation theory, and use instead non-degenerate perturbation theory, if physical arguments can be used to find the particular linear

combinations of degenerate eigenfunctions that do not mix under the

Consider the problem that we have been


and assume that y> and xp\ are known a priori. Then E\ and E\,
which are usually what we are really interested in finding, can be evaluated
simply by applying the equation (9-13) of non-degenerate perturbation
theory. That is, equation (9-13) applied to ip gives
application of the perturbation.
treating,

El

- E = JV vfl dx = j^= [ft +


=

f 11

y>*t>yi

*>12

dx

f*vyj 2

Wtlv

dx

J|

[% + %]

y*fVi dx

dx

y)*c^ dx

THE TREATMENT OF DEGENERACIES

Sec. 4]

283

and the same equation applied

in agreement with equations (9-25),

to

ft g ives
E2

-E = vSVS dx =
J
J
=

f*vrp x dx

= iKl - 12 = v n - v 12

-j= [y*

- y*> -1= [vi - vJ

f*vy 2 dx

21

ip* v/,Pi

dx

d*

ip*vrp 2

dx

U 2 2]

also in agreement with equations (9-25).

Non-degenerate perturbation theory can be applied to the particular


linear combinations yl and y>1 of the degenerate eigenfunctions y> x and
f2 because those particular combinations are not mixed by the application
of the perturbation. We can see this by evaluating the integrals that appear
in the coefficients which, according to the equation (9-14) of non-degenerate
,

perturbation theory, determine the mixing.

rpfvxpl

dx

Of +

i[^ii

wt vxPi dx

"is

y>f\v -j= |>i

ip*vrp 2

"21

dx

For

instance,

V2]

dx

y*vfi dx

y>*vip 2

dx

^22]

Similarly,

[xpfvipl

dx

So the perturbation does not mix y\ and


tion theory can

=
ip

2,

and non-degenerate perturba-

be applied directly to these particular linear combinations,

even though they are degenerate before the application of the perturbation.
In subsequent chapters we shall consider a number of examples of the
physical arguments which are used in

quantum mechanical systems

to find

the particular linear combinations of degenerate eigenfunctions that are

not mixed by the application of a perturbation. (Cf. Chapters 11, 12, and
13.) We consider here an example of a physical argument which can be
used in a classical system. In one of the higher frequency modes of a
circular drum head, the drum head vibrates with a nodal line lying along
a diameter. This mode is degenerate because the same frequency is
obtained for all orientations of the nodal line. But there is only a twofold degeneracy because there are only

vibrations whose nodal

lines

are

two independent vibrations

perpendicular.

the

These independent

PERTURBATION THEORY

284

[Ch. 9

degenerate vibrations are indicated in figure (9-5). All other vibrations


at this frequency can be obtained by linear combinations of these two.
In particular, other sets of two independent degenerate vibrations, with
perpendicular nodal lines of different orientation, can be obtained by
appropriate linear combinations. In the absence of a perturbation, all
these sets are equivalent.
Node

Node

Node

Node

Two

Figure 9-5.

independent degenerate vibrations of a circular drum head.

Figure 9-6. Applying a perturbation to a circular

Figure 9-7.

Now

The

drum

head.

results of applying a perturbation to a circular

drum

head.

imagine applying a perturbation by fixing a small weight to the


at some position other than its center, as indicated in figure

drum head
(9-6).

Because of the asymmetry introduced by the perturbation, the two

new
The forms of these new vibrations, as well as their frequencies,
can be found by a calculation using the procedure of degenerate classical
perturbation theory. The vibrations which are found are indicated in
previously independent vibrations are mixed together to form two
vibrations.

figure (9-7).

It is also

found that the perturbation removes the degeneracy

because the weight lies along the nodal line for one vibration and therefore
has no effect on the frequency of that vibration, while it does effect the
frequency of the other vibration.

TIME DEPENDENT PERTURBATION

Sec. 5]

THEORY

285

After gaining some experience with these problems, it is possible to tell


from a priori physical arguments what the form of the perturbed vibrations
must be. This makes it possible to eliminate the complicated procedure of
degenerate classical perturbation theory by taking the set of independent
degenerate unperturbed vibrations to be the particular set for which one
nodal line runs through the weight. Then the application of the perturbation does not mix the vibrations because they have the same form both
before and after its application, and the non-degenerate classical perturbation theory can be used. This considerably simplifies the calculation of the

frequency

shift

produced by the perturbation, which

is

the quantity that

is

usually of interest.

5.

Time Dependent

Perturbation Theory

We now develop a theory for treating perturbations which are functions


of both position and time. This is an important case for several reasons,
one being that time dependent perturbation theory provides the only useful
method

for solving the Schroedinger equation for a time dependent

r).f Thus we consider a time dependent potential


which can be decomposed as follows

potential V(x,

V'(x,

where V(x)

is

i)

V(x)

equation for V(x) are the

set

The

t)

(9-29)

v(x, i)

a time independent unperturbed potential and

small time dependent perturbation.

V'(x,

v(x,

t) is

solutions to the Schroedinger

of unperturbed wave functions

Yn(x, = e~ m^ Wn{x)
i)

(9-30)

where the En and ip n (x) are the unperturbed eigenvalues and eigenfunctions.
Assume that a solution to the Schroedinger equation for V'(x, t) can be
written

= 2>(01 (*,
r

'H*.

(9-31)

where the coefficients aJJ) are functions of time. Different solutions will
have different sets of coefficients, but here we shall not use a second subscript to indicate this explicitly.

02xp

2M

dx%

fixir,

+ V'W

=
-ih
dt

exception is a time dependent potential of the form V(x, t) = Vt(x) + F2 (f).


form only, the Schroedinger equation can be separated in the manner of section
Chapter 7, by assuming a solution 'Ffo t) = y>(x) <j>(t).
t

One

For

this

4,

fc2

Substitute (9-31) into the equation

PERTURBATION THEORY

286

which

it is

supposed to

[Ch. 9

This gives

satisfy.

"

2, a n

9 F

3T

2m

ax2

at

2v

The bracket vanishes because


complex

conjugate

T m = e~

lEmt/n

ip

of some

m and
,

the

equation for the potential V.

at

are solutions to the Schroedinger

Multiply the remaining terms by the

wave function

unperturbed

particular

Then, evaluating the Y, we

integrate over all x.

have

J oo

Since the

y> n

are orthogonal

J go

ut

and normalized,

= - 7 2 "-(^-^--^""fm.

(9-32)

where we have extended the definition of the matrix element v mn (equation


9-6) to include time dependent perturbations. This is a set of coupled
first order ordinary differential equations, one for each m, which determine
the a n (t). The details of the solution of these equations depend on the
details of the particular problem at hand. We consider here a simple but
illustrative case.

Assume a perturbation of the form


v(x,

t)

t<0
>

0,

v(x),

This

is

a perturbation v(x) which

is

(9-33)

"switched on" at

0.

For

this

case the set of unperturbed wave functions (9-30) are exact solutions for

<

t
0. Next assume that the wave function for the particle is known to
be equal to a single one of these wave functions, say
k (x, t), for t < 0.
This amounts to assuming that the total energy of the particle is known to
be precisely Ek for t < 0. This does not conflict with the uncertainty

principle

At AE >
because in the

infinite

time before

h
it

(9-34)

would be possible

to

measure

In terms of equation
(9-31), this assumption provides the following set of initial conditions for
the energy of the particle with perfect precision.
the a(t) att

0.

.(0)

0,
t

1,

n^k
n

(9-35)
,

THEORY

TIME DEPENDENT PERTURBATION

Sec. 5]

[We assume that the ajf) do not change discontinuously at t


assumption will be justified by the results of the calculation.]
like to find the perturbed

wave function T'(z,

t)

287

0.

We

This

would

for the particle at a time

a n (t) for t > 0.


Let us require that the perturbation v(x) be small enough, or that the
time t be short enough, that
t

>

0.

To do

this

we

shall evaluate the

a n (t)

Then we may
for n

k.

neglect

all

< 1,

^ k,

1,

>

(9-36)

k,

terms in the right side of equation (9-32) except

This gives

~ --a

daji)

-i(E h

k (t)e

-Em)tlh V

(9-37)

dt

To

m=

evaluate ak (t), set

Then

k.

<tafc(0

dt

a*(0t>**

or

da k(t)

fl*(0

Integrate both sides from

independent of

for

>

to

t'

>

all

- tW

In a k (t)

which

remembering that the v mle are

0,

This gives

0.

is

a k (t')

In

^ - 7 Vat',

According to (9-35), ak (0)

=1. So we

a k (t)

>0

_%(0)J

~ e - iVatin

find
t

>

(9-38)

where we have dropped the primes to simplify the notation.


Next evaluate the ajf), n = k, by setting m = n in (9-37) and by making
the additional approximation that ak {i) = 1. We have
da(t)

_ _ 1 e -nB -B.)tiH v
t

dt

n=k

or

da n(t)

~ - - v nk e- i(E*- E wn dt,

n^k

PERTURBATION THEORY

288
Integrate

from

to

>

t'

to obtain

-i(.E k -E)t/h

(0

Ejs.

From

(9-35),

an (0)

0,

[Ch. 9

so

y-nEh -Bn )tm _

a n (t)

in>

B _^ k

(9-39)

-C<i

&Tf.

where we have again dropped the primes.


(9-31), (9-38), and (9-39), we find

+ 1 E.-E,

[e

Evaluating T'(,

-X>-BMH_

l]g-./ y^a;)

t)

from

(9_4Q)

Note that the energy Ek + vkk appearing in the exponential of the first
term is exactly the perturbed energy E'k = Ek + vkk which would be
predicted for a completely time independent perturbation equal to v(x).
It is of interest to consider the quantity a*{i) a n {t). This real function of
Multiplying
t is the square of the magnitude of the coefficient ajf).
conjugate,
we find
(9-39) into its complex
,

sin

]%Lp*

2h

al(t)a n {t)~?*&*-

(9-41)

&

(^J

This quantity oscillates in time between zero and ^v*k v nlc \{En
with frequency v
sin 2 [(En
fixed t.

(En

Ek)/h. We plot in figure


- Ek)j2hf as a function of

Ek)2

(9-8) the factor

- Ek)tl2h]l[(En
(En - Ek)\2h for
Now the wave function describing the particle initially contained

Wk

(x, t). The perturbation v(x, t) has


from other eigenstates over a whole
range of the quantum number n. However, we see that the most important
contributions come from those n which correspond to eigenvalues En
lying within a range centered about Ek and of width AE, where

only the eigenstate wave function

the effect of mixing in contributions

AE/in

it it

or

AE~2nhlt
In section
instant

t is

8,

Chapter

7,

we showed

(9-42)

that the value of a*(t) a(t) at

any

equal to the probability of finding the particle in the eigenstate

Sec. 5]

TIME DEPENDENT PERTURBATION

THEORY

289

n at that instant. (The argument actually considered the a n to be constants,


but inspection will demonstrate that it is equally valid when the a n are
functions of t .) Thus at any time t there is a certain probability of finding
the particle in final eigenstate n which is different from the initial eigenstate
k, and with total energy E different from the initial total energy Ek This
appears to be a violation of the law of conservation of energy by an
amount En Ek which may be large compared to the energy vkk supplied
.

2ir

IT

3t

En~ Ek
2%
Figure 9-8.

The

plot of a function

which

by the perturbation. However,

arises in

time dependent perturbation theory.

in the time interval

to

the probability

of finding the particle with energy En is important only when EH Ek is


at most equal to about A.E, where t and A.E are related by equation (9-42).
According to the uncertainty principle (9-34) any measurement of the total

energy of the particle which


uncertain by an
A.E. This

amount of

removes the

is

must be
which is comparable to
and provides an example of this statement
carried out in this time interval

the order of

difficulty

hjt,

of the uncertainty principle.


Consider equation (9-41) for small values of

>

0.

The equation

says that the probability of finding the particle in a particular eigenstate

is

proportional to the square of

linear dependence

t.

This statement

is

in contrast to the

that might be expected intuitively.

However,

always based on our experience with systems in the


In that limit the resolution of any experimental apparatus

physical intuition
classical limit.

on
is

PERTURBATION THEORY

290

[Ch. 9

so large compared to the separation of the eigenvalues, or even to the


width of the range AE, that it is not possible to measure a*(t) a n (t) for a
single value of n. All that can be measured classically is the total probability of finding that the particle has made a transition from the initial
is

eigenstate

to

some other

the transition probability

which

Pk =
To

evaluate this,

we assume

is

dNn

dNn

dEn

per energy interval

that" there are

is

express this in terms of

defined as

2n *(0 (0

spaced final eigenstates in the range

Then

We

final eigenstate n.

Pk

AE;

the

(9-43)

a large number of closely


number of final eigenstates

fc

a*n (t)a n (t)dN n

=\

J -oo

75ifJ-

00

Owing

{t)^dE n

J -oo

Evaluating a*(t) a n (t) from (9-41),

ft

at{i)a n

dhi n

we have

oo

-^

r( K

sin"
*
V nk v nkPn-

2H

dE r

(^J

to the factor sin 2 [(En

tribution to the integral

dNjdEn .
integral over

the density of final states p n

summation over n can be approximated by an


That is,
the

- Ek)tj2h]j[(En - Ek)j2hf, most of the con-

comes from the range AE.

If

we assume

that the

matrix element v nk and the density of final states p n are both slowly varying
functions of n in that range, we can write

~ir

Pn

i-

The quantum number n now

dE n

2h

refers to

00

V n!e V nlc

j f= f

a typical final eigenstate in the

initial eigenstate k.

X - eA

2h
(

neighborhood of the

- Ek

\E n

sin"

Let
sin

Z=
Z

(En

Ek)tj2fr;

then

dZ

which gives

D
**

^n

7r

"7" v nk v nkPrf

(9-44)

EXERCISES

291

Ch.

9]

The

transition probability

rate

Rk =

dPjdt

is

proportional to

independent of

is

Rk

t,

t,

as expected.

The

transition

since

~T V*nlPnlcPn

(9-45)

This important formula

is

often called Golden Rule No. 2.

It is

very

of very general applicability.! In any situation


in which transitions are made to an essentially continuous range of final
states under the influence of a constant perturbation, the transition rate
widely used because

it is

can be evaluated from this formula. Note that we have here a good example
of the use of quantum mechanics in the evaluation of transition rates.
The ability to do this is one of its most important advantages over the

quantum

old

theory.

BIBLIOGRAPHY
Bohm,

D.,

Quantum Theory, Prentice-Hall, Englewood Cliffs, N.J., 1951.


and E. B. Wilson, Introduction to Quantum Mechanics, McGraw-

Pauling, L.,

New York, 1935.


Quantum" Mechanics, McGraw-Hill Book Co., New York, 1955.
Sherwin, C. W., Introduction to Quantum Mechanics, Henry Holt and Co.,
New York, 1959.
Hill

Book

Co.,

Schiff, L. I.,

EXERCISES
1.

Use perturbation theory

eigenfunction

y> x (x)

to calculate the first eigenvalue

E1

and the

first

of the potential

oo,

V(x)

<

-a/2

a/2, x

<

<

>

a/2

a/2

a/2

where

8 is small relative to

calculation of section

Ev

Compare with

the results obtained in the

3.

Use perturbation theory to calculate the first eigenvalue Ex for the potential
4, Chapter 7. Compare with the results of the numerical integration
performed in that exercise, and also with the results of the analytical treatment
2.

of exercise

of exercise

6,

Chapter

8.

Except for certain pathological cases, no degeneracies arise in problems


involving one particle moving in one dimension. In order to obtain a simple
3.

We shall use it in Chapters 1 5 and 1 6, and use a closely related formula in Chapter 1 3

PERTURBATION THEORY

292

[Ch. 9

example of the application of degenerate perturbation theory, consider one


particle

moving

in the

V{x,y)
T

two dimensional
x

<

>

-a/2, x

-a/2

0,

infinite

<

<

square well potential

a/2,

<

-a/2

a/2,

-a/2, y

<

<

>

a/2

a/2

first section of the next chapter, set up the time independent


Schroedinger equation for this potential. Separate this partial differential

After consulting the

equation into two ordinary differential equations by the usual method, making
use of the fact that V(x, y) can be written V(x) + V(y). Since these equations,
and the conditions on y> at the edges of the well, have the same form as for a
one dimensional infinite square well, their solutions can be written immediately.

Note

that there are degeneracies in almost

all

the eigenstates.

Now

consider

the application of the perturbation

Make
effect

>

d,

0,

otherwise

a/4,

>

a/2,

a routine degenerate perturbation theory calculation to investigate the


first two degenerate eigenstates.

of this perturbation on the

4. Take the linear combinations of the unperturbed degenerate eigenfunctions


found in exercise 3, and show that if these linear combinations are used in
non-degenerate perturbation theory the energy shifts which are obtained are
the same as those obtained in that exercise. Show also that in non-degenerate
perturbation theory these linear combinations of the unperturbed eigenfunctions
are not mixed by the perturbation, even though they are degenerate before the

perturbation

is

<

0,

At

5.

dimensional

a/2.

At

applied.

an electron

infinite
t

0,

is

known

to

be

in the n

eigenstate of a

one

square well potential which extends from x = a/2 to


a uniform electric field E is applied in the direction of

The

electric field is left on for a short time t and then removed.


Use time dependent perturbation theory to calculate the probability that the

increasing x.

electron will be in the n

2, 3,

probabilities as a function of

used.

t.

or 4 eigenstates at

Hint :

Some

>

t.

Make

plots of these

of the results of exercise

can be

CHAPTER

10

One-Electron

Atoms

I.

Quantum Mechanics for Several Dimensions


and Several Particles

Up to this point we have restricted our treatment of quantum meqhanics


to a single particle

moving

in a one dimensional space.

tion for the discussion of atoms,

we must

Now,

in prepara-

generalize to the case of a system

of particles moving in a three dimensional space.

Let us consider

N particles

rectangular coordinates, x, y,
equation for this case, using the

whose positions are defined by a


z.

We

first

set

of

develop the Schroedinger

same technique as

in section 8,

Chapter

7.

We write the classical expression defining the total energy of the system:

+ Pi + p!)
i^-iPi
2m
+ V(x y %,..., x

i=i

x,

x,

},

y s , z jt ...,x N y N
,

t)

=E

(10-1)

The quantity^ is the x component of the momentum of the/th particle,


py and pZj are its y and z components, and rrij is the mass of that
particle. Thus the/th term of the summation is the kinetic energy of the
y'th particle, and the value of the sum is the kinetic energy of the system.
The potential energy of the system depends, in general, on the coordinates
of each particle, as well as on the time t. The quantities x jt y } z j are the
,

three coordinates of the


quantities
(cf.

px

7-77).

i=i

,pz

y'th particle.

and E, by

Now we

replace the dynamical

their associated differential operators

This gives*

+ -^
+ I +
d
dzy

2mAdx*

y)

293

V(x,,
y x ...,z N ,i)
}

ihdt

(10-2)}
v

ONE-ELECTRON ATOMS

294

[Ch. 10

Multiplying this operator equation into the equality

Y(xly ...,z N ,t)=


we obtain

(*!, ...,z N ,t)

(10-3)

the Schroedinger equation for the system,

2 3=i

v;y + vv = m

(io-4)

9(

2m.,-

where we use the symbol

V ? s ll + |l + Jl
3x^
a^ 9z^
which

( 10

-5)

called the Laplacian operator in rectangular coordinates of the

is

y'th particle.

The wave function


3iV~

variables

the time.

The

T^,

is

a function of

3 spatial coordinates for each of the N

particles, plus

for the system,

significance of this

wave function

t),

specified

is

by a natural
For

extension of the interpretive postulates (7-21), (7-31), and (7-80).


instance,
If,

we postulate

at the instant

that

a measurement

t,

is

made

to locate the particles in the

z N t), the probability


system associated with the wave function ^F^i,
x
P(xlt
,z N t) dxt
dzN that the
coordinate of the first particle will
be found between x 1 and x x + dx x etc., and finally that the z coordinate
.

of the Nth particle

P{xx

will

..,z N ,t)dx 1

be found between zN and

N+

dz N ,

is

dz N

= T*^,

,z N ,i)

T(xx

dxx

dz N

and
extensions
the
obtained
by
obvious
of
wave
function
be
the
(10-3) can
properties developed in Chapters 7, 8, and 9. As an example, it is easy to
show by the usual method that, when the potential V does not depend on
time, there are solutions to the Schroedinger equation which can be
Essentially all the properties of the Schroedinger equation (10-4)

written

Y(xx ...,z N ,t)=


,

xp( Xl ,

...,z N)

<Kt)

(10-6)

where
tft)

and where y(x2

z ) is

equation

e~ imin

a solution to the time independent Schroedinger

2 - 2m, V;fy +

3=i

(10-7)

Vy>

Erp

(10-8)

THE ONE-ELECTRON ATOM

Sec. 2]

295

Note that the time independent Schroedinger equation


case

we

are

now

wave function Yfo,


z N i) is denned
more than three dimensions. This makes it
the

function

is

in the general

is,

Also note that

discussing, a partial differential equation.


.

in a non-physical space of

quite clear that the

wave

only a computational device.

The One-Electron Atom

2.

The procedures involved in solving the time independent Schroedinger


equation (10-8) are best illustrated by taking a specific example. In this

we

our treatment of one of the simplest the onemost interesting example because
of the unique importance of the one-electron atom in atomic theory.

section

shall begin

electron atom.

The reader

This

is

also perhaps the

will recall that

feature of the old

Bohr's one-electron atom

quantum

We

theory.

electron atom is equally important in


Above and beyond its own intrinsic

system

is

fundamental because

mechanical treatment of

all

it

is

the

most

shall see that the role

the theory of
interest, the

significant

of the one-

quantum mechanics.

theory of this simple

forms the foundation of the quantum

multi-electron atoms.

The one-electron atom is a three dimensional system containing two


a nucleus and an electron, moving under the influence of a mutual
Coulomb attraction and bound together by that attraction. Let the mass
of the nucleus be m 1 and its charge be +Ze (Z = 1 for neutral hydrogen,
Z = 2 for singly ionized helium, etc.). Designate the mass of the electron
by w 2 and its charge by e. Then the (Coulomb) potential energy of the
particles,

system

is

V(xlt

=.

z2 )

Ze2

V(z 2

where xlt yx

zx

xxf

(y2

Vl f

(z 2

-zf

(10-9)

are the rectangular coordinates of the nucleus

and

.r

2,

y 2 z 2 are the rectangular coordinates of the electron. The denominator


of equation (10-9) is the distance between the two particles. The time
,

independent Schroedinger equation for

/9

2m x \

yj T

dx\

y>T

dy\

9Vr\
dz\ /

this case is

d ipi)

dx\

dy\2

d4-

V{xx ...,

z 2 ) VT

2m 2

d fi

ft

(9
v

fi

= E T yr

The reason for using the subscript T will become apparent


By making a change of variables, this equation goes over

(10-10)

later.

into a

form

in

ONE-ELECTRON ATOMS

296

[Ch. 10

can be separated into two parts. One part represents the transmotion of the center of mass of the system, and the other part
represents the motion of the two particles relative to each other. The

which

it

lational

new

variables are

= m1x + m
1

x2

= m 1 y1 + m z y2
m1 + m

(10-11)

_ m xzx + m 2z
m1 + m2

*-y

Figure 10 I.

The one-electron atom described

in

rectangular and spherical coordinates.

the rectangular coordinates of the center of

mass of the system, and

r,0,4>
the spherical coordinates of the electron relative to the nucleus.

These coordinates are


easy to see that

r, 8,

<f>

illustrated in figure (10-1).

are related to xx
r sin 6

cos

r sin 6 sin
r

cos 6

d>

<f>

z2

#2

/ 2

zx

z2

From

this figure it is

by the equations

xx

yx

(10-12)

Using equations (10-11) and (10-12), equation (10-10) can be transformed

THE ONE-ELECTRON ATOM

Sec. 2]

to the

new

This

variables.

is

2( mi

dy?
3f

fi

+ m %)\3x2

actually quite a tedious task,

3z
dz

'

T^ |5 + -T^r-M**
r

sin

3<

30

r sin

where

and so we

They are

present here only the results.


(d

297

J
/

+V
I?))
30/1

= _2MbL_
m +m

lu, \r
2/*

a nd

F(r)

dr

dr I

- ***

(10-13)

= -

The quantity jM is the reduced mass.


The separation of this time independent Schroedinger equation
effected

by

looking for solutions of the


ip

T (x, y,

z, r, 0,

<j>)

y>

Substitute this into equation (10-13),

is

form

CM (x,

y, z) ip{r, 0,

(10-14)

<f>)

and divide through by Vt

VcuV-

This gives
h

[_
L

2(m,

pVcM + g>cM + 3>cm V


2
2
I.

w OM \ dx2

m,)

3w

2
2/if\r dr\

dz

dr'
'

3> +

H
r

sin

3<f>

dfp\\,
- }
sin a
-r
2
36/
r sin 6 36\
3 (

The first term on the left is a function only of x,

ET

variables

r, 0,

constant.

y, z

V(r)

JBi

and the second term is

sum

of the two terms is equal to a


even though the variables x, y, z are independent of the
it is clear that each term must separately be equal to a
(f>,

a function only of
constant

r, 0,

Since the

<f>.

Thus we obtain two equations,

2(m x

+ m )y> cu \

dx2

By 2

dz

and

2i*y>\i*dr\

3r/

sin

6 3 </> 2

sin636\

where
-Ecm

+ E = ET

3d!)

ONE-ELECTRON ATOMS

298

Consider the

first

+m

V
2)

5ck

write

ycM

2(m x

we

equation, which

3y

[Ch. 10

"

ycM
dz
5

mass
whose
coordinates
are
those
of
the
center
of
mass
of
the
atom,
2>
i
and which is moving with total energy ECM in a region of zero potential.
The equation describes the translational motion of the atom. Consider
next the second equation, which we write
This

is

the time independent Schroedinger equation for a particle of

m +w

r
2

dr\

2fj.\r

Br)

sin

(sin

The curly bracket is V 2 ^j, where V


r, 8,

2
B<f>

sineae\

coordinates

is

This equation

<f>.

0^)1+
del)

= Ew

V(r)w
y,v

(10-15)
y
'

the Laplacian operator in the spherical


is

the time independent Schroedinger

equation describing the motion of the electron relative to the nucleus.

E = ET ECM where ET is the total energy of the atom and


CM is its energy of translational motion, E is the total energy of relative
motion that is, the total energy of the atom in a frame of reference in
which the center of mass is stationary. Equation (10-15) is formally
Since

"

an electron

identical with the time independent Schroedinger equation for

of mass

[i,

infinitely

instead of mass

x,

moving with coordinates

r, 8,

<f>

about an

massive nucleus which, consequently, remains stationary

section 4, Chapter
It is

(cf.

5).

apparent that in equation (10-14) we have decomposed the total

eigenfunction ^t mto tne product of an eigenfunction y CM specifying


the motion of the center of mass of the atom, times an eigenfunction y>,
,

specifying the

position

is

motion of the electron

of one particle relative

3.

relative to the nucleus.

possible whenever the potential

V depends only on

This decom-

the coordinates

to the other.

Separation and Solution of the Equation of Relative Motion

The

translational

motion of the atom

Consequently, we shall assume that

we

is

of no interest to us here.

atom from a frame


Then we need consider

are viewing the

of reference in which the center of mass

is

at rest.

only the time independent Schroedinger equation (10-15).

V = Ze2 jr

of a

Coulomb

any function of the radial coordinate

is

potential

For the case

and, in fact, whenever the potential


r only,

it is

possible to separate

THE EQUATION OF RELATIVE MOTION

Sec. 3]

299

To do

the equation into three ordinary differential equations.

for solutions of the

this,

we look

form
y{r, 6,

</>)

(10-16)

J?(r)0(0)O>(</>)

and dividing through by

Inserting this into equation (10-15)

R@Q>h

l2fx,

we obtain
,
l

r
r

Rdr\

>/M
J?\

d 2 <t>

sin

0<P d<4

sin 6

00 d6\

sin

d&\

del

+ |[^-nr)] =
Next multiply through by

l#<b
dd>

sin

=
~

d d /

sin 2 d

dR\

and transpose. This gives


d

sin 6

d0\

/
fl

dr\

dBJ

dd\

dr)

sin

0[E

F(r)]

and the right side only on r and 6,


Since the left side depends only on
both must equal a constant, which we shall call m 2 Thus we obtain two
<f>

equations,

= -m
d<

(10-17)

<P

and
1

dUdR \

sin

we

transposing,

l(
Rdr\

d0\

sin00d0\

By

dl.

d_R\

dr)

2^ 2rE
- r\E
h

</0/

.-.

F(r)J

2
m

sn 2

rewrite the latter as

=Jn^

2j^
h

sin

1_ .d

d&\

sin00d0\

dO

According to the usual argument, both sides of the equation must equal
a constant a. Consequently,
1

d (

sin a
.

sin0d0\

d@\

dO/

m 2 = n
a0
2

(10-18)

sin

and
2

+
-M*t)
r* dr\
dr I
The problem has been reduced

%n

-v

^R -

4
r

<

io - i9 >

to that of solving the three ordinary

differential equations (10-17), (10-18),

and

(10-19).

We

shall find that

ONE-ELECTRON ATOMS

300

[Ch. 10

equation (10-17) has acceptable solutions only for certain values of m.

Using these values of m in equation (10-18),

turns out that the equation

it

a. With these values of


a in (10-19), the equation is found to have acceptable solutions only for
certain values of the total energy E. Thus the energy is quantized.
Consider equation (10-17). The reader can easily verify that it has a

has acceptable solutions only for certain values of

solution

O(^)

Now we

must, for the

im *

time, explicitly consider the requirement that

first

the probability density

and probability

flux

have

single-valued, for all values of their variables.

definite values,

i.e.,

be

This must be the case

because they are measurable quantities describing the behavior of a real

We may ensure that it will be so by demanding that the


must be single-valued. This requires the function $(<)
to be single-valued, and it must be considered explicitly because the
azimuthal angles and 2v are actually the same angle. Thus we must be
physical particle.
eigenfunction

ip

sure that
<D(0)

This

= 0(2tt)

is

or
1

This

is

true only

cos miTT

|m|

that

is,

if

sin m2rr

if

0, 1, 2, 3,

has integral values.f

Thus the

(10-20)

set

of functions which are

acceptable solutions to equation (10-17) are


<D m

where
is

m assumes the values

(# = e**

specified

by

(10-20).

(10-21)

The quantum number

used as a subscript to identify the form of a particular solution.


In working with equation (10-18), it is convenient to change variables

from

8 to

where
I

t In

a more rigorous treatment

it is

cos

(10-22)

not necessary to demand that y be single-valued,

but only that the measurable quantity v>*V be single-valued. It is evident that this would
admit either integral or half-integral values of m. However, in the present context the
half-integral values can be ruled out because they lead to solutions of the 0(0) equation

On the other hand,


a quantum corresponding to m do arise in connection with
functions, related to the <b(<f>), that are used to describe "spin." This will be discussed

for which the probability flux has a physically impossible behavior.


half-integral values of

in the next chapter.

THE EQUATION OF RELATIVE MOTION

Sec. 3]

Then

301

the equation reads

+
a-a^)
m

d_

This differential equation

m
-

0(f)

(10-23)

e)

not directly solvable by the power series


relation involves more than two terms

is

method because the recursion


(cf.

But, by considering the behavior of the differ1, the extreme values of the variable, it is found
solution can be obtained for the function F() denned by
= (1 - 2 W/2 F()

section 6, Chapter

that a

power series

8).

ential equation at f

0(f)

After solving the differential equation for F(f),

found that solutions

it is

for 0(f), which have the acceptable behavior of being everywhere finite,
exist only when F(g) is a polynomial. This happens only when a has the

value

where

/ is

1(1

1)

1,

\m\

2, |m|

(10-24)

one of the integers


I

\m\, \m\

3,

(10-25)

an acceptable

Corresponding to each value of the quantum number I there is


solution

To

.Fi(f).

identify the particular solutions to equation (10-18),

which are written


*,(!)
it is

= (l-a W/2 ^(D

(10-26)

quantum numbers

necessary to use as a label the two

and m. The

functions 0[ m (f) are known in the mathematical literature as associated


Legendre functions. Some examples of these functions will be presented in
section

5.

Setting a

\_d_
r

/(/

1) in

^^ +

equation (10-19),

{_l(Ll)

;
dr^

In terms of the

new

we have

F(r)]

i?

(10-27)

variable

(10-28)

2/Sr

where
2
fl2
/3

and

_
=-_ IjxE

(10-29)

also using

= fiZe

(10-30)

h2p

the equation becomes

dR(p) \
I U im)
+
t

2
p dp\

dp

L i _ Sl+Ji + rU) ,
14

p)

awl)

ONE-ELECTRON ATOMS

302

[Ch. 10

Again it turns out that the differential equation is not directly amenable
to a power series solution. However, by considering the behavior of the
equation for p - oo, it is found that a series solution can be obtained for
the function G(p) defined by
R( P)

e- pl2 p G(p)
l

In order that G(p) be a polynomial, which

from diverging

where n

is

as p -- oo,

it is

is

necessary to prevent R(p)

found that y must have the value

1./

2,

(10-32)

one of the integers


n

The functions

= l+

/+

3,'...

(10-33)

that are acceptable solutions to equation (10-19),

which are

written

Rnlp)

e- pl2 p

G n (p)

carry as labels the quantum number n and the

(10-34)

quantum number /. Examples

of these associated Laguerre functions are given in section

5.|

The reader will note that the procedures involved in solving equations
(10-18) and (10-19) are essentially the same as those which were employed
in the solution of the simple harmonic oscillator equation. Thus he should
not have too

much

difficulty in filling in the details himself,

using the

outline presented above.

Quantum Numbers,

4.

Eigenvalues, and Degeneracy

According to equation (10-16), the eigenfunctions are

VmJT,
All three

6,

quantum numbers

<f>)

= R nl (r) @ [m (6) <B m (0)

(10-35)

are required to identify the eigenfunctions

because their mathematical form depends on the values of each of the

quantum numbers. That three quantum numbers

arise is

a consequence of

the fact that the time independent Schroedinger equation contains three
t

Note

that for a given set of

quantum numbers

Ri(p) to equation (10-31), even though this

is

there

is

only one acceptable solution

a second order differential equation and

two particular solutions. The reason is that one of the particular solutions
not acceptable as it becomes infinite within the range of the variable p. The same
situation obtains concerning the single acceptable solution & im () to (10-23).
therefore has
is

QUANTUM NUMBERS,

Sec. 4]

303

Gathering together the conditions which the

independent variables.

quantum numbers

EIGENVALUES, DEGENERACY

satisfy,

\m\
/

we have

= 0, 1, 2, 3,
= \m\, \m\ + 1, \m\ + 2,
= + 1,1+2, 1+3, ...
.

(10-36)

These conditions are more conveniently expressed in the form


n

/=

1, 2, 3, 4,

m =-/,-/+
It is

...,-

0, 1, 2,

1,

(10-37)

..., 0,

...,+/-

1,

+/

easy to see that equations (10-36) and (10-37) are equivalent.!

Now

let us combine equations (10-29), (10-30), and (10-32) to obtain


an expression for the allowed values of total energy for the atom containing
a single bound electron. This gives

E=
which

vzv =

h p _
=
_n^

2/j,

2fi

flyzv

hY

2/*W

is

pZ 2 e*
=-^-i>

1.2,3,4,...

(10-38)

2n n

Comparing this equation with (5-14), we see that the eigenvalues predicted
by quantum mechanics are identical with those predicted by the Bohr
theory. J

Schroedinger's derivation of this equation (1926) constituted the

convincing verification of the validity of the theory of quantum


mechanics.
first

It is important to note that, whereas the form of the eigenfunctions


depends on the values of all three quantum numbers n, 1, m, the eigenvalues
depend only on the quantum number n. Since for a given value of n there

are generally a
t

number of different possible values of / and m,

According to equations (10-36), the

minimum

value of \m\

is 0.

1, 2,

value

/ is

/ is

= + 1,
1.

the one satisfying the relation n


I
possible values of / are /
0, 1, 2, ...,
is

apparent

equal to \m\, and the


and the minimum value
= 1. Since n increases by integers without limit,
For a given n, the maximum value of /
3, 4, ...
value of

Thus the minimum value of

of , which is equal to /+1, is


the possible values of n are n =

which

minimum

it is

that

is,

1.

Finally, for a given

Consequently the
/,

the largest value

can assume is \m\ = I. Thus the maximum value of m is +/ and the minimum
/, and it can assume only the values m =/,/+ 1, ..., 0, ...,+/ 1,

\m\
is

+/.
%

the

Remember
finite

that in (5-14)

nuclear mass.

m will be replaced by

fi

when proper account

is

taken of

ONE-ELECTRON ATOMS

304

that there will be situations in which

two or more completely

[Ch. 10
different

eigenfunctions are degenerate because they correspond to exactly the same

The degeneracy with respect to m arises whenever the potential


depends only upon r, while the / degeneracy is a consequence of the particular form of the r dependence of the Coulomb potential V = Ze^jr
(cf. section 7, Chapter 5). From equations (10-37) it is easy to see how many
eigenvalue.

degenerate eigenfunctions there are which correspond to a particular

TABLE
Possible Values of

1,2,3

for n

Number

(10-1)

and

-1.0.+1 -2, -1,0, +1, +2

-1,0, +1

of

degenerate
eigenfunctions
for each

Number

of

degenerate
eigenfunctions
for each n

eigenvalue

and

3 are

(10-37),
1.

2.
3.

5.

The possible values of the quantum numbers for n = 1,2,


shown in table (10-1). From this table, or from equations

En

it is

apparent that:

For each value of n, there are n possible values of /.


For each value of /, there are (21 + 1) possible values of m.
For each value of n, there are a total of n2 degenerate eigenfunctions.

Eigenfunctions and Probability Densities

The eigenfunctions

are

WmJr,

8,

</>)

= R nl(r) e im (0) <D

TO (<)

and we know that

im(0)

R nl (r)

=
=

Jm4'

(polynomial in cos 6)
e-te<**rt* (polynomial in r)

AND

EIGENFUNCTIONS

Sec. 5]

All the

bound

PROBABILITY DENSITIES

state eigenfunctions

305

have basically the same structure,


/ the polynomials in r and in

except that with increasing values of n and


cos 6

become

increasingly

eigenfunctions for

more complicated.

unbound

In addition, there exist

an ionized hydrogen atom plus


which have a somewhat different

states (e.g.,

E>

an unbound electron in which

0)

we shall not be concerned at present with situations in


which unbound state eigenfunctions are needed. We list below the bound

structure, but

state eigenfunctions for the first three values

?ioo

/Z\'

4yj2TT\a

Vaii

-Zr/a

<>

cos

'

a,

= j/_zfZr e -Zr/2a sJn g e i4.


8V

^300

n.

-7=|-J
yJiT \a /

l/Z^Zi
Zr - Zr/Za
e

210

of

77"

_L/ZJV

18 +

-Zr/3a

(10-39)

e~
aly/TT \a

olyjir \a

V-320

-^
81^077

V32i

e
- ^f
\a
al

= rrV (-f
81^/77 \a

77^7=
1627"-

-ZH3a<,

cos

sin Q e

V C S2 6

i*

~D

^322

' Zrl3a

ZrlZa '

~ Zrl2a

sin

cos d ei *

aj

/Zf
ZV
i~ e -^r/3o sin 2 0e 2i*

These are written in terms of the parameter


a

= Ar =

0.529

X l(T 8 cm

(10-40)

lie*

which the reader will recognize as the radius of the smallest orbit of a
Bohr hydrogen atom.

ONE-ELECTRON ATOMS

306

Figure 10-2.

The

radial probability densities of

the one-electron atom for n

[Ch. 10

I,

2, 3.

AND

EIGENFUNCTIONS

Sec. 5]

The eigenfunctions (10-39)


/so

That

are normalized.

307

is,

f2ir

'"lm

J Jo
J
Jo

J
>o

where

/V

PROBABILITY DENSITIES

sin 6 dr

dd

(r, 0,

d<f>

<f>)

Vnlm (r,

the unit of

is

0)r sin 6 dr dB

6,

d<f>

(10-41)

volume in the system of spherical


Also the
all space.

coordinates, and where the integral extends over


eigenfunctions-are orthogonal, so that
Cn f 2ir

/*oo

Wtvm-ir,
Jo

unless ri

An

6,

<j>)

Wnl Jr,

<f>y sin 6 dr dd

6,

Jo Jo

n,V

and m'

I,

d<]>

(10-42)

m.

is best obtained by considering


form of the corresponding probability density functions ^*imim =

the

interpretation of the eigenfunctions

R%iimmRni imm- As these are functions of three variables, we cannot


plot them directly. However, an adequate understanding of their three
dimensional behavior can be had by discussing separately their dependence
on each variable. We consider first the r dependence in terms of the

Pn i(r),

radial probability density

defined by

n27T W*m m Wm m r2 sin

dr dd

d<f>

n2ir
The

R nl

integrals over 6

0, m and

Om

and
is

<f>

are equal to unity because each of the functions

separately normalized.

(They are also separately

orthogonal.) Thus

P nl (r) dr =

Now

r*R*nl (r)

R nl (r) dr

(10-43)

ytimWnnJ sm dr dd d<j> is the probability of finding the electron


volume element r 2 sin 6 dr dd dj>. Since P nl (r) dr is obtained from
quantity by integrating over the entire range of 6 and
Pn i(r) dr
2,

in the
this
is

<f>,

the probability of finding the electron anywhere with a radial coordinate

between

and

In figure (10-2),

dr.

as a function of rj{a jZ) for the cases n


/

1,

2,

is

plotted in units of Z/a

and

3,

and

all

the possible

Pn i(r)

does

see that the radial probability densities for each eigenstate

have

The radial
quantum number m.

values for these values of n.

not depend on the

We

Pn i(r)

probability density

appreciable values only in reasonably restricted ranges of the r axis.

Thus, when the atom

is

in

an

eigenstate, there is a high probability that

the radial coordinate of the electron will be found within a reasonably


well-defined range;

that

is,

the electron

would quite probably be found

within a certain shell limited by two concentric spherical surfaces centered

ONE-ELECTRON ATOMS

308

A study of these figures will demonstrate that the character-

on the nucleus.

of these shells

istic radii

although there

n,

quantitative

[Ch. 10

is

is

determined primarily by the quantum number


/ dependence.
This can be seen in a more

a small

way by using

the expectation value of the radial coordinate

of the electron to characterize the radius of the

shell.

The expectation value

Wmnf

sin 6

obtained by evaluating the integral

is

Too

*\i

/*oo

rP nl (r) dr
Jo

result

Civ

y*nlm

Jo Jo

Jo

dr dd

d<f>

is

r nl

The

r-ir

values of

apparent that

Kl

^{l+J

+
n

1)1

(10-44)

T^ are indicated in figure (10-2) with small arrows.


T^ depends primarily on n, since the / dependence is
l

It is

sup-

pressed by the factor J and the factor l/ 2 It is interesting to compare this


expression with equation (5-12), which gives the radii of the circular
.

orbits of a

Bohr atom. The equation


r
'Bohr

is

n2a

Quantum mechanics shows that the T\ are of approximately the same size,
and have approximately the same n dependence, as the radii of the circular
Bohr orbits.
x

All electrons in eigenstates with


radial probability densities

common

n values have roughly similar

and approximately the same expectation value

for the radial coordinate, independent of the values of

to the

commonly used

or m. According

terminology, such electrons are said to be in the

Each shell also has the property that the associated eigenhave exactly the same value En
Some insight into the reason why the eigenvalues depend only upon n
may be obtained by computing the expectation value of the potential
energy of the atom. This is
same

values

shell. f
all

y=\
Jo Jo Jo

)y

Wtimi
\

nlm r sin 6 drddd<]>

-zWi

Evaluating (1/r) leads to the result


,Ze

t Since at present we are considering one-electron atoms, this terminology refers to


different atoms in each of which the electron is in the same shell. However,

two or more

is also used for multi-electron atoms


be in the same shell.

this

terminology

can

literally

in

which two or more electrons

AND

EIGENFUNCTIONS

Sec. 5]

Since this does not depend

PROBABILITY DENSITIES

on / or m, the expectation value of the

309
potential

energy,

/rV
does not
because

This is reflected in the behavior of the total energy E,


a characteristic property of any potential of the form
that the total energy of a particle moving in the potential is

either.
it

V(r) oc \\r

is

always equal to exactly one-half

we

Finally,

call attention to

its

average potential energy.

a point which will later be seen to be of

considerable importance.

Inspection of the eigenfunctions (10-39) will

verify that for values of r

which are small compared to a [Z (where the

exponential term
functions

is

slowly varying), the radial dependence of the eigen-

is

Vmm^r

r^O

1
,

(1(M5)

Therefore the radial dependence of the probability density for small r

VtlmVnlmCCr

As a consequence,

21
,

-*

is

(10-45')

f* lm f nlm in a volume element near r = is


and decreases very rapidly with increasing /.

the value of

relatively large only for

0,

Consider again the probability density,

From

equation (10-21),

we have

Thus the probability density does not depend on the coordinate

<j>.

The

three dimensional behavior of y)* lm ip n i m is therefore completely specified


by the product of the quantity R*i(r) Rnl {r)
Pnl (r)/r 2 and the quantity

m (0)

factor.

lm (0), which plays the role of a directionally dependent modulation


The form of 0?m (0) ;m (0) is conveniently presented in terms of a

polar diagram as
the point

shown

The origin of the diagram is at


and the z axis is taken along the direction
measured. The distance from the origin to the

in figure (10-3).

(the nucleus),

from which the angle d is


curve, measured at an angle 6, is equal to the value of @?ro (0) &lm (0) for
that angle. Such a diagram can also be thought of as representing the
complete directional dependence of if*i m if nlm by visualizing the three
dimensional surface obtained by rotating the diagram about the z axis
through the 360 range of the angle
The distance, measured in the
direction specified by the angles 6 and
from the origin to a point on the
surface, is equal to 0?m (0) !m (0) 0*(<) O m (> for those values of 6 and
<f>.
</>

<f>.

ONE-ELECTRON ATOMS

310

In figure (10-4)

we

illustrate

form of the function *m (Q)

Figure 10-3.

a typical example of the dependence of the

!m (0) on

1=3, and

of polar diagrams for

[Ch. 10

the

quantum number m, by a

the seven possible values of

set

for this

polar diagram of the directionally dependent modulation factor of the

probability density.

,,Z

3,

Figure 10-4.

m=
Polar diagrams of the directionally dependent modulation factors of the
From L. Pauling and E. B. Wilson,
I. 2, 3.
/
0,
3; m

probability densities for

Introduction to

Quantum Mechanics, McGraw-Hill Book Co.,

value of

i.e.,

/,

\m\

0, 1, 2, 3.

Note

the

way

New
in

York, 1935.

which the region of

?m 0j m (and therefore of vJ, m VniJ snifts from the z


axis to the plane perpendicular to the z axis as the absolute value of
increases. Some important features of the dependence of 0*m (0) ; m (0)
concentration of

AND

EIGENFUNCTIONS

Sec. 5]

PROBABILITY DENSITIES

on the quantum number / can be indicated by presenting a


plots with \m\ = / and / = 0, 1, 2, 3, 4, as illustrated in figure

=m=

311

set

of polar

(10-5).

For

which is the ground state of the


and the probability density is
atom, iptimWmm depends on neither 6 nor
spherically symmetric For other states, the concentration of probability
density in the plane perpendicular to the z axis, when \m\ = I, becomes more
the eigenstate with n

1,

0,

<f>

and more pronounced with increasing

Figure 10-5.

m=

= o,

3,

The

set

1,

m=

+3

= 4,

2,

m=

m=

+4

Polar diagrams of the directionally dependent modulation factors of the

probability densities for


Introduction to

m=

/.

0,

I,

2, 3, 4;

\m\

/.

From

Quantum Mechanics, McGraw-Hill Book Co.,

L.

Pauling and

New

E. B.

Wilson,

York, 1935.

of quantum numbers nlm determines the complete three

dimensional behavior of y* im y n;m with the behavior as a function of the


,

direction relative to the z axis depending only


set

of quantum numbers,

and *m (0)

!m (0)

R*

{r)

R n i(r)

on

has zeros for several values of

density for an electron

and m. For a

typical

has zeros for several values of

bound in the Coulomb


and conical nodal

6.

r,

Thus the probability

potential of a nucleus has,

on which its value


These surfaces are analogous to the nodal points in the probability
density of a particle bound in a one dimensional potential, and they are a
consequence of the fact that the wave functions for a bound particle must
be standing waves. Of course, it is not possible to perform a sequence of
measurements on the location of an electron in a particular eigenstate
with quantum numbers nlm, and thereby locate the nodal surfaces and
in general, a set of spherical
is

surfaces

zero.

determine the direction of the

z axis.

This would amount to finding a

preferred direction in a space which should be spherically symmetric

because the potential V = Ze?/r is spherically symmetric. It cannot be


done because any eigenstate, except the spherically symmetric eigenstate

1,

=m=

0, is

degenerate with several other eigenstates with the

ONE-ELECTRON ATOMS

312

[Ch. 10

same value of n and so cannot be investigated separately. All that can be


measured experimentally is the average probability density for the whole
set of states which are degenerate with each other. For instance, it would
be possible to make measurements on the probability density for atoms
whose energy is E2 The results would be
.

il>200V200

ftl -1V21-1

= T^(-)3e
12877 \a

" Zr/a

r(
L \

VzloWilO

- "I
+
a /

V21lV21l]

(%
a

sin2 6

* sin2 d

cos2 e )

(10-46)
128t;
2877

\a /

This is a spherically symmetric distribution, and it cannot be used to


determine the direction of the z axis. Thus there is no contradiction of the
fact that this direction

On

was

initially

chosen in a completely arbitrary way.

the other hand, consider a situation in which the orientation of the

z axis is

not arbitrary because there

is

a preferred direction defined by an

external magnetic or electric field applied to the atom.


as

we

shall see later, the eigenstates are

In such a field,
not degenerate and measurements

of the probability density of a particular eigenstate could be performed.


In fact, such measurements could be used to determine the direction of the
external

We

field.

shall

show

related to the

and to

its z

quantum numbers

shortly that the

component

L by
z,

and

are also

momentum L of the electron,

magnitude of the orbital angular


the equations

L = J 1(1 + l)h
(10-47)

L z = mh

We mention this now because it is

an important clue to the interpretation


I an d m.
Consider the case m = I.

of the dependence of wtimWnim on

Then L z

angular

momentum

Ih,

which

is

almost equal to

For a Bohr atom

axis.

would

lie

L=

Vl(l

1) h.

In this case the

vector must point nearly in the direction of the z


this

would mean that the

orbit of the electron

nearly in the plane perpendicular to the z axis.

With increasing

> Vl(l

values of/, lh
+ 1) h and L z -^ L, so the angle between the angular
momentum vector and the z axis decreases. In terms of the Bohr picture,
this

demands

the z axis.

that the orbit

An

lie

more nearly

in the plane perpendicular to

inspection of the polar diagrams will demonstrate the

correspondence between these features of fnimWnim an ^ tne picture of a


Bohr orbit. For
have L z = 0, and the angular momentum

m=0we

ANGULAR MOMENTUM OPERATORS

Sec. 6]

In a Bohr

vector must be perpendicular to the z axis.

mean

313

atom

Some

that the plane of the orbit contained the z axis.

would

this

indication of

behavior can be seen in the polar diagram for / = 3, m = 0.


Although there are many points at which the quantum mechanical
theory of the one-electron atom corresponds quite closely to the Bohr
theory, there are certain striking differences. In both theories the ground
state corresponds to the quantum number n = 1
But in the Bohr theory
the orbital angular momentum for this state is L = nh = h, whereas in
this

quantum mechanics L

V/(/

1)

since

when n

1.

There is an overwhelming accumulation of experimental evidence which


shows that the quantum mechanical prediction is the correct one. It
would be very difficult to modify the Bohr theory in a way which would

momentum

allow for zero angular

such a state

states, since the orbit for

would be a radial oscillation passing directly through the nucleus. In


quantum mechanics the electron does not move in a well-defined orbit, and
so no problem arises.

6.

Angular

Momentum Operators

In the next sections

we

shall discuss the orbital angular

of the eigenstates of a one-electron atom.

We

shall start

lating the expectation values for this quantity, but

possible to obtain a very

much more

we

the theory of

shall find that

precise evaluation.

momentum

In classical mechanics the angular

the origin of a certain coordinate system,

is

it is

In the course of

new and broader insight into the general


quantum mechanics will be obtained.

this discussion a

momentum

by simply calcu-

properties of

of a particle, relative to

a vector quantity

which

is

given by the equation

L=
The vector

r is the position

vector p is the linear


in figure (10-6).

vector

is

of the particle relative to the origin, and the

The

"cross product"

do the

x, y, z

r, p,

pis defined to mean that the


p, and
have the same relative spatial

and L

is

defined so that the magnitude of the

is

L=
r

rX

axes of a right-handed coordinate system.

Furthermore, the cross product

where

(10-48)

perpendicular to the plane formed by the vectors r and

orientation as

momentum of the particle. These vectors are indicated

points in the direction so that

vector

rp sin 6

and p are the magnitudes of

and

(10-49)
p,

and

is

the angle between

ONE-ELECTRON ATOMS

314
these

two

From the

vectors.

three rectangular

definition of r

components of

p, it is

easy to

show

x, y, z are the

that the

are

Lx = yp z - zpv
Lv = zPx xPz
Lz = XPV - VPx
where
of p.

[Ch. 10

components of

r,

(10-50)

and px p y p % are the components


,

*-y

Figure 10-6.

moving

The

position, linear

momentum, and

angular

momentum

vectors of a

particle.

In order to discuss the dynamical quantity angular

momentum

in

quantum mechanics, we construct the associated operators by replacing


px py p z with their quantum mechanical equivalents ih djdx, ih djdy,
ihdjdz (cf. 7-77). Thus the operators for the three components of
,

angular

momentum

are

"*

LVos

= in\z

L,

= in\x

V
It is
r, d,

By/

dz

dx

(10-51)

dz!

y
dy

dx)

desirable to transform these expressions into the spherical coordinates


(f>.

When

this is

L*

done,

it is

= ^( sin

found that

4>

jig

l.op = ^(-cos $
L ** = - ih

cot 6 cos

^ + cot

4>

sin

^7/
4>

2\

(10-52)

ANGULAR MOMENTUM OPERATORS

Sec. 6]

We

shall also

L, which

315

be interested in the square of the magnitude of the vector

is

L2 =

I?B

4 + L\

(10-53)

In spherical coordinates the associated operator


a (

-sin0 30

sin
s

dd!

is

'

sin

(10-54)

2
0d< J

Let us calculate the expectation values of the components of L, and of


the square of

its

magnitude, for an electron in the eigenstate nlm of a

According to equation (7-80), extended to three

one-electron atom.f

dimensions, they are given by

-J7J

-///e

Tr2 sin

1"

6 dr dd dd>

iE tl!i
iE tl!i y>*
rZ sin 6 dr
y, nlm
nlm L Xo e-

dd

d<f>

So

Lx

=
J J J

2
'*
v imL *op^''" sin 6 dr dd d ^

Similarly,
L~v

2
vti m Ly 09 W n m r sin 6 dr dd dj>
i

J J J

-J7M

nL z

-HI*

w nlm r2

L2oV Wm m r2

sin 6 dr

sin 6 dr

dd

dd

(10-55)

d<f>

dcf>

where the integrations are over the entire range of the variables.
evaluate these integrals, we must first evaluate the quantities

L Xov y> n m
i

In doing

this,

we

Consider

first

the quantity

LVop ip nlm

L^rpnim,

shall find that the last

Lt Zov ,
y>ni m

L,

To

ov %p nlm

two lead to an

interesting result.

dyw
= ~ lh ~^T
.-fc

of a
t Strictly speaking, we shall calculate the components and squared magnitude
which is the orbital angular momentum of the electron plus the orbital angular

vector

of the nucleus, relative to the center of mass of the atom. This is because
y> nlm actually describe the motion of an electron of reduced mass n,
which is moving about a fixed nucleus in such a way that the orbital angular momentum
of the reduced mass electron is equal to the total orbital angular momentum of the atom.

momentum

the eigenfunctions

Cf. section 2.

ONE-ELECTRON ATOMS

316

[Ch. 10

Since

Vnlm

Rnl(r) l m (Q)

m(<t>)

we have

- ih dJ = R

nl (r)

lm (Q)

-ih

d m (4>)
dcj>

According to equation (10-21),


0>

m (<)

im*

So

dO n (4>)

ime

im<D m (>

dcp

Thus
-fc

Pynlm
d<j>

:mM

Bj

(r)0 ;m(0)O m()

and

Lz

nr>

V>nlm

mkip nl

(10-56)

Next evaluate
~

= R nl (r)(-h

L%y> nlm

2
)

m (4>) d_( sin


sin9 d6\

Differentiating equation (10-21),

dQtJO) )

dd

i m (8) d

sin

(<f>)

dtf

we have

= -m m
2

(<f>)

d<f>

So

L 2 p y> nlm = R nl(r)<& m(<l>)(-ti )

Now

This

dO

is

(6)

dd

a solution to the

0,
I

sin

differential

m]

equation

is

d (

Lsin

a function which

Jm (0) is

(10-18).

dL ne d& lm

sin

d lm (6)\

sindddX
sin
dV \

dd

sin

lm(0) =-/(/+

l) m (0)

where we have evaluated a from (10-24). Therefore


8

i< opVi

= RnMQj&X-Pti-Ki + i)0(fl)]
= /(Z+l)A w! (r)0, m(0)(D m (^
2

and we have
i?opVi

Kl

(10-57)

l)*Vi

we see that LZo y> nlm is equal to mhy nlm Thus the
of operating on the one-electron eigenfunction y nlm with the
is simply to multiply that eigenfunction by the
differential operator L
In equation (10-56),

result

2op

ANGULAR MOMENTUM OPERATORS

Sec. 6]

317

constant mh. Similarly, equation (10-57) shows that the result of operating
on the one-electron eigenfunction y> n[m with the differential operator L\
v

is

multiply that eigenfunction by the constant

to

results are certainly

/(/

\)h 2

These

not typical of the behavior of differential operators.

we operate on the function /(x) = x2 with the differential


operator djdx, we obtain a very different function/'(z) = 2x. As another
example, the reader may verify that the result of operating on y nlm with
the operators Lx
or L Vo is to produce new functions of r, 6,
in which
these variables enter in a different manner from the way they enter in the
For

instance, if

cj>

function

That

y> nlm .

is,

L Xop y>nim

Lv ot,Wmm

# (constant)y) Blm
^ (constant)^

Using equations (10-56) and (10-57),

L2

from equations

(10-58)

trivial to

it is

evaluate

T and
z

These values are

(10-55).

n^z^Wmmf
//J' nlrn^Zo

Lz = mh\

v*nlm y nlm r

sin 6

dr dd

d<f>

sin 6

dr dd

d<f>

T = mh

(10-59)

and
2

=
J J J

ftinfiovVnimr* sin 6 dr dd

L2 =l(l +

l)k

!?=/(/+

l)h

jjjy>*nlmWnlm r

d<f>

sin d

dr dd

d<f>

(10-60)

where we have employed the normalization condition (10-41). Let us


compare these results with the relations (10-47) which were stated without
proof in the previous section. These relations are

Lz = mh
L =
2

It is

/(/

(10-61)

\)h

(10-62)

apparent that equations (10-59) and (10-60) are consistent with


do not prove their validity, as the latter pair make

(10-61) and (10-62) but

very

much

stronger statements about the values of

do the former. In order


of equations (10-61) and (10-62), it

eigenstate nlm, than

Lz and

I?, for the

to complete the proof of the

will be necessary to make a


short digression into the general properties of the theory of quantum

validity

mechanics.

ONE-ELECTRON ATOMS

318

[Ch. 10

Eigenvalue Equations

7.

Consider a particle moving under the influence of a time independent

For simplicity, consider the motion to be one dimensional.


Let / be any dynamical quantity useful in describing the motion of the
particle. Then the average value of/, obtained in a series of measurements
made when the particle is in a state of motion described by the wave
function W, is the expectation value /. According to equation (7-80),
potential.

it

has the value

f^yVfeJVdx
=JV
where the integral is taken over the entire range of the variable. The
quantity /provides a partial specification of the behavior of/ To obtain
a more complete specification, we must learn something about the
fluctuations

of/ This can be done by

studying the quantity

A/ 2 which is
,

the expectation value of the square of the difference between the results of

a particular measurement

That

of/ and

the average of these measurements /.

is,

A/ 2 = (/-/)2

(10-63)

It is clear

that information relating to the fluctuations in /is provided

A/

instance, consider a case in

2
.

(e.g.,

For

by

which /can have only a single value


then/ will always be

the total energy of a particle in an eigenstate);

observed to be exactly equal to the value

However, if/

fluctuates

greater than zero.

To

about

evaluate

/,

and (/ ff

its

average value

A/

2
,

/,

will

then (/ /)

we apply equation

be zero.
2

will

be

(7-80) to (10-63)

and obtain

Y^

A/ 2 =jV*[(/-/)2] op
Since

/ is

a constant, the operator [(/

/)

2
] op

is

simply (/op

/)2

we have

A75 =JV*(/op-/fFiz

vp*
'*(i1p-2/op/+/)Ydx

= JV*/VF dx -

2/jVi/opY dx

+ f JV*T dx

and

EIGENVALUE EQUATIONS

Sec. 7]

319

Using equation (7-80) on the first two terms of the right


on the third term, we obtain

side,

and using

the normalization condition

W=P-w+f
or

&p=p-f
Thus the quantity A/2

is

(10-64)

equal to the difference between the expectation

value of the square of/, and the square of the expectation value of/.

Now let us assume that the particle is in a certain eigenstate so that


Yix,

t)

e~ iEi ' n y(x)

Then we have

je

iEt/n

y>*fov e-

iEtin

y>dx

According to equation (7-80), /op never contains derivatives with respect


to time. Consequently the term er iEtltl can pass through op as if it were a
constant, and we have

f=L

iEt!n

e-

iEil 'l

f*fov xp dx

or

f=\f*fo P fdx
Similarly

we can show

(10-65)

that

r*=\xp*flvW dx

(10-66)

when

the particle is in an eigenstate.


For a given dynamical quantity/ and a given eigenfunction

evaluate

A/2 >
A/2 =

0.
0.

A/2 from

the equations above.

In general,

However, as mentioned previously, situations

is

we can

ip,

found that

exist in

which

We shall now show that this happens when the relation between

the dynamical quantity /and the eigenfunction

/opV
where
have

it

F is

a constant.

/=

dx

y)*Fy)

dx

so

f=

such that

Fy>

(10-67)

If this is the case,

ip*fovW

ip is

i72

then in equation (10-65)

=F

rp*y>

dx

=F

we

ONE-ELECTRON ATOMS

320

[Ch. 10

we have

In equation (10-66)

f=

V^fopf dx

W*fopFf dx

f*fopfopf dx

=F

ip*fopf

dx

= FF

so

Then, from equation (10-64), we find

A/1 =/2 - f = F 2 - F 2 =

(10-68)

2
Thus, when the relation (10-67) is true, the quantity A/ which is a measure
of the fluctuations in the dynamical quantity /, is equal to zero. In any
normal case this is adequate to prove that/can have only the single value
,

f= /=

F. In exceptional cases equation (10-68) alone does not constitute

a proof of this statement, but

p -f =
However,

it is

it is

also necessary to
for any integer

0,

easy to see that

when equation

show

that

(10-69)

(10-67)

is

true the steps

immediately preceding equation (10-68) can be repeated k times to show


that the requirement (10-69) is satisfied. Thus we conclude that when
the relation between

f and rp

the dynamical quantity

f can

is

foW =

(10-70)

Fy>

have only the

definite, precise value

f=F

(10-70')

In the mathematical literature an equation of the form (10-70) is


an eigenvalue equation, the function y> is called an eigenfunction of

called

the operator fop,

and the constant Fis

called

an eigenvalue of the operator.

We have come across most of this terminology before in connection with a


Consider the time independent Schroe-

particular eigenvalue equation.

dinger equation (10-8), and rewrite

it

fc2
"J

i=i

Comparing the

2m,-

curly bracket with the classical definition of the total energy

of the system (10-1),

we

see that this equation

e op y

Ey>

may be

written

(10-71)

EIGENVALUE EQUATIONS

Sec. 7]

321

where is a constant numerically equal to the total energy of the system,


and where the total energy operator e op is obtained from the corresponding
dynamical quantity by replacing^, with ihdjdxj, etc. This operator
is often called the Hamiltonian.
For one particle in one dimension it is
simply
e ov

+ V
2m i
h2

= -

d*

T/

dx'

In the process of studying the particular eigenvalue equation (10-71),

we have

learned

common to
y>

and

its

finiteness,

all

much about mathematical

properties which are actually

such equations. For instance,

we know that,

and single-valuedness, there

will

(10-71) only for a certain set of eigenvalues


eigenfunctions

rp

x,

ip

ip 3 ,

2,

eigenvalue equation (10-70).


(10-71),

if

the function

derivative are required to satisfy conditions of continuity,

we concluded

ip

n,

be solutions to equation

Ex E2 E3
,

The same

....

is

total energy are the eigenvalues

of the subject,

is

En

and

In interpreting the significance of equation

operator appears in the equation.

we

that the set of eigenvalues comprise the only

physically possible values of the dynamical quantity

eigenvalue equation,

true of the general

(That

Ex E

2,

is,

.)

whose corresponding

the allowed values of the

If we

extend this to the general

obtain a statement which, in advanced treatments

usually taken as the fundamental postulate of the theory

of quantum mechanics:

A measurement of the value of a dynamical quantity f can yield only one


of the eigenvalues Fn of the corresponding operator /op where fop is obtained
from f by replacing x with ih djdx, etc., and where the relation between
,

Fn and fop

is

fovVn
in

which the eigenfunction

and

continuity, finiteness,

ip

= Fn y n

and

its

derivative satisfy

conditions of

single-valuedness.

This one-sentence postulate, and the postulates which provide a physical


interpretation

of the eigenfunctions, contain the complete theory of

quantum mechanics

for time independent

phenomena.

The theory

is

extended to include time dependent phenomena by introducing the


operator

ih djdt.

We may illustrate the theory developed in this


examples.

Consider

first

the traveling
y>

We know

that this

is

section with

some simple

wave eigenfunction
iKx

an eigenfunction of

(10-72)
e op , as the function (10-72)

ONE-ELECTRON ATOMS

322

was obtained by solving (10-71)


is

V=

for the case

associated with this eigenfunction

is

E=

h2

0,

and that the eigenvalue

K2 j2m.

The

function (10-72)

an eigenfunction of the momentum operator pop

also

[Ch. 10

= ih d/dx

since

P opW

ih

dx

V>

ih

dx

iKx

= ihiKeiKx = hKy

or

= pf

PotW

where the eigenvalue is P = hK. We see that this function is simultaneously


an eigenfunction of two operators. Therefore, according to equation
(10-70'), a particle whose motion is described by this function must have a
definite value of total energy E = h 2 K 2 \2m and a definite value of momen-

tum

function

E=

hK.

y>

K 2 /2m, and

h2

value

we may show

Similarly,

that the traveling

wave

eigen-

= e~~iKx is an eigenfunction of eop associated with the eigenvalue

P = hK.

also

is

function has the definite


definite value

an eigenfunction of pov associated with the eigenis described by this


value E = h2 K 2 j2m for its total energy, and the

Therefore a particle whose motion

P = hK for

its

momentum.

Next consider the standing wave eigenfunction

We know

that this

eigenvalue

E=

h2

is

also

K /2m.
2

sin

Kx

(10-73)

an eigenfunction of e p associated with the


However, it is not an eigenfunction of pop

since
Por,y>

ih

sm Kx =
dx

Although the energy of a

particle

function has the definite value


value.

ihK cos Kx ^ (constant)y

E=

h2

whose motion

is

described by this

K2/2m, the momentum has no definite

According to the postulate stated above, a measurement of the

momentum of the particle can yield only one of the eigenvalues of the
momentum operator. For a given value K there are just two eigenvalues of
= hK. A measurement of the momentum of
jp op namely P = hK and P
,

the particle

would

yield either of these

two values and,

in a large

number

of measurements, they would occur with equal probability.! It is evident


that these properties of the standing wave eigenfunctions and traveling
wave eigenfunctions are in complete agreement with the interpretation
presented in the
t This

may

is

verify.

first

section of Chapter

a consequence of the fact that p

8.

for this eigenfunction, as the reader

ANGULAR MOMENTUM OF ONE-ELECTRON ATOMS

Sec. 8]

Angular

8.

Momentum

323

of One-Electron-Atom Eigenfunctions

We return now to the functions y nlm which arise in treating the problem
of a one-electron atom.

Being solutions to the time independent Schroe-

dinger equation for this problem, the

y nlm

are necessarily eigenfunctions

of the total energy operator. According to equation (10-38), the associated


eigenvalues are

En = aZ 2 e4/2^2 2 From
.

we
2

equations (10-56) and (10-57),

see that these functions are also eigenfunctions of the operators for the

component of the

orbital angular

momentum L Zo

and for

its

square

L\ v with associated eigenvalues mh and /(/ + l)/j2 However, equations


(1058) state that they are not eigenfunctions of the operators for the
x and y components of orbital angular momentum. Also, it is easy to
show that they are not eigenfunctions of any of the linear momentum
.

operators.

Therefore,

the total energy of the


orbital angular

when an
atom is

momentum

is

electron
definitely

its z

component

is

by y n!m
the magnitude of the total

in a state described

En

Vl(l

1)

(10-75)

is

no

(10-74)

definitely

L = mh
but there

definitely

L=
and

is

definite value of the

x or y components of

orbital angular

momentum, or of any component of linear momentum.


The fact that rp nlm does not describe a state of definite linear momentum
any physical picture of the motion of an
and z component of orbital
angular momentum calls to mind some sort of rotation in which linear
momentum could not be constant. However, the fact that y nlrn does not
describe a state with a definite x and y component of orbital angular
is

certainly reasonable, since

electron in a state having a definite magnitude

momentum

somewhat mysterious because, in classical mechanics, the


momentum vector of an isolated system moving under the
influence of a spherically symmetrical potential V(r) would be fixed in
space, and all three components would have definite values since there
would be no torques acting on such a system. The quantum mechanical
result is a manifestation of an uncertainty principle relation which can be
shown to exist between any two components of orbital angular momentum.
Because the z component of orbital angular momentum has the precise
value mh, this relation requires that the values of the x and y components
be indefinite. Upon evaluating J7X and L~ it is found that both of these
v
quantities are equal to zero. Thus the orientation of the orbital angular
is

orbital angular

ONE-ELECTRON ATOMS

324

[Ch. 10

momentum

vector must be constantly changing in such a way that its


x and y components rapidly fluctuate about an average value of zero.
Many of the properties of the orbital angular momentum can be conveniently represented in terms of vector diagrams. Consider the set of
states

having a

common value of the quantum number For each of these


momentum vector, in units of
/.

states the length of the orbital angular

It,

= V + 1). In the same units, the z component of this vector is


LJh = m which, according to equations (10-37), may assume any integral
value from Ljh = to LJh = + depending on the value of m. The
is

L\h

/(/

1,

Figure 10-7. Illustrating the angular


the five possible states for /
2.

momentum

vectors precessing about the z axis in

case of

momentum
values of

is illustrated in figure (10-7). The figure depicts the angular


vectors of the five states corresponding to the five possible
for this value of /. Since L x and L fluctuate about their average

values of zero, these vectors must precess randomly in the conical surfaces
surrounding the z axis. The actual spatial orientation of the angular
momentum vector is known with the greatest precision for the states with

m=

1;

but even for these states there

is

some uncertainty

since the

vector can be anywhere on a cone of half-angle cos-^ljV 1(1


In
1)].
the classical limit / -> oo, and this angle becomes vanishingly small.

Thus, in

this limit, the

constrained to

angular

momentum

vector for the states

m = 1

almost along the z axis and is therefore essentially


fixed in space in agreement with the behavior predicted by the classical
is

lie

theory.

Since

Lz = mh

was chosen
component

with

arbitrarily,

m
it

an

integer,

and

follows that

since the direction of the z axis

quantum mechanics

requires the

any direction of the orbital angular momentum to be an


integral multiple of h. This is known as spatial quantization. If the reader
will review the derivation of the eigenvalue equations (10-56) and (10-57),
in

Ch.

EXERCISES

10]

325

will be evident that both spatial quantization and the quantization of the
magnitude of the orbital angular momentum according to the rule
it

L = V /(/ +

1)

h occur not only for the

Coulomb

potential V(r)

= Ze

but for any potential which depends only on the radial coordinate

/r,

r.

BIBLIOGRAPHY
Leighton, R. B., Principles of Modern Physics, McGraw-Hill

Book

New

Co.,

York, 1959.
Pauling, L.,
Hill

Book

Schiff, L.

I.,

and E. B. Wilson, Introduction


Co.,

New

to

Quantum Mechanics, McGraw-

York, 1935.

Quantum Mechanics, McGraw-Hill Book Co.,

New York,

1955.

EXERCISES
1

one
2.

Write an expression for the probability flux vector for a system containing
a three dimensional space.

particle in

Separate (10-10) into an equation in the rectangular coordinates of the

center of

mass of the atom,

x, y, z, defined

by (10-11), and an equation

in the

rectangular coordinates of the electron relative to the nucleus, X, Y, Z, defined

by

X = x - xu Y = y - yu Z = z - z v
2

3.

Transform the equation

and thereby

in X, Y,

Hint:

verify (10-15).

Z'mto an equation in the r, 6, j> of (10-12),


Consult some reference on intermediate

calculus for an explanation of the use of Jacobians.


4.

Solve equation (10-23) by the procedure indicated in the paragraph

containing that equation.


5.

Solve equation (10-31) by the procedure indicated in the paragraph

containing that equation.

Consider a rigid rod of length R and negligible mass, with one end fixed
way that it can rotate freely and the other end carrying
a particle of mass M. Adapt the time independent Schroedinger equation (10-1 5)
to this system. Solve the equation and show that the eigenvalues are
6.

at the origin in such a

hH{l

1)
I

2/

where /

MR

is

the

moment

0, 1, 2, 3,

of inertia about whatever axis the rod

may

Note that there is no zero point energy for this system because all the
coordinates are free to assume all possible values. Compare these results with
rotate.

the results of exercise

is

3,

Chapter

5.

Evaluate the expression analogous to (10-46) for n


spherically symmetric.
7.

3,

and show that

it

ONE-ELECTRON ATOMS

326
Verify equations (10-50).

8.

Use some of the equations obtained in exercise

9.

and

3 to verify equations (10-52)

(10-54).

10.

on

[Ch. 10

their values.

and Ap 2 for the eigenfunction y


Hint: Use box normalization.

Evaluate/,

sin

Kx, and comment

II

CHAPTER

Magnetic Moments,
Spin, and
Relativistic Effects

Orbital Magnetic

I.

In the

first

Moments

part of this chapter

we

shall discuss

some experiments which

provide very convincing evidence for spatial quantization.

momentum

related quantity

we

Actually

measure the z component of the angular


vector L, but instead measure the z component of a closely

the experiments

do not
(t,

directly

the magnetic

moment vector. To begin the discussion,


moment (*j of an atomic electron.

shall evaluate the orbital magnetic

This will be done by using a combination of simple electromagnetic theory,


the Bohr theory, and quantum mechanics. A completely quantum mechan-

not be presented because such a treatment would


more advanced knowledge of electromagnetic theory than has

ical calculation will

require a

been assumed in this book, and also because, owing to the conceptual
simplicity of the description of the atom provided by the Bohr theory,
there are advantages in using that theory when introducing something
new. Of course, the Bohr theory does not always yield quantitatively
correct results but for the calculations we shall perform here it does.
Consider, then, an electron of mass m and charge e moving with

Bohr orbit of radius


loop constitutes a current of magnitude
velocity v in a circular

r.

e
= -f~
r
c 2nr
c r

e
-

This charge circulating in a

327

(11-D

MAGNETIC MOMENTS,

328

RELATIVISTIC EFFECTS

SPIN,

[Ch.

where r

is the orbital period of the electron.


It is well known from
elementary electromagnetic theory that such a current loop produces a
magnetic field which is the same at large distances as that of a magnetic

dipole located at the center of the loop and oriented perpendicular to its
For a current / in a loop of area A, the magnitude of the magnetic

plane.

moment

of the equivalent dipole

is

P-i

iA

(11-2)

L
Figure

l-l.

The

quantities associated with an electron

The magnetic moment is a measure of the


dipole and is defined as

moving

in a circular

Bohr orbit

strength and orientation of the

V-^qh

(11-3)

where q

is the magnetic strength of each of the poles and 8 is a vector


extending from the negative (south) pole to the positive (north) pole.
For a current loop produced by a negatively charged particle, the direction
of {Aj is opposite the direction of the angular momentum vector L, whose

orientation and magnitude are given by the equation

L=

rXp = wrXv

(11-4)

These quantities are illustrated in figure (11-1). Evaluating / from equation


(11-1) and A for the circular Bohr orbit, we find from (11-2) that the
magnitude of the magnetic moment is

Vi

e
~

w =
v

evr

,.

(11-5)

2c

c 2-nr

Since r and v are perpendicular in a circular orbit, the magnitude of the

angular

momentum

is

simply

L=

mrv

(1

1-6)

Sec.

ORBITAL MAGNETIC MOMENTS

I]

From

these

magnetic

two equations we

329

see that the ratio

of the magnitude of the

moment to the magnitude of the angular momentum is a constant,

This constant

fi t

evr

2cmrv

(11-7)
line

usually written gifi B jfi, where

is

[i

ehj2mc

0.927

X lO" 20 ergs/gauss

(11-8)

and

gi=l

(H-9)

The quantity (iu forms a natural unit for the measurement of atomic
magnetic moments and is called the Bohr magneton. The quantity g
is called the orbital
g factor. It is introduced, even though it appears
redundant, because we shall later come across a g factor which is not unity.
t

In terms of these quantities,

we may

rewrite equation (11-7) as a vector

equation specifying both the magnitude of

(x,

and

its

orientation relative

to L:

|t,=

-^L

(11-10)

The

ratio of

fi

an

L does not depend on the size of the orbit or on the


By making a calculation similar to the one above for
it can be shown that (i \L is independent of the shape

to

orbital frequency.
elliptical orbit,

That

of the orbit.

completely independent of the details of

this ratio is

the orbit suggests that

its

value might not depend on the details of the

mechanical theory used to evaluate


evaluating

dividing by the
ratio of

(i l

to

it,

and

this is actually the case.

quantum mechanically (which

fx l

is

will

Upon

not be done here), and

quantum mechanical expression L = Vl(l+ 1) ft,


found to have the same value that we obtained from

the
the

Bohr theory. Granting this, the reader will accept that the correct
quantum mechanical expressions for the magnitude and the z component
of the magnetic

(i x

moment

vector are

^L=_ ^
gl/*B T

SiPb

V/(/

+ \)h= g^v-JIV +

1)

(H-ll)

and

K.

= -

^L = - S-^mh = 2

gl!a

Bm

(The minus sign follows from the minus sign in equation 11-10.)

(11-12)

MAGNETIC MOMENTS,

330

2. Effects

SPIN,

RELATIVISTIC EFFECTS

[Ch.

of an External Magnetic Field

Using the definition


with magnetic

(1 1-3), let

moment

us evaluate the forces acting on a system

placed in an external magnetic

field H, which
uniform in the region of the equivalent dipole. This is illustrated in
figure (11-2). In this field there is a force parallel to the direction of
and of magnitude qH acting on the positive pole, and a force of the same
|jl,

is

H
Figure

1-2.

magnetic

moment

in a

uniform external magnetic

field.

magnitude, acting in the opposite direction on the negative pole.


is no net translational force, but there is a torque of magnitude

T = qH - sin
where
to both

There

+ qH - sin = qd H sin = ^H sin 6

|ij

[L and H. The torque vector is perpendicular


and H, and it points in the direction which would tend to align
with H. Its magnitude and direction can thus be expressed by the

the angle between

is
(jl

equation

T=fijXH

(11-13)

H, the system with magnetic moment jx,;


AE which depends on the angle
The energy A.E can be evaluated by summing the

In an external magnetic

field

has an orientational potential energy

between the two vectors.


potential energy of each of the poles of the dipole, if we define the zero of
potential energy to be at its center. For a magnetic field which is uniform
in the region of the dipole, this

is

AE = -zqH - (~z)(-q)H = -2 - cos qH = -^Hcos

(11-14)

EFFECTS OF

Sec. 2]

where

z is the

AN EXTERNAL MAGNETIC

coordinate of the positive pole, and

FIELD

is

331

the coordinate

of the negative pole, both measured along an axis parallel to H.

can be written in terms of the "dot product" notation

AE=
Equation

(1

-|a,-H

(11-15)

1-14) can be taken as the definition of this notation.

AE

orientational potential energy

minimized when

is

This

as

The

and H are
mechanism, the
(a (

However, in the absence of some dissipative


cannot obey the natural tendency to align itself with the vector
H. Instead, (i will precess around H in such a way that 6 and A remain
aligned.

vector

(jL

constant.

This precessional motion

is

a consequence of the fact that the torque

is

always perpendicular to the angular

momentum

vector

since,

from

we have

equations (11-10) and (11-13),

_^ LXH

T=

h
Substituting this into Newton's law of motion for angular

momentum,

dt

we

obtain the equation of motion of the vector L:

-=-^LXH
From

the equation

we

(11-16)

dt
see that the

change

in

in the time dt

is

dL= _?LxH<il
h

These vectors are illustrated in figure (ll-3).f Since dL is perpendicular


to L, its magnitude L remains constant. In the time dt the tip of the vector
L moves through an angle

dL

gll*B

L sin
L, and also
magnitude

Thus

|a ; ,

LH sin 6 dt _

HL sin

precess about

(11-1),

with an angular frequency of

H = b h
H

dt
t This figure, which shows

H dt

gl!*B

L precessing about H,

which shows the electron circulating

in

should not be confused with figure

its orbit.

MAGNETIC MOMENTS,

332

As

the sense of the precession

is

RELATIVISTIC EFFECTS

SPfN,

in the direction of

[Ch.

H, we may write

Wi = &H

(11-17)

phenomenon is known as the Larmor precession, and w L is called the


Larmor frequency. It is quite analogous to the precession of the spin
This

Figure 11-3.

lustrating a magnetic

moment

precessing about an external magnetic

field.

angular
(11-16)

momentum
is

of a top in a gravitational

identical in

field.

form with the equation of a

An experiment capable

In fact, equation

top.

of detecting spatial quantization of the magnetic

moment of an atom must be able to measure /u, or some quantity which


depends on /x, Equation (11-15) suggests placing the atom in a magnetic
field oriented along the direction of the z axis and measuring the potential
energy A =
H = ^ H by observing, for instance, the effect of
the magnetic field on the optical spectrum of the atom. Such an experiment
will be described in Chapter 13 under the heading Zeeman effect. Here
we shall consider instead a different technique which involves sending a
,

(x,

AN EXTERNAL MAGNETIC

EFFECTS OF

Sec. 2]

beam of atoms through a magnetic


amount which depends on fi

FIELD

field that deflects

333

each atom by an

We

have seen above that in a uniform magnetic field there is no net


on the magnetic moment. However, if the
is not uniform there will be a translational force. Consider a magnetic
which is non-uniform, but does not vary with y and is symmetric

translational force acting


field
field

with respect to the yz plane.f In the yz plane the

field

then has a com-

ponent only along the z axis. Assume that a magnetic dipole producing
a magnetic moment (x is straddling the yz plane with \l at an angle 6

paper

axis into

-<7

Figure

1-4.

magnetic

moment

a non-uniform external magnetic field.

in

from the

z axis, as indicated in figure (1 1-4).

the field

is

which

is

negative poles the field in the z direction

,
(H z) +q
x

(H z )_ g

where x

(<5/2) sin

At

the center of the dipole

entirely in the z direction.

,,
=H
+
.

oH,

^ x

is

oH

^
z

+
,

ox

At

the positive

and

approximately
z

oz

d
d
= H + -^(-x)
+ -^(-z)

ox

and

oz

cos

(<5/2)

6.

Thus

dH,
Fz = q(H z) +a - q{H z )_ q = -?qd sin

dx

the net z force

is

dH,
+ ?qd cos
dz

or

dH z
Since

\}. l

is

not fixed in space, we must evaluate the average of

dH

ox
t

Such a

dH z

field is

shown

in figure (11-5).

dU z
oz

z.

It is

MAGNETIC MOMENTS,

334

In the classical theory

quantum mechanics
theories /If

RELATIVISTIC EFFECTS

= because of
= {g^^h) Lx =

7*7"

p~i

since

[i t

SPIN,

it

the

Larmor

because

0.

In

precession.

Lx =

has a fixed value. Consequently

[Ch.

In both

we have

- dH*u

(11-18)

similar calculation shows the average net x component of force to be


F~x = (dHjdz) n but this vanishes because dHJdz = near the yz plane
owing to the symmetry of the field.
%

The Stern-Gerlach Experiment and Electron Spin

3.

In 1922 Stern and Gerlach measured the possible values of /i^ for Ag
atoms by sending a beam of these atoms through a non-uniform magnetic

Figure

1-5.

An

apparatus used to measure the z component of the magnetic

moment

of an atom.

A schematic diagram of their apparatus is shown in figure (11-5).


beam of neutral atoms is formed by evaporating Ag in an oven O.
The beam is collimated by the diaphragms D and enters a magnet N-S.
The cross sectional view of the magnet shows that it produces a field of the

field.

type described in the last section.

Since the atoms are neutral, the only

by equation (11-18).| The force


on the value of p h
acting on each atom of the beam
according to the
components
into
analyzed
for that atom, and the beam is
metallic
plate P,
a
atoms
strike
deflected
The
various values of /i
trace.
visible
leave
a
and
upon which they condense
According to classical theory, p h can have any value from p to
The predictions of quantum mechanics, as summarized by equation
(11-12), are that (i can have only the discrete values
force acting

is

the force

specified

therefore depends

fj, t

/^

-gif*B>n

ionized atoms because they would


Hevjc, acting directly on the charge, which is very much larger

fThe experiment cannot be performed with


experience a force

F=

(11-19)

than the force acting on the magnetic moment.

STERN-GERLACH EXPERIMENT

Sec. 3]

where

is

SPIN

335

one of the integers

m = -1,-1 +
Thus

AND ELECTRON

1,

0,

...,+/-

1,

+/

(11-19')

beam would be spread


quantum mechanics predicts that the
up into several discrete components.

classical theory predicts that the deflected

into a continuous band, while

beam would be split


Furthermore, quantum mechanics predicts that this should happen for
all orientations of the magnet, i.e., for any choice of the direction of the
deflected

Stern and Gerlach found that the beam of Ag atoms is split into
two discrete components one component is bent in the positive z direction,
and the other component is bent in the negative z direction. The experiment was repeated using different orientations of the magnet and several
other kinds of atoms, and it was always found that the deflected beam is

2 axis.

split into

two, or sometimes more, discrete components. These results are,

qualitatively, very direct experimental

proof of the existence of

spatial

quantization.

However, they are not in quantitative agreement with equation (11-19).


According to that equation the number of possible values of n is equal to
the number of possible values of m, which is (21 +1). Since / is an integer,
this is always an odd number. Also, for any value of / one of the possible
values of m is zero. Thus the fact that the beam of Ag atoms is split into
l

only two components, both of which are deflected, implies either that

something

is

wrong with the Schroedinger theory of quantum mechanics,

or that the theory

is

incomplete.

The theory is not wrong, but, as it stands, it is incomplete. This is shown


most clearly by an experiment performed in 1927 by Phipps and Taylor,
who used the Stern-Gerlach technique on a beam of hydrogen atoms.
is particularly significant because the atoms contain a
and the theory we have developed makes the unambiguous
prediction that, the atoms being in the ground state, the quantum number
/ will have the value / = 0.
Then there is only one possible value of m,
namely m = 0, and we expect that the beam will be unaffected by the
magnetic field since //j will be equal to zero. However, Phipps and Taylor
found that the beam is split into two symmetrically deflected components.
Thus there is certainly some magnetic moment in the atom which we have

This experiment
single electron

not hitherto considered.

One possibility would be a magnetic moment due to a spinning motion


of the charged nucleus about its own axis. The magnitude of such a
magnetic moment would be of the order of ehjlMc, where
is the mass

of the proton. But the magnetic moment measured experimentally (from


the size of the splitting) is of the order of \i B = ehjlmc, which is a thousand

MAGNETIC MOMENTS,

336

RELATIVISTIC EFFECTS

SPIN,

[Ch.

times larger. Therefore the nucleus cannot be responsible for the observed

magnetic

moment

its

source must be the electron.

This leads us to the reasonable assumption, which

supported by

is

other evidence to be discussed later, that an electron has an intrinsic

magnetic moment

[i s

We

called the spin.

Sz
s

due

of the spin angular

and

to the existence

of an

angular momentum
S and 2 component
two quantum numbers

intrinsic

also assume that the magnitude

momentum,

are related to

by the equations

Sz

=m

Vs(s
s

+\)h

(11-20)

(11-21)

which are identical with those for orbital angular momentum, and that
the relation between the spin magnetic moment and the spin angular
momentum is of the same form as the relation for the orbital case; i.e.,

_gs^B s

iLs=

(u _22)

K=
The quantity gs

is

cally deflected

g factor. Now, from

called the spin

observation that the

beam

of hydrogen atoms

components,

(11-23)

-gsf^B^s

is split

apparent that

it is

/a,

the experimental

into

two symmetri-

can assume just two

values which are equal in magnitude but opposite in sign. If

we make

the

assumption that the possible values of m s differ by unity and range


from s to +s, as is true of the quantum numbers for orbital angular
momentum, then we conclude that the two possible values of m s are
final

m
and that

=-h+\

(11-24)

s has the single value

By measuring

the splitting of the

to evaluate the magnetic force

in traversing the magnetic field.

(H-25)

beam of hydrogen atoms, it is possible


which the hydrogen atoms experience
From equations (11-18) and (11-23),

this is

Fz= -

~T-l*Bgsm s
dz

Since

ix

is

known and dHjdz can be measured,

evaluation of the quantity


is

gm
s

this

equation allows the

Within the accuracy of the experiment

found that

gs m s

it

STERN-GERLACH EXPERIMENT AND ELECTRON SPIN

Sec. 3]

Since

we have concluded

that

m =
s

gs

\,

Thus the spin g factor for an electron


O n lne other hand, \i s and S
are,

this implies

factor.j

337

(11-26)
twice the value of

is

because the relative orientation of either

orbital

its

and L
pair of vectors depends only

are antiparallel, just as

{*,

on the fact that an electron is negatively charged.


The addition of electron spin to the quantum mechanical theory requires
the addition of the quantum number m s to the set n, /, m, in order to
identify completely the mathematical form of an eigenfunction. However,
it is not necessary to indicate specifically the value of the quantum
number s since it always has the same value, J. Thus an eigenfunction
must carry four subscripts, and it is written
(11-27)

Wnlm.m,

where we use the symbol m for the symbol we have previously written as
m in order to have a symmetrical notation. Such an eigenfunction contains
the information that the z component of the spin is +f h when m = +,
and is \h when m = \. The eigenfunctions y m m can be written
t

as a product of a spatial eigenfunction

y nlmi times a

spin eigenfunction

However, the mathematical structure of the spin eigenfunctions is


m>
quite different from that of the spatial eigenfunctions. The xp nlm are
<T

the continuous functions of the continuous variables

r, 6, <j>, which we
have studied; but the "spin variable" can only have two values corresponding to the two possible values of m s so the a m are not continuous
functions. There are several ways of representing these functions, the
most commonly used being the Pauli matrix representation in which they
are written as two row, single column matrices. This representation is
,

very closely related to the Heisenberg matrix formulation of

quantum

mechanics mentioned in section 1, Chapter 7, and is treated in advanced


books on quantum mechanics. Since for our purposes it is possible to
avoid situations in which it is necessary to know the explicit form of the
spin, eigenfunctions,

As

is

implied by

they will not be treated in this book.

its

name, the spin angular

of as analogous to the angular

momentum

momentum S

can be thought

of a particle spinning about an

axis through its center. Using this in the Bohr theory, we can make an
analogy between the motion of a spinning electron around its nucleus
and the motion of a spinning planet around its sun. However, this analogy
must not be taken too seriously as electron spin is an essentially non-classical
t

is

Very accurate experiments by


actually very slightly different

Lamb
from

(1947), using different techniques,


2.

Its precise value

is^ s

show

2.00232.

that

338

MAGNETIC MOMENTS,

SPIN,

RELATIVISTIC EFFECTS

[Ch.

The reason is that the quantum number s which specifies the


magnitude of the spin angular momentum has the fixed value \. Therefore
we cannot take S to the classical limit by letting s oo, as we did in the
last chapter for L by letting / > oo. Another way of saying the same thing
quantity.

>-

magnitude of the spin S

V\{\

is

that in the classical limit the

is

completely negligible, and so spin must be essentially non-classical.

Although electron spin

is

1)

completely compatible with the Schroedinger

theory of quantum mechanics, the theory does not predict


it

must be introduced

associated with the fact that the

which ignores
developed a

its existence and


The reason for this is
Schroedinger theory is an approximation

as a separate postulate.

relativistic effects (cf. section 2,

relativistic

Chapter

7).

In 1928 Dirac

theory of quantum mechanics. Utilizing essentially

same postulates as the Schroedinger theory, plus the requirement


quantum mechanics conform with the theory of relativity, Dirac
showed that an electron must have an inherent angular momentum and

the

that

magnetic

moment

with exactly the properties described above.

This put

on a firm theoretical foundation. A quantitative treatment


of the Dirac theory would not be appropriate in this book, and we shall
continue discussing spin as an ad hoc addition to the Schroedinger theory.
However, some of the more interesting features of the Dirac theory will be
electron spin

described qualitatively in the following chapters.

4.

The Spin-Orbit

Interaction

The existence of electron spin actually was postulated first by Goudsmit


and Uhlenbeck in 1925. These gentlemen were trying to understand why
certain lines of the optical spectra of hydrogen and the alkali elements are
composed of a closely spaced pair of lines. This is the fine structure,
which had been treated by Sommerfeld in the old quantum theory as due
to a splitting of the atomic energy levels because of a small contribution to

the total energy resulting

from the

relativistic variation

The

of electron mass

Sommerfeld
by equation (5-22), were in good numerical agreement
with the fine structure of the hydrogen spectrum. But they were also in
good agreement with the fine structure of the alkali element spectra.
This was perplexing because in the old quantum theory the spectroscopically active electron of an alkali element (an element with a single valence
electron) would be moving in an orbit of large radius at relatively low
with velocity

(cf.

section 7, Chapter

5).

results of the

theory, as expressed

be negligible. Consequently
Sommerfeld's explanation of the

velocity, so that relativistic corrections should

doubt arose concerning the

validity of

origin of the fine structure.

In considering other possibilities, Goudsmit

THE SPIN-ORBIT INTERACTION

Sec. 4]

339

and Uhlenbeck proposed that an electron has an inherent angular


momentum and magnetic moment, whose z components are specified by a
quantum number m s which can assume the two values \ and +\.
The splitting of the atomic energy levels could then be understood as due
to an orientational potential energy of the inherent electron magnetic
moment in the atom's own magnetic field, which would be either positive
or negative depending on the sign of m s
.

we

In this section

shall evaluate the orientational potential energy,

+Ze

Figure
in

1-6.

a circular

An

electron moving

Bohr

orbit,

as,

Figure

seen

1-7.

in a circular

An

electron moving

Bohr

orbit, as seen

by the electron.

by the nucleus.

by a calculation which involves a judicious


mixture of the Bohr theory and the theory of relativity. The justification
of this calculation is found in the fact that its results agree with those of a
called the spin-orbit energy,

rigorous calculation using the Dirac theory.

Consider an electron of charge


nucleus of charge

+Ze,

as

shown

moving

electron with respect to the nucleus be

nucleus be

around

it

r.

From the point of view

with velocity v. This

nucleus moving with velocity

According to Ampere's law,


position of the electron,

is

on

is

and

this

its

around a

position relative to the

shown

is moving
The charged

in figure (11-7).

a current element

Zev/c

produces a magnetic

field

Ze

vX

which, at the

convenient to express this in terms of the electric

the electron.

orbit

is

H== iXj
It

v,

Bohr

Let the velocity of the

of the electron, the nucleus

v constitutes
j

in a

in figure (1 1-6).

According to Coulomb's law,

E= +Ze

field

acting

MAGNETIC MOMENTS,

340

From

the last two equations

RELATIVISTIC EFFECTS

SPIN,

[Ch.

we have

H=--vXE

(11-28)

From

a study of the behavior of Maxwell's equations under the appli-

cation of a Lorentz transformation,

general validity; any particle


electric field

Figure

1-8.

E will

it

can be shown that

moving with

experience the magnetic field

The frames of reference used

moment

equation

is

of

an arbitrary

H given by the equation.

in calculating

According to equation (11-17), the magnetic


precession of the spin magnetic

this

velocity v through

the Thomas precession.

field will

cause a

Larmor

with angular frequency

(11-29)

where we have replaced g by gs to obtain the correct value of the spin


magnetic moment. The magnetic field will, according to equation (1 1-15),
also produce an orientational potential energy
t

^E =

-{*s

gst*B

S-H

(11-30)

Both of these quantities have been evaluated in a frame of reference in


which the electron is at rest, but what we are really interested in is their
values in the normal frame of reference in which the nucleus is at rest.
Although it is not obvious that transforming back to the normal frame of
reference will change the apparent values of w L or AE, this really happens
due to a kinematical effect of the theory of relativity called the Thomas
precession.

Consider the situation as seen by an observer in the frame of reference


which the nucleus is at rest. Figure (11-8) illustrates this. The
electron is momentarily at rest in the frame
at the instant tx and
xy, in

x^

THE SPIN-ORBIT INTERACTION

Sec. 4]

momentarily

at rest in the

341

frame x 2 y2 at the

slightly later instant

t2 .

The axes of both xy and x2y 2 have been constructed parallel to the axes
of x x y x as seen by an observer instantaneously moving with the electron
in x x y x Nevertheless, we shall show that the observer in xy sees the axes of
,

own axes. He

frame
which the
electron is instantaneously at rest are precessing, relative to his own set of
axes, as the electron goes around the nucleus
even though the observers
instantaneously at rest relative to the electron contend that each set of
axes x n+1 y n+1 is parallel to the preceding set x n y n

x 2 y2 rotated

slightly relative to his

xjy3 rotated even more,

Thus he

etc.

sees the axes of the

sees that the set of axes in

Figure
in

1-9.

The frames

of reference used in calculating the

Thomas

precession, as seen

the xjrj frame.

Figure (11-9) shows xy, xx yx and x 2y2 from the point of view of the
observer in x t yv Since the electron is moving with velocity v relative to
,

the nucleus, the axes xy are

moving with velocity

negative x x axis relative to x x y x

As

v in the direction of the

seen in x-$ x the electron


,

is

accelerating

toward the nucleus with acceleration a in the direction of the positive


tx is very small, the change in velocity of
axis. If the time interval t

yx

the electron in that interval

is

dv

a(t 2

t-j)

a dt

(11-31)

and this will be the velocity of x$2 as seen by xxyv Now let us use equations
(1-16) to evaluate the components of VOJ the velocity of x 2 y 2 as seen by
xy.

These give
V

dv x

vdv
1

dvjl V

=
1

vx

%=

VndV.,

- -v-0

dvjl

342

MAGNETIC MOMENTS,

SPIN,

RELATIVISTIC EFFECTS

[Ch.

Using equations (1-16) to evaluate the components of V, the velocity of xy


by x$2 we have

as seen

dvl

vhx =

dvy vy
c

"

vv

n =

1-10.

The reader may


of

V;,,

dv
c

= -Jl-

dv 2
c

dv

dvv vy

Figure

dvv

dv 2

1.

An

exaggerated illustration of the Thomas precession.

verify that the

magnitude of V

is

equal to the magnitude

in conformity with Einstein's postulate of the equivalence of

inertial frames.

and the x

Now

axis of the

let

us calculate the angle between the vector V

xy frame.

It is

*y-3
The angle between the vector

and the x

e.-

axis of the

x$2

frame

is

dv
I.

dv 2

Figure (1 1-10) shows the x^y2 and xy frames from the point of view of xy.
According to the postulate of equivalence, V and V6 must be exactly
opposite in direction. Since the angles between the x axes and the relative

THE SPIN-ORBIT INTERACTION

Sec. 4]

343

x$2 frame appears to be rotated


xy frame. The angle of rotation is

velocity vectors are not the same, the


relative to the

dd

6h

dv

ft,

-fA
d

dv<

As dv

is

differential,

we may
d6

Now

first

we may

Then

-tV"

v represents the velocity of

Therefore
the

neglect dv 2 jc 2

an electron in an atom, so v2 /c 2

two terms in the binomial expansion of

d6

<

10~ 4

obtain a very good approximation to dd by taking only

-M

dv
v

dvv2
2vc

dv
~2#
v

Vl

v 2 \c 2

We

have

4-4

dv
v

va dt

where we have evaluated dv from equation (1 1-31). The axes in which the
electron is instantaneously at rest appear to precess, relative to the nucleus,
with angular frequency

oT =
This frequency

is

called the

dd/dt

va\2c 2

Thomas frequency. Inspection of the figures


by the vector equation

will verify that the sense of the precession is given

Ui

T=~^**

Relative to frames in which the electron

moment

is

(11-32)

at rest, its spin magnetic

precesses with angular frequency u> L

But these frames are

themselves precessing with angular frequency (o T relative to the normal

frame in which the nucleus is at rest. Consequently the spin magnetic


moment is seen in the normal frame to precess with angular frequency
(11-33)

According to equations (11-28) and (11-29),

Wr =

gs/*B

ch

vx E

MAGNETIC MOMENTS,

344

Evaluating

pi

2eA

evaluate <a T

(11-34)

mc 2

we may use Newton's law

[Ch.

we have

(11-8) and (11-26),

vxE=--^-vXE

chime

To

RELATIVISTIC EFFECTS

SPIN,

B and gs from equations

to express the acceleration of

the electron in terms of the electric field

= eE/w

With

this,

equation (11-32) becomes


<o r

Thus the precessional frequency

o>=

-^vXE
2mc*

in the

(11-35)

normal frame of reference

--^VXE
+ -^VXE= --i-vXE
mc
2mc

is

(11-36)

2mc

first derived in 1926 by Thomas.


Comparing equations (11-36) and (11-34), we see that transforming the
spin precession frequency from the frame in which the electron is at rest
to the normal frame reduces its magnitude by exactly a factor of 2. The
same is true of the orientational potential energy since, as can be seen from
equations (11-29) and (11-30), the magnitude of that quantity is pro-

This equation was

portional to the magnitude of the precession frequency.

Therefore the

value of the orientational potential energy in the normal frame of reference,

which we

shall designate

by the symbol

A s
It is

we

AEa

L is
,

i^S-H

(11-37)

convenient to express this in terms of the quantity

use equation

(1

L.

To

this

end

1-28) to write

H=

- ivx E
c

and express

in terms of the force

acting

on the

electron, using

-eE = F
In the

Coulomb

potential of the nucleus or,

spherically symmetric potential V(r), the force

dV(r) r

dr

more
is

generally, in

any

THE SPIN-ORBIT INTERACTION

Sec. 4]

Combining

we have

these three equations,

H=

345

v X

dV(r)

dr

ec r

Multiplying the numerator and denominator by m, and using equation


(11-4),

we

obtainf

H = J_ldKr) L
emc

(U38)

dr

Thus equation (11-37) can be written

^-i^S'

=
A SL
s L
.

gs and

Evaluating

[x

B we
,

2,
lemch

dr

obtain

i^H^S-L
= _LAESL
S L
2 2

(11-39)

2m

dr

Although our calculation of the spin-orbit energy was essentially based on


Bohr theory, the results, as expressed by equation (11-39), are in
exact agreement with the results of a relativistically correct quantum
the

mechanical treatment.
Let us make an estimate of AZs s L for a typical state of a one-electron
atom, to check whether it is of the same order of magnitude as the observed
fine structure splitting of the corresponding energy level.
Consider
hydrogen in the n = 2, / = 1 state. The potential V(r) is
.

V(r)

= -e

2 /- 1

so

dr

and

A s

-^-S
2m
1_

'

2 2

The magnitude of S L

for the n

Since

approximately h 2 The average value of


approximately 1/2 3 Aq, where a = fr 2 /me 2 So

2 state

is

As-lI

~ ^^

S L can be

is

yyPp^

~ 10-

16

ergs

either positive or negative,

From

~ lO^ev

m = +\ or J, the energy level is split by about 2


t

1/r

p2

the definition of the cross product


r

it is

= v X

depending on whether

10~ 4 ev. Comparing

apparent that
r

MAGNETIC MOMENTS,

346

with the energy of the n

this

we

ev,

RELATIVISTIC EFFECTS

SPIN,

2, /

Ch.

level of hydrogen, E2 = 3.4


about one part in 104 in good

see that the predicted splitting is

agreement with the splitting required to explain the fine structure of the
lines of the hydrogen spectrum associated with this level.
estimate the magnitude of the magnetic field

It is also interesting to

acting
state

on the spin magnetic moment of an electron

From

of hydrogen.

where
Fs

equation (11-30)

~ Pb ~ 10

-20

in the n

2, /

H
1

we have

erg-gauss

-1

Therefore

H ~ wl(T
10

This

is

a strong magnetic

Total Angular

5.

16

ergs

erg-gauss

lrv4
10*
gauss

field.

Momentum

In the absence of the spin-orbit interaction, and of any external torques,


the magnitudes and z components of the spin and the orbital angular

momentum

vectors

quantum numbers

L and S would have the fixed values specified by the


m m .| But in actuality there is a strong magnetic

/,

determined by L, that produces a torque


orientation of which is determined by
S. We have seen in section 2 that such a torque will not change the
magnitude of the vector S. Nor will the reaction torque acting on L
change its magnitude. But this torque does provide a coupling between
L and S which makes the orientation of each vector dependent on the
orientation of the other. As a result, neither of the vectors will have a
field,

on

the orientation of which

the spin magnetic

fixed z

is

moment, the

component because both

will precess

J
which

is

their vectorial

momentum

sum.

=L+

about the vector

The vector J

(11-40)
is

called the total angular

vector.

magnitude and all three components


of the total angular momentum vector would be fixed if there were no
external torques acting on the atom. This is only partly true in quantum
mechanics because, as the reader might guess, this theory predicts that
According to

classical physics, the

t It is not necessary to specify the value of the quantum number


have the value s = i.

s since it

can only

TOTAL ANGULAR MOMENTUM

Sec. 5]

the magnitude and z

component of J

those

we used

vector,

it

angular

are fixed, whereas

By

ponents have no definite value.

347
its

x and y com-

using techniques closely related to

in studying the properties of the orbital angular

momentum

can be shown that the magnitude and z component of the

momentum

total

quantum numbers j and

vector are specified by two

according to the equations!

j=Vj(j+i)h
Jz

(11-41)

mfi

where

m,

-j

-j,

... ,j

1,

(1

1,7

1-42)

Let us determine the possible values of the quantum number j.

Taking the z component of equation (11-40), we have


/,

which

=L +S
z

is

m-h

mfi

mfi

or

m =m +m
(11-43)
the maximum possible value of m is
and the maximum possible
of m is s = \, the maximum possible value of m, is
j

Since

value

/,

K)max = +

According to equation (11-42),

of/

this is also

equal to the

maximum possible

common

with the other quantum numbers we have been


discussing, the possible values of j differ by integers. Therefore these
value

In

values must be

members of the
J

To determine

series

25*

2>'

'

'

2>

5
2>

the point at which this series terminates,

we may

use the

vector inequality
|L
the validity of which

may

S|

easily

>

||L|

|S||

be demonstrated by constructing some

simple vector diagrams. This inequality gives

|J|>||L|-|S||
or

Vj(j

l)

> |v7(/ +\)h - Vm +

i) h\

t In the Heisenberg matrix formulation of quantum mechanics it is proved that all


angular momentum vectors (orbital, spin, and all combinations) obey quantization

equations of exactly the same form.

MAGNETIC MOMENTS,

348

We

leave

the two

it

SPIN,

RELATIVISTIC EFFECTS

an exercise for the reader to show that when

as

members of the

of j

only

series

= /+*,/-*
When = 0, L = 0, and

(n-44)

satisfy this inequality.

[Ch.

the only possible value

is

j=h
The content of equations

(11-42), (11-44),

(11-44')

and (11-44') can be formally

represented in terms of the rules of vector addition,

if

we

construct a

II
1

'

CO

OJCVJ
II

set

of vectors whose lengths are proportional to the values of the quantum


As an example, consider 1 = 2. Then the two possible
/, s, and/'.

numbers

values of j are f and

For/ =

f.

For/

m = f, \,

+\, +f.

ntj

= f, f, i,

+i, +f, +f.

Vector addition diagrams for the

two values of/ are shown in figure (11-11). The interpretation of these
diagrams needs no discussion. The diagrams represent only the rules for
adding the quantum numbers / and s to obtain the possible values of the
quantum numbers j and w If the relation between the magnitude of an
angular momentum vector, such as L, and its associated quantum number
.

were

L=

Ih

1 = Vl(l + 1) H, these diagrams would also


of L and S to obtain J and Jz Although this is not

instead of

represent the addition

true, such diagrams are often presented in elementary discussions of

atomic structure as a simplified representation of the addition of the


vectors. This is a useful representation to keep in mind, but only if it is

TOTAL ANGULAR MOMENTUM

Sec. 5]

also kept in

mind

that

349

an approximation, and that a more accurate

it is

representation of the addition of these vectors would have an appearance


similar to that

The

shown

in figure (11-12) for the case

figure attempts to

vector J, which in turn

When the

is

the vectors

Illustrating

and

the

L and S

2,j

f, ntj

f.

precessing about the

precessing about the z axis.

spin-orbit interaction

quantum numbers

Figure 11-12.

show

is

considered,

we may no

to identify the function

L and S

y>

longer use the

because

it is

no

vectors precessing about the J vector, and the

J vector precessing about the z axis.

longer an eigenfunction of the operators associated with the dynamical

Lz and Sz That is to say, L z and Sz do not have definite values


and so we cannot characterize y> by definite values of m and m s But
in losing m and m s as good quantum numbers, we gain an appreciation of
the fact that the quantum numbers j and m s can be used instead. As
quantities

the definite values vj(j + 1) h and m s h, it is apparent that


must be an eigenfunction of the operators associated with these dynamical

J and Jz have
ip

quantities.

Therefore the eigenfunctions can be written


Vniim,

(H-45)

where the four quantum numbers identify the form of the function.
Furthermore, these quantum numbers can be used and the functions (1 1-45)
are proper solutions to the Schroedinger equation, even if there were no
spin-orbit interaction, since we know that / and Jz would certainly still
have definite values because there are no external torques acting on the
atom.
This brings up the interesting point that there is a redundancy of
quantum numbers if the spin-orbit interaction is neglected. That is,
n, I, m m s ,j, ntj would all be good quantum numbers, but, since there are
only three spatial coordinates and one spin coordinate, only four quantum
t,

MAGNETIC MOMENTS,

350

RELATIVISTIC EFFECTS

SPIN,

[Ch.

numbers are needed. In other words, if we ignore the spin-orbit interaction, we know from (11-27) that there are solutions to the Schroedinger
equation, written
(11-46)

Wnlm.m,

which are eigenfunctions of the operators associated with: the total


energy, the magnitude of the orbital angular momentum, the z component
of the orbital angular momentum, and the

component of the spin angular

momentum.

However, there are also solutions to the Schroedinger


equation, written as (11-45), which are eigenfunctions of the operators
associated with
the total energy, the magnitude of the orbital angular
momentum, the magnitude of the total angular momentum, and the z
component of the total angular momentum. The relation between the
y> nlrn m and the f nUm can best be understood by considering the procedures involved in evaluating the spin-orbit energy AEa t by the methods
:

of perturbation theory.
In using these methods

we

first

solve the

Schroedinger equation,

and obtain a set of eigenfunctions,


say the xp nlm m which we found first. Then these unperturbed eigenHowever, the y>m mim are
functions can be used to evaluate Ais s L
degenerate since the total energy of the state specified by the quantum
numbers nlm m s depends only on the quantum number n. Therefore it

neglecting the spin-orbit interaction,

is

necessary to follow the procedures of degenerate perturbation theory,

which require a preliminary investigation of the way that the perturbing


spin-orbit interaction mixes those eigenfunctions that are degenerate with

each other.

From this we

obtain a

new

of certain linear combinations of the


it is

found that the new

shall not bother to write

of the

know

yj nlm

set

of eigenfunctions which consist

In carrying this through,


nlmim
set are precisely the eigenfunctions f nUmj
set

ip

down

the

y> nljm

since they are quite complicated

their explicit form.

It is

more important

and we

that

shall

we make

that the set of degenerate unperturbed eigenfunctions

fn

not need to

sure that the

reader understands the physical significance of these results.


is

We

in terms of linear combinations

The point
mi m 3

cannot

be used directly in a perturbation calculation of the spin-orbit energy,


because in the presence of this small perturbation L z and S z have no
definite values. Thus the effect of the perturbation is to mix completely

m and m s forming a new set of functions.


new set of functions are just the rp n im which describe the fact that J
and/z have definite values even in the presence of the spin-orbit interaction.

functions with different values of

This

.,

This discussion suggests that the evaluation of the spin-orbit energy

might be considerably simplified by


degenerate unperturbed eigenfunctions

initially
ip

nUm

..

starting with the set of

Since J and

Jz have

definite

TOTAL ANGULAR MOMENTUM

Sec. 5]

351

values whether or not the spin-orbit interaction

is

present,

it is

apparent

that the application of this perturbation cannot change their values.

Consequently the perturbation cannot produce a large mixing of the


Vnijm' even though they are degenerate. In this case non-degenerate
perturbation theory can be used directly (cf. section 4, Chapter 9).
According to that theory (equation 9-13 and the following discussion
extended to three dimensions), the approximate value of the shift in the
energy of the state specified by the quantum numbers nljrrij, due to the
presence of the spin-orbit interaction energy Ais g L is
,

A s -l
which

&ES

H
Using

L in that state.

we have

aeTl
This

vIi*m,(AEi.L)opV'w/ asin e dr dd

equal to the expectation value of

is

(1 1-39),

=J J J

JjJ^ (dv 7 ^r s

L^-<

r2 sin d

dr de d +

is

As

'

=
2rf?

J J J

f *" mi
~r

~d?

(S

8
L)opV '""""'" sin 6 dr d6 d t

'

(11-47)

For a Coulomb potential V(r)


becomes

= -Ze /r,
2

so dV(r)jdr

= Ze^jr 2

and

itself,

and

this

AE,. L

To

Ze 2

evaluate (S

}))

Vtiim,

(S

L) op y Wmi r 2

sin

dr dd dj>

L) op consider the equation


,

=L+

Taking the dot product of this equation times


employing the fact that L S = S L, we have
defining J.

JJ = LL + S-S + 2S-L
So

\(3

S-L =

K-/

SL
This

-L L-

S S)

is
2

- -S
2

(11-48)

Therefore
(S

L) 0P

W -L*- S%
2

i(J"ov

- L% -

S%)

MAGNETIC MOMENTS,

352

RELATIVISTIC EFFECTS

SPIN,

[Ch.

Then we have
Ze 2
L

Now

4m 2 c 2

to the eigenvalues j(j


2

(J o P

( J2
P

-L

W+

*)&,

S 20V) Wnlimj

L2P

an eigenfunction of yp

is

v'nwm.

Vllimh

J J J

^ dd H

L^, and Sfp corresponding


and |(| + l)^2 .f Therefore
,

!)#*>

[j(;

S Op)VWm/ Sin

1)

/(/

sin

1)

*]Vh,

and we have

As

L=

2
Ze,2fc2
/r
"

''
D +1) -

/(/+1)

f]

x jjjvtum,
The

and
Thus

integrals over

properties of the

y>,

^^ = ri

IJO'

Am^c*

The

1)

^+

1)

I]

Note that when

"'

this

(11-49)

- Rm(ry dr

Jf**^) r

It gives

Rtir)RJrydr=r-*=
3

dr dd dj>

both give unity owing to the normalization

<

can be evaluated.

integral over r

V>m im

'"

aln%l

equation says r~ 3 -

oo.

)(/

For

1)

this value

of

/,

- /(/+ 1) |] = 0. Thus AEa L appears to be


= 0. However, this is merely the result of evaluating
r~ s in perturbation theory. A more accurate evaluation of this quantity
shows that is very large, but finite, for = 0. Since [j(j + 1) /(/ + 1)
] is strictly zero in this case, we have
=

| and [j(j+ 1)
indeterminate for /

it

AS
For other values of

/ it is

^ =
S'

0,

z
[j(j
4m 2 c2 a 3

+
n

l)

l(l

=
+

(11-50)

+ l) \){l + 1)
i(l

In terms of thefine structure constant a


e^\hc
value of the unperturbed energy of the state,
t All these functions are eigenfunctions of 5 p since
,

'

Hi +

DA*.

1/137,

f]

and the absolute

En = mZ2e4/2/j

S 1 can have only

2
,

this

the single value

Sec. 6]

RELATIVISTIC

can be written

^ = Z
*'

The value of

|E W | [j(j

2n

353

- /(/ +
+ |)(/ +

1)

1(1

1)

= +
I

the situations in which

AES

AS

\ or j

is

f]

1)

on the value of/ The two


which correspond respectively to

=
/

\,

roughly parallel to

L for these possibilities,

+
L

=
_

we

+^
1)

L"Jf

1)(/

or antiparallel to L.

find

Z2|J

,
n(2l

= + i'#0

= _

'

(11-51)

7 2 IF
l2
z
^"^

n/(2/+l)

6.

the spin-orbit energy depends

possibilities are

Evaluating

CORRECTIONS

Relativistic Corrections for

In section 4 of this chapter

|,

#o

One-Electron Atoms

we estimated

the shift in the energy of a

hydrogen atom, due to the spin-orbit interaction, to


be about 10~4 times the energy of the level. In section 7 of Chapter 5
we estimated that the shift in the energy of such a level, due to the relativistic dependence of mass on velocity, is also about 10~ 4 times the energy
of the level. The energy shifts due to relativistic effects other than the spin
are thus quite comparable to those due to the spin-orbit interaction for
the hydrogen atom, and we cannot expect to be able to compare the
typical level of the

predictions of equation (1 1-50) or

(1

1-51) with the experimental data until

A complete treatment of the


can only be given in terms of the Dirac theory; but
which are almost complete can be obtained with the Schroedinger

we have taken

these effects into account.

relativistic effects

results

theory by treating the additional relativistic corrections as a perturbation.

The

starting point of Schroedinger

quantum mechanics

writing the classical expression for E, which


kinetic energy

T and

the potential energy V.

is

defined as the

That

consists in

sum of

the

is,

E=T +V = ^-+V

(H-52)

relativistically correct since,

according to equations

2m

This equation
(1-22')

and

is

not

(1-25),

where we use

for the rest mass.

Let us write the

expression for E, and then evaluate

it

relativistically correct

in the limit of small values of p.

MAGNETIC MOMENTS,

354

RELATIVISTIC EFFECTS

SPIN,

[Ch.

This procedure gives

+V = (cV + mV) H - mc

nei

uei

= mc 2
[(' +

Re

=i

Enel

Tnei

;&

+ F

2mV

8mV

2m

+ V

\\4

,2

m&

+V

8m 4c 4

+ K

We have taken the first three terms in the binomial expansion.


this

for

with equation (11-52),

E is

we

Comparing

see that the error in the classical expression

approximately

(E

AE Rel

8mV

E 2 + V 2 - 2EV

- Vf

2mc 2

2mc

2mc
(11-53)

If we use the eigenfunctions y> ljm .,

we can use non-degenerate

perturbation

E of the corresponding state

theory to evaluate the shift in the total energy

due to the addition of the term AE^, which may be thought of as an


addition to V. According to the theory, the approximate value of the
energy

shift is

AReI

"wJJ

Now

y=

wt lims(E

-Ze 2 jr

+V

2EV) 0V y> nljm r2

E = En

and

a constant

So we obtain

zv

El
A Be = 2
i

rcc

2mc 2 JJJ

2mc

2T

Tn " mt

'" Jmi

J J JvXw, ; V /

sin

rfr

d6

d<f>

Evaluation of the integrals yields

Aaei

z2

31^
|.

...

P
i

A)

(11 _54)

Comparison of equations (11-51) and (11-54) shows that the energy shift
calculated is, in fact, comparable in magnitude to the energy

we have just

RELATIVISTIC

Sec. 6]

CORRECTIONS

355

due to the spin-orbit interaction. Since both quantities have the


same Z dependence, this is true for all one-electron atoms.
Adding the two energy shifts to the unperturbed energy, we obtain an

shift

expression for the total energy,

E=En + AE

+ ABel

of the state specified by the quantum numbers

(21

!)(/

+1)

It is

nljrrij.

21

An

j=l + hJ^0

-*4^(rh-h)\
+
and

For both values of j

For

0,

Ag

this

can be written

and

E = En + ABel

which

is

^Z 2e 4

'-

"Sbl'+^-s)}2/rV

<u - 56)

where we have evaluated En using the reduced electron mass fi.


Let us compare these results with the results of Sommerfeld's relativistic
Bohr theory of the energy levels of a one-electron atom, given in equation
,

(5-22).

This

is

Shi'
gZV

E_

2h

J,

ri

t
+ vii-s;)}

(11 - 57>

The quantum number n e appearing in this expression can assume the values
n e = 1, 2, 3,
n.
Since the quantum number j can assume the
.

values j

n
\,\ the quantity j
\ can assume the values
n. Thus, for /
1, 2, 3,
\
0, the values of total energy predicted by our calculations are in exact agreement with the values predicted

i,

-f ,

f
.

by Sommerfeld. For
For /

0,y

value of j

is

there

i; otherwise y

is

=/

and the maximum value

is

a discrepancy.

J.

Since

But

0, 1, 2,

!+=

\.

this
,

discrepancy
1,

the

is

minimum

MAGNETIC MOMENTS,

356

SPIN,

RELATIVISTIC EFFECTS

[Ch.

removed by the Dirac theory, which shows that there is one more relativistic correction to the energy, whose existence we could not have guessed
because it has no classical analog. This correction is zero for / ^ 0. For
its value is such that, when it is added to the right side of (1 1-56),
/ =
the resulting equation has the same form as equation (11-55). Therefore
the results of a complete relativistic treatment by the Dirac theory can be
expressed by the single equation

=
These

/.ZVf.

-fcl

results are in exact

theory.

+
,

ZV/

(jTi-J)
1

3\\

, co ,
(11 - 58)

agreement with the predictions of Sommerfeld's

Since the Sommerfeld theory was based on the

Bohr

theory,

it is

only a very rough approximation to physical reality. In contrast, the Dirac


theory represents an extremely refined expression of our understanding
of physical reality. From this point of view the agreement between equa-

and (11-58) is one of the most amazing coincidences to be


found in the study of physics.
The energy levels of the hydrogen atom, as predicted by the Bohr,
Sommerfeld, and Dirac theories, are shown in figure (11-13) for the first
tions (11-57)

three values of n.

In order to

make

the energy level splittings visible, the

displacements of the Sommerfeld and Dirac energy levels from those given

by the Bohr theory have been exaggerated by a factor of (137) 2 = 1.88 x


104 Thus the diagrams would be completely to scale if the value of the
fine structure constant were 1 instead of 1/137. The values of the quantum
number w which specifies the spatial orientation of the total angular
momentum vector, are not shown on the Dirac energy level diagram
because the energy of the atom does not depend on this quantum number
(unless the atom is in an external electric or magnetic field). Although we
have not mentioned it before, a quantum number analogous to w is
required in the Sommerfeld theory to specify the spatial orientation of the
plane of the orbit. Since the energy of the atom does not depend on this
quantum number (if the atom is not in an external field), its values are not
indicated on the Sommerfeld diagram. Also not shown are the energy
levels of hydrogen measured experimentally by optical spectroscopy,
which are in extremely good agreement with the levels of the Sommerfeld
and Dirac theories.
The only essential difference between the results of these two theories
is that the Dirac theory predicts that for most levels there is a degeneracy,
.

in addition to the trivial degeneracy with respect to spatial orientation

mentioned above, because the energy depends on the quantum numbers n


and j but not on the quantum number /. Since there are generally two
values of / corresponding to the same value of/, the Dirac theory predicts

RELATIVISTIC

Sec. 6]

CORRECTIONS

that most levels are actually double.

mentally in 1947 by Lamb,

357

This prediction was verified experi-

who showed

that for n

are two levels, which actually do not quite coincide!-

above the
and the n

2 and j
The / =

\ there

level lies

= level by about one-tenth the separation between that level


= 1,1 = 1, j = f level. This is called the Lamb shift. It can

Bohr

Sommerfeld

Dirac

n = 3,i=2,/ = 5/2

= 3, i = 1, > = 3/2 and n = 3, 1 = 2, = 3/2


^n = 3, i = 0,/=l/2and n = l, 1= l,j= 1/2
j'

n = 2,

= l,;=3/2

n = 2,

= 0,/ = l/2and n = 2,

(=1, J

= 1/2

-10

-13.6ev^
-15

n=U = 0,/ =
Figure
Dirac.

l/2

1-13. Energy levels of the hydrogen atom according to Bohr, Sommerfeld, and
The displacements from the Bohr energy levels are exaggerated by a factor of (1 37) 2
1

be understood quantitatively in terms of the theory of quantum electrodynamics, as can the slight departure of g s from 2 mentioned in the footnote
on page 337, but we shall not be able to discuss this theory here.
We close our treatment of the one-electron atom by mentioning one

which influences the atomic energy levels. This is the hyperfine


which is due to an interaction between the magnetic field produced by the moving electron and an inherent magnetic moment of the

final effect
splitting,

Lamb's experiment involved measuring the frequency of quanta absorbed in


between the n 2, I = 0,j = i and n = 2, /=l,y = J levels. The energy
difference between these levels is so small that the frequency is in the microwave radio
range. Since measurements of radio frequencies can be made very accurately, it was
possible to measure the energy difference to five significant figures!
t

transitions

MAGNETIC MOMENTS,

358

SPIN,

RELATIVISTIC EFFECTS

[Ch.

As nuclear magnetic moments are smaller than atomic magnetic


moments by three orders of magnitude, the hyperfine splitting is smaller
than the spin-orbit splitting by the same factor. Nevertheless, we shall
nucleus.

see in Chapter 13 that this effect can be described quantitatively in terms

of Schroedinger quantum mechanics. In fact, every aspect of the behavior


of a one-electron atom can be explained in detail by the theories of modern
physics. As we have seen, these explanations can be quite complicated.
Fortunately the degree of complication does not increase in proportion
number of electrons in the atom.

to the

BIBLIOGRAPHY
Leighton, R. B., Principles of

Modem

Physics,

McGraw-Hill Book Co.,

New

York, 1959.
Ruark, A. E., and H. C. Urey, Atoms, Molecules, and Quanta, McGraw-Hill
Book Co., New York, 1930.
White, H. E., An Introduction to Atomic Spectra, McGraw-Hill Book Co.,
New York, 1934.

EXERCISES
1.

Use Ampere's law

to evaluate the magnetic field produced

by a

circular

current loop at a point on the axis of symmetry far from the loop. Then evaluate
the magnetic field produced at the same point by a magnetic dipole located at
the center of the loop and lying along the axis of symmetry.
fields are the

magnetic
to

show

same

moment

that

if this

if

the current in the loop

and

its

Show

that the

area are related to the

of the dipole by equation (11-2). Next extend the argument


relation

is

obeyed the fields are the same at all points

the loop or dipole. Finally, extend the argument to

show

that this

is

far

from

true for a

plane current loop of arbitrary shape.


2. Evaluate ji^L for an electron moving in the n = 2, n = 1 elliptical orbit
e
of an atom of the type treated in the Sommerfeld theory. Compare with equation
(1 1-7). Hint: Use some of the equations derived in exercise 3, Chapter 4, and
exercise 4, Chapter 5.
3. The first footnote of section 3 states that a Stern-Gerlach experiment
cannot be performed with ionized atoms. It might be thought, though, that the
experiment can be done if the axis of the magnetic field is bent to follow the
curvature produced by the Hev/c force, and if a beam is used which has such
small transverse dimensions that the particles never get far from the axis, because
then all the particles of the beam will feel essentially the same Hev/c force.
However, the uncertainty principle shows that it is impossible to produce a
beam in which the transverse dimensions remain sufficiently small. Prove" this.

Ch.

EXERCISES

II]

359

4. The Thomas precession can also be described in terms of a time dilation,


between the reference frame in which the nucleus is at rest and the reference
frames in which the electron is instantaneously at rest, which leads to a disagreement between an observer at the nucleus and the observers at the electron
concerning the time required for each to make a complete revolution about the

other.

Work

equation
5.

out the details of this description, and compare the results with

(1 1-32).

Verify equation

to that equation.

(1 1-44),

and

also verify the vector inequality

which leads

CHAPTER

12
Identical Particles

I.

Quantum Mechanical
One

final topic

discussed before

of the theory of quantum mechanics remains to be

we can study

how

the structure of multi-electron atoms.

It

quantum mechanical
of a system containing two or more identical particles. The

concerns the question of


description

Description of Identical Particles

to give a proper

nature of the question can best be illustrated by a specific example.

Consider a box containing two electrons. These two identical particles

move around
scattering

in the box,

from each

bouncing from the walls and occasionally


In a classical description of this system, the

other.

electrons travel in sharply defined trajectories so that constant observation

of the system would allow us, in principle, to distinguish between the two

That is, if at some instant


and the other electron 2, we could constantly
observe the motion of the electrons without disturbing the system and
would be able to say which electron is 1 and which electron is 2 at any
later instant. In a quantum mechanical description this could not be done
because the uncertainty principle would not allow us to observe constantly
the motion of the electrons without completely changing the behavior of
electrons even though they are identical particles.

we label one of the electrons

An

is that in quantum mechanics the


wave function associated with each particle will lead
to an overlapping of these wave functions which would make it impossible
to tell which wave function was associated with which particle. A good
example is the two electrons in an He atom. When this system is in its
ground state the wave functions of the two electrons overlap completely.
Of course, there is also overlap of the wave functions associated with the
electron and proton of an H atom, but this does not lead to any problems
in distinguishing one particle from the other because the particles are not

the system.
finite

equivalent statement

extent of the

identical.

360

Sec.

DESCRIPTION OF IDENTICAL PARTICLES

I]

We

see that there

361

a fundamental difference between the classical and

is

quantum mechanical description of a system containing identical particles.


A correct quantum mechanical theory of such systems must be so formulated as to take explicit account of the indistinguishability of identical
particles.

When

done, the theory leads to the prediction of some

this is

very interesting

These

effects.

indistinguishability

have no

effects

analogy because

classical

a purely quantum mechanical situation.

is

way of

Let us look for a

wave functions describing the two

writing

box which

identical particles in the

To

will contain a

mathematical expression

we shall assume
two particles. Then the particles
will bounce between the walls of the box and their wave functions will
certainly overlap, but they will not scatter from each other. We shall show
of the ideas developed above.

that there are

no

simplify the argument,

interactions between the

later that the results of the following discussion are of quite general validity

The time independent Schroedinger equation

despite this simplification.


for the system

is

&_ / dVr

2m\dx\

yr

ip

T\

dz\)

dy\

tf_ l d

yT
2m\dx\

yr

~dyj

+ VT

tp

tp

T\

~dz\)

= ET y T

(12-1)

where

xi>
x2

We

have written

and m 1

=m =
2

with the labels

m=
=
Vi> z
z =
t/

2,

the

ine coordinates of particle

the coordinates of particle 2

this directly

and

either particle

m. Since
1

mass of

from equation (10-8)

for the case

N=

equation specifies the identity of the particles


in using it we stand a chance of violating the
this

2,

requirements of indistinguishability. This will be discussed more carefully


later.

In equation (12-1),

y T (*1;

VT (x,
Since

=
=
z
ET =

the eigenfunction for the whole system

2 2)

the potential energy for the whole system

2)

the total energy for the whole system

we have assumed

that there

is

no

the potential energies of each particle in


the box.

Each

its

is

simply the

two

sum of

interaction with the walls of

depend only on the coordinates of one


are identical, the two potential functions

potential energy will

particle and, since the particles

are the same.

interaction between the

energy of the whole system

particles, the potential

Thus

VT (xx

z 2)

V(xx yx zj
,

V(x 2 y 2 z^
,

(12-2)

IDENTICAL PARTICLES

362

We leave it

as

[Ch. 12

an exercise for the reader to show that for the potential

(12-2) solutions exist to equation (12-1) of the form

frixi, ... ,z 2 )

where

y)(x lt

y lt

and y{x2 y2

z-J

y2

(12-3)

2)

z 2 ) satisfy identical time independent


These equations have three spatial coordinates
,

Schroedinger equations.

v(*i, Vi, i) y(x2 ,

as independent variables.

Consequently, each of the eigenfunctions

quantum numbers to specify the mathematical form of its


dependence on the three spatial coordinates. In addition, one more
quantum number is required to specify the orientation of the spin angular
momentum vector of the particle.f However, we shall shorten the notation
to designate
by using a single quantum number from the series a, /?, y, d,
a particular set of the four quantum numbers required to specify the state
of a particle. Then a particular eigenfunction for particle 1 would be
requires three

written

We

by writing

further shorten the notation

this

V(l)

This eigenfunction contains the information that particle


described by a.
/?

An

eigenfunction indicating that particle 2

is

in the state

is

in the state

would be written

Then
1 is

the total eigenfunction

and

in the state a,

rp

We write

ip

T (xx

particle 2

T{xx

is

z 2 ) for the case in

in the state
z2 )

/?,

which

y>(l) V/,(2)

this

V = %(1) V&)

An
is

eigenfunction indicating that particle

in state

particle

is

(12-4)
is

in state

/?

and

particle 2

a would be
Vft

It is interesting to

Ytfl)

(12-5)

%(2)

evaluate the probability density functions for systems

described by the two eigenfunctions (12-4) and (12-5). These are

V> =

V?(l) W*( 2 ) V*(l) W(2)

(12-6)

tftt) V.*(2) V,(l) V(2)

(12-7)

and
v;V/,

Now,

since the

two

identical particles are indistinguishable,

we should

be able to exchange their labels without changing a physically measurable


t Particles other than electrons also

have

spin.

SYMMETRIC

Sec. 2]

AND ANTISYMMETRIC EIGENFUNCTIONS

quantity such as the probability density.

on

From

V*(l) Wp( 2 ) V(!)

V/>(2)

->

y(l)

ip*(2)

y> a (2)

^(1)

1-2
2-1

equations (12-6) and (12-7)

and so a

Let us carry out this operation

We find

<3W

vZvVrt

363

it

is

apparent that y>*pya p

y^w,.

^ W*aWfia

->

relabeling of the particles actually does change the probability

density function calculated


true for (12-5). Therefore

from the eigenfunction

we must conclude

The same

(12-4).

is

that these are not acceptable

eigenfunctions for the description of a system containing two identical


particles.

has been

2.

The suspicion which we expressed

after writing equation (12-1)

justified.

Symmetric and Antisymmetric Eigenfunctions

It is possible to construct an eigenfunction which satisfies the time


independent Schroedinger equation (12-1), and yet has the acceptable
property that its probability density function is not changed by a relabeling

of the particles.

two

In fact, there are two ways of doing

linear combinations of

y^

Consider the

this.

and y

fi(X ,

V>s

V/.J

(12-8)

Va

= -j= [V -

VVJ

(12-9)

-t=

[>*

and

done the exercise suggested below equation (12-2), it


be apparent that y^ and yfia are degenerate. If not, it should still be
apparent since the total energy of a system containing a particle in state a
If the reader has

will

and a
if

particle in state

fi

will

not depend on which particle

the particles are identical.

This

is

is

in

which

called exchange degeneracy.

state,

In the

we have shown that any linear combination of


degenerate eigenfunctions will be a solution to the time independent
Schroedinger equation. Therefore the two new eigenfunctions ys and y>A
footnote on page 189

are both solutions to equation (12-1).

The

eigenfunctions

s and

rp

rp

are,

of course, also degenerate. The probability densities for these two eigenfunctions are

YsWs

= KviVM +

vAVa.)

vlv>)

(12-10)

vIwa

wtv>pJ

- Kvw* +

vivn>

(i 2 - 11 )

Kyvf

and
i(y*P y> x p

IDENTICAL PARTICLES

364

From
y>p x

these

it is

[Ch. 12

show that y>s and %pA are normalized if y)ap and


Furthermore, since (12-4) and (12-5) show that
~* Vajs when we exchange the labels of the particles, we

easy to

are normalized.

Vxp - Vpa an<i V>pa


see by inspection that both
tion.

This

fsVs an d WaVa

Thus the eigenfunctions

ability flux.

are

unchanged by

this

opera-

also true for other measurable quantities such as the prob-

is

ip

s and

xp

as well as the

wave

functions associated with these eigenfunctions, provide an acceptable


description of a system containing two identical particles.

Only region

in

1/^(1) or i/^(2)

is

Figure 12-1.

which
non-zero

Only region

in

\p (1) or \p (2) is

Illustrating the classical limit for a

Although the

which
non-zero

system of two indistinguishable

particles.

numbers 1 and 2 do appear when we write in full the expressions for ys


and tpA this labeling does not violate the requirements of indistinguishability because the value of any physically measurable quantity, which can
be obtained from the eigenfunctions, is independent of the assignment of
,

As far as the eigenfunctions themselves are concerned, the


operation of exchanging the labels of the particles leaves one unchanged

the labels.

and

multiplies the other

by 1. By inspection we

see that

(12-12)

\Z\

Va-^-Wa
The eigenfunction

s is called a symmetric eigenfunction, and the eigenan antisymmetric eigenfunction. The behavior of y>A
not objectionable since an eigenfunction is not physically measurable.
An instructive interpretation of the role played by the first and second

function
is

y>

is

y>

called

parentheses in the expressions (12-10) and (12-11) for

fsVs

obtained by considering these expressions in the classical

aiK*

limit.

WaWa

IS

In this

one particle can be sufficiently localized that


do not overlap. That is, there
will exist two non-overlapping regions of space A and B, as indicated in
figure (12-1), such that y>J,l) or y>J2) is non-zero only when the coordinates
of particle 1 or particle 2 are in A, and such that y/1) or ^(2) is non-zero
only when the coordinates of particle 1 or particle 2 is in B. Then the
limit, the eigenfunctions for

two

different one-particle eigenfunctions

THE EXCLUSION PRINCIPLE

Sec. 3]

two-particle eigenfunction yi^

ordinates of particle
But,

if this is so, y>p a

are in

if>p(l)

ya(l)

365

y>p(2)

will vanish unless the co-

and the coordinates of particle 2 are in B.


ya(2) must vanish. Thus, if there is no region

which both of the one-particle eigenfunctions have non-zero values,


either one or the other of the two particle eigenfunctions ya/3 or y>^ will
always be zero. Then the second parentheses in the expressions for
ipsfs and WaVa w iH always vanish, and we have
in

Wsfs

We

VaWa

Kv*/>V*/j

see that in the classical limit the

(12-13)

V*Vp)

quantum mechanical

expressions for

the probability densities of the symmetric and antisymmetric eigenfunctions

are equal.

for

y>*ip

moment's consideration should demonstrate to the reader

common

that their

s and

value (the

y>*ipA ) is

first

parentheses in the general expressions

completely analogous to the classical expression

for the probability density of a system containing two identical particles,

and the other in the state /?, but it is not known which
which state. The quantity which is responsible for the
difference between f*y) s and f*yA when the one-particle eigenfunctions
if

one

is

in the state a

particle is in

overlap (the second parentheses in their general expression) contains terms

which correspond to some sort of interference between the two-particle


eigenfunctions xp^ and y>^. The presence of this quantity is a result of the
quantum mechanical requirements of indistinguishability.

3.

The Exclusion

We

T^,

now prove that,


N t) describing

shall
.

Principle

if at

some

certain instant the

a system of

wave function

identical particles

is

either

symmetric or antisymmetric with respect to exchanging the labels of any


two of the particles, then under any circumstances the wave function must
always preserve the same symmetry. To do this we consider the Schroedinger equation as a differential equation governing the time dependence
of *F(x lt
, z
N i), which can be written
.

eop^K, ...,zN ,t)

ih

d^ (*i.--

**.*)

at

This equation allows us to evaluate the wave function at time


we know it at time /, because

^i, .-.,%, +
t

dt)

= Y(x

t,

...,zN ,t)

+ gyfa '--'^

dt, if

t)

dt

dt

= TK, ...,zN ,t)+ ^T^, .... zN


in

t)

dt

IDENTICAL PARTICLES

366

Now the

[Ch. 12

derivative terms of the total energy operator

are obviously symmetric with respect to exchange of the labels of any


particles since such

This

on

is

will leave the

also true of the potential energy term since

the labeling of the identical particles, even

between the
(e

an exchange

j>lih)

^fa,

,z N ,t)

^(a^,

if

be symmetric and,

will

+
+

if

if

there are interactions

Since eop is a symmetric operator, the term


,z N t).
has the same symmetry as Yfo,

particles.

Consequently,

two
summation unchanged.
its value cannot depend

z N , t)

W (xv

is
,

symmetric,

z N , i) is

Y^,

antisymmetric,

zN,

Y^,

t
.

+
.

di)

,z N

be antisymmetric. By continuing this integration process to


2 dt, etc., we prove that the symmetry character of the wave function
t
must be preserved. If we consider the particular case of an eigenstate
t

di) will

wave function

W(xx ...,zN ,t)


,

e-

imlh

vte, ..-,%)

symmetry character of the eigenfunctions is the same


character
symmetry
of the wave functions,
as the
For a system of two non-interacting identical particles, we know that
the eigenfunction must be either symmetric or antisymmetric. We have
it is

just

clear that the

shown

that, in the

more general

case of a system of several identical

which are interacting with each other, the eigenfunction may


be either symmetric or antisymmetric. That is, we have shown that it is
consistent for the eigenfunction always to be symmetric or always to be
particles

The question then

antisymmetric.
real

arises:

Do

eigenfunctions describing

systems of several interacting identical particles have a definite

symmetry, either symmetric or antisymmetric, and, if so, which one?


The answer to this question is known. However, it is provided not by the
theory of quantum mechanics, but by experiment.
As a result of an analysis of experimental information concerning the
energy levels of certain atoms, in 1925 Pauli enunciated the exclusion
principle:

In a multi-electron
the

same quantum

atom

there can never be

more than one

electron in

state.

was soon established from the analysis of other experimental data that
and not, specithe exclusion principle operates in any system containing
fically, of atoms
It

the exclusion principle represents a property of electrons


;

electrons.

Now

consider the two-particle antisymmetric eigenfunction

THE EXCLUSION PRINCIPLE

Sec. 3]

367

(12-9) for the case in which both particles

and 2 are

in the

same

state a.

It is

Wa

The eigenfunction vanishes

~(^

v>

identically;

if

J=

two

(12-14)

Thus the property of electrons expressed by the exclusion

principle

exactly the property of the antisymmetric eigenfunctions,

and we

state.
is

by
same quantum

particles are described

the antisymmetric eigenfunction, they cannot both be in the

conclude that:

system containing several electrons must be described by an antisym-

metric eigenfunction.

We know
taining

the form of the antisymmetric eigenfunction for a system contwo non-interacting particles. Note that it can be written as a

determinant

= J_/%(1)

Wa

VJ>
(12-15)

V2"Wa) w(2>
since,

upon evaluating the determinant,


V>a

= j= IvJX) fi>(2) -

?/,(!)

this is

%(2)]

-i=

[VW

y,J

For a system containing


non-interacting particles the antisymmetric
eigenfunction can be written, in a completely analogous form, as

%(2)

^.(1)

Wa

(12-16)

\/W
W(l)

This

is

called a Slater determinant.

consists of

sums and

V>XN)j

Vv (2)

differences of

Since the value of this determinant

Nl terms comprised of products of

one-particle eigenfunctions, such as if>x (l) y> (2)


y>v (N) and similar
terms in which the particle labels are permuted, rpA is simply a linear
combination of degenerate eigenfunctions. Therefore %pA will also be a
fi

solution to the time independent Schroedinger equation.

IjVNl

is

such that

y>

particle eigenfunctions is

The constant

be normalized if each of the products of onenormalized. Because a determinant vanishes if

will

IDENTICAL PARTICLES

368

any two rows are


of vanishing
It is

such a

if

identical,

particle

possible to formulate

way

apparent that

it is

more than one

A has
same

y>

in the

is

[Ch. 12

the required property


state (e.g., if

most quantum mechanical

/?

a).

calculations in

that the interactions between the identical particles are con-

sidered perturbations.

Then, for unperturbed eigenfunctions, functions


However, if the interactions between

of the form (12-16) can be used.

the particles cannot be ignored, the antisymmetric eigenfunctions cannot

be written as (12-16), since y(l) y^(2)


y> u (N), and all the other terms,
will not be solutions to the time independent Schroedinger equation.

In this case antisymmetric eigenfunctions which vanish identically

same

if

two

be obtained by taking sums and


differences of the N\ degenerate solutions to the time independent Schroeidentical
dinger equation which will always exist for a system with

particles are in the

state

can

still

particles.

In this and the following chapters

we

shall study the multi-electron

systems whose analysis led Pauli and others to the conclusion that such
systems must be described by antisymmetric eigenfunctions. We shall
see that this property of electrons plays an extremely important role in

Atoms,
and therefore the entire universe, would be radically different if electrons
were such as to require a description in terms of symmetric instead of
governing the behavior of systems containing these particles.

antisymmetric eigenfunctions.
The symmetry character of particles other than electrons

is

also a

question which can be settled only by experiment. It is found that systems


atoms), or neutrons (uncharged particles of
of protons (the nuclei of

approximately proton mass), or certain other particles, must also be


described by antisymmetric eigenfunctions. Thus an eigenfunction describing a system of non-interacting protons must be of the form (12-16).

On

it is found that systems of quanta, or certain other


must be described in terms of symmetric eigenfunctions. Also,
systems containing particles which are tightly bound aggregates of an
even number of antisymmetric particles (e.g., a system containing alpha
particles which consist of two neutrons and two protons) must be described
by eigenfunctions which are symmetric with respect to exchange of the
labels of any two of the aggregate particles, because such an exchange is
equivalent to an even number of exchanges of the labels of antisymmetric
particles, and the net effect is to multiply the eigenfunction for the system
by an even power of 1 A symmetric eigenfunction for a system of N
non-interacting particles can be obtained by taking the sum of the Nl

the other hand,

particles,

y,(N) and the


terms that can be formed from the product y> a (l) %Pp(2)
similar terms obtained by permuting the particle labels, and then dividing

by the normalization constant

VWl

If the particles are interacting,

PROPERTIES OF ANTISYMMETRIC EIGENFUNCTIONS

Sec. 4]

symmetric eigenfunction can

still

be constructed by taking the

369

sum of

the TV! degenerate solutions to the time independent Schroedinger equation.


It is interesting to

note that there must be some connection between the

symmetry character of a

particle

and

its intrinsic

spin angular

momentum,

since all the antisymmetric particles also have half-integral spin just as the

However,

electron has, while

all

the reason for this

not completely understood at present. Consequently,

is

the others have zero or integral spin.

the symmetry character of a particle


like

4.

must be considered a basic property,


mass, charge, and spin, which can be determined only by experiment.

Additional Properties of Antisymmetric Eigenfunctions

As mentioned

in section 3,

Chapter

11,

it is

often convenient to write

the complete eigenfunction containing both the spatial

and spin coordin-

ates of a particle as a

function.

This

product of a spatial eigenfunction and a spin eigenpossible providing there are no interactions explicitly

is

involving the spin

(e.g.,

the spin-orbit interaction), so that

remains constant and m s is a good quantum number.


example, the one-electron-atom eigenfunctions f nlmim

its z

component

As a particular
may always be

written

Ynlmim s
since, if the

Wm mi m

are valid,
s

T nlmi

the general case in which this restriction

is

eigenfunction for a particle carrying the label


V(l)

must be a good quantum number. In

V0)

satisfied,
1

we may

write the

as

(12-17)

*..(!)

where the single quantum number a stands for a set of three spatial
quantum numbers. Now consider a system containing two particles in
which there are no explicit interactions, spin or otherwise, between the
particles, Then, according to (12-4) and (12-17), we may write the eigenfunctions describing these two particles as the product
V>i,

y.(l) y*(2)

%(1)

v(2)

a ms

(\)

a ms p.)

Using a notation analogous to that defined by (12-4),


Wa0

y>a b <y msa m, b

this

can be written
(12-18)

not an acceptable eigenfunction for a system of two electrons


an antisymmetric character) because it is
not antisymmetric with respect to an exchange of the numbers 1 and 2
This

is

(or other identical particles of

which label the

electrons. However, it is possible to construct a properly


antisymmetrized eigenfunction by taking a symmetric or antisymmetric

IDENTICAL PARTICLES

370
linear

[Ch. 12

combination of spatial eigenfunctions times a linear combination

of spin eigenfunctions of the opposite symmetry, so that the product which

forms the complete eigenfunction is antisymmetric. Consider the spin


eigenfunctions. Since for an electron m s can assume just two values

and

From

+ |,

these

there are four possibilities

we can

construct four different spin eigenfunctions,

-j=( a +K-K

a -A+vd

singlet

a -H+\d

triplet

a +X+H
(12-19)

j=(o+ii-H

All four are normalized.

The

one, which

first

is

called the singlet spin

obviously antisymmetric with respect to an exchange of


the particle labels. The last three form the so-called triplet of spin eigenfunctions. They are all obviously symmetric with respect to an exchange
eigenfunction,

of the labels.

we

is

To

obtain complete eigenfunctions which are antisymmetric,

multiply the antisymmetric singlet spin eigenfunction by a symmetric

spatial eigenfunction,

4(Va,+
and multiply each of the symmetric

(12-20)

V>

triplet spin eigenfunctions

by an

antisymmetric spatial eigenfunction,

j=(V,*-YO

(12-21)

So there are the following four different antisymmetric complete eigenfunctions for a system containing two non-interacting electrons
1

"/|(Va&

Va

Vba)

-]=(? A- A
l

ff-H + M)

=V
1

-^(Vai.

(12-22)

-friWab-

ViJ-7=( tf +H-W

V>ba)<*-X-<4

a -H + %)

PROPERTIES OF ANTISYMMETRIC EIGEMFUNCTIONS

Sec. 4]

The

371

and the others describe the three


and degenerate.
A physical interpretation of the singlet and triplet states can be obtained
by evaluating, for each state, the magnitude and z component of the total
spin angular momentum vector S', which is
first

describes the singlet state,

All four are normalized, orthogonal,

triplet states.

S'

sum

the

do

Sj

S2

(12-23)

momenta of the two electrons. We cannot


because we have not studied the properties of the
we shall merely state the results and give some

of the spin angular

this in detail here

spin eigenfunctions ;

s'=l
S

= ^||*2=
s'

Figure 12-2.

arguments for

s,

=H
o

Vector addition diagrams for the quantum numbers


their validity.

The reader

st

i, s 2

should, however, have no

|.

diffi-

culty in accepting the results, as they are completely analogous to those

governing the addition of the spin and orbital angular momentum vectors
to form the total angular momentum vector. It is found that the magnitude

and z component of the


by the equations

total spin

S'

=
=

m,

s'

S'

angular

vs'(s'

momentum

vector are given

l)h
(12-24)

m'h

where

s'

and where
0,

(12-25)

and triplet states, respectively. Equation (12-25) can be


proved by the same kind of arguments as were used in proving equation
(11-44). These rules can formally be represented by diagrams in which

for the singlet

two vectors of length

momentum
This

is

illustrated in

\, which must not be interpreted as angular


added to form a vector of length s =
or 1.
figure (12-2). For the singlet state, m' = 0. The
s

vectors, are

372

IDENTICAL PARTICLES

[Ch. 12

quantum number for the three triplet states is shown in


The properties of the states in which the spin eigenfunction
is <y
+ yt+ A or O-14-14 are easy to interpret. In these states, the z components
of the spin vectors of the two electrons are of the same sign. Thus the
spins combine into a total spin with z component for m' = 1 which is
s
twice that of a single electron. The quantum number specifying the
magnitude of the total spin vector for these two states would then
necessarily be s' = 1. For the states in which the spin eigenfunction is
value of this
table (12-1).
i

(1/V2)(e7

+H_ H

a_ H+ii ) or (l/V2)(o- _ K
+K

TABLE
Quantum Numbers
Spin eigenfunction

a_,

A+}4 ),

it

is

clear that

(12-1)

for the Singlet and Triplet States


s'

m's

triplet

+1

triplet

triplet

Designation

singlet

c +14+14
1

-'A-'A

we must have m' s


number

0.

Although

it is

not so apparent

-1

why

specifying the magnitude of the total spin vector

the antisymmetric state

and

s'

the

quantum

is s'

for

for the symmetric state, these results

The triplet and


which the spin vectors
of the two electrons are parallel and antiparallel, respectively. Although
it is not literally true, this description certainly contains an essence of
truth and provides a useful physical picture.
If the spin vectors of the two electrons are "parallel," the spin eigenfunction is symmetric and, to make the complete eigenfunction antisymmetric, the spatial eigenfunction must be antisymmetric. Let us consider
such a situation for a case in which the spatial coordinates of the two
x2 y1
z 2 Then y> a {\)
electrons are very similar, i.e., x l
y2 z x
fa {2)
and y 6(l)
Consequently
y(2).
follow directly from the theory of spin eigenfunctions.t
singlet states are often

spoken of as the

states in

Wat

These

Vo(l) Vk(2)

results are certainly also reasonable

from an

intuitive point of

view because

mean that all three symmetric spin eigenfunctions correspond to the same value of
quantum number s' = 1, i.e., to the same physical situation in which the electron

they

the

^ Vo(2) V&0) = Vi.0) Va(2) = fta

spin vectors are "parallel."

THE HELIUM ATOM

Sec. 5]

373

In this case the value of the antisymmetric spatial eigenfunction

~j= (fab

As a

result,

Vba)

^ ~j= {fab ~

fab)

the probability density will be very small

electrons are close together.

Since there

is

is

when

the two

only a very small chance of

finding the two electrons close together, the two particles act as
repel each other.

because

This has nothing to do with a

we assumed

Coulomb

if

they

repulsion

at the very beginning of this section that there

is

no

between the two electrons. Furthermore, it is easy to


the spatial eigenfunction describing the two electrons is

explicit interaction

show

that, if

symmetric, the probability density


they are close together.

Thus,

if

twice as large as average

is

when

the spin vectors of the two electrons

are "antiparallel," the spin eigenfunction

is

antisymmetric, the spatial

eigenfunction must be symmetric, and the two particles act as

if

they

a large chance of finding them close


together. We see that the requirement that the complete eigenfunction
be antisymmetric with respect to an exchange of the labels of the two
attract each other because there

is

and spatial coordinates.


They act as if they move under the influence of a force whose sign depends
on the relative orientation of their spins. This is called an exchange force.
It is a purely quantum mechanical effect and has no classical analog.
However, it does lead to directly observable results, such as those described
electrons leads to a coupling between their spin

in the next section.

The Helium Atom

5.

The

He

simplest real system containing

atom.

show

two

identical particles

is

the neutral

On the right side of the energy level diagram in figure (12-3), we

the energies of the ground state

and the

first

four excited states of

He. The energy of the ground state is obtained experimentally by measuring the energy required to ionize He to form He+. The sum of this ionization energy and the ground state energy of the one-electron atom He+,
which is 54.1 ev, is just equal to the energy of the ground state of He.f
The energies of the excited states relative to the ground state are obtained

from an

To

analysis of the

treat

He

He

spectrum.

with the theory of quantum mechanics,

we assume

first

t The magnitude of the energy of the ground state is thus denned as the energy required
to ionize completely the He atom to form He 2+ In spectroscopy it is, however, conventional to define the energy of the ground state as the energy required to ionize singly
.

the atom; a spectroscopist

would put

E=

at the value

we

label

E=

54.1

ev.

IDENTICAL PARTICLES

374

[Ch. 12

that the two electrons are

Coulomb

potential,

moving under the influence of the same nuclear


but that they are non-interacting. The interaction

between the two electrons is then taken into account by means of perturbation theory. Thus, we consider an unperturbed system whose energy
u

-10

-20

-30

-40

-50

.N

-60

-70

-80

-90

-100

-110
Figure 12-3.

Right:

An energy

level

diagram showing the ground state and the first


Left: The energy levels predicted by ignoring

four excited states of the helium atom.


the electron-electron interaction.

states are given

we

by the sum of two expressions of the form (10-38). That is,

take

E = - Z 2e4/2^ 2f - aZ e4/2A 2l
= 1, 2, 3,
Z = 2. In
2
2

/M

(12-26)

where n x

1, 2, 3,

rt

the

ground

state

375

THE HELIUM ATOM

Sec. 5]

quantum numbers n x and n 2 are both equal to 1. In


and the
one of the quantum numbers is equal to
the
approxithis
in
states,
these
for
predicted
energies
2.
The
other is equal to
mation, are shown on the left side of the energy level diagram of figure
of this system, the

first excited state,

poor descripis apparent that the unperturbed system provides a


that
the intermeans
This
system.
actual
the
of
levels
energy
of
the
tion
behavior
the
in
role
important
an
play
electrons
the
two
between
actions
(12-3). It

of the

He

atom.

The interactions consist primarily of the Coulomb repulsion which the


two negatively charged electrons exert on each other. It produces a
positive potential energy -which raises the total energy of the atom. In
addition, there are magnetic interactions involving the spins of the elec-

trons; but they are very small

and can be

neglected.

to the interaction of their charges

compared

Thus we take

for the perturbing potential the

quantity
v

(- eX~ e >

i!

(12-27)

r 12

r 12

between the two electrons. For unperturbed


use the antisymmetric eigenfunctions (12-22)
and, for each energy state of the unperturbed system, there will, in general,
be a number of degenerate unperturbed eigenfunctions. The degeneracy
of the energy
is due to exchange degeneracy, and also to the degeneracy

where

r 12 is the distance

we must

eigenfunctions

states of the individual electrons

and w r
For the ground

with respect to the quantum numbers

unperturbed system, there is, however,


only a single eigenfunction. The lm degeneracy is absent because when
n x = 2 = 1 the other spatial quantum numbers can only have a single
Furthermore, since all the spatial
value /x = m = /2 = m, =0.
quantum numbers for both electrons are the same, we see from equation
state of the

(12-22) that exchange degeneracy

is

absent as there

is

only one possible

antisymmetric eigenfunction

The ground

state of the

unperturbed system can only be the singlet

state

because the spatial part of the antisymmetric eigenfunctions, the use of


which is required by the exclusion principle, vanishes identically for the
triplet states

are the same.

the

when all the spatial quantum numbers for the two electrons
Of course, the fact that all the spatial quantum numbers are

same does not contradict the exclusion

quantum numbers

are different.

We

principle because the spin

see that the properties of spin enter

376

IDENTICAL PARTICLES

[Ch. 12

even in the properties of the spatial part of the unperturbed


However, when we use the eigenfunction to calculate the

implicitly,

eigenfunction.

from the perturbation, it is not necessary to consider


Because the perturbation does not involve spin, the spin
eigenfunctions enter only as a product in the term tp*vy> which must be
energy

shift resulting

spin explicitly.

evaluated in the calculation. Since the spin eigenfunctions are normalized,


this

product

is

function,

Consequently in the present calculation

equal to unity.

we may simply ignore

the spin part of the complete antisymmetric eigen-

and take for the properly normalized unperturbed eigenfunction

Waa

= U^e-^-e-*""*

Vioo(l) Vioo(2)

where we have evaluated

ip
lm from (10-39). Then, according to the three
dimensional extension of equation (9-13), the energy shift due to the
perturbation is

E'

- E~
J J

v4>(!)

Vi*oo( 2 )

ViooC 1 ) Vioo( 2 )

dri dr2

r 12

where

drx
dr2

=
=

r\ sin 6 1
rjj

sin

This integral has been evaluated.

dr x dd 1

dr 2 dd 2 dj> 2

It gives

E '-E~

5
-Z>^
2

Combining

this result

(-2Z 2 +
\

The value of E'

2/j

with (12-26) for n x


E' =

calculated

from

dj>

this

n2

1,

we have

Z)^

(12-28)

/2/7 2

equation for

Z=

2 agrees with the

energy of the ground state of He to within 5 percent. Perturbation theory


gives a reasonably accurate result, even though the perturbation is ob-

Much more accurate results can be obtained by using


a modification of perturbation theory in which the error is of the order of
the cube of a small quantity. Expression (12-28) also predicts the ground
viously not small.

state energy of a two-electron ion with Z > 2, such as Li+, Be 2+ B 3+


The agreement improves with increasing Z because the Coulomb interaction between the nucleus and each electron increases more rapidly than
the Coulomb interaction between the two electrons, so the perturbation is
smaller and the perturbation calculation more accurate.
,

Next consider the

first

excited state of the unperturbed system for He.

THE HELIUM ATOM

Sec. 5]

This

is

377

a highly degenerate state because there are four degenerate one2, and there are also exchange degeneracies.

electron states for n

However, much of the degeneracy is removed when the interaction between


the two electrons is applied as a perturbation. Before considering this
quantitatively, it is worth while to describe what happens qualitatively.
The increase in the energy of the atom produced by the Coulomb
repulsion between the two electrons depends on the average distance
between the two electrons. This, in turn, depends on the quantum numbers
of the spatial part of the eigenfunction, and also on its symmetry character.
Let us consider first the dependence on the spatial quantum numbers,
ignoring temporarily the symmetry properties and the requirements of

Then we may

indistinguishability.

numbers

for the

two electrons

list

in the first

the possible sets of

quantum

excited state of the unperturbed

system as follows:
!

nx
"i
"i

=
=
=
=

1,

1,

1,

1>

= 0, m h = 0;
= 0, m h = 0;
= 0, m h = 0;
= 0, m h = 0;

=
=

n2
n2

=
=

n2
n2

2,

12

2,

2,

2,

Since the probability density of electron

=
=
=
=

is

0,
1,
1,
1,

mh
mh
mh
mh

=
= -1
=

spherically symmetrical,

the average distance between the two electrons depends only


probability densities for the

two

particles.

on the

radial

These quantities, denned in

equation (10-43), are specified completely by the quantum numbers n

and

By

/.

inspecting figure (10-2), the reader will see that the average

distance between the two electrons

tum numbers

0;

1,

2, /2

1.

/jj

0; n 2

is

somewhat

2, /2

larger for the set of quan-

than

it is

for the set n 1

Since the perturbing potential

is

I,

inversely pro-

portional to the distance between the two electrons, the total energy of the

atom
2

will

= 2,
= 0.

/2

be increased more for the state specified by n 1 = 1, 4


1 than for the state specified by Hj = 1, lx = 0;
n2

= 0;
= 2,

l2
Thus the degenerate energy level for the first excited state of the
unperturbed system is split into two components by the perturbation. If
we now take into account the requirements of indistinguishability, we shall
see that each of these components is split again into two components.
This is a result of the fact that each of the states, such as the one specified

by the quantum numbers nr

1, lx

0; n 2

2, l2

0, is

comprised of

four exchange degenerate states corresponding to the four antisymmetric

complete eigenfunctions (12-22) which must be used to describe the system


of two electrons. The splitting occurs because the average distance between
the two electrons is smaller in the singlet state than it is in the triplet states,
so the increase in the energy of the atom is larger for the singlet state than

378
it is

IDENTICAL PARTICLES

The net

for the triplet states.

increase the energy of the

first

effect

of the perturbation

[Ch. 12
is

thus to

excited state of the unperturbed system

and split the energy level into four components. In the energy level
diagram of figure (12-3), the reader will see that this conclusion is qualitaagreement with the experimentally observed energy levels of He.
Let us use perturbation theory to evaluate quantitatively the effect

tively in

of the perturbation. This will not be so

difficult as the

reader might guess,

for the following reasons:


1

the

As in the ground state calculation, we may neglect the spin part of


antisymmetric complete eigenfunctions and use for unperturbed

eigenfunctions only the spatial part.


2. Since the total angular momentum of the atom (the sum of the orbital
and spin angular momenta of the two electrons) cannot change if there are
no external torques, and since the perturbation does not cause interactions
between the spin angular momenta and the orbital angular momenta, it

momentum nor the total


can change as a result of the application of the
Therefore the perturbation cannot induce appreciable

follows that neither the total spin angular

momentum

orbital angular

perturbation.

mixing between the degenerate states specified by n x = 1, /x = 0; n 2 = 2,


/2 = 0, for which the magnitude of the total orbital angular momentum

and the degenerate states specified by n x = 1, /x = 0; 2 = 2, /2 =


for which the magnitude of the total orbital angular momentum
is 0,

V 1(1 +
these
3.

1) h.

two
For

sets

Consequently

it is

1,
is

possible to treat the degenerate states for

of quantum numbers separately.

simplicity,

we

shall actually carry

for the degenerate states specified

by n 1

through the calculation only

\,

/x

0;

n2

2,

l2

0.

For these states there is no spatial quantum number degeneracy since


m and m can have only the single value m = m u = 0. There is,
however, an exchange degeneracy which arises because the spatial part
of the singlet and triplet eigenfunctions,
l

~j=

[Wai>

+ W] =

-j= [Vioo(l) %oo( 2 )

VaooC 1 ) Vioo(2)]

(12-29)

L>ioo(l) Wwoi 2 )

^ooC 1 )

Vioo( 2 )]

(12-29')

and
-j= [fab

Wba]

-ft

correspond to the same total energy for the unperturbed system.


4. As stated in 2, the total spin angular momentum of the atom cannot
change as a result of the application of the perturbation. For this reason
the perturbation cannot induce appreciable mixing between the degenerate
eigenfunctions (12-29) and (12-29') because the first corresponds to the

THE HELIUM ATOM

Sec. 5]

379

which the two spin vectors are "antiparallel" and the


magnitude of the total spin angular momentum is 0, while the second
corresponds to the triplet states in which the two spin vectors are "parallel"
singlet state in

and the magnitude of the


Therefore

it is

total spin angular

momentum

is

^1(1

1)

fr.

due to the perturbation

possible to evaluate the energy shift

by applying ordinary perturbation theory to each of the eigenfunctions,


even though they are degenerate, just as we did for the degenerate eigenfunctions

y nl j m when

calculating the spin-orbit energy shift.

Thus we use

the three dimensional extension of (9-13) with the pertur-

bation (12-27) and the unperturbed eigenfunctions (12-29) and (12-29'),

and obtain for the energy


E'

- E~i

due to the perturbation the expression

shift

f f [ Vl*oo(l) %*oo(2)

2J

X
corresponds to the

~E+

V2*oo(l) Vi*oo(2)]

r 12

where E'+ corresponds to the

Ol(K)(l) ^20o( 2 )

VW

1)

VlOo( 2 )]

drl dr2

by (12-29) and E'_

described by (12-29'). This

triplet state

singlet state described

is

flooil) V20o(2)
r iWV^ %oo(2) ~
rv

2 J J L

1 00

'

'

""""

t.

'

'

12

V2*oo(l)

Vlooi 2 )

V200C 1 ) Vioo(2)

dr x di

'12
2

Vioo(l) V2oo(2)

J/J[-

^2oo(l) Vioo(2)

r 12

VJ 2oo( 1 ) Vioo( 2 )

^100(1) VaooC 2 )

^12

d 7l dT2

(12-30)

The reader should check

these results by performing a routine degenerate


perturbation theory calculation, as described in section 4, Chapter 9, for

the case of two degenerate eigenfunctions

This

is

y100 (l) V2oo(2) and V2oo(0 Vioo(2)easy to do because the calculation carried out in that section

can be immediately applied to the present problem. Equations (9-26)


and (9-27) show that (12-29) and (12-29') are exactly the right linear
combinations of y100 (l) V2oo( 2) and
1)
yioo(2), and that (12-30) is the

vW

correct expression for the perturbed energy.

Let us compare the results of our calculation with equations (12-10)


and (12-1 l),and with the interpretation which was given to these equations.
This comparison will identify the first integral in equation (12-30) as an
expression which, from classical considerations, would be written for the
expectation value of the perturbation in a system containing two identical
particles in the one-particle states n = 1, / =
and n = 2, / = 0. The
same comparison identifies the second integral as a manifestation of the

380

IDENTICAL PARTICLES

quantum mechanical requirements of

indistinguishability.

[Ch. 12

This integral

takes into account the effective attraction or repulsion in the singlet or


triplet states
it

by increasing the energy

for the latter.

The second

shift for the

integral
n,

I:

is

n,

former and decreasing

called the exchange integral,

and

Spin

singlet
1, 0; 2, 1

"

triplet

singlet
1, 0; 2,

triplet

1;2

//

1,

i;i

singlet

0; 1,

Unperturbed

Perturbed

Perturbed

(ordinary

(ordinary
plus

integral

exchange

only)

integrals)

Figure 12-4.

and the

first

Illustrating

the effects which contribute to the energies of the ground state

four excited states of the helium atom.

the

first

we

evaluate

integral
y100

is

called the ordinary integral.

and

y^ from (10-39),

To

evaluate these integrals

which gives

-Zr/a

Vioo

By

V2O0

J7r\aj

substituting these into (12-30),

4J277W

Zr/a

and carrying through the

integration,

values for E'_ and E'+ are found. These values agree with the experimentally

observed energies of the

first

two excited

states

of

He

to about

percent.

Sec. 6]

THE FERMI GAS

The perturbation

381

calculation

is

more

accurate for the excited states than

ground state because the average distance between the electrons


is larger and the perturbation is weaker.
A perturbation calculation for the unperturbed states, in which one set
of quantum numbers is n = 1, / =
and the other set is n = 2, I = 1,
leads to results which are similar to (12-30) except that the numerical
value of the ordinary integral turns out to be somewhat larger. The
important features and results of the theoretical description of the first
few energy levels of He are represented schematically by the set of energy
level diagrams shown in figure (12-4).
Not shown in the diagram is a fine structure splitting of the triplet
energy levels, the existence of which is inferred from the spectrum of He.
It is due to the very small interactions involving the spins of the electrons,
and other relativistic effects, that split each of the triplet levels (except
those in which the total orbital angular momentum is zero) into three
levels with a separation about 10~ 4 times the separation between associated
singlet and triplet levels. A theoretical treatment of this effect can be
given in terms of the equations developed in the previous chapter. There
is also a hyperfine splitting of the levels, due to extremely small interactions
involving the spin of the nucleus, which produces a separation of about
10~ 3 times the separation produced by the fine structure splitting, and
which can be treated by methods to be developed in Chapter 13. But we
for the

shall

Of

not treat either of these


particular interest

is

effects here.

the absence of a triplet energy level associated

with the singlet ground state

level,

which

in

our treatment we have shown


This is quite true, but

to be a consequence of the exclusion principle.


historically the

argument was made in the opposite order. The experi-

mental fact that the He spectrum shows this triplet to be absent provided
the primary evidence that led Pauli to the discovery of the exclusion
principle.

6.

The Fermi Gas


All particles which

are given the generic

must be described by antisymmetric eigenfunctions


in honor of a man prominent

name Fermi particles,

in developing the theory of their properties. In the last section

we presented

a detailed discussion of a system containing two Fermi particles.

we shall discuss, in somewhat


number of these particles.

less detail,

Here

a system containing a large

Consider a system in which a "gas" of freely moving, non-interacting,


Fermi particles is confined to a certain region. Such a system, which is

IDENTICAL PARTICLES

382

[Ch. 12

Fermi gas, can be used to describe some of the important features


of the behavior of the electrons confined to a multi-electron atom. A
Fermi gas can also be used to describe certain features of the motion of
the protons and neutrons confined to a nucleus. Possibly the best physical
example of a Fermi gas is the conduction electrons in a metal. These

called a

which are responsible for the electrical conductivity of the metal,


it quite freely because the forces exerted by the various
through
move
atoms tend to cancel, and the forces which the electrons exert on each
other are small. However, the conduction electrons are confined to the
metal by non-canceling attractive forces at the surfaces. In this section

electrons,

shall develop some of the properties of a Fermi gas, such as its total
energy and the energy distribution of the individual particles, and shall
illustrate their significance for the conduction electrons. We shall see that

we

by the exclusion principle.


moving independently in a potential
we define as zero) inside a certain region, and

these properties are very directly influenced

Each

particle of a

Fermi gas

is

which is a constant (that


which becomes larger than the total energy at the boundaries of the region.
In most cases it is an adequate approximation to assume that the potential
becomes infinitely large at the boundaries. Then the identical Schroedinger
equations describing the motion of each of the identical particles assume
a particularly simple form,

+ & + )_*,
_(
2m\dx
dy
2

in

which y(x,

y, z)

y,(x, y, z)

2-31>

at the boundaries of the region.

The reader may

equation can be written

easily verify that solutions to this

in

dz

X{x) Y(y) Z(z)

(12-32)

which

_tf_#J^ =
2m
and

similarly for Y(y)

and

Z(),

(12_ 32

ExX{x)

dxr

and where

E=E

+ Ey + E

not restrict the generality of our calculation to assume that the


the Fermi gas is a cube of side a. Then, defining the
containing
region
coordinates
at the center of the cube and the axes parallel to
of
the
origin
It will

its

edges, the condition

X{x)

and

similarly for Y(y)

on X(x)

0,

is

<

-a/2,

>

a/2

(12-33)

and Z(z). Equation (12-32') and condition (12-33)


and (8-90), which describe the problem of a one

are identical with (8-89)

.:

THE FERMI GAS

Sec. 6]

dimensional
values of

Ex

383

2fc2

2,

1,

3, 4,

The

1, 2, 3,

three spatial

4,

Ey and E

energy of one of the identical particles

total

+; + !)
^("2
2ma

=
where n x

(12-34)

Similar expressions obtain for

Thus the allowed values of


is given by the equation

= ~^i

Ex
where nx

Consequently the allowed

square well potential.

infinite

can be written directly from (8-96), and we have

(12-35)

ny

1,

2, 3, 4,

quantum numbers n x n y n z
,

nz

specify the

1,

2, 3, 4,

quantum

states

of the particle.
It is

useful to calculate

number of quantum
within the limits

section 6, Chapter
in

from

this

equation the quantity N(E) dE, the

which the

states for

energy of the particle

total

is

E + dE. The procedure is similar to that used in


2.
We imagine a uniform cubic lattice constructed

to

one octant of a rectangular coordinate system in such a way that the

three coordinates of each point of the lattice are equal to a possible set of

quantum numbers n x ny n z Then the number of possible sets of quantum


numbers is equal to N(r) dr, the number of points contained between
concentric shells of radii r and r + dr, where
,

^r

E =
By

construction, N(r) dr

is

just equal to the

N(r) dr

- 477^ dr

(12-36)

volume enclosed by the

shells

According to equation (12-36),

dE

==

dr

ma

and
E'

ma

so

'

2
Since N(r) dr

irH

7T

N(E) dE, we have


JV()

dE

-g-il E* dE
n

(12-37)

IDENTICAL PARTICLES

384

where we write
particle of mass

V=

equation (12-37)
states in the

cP for the

volume of the cubical region

m and kinetic energy E is confined.!


is

the correct expression for the

E to E + dE for

energy range

It

in

[Ch. 12

which the

can be shown that

number of quantum

the general case of a particle

V and any arbitrary shape.


The eigenfunction (12-32) describing the motion of one of the identical
particles is a standing wave with nodes at the boundaries of the region.
The eigenfunction describing the entire system of particles is composed of
products of such standing wave eigenfunctions and, to satisfy the exclusion
principle, it must be written in the antisymmetrized form (12-16). If we

confined to a region of volume

discuss only the energetics of a

Fermi

gas,

it

will

not be necessary to deal

&
Without exclusion

With exclusion

principle

principle

Figure 12-5.

Illustrating

fat

the effect of the exclusion principle.

with this eigenfunction. However,

it

will

be very necessary to take account

This can be done simply by imposing the

of the exclusion principle.

no more than two particles have the same spatial


quantum numbers n x nv n z The exclusion principle will be satisfied
because one particle can have the spin quantum number m s = \ and
the other can have m s = +|.
auxiliary requirement that
,

It is

easy to see that the exclusion principle plays an important role in

determining the

minimum

total

energy of the Fermi gas. Consider the


two diagrams of figure (12-5).

situation illustrated schematically in the

The diagram on

the left represents the occupied

quantum

states

when a

system of eight identical particles, which do not obey the exclusion


principle, has the

minimum

possible total energy.

All the particles

would

be in the state of lowest energy. The other diagram represents the occupied
quantum states for the minimum possible total energy of a system in which

do obey the exclusion principle. Only two particles can be in


same spatial state, and so the total energy of the system is considerably
larger. The diagrams also illustrate the way in which the number of
For a system
quantum states per unit energy increases according to E
of N Fermi particles, the lowest total energy is achieved when there are
and E = Et
two particles in each of the energy states between E =
the particles
the

l/i

t Since the potential energy


kinetic energy.

is

zero in the region, the total energy

E is

equal to the

THE FERMI GAS

Sec. 6]

The

particles

fill

385

up

the states

to the

Fermi energy

to the

number of particles

That

in the system.f

Ef

This energy

dE from

evaluated by requiring that the integral of 2N(E)

to

may be

Et be equal

is,

N=

2N(E) dE
Jo

The energy

Figure 12-6.

The

distribution for a system of Fermi particles at zero temperature.

factor of 2 enters because there are

state.

Evaluating N(E)

dE from

two

(12-37),

particles per spatial

2 H m 3A vC E<i
T

quantum

we have
,

Jo

Thus

Ef =

2m \F/

or

E f = 3V*i
where p

is

the

number of

particles with energies

gas

is

just

(12-38)

particles per unit

volume.

The number of

E and E + dE in a minimum energy Fermi


E < Er As shown in figure (12-6), there are no

between

2N(E) dE for

particles with

2m

_2_ JA

E > Ef We leave it
.

as an exercise for the reader to prove

that the average energy of the particles in this energy distribution

= \Et
Thus the

total

energy of the Fermi gas

is

(12-39)

is

E = NE = %NEt
t

(12-40)

t The use of an integral is justified if we are treating a system containing a large


number of particles. Then the integral is a good approximation to the sum over the
discrete energy states which, strictly speaking, should be used.

IDENTICAL PARTICLES

386

[Ch. 12

should be noted that, for Fermi particles of a given mass, Ef and E


depend only on the density of the Fermi gas.
Let us evaluate some of these quantities for the Fermi gas of conduction
electrons in a metal. Consider Cu, which has one conduction electron
per atom. In this case p is equal to the density of atoms, which is Avogadro's
number times the density of Cu divided by its atomic weight. Thus
It

Figure 12-7.
states

The

10 23

x 9/64

10 22

cm"3

potential experienced by the conduction electrons in copper, and the

which these electrons

Using

this value

of

p,

fill.

and

setting

equal to the mass of the electron,

equation (12-38) gives

>
This

is

10

x 10" 12

ergs

~ 7 ev

the energy of the highest filled state relative to the

potential.

the metal

measured

The energy of

this state relative to the potential

just the quantity

is

W, appearing in equation
For Cu it is

bottom of the
energy outside
(3-8), that

is

in the photoelectric effect.

PK~4ev
Knowing Ef and W, we may draw

a diagram representing the potential

energy in which the conduction electrons move, and also the energy states
which they fill, as illustrated in figure (12-7). For a 1 cm3 sample of Cu
there are

~4

10 22

filled

energy states!

This fact

justifies the

approxi-

mation discussed in the footnote on page 385. The approximation of


assuming that the potential is infinitely large at the boundaries is justified
by the fact that the highest filled state is far below the top of the potential.
The conduction electrons of Cu will have a minimum total energy when
the total energy content of the metal

temperature

is

at absolute zero.

of the conduction electrons

is

itself is

But even

at

minimum,

T=

E = 3Ef/5 ~ 4 ev.

This

i.e.,

when

the

the average energy


is

a direct result of

Sec. 6]

THE FERMI GAS

the exclusion principle.

obey

387

For a system composed of particles which do not


gas of distinguishable particles),

this principle (e.g., a classical

the particles will be in the ground state

when

For such

their energy will be essentially zero.

the system

is

particles to

at

T=

0,

all

and

have an average

energy of 4 ev, the energy must be supplied by thermal agitation and the
required temperature satisfies the equation

%kT

= f Ef ~

4 ev

where k is Boltzmann's constant. This equation gives the high temperature

T~

10 4

which can be taken as a measure of the temperature equivalent of the


average energy of conduction electrons.
The fact that Fermi energy Ef is very much larger than thermal agitation
energy at normal temperatures,

(ffcZW*
leads to

some

~ 0.025 ev
For a Fermi gas

interesting consequences.

the states below the Fermi energy are completely

of the metal containing the Fermi gas

is

filled.

raised, there

the electrons to be excited to higher energy states

thermally agitated atoms.

by

at

T=

is

a tendency for

collisions with the

But, according to the exclusion principle,

make a

all

As the temperature

it is

which is not
filled.
Since at normal temperatures the thermal energy which can be
supplied to an electron is quite small compared to the Fermi energy, only
the electrons whose energies are very near Ef can be given enough energy
to allow a transition to an unfilled state at E > Ef Electrons with energies
appreciably less than Ef cannot be excited because there is not enough
only possible for an electron to

transition to a state

thermal energy available. Consequently, only a small fraction of the total

number of conduction

electrons can take part in the thermal excitation,

and the contribution of the electrons to the heat capacity of the metal
will be very small. For this reason the heat capacity of a metal at normal
temperatures is observed experimentally to be just what would be calculated
from the thermal motion of the atoms alone. When this explanation was
first proposed, it answered a long standing puzzle because, in classical
physics, the heat capacity arising from the thermal motion of the conduction
electrons should be quite comparable to fhat arising from the thermal
motion of the atoms.
Figure (12-6) shows the energy distribution for the particles of a Fermi
gas at T = 0. Energy distributions for a Fermi gas at several other values
of Tare shown in figure (12-8). Note that at each temperature the drop
to zero occurs in an energy range of width about equal to kT, because

388

IDENTICAL PARTICLES

[Ch. 12

particles of energy much less than Ef


kT cannot be excited so the
energy distribution of these particles cannot be different from the T =

distribution.

The curves

in this figure are plotted

Jf(E) dE

The quantity N(E)is


proportional to

l/i
.

2N(E) F(E) dE

(12-41)

the density of states for Fermi particles (12-37)

The quantity F{E) is

The energy

Figure 12-8.

from the equation

and is

the Fermi probability distribution

distributions for a system of Fermi particles at several tem-

peratures.

which

specifies the probability that a Fermi particle, in a system in


thermal equilibrium at temperature T, will be in a state with energy E.
It is

F
where

is

W-

eiE - E !)lkT

a constant which depends on

when &r< Ef
The role of

(12-12)
1

T in

such a way that

E = Ef

the Fermi probability distribution in the description of a

system of Fermi particles

is

analogous to that of the Boltzmann probability

distribution (2-23) in the description of a system of classical distinguishable


particles.

The

latter probability distribution

P()

,E/kT

(12-43)

specifies the probability that a particle of such a system, which is in thermal


equilibrium at temperature T, will be in a state with energy E. The

difference
reflects

between the probability distributions (12-42) and (12-43)


behavior of Fermi particles and

the difference between the

THE FERMI GAS

Sec. 6]

classical particles

and

a Fermi gas with

389

primarily an effect of the exclusion principle.

is

kT > Ef

the thermal excitation

is

very large,

all

For
the

particles are highly excited, and the states are very sparsely occupied.

Then

the chance that a Fermi particle would be prevented from

was already
no longer important. But

transition to a state, because that state

the exclusion principle


the constant
is

is

in equation (12-42) is large

filled, is

making a

so small that

in the limit

kT > Ef

and negative and the term

compared to the exponential. Then the probability distribuF{E) and P{E) are identical except for a multiplicative constant

negligible

tions

which only affects their normalization.


To complete this discussion, we mention that particles which must be
described in quantum mechanics by symmetric eigenfunctions are given
the generic name Bose particles. Even though they are indistinguishable,
these particles do not obey the exclusion principle because a symmetric
eigenfunction, such as (12-8), does not vanish if all the quantum numbers
of two or more particles are the same. The probability that a particle
in a Bose gas in thermal equilibrium at temperature T will be in a state
with energy E is given by the Bose probability distribution,

B
In general, the constant

W-

e,

E - Eo

L_

depends on

in

(12-44)
t

such a way that the Bose

probability distribution becomes identical, within a multiplicative con-

Boltzmann probability distribution at high temperatures.


However, the two probability distributions differ at low temperatures
where a large number of particles occupy the same state since, even though
they both describe particles which do not obey the exclusion principle,
the particles are indistinguishable in the Bose case and distinguishable
in the Boltzmann case. The He atom is an example of a Bose particle
because it is a tightly bound aggregate containing an even number of
stant, with the

antisymmetric particles. The extremely interesting properties of a system

of He atoms at low temperatures (liquid He) are due primarily to the


symmetry character of these atoms. The quantum is another example of a
Bose particle. In this particular case the constant in equation (12-44)
is identically zero, and the probability distribution is

B() quan ta

By setting E

J_

(12-45)
^

equation can be used to evaluate the average energy


v in an enclosure containing a large number
of quanta in equilibrium at temperature T. The resulting expression is
identical with Planck's equation (2-31) for the average energy contained
carried

hv, this

by quanta of frequency

IDENTICAL PARTICLES

390
in electromagnetic standing

density of a black

body

waves

in such

[Ch. 12

an enclosure. Thus the energy

cavity can be evaluated

by thinking of

it

as con-

taining a system of standing waves, or as containing a Bose gas of quanta.

In Chapter 2

we showed with a simple system how

probability distribution
ics.

is

developed from the theory of

The Fermi and Bose

the Boltzmann

statistical

mechan-

probability distributions are developed from

a variant of that theory, called quantum statistical mechanics, which differs

from the

classical theory only in the definition

of an experimentally

dis-

tinguishable division of the total energy of the system and, for Fermi
particles, in the application

of the exclusion principle. This

is

an important

quantum mechanics which we shall not treat here only


would divert us from its even more important application to

application of

because

it

the study of multi-electron atoms.

BIBLIOGRAPHY
Bohm,

Quantum Theory, Prentice-Hall, Englewood Cliffs, N.J., 1951.


Quantum Mechanics, McGraw-Hill Book Co., New York, 1955.
Sherwin, C. W., Introduction to Quantum Mechanics, Henry Holt and Co..
D.,

Schiff, L.

New

I.,

York, 1959.

EXERCISES
1.

2.

Show

that equation (12-3)

is

a solution to (12-1).

Using the procedure suggested immediately below (12-30), show that the

expression agrees with the results of a routine degenerate perturbation theory


calculation.

Consider two electrons moving in a narrow impenetrable walled tube of


Write proper eigenfunctions for the ground and first excited states
of this system, assuming that the electrons are non-interacting. Evaluate the
energy of the ground state and first excited state of the system in this approximation, and discuss the degeneracy of these states. Now consider the perturbation due to the Coulomb interaction between the electrons. Write equations
3.

length a.

for the energy shifts of the ground state


shifts.

and

first

excited state,

Explain the physical origin of the splitting of the

and evaluate these


excited state by

first

the application of the perturbation.


4.

Verify equations (12-31) and (12-32).

5.

Verify equation (12-39).

Repeat the calculation of exercise 3, Chapter 2 for a system of Fermi


and also for a system of Bose particles. Compare the first calculation
with equation (12-42) and the second with (12-44), and compare both with
6.

particles,

(12-43).

CHAPTER

13

Multi-electron

Atoms

Introduction

I.

A multi-electron atom
Z electrons of charge

by

contains a nucleus of charge

+Ze

surrounded

Each electron moves under the influence


of an attractive Coulomb force exerted by the nucleus and repulsive
Coulomb forces exerted by all the other Z 1 electrons, as well as certain
other forces involving the spin angular momenta which are usually weaker
than those involving the charges. To find an exact solution of the extremely
complicated
impossible.

e.

Schroedinger equation for such a system

is

essentially

Nevertheless, there exists a completely adequate theoretical

description of multi-electron atoms.

In section

by

first

5,

Chapter

12,

we

treated a multi-electron

atom with

Z=

considering an unperturbed system describing a simplified

atom

in which the interaction between the electrons is ignored, and then improving the accuracy of the description by using perturbation theory to

evaluate the effect of this interaction.


larger values of

atom

The modus operandi

for

atoms with

also consists in taking a simplified description of the

an unperturbed system, and then using perturbation theory to


However, if there are a number of electrons,
their interactions cannot be completely ignored in the unperturbed system
because the effect of these interactions is much too large to be treated as a
perturbation. A useful unperturbed system must take into account the
main effects of the Coulomb interactions between the electrons. On the
other hand, the unperturbed system must not be so complicated that the
as

evaluate the corrections.

Schroedinger equation describing the system is insolvable. In effect, the


requirement means that in the unperturbed system the electrons

latter

391

MULTI-ELECTRON ATOMS

392

only

must be non-interacting

[Ch. 13

then can the Schroedinger equation for

Z equations which can


because each of them involves the

the unperturbed system be separated into a set of

be solved without too

much

difficulty

coordinates of a single electron.

An unperturbed system with the required properties can be obtained


by assuming that each electron moves independently in a spherically
symmetrical net potential V(r), where r is the radial coordinate of the
electron with respect to the nucleus.

the spherically symmetrical attractive

This net potential

Coulomb

is

the

sum of

due to the
nucleus and a spherically symmetrical repulsive potential which represents
the average effect of the repulsive Coulomb forces between the electron
and its Z 1 colleagues. The average repulsive potential depends on
potential

the average radial probability densities of the electrons.


probability densities are not

known

Of

course, these

until after the equations

the behavior of the unperturbed system have been solved.

governing

This might

seem to be an intractable situation, but actually it is not. It can be handled


by demanding that the average repulsive potential be self-consistent. That
if

is,

we calculate

the probability densities for the correct average repulsive

and then evaluate an average repulsive potential from these


probability densities, we demand that the potential with which we end up
must be identical with the potential with which we started. This condition
potential,

is sufficient

Two
2,

to determine the correct average repulsive potential.

different self-consistent treatments will be presented.

In section

the electrons in the unperturbed system will be treated as a Fermi gas.

Then

it will not be necessary to solve the Schroedinger equations for the


unperturbed system explicitly, because many quantum mechanical properties of a system of non-interacting electrons have already been included in

the derivation of the equations specifying the properties of a Fermi gas. In


section 3, we shall describe a more accurate treatment of the unperturbed
system which involves solving the Schroedinger equations for the spherically symmetrical potentials. The results of this treatment will provide the
basis for a discussion of the

ground

state of multi-electron

atoms, and of

Then we shall use perturbation theory


effect of the Coulomb interaction between

the periodic table of the elements.


to correct for the fact that the

the electrons

is

only approximately represented by the spherically sym-

metrical repulsive potential, and also to correct for other interactions

such as those involving spin. This will allow us to give a description of


Finally, we shall use time dependent perturbation theory to evaluate the probability that an atom will make a
the excited states of atoms.

from one excited state to another. From all these considerations,


be able to understand the optical and X-ray spectra of multielectron atoms.
transition

we

shall

Sec. 2]

2.

THE THOMAS-FERMI THEORY

393

The Thomas-Fermi Theory

Thomas and Fermi independently developed a simple


The theory is based on
treating the electrons as a minimum energy Fermi gas confined to a region
by a spherically symmetrical net potential energy V(f). Then equation
About

1928,

theory of the ground state of multi-electron atoms.

(12-38) can be used to relate the depth of the potential, at any value of

r,

Fermi gas for that value of r. This relation is obtained


by assuming that the depth of the potential is everywhere such that the
energy levels are filled to the very top. That is,
to the density of the

Et =-V{r)

(13-1)

This condition assures that the total energy of the atom


Solving (12-38) for the density p

p
This procedure

= JH-

is

minimum.

\_-V(r)-p

(13-2)

the distance in which the value of V(r) does

is justified if

not change significantly

is

we then have

large enough,

compared

to the de Broglie

wavelengths of the electrons, that a number of electrons can be localized

volume in which V(f) is essentially constant. The condition is satisfied


atoms of large Z, except very near the center and very far from the
center. The theory does not apply for atoms of very small Z because the

in a

for

condition

is

not

satisfied

anywhere.

obtained by requiring that the electron density be


such as to reproduce the net potential V{r). The potential is evaluated in
terms of p by using Gauss' law. This law states that the integral over any
Self-consistency

is

component of the electric field E times the


element of surface area dS is equal to 4w times the charge contained within
the surface. Consider a spherical surface of radius r centered on the
closed surface of the normal

Then we have a completely spherically symmetrical situation


which Gauss' law has the simple form

nucleus.
in

477r

The

left side

= 4n Ze +

of the equation

is

-ep{r')A7Tr' dr'\

(13-3)

the area of the surface times the electric

which is constant in magnitude and everywhere normal


to it. The first term in the bracket is the nuclear charge. The integral
is the total charge due to the electrons contained within the surface;
2
ep(r') is the charge per unit volume, and 4nr' dr' is the volume element.
field at the surface,

MULTI-ELECTRON ATOMS

394

[Ch. 13

Now the electric field is related to the force acting on an electron, and thus
on the

to the net potential V(r) acting

_ e =

by the equations

electron,

= _^(!)

/r

(13
V

dr

When we

_ 4)

write this as

rf[-Hr)]

dr

equation (13-3) becomes


r2

4zW] = _ Ze2 + 47re2


dr

p(r')r'

dr'

Jo

Differentiating both sides with respect to r gives

d
I=im=A^
U
y
dr\
dr
p{r

dr

'

Using equation (13-2) to express

p(r) in terms of

L-^)])
hT^Y
dr\
r

It is

dr

4^

V(r),

we have

L-V(r)T

(13-5)

5ttFi

convenient to rewrite this in terms of the variable

x,

defined by the

equation

and

to

make

the substitution

-V= X

(13-7)

In terms of these quantities, the differential equation (13-5) becomes

ud
**!&-&)
'

(13-8)

dx2

To
know

obtain an explicit solution to this differential equation, we must


"boundary conditions" for the function %(x). They are obtained

the

from the following considerations. For r>*0, it is clear that the net
potential energy of an electron approaches the Coulomb potential energy
due to the nucleus alone. That is,
V(r)

From

->

-Ze2 /r,

equations (13-6) and (13-7),

we

X (x)^l,

r-+0

see this

x^O

means
(13-9)

THE THOMAS-FERMI THEORY

Sec. 2]

For

->

charge

co,

+Ze

395

an electron experiences a potential energy due to the nuclear


shielded by the charge (Z \)e of the other electrons.

Thus

e?jr,

F(r)

>-

r~>oo

or
rF(r)

-e\

CO

1.0

0.8
0.6
0.4

-\

0.2

o.i

-L

-0.08

0.06
0.04

0.02

nm

The

Figures 13-1.

We make

plot of a function

which

14

12

16

arises in the

the approximation that this condition

rV(r)>-0,

>Nl

10

r->

oo

Thomas-Fermi theory.

is

(13-10)

This
which amounts to assuming that e2 is very small compared to Ze2
approximation is another reason why the Thomas-Fermi theory is not
accurate far from the center of an atom. Using equations (13-6) and
.

(13-7) to write (13-10) in terms of %(x),

x (x)-+0,

we have

*^co

(13-11)

The

solution of equation (13-8), with the boundary conditions (13-9)


and (13-11), has been carried through by several people. Numerical
methods are required, and so the results must be displayed as a table or
a plot. Such a plot is shown in figure (13-1).
The function %(x) determines the net potential V(r) and the electron
density p(r).

In the next section these quantities will be plotted for a

particular atom,

theory.

It will

and compared with the predictions of a more accurate


that, even for an atom of fairly small Z, the

be found

MULTI-ELECTRON ATOMS

396

Thomas-Fermi theory

good

gives a

[Ch. 13

description of the average radial

we

behavior of the net potential and the electron density. Here

shall

only

equation (13-6) says the factor connecting the


and the dimensionless variable x is proportional to

call attention to the fact that

radial coordinate r

Z~ ^.

Since the same dimensionless variable

means

Z~

to

that for increasing

The Thomas-Fermi theory

X/i
.

is

used for

atoms, this

all

the radial scale of p(r) contracts according


predicts that the "size" of

atoms

decreases slowly with increasing Z.

The Hartree Theory

3.

A more accurate theory of the unperturbed system for the ground state
atom was developed by Hartree

of a multi-electron
following years.

system of

moving independently
The equation is

=^~
2m
where Vf
electron

fr
is

we

ET

is

is

the potential energy of

the total energy of the atom,

y>

WriTi* #1,

<i,

rv

&,

9 t>

rz , 6 Z ,

ignored, except that the exclusion principle


will

(13-12)

and
T is the eigenfunction describing the motion of all
That is,

The eigenvalue

Z electrons.

which

E T rpT

equation 10-15), and where F/^)

the eigenfunction

Spin

1 VJLrfrpr =
I= V?Vr + i=i

the Laplacian operator in spherical coordinates for the rth

is

(cf.

the rth electron.

its

and the

in spherically symmetrical

electrons

net potentials.

in 1928

solving the Schroedinger equation for a

It consists in

be described

expect that there

Wt
In fact there

is

is

<f>

(13-13)

z)

is satisfied

in a

way

Since the electrons are non-interacting,

later.

a solution to equation (13-12) of the form

v( r i

0i> i>i) V>( r 2>

such a solution.

e 2,

<t>z)

The equation

V>(

r z>

e z>

(13-14)

<t>z)

splits into

one-particle

Schroedinger equations such as

^
2m

VM'*

0*

&)

Kfo) Vfo.

6i +<)

= EM*,

Oi,

4>d

(13-15)

where

ET

=iE
=
t

still

(13-16)

Vt (rt) are not known. However, the problem


be solved by the following self-consistent treatment.

Initially the potentials

can

THE HARTREE THEORY

Sec. 3]

397

A reasonable guess about the form of the V^r,) is obtained by setting


for all
VAr ) = V(r ),
(13-17)

1.

where V(r^
for the

from the Thomas-Fermi theory


With these common values of V^rJ, all

the net potential evaluated

is

atom with

electrons.

the one-particle Schroedinger equations (13-15) are identical.

These equations are solved by numerical integration, and the one-

2.

y> a (r 6
<^), tpfa, Q t <f> ( ), ..., yfc, it &), ... are
found. They are listed in order of increasing energy of the corresponding

particle eigenfunctions

t,

Each of the quantum numbers a, fi,


i,
stands for
of spatial and spin quantum numbers for one electron.
3. To obtain the ground state of the atom, the one-particle states are
filled in such a way as to minimize the total energy (13-16) and yet satisfy
the exclusion principle. That is, the one-particle states are filled in order
of increasing energy with one electron in each state. Then the oneeigenvalues.

a complete

set

particle eigenfunctions describing the

v(i> 0i,

<M, Vnfo,

K 4*1

Z electrons
win,

#i,

&),

are

V;( r z. e z, 4>z)

4. Next a more accurate estimate of the potentials Vt (r ) is calculated


by integrating the equation obtained by combining Gauss's law (equation
13-3) and the relation (13-4) between the electric field and the potential
energy. In calculating the potential acting on a particular electron, say
(

the A:th electron, the quantity p(r) in equation (13-3)

is set

equal to the

sum

of the quantum mechanical probability densities for all the filled states,
except the state k occupied by the kth electron, averaged over a spherical
surface of radius

p(r)

r.

That

-4
Artr Jo

-Jo

is,

2>.*(i-. e .#V.(-,0,#r
i=x

sin0d0ty

(13-18)

Thus the potential acting on the kth electron is evaluated from a spherically
symmetrical charge distribution due to the nucleus plus the average radial
dependence of charge carried by all the electrons except the A:th. The
Vii r i) which are obtained will, in general, be somewhat different from
the initial estimates taken in step

2 and using the

1.

If they are different, the entire

5.

new set of Vfr,).

procedure

is

repeated, starting at step

After several cycles (2>-3>4>-2->-3-*

is found that the K,-(/" -) obtained at the end of a cycle are essensame as those used in the beginning. Then this set of Vi(rt) is
the set of self-consistent potentials, and the eigenfunctions, calculated
from these potentials are the set of one-particle eigenfunctions which
describe the motion of the electrons in the multi-electron atom.

tially

) it

the

MULTI-ELECTRON ATOMS

398

is

[Ch. 13

There are several obvious approximations in the Hartree theory. One


that the charge distribution due to each electron is effectively treated as

were spherically symmetrical. The justification of this will be discussed shortly. Another approximation is that antisymmetric eigenfunctions, such as (12-16), are not used to describe the system of Fermi
if it

particles.
satisfied

This

is

not a serious

by means of the

electron can be in each

fault,

because the exclusion principle

is

auxiliary requirement of step 3 that only one

quantum

state.

modification of the Hartree

theory to include antisymmetric eigenfunctions has been given by Fock


(1930).

In comparing the results of the two theories,

found that the

it is

error introduced by not using antisymmetric eigenfunctions

Consequently, most calculations have been


theory because

the original Hartree

Even

so, the calculations

relatively easier to handle.

it is

small.

is

made with

are very lengthy and, as yet, results have been obtained only for a limited

number of atoms. However,


tion of multi-electron
It is

found that

all

the essential features of the Hartree descrip-

atoms are quite well understood.

V(r) in

which

independently.

V (r )

the self-consistent potentials

so that, to a good approximation,

we may speak

atom

the electrons of the multi-electron

all

The

are very similar

of a single net potential


are

moving

one-particle eigenfunctions for these electrons in this

spherically symmetrical potential are closely related to the one-electron-

atom eigenfunctions

discussed in Chapter 10.

V-im.m.fo

The eigenfunctions

m m
t

0. 4>)

are labeled

In fact, they can be written

Rm(r) , mi (0) <M#f..

by the same

set

(13-19)

of quantum numbers

as are used for the one-electron-atom eigenfunctions,

quantum numbers are related to each


(10-37). The spin eigenfunction a m

n,

I,

and these

other just as specified by equations


is

same function

exactly the

as in

the case of a one-electron atom. Furthermore, the functions @ lm (6)


and Q> m (4>) are al so exactly the same. The reason is the Schroedinger

equation (13-15) for an electron in a spherically symmetrical net potential


is just equation (10-15). This equation leads directly to equations (10-17)

and (10-18), and

to their solutions

Q lmi (9)

and

<J>

mi (<).

and

Consequently,

dependence of the
one-electron-atom eigenfunctions applies directly to the 6 and dependence
all

the discussion in Chapter 10 concerning the

<j>

<f>

of the multi-electron

atom

As an example, equation

one-particle eigenfunctions.

(10-46) shows that the

sum of the

all

possible values of

is

spherically symmetrical.

certainly also true for the case of n

true for any given n

the

same statement

and

/.

From

2,

0,

and

it

probability

and
2,1
1
This statement is

densities for the one-electron-atom eigenfunctions with n

can be shown to be
we conclude that

the discussion above,

also applies to the one-particle eigenfunctions for a

THE HARTREE THEORY

Sec. 3]

multi-electron atom.
state, the

means
with

Now, when

a multi-electron

atom

is

in

lowest energy one-particle states are completely

that for almost all values of n

all

399

possible values of

for these electrons

is

and

Since the

its

ground

filled.

This

there are electrons in states

sum of

the probability densities

spherically symmetrical, their total charge distribution

Figure 13-2. The radial probability densities


atom according to the Hartree theory.

for the filled

quantum

states of the argon

also. At most, only a few electrons in the highest energy


which the states with all possible values of m might not be filled,
can contribute to any asymmetries in the charge distribution. This is the

is

symmetrical

states, for

justification for treating the charge distributions as spherically symmetrical

in the Hartree theory.

The functions R nl {r) determining the ""radial behavior of the eigenfunctions are not the same as for the case of a one-electron atom because
the net potential V{r), which enters in the differential equation for these

same r dependence as the Coulomb potential


Typical examples of the radial behavior of the
one-particle eigenfunctions are shown in figure (13-2). In this figure we

functions, does not have the

of a one-electron atom.

MULTI-ELECTRON ATOMS

400

[Ch. 13

A atom (Z = 18)
+ l)Pnl {r), where

plot the results of a Hartree theory calculation for the


in

terms

P nl (r)

of the

quantities

2(2/

Y)r

R2Jr) =

2(2/

by equation (10-43). The


one-particle quantum
completely
fill
the
18
possible
atom
electrons of
there are 2/ + 1
figure.
Since
values
shown
in
the
for
and
/
states
the n
possible values of m for each /, and for each of these there are 2 possible
is

the radial probability density defined


this

values of

m
=

s,

there are 2 states with n

=
=

1,

0;

2 with n

2, /

.0;

and 6 with n = 3, / = 1.
2,
1
These states are the ones which are filled in the ground state of the atom
because, as we shall see later, they are the states of lowest energy. The
6 with n

quantity 2(2/

2 with n

X)P nl {r)

is,

3,

0;

therefore, the radial probability density for

the state with

quantum numbers n and

in that state.

It is

/,

times the

number of

electrons

the total radial probability density for the electrons

and the sum of these quantities gives P(r), the total radial
which is the probability of finding some
electron with a radial coordinate in the region of r. Thus
in that state,

probability density for the atom,

P(r)

= Z2(2l+l)P nl(r)

(13-20)

In the same
plotted for the A atom in figure (13-3) as P(r) H
we plot P(r) T_ F which is the total radial probability density for
the A atom evaluated from the Thomas-Fermi theory. We see that the
Thomas-Fermi theory gives a good description of the average behavior

This

is

figure

of the total radial probability density predicted by the Hartree theory,


except that for large values of r it does not decrease rapidly enough.
The Thomas-Fermi theory is also inaccurate for small r, but it is not

apparent in a plot of P(r ) because this


small

r.

It is

is

dominated by the

factor for

measurement
electrons from the

possible to obtain a fairly direct experimental

of the total radial probability density by scattering


atom. The angular distribution of the scattered electrons depends, in a
manner which will be described in Chapter 15, on the total radial prob-

atomic electrons.! It can also be measured by


from the atom, using a technique which will be explained
in Chapter 14. Although neither of these measurements is accurate enough
to resolve the oscillatory structure of P(r) H they do confirm its average
ability density of the

scattering X-rays

behavior.

potential V{r) in

we

dependence of the net


which, according to the Hartree theory, each electron

In figure (13-3)

also indicate the radial

t The reader will recall that this is quite different from the information provided by
scattering alpha particles from an atom. The reason is that in the electron experiments

the mass of the scattered particle equals the mass of an atomic electron, whereas in the
alpha particle experiments the mass of the scattered particle is so large compared to the
mass of an atomic electron that it essentially cannot be scattered by these electrons.

THE HARTREE THEORY

Sec. 3]

401

of the A atom is moving. This is given by plotting the radial dependence


of the effective charge Z(r) defined by the equation
V(r)

The behavior of

Z(r)

is

energy of an electron

is

= -

e Z(r)

(13-21)

For r -> 0, the potential


dominated by the Coulomb attraction of the

easy to understand.

22 i

Figure 13-3. The total radial probability densities of the argon atom according to the
Hartree and Thomas-Fermi theories, and the effective charge of that atom according to
the Hartree theory.

nucleus of charge
increasing

+Ze. So V(r)^ e*Z/r, and Z(r)->-Z

18.

With

Z(r) decreases because the magnitude of V(r) decreases more


rapidly than the Coulomb potential due to the nucleus, as the electrons
r,

with radial coordinates

less than r shield part of the nuclear charge. For


V(r)-> e 2 /r, and Z(r) 1, because the nuclear charge +Ze is
shielded by the charge (Z l)e of the other electrons in the atom.

r^-

co,

By inspecting the

plots of

P nl (r) we

see that, for all the electrons in states

MULTI-ELECTRON ATOMS

402

common

with

values of the

[Ch. 13

quantum number n, the probability densities


same range of r. All these electrons are
terminology we have used before in connection

are large only in essentially the


said to be in the

same shell

Furthermore, the range of

with one-electron atoms.

which the

in

probability densities are large (the "thickness" of each shell)

restricted

is

enough that Z(r) has a reasonably well-defined value in that range. These
circumstances form the basis of a very approximate description of the
motion of the electrons in a multi-electron atom, in which all the electrons
in a particular shell are considered to be moving in a Coulomb potential,

VJr)

Z=

e2

(13-22)

where Z

is

a constant equal to Z(r) evaluated at the average value of r

As an example, we

for the shell (the "radius" of the shell).


figures (13-2)

and (13-3) for the

atom:

In the approximation (13-22), the one-electron


the expectation value of

Z by Z.

This

r,

the total energy,

to within a factor of 2 or 3 for the n

the approximation

many

picture of

is

3 shell of the

useful because

it

can be used

shell,

16,

find

but

it is

A atom.

if

from the

Z3 ~

8,

atom equations

etc.,

quite accurate for the n

is

2.5.

specifying

we

replace

only accurate
Nevertheless,

provides a readily understandable

of the basic results of the Hartree theory.

In the Hartree theory, shells for small n are of very small radii because
for these shells there

Coulomb
is

is

little

smaller than that of the n

1/Z X

shielding,

attraction of the nucleus.

and the electrons

feel the full

= shell
H atom by about a factor of

In fact, the radius of the n

shell

of the

This can be understood in our approximate description from


The electrons in these inner shells are in a region of

equation (10-44).

large negative potential energy, so their total energies are correspondingly

large

and

The

negative.

results of the

Hartree theory show that the magni-

tude of the total energy of an electron in the n

of an electron in the n

shell

of the

shell

atom by a

is

larger than that

factor of about Z\.

This can be understood by referring to equation (10-38).

On

the other

hand, electrons in shells of large n are highly shielded from the nuclear

The

electrons in

these outer shells feel only a small net attraction to the nucleus.

The radius

Coulomb

attraction

by the electrons

in the inner shells.

of atoms of different atomic number can be estimated from equation (10-44), using the appropriate values of Z n This
equation predicts an increase in radius, with increasing n value of the
of the outermost

shell

outermost
less

shell for

rapid than 2

atoms of increasing atomic number, which is somewhat


because Z n slowly becomes larger with increasing

atomic number since the shielding of the nuclear charge is not perfect.
Actually the Hartree theory does predict an increase in the radius of the

THE HARTREE THEORY

Sec. 3]

outermost

shell,

but one which

is

403

much

less

rapid than 2 ,t because the

electrons in the outermost shell do have a small probability, which we have


ignored in our approximate description, of being found in the region near
the center of the atom. Although the probability is small, it is important

because near the center the electrons

feel a

to decrease the radius of the shell.

This

very strong attraction that tends


effect

becomes more important

with increasing atomic number and, along with the effect of the increase
in Z n leads to the important consequence that according to the Hartree
,

theory the radius of the outermost shell is roughly the same for all atoms.
Finally, we can see from our approximate description why the Hartree
theory predicts that the total energy of the electrons in the outermost shell

comparable to that of an electron in the ground state of the


H atom. It is because in all atoms Zn for the outermost shell is always of
the order of unity, so the potential energy is comparable to the potential
of all atoms

is

energy acting on the electron in the


the Hartree theory mentioned

in this

atom. Each of the predictions of


paragraph are borne out by the

observed properties of atoms.


In a one-electron atom,
the

same

all

the electrons in a certain shell have exactly

total energy, if the very small effects like spin-orbit interaction

are ignored. J The reason is that in the one-electron atom the total energy
depends only on the quantum number n, and so is the same for all electrons
in the

same

shell,

even though they

may have

different values of the

In a multi-electron atom this is not true. This


statement should not be surprising because we have mentioned on several
previous occasions that the fact that the total energy of a one-electron

quantum number

/.

atom does not depend on

/ is

a consequence of the fact that the potential

In a multi-electron atom the electrons are


moving in a net potential V{r) which is not exactly proportional to 1/r,
and the total energy Enl of these electrons depends on / as well as on n.
However, since we are ignoring spin-orbit interactions and certain other
effects, the total energy does not depend on the quantum number m s which
determines the spatial orientation of the spin, nor on the quantum number
m which determines the spatial orientation of the "orbit." The important
features of the dependence of E nl on n were described approximately
is

exactly proportional to

1/r.

paragraph in terms of the energy equation for a one-electron


replaced by Z. That is, we took

in the previous

atom with

E nl ~ -ftZie*l2hW

(13-23)

of a
t This statement does not contradict the Thomas-Fermi theory prediction
gradual decrease in the overall radius with increasing atomic number.
number of different
X By this we mean that the total energies of the electrons in a
one-electron atoms are the same if the electrons have the same values of the quantum

number

n.

MULTI-ELECTRON ATOMS

404

For small

becomes

n, this

with increasing n much more


atom because of the dependence

less negative

rapidly than for the case of a one-electron

of Z n on n. For large
n dependence similar

Now,

Zn does not change rapidly,

the difference
increasing

Enl has a less rapid

show

that

Enl

is

indicated by the approximation (13-23).

is

is

largest for

This difference

/.

so

to that of the total energy of a one-electron atom.

the results of the Hartree theory

negative than

of

n,

[Ch. 13

the dependence of

is

7=0

actually

more

Furthermore,

and diminishes progressively with

quite appreciable.

In fact, for large values

Enl

on / can be stronger than it is on n.


The reason for the dependence of the total energy on / is easy to understand in terms of the behavior of Z(r) and the probability density
n,

Wtim im ,Wnim im! in the region near r

0.

In (10-45')

we pointed out

that

V>%imi

<* r \

m V>mm m
t

-*

(13-24)

This was with reference to one-electron-atom eigenfunctions, but it is


equally true for multi-electron-atom eigenfunctions because the term
l)/r 2 in the differential equation (10-27), which governs the
/(/
radial behavior of the eigenfunctions, completely dominates the term

(2^/A 2 ) [E

V(r)] for small r.f Consequently, in the region near r


the
exact form of V(r) is unimportant and all eigenfunctions for spherically sym-

metrical potentials have the r l radial behavior of (10-45).

Now

consider

two electrons in the same shell of a multi-electron atom, one with 1 =


and the other with / = 1. According to (13-24), there is some chance of
electron in the region of small r, but only a very small
/ =
chance of finding the / = 1 electron in that region, since r
1
and
r2
for small r. Although the /
electron will spend only part of
its time near r = 0, this makes an appreciable difference in the average

finding the

value of

its

potential energy because Z(r)

becomes very

so the net potential energy V(r) becomes very negative.


potential energy of the

An

more

electron will be

average potential energy of the

electron.

instructive interpretation of the first of these terms

the substitution u

rR

large near r

Thus

the average

negative than the

Similarly, the average

can be obtained by making

in equation (10-27). This gives

d'u

1(1

1)

2(i

or

which is seen to be equivalent to the Schroedinger equation for motion in one dimension,
with the term /(/ + l)h 2 /2fir 2
L*/2fir 2 added to the potential V(r). This term is often

called the centrifugalpotential, for reasons

which can be seen by considering the classical


energy conservation equation for a particle of mass ji moving under the influence of a
potential V(r). As the particle will move in a plane containing the origin, it can be

THE HARTREE THEORY

Sec. 3]

potential energy of the

405

electron will be

more

negative than that for

>

4
2
r for small r. On the
2 electron in the same shell because r
electron is larger than
other hand, the average kinetic energy of the /
1 electron, since it spends more time in the region of
that for the /

an

larger negative potential energy.

It

happens that these two

balance out for a potential in which Z(r)


\\r

Coulomb

dent of

on

n.

/.

is

effects exactly

a constant; thus, for the exact

potential of a one-electron atom, the total energy

But

in all other cases the total

In a multi-electron atom, the

energy depends on

is

indepen-

as well as

dependence of the average potential

dependence of the average kinetic energy, and the


total energy Enl which is the sum of the average potential and kinetic
than it is for /= 1, more negative
energies, is more negative for / =
=
high Z multi-electron atoms, in
/
is
1=2,
etc.
For
than
it
for
for
1
which Z(r) increases particularly rapidly with decreasing r in the region
near r = 0, and for outer shells of large n in which the n dependence of Enl
is particularly weak, the / dependence of Enl is actually stronger than the
n dependence. We shall see that this plays an important role in the strucenergy dominates the

ture of multi-electron atoms.

All the electrons in a particular shell have probability densities which

same form in the region where their values are


However, the probability densities are significantly different in the
region near r = 0. As we have seen, the result is that the total energies of
the electrons in the same shell are not all the same. For this reason, it is

are of approximately the


large.

convenient to speak of each shell as being composed of a number of

one for each value of /. All the electrons in the same subshell
have the same quantum numbers n and /. Therefore, they have exactly
the same total energy (in the approximation in which we neglect spin-orbit
interactions and other effects). Also they have exactly the same radial
subshells,

probability densities

Pnl (r).

described by the coordinates

r, 6,

=
Also the orbital angular

and

the equation is

^J +i4fJ-\+V(r)

momentum

of the particle

L = *r

is

a constant,

dd
dl

so the equation can be written

which

is

seen to be the classical energy conservation equation for motion in one dimenterm L 2 /2(ir 2 added to the potential V(r). This positive term acts like a

sion, with the

repulsive potential, tending to keep the particle

value of L, the stronger

is this effect.

away from

the origin.

The higher the

MULTI-ELECTRON ATOMS

406

tu

cn

0)

<

a s

00

to

h-4
PO-

CO

CO
in

US

CU

CO
CO

02

(2

00

W3

S
lO

CO

S-,

J w
s

3
S3

CO

CO

T3

00

ft

U5

PQ

[Ch. 13

CO

00

>.

CO

t>.

a
>i

$3
a

-4

t-

to

PQ

CO

Ph
co

<^
to

f^

B ^
O

to

00

^8

ac-

t's

"W

r_
fl

in

,n<

O-

(>

e>

**-.

88

*-.

00

CO

uj
-i

n
&

lO

eg

.e

to

fl

CO

c^-

CO

1
9

r2

<u

a.

**.

25

>

in

e2
CO

S3

s
(M

<M

U5
C

GQ

be

CM

U3

c8

(3

ffl

CO
88

3
us
CO

Ph
Oi

S3

(8

CO

>

00

in

CO

THE PERIODIC TABLE

Sec. 4]

407

The Periodic Table

4.

of the most important properties of the elements are periodic


which specifies the number of electrons
functions of the atomic number

Some

in

This can be demonstrated by constructing

an atom of the element.

a periodic table of the elements such as the one presented in table (13-1).
Each element is represented in the table by its chemical symbol, and also

by

(The entries sometimes appearing below the


quantum numbers for the ground state of an

atomic number.

its

chemical symbol

specify the

atom of the element in a way which will be explained later.) Elements


with similar chemical and physical properties are in the same column.
For instance, all elements in the first column are alkalis and have a valence
of

all

elements in the last column are noble gases and have a valence

We

assume that the reader is at least partially familiar with the


periodic properties of the elements from his study of elementary chemistry.
of

0.

Our

task here

is

to interpret these properties in terms of the theory of

multi-electron atoms.

This interpretation
filled

is

based upon knowing the way in which the outer


atoms are ordered according to energy. In

subshells of multi-electron

be obtained entirely from the Hartree

principle, this information could

theory calculations described in the last section.

However, these calcula-

number of elements.
we need must be obtained from
experiment. This can be done by studying the spectrum emitted when
excited atoms of an element make transitions to their ground states.
In such transitions an electron drops from some high energy one-particle

tions have been carried through for only a limited

Consequently,

much of

state to the lowest

highest energy

From an

energy state which

quantum numbers n and


ground

on a large
which will be described
filled

then the

of

state

of the atom.

found

it is

later, the

it is

possible to

is

This

the highest energy

By making such an

that, except for

analysis

minor deviations

ordering according to energy of the

shown in table (13-2). The first column identifies


quantum numbers n and /. The second column does

subshells

the subshell by the

is

this one-particle state.

to identifying an outer subshell which

number of elements,

the

unoccupied. This state

one-particle state for the ground state of the atom.

filled

subshell filled in the

outer

is

analysis of the spectral lines which are emitted,

identify the

amounts

the information

is

same thing except

as

that n

and

are given by the spectroscopic notation.

commonly used in discussing the spectrum and energy


levels of atoms. The number gives the value of n, and the letter gives the
value of /according to the scheme shown in table (13-3). The third column
This notation

of table (13-2)

is

is

2(2/

1),

which

is,

as mentioned in the last section, the

MULTI-ELECTRON ATOMS

408

[Ch. 13

number of possible quantum states for the value of / characteristic of the


subshell. Thus this column gives the maximum number of electrons that
can occupy the subshell without violating the exclusion principle.

TABLE

(13-2)

The E nergy Ordering of the Outer

Filled Siibshells

Quantum

Designation

Capacity

numbers

of

of subshell

nj

subshell

6,2
5,3
7,0

6d

10

5/

14

7.$

6,1

6p

5,2
4,3

5d

10

4/

14

6,0

6s

5,1

5f

6
10
2

2(2/

4,2
5,0

4d

4,1
3,2

Ap
3d

10

4,0

4s

3,1

lp

3,0

3s

2,1

2/>

2,0

2s

1,0

Is

5s

1)

t
Increasing energy
(less negative)

Lowest energy
(most negative)

TAB LE H3-3}

Spectroscopic
s

notation

Many

of the features of this experimentally observed ordering of the


according to energy have also been explicitly obtained
by Hartree theory calculations. For instance, calculations for the
atom
outer

filled subshells

show

that the energy of an electron in the 4s subshell should, in fact, be

THE PERIODIC TABLE

Sec. 4]

409

lower than the energy of an electron in the 3c? subshell of this atom. Further-

more,

the features which are observed experimentally can be understood

all

at least qualitatively in terms of the description of multi-electron

atoms

provided by the Hartree theory. In our discussion in the last section we


found that this theory predicts that the energy of the subshells becomes
more negative with decreasing values of n and with decreasing values of /.

We see this immediately in table (13-2).


subshell in the n

shell,

The

\s subshell,

which

The two

has the lowest energy.

is

the only

subshells of

2 shell are both of higher energy and, of these, the 2s subshell is


of lower energy than the 2p subshell. In the n = 3 shell the subshells
3s, 3p, 3d are also ordered in energy according to the predictions of the

the n

Hartree theory. However, the energy of the 4s subshell is actually lower


than the energy of the 3d subshell because, for reasons described in the
last section, the / dependence of the energy E nl of the subshells becomes
more important than the n dependence for outer subshells with large

Continuing up the

values of n.
shells

always

satisfies

list,

we

see that the ordering of the sub-

the rule that, for a given n, the subshell with the

lowest /has the lowest energy and that, for a given


lowest n has the lowest energy. Near the top of the

Enl

is

so

subshell

much

is

/,

the subshell with the

list

the

dependence of

stronger than the n dependence that the energy of the Is

lower than the energy of the 5/ subshell.

It

should be emphasized

that table (13-2) does not necessarily give the energy ordering of all the

any particular atom, but only the energy ordering of the


which happen to be the outer subshells for that atom. For
instance, the energy of the As subshell is lower than that of the 3d subshell
for K atoms and the next few atoms of the periodic table, but for atoms
subshells in

subshells

near the end of the periodic table the 3d subshell

is

of lower energy than the

As subshell because for these atoms they are inner subshells

dependence of

Enl is

so strong that

it

dominates the

and the n

dependence.

The characteristics of an atom depend on the motion of its electrons.


The motion of an electron is specified by the set of four quantum numbers
which specify its quantum state. However, in the approximation represented by the Hartree theory, only the quantum numbers n and / are
important. Therefore, in this approximation an atom can be characterized
by specifying the n and / quantum numbers of all the electrons, i.e., by
specifying the subshells occupied

by the various

electrons.

This

is

called

the configuration. The ordering according to energy of the outer filled


subshells being known, it is trivial to determine the configuration of any

atom

in

its

ground

subshells in such a

state

way

because in this state the electrons must

as to minimize the total energy of the

fill

all

the

atom and yet

not exceed the capacity 2(2/ + 1) of any subshell. The subshells will be
Consider
filled in order of increasing energy, as listed in table (13-2).

MULTI-ELECTRON ATOMS

410
first

the

H atom.

atom both

The

single electron occupies the Is subshell.

He atom

the ground state of the


figuration of

is

written

The configuration of He

which

it

Chapter

He:

Is

number of
on the chemical symbol
the atom. In the 3 Li atom one of the electrons

for

the superscript

The 4 Be atom completes the

In the six elements from

10

Ne

the additional electrons

The configurations of B and


5

10

will

table (13-1)

is

and has the configuration

ls 2 2s 2

Be:

to

Is subshell is only 2.

lsW

Li:

2s subshell
4

The reader

The con-

must be in the 2s subshell because the capacity of the


The configuration of this atom is

subshell.

12).

subshell designation specifies the

contains;

specifies the value of

5,

For the He

in our discussion of

written

is

on the

superscript

electrons

in section

The

we saw

electrons are in the Is subshell (as

[Ch. 13

10

Ne

fill

the 2p

are

B:

ls^s^p 1

Ne:

ls 2 2s 2 2p e

note that the periodic table of the elements presented in


divided vertically into a series of blocks with each

row

labeled by the subshell which, according to table (13-2), the elements of

the

row are

tion of any

With this information,


atom using a procedure that

filling.

it is

easy to write the configura-

will

become apparent by com-

parison with the configurations listed above. The configuration of certain

atoms, for which the

last

few electrons are observed to be in different

would be predicted by this scheme, are indicated by entries


below the chemical symbol of the element. To illustrate the way this
information is used to write the configuration of an atom, we present
subshells than

several typical examples


ls22s 2 2p*3s 2 3p e4s 1

23y
24

Cr

Tc

"Ru
46p d

2
2
l5 2i 2/j 6 35 2 3/45 2 3rf 3
2

\s 2s 2p*3s 2 2pHs 1 3d h
1

1s
1

2s 2 2/3 s 2 3/> 8 4s 2 3d 104/5s 2 4</ 6

2s 2 2/3s 2 3/Ms 2 3d 104/> 6 5.s 1 4rf 7

2s 2 2p s 3s 2 3p s 4s 2 3d 10 4p 6 4d 10

57

La

l5 25 2 2/35 2 3/j 6 4j 2 3c/ 104/5j 2 4rf 10 5/65 2 5</ 1

58

Ce

lj 2 2s 2 2/3s 2 3/45 2 3rf 10 4/5s 2 4rf 10 5/7 6 6.y 2 5d 1 4/' 1

59 Pr

s 2s*2p3 s 2 3pHs 2 3d

w4p & 5s 24d la 5p%s 2

4f

THE PERIODIC TABLE

Sec. 4]

We

411

see that in certain cases the actual configurations observed for the

elements do not

strictly

adhere to the predictions of table (13-2).

instance, this table says that the energy of the

the energy of the 4? subshell


24

these subshells are

is

For

greater than

filling.

Yet

in

29

Cu, one of the electrons that could be in the 4s subshell


actually in the 3d subshell. Similar situations are observed to occur

Cr, and also in

is

when

3d subshell

4d and 5s subshells. In 43 Tc the 5s subshell is filled in the normal


manner. But in 45 Rb there is only one electron in the 55 subshell; in
46
Pd both electrons have left the 5s subshell and moved to the Ad subshell.
The 78 Pt and 79 Au configurations show that the same kind of thing can
happen for the 5d and 6s subshells. From these circumstances we conclude
that the energy differences between the 3d and 4s, the 4d and 5s, and the
5d and 6s subshells must be so small while they are being filled that,
although generally the ordering of these subshells is as shown in table
(13-2), in certain cases the ordering can actually be reversed. Configura57
La and in the
tions which disagree with table (13-2) are also observed in
lanthanides (Z = 58 to 71) more commonly called the rare earths.
The table predicts that after the completion of the 6s subshell the 4/'
subshell should fill. But in 57 La, and several rare earths, there is one 5d
electron. A similar situation occurs in the group of elements following
89 Ac, which are called the actinides (Z = 90 to 103).
From the same
mean that the
observations
to
argument as above we interpret these
between the
subshells,
and
energy differences between the 5d and 4/
are being
these
subshells
while
6d and 5/ subshells, are very small
for the

filled.

On

the other hand, certain predictions of table (13-2) are always obeyed.
is exceptional for elements in the first two
columns of the periodic table (13-1), we conclude that

Since none of the configurations

and the
every

last six

subshell

is

always of higher energy than the preceding

subshell while these subshells are being

filled,

and that

s or

in these circum-

is always of higher energy than the preceding p


Therefore there must be large energy differences between the

stances every s subshell


subshell.

subshells concerned while they are being filled. In fact, the energy differences
between every 5 subshell and the preceding p subshell are particularly
large, and it is easy to understand why. Since for a given n the energy of a
subshell becomes higher with increasing /, an s subshell is always the first
subshell to be occupied in a new shell. -.Consequently, when an electron
is added to a configuration with a completed p subshell and goes into the

subshell of next highest energy,

an

s subshell,

which according to table (13-2) is always


first one in a new shell. Compared

the electron will be the

to the electrons in the preceding subshell,


will

be considerably larger,

its

its

average radial coordinate

average potential energy will be considerably

MULTI-ELECTRON ATOMS

412

[Ch. 13

and its total energy will be considerably higher much


higher than for the usual increase in total energy in going from one subshell

less negative,

to the next.

The fact

is a particularly large energy difference between every


and the preceding p subshell has some very important consequences. Consider atoms of the elements 10 Ne, 18 A, 36 Kr, 54 Xe, 86 Rn
in which ap subshell is just completed. Because of the very large difference
between the energy of an electron in the p subshell and the energy it
would have if it were in the s subshell, the first excited state of these atoms
is unusually far above the ground state.
As a result, these atoms are
particularly difficult to excite and are very inert.
Furthermore, they
produce no external electric field because they consist of sets of completely
filled subshells and, therefore, have charge distributions which are spherically symmetrical.! Also we shall see later that atoms with completely
filled subshells produce no external magnetic fields. Such atoms cannot
interact with other atoms to produce chemical compounds, and they

that there

s subshell

constitute the noble gas elements.

for this

atom

the

does not contain a

In addition,

He

is

a noble gas because

an 5 subshell (even though it


has an unusually high first excited

first

unfilled subshell is

filled

p subshell) so it

atom consists of completely filled subshells


and so produces no external electric or magnetic fields. That 2 He is a
noble gas is indicated by its being listed in the last column of the periodic
state,

and

also because the

table instead of the second column. An element such as 20 Ca is not a


noble gas, even though it consists of completely filled subshells, because
in
is

excited state an electron goes to a 3d subshell so the excited state


not far above the ground state and the atom is not particularly

its first

inert.

Another aspect of the particular inertness of the noble gases can be


obtained by plotting, for the various elements, the measured values of the
magnitude of the total energy of an electron in the highest energy filled
subshell.

This

is

the atom, which

such a plot.
value which

equal to the energy required to remove the electron from


is

the ionization energy of the atom.

Figure (13-4) shows

We see that the ionization energy oscillates about an average


is

essentially

independent of Z, in agreement with our earlier

conclusion that the total energy of the electrons in the outer shells

is

f There is obviously no external monopole electric field (the field produced by a single
charge) for any neutral atom. Also, there is no external dipole electric field (the field

produced by two displaced charges of opposite sign) because of the symmetry of the
However, there is an external quadrupole electric field (the field
produced by two identical dipoles arranged along a line with one charge of the same
sign from each dipole coinciding) unless the charge distribution of the atom is spherically
charge distribution.

symmetrical.

Cf. equation (10-46).

THE PERIODIC TABLE

Sec. 4]

413

roughly the same throughout the periodic table. However, the oscillations
are quite pronounced, and it is apparent that the total energy of an electron
in the highest energy filled subshell of a noble gas

atoms are very

We

Li,

25

r?

20

considerably

more

difficult to ionize.

also see that the ionization energy

elements

is

These electrons are very tightly bound, and the

negative than average.

u Na,

19

K,

37

55

Rb, and

is

87

Cs

particularly small for the

Fr has not been measured).

15

Kr

Xe

<D

0)

% 10

Na

Li

Rb

20

10

L
30

Cs

40

J
60

50

70

80

90

100

ZFigure 13-4.

These are the

alkalis.

The

ionization energy of the atoms.

They contain a

single

weakly bound electron in an

Alkali elements are very active chemically because they are

s subshell.

quite eager to get rid of the weakly

bound

electron

and

revert to the

more

arrangement obtained with completely filled subshells. These


elements have a valence of +1. At the other extreme are the halogens,
stable

F,

17

C1,

35

Br,

53

I,

85

At, which have one less electron than

is

required

These elements are very prone to capture an


p
electron, and they have a valence of 1
For the first three rows of the periodic table (13-1), the properties of
the elements, such as valence and ionization energy, change uniformly
from the alkali element with which the row begins to the noble gas with
which it ends. In the fourth row of the periodic table this situation is no
longer always true. The elements 21 Sc through 28 Ni, which are called the
to

fill

their

subshell.

have quite similar chemical properties and also almost


same ionization energies. These elements occur during the filling of
the 3d subshell. The radius of this subshell is considerably less than
that of the 4* subshell, which is completely filled for all the first transition

first transition group,

the

MULTI-ELECTRON ATOMS

414
24

group except
from outside

Cr.

The

[Ch. 13

4s subshell tends to shield the 3d electrons


and so the properties of these elements are all
quite similar, independent of exactly how many 3d electrons they contain.
The chemical properties of 29 Cu are somewhat different from those of
the first transition group because it has only a single 4? electron in the
filled

influences,

outermost subshell. To a lesser extent this is also true for 24 Cr. The
element 30 Zn consists of a set of completely filled subshells and so is
somewhat more inert, as can be seen from its ionization energy. Similar
transition groups arise in the filling of the Ad and 5d subshells.
An extreme example of the same situation is found in the rare earths,
58

Ce through

filling.

T1

Lu.

This subshell

These are the elements in which the 4/' subshell is


lies deep within the 6s subshell, which is completely

the rare earths. The 4/ electrons are so well shielded from the
outside that the chemical properties of these elements are almost identical.

filled in all

The same thing happens


the 5/ subshell

We

finish

is filling

in the actinides, 90 Th

through

103

?.

In this group

inside the filled Is subshell.

our discussion of the ground state of multi-electron atoms

and the periodic

table by calling attention to two points. First, the study


of the electronic structure of atoms gives no indication of why only 92
elements are found in nature. This implies that the instability of atoms
with Z
92 must be due to an instability of the atomic nuclei. We shall

>

Chapter 16 that this is indeed the case. However, it is possible to


produce such nuclei artificially in certain nuclear reactions. It is found
that they have absolutely no difficulty in collecting electrons and forming
atoms which are perfectly stable as long as the nuclei remain stable.
The members of the actinide group with Z > 92, which are called the
transuranic elements, are formed in just this way.
The second point we emphasize is the importance of the antisymmetric
see in

symmetry character of electrons the property of electrons which causes


them to^obey the exclusion principle. If they did not obey this principle,
all the electrons of a multi-electron atom would be in the \s subshell
because it is the subshell of lowest energy. If this were the case, all atoms
would have very small radii, very high first excited states and ionization
energies, and^no external electric or magnetic fields. All elements would
be extremely inert and there would be no chemical compounds, no life,
and no readers or writers of books on modern physics

5.

Excited States of

Atoms

In the previous sections

we have

described qualitatively the energies

of electrons in the subshells that are being

filled in

various atoms.

Now

Sec. 5]
let

EXCITED STATES OF ATOMS

415

us consider a quantitative example of the energies of electrons in

the subshells of a particular atom.

Consider the typical atom

26

Fe,

all

whose

shown in the energy level diagram of figure (13-5).


The diagram is plotted on an energy scale which is linear from to 10 ev,
and logarithmic for more negative energies. The measured values of the
electron energies are

total
26 Fe

energy

atom

in

Enl
its

of an electron in the various occupied subshells of the

ground

state are plotted,

and are also written to the

right

-5
4s

3d

-10

-50
-c
>

-7.9 (-6.58)
-8.2 (-10.3)

3p

56 (-62.5)

-100 -

3s

94 (-94.8)

-500
2p
-1,000 _

2s

-710 (-721)
-850 (-827)

-5,000
Is

7130 (-7115)

-10,000

Figure 13-5.

An energy

level

diagram for the

filled

subshells of the iron atom.

levels. These values are obtained from an analysis of the optical


and X-ray spectra produced by the atom. The procedures involved in
such analyses will become apparent later. Shown in parentheses are the
values of E nl obtained from a Hartree theory calculation for the 26 Fe
atom. The reader will note that the agreement between theory and experiment is generally quite good. He will also note that these data confirm
many of our previous statements regarding the energies of the various
subshells. For instance, the energy difference between each p subshell
and the s subshell of higher energy is seen to be particularly large. Also
the energy difference between the 3d and As subshells is seen to be very
small. It is also interesting to note that the energy of a 3d electron in the
26
Fe atom is actually slightly more negative than that of a 4* electron,
despite the fact that the observed configurations of this atom and the others
in the fourth row of the periodic table show that generally the 4s subshell

of the

MULTI-ELECTRON ATOMS

416

[Ch. 13

before the 3d subshell. This is not at all paradoxical, and it is predicted


by the Hartree theory. Because of repulsive effects, the presence of an
electron in a particular subshell changes somewhat the electric field produced by the other electrons in that subshell. When the energy difference
between two subshells is very small, this can lead to the situation found
in the 26 Fe atom.
Figure (13-5) shows only the subshells that is, the one-particle energy
levels
which are occupied in the ground state of the atom. There are
fills

also

many unoccupied

energy levels

one-particle

of higher energy.

Just as for the H atom, there are an infinite number of discrete one-particle
levels with E nl < 0, and a continuum of levels with Enl > 0.
In an
1

excited state of the atom,

some of

its

electrons go to the higher energy

one-particle levels.

Fortunately, the
tions

is

number of electrons in an atom which make such

transi-

small, because of the nature of the processes responsible for the

an atom. For instance, in a discharge tube atoms are excited


by the bombardment of energetic electrons. One of the energetic electrons
collides with an atom, knocking one or two atomic electrons into levels
excitation of

improbable that many atomic


Another typical process involves
the excitation of atoms by the absorption of quanta. At most, one quantum is absorbed by any particular atom and, particularly for higher energy
quanta, the quantum tends to interact with only one or two electrons.
Thus the excited states of an atom are not so complicated as they could be.
The configuration of an atom in an excited state differs from its ground
state configuration only in that one or a very few electrons are in higher
energy subshells, and there are corresponding vacancies in the subshells
from which they came. For simplicity, we often speak as if only one
electron in an atom can be excited. Even so, it is apparent that there are
still a very large number of different kinds of configurations for an atom
In such a process

of higher energy.

it is

electrons are simultaneously excited.

in

its

excited states.

However, it is possible to divide the different kinds of excitations into


two types. The first type corresponds to excitations in which an electron
in one of the higher energy subshells occupied in the ground state of the

atom

is

excited.

Since the subshells of higher energy are also the outer

subshells, this type corresponds to the excitation of

electrons of the atom.

one of the outer

The second type corresponds to the excitation of an

is one of the inner


examples of these two types of excitations are illustrated
26
in figure (13-6) for a Fe atom. [These diagrams employ the same pseudologarithmic energy scale as figure (13-5).] In an excitation of the first type,
the excited electron may make a transition to any of the discrete or

electron in one of the lower energy subshells, which


subshells. Typical

Sec. 5]

EXCITED STATES OF ATOMS

417

continuum one-particle

levels of energy higher than the energy of its


not true of an excitation of the second type. With
the exception of the 3d level, all the levels through the 4s level contain
their complete complement of electrons.
Consequently the exclusion

original level.

This

is

demands that the excited electron go to one of the one-particle


above the As level, or possibly to the 3d level. This means that the
energy required to produce an excitation of the second type is much higher
principle
levels

3p
3s

2p
2s

It

Figure 13-6.

Illustrating excitations of

the

first

type (optical excitations) and excitations

of the second type (X-ray excitations).

than that required to produce an excitation of the first type. The latter
is typically only a few electron volts, but an excitation of the
second type for the case illustrated in the diagram requires more than 800
energy
ev.

After an excitation of the second type, the very highly excited atom

will eventually return to its

ground state. It does this by emitting quanta.


These are the very high energy quanta which form the X-ray spectrum
of the atom. The much lower energy quanta which are emitted when
an atom returns to its ground state after an excitation of the first type

form the

optical spectrum of the atom (the spectrum in and near the visible
X-rays and X-ray spectra constitute the subject of the next
chapter. In this chapter we discuss only optical spectra. Consequently

range).

we

consider here only excitations of the

optical excitations.

first type, which are often called


These excitations are produced, for instance, by

exposing atoms to low voltage electrical discharges or flames.

MULTI-ELECTRON ATOMS

418

6.

Atoms

Alkali

In an optical excitation, usually


filled

[Ch. 13

it is

only the electrons in a partially

subshell that are excited, because the energy of these electrons

is

usually considerably less negative than the energy of the electrons of any

completely

There are exceptions to

filled subshell.

of an exception

is

26

Fe, for which the energy of

pletely filled 4s subshell

is,

as

we have

In an

is

example

in the

com-

filled

3d

subshell.

However,

normally obeyed.

atom of an

alkali

element the rule

contain a set of completely

An

seen, actually slightly less negative

than the energy of an electron in the partially


the rule

this rule.

an electron

filled subshells,

is

always obeyed. These atoms

the one of highest energy being

subshell, plus a single additional electron in the next s subshell.

As

mentioned before, the energy of the electrons in a filled p subshell is


particularly negative with respect to the energy of an electron in the next
s subshell. Consequently these electrons are not excited in any of the low
energy processes which lead to optical excitations. In essence, an alkali
atom consists of an inert noble gas core plus a single electron moving in an
external shell. The analysis of the optical spectrum of an alkali atom in
terms of

its

excited states

is

relatively simple

because the excited states

can be described completely by describing the single optically active


electron, and the core containing completely filled subshells can be ignored.

The total energy of the


atom is a constant plus

core does not change, so the total energy of the


the total energy of the optically active electron.

convenient in discussing the excited states of an alkali atom to define


the zero of total energy in such a way that the total energy of the atom is
It is

equal to that of the optically active electron.

Using

this definition,

we

present in figure (13-7) diagrams showing the energies of the ground


3
n Na,
state and the first few excited states of the alkali elements Li and

obtained from an analysis of the spectra of these elements, and also the
energy levels of 1 H for n = 2, 3, 4, 5, 6. Each energy level is labeled by
the

quantum numbers n and

configuration.

which

of the optically active electron,

These diagrams do not show a

fine structure

i.e.,

by

its

splitting,

be discussed later.
level diagrams also do not show the results of Hartree
calculations which have been made for these atoms, since the discrepancies
between theory and experiment are generally too small to be indicated.
The Hartree theory works particularly well in calculating the energy levels
will

The energy

of the optically active electron of an alkali element because the net potenV(r), due to the nucleus plus the electrons of the core, actually is
spherically symmetrical, as assumed in the theory. In terms of the Hartree

tial

ALKALI ATOMS

Sec. 6]

theory,

and

it is

419

easy to understand the structure of these energy level diagrams

diagram for 1 H. The dependence of the energy


its quantum numbers n and / is just as
we have described on previous occasions. For a given n, the energy is
their relation to the

of the optically active electron on

/-*^

-1

6
5

_ZeT

__

_. 6d_ _6f_

-t

~4

*Na

Li

5P

5d

5/

6d
5d

4p

4d

4/

5P

Ad

4s

-2

2.

5/

5s

3d

IF

_6.

3d

17

-3 -

IF
2

"2p"

-4 -

-5"5T
2s

-6

Some energy

Figure 13-7.

most negative

levels of

hydrogen, lithium, and sodium atoms.

for the smallest value of

because the electron spends more

time near the center of the atom, where

it

feels the full

nuclear charge.

In the ground state of the 3 Li atom, the optically active electron


2s subshell and
electron in a
electron

is

its

shells

is

in the 2p subshell

more

is

in the

about 2 ev more negative than an n

atom. But, in the

negative than an n
As level

energy

and

first

its

energy

2 electron in 1 H. In

negative than the 3d level.

with large values of n, the

excited state, the optically active


is

only about 0.2 ev

more

u Na the / dependence makes the

However, for the large radii subdependence becomes less important and

MULTI-ELECTRON ATOMS

420

the energy levels of the optically active electron

energy levels of an electron in a X


nuclear charge +Ze by the charge

become very

[Ch. 13

close to the

H atom, because the shielding of the


(Z l)e of the electrons in the core

of the alkali atom becomes essentially complete for an electron in a subshell

of radius large compared to the radius of the core. This occurs for smaller
u Na because the radius of the core is smaller for
the former atom. Putting it another way, although y*ip oc r 21 near r =
for all n, (10-39) shows that the proportionality constant decreases with
values of n in 3 Li than in

increasing n so that y>*%p

the

becomes smaller near

dependence becomes

less

0.

The

result is that

important when n becomes very

TABLE

large.

(13-4)

The Quantum Defects for the Sodium Atom

Note

i"

1.35

0.86

0.01

~0

going from the inner filled subshells, through the outer filled
and then through the unfilled subshells of a particular atom, the

that, in

subshells,
/

Subshell

dependence of EnX first increases in relative importance, is of maximum


importance in the outermost filled subshells, and then decreases in

relative

relative importance.

The energy

levels

of the

atom are given by

E=

the equation

-Rch\n 2

where R is the Rydberg constant. For an


can obviously be written

alkali

atom a

similar equation

E=-Rch/n' 2
if

the

n',

which are not

integers, are given the

quantities are related to the

(13-25)

proper values.

quantum numbers n by

ri

These

the equation

(13-26)

By fitting equation (13-25) to the


ft are called quantum defects.
observed energy levels of various alkali elements, it is found that for a

The

given element the

quantum defects

with the same quantum number


listed in table (13-4).

These

are essentially the


/.

The quantum

same

results explain the basis

formula (5-5) for the alkali elements.

for all subshells

defects for

u Na

are

of Rydberg's series

The

ATOMS

ALKALI

Sec. 6]

421

of the optical spectra emitted by alkali elements show a fine


which indicates that all energy levels are double, except

lines

structure splitting

due to a spin-orbit interaction acting on the


relativistic effects, which are just as
important as the spin-orbit interaction in the case of a one-electron atom,
are generally quite negligible. We can see this by using equation (5-13) to

those for

This

0.

is

The other

optically active electron.

estimate the average velocity of an optically active electron.

equation

we must

replace

ZM

by the quantity

In this

defined by (13-22).

As

roughly equal to unity for the optically active electrons of all atoms,
(5-13) shows that the average value of vjc is roughly equal to l/ times
the average value of this ratio for an electron in the ground state of the

is

atom; that

is,

u/ c

(l/)

lO

-2
.

Since the

quantum number n

for

an atom increases with increasing Z,


relativistic effects other than the spin-orbit effect, which are of the order
of (vjcf, are negligible for the optically active electrons of all atoms except
the optically active electron of

those of very small Z.|

The

of the energy levels of an alkali element due to a spin-orbit

shift

considering equation (11-47).

be understood by
immediately
to an
That equation leads

equation analogous to

This

interaction acting

AETi =

on the
(1

optically active electron can

1-49).

DO +
TTi
4m c

1)

"

Kl

is

1)

*(s

D] [\r

-^\
dr

(13-27)

where

Just as for a one-electron atom,

when

the spin-orbit interaction

is

included

no longer good quantum numbers and the eigenfunctions


describing the behavior of the optically active electron must be those
which are specified by the quantum numbers n, I, j, m^ These quantum
numbers obey the same rules as before. Specifically,
l

and

are

*=*
= /+ior/-l,
=o
j = h

/^0

(13-29)

For
/,

AS

0,

f These

For a

equation (13-27) shows that

&E8

0.

For other values of

and the other negative,


L assumes two different values, one positive
relativistic effects are

Is electron in the

82

not so negligible for electrons in the inner subshells.

U atom, the average value of vjc is about 0.5.

excitations

it

However,

this

energy of the core, and therefore in optical


affects only the definition of the zero of total energy.

affects only the fixed value of the total

MULTI-ELECTRON ATOMS

422

[Ch. 13

according to whether j = I + J or j = / \. Except for / = 0, each


energy level is split into two components, one of slightly higher energy
for the spin

and

of the

momenta "parallel" and one of slightly


momenta "antiparallel." The magnitude

orbital angular

lower energy for these angular

splitting is proportional to the expectation value of (1/r) (dV/dr).

Since both 1/r and the derivative of the net potential V(r)
for small

r,

the expectation value

is

become

large

very sensitive to the behavior of

0. In fact, it is so sensitive that it is not possible


y>i im and V(r) near r
to obtain accurate values of the splitting from Hartree theory calculations.

Nevertheless, the Hartree theory certainly predicts the general features of

the experimentally observed values of fine structure splitting.

TABLE
Spin-Orbit Splittings
Subshell

(13-5)
in

the Sodium

Ap

3P

Atom

5p

6p

Spin-orbit
splitting

21.3

For a given

x 10-* ev

x 10-*

7.63

3.05

x 10~ 4

1.74

x 10~ 4

alkali element, the theory predicts that the splitting will

decrease with increasing n because, with increasing values of this

quantum

number, the eigenfunction ip nUm becomes smaller in the region near


r =
where (1/r) (dVjdr) is large, so the expectation value of this quantity
.

decreases.

This behavior

measured energy

levels of

is

seen,

n Na

for instance, in the experimentally

associated with an electron excited to the

The splittings of these energy levels are listed in table (13-5).


we showed that according to the Hartree theory the dependence
of the net potential V(r) on r, near r = 0, becomes more pronounced with
increasing Z. Thus the Hartree theory predicts that dVjdr increases

subshells.

In section 3

with increasing

Z near r

0,

increase with increasing Z.

and therefore that the spin-orbit

splitting will

This behavior can also be seen in the experi-

mental data. In table (13-6) we list for various alkali elements the observed
splittings of the energy levels for an electron excited to the first p subshell.

Note

that the spin-orbit splitting

becomes quite

large for large values of Z.

Figure (13-8) shows a few of the lowest energy levels of the


Levels for which

quantum numbers except n

n Na

atom.

same are
grouped in columns, and the value of n for each level is labeled by the
configuration of the optically active electron. The quantum numbers
/ and j are given according to the standard spectroscopic notation by a
symbol such as 2P^A which is read "doublet P one-half." The letter
specifies the value of / according to the scheme we have become familiar
the

all

are the

ALKALI ATOMS

Sec. 6]

423

TABLE
Spin-Orbit Splittings

Element

Li

n Na

Subshell

2f

2p

(13-6)

Number

in a

19

Atoms

of Alkali

37

55

Rb

Cs

Ap

5p

6P

72 x 10" 4

295 x 10- 4

687 x 10"4

Spin-orbit
splitting 0.42

x 10" 4 ev 21 x 10" 4

with, except that

it is

conventional to use capitals.

of/ The superscript is equal to (2s + 1)


equal to the number of components into which

value

Figure 13-8.

between these

Some energy

P%

(2

subscript gives the

1),

and

the levels are split

is

also

by the

sodiunvatom, and the possible transitions

levels of the

X/i

levels that are

not

split.

Since

= +

/
\ and negative fory
\,
level should be slightly higher than the associated 2 P^ level, and

the spin-orbit energy


2

The

levels.

spin-orbit interaction, except for

each

similarly for the

is

D 5A

positive fory

and

and
3/i

for the

2
A and F

F-,

h/i

levels.

This

MULTI-ELECTRON ATOMS

424
splitting

is,

of course,

much too

however, we have indicated

on the scale of the diagram;


and 3p 2P,A levels.

small to be seen
for the 3p

it

We have also indicated in figure

3/i

(13-8) the possible transitions between

u Na

energy levels which produce the spectral lines of


with the few levels shown.

[Ch. 13

that are associated

If the reader visualizes the

complete

set

energy levels and transitions, he will certainly be impressed with


complexity, and with the tremendous

of
its

amount of work expended by

Rydberg, and others, in obtaining energy level diagrams from an analysis


of the observed spectra. The analysis became possible only after Rydberg
discovered that the spectral lines could be grouped into various series.
For alkali elements there are four important types of series. These arise

from the

transitions indicated below, f

"Sharp"

n^A ~+ n

series:

"Principal" series:

"Diffuse" series:

"Fundamental"

In a particular series, n

doublet of the

for
3

X/i

3.

and

u Na

is

is

VzM

"*

~* n o 2p
*AM

^AM

-+ n *

the

first

is

doublets of lines
doublets of lines

"o S14

and n

fixed

spectrum

2p

PHM

series:

D WA

D WA

triplets

of lines

triplets

of lines

variable.

member

The

intense yellow

of the principal series

That is, the two lines are due to the two transitions 3 2P,A -*
2
P - 3 2 S, All the lines of the sharp and principal series
lyi

are doublets since they involve transitions from a single


closely spaced pair of 2 P, and 2 P\A
A

levels,

or vice versa.

S,

The

level to

lines

of the

and fundamental series might be expected to be quartets since the


and final energy levels are always both members of a pair of levels.
However, it is observed that the transitions n 2 D > n 2P,A and n 2 FA >
n 2 D do not occur. From an analysis of many spectra, it is found that
diffuse

initial

byi

3/i

these are particular examples of a general selection rule for the

number j. Transitions only occur

A/=0, 1
Because of
triplets.

this rule the lines

In figure (13-8)

on the quantum number

(13-30)

of the diffuse and fundamental series are

we can
/.

quantum

for which

also see the operation of a selection rule

Transitions are indicated in the diagram only

when

A/=l
From

a study of the spectra of the alkali elements,

(13-31)
it is

t The series were named in the early days of spectroscopy for


reasons. These names are the origin of the scheme by which the
/

0, 1, 2, 3,

are represented by s,p, d,f.

found that

this

somewhat obscure
quantum numbers

ATOMS WITH SEVERAL OPTICAL ELECTRONS

Sec. 7]

425

on / is obeyed whenever there is a single optically


Thus equations (13-30) and (13-31) are both also obeyed

selection rule

active

electron.

in the

case of the

atom.

We

of selection rules in

shall discuss the origin

section 12.

Atoms with

7.

An atom

Several Optically Active Electrons

of a typical element contains a core of completely

filled

subshells surrounding the nucleus, plus several electrons in a partially

Since any of the electrons in a partially

filled subshell.

may

an optical

participate in

active.

It is

filled

subshell

excitation, all such electrons are optically

apparent, then, that a description of the excited states that

give rise to the optical spectrum of such

the behavior of

an atom requires a description of

the optically active electrons.

all

number of

Despite the fact that

which determine the behavior of


these electrons, it is possible to understand their influence on the energy of
the excited states by treating them one at a time, in order of decreasing
there are a

importance, with each


already treated.

As

different effects

new

effect

considered a perturbation on the effects

we consider each of
moving independently in the Hartree

a starting point of this perturbation treatment,

the optically active electrons to be

net potential V(r) due to the core plus the other optically active electrons.
In this Hartree approximation the energy of each optically active electron

quantum numbers n and /, and the dependence


quantum numbers is very similar to that of
a single optically active electron in an alkali atom with the same core.f
The total energy of the atom is a constant plus the total energy of the
optically active electrons. Consequently, the energy of the atom is deteris

completely specified by

of this energy

Enl on

mined completely

its

the two

in this approximation

by the configuration of the


by the n and / quantum numbers of these
electrons. Now, associated with each configuration of the optically active
electrons are generally a number of different possible sets of quantum
numbers m and m s Since the energy does not depend on these quantum
numbers, there is usually a quantum number degeneracy for each conoptically active electrons,

i.e.,

Also, there is generally an exchange degeneracy because the


energy does not depend on which of the identical particles has a particular
set of quantum numbers. We see that in this approximation a number of
degenerate energy levels are associated with each configuration. However,
many of these degeneracies are removed when the perturbations are taken
figuration.

The reason

is

that V(r)

is

very similar to the net potential due to the core alone.

MULTI-ELECTRON ATOMS

426
into account.

The

situation

is

[Ch. 13

analogous to the case of an alkali atom in

which some of the degeneracies associated with each configuration of the


optically active electron are

removed when the

spin-orbit interaction

is

treated as a perturbation.

The perturbations which we must include correct for the effects that are
The two most important correc-

ignored in the Hartree approximation.


tions are for:
1.

the fact that the Hartree potential acting

on each

optically active

Coulomb

electron describes only the average effect of the

interactions

between that electron and all the other optically active electrons, and
2. the magnetic interactions between the spin and orbital angular

momenta

of each optically active electron.

There are also

relativistic corrections, corrections for interactions

between

the spins of the optically active electrons because of magnetic interactions


between the associated magnetic moments, and certain other corrections;

of them are very small and can generally be neglected.


have seen one manifestation of the residual Coulomb interaction
2
effect (correction 1) in our treatment of the He atom. In this atom the
energy level for the \s2s configuration is split into two levels, one corresponding to the triplet states for which the energy is lowered, and one

but

all

We

corresponding to the singlet state for which the energy


is

true of the \slp configuration.

The reader

is

raised.

The same

will recall that this

happens

because the requirement that the complete eigenfunction be antisymmetric


introduces an interaction between the relative orientation of the spins of

The average distance


when the spins are

the electrons and their relative spatial coordinates.

between the two electrons


"parallel1 than it is in the
'

is

larger in the triplet states

singlet state

when

the spins are "antiparallel."

Consequently the positive Coulomb repulsion energy is smaller in the


triplet states, for which the magnitude of the total spin has the constant
value 5"

Vl(l

1) h,

than

it

is

in the singlet state, for

which the

magnitude of the total spin has the constant value S' = 0. Of course,
the magnitudes of the individual spin angular momentum vectors are not
changed by this effect of the residual Coulomb interaction. It is shown by
an analysis of the experimentally observed spectra, and can also be under2
stood theoretically from calculations similar to those we made for He,
active,
optically
that this effect is important in all atoms with two or more
electrons. That is, for such atoms there is a tendency for the spin angular

momenta

of the optically active electrons to couple in such a

magnitude of the total spin angular


energy

The

is

momentum

S'

usually smallest for the state in which 5"

residual

Coulomb

is

is

way

constant,

that the

and the

largest.

interaction effect also produces a tendency for

ATOMS WITH SEVERAL OPTICAL ELECTRONS

Sec. 7]

the orbital angular


in such a

L'

is

the

way

constant.

momenta of

427

the optically active electrons to couple

that the magnitude of the total orbital angular

We

momentum

did not see an example of this in the excited states of

He atom because

we studied one electron is a


most only one non-zero orbital

in all the configurations

Is electron and, consequently, there is at

momentum present. However, it is easy to understand the origin


of the forces tending to couple the orbital angular momenta by considering

angular

the polar plots of the probability density for an electron


spherically symmetrical potential

[cf.

the z axis so that the orbital angular

Figure 13-9.

maximum

Illustrating

and

momentum has

why two optically active


momentum vector.

its

moving in a
Choosing

(10-5)].

maximum possible

electrons tend to couple to give a

total orbital angular

component along

we

figures (10-4)

that axis,

and considering the corresponding polar

see that the probability density of the electron

is

plot,

concentrated near the

plane perpendicular to the orbital angular momentum vector. Thus the


charge density produced by the electron is largest near this plane. Now,
considering two or more electrons, itis apparent thatthe repulsions between
their charge densities will
their orbital angular

have the

momentum

effect

of producing interactions between

vectors.

acting between the orbital angular

These interactions are torques


vectors which do not tend

momentum

to change the magnitudes of the individual orbital angular momentum


vectors but only to make them precess about the total orbital angular

momentum

way that its magnitude L' remains constant.


Which of the possible values of L' corresponds

vector in such a

The question then

arises:

to the state of smallest energy?

not so easy to answer because


However, the physical basis of
the one which is usually most important can be understood by considering
two electrons in a Bohr atom, as illustrated in figure (13-9). Because of the
Coulomb repulsion between the electrons, the most stable arrangement is
This

is

there are several opposing tendencies.

obtained when the electrons stay at the opposite ends of a diameter.


In this state of minimum energy, the electrons rotate together with individual orbital angular momentum vectors parallel, and with the magnitude
of the total orbital angular momentum vector L' a maximum. This

MULTI-ELECTRON ATOMS

428

[Ch. 13

confirmed by an analysis of the spectra produced by atoms


is, for such atoms there is a
tendency for the orbital angular momenta of the optically active electrons
conclusion

is

with several optically active electrons. That


to couple in such a

momentum
in

which 11

L'

is

way

that the magnitude of the total orbital angular

constant,

and the energy

is

usually smallest for the state

is largest.

In contrast to the tendencies produced by the residual

Coulomb

inter-

action, the spin-orbit interaction (correction 2) produces a tendency for

the spin angular

couple with

its

momentum

own

vector of each optically active electron to

orbital angular

momentum

vector in such a

way

as to

leave the magnitudes of these vectors constant while they precess about

angular

their resultant total

magnitude

/.

We know

momentum

that this

is

due

vector,

which

is

to torques arising

of constant

from the

inter-

moment connected with the spin angular momentum


magnetic field connected with the orbital angular momentum.
know that the energy is smallest for the state in which / is

action of the magnetic

and the

We

also

smallest.

8.

LS Coupling

Because the residual Coulomb and spin-orbit interactions tend to


produce effects which are in opposition to each other, atoms in which both
interactions are of comparable importance are very difficult to treat.
Fortunately, such atoms are in the minority. For atoms of small and
intermediate Z, the effects of the residual

Coulomb interactions

are usually

compared to those of the spin-orbit interactions. We can


see evidence of this by comparing the energy difference (~1 ev) associated
with the residual Coulomb interaction coupling of the two spin angular
quite large

in the first excited states of 2 He with the energy difference


(/~10~ 4 ev) associated with the spin-orbit interaction coupling of the spin

momenta
and

orbital angular

momenta

small and intermediate


is

sometimes

who

first

in the first excited states of 8 Li.

are said to be examples of

also called Russell-Saunders coupling after

LS

Atoms of

which
two spectroscopists
coupling,

explained the structure of the energy levels of these atoms.

In studying coupling of this type, the perturbation due to the residual


Coulomb interactions is treated first since it is the most important, and
the perturbation due to the spin-orbit interactions is temporarily ignored.

Then

the individual spin angular

electrons couple to

form a
S'

Sx

total

S2

momenta

S, of the optically active

momentum

spin angular

S',

where
(13-32)

COUPLING

LS

Sec. 8]

429
satisfying the quantization equation

and where S' has a constant magnitude


S'

Vs\s'

1)

(13-33)

Also the individual orbital angular momenta L 4 of the optically active


electrons couple to form a total orbital angular momentum L', where
L'

=L +L +
2

and where L' has a constant magnitude


L'

L,

(13-34)

satisfying the quantization equation

+l)h

Vl'(l'

(13-35)

These vectors couple in such a way that all their magnitudes S{ and L t
also remain constant. Because of the perturbation, the energy of the atom
depends on S' and L'. Then the quantum states of the same configuration,
but associated with different values of S' and L', no longer have the same
energy.

has the

The

state

minimum

The dominant
account as a
included in

with the
residual

first

LS

maximum

possible values of S'

and L' usually

energy.

Coulomb

interactions having been taken into

perturbation, the smaller spin-orbit interactions are

coupling as an additional perturbation.

This

considering a spin-orbit interaction between the angular


vectors S'

and

L'.

done by

This interaction couples these two vectors in such a

that the magnitude /' of the total angular

J'
is

is

momentum

constant, and S'

L'

way

momentum
(13-36)

S'

and L' remain constant. The magnitude

J'

is

also

quantized according to the usual equation


J'

Vj'ij'

+l)h

(13-37)

As a

result of this additional perturbation, the energy of the

also

on

J'.

minimum

The

state

minimum

with the

atom depends

possible value of J' has the

energy.

Figure (13-10) illustrates the

LS coupling

way

the various angular

momentum

normally the one of


minimum energy for two optically active electrons with quantum numbers
d = 1, .si = i and /2 = 2, sa = . The spin angular momenta S2 and S2v
precess about their sum S', and this vector has its maximum possible
The orbital angular momenta Lx
magnitude (corresponding to s'
1).
vectors

combine in

in the state

which

is

and L2 precess about their sum


magnitude (corresponding to /'

sum J',
to j'

which also has its maximum possible


Also S' and L' precess about their
but this vector has its minimum possible magnitude (corresponding
Finally, J' precesses about the z axis in such a way that its
2).
L',

3).

MULTI-ELECTRON ATOMS

430

component J't along that

axis

is

a constant which

satisfies the

[Ch. 13

quantization

equation
J'z

(13-38)

m'fi

where

m;=-j',-j'+l,...,+j'-h+j'
Figure (13-10)

is

drawn

of the total angular

for

m\

momentum

=/. The
J'

and

(13-39)

quantization of the magnitude

its z

component

J'
z

is

a necessary

consequence of the absence of external torques acting on the atom

Figure 13-10.
in

Illustrating

the coupling between the various angular

a typical IS coupling state of

momentum

vectors

minimum energy.

Figure (13-10) illustrates only one of the quantum states that can be
formed in LS coupling by two optically active electrons with quantum
numbers lx = 1, sx = \ and /2 = 2, s2 = |. In fact, there are twelve
different sets of states, with different quantum numbers s', I',]', that can be
formed by these two electrons. And each of these sets consists of 2/" + 1

quantum states corresponding to the 2j" + 1 different possible


values of m'j. The rule specifying the possible values of m,' is expressed by
equation (13-39). The rules specifying the possible values of s', l',j' are
different

conveniently expressed with reference to vector addition diagrams employ-

ing vectors whose lengths are proportional to the values of the

quantum

LS

Sec. 8]

COUPLING

431

numbers. For the two electrons in question, these diagrams have the form

shown
s\

in figure (13-1

l',j'

shown

1).

The reader may

verify that the possible values of

in the vector diagrams agree with those obtained

from the

equations

=
=

V
j'

s2

\k

+
+

hi \li-k\

\s'-i'\,\s'-r\

= H\{.

h=2

s'=l

+k
+

'i

i,...., s

'

(13-40)

i'

h=2

'->

Jl=l

U=2

s'=0

f=2

l'=3

r=\

s'=l

s'=\
s'=l

f=3

r=3

I'

S '=l

1=3
l'=2

;=>t>

i'=2

/=4

/=3

/=3

j'=2

/=2

/=1

j'=2

\_

-/
'-l, /'=3

i'=l,

f=3,/=3

Figure

13-11.

/2

2, s 2

Since sx

s'=0,

/=1

/=o

\-

f-2

i'=2

s"=0,

-f|-

C=2,/=2

s'=0,

i'=l,/=l

Vector addition diagrams for the quantum numbers

I,

st

J.

s2

$,

the

first

equation gives
s'

0, 1

(13-40')

These equations can be proved by the same kind of arguments as were


used in proving equation (11-44). The equation for s' is, of course,
the

same

as (12-25).

Equations (13-40) are the rules specifying the possible values of s', l',f

MULTI-ELECTRON ATOMS

432

LS

for the general case of

coupling of any two electrons.

[Ch. 13

For any

electrons, the rules can be written


s'

+s +

jsj

+ sN min
\

|si

s2

+ 1,
s + s +
+ lA'Imin + *>
k + k +
+
.

I'

|li

+h+

'

'

Iwlmin'

/'|

'

+U+

I'l

'

'

f=

s'

/'|, Is'

1,

2.

s'

sN

In

(13-41)

Since the magnitudes of the vectors

= 0,1,2,3, ..., N/2,


= \,%,i,l,...,Nj2,

s'

equation gives

s N are all , the first

s'

N=

which reduce to (13-40) for


s1; s 2 ,

sN min

N even
for N odd
for

(13-41')

As an example of the second equation, consider three electrons with


quantum numbers lt = 1, /2 = 2, /3 = 4. For this case the minimum value
of

/'

is

|l

12

l3

7.

|li

+ h + Islmin =

As

min

to k

and the maximum value of

1,

another example,

if

= 1, 2 = 1, 3 = 1,
+ h = 3- We shall
/

lx

/' is l
x

then

/'

/2

+h

ranges from

soon see

that, for

two or more electrons with the same n


and / quantum numbers, the exclusion principle puts certain restrictions on
the possible values of s', l',f in addition to those indicated by equations
configurations in which there are

(13^0) and (13-41).


In the Hartree approximation the eigenfunctions y> T describing the
system of optically active electrons are labeled by the quantum numbers
of n, I, m m s of each of the electrons. In this approximation, all the %p T
associated with a particular configuration of these electrons are degenerate.
x

In the next approximation, the residual Coulomb and spin-orbit interactions of the optically active electrons are treated in LS coupling as
perturbations. These perturbations thoroughly mix the degenerate ip T

form an equal number of new eigenfunctions ip'T, which are certain linear
combinations of the y> T .-\ Each of the %p'T describes a state in which the
optically active electrons couple so as to yield a particular set of S', L',
Consequently the f'T are labeled by specifying the configuration of
J', J'z
to

quantum numbers s', V, j', m-.


LS coupling on the values of
on the configuration, most of the degeneracy is

the optically active electrons plus the

Since the energy of the atom depends in


S', L', J', as well as

removed.
t This

is

The

eigenfunctions

T have only a

tp'

trivial

degeneracy with

completely analogous to the mixing of degenerate one-electron eigenfunctions


(cf. section 5, Chapter 11). As in that case, it will not be

by a spin-orbit perturbation
necessary for us to

know

the exact form of these linear combinations.

Sec. 8]

COUPLING

LS

433

quantum number m'j which specifies the spatial orientation of


momentum vector. The energy levels of the atom are

respect to the

the total angular

LS coupling by the configuration plus s', I', ;", using the spectroscopic notation introduced previously. If we are not interested in the value
of m'j, the same notation can also be used to label the quantum states.
labeled in

For

instance, the energy level

and quantum

/
^-

states

of an atom in which

*=2

,'=2

l=l

j'=3

^F
i

/
/

ndn'p

y'=?,l,0

.'-1

y^Pl

-J^JLQ^

T=l

_L^i__ _H_L
)'=

3,2,1

^0
3l) 3
2

\\

/=
r=

4,3,2

%
Figure 13-12.

The

splitting of the

energy levels

in a typical

LS coupling configuration.

have the quantum numbers = 1,


and couple in such a way that s' = 1, /' = 1,
2, are identified by the symbol 1 sip 3P2
The first part of the symbol
gives the configuration, and the last part gives the other quantum numbers
according to the scheme described in section 6.
The splitting of the single degenerate energy level of a particular
configuration, say the nd n'p configuration of an atom with two optically
active electrons, due to the residual Coulomb and spin-orbit perturthe

=
/=
/

two optically
and n = 2,

active electrons

1,

bations

is illustrated in figure (13-12). For obvious reasons


we cannot
present explicit equations from which the energies of all these levels can
be evaluated. However, we can write an equation which gives the

dependence of the spin-orbit interaction energy. This is the effect which


produces the fine structure splitting of the levels for s' = 1, and a given
/', into triplets of levels.
Consider equation (13-27), which is
AEg.r.

'

2
4mb(r^r)
c

'

1)

/(/

1)

s(s

1) ^

This equation was developed for an atom with one optically active electron,
but the same equation written with primes gives the interaction energy of

434

MULTI-ELECTRON ATOMS

the total spin and orbital angular

momentum

and L' of any


atom
in which
any
energy of the atom

vectors S'

which these vectors can be defined, that is,


is valid. Thus in LS coupling the total
contains a term

atom

LS

in

coupling

AEs

..

K[j'(j'

1)

/'(/'

[Ch. 13

1)

s'(s'

(13-42)

1)]

where the quantity

K=

dV{r)

4m V

dr

has the same value for all energy levels of a configuration with common
values of s' and /'. Let us calculate from this equation the separation in

TABLE

(13-7)

Fine Structure Splittings in the Calcium

Atom
Ratio

Con-

4s4p
4s3d

P1; 3P
P1 3P

3d3d

D 2 3 D1
3
D 2 3 D

3d4p

Levels

Separation

Levels

figuration

4
16.7 x 10" ev
4
64.9 x 10" ev

x 10" 4 ev
4
33.1 x l(T ev
16.9

VPi
P2> 3P1
*D 3 3 D 2

D3 3D
,

Separation

Exp.

Theo.

33.3 x l(Hev
131.2 x 10" 4 ev

1.99

2/1

2.02

2/1

26.9 x 10" 4 ev

1.59

3/2

49.6 x 10~ 4 ev

1.50

3/2

energy between the adjacent levels of a configuration with common values


of s' and /', which are called energy levels of the same multiplet. If the
quantum number associated with the level of lower energy is/, the quantum

number

associated with the level of higher energy

separation
K[()

is j"

1,

and the

is

- s'(s' +

1)0"

2)

/'(/'

1)]

+ 1) - s'(s' + 1)] - K[j'(j' + 1) = K[(j' + l)(j' + 2) -/(/' + 1)] =

/'(/'

2K(f

1)

1)

(13-43)
separation in energy of adjacent levels of the same multiplet is
proportional to the total angular momentum quantum number of the
level of higher energy. This is called the Lande interval rule. In table
(13-7) we present some energy separations between adjacent levels of

The

20
Ca atom, and
multiplets which are observed in the
of these separations with the ratios predicted from the

compare the ratios


Lande interval rule.
Since we compare ratios, the quantity 2K cancels out. The excellent
agreement between the experimentally observed and theoretically predicted ratios, which is typical for atoms of small and intermediate Z,
provides convincing evidence of

LS coupling in

these atoms.

Sec. 8]

COUPLING

LS

435

In discussing the coupling of the angular


active electrons,

we have not

momenta

of several optically

yet considered the effect of the exclusion

But, for configurations in which two or more of these electrons


have the same n and I quantum numbers (are in the same subshell), the
exclusion principle cannot be ignored because it imposes important
restrictions on the way in which their angular momenta can couple.
To determine these restrictions, we consider the Hartree approximation
principle.

in

which

electrons.

n,

I,

m m

Then

f,

are

good quantum numbers

for the optically active

the exclusion principle states simply that

can have the same

set

of four quantum numbers.

We

no two electrons

shall see that this

/', w ',y", wj which


components of S', L', J', and which also
are all good quantum numbers in the Hartree approximation. Now,
although in LS coupling the 2 components of S' and L' are changed by
the application of the residual Coulomb and spin-orbit perturbations,
S', L', J', J'
are not changed.! Consequently, any restrictions we find in
z
the Hartree approximation concerning the quantum numbers s', l',j", m'j

limits the possible values of the

quantum numbers s', m's

determine the magnitudes and

will

be equally applicable in

Consider

first

LS

coupling.

a configuration ns 2

same quantum number n and so are

in

which two

s electrons

have the

same 5 subshell. Since for


it is only possible that m
0, these two electrons actually have the
/ =
same values of the three quantum numbers n, I, m To satisfy the exclusion
principle, their m s quantum numbers must be different; one electron
must have m s = + and the other must have m s = J. According to
(12-23), (12-24), and (11-21), m's is equal to the sum of the m s quantum
numbers of the two electrons. Thus the only possible value for this
quantum number is m's = + J \ = 0. Similarly, m[ is equal to the sum
of the m, quantum numbers of the two electrons, and so has only the
single possible value m\ =
+ = 0. According to (1 1-43), m- = m[ +
m'g so the only possible value of this quantum number is m'- =
+ = 0.
Since the three quantum numbers m's m{, m'j can have only the single
value 0, it is apparent from equations such as (12-24) that the quantum
numbers /, /',/ also can only have the single value 0. Thus the two electrons can couple to form only a single quantum state. This is the state
X
SQ in which the spin vectors of the two electrons are "antiparallel."
If it were not for the exclusion principle the m s quantum numbers of the two
electrons could be equal, and there would be the three additional S S
1
states in which the spin vectors are "parallel." These conclusions should be
in the

t In the presence of the perturbations, S'z and L' have no definite values and m', and
z
ml are not good quantum numbers. However, since S', L', J', J'z have definite values
whether or not the perturbations are present, it is apparent that the application of the
perturbations cannot change their values.

MULTI-ELECTRON ATOMS

436
familiar to the reader because
Is 2

ground

we have already used them

state configuration of the

He

Consider next the configuration np 2

[Ch. 13

in the case of the

atom.

in which two optically active


same p subshell. This case is more complicated, but
it can be handled by the same procedures used above.
In table (13-8) we
list all the possible sets of values of m and m for the two electrons which
s
,

electrons are in the

Possible

Number
1

2
3

4
5

6
7
8

TABLE (13-8)
Quantum Numbers for an np 2

m,

m.

m,

m.

+1
+1
+1
+1
+1
+1
+1
+1

+2-

+1

~"21

+h
+i
+i
+i

2
1
2"

+1
+*

15

do not

+2

+1
+1

+2
+1
+1

+1

-1

+1

-1

2
1

+1

-1
+1

-1
-1
-1

-1
-1

-1
-1

+i
+i
+i

-1
-1
-1

14

+1

+2

>

11

13

-1
-1
+1

mi

2
1
2

12

+1

+i

10

mi

m'

2-

-l
-l

"2

Configuration

+1

+1

2
1

_1

-1

_i

-2

For each

violate the exclusion principle.

sponding values of the quantum numbers m'K

-1

-1

set

we

also

m[, m'}

list

-2
-2

the corre-

There are 15
different sets of m and m s quantum numbers for the two electrons which
satisfy the exclusion principle, and a number of others, such asm, = +1,
m s = +J; i= +1, m s = +J, which are ruled out because they violate
this principle.
The problem now is to identify the allowed quantum
states, specified in Table (13-8) in terms of the quantum numbers m'
s
,

m'} with the


numbers s', l',j".
m'j,

specification of these states in terms of the


First

we

use equations (13-40), with

lx

l2

quantum
= 1, and

find that the various possible combinations of*', l',j' values are only those

which are

specified in the following list by the spectroscopic notation:


3
3
3
PV 3P2 *D lt 3 D 2 3 D3 The *Da quantum states are
^o, !/>!, *D S 5l7 P
immediately ruled out because, for these states with j" = 3, equation
(13-39) shows that there would necessarily be wj values of +3 and 3,
but there are none in the table. Since there are no 3 > 3 states, there can be
,

no

D 2 or 3 Z)

states.

All these states correspond to S'

and L' vectors of

Sec. 8]

COUPLING

IS

the same magnitude

437

they are states of the same multiplet, and they stand

must
and
be quantum states with s'
and
m's
s',
=
the
be
satisfied
only
by
These
requirements
can
m{
V
V
or

fall

Now,

together.

entry

number

quantum
values

states

s'

in table (13-8) says that there

2, since

s'

D2

1, 2.

such states corresponding to the five


Entry number 2 says that there must be

This requires the presence of the states

There are

= 2, 1,

m'j

states with
3

/'

0,

and

/'

five

1.

PX 3 P2 For 3P there is one quantum state corresponding to wj = 0,


S
P1 there are three states corresponding to m- = 1, 0, 1. For 3P2

3
,

For

= 2, 1, 0, 1, 2, The number of
+ 1+3 + 5=14. Only a single state
because all the other m'^ values
is left, and this must be a state with wj =
there are five corresponding to
states

we have

identified so far

of the table have been used.

quantum
quantum

then that this must be the single

S 1 D 2 3P 1 2 Since
quantum numbers s', l',j',

expressed in terms of the

LS

It is clear

Thus, in the Hartree approximation the possible


2
states for two electrons with the configuration np are only those
state

associated with the symbols

in

m,'

is

these restrictions are

they are equally valid

coupling.

In table (13-9)

we

list

the

quantum

states

which are allowed by the

exclusion principle for configurations containing several electrons in the

same subshell. Each symbol gives the values of the quantum numbers
and /' of an allowed multiplet of quantum states. The possible values
j'
of
and m'j for the states of any multiplet can be determined in terms of
Entries are given for cons' and /' from equations (13-39) and (13-41).
ranging
from
no
electrons
in
the
subshell
up to the maximum
figurations
number of electrons consistent with the exclusion principle. For no
1
For
electrons, s' = /' =j' = 0, which is described by the symbol S
one electron in any subshell, /' = / and, since s' = \, the allowed states
comprise a doublet. The allowed states for the other configurations are
determined by calculations similar to those we used in determining the
allowed states of the ns 2 and np 2 configurations, or by more elegant
calculations based on the mathematical theory of groups.
s'

It is particularly

about the

half-filled

interesting to note the symmetries in table (13-9)

subshell configurations.

greatest for this configuration,

a subshell

is filled

and the

The number of

states is

states for a configuration in

which

except for k electrons are exactly the same as the states

for the configuration in

which there are

just

k electrons in the subshell.

a striking demonstration of the effect of the exclusion principle


because, if it were not for this principle, the number of states would increase

This

is

monotonically as the number of electrons in the subshell increased.


There are several interesting consequences of this property. The most
important is that when a subshell is completely filled the only allowed

MULTI-ELECTRON ATOMS

438

TABLE
Possible

Quantum

(13-9)

States for Configurations Containing

Several Electrons in the

ns 1

3
l

np"

S,

tip

P,

tip*

tip

Subshell

ns 2

tip

Same

*S

ns

tip

[Ch. 13

S,

*S

nd"

S,

nd 1
nd 2
nd 3
nd 1

*S
3

np e

D, *G
2

>,

P,

D, F,

G,

P,

*P,

*F

3
D, 1 G, 1 S, 1 D, 1 G, 1 H, H
P, 3 F, 3 P, 3 D, 3 F, 3 G, 3
2
i
D, 2 P, 2 D, 2 F, 2 G, 2 H, 2 S, 2 D, 2 F, 2 G, 2 I
P, i F, i D,
1
3
S, 1 D, >G, 1 S, 1 D, 1 G, 1 H, H
P, 3 F, 3 P, 3 D, 3 F, 3 G, 3
2
D, 2P, 2 D, 2 F, 2 G, 2
*P, i F
1
3
S, 1 D, G
P, 3 F
X

nd*

S,

S
i

nd e
nd 1
nd 3

Z>

nd'

nd 1 "

Z)

*S

same as when the subshell is completely empty. This is the


1
5 in which S', L', and J' all vanish. We proved this for
the s subshell, and it is easy to prove for any other subshell. Let us
consider the np 6 configuration and construct table (13-10), which is
state is the

quantum

state

Possible

Number

Wj

TABLE (13-10)
Quantum Numbers for an np 6

mi

similar to the table

we used

mi

completely

filled

that the total spin

and

is

m m
t

tn's

m\

in',

+1-1-40

up 2 configuration. The table

obviously the single state

orbital angular

momentum

of a

is confirmed by analysis of the optical


Furthermore, if there is no net spin or orbital

subshell are zero

spectra of noble gas atoms.

angular

m m

in studying the

has only a single entry, and this entry

The conclusion

-i -1

+|

+1 +| +1 -i

Configuration

momentum,

no net magnetic moment, so this conby Stern-Gerlach experiments on noble

there can be

clusion can also be confirmed

gas atoms

ground

439

COUPLING

LS

Sec. 8]

or by the fact that the magnetic moment of

47

Ag, which in the

a single 5s electron outside a set of completely filled


measured in these experiments to be the same as the magnetic

state contains

subshells,

is

moment

We

of 1 H, which in the ground state contains a single Is electron.


are now in a position to understand the various features of a diagram

showing the optically excited energy

levels

rH

a,"
iH

of
CO

r5^ ,P

JP

of

CO

atom of small or

of a typical

ja.

CO

jp

CO

r-l

CO

j
CO

r\

U
1

1 _

~2p5s~~

-2 _

2p4s~

2p4d_

2p4p

2p3d

-3

~2p3p

-4 -

2p3s~~

-5

-6

-7

-8
-9
-10
-11

2>

-12

Figure 13-13.

Some energy

levels of

the carbon atom.

Consider the energy level diagram for 6 C presented in


state of this atom has the configuration
ls*2s 2 2p 2 so there are two/? electrons which are optically active. The zero
of the energy scale is so defined that the magnitude of the total energy of

intermediate Z.
figure (13-13).

The ground

atom

ground state is equal to the energy required to ionize singly


Consequently, the diagram is^directly comparable with energy
X
level diagrams for alkali atoms and H, in which the zero of energy is
the

in its

the atom.

defined in the same way.

The energy levels

are labeled by the configuration

of the two optically active electrons, and by the spectroscopic symbol


specifying s', /', /". The fine structure splitting according to the value of
y"

is

not shown because

it is

much too

small to be seen on the scale of the

MULTI-ELECTRON ATOMS

440

diagram.

Consider

[Ch. 13

the average energy of the levels of the various

first

configurations.

In the configuration of lowest average energy, 2p 2 both


electrons remain in the same subshell that they are in for the ground state
,

of the atom.
subshell

In the other configurations, one electron remains in that


is in a subshell of higher energy. Note that the average

and one

energies of the configurations

depend on the n and

quantum numbers of
same way as if

the electron in the higher energy subshell in essentially the


this electron

were the single optically active electron in an

alkali

atom.

The

configuration of lowest average energy provides an example of an np 2


configuration for which the exclusion principle allows only the 1 5 1 2

and

,i,2

The

states.

-P

,i,2

states are

of lower energy than the

S and

because they correspond to a larger value of s', and the 1 D 2


states are of lower energy than the
state because they correspond to a
larger value of /'. Note that the s' dependence is stronger than the /'
1

Z> 2 states

dependence.

almost always found that the energy associated with


interaction coupling of the spin angular momenta
larger than the energy associated with the residual Coulomb interaction

the residual
is

It is

Coulomb

coupling of the orbital angular momenta.

energy levels for the


the one for the

-P

,i,2

Of

the three closely spaced

would be seen on a

states that

of lowest energy because

state is

smallest value of7".

The exclusion

Thus the ground

state of the

larger diagram,

corresponds to the

it

atom

is

the state 2p 2

principle does not apply to the configurations other than

the one of lowest average energy, because in these configurations the


electrons are in different subshells.

Consequently,

all

the states predicted

by equation (13-40) are found. In the 2p3s configuration the energy

maximum

corresponding to
figuration.
/'

s' is

level

lowest in energy, just as in the 2p 2 con-

Deviations from this rule, and from the rule that the

maximum

gives the lowest energy, are seen in the configurations of higher average

energy, but there are


gives the

minimum

no deviations from the rule that the minimum j"


Not shown in the diagram are a few energy

energy.

which have been identified as belonging to the configuration 2s2/j3


The optical spectrum of the 6 C atom can be constructed from its energy
level diagram by evaluating the energy and frequency of quanta emitted
in all possible transitions which do not violate the following selection rules
levels

1.

Transitions can occur only between configurations which differ in

quantum numbers of a

the n

and

more

electrons cannot simultaneously

2.

Furthermore, the

configurations

must

single electron.

differ

by one unit

A/=l
This

is

the

same

This means that two or

make transitions between subshells.


quantum number which is different in the two

rule as (13-31).

(13-44)

Sec. 9]

JJ

COUPLING

441

Transitions can occur only between states in these configurations


for which the differences in the s', l',j' quantum numbers are
3.

As'

A/'=0, 1
A/'

0,

(13-45)
(but not;"

->;"

0)

The first equation prohibits transitions between singlet and triplet states,
and vice versa. Nevertheless, transitions are observed between the

because all excitations of atoms to


population of the 2p 2 D 2 states and,
when these states are highly populated, the total number of transitions
3
per second to the 2p 2 -P ,i,2 states becomes appreciable even though the
probability that any single atom will make this transition is extremely

2p

2 1

Z) 2 states

and the

2/>

2 3

,i,2

states

singlet states eventually lead to the

small because

above
is

the

which

9. JJ

is

it

The second of the equations


The third equation
selection rules apply in any atom

violates the selection rule.

similar to (13-31) but not quite so restrictive.

same rule as (13-30). All these


is an example of LS coupling.

Coupling

Atoms of

small Z, such as

C, are

known

to be

good examples of

LS

coupling because the energy levels are grouped in the characteristic LS


multiplets with ay" dependence of the energy that is very small compared to

LS

JJ coupling

coupling

Qi,H) 2

/
s-

3p

/
/

_
6

u Si

2p3s

3pAs

32
Ge
4p5s

\
j0

Sn

in fie

6p7s
Figure 13-14.

from LS to
McGraw-Hill Book Co.,

Illustrating the transition

Introduction to Atomic Spectra,

J) coupling.

New

From H.

E.

White,

York, 1934.

and /', and because the Lande interval rule (13-43)


However, with increasing Z the j' dependence
is accurately obeyed.
rapidly becomes larger, and the Lande interval rule is obeyed less accurately. This is shown schematically in figure (13-14), which indicates the
the dependence

on

s'

MULTI-ELECTRON ATOMS

442

[Ch. 13

disposition of energy levels in the 2p3s configuration of 6 C, and of


the levels in corresponding configurations of higher
atoms that are in

same column of the periodic table. Significant departures from LS


coupling are often seen in atoms of large Z because, as we saw in section
6, the spin-orbit interactions become large in these atoms.
In fact, for
some of the largest Z atoms the perturbation due to the spin-orbit interactions becomes appreciably larger than the perturbation due to the residual
Coulomb interaction. These atoms are said to be examples of JJ coupling.
In coupling of this type, we assume first that the dominant perturbation
due to the spin-orbit interactions causes the spin and orbital angular
the

momentum

vectors of each optically active electron to couple, forming a

momentum vector
L and S, couple to form

total angular

electron

for that electron.

Thus, for the

/th

J*

The magnitude of this vector

is

Ji

The energy of

L,

(13-46)

S,.

quantized according to the usual equation

VjAji

I)

(13-47)

ft

atom then depends on the quantum numbers j of all


is smallest when these quantum numbers have their
smallest possible values. Next, the weaker perturbation due to the Coulomb

interactions

the

and

the electrons

taken into account. This perturbation couples the individual

is

momentum

total angular

vectors to

form the

total angular

momentum

vector
J'

31

J2

J,

(13-48)

which obeys the same quantization equations (13-37), (13-38), and

LS coupling. The

energy of the atom then also depends, but


on the quantum number/. However, it is not possible
to make a general statement about the values of the /' which yield the
smallest energy. The possible values of the/ and of/" are given by obvious
adaptations of equations (13-41). These quantum numbers must be used
to identify the quantum states and energy levels. t In figure (13-14),
the four levels of the 6p7s configuration of 82 Pb are labeled with the
spectroscopic notation for these quantum numbers. This atom is a fairly
good example of JJ coupling because the (f, %) 12 levels, for which j\ f
and j2 = \, are of nearly the same energy and are well separated from the
(4> l)i,o levels, for which j\ = j2 = \. The subscript gives the value of/'.
Note that the smallest/' gives the smallest energy in one pair of levels and
the largest energy in the other pair. The exclusion principle operates in
(13-39), as in

to a lesser extent,

Of course,

of //coupling.

s'

and

/'

are not

good quantum numbers

in

an atom which

is

an example

THE ZEEMAN EFFECT

Sec. 10]

443

JJ coupling, for configurations in which two or more electrons are in


the same subshell, in much the same way as in LS coupling. Selection
rules for JJ coupling are the same as those for LS coupling except that the
restrictions on As' and AV are replaced by the restriction
A/
for the electron

all

(13-49)

which makes a transition between


A/,

for

0,

subshells,

and

(13-49')

the other electrons.

We do not give a detailed discussion of JJ coupling here because


examples of it are not very frequently found in the behavior of the optically
active electrons in atoms. Most atoms are either good or fair examples
of LS coupling However, examples of JJ coupling are very often found
in the behavior of the protons and neutrons in nuclei. We shall study this
Chapter 16, which concerns the nuclear shell model. The
model explains many properties of nuclei by assuming that the motion of
the protons and neutrons is the same as the motion of the electrons in an
in section 8,

atom, except that the net potential V(r) has a different

dependence, and

that these particles experience a very large spin-orbit interaction.

10.

The Zeeman

Effect

was observed by Zeeman that placing an atom in a magnetic


lines which it emits into several components.
This is called the Zeeman effect. For fields up to a few thousand gauss,
the splitting is proportional to the strength of the magnetic field and is
small compared to the fine structure splitting of the spectral lines of the
atom. This means that in the presence of a magnetic field the energy levels
of the atom which are involved in the emission of the spectral lines must
In 1896

field splits

be

it

each of the spectral

components. In certain special cases it was possible


splitting of the energy levels in terms of a
classical theory developed by Lorentz. But, in the general case, even a
qualitative explanation of the observed splittings could not be given until
the development of quantum mechanics and the introduction of electron
split into several

to understand the

Zeeman

spin.

In terms of the modern theory the Zeeman effect is easy to understand.


Except when it is in a 1 S state, an atom will have a net magnetic moment
(x due to the spin and orbital magnetic moments of its optically active
electrons.!
t

The other

According to equation

(1

1-15),

when

electrons are in completely filled subshells

this

magnetic

moment

which do not contribute

to

is

|i.

MULTI-ELECTRON ATOMS

444

magnetic

in the external

field

will

it

[Ch. 13

have an orientational potential

energy

A=

-(ji-H

or

AE=-/j, k H
where

/j,

the

is

be

level will

component of p along the


/n

To

direction of H.

Each energy

energy levels corresponding to the several

split into several

possible values of

(13-50)

calculate these values,

we

evaluate

(11-10) and (11-22) for each optically active electron.

fl

That

y.
is,

by using

we

take

2ii + 2Zs

where the summation goes over all the optically active electrons, and
where (11-9) and (1 1-26) are used to evaluate g and gs Now let us assume
that the atom in question is an example of LS coupling. Then
.

H
Note that

(jl

is

= _

(13-51)

2S']

not antiparallel to

J'

unless 1/

[L'

L'

S'

or S' =0, because the orbital and spin g factors are different.
coupling, L' and S' precess about J' because of the Larmor pre-

In LS
cession of S' in the atomic magnetic field associated with L', with a
precessional frequency m l which is proportional to the strength of
It is apparent from equation (13-51)
with the same precessional frequency.
In the presence of the external magnetic field H, there will also be a
precession of \l about the direction of this field with a precessional frequency which is proportional to its strength. Thus the motion of ^ is

the atomic magnetic field


that

(x

must

also precess

quite complicated.

(cf.

1-29).

about

However,

J'

if

tne external field

is

weak compared

to

about the external field will be slow


compared to its precession about J', and it is not so complicated to
evaluate the orientational potential energy &E. We have seen in section 4,
Chapter 11, that the strength of an atomic magnetic field is typically of
the atomic

field,

the precession of

4
the order of 10 gauss.

[>.

which is weak
Consider, then, the case of an external magnetic field
4
about J'
rapidly
more
much
precesses
(x.
~10
Since
gauss.
compared to

Sec. 10]

THE ZEEMAN EFFECT

than about H, we

component of
fij,

jji

may

^ H by

evaluate

in the direction of J',

in the direction of

445

H. That

finding /n r which is the average


and then taking the component of
,

is,

-V-- 3 '-

/uJ

(L'

2S')-(L'+S')
J'

and

r g-J?
j'//

rr

where we have chosen the


the dot product gives

gg (L'

j'

S'

J'

IS' 2

(i/

+ sy

31/

S')

1
J'

Writing equation (11-48) with primes,

2S>)

be in the direction of H. Evaluating

z axis to

m = - h (U
31/

|(J'

we have

L'

S'

2
)

So

B= -Pg (L' +
2

ft

2S' a

^ H=
Then according

^B

(3J'

+S' 2

"T

fJ'

-L

2}s

fL'

|-5' 2) li-

j' 2

2
)

JT ,

to equation (13-50), the orientational potential energy is

a
=
AE
i7

S' - L
+
is
^
L_ OJ'
^
= VbH
H
2

~f* B

2
)
>

j>

(13 _52)

The value of
to

this term is of the order of


p B H, which is small compared
the other effects that contribute to the total energy of the atom.
instance, the spin-orbit interaction energy is large compared to
this

all

For

term because it is of the order of ju times the strength of the atomic


b
magnetic field, which is large compared to H. Consequently, we may use
perturbation theory to calculate AE, the contribution of AE to the total
energy of the atom. Using the LS coupling eigenfunctions y' which
are
T
labeled by the quantum numbers s', l',j', m', we havej
,

AE =

y>'*

J
t

Although the eigenfunctions

we may use

yi'
T

AEoWt dr

are degenerate with respect to the

(13-53)

quantum number

the ordinary perturbation theory expression (13-53) to evaluate


because J', has a definite value whether or not the perturbation is present,
so the perturbation cannot produce a large mixing of the eigenfunctions.
m-j,

MULTI-ELECTRON ATOMS

446

[Ch. 13

where the integration is taken over all the coordinates of all the optically
active electrons, and where AEop is the operator formed from AE by replacing S' 2 L' 2 J' 2 J'z with their associated operators. Since
,

function of
/(/'

1)#

2
,

all

these operators with eigenvalues

m'jh, respectively,

s'(s'

T is an eigen-

tp'

\)h 2 ,

/'(/'

l)^

2
,

equations (13-52) and (13-53) immediately

give

pbH

\mr +

m
mr m

+
+

As'

vn

- iv +

"V

+%
2

p%(g=%)-

-#:

-%

-H

+H
*S H (g

Figure 13-15.
transitions

It is

2).

No magnetic

Magnetic

field

field

The Zeeman

between these

splitting of a typical pair of

energy

levels,

and the possible

levels.

convenient to write this as

A =

(13-54)

ftBHgm',

where

j'(j'

1)

+ 1) mr + 1)
s'(s'

I'd'

1)

(13-54')

Lande g factor. Note that g = 1 = g for


= 0, as would be expected. From equation
(13-54), we see that in the presence of an external field of strength
each energy level will be split into 2/' + 1 components, one for each of
the values of wj, and that the magnitude of the splitting will be different
for levels with different Lande g factors.
We illustrate equation (13-54) by the energy level diagram of figure
2
2
(13-15), which shows the splitting of P,AM and S,A energy levels in a

The quantity g is
s' = 0, and g = 2

called the

=g

for

/'

THE ZEEMAN EFFECT

Sec. 10]

447

weak magnetic field. Also shown are the possible transitions between the
P ?/i and the 2 S,A states (which, of course, are in a different configuration).
Transitions do not occur if they violate the selection rule

lyi

Am;.

Even with

0,

(but not m}

-> m;

if

A/ =

0)

(13-55)

Zeeman

effect splits each of the


which can be quite complicated
for lines arising from transitions to or from states of large y". However,
all spectral lines arising from transitions between singlet states are split
into a simple pattern of two components symmetrically disposed about
a third component which has the same frequency as the single zero field
for singlet states so all the g factors have the same
line, because s' =
value, g = 1. By drawing a diagram similar to the one in figure (13-15)
for transitions between singlet states, the reader should show for himself
that all the transitions which satisfy (13-55) lead to only three spectral
lines, even though there may be more than three possible transitions,
this

simplification, the

spectral lines into a pattern of components

because there are only three possible energy differences for these transitions.

We mentioned before that in certain special cases it was possible to explain


the

Zeeman

effect in

terms of a classical theory. The reader can probably

guess that these were the cases in which the spectral lines are split into
three components, because these cases correspond to transitions between
states in

role as s'

which the non-classical phenomena of spin does not play its usual
= 0. These cases are called the normal Zeeman effect; the other,

and more

general, cases are called the anomalous Zeeman effect. This


terminology was introduced long before the modern theory provided a
complete understanding of the Zeeman effect and, from the point of view
of this theory, it is not very appropriate.

An

H, which is weak compared to the atomic


couple S' and L' to form J', cannot disturb this
coupling and only causes a relatively slow precession of J' about the
external magnetic field

magnetic

fields that

direction of H.

However,

if

overpowers these

is

strong compared to the atomic magnetic

and destroys the coupling of S' to L'.


In this case S' and L' precess independently about the direction of H.
This is the so-called Paschen-Bach effect, which is observed for external
fields, it

fields that are

fields

comparable to or larger than

= _ M L +
/

(jl

~104

2S')

and
/,

=_?(l; +
n

2s;)

gauss.

In this case,

MULTI-ELECTRON ATOMS

448

where we have chosen the

z axis in the direction

AE = &(L', +

[Ch. 13

of H. Then

2S'.)

From

we immediately

this

obtain

A = fiB H(ml +
external field
the selection

(13-56)

good quantum numbers for an atom in a strong


and
because L'z
S'z have definite values. It is observed that
rules for these two quantum numbers are

Of course, m' and m's


l

2m's )

are

Am;
Am;

=
=

(13-57)

0,

All spectral lines are split by the Paschen-Bach effect into three

ponents, just as in the normal

Zeeman

com-

effect.

The Zeeman effect plays an important role in experimental spectroscopy. By analyzing the Zeeman splitting of the spectral lines of an atom,
the spectroscopist determines the Zeeman splitting of the energy levels of
the atom. This can be used to identify the quantum number j' of each
level because 2/" + 1 is equal to the number of components into which the
Furthermore, the magnitude of the splitting between any
being known,
two components gives the value of fiB Hg and, \i B and
the
of g depends
Since
value
level.
energy
of
for
the
gives
value
this
the
g
on s', /', j", it can be used to confirm the assignment of these quantum

level is split.

numbers. The quantum number s' is assigned by counting the number


of levels of the multiplet of which the level in question is a member. This

number
number

is 2s'
/'

(except that

when

s'

for the levels of a multiplet

>
is

/' it is

21'

1).

The quantum

assigned, after assigning

s',

by

analyzing the separations between adjacent levels in terms of the Lande


interval rule (13-43), and finding the value of/' which, in consideration of
the restrictions imposed by the third of equations (13-41), leads to the
values of /" that give the observed separations.! This procedure also
evidently leads to the assignment of the

numbers

/'

quantum numbers/. The quantum

for all the levels of a particular configuration being

known

can be grouped into configurations by the similarity of their


energies), the / quantum numbers of that configuration are identified by
using the second of equations (13-41). Identification of the n quantum
numbers associated with the various / quantum numbers is not difficult,

(the levels

if

the n

quantum numbers of

the ground state configuration are

f This is applicable only in the usual case in which the


coupling.

atom

is

known,

an example of

LS

HYPERFINE STRUCTURE

Sec. II]

by making use of the


values of

449

energy of the subshells with

fact that the

increases monotonically with increasing n.

The

common

identification

of the n quantum numbers of the ground state configurations of the atoms


is

based on the same

fact.

Hyperfine Structure

II.

When an

interferometer

atom with extremely high

is

used to measure the spectrum of a typical

resolution,!

many of

the spectral lines are

found to consist of several components. The separation between these


components is typically three orders of magnitude smaller than the separation characteristic of the fine structure of spectral lines. These very closely
spaced components constitute the hyperfine structure of the spectral lines.
They imply the existence of a hyperfine splitting of the energy levels of the
atom.J
Hyperfine structure was discovered in 1891 by Michelson. It was first
explained in 1924 by Pauli as a result of the orientational potential energy
of a magnetic dipole moment associated with the atomic nucleus, in the
magnetic field associated with the motion of the atomic electrons. The
nuclear magnetic dipole moment (/.; is due to a nuclear spin angular momentum I. The magnitude and z component of I are quantized according to
the equations

/= Vi(i +
L = mfi

1)

h
(13-58)

where

m = +
t

/,

1,

+i

1,

+/

These are the same as the quantization equations for the electron spin S.
However, the equations relating the nuclear magnetic dipole moment to
the nuclear spin,

Pit

are not quite the

same

= giP-X m

as (11-22)

and

t Interferometers are described in section 3,

We

(13-59)

H
i

(11-23),
Chapter

which

relate the electron

1.

all atoms have


same atomic weight, as well as the same atomic number. For a source containing
atoms of the same atomic number but a variety of atomic weights (a variety of "isotopes"),

are considering only the spectrum emitted by a source in which

the

there

is

an additional multiplicity of the spectral lines, but this is primarily a trivial


from the slight dependence of the reduced electron mass ji on the atomic

effect arising

weight.

MULTI-ELECTRON ATOMS

450

magnetic dipole
of unity, but
quantity

ft

moment to

N which

fi

some

nuclei

called the nuclear magneton,

is

The factor gt is of the order


and negative for others. The

the electron spin.

positive for

it is

_
=

[Ch. 13

eh

eh

2Mc

1836 2mc

is

1836

defined as
...

uB

(13-60)

mass of the nucleus of a X H atom, and m is the mass of an


electron. Since the nuclear magneton ft N is three orders of magnitude
smaller than the Bohr magneton fi B nuclear magnetic moments are smaller
than electron magnetic moments by the same factor. Consequently the
where

is

the

moment

orientational potential energy of the nuclear magnetic


structure)

is

potential energy of the atomic magnetic

moment

(fine structure),

both magnetic moments experience comparable magnetic

The

interaction of the nuclear magnetic dipole

magnetic

(hyperfine

three orders of magnitude smaller than the orientational

field

couples the nuclear spin angular

the total angular

momentum

vector

J of

because

fields.

moment with the atomic


momentum vector I to
Because of

the electrons.!

this

coupling, both vectors precess about the resultant grand total angular

momentum

vector

=I+

(13-61)

The magnitudes of I and J remain constant, but their z components do not.


Consequently, when the hyperfine structure interaction is considered,
m and m s are not good quantum numbers. However, the two quantum
numbers which specify the magnitude and z component of F according to
t

the equations

F=Vf(f+
J J

\)h
(13-62)

F = mf h
z

where

mt =

-/,

-/+1,...,

+/-1.+/

good quantum numbers and can be used to identify the quantum


and the energy levels. For instance, the term in the total energy of
the atom which gives rise to the hyperfine splitting of the energy levels
can be written
are

states

A,
This
is

,
.

K[f(f +

1)

- j(j +

1)

i(i

completely analogous to equation (13-42);

is

1)]

(1

3-63)

equation (13-63)

obtained by evaluating the orientational potential energy of the nuclear

magnetic
t

To

moment

in the atomic magnetic field,

simplify the notation, in this section

associated

quantum numbers.

we

shall not

which

is

proportional

put primes on J or on the

HYPERFINE STRUCTURE

Sec. II]

to the expectation value of I

J,

451

and equation (13-42)

is

obtained by

evaluating the orientational potential energy of the electron magnetic

moment

in the atomic magnetic field,

tion value of

which

is

proportional to the expecta-

L.

The term Ax j assumes different values according to the value of the


quantum number/. The possible values off are determined by an equation
analogous to the third of equations (13-41), and there are 2/ + 1 of these
values (except that when i > j there are 2/ + 1 values). Thus the energy
levels will generally be composed of 2i + 1 hyperfine structure components.
Consequently, the quantum number specifying the magnitude of the spin
of an atomic nucleus can be determined by simply measuring the hyperfine
structure of the spectrum of the atom, analyzing the spectrum into an
energy level diagram, and then counting the number of hyperfine structure
components of the energy levels. It is found that the quantum number i is
integral (0, 1 2,
.) for nuclei of atoms in which the atomic weight A
is even, with i =
when the atomic number Z is also even i is half-integral
) for nuclei of atoms in which the atomic weight A is odd.f
(I, f f
The quantity K in equation (13-63) depends on the magnetic field at
the nucleus and on the nuclear magnetic moment. If the Hartree eigen.

functions describing the motion of the atomic electrons are available,


the magnetic field can be calculated to an accuracy of about 10 percent.

measured value of Aj. s can be used to


moment.
As stated above, it is found that for all nuclei the magnitude of the dipole
magnetic moment vector is of the order of one nuclear magneton (i.e.,
Then, within

this accuracy, the

evaluate the magnitude and sign of the nuclear magnetic dipole

1), but that the vector is parallel to the spin vector in some nuclei
|^|
and antiparallel in others (i.e., g is + or ).
Measurements of hyperfine structure also give information about the
electric quadrupole moment Q of the nucleus of the atom producing the
spectrum. This quantity specifies the extent to which the charge distribution of the nucleus departs from spherical symmetry. In all cases, the
charge distribution is that of an ellipsoid with an axis of symmetry along
the direction about which the nuclear spin angular momentum vector
t

precesses

Q<

for

for
it is

Q>

the ellipsoid

is

elongated in the direction of the axis,

flattened in the direction of the axis,

charge distribution has spherical symmetry.J


will

For

and

have an orientational potential energy in an atomic

=fc

for
0,

Q=

the

the nucleus

electric field

which

t We are refering to the integer A which is closest to the atomic weight of a particular
"isotope" of an atom. With this stipulation, the atomic weight is always very close to

an
%

integer.

A precise definition of Q

is

given in section

9,

Chapter

16.

452

MULTI-ELECTRON ATOMS

[Ch. 13

not spherically symmetrical. In a typical case, the magnitude of this


energy is comparable to the magnitude of the orientational potential energy
of the nuclear magnetic dipole moment in the atomic magnetic field.
is

However, the dependence of the energy on the quantum numbers specifying


the orientation of the nucleus in the

atom

is

quadrupole and magnetic dipole interactions, so


these
is

two

effects experimentally.

different for the electric


it is

present, the hyperfine splitting of the energy levels

from the predictions of equation


dipole interaction.

atomic

possible to separate

When the electric quadrupole interaction


(13-63),

which

is

observed to depart

treats only the

By analyzing these departures, and then

magnetic

calculating the

from Hartree eigenfunctions, it has been


and the approximate magnitude of the electric
quadrupole moment of a number of nuclei. It is found that Q =
for
all nuclei with
= or i = J. This can be understood from symmetry
considerations; for instance, when
= there is nothing to define an
axis of symmetry about which an electric quadrupole moment can be
measured. For nuclei with i > 1, Q is usually measurably different from
occurs most often. The nucleus of the 71 Lu atom has one of
0, and Q >
the largest known quadrupole moments. The charge distribution of this
nucleus is elongated in the direction of the axis of symmetry, and the ratio
of the semimajor axis of the ellipsoid to its semiminor axis is 1.20 0.05.
Techniques for measuring the magnetic dipole moments of nuclei from
an analysis of hyperfine structure are limited in their accuracy because
they depend upon calculating the strength of the atomic magnetic field.
In the years following 1933, Rabi and others developed several different
techniques which are extremely accurate. All of them involve measuring
the orientational potential energy of the dipole in an external magnetic
field of known strength. For atoms in which the total magnetic dipole
moment produced by the atomic electrons is zero (an atom in a 1 S state),
it is possible to measure the nuclear magnetic dipole moment by sending a
beam of atoms through an apparatus similar to that used by Stern and
Gerlach. In other cases the energy of a nuclear magnetic moment in an
external magnetic field can be investigated by using a field which is strong
enough, compared to the atomic magnetic field, to destroy the coupling of
I and J. In the most accurate of these techniques the magnetic moment is
measured, not in terms of the deflection of an atomic beam, but in terms of
the energy of the radiofrequency quanta which produce transitions between
two energy levels that are split by hyperfine splitting in an external magnetic field of known strength. These measurements, and the quantum
mechanics upon which they are based, form a very interesting subject which
we reluctantly pass over in order to continue the discussion of atomic structure and spectra.
electric field at the nucleus

possible to evaluate the sign

('

/'

TRANSITION RATES

Sec. 12]

12.

AND

SELECTION RULES

453

Transition Rates and Selection Rules

Consider an isolated atom in an excited state of energy E v The atom


make a spontaneous transition from that state to a state of
lower energy Ek emitting a quantum of electromagnetic radiation. We
would like to calculate a quantity called the transition rate for spontaneous
will eventually

which is the probability per unit time that the atom will make
such a transition. This calculation will enable us to understand the selection
rules governing transitions between any two energy states, as well as the
lifetimes of these states and the widths of the spectral lines emitted in such
emission,

transitions.

The theory

how

that has been developed so far in this

to treat spontaneous emission, but

it

does

related process of absorption. In this process,

magnetic radiation makes


state of higher energy E

book does not

tell

us

how

tell

us

to treat the

an atom acted on by

a transition from the state of energy

electro-

Ek

to the

absorbing a quatum. The electromagnetic


radiation acts as a small perturbation on the atom, and time dependent
perturbation theory can be used to calculate the probability per unit time
t

that this perturbation will produce such a transition.

per unit time

is

We

the transition rate for absorption.

this transition rate

and then

establish a relation

This probability

shall first calculate

between

and the

it

transi-

tion rate for spontaneous emission.

An atom
between
its

its

in a field of electromagnetic radiation experiences interactions

magnetic moments and the magnetic

electric charges

and the

field,

and

The magnetic

electric field.

also between

interaction

is

compared to the electric interaction, and we


shall consider here only the latter. Although the electric field is described
by a vector E, it is possible to treat separately the interaction produced by
each component of this vector. Each component, such as Ex is an oscillatory function of time and also of position. However, the wavelengths
which are characteristic of the absorption or emission optical spectrum of
an atom (A
10 3 A) are so large, compared to the dimensions of the atom
(r
1 A), that at any instant Ex is essentially constant over any region of
space occupied by the atom. Consequently, we can ignore the dependence
on position and describe Ex at a given atom by the function
usually completely negligible

Ex =

2Ex (v) cos

where the amplitude 2Ex\v) depends,


is

(13-64)

2-nvt

on

in general,

the frequency

v.

It

convenient to write this as|

Ex =

Ex(v)(e2 " in

- 2 " in

(13-64')

t The procedure for proving the equality of (13-64) and (13-64')


footnote on page 172.

is

indicated in the

454

MULTI-ELECTRON ATOMS

The

[Ch. 13

of the atom have a potential energy in the electric


Let us define the zero of potential energy and the origin of the x
axis to be at the center of the atom, and calculate the total potential
energy of the atom due to the x component of the electric field by summing
field

electric charges

Ex

the potential energy of each charge.


vx

This gives

= -Ex 2

-ex,

(13-65)

The summation

is

taken over each electron of charge

x but the charged nucleus does not contribute as


Evaluating Ex we have the perturbing potential

its

e and

coordinate

x coordinate

is 0.

vx

which

We
is

is

an

(e3 <

e-**)

^ ex,

(13-66)

oscillatory function of time.

on an atom, which at t =
by solving equation (9-32), which is

find the effect of this perturbation

in the

quantum

state

of energy

^
at

with the

initial

Ek

a
=-{l
n

n (t)e-

i(EMI *v Xmn

conditions for a n (t) at


a n (0)

n^k

0,

1,

The quantity
is the matrix element of
the states n and m. That is,
t?,.

v Xmn

where the integration


element dr.

The

(13-67)

the perturbation v x between

=jwtv x f n di

taken over

is

(13-68)

all

the coordinates of the

volume

quantities a n {t) are the coefficients in the expansion

^'{x,i)

= ^a n{t)W{x,t)
n

of the perturbed wave function in terms of the unperturbed wave functions,


and the combination a*(t) a n (t) is equal to the probability of finding the

atom in

the

quantum

state

of energy

En at time

t.

Now,

if

we

require that

the perturbation be small enough that

a n {t)

we may
for n

neglect
k.

all

< 1,

^ k,

1,

>

(13-69)

k,

terms on the right side of equation (13-67) except

This gives

da m(t)

~- a (/)e-B
t

dt

*- B

->"i;,

Sec. 12]

TRANSITION RATES

By

m=

setting

I,

we

AND

SELECTION RULES

455

obtain the differential equation for the coefficient

which we are particularly interested. In this equation we may make


the approximation that a k {t) = 1. Then it reads

(t),

in

daM ^
Upon

evaluating v x

*B&1 ~

daM ^ _

_l

-nE*-Ei)t/h

dt

and doing a

rearranging, this

little

L* ex jfk dT [ e--B* + w +

hJi

dt

becomes

*aBi--B*-*W]0( v

or
ry(E,-.Et + Av)t/

_j_

HE,-Ek -hv)tlH-\0v

d?

where
/*.

=
J

V*

Integrating the differential equation

conditions (13-68),
*('')

2 ex % dT

(13-70)

>

from

to

t',

with the

initial

we obtain
e

/"*

i(E,-Et +ht)t'Jll

- +
ft

hv

_ ai(Ei-E -hv)t'/h'
E,-Ek - hv

(13-71)

Let us consider the behavior of this equation. Since the perturbation is


small, the quantity px E%(v) will be small. As the absolute values of the
numerator of each of the two terms in the bracket are always in the range
to

2,

the entire right side

is

small except for frequencies for which the

denominator of one of the terms becomes small. Thus

a,(t') is

appreciable

when

only

,-^11^0
Since

E > Ek
t

this

can be

satisfied

only for the

radiation satisfies approximately the relation


precisely Einstein's relation

quantum absorbed

From
(13-71)

may

E Ek =
h

hv.

This

is

between the frequency and the energy of the

in the transition.

these arguments

be neglected.

we

see that the first term in the bracket of

Doing

this,

then multiplying the remaining

complex conjugate, and finally using the relation discussed


the footnote on page 172, we find

equation into
in

sign.

Thus the proba-

of finding the atom in the state of energy E t which is at {t') a (t'),


appreciable only when the frequency of the perturbing electromagnetic

bility
is

(13-72)

its

~(E t
a*(t) ai (t)

-E - hv)
k

2h

= 4n*u x
(,

-E - hvf
k

E x(vf

(13-73)

4 56

MULTI-ELECTRON ATOMS

[Ch. 13

where we have dropped the primes to simplify the notation. The reader
should compare this with equation (9-41), which expresses the result of
a perturbation that is "switched on" at / =
and remains constant
afterwards. Both equations have a t 2 dependence for small t. In equation
(9-41) this goes over to a
intuitively,

(13-73).

when

dependence, which

the equation

To correspond

is

is what would be expected


The same is true for equation

integrated.

with physical

integrated over v because an

atom

reality, this equation must be


never subjected to electromagnetic

is

radiation of perfectly sharply defined frequency, but instead to electro-

magnetic radiation which covers a band of frequencies. The integration


can be taken from + oo to oo, since the integrand is small except for a
restricted range of v. The calculation is the same as that preceding
equation (9-44), and the results are

1 *
P^-Ti^^M^ft
h

(13-74)

PXa is the probability that an atom initially in the state of


and subjected to the x component of the electric field of a
continuous spectrum of electromagnetic radiation, will be found in the
state of energy E v The quantity E%v lk) 2 is the square of the value of E%v)
The quantity
energy

Ek

for a frequency satisfying the relation

-E

hv lk

The contribution of the x component of the


rate for absorption

is

(13-75)
electric field to the transition

thus

dP

Rx *

^~ir~J aO-^w
1

(b-76)

Similar expressions are obtained for the contributions from the y and z

components. The complete transition rate for absorption

* -

J*

tej*. a

E>J + < ^M*i? +


k

Let us assume that the radiation

is

It is

^ E%v

lk

)\nl ikfi Xa

+ fi*

all

lkl Vik

equal,

K{v lkf]

p,

the amplitudes of

(13-78)

the time average of

volume of the electromagnetic

Taking the time average of both sides of equation


P

= -E
An

(13-77)

and

+ ^.J

convenient to express this result in terms of

the energy content per unit

lk

Then

unpolarized.

the three components of the electric field are

* ^

is

(2-4),

we

radiation.

obtain
(13-79)

where

AND

TRANSITION RATES

Sec. 12]

~E

is

457

SELECTION RULES

the time average of the electric field, and where

deleted the subscript for transverse.

components of the

electric field are

assumed

same is true
That is,

to be all equal, the

of the time averages of the squares of these components.

pi

p2

we have

Since the amplitudes of the three

p2

Then, since

H + E\ + El

2 =
we have

By squaring both

(13-80)

3E]

sides of equation (13-64),

obtain a relation between | and E%(v lk )

El

and averaging over

/,

we

2
:

(13-81)

2Ex (v lkf

Then, from equations (13-78) through (13-81), we obtain

- ft>n)[M2j*x a + <* Pv a + rfjtij


Itt

Ru
where we have

(13-82)

explicitly indicated that the

at the frequency v lk

energy density

is

to be evaluated

This equation gives the desired expression for the

transition rate for absorption.

We

have seen that there is an appreciable probability of a transition


from the state of energy Ek to the state of higher energy E lf with the
absorption of a quantum, only when equation (13-72) is satisfied with
the sign. However, the equation can also be satisfied with the + sign
for a transition from the state of energy E to the state of lower energy
Ek if a quantum is emitted. This phenomenon, in which the perturbation produced by an electromagnetic field induces a transition where
x

a quantum is emitted, is called induced emission.


for induced emission R kl can be evaluated by starting
and interchanging k and / everywhere except in the
and in the symbol v lk In this case the first term in the
.

(13-71)

is

the one which contributes;

The
at

transition rate

equation (13-69)

condition

E > Ek
l

bracket of equation

but otherwise everything

is

the

same and we immediately obtain


2tt

R ki-Jp #")l>i!> + ft*A, +


It is

easy to see from equation (13-70) that

ftj>.

(13-83)

458

MULTI-ELECTRON ATOMS

So, comparing (13-83)

and (13-82), we

Rm

[Ch. 13

see that

= R

(13-84)

The

transition rates for absorption and induced emission are equal. This
an example of the principle of microscopic reversibility, which is very
important in the theory of statistical mechanics.
is

The process of induced emission must not be confused with the process
of spontaneous emission. According to equation (13-83), the transition
rate for induced emission

is proportional to the energy density p~ of the


electromagnetic radiation applied to the atom. If no electromagnetic
radiation is applied, there will be no induced emission. However, there

be spontaneous emission, whether or not electromagnetic radiation is


applied to the atom, because the transition rate for this process does not
involve p. The transition rate for spontaneous emission Skl is an inherent
will

atom and

characteristic of the

which the atom


evaluation of

is

Skl

is

not influenced by the environment in


now to our original purpose the

placed. Let us return

This can be done in two quite different ways.

ment

is

The more rigorous

based on the theory of quantum electrodynamics.^

some years before


chanics was developed, which leads to the same results and is
the other treatment, given by Einstein

The

idea

treat-

We present here
quantum me-

much simpler.

to establish a relation between the transition rates for absorp-

is

induced emission, and spontaneous emission, from considerations of


thermodynamic equilibrium. Imagine a cavity containing a large number
of atoms which are emitting and absorbing electromagnetic radiation.
tion,

When
1.

this

system

is

in equilibrium at temperature T:

The electromagnetic

radiation must have the black

body

spectral

distribution characteristic of that temperature.


2.

The

total

number of atoms

in each energy state

must be distributed

according to the Boltzmann probability distribution for that temperature.


3. The total number of atoms in each energy
of the second condition, remain constant.

state must, as

a consequence

Condition 3 requires that the total number of transitions per unit time
the state of lower energy Ek to the state of higher energy E equal

from

the total

number of

number of

transitions in the opposite direction.

transitions per unit time

higher energy state

k is

the total

t Cf. footnote

from the lower energy

the total

state to the

is

*"*

where

Now

= Nk R

number of atoms

on page 462.

lh

in the lower energy state,

(13-85)

and where

AND

TRANSITION RATES

Sec. 12]

the transition rate for absorption


that one of these

atoms

will

According to (13-82),

lk

R lk

make a

SELECTION RULES

459

gives the probability per unit time

transition to the higher energy state.

can be written

R lk = B lk p(v lk)

(13-86)

This defines the proportionality constant B lk between the transition rate


for absorption and the energy density, which is called Einstein's coefficient

of absorption.^ Using

this quantity,

equation (13-85) becomes

^b = N B
k

The

total

number of transitions per

to the lower energy state

lk p{v lk)

(13-87)

from the higher energy

unit time

^m = #,[* + S

(13-88)

kl ]

where

is

the total

number of atoms

transition rate for induced emission,

From

spontaneous emission.

(13-83)

Einstein's coefficient

Skl
which defines
coefficients

Einstein's

B lk Bkl
,

and

Rkl is the

in the higher energy state,

and Skl

we can

Rkl = Bkl p(v


which defines

state

is

is

the transition rate for

write

(13-89)

lk )

of induced emission

=A

Bkl We
.

also write

(13-90)

kl

The
of spontaneous emission A kl
depend only on the inherent characteristics

coefficient

A kl

all

of the atoms, and equation (13-90) emphasizes the fact that we assume
the transition rate for spontaneous emission to be independent of the
intensity of the electromagnetic radiation applied to the atoms. From
the last

two equations, we may write equation

(1

^m = WmpCv*) + A
Then we

satisfy condition 3

3-88) as

by equating Ykl and jV

N* _
N
t

B kl p(vik)

+A

(13-91)

kl ]

11c ,

and obtain

kl

(13-92)

B lk p(v lk )

The quantity Nk jN can be evaluated by applying condition 2. The


Boltzmann distribution (2-23) says that at temperature T the probability
that an atom will have energy Ek is proportional to e - E* IK r and the
probability that an atom will have energy E is proportional to e - E IKT
l

'

'

At

the time he developed this treatment, Einstein wrote equation (13-86) as a


know equation (13-82).

postulate because he did not

MULTI-ELECTRON ATOMS

460

K is Boltzmann's constant.

where

of atoms with these two energies

According

to

e-

(E,

(13-93)

'

equation (13-75), this

this

-Ek )IKT

E /KT

Hk
Combining

Consequently, the ratio of the number


is

-E k /KT
:t

ly k

[Ch. 13

is

hv lk /KT

(13-94)

with equation (13-92) gives

fJivit/KT

Bklf>( v lk)

+ Am

Bi k p{v lk)

Solving for

p~(v lk),

we obtain
pXv lk )

Finally,

we apply condition

tral distribution of

f^

(13-95)

by equating

p~(v lk)

to the black

body

spec-

energy density (2-31). In terms of the present notation

this is

Then a comparison of equations (13-95) and (13-96) shows that


Einstein coefficients must be related to each other as follows

B a = Bkl
and

An =

the three

(13-97)

^B

(13-98)

kl

Equation (13-97) shows that this treatment is consistent with equation


(13-84). Equation (13-98) provides the desired connection between the
transition rates for absorption and spontaneous emission. Using (13-83),
(13-89), and (13-90), we obtain at last the transition rate for spontaneous
emission

Su
The

~ Ifo?

^^ + ^A

zJ

(13 " 99)

is seen to be proportional
cube of the energy) of the emitted quanta.
with the transition rates for induced emission and absorption,

transition rate for spontaneous emission

to the cube of the frequency (or

In

common

it is

also proportional to the

fi x

and

their

sum

of the products of quantities such as

complex conjugates. The quantity

[x

Xui

is

called the matrix

AND

TRANSITION RATES

Sec. 12]

SELECTION RULES

461

element of the x component of the electric dipole moment of the atom.


The quantities fi and /e are given corresponding names. This nomenclature

is

used because the term

2 ex

that occurs in the integrand of

is the x component of the electric dipole


fi x
The matrix element of J exi * s its average value

equation (13-70) defining

moment

of the atom.

calculated with a weighting factor ip*ip t

the probability density y)*y> l of the initial


bility density y>*y k

of the

final

quantum

This weighting factor

quantum
state,

state,

is

neither

nor the proba-

but instead some sort of

"mixed probability density" which involves the eigenfunctions of both the


initial and the final states. The probability of an atom radiating a quantum
depends, then, not only on what the atom does in the initial quantum state
before it radiates, but also on what it is going to do in the final quantum
state after it radiates. The reason is that when the atom radiates it is
actually not in either the initial state or the final state, but in some linear
combination of the two.
This point of view can be used to give a physical picture of what happens
in a radiation process. Consider an induced transition of an atom from
the initial state of energy E to the final state of lower energy Ek accompanied by the emission of a quantum. According to (7-45), the wave
function describing the atom can always be written
,

Y = 2n a n e~

iEnt/n

Before the application of the electromagnetic

atom
non-zero, and

transition, the

is

is

in the eigenstate of energy

Y=
The corresponding

a e-

field

which induces the


Only the coefficient

iB ' t "'y>
l

probability density

Y*Y =

(13-100)

Wn

is

ata lV Vl

This quantity specifies the charge distribution of the atom.

(13-101)
It is

constant

Consequently the electric dipole moment (as well as every other


feature of the charge distribution) is constant in time. According to
classical physics, the atom can radiate only if its electric dipole moment
in time.

some other feature of the charge distribution) changes in time. Thus


we can understand, in a sense even classically, why the atom cannot
radiate as long as it is in an eigenstate. Now, when the electromagnetic
(or

produces a perturbation which mixes contributions from


wave function for the eigenstate of energy Ek into the wave function
the atom. The coefficient ak increases with time from its initial value

field is applied, it

the
for

MULTI-ELECTRON ATOMS

462

of zero, and the coefficient a decreases.

Consider the wave function for


It has the form

the

atom when both ak and a

The corresponding

W*W =

a*a

The

last

probability density

^p*rp l

are non-zero.

a e -uw fy

Y=
+

e-

iE " tlh

xp

(13-102)

is

*
fl

aiC

-(,-^),/

Etc

(13-104)

reduced to zero. Then the probability density

Y*F =

(13 _ 103)

Vi

with the frequency

oscillate

Ei

is

aic

a?a k y>tfk

two terms of F* F

Eventually a l

,V*rwy ft +
,

[Ch. 13

is

a*
k a k y>: Wk

With such a probability density the atom cannot

But it is apparent
by the wave function
(13-102). In this state the probability density (13-103), and consequently
the charge distribution, have components which oscillate with frequency

that

it

can radiate when

(13-104) that

is

it is still

radiate.

in the state described

exactly equal to the frequency of the

magnetic radiation which

is

quantum of electro-

emitted during the transition.

These arguments provide a good physical picture of the process of


induced emission and absorption. It would appear that they do not apply
to the process of spontaneous emission because in this process there is
apparently no perturbing electromagnetic field to produce the mixing of

wave functions that leads to an oscillating probability density. Howquantum electrodynamics removes this difficulty. In
this theory both the electromagnetic field and the atom are quantized.f
Then it is found that there are interactions between these two quantized
systems, even when the electromagnetic field is in the state of lowest
energy, because some electromagnetic radiation must be present in this

the

ever, the theory of

has a zero point energy like any other


quantized system executing simple harmonic oscillations.

state as the electromagnetic field

The electromagnetic

field

was quantized

in the very simple theory of

Planck and

comprised of quanta of energy E = hv. But a more


complete theory is needed because it is necessary to know how an electromagnetic field
in the quantum limit (containing few quanta) interacts with an external system of
Einstein,

which

charges, just as

states that

it is

it is

necessary to

know how an

electromagnetic

field in

the classical

quanta) interacts with an external system. The classical limit


is described by Maxwell's theory; the quantum limit is described by the theory of
quantum electrodynamics. In the latter theory, Schroedinger's quantum mechanics is
applied to a system of quanta, but the quantization of this system is considerably more
limit (containing

many

involved than the quantization of a system of particles since the number of quanta
in a system can change when it interacts with an external system.

AND

TRANSITION RATES

Sec. 12]

If the eigenfunctions for

elements between

all

SELECTION RULES

an atom are known, the

electric dipole

pairs of excited states can be obtained

Then

integrals such as (13-70).

463

matrix

by evaluating

the transition rates for spontaneous

emission between these states can be calculated from equation (13-99).


These transition rates immediately lead to theoretical predictions of the
relative intensities of the spectral lines emitted in transitions

between the
found that the agreement between these predictions
and the experimental data is quite good. The transition rate for spontaneous emission varies widely from one case to the next. For the transition
of the X H atom from its first excited state to its ground state, this transition
rate has the value Skl
10 8 sec -1 This is typical of the orders of magnitude encountered, except that the transition rates between certain pairs of
states are essentially zero. These are the transitions for which the spectral
lines are observed to be absent or extremely weak. The transition rates
are predicted to be zero in these cases because the integrals involved in
various states.

It is

the electric dipole matrix elements vanish identically.

Thus

the selection

rules are a set

of conditions on the quantum numbers of the eigenfunctions


of the initial andfinal states such that the electric dipole matrix elements are
zero when calculated with a pair of eigenfunctions whose quantum numbers
violate these conditions.

As a

we shall carry out part of the calculation of the


matrix elements for an atom with one optically active

simple example,

electric dipole

electron.

In this case the core of

and

filled subshells

does not change in the

necessary only to evaluate the matrix elements of e


times the three rectangular coordinates of the optically active electron.

transition,

it is

Furthermore, in evaluating these matrix elements

we need

only to write

the eigenfunctions for the optically active electron. All the other factors in

atom immediately integrate to unity


because of the normalization conditions. The rectangular coordinates are
the complete eigenfunction of the

related to the spherical coordinates, in terms of

which the eigenfunctions

are written, by the equations

r sin 6 cos
y = r sin 6 sin
z = r cos 6
x

<f>

(13-105)

<f>

Using these coordinates, we have

-"*'',;,.,

) J J

Vti-mf

sin 6 cos

nlm

f sin 6 dr dd

d<j>

e f

Jo

R:. v R nl r 3 dr f*^-,,,, sin 2 6 dd


Jo

P $: <D mi cos

Jo

<f>

d<f>

MULTI-ELECTRON ATOMS

464

[Ch. 13

Consider the third integral,


2

r^mi

-..

$d

cos

<f>

Jo

f 'e-".'V"*cos

d<f>

<>0

C 2n

-i(m',-mi + l)*

_|_

-i(m,-ro! -1)0-1

JJ

2 Jo

We leave it as an exercise for the reader to

show

that this integral vanishes

unless

m
If

it

does vanish, so does

evaluating
upon
o
X

u,.
'

y n j'm,,nlm(

-m

'

For *n

'

Since the

^^./^

-T

is satisfied.

Thus
m[

obtained

we have
f2ir

'- Sin0COS0de

*-.i

'0

^m,^

-m =

(13-107)

is

the complete selection rule for

-m =
l

Am,

0,

transitions that satisfy equation (13-107) are


direction.

The

spectral lines

from

The

is

(13-108)

in agreement with the second of equations (13-57),

for one or several optically active electrons.

is

transition rate 5. rm;w!mi will be non-zero only if either (13-106)

or (13-107)

This

Imj^Imj

selection rule

are orthogonal, the third integral vanishes unless

mj

The

(13-106)

fir

foo

/W..= Jo

= 1
The same

>

which

is

spectral lines

valid

from

linearly polarized in the

transitions that satisfy equation

(13-106) have components, which are linearly polarized in both the x


and y directions, that combine to form circularly polarized radiation.

These results are verified by measurements of the polarization of the


radiation emitted in the Paschen-Bach effect by the different spectral lines
corresponding to different values of A/j.
The selection rules for / can be obtained by a straightforward evaluation
of the integrals over dd. However, this is quite tedious, and part of the
can be obtained in a simpler way from considerations of the
In section 4, Chapter 8, we defined the
parity of a one dimensional eigenfunction. This is the quantity which
describes the behavior of the eigenfunction when the sign of the coordinate
is changed.
The definition can be immediately extended to three
dimensions. That is, eigenfunctions which satisfy the equation
results

parities of the eigenfunctions.

f{-x, -y,

z) =

+ip(x, y, z)

(13-109)

AND

TRANSITION RATES

Sec. 12]

SELECTION RULES

465

are said to be of even parity, and eigenfunctions which satisfy the equation

y(-x, -y, -2)


are said to be of odd parity.

not

It is

-ip(x, y, z)
difficult to

(13-110)

show

that

most

eigen-

functions which are solutions to Schroedinger equations with potentials

whose values are unchanged when the

signs of the coordinates are changed


even or odd).f
It is convenient to express equations (13-109) and (13-110) in terms
of spherical coordinates. By considering the diagrams in figure (13-16)
we see that, when the signs of the rectangular coordinates are changed,
the behavior of the spherical coordinates is
will

have a

definite parity (either

r~>r,

O-^tt-9,

<f>~+TT

(13-111)

<f>

t Write the time independent Schroedinger equation as

^op'*. y,

Change the

z )v(x , V, 2 )

Exp(x, y, z)

signs of the coordinates:

^opC

-x

ll>

z)y( x

y,

~ z) =

Ey>(x, y,

z)

Since e op contains only V(x, y, z) plus terms like d z /dx 2 ,e ,(x, y,


z) = e op (x, y, z)
0T
z) V{x, y, z). If this relation holds, the previous equation reads

when V(x, y,

^opfo
Since both

y(x, y,

y,

z)v( *, y,

z) and

- z) =

y)(x, y, z) satisfy

Ey>(x, y,

the

same

differential equation,

it

must

be that

z) =

y(x, y,
where

AT is

a constant.

Now

Ky>(x, y, z)

change the signs of the coordinates

in this

equation and

obtain

xp(x, y, z)

Kyi(x, y,

Multiplying this equation into the previous equation,

K"

z)
we

obtain,

or

K=

so

V( *. y, -2)
This proof breaks

down

if

there are

w(x,y,z);

Q.E.D.

two or more linearly independent eigenfunctions


same value of E, i.e., if there is degeneracy.

that satisfy the differential equation for the

But even in this case the eigenfunctions will have definite parities if they are approchosen linear combinations. The y>mm are examples of degenerate eigenfunctions with definite parities, as shown by equation (13-112). An example of a
degenerate eigenfunction without a definite parity is the traveling wave (15-64).
priately

MULTI-ELECTRON ATOMS

466

Now,

if

[Ch. 13

the reader will inspect the eigenfunctions for the hydrogen

(10-39), he can verify that for those

and

which are

atom

listed

<M* +

4>)

(-i) |mil <i> mi (<)

^im^ -

e)

(-i) l+M lmi (d)

Thus
Rm(r)

.,(*

~ ^mfr +

ft

VW."" ~

>"

(-l) l+lmi K-l)

MR

nl (r)

@ lmi (0) $(<)

or

This

true for all the hydrogen

is

transformation (13-1

1)

= (-V V>m mi(r,


l

ft

atom

6,

(13-112)

In fact, since the

eigenfunctions.

does not affect the coordinate

0)

r,

equation (13-112)

(*,y,z)

(-x, -y, -z)

Figure 13-16.

is

Illustrating the parity operation.

true for all eigenfunctions for spherically symmetrical potentials because

R nl (r) Q tmi (0)

^>
m i(ft where
Thus the
atom.
(c) are the same as for the hydrogen
{m (6) and O mi
determined
potentials
is
parity of all eigenfunctions for spherically symmetric
all

such

eigenfunctions

by ( 1)'; the parity

Now
dipole

can

even if I

is

be

is

written

even,

and odd

if

I is

odd.

consider the matrix element of the x component of the

moment, which

electric

is

^ m
The x component of the

;, te;

V-'i'mj

electric dipole

2 ex, Wnlmi dr
moment

is

of odd parity since

I(-^)= -lex,
y and z components. If the eigenfunctions
and ip nlm are of the same parity (both even or both odd), the
This

is

also true of the

y>

rm

entire

TRANSITION RATES

Sec. 12]

AND

SELECTION RULES

467

uy
integrands
of m_
and u,*n' V m i,nlmi will be of odd xparity.
J

''n'l'm^nlmi
n'l'mi,nlmi
,

'

'

'

it is obvious that the integrals will vanish identically


because the contribution from the point (r, 6, <f>) will be canceled by

If this is the case,

the contribution from the point


rates will

(r,

be zero. Therefore there

relative parity of the initial

and

is

d,

</>),

and the

transition

the following selection rule for the

final eigenfunctions in

an

electric dipole

transition

The
Equation (13-112) allows
the

quantum number

/,

parity

this to

must change.

(13-113)

be expressed, in terms of the change in

as

A/=

1, 3, 5,...

(13-113')

but not all, of the selection rule (13-31). The rest of it can be
by considering (13-108), which is &m = 0, 1. This selection
rule makes it possible to understand why transitions are not allowed when
A/ = 3, 5,
since in some of these transitions Am, > 1.
As
mentioned before, the selection rule (13-31) can also be obtained by a
This

is

part,

justified

straightforward evaluation of the matrix elements.

To calculate the selection rules for the quantum number/, it is necessary


know the exact form of the eigenfunctions in the presence of the spinorbit interaction. Therefore we can only assure the reader that calculations
to

with these eigenfunctions do verify the selection rule (13-30). The same
true of the selection rules (13-45) for the quantum numbers of an atom
with several optically active electrons that is an example of LS coupling.

is

But one of these, As' = 0, has such a simple interpretation that we must
mention it. The selection rule states that the coupling of the spin angular
momentum vectors cannot change in the transition. The reason is
simply that the interaction between electromagnetic radiation and spins
takes place through the interaction of the magnetic field and the magnetic
moments. These interactions are so very weak that the coupling of the
spins is normally not affected.
However, faint spectral lines that violate the selection rule for s' can
be seen in certain special circumstances (cf. the discussion at end of

Thus

this selection rule does not absolutely prohibit transitions


but only makes these transitions very unlikely. If the
transition cannot take place by the normal means of an interaction in-

section

8).

that violate

it,

volving the electric dipole

an

moments and

the electric fields, which

is

called

electric dipole interaction, there is

a very small probability that it can


take place through an interaction of the magnetic dipole moments and the
magnetic fields that are also present in electromagnetic radiation. In

fact,

there

is

a very small probability of violating, by means of such

468

MULTI-ELECTRON ATOMS

[Ch. 13

magnetic dipole interactions, any of the selection rules quoted in this


chapter. This can also be done by means of electric quadrupole interactions,

which represent violations of our assumption that the wavelength


is so long compared to the dimensions of the atom that

of the radiation

any variations of the

electric fields over the atom can be neglected. Nevergoing through electric dipole interactions are by far
the most important for atoms.

theless, transitions

Lifetimes and Line Widths

13.

Consider a system of atoms in which N(0) have been excited at time


by some process to a state of energy E Let us calculate the number
N(t) remaining in that state at some later time t, assuming that the transition rates for spontaneous emission Skl are known for all k for which
t

Ejc

<E

t,

and that these

much larger than the


because the average density p~ of

transition rates are very

Rkl

transition rates for induced emission

electromagnetic radiation

is

very small.

Then

the probability per second

atoms in the state of energy E will make a transition out of


that state into any of the states of lower energy E is the sum of the S
k
kl for
all k for which E < E
Call
this
probability
X;
then
k
that one of the

= 2s

In a time interval dt at

the total

t,

(13-H4)

number of such

transitions will be,

on

the average,

N(i)X dt
since during that interval there are approximately N(t)

and each one has a transition probability 1


the decrease in N(t) during that interval. Thus
state

dN(t)

-N(t)X

dt.

atoms

in the initial

This quantity

is

just

dt

or

-^=-Xdt

(13-115)

jv(o
Integrating

from

to

/',

we have

rd
Jo

-x!"dt

N{t)

Jo

which gives
In \_N(t')

iV(0)]

-At'

Sec. 13]

LIFETIMES

AND

LINE

WIDTHS

469

or

N(0)

So
N(t')

e~ u

'

N{0)
or
N(t)

N(0) e

-U

(13-116)

where we have dropped the primes to simplify the notation. The number
of atoms remaining in the initial state decreases exponentially from the
original value JV(0) and, at the time t = 1/A, the number remaining is
reduced by a factor e~ x The lifetime t of this state is defined as the average
time an atom spends in the state before making a transition. This is,
.

/*00

00

tN(i)dt
J*
_o

N(0)te'

xt

dt

Jo

N(0)e- Xi dt

N(t)dt
Jo

Jo

Since this ratio of integrals has exactly the same form as (2-19), we may
use the results of the calculation following that equation to write im-

mediately

t=1/A
which

(13-117)

is

IS J

(13-118)

- k

Because r generally involves a sum over several Skl the variations in


those quantities tend to average out and yield lifetimes which are of the
10" 9 sec for many states. However, there are a large number
order of t
,

of exceptions.

There is a very interesting relation between the lifetime of an excited


of an atom and the width of the spectral lines arising from transitions
involving that state. Spectral lines are not perfectly sharp; instead they
have finite widths that can be measured, under favorable circumstances,
by using equipment of the highest resolution. Usually the width is due
state

primarily to a

random Doppler shift of the frequencies of the emitted


random thermal motion of the atoms in the

quanta, resulting from the

Doppler broadening can be reduced by


using low temperature sources, and when this is done it is found that for
lines originating from states of short lifetime a measurable width remains.
This is a manifestation of the uncertainty principle relation (6-28), which
spectral source.

However,

this

MULTI-ELECTRON ATOMS

470

says that a measurement of the energy of a system that


time At must be uncertain by an amount AE, where

[Ch. 13

carried out in a

is

~h

AE At

Consider a transition from some excited state to the ground state of the
atom. A measurement of the frequency of the quantum emitted in this
transition constitutes a measurement of the energy of the atom in the
excited state. Furthermore, the measurement obviously must be carried
out within a time comparable to t, the lifetime of the excited state. Thus
the uncertainty principle

demands

that the energy of the

quantum can

have any value within the range

AE~hJT
Then

its

(13-119)

frequency can have any value within the range

Av

^^l = Jhr

and the associated

(13-119')

2-nr

spectral line will have a corresponding width.

natural broadening can be observed experimentally

if

This

the pressure in the

At high pressures the time between collisions of


is low.
atom with other atoms becomes smaller than the lifetime of
the excited state and, since it is no longer possible to carry out an undisturbed measurement for a time comparable to t, the width of the line will
spectral source

the emitting

be larger than

its

natural width.

This

is

called collision broadening.

If a spectral line is emitted in a transition


final state

of very

much

width of the spectral

longer lifetime

(e.g.,

line emitted in the transition is

lifetime of the initial state.

The

initial state to

due

entirely to the

In these circumstances the natural width of

the spectral line, expressed in energy units,

of the

from some

the ground state), the natural

is

said to be equal to the width

idea that a state has a non-zero width

if it has a
comes into quantum theory directly through the uncertainty
principle, as we have seen, and certainly is not in conflict with the statement
made in Chapter 7 that the energy of a system of bound particles, such as
an atom, can be equal only to one of a particular set of discrete (i.e., zero

initial state.

finite lifetime

width) energies.

We

could

possibility of transitions

mation,

all states

have

make

between

this

statement because

states in

infinite lifetimes or

we ignored

the

Chapter 7 and, in that approxizero widths.

The idea of a non-zero width for a finite lifetime even comes into classical
theory. Consider any resonant system, such as an electrical circuit containing inductance and capacitance. If it contains absolutely no resistance,
the theory of electrical circuits predicts that the circuit has a perfectly

sharp resonant frequency v and that,


frequency,

it

will accept

if driven by some source of variable


power only when the source frequency is exactly

LIFETIMES

Sec. 13]

equal to v

AND

But, of course,

LINE

WIDTHS

all circuits

471

have some resistance. In

the theory predicts that the resonant frequency


that the

this case,

not so well defined, and

is

power P accepted from the source depends on the source frequency

v according to the relation

^~

P(v) c
(v

The quantity

P(v),

-vf+

(13-120)

(Av/2) 2

which is plotted in figure (13-17), is the familiar


full width at half maximum Av is a simple function of

resonance curve. The

Figure 13-17.

resonance curve.

the resistance that increases with increasing resistance.

accepts a certain

turned

off,

amount of energy from

the energy content

Now,

if the circuit

and the source

is

then

of the circuit will decrease in time by

The theory of

dissipation in the resistance.


this decrease follows the

the source

electrical circuits

shows that

law
E(t)

= E(0)e-

tlT

(13-121)

where
T

(13-121')
2-rr

Of

course,

all

Av

these equations for electrical circuits have been verified

experimentally.

Now

consider an

atom

in

its

electromagnetic radiation which

ground
is

state,

exposed to a source of

emitting quanta of

The atom can absorb a quantum only if its energy is equal


one of

its

frequencies.

excited states but, since these states have non-zero widths,

quantum

possible for a
if its

all

to the energy of

energy

is

it is

to be absorbed in a transition to a particular state

within a range equal to the width of the state.

In fact,

quantum theory shows

that the transition rate

particular state depends

on the energy hv of the quantum according

for absorption to a

to

the relation

R(hv) oc
'

(hv-hv f + (hAvl2f

(13-122)'
K


MULTI-ELECTRON ATOMS

472

[Ch. 13

is the mean energy of the state and h Av is its full width at half
maximum. Furthermore, the theory shows that the width h Av is related

where hv

to the lifetime r of the state

by the equation
r

The width h Av of a
(13-122)

oc

is

the energy of the state

Eo f

and V
r

is

by the symbol T, and

state is usually designated

(e

(13-123)'

usually written in terms of the energy

is

R(E)
where

2nh Av

(r/2)

is its

of the quantum as
(13-124);
l

width.

Then
(13-125)

the relation between the lifetime t of a state

the transition rate for spontaneous emission

is

and

its

width V.

Since

proportional to the transition

former also has the functional dependence of


Therefore the natural width of a spectral line emitted in a

rate for absorption, the

(13-124).
transition

width

is

Note

from an

initial state

of width

to a final state of much

narrower

r, as has already been mentioned.


that equation (13-125)

is

in agreement with the limit set

uncertainty principle relation (6-28).

Also note that the

by the

results (13-122)

and (13-123) of the quantum theory for resonant absorption of quanta


by a state, which subsequently radiates, are essentially identical with the
results (13-120) and (13-121') of the classical theory for the resonant
absorption of energy by a system, which subsequently dissipates. The
natural spectral line shape predicted by (13-124) has been verified experimentally. Equation (13-125) cannot be tested by experiments measuring
both t and T because h is so very small that only one or the other is
measurable. Nevertheless, it is assumed to be valid and is often used to
predict the value of V from the measured value of r, or vice versa.

BIBLIOGRAPHY
Herzberg, G., Atomic Spectra and Atomic Structure, Dover Publications,

New

York, 1944.
Leighton, R. B., Principles of Modern Physics, McGraw-Hill Book Co., New
York, 1959.
Richtmyer, F. K., E. H. Kennard, and T. Lauritsen, Introduction to Modern
Physics, McGraw-Hill Book Co., New York, 1955.
White, H. E., An Introduction to Atomic Spectra, McGraw-Hill Book Co.,
New York, 1934.

Ch.

EXERCISES

13]

473

EXERCISES
Evaluate V(r) and p(r) in the Thomas-Fermi theory for the

and compare with

Use the approximation (13-22), the Z


atom equation (10-44), to estimate the
of the A atom. Compare with figure (13-3).
2.

electron

3.

A atom (Z =

18),

figure (13-3).

for the

atom, and the one-

radii of the n

1, 2,

and

3 shells

Verify that the vector diagrams of figure (13-1 1) are equivalent to equations
Also construct a similar set of vector diagrams for / x = 2, s 1 = |;

(13-40).
'2

==

A.

-^

^2

==

Z'

Carry through a calculation as in table (13-8) for the configuration np z

and thereby
5.

verify the results for that configuration listed in table (13-9).


is an example of LS coupling, and an nd n'd
atom in which both the ordering of the energy levels
and the relative strengths of the s' and /' dependence of the

Consider an atom which

configuration of that

according to

s', l',j',

Draw a schematic energy level diagram for this configuration,


using the same (exaggerated) scale for the fine structure splitting of all the levels
energy, are normal.

within a given multiplet. Label each level with the spectroscopic notation.
6.

Indicate which levels of the energy level diagram of exercise 5

absent
7.

if n'

On

the energy level diagram of exercise

exaggerated) scale the

Zeeman

effect splitting

6,

Count the
quantum states

total

number of components,

draw

to the

same

of each level into

components under the influence of a weak magnetic


8.

would be

n.

i.e.,

(highly

its

various

number of

different

field.

the total

in the configuration, obtained in exercise 7.

Show

that this

equals the degeneracy of the configuration before the application of the perturbawhich is the product of degeneracy factors 2(2/ + 1) for each of the two

tion,

optically active electrons of the configuration.


9. In an atom which is an example of LS coupling, the relative separations
between adjacent levels of a particular multiplet are 4:3:2:1. Use the procedure
outlined at the end of section 10 to assign the quantum numbers s', l',j' for
these levels. Repeat for multiplets with relative separations 7:5:3 and 9:7:5:3.
10.

Verify equation (13-74).

11.

Verify equation (13-112).

By a straightforward evaluation of the electric dipole matrix elements for


the eigenfunctions (10-39), show that the selection rule (13-31) is valid for
n = 2 -* n = 1 transitions in the
atom.
12.

Evaluate a few of the electric dipole transition matrix elements for a one
dimensional charged simple harmonic oscillator by using the eigenfunctions
13.

Show that their values agree with the selection rule An


pare with the discussion of section 8, Chapter 5.
(8-118).

1. Com-

14. Consider a gas of H molecules at room temperature and pressure, in


2
which the H atoms are making spontaneous transitions from the first excited

MULTI-ELECTRON ATOMS

474
state to the
is

10

sec

ground
-1

Use the

state.

Use the

[Ch. 13

fact that the transition rate for that transition

to estimate the natural broadening of the resulting spectral line.

relation of elementary kinetic theory connecting the average distance


between collisions to the density of the gas and the projected geometrical cross
sectional area of the molecules, the perfect gas law, and the law of equipartition
of energy to estimate the collision broadening. Use the law of equipartition
of energy and the classical Doppler shift equation to estimate the Doppler

broadening.

CHAPTER

14

X-rays

I.

The Discovery

of X-rays

Except for a brief description of the Compton effect, and a few other
we have postponed the discussion of X-rays until the present
chapter because it is particularly convenient to treat X-ray spectra after
remarks,

Although this ordering may have given the reader


a distorted impression of the historical importance of X-rays, this impression will be corrected shortly as we describe the crucial role played by
treating optical spectra.

X-rays in the development of modern physics.


X-rays were discovered in 1895 by Roentgen while studying the phenomena of gaseous discharge. Using a cathode ray tube (cf. Chapter 3)
with a high voltage of several tens of kilovolts, he noticed that salts of

barium would fluoresce when brought near the tube, although nothing
visible was emitted by the tube. This effect persisted when the tube was
wrapped with a layer of black cardboard. Roentgen soon established
that the agency responsible for the fluorescence originated at the point at

which the stream of energetic electrons struck the glass wall of the tube.
Because of its unknown nature, he gave this agency the name X-rays.
He found that X-rays could manifest themselves by darkening wrapped
photographic plates, discharging charged electroscopes, as well as by
causing fluorescence in a

number of

different substances.

He

also

found
low

that X-rays can penetrate considerable thicknesses of materials of

atomic number, whereas substances of high atomic number are relatively


opaque.f Roentgen took the first steps in identifying the nature of X-rays
t This property
tissue,

is the basis of the technique of X-ray photography of bones in living


a technique which was used by surgeons only a few months after Roentgen's

discovery.

475

X-RAYS

476

by using a system of

slits

to

show that

(1) they travel in straight lines,

that (2) they are uncharged, because they are not deflected

magnetic

[Ch. 14

by

and

electric

or

fields.

The discovery of X-rays aroused the


performed a

and many
Haga and Wind

interest of all physicists,

joined in the investigation of their properties.

In 1899

diffraction experiment with X-rays

which showed
from the size of the
diffraction pattern, their wavelength could be estimated to be X -~ 10~ 8 cm.
In 1906 Barkla proved that (4) the waves are transverse by showing that
they can be polarized by scattering from many materials.
There is, of course, no longer anything unknown about the nature of
X-rays. They are electromagnetic radiation of exactly the same nature as
visible light, except that their wavelength is several orders of magnitude
shorter. This conclusion follows from comparing properties 1 through 4
with the similar properties of visible light, but it was actually postulated
by Thomson several years before all these properties were known. Thomsingle

slit

that (3) X-rays are a wave motion phenomenon, and,

son argued that X-rays are electromagnetic radiation because such radiation would be expected to be emitted from the point at which the electrons
strike the wall of a

cathode ray tube.

At

this point, the electrons suffer

very violent accelerations in coming to a stop and, according to classical

electromagnetic theory,

magnetic radiations.
tion of X-rays

In

common

is

all

accelerated charged particles emit electro-

We shall see later that this explanation of the produc-

at least partially correct.

with

other

electromagnetic radiations,

X-rays exhibit

The reader will recall


that the Compton effect (cf. Chapter 3), which is one of the most convincing
demonstrations of the existence of quanta, was originally observed with
particle-like aspects as well as wave-like aspects.

electromagnetic radiation in the X-ray region of wavelengths.

2.

Measurement

of X-ray Spectra

The problem of measuring the wavelength spectrum of the radiation


emitted from an X-ray tube seemed, at

first, a very difficult one indeed.


Roentgen had shown that the index of refraction of all substances for
X-rays is very nearly unity. Consequently, a prism spectrometer is not

a useful instrument for X-rays.

An

alternative

is

to use a diffraction

However, the wavelength of X-rays is so small compared to the


line spacing of the best ruled gratings which can be produced that these
devices would also appear to be useless. The problem was solved in 1912
when Laue suggested that a crystal might be used as a diffraction grating
for X-rays because the atoms of a crystal are arranged in a regular lattice,
grating.

MEASUREMENT OF X-RAY SPECTRA

Sec. 2]

with a separation of the lattice planes that

wavelengths of X-rays.

is

This suggestion was

477

quite comparable to the

put into practice by


and Knipping, and the technique was perfected by the Braggs.
Consider a crystal of NaCl. The macroscopic characteristics of such
a crystal strongly suggest that on a microscopic scale the atoms are
arranged in a regular cubic lattice. A two dimensional section of this
lattice is indicated in figure (14-1). The figure also attempts to indicate
first

Friedrich

the density distribution of the electrons within each atom.

study of multi-electron atoms,

we know

that there

is

From our

a region of very high

##*#

## + #

##**#
A two

Figure 14-1.

dimensional section of a sodium chloride crystal.

electron density quite near the center of the atom,


in this region are very tightly

bound

and

that the electrons

Now

consider a beam
of X-rays incident upon the crystal. The beam penetrates into the interior
of the crystal; a fraction of it goes all the way through; but another
fraction is scattered by the atoms of the crystal. There are several mechanisms which produce this scattering. All of these mechanisms will be
discussed later, but at present
the X-rays are scattered with

we

to the atom.

are interested only in the one in which

no change

in wavelength and no change in


phase from the tightly bound electrons near the center of the atom. Each
atom scatters X-rays by this mechanism into all angles, although the

scattered intensity

is largest in the direction of the incident beam.


In the region exterior to any atom, the intensity pattern of the X-rays

which
center.

it

scatters

For

is

the

same as

this reason,

it is

if all

the scattering occurs exactly at

its

possible to analyze the scattering by the

model crystal consisting of a set of point scattering


same relative locations as the centers of the atoms of the

crystal in terms of a

centers with the


real crystal.

spaced

These scattering centers consequently lie on a

lattice planes.

set

of uniformly

X-RAYS

478

[Ch. 14

Imagine that a beam of monochromatic X-rays is incident on the model


an angle 6 measured from its surface. Part of the beam is
scattered from the scattering centers of each lattice plane. Let us consider
one such lattice plane. In the language of wave motion, the incident beam
of X-rays can be described as a succession of wave fronts (surfaces of

crystal at

constant phase) moving over the scattering centers of the

lattice plane.

Each time a wave front crosses a scattering center the scattering center
re-emits a spherical wave, with no change in phase, that moves out in all
directions. Figure (14-2) shows several incident wave fronts and one set

Figure 14-2.

beam from

Huygens' wavelet construction for the

partial reflection of an

X-ray

a lattice plane of a crystal.

Huygens' wavelet construction that the re-emitted waves combine to produce a set of reflected wave
fronts, several of which are shown in the figure. These wave fronts constitute a beam reflected from the lattice plane. As can be seen from the
of re-emitted waves.

It is

apparent from

figure^the angle of reflection

is

equal to the angle of incidence.

This result allows us to deal with


sisting

of the

ing surfaces.

this

an even simpler model

crystal con-

of lattice planes which are assumed to be partially reflectFigure (14-3) illustrates the incident beam passing through

set

the crystal and being partially reflected from each lattice plane.
figure the incident

and

reflected

beams

In this

Emerging
beams. These beams inter-

are represented as rays.

crystal is a set of reflected


each other, and the interference is destructive unless each beam is
in phase with all the others. For a given lattice plane spacing d, this

from the face of the


fere with

condition

is

reflection 6.

satisfied

only for certain values of the ang^e of incidence and


these angles, consider a typical pair of reflected

To determine

and r + 1. Since there is no phase change on reflection, these


will be in phase if the difference between the path length of the
radiation following beam r + 1 and the path length of the radiation

beams
beams

Sec. 2]

MEASUREMENT OF X-RAY SPECTRA

following

beam

r is

equal to an integral

they will be in phase

479

number of wavelengths. That

is

if

nX,

2, 3,

1,

Since
djd

sin 6

and
eld

cos<f>

cos (180

r+

s\

<A

Figure 14-3.

IS

r\r

20)

1
/

Illustrating the partial reflection of an

X-ray beam from each

lattice plane

of a crystal.

this

condition can be written

A
sin 6

This

cos (180

2d)

nX

sin 6

is

-4-n (1 -

cos 26)

= -A- (1 -

sin 6

2 sin 2 d)

nX

sin 6

or

2d
This condition

nX,

1, 2, 3,

(14-1)

Bragg 's law. For an incident beam of X-rays of


beam will be found only if 6 is one of the solutions
condition. The solution corresponding to n = 1 gives the so-called

wavelength
to this

sin d

X,

is

called

a reflected

first order reflection;

2 gives the second order

intensity of the first order reflected

beam is

reflection, etc.

The

greater than that of the second

order, the second greater than that of the third, etc., because the intensity
of the radiation scattered by atoms decreases with the increasing scattering

X-RAYS

480
angles corresponding to increasing values of n.

necessary to use the

first

For

this

reason

[Ch. 14
it is

often

order reflection although the higher orders give

better spectral resolution.

Figure (14-4) shows an X-ray tube and crystal spectrometer. Following

modern X-ray tubes are operated at high


vacuum, and the electron beam is produced by a heated filament F.
A high voltage applied between the filament and the anode accelerates
the electrons into the anode A, which usually must be water cooled. For
maximum X-ray yield, a heavy element such as tungsten is used for the
the design of Coolidge (1913),

An

Figure 14-4.

anode

however,

it is

apparatus used to measure X-ray spectra.

possible to use almost

emitted from the anode are defined into a

and

reflected

from the

crystal C.

any

solid substance.

beam by

For a given angle

X-rays

lead diaphragms D,
6,

only those wave-

The reflected beam


is defined by a second set of diaphragms D' and is detected by an ionization
chamber /. The X-rays enter through a thin window and pass between two
lengths are reflected for which Bragg's law

is satisfied.

producing ionization of the gas contained in this region (ionization


will be discussed later). Ions are attracted to the insulated plate
because of the electric field produced by the battery B, a current consequently flows through the galvanometer G, and the strength of the current
is proportional to the intensity of the reflected X-ray beam. By measuring
the galvanometer current as a function of the angle 0,t the spectrum of

plates,

by X-rays

by the anode is measured as a function of that parspacing d is known, this can be converted into a
a function of wavelength by using Bragg's law (14-1), since

the X-rays produced

ameter.

If the lattice

spectrum as
it is

easy to identify the order of the reflection,

n.

The lattice spacing d can be obtained for a simple cubic crystal like
NaCl from a knowledge of Avogadro's number N, the density p of the
is the molecular weight of NaCl,
crystal, and its molecular weight. If

t It

is

necessary to rotate both the ionization chamber and the crystal to keep both

angles equal to

6.

MEASUREMENT OF X-RAY SPECTRA

Sec. 2]

W grams.

then vV molecules weigh

and

in that

volume there

will

be

They

IN

will

atoms.

arranged at the corners of a cube of length

p
Substituting the

known
d

2.820

occupy a volume W\p cm3


For NaCl, these atoms are
Thus
,

2N

values of N,

d.

481

W, and p

x 10- 8 cm

for

gives

NaCl

This value has been verified by using Bragg's law to determine d in terms
of the known wavelengths of certain X-rays. These wavelengths were

from measurements using a ruled diffraction grating of


line spacing. Although the line spacing of any ruled
grating is very large compared to X-ray wavelengths, ruled gratings can
still be used if the X-rays are incident at a grazing angle because, under
actually obtained
directly

measurable

these circumstances, the effective

This technique

(i.e.,

projected) line spacing

is

very small.

and it is not appropriate for the routine


measurement of X-ray spectra. However, it is valuable for calibration
purposes, and it forms the basis of one of the most accurate experimental
determinations of the value of Avogadro's number.
is

very

difficult,

X-rays are extremely useful in studying the properties of crystals.

By bombarding

a crystal with a beam of X-rays of known wavelength


composition, and determining the angles at which Bragg reflections occur,
it

is

possible to determine the structure

and dimensions of the

crystal

As an example, such measurements confirm our supposition


that the atoms of NaCl are arranged in a regular cubic lattice. However,
we shall not describe the results of such experiments because we are
more interested in describing the results of the converse experiments,
in which the wavelength composition of an X-ray beam is measured by
using a spectrometer with a crystal of known properties. Typical results
lattice.

are indicated in figure (14-5), which represents the energy content per unit
wavelength interval emitted by an X-ray tube with a 74
anode. The
wavelength spectrum is shown for two different values of accelerating

two different values of the kinetic energy E eV o$


upon the anode. It is apparent that the observed
spectrum can be decomposed into a set of sharp lines superimposed on a
continuum. This decomposition is convenient because measurements
using a variety of anode materials and accelerating voltages demonstrate
that the properties of the continuum spectra are quite different from those
voltage V,

i.e.,

for

the electrons incident

of the line spectra.

Some

of these properties are described below.

X-RAYS

482

[Ch. 14

V=80kv

40 kv
0.2

0.4

0.8

0.6

1.0

1.2

1.4

1.6

X(xl0~ 8 cm)
Figure 14-5.

Typical X-ray spectra.

Continuum Spectra

(I)

The shape of the continuum spectrum depends only on the energy of


electrons incident on the anode, and not on the nature of the anode.

the

In particular, the short wavelength cut off Xq is inversely proportional


cjl
to the incident electron kinetic energy E. The cut-off frequency v
2.

follows the empirical law


hvv

where h

is

=E

(14-2)

in excellent numerical agreement with the value of Planck's

constant obtained from the photoelectric effect, etc.


3. The total X-ray energy per incident electron, W, which is proportional
to the integral over X of the continuum part of the spectrum, obeys the
empirical equation

W = kZE
where

~ 0.7

10~ 4

for

(14-3)

W and E in Mev

and

Z is
The

the atomic

number of the anode

material

fraction of electron kinetic energy converted into X-ray energy

W\E

= kZE

is

(14-3')

X-RAY LINE SPECTRA

Sec. 3]

For the

typical case

Z=

483

E=

90 and

Mev, WjE

0.05

is

only about 0.3

percent.

Line Spectra

(II)

1.

The wavelengths of the lines depend only on the nature of the anode
on its atomic number Z), and not on the energy of the incident

(in fact,

electrons.
2.

However, a

line

cannot appear in the spectrum unless the incident

electron kinetic energy

E satisfies

the relation

E^?ET =
where
3.

is

The

hv

hc\l

(14-4)

the frequency and X the wavelength of the line in question.

total

X-ray energy emitted in a particular

line increase

with

increasing incident electron kinetic energy according to the empirical

equation

Ioz{E- ET) n

(14-5)

where
n

The

origins of the

properties.

3.

For

this

r+j 1.5

continuum and
reason

it

is

line spectra are as different as their

convenient to discuss them separately.

X-ray Line Spectra


The

origin of the X-ray line spectra

may

be understood

if

we

a particular one of the processes by which an energetic electron

down

as

it

passes into the anode of an X-ray tube.

consider
is

This process

slowed
is

the

from the incident electron to the atoms of the anode by


Coulomb interaction collisions between the incident electron and the
transfer of energy

electrons of the atoms.

An

upon an atom can transfer an appreciable


an atomic electron only in a close collision
This statement may be justified and made quantitative,

energetic electron incident

fraction of

its

kinetic energy to

with that electron.

by considering the Coulomb interaction potential energy between two


electrons
V(r)

e 2 /r

Evaluating from this equation the separation distance r at which the


is V(r)
10 2 ev, we find r
10 -9 cm. Since the energy

potential energy

from the incident to the atomic electron cannot be bigger than


the maximum potential energy which they experience, we see that energy
transfer

X-RAYS

484
transfers greater than 100 ev occur only in collisions in

of closest approach between the two electrons


distance

is

is less

[Ch. 14

which the distance

than 10 -9 cm. This

about 10 percent of an atomic "diameter." Such close collisions

are not very frequent because the atomic electrons in the outer shells can

move out of

the

way of

the incident electron and, although the atomic

electrons in the inner shells cannot

do

this as

bound to the
come near
Almost all the

they are tightly

center of the atom, the probability that the incident electron will
these electrons

is

small because the inner shells are small.

an energy transfer of a few electron volts or,


few tens of electron volts. In such collisions, it is energetically
possible to excite only an atomic electron in one of the outer shells. The
electron can be excited into a bound state, or into a continuum state
collisions are associated with

at most, a

atom

leaving the

ionized.

some other

In either case, the electron, or

drop back down through the various excited states


to the ground state, and the atom will emit lines of its optical spectrum.
In a small energy transfer collision there is not enough energy available to
excite an atomic electron from an inner shell. These electrons are in
highly negative energy levelsf and, if they are to be excited, they must be
excited all the way up to one of the slightly negative bound energy levels,
or to one of the positive continuum energy levels, because all the intervening
energy levels already contain their complete complement of electrons
(cf. section 5, Chapter 13).
However, there are occasional collisions in which the energy transfer
is a big fraction of the energy of the incident electron.
Providing this
energy is large enough (cf. equation 14-4), it is possible to excite an atomic
electron of an inner shell from its highly negative energy level to one of
electron, will eventually

the slightly negative or positive energy levels.

very highly excited state.


state,

and

in this process

members of

it

The atom

This leaves the atom in a

will eventually return to its

ground

emits a set of very high energy quanta which are

spectrum. As an example, let us assume that


removed from the n = 1 shell. In the first step of
the de-excitation process an electron from one of the shells of higher n
drops into the hole in the n = 1 shell; for instance, an n = 2 shell
electron could drop into the hole. This would leave a hole in the n = 2
shell, but the excitation energy of the atom would be considerably reduced
because an electron in the n = 2 shell is much less tightly bound than an
electron in the n = 1 shell. Energy is conserved by the emission of a
quantum which has an energy equal to the decrease in the excitation energy
of the atom. In the next step of the de-excitation process an electron from
a shell of n > 2 drops into the hole in the n = 2 shell, a hole is produced
in the shell from which it came, and a quantum of the appropriate energy is

an electron

its

X-ray

line

is initially

For example, the energy of an electron

in the

innermost shell of

74

is

7 x

10 4 ev.

X-RAY LINE SPECTRA

Sec. 3]

485

is that a hole jumps


These steps continue until the hole has worked its
way to the outermost shell, where it is filled by the electron initially ejected

emitted.

Note

that the net effect of each of these steps

to a shell of higher

from the n =
The energy

n.

some other

or

shell,

electron.

an atom which are involved in the emission of its


X-ray spectrum can be conveniently represented in terms of an energy
level diagram. We present in figure (14-6) a diagram for the 92 U atom
showing all its X-ray energy levels through n = 4. In this diagram the
levels of

K series

10

*I

series

.L,

11= -

-L !r
-illl

'

'

/Mi

series

Mm

Af

'My
'

* (- .

"-I

10'

'-

'

'

uU

'

'

3
3

3
3
3

_5 i

-\ - -\

.*-JV,
'-

'

'

'i

'

'

'V

10'

Figure 14-6.
transitions

The higher energy X-ray

between these

levels for the

uranium atom, and the possible

levels.

energy of the atom in

its ground state is denned as zero. Each level gives,


on a logarithmic scale, the energy of the atom when one electron of the
indicated quantum numbers n, I, j is missing.f This can be thought of,
then, as an energy level diagram for a hole with quantum numbers n, l,j.
Since an electron of negative energy is missing, the energies of all these
levels are positive.
The energy levels are also identified by a notation

commonly used

in

X-ray spectroscopy.

In this notation the value of

t Note that the quantum numbers n, I, j, which completely specify the energy of an
atom with one electron in a subshell, also completely specify the energy of an atom with

one hole in the subshell. Because of the coupling of the individual spin and orbital
angular momenta required by the exclusion principle for the electrons in a subshell, the
magnitude of the
hole

is

the

the hole.

angular

same

and spin angular momenta of a subshell which has one


the subshell contained only the single electron corresponding to

total, orbital,

as

if

In either case the

momentum

is

quantum number

not written as

it

specifying the magnitude of the spin

can only be

J.

Cf. discussion of table (13-9).

X-RAYS

486

[Ch. 14

quantum number n is given according to the scheme shown in table


The Roman numeral subscripts in figure (14-6) label the levels
according to decreasing energy, but do not specify the values of / and j.

the

(14-1).

Because an X-ray energy

level

diagram gives the energy

the energy ordering of the levels with respect to the


n,

I,

is

levels

of a hole,

quantum numbers

the reverse of that found in an optical energy level diagram giving

the energy levels of an electron.

atom with one

electron of

This

is

true because,

quantum numbers

atom with a corresponding hole


more energy must be given to the atom in order

the energy of an

is

as

to

TABLE

if

the energy of an

n, /,/is particularly negative,

particularly positive,

remove the

electron.

(14-1)

rhe Spe ctroscopic Nota tion for n


n
Spectroscopic

notation

Keeping

this difference in mind, all the familiar features of the dependence


of atomic energy levels on the quantum numbers n, I, j can be found in

an X-ray energy

level

diagram.

In fact, they are

more pronounced,

particularly the splitting of the energy levels according to the value of j.

In the

of

shell

92

this splitting is

more than 2000

and

ev,

it is

larger

than the splitting according to the value of /. In such a case, it is hardly


appropriate to describe they dependence of the energy levels by the optical
energy level terminology "fine structure splitting." This strong j depend-

atoms except those of

which

is

characteristic of the inner shells of all

very low Z,

is

partly due to the usual increase in spin-orbit splitting with

ence,

decreasing

n,

but primarily due to other

relativistic effects

which become

very large for the high velocity electrons that populate these shells

(cf.

section 6, Chapter 13).

As we have
X-ray

indicated,

it is

convenient to think of the production of

terms of the creation of a hole in one of its higher


and the subsequent jumping of the hole through its lower

line spectra in

energy

levels,

energy

levels,

with the emission at each

carrying the excess energy.

jump

of an X-ray quantum

The frequency v of the quantum bears

the usual

which it carries; E = hv. However, not all


transitions are possible. There are the following set of selection rules for
the change in quantum numbers of the hole
relation to the energy

A/= 1

X-RAY LINE SPECTRA

Sec. 3]

487

Note that these are the same as the selection rules (13-30) and (13-31)
which govern the one optically active electron in an alkali atom.f In the
X-ray energy level diagram for 92 U, we have shown the transitions that
obey (14-6). The totality of X-rays which are emitted in these transitions
(plus a few which are observed to be emitted very infrequently in violation
of these rules) constitute the X-ray line spectrum of the atom. It is conventional to group those lines produced by transitions from the K shell
into the so-called K series, from the L shell into the L series, etc. X-ray
energy level diagrams are obtained experimentally from an analysis of
measured X-ray spectra, and the analysis is greatly facilitated by this
grouping.

One

between X-ray and optical spectra is that X-ray


from element to element; there are none of the
abrupt changes from one element to the next which occur in optical spectra.
The reason is that the characteristics of X-ray spectra depend essentially on
the binding energies of the electrons in inner shells. With increasing atomic
number Z, these binding energies simply increase uniformly, owing to the
higher nuclear charge, and are not affected by the periodic changes in
the number of electrons in the outer shell. The regularity of X-ray spectra
was first observed by Moseley. In 1913, shortly after the invention of the
crystal spectrometer, he made a survey of X-ray spectra and obtained data
for a number of elements on the wavelengths of two prominent lines,
Kx and Lx The first line is emitted in transitions from the K shell to the L
shell, and the second is emitted in transitions from the L shell to the
shell. (Actually, Ka and L are each comprised of two lines, but Moseley
x
could not resolve this structure.) Moseley found that these wavelengths
could be fitted to within the experimental accuracy by the simple empirical
striking difference

spectra vary smoothly

formulas

For

For L a

1/A

= CK J,Z -

1/A

= Cia(Z -

(14-7)

l)

7.4)

(14-7')

CKa and Cia are constants having values of the order of the Rydberg
constant for an infinitely massive nucleus, R x defined by (5-15). Moseley
where

on the basis of the Bohr theory of the atom, which


had been proposed in the same year as his experiments. According to
equation (5-14), the Bohr theory predict* that the energy of an electron
in the shell of quantum number n is
interpreted these results

En

=-R x hc
rr

(14-8)

t This is a manifestation of the symmetry, described in the previous footnote, between


the properties of a subshell with one electron and a subshell with a corresponding hole.

X-RAYS

488

[Ch. 14

where we have introduced R^. As the excitation energy of an atom with a


electron missing from the shell of quantum number n is equal to \En \,
the energy E of the X-ray quantum emitted when the hole jumps from
the n t shell to the n t shell

E
This energy

is

RJicZ\\\n\

1/n

2
)

quantum by

related to the frequency v of the

is

E=

hv

the equation

hcjk

so

hcjk

RJicZ\\\n\

1/A

RZ^l/nf

1/n

2
)

or

For

nt

a,

and n t

2; for

Lx

nt

1/n ,)

2 and n t

(14-9)

3.

Therefore the Bohr

theory predicts

For
For

Lx

\\l

1/A

=
=

[R^l/1 2

1/2 )]Z

(14-10)

(14-10')

[KJ1/2

1/3 )]Z

In each case the value of the constant in square brackets agrees well with
the constant

in equations (14-7)

equations contain (Z

1)

and (Z

and

(14-7').

7.4),

instead of Z, was explained by

That these empirical

Moseley as due to shielding of the nuclear charge by atomic electrons


(which was not accounted for in Bohr's derivation of equation 14-8).
According to Moseley, the one electron remaining in the K shell during
the transition of the hole from the K to L shells shields the nucleus so that
the effective Z is reduced to (Z 1) for the emission of the Kx line.
Similarly, the nine electrons remaining in the K and L shells shield the
shells. However,
nucleus during the transition of a hole from the L to
this shielding is not perfect, and the effective Z for the emission of the Lx
line is reduced only to (Z 7.4). Moseley's description of shielding is
closely related to the description connected with equation (13-22), and

both are only approximations to the more accurate description provided


by the Hartree theory. Furthermore, later measurements have shown that
the empirical formulas (14*7) and (14-7') are also only approximations.
Nevertheless, Moseley's work was an important step in the development
of modern physics. His simple and successful application of the Bohr
theory to X-ray line spectra provided one of the earliest confirmations of
that theory.

By using

the empirical formulas (14-7)

and

(14-7')

to

determine Z, Moseley established unambiguously the correlation of the


nuclear charge of an
elements.

He

atom with

its

ordering in the periodic table of the

found, for instance, that the atomic number of

27

Co

is

less

X-RAY CONTINUUM SPECTRA

Sec. 4]

489

than that of 28 Ni even though its atomic weight is greater. He also showed
that there were gaps in the periodic table, as it was then known, at Z = 43,
61, 72,

and

Elements of these atomic numbers have

75.

all

subsequently

been discovered.

4.

X-ray Continuum Spectra

The transfer of energy from an incident electron to the electrons in the


atoms of an X-ray tube anode is the process which is primarily responsible
for the slowing down of the incident electron, and it also leads (in the
infrequent case of a large energy transfer) to the emission of the X-ray line
spectra. But this process is not the only one which the incident electrons
undergo. Another is the Coulomb interaction scattering of these electrons
atoms in the anode. In a single
an incident electron suffers large accelerations and
can be scattered through a large angle. Although the electron does not
in close collisions with nuclei of the
collision of this type

lose energy to the nucleus because nuclei are so massive that they

do not
sometimes does lose energy by electroThis radiation constitutes the continuum X-ray

recoil appreciably, the electron

magnetic radiation.
spectrum.f

Both classical theory and quantum theory predict the emission of electromagnetic radiation by a scattered electron. Classically, small amounts of
radiation are emitted at

all times in every scattering because Maxwell's


equations lead to the conclusion that charged particles radiate whenever
they accelerate. In the quantum theory this is not what happens. Instead,

large

amounts of radiation (quanta) are emitted occasionally in a scattering


The situation can be compared to that which exists for the atom

process.

the classical theory predicts the constant emission of radiation by the


accelerated electron circling the nucleus, while the quantum theory

when the electron changes


In the case of the bound atomic electron, we know that

predicts the occasional emission of radiation

quantum

states.

conclusions drawn from the

drawn from

quantum theory

are always correct

the classical theory are generally incorrect.

and those

In the case of

the unbound scattered electron,

it is also true that the quantum theory


always leads to the correct conclusions. However, in this case the classical
theory does not lead to conclusions which are generally incorrect. In

theory is capable of describing quite well certain features


of the production of continuum spectrum X-rays. The contrast between

fact, the classical

t X-rays of the

continuum spectrum are often

word means braking

radiation

(i.e.,

called Bremsstrahlen.

deceleration radiation)

the theoretical explanation of their origin.

and

is

name

This

German

derived from

X-RAYS

490

[Ch. 14

the two cases reflects the fact that the differences between a quantized

and

a non-quantized theory are not so fundamental in the case of an unbound


particle as they are in the case of a

bound

particle.

According to equation (2-7) of the classical theory, a particle of charge


ze, suffering an acceleration a, will emit energy in the form of electromagnetic radiation at the rate

R=

l*
3

When

(14-11)

by a nucleus of charge Ze, it experiences a


and, if the mass of the particle
2
Consequently the rate
is m, its acceleration is proportional to zZe jm.
of emisssion of energy in electromagnetic radiation can be written
this particle is scattered

Coulomb force which is proportional to zZe2

oc s2 a 2 oc 4

/w 2

(14-12)

factor 1/m 2 predicts that the emission of electromagnetic radiation

The
very

much more important

heavy

particles.

This

is

for electrons than

true;

it is

is

for protons or other

the energy emitted in continuum spectrum

X-rays by a heavy charged particle in traversing matter is almost immeasurably small. The factor Z 2 predicts that anodes of high atomic

much more

number

are

verified

by experiment.

efficient

By

producers of these X-rays.| This

is

also

evaluating the proportionality constant in

and then integrating over the (average) trajectory of the incident


coming to a stop in the anode, it is possible to estimate the
total continuum spectrum X-ray energy JFthat is radiated. The results are
in reasonable agreement with the empirical equation (14-3).
The classical theory is also capable of giving a good account of the
polarization of the continuum spectrum X-rays. This theory predicts that
the X-rays, emitted in the direction r during an acceleration a, are polarized
with the electric vector perpendicular to r in the plane defined by r and a.
Now the higher energy X-rays are all emitted in "head-on" collisions in
(14-12),

electron in

always approximately antiparallel to i, the direction of the


incident beam of electrons. Consequently, such X-rays have a large net
polarization. In the collisions which lead to the emission of lower energy

which a

is

X-rays, the direction of a


polarization

is

is

not so constrained and, as a

result, the net

small.

Classical electromagnetic theory even gives

an accurate prediction of
continuum

the angular distribution of the X-ray energy emitted in the

The

dependence also shows that the radiation emitted

atomic electron

(i.e.,

for

Z=

1) is negligible.

in scattering

from an

X-RAY CONTINUUM SPECTRA

Sec. 4]

491

For incident electrons of non-relativistic


by equation (2-6) is

spectrum.

energies, the angular

distribution predicted

2
1(6) oc sin 6

where
which

is

(14-13)

the angle between r and the average direction of acceleration,


i.|
For incident electrons of relativistic energies, certain

is

relativistic

transformation effects shift the angular distribution in the

direction of

i.

The one aspect of the emission of continuum spectrum X-rays that


cannot be explained by classical theory is the shape of the spectrum; this
can be explained only by quantum theory. Consider the cut-off frequency
(equation 14-2), which is certainly the most striking feature of the spectrum.

v
Figure 14-7.

thin

>-

=E/h

anode X-ray continuum frequency spectrum.

This cut-off exists because the highest energy quantum that can be emitted
by an electron of kinetic energy E has the energy hv = E, where v is
its frequency.
In fact, this was one of the earliest pieces of evidence in
support of Einstein's theory of quanta. Unfortunately, an explanation
of the other features of the spectrum cannot be given by such elementary

quantum
subject
results

theory, but only in terms of

quantum electrodynamics. This

beyond the level of this book, and we can only describe the
of the theory. For electrons of energy E incident upon an anode
far

is

which is so thin that their energy remains essentially constant in passing


through it, the spectrum is predicted by the theory to have the simple
shape shown in figure (14-7). The quantity
unit frequency interval.

I(v) is the

X-ray energy per

For comparison with the data presented

in

we convert this to I(X), the X-ray energy per unit wavelength


The relation between two quantities such as these has been derived

section 2,
interval.

before

(cf.

equation 2-27).

It is

IW =

I(v)

t This angular distribution has the same form as the angular distribution of the
energy emitted in radiofrequency radiation by electrons accelerated back and forth
along a linear antenna. From the point of view of the classical theory, this process is
essentially the same as that involved in X-ray emission.

X-RAYS

492

Now,

since
I(v)

I(X)

[Ch. 14

~ constant,

<

>

cjv

E/h

can be written
I(X)

This spectrum

is

~ constant/A

ch/E

The predicted wavelength


good agreement with measurements

plotted in figure (14-8).

spectrum for a "thin" anode

is

X = ch/E
Figure 14-8.

A thin

in

anode X-ray continuum wavelength spectrum.

made for thin anodes. It is also in agreement with the more


measurements using "thick" anodes in which the incident electrons
are brought to a stop. We show this in figure (14-9) by generating a
thick anode spectrum from the sum of several thin anode spectra. Four
that have been
typical

Figure 14-9.

Illustrating

the composition of a thick anode X-ray continuum wavelength

spectrum.

values of

are used to represent the range of electron energies present

The dotted line shows the results of this approximate


The dashed line indicates the results of an accurate calculainvolving integration over the electron energy. Agreement with

in the anode.
calculation.

tion

experimental data, such as presented in section

2, is excellent.

THE SCATTERING OF X-RAYS

Sec. 5]

The Scattering

5.

493

of X-rays

Having discussed the nature and production of X-rays, we now turn


our attention to their interaction with matter. Consider a beam of X-rays
passing through a foil and into an X-ray detector, as indicated schematically
in figure (14-10). A quantitative measurement of the interaction of the
X-rays with the material contained in the foil is obtained by measuring
the intensity of the attenuated beam striking the detector, and then, by

removing the

measuring the intensity of the incident beam. The


of the attenuation of the foil for these X-rays,

foil,

results are expressed in terms

which

is

defined as the fractional decrease in the intensity of the beam.

Incident

Attenuated

beam

beam

Figure 14-10.

An

Detector
jl

apparatus used to measure the scattering and absorption of a beam

of X-rays.

In passing through the

foil, X-rays can be completely absorbed from the


beam, and they can also be scattered. If the angle subtended at the foil
by the detector is very small, then essentially all the scattered X-rays will
miss the detector and the measurement will give the attenuation due to
both absorption and scattering.! In such measurements it is found that

the attenuation increases with the thickness of the foil and, generally,
with the atomic number of the material which it contains. The attenuation
also depends

on the energy of the X-rays, but the dependence

easy to describe.
the absorption

In

this

Several very interesting

and

section

scattering processes

we

shall

discuss

phenomena

is

not so

are involved in

which produce the attenuation.

those

involved

in

the

scattering

process.

Consider the interaction of a

atoms of the

foil.

We make

beam of X-rays with

1. The wavelength of the X-rays


dimensions of the atoms

By

varying this angle

the electrons in the

the following assumptions.


is

small compared to the characteristic

< 10"

cm

(14-14)

it is possible to separate experimentally the attenuation


due
from the attenuation due to scattering, and measure these quantities
independently. The attenuation due to scattering can also be measured separately by
an experiment in which the scattered X-rays are detected.

to absorption

X-RAYS

494
2.

The energy of

quantum of

the X-rays

is

large

[Ch. 14

compared

to the

binding energy of the atomic electrons

/n>>B.E.
3.

rest

The energy of a quantum of the X-rays


mass energy of an electron
hv

The

(14-15)

let

small compared to the

<^mc 2

(14-16)

be discussed

of these assumptions will

validity

moment

is

us explore their consequences.

The

first

later.

For the

assumption implies

by the electrons in
same atom, will be essentially random, or incoherent.
It rules out interference effects between different atoms (such as those
discussed in section 2), as well as interference effects within the same atom.
If this assumption is valid, the interaction of any electron with the X-rays
is independent of the interactions of all the other electrons in the foil, and
we therefore need to consider only one of these interactions. The second
assumption allows us to ignore the binding of the electrons in the atoms and
consider them to be free. The third assumption allows us to ignore the
quantum aspects of X-rays which manifest themselves in Compton
that the phase relations between the X-rays scattered
different atoms, or the

Making these assumptions, we

give a classical electromagnetic

treatment of the scattering process due to

Thomson (ca. 1900). In this


beam of electromagnetic

scattering.!

treatment the X-rays are considered to be a

waves incident upon a


with the electron

is

single free electron.

The

sinusoidally oscillating electric field incident

with the magnetic


strength

and

interaction of the X-rays

then due to the interaction of


field is negligible.)

We

upon

its
it.

charge with the


(The interaction

describe the instantaneous

direction of the electric field at the location of the electron

by the vector E. This electric field acting on the electron produces a


force F = eE, and an acceleration
a

where E, F, and a

F//M

all oscillate

incident electromagnetic waves.

= -eE/w

(14-17)

sinusoidally with the frequency v of the

Under

the influence of this acceleration,

v, and
same frequency as,
and in phase with, the incident waves. The energy emitted in these waves
Thus the electron
is supplied by the incident electromagnetic waves.
absorbs energy from the beam of X-rays and scatters it in all directions.
The rate of emission of energy by the accelerated electron, that is, the

the position of the electron will oscillate sinusoidally with frequency


the electron will emit electromagnetic waves of the

t Equation (3-17) shows that if (14-16) is valid the


scattering of X-rays from a free electron is negligible.

change of wavelength in the

THE SCATTERING OF X-RAYS

Sec. 5]

rate at

2=

which the electron scatters energy,

given by equation (14-1 1) with

-1:

S
where here we write

5 for

E changes

to deal with

Using (14-17),

this

|e*/if^c

constantly in time,

its

feV/c3

scattered.

S
As

is

495

does

also,

becomes
(14-18)

and

it is

most convenient

time average,

S = fe^/mV3

= fe^/m2

*:

(14-19)

The quantity S is the average energy emitted in all directions per second
by the electron in the form of electromagnetic radiation, and we define
be the intensity of the scattered radiation.
/, the intensity of the incident radiation. We define
this quantity to be the average energy carried by the beam per second
across a 1 cm2 area oriented normal to the direction of the beam.f
Equation (2-5) shows that the relation between / and the average energy
this quantity to

Next

density

let

p~

us evaluate

is

I=pc
According to equation

(2-4),

p~

(14-20)

is

12/477

where we delete the subscript for transverse. Thus / may be written

I=cE?\bn

Now

(14-21)

comparing equations (14-19) and (14-21), we see that both

and /are proportional


other. That is

to

E2

Sand /are

Therefore,

proportional to each

S = a TI

(14-22)

where a T is the proportionality constant. Evaluating a T from equations


(14-19) and (14-21), we have
8ir

= -t (e2/wc2 =
2

6.66

x 10" 25 cm 2

(14-23)

This quantity, called the Thomson cross section of the electron, describes
a characteristic property of a free electron. Through equation (14-22),
aT tells what fraction of the intensity of a beam of X-rays is scattered into
all directions by a free electron due to the Thomson scattering process.
t

Note that the dimensions of S are

cm~*-sec _1

ergs-sec

-1
,

while the dimensions of I are ergs-

X-RAYS

496
It is

of

cm3

[Ch. 14

reasonable to call a T a "cross section" because it has the dimensions


But the name has a deeper significance that can be understood if
.

we imagine

the electron at the center of an imaginary disk oriented normal


beam, and of area equal to a T cm2 Then, if all the radiation in the
beam hitting the disk were scattered, the ratio of the scattered intensity 5
to the incident intensity I would be just a T in agreement with equation
(14-22). This is easy to see by referring to figure (141 1) (in which the
to the

Figure 14-1

1.

The geometrical interpretation of

a cross section.

area of a T is exaggerated by about a factor of 1023 !). The average energy


incident per second on the 1 cm2 area is equal, by definition, to the incident
intensity

/.

The

fraction of this energy that hits the imaginary disk

cm2 /l cm2

is

a T If all this radiation were scattered,


the energy scattered per second, which is the scattered intensity S, would
be a T I. This proves our point. The geometrical interpretation of a cross
section is convenient for visualization, and even calculation, but it should

precisely the ratio

Figure

14-12.

aT

Illustrating

the relation between the vectors which determine the

polarization of scattered X-rays.

literally as to lead to the conclusion that there actually


2
a well-defined distance r, satisfying the equation 27rr = a T such
electron
are
scattered,
distance
of
the
within
this
that all X-rays passing
not
scattered.
X-rays
are
whereas all other
Classical electromagnetic theory also makes predictions concerning the

not be taken so
exists

angular distribution of the X-rays scattered by an electron in the Thomson


scattering process. Consider figure (14-12). The direction of the incident

Sec. 5]

THE SCATTERING OF X-RAYS

497

beam of X-rays

is the vector i. The vector E, which is perpendicular


to i,
the direction of the electric vector of a typical polarization component
of the beam, and the acceleration a which this component gives to the

is

electron

lies in the opposite direction.


The angular distribution of the
radiation emitted by the electron suffering this acceleration is given by

(14-13), where d is the angle between the direction of acceleration a and


the direction of emission of the radiation r. This angular distribution may

Figure 14-13.

The element

of solid angle between

and

d.

also be written in terms of the angle

between r and i, and the angle


on the figure) specifying the azimuthal orientation of E.
Doing this, and then averaging over all possible azimuthal orientations
to
treat the case of an unpolarized incident beam,
the angular distribution is
(not indicated

obtained as a function of the angle 0.


cc (2

5(0)

where S(0)

is

It is

sin 2

0)

(14^24)

the average energy emitted per second per unit solid angle
The integral of S(@) over all solid angles is equal to the

at the angle 0.f

average total energy emitted per second, or the intensity


of the emitted
Let us write (14-24) as

radiation S.

S(0)

and evaluate the

K(2

sin 2

0)

(14-24')

integral:

S=
J

The element of solid


cones of half-angles
the sphere. This

is

5(0) dQ.

angled is equal to the area intercepted on a sphere by

and

d, divided by the square of the radius of

illustrated in figure (14-13).

The area

is 27rr 2

sin

'

so
,

n=

dil

f Since solid angle

is

2ttt

sin

277 sin

dB

dimensionless, the dimensions of S(@) are ergs-sec

(14-25)
1
.

X-RAYS

498

[Ch. 14

Consequently

S=

2ttK f*(2

sin

0)

sin

d&

(14-26)

Jo

According to equations (14-22) and (14-23),

{^Imcfl

S=
By

(14-27)

carrying out the integration of (14-26), and equating the resulting

we

expression to (14-27),

S{&)

evaluate K.

With

i(e 2 /wc 2 ) 2 (2

this value, (14-24')

sin2

0)/o

becomes
(14-28)

Denning the symbol

da T jdQ.

=
=

i(e

/wc 2) 2 (2

3.98

10- 26

(2

sin 2

equation (14-28) can be written

S(@)

0)

sin 2

0)

cm2 /unit

solid angle

(14-29)

(14-30)

dQ,

The quantity da T \dl, which also has the dimensions cm2 is called the
differential Thomson cross section of the electron. It tells, through equation
(14-30), what fraction of the intensity of a beam of X-rays is scattered in a
Thomson scattering process by a free electron into a unit solid angle at
the angle 0just as a T specifies, through equation (14-22), what fraction
of the intensity is scattered in that process by a free electron into all solid
,

angles.

It is possible to give

a geometrical interpretation to a differential

which we gave above for a cross section.


However, we again caution the reader against taking these interpretations

cross section similar to that

too

literally.

Having completed our discussion of the Thomson scattering of X-rays


by a free electron, let us return to the more realistic problem of the
scattering of X-rays by an atom. If assumptions (14-14) and (14-15) are
satisfied, this problem is trivial because each of the Z electrons of the
atom scatters X-rays independently, and the total intensity scattered by
the atom is Z times the intensity scattered by a single electron. Evaluating
this intensity from equation (14-22), we obtain for the total .?

S = Za T I
Defining

= Za T

(14-31)

S=as I

(14-32)

aSa
this

can be written

THE SCATTERING OF X-RAYS

Sec. 5]

The quantity

<r

Sn

is

499

the scattering cross section per atom.

given by (14-31) only

if

It

has the value

the assumptions mentioned above are satisfied.

If they are, the differential scattering cross section

per atom

is

also simply

do?
^darp
~
= Z
di\

This

related to the total

is

S(Q) by the equation

S(&)

From

(14-33)'

dQ.

1
^f
dil

these scattering cross sections per

atom

(14-34)
it is

easy to calculate the

predicted attenuation, due to scattering, of a foil containing a

number of atoms per

unit volume.!

I"

comparing

this prediction

known
with the

vresults of the experiments described

at the beginning of this section,


agreement to within 10 or 20 percent is found for atoms with Z < 15
and for X-rays of wavelength near I = 0.2 x 10~ 8 cm. This formed the
basis of experiments performed by Barkla in 1909, which constituted the
first measurements of Z, the number of electrons in an atom.
The quantum energy of X-rays of wavelength X = 0.2 x 10~ 8 cm is
hv = 6 x 10 4 ev. As this wavelength and energy satisfy both assumptions (14-14) and (14-15) for atoms of small Z (and also satisfy assumption 14-16), we can understand why the cross sections per atom measured

with these X-rays agree, approximately, with the predicted values.

We

can also understand why they do not agree for measurements made with
X-rays of longer wavelength and atoms of larger Z. For these cases, the
X-ray wavelength is comparable to, or larger than, the characteristic
dimensions of the atoms, and also the electrons do not act as if they were

For both these reasons the electrons of the atom do not interact
beam of X-rays. In the limit of very long
wavelengths and very low quantum energies, the interactions are the very
opposite of being independent because, under these circumstances, the
bound electrons of the atom are all oscillating under the influence of
free.

independently with the incident

the incident electromagnetic waves with exactly the same phase, and the
electromagnetic waves emitted by any one electron are completely in
phase with those emitted by all the others. Because of this phase coherence
the scattered intensity will be very

much

larger.

It is

easy to evaluate this

by thinking of all the Z electrons oscillating together as one "big electron"


of charge Ze and mass Zm. According to equation (14-17), the acceleration of this object will be the
to (14-1

1)

the energy which

single electron
t

same
it

by a factor of

Such a calculation

will

as that of a single electron, but according

radiates will be larger than that radiated

Z2

Thus,

by a

in this limit the actual scattering

be carried out at the end of section

7.

500

X-RAYS

cross section per atom,


the electron,

The

will

(x,^,

a T andZtimes
,

be

Z 2 times the Thomson cross

[Ch. 14

section of

larger than the prediction of equation (14-31).

actual differential scattering cross section per atom, dosJdQ., will also
times larger than the differential Thomson cross section of the

Z2

be

da T jdQ,, and Z times larger than the prediction of (14-33).


For X-rays whose wavelengths and quantum energies are such that the
scattering from the various atomic electrons is neither completely coherent
nor completely incoherent, the situation is much more complicated, but it
can still be treated by classical electrodynamics. Assuming that the distribution of electrons in the atom is spherically symmetrical, the amplitude
of the electromagnetic wave scattered by each volume element is evaluated,
and then the contributions from all the volume elements are summed by
adding amplitudes with relative phase factors that depend on the location
of the volume element, the angle of scattering, and the wavelength. By
electron,

squaring the net amplitude, the total intensity scattered at the angle in
is found.
Dividing this by the incident intensity, the differential

question

scattering cross section per

atom

is

obtained.

^=^|F(x)|
dQ.

It

can be written

(14-35)

dQ.

where
F(X)

"pWCsin xrlxr)4*r* dr

(14-35')

with

= 2K sin (0/2)

and

K=
In these expressions X

and

is

277/2

the X-ray wavelength,

p(r) is the electron density

charge.

is

the scattering angle,

of the atom in units of the electronic

Since

p(r)4Trr

dr

Jo

we

see that the function F{%),

takes on

its

maximum

This condition

is

which

is

value, the atomic

satisfied

when % >

0,

called the atomic

number Z, when

form

factor,

sin %r\%r

which happens for

all

1.

when

K 0,

> 0.
and which also happens for
The behavior of F(%) for K-+0, or A

>-

of the next to

last

paragraph.

Its

confirms the conclusions

oo,

behavior for

--

shows the reason for

the small angle peaking of the scattered intensity mentioned in Section 2.

Figure (14-14) compares the differential scattering cross section per


0.71 x 10~ 8 cm incident

atom, measured with X-rays of wavelength A

THE SCATTERING OF X-RAYS

Sec. 5]

501

upon 6 C atoms, with the prediction of equation (14-33). The discrepancy at


small angles between the measured values of das jdQ. and the predictions
of equation (14-33) provide convincing evidence of coherent effects in

To compare

the scattering process.

das

the measured values

jdl with

the predictions of equation (14-35), which takes these coherent effects


into account,
in the

it

necessary to

is

form factor

Hartree

theory of the atom

obtained.

know

the electron density that enters

If the electron density

(14-35').

is

used, quite

d<Tc

for

do

x = 71 * 10

agreement

is

and (14-35') can be used to

Alternatively, equations (14-35)

obtained from the

satisfactory

cm z = 6
'

1
1

4-

.2

I
\

ail

-^is.
i

60

30

90

120

150

180

e
The differential scattering cross section per atom for A =
upon carbon. From A. H. Compton and S. K. Allison, X-rays
Experiment, D. Van Nostrand Co., Princeton, N.J., 1935.
Figure 14-14.

0.71

incident

in

X-rays

Theory and

an atom, p(r), from the measurements of


das jdO.. X-rays allow us to "see"
structure of bones in living tissue, but also the

extract the electron density of


its

differential scattering cross section

not only the invisible

invisible structure of atoms.

By

integrating

das

jdQ.

from equation (14-35) over

all solid

angles, the

predicted value of the scattering cross section per atom, a s is obtained as


an expression involving the electron density p{r). For cases in which this
,

expression can be evaluated because p(r)

is

known, the

results are in

reasonably good agreement with experiment, as long as X is greater than


about 0.2 x 10~ 8 cm. Experimental data are shown in figure (14-15).

The ratio of the observed

scattering cross section per

atom to

the prediction

of equation (14-31) is plotted as a function of the X-ray wavelength. The


curves are labeled with the atomic number of the atom. Note that these
data provide examples of the fact, mentioned above, that the observed
scattering cross sections per
0.2 x 10~ 8

cm

owing to the

atom agree with equation (14-31)

for X

but become larger than this prediction with increasing X

effect

of coherences.

Also note that the observed scattering cross sections per atom become

X-RAYS

502

[Ch. 14

smaller than the predictions of classical electromagnetic theory (with or


without coherences) when X becomes smaller than ~0.2 x 10~ 8 cm.
For X-rays of these wavelengths, the quantum energy hv is no longer very

small compared to the rest mass energy of an electron,

0.4

0.6

X
Figure 14-15.

wavelengths.

Compton and

The

is

1.0

atom

labeled by the atomic

K. Allison, X-rays

in

1.2

5.

1.4

10 s ev.

1.6

10~ 8 cm)

scattering cross sections per

Each curve
S.

0.8
(in

mc 2

for several

number

atoms and

of the atom.

a range of

From A. H.

Theory and Experiment, D. Van Nostrand Co.,

Princeton, N.J., 1935.

This violates assumption (14-16), and so the scattering can no longer be


treated by classical theory; instead, quantum theory must be used. We
have already discussed the energetics of the quantum theory for scattering
of X-rays by free electrons. The reader will find it under the heading
The Compton Effect in section 8, Chapter 3, and it would be appropriate

him to review the discussion there.


The wavelength spectra presented in Chapter 3 are for the case of
scattering from 6 C. For this small Z atom, the data show that the probability of scattering with an increase in the X-ray wavelength (Compton
scattering) is quite comparable to the probability of scattering with no
change in wavelength (Thomson scattering), even for incident wavelengths
at this point for

THE SCATTERING OF X-RAYS

Sec. 5]

x 10~ 8 cm

as large as 0.7
10 4 ev.

This

503

quantum

that correspond to

energies as

energies.

Thomson

scattering is very large compared to its value for


of the coherences, but the cross section for Compton scattering

same

as

low as

not very typical of the situation for such low X-ray


For atoms of intermediate or large Z, the cross section for
is

for

it is

is

because

much

the

C since Compton scattering takes place only on essentially


For X-rays of low energy,

free electrons.

this

means

that

Compton

scattering can involve only the electrons of relatively small binding energy
in outer shells of the

atom. As the energy of the X-rays increases, more

and more electrons of the atom become available for Compton scattering.
However, for intermediate energies, Thomson scattering can still occur
from the very tightly bound electrons in inner shells even from the

quantum point of view

because these electrons are inseparably bound to

a very massive nucleus and cannot absorb energy by recoiling, so the


quantum is scattered with no change in energy or wavelength. For

X-rays of energy large compared to the binding energy of the electrons in


atom (~105 ev for 92 U), all the electrons of

the innermost shells of the


the

atom

only

are essentially free in interactions with the X-rays.

Compton

In this limit

scattering can take place.

apparent that the dependence of the scattering cross section for a


atom on the X-ray energy would be very difficult to calculate in
the energy range we are now discussing.! In fact, calculations of the cross
It is

typical

section for

Compton

scattering have been carried through only for a

single completely free electron.

who

1928 by Klein and Nishina,

These calculations were performed in


relativistic theory of quantum

used Dirac's

mechanics, and even the results of the calculations are complicated. They
are

4a2
dan

la

da

cr.t.

a) 2

4a 2

.1

(3(1
a,

-2

O
2a sin'

(14-36)

+
+

sin

(1

cos 2 0)1

2a

sin

"

2a

In (1

2a

2a)

a
1
.(1

+ 3a
+ 2a)

In (1
2

2a)"

(14-37)

2a

where

hv

mc
f

hv
2

0.51

Mev

quantum energy
electron rest mass energy

However, the scattering cross section can be measured. Some data

in section 7.

will

be presented

504

The

X-RAYS

[Ch. 14

Compton scattering from a free electron are plotted


and (14-17) from these two equations. For Compton
of X-rays of quantum energy large compared to the binding

cross sections for

in figures (14-16)

scattering

Figure 14-16.

The

differential scattering cross section for

Compton

scattering from a

free electron.

energy of electrons in the innermost shell of the atom, an approximate


atom can be obtained by setting

estimate of the scattering cross sections per

do Sa

= z da s

(14-38)

(14-39)

JD

dQ.

and
Sm

Za s

where das/dQ, and as are given by equations (14-36) and (14-37).

0.01

0.1

10

100

1000

a
Figure 14-17.

The

As Compton

scattering cross section for

scattering

is

Compton

scattering

from a free electron.

described in terms of quanta,

it is

convenient

in considering this process to express the intensities of the incident

scattered X-rays as fluxes of quanta.

Thus we

and

redefine /to be the average

number of quanta

cm2

AND

PHOTOELECTRIC EFFECT

Sec. 6]

PRODUCTION

505

beam per second

across a

PAIR

carried by the incident

to be the average

number of quanta

scattered per second in all directions,

and redefine 5(@)

to be the average

number of quanta

area oriented normal to the beam, redefine

scattered per second per unit solid

measured from the direction of the incident beam.


from the previous ones only by a factor equal
the energy carried per quantum. Since for Thomson scattering

angle at the angle

These definitions
to hv,

differ

the energy of the scattered

quantum,

may

all

quantum is equal to the energy of the incident


we have written for Thomson scattering
equations involving the new definitions simply

the equations which

be transformed into

by dividing both sides by the factor hv. Of course,


the form of the equations, and we still have

this

does not change

S = o sJ
5(0)

(14-40)

^/

(14-41)

all

where for Thomson scattering the cross sections have exactly the same
values as before. For Compton scattering the cross sections of the previous
paragraph also satisfy equations (14-40) and (14-41) when the new definitions are used.

new definitions allow all scattering


absorption processes) to be discussed in the same

Consequently, these

processes (and also

language and, for

all

this reason,

we

shall henceforth use these definitions, or

equivalent ones, exclusively.

We
i.e.,

close this section

by pointing out that the Klein- Nishina equations,

(14-36) and (14-37),

reduce to the free electron

show

the free electron

Thomson

Compton

cross sections

cross sections in the limit of very low

X-ray quantum energies. Furthermore, equation (3-17) shows the wavelength increase in Compton scattering from a free electron becomes
negligibly small for very long wavelength X-rays because the wavelength
change cannot be larger than 0.048 x 10~ 8 cm. Thus X-rays of very low
quantum energies do not lose energy in Compton scattering. From both
points of view,

Thomson

scattering acts as the

scattering for X-rays incident

6.

The Photoelectric

Effect

upon a

low energy limit of Compton

free electron.

and Pair Production

In section 7, Chapter 3, we presented an elementary discussion of the


photoelectric effect for electromagnetic radiation in the optical frequency
range.

The reader should review

photoelectric effect

is

that discussion because the optical

the limit of the X-ray photoelectric effect as the

X-RAYS

506

quantum energy becomes very small compared


energy,

in the limit hv

i.e.,

In this limit hv

< mc

[Ch. 14

to the electron rest

mass

2
.

comparable to the binding energy of atomic electrons

is

in outer shells, but

it

is

much

very

electrons in inner shells.

Thus the

smaller than the binding energy of


optical photoelectric effect can lead

only to the ejection of optically active electrons.

In contrast, the X-ray

photoelectric effect for very large hv can lead to the ejection of any electron.

The fundamental

Compton

difference between the photoelectric effect

and the

former process the incident quantum is


completely absorbed by the atom, whereas in the latter process the
quantum scatters from an essentially free electron and goes off suffering
only a reduction in energy. We leave it as an exercise for the reader to
demonstrate that in the Compton effect it is not possible to conserve both
total relativistic energy and momentum if an essentially free electron
absorbs all the energy of a quantum. On the other hand, the complete
effect is that in the

absorption of the incident quantum in the photoelectric effect does not


violate

energy-momentum conservation because

electron concerned
static forces to the

is

not essentially

it

plays

the recoiling nucleus can take

enough

the

quantum and

From

is

Since the nuclear mass

ing

in this process the

but instead

is

bound by

to recoil in such a
is

way

as to conserve

very large, and since/?

= y 2ME,

up the required momentum without absorb-

kinetic energy to affect significantly the energy balance

this

atomic
electro-

massive nucleus. This allows the nucleus to participate

in the process, and the role

momentum.

free,

between

the electron.

argument we can

see that the probability of a photoelectric

absorption on a particular atomic electron should increase with decreasing

quantum energy, since for low energy quanta the electron is relatively
more tightly bound to the nucleus and, therefore, it is easier for the nucleus
to absorb the required momentum. However, when the quantum energy
becomes smaller than the binding energy of the electron, the probability
of photoelectric absorption on that electron abruptly drops to zero
because there is no longer enough energy available to eject the electron
from the atom. This behavior can be seen in the measured cross section
for photoelectric absorption by 82 Pb atoms, which is presented in figure
(14-18). The probability for photoelectric absorption is expressed in terms
of a photoelectric cross section per atom, a PE defined such that A, the
average number of quanta absorbed by this process per second per atom in
a beam of / quanta per second per square centimeter, is
,

A = o pE i

(14-42)

an example of the use of a cross section to describe absorption


processes. For the reasons mentioned above, the photoelectric absorption

This

is

PHOTOELECTRIC EFFECT

Sec. 6]

AND

PRODUCTION

PAIR

507

cross section increases rapidly with decreasing

quantum energy but shows


sharp breaks at quantum energies equal to the binding energies of the
electrons in each of the subshells of the atom. These energies are labeled
with the X-ray subshell designations introduced in section
tending to smaller

show

quantum

3.

that the photoelectric absorption cross section continues


tJ

10

mL

-a

1^

-18.

ex-

its

general

Ij
i

10

hv
Figure

Data

energies than those presented in figure (14-18)

10

(ev)

The photoelectric absorption

cross section per

atom

for lead.

with decreasing energy, except for breaks of diminishing severity at


and
subshells. Such data provide a direct measure
of the binding energies of electrons in the various subshells of an atom
rise

the energies of the

and can be used to confirm or supplement the binding energy values obtained from an analysis of the X-ray line spectrum of the atom.

We

not be able to calculate the cross section for photoelectric


Such a calculation requires the theory of quantum electrodynamics, and we can only present some results. The predicted cross
section per atom for photoelectric absorption on a A' shell electron is
shall

absorption.

a pK =<r T 4V'2(l/137) 4
for hv
that,

>

binding energy of

K electron,

in the energy region indicated,

Z s (#wc*/A>') H

but hv

<

shell

mc

2
.

(14-43)

Experiment shows

absorption accounts for

X-RAYS

508

[Ch. 14

and that equation


The very strong Z
dependence of a PEa reflects the fact that a K shell electron becomes
much more tightly coupled to the nucleus with increasing Z, and so the
nucleus finds it much easier to absorb the momentum which it must
almost 90 percent of
(14-43)

in quite

is

all

photoelectric absorptions,

good agreement with the

data.

absorb in a photoelectric process.


For X-rays of quantum energies in excess of two times the electron

rest

mass energy, there is a mechanism in addition to the photoelectric effect


through which the X-rays can be completely absorbed in interaction with
matter. This is the process of pair production, which consists of the complete disappearance of a quantum and the subsequent materialization of
part of its energy hv into the rest mass energy of a pair of electrons. If
hv = 2mc 2 the materialization of the two electrons uses up all the available
energy. In the general case hv > 2mc 2 and the excess energy hv 2mc 2
appears in the form of kinetic energy of the two electrons. One of these
electrons is of the usual variety with mass m and charge e. The other has
the same mass m but has the positive charge +e and is called a. positron.
Thus charge is conserved in the pair production process because the quantum is uncharged and the total charge of the electron and positron is e
,

+e =

0.

Pair production cannot take place in free space because, as the

reader can easily show, the disappearance of a quantum, followed by the

appearance of an electron and positron, cannot conserve both total relativistic energy and momentum. However, the process can take place in the
Coulomb field of an atom because there can be interactions between the pair
of charged particles and the nucleus. These interactions allow the massive
nucleus to absorb whatever momentum

is

required to conserve

without affecting the energy balance. In the next section

more of

we

momentum,

shall describe

and some of the theory, of this extremely interesting


But it seems best to complete our discussion of the attenuation
of X-rays by matter before going into these subjects.
The cross section per atom for the absorption of X-rays by the pair
production process can be calculated only with the aid of the Dirac theory.
the details,

process.

The

results are

pr
FR"

=o T /(a)
8tt

(14-44)

137

where a = hv/mc2 The Z 2 dependence of the pair production cross section


per atom, a PR is a manifestation of the fact that the required momentum
transferring interaction between the nucleus and the pair of charged
.

particles

/(a) has

is

more

much

the

readily achieved for

same form

a graph of /(a) for the

82

atoms of large Z. The function

for all atoms.

We

Pb atom. This shows

present in figure (14-19)

that

aPRa rises slowly from

Sec. 7]

CROSS SECTIONS

AND ATTENUATION

COEFFICIENTS

509

2
1.02 Mev and, for
a value of zero at the threshold energy hv = 2mc
very large hv, approaches a constant value as /(a) approaches

.,

/(oo)}
JK

28, 183

In
A

(14-45)

T7

27

We

cannot write an equation for f(x) itself, as this function is obtained


from a numerical integration. In an incident beam of / quanta per second
12

^^-
10

$.6
*-

^~

rt~)

4
2

i^i

ill

il

1000

100

10

10,000

a
Figure 14-19.

From W.

The

which arises in the theory


Quantum Theory of Radiation, Oxford University

plot of a function

Heitler, The

of pair production.
Press,

London, 1944.

per square centimeter, the average number A of quanta absorbed per


second per atom by the pair production process is related to the cross
section

a PR by the usual

relation

A = o PR

7.

(14-46)

Total Cross Sections and Attenuation Coefficients

We have now discussed all the scattering and absorption processes


which contribute in a significant way to the attenuation of a beam of
X-rays passing through matter. Let us next see how these processes
combine to produce the total attenuation observed in an experiment, such
5. The average number A of
quanta which are either scattered or absorbed per second per atom, in a
beam of / incident quanta per second per square centimeter, can be
evaluated by adding equations (14-40), (14-42), and (14-46):

as that described at the beginning of section

A=

<*sj

where
<*a

peJ + OprJ = oj

as

+
,

,..

<*

PEa

o-pie,

(14-47)

510

X-RAYS

The quantity aa

is

called the total cross section per atom.

fraction of the incident

[Ch. 14

It specifies

beam of X-rays undergoes any of

the

what

beam

attenuating interactions with a given atom.


cross sections

Figure (14-20) shows the


aSa aPEa a PRa and aa for the 82 Pb atom. For 82 Pb, the energy
,

hv
Figure 14-20.

The

(ev)

scattering, photoelectric, pair, and total cross sections per

atom

for lead.

ranges in which each of the three processes makes the most important
contribution to aa are approximately
Photoelectric effect:
Scattering:

Pair production:

Because of the different

hv
5
5

<

< hv <
ev < hv

10 s ev

10

10 5 ev
5

10 6 ev

dependencies of the cross sections for these


which they dominate aa are quite different

processes, the energy ranges in


for

atoms of low atomic number. For 13 A1 these energy ranges are approxi-

mately
Photoelectric effect:
Scattering:

Pair production

hv

<

5
4

10 4 ev

< hv <
ev < hv

10 ev

10 7

10 7 ev

CROSS SECTIONS AND ATTENUATION COEFFICIENTS

Sec. 7]

Using the
/

beam of X-rays

cm, containing

figure (14-21).

The

number of quanta
the
flux

foil.

and

atom aa

total cross section per

attenuation of a

flux (or intensity) of the

crossing

cm2

beam, which

cm

I(x)

the attenuated flux emerging

then

is

the average

beam

per sec, decreases as the

Call the flux after penetrating x


I(t) is

us calculate the total

through a foil of thickness


Consider the situation shown in

in passing

cm3

atoms per

let

511

penetrates

1(0) is the incident

from the

Consider a

foil.

tern

A beam

Figure 14-21.

of X-rays passing through a

foil.

in the lamina is N
where S is the area of the foil
in cm 2 The probability of a quantum being removed from the beam by
one of these atoms is equal to the probability of the quantum striking an
imaginary disk of area aa cm 2 This is equal to aa cm2 /5 cm 2 the ratio of

lamina of thickness dx at

The number of atoms

x.

times the volume of the lamina;

it is

NS dx,

the area of the disk to the area of the lamina.

The

net probability of a

quantum being removed from the beam by any of the atoms in the lamina
is equal to this ratio times the number of atoms; i.e., (ajS)NS dx = a Ndx.
a
As an average of I(x) quanta are incident upon 1 cm 2 of the lamina in 1 sec,
the average number of quanta removed per cm2 per sec from the beam by
the lamina is I(x)aa N dx. The number of quanta per cm 2 per sec emerging
from the lamina, I(x + dx), is equal to the number incident minus the
number removed so
;

I(x

dx)

I(x

dx)

I(x)

I(x)aa Ndx

or
dl(x)

I(x)

~I(x)aa Ndx

(14-48)


X-RAYS

512

We

find I(t)

by integrating

this expression

dl(x)

[Ch. 14

over x:

= aNdx

I(x)

Jo l(x)

[in /(*)]'

Jo

= -aa Nt

7(0

In

-oNt

1(0)

7(0)

7(0
2.0

1.8

1.6

1.4

/^

l(0)e-"

^^_

S*^

1.2

^0.8
b

**>/

0.4
0.2

\\

A1

n
10 5

io 6

The attenuation

io 7

hv
Figure 14-22.

V/v
^^
^^^-^1
\ \\
\

0.6

(14-49)

i. 1.0

10 8

10 s

(ev)

coefficients for several

atoms and

a range of energies.

The flux of the attenuated beam decreases exponentially with increasing


The quantity aa N, which is called the attenuation coefficient, has the
dimensions cm -1 and is the reciprocal of the thickness of foil required to
attenuate the beam by a factor e. This thickness is called the attenuation
t.

length A.

That

is

A=
Of

course, the attenuation coefficient has the

the total cross section.

of

\jaa

82

Pb,

50

Sn,

and

13

(14-50)

same energy dependence as

Figure (14-22) shows the attenuation coefficients

A1 for X-rays of

relatively high energy.

8.

AND OTHER

POSITRONS

Sec. 8]

ANTIPARTICLES

513

Positrons and Other Antiparticles

The

first

experimental evidence for the pair production process, and the


was obtained in 1933 by Anderson during an

existence of positrons,

investigation of the cosmic radiation.

This radiation consists of a flux


of very high energy X-rays and charged particles incident upon the earth
from extra-terrestrial sources. Anderson was using a cloud chamber
(a device filled with supersaturated

formed by charged

vapor that condenses on the ions

particles traversing

it

and makes the paths of these


a

particles visible) containing a thin lead plate, with the entire apparatus in

magnetic

field.

Upon

exposing this apparatus to the cosmic radiation,

it

was found that very infrequently a pair of charged particles were ejected
from some point in the lead plate. These events were assumed to be the
result of the interaction of an X-ray quantum in the lead because no charged
particle was seen to strike the point of ejection, whereas a quantum, being
uncharged, could strike the point of ejection without being seen. The two
charged particles ejected in these events were bent in opposite directions
by the magnetic field. Therefore their charges were of the opposite sign.

From other considerations it could be shown that the magnitudes of these


charges were equal to one electronic charge and that the masses of the
particles

were at

least

the discovery of these

approximately equal to one electronic mass. Since


few electron-positron pairs, countless others have

first

been observed in many different experiments. In these experiments


has been established with great accuracy that the rest mass, charge, spin,
symmetry character, and magnetic moment of the positron and electron
are equal, except that the signs of the charge and magnetic moment of the

it

positron are opposite those of the electron.

The discovery of the pair production process explained the origin of


a discrepancy between the then current theory of X-ray attenuation and
the measured attenuation coefficients of several materials for 2.6 Mev
X-rays

(gamma

rays obtained from a radioactive source).

originally did not include pair production,


coefficients

were too small;

As

the theory

the predicted attenuation

with the inclusion of the pair production

good agreement is now obtained between experiment and theory.


However, the real importance of Anderson's discovery was in the beautiful
confirmation which it provided for Dirac's relativistic theory of quantum
process,

mechanics.

The Dirac theory

leads to the prediction that the allowed values of

total relativistic energy

E for

a free electron are

E = Vc p z +
2

where

is

the electron rest mass.

(mc2f

(14-51)

These are simply the solutions for

X-RAYS

514

[Ch. 14

of equation (1-25). But the solution with the minus sign corresponds to
a negative total relativistic energy a concept just as foreign to relativistic

mechanics as a negative total energy is to classical mechanics. Instead


of just throwing away the negative part of equation (14-51) on the
grounds that it is not physically realistic, Dirac pursued the consequences
of the entire equation. In doing this he was led to some very interesting
conclusions. Consider figure (14-23), which is an energy level diagram

+mc 2

Higher

levels,

Lowest

level,

corresponding to

corresponding to

>

p =

Highest

Lower

Figure 14-23.

The energy

make

level,

levels,

corresponding to

corresponding to

p =

>

levels of a free electron according to Dirac.

representing equation (14-51).

energy levels

If the indicated

exists, all free electrons

transitions into these levels,

continuum of negative

of positive energy should be able to

accompanied by the emission of quanta

of the appropriate energies. This obviously disagrees with experiment


because free electrons are not generally observed to emit spontaneously

2mc 2 However, Dirac pointed out that this diffiquanta of energy hv


culty can be removed by assuming that all the negative energy levels are
normally occupied at all points in space. Then the exclusion principle
would prevent a free electron from dropping into any of these levels.
According to this assumption, a vacuum consists of a sea of electrons in
negative energy levels.

This does not disagree with experiment.

For

assumed

to be

instance, the negative charge could not be detected, as

it is

uniformly distributed and therefore exerts no force on a charged body.


Similar considerations will demonstrate to the reader that all the "usual"
properties of a sea of negative energy electrons are such that

would not be apparent

in

any of the "usual" experiments.

its

presence

However,

POSITRONS

Sec. 8]

Dirac's theory of the


certain

new

AND OTHER

ANTIPARTICLES

515

vacuum is not completely vacuous because it predicts

properties which can be tested

The energy

by experiment.
diagram for a free electron suggests the

level

possibility of

an electron in a negative energy level by the absorption of a


quantum. Since all the negative energy levels are assumed to be occupied,
the electron must be excited to one of the unoccupied positive energy
exciting

The minimum quantum energy required for this process


= 2mc 2 and the process results in the production of an

levels.

ously hv

in a positive energy level plus a hole in a negative energy level.


strate

below that a hole

obvi-

is

electron

We demon-

in a negative energy level has the properties of a

positron of positive energy.

Consequently this is the pair production


process observed experimentally by Anderson three years after its theoretical prediction by Dirac.t
Consider the excitation of an electron from a level of total relativistic
to a level of total relativistic energy
The energy of
1
2

energy
the

+E

quantum required

hv
If

we

= +E - (-EJ = E + E
2

energy

(14-52)

use the symbol e~ to designate the electron

its total relativistic

is

left in

the level

+E

2,

is

E.-

= E2

and, according to equations (1-22') and (1-24),

its

kinetic energy

T - = E - - mc 2
e

is

(14-53)

where m is its rest mass. These relations are illustrated in the energy
level diagram of figure (14-24). Now, there will be a positive total relativistic energy, which we call E +, associated with the absence of an electron
e
in the level of total relativistic energy

two quantities

Ev

The

relation

between these

is

E e+ = -(-,) = E x
From

figure (14-24)

we

E can be decomposed as follows:


E = mc T +

see that

which may be taken as a definition of the positive energy

T +. Thus
e

E e+ = -{-mc 2 - 7~= mc 2 + 7>


or

7>
t In section 6

we mentioned

that

= E e+ - mc

(14-54)

momentum

to take place in the presence of a nucleus.


affect the

conservation requires pair production


However, the massive nucleus does not

energy balance, so an energy level diagram for a free electron can be used to

describe the process.

516

X-RAYS

[Ch. 14

By comparing this equation with (14-53), we identify Te+ as a kinetic


and mc 2 as a positive rest mass energy. There are, then, a positive

energy,

mass

E +,

Te+,

and a

positive

associated with the hole in a negative energy level.

Conse-

total relativistic energy


rest

quently the hole has

all

a positive kinetic energy

the mechanical properties (rest mass, inertia, etc.)

of a particle of rest mass equal to one electronic mass.


there will be a positive charge

+e

apparent that

I i

It is

associated with the absence of an

Eh"

+
i

u
E

o
B
1
1

'

+
'
'

An energy

Figure 14-24.

electron of charge e.

magnetic
viously
is

the

moment

filled

same

level

'

diagram illustrating pair production.

There are also a spin angular

momentum and

associated with the absence of an electron in a pre-

energy

level.

The magnitude of the

spin angular

as for a level containing only a single electron

(cf.

momentum
footnote on

page 485). The magnetic moment is of the same magnitude as for an


electron, but the sign is opposite because the charge associated with the
hole

is

opposite that of an electron.

We

see that in all respects the pro-

perties associated with a hole in a negative energy level are identical with

the experimentally observed properties of a positron, which

we have

designated by the symbol e+.

A positron is a completely stable particle as long as

it is kept away from


However, a positron passing through matter will eventually
collide with an electron and the two particles will annihilate. In terms of
the Dirac theory, this can be thought of as a transition in which the electron
drops into the hole that constitutes the positron, with the liberated energy
emitted in the form of quanta. Annihilation normally occurs when the

electrons.

Ch.

EXERCISES

14]

517

is moving very slowly since then the probability of colliding with


an electron is higher. If annihilation occurs on a free electron, two quanta
must be emitted in opposite directions in order to conserve momentum.
If the electron is bound to a nucleus, one quantum annihilation is possible.

positron

The average time required

for a slow positron to annihilate in

82

Pb

is

~10-10 sec.
The

and positron are but one example of a particle and its


The existence of a number of other examples is predicted
by theory and has been verified by experiment. For instance, the antiproton, a particle of charge e and mass equal to the proton mass M,
was observed in 1955 by Segre and his co-workers. Since the proton rest
mass energy Mc 2 = 931 Mev is very large, antiprotons can be generated
only in the cosmic radiation or in high energy particle accelerators. The
electron

antiparticle.

Segre experiment produced antiprotons in collisions between stationary


protons and protons of 6200 Mev kinetic energy obtained from a large
particle accelerator.

BIBLIOGRAPHY
Compton, A. H., and

S. K. Allison, X-rays in Theory and Experiment, D. Van


Nostrand Co., Princeton, N.J., 1935.
Evans, R. D., The Atomic Nucleus, McGraw-Hill Book Co., New York, 1955.

Richtmyer, F. K., E. H. Kennard, and T. Lauritsen, Introduction to Modern


Physics, McGraw-Hill Book Co., New York, 1955.

EXERCISES
1.

Verify equation (14-24).

2. Verify
3.

equation (14-28).

Prove that a

stated in section
4.

free electron

cannot absorb

all

the energy of a quantum, as

6.

Prove that pair production cannot take place in free space, as stated in

section 6.
5.

Using the data of

to attenuate a
6.

Show

quantum

beam of

figure (14-20), calculate the thickness of


10 4 ev X-rays by a factor of 100.

that the attenuation length

is

Pb required

just equal to the average distance a

will travel before being scattered or absorbed.

7. A slow positron moving through matter can be captured by a free electron,


forming an atomic-like system of transient existence called positronium. Evaluate

the spectrum of electromagnetic radiation emitted in processes of this type.

CHAPTER

15
Collision

Theory

I.

Introduction

We

have been able to go through a

fairly

complete discussion of the

atom in terms of the quantum mechanical theory of the motion of particles


bound in a potential, because most of the understanding of atoms has come
from interpreting experimental measurements of the spectra which they
emit in terms of the energy levels of their bound atomic electrons. For
the nucleus this

is

not the case.

From the very first the understanding of its

come largely from interpreting the results of experiments


in which unbound particles collide with nuclei. As an example, the reader
will recall that the very discovery of the nucleus came from the alpha parproperties has

ticle

scattering experiments.

nucleus,

we must now extend

In preparation for our discussion of the

quantum mechanical
unbound particle with a potential.f
previous chapter we described the processes which occur in the
to three dimensions the

theory of the collision of an


In the

quanta with atoms. This description was


probably not very satisfying to the reader because we were generally able
to state only results. A rigorous treatment of the scattering and absorption
of quanta is not possible at the level of this book as it requires quantum
electrodynamics and relativistic quantum mechanics. In this chapter
we shall, however, consider only the scattering and absorption of particles
and shall ignore spin and other relativistic effects. This will be adequate for
most of our discussion of the nucleus and will allow us to treat the problem
rigorously with the "old familiar" Schroedinger theory. At first we shall

collision of electromagnetic

t Collisions in

one dimension were treated

518

in the first sections of

Chapter

8.

LABORATORY-CENTER OF MASS TRANSFORMATIONS

Sec. 2]

519

we shall give an approxiand then an exact treatment. Finally we

consider scattering. After certain preliminaries,

mate treatment of

this process,

shall treat absorption.

2.

Laboratory-Center of Mass Transformations


In working with the Schroedinger equation for a system of two particles

of mass

and

interacting with each other through a

mutual potential
problem can be reduced from one
involving six spatial coordinates to one involving three by transforming to
a frame of reference in which the center of mass of the system is at rest.
In this center of mass (CM) frame the system can be described in terms of
the motion of a single fictitious particle of reduced mass
1

energy V(r),

we have found

that the

A*

-^^-

(15-1)

ml + m 2

interacting with a center of force fixed at the origin.

The

distance

this particle to the origin is equal to the separation of the

from

particles

x and
2 and the potential energy V(r) of the fictitious particle
equal to the mutual potential energy of the two particles. Furthermore,

of mass
is

two

the total (and kinetic) energy of the fictitious particle in the

CM frame is

equal to the total (and kinetic) energy in that frame of the system of two
particles.f In the
frame the system obtained by replacing one of the

CM

by a particle of mass (x, replacing the other by a particle of such


large mass that it must remain fixed at the origin, but otherwise making no
particles

changes in the system or the equations governing


equivalent to the original system.

work with because one of

its

The

its

is

is

completely

much

easier to

particles remains fixed and, consequently,

can be ignored except as a source of the potential

Now,

behavior,

equivalent system

V(r).

we have never
had to bother with the fact that the
frame is generally moving, since
this motion only adds a constant and non-quantized translational kinetic
energy to the system which does not affect the energy differences that are
observed in measurements of spectra. But, in treating scattering problems,
we cannot ignore the motion of the CM frame because it affects the
observed scattering angles and cross sections. In almost all experiments
these quantities are measured in a frame of reference, called the laboratory
(LAB) frame, which is moving with respect to the CM frame. The reason
is that in almost all scattering experiments one of the particles, and not
in evaluating the energy levels of such a system,

CM

t Cf. section 2,

Chapter

10,

and section

4,

Chapter

5.

COLLISION THEORY

520
the center of mass,

[Ch. I5_

stationary in the laboratory. Let us develop

is initially

the relations between the description of a scattering process in these

frames.

we

In doing this

shall deal

with

momentum

two

vectors of precise

magnitudes and directions however, since the locations of the particles


need not be specified, this will not violate the uncertainty principle.
;

Assume

that the particle of

mass

2 is

initially stationary in the

LAB

of m 1 with respect to m 2 is \ 1L as shown


in figure (15-1). After the collision, m 1 moves off with velocity \'1L at the
frame, and that the

initial velocity

V 1L

Before

Figure 15-1.

LAB
and

between two

collision

particles, as

seen

in

the

LAB

frame.

scattering angle 6 L measured from its initial direction of motion,


In the
frame
2 recoils in some direction with some velocity.

CM

has the different appearance shown in figure (15-2), because


frame is moving with velocity V with respect to the LAB frame.

this process

the

CM

By

differentiating equations (5-16) with respect to time,

and multiplying

each of the resulting velocities by the mass of the associated

CM

frame

mj

vie

particle,

it is

V2c "*2

Before

Figure 15-2.

between two

collision

particles, as

seen

in

the

CM

frame.

CM

frame the total momentum must vanish.f


and w 2 must therefore always be equal and
opposite in this frame. Furthermore, the magnitudes of the momenta
cannot change because energy is conserved. Therefore the result of a
collision is simply to rotate the line along which these vectors are directed
scattering angle 6 C without changing their magnitudes.
through the
easy to see that in the

The momentum

vectors of

CM

Thus

v ic
t This

is

ic

v zc

also true relativistically. Cf. exercise

8,

(15-2)

V 2C

Chapter

1.

LABORATORY-CENTER OF MASS TRANSFORMATIONS

Sec. 2]

An

521

CM

frame is obtained by
by fi, with the initial velocity of fi equal to the relative
velocity of the two particles (because the distance from fi to the origin is
always equal to the distance between m 1 and m), and by letting m 2 -> oo.
equivalent description of the process in the

replacing

This

is

We

indicated in figure (15-3).

v*c

and

it is

have
(15-3)

v il

apparent that

(1^4)

CM

frame

"

v" c

\_

Before

Figure 15-3.

the

CM

To
to the

An

After

equivalent description of a collision between

two

CM

determine the value of V, the velocity of the


frame with respect
LAB frame, we use the classical velocity transformation equations

Vic

v il

~ v

= V

v ic

>

Their use, and the use of velocity independent masses,

we

particles, as seen in

frame.

are not treating relativistic problems.

momentum

vanish in the

CM frame

(15-5)
is

justified

is

m lVlc = m 2V2C
This

because

Now the condition that the total


(15_6)

is

m l vlL = ( m l + m d V
or

V=
To

mi

mx + m 2

find the relation between 6 C

after the collision

form of the

from the

is

(15-7)

and 6L we transform the velocity of


The vectorial
,

CM frame to the LAB frame.

classical velocity transformation

ii = V +
This

v 1L

illustrated in figure (15-4).

equation which

we

use

is

(15-8)

vi c

Taking transverse components, we have

v il sin e L

v ic sin

0c

(15-9)

Longitudinal components give


v'1L

cos 6 L

= V + V1C cos 6 C

(15-10)

COLLISION THEORY

522

[Ch. 15

Dividing equation (15-9) by (15-10), we find


y,V sin

v'ic

sin 6 r

cos

^/"ic

cos 6c

But
1C

or

W"ic

Figure 15-4.

the

LAB

=m

lm 2

Illustrating the transformation of a velocity

vector from the

CM

frame to

frame.

Therefore
sin d c

tan0,

mjrriz

Now

let

(15-11)

cos 6 C

us find the relations between the scattering cross sections and

two frames of reference.


and of the incident and scattered
fluxes, may be directly adapted from equations (14-40) and (14-41) and
the definitions immediately preceding those equations, if we rewrite them
differential

The

scattering cross

sections for the

definitions of the cross sections,

in terms of a single incident particle instead of

That

is,

we

the incident particle will be carried by the


to the

many

incident quanta.

define the incident flux / to be the probability per second that

beam

direction,

and

beam across a

define the scattered fluxes

cm2

S and

area normal
S(0) to be,

respectively, the probability per second that the incident particle will be

and the probability per second that it will be


an angle 6 from the direction of the
quantities
are
related by the scattering cross section
beam.
These
incident
a and the differential scattering cross section dajdD. as follows
scattered into

all

angles,

scattered into a unit solid angle at

al

(15-12)

(15-13)
ail

Sec. 2]

LABORATORY-CENTER OF MASS TRANSFORMATIONS

523

In the present context we simplify the notation by dropping the bar


denoting average, and the subscript S for scattering. We also drop the
subscript a denoting cross sections per atom, as it is no longer necessarily
appropriate.

However, we must add a subscript

C for center

of mass, or

for laboratory, to all these quantities in order to designate the frame of


reference in which they are defined.
do not need to use some other

We

symbol to distinguish between the quantities defined in the two alternative


frame descriptions of the scattering process, since henceforth we shall

CM

be concerned only with the description of a particle of mass


n scattering
from the potential V(r) associated with the center of force fixed at the
origin.

From

the Schroedinger theory

cross sections for this description in the

we

shall calculate the scattering

CM frame.

These cross sections


can then be transformed to the LAB frame, for comparison with experiment, by noting that the probability per second for the incident particle
to be scattered must be the same whether measured in the
frame or the
LAB frame. That is,

CM

Using equation (15-12),

this

becomes
a L IL

a C IC

II

Ic

But
(15-14)

because the motion of the particle /j, relative to the origin before scattering
(which determines Ic ) is identical with the motion of the particle m
x
relative to the particle m 2 before scattering (which determines I ). ThereL
fore

oc

(15-15)

Similarly, the probability per second for the incident particle to be scattered

into the solid angle dD. L

dQ. c measured in the

LAB frame must be the same as


be scattered into the solid angle

measured in the

the probability per second that

it

will

CM frame, if the same physical range of directions

described by both solid angles.

That

S L (8 L) dQL

is,

= S C (6 C) dQ c

so

dtt L

L
dl c

and

da L

dac dQ.n

dl L

dQ. c dQ. L

is

COLLISION THEORY

524

[Ch. 15

This can also be obtained by formally differentiating equation (15-15).


The elements of solid angle have been evaluated in equation (14-25).

They are
dQ. c

dd c

2tt sin d

and

dQ L =

L dd L

2tt sin

so

d&c

sin6 c d0 o

dCl L

As an

sin 6 L

(15-161

dd L

should use (15-11) to evaluate (15-16), and

exercise, the reader

obtain

da L

dQ L

3.

doc

2(m 1 /m 2) cos dgf4


(m 1 /m 2 ) cos C

+ (mjmtf +

[1

dQ c

n 5 _ 17 )

The Born Approximation


In this section

we

shall

develop a method, due to Born, for calculating

the approximate values of dajdD.

and

(The subscript

a.

C is

to be under-

stood because here, and throughout the remainder of this chapter,


sections are defined in the

CM frame.)

The

first

step

is

to give a

all

cross

quantum

mechanical description for the motion of a particle of the incident beam.


Equation (8-4), and the subsequent discussion, show that in one dimension

ah eigenfunction describing a

mass n traveling with velocity


which we call here the z axis, is

free particle of

v in the positive direction along

an

y>(z)

axis,

AeiKz

(15-18)

where

K=

IwjX

lirplh

(15-18')

nv\h

and where A is a constant. The reader may show by substitution that the
wave eigenfunction (15-18) is also a solution to the three dimensional time independent Schroedinger equation for a free particle,

traveling

_^!
1(i

.dx

dV =

<Py)

ip
2

E%p

(15-19)

a?.

~dy~

where

E = />2 /2^ =

h 2 K 2 l2/j,

(15-19')

In three dimensions, equation (15-18) describes a particle which is definitely

known

to be

moving

parallel to the z axis with velocity

coordinates are entirely

of x and

y,

and whose

unknown since rp*{z) ip{z) is

z coordinate is also entirely

y>*(z) y>(z)

A*e-

iKz

iKz

Ae

v,

whose x and y

obviously independent

unknown

A*

since

(15-20)

Sec. 3]

THE BORN APPROXIMATION

Thus the

particle is

525

moving somewhere in a beam, parallel to the z axis,


and longitudinal dimensions. Of course, this is
not physically realistic since all beams are always limited in their transverse dimensions by diaphragms of finite aperture and in their longitudinal
dimensions by the finite length of the apparatus. On the other hand, the
dimensions of real beams are extremely large compared to the characteristic
atomic or nuclear dimensions. Therefore equation (15-18) provides an
of

infinite

transverse

accurate description of the incident particle in the region of importance


where the atomic or nuclear potential which produces the scattering has
any appreciable value.

Figure 155.

The

spatial

dependence of the

real part of an eigenfunction in

box normali-

zation with periodic boundary conditions.

The

unrealistic aspects of equation (15-18) are, however, the origin of

certain problems concerning the normalization of the eigenfunction.

In

Chapter 8, we showed that these problems can always be handled


and can usually be ignored. The present calculation provides an example
of a case in which they cannot be ignored we must use a three dimensional
extension of the technique of box normalization with periodic boundary
section

1,

conditions.

where

is

We

set

A =

(15-21)

the edge length of a very large cubical box surrounding the

region of the scattering potential,


variables to

Ir**

lie

within the box.

and we restrict the range of the spatial


Then the eigenfunction is normalized

because
ip*ip

A* A

= A2 = L

(15-22)

and
ip*ip

dr

=L

\dr

IT 31*

where dris the volume element, and where the integration is now taken only
over the volume of the box. We furthermore demand that the eigenfunction
and its spatial derivative in the direction normal to the wall have the same
values at corresponding points of the opposing walls of the box. The real
(or imaginary) part of y> would then typically have the behavior plotted in
figure (15-5) as a function of

one of the

spatial variables, holding the other

COLLISION THEORY

526

[Ch. IS

two constant. Using periodic boundary conditions, the eigenfunction


be completely periodic, with period L, in

all

three directions.

Its

will

behavior

repeats indefinitely in the adjacent boxes (just as a scene observed within


a cube with mirror walls repeats indefinitely), and we are justified in considering what happens only within a single box that is, in restricting

the ranges of the variables to the box.f

In most cases of physical interest the scattering potential

symmetrical function V(r).


necessary restriction.

It

We assume this
is

the rectangular coordinates x, y,

where
which

K
is

is

Lr ZA eiKz

beam can be

L- 3AeiKrcose

K directed

a vector of magnitude

the direction of the z axis, and where r

to the point

(r,

d,

<f>).

free particle traveling in

is

r,

6,

<j>

instead of

z axis.

This means

written

Z,-e iK-r

(15-23)

along the beam direction,


is a vector from the origin

In this form the normalized eigenfunction for a

some other
y,

where K'

not a

cos

for a free particle of the incident

it is

Define the origin to be at the center

z.

of the potential and the polar axis to be along the

y>

a spherically

then obviously convenient to describe the

incident particle in terms of the spherical coordinates

and f

is

to be true, although

direction can be written

z,-?V K

''

(15-24)

a vector in the direction in question of magnitude equal to the

value of K' appropriate to the reduced mass and velocity of the particle.

The

validity of equation (15-24)

can be

verified

by the same arguments

as were used for (15-23).

Now

consider a particle in the incident

potential V(r).

We

want

the particle will be scattered in

can treat

this as

beam impinging upon

the

to calculate the probability per second that

some direction.

a perturbation problem:

If V(r) is not too strong,

What

is

we

the rate at which a

constant (in time) perturbation V(r) induces transitions from the

initial

eigenstate associated with the free particle eigenfunction (15-23) to a


final

eigenstate associated with a free particle eigenfunction (15-24)?

Since the final eigenstate

is

in

an

essentially continuous range of final

eigenstates because the possible eigenvalues E'

fr

K'2 /2[i are almost

t Another consideration, which serves to justify box normalization with periodic


boundary conditions, is that the set of eigenfunctions satisfying these conditions forms
a complete set of orthogonal functions in terms of which almost any function can be
expanded. Cf. section 1, Chapter 8.

Sec. 3]

THE BORN APPROXIMATION

527

continuously distributed even with box normalization, the answer is


given, approximately, by the three dimensional extension of Golden Rule

No. 2 (equation 9-45). This

is

2n

R K-~^

(15-25)

^'K^K'K/OK'

where we have used the vectors K and K', instead of quantum numbers,
to label the initial and final states, and where v
K K is the matrix element of
the potential taken between these states. That is,
.

k-k

= jiL-X/r-'rWirWdT =

- '>L- 3 jV(ry (K K r rfT

= L~ VKK
3

(15-25')

with
FK' K

=jV(ry'(K K

'

d7

The quantity p K of (1 5-25) is the number of possible eigenstates per unit


energy interval for the particle associated with the final eigenfunction.
As we have employed box normalization with periodic boundary condi,

tions (required because the eigenfunctions appearing in v


K

K must be

normalized), the density of final states p


K will have some finite value
since the boundary conditions impose restrictions on the possible
de Broglie wavelengths. As an example, consider K' parallel to one edge of
,

Then the real (or imaginary) part of y would typically have the
appearance shown in figure (15-5), with the distance d equal to the
de Broglie wavelength X = 2-n\K' The periodic boundary conditions can be
the box.

propagation parallel to one edge of the box only if L contains


exactly an integral number of wavelengths of the traveling waves. Compare this with the case of free particle standing wave eigenfunctions in a
box with impenetrable walls, which we treated in section 6, Chapter 12.
In that case, the boundary conditions demand that y> have nodes at the
satisfied for

For the propagation direction parallel to one edge of


the box, the condition can be satisfied if L contains either an integral
number of wavelengths or a half-integral number of wavelengths. Conwalls of the box.

sequently, in every wavelength or energy interval there are

two times as
allowed wavelengths in the standing wave case as there are in the
traveling wave case. However, for each possible wavelength there are two
separate traveling waves, one propagating in one direction and another
propagating in the opposite direction. The factors of two cancel out, not

many

only for propagation in directions parallel to the edges of the box but also
and the number of possible eigenstates

for propagation in all directions,

COLLISION THEORY

528

per unit energy interval

therefore the

is

same

in

both

use equation (12-37) which, in our present notation,

dE'

cases.

[Ch. 15

Thus we may

is

u 3A T 3 F' 1A dF'

f*

(15-26)

,-f

where

I?

volume of the box

the

V,

and
E'

h 2 K' 2 l2p

Therefore

^L

PK
This

is

'

not quite what we want because

we want p K
within some

ciated with K', whereas

when

hK'

~ 2V/i3 2

that vector lies

>,

f,L*K'

it is

2-n*P

the density of

all states

asso"

the density of states associated with

Now

certain range of directions.

it is

clear that for a spherically symmetrical potential V(r) the scattering angular

distribution will not

depend on the azimuthal angle

appropriate to consider together

all final states

<j>.

Consequently,

it is

associated with vectors

K' whose directions lie anywhere within the angular range 6 to 6 + dd.
The density p K of these states is smaller than p K by a factor equal to
the ratio of the solid angle dQ.

8 to 6
6.

That

2tt sin 6

d6 contained within the range

d6 to the total solid angle 4n contained within the entire range of


is,

Pk'

dQ.

Pk-

^
477

so

Pr

_ fiL K'
=
S7T

Using

this in (15-25),

we have

R K =-

(15-28)

ti

^k l-Kk^k^F
3

877

Now let us

( 15

- 29

/P

calculate the probability per second that the particle in the

initial eigenstate

associated with the vector

normal to the direction of K. This

is

will cross

the incident flux

/,

and

cm2
it

area

can be

evaluated as in the steps preceding equation (8-7) by using the three

dimensional extension of the definition (7-31) of probability flux. Alternatively, we can use (8-8) and equate / to the product of the velocity of
the particle and the probability density. That

I=vy>*y>

is,

(15-30)

THE BORN APPROXIMATION

Sec. 3]

With equations

(15-18')

and (15-22),

Next, divide

RK

becomes

this

Kb

L~

529

(15-31)

by the element of solid angle dQ, to obtain the rate of

transitions per unit solid angle into the final states associated with the

vector K'.

Then we have

the probability per second of scattering into a

unit solid angle at the angle

which

d,

S(d), the scattered flux.

is

S(0)~^^Vk k Fk K
*.

According to equation (15-13), the

da

S(6)_

dQ
But K'

pv'jh

fxv\h

juL~

KhL~

47T

because

K,

(15-32)

differential scattering cross section is

ju

Thus

K'

KK KK

2
/*

equation

(15-4)

shows

v'

v.

Therefore

where
,

ilK - K yt
'

^K K

! V(r)e

dr

(15-33')

with the integration taken over a very large box surrounding the scattering
This is the Born approximation for dafdQ.. Note that the size

potential.

of the box has dropped out since

does not appear in (15-33), and since

come only from the small


region in which V(r) has any appreciable value and therefore the value of
the integral is independent of its limits.t
contributions to the integral in (15-33') will

It is possible to

carry out part of the integration of (15-33') immediately.

Set

X = K

K'

(15-34)

define a set of spherical coordinates r, 0, O with an origin at the center


of the potential and polar axis along the direction of %. (They should not

and

be confused with the spherical coordinates


along the direction of K.) Then

(K

K')

=X

r,

6,

%r cos

and
dr
t

sin

We use this fact in equation (15-35).

dr

d@ dO

<f>

whose polar

axis lies

COLLISION THEORY

530

[Ch. 15

so

>K'K

V(r)e ixrcos

dr

sin

d@

(15-35)

d<D

Jo Jo Jo

VKK =

V(ry* ,C0B

2nr 2

dr

sin

d0

Jo Jo

The

integral can be evaluated

iyj cos

0. The result

by making the change of variable

is

_r no
Jo

^k-k=
K'K
which

Z=

2nr 2 dr

(15-36)

'X r

is

Vkk

K(r)

^^47rr

Figure 15-6.

dr

(15-37)

Jo

xr

illustrating the relation

between the vectors which enter

in

the Born

approximation.

% in terms of the scattering angle

Finally, let us express

6.

Consider the

vector diagram of figure (15-6), which illustrates the relation (15-34).

From

this figure

apparent that

it is

Now

compare the

= 2K sin ((9/2)

results of

(15-38)

our calculation, (15-33), (15-37), and

(15-38), with the equation (14-35) for the differential cross section for

The similarity is no accident


an alternative derivation of the Born approximation
which leads to exactly the same results as we have obtained, but which is
coherent scattering of X-rays by an atom.

because there

is

completely analogous to the derivation of (14-35). This derivation


involves calculating the amplitude of the wave scattered in a given direction

by every volume element in the region of the


phases into account,

added, and the result

4.

Some

all
is

potential.

Taking the

relative

these contributions to the total amplitude are

squared to obtain the total scattering intensity.

Applications of the Born Approximation

Let us evaluate dajdQ. in the Born approximation for two illustrative


cases.

Consider an attractive Gaussian potential


V(r)

= - F e-

(r/is >

(15-39)

SOME APPLICATIONS OF THE BORN APPROXIMATION

Sec. 4]

which

is

illustrated in figure (15-7).

531

We have

= -V

Vkk.

Upon

f
Jo

'e-W*?^4Trr 2 dr
%r

integration this gives

VKK =

A V R 3 e- lA

-{27rf

-v.

Figure 15-7.

An

attractive Gaussian potential.

Substituting into equation (15-33),

we

obtain

da^jjtJ {2n?VlR*e-* R
\2

da

\2ttH

'

or
(15-40)

US'

Figure

15-8.

The

h4

differential

\R

scattering

cross

section

an

for

attractive

Gaussian

potential.

The differential
This

scattering cross section has the

a Gaussian-like function with a


-1
e
times its maximum value at an angle
is

4K 2 R2

form shown in figure

maximum

at 6

0' satisfying

sin2 (6' 12)

0.

(15-8).

It falls

the relation

to

COLLISION THEORY

532

For energies high enough that

KR =

(V'2i,E\K)R

>

1, 0' will

[Ch. 15

be small and

we have

AK2R\d'j2f

or
d'

(15-41)

KR

-V

Figure

An

59.

attractive square well potential.

Consider next an attractive square well potential,

V(r)

-V

r<R

(15-42)

>R

0,

Here

illustrated in figure (15-9).

sin vr

4-

Airr*

dr

Xr

and we obtain upon integration

(xRfi

So

da^,

dQ

V \
\2ttW

16jt2j/2q6

[si n

X*

XR cos xRf

'-*
(xRf

or

{sin

[2KR

sin (0/2)]

- 2KR sin (6/2) cos


[2KR

sin (0/2)]

\2KR

sin (0/2)] }

(15-43)

The form of

this

differential

scattering cross section

figure (15-10).

At

but [ sin

0,

%R

X*

is

indicated

- X R c os %Rfl(x RT =

in
1/9.

SOME APPLICATIONS OF THE BORN APPROXIMATION

Sec. 4]

Consequently da/dQ. has a


increasing angle, reaching

has

its first

maximum

finite

non-zero root. This

at 6

when

zero

first

its

sin

=
%R

0.

It

533

drops with

%R cos %R =

is

%R

4.49

or

2KR sin (0'/2) =

4.49

^-Vb 6 1

Figure 15-10.

The

differential scattering cross section for an attractive

square well

potential.

For high
da/dQ is

energies,

KR >

1, 6'

< 1,

and the value of

6 at the first zero of

4 49
0'~

(15-44)

KR

For a sharp edge

potential, such as (15-42), da/dQ. has the characteristic

behavior of an optical diffraction pattern.

For a potential with a smooth

edge, such as (15-39), the similarity between da/dQ, and an optical diffraction pattern

is,

perhaps, not so obvious because the secondary

are not present (or at least are weak).

wave

maxima

Nevertheless, the scattering of a

function, in passing through a small region in which the potential

changes in any way,

is

completely analogous to the scattering of an

electromagnetic wave, in passing through a small region in which the index

of refraction changes in an equivalent way


cases an angular distribution

concentrated at angles

less

is

(cf.

equation 8-56).

In both

obtained with most of the scattered intensity

than
0,

L = J_A
~ KR
2ttR

This angle decreases with increasing

K (increasing energy of the particle

or increasing frequency of the electromagnetic quanta), and the angular


distribution

The

becomes more strongly peaked forward.


two potentials we have considered

scattering cross sections for the

COLLISION THEORY

534

can be evaluated from their

scattering cross

differential

[Ch. 15

sections

by

calculating

(15-45)

dl

where the

integral

is

taken over

da

all solid

angle.

(This obvious equality

follows from the definitions of the quantities involved.)

We

shall

not

actually carry this out because the important characteristics of the scattering cross sections are easy to see qualitatively.

For both

potentials,

a decreases with increasing K because the angular region in which dajdQ.


has an appreciable value becomes smaller.
In closing,

we must discuss the range of applicability of the Born


The condition of validity of the perturbation theory

approximation.

underlying the approximation

is

(9-36),

which

states that the

amplitude

of the wave function for the scattered particle is small compared to the
amplitude of the wave function for the incident particle. It is also necessary that the free particle wave function for (15-23) be a reasonable

wave function and that the free particle


wave function for (15-24) be a reasonable representation of the scattered
wave function, in the region of the scattering potential where VKK is
representation of the incident

evaluated.

These conditions

incident particle

is

potential, that

if

is,

large

will generally

compared

be true if the energy E of the


magnitude of the scattering

to the

E>\V(r)\,

(15-46)

for all r

is a small perturbation which can


However, (15-46) is neither necessary
nor sufficient. For instance, the Born approximation turns out to be
fairly accurate for the scattering of low energy electrons from atoms for
reasons to be discussed later, and the Born approximation is not accurate

because then the scattering potential


usually produce only a small effect.

in certain cases for the scattering of high energy particles.


to give criteria

more

useful than (15-46), but this

is

It is

possible

complicated and will

not be done here.

5.

Partial

Wave Analysis

In the previous section we developed an approximation method useful


mainly for scattering at high energies. In this section we shall develop a

method which is exact at all energies, but useful mainly for scattering at
low energies. This is the method ofpartial wave analysis, originally applied
in the nineteenth century by Rayleigh to the scattering of sound waves,

PARTIAL

Sec. 5]

WAVE

ANALYSIS

535

and adapted

in 1927 by Faxen and Holtsmark to the scattering of quantum


mechanical waves.

Consider the scattering, by a potential

beam which

V(r),

of the wave function

upon the potential.


The incident wave passes over the potential, and scattered from it is an
outward moving wave. In the immediate vicinity of the potential both
the incident and scattered waves generally are strongly distorted by the
associated with a particle in a

is

incident

O
Figure 15-11.

Illustrating incident

and scattered waves as seen on a scale large compared

to the size of the scattering potential.

potential.

make

In fact, the waves

may become

so tangled that

it is difficult

to

a distinction between the incident and scattered waves.

However,
at large distances from the region in which the potential has any appreciable value these complications are no longer present, and the scattering
has the simple appearance represented in figure (15-11) in terms of a set
of nodal surfaces. If we do not look too closely, we see a plane wave
incident
it.

An

upon

the potential

and a spherical wave scattered outward from

eigenfunction for this can be written

iKx

00

(15^7)

where we have introduced spherical coordinates r, 6,


with origin at the
center of V(r), and z axis in the propagation direction of the incident wave.
The first term is the eigenfunction (15-18) for a plane wave moving in the
direction of increasing z, which represents the incident particle. (Here it
(f>

not be necessary to normalize the eigenfunction.) The second term is


an eigenfunction for a spherical wave moving in the direction of increasing
r, which represents the possible ways in which the scattered particle can
move. The directional dependence of its amplitude is given by f(d);
there can be no
dependence because of the symmetry of the scattering
potential. The 1/V provides for the necessary inverse square law decrease
in the intensity of the scattered wave. Later we shall verify that the second
term is a solution to the time independent Schroedinger equation in the
region r -> oo where V{r) = 0. For the present we shall assume this and
will

<t>

COLLISION THEORY

536

[Ch. 15

evaluate the incident flux, the scattered flux, and the differential scattering
cross section for the eigenfunction (15-47).

In evaluating the fluxes,

we may treat the two terms

of (1 5-47) separately

between the incident and scattered waves


in the regions where the fluxes are measured. This is true because the
incident flux is measured at large negative values of z where the second
term is negligible owing to its 1/r dependence, and because the scattered
flux is measured outside the transverse limits of the incident beam (which
since there are

no

interferences

Thus the incident

are always finite in a real scattering experiment).

is,

where
bility

to

its

hKjfx

is

v(e

iKz
iKz
)*(e
)

ve- iK2eiKz

(15-48)

This

the velocity of the incident particle.

is

the proba-

2
per second that the incident particle crosses a 1 cm area normal
direction of motion. Similarly, the probability per second that the

scattered particle crosses a

m
e

cm2

area normal to

iKr

iKr

its

direction of

r J

motion

is

f(6)\=vf*(d)f(d)

The

flux

in analogy to equation (15-30),

probability per second that

it

crosses

dA cm2

is

dA

But dA/r 2

is

equal,

by

definition, to the

element of solid angle dD., so the

probability per second for scattering into the solid angle dD.

is

vf*{6)f{6)dD
Dividing by dD,

we have

unit solid angle.

This

is

the probability per second for scattering into

S(d), the scattered flux.

S(d)

The

Thus

vf*(6)f{d)

scattered flux can also be evaluated

(15-49)

from the three dimensional extenmethod we have

sion of the definition (7-31) of probability flux, but the

used

is

simpler.

Now
dajdD

so

S(6)II

we have
do
dQ.

vr(8)f(&)

= p{Q)m

(15 _ 50 )

and the calculation of da\dO. is seen to be a matter of calculating f(6).


This is done by matching the incident wave with the scattered wave in
such a

way

as to obtain a solution to the Schroedinger equation, with the

PARTIAL

Sec. 5]

WAVE

scattering potential

preliminary step,

ANALYSIS

V(r), that

we must

537

behaves

find out

first

As a

like (15-47) for r->-oo.

how to decompose

the plane

wave

described by
e

and associated with the

linear

iK

iKr<x*e

momentum

(15-51)

Kh, into a

set

each associated with the orbital angular momentum Vl(l

do

of partial waves,

1) h.

We must

because the Schroedinger equation for a spherically symmetrical


potential yields most easily a set of solutions of definite orbital angular
this

momentum.

Now the

free particle eigenfunction (15-51) is a solution to

the time independent Schroedinger equation (10-15) for the case V(r)

We

have seen that solutions to


y(r, 0,

But, since (15-51) has no

(f>

<f>)

R(r) 0(0) <D(0

dependence,

solutions which are independent of

equation (10-17) for

m = 0,

which

0.

equation have the general form

this

is

it

suffices here to consider

Therefore

<f>.

(<f>)

(15-52)

1.

only

we take the solution to


Then 0(0) is an accept-

able (never divergent) solution to equation (10-23), with the condition

m=

These are the functions 0, o(0). They are called


0.
Legendre polynomials and are written Pj(cos 0). The first few are
(10-24), for

(cos 0)

P^cos

0)

=
=

cos
2

_
P2(cos 6) =

3 cos

_
.
=
P3 (cos 0)

5 cos

0-1
(15-53)

3 cos

The reader should compare these functions with the


dependence of
The Legendre polynomials are normalized to
(10-39) for m = 0.
2/(2/ + 1). This is represented by the first of the equations

cos0=l

fcose=-i

Pi-(cos 0) P,(cos 0) d(cos 0)

97

_l_

,^

(15-54)

/ft

0>

The second equation represents the orthogonality of these functions.


The function R(r) is an acceptable solution to equation (10-27) with
V(r)

0.

That equation

is

H("S) + (-^^}

COLLISION THEORY

538

[Ch. 15

E > 0, so E is not quantized and the quantum number


These solutions are written j\(Kr) and are called spherical
Bessel functions. The first few are listed below.
In the present case

n does not

arise.

UKr)
h(Kr)

sin

Kr

Kr
sin

Kr

cos

Kr

(Krf

Kr
(15-55)

h(Kr)

sin

.(Krf

J s (Kr)

cos

Kr

(Krf

15

(Krf

verify that for r

15

Kr

sin

_(Krf

The reader can

Kr

Kr

.(Krf

J_" cos

Kr

they have the same

Kr

dependence as

(10-45), in agreement with the remarks following (13-24). We shall


Finally,
verify their behavior for r -> oo in the following paragraph.
we see that for a free particle a particular one of the solutions (15-52)

can be written
(15-56)

;,(*>) itfcosfl)

This solution

is

associated with a definite orbital angular

the origin of coordinates, of magnitude Vl(l


in the

form (15-52)

for this case

1) fi.f

about

A general solution

is

aJiKr)

where the a are arbitrary constants.


t

free particle eigenfunction

momentum

P,(cos 6)

Now

(15-57)

the particle described by the

(15-51) has the definite linear

momentum

Kh. But it can have any orbital angular momentum about the origin since
its impact parameter with respect to the origin is completely unknown.
Therefore we certainly cannot describe the particle by a single one of the
eigenfunctions (15-56). However, we can describe it by the linear combination (15-57), since this contains all possible orbital angular momenta,
a general solution to the Schroedinger equation for the case
of a free particle. Thus we can write

and

since

it is

rco.

flJ,(K>) P,(cos 0)

1=0
t

The reason

is

that

it

satisfies the Z, p

eigenvalue equation (10-57).

(15-58)

WAVE

PARTIAL

Sec. 5]

To determine

P r (cos

the a lt

ANALYSIS

we

539

multiply both sides of this

and integrate over the

d) (/(cos 6)

P,.(cos 6)e iKrmse d(cos 6)


J -I

= f
=

aJ^Kr)

According to (15-54),

upon

this picks

P,.(cos 6)

P^cos

0) d(cos 6)

J-l

yielding,

equation by

entire range of cos 6

sum

out of the

the single term for

/',

transposition,

21'

a rj r (Kr)

P,.(cos 6)e

iKr cos '

d(cos 6)

dv

Integrating by parts,

we have

J-K.T cos

21'

a vJv(Kr)

P,.(cos 6)

0'

iKr

-l

ir cos 6

/.

dP r (cos

iKr

6)

d(cos 6)

d(cosd)
du

Integrating by parts again,

21'

avJviKr)

we

obtain
J.K.T cos 6'

'dP^cos

iY(cosfl)-

iKr

d) e

d(cosd)

l
iKr
cos 0'

(iKrf J-i
d(cos 6)

J-i (iKr)
(iKrf2

d(o
d(cos

6f

In this form the determination of a can easily be effected by taking


v
r -* co. Since the a are constants independent of r, the values obtained
l

For r - co, the last two terms of the equation


above will be negligibly small compared to the first term, since they are
both proportional to r~ 2 while the first term is proportional to r~K Neglectwill

be valid everywhere.

ing these terms,

2V

To

we have

a UlK
vj v (Kr)

P .(cosd)eiKrcose

iKr

evaluate this expression,

P,.(cos 6)

which can be

verified

we use

-> co

(15-59)

COS0= -1

the relations

8=1

1,

cos

(-If,

COS0

= -1

by comparison with equations (15-53). Sincef


(-1)''

t This relation

depends upon the equality e"

iv "

cos z

sin

z.

COLLISION THEORY

540

[Ch. 15

equation (15-59) gives

21'

a vj v (Kr)

[e

iKr

iKr

il

e '"e-

-'ir/2

,i(Kr-lV/2)

r-+co

],

iKr

-l(Kr-Vi,im

_> oO

Butf
(15-60)

andf

Ur - )

sin

(15-61)

2i

Therefore this can be written


2
21'

a vj v (Kr)


Kr

v sin(jK>
-

2i'

is

consistent with the behavior as r -* oo of

Note that equation

(1

5-62)

theyV(^") listed in (15-55) since, for


.

Jv(Kr)

Z'tt/2)

sin(Kr
^

-+ co

(15-62)

these functions,

all

J'77/2)

'-

-+ oo

(15-63)

Kr

by neglecting the terms of (15-55) which vanish more


and writing sin {Kr /V/2) = sin {Kr) cos {I'-njl)
(/V/2). Finally, we substitute (15-63) into (15-62) and solve

as can be verified
-1

rapidly than r

cos (Kr) sin


for a v

We

obtain
av

This

true for all

is

Consequently we

and

r.

It is also

may

use

it

(2V

'

l)i

true for all values of the

quantum number.

to evaluate the constants in equation (15-58)

find
00

jKrcose

This

is

= ^ (21 +
1=0

i)i jJ(Kr)

P (cos

the desired decomposition of the plane

(15-64)

6)

wave associated with the

momentum Kr into a set of partial waves,


orbital angular momentum Vl(l + 1) h. The

each associated with

linear

terms j\(Kr) P^cos 6)


specify the radial and angular dependence of the /th partial wave. The
plane wave elkrcos6 is obtained by mixing together the totality of partial

the

waves, giving each an amplitude (21

An

and a phase factor

term (21 + 1) can be


argument. Consider an imaginary

instructive interpretation of the amplitude

obtained from the following classical


t

1)

These relations depend upon the equality e iz

cos z

/'

sin z.

PARTIAL

Sec. 5]

WAVE

ANALYSIS

541

surface, perpendicular to the direction of propagation of the plane wave,

and containing a

set

of circles of radii

d
centered

on the

intercept of the z axis

in figure (15-12).

(15-65)

II

and the

surface.

Classically, the orbital angular

origin of coordinates of a particle

This

is

illustrated

momentum about

moving parallel to the z axis

is

the

equal to

its

Figure 15-12.
in

Illustrating the classical interpretation of the amplitude term which arises


the decomposition of a plane wave into partial waves.

impact parameter d times its linear momentum/) = KH. Thus all particles
passing through a ring of inner radius d and outer radius dl+1 will have
angular momentum between (l(K)Kh
IH and [(/ + \)jK\Kh = (/ + \)H.
In the classical limit, / is large compared to 1 so all particles passing through
t

the ring can be characterized as having the angular

momentum Ih. Now a

which is a member of a uniform beam (one which would be


described by eikrcose) is equally likely to have any impact parameter.
Therefore the probability that it passes through the ring is just proporparticle

tional to

its

area A. This

is

^=^ i-fl = J[(i + D


+

-!2]

A.

A = -k V 2 +
K

2l

[2/ + 1]
n=
K
2

(15-66)

or

a:

(21+

1)

So (2/ + 1) is the relative probability that a particle moving in a uniform beam will, in the classical limit, have an orbital angular momentum
about the origin of IH, which is the orbital angular momentum associated

COLLISION THEORY

542

quantum number

with the partial wave of

V7(/

1)

[Ch. 15

in the classical limit

where

h -> m.

Let us continue with the problem of determining the function f(8) in


the expression
rp(r,

6)

iKrmse

+f(B) -

iKr

-* oo

(15-67)

which, according to (15-47), describes the eigenfunction for the scattering

of a particle by the potential V(r), at positions far from the region where
V(r) has

any appreciable value. Now,

having no

(f>

From

for the potential V(r).

form

is

similar to the

the discussion above,

form of the solution (15-57)

V(r, 6)

The angular dependence

= |

of constants

because the

is

R (Kr)
t

different,

b.RlKr) P

exactly the

is

equation (10-23) which governs


set

at all positions,

a solution,

evident that

0.

and the

its
is,

(15-68)

(cos 6)

same because the


dependence

radial

That

differential

is

But the

different too

satisfy (10-27), i.e.,

dr!

dr \

it is

for V(r)

never contains the potential.

it

12 Uf) + [- !H!>
+
2
r

8) is

rp(r,

dependence, to the time independent Schroedinger equation

| [ - vtdR =
ft

(15-69)

= which is satisfied by the j\(Kr).


However, the Ri(Kr) are essentially different from the j\(Kr) only for
0. For large r, V(r) = 0, and both R^Kr) and
small r, where V{r)
ji(Kr) satisfy the same differential equation. Thus, for large r these functions are-essentially the same and, in particular, for r co the R {Kr)
have the form
instead of this equation with V(r)

>-

R
->

, .
(Kr)

sin

(Kr

\n\2
'

d,)
-,

r -- oo

/1C _.

(15-70)

Kr

R (Kr)

can at most differ from the corresponding


by a phase shift d which arises from the different r dependence of the two functions in the region where V(r) ^ 0.
It is not difficult to prove this by showing that the functions (15-70) are

For

oo,

each of the

ji{Kr) of equation (15-63)

solutions to (15-69) for

r-*

with increasing r faster than

co,
\jr,

providing only that V(r) approaches zero

but we shall not do

this here.

Instead,

shortly we shall study some graphs of j (Kr) and R {Kr), for typical potentials V(r), which will make it apparent that (15-70) is correct, and will
t

also

show the

significance of the phase shifts.

WAVE

PARTIAL

Sec. 5]

By equating

ANALYSIS

543

we

obtain

a iKr

2 (21 +

l)i

and (15-68), and

the right sides of equations (15-67)

writing eikrcose in terms of (15-64),

MKr) P

(cos 0)

oo

= MjC&) /'.(cos 0),


=

+f(6)

oo

Evaluating ^(Ar) for r-^ oo from (15-63), and .R/Ar) in the same limit

from

(15-70), gives

| (2/ +

sin(Xr-M2)
i),.

Kr

+ f{Q)

Q)

Kei

Kr

^^ ^(Kr-Kr

1.12

6Q

oo

This equation determines both/(0), and the entire set of b lf in terms of the
phase shifts. To see this, we use (15-61), and a similar equation, to write
the sine functions as complex exponentials.

Then, after multiplying

through by Kr, we have


~

2 (2i +
1=0

iy

i(.Kr-ln/2)

-i(Kr-lirl2Y

P,(cos0)+/(0)KViKr

2x
'gHKr-lirlZ + dd

lb
1=0
(21

l)i

P (cos

- n!r / 2

coefficients

bi

oo

0),

2i

+ 2 _
L

-HKr-lvl2 + diY

li

Grouping together the

t
.

of eiKr and e~ iKr gives

P (cos6)+f(d)K\e iKr

-ii7r/2 iSi

2i
(21

l)i

bi

eillr/2

2i

ii,i2

-i,

P (cos
;

2i

0) } e

- iKr

0,

oo

This can be satisfied for arbitrary r only if the coefficients of both eiEr
and e~ iKr vanish separately. The latter condition gives

P,(cos 0)
1

24

Since this must be true independent of the value of

each P^cos

0)

must be zero. For a


(21

2x

1)

typical

jj^/g

_ i en W

li

0,

the term multiplying

this requires

j2

-u

li

or

= (21 +

1)<V*'

(15-71)

COLLISION THEORY

544

Substituting this into the coefficient of e iEr ,


coefficient vanish,

and demanding that the

we obtain

P (cos
1

[Ch. 15

2i

0)

-f(6)K

2i

or

ffl
which

= r^

2(2/

2<a

l)(e

l)e"' sin 3,P,(cos 6)

'

l)P,(cos 6)

(15-72)

is

/(0)

= 7:

2(2/

(15-72')

where we have used (15-60) and an equation similar to (15-61). Since


the b are constants, they have the values specified by (15-71) everywhere,
even though that equation was obtained at r -* 00. A similar statement
-* 00.
is not necessary for (15-72') as the function
f(6) is defined only at r
t

Having the expression (15-72') for f(d), the scattering cross


can be obtained immediately. According to (15-50), da\di

sections

is

^=/*(0)/(0)
ail

= 4 2 (2/' +
K U'=o

l)e~

iS >-

| (2/ +

sin d v P,.(cos
0)]f
'

l)e

iS
>

sin

<5

P (cos 0)]

u=o

(15-73)

where we have used


to be

summed
a

/'

and

to indicate specifically that the

independently. According to (15-45), a


f

dQ

dQ.

!
J

dQ

2tt sin 6

dB

2n

*
\

J-i

two

series are

is

dQ

d(cos 0)

is evaluated by multiplying the two series of (15-73) and integrating


each of the resulting terms over cos 0. By applying (15-54), the reader

This

may

verify that all terms for

S
K 2
j

We

(2/
=o

I)

/'

sin

vanish, and that the result

=I
-41=0
2/

*,

see that the scattering cross sections

(2/

1) sin

is

3, (15-74)

da/dQ and a are completely

determined by the value of


and the values of the phase shifts d
The phase shifts are evaluated by solving equation (15-69) for each
t

and comparing the phase of R^Kr), at some large value of r, with the phase
of j\(Kr) at the same value of r. This is illustrated in figure (15-13) for a
typical / and three different forms of the scattering potential V(r). In
these curves the function

u,(Kr)
is

plotted versus

r.

This

is

r^
DfI
rR
(Kr)
t

,
,n
(15-75)

convenient because the amplitude of u (Kr)


t

PARTIAL

Sec. 5]

behaves like
including

rr

= 0.

WAVE
l+1
,

ANALYSIS

545

as r -* 0, so that

The reason

is

R (Kr)
t

it

goes to zero at r

cc r\ as r

0, in

for all

common

/,

with

the radial dependence of the eigenfunctions for any spherically symmetrical


potential

[cf.

(10-45)

and the remarks following

(13-24)].

In addition,

the function u (A>) satisfies a differential equation


(

dr 2
No

{-^+![*-nr)]}-0

(15-76)

potential

Attractive potential

<-*,

-1

/?
R

Repulsive potential
*|

r*

The

Figure 15-13.

radial

dependence of an eigenfunction for three different forms of

scattering potential.

which has a simpler form than the equation (15-69) for R {Kr). [The
equivalence of (15-69) and (15-76) can be verified by substituting (15-75)
t

into the latter.] In figure (15-13) the top curve

no

potential: V(r)

any

0, all

r.

Here

shows u (Kr) for the case of


t

R (Kr) = j (Kr), there is no phase


t

shift

and the phase of u (Kr) at large r provides the zero of phase


with respect to which the phase shifts that arise when V(r) ^
can be
measured. The middle curve shows u (Kr) for the case of a potential:
V{r) < 0, r < R and V(r) = 0, r > R. With such an attractive potential,
for

/,

[E

V(r)]

> E in the region of the potential and,

according to equation
be larger in that region than
it is in the case of zero potential.
Consequently, u^Kr) oscillates more
rapidly for r < R. For r
R, its behavior is unchanged except that its
phase is shifted. We see from the curve that: With an attractive potential
u (Kr) is "pulled in" by its behavior at r
R, its phase is advanced, and
(15-76), the

magnitude of (Pu^Kr^dr 2

will

>

<

COLLISION THEORY

546

[Ch. 15

The bottom curve shows u (Kr) for the case


> 0, r < R and V{r) = 0, r > R. Here
[E V(r)] < E in the region of the potential, and the magnitude of
d^u^Kr^jdr 2 is decreased in this region. The result is that: With a repulsive
potential u (Kr) is "pushed out," its phase is retarded, and the phase shift
the phase shift

is

positive.^

of a repulsive potential

V{r)

negative.^

is

Some

6.

Applications of Partial

In this section

we

Wave Analysis

shall calculate the scattering cross sections for

some examples of

quite different potentials in order to provide

wave

and to

analysis,

illustrate further the significance

An

of phase

shifts.

AAA

Mil
i/u V

Figure 15-14.

two

partial

DvUI

V u

eigenfunction and the scattering potential for the scattering of hard

spheres.

Consider

by

first

the scattering, in the classical limit, of a small hard sphere

a large hard sphere (the scattering of a ball bearing

The

by a

potential V(r) can be represented as an infinite repulsion

billiard ball).

commencing

R, where the interaction radius R is the sum of the radii of the two
R is the same as at
spheres. The boundary condition for the u (Kr) at r
at r

the edge of an infinite square well potential


it

demands

that each u t (Kr) vanish at r

(cf.

section

R.

The

typical u t (Kr), are plotted in figure (15-14).

interaction radius

length X

is

5,

Chapter

and
and a

8),

potential V(r),

In the classical limit, the

extremely large compared to the de Broglie wave-

2tt/K, so

R > IttJK
t

The phase

potential, only
shift greater

an attractive potential, and negative for a repulsive


not so strong as to make the magnitude of the phase

shift is positive for


if

the potential

is

than 180. Cf. section

6.

SOME APPLICATIONS OF PARTIAL WAVE ANALYSIS

Sec. 6]

547

or

KR >
and m (A>) has been pushed out by a
;

clear, then, that the resulting

value in

its

phase

(15-77)

large

number of wavelengths.

It is

of u (Kr) might well have any

shift d t

range of definition 180 to +180, and that for similar

randomly distributed over this range.


However, the statement is not true for all u (Kr). For values of / greater
than a certain /max u (Kr) is not significantly changed at any r by the
presence of the potential V{r), and the phase shift d is essentially zero.
The reason is u (Kr) = rj\(Kr) in the absence of the potential, and it is a
property of the spherical Bessel functions ji{Kr) that their values are very
u,(Kr) the values of d l will be

small for

r less

than r max

the location of their

Ijk,

first

maximum.

[This

property can be verified for the j,(Kr) of (15-55).] For sufficiently large
r max
its
/

>

-K>

presence cannot produce any significant change.

raax

/,

= rj\(Kr) is negligible in the region of the potential, and

so u i(Kr)

good estimate of

can be obtained by setting

~R

IK~

r
'max

lmax

~KR

/
'max/ JV

"

or

The same

(15-78)

estimate can be obtained from the classical argument connected

with equation (15-65) by saying that a particle of angular

momentum

greater than /max /r will have an impact parameter greater than lma,jK, and
thus, if Ima
jR, its impact parameter will be greater than the radius

jK =

Classically, such a particle would not hit the potential


motion could not be affected by its presence. Quantum mechanically, the potential also has no appreciable effect and the phase shift is
negligible for the function u^Kr) of corresponding angular momentum.
For the problem at hand, it happens to be particularly easy to calculate
an accurate approximation to the scattering cross section,

of the potential.

and

its

<7

We
sum

set

<5;

for

cuts off at

>

= ^2(2>+l)sin

max and

/max ,

be random for

let d,

and we may replace

sin 2

<5,

by

<

its

/max .

Then

the

average value

J.

This gives

4_

Jmax

o~
T(2/+l)
K*2i=o
Furthermore, since / max

we may approximate
a

~ 2H P
K

Jo

is

large

by an

it

many

maX
(2 /

terms contribute to the sum, and

integral over

1)

di

~ "kL
K

dl.

Thus

[""21
Jo

dl

^I^t

K2

COLLISION THEORY

548

Using (15-78),

this

[Ch. 15

becomes
a 2i 2ttK2 R 2 /K 2

or

The

~ 2nR

(15-79)

differential scattering cross section is,

calculate as

however, not so easy to

necessary to determine the exact behavior of the d for

it is

values of

in the

neighborhood of /max

This

is

because dajdQ, involves

.SIS

8-

Figure 15-15.

The

differential scattering cross section for the scattering of hard spheres.

interferences between terms of different


sensitive to the values of d t

We

calculation in figure (15-15).

/ (cf.

equation 15-73) and

The

differential

larger angles dajdQ.

is

very

scattering cross section

contains a strong peak at very small angles, with a width

At

is

therefore present only the results of the

The

isotropic.

6'

~ IjKR.

integral (do/dQ) dQ. over the

region of the peak has the value ttR 2 and the integral over the isotropic
,

region has the same value.

The peak

represents a diffraction effect which


It cannot be
immeasurably small;

the theory predicts to be present, even in the classical limit.

observed experimentally in

this limit

because

6' is

measurement of the scattering cross section yields the projected


geometrical cross sectional area ttR 2 and not 2ttR 2 because the diffraction
scattering is missed. However, the diffraction scattering can be observed
in experiments involving the scattering of nuclear particles by nuclei,
for which 6' is not so small. Another example is the scattering of electromagnetic radiation from a reflecting sphere of radius R large compared to
XjR is too small. The differenthe wavelength, but not so large that 8'
tial scattering cross section is found to have the form shown in figure
2
Half
(15-15) and the scattering cross section is found to be a = 2ttR
of this arises from the reflection of radiation hitting the sphere, and half
a

classical

SOME APPLICATIONS OF PARTIAL WAVE ANALYSIS

Sec. 6]

549

from the diffraction of radiation, passing near the sphere, into the
shadow on its dark side.
Partial wave analysis provides an exact procedure for solving scattering
problems at all energies. However, its application at high energies is

arises

K is

usually a very formidable task because


is

the value of r

large, so

KR >

beyond which the scattering potential V(r)

1,

is

where

negligible.

This means that the expression (15-73) for dajdQ. will contain contributions

KR there will be an
/ through /max
At high energies it is generally much easier to use
the Born approximation, and this approximation is often of sufficient
accuracy. The real utility of partial wave analysis is found at low energies
where KR is not large. In this region the Born approximation normally
cannot be applied because it is very inaccurate, but partial wave analysis
is both exact and easy to apply. We shall illustrate this point by calculating
the scattering cross sections for a low energy particle, incident upon a
from many terms,
appreciable phase

since for all

shift.

simple potential, in the limit

Equation (15-78)

KR <

appreciable in the limit


spectroscopy, this

is

KR <

us that only the phase shift for

tells

called

will

be

In terminology adopted from atomic

S-wave scattering. For this case, the expressions

(15-73) and (15-74) assume the simple forms

lZ = hi
dQ
K*

{'" sin 8
o

po(cos

B)}{e

id

sin d

(cos 0)}

or

da

d~n~

sin

2
<5

KR < 1

(15-80)

KR<1

(15-81)

'

and

a
because sin d

1.

We

is

^Jo,
K

essentially zero for

see that

da/dQ

times dajdQ, in the limit

is

isotropic,

1, 2, 3,

and a

...
is

and because

(cos 6)

the total solid angle

4n

KR < 1

Let us calculate, in this

limit,

a for the attractive square well potential,

V(r)=-

V>

<R

(15-82)

which is plotted in figure (15-16). The phase shift <5 that this potential
produces is evaluated by finding a solution to equation (15-76) for / =
which goes to zero at r = O.f The equation is
i}L
dr*
t Cf. equation (15-75).

It [
2t
h

v{ry\u

(15-83)

COLLISION THEORY

550
In the region r

<

R,

[Ch. 15

can be written

it

d 2u
dr 2

+ Ku =

(15-84)

where

K = V2M (E + V
The general

solution to this familiar equation

= A sin KgT + B cos K

(15-84')

)/ti

is

r,

<R

Figure 15-16.

To

An

attractive square well potential.

obtain the required behavior at

u
In the region r

>

=A

sin

0,

we
r

r,

B=

set

and take

<R

(15-85)

R, equation (15-83) can be written

^L + K 2 u=0
2

(15-86)

dr

where

K=
The general

solution to this

is

= C sin Kr + D cos Kr,

Both terms are needed


dujdr be continuous at

(15-86')

y/2/lElh

r> R

(15-87)

in order to satisfy the requirement that u


r

R.

Before applying these requirements,

and
it is

convenient to rewrite (15-87) as


u

/"sin (Kr

),

r> R

(15-88)

The equivalence of (15-87) and (15-88) can be seen by expanding the


latter to obtain u = Fsin Kr cos d -\- F cos Kr sin d
and setting F cos
= C, F sin = D. Furthermore, by considering the function R (Kr)
= u (Kr)/Kr, where u (Kr) is given by (15-88), and comparing this at
of (15-88)
r-> oo with (15-70) for / = 0, it is seen that the quantity
(5

<5

<5

is

precisely the

S-wave phase

shift.

Now

apply the requirement r that

SOME APPLICATIONS OF PARTIAL WAVE ANALYSIS

Sec. 6]

(15-85) and (15-88) give the same value of u at

A
The requirement
at r

cos

K R = FK cos (KR +

Dividing (15-89) by (15-90),

R. This yields
(1

5-89)

and (15-88) give the same value of du/dr

that (15-85)

R yields

AK

K R = F sin (KR +

sin

551

(1

5-90)

we obtain

tan

KR=

tan

(KR

<5

or

K
IS

tan"

Ka R = KR +

tan

so
d

Knowing

the phase

tan

shift,

~K

tan

we can

KR

(15-91)

calculate the scattering cross section

from equation (15-81). The calculation is simple if we use the


conditions
KR < 1 and K/K < 1, in the limit of very small E, and assume that
K R does not happen to have such a value that tan K R > 1. Then
(15-91) becomes
S

~ tan K R - KR~ sin


Kn

(S

or

;mS

^KRhKnKR R -l)

and (15-81)

gives

This approximate equation predicts a -> oo


for K R = tt/2, ln\2,
where tan K R -> oo. Of course, the equation
cannot be trusted for these
K R since they violate the assumption from which it was derived. The exact
equation (15-91) shows that for these
R the phase shifts <5 = K R and,

from

(15-81),

very large but

Note
very

we

see that for these

finite

values

that this can be written a

much

larger than

nR\

<5

the scattering cross sections have the

[4/(KRf]nR^ so

47r/X 2 -

(15-93)
that, since

KR <

1,

is

the projected geometrical cross sectional area


of the scattering potential. When a takes on
its maximum possible value
for S-wave scattering (15-93), the
scattering cross section is said to be at
an S-wave resonance. Resonances for other
partial

waves occur

if

E is

COLLISION THEORY

552

high enough that there are large phase

when dt

7r/2, 3tt/2,

shifts for /

>

0.

For

[Ch. 15

instance,

the scattering cross section takes on a particularly

and is said to be at a P-wave resonance.


The approximate equation (15-92) does correctly predict

large value

KqR which

that a -*

for

K R = K R.

For these
KgR, the S-wave phase shift is d = n,2ir,
,t and we find directly from
(15-81) that a 0. This behavior of the scattering cross section can be
seen in the Ramsauer effect, which is the term used to describe the extremely
are solutions to the equation tan
.

Figure 15-17.

The

radial

dependence of an eigenfunction for two attractive square well

potentials.

small value of a measured in the scattering of electrons from atoms of a

E of about 1 ev. At low energies only the S-wave


by the quite sharply defined attractive potential that a noble
gas atom exerts on a passing electron, and at about 1 ev the S-wave is
pulled in by just such an amount that d c n and a
0.
Figure (15-17) shows plots of the function u (Kr) for an attractive
square well potential of radius R which produces a phase shift <5 = tt/2,
and for a deeper attractive square well potential of the same radius which
produces a phase shift d = n. Note that this figure, and figure (15-13),
show that the amplitude of u^Kr) is generally smaller in the region r < R
than in the region r > R, but that the amplitude is equal in the two regions
when the partial wave is in resonance. It would be instructive for the
reader to compare this, and the other properties of three dimensional
scattering potentials, with the properties of one dimensional scattering
potentials discussed in sections 3 and 4 of Chapter 8. It would also be
instructive to compare the condition K R = tt\2, for the lowest energy
S-wave resonance in the potential (15-82), with the condition that a new
noble gas at an energy

is

affected

7=0 state is just bound in that potential.


t 8

is

not of physical interest because

it

This can be done by using the

corresponds to no scattering potential.

SOME APPLICATIONS OF PARTIAL WAVE ANALYSIS

Sec. 6]

553

solution to the Schroedinger equation for the potential (8-61), since

same form

exactly the
/

state in the potential (15-82).

show
it

as the Schroedinger equation (15-83) for a

when

that

also has

higher

The

the potential binds an

results of this

has

comparison

will

near zero total energy

state

an S-wave resonance near that energy. This

values and for potentials of any shape.

it

bound

is

also true for

The reason

for this

At a low energy resonance the eigenfunction


has a particularly large value in the region of the potential, and this is
exactly the condition that characterizes a bound state.
Furthermore,
relation

is

easy to understand.

since for a high energy resonance the eigenfunction also has a particularly
large value in the region of the potential, the eigenfunction for such a case

has a certain analogy to the eigenfunction for a bound


reason

it is

state.

For

this

often said that a potential has virtual states, or virtual levels,

which it exhibits resonances, even though these energies


and correspond to situations in which the particle is actually

at all energies for

are

all

positive

unbound.

For KR < 1, only the S-wave is affected by the scattering potential,


and do/dCl is isotropic. With increasing KR, partial waves of higher
angular momentum are affected; for KR = 1 the P-wave will be significantly affected and the D-wave will be slightly affected. When these
partial waves begin to contribute to the scattering, equation (15-73)
shows that do/dO. will no longer be isotropic. As more and more partial
waves contribute to da/dQ, its structure becomes more and more pronounced. This can be seen by comparing the isotropic do/dO., which
we have just obtained for an attractive square well in the limit KR < 1,
with the highly non-isotropic da/dD. for such a potential in the high energy
limit
4.

KR >

The

1,

which we obtained by the Born approximation in section

differential scattering cross section do/dO,

KR, but

not only exhibits more

becomes more sensitive to the


exact form of the scattering potential V(r). As an example, compare the
complete independence of the structure of the isotropic da/dQ. on the form
of V(r) in the limit KR < 1 with the strong dependence shown in the limit
KR > 1 for the two potentials treated in section 4.
This situation simply reflects the fact that the details of the form of the
structure with increasing

it

also

scattering potential can be important only if the de Broglie wavelength of

the scattered particle, X


size

2tt]k,

is

small enough, in comparison with the

of the potential, R, to allow the particle to resolve these details. These

conclusions are of great practical importance in nuclear physics where

form of V{r) is not known a priori but must be inferred from analyzing
measurements of dajdD.. To study the details of V(r), it is necessary to
measure da/dQ for high energy particles. The analysis of the measurements is usually quite difficult (unless the energy is so high that the Born
the

COLLISION THEORY

554

[Ch. 15

because many partial waves enter


However, it is not impossiblethanks to high
speed electronic computing machines.
We end this discussion by mentioning a very important scattering
potential which cannot directly be treated by the method we havedeveloped.

approximation

is

sufficiently accurate),

into equation (15-73).

This

is

the

Coulomb

potential V{r) oc

\jr.

As

this potential

does not

approach zero with increasing r faster than 1/r, the expression (15-70)
is not valid.
It is possible to modify the method in such a way that the
potential can be treated, but because this is complicated we shall only
present the results. It is found that for a particle of reduced mass //
and initial velocity v, scattering from a Coulomb potential V(r) = zZe2 jr,
the differential scattering cross section

do
dQ.

is

(*%-*
\2uvV smUC'^

(,5-94)
y
'

\2/uvV sin 4 (6/2)

Comparison with equation (4-18), and a moment's consideration, will


demonstrate that these results of the quantum mechanical theory are
identical with the results of Rutherford's classical theory of the scattering

by a Coulomb

potential.

This remarkable situation

is

true only for a

potential corresponding to an inverse square law of force,


in the following

way.

From

dimensional analysis,

it

and

it

arises

can be shown that

if

on a particle varies according to r n then the scattering


cross section must vary according to h i+2n
For the inverse square law
n = 2, the scattering cross section is independent of the value of
Planck's constant h, and this requires correspondence between the results
of the classical and quantum mechanical calculations of the scattering
cross section. In fact, for the Coulomb potential even the Born approximation yields a dajdD. which is identical with (15-94), as the reader may
verify with little difficulty. This is why the Born approximation is fairly
accurate for the scattering of electrons from atoms at energies as low as
^100 ev. The approximation is good in the outer shells of the atom
because there the energy of the scattering electron is large compared to the
non-Coulombic potential; it is good in the inner shells, although the
energy is not large compared to the potential, because the potential is
the force exerted

essentially

7.

Coulombic.

Absorption

The theory we have developed up to this point treats only the scattering
produced by the potential V{r) representing the interaction between the
incident particle and the target particle. But there are also processes in

ABSORPTION

Sec. 7]

555

which the incident particle is absorbed by the target particle. An example in


atomic physics would be the capture of an incident electron by a target
atom, leading to the formation of a stable negative ion or the re-emission
of an electron of reduced energy. In nuclear physics there are many
examples of the capture of an incident nuclear particle by a target nucleus,

some nuclear reaction. In this section we shall outline the


procedures used for treating absorption processes in the Schroedinger
theory. It is to be understood that we use the word "absorption" to mean
leading to

all

no energy

processes except scattering with

in the

loss in the

CM frame (elastic

in particular, absorption includes scattering with energy loss

scattering) ;

CM frame (inelastic scattering).

as throughout this chapter

The word

"scattering"

is

used here,

and the remainder of the book, to mean only

CM

no energy loss in the


frame.f
have seen in equation (15-47) that the amplitude of the radially
outward traveling waves, scattered at the angle 6, is proportional to the
scattering with

We

quantity f(d). According to (15-72), this

M=^I

is

(2/

2iK i=o

lXf,

- W(cos 0)

(15-95)

where we write
r\%

2i6 >

(15-96)

Equations (15-68) and (15-70) show that the total eigenfunction, describing
the incident plane waves plus the scattered waves, can be written as radial
standing waves. These radial standing waves can be decomposed into
radially inward and radially outward traveling waves. In doing this, it is
found that the rj of the relation (1 5-96) specify the amplitudes and phases
of the radially outward traveling waves relative to the radially inward
traveling waves. For instance, when r] = 1 or d =
for all /, the inward
and outward traveling waves have the same amplitudes and phases, and
they combine to give a total eigenfunction describing incident plane waves
plus no scattered waves. If r] = e 2ie = 1 for some /, because d
0,
the inward and outward traveling waves have the same amplitudes but
different phases, and they combine to give a total eigenfunction describing
incident plane waves plus some scattered waves.
l

'

To

allow for absorption as well as scattering, the relation (15-96)

is

broadened to read

= A ie \
A, <
(15-97)
If A <
for some the amplitude of the associated
2i

tj,

where the A are real.


1
/,
outward traveling wave, which is proportional to
t

x,

will

be smaller than

Note that in the LAB frame there is always at least a small energy loss in

due to the

recoil

of the struck

particle.

'

'scattering"

COLLISION THEORY

556

[Ch. 15

the amplitude of the associated inward traveling wave, which is proporThen there is a net inward probability flux, proportional to
tional to 1
.

the difference between the inward and outward intensity, i.e., proportional
1
This implies absorption. The absorption cross
r\*r\i.
to 1
Af

a A can be evaluated in terms of the r) by using the three dimensional


extension of the definition of probability flux (7-31) to calculate the net
section

inward flux integrated over all angles. The scattering cross section as
can be evaluated just as we have done before. The results are

It is

TiI&+W-riivd
1=0

OA

= ir,lW +

- vd

tfXi

easy to verify that (15-99) reduces to (15-74),

total cross section

aT

is

Using (15-98), (15-99), and a


oT

=f*(6)f(d)

little

I (21
fj
K i=o

The differential scattering


from (15-50) and (15-95).

~dh2

if

~98 )

(15-99)

A =
x

for all

/.

The

defined as

aT

do5

iXi

( 15

as

aA

we

algebra,

(15-100)

1) (l

find

dasldQ.

cross section

(15-101)

LI

is

obtained immediately

It is

= -L ( | (2 +
4K \i'=o
/'

1)07*

l)P,.(cos 0))
I

X (f (21

l)(m

- IWcos 6))

(15-102)

Differential absorption or total cross sections are not defined.


These equations show that the absorption and scattering cross sections

are not independent.

If there is absorption there

because, in order to have


for

all

not

true.

it is

/,

and

if

no

scattering,

so there can be

possible to have r?*?^

1,

and

must be

scattering

However, the converse

is

necessary that

no absorption.

If there is scattering there is

it is

rft

r\ x

not necessarily absorption because


1.
have rft
1, rj t

still

the interaction in the classical limit


target sphere, and assume that the
large
sphere
with
a
incident
of a small
R, where the interaction radius
incident sphere is absorbed whenever r

As a very simple example, consider

R is the sum of the radii of the two

problem of the perfectly


absorbing sphere can be treated similarly to the problem of the perfectly
reflecting sphere given in the previous section. In the present case, the
spheres. This

ABSORPTION

Sec. 7]

total eigenfunction

KR, and

557

can contain only inward traveling waves for /


/max
normal mixture of inward and outward

will contain the

traveling waves for

>

Thus

/max .
Tjt

Tjt

=
=

0,

<

1,

>

max
max

= KR

'max

In both equations (15-98) and (15-99) the sums have contributions only
for /
/max , and

<

'max

"a

As

in the

the sum.

work preceding
The result is

(15-79),

<*a-<*b=
and we

(21

we may

1)

use an integral to approximate

= jp- = wH

Ki

(15-103)

also have

aT

The absorption cross


area nR2 as would be
,

section

oA

is

a8

is

(15-104)

the projected geometrical cross sectional

expected. But there

of equal magnitude. This

2-nR?

is

also a scattering cross section

the diffraction scattering of incident particles

passing near the absorbing sphere into the shadow on

its far side, and it is


found in the case of the reflecting
sphere. As we have said before, the diffraction scattering cannot be
observed in the classical domain because the angle d'
1/KR, within
which it occurs, is much too small. However, diffraction scattering is very
observable and very important in the nuclear domain.
In more typical examples of problems involving absorption, it is not
possible to evaluate the r) by qualitative arguments, and these quantities
must come instead from the solution of an appropriate Schroedinger

exactly the

same phenomenon that

is

equation.

write a Schroedinger equation for a


which quantum mechanical probability is absorbed, even
though we have so far always found it to be conserved. This is done by
using, as an interaction potential, the complex function
It actually is possible to

situation in

V(r)

iW(r)

(15-105)

<

where V(r) and W(r) are both real, and for absorption W(r)
everywhere [for emission W(r) > 0]. There are several ways of demonstrating
that the presence of an imaginary term in the interaction potential leads
to absorption [or emission]. For instance, by starting at equation (7-26),
and carrying through the subsequent calculation with a complex potential

COLLISION THEORY

558

[Ch. 15

V(x) + iW{x) and a small region x x to x2 the reader may show that the
quantum mechanical probability absorbed per unit time in this region is
-(2Winy*Y(x2 - xj), where
and Y*T are evaluated at some typical
point in the region. The rate of absorption of probability per unit length is
,

then

R= -

2W

T*T

(15-106)

In three dimensions the same term gives the rate of absorption of probability per unit volume. Within recent years complex potentials have been

widely used in describing the interactions between nuclear particles and


nuclei.

BIBLIOGRAPHY
Bohm,

D.,

Schiff, L.

Quantum Theory, Prentice-Hall, Englewood Cliffs, N.J., 1951


Quantum Mechanics, McGraw-Hill Book Co., New York, 1955.

I.,

EXERCISES
1.

Carry through the calculation leading to equation (15-17).

2.

Verify that (15-18)

3.

Use the Born approximation

potential V(r)

is

a solution to equation (15-19).

to calculate do/dCl for the screened Coulomb


(zZe 2 /r)e~ r ' d , which provides a useful approximation to the

potential between a charged particle

radius

Then

of the atom.

the normal

Coulomb

let

and a neutral atom if d is set equal to the


<x>, and
show that da/dQ approaches

differential scattering cross section (15-94)

when

V(r)

approaches the normal Coulomb potential.

For the Bessel functions listed in (15-56), verify that, as stated in


and 6: (a), y,(A>) * r r -0; (b), j (Kr) = sin (Kr - h/2)/Kr,

4.

(c),

the

first

maximum

of jiiKr)

5.

Verify equation (15-74).

6.

By

is

using the procedure indicated in section

the lowest energy S-wave resonance


just
7.

located at rmax

bind an

Use

partial

hard spheres with

6,

II

oo;

K.

prove that the condition for

the same as the condition that the potential

near zero total energy.

state

wave

is

sections

and a for the scattering of


=0.1, considering only S-waves, and also with KR = 0.5,
and P-waves. Plot the da/dQ. obtained in both calculations.
analysis to calculate da/dQ.

KR

considering only SHint: For r > R it

is

necessary to use the general solution to the differential

(Kr) + P,,(Ar), where each n t (Kr) can be generated


j
from the corresponding j (Kr) by everywhere replacing cos by sin and replacing
sin by cos.

equation in

r.

This

is

8.

Derive equation (15-106).

CHAPTER

The Nucleus

I.

Introduction

It is fair to say that the properties of the atom are completely understood.
The nature of the forces acting between its various parts is known, and
quantum mechanics provides a complete technique for evaluating the
effects of the forces. For the nucleus this is not the case. Despite great

progress in recent years, the nature of the forces acting between


parts

is

not completely known.

quantum mechanics

will

Furthermore,

all its

exist

its

various

not even certain that

provide a completely adequate technique for

evaluating the effects of these forces

does not yet

it is

when

they finally are known. There

a well-established theory of the nucleus through which

properties can be understood. However, there

do

exist a

number of

models, or rudimentary theories of restricted validity, each of which can


explain a certain limited range of
the study of the nucleus

is

able consequence of this situation

lack

much

its

properties.

At

the present stage,

largely the study of these models.


is

An unavoid-

that a discussion of the nucleus

must

of the coherence which characterizes a discussion of the atom.

On the other hand, it is just the fact that everything about the nucleus is not
yet understood that

makes the

subject particularly interesting.

Let us introduce our study of the nucleus by recapitulating the pertinent


information which

and by
1.

we have

filling in certain

We

have

learned incidental to our study of the atom,

other information.

briefly described

(Chapter 4) Becquerel's discovery of the

and the identification of the alpha particles


as doubly ionized He atoms. The beta particles were identified as fast
electrons from a measurement of their charge to mass ratio. The gamma
rays were assumed to be electromagnetic radiation, and this was confirmed
radioactivity of heavy atoms,

559

THE NUCLEUS

560

by showing that they can be


that radiation.

diffracted

from a

[Ch. 16

crystal like other

forms of

After Rutherford's discovery of the nucleus from an


it became apparent that
heavy atoms. Later in this
study alpha, beta, and gamma emission by such nuclei,

analysis of alpha particle scattering (Chapter 4),


radioactivity involves the unstable nuclei of

chapter

but

first

ground

we shall
we shall

concentrate on the properties of stable nuclei in their

states.

We

have learned (Chapter 4) that the mass of a nucleus is only


than the mass of the corresponding atom. Thus the nuclear
mass is approximately equal to the integer A times the mass of an H atom,
or approximately equal to A times the mass of a proton, the nucleus of
an H atom. The integer A is the integer closest to the atomic weight of
2.

slightly less

We

have also learned that the charge of a nucleus is exactly


number Z times the negative of the charge of an electron, or exactly Z times the charge of a proton. The periodic table of the
elements shows that A is roughly equal to 2Z, except for the proton for
which A = Z = 1.
3. Analysis of alpha particle scattering from nuclei of low A (Chapter 4)
showed that the radii of such nuclei are of the order of 10-12 cm, where the

the atom.t

equal to the atomic

radius

is

from the center of the nucleus at which


on the alpha particle first deviates from a Coulomb
Analysis of the rate of alpha emission from radioactive nuclei

defined as the distance

the potential acting


potential.

(Chapter 8) indicated that the radii of these nuclei, defined in


same way, are ~9 x 10~13 cm. Note that the ratio of the density of a
nucleus to the density of bulk matter in the solid state is
of high

the

[volume of nucleus]

density of nucleus
density of solid matter

[volume of atom]

-1

[(10

-12

3
) ]

-8

-1

[(10

-1
nl2

-1

-3

So the density of a nucleus is ~10 gm-cm


4. The word "radius" implies that the nucleus has a spherical shape.
12

We

have seen that the study of atomic hyperfine structure (Chapter 13)
shows the charge distribution of nuclei is generally not spherical, but
instead forms an ellipsoid of revolution with an axis of symmetry along
the direction about which the nuclear spin angular momentum vector
precesses. As a consequence, nuclei have a static electric quadrupole
moment Q. We have learned that for nuclei of spin angular momentum

quantum number

>

1,

there are cases of

Q>

(ellipsoid elongated in

the direction of the symmetry axis) as well as cases of


flattened in the direction of the
t

symmetry

axis),

but for

For any particular isotope of an atom, the atomic weight

integer.

The term "isotope"

will be explained in section 4.

is

Q<

or

(ellipsoid
\,

Q=

0.

always very close to an

THE COMPOSITION OF NUCLEI

Sec. 2]

561

However, the departures from sphericity are never severe;

in extreme

cases the ratio of the semimajor axis to the semiminor axis

Thus

it is

is

~1.2.

often possible to ignore these departures and assume a spherical

Nuclei do not have static electric dipole moments, for this would
correspond to a constant displacement of the center of the charge distribution from the center of the mass distribution, which cannot happen.
shape.

5.

Nuclei do have

static

magnetic dipole moments. This

is

also

shown

in

We

have seen that


the magnitudes of these magnetic dipole moments are ~jMjv = eh/2Mc <~
10~ 3
is the proton mass, and fi B
i"b> where fi N is the nuclear magneton,
is the Bohr magneton. Hyperfine structure also shows that the signs of the
magnetic dipole moments (the relative orientation of the magnetic dipole
the study of atomic hyperfine structure (Chapter 13).

moment

vector and the nuclear spin angular

positive in
6.

The

some

cases

spin angular

and negative

momentum quantum number

able from atomic hyperfine structure.


i is

momentum
7.

is

given in terms of

vector) are

is

directly

measur-

As we have learned (Chapter 13),


when Z is also even, and i is half-

when A is even, with / =


when A is odd. The magnitude

integral

integral

momentum

in others.

/ of the nuclear spin angular

by the usual

relation /

Closely related to the question of nuclear spin

is

Vi(i

1) h.

the question of the

symmetry character of the eigenfunction for a system containing two or


nuclei of the same species (Chapter 12). This is studied by analyzing
the spectra of diatomic molecules containing two identical nuclei and, as
we have stated, it is found that nuclei of integral spin quantum number
(even A) are of the symmetrical type and nuclei of half-integral spin
quantum number (odd A) are of the antisymmetrical type.

more

2.

The Composition of Nuclei

With one exception, A <~ 2Z for all nuclei. (The exception is the proton
which A Z = 1 .) From this it is immediately apparent that a nucleus
with a given value of A cannot be composed of A protons alone, because
A/2. Before 1932 it was generally
its Z would be Z = A instead of Z
assumed that a nucleus is composed of A protons and (A Z) electrons.
This gives the correct charge and also allows the mass to be correct.
However, it was realized that the assumption that electrons are contained
for

in nuclei presents serious difficulties.

Although the assumption made it easy to understand how electrons


could be ejected from nuclei in the beta emission process, there were problems with the magnitude of the zero point energy necessary to confine a
particle of

mass

as small as

an electron in a region of space as small as

THE NUCLEUS

562

By

the nucleus.

directly applying the uncertainty principle,

momentum from

the equation

E=

[Ch. 16

and evaluating

cp which obtains for highly relativistic

particles, the reader may show that the kinetic energy of an electron
confined to a region of dimensions 10~ 12 cm must be > 100 Mev. It was

with the fact that the typical kinetic energy of

difficult to reconcile this

an

electron ejected in the beta emission process

It

was

is

~1

Mev.

also difficult to reconcile the observation that all nuclear magnetic

dipole

moments

are three orders of magnitude smaller than the magnetic

dipole

moment

of an electron, under the assumption that electrons are

contained in nuclei.

The insurmountable

difficulty

sideration of nuclei in which

the

N atom with A =

spin angular

14,

Z=

with the assumption came from a con-

is

7.

even and

Z is

odd,

e.g.,

the nucleus of

In conformity with the usual rules, the

momentum quantum number

of this nucleus

observed to

is

1), and the symmetry character is observed


to be symmetrical. But, if it contains 14 protons and 7 electrons, its spin
quantum number would necessarily be half-integral and its symmetry
character would necessarily be antisymmetrical, because both the proton
and the electron have spin quantum number i = \ and are of antisymmetrical symmetry character, and because 14 + 7 = 21, an odd number. It is
easy to see from the rules for the addition of spin and orbital angular
momentum, presented in Chapter 13, that a system containing an odd
number of particles with half-integral spin quantum number can only have

be integral

(specifically,

momentum (the nuclear spin angular momentum) corresponding to a half-integral quantum number. It is also easy to see, from
the discussion in Chapter 12, that the symmetry character of a system
containing an odd number of particles of antisymmetrical symmetry
character must be antisymmetrical.
Some years before its discovery in 1932, Rutherford suggested the
total angular

existence of a particle having the

same mass, symmetry

spin quantum number


we now call the neutron. One motivation

character,

as the proton, but zero charge. This

is

for this suggestion

and

the particle

was

that the

problems outlined above would not arise if nuclei were assumed to be


composed of protons and neutrons. In particular, the spin and symmetry
character of the
nucleus would agree with observation if it contained
7 protons and 7 neutrons. This would also, of course, be in agreement
with the mass and charge required for A = 14 and Z = 7. A number of
people tried to devise experiments which would detect the neutron. But

this was difficult because, being uncharged, the neutron cannot directly
produce ionization, and most devices for detecting particles depend upon

ionization.

We shall describe the sequence of experiments and interpretations leading

THE COMPOSITION OF NUCLEI

Sec. 2]

to the discovery of the neutron, as

and

it

is

563

of great historical importance

good examples of some of the early techniques of


nuclear investigation. In 1919 Rutherford bombarded a target of N nuclei
with 7.7-Mev alpha particles from a radioactive source and found that
charged particles of long range were emitted from the target. The experialso provides

mental apparatus, employing a ZnS scintillation screen (cf. section 3,


Chapter 4) as a charged particle detector, is indicated schematically in
figure (16-1). The range of a charged particle is the thickness of some
standard absorbing material which it can traverse before losing all its
kinetic energy by ionizing the atoms of the absorbing material, and coming

gas

An apparatus used

Figure 16-1.

ZnS

Absorber

Source
in

the

first artificially

produced nuclear reaction.

This quantity increases monotonically with the velocity of the

to a stop.

and decreases monotonically with its charge. The range of the


particles emitted from the N nuclei of the target was so long that they could
unambiguously be identified as protons, since these particles have much
particle

longer ranges than alpha particles of comparable energies because they


have higher velocities and smaller charges. Rutherford's experiment
constituted the

first artificially
2

The equation
2

He4

A =
A =

14;
1

+ N 14 -* 8
7

represents the reaction

17

2,
8

A =
17

is

4;

N 14 is

(16-1)

and introduces some

the bombarding particle, an alpha particle which

Z=

with

and

is

He4

produced nuclear reaction

useful notation:

is

the

He

nucleus

N nucleus with Z = 7,
O nucleus with Z = 8, A = 17;

the target nucleus, an

the residual nucleus,

an

H 1 is the product particle, a proton which is the H nucleus with Z =

I.

Several other reactions of essentially the same type were investigated in


succeeding years, but in 1930 a reaction of a different type was discovered

by Bothe and Becker. They used an arrangement


in figure (16-1), except that the

ZnS

one shown
was replaced by a

similar to the

scintillation screen

Geiger counter. This is a chamber filled with gas and containing a fine
wire held at a high positive voltage with respect to the walls. Any ionizawithin the chamber liberate one or more
which are accelerated so strongly in the direction of the wire
that they produce additional ionization by collisions with gas atoms.
This process repeats, and an avalanche of electrons rapidly builds up.
tion processes occurring
electrons,

THE NUCLEUS

564

Enough charge

is

collected

by the

wire, as the result of a single initial

ionization, to produce an easily detectable electric pulse.


particles of kinetic energy 5.3

bombardment of Be or B

very thick absorbers with

assumed these

rectly

Using alpha

Mev, Bothe and Becker found


uncharged

nuclei
little

[Ch. 16

particles,

that in the

capable of traversing

attenuation, were emitted.

to be very high energy quanta

They

(gamma

incor-

rays).

Geiger counter can detect quanta since they can produce electrons in
the filling gas by the photoelectric effect, the

Compton

effect,

or pair

production.

BeorB

Source

Hydrogenous

Absorber

Figure 16-2.

Further work on

They

An

apparatus used

this reaction

in

Geiger

counter

layer

the discovery of the neutron.

was done

in

1932 by

I.

Curie and Joliot.

discovered, quite accidentally, that placing a layer of hydrogenous

material in front of the Geiger counter caused the counting rate to increase

was soon

was due to the ejection of protons from


its bombardment by the uncharged
particles. Protons, being charged, are counted much more efficiently by a
Geiger counter than quanta or other uncharged particles. The apparatus
used by Curie and Joliot is indicated in figure (16-2). By measuring
the range in the absorber of protons ejected from the hydrogenous layer,
It

realized that the effect

the hydrogenous layer as a result of

the velocities or energies of these protons could be determined from the

known

Ev

range-velocity relation.

~ 4.5 Mev

for

Be

Curie and Joliot found the proton energy


Following Bothe and Becker, they

target nuclei.

assumed that quanta incident upon the hydrogenous layer ejected protons
by a process analogous to Compton scattering. They also evaluated E,
the minimum energy of the quanta required to eject a 4.5-Mev proton.
This gave Ey
50 Mev, as the reader may verify by using the equations of
section 8, Chapter 3 (with the electron mass replaced by the proton mass).
Shortly after this experimental development, Chadwick demonstrated
that its interpretation was untenable, and that the uncharged particles
emitted from the Be or B nuclei under alpha particle bombardment were
not quanta but the long-sought neutrons. He showed first that the number
of protons ejected from the hydrogenous layer was several orders of
magnitude smaller than would be predicted by the Klein-Nishina equation
(14-36 with the electron mass replaced by the proton mass), if the incident

NUCLEAR

Sec. 3]

AND THE

SIZES

OPTICAL MODEL

He showed

uncharged particles were 50-Mev quanta.

50-Mev quantum

energetically possible for a


in

which a 5.3-Mev alpha


This calculation

nucleus.

particle
is

is

565

next that

it is

not

to be liberated in a reaction

captured by a Be nucleus to form a

based upon a knowledge of the mass of the

C nucleus and

alpha particle, the mass of the Be nucleus, the mass of the

mass and energy. We


For the present, suffice

Einstein's relation (1-23) between

shall

such calculations in section

it

4.

do several

to say that a

showed that all the experimental results were consistent


was assumed that the reaction involved in the case of a Be target

similar calculation
if it

nucleus

is

He4 + 4 Be 9 -> 6 C 12 +

A =

1.

explain the experiments performed with a

where

is

the neutron with

Z=

0,

(16-2)

would
With this

similar reaction

target nucleus.

assumption, the ejection of protons from the hydrogenous layer could be


understood as a simple collision between an incident particle, the neutron,

and a stationary

free particle of equal mass, the proton.t

minimum neutron

energy required to eject a 4.5-Mev proton

Mev, and

this is in

masses of the

good agreement with the

entities

is

Then

the

En =

4.5

imposed by the

restrictions

appearing in (16-2) plus the mass-energy relation.

Since the discovery of the neutron, there has never been any doubt that

a nucleus

composed of Z protons and {A Z) neutrons. Thus a nucleus


(A Z) = A nucleons, a word used for both protons and

is

Z+

contains

neutrons.

Nuclear Sizes and the Optical Model

3.

Some of the

first

came from alpha

evidence for the sizes of nuclei, other than that which

particle scattering

and emission, was obtained from the

Two nuclei are


common values of A,

differences in the total binding energies of mirror nuclei.

said to

and

if

form a

Z=

{A

Some examples

set

- Z) +

In each

set,

N 15

for one

if

they have

and {A

and 2 He3

and

Li 7 and

u Na23 and

15

one nucleus

3
,

is

12

Be 7

5
,

Mg23

17
,

kinetic

C1 35 and

18

A35

unstable and eventually decays by the emission

differences in the total binding energies of the

The

for the other.

B U and 6 C n

of an electron or a positron, forming the stable

=Z+

Z)

are

W
7

of mirror nuclei

and binding energies of a proton

completely negligible compared to 4.5 Mev.

two
in a

member of

the

set.

The

nuclei of each set can be

hydrogenous compound are

THE NUCLEUS

566

[Ch. 16

obtained directly from measurements of the energies of the electrons or


positrons which are emitted in the decay, and can also be obtained in-

some cases, from measurements of the masses of the nuclei


and the mass-energy relation.

directly, in

Now it is clear

that, in addition to the repulsive

Coulomb

forces acting

between each pair of protons in a nucleus, there must be attractive forces


of some other type acting between all the nucleons. If it were not for these
so-called nuclear forces, neutrons would not be bound in nuclei and, in
addition, all nuclei for Z
2 would break up by Coulomb repulsion.f
It was assumed at an early date, and has subsequently been verified by
several different kinds of evidence to be presented later, that the nuclear
forces acting between pairs of nucleons are independent of whether the

nucleons are protons or neutrons. With this assumption, the total binding
energies of two mirror nuclei will differ only

by the additional Co'ijlpmb

repulsion energy A.EC of the nucleus with the extra proton.

assumed that

all

extra proton,

is

A c =

is

this

it is

also

uniformly distributed over a sphere of radius R, the addi-

Coulomb repulsion energy can be


statics. The result is

tional

where
Using

If

the charge due to the protons, including that due to the

the atomic

evaluated from classical electro-

7e
-
6

(16-3)

number of the nucleus without

the extra proton.

equation to analyze the observed differences in total binding

number of sets of mirror nuclei with A ranging from 3 to


was found that values of R so determined could be represented by

energies of a
38,

it

the equation

R=
where
recent

r is a constant.

work on

this

The

(16-4)

A*>

~ 1.4 X

analysis gave r

10~ 13 cm. However,

problem has shown that quantum mechanical cor-

rections to this simple classical calculation reduce r somewhat.

currently quoted

The quantum mechanical


12-30).

decrease in

1.2

10" 13

calculation uses a

tion for the extra proton


(cf.

The value

is

and takes

cm
more

(16-4')
realistic

charge distribu-

into account the exchange integral

Both factors tend to lower AEC thus requiring a compensating


,

The nuclear forces are definitely not gravitational in nature. From the data presented
1 on the charge and mass of nuclei, the reader may easily show that gravitational forces are too weak to overcome the Coulomb forces by about 35 orders of
t

in section

magnitude.

NUCLEAR

Sec. 3]

SIZES

AND THE OPTICAL MODEL

567

There are several other methods of measuring the size of the nuclear
We shall consider the one which is the most straightforward and accurate. It is the scattering of electrons from nuclei at
charge distribution.

hundred Mev. As nuclear forces do not act between


is due
to the Coulomb interaction between the electron and the nuclear charge
distribution.
Therefore measurements of electron scattering should
provide information about the nuclear charge distribution. The method
can be thought of as the use of an "electron microscope" of extremely high
resolution to "look at" the charge distribution. The resolution is extremely
energies of several

nucleons and electrons, the scattering of an electron from a nucleus

Target

foil

Electron

beam
'

Collimator

^\

Co

or

Magnet

Figure 16-3.

An

apparatus used to measure the scattering of high energy electrons by

nuclei.

high because the de Broglie wavelength X of the electrons

is

extremely

Using the expression p = Ejc to evaluate the momentum of the


electrons, which are highly relativistic because their energy is very large
compared to their rest mass energy m c 2 = 0.51 Mev, we find for 500-Mev
10~ 12 cm as an estimate of
electrons I = 2.4 x 10~ 13 cm. Taking R
the size of the charge distribution, we see that for such electrons 1 <~ RjA
and KR = (2ttIX)R <~ 20. According to the conclusions drawn in the
short.

preceding chapter, the differential cross section for the scattering of such

from the nuclear charge distribution should be very sensitive to


and even to the details of its shape.
shall not attempt to describe the machines which are used to accel-

electrons
its size,

We

erate electrons (or other particles) to the high energies required in these

experiments since, in the space available, we could not do justice to the


subject. f

However,

it

is

worth while to describe the other aspects of

the electron scattering experiments.

Hofstadter and colleagues


electron

beam

is

The apparatus used

passes through a thin

the nuclei to be investigated,J and

it is

foil,

the

work by
The collimated

in recent

indicated in figure (16-3).

atoms of which contain

stopped in an insulated metallic cup.

t Descriptions can be found in a number of elementary textbooks.


An electron of several hundred Mev has such a high relativistic mass that the

scattering

produced by the atomic electrons

is

completely negligible.

THE NUCLEUS

568

The

total charge collected

by the cup

[Ch. 16

a measure of the incident

is

flux.

Electrons scattered into a certain angular range at 6 enter a magnet, which


bends them in a semicircle in the plane perpendicular to the paper. Upon
leaving the magnet, they enter a scintillation detector similar to the old

ZnS

human

eye detector, but employing a

more

efficient crystal

and a

10"

\
10"

10"

10"

.Sic

\A

10"

10' -34

30

40

50

60

80

70

90

Figure 16-4.

upon carbon.

The differential scattering cross section for 420-Mev electrons incident


From R. Hofstadter, Annual Review of Nuclear Science, Vol. 7, Annual

Reviews, Stanford, 1957.

photoelectric

magnetic

cell.

field is a

The radius of curvature of the electron path in the known


measure of the

momentum

or energy of the scattered

electrons.
It is

found that by

energy loss in the

measured for 420-Mev


6

C 12

is

t This

shown
is

far the

CM

LAB

likely process

The

frame and the

LAB

is

scattering with

no

differential scattering cross section

electrons, incident

in figure (16-4).f

actually dajdQ. in the

between the

most

frame.

The

on the

typical small

nucleus

points with accuracy estimates are

frame, but for electron scattering the difference

CM frame

is

quite negligible.

NUCLEAR

Sec. 3]

SIZES

AND THE

OPTICAL MODEL

569

They are evaluated from the


number of full energy electrons
each angle, and the number of nuclei per

the experimental measurements of da/dQ..

measured incident

flux,

the measured

scattered per unit solid angle at

cm2

in the target foil.

That absorption processes are observed to be improbable verifies our


idea that the scattering should be explicable on the basis of a real scattering
potential representing the Coulomb interaction between the electron and
the nuclear charge distribution. That the electron energies are high
suggests the additional idea that the Born approximation should be useful
in the analysis of these experiments. It actually is found that for nuclei
of small A (and therefore small nuclear charge and scattering potential)
the Born approximation is capable of giving a reasonably good account of
the measured da/dQ if cognizance is taken of the fact that the electrons
are highly relativistic. This is done by using free particle eigenfunctions
obtained from the Dirac equation, instead of the Schroedinger equation,
to evaluate the matrix elements entering in the Born approximation.
The results of this relativistic Born approximation are

where

E is

da

(Ze2 \2 cos2

dD

\2e)
\2E)

(0/2) \F( X )\

si
sin*
(6/2)

The form

the electron energy.

(16-5)

Z2

factor F{%)

is

"

F(X)

r
f p(
Jo

)!!L2T Airr2 dr

(16-5')

with

where

= 2K sin (0/2),

scattering
is

p(r) is the nuclear

as in the case of the


(cf.

* = 7^4

charge density in units of proton charges.

form factor that

arises in the discussion of

equation 14-35), at small angles F(%)

essentially unity at small scattering angles.

writing

2E

en

~ 2mc

for the highly relativistic

~ Z.

Just

X-ray

Also, cos 2 (0/2)

Keeping these points in mind,


electrons, and then comparing

equation (16-5) with equation (15-94), we see that at small scattering


angles da/dO. of the former equation is equal to the Coulomb differential
scattering cross section for a point charge nucleus.
angles,

dajdQ.
2

is

reduced below this value because

<

The

At

large scattering

\F(%)\

< Z2

and

term describes the effect of the finite


extension of the nuclear charge; the second term describes a relativistic
effect which may be attributed to the electron spin.
The behavior of dajdQ, as calculated from equation (16-5), depends in a
quite sensitive manner on the assumed form of the nuclear charge density

because cos (0/2)

first

THE NUCLEUS

570

[Ch. 16

Thus a comparison of the calculated and measured values of dajdQ


allows a determination of this quantity. The dashed curve in figure
(16-4) shows the best fit obtained for dajdQ calculated in the Born

p{r).

The

approximation.

from a phase

shows dajdQ. calculated


using the same p(r) as was used in the Born

solid curve in the figure

shift analysis

approximation calculation.

Of

course, the phase shift analysis

upon the Dirac equation, and not upon

2
r in 10

Figure 16-5.

In figure (16-5)

same

p(r) is

we

7,

based

cm

The nuclear charge density of carbon. From

of Nuclear Science, Vol.

the

3
_13

is

the Schroedinger equation.

R. Hofstadter, Annual Review

Annual Reviews, Stanford, 1957.

plot the p(r) used in these calculations.

found

in the analysis of dajdQ.

measured

of electrons of several different energies from the

As

essentially

in the scattering

and as the

nucleus,

analyses are sensitive enough to distinguish variations in, for example,


p(0) of perhaps

10 percent, there

is

considerable confidence that this

must be quite close to the actual charge density of the nucleus.


For a given electron energy, the differential scattering cross sections
measured for nuclei of larger A are found to be similar to those measured
p(r)

for C, except for a decrease with increasing


first

minimum

This

means the
is

partial

in the angle at

found, and a development of subsidiary

The experience gained

larger angles.
this

is

size

which the

maxima

at

in the previous chapter tells us that

of the charge distribution increases with increasing A.

seen in figure (16-6), which plots charge densities obtained by a

wave

analysis of the differential scattering cross sections

for electrons of several energies

The forms of

and

for a

number of

measured

different nuclei.

For inand heavier nuclei,


is known to be constant only to within 10 or 20 percent.
However, it is
assumed that the central density is constant, and that the form of p(r)
p(r) given in the figure are

not completely unique.

stance, the charge density in the central region for Ca,

NUCLEAR

Sec. 3]

AND THE

SIZES

OPTICAL MODEL

571

can be described by the empirical equation

With this assumption, the parameters a and b are accurately determined


by the analysis of the scattering measurements to be

a=
b

l.01A 1A

10" 13

cm
(16-7)

0.55

x 10" 13 cm

0.10

E 0.08

a>

0.06

0.04

0.02

rin 10
Figure 16-6.

The nuclear charge

Review of Nuclear Science, Vol.

7,

"cm

density of several nuclei.

From

R. Hofstadter, Annual

Annual Reviews, Stanford, 1957.

The parameter
it

p(0) is not determined very accurately by this analysis, but


can be simply by demanding that the integral of the charge density over

volume equal Z. Figure (16-6) consists of plots of equation


(16-6) using the values of the parameters so determined.
the nuclear

We
1.

draw

the following conclusions

The charge

density of nuclei, which

tion of the protons in the nuclei,

region and
2.

The

falls fairly

The

must be essentially the distribuapproximately constant in the interior

rapidly to zero at the surface.

half-density radius a increases with increasing

A'A law.
3.

is

fall-off distance

is

according to an

approximately the same for

all nuclei.

THE NUCLEUS

572

[Ch. 16

The interior charge density p(0) decreases slowly with increasing A.


If we assume that the distribution of protons in nuclei is approximately the same as the distribution of neutrons (there is good evidence for
4.
5.

this

assumption), the charge density p(r) of a nucleus

density p

M {r)

is

related to

its

- PM (r)

p(r) cc

(16-8)

slow decrease of p(0) with increasing A


Z/A with increasing A (Z/A

If so, the

with the decrease in

1/2.5 for

density p

M (0)

mass

by the equation

completely consistent

is

1/2 for

~ 240),

is

and with the assumption that the


approximately the same for all nuclei.

~ 40;

interior

ZjA
mass

coefficient of A Vz in the expression for a is 1.07, and


equations (16-6) and (16-7) are actually consistent with (16-4)
and (16-4') because, within the accuracy of the analyses, both lead to

Although the

6.

not
the

1.2,

same root-mean-square

criterion for comparison.

radius,

which can be shown to be the proper

Consequently,

this provides support for the


assumption of the equality of the forces between various nucleons, which
is the basis of the mirror nuclei analysis that leads to (16-4) and (16-4').

Let us turn now to measurements of nuclear sizes from experiments


which use nucleons as probes instead of electrons. Certain contrasts
between these experiments and the electron experiments immediately

One

arise.

forces,

is that a nucleon interacts with a nucleus by means of nuclear


and not by means of Coulomb forces (although there is also a

Coulomb

interaction if the nucleon

is

a proton).

of nucleons with nuclei involve a different

measure a

set

Since the interactions

of forces, these interactions

different set of attributes of the nuclei,

and

it

should not be

the measurements yield nuclear sizes which are not the

same
measured in the electron experiments. Another point is that the
interaction between a nucleon and a nucleus must be described by a
complex interaction potential. A real potential is not adequate because
surprising

if

as those

there

is

appreciable absorption

(cf.

section

experiments show that, when a nucleon

is

7,

Chapter

incident

upon

15)

probability for absorption, including scattering with energy loss,


quite

since the

a nucleus, the
is

comparable to the probability for scattering with no energy

generally
loss.

In fact, the interpretation of the earliest experiments of this type was

based on the assumption that, in its interaction with a nucleon, a nucleus


behaves like a sphere of radius R which is completely absorbing. The
experiments were measurements of the total cross sections a T of nuclei
for neutrons of energies near 20 Mev. For a neutron of this energy, the
de Broglie wavelength is X
6 x 10 -13 cm. Using i!~5x 10~ 13 cm

as a reasonable estimate of the radius of a nucleus of intermediate A,

Sec. 3]

NUCLEAR

it

OPTICAL MODEL

573

find KR = {^.7tJX)R
5.
We have then the case of a sphere which
assumed to be completely absorbing, with large KR. Consequently
should be possible to apply equation (15-104), which is

we
is

AND THE

SIZES

aT

The

total cross section

aT

attenuation technique that

is

= 2nR

(16-9)

measured by

used for X-rays

essentially the

same beam

Chapter 14),
with a neutron detector consisting of a mixture of some hydrogenous
compound and some compound that scintillates. Neutrons entering the

Figure 16-7.

The

is

12

16

20

24

28

(cf.

32

section

40

36

total cross sections of several nuclei for

detector collide with protons of the hydrogenous

5,

26-Mev neutrons.

compound,

transfer

energy to the protons, and these charged particles subsequently produce


scintillations.
A number of different measurements have been made;

shows some data for 26-Mev neutrons. The solid curve is


equation (16-9), using R
1.55^ x 10" 13 cm. However, this is no
longer accepted as a reliable estimate of the nuclear size. More sensitive
figure (16-7)

measurements of the interaction between nucleons and nuclei show that


the analysis based

KR ->

oo is

much

upon a completely absorbing sharp-edged sphere with


too crude to yield accurate results.

These more sensitive measurements are of the differential scattering


cross sections of nuclei for nucleons. In the past few years many accurate

measurements have been made for both neutrons and protons, and for a
and nuclei. The experimental arrangements used are
basically the same as that used in the electron scattering experiments.
For protons, the energy sensitive detector can be a magnet plus scintillavariety of energies

tion counter, or simply a scintillation counter using a crystal such as Nal


is a linear function of the proton energy. Some data for

whose response

THE NUCLEUS

574

0"

Figure 16-8.

protons.

The

The

25"

50

75"

[Ch. 16

100 125 150 175

differential scattering cross sections of several

nuclei for

points are experimental and the curves are theoretical

Glassgold, Revs. Mod. Phys., 30, 419, (1958).

fits.

I7-Mev

From A.

E.

NUCLEAR

Sec. 3]

AND THE

SIZES

OPTICAL MODEL

17-Mev protons are shown by the points

575

Following
frame of the observed

in figure (16-8).

convention, this figure plots the ratio in the

CM

differential scattering cross section (scattering

with no energy loss in the

CM frame) to the Coulomb differential scattering cross section for a point


charge nucleus (equation 1 5-94).

da/dQ

It

should be noted, then, that the observed

actually decreases rapidly with increasing 6

owing

-4
to the sin
0/2

dependence of the Coulomb da/dQ.

| -9

Imaginary

s -55

CL

Figure 16-9.

The

real

Real

and imaginary parts of the optical model potential for the scatter-

ing of 17-Mev protons.

The solid curves are the best fits to the data obtained from a partial
wave analysis using equation (15-102). In this analysis the interaction
potential is taken to be of the form

The first term is a complex

(r-a)/ b

ydr)

(16-10)

interaction potential representing the refracting

and absorbing properties of the nucleus for the incident proton. The
second term is the real Coulomb potential acting upon the proton and
is evaluated from the nuclear charge density determined by electron
scattering. For the analysis of the 17-Mev proton data, the following
parameters were used in the first term

V = -55Mev
W = -9 Mev

a=

1.25

0.65

The

real

,A

x 10"13 cm

(16-10')

x 10" 13 cm

and imaginary parts of the interaction potential are indicated

in

figure (16-9).
It is

evident that a negative

the nucleus
tial

is

for

an unbound proton moving through

reasonable in light of the fact that the real part of the poten-

must, by definition, be attractive for a bound proton moving in the


It is even more evident that a negative
Q is reasonable since, as

nucleus.

THE NUCLEUS

576

[Ch. 16

indicated in (15-106), to have absorption the imaginary part of the

must be negative. But the very form of the interaction potential


it is only an approximation to the true potential
acting on the proton moving through the nucleus. This is because the
true potential must certainly be a function of all the distances between
the proton and the nucleons of the nucleus (or at least a number of these
distances), and not just a function of the single distance between the
proton and the center of the nucleus. Thus the approximation replaces
the very complicated true potential by a much simpler potential which,
potential

makes

it

evident also that

50 100 200

500

(Mev)
Figure

16-10.

The approximate energy dependences of the depths of the

real

and

imaginary parts of the optical model potential.

presumably, represents the average

effect

of the true potential

a procedure

quite similar to that used in the Hartree treatment of the atom.

This

approximation is called the optical model of the nucleus. It was originally


used around 1930, with a purely real potential. This was not adequate,
and so the optical model fell into disfavor. However, with a complex
potential it has experienced a very successful revival at the hands of
Feshbach, Porter, Weisskopf and Saxon in the years after 1952.
As is true for any model, the justification of the optical model is that
it

provides a simple picture of certain processes and enables a large

number

of experiments involving these processes to be explained in terms of a


small number of adjustable parameters. For instance, the set of parameters
which provide an acceptable fit to the 17-Mev proton differential scattering
cross sections also provides an acceptable fit to neutron total cross sections
at a similar energy. In fact, the single set of parameters a and b quoted in
(16-10') is found to give an acceptable fit to the scattering and total cross
sections for both protons and neutrons over a wide range of nuclei and

At a given energy, a single set of parameters V and


Q also is
adequate to explain both cross sections for both particles and all nuclei.
do exhibit a dependence on the energy E of the
However V and
proton or neutron, as indicated in figure (16-10). There are still certain

energies.

NUCLEAR

Sec. 3]

AND THE OPTICAL MODEL

SIZES

ambiguities in the values of

V and

some

since, in

577

cases, different values

of these parameters can give equally acceptable fits to the data. But the
general behavior of their energy dependence is fairly well established, as

For

are their values at low values of E.

E is

instance,

known

it is

that

when

2 or 3 Mev,

can only be 1 or 2 Mev, because very pronounced


resonances are observed in the scattering and total cross sections at these
values of E. Resonances (cf. section 6, Chapter 1 5) involve constructive
interferences between the radially inward traveling and radially outward
traveling components of some partial wave. Therefore they can occur only
if the outward traveling component is not absorbed and, as shown by

equation (15-106), small absorption means small

and the rate of absorption


of particles in the complex optical potential, which is of a form different
from (15-106). Consider a particle of mass
moving with total energy
E along the x axis through a medium of constant optical potential
(V + iWo)- The reader may verify by substitution that a solution to
There

is

an

between

interesting relation

the time independent Schroedinger equation for this situation


ip(x)

-S/2A
e~

is

(16-11)

where

(3-

AM
2

\W

//

(E+ini)

+ 1-1

(16-11')

and

K 2 = ^( +
Note

that in the limit

V,\)

\W

where

E+\V
\I(E

|K

|)

(16-11")

is

negligible

compared

to

1,

equation (16-11") reduces to the usual expression for K. When this


quantity is only small compared to 1, equation (16-11') reduces to

2(E+\V

A:
2|wg

To

identify

\)

(16-12)

A, calculate the probability density


y>*(x) yi(x)

-iKx -x/2A iKx -x/2A __ -x/A


a
e~

We see that the probability density dies out exponentially with an attenuation length equal to A. Now let us evaluate A for E = 20 Mev, using the
values of V and W taken from figure (16-10). We have: E = 20 Mev,
v = 45 Mev, W = 10 Mev, M = one proton mass. For these values
|

(16-12) gives

A=
This

is

3.3

lO" 13 cm,

E=

20

Mev

(16-13)

about one-half the average distance across a nucleus of average

THE NUCLEUS

578
radius.

Thus the probability density associated with a

energy following such a path through a nucleus

is

[Ch. 16

particle of this

attenuated by about

a factor of e 2 and the particle has only about a 10 percent chance of not
being absorbed. Nuclei are quite opaque to nucleons of energies in the
range 20 Mev (and higher). | The situation is quite different at E = 2 Mev.
,

At

this

energy

F =
|

50 Mev,

A=

20 x

|I

=
rr
WA
0]

10" 13

1.5

Mev, and we obtain

E=

cm,

a^l.07A*x

10

-13

Mev

(16-14)

cm

a^l.25A H x 10- 13 cm
Figure 15-11.

comparison between the nuclear mass distribution and the nuclear

interaction potential.

This

about three times the average distance across a nucleus of average


The probability density for a particle of this energy following such
a path is attenuated only by a factor of about e /s and the particle has
about a 75 percent chance of not being absorbed. Nuclei are fairly
transparent to nucleons of energies in the range 2 Mev (and lower).
Let us compare the results of the experiments that measure the charge
or mass distribution with the results of the experiments that measure the
distribution of the real or imaginary interaction potentials. This is done,
for a typical nucleus, in figure (16-1 1). The figure plots the mass distribuis

radius.

tion,

and the absolute value of the

real

interaction potential.

It

is

extremely interesting to note that the spatial distribution of the nuclear


interaction potential exerted
t This

is

by a nucleus on a passing nucleon extends

part of the explanation for the large values of

20-Mev neutron

total cross sections with equation (16-9);

found by analyzing the


is high enough

the opacity

that there is appreciable probability of absorption even when the neutron passes through
the outer nuclear surface. The rest of the explanation is that (16-9) assumes that the

de Broglie wavelength of the neutron is negligibly small. Actually it is not, and, since
the neutron can be absorbed if it passes within a de Broglie wavelength of the nucleus,
this also increases the

from

(16-9).

absorption probability and, therefore, the value of

R calculated

NUCLEAR

Sec. 3]

beyond the
<5

SIZES

AND THE OPTICAL MODEL

579

spatial distribution of the nucleons in the nucleus

which, for a typical nucleus,

is

only about

by a distance

x 10~ 13 cm.t This means

that nuclear forces must be of very short range. Thus nuclear forces are
certainly very different from the long range gravitational and electric
forces of

common experience. An electron passing 2

the surface of a nucleus feels an appreciable

passing 2 or 3
essentially

x 10~ 12 cm from

no force

at

Figure 16-12.

Illustrating the

and a proton.

Of

Coulomb

the surface of the

x 10~8 cm from

force ;

a neutron

same nucleus

feels

all.

typical nucleus

composition of the total

course, a proton passing 2 or 3

nucleus does feel a

or 3

Coulomb

force.

10~ 12

This force

real potential acting

cm from
is

between a

the surface of the

strong, but not so strong

would feel on passing a little closer. An


comparison of the relative strengths of nuclear forces and
Coulomb forces can be obtained by plotting the real nuclear potential,
and the Coulomb potential acting on a proton, for a typical nucleus and a
typical proton energy. This is done in figure (16-12). Because of the
as the nuclear force the proton
instructive

extension of the nuclear charge distribution, the Coulomb potential


does not continue its Ze2 /r behavior for
less than the nuclear radius.
For 20-Mev protons incident upon 92
nuclei, the maximum positive

finite

jr,

113
but more probably i!~l x 10~ 13 cm oc A". This
;
t The figure implies 6 oc A
not inconsistent since the values of a for the mass and (particularly) for the potential
distributions are not known with too much certainty, so that S is the relatively small

is

difference between

two large uncertain values.

THE NUCLEUS

580
value of the total real potential

is

about

Mev and

the

[Ch. 16

minimum

nega-

example of a large Z
nucleus, we see that when r is less than the nuclear radius the magnitude
of the real nuclear potential is somewhat larger than the magnitude of the
tive

value

is

about

45 Mev. Even

for this extreme

Coulomb potential.

4.

Nuclear Masses and Abundances

The detailed determinations of nuclear density distributions obtained


from the recent electron scattering experiments show that the density in
the interior of a nucleus is approximately the same for all nuclei. However,
this was known at an early date because all the analyses of nuclear sizes,
based on the assumption that the nucleus is a sharp-edged sphere of radius
R and uniform interior density, lead to the result R oc A!/z for which the
density p = mass/volume oc AI(A'A ) 3 is independent of A. In this section
we shall study, primarily, the measurements of nuclear masses. Among
other things, these measurements show that over a very broad range the
binding energy per nucleon is approximately the same for all nuclei.
This was also known at an early date and, along with the approximate
equality of the nuclear density,

it

led to the assumption that all nuclei

contain essentially the same material in essentially the same state, and differ

from each other only in the amount of this material they contain. The
assumption formed the basis of the very successful liquid drop model of
the nucleus, which we shall discuss in the next section.
We know that the mass of a nucleus is almost equal to the mass of the
corresponding atom. The masses of atoms of a particular Z, but possibly
a mixture of A, can be obtained to an accuracy of several significant figures
by the chemical techniques and a knowledge of Avogadro's number.
But, for the more accurate determinations needed in the study of nuclei, it
is necessary to use the physical techniques of mass spectrometry or energy
balance in nuclear reactions. Both give information about the masses of
atoms of a particular Z and A. From these masses, the masses of the
corresponding nuclei can be evaluated by subtracting Z times the electron
mass. In doing this, the mass equivalent of the electron binding energy is
neglected. However, even in extreme cases, the error thereby introduced
is less than 20 percent of the experimental uncertainties in the mass
measurements.!

The first mass spectrometer was built in 1911 by Thomson. It used the
same configuration of crossed electric and magnetic fields as the electron
t

For

binding energy is 2 x 10 5 ev
quoted as 238.1234 0.0010 mass units.

S2JJ23S t h e tota j e i ec tron

The mass of

the

atom

is

~ 0.0002

mass

unit.

NUCLEAR MASSES AND ABUNDANCES

Sec. 4]

charge to mass ratio measurements

(cf.

581

section 2, Chapter

3).

beam of

atoms, ionized by electron bombardment, was passed through these fields


which analyzed it into components according to the different charge to
mass ratios present in the beam. It was usually possible to determine the
degree of ionization of a component and thus determine

somewhat low accuracy of the apparatus.

the

Thomson

its

With

mass to within
this

apparatus

discovered the existence of isotopes. Using a source containing

Source
Region of
parallel

Boundary of

E
_

to'

region of

perpendicular

paper

to

paper

Photo
plate

Figure 16-13.

An

apparatus used to measure atomic masses.

component with mass corresponding


component with A = 22. A number
weaker
associated
and
an
20,
of tests proved that these were both due to a noble gas, and this could only
be Ne, which has a chemical atomic weight of 20.18. Thomson interpreted
his results to mean that there are two chemically indistinguishable species
of Ne atoms, called isotopes, one with A = 20 and relative abundance of
about 91 percent, and one with A = 22 and relative abundance of about
a mixture of noble gases, he found a
to A =

9 percent.
exactly the

They are chemically indistinguishable because they have


same structure of atomic electrons since their nuclei have the

same Z, but they

are physically distinguishable because they have different

have different A. The nuclei of the Ne isotopes


the second occurs with relative abundance of
about 0.3 percent and could not be detected with Thomson's apparatus.
A typical example of a more recent mass spectrometer is the type designed

masses since
are

10

Ne 20

their nuclei

10

Ne 21

10

Ne22

by Bainbridge (1933) and illustrated in figure (16-13). The source produces ionized atoms with charge +Ze,mass M, and a spectrum of velocities.
These atoms pass through a region of crossed electric and magnetic fields
which act as a velocity filter, passing only those with velocity v for which

ZeE
The terms on

the left

and

HZevjc

right are the electric

and magnetic

forces,

THE NUCLEUS

582

Atoms of velocity

respectively.

magnetic

EcjH

then enter a region of uniform

are bent in a semicircle of radius R,

field,

The

graphic plate where they produce an image.

image

is

2R, where

Solving for

the right

the

is

and

on a photofrom S2 to the

fall

distance

the equation

satisfies

HZevjc

The term on

[Ch. 16

Mv*/R

mass times the

centrifugal acceleration.

M, we have

= RZeH = RZeH*

M
c

Ec

The mass can be determined from absolute measurements of the


on the right side of the last equation. In addition, much use is

quantities

made

of hydrocarbon molecules to calibrate the apparatus over a wide

range of masses in terms of the carefully standardized masses of 1 1 and


6 12
With these techniques, extremely accurate measurements can be
C
.

A40 is quoted as
= 39.975022 0.000029 amu

made. As an example, the mass of

18A 4o

An amu

is

one atomic mass

unit.

80 i6

18

It is

defined such that

16.000000

amu

(16-15)

(Note that this differs slightly from the chemical mass scale which uses the
normally occurring mixture of s O isotopes as a standard.)
Mass spectrometers, using detectors which are more linear than photographic plates, can provide accurate determinations of the relative abundances of the various isotopes. As an example, the abundances of the
normally occurring mixture of

i6

17

isotopes are

known

to be

= 99.759%
= 0.037%
18 =
0.204%

Another technique of mass determination, which provides an accurate


supplement and check for the technique of mass spectrometry, is the study
of the energy balance in nuclear reactions. Consider a nuclear reaction
such as (16-1),
2

the one

first

He4

+ N 14 ->7

observed by Rutherford.
a

where a is the bombarding

17

is

The general

+ A-+B +

particle,

+ H

case

the target nucleus,

may

be written
(16-16)

B is the residual

NUCLEAR MASSES AND ABUNDANCES

Sec. 4]

583

and b is the product particle. As mentioned in section 9, Chapter 1


found that in these reactions mass and kinetic energy are not separately
conserved, but instead in any frame of reference there is conservation of
total relativistic energy, E = T + mc 2 where T is kinetic energy and m
is rest mass. For the reaction (16-16), the conservation of total relativistic
energy in the LAB frame of reference reads

nucleus,
it is

(J.

where
where

+ ma c*) + m A c* =

+ mB c*) +

(TB

+m

<Jb

c*)

(16-17)

Ta and ma are the kinetic energy and rest mass of a, and so forth, and
TA = since A is stationary in the LAB frame. Because there

o
Before

Figure 16-14.

After

Illustrating

the kinematics of a nuclear reaction.

can be an exchange of energy between kinetic energy and


in these reactions,

be greater, or

it is

less,

than the

TB + T Ta is called
b

the

initial

kinetic energy

of the reaction. That

Q = TB
From

rest

possible for the final kinetic energy

Ta

mass energy

TB + T

The

to

difference

is,

+T-T

(16-18)

(16-17) this can also be written

Q=

+ m A - mB - w )c2

{ma

(16-19)

We see that a measurement of the Q of a reaction gives information about


The Q can be
However, for TB this is usually
difficult and, fortunately, can be avoided by using a relation that comes
from the conservation of momentum. Consider figure (16-14), which

the rest masses of the entities involved in the reaction.

measured by measuring

Ta Th
,

and

TB

LAB frame. By
momentum, an equation is obtained
eliminate TB from (16-17). This is quite easy

represents the kinematics of the reaction as seen in the

applying the conservation of linear

which makes

it

possible to

in the limit

Tjma c 2
where the

<

TJm.c 2

classical expressions

<

such as

1,

TB \mB c* <

Ta = \ma t%

and pa

= ma va can be

THE NUCLEUS

584

we

Consequently,

used.

shall leave

it

as

[Ch. 16

an exercise for the reader to show

that in this classical limit

=T

(l
\

This

is

- Ta (l
+ ^)
m B>
\

Tm m
-^]m Bi mB (T
a

)y>

cos 6

(16-20)

of sufficient accuracy for the analysis of nuclear reactions at the

most experiments. Equation (16-20)


can be determined by a measurement of Ta and Tb

energies which have been used in

shows that the


[Even

if

the reaction

is

being studied in order to find the value of

terms of the other masses, in practice this quantity would already be

with sufficient accuracy for use in (16-20)

and

mJm B <

in typical cases.]

particularly since

The measurements of Ta and

mB

known

b /m B

in

<

use the

techniques employed in the scattering experiments.


In equation (16-19), the masses refer to the rest masses of the nuclei

and B, and to the rest masses of the completely ionized nuclear particles
a and b. However, to the accuracy of the approximation in which the mass
equivalent of the electron binding energy

is

ignored, this equation can also

be considered to read

Q = (Ma +
where the large

refer to the

M A - MB - M

b )c

rest

mass energy of an

+ Z )mc

+ Z A )mc

from the

last

This procedure

Za + Z A
must be

electron.

(16-21)

masses of the neutral atoms.

obtained from equation (16-19) by adding (Za

terms and subtracting (Z B

=ZB + Z

to the

two, where

This
first

mc 2

is

is

two
the

valid since the relation

is

(16-22)

true in any nuclear reaction in order to have conservation of

charge. Rutherford's reaction (16-1) provides a

good example of the use

of (16-20) and (16-21) to supply a connection between the atomic


masses involved in a reaction. The measurements show that for this
reaction

Q =

equivalent,
units, or

1.18 Mev. This energy can be converted to its rest mass


0.00126 amu, by dividing by c 2 and converting to the proper

by using the

relation
1

amu <->

which comes from evaluating the


1 amu.
Then we have

0.00126

rest

Mev

(16-23)

mass energy of a

Mmu M
atomic masses, say M

amu

This allows one of the

931.14

terms of the other three.

2He 4

particle of rest

SQl ,

ao n, to

The

Mm

mass

be determined in

analysis of the energy balance in a large

NUCLEAR MASSES AND ABUNDANCES

Sec. 4]

585

number of reactions has provided measurements of masses which accurately check the measurements by mass spectrometry. Furthermore, the
agreement between these two methods provides excellent confirmation of
the relativistic theory of mass and energy, upon which the energy balance
analysis

is

based.

many atomic masses that have


been measured, and also the mass of the neutron. Note that
z A the
In table (16-1) are listed a few of the

TABLE

(16-1)

Atomic Masses and Binding Energies


Binding Energy (Mev)

V
w
H

Mass (amu)

1.008982 (3)
1.008142 (3)

2.014732 (4)
3.016977 (11)

He3
2
He4

Be 9

8Q16

16

16.00000

29CU 63

29

63

62.94826

50

50
74
92

120

119.94012

184

184.0052

238

238.1234

Sn 128

74\\tt84

92IJ238

Total

Per Nucleon

(AE)

(AEIA)

2.22

1.11

7.72

2.57

4.003860 (12)
9.01494 (16)

28.3

7.07

58.0

6.45

(0)
( 20)
(72)
(11)
(10)

127.5

7.97

552.1

8.75

mass of an atom of given Z and A,

1020

8.50

1476

8.02

1803

7.58

is always quite close to A amu. This is a


= 16.00000. If the amu
consequence of defining the amu such that
a 16
=
which
might
1.00000,
at first seem more
were defined such that
1X
close
A
amu
for
atoms of large A.
would
not
be
to
natural, then
zA
The reader can demonstrate this by doing a little arithmetic.
The same procedure will also demonstrate the much more fundamental
point that the mass of an atom is less than the mass of its constituent
parts.
For example,
2A = 4.003860 amu, whereas the mass of its
constituent parts is the mass of two 1 1 atoms plus the mass of two neutrons, i.e., 2Mia + 2M01 = 4.034248 amu. This mass deficiency of the

atom

is

due to a mass deficiency of its nucleusf resulting directly from the


We can see this by considering

equivalence between energy and mass.

t In the comparison just made, two electron rest masses are included in both
2i4
and 2M + 2M0jl and the mass equivalent of the electron binding energy is negligible.
Thus the mass deficiency can only be due to the nucleus.

THE NUCLEUS

586

[Ch. 16

any one of the four nucleons in the 2 He 4 nucleus. Since the nucleon is
stably bound to the nucleus, it must be moving in some sort of an attractive
real potential representing the net attraction of the other three nucleons.

Furthermore, to be bound
situation

is

it

must have a negative energy E < 0. The


The minimum energy required to

depicted in figure (16-15).

remove the nucleon from the nucleus, leaving it a free nucleon at r ->- oo,
is \E\. Conversely, if a free nucleon comes in from r -> oo and combines
with the other nucleons to form the nucleus, its energy must decrease by

Attractive potential

Figure

nucleon

the

16-15.
in

schematic representation of the potential and total energies of a

a helium nucleus.

amount

\E\.

(The excess energy

or by some other mechanism.)

will

be carried off by a

gamma

ray,

Since the same situation obtains for the

other nucleons of the nucleus, we see that when a dispersed system of free
nucleons combines to form a nucleus the total energy of the system must
decrease by an

amount AE,

the binding energy of the nucleus. According


in the total energy of the system must

to equation (1-24), the decrease

AE

AM in mass, where
AM = AE/c
(16-24)
deficiency AM
4.034248 - 4.003860 = 0.030388

be accompanied by a decrease

its

For 2 He4 the mass


amu. Therefore the binding energy
,

we have used

(16-23).

This figure

is

is

is

AE = AMc 2 =

listed in the

28.3

Mev, where

next to last column of

The last column lists AEjA, the binding energy of the nucleus
number of nucleons it contains. For 2 He 4 this average
binding energy per nucleon is 28.3/4 = 7.07 Mev.
One of the most important features of a nucleus is its average binding
table (16-1).

divided by the

energy per nucleon. This quantity

is

plotted as a function of

in figure

The points are the data obtained from the measured masses in
the manner just described. The smooth curve is obtained from an equation
to be described later. Note that AE/A at first rises rapidly with increasing

(16-16).

A, and then becomes reasonably constant at a value

AE/A

~ 8 Mev

(16-25)

Sec. 4]

NUCLEAR MASSES AND ABUNDANCES

However,
about 8.7

Mev

587

not completely constant at this value;

it maximizes at
and then drops slowly to about 7.5 Mev for
A
240. One consequence of this slow drop in AE/A is the famous
phenomenon of nuclear fission, in which a large A nucleus, such as 92 U 238
splits into two intermediate A nuclei.
This happens because the final
state is more stable than the initial state, since the average binding energy
it

is

for

~ 60

9r-

20

40

60

80

100

120

140

160

180

200

220

240

A
Figure 16-16. The average binding energy per nucleon for a number of nuclei. The
smooth curve is from the semi-empirical mass formula. From R. B. Leighton, Principles
of Modern Physics, McGraw-Hill Book Co., New York, 1959.

per nucleon increases from the value

A
A

nucleus to the value

~8.5 Mev

~7.5 Mev

characteristic of the large

two intermediate
The energy liberated is about 1 Mev per nucleon, or about 1 Mev
per nucleon x 200 nucleons = 200 Mev in total.t Alpha particle emission
is a special case of fission in which one of the final nuclei is 2 He 4 and it
takes place for the same reason. The phenomenon of nuclear fusion
consists of the combination of two or more nuclei of very small A to form
characteristic of the

nuclei.

a larger nucleus which has a higher average binding energy per nucleon,
and therefore is more stable, because its value of A is nearer the value
14
Btu, which does not seem like much energy in terms of this
t 200 Mev = 3 x 10
macroscopic unit. However, the mass, 200^14^ = 200 x 1.7 x 10~ 24 gm = 7 x 10~ 25
lb, of the system emitting this energy is not very large either. The energy emitted per
pound, 3 x 10" 14 /7 x 10- 25 = 4 x 10 10 Btu/lb, is phenomenal; it is about 10" times

larger than the energy emitted per

pound from burning

coal.

THE NUCLEUS

588

A =

AE/A

60 for which

maximizes.

that only a few nuclei near

A =

It

might seem from

60 would be

this discussion

This

stable.

[Ch. 16

not true

is

because there are other factors which tend to inhibit fission and fusion.

Except for A 5 and A


beyond 200, which are

may

lifetimes that they

A from 1 to
have such immeasurably long

there are nuclei for all values of

8,

either stable or

be considered

stable.

100

90

.jr

80

pip
70

60
IIBM^I

50

^
-^

40
30

20
10

\?

10

40

30

20

60

50

80

70

90

100

110

120

130

140

N = (A - Z)
The

Figure 16-17.

distribution of stable nuclei.

Z and (A Z)
have seen that this information can be
obtained from the mass spectrometer measurements. For instance, these
measurements show that for Z = 8 there are stable nuclei for A = 16, 17,
It is

interesting to consider the distribution of the

We

values of these stable nuclei.

The data are plotted in figure (16-17). Each


whose abscissa is (A Z) = N,
the number of neutrons in the nucleus, and whose ordinate is Z, the
number of protons in the nucleus. Note that:
18 (the isotopes of

stable nucleus

1.

for

is

For each

Z=

isotopes

there are generally several values of N.

10 there are stable nuclei for


10

Ne20

10
,

For small
2b. For large
2a.

O).

indicated by a square

Ne 21

Z
Z

10
,

there

Ne 22
is

N=

10, 11,

and

12;

For

instance,

these are the

Z = N.
due to Coulomb

a tendency for stable nuclei to have

stable nuclei

have

Z<

N.

This

is

THE LIQUID DROP MODEL

Sec. 5]

589

repulsions between protons which produce a positive energy proportional

Z2

Coulomb effect were operating, the most stable nucleus


would be obtained for Z = 0, N = A. However, the data
show that for small Z, where the Coulomb effect is small, the tendency
for Z = N dominates and stability is obtained for Z
N. For large
Z the Coulomb effect becomes important and stability is obtained for
to

If

only this

with a given

Z < N.

At

Z=

Z/N =

82,

0.65.

There is also a tendency for stable nuclei to have even Z and even N.
This can be seen from table (16-2), which lists the number of stable nuclei
3.

TABLE

(16-2)

The Distribution of

Stable Nuclei

Number of

Stable

Nuclei

Even

Even

Odd

Odd

Even

Odd

57

Odd

Even

53

166

Even

Odd

of various types. The stable nuclei of odd


7

Nu

i9

40

so

57

La

have even Z
be explained in section

The preceding
free to

i3 8;

Z and

odd

N are H
1

2
,

Li 6 ,

B 10

i76

7i

and Lu
The reasons for tne tendency to
and even N, and also for the tendency to have Z = N, will

23

7.

discussion has been written as

make any adjustments

in

if

a nucleus of given

is

and N, consistent with the condition

Z+ N=

A, which are required to achieve the condition of stability.


not poetic license. Such adjustments can occur by the emission of
electrons or positrons from the nucleus (beta emission), or by the capture
This

is

of atomic electrons in the nucleus. These processes, which will be discussed


in section 11, allow for changes in

constant; that

5.

is,

and

such that

Z+N

remains

they convert protons to neutrons or vice versa.

The Liquid Drop Model and the Semi-empirical


Mass Formula

We shall now use the liquid drop model of the nucleus, and the information
obtained from the study of the distribution of Z and
values for stable

THE NUCLEUS

590

[Ch. 16

nuclei, to obtain a formula for the masses of these nuclei. This model is
based upon the facts that for all nuclei, except those of very small A, the
interior densities are approximately the same and the binding energies
are approximately proportional to the masses of the nuclei {iS.EjA

Both

can be compared with certain others


concerning macroscopic drops composed of some incompressible liquid.
For all such drops the interior densities are the same and the heats of
constant, so Ais oc A).

facts

vaporization are proportional to the masses of the drops.

comparison

is

required to disperse the drop into


it is

The

latter

significant because the heat of vaporization is the energy


its

constituent molecules,

analogous to the binding energy of the nucleus.

and therefore

In developing the

mass formula, we shall use the model to suggest other analogies between
a nucleus and a liquid drop.
The formula contains the sum of six terms,

MZ A =ZMZ,A)
=

(16-26)

The

first

term

is

the

mass of the constituent parts of the atom,

(16-27)
/ (Z, A) = 1.008142 Z + 1.008982 (A - Z)
where the coefficient of Z is the mass of the 1 H 1 atom in amu and the
coefficient of {A Z) is the mass of the neutron n 1 in amu. The remaining

terms correct for the mass equivalents of several


to the total nuclear binding energy.

effects

which contribute
is the term

Of most importance

f (Z, A) =

-M

(16-28)

which accounts for a binding energy essentially proportional to the mass,


or volume, of the nucleus. A similar term would be present for a liquid
drop. The next term is
(16-29)
f2 (Z, A) = +a*A*
a positive correction proportional to the surface area of the nucleus.
In a liquid drop this term would represent the effect of the surface tension
energy; it arises from the fact that a molecule at the surface of the drop
experiences attractive forces only from one side, so its binding energy is
less than the binding energy of a molecule in the interior.
Therefore
simply setting the total binding energy proportional to the volume of the
drop overestimates the binding energy of the surface molecules, and a
correction must be made which is proportional to the number of such
molecules, or to the surface area. The term

f3(Z,A)=+a 3

(16-30)

THE LIQUID DROP MODEL

Sec. 5]

591

Coulomb energy of the charged nucleus, which is


assumed to be a sphere of radius proportional to A y%
A similar term
would be present for a charged liquid drop. The last two terms bring in
accounts for the positive

properties specific to a nucleus.

have

Z=

For one thing, there


N. This can be accounted for by the term

UZ,

A)

is

the tendency to

= +a 4 (Z ~f l2f
A
- Z), or 2Z = A, but

(16-31)

zero for Z = N = {A
which is otherwise
and increases with increasing departures from that condition.
The form used in equation (16-31) is about the simplest one having these
properties, but there is also some theoretical justification that will be presented in section 7. The tendency to have even Z and even N is accounted
for by the term

which

is

positive

- Z) = N even
[Z even, {A - Z) = N odd
\Z odd, (A-Z) = N even
Z odd, (A-Z) = N odd

-f(A),

f5(Z,A)

0,

+f(A),

even,

{A

(16-32)

which decreases the mass if both Z and


are even, and increases it if
both Z and
are odd. The form of the function/^) is determined by
fitting the data. It is found that the best fit for a simple power law is

obtained with

f(A)

a5 A-*

(16-32')

Gathering together equations (16-26) through (16-32'), we have

MZ A = 1.008142 Z + 1.008982 (A - Z) - a A + a^A


x

a s Z 2A~ 1A

ai (Z

AfrfA- 1

\a 5 A-'A

(amu)

(16-33)

This

is called the semi-empirical mass formula because the parameters a


x
through a5 are obtained by empirically fitting the measured masses.
A formula of this type was first developed by Weizsacker (1935). Determinations of the parameters have since been made on several occasions.

Green (1954) found

= 0.01692
= 0.01912
a3 = 0.000763
a4 = 0.10178
a5 = 0.012
ax
a2

(all in

amu)

(16-33')

THE NUCLEUS

592

Using these parameters, quite a good

fit

[Ch. 16

obtained for the measured

is

masses of all stable nuclei except those of very small A the discrepancy is
generally less than several thousandths of an amu. A comparison is shown
;

smooth curve shows the average bindfrom equation (16-33), and the points
show the same quantity evaluated directly from the measured masses.
The semi-empirical mass formula is extremely useful because it describes
fairly accurately the masses of several hundred stable nuclei, and many
more unstable nuclei, in terms of only five parameters.

in figure (16-16).

In that figure the

ing energy per nucleon evaluated

6.

Magic Numbers
The

drop model gives a good account of the average behavior of


and therefore in regard to stability. However,
show significant departures from
nuclei with certain values of Z and/or
this average behavior by being unusually stable. These values of Z and/or
are the magic numbers
liquid

nuclei in regard to mass,

2, 8, 20, 28, 50, 82,

The

situation

is

analogous to the unusual

126

stability

(16-34)

of the electronic systems

of the noble gas atoms containing a magic number

But

electrons.

stability are

2, 10, 18, 36, 54,

86 of

in the nuclear case the effects indicating the unusual

not so pronounced as in the atomic case, and it is necessary


them in order to demonstrate conclusively the "magic"

to consider all of

character of the numbers (16-34).

Mayer

These considerations, largely due to

(1948), run as follows:

1
There is a tendency for nuclei to prefer magic Z and/or N. This can
be seen from consulting figure (16-17). For example, there are 6 stable
isotopes for Z = 20, whereas the average number of stable isotopes in

that region of the periodic table

is

about

stable isotopes, although the average

N=

There are 7 stable nuclei having


in that region of
is only about

3.

For

2.

number

Z=

50 there are 10

is about 4.
though the average number
number of similar examples can

in that region

82, even

be seen.
2.

Figure (16-16) shows that the average binding energy per nucleon
equal to 2 and 8, than
with Z and/or

significantly higher, for nuclei


is

for neighboring nuclei.

Z=N=
stability

These

2.

more

effects

sensitive

is
it

The outstanding example is 2 He4 for which


are even more pronounced if a measure of
,

than the average binding energy per nucleon

is

energy of the "last" neutron or proton


b the binding
in the nucleus, which is the minimum energy required to separate a
considered.

This

is

MAGIC NUMBERS

Sec. 6]

593

neutron or a proton from the nucleus. As an example, the binding energy


of the last neutron in 2 He 4 (the energy required to produce the reaction
2
He 4 -* 2 He 3 + n 1) is 20.6 Mev. The binding energy of the last proton
in

He 4

19.8

is

Mev. These are abnormally

high.

Figure (16-18) shows

a plot of the binding energy of the last neutron for a

The

nuclei.

ordinate

is

abscissa

number of neutrons

the

number of

heavier
the

in the nucleus;

the difference between the actual value of the binding energy of

and the value predicted by the semi-empirical mass


is a smooth function of N (except for the

the last neutron

The

formula.

is

predicted value

+3

+2
3

+1

T
bq

.*

BS

_2

-3

20

28

40

50

1
1

60

80 82

100

120

126

140

N
Figure 16-18.
J.

The binding energy

of the last neutron for a

number

of nuclei.

From

A. Harvey, Phys. Rev., 81, 353 (1951).

effects

of term 16-32), and

intermediate
nuclei with

to

N=

around

28, 50, 82,

large energy required to


3.

it

decreases slowly from around 7

Mev
126

remove

for large N.
is

The unusual

36

17
,

Kr87 and 54 Xe 137


,

of

their last neutron.


less

evidence, such as the fact that for three of the four


8

for

demonstrated by the exceptionally

There are a number of other somewhat

neutron emitters,

Mev

stability

convincing pieces of

known spontaneous

N equals a magic number plus

1.

This implies an unusually small affinity for the extra neutron.

The analogy between nuclear and atomic magic numbers prompted


people to look for an explanation of the nuclear phenomenon
analogous to the explanation of the atomic phenomenon. The reader will

many

key point in that explanation is the formation of closed


by the electrons moving independently in the atomic potential. But,
when the nuclear magic numbers were first being discussed seriously
(1948), it seemed very difficult to understand how any explanation based
upon independent particle motion could be valid for the nucleus. The
reason was that the very successful liquid drop model had been dominant
for a number of years, and it seemed basic to this model that a nucleon in a
recall that the
shells

THE NUCLEUS

594

[Ch. 16

its neighbors. With strong interactions,


would be constantly scattered in traveling through the nucleus
and would follow an erratic path, resembling Brownian motion much
more closely than the motion of an electron moving independently

nucleus interacts very strongly with

the nucleon

through

its

orbit in a closed shell.

From our present

point of view there

is,

however, no problem in under-

standing the independent particle motion of nucleons in a nucleus.

In

a very reasonable extrapolation of the optical model (1952).


The calculations recorded in equations (16-13) and (16-14) show that
fact, it is

A of a nucleon moving through a nucleus increases


by about a factor of 6 as the total energy of the nucleon decreases from
E = + 20 Mev to E = +2 Mev, and that A is several times nuclear
the attenuation length

dimensions at the lower energy.


through E
to E < 0, predicts

sponding to

E<

0,

An

extrapolation of this behavior,

that, for the

the attenuation length

is

that the

bound nucleons can move

nucleus.

Putting this in other words, a nucleon

with

E>

must have a

corre-

quite independently through the

moving through a nucleus

certain probability of suffering collisions, which

involve energy loss in the frame in which the

moving

bound nucleons

very long and, therefore,

CM of the nucleus

is

at rest,

complex potential. However, the energy dependence for E >


of the real and imaginary parts of this potential, indicated
by figure (16-10), makes it easy to believe that for E <
there is no
imaginary part to the potential, so that there can be no collisions with
energy loss in the
frame of the nucleus. But, since essentially all
collisions between a nucleon and another nucleon involve energy loss in
this frame, it follows that if the potential is purely real for the bound
nucleons there can be no such collisions between these nucleons, and that
they must therefore move independently through the nucleus.
because

it is

in a

CM

7.

The Fermi Gas Model

It is comforting to realize that the optical model analysis, of the experimental data concerning the motion through the nucleus of nucleons with
should move independE > 0, strongly suggests that nucleons with E
ently through the nucleus. But it is even more comforting to realize that

<

why this should happen. The


based on the Fermi gas model of the nucleus, which had been
in existence for a number of years, but which had not been held in very
high regard during the period of dominance of the liquid drop model.
This model is essentially the same as the Fermi gas model of the conduction
there

is

a simple theoretical explanation of

explanation

is

electrons in a metal considered in section 6, Chapter 12.

It

assumes that

Sec. 7]

THE FERMI GAS MODEL


move

the neutrons and protons

595

in

an

attractive potential of nuclear

dimensions, and that in the ground state these nucleons

fill the energy


such a way as to minimize the total energy but not violate the
exclusion principle. (Remember that both neutrons and protons are Fermi
particles with spin quantum numbers i
\, just like electrons.) Figure

levels in

(16-19) indicates the energy levels filled by the neutrons in the ground state

of a nucleus. Since a proton

is

we must imagine
energy levels

from a neutron, the excluon the two types of nucleons, and

distinguishable

sion principle operates independently

a separate and independent diagram representing the

filled

by the protons in the ground

state

Figure 16-19. A schematic representation of the energy levels


the ground state of a nucleus.

filled

of the nucleus.

by the neutrons

in

immediately apparent from these diagrams why there can be essentially


between the nucleons when the nucleus is in its ground state.
The point is that all the levels which are energetically accessible are already
It is

no

collisions

and so there can be no collisions except those in which two nucleons


of the same type exactly exchange energy. But the net effect of such an

filled,

exchange of two indistinguishable particles is the same as if there had been


no collision at all. We can understand, then, why for the ground state
of the nucleus the potential is real, and the nucleons are moving independ-

We can also understand how the exclusion principle can be effective


even in inhibiting collisions between a nucleon of low positive energy and
the bound nucleons of a nucleus. In fact, calculations based on the Fermi
gas model have shown that the decrease in the imaginary part of the optical
ently.

model potential with decreasing energy (the energy dependence of


in figure 16-10) is due primarily to the exclusion principle.
Furthermore, the Fermi gas model provides a good estimate of the depth
of the real potential for the ground state of a nucleus. This is obtained
by evaluating the Fermi energy (12-38), which is

in terms of the nucleon mass


and the nucleon density p. Let us consider
the Fermi gas of neutrons in a uniform spherical nucleus of radius

R=

A lA

THE NUCLEUS

596

For a

number of neutrons

typical nucleus the

N=

[Ch. 16

is

0.6/4

Thus

N
iirr^A
gives

0.6A

0.45

1.33ttt$A

nr\

* A

Eh

Ef

-vn
Figure 16-20.

and

the

between the depth of the

Illustrating the relation

energy, and the binding energy of the

Fermi energy

last

potential, the Fermi

neutron.

is

Ef =

A7TYs

h\0A5f

(16-35)

2M; .M_3

Using a radius constant r = 1.25 x 10~ 13 cm consistent with the optical


model potential radius of (16-10'), and evaluating the other parameters,

we

obtain

Ef

~ 34 Mev

relations between the depth of the potential V the Fermi energy Et


and the binding energy Eb of the last neutron are shown in figure (16-20).
As mentioned above, Eb is measured to be roughly 6 Mev for a wide range
of nuclei. Thus the Fermi gas model predicts

The

in satisfactory

V = Ef + Eb
\

^ 34 + 6 = 40 Mev

agreement with the value of about 50

by extrapolating to
(extrapolating the

E<

the values obtained

Mev

that

from the

is

obtained

optical

model

of figure 16-10).

= finds a simple explanation in


The tendency for nuclei to have
terms of the Fermi gas model. Consider a nucleus of very small Z, where

this effect

dominates because the Coulomb potential is very small. In


two independent Fermi gases, the neutrons and the

this nucleus, there are

THE FERMI GAS MODEL

Sec. 7]

moving

protons,

597

same nuclear

in the

The energy

potential.

levels

of these

systems are indicated in figure (16-21). Except for the effect of a very small
positive

Coulomb

potential acting

on the proton system, the energy

levels

of the two systems are identical because for both systems the nuclear
potential

is

the same. It

the nucleus

is

clear, then, that for

is

minimized

if

a given

the levels are filled with

the total energy of

Z=N

since, if this

condition were violated, nucleons would occupy levels which are of higher

Neutrons
Figure 16-21.

protons

'

'

Protons

schematic representation of independent Fermi gases of neutrons and

in a nucleus.

energy than necessary.

quantitative treatment of this argument leads

to the term (16-31) used in the semi-empirical

mass formula to express


According to the equations preceding (16-35),
the Fermi energy for the neutron and proton systems can be written
the t tendency for

Z=

N.

Efn =

C(N\A)K,

Eu =

C(ZjAfA

where C is a constant. Now use (12-40) to evaluate the total energy of the
neutron and proton systems, as measured from the bottom of the potential,
in terms of E and E
This gives
f
f
.

ZA

E =C'^where C"

terms, which

3C/5.

is

The energy of the nucleus

a function only of Z and

C'A-K[N iA

already know, the

condition (16-37),

minimum

is

the

sum

of these two

is

Z*]

minimum of the

found for

(16-36)

since

N+Z=
As we

is

E(Z, A)
This

'

(16-37)

function (16-36), subject to the

N = Z = A/2.

In departing from this

the energy of the nucleus increases, and the corresponding

THE NUCLEUS

598
correction in the semi-empirical
increase.

mass formula

will

[Ch. 16

be proportional to

this

Thus

/4(Z, A)

oc E(Z,

A)

(Z, A) min

C'A~ 2A {N

iA

2[A/2]*}

Z*

2|>4/2]}

Now let

D^*^ = N-* = -Z
2

Then

/4(Z, A)
Use Taylor's

oc

A~ 2A {[AI2 + DfA +

series to

expand the

keeping terms through

Z>

2
.

(-D)]*

two quantities

in the curly bracket,

This gives

*A

+ D

first

[All

[C

-H

5
+ -D

L2.

3 3 2 L

^HMfr+f^

L2

2.

5 2

(-D)2

33

so
'

/4(Z, A)

oc

A-\2

5 2

+ --D
33

cc

A -H

%
2

^4

L2J

- lA

3 3

oc

2
Z2
^-Z) 2 = ^/ ~ >

in agreement with (16-31).

The Fermi gas model


of nuclei to have even

also gives a partial explanation for the tendency

Z and even N.

Consider the even Z, even


nucleus
represented in figure (16-21), and imagine constructing the nucleus of
next higher A by adding one nucleon. To minimize the total energy, the

go into the lowest available energy level. This might be either


level, but let us assume that it is a neutron
level, as shown in the figure. Then this nucleus will contain one additional
neutron. Now construct the next nucleus in the series by adding yet
another nucleon. The nucleon will also be a neutron, since there is still one
vacancy in the neutron level and this level is still of lower energy than the
unfilled proton level. We see that for only one of the three nuclei involved
in this dicussion is
or Z odd, and for none of them are both N and Z odd.
This explanation is indicative of what is happening, but it is not the whole
story. The mass measurements, which demand the term (16-32) in the
semi-empirical mass formula, show that there is also a pairing energy that
nucleon

will

a neutron level or a proton

THE SHELL MODEL

Sec. 8]

599

increases the binding energy of a nucleus

when

NorZ becomes even.

The

origin of the pairing energy lies in the fact that the force between

two
an energy
level, the attractive force which it exerts on the nucleon that was already
there produces a negative potential that lowers the energy of both these
nucleons

nucleons,

is

attractive.

i.e.,

When

the second nucleon

added

to

lowers the energy level in question.

Neutrons

Figure 16-22.

is

Illustrating

why

there are

Protons

'

more neutrons than protons

in

a nucleus of

large Z.

we present figure (16-22), which


form of the potentials acting on the protons and neutrons in

Before leaving the Fermi gas model,


indicates the

a nucleus of intermediate or large Z.

Because of the Coulomb potential,

on the protons is elevated. However, the highest


essentially the same energy for both the neutrons and

the total potential acting


filled level

must be

the protons

if

at

the nucleus

explanation of the reason

is

to be stable.

why

there are

This provides a very graphic

more neutrons than protons

in a

nucleus of large Z.

8.

The

Shell

Model

The Fermi gas model establishes the validity of treating the motion
of the bound nucleons in a nucleus in terms of the independent motion of
each nucleon in a common net potential V(r), which represents the average
attraction of all the other nucleons. In this model, certain quantum
mechanical properties of a system of non-interacting Fermi particles
are brought in by the use of equation (12-38), and the Schroedinger equais not solved explicitly. This is adequate for the description
of certain average properties of the nucleus but, for a more detailed
description capable of accounting for the magic numbers, it is necessary
to bring in more of the important quantum mechanical properties of such

tion for V(r)

a system by actually solving the Schroedinger equation for

V(r).

The

THE NUCLEUS

600

[Ch. 16

treatment of the nucleus employing this procedure is called the shell


model of the nucleus. The reader will note that the shell model of the
nucleus is related to the Fermi gas model of the nucleus in much the same

way

that the Hartree theory of the

theory of the atom.

He

will also

atom

is

related to the

Thomas-Fermi

note that these nuclear models are

cruder than the corresponding atomic theories.

much

In the self-consistent

atomic theories the net potential V(r) is completely determined by the


theory; in the nuclear models the radius and exact form of the net potenV(r) must be inserted ad hoc.
The procedure of the shell model

tial

is

to find the neutron

and proton

energy levels for the potential V(r) of a particular nucleus by solving the
Schroedinger equation for this potential, and then to construct the nucleus

by

these levels in order of increasing energy with

filling

Z protons.

N neutrons

and
depend

Just as in the Hartree theory, the energies of the levels

on the quantum numbers n and

/,

and each

corresponding to the 2 possible values of

level has a capacity 2(2/

and the

(21

1)

1)

possible

was hoped that a form for the potentials V(r) of the


various nuclei could be found in which the ordering and spacing of the
energy levels would be such that an unusually tightly bound level would
completely fill in those nuclei having magic N or Z. However, it was
found that there is no reasonable form for the V(r) which leads even to the
values of

It

ffij.f

proper ordering of the energy

The

levels.

was provided by Mayer, and independently by Haxel,


Jensen, and Suess (1949). These workers proposed that, in addition to
solution

the potential V(r), there

portional to

this interaction

is

a strong inverted spin-orbit interaction, pro-

means that
would be predicted
the net nuclear potential and

L, acting on each nucleon in a nucleus. Strong

much (about 20

is

times) larger than

by using equation (13-27), equating V(r) to


m to the nucleon mass. Inverted means that the energy of the nucleon is
decreased when S L is positive and increased when it is negative, so that
the sign of the spin-orbit interaction is inverted from the sign of the
magnetic spin-orbit interaction experienced by an electron in an atom.
However, as the magnitude of the spin-orbit interaction is proportional
to S L just as it is for an electron, the magnitude of the spin-orbit splitting

will

be approximately proportional to the value of the quantum number

just as

it is

for

The quantum numbers

momentum

/,

an electron.}

vectors,

n,

/,

mu m

s,

as well as the other

quantum numbers, angular

and associated terminology, have the same

significance as in the

theory of the electronic structure of atoms.


t

This

is

the

the potential

quantity y

Lande

is
1

is

interval rule (13-43),

which depends only on the assumption that

L, plus the statement that since j


approximately proportional to /.

proportional to

=
/

\ the

Sec. 8]

THE SHELL MODEL

601

S>

CI

4
?

14
2
6

4
8
10

Figure 16-23.

184

126

126

112
110
104

100
92

70
68

64
58

10
2

50
40

38
32

28

28

20

20

16

14

2Pi/s-

2p

3d

Is

,/2

Without

With

S-L

S-L

The ordering

-3d 5

shell

c
o

2sv

Is

1:

184
168
164
162
154
142

=cr~

+
16

Id

r-t

50

of the highest filled nucleon energy levels according to the

model.

The net

effect

figure (16-23).

of the proposed spin-orbit interaction

The part on

the

left

is

indicated in

shows the ordering and approximate

spacing of the energy levels which nucleons are filling in nuclei with
form of square wells with rounded edges, such as are

potentials V(r) in the

used in the optical model

the part

on the

right

shows the way these levels

THE NUCLEUS

602

[Ch. 16

by the spin-orbit interaction. This is not an energy level diagram


any particular nucleus; it is a diagram giving the order in which levels
appear as the radius of the potential increases with increasing A. Thus
it gives the order in which the highest energy levels of the nucleus are filled.
The diagram is analogous to table (13-2) giving the ordering of the highest
are split
for

filled electronic

energy levels of the atoms, except that

it

also indicates

the relative magnitudes of the separation between adjacent levels.

the presence of the spin-orbit interaction,

and

quantum numbers, and n, /, j, m s must be used.


the levels depend on j as well as on n and /, the

are

In

no longer good

Also, the energies of


larger j corresponding

Each of these levels


number of possible values of m,. This
is shown in the first column on the right in the figure. The second column
shows the total capacity of the levels up to and including the level in
question. The third column shows the same thing for each level which
to the smaller energy since the interaction

has a capacity

lies

(2/

1)

will

inverted.

equal to the

unusually far below the next higher

which

is

level.

Since these are the levels

be unusually tightly bound, we see that the

shell

model with

strong inverted spin-orbit interaction predicts precisely the magic numbers


(16-34).f

Within the past few years, completely independent verification of the


existence of a spin-orbit interaction has been obtained

from experiments
It is found that

involving the scattering of low energy nucleons by nuclei.

a scattered nucleon can have

its

spin angular

momentum

polarized as a result of interacting with the nucleus,

vector partially

and

this

explained only by the presence of a spin-orbit term in the


describing the interaction.

can be

potential

Furthermore, the sign of the polarization

shows that the sign of this term is the same as that required by the shell
model, and the amount of polarization shows that its strength (the spinorbit potential

is

the shell model.


for nucleons

is

several
It

Mev

deep) agrees with the strength required for

should be emphasized that the spin-orbit interaction

not electromagnetic in origin. Instead

nuclear forces, whose origin

is

it is

an attribute of

not well understood.

The spin-orbit shell model can do much more than explain the magic
numbers and their consequences. For instance, it can also explain the
spin angular momentum of the ground states of almost all the nuclei.
t It should be mentioned that the eigenfunctions obtained in solving the Schroedinger
equation for the potential V(r) show that the probability densities for nucleons in the
eigenstates of this potential do not have the property of being large only in a fairly
restricted range of r. Thus in the nuclear shell model there are no "shells" of fairly

which the nucleons are moving. The contrast in this point


between the behavior of the nucleus and of the atom is due to the difference in the form
of V(f) for the two systems.

restricted thickness in

THE SHELL MODEL

Sec. 8]

603

Consider nuclei for which both


and Z are magic, such as 8 16 ^Ca40 ,
208
Pb
According to the model, they will contain only completely filled
subshells of neutrons and protons, and the exclusion principle therefore
,

82

demands
and

that for both the neutron

orbital angular

and

momenta couple

the proton systems the total spin

agreement with the


For nuclei such as 7 N 15 8 17
19
K 39 82 Pb 207 and 83 Bi 209 which contain a magic number of nucleons of
one type and a magic number plus or minus one of nucleons of the other
observation that
,

to zero,| in

for these nuclei.

type, the exclusion principle

that the total angular momentum


momentum of the extra nucleon or hole.|

demands

of the system be the total angular

shows that these nuclei should have i = \, f, f, |, and


agreement with observation. Now consider nuclei
f
for which
and/or Z are further away from the magic numbers. These
nuclei will contain subshells with more than one nucleon or hole, and the
problem of how the spin and orbital angular momenta of the nucleons in
these subshells couple is analogous to the same problem for the electronic
structure of an atom. However, there are important differences. Most
atoms seem to be examples of LS coupling, but most nuclei seem to be
examples of JJ coupling. Of course, this is simply a consequence of the
strong spin-orbit interaction in nuclei. It appears that another consequence

Thus
,

figure (16-23)

respectively, also in

that the coupling in nuclei is much simpler than in atoms. Consider


any nucleus with even N and even Z. As mentioned in section 1, all such
nuclei are found to have i = 0. It is clear that this means that, whenever
there are an even number of nucleons in a subshell, the total angular
is

momenta of these nucleons


the JJ coupling

simply couple to zero. If it

is

now assumed that

strong enough that the addition of one

more nucleon to
the subshell does not destroy the coupling of the nucleons that were already
is

there, the total angular

momentum

of the whole subshell will be that of

"odd" nucleon. With this assumption, the entire angular momentum


of an odd A nucleus will be due to the total angular momentum of the single
odd nucleon in the highest occupied energy level, and i will be equal to the
value ofj for that level. With only one or two exceptions, this assumption
allows the observed values of i for all odd A nuclei to be explained in
the

terms of figure (16-23). In doing this

it is, however, necessary to allow for


occasional interchanges of the ordering of certain closely spaced energy

levels,!

The

because of the

effect

of the pairir^g energy.

spin-orbit shell model, plus the additional assumption just

mensomewhat less satisfactory explanation of the magnetic


moments of odd A nuclei. The prediction is that the magnetic moment of
the nucleus is due to the magnetic moment of the single odd nucleon
tioned, provides a

t Cf. discussion associated with table (13-9).


X Recall that this also happens for the atomic energy levels.

THE NUCLEUS

604

[Ch. 16

because the contributions of the other nucleons cancel one another. Now it
is easy to calculate the magnetic moment of a single nucleon of given j in
terms of its orbital g factor (which is 1 for the proton and for the neutron),
and in terms of its spin g factor (which is such as to give z components of

intrinsic magnetic moment


2.79,m y for the proton and
1 .91^
v for the
neutron, where the sign gives the orientation relative to the spin). For

protons, two values are found corresponding to the two possibilities

= +

I
I
f r neutrons, two other values are found.
I an d j
i
plot of these two values of magnetic moment as a function of j forms the

'>

When

so-called Schmidt lines.

neutron Schmidt
it

lines are

N nuclei,

compared with those of even Z, odd

observed values, but that the agreement

N nuclei,

is

far

from

perfect.

of the

This shows

only an approximation to assume that an even number of nucleons

it is

always couple to give zero total angular


is

Z, even

seen that the predictions agree with the general trend

is

that

compared
and the

the proton Schmidt lines are

moments of odd

with the observed magnetic

good enough

spin angular

momentum

because this quantity

is

quantized

so, if there is

number of nucleons occasionally have a


angular momentum, there must also be a corresponding

some probability
non-zero total

momentum. The approximation

to lead to the prediction of correct values of the nuclear

that the even

odd nucleon have


compensate and keep the

probability that the

exactly the right total angular

momentum

total

to

system constant. But,

which

not quantized, can be appreciably different from the predicted


Apparently it does happen.

The
The

momentum of the
moment of the system,

angular

happens, the magnetic

is

value.

9.

if this

Collective Model

model

shell

nucleus

move

is

based upon the idea that the constituent parts of a

The

independently.

liquid

drop model implies

just the

opposite, since in a drop of incompressible liquid the motion of any

constituent part depends on the motion of

The

all

the neighboring parts.

between these ideas is indicative of the current state of the


theoretical understanding of the nucleus. It also emphasizes that a model
provides a unified description of only a limited set of phenomena, completely without regard to the existence of conflicting models used for the
conflict

description of other sets.

Contrast this with a theory, such as relativity

theory or quantum theory.


large set of
sets

phenomena and,

of phenomena,

it

theory provides a description of a very


border lines between this set and other

at the

fuses without conflict into the theories used for

the description of the other

sets.

THE COLLECTIVE MODEL

Sec. 9]

As

in every scientific field,

physics proceeds

on two

605

the theoretical development of nuclear

levels.

On

one

level,

people develop models,

which correlate experimental results and provide simple descriptions of the


phenomena involved. Then they constantly attempt to broaden the range
of applicability of each model, to remove any apparent conflicts between
the various models, and finally to fuse these models into a single unified
model. On another level, people attempt to develop theories, which
explain the properties of the models from the fundamental laws of nature.
a theory of the nucleus would be able to explain the
of the parameters used in the semi-empirical mass
values
numerical
the laws of force between nucleons.
terms
of
in
formula,

As an example,

parallel development of models and theories was not very evident


our discussion of atomic physics, because that is a field in which there
are now acceptable theories. At this stage, the models lose much of their

The

in

importance and so are used in a discussion only if they are of particular


historical interest, or are of pedagogical value, or provide convenient
calculational tools. An example is Bohr's description of atomic structure,
a model which

we discussed at length for precisely these three reasons.


we shall describe some of the attempts which have been

In later sections

and theories of nuclear forces. In


the present section we shall describe an example of one of the attempts to
fuse two apparently contradictory models into a unified model. This
been
is the collective model of A. Bohr and Mottelson (1953), which has
the
and
model
shell
features
of
the
quite successful in combining certain

made

to develop theories of the nucleus

The model assumes

liquid

drop model.

move

independently in a real potential.

that the nucleons in a nucleus

However, the potential

is

not

the static spherically symmetrical potential V(r) used in the shell model;
instead it is a potential capable of undergoing deformations in shape.
These deformations represent the collective motion of the nucleons in the

nucleus

the type of motion associated with a liquid drop.

As

in the shell

which are split


by the same spin-orbit interaction, and form a core containing an even
number of protons and an even number of neutrons, plus (for odd A
model, the nucleons

fill

the energy levels of the potential,

odd nucleon. As is also true in the shell model, the odd


nucleon plays an important role in determining the properties of the
entire nucleus. However, in the collective model the odd nucleon is not

nuclei) a single

entirely responsible for these properties because the core is not inert

and

can have angular momentum, etc. The instantaneous shape of the core
Its departure
is described by the instantaneous shape of the potential.
nucleon
and, in
from spherical symmetry affects the motion of the odd
latter
The
core.
of
the
turn, the motion of this nucleon affects the shape
coupling

is

due to what can be described as a centrifugal reaction exerted

606

THE NUCLEUS

[Ch. 16

by the odd nucleon on the wall of the potential. The net effect is that a
"tidal wave" circulates around the surface of the core, following the
motion of the odd nucleon. As the reader might guess, calculations based
on this model are very complicated. Consequently we shall only describe
the results.
It is found that the collective model preserves all the features of the shell
model which agree with experiment, and also provides new features which

allow this agreement to be considerably extended. For instance, in this


model part of the total angular momentum of the nucleus is carried in
the form of orbital angular momentum by the tidal wave circulating around
the surface of the core.

and

This moving deformation constitutes a current


moment proportional to the associated orbital

gives rise to a magnetic

angular momentum. The proportionality constant is, however, smaller


than the usual expression (11-7 with the sign changed and the electron
mass replaced by the proton mass) by a factor ZjA, which is the ratio of
the number of charged nucleons to the total number of nucleons. Consequently, that part of the spin angular momentum carried by the core
does not contribute the same magnetic moment as it otherwise would,
and the change is exactly what is required to explain the discrepancy
between the Schmidt lines and the measured nuclear magnetic moments.

Another feature which can be explained quite nicely in terms of the


model is the nuclear electric quadrupole moment Q. We have

collective

given a qualitative definition of this quantity on previous occasions

(cf.

moment about

the

section

Quantitatively, the electric quadrupole

1).

z axis is defined

Q =jp(x,

by the equationf

y, z)[3z

(x 2

z )]

dr

Z[_3z~

(x

?)]

(16-38)

where p(x, y, z) is the nuclear charge density in units of proton charges,


and where the integral is taken over the nuclear volume. The value of

Q is Z, the number of protons in the nucleus, times the average over their
2
charge distribution of the quantity [3z 2
(x 2
z 2 )].
It is clear
y

then

that

Q=

that case x 2

must

at least

if

p(x, y, z)

spherically

symmetrical,

since

in

2
2
If p(x, y, z) is not spherically symmetrical, it
y = z
have an axis of symmetry along the direction about which
.

the nuclear spin angular

taken as the

is

z axis,

momentum

vector precesses.

In (16-38) this

and the equation immediately shows

p(x, y, z) is elongated in the z direction so that z 2


t See also the footnote

on page

646.

>

x2

that

2
,

Q>

is

if

and that

Sec. 9]

Q<

THE COLLECTIVE MODEL


flattened in the z direction so that z 5

if p(x, y, z) is

Figure (16-24) shows a plot of

odd A

stable

607

known

the presently

all

< a? =

measured under circumstances in which

nuclei,

+ 14

71

is

175

Lu

+ 13

y*.

values of Q, for

+ 12 i

+ 11

+ 10 +9 -

+8
/

+7 -

t
I

+6

o +5

<

11

T
O-

CO

+4

+3

+2
'I

X
/

*J\1
v v

*"xJ

XI

'!

\l

-1

'

+1 ~

1
i

-2

1*

v
1

X
X

-3

-4

fill
2

20 28

50

82

126

40

50

60

70

80

90

1
1

10

20

30

ZorN
Figure 16-24.

The quadrupole moments

100 110 120 130 140

of stable odd

nuclei.

maximized because the nuclear spin angular momentum vector has


maximum possible component along the z axis (m t = i). The abscissa
is Z for odd Z, even JV nuclei, and the values of Q for these nuclei are
its

indicated with crosses;

and the values of

the abscissa

is

for even Z,

odd N nuclei,
The ordinate

for these nuclei are indicated with dots.

THE NUCLEUS

608
is

the ratio of

Some

to the square of the nuclear charge distribution radius.

of the features of this plot agree qualitatively with what would

be expected from the

Q<

[Ch. 16

shell

an odd Z, even

for

The reason

model. For example,

this

model would predict

N nucleus with Z equal to a magic number plus

that such a nucleus contains only completely filled proton


which the charge distribution is spherically symmetrical,
plus one proton moving in an "orbit" near the x-y plane, for which
1

is

subshells, for

<

x2

even

nucleus with

z2

quadrupole

2
.

The

shell

model would

moment would be due

orbit near the x-y plane.

also predict

equal to a magic

entirely to a

Thus the

fact that

positive to negative values each time the

Q>

for

number minus

an odd Z,

one, since the

proton hole moving

in

an

goes through zero from

number of odd protons goes

through a magic number, and also the concomitant fact that Q goes
through zero in the opposite direction somewhere between each pair of

magic proton numbers, can be understood on the basis of the shell


model.| However, the magnitudes of Q cannot be understood in terms of
this model. According to the shell model, the quadrupole moment of
nucleus is entirely due to the odd proton since the core
any odd Z, even
containing an even number of nucleons has zero total angular momentum
and must, therefore, have a charge distribution which appears to be
spherically symmetric because there is no way to define a symmetry axis.

Now

is

it

apparent from equation (16-38) that the

maximum

possible

due to a single odd proton is about Q = R 2 where R is the


charge distribution radius. But figure (16-24) shows that Q is observed
nuclei. Also,
to be very much larger than this for many odd Z, even
nuclei
in the shell model the quadrupole moment of all even Z, odd

value of

would be very small because

it

could arise only from the very small

displacement of the center of the charged core from the center of the
entire nucleus. However, the figure shows that the value of Q for odd

nuclei

Z, even

is

observed not to depend significantly on whether they are odd


The collective model provides an explanation

N or even Z, odd N.

It leads to large enough values of Q between the


magic numbers simply because the core can be deformed so that many
protons contribute to the total quadrupole moment. As the deformation
can be due to either an odd proton or an odd neutron, the collective
model explains why the values of Q are similar in both cases.

for all these features.

t Note that Q does not behave this way at 20. This, and other recent evidence, has
been interpreted to mean that 20 is somewhat less magic than the other magic numbers.
Also note that at 40 (corresponding to the filling of the ip\$ level) and 14 or 16 (corresponding to the filling of the 3<% or 3si^ levels) Q behaves as it does at magic numbers.
However, 40 and 14 or 16 are not considered magic because none of the other usual
effects is observed.

ALPHA DECAY AND

Sec. 10]

FISSION

609

we have confined our

attention to nuclei which are

Alpha Decay and

10.

Up
stable

to this point

at

Fission

have immeasurably long

least in the sense that they

The

qualification

with

A much

is

A
AEjA maximizes,

~ 60,

larger than the value

energy per nucleon

we saw

necessary because, as

which the average binding

at

are energetically unstable to fission.

Let us

now

decay.

In this particularly important special case of

discuss this interesting decay process, starting with alpha


2

spontaneously emits an alpha particle


greater than the

is

mass

m 2A

sum of

its final

He 4

mass

The energy equivalent of

because

fission,

its initial

mz _ 2A _ i and the

the mass difference

it

Mev

- (M z _

\_M Z A
,

If the semi-empirical
this equation,

or 6

carried

away

can be written

atomic masses as|

mz

alpha particle

is

the electron binding energies, this alpha decay energy

of

a nucleus

mass

Ignoring the mass equivalents of

as kinetic energy of the alpha particle.

in terms of

lifetimes.

in section 4, all nuclei

mass formula

predicts that

2 , 4 )]c

(16-39)
in

slowly with A, reaching values

However,

for the heaviest nuclei.

used to estimate the masses

is

E increases

accurate enough to be useful because

2 ,^_ 4

this prediction is

ignores magic

it

number

not

effects,

which are very important for nuclei of particular interest in alpha decay.
Figure (16-25) shows the values of E for the nuclei in the alpha emitting
range, obtained from direct measurements of the kinetic energy of the
alpha particle with range or momentum analysis techniques, and/or from

We see that the


E rises to a peak about 4 Mev above the general trend,
general trend slowly increases with A as predicted by the

the use of equation (16-39) with measured masses.

alpha decay energy

and that the

The peak occurs

semi-empirical mass formula.


is

easy to understand why.

N=

126.

The

is

82

because

Pb208 which has magic


,

about 4

[Figure (16-18) shows that the increase

Mev

and

it

Z=

82

increase in alpha decay energy arises because

the binding energy of the daughter

and about 2

In emitting an alpha particle, this parent

nucleus decays to the daughter nucleus

and magic

just at 84 Po 212

is

is

We

magic]

Mev higher than average.


Mev because Z is magic

about 2

also see

from

figure (16-25)

that the energies of the known alpha emitters range from 8.9 Mev for
84p 2i2 t0 4i ]yj ev for 90^^232
te, then, that all these energies are far

below the maximum energy of the Coulomb barrier acting on the alpha
particle, which is about 30 Mev. Therefore the alpha particle can escape
t

There

is

no spontaneous emission of

"n 1 ,

1
,

2
,

or

He 3 from

the naturally

occurring nuclei because the values of E, calculated from equations similar to (16-39),
are negative for

all

these processes.

THE NUCLEUS

610

[Ch. 16

the nucleus only by the process of barrier penetration discussed in section


3,

Chapter

8.

In equation (8-58) of that section,

T=

_2

its

the expression

v'ewrtw-^J*

an alpha

for the probability that


potential barrier V(r), if

we obtained

particle of

kinetic energy far

(16^10)

mass
will penetrate the
from the nucleus is E. This

10i

Cf

Fm

260

The alpha decay energy for the nuclei in the alpha emitting range.
Perlman, A. Ghiorso, and G. T. Seaborg, Phys. Rev., 77, 26 (1950).

Figure 16-25.
I.

equation

is

From

actually valid only in the case of emission of alpha particles

of zero orbital angular momentum because only for that case is the time
independent Schroedinger equation (15-76), in the radial coordinate r,
the same as the time independent Schroedinger equation, in the rectangular
coordinate x, which was used in the derivation of (16-40). For the emission
of alpha particles of non-zero orbital angular

of u in equation (15-76)

is

more

negative;

momentum,
this

the coefficient

has the same

effect as

'

ALPHA DECAY AND

Sec. 10]

making the

FISSION

611

and therefore the probability

potential barrier V(r) higher,

T is reduced. However, the reduction is negligible if


momentum is not too large, and usually it can be ignored.

of penetration

angular

the

Let us evaluate equation (16-40) by using the slightly simplified form


In the alpha decay of a parent nucleus

for V(r) indicated in figure (16-26).

of atomic number Z, the potential

is

2(Z

2)e

V(r)

\vw

E
,

Figure 16-26.

The

integral of (16-40)

radius

r'

r'

slightly simplified alpha particle-nucleus potential.

is

taken from the nuclear potential radius

to the

satisfying the equation

2(Z

2>2

Thus

T=
Changing

to the variable

2V2
\/iM
n

fr
J

the integrand becomes sin 2 p

Since Rjr'

may

is

fairly

2(Z

and the

_
2)e

integral immediately gives

-^4-- (?)-(f)(-f) i
,

-2
e

2(.Z-2)e'

I:

p defined by

COS p

T=

'

RV

small compared to

be expanded. The result

T~ e

1,

the terms in the square bracket

is

V 2MB

fM-

H
(f) ]

THE NUCLEUS

612

Evaluating

r',

becomes

this

[Ch. 16

e<-

which gives

1
when
As in

is

~g

2.97(Z -1)

written in units of

(8-59),

we

shall

A li A -3.95(2 -2)E~ A

Mev and R

now assume

is

(16-41)

written in units of 10~ 13 cm.

that there

is

one alpha particle with

energy E, which makes vj2R attempts at the barrier per second, where

There is little on which to base this


assumption because little is known about what an alpha particle is doing
in a nucleus before it is emitted. However, it does allow us to write for
the decay rate X, which is the probability per second that the nucleus will
decay by alpha emission, the simple expression
is

its

velocity inside the nucleus.

(16-42)

2R
Substituting for

In X

J from

(16-41),

and taking logarithms, we have

~ In + 2.97(Z - 2JA R Vi

3.95(Z

2)E~ H

(16-43)

2/v

This makes

it

apparent that X

the coefficient of

is

in (16-42).

form of
show the

actually insensitive to the exact

In figure (16-27), the points

measured values of In X for all the naturally occurring alpha emitters,


and the curve shows the values of this quantity predicted by equation
Equation (16-43)
(16-43), using R = 8 x 10~ 13 cm and Mv 2 /2 = E.
can be plotted, approximately, as a function of (Z 2)E^A alone because
there is essentially no change in In (v/2R), and very little change in

A R'A over the restricted range of nuclei involved. The only


change comes from 3.95(Z 2)E~ A which varies because E varies
from about 4 to about 9. We see that this leads to a variation of about
2.97(Z

2)'

55 in In

X,

corresponding to a variation

in X

of about a factor of e 55

10 25

The correlation between the extremely rapid variation of X and the slow
variation of E was first pointed out by Geiger and Nuttall (1911) from a
compilation of the experimental data. They devised an empirical formula
giving the essential features of the energy dependence of the theoretical

equation (16-43), which is due to Gamow, Condon, and Gurney (1928).


Now consider a system containing N(0) unstable nuclei of the same
species at time t = 0. If the probability per second that any one of the

ALPHA DECAY AND

Sec. 10]

nuclei will decay

time

is

FISSION

the decay rate

A,

the

613

number of

nuclei remaining at

t is

N(t)

and the

N(0)e' u

(16-44)

lifetime r, or average time required for

one of the nuclei to decay,

is

(16-45)

1/1

These relations are (13-116) and (13-117), which were obtained for the
decay of a system of excited atoms by the emission of quanta. But a

30

40

35

(Z-2)E-% {Em
Figure 16-27.

An series;

The

relation

4n

between the alpha decay

2 series;

of Modern Physics, McGraw-Hill

A=

Book

Co.,

4n

New

45

50

Mev)

3 series.

rates and alpha decay energies.

From

R. B. Leighton, Principles

York, 1959.

review of their derivation will show that they apply equally well to any
system where a constant probability per second of decay can be defined.

more

In a

typical system, there are several different nuclear species

For instance, 92 U 234 decays by


alpha emission into
Th
this nucleus decays by alpha emission into
88
Ra 226 this nucleus, in turn, decays by alpha emission into 86 Em222
and this nucleus decays by alpha emission into 84 Po 218 Consequently,

which decay successively into each other.


90

230

a system, initially

filled

with

92

U 234

will eventually contain a

mixture of

all

The decay of such a radioactive series is indicated schematifigure (16-28). The symbol N (t) is the number of parent nuclei

these nuclei.
cally in

is the probability per second for


t, and the symbol X
decay of these nuclei. Similar symbols are used for the numbers and decay

in the system at time

rates of the first daughter nuclei, the second daughter nuclei,

and so

forth.

THE NUCLEUS

614

[Ch. 16

In this system, the relation (16-44)

is valid for the parent only. However,


analogous relations for the daughters can be obtained from the following
set of equations

dN

(t)

= -N (OAo

dt

dNjft)

=N

(t)^

(16-46)

NjHOXj,

dt

^ = Nmi ~ N

dJ

2 (t)X,

dt

(t)

-M>

Figure 16-28.

Ntft)

Illustrating

-^->(

2 (t)

the decay of a radioactive series.

The negative term on the

right of each equation represents the loss of


due to decay into the next nuclei of the series; the
positive term represents the gain of nuclei per second due to decay from

nuclei per second

A number

the previous nuclei of the series.

of interesting solutions exist

for this set of coupled differential equations, depending

conditions and the constants

One

solution of particular importance

small compared to

very

much more

all

the other X.

slowly than

left sides

of

initial

for the case

when

is

very

the parent of the series decays

daughters and, after sufficient time,

all its

all

is

Then

a condition of equilibrium will obtain.


constant, so the

on the

X.

In equilibrium,

all

the N(t) are

the equations (16-46) go to zero.

(This

can be exactly true for all the equations except the first. For the first it
can obviously be only an approximation; but it is a very good approximation if X is very small as assumed.) Applying this condition, we have

= N (t)X = ^(OAi -

NjUt)^.
JV2 (0A 2

which gives

N (t)X =
The equilibrium obtains
independent of the

after a time

initial

= N (t)X =

N^t)^

(16-47)

>

such that Xt
1 for all Xt except X t,
This case is important because the

conditions.

dependence of E, and therefore of

X,

on

is

usually such as to satisfy the

ALPHA DECAY AND

Sec. 10]

FISSION

615

requirement that X be very small compared to all the other X [cf. the part
of figure (16-25) between A = 240 and A = 210]. Note that the relation
(16-47) shows that in equilibrium the number of nuclei of a particular
species
relation

is
is

inversely proportional to the decay rate of that species.

often used in determining the values of the X

of the N(t) and one

We

now

known

This

from measurements

X.

understand how alpha emitting nuclei with


very short lifetimes can be found in nature. For example, 84 Po 212 with
lifetime of about 10~ 6 sec, can be extracted from certain naturally occurare

in a position to

ring minerals

which presumably have been in existence for millions of

Of course, the reason is simply that the short lifetime alpha emitters
members of radioactive series with parents of long lifetime, and are in

years.

are

equilibrium in these

There are three such series which occur


whose parent is 9 Th 232 with lifetime r = 2.01 x
10 10 years, the An + 2 series whose parent is 92 U 238 with lifetime t =
6.52 x 10 9 years, and the An + 3 series whose parent is 92 U 235 with lifetime
t = 1.02 x 10 9 years. The names characterize the values of A for the
nuclei which are members of the series. For instance, the parent of the
4/3 + 3 series has A equal to 4 times an integer plus 3, where the integer is
58. Since each alpha decay reduces A by 4, all the daughters of this series
will also have A equal to 4 times some smaller integer plus 3.
In addition to the An, An + 2, and An + 3 series, there is evidently also
room for a An + 1 series. Actually there is such a series, and its parent is
93
Np 237 with lifetime t = 3.25 x 106 years. The series can be produced
naturally:

series.

the An series

artificially by using a nuclear reaction to make the parent, but it is not


found in nature since the lifetime of the parent is very short compared to

the age of the earth (estimated from certain geological and cosmological
evidence to be of the order of 10 9 years), and therefore any parent nuclei
initially present have decayed away. In this connection note that figure

(16-25) shows that the alpha decay energies of the parents of the three
naturally occurring series

than

Mev

happen

to be particularly low.

If they

were

less

would have lifetimes short compared


to the age of the earth, all the alpha emitters would have long since decayed
away, and the naturally occurring elements would stop at Z = 82 instead
of Z = 92. The same figure makes it clear why the naturally occurring
elements do stop at Z = 92: it is because the alpha decay energies of all
nuclei with Z > 93 are large enough to le.ad to lifetimes short compared
1

higher, these parents too

to the age of the earth.

Finally, a mental extrapolation of figure (16-25)


an actual calculation of the alpha decay energies from the
measured masses, shows that the corresponding elements are apparently

to

Z<

82, or

stable to alpha decay because the alpha decay energies are so small that

the lifetimes are immeasurably long.

Strictly speaking, a

number of

THE NUCLEUS

616

[Ch. 16

Z 82 are unstable, but their lifetimes are so long that decay


cannot be observed.
Alpha decay occurs essentially because the total Coulomb repulsion
nuclei with

energy of the protons in the nucleus

decreased by the emission of a

is

fragment containing two of these protons (the alpha


larger decrease in this energy

is

containing a larger fraction of the protons.

99
A

Figure 16-29.

particle).

But an even

obtained by the emission of a fragment


If the nucleus splits into

w9

schematic representation of the steps involved

two

in a fission process.

Coulomb repulsion energy of the system is obviously


minimized when the parts contain equal numbers of protons. The calculation performed in section 4 shows that, when a nucleus of large Z
undergoes such a fission process, the decrease in the energy equivalent of
the total mass of the system is about 200 Mev. This is almost entirely due
parts, the total

6 Mev

V(0)

\Surface

I i

tension

\
~2200 Mev

^W

Coulomb,
repulsion

V(=o)

Figure 16-30.

An energy diagram

to the decrease in the total

of this energy

is

carried

Coulomb

away by

for a fissionable nucleus.

repulsion energy of the system.

Most

the two fission fragments as kinetic energy.

steps which are involved in a fission process are indicated schematiby the set of drawings in figure (16-29). These drawings define
(somewhat unprecisely) a parameter s characterizing the progress of the

The

cally

process.

It is

convenient to use this parameter and, in particular, tplot


on s. This is shown in

that part of the energy of the system which depends

by the function V(s). Starting at small s, there is relatively


change in the Coulomb repulsion energy with increasing s, but the

figure (16-30)
little

ALPHA DECAY AND

Sec. 10]

FISSION

617

surface area of the nucleus increases rapidly. From the clearly appropriate
point of view of the liquid drop model, the increase in surface area produces an increase in the surface tension energy. Thus V(s) increases with

increasing s for small

As

s.

s continues to increase, the surface tension

"neck down" and eventually to split. After it


the surface tension energy no longer changes with s, and V(s) de-

effect causes the nucleus to


splits,

creases with increasing

s, following the decrease in the Coulomb repulsion


energy of the two fragments. Since V(s) first goes up and later comes
down, it necessarily must pass through a maximum. Calculations based

on the

drop model show that for a

liquid

maximum

typical nucleus of large

this

about 6 Mev above K(0). We already know that K(0) is


about 200 Mev above K(oo).
Figure (16-30) shows that nuclei are normally stable to decay by fission
since they are sitting, with total energy E = K(0), at the bottom of the
is

depression in the potential V(s).


penetration but, because the mass

The process can take place by

M entering

(16-40) for the barrier penetrability

penetration

extremely small.
this spontaneous fission process,

theless,

energy

is

spontaneous
is

fission

is

in the

barrier

exponent of expression

very large, the probability of barrier

For instance, if 9 2 U 238 decayed only by


its lifetime would be ~10 16 years.
Never-

can actually be observed, because so much


it is very easy to detect.
One

liberated in the fission process that

technique often used involves collecting the large number of ions produced
fission fragments as they come to a stop in the gas filling of a device

by the

called an ionization

process of

chamber.

much more importance

is induced fission.
Usually this is
brought about by the nucleus capturing a low energy neutron. As the
binding energy Eb of the last neutron in a nucleus of large Z is around

5 or 6

Mev,

energy to put

in favorable cases the capturing nucleus receives


it

over the top of the fission barrier.

Very often

enough

this

energy

actually does go into vibrations of the type which lead to fission. Induced
fission is perhaps the best example of the collective motions that
are
implied by the liquid drop model. The process is illustrated in terms of

an energy diagram in figure (16-31). For 92 U 233 and ^U 235 the binding
Eh is enough to make the total energy E higher than the top of
the barrier. For 92 U 238 it is not quite enough, and the neutron must also
carry in about 1 Mev of kinetic energy. The difference between the two
,

energy

cases reflects the fact that

capture a neutron,

when

is

about

b,
1

for the nuclei

Mev

formed when

greater than

92

U 233 and

for the nucleus

92

u 235

formed

92(j238

(16-32)

is

ca ptures a neutron. The reason is that the pairing energy


negative in the first case and zero in the second case. Fission

can also be induced by the capture of a gamma ray of sufficient energy, or


by the capture of a high energy proton (a low energy proton will not be

THE NUCLEUS

618

captured because

about 15

The
arises

Mev

it is

repelled

from the nucleus by a Coulomb

[Ch. |6

barrier

high).

produce power in a chain reaction


3 neutrons are emitted, on the
process. This can be understood by considering

possibility of using fission to

from the circumstance that 2 or

average, in each fission


figure (16-32).

The

figure plots the

An energy diagram

Figure 16-31.

Figure 16-32.

are the

most

Z and N

Illustrating

values of the nuclei which

illustrating induced fission.

why neutron

emission occurs

in fission.

stable for each value of A, as in figure (16-17).

These nuclei

are represented by the shaded line of stability. The large dot indicates the
fissioning nucleus. The two small dots indicate the two fragments which
result

from the

fission.

These fragments have the same

fissioning nucleus but, since their values of

ZjN

are smaller

ratio as the

by a factor of

ZjN ratios are smaller than those of stable nuclei with these values
A the fission fragments have too many neutrons. Most of the necessary adjustment in the ZjN ratio takes place by the beta decay process to

2, their

of

be discussed in the next section, but part of it is achieved simply by emitting


a few neutrons. These neutrons typically have kinetic energies of several
Mev. However, they rapidly lose energy by processes involving scattering

ALPHA DECAY AND

Sec. 10]

FISSION

619

with energy loss in the material containing the fissioning nuclei and are no
longer able to induce fission in 92 U 238 This is why that isotope is usually
.

not directly useful in a chain reaction. The slowing down of the neutrons
emitted in fission is, however, not undesirable since, when their energy

becomes so low that


of the

92

de Broglie wavelength

their

is

larger than the radius

by these nuclei increases


with increasing wavelength or decreasing energy.| For a chain reaction
nuclei, the cross section for capture

10

t lO"

10~ 2

10"

10"

10"

60

80

100

120

A
Figure 16-33.

The

140

160

180

distribution of the masses of fragments emitted in fission by thermal

neutrons.

it is

important to capture as

many of the

neutrons as possible; the reaction

one of the neutrons emitted in each fission


is later captured. In reactors, advantage is taken of the energy dependence
of the capture cross section by scattering the neutrons from some material
of low atomic weight so that they lose energy by recoil and eventually
come into thermal equilibrium at the very low kinetic energy %kT~ 0.025
ev. This is not done in a bomb because it takes too long (~10~ 6 sec).

is

self-sustaining only if at least

amount of

92

U 238

can be used indirectly to


These nuclei capture neutrons and, instead of fissioning,
the resulting unstable nuclei 92 U 239 undergo two successive beta decays,
turning into the stable nuclei 94 Pu 239 which have the same fission properties
In breeder reactors a certain

produce

as

fissions.

92
U 233 and 92 U235
We have simplified our discussion of fission by speaking as if the process
.

always leads to two equal fragments.


infrequently.

Actually this happens only very

Figure (16-33) shows the distribution in the

t This should be apparent from the discussion of the previous chapter.

values of

THE NUCLEUS

620

from

distribution

typical.

to

N=

is

U 235 by

fission of 92

the fragments

neutrons of thermal energy. This

Since one peak occurs near

50, while the other occurs

near

[Ch. 16

values corresponding

values corresponding to TV

82,

thought that these bimodal distributions arise from the influence of


the magic numbers. However, the details of their origin are not well

it is

understood.

II.

Beta Decay

We

have also simplified our discussion of the radioactive series by


fact, an important
feature of these series. This is illustrated by the processes occurring in the
An series, plotted in figure (16-34). In addition to alpha decay processes,
ignoring the processes of beta decay, which are, in

90^h 228

90
d-

^R a 224

88

/ \J
89^ c 228

86

E m 220

85

At

N
84

84
*;

82

82

j>b

208

212

&

\j/1
j

.j

208

/ X? >/\

/ \}/ ^

/ Jy \V
81

216

pj*

M/

Th 232

88jla 228

*/

86

90

B AW

82
pb 212
*Bi 212

80

78
124

126

128

130

132

134

N
Figure 16-34.

138

136

The decay processes occurring

in

is

142

144

the 4n series.

there are beta decay processes in which a nucleus Z,

charged electron and

140

*-

transformed into the nucleus

emits a negatively

Z+

1,

A.

Thus,

by 1, N decreases by 1, and the nucleus


becomes the next nucleus lying on a line of slope 135 in the Z versus
plot. (There are also other types of beta decay processes that will be
described later.) In alpha decay processes, Z decreases by 2,
decreases
in these processes

increases

Sec.

by

2,

BETA DECAY

1]

621

and the nucleus becomes the next but one nucleus

slope 45 on the

The reason why

versus

on a

line

of

these beta decay processes occur in the An series, as well

as in the other series, can be understood

Z and

on a plot of the

lying

N plot.
by superimposing figure (16-34)
most stable for

JV values of the nuclei that are the

each value of A, as shown in figure (16-35). As the slope of the line of


stability is appreciably less than 45, beta decay processes must occur in
100

80

Line of stability

~\ rf

vim
An

r-

series

60
*
f
is!

40

20

20

J_
40

60

80

N
Figure 16-35.

Illustrating

why

beta decay occurs

120

100

140

160

*-

in

the decay of a radioactive series.

order that the average slopes of the radioactive series be essentially the
same -as the slope of the line of stability. The necessity of this simply
reflects the

requirement that each

member of

the radioactive series be

"as stable as possible."

The decay of a radioactive series involves an interplay between the two


competing processes of alpha decay and beta decay. Each member of the
series will decay by the process having the largest decay rate. (The last
statement is essentially correct if one decay rate is very much larger than
the other. This is usually the case, but sometimes the decay rates are
comparable, both processes occur, and the series branches. Cf. 84 Po 216
and 83 Bi212 in the An series.) We have seen that the decay rate in alpha

decay depends strongly on the energy available for the decay.


see that this

To

Mz A

is

discuss the energy available in beta decay,


,

We

shall

also true in beta decay.

the mass of

atoms of a given

Z and A,

it is

convenient to plot

as a function of

for fixed

A. Typical plots are indicated in figure (16-36) for odd A, and in figure
and/or magic Z, the masses are
(16-37) for even A. Except near magic
described quite well by the semi-empirical mass formula (16-33). For odd

arises

ZiA lie on a slightly skew parabola. The parabolic shape


from the Z dependence of the term ( 1 6-3 1 ) the skewness arises from

A, the values of

THE NUCLEUS

622

[Ch. 16

the Z dependence of the terms (16-27) and (16-30). For even A, there are
two parabolas corresponding to the two possible values of the term (16-32).

The

odd Z, odd N;

elevated parabola corresponds to

Odd

the depressed

A
=Q

f6 (Z,A)
Beta stable^

Figure 16-36.

L_l

(for

L_J

all

Z)

L_^

*-

The masses of atoms with

a given

odd value of A.

Even

f(Z,A)= +f(A)
(for

odd Z)

f5 (Z,A)=-f(A>
'

(for

even Z)

Beta stable

X
Figure 16-37.

The masses of atoms with

a given

even value of

A.

parabola corresponds to even Z, even N. From the plot for odd A, we see
that for a given odd value of A there is only one nucleus which is stable
to beta decay the nucleus of the atom of smallest mass. (Very rarely an

apparent exception to

this rule is

found when the difference between the

Sec.

BETA DECAY

1]

623

masses for two adjacent values of Z at the bottom of the odd A parabola
happens to be extremely small.) From the even A plot, we see that there
are generally two nuclei with a given even value of

A which are stable to

beta decay, and that these two stable nuclei have even values of
differ

by

Figure (16-38) indicates a plot of

example

A =

found for

is

Z which

Occasionally there are three beta stable nuclei for even A.

2.

MZiA representing

this situation.

An

96.

stable

Figure 16-38.

even value of

Illustrating a case in

which there are three beta stable nuclei for a given

A.

Nuclei of given values of A, whose values of Z are such that they are not
stable,

radioactive series.
electron

by

1.

Z values

can change their

decay processes. One


is

is

As

to attain stability

by three

different beta

the process of electron emission that occurs in the

stated above, in this process a negatively charged


1 and N decreases
and positron emission.

emitted by the nucleus, so that Z increases by

The other two processes

are electron capture

In the former the nucleus captures a negatively charged atomic electron,

and in the latter it emits a positively charged positron. In both, Z decreases


by 1 and N increases by 1 The energy available for electron emission by
.

the nucleus Z,

is

E = [MZ A ,

The energy

ZA
,

is the electron rest mass.

fraction of
1

Mev. Of

by that nucleus

Mev

to

course, in

MZ_

(16-49)

by the nucleus

is

2m]c*
(16-50)
X A
The energy E ranges from a small
,

more than 10 Mev and


all

is

MZ_ A ^

available for positron emission

E = [MZ A where

(16-48)

]c*

available for electron capture

E=
The energy

Mz+1 A

is

typically of the order of

these processes the available energy arises

from

the differences in the masses of the nuclei involved. These three equations
are written in terms of the masses of the atoms, such as

because

it is

convenient and conventional to do

so.

Mz A

simply

A moment's reflection

THE NUCLEUS

624
will

[Ch. 16

demonstrate to the reader that, although these equations employ the

usual approximation of neglecting the mass equivalents of the binding


energies of the atomic electrons, the rest

mass energies of all the

electrons

involved in the processes are properly taken into account. Note that there
is

a range in which the difference in masses between two atoms is such


is possible but positron emission is energetically

that electron capture

However, if
ZA is greater than Z_ 1A by more than 2m,
both electron capture and positron emission can take place. In these
circumstances positron emission is usually the more important process
forbidden.

Electron

Z,A

Z + l.A

Initial

Figure

16-39.

state

Final state

The beta decay process under the assumption

that there are

two

particles in the final state.

for nuclei of small Z.

For nuclei of intermediate and

large Z, the

wave

functions of the atomic electrons are pulled in to the positively charged

and the wave function of a positron

nucleus,
effects

tend to

make

electron capture relatively

cularly for the case of a

Now

let

is

repelled.

us take up the interesting question of what happens to the

these processes, electron emission.

Consider the most

A nucleus Z,A, which we

stationary in the initial state, emits an electron


in figure (16-39).

and

fact, since the

all

little

common

of

assume to be

recoils, as indicated

Since there are just two particles in the final state, there

should be only one way in which the total available energy


carries

parti-

K shell electron, electron capture usually dominates.

available energy in beta decay processes.

In

Both of these

more probable and,

E can be shared.

very low and it


Thus the electron should carry away essentially
the form of kinetic energy. However, measurements

nucleus

is

so massive,

its

recoil velocity

is

kinetic energy.

the energy

E in

an early stage of the study of radioactivity, using momentum


analyzing magnets and other techniques, showed that the electrons are
emitted with a spectrum of kinetic energies Te as indicated in figure

made

at

(16-40).f

For a number of

years, the fact that electrons are emitted in the beta

some cases there are also a few groups of monoenergetic electrons in the measured
They come from a process, called internal conversion, which has nothing to do
with beta decay, and which will be explained in the next section.
t In

spectra.

Sec. II]

BETA DECAY

625

decay process with a spectrum of energies was very mysterious and very
disturbing. Electrons emitted at the end point T ax of the spectrum carry
away all the available energy E, because Tf ax is observed to be equal to

within experimental accuracy. But the more typical electrons carry


away much less energy than the energy E which, the mass differences
show, must be released in the process, and it would appear that some of
this energy has vanished
A number of suggestions aimed at resolving
this dilemma were made. For instance, it was suggested that the electrons
!

emitted from the nucleus could lose energy by collisions with the atomic

Figure 16-40.

typical beta decay electron kinetic

energy spectrum.

electrons in emerging from the atom. If this were the case,


and if the
electron emitting atoms were enclosed in an isolated system,
the lost
energy would eventually be degraded into the form of heat,
as would

the energy actually carried

energy

E would

performed by

electrons, so that the entire

But experiments
and Wooster (1927), using a calorimeter with very
showed that the energy deposited in the calorimeter per
in.

the system.

Ellis

thick lead walls,

electron emission
less

away by the emitted

eventually be deposited

is

equal to the average energy of the spectrum, which

is

than half of E.

Another suggestion was that the electrons of kinetic energy lower than
are emitted in processes m which the final nucleus Z
+ I, A is left
with excess energy in one of its excited states, and that these excited
states
are so close together that the spectrum of emitted electrons
appears
continuous. The excited nucleus would subsequently re-emit this energy
in the form of a gamma ray. However, this was not
tenable because there
are a number of cases in which electrons are emitted with
a spectrum of
energies but no gamma rays are emitted, and also because other
evidence
shows that the excited states are not so closely spaced as would be required.
The situation was grave enough that people were beginning seriously to
consider abandoning the law of conservation of energy, until Pauli
(1931)
proposed a less repugnant alternative.
l

THE NUCLEUS

626

[Ch. 16

Pauli postulated that there is a particle called the neutrino which is


emitted along with the electron in each electron emission process, but

which
1

not observed because

is

Neutrinos interact extremely weakly with matter.

This property allows the neutrinos, and the energy they carry, to pass

through the thick lead walls of the

Ellis

and Wooster calorimeter. Pauli

also postulated that:

3.

Neutrinos are uncharged.


Neutrinos have spin angular

4.

Neutrinos have zero rest mass.

2.

The second property

is

momentum quantum number

J.

required in order that charge be conserved in the


The third property allows total angular

emission process.

electron

Consider a nucleus Z,A


become the nucleus Z + I, A and assume, for
example, that A is even. Then the spin angular momentum quantum
numbers i of both the initial and the final nuclei are integral (cf. item 6 of
section 1). If only the electron with s = \ is emitted, it would be impossible

momentum

to be conserved in this process.

emitting an electron to

momentum

for the final total angular


initial integral total

angular

of the system to be equal to the


the sum of a half-integral

momentum, because

an integral angular momentum (the


However, if a neutrino with
final nucleus) can only be half-integral.
s = | is also emitted, there are two half-integral angular momenta which
can combine with the integral angular momentum to form an integral
angular

momentum

total angular

(the electron) plus

momentum equal to

the initial value.

The fourth property of

neutrinos was postulated by Pauli to agree with the observation that the
end point J ax of the electron kinetic energy spectrum is equal to the available energy E. When an electron happens to be emitted at the end point,
it

carries

away

all

the available energy,

and none

is left

for kinetic or rest

mass energy of the neutrino.


In general, however, the electron will be emitted with kinetic energy Te
than the available energy E, and the neutrino will be emitted with

less

non-zero kinetic energy

T,

With

such that

+T = E = TT X

(16-51)

Pauli's postulate, there are three particles in the final state of the

emission process, very little of the available energy is taken by the recoil
of the massive nucleus, and therefore essentially all of it is divided between
the electron and neutrino. As there are an infinite number of ways in

which

this division

can be made,

it is

evident that the postulate allows

the distribution of electron kinetic energies

spectrum.

Te

to

form a continuous

BETA DECAY

Sec. II]

The

627

postulate also forms the basis of the theory of the shape of this

spectrum. In

its

most elementary form due to Fermi

(1934), the theory

is

Consider a process in which a system makes a transition


from an initial state, consisting of the nucleus Z,A, to a final state, consisting of the nucleus Z + l,A plus an electron moving off with momenquite simple.

tum p e and a
electrons

momentum

neutrino moving off with

do not

exist inside nuclei,

neutrinos do either.

and there

transition of

p.

We know

no reason

we must imagine

Therefore

neutrino to be created in the transition

an optical

is

just as a

that

to believe that

the electron and the

quantum

is

created in

an atom. Then the eigenfunction for the

initial

state is

Wi

(16-52)

Vz.a

the eigenfunction describing the nucleus Z,A,


the final state

and the eigenfunction

for

is

YV

(16-53)

Vz+i.AWeWv

where we assume that in the final state there are no interactions between
the nucleus, the electron, and the neutrino, so that the eigenfunction for
the system may be written as a product of the eigenfunction y> z+i,A f r tne
nucleus Z + I, A, the eigenfunction xp e for the electron, and the eigenfunction

y> v

for the neutrino. If there are

for the electron


(15-23).

That

and the neutrino are

is,

VK

L-?

K =

jtjh

xp e

no interactions, the eigenfunctions


waves of the form

free particle plane

(16-54)

where
e

(16-54')

and
y> v

L-K**x

(16-55)

where

K =
v

pjh

(16-55')

from the origin of coordinates to the


Z,A and Z + 1 ,Aj These eigenfunctions are normalized, using periodic boundary conditions, in a box of edge length L.
We next use the three dimensional extension of Golden Rule No. 2 (equation 9-45) to calculate the rate at which the system will make transitions
from the initial state to the final states. This says that the transition rate is
and where

r is the position vector

center of the nuclei

approximately

^v*vfiPf

(16-56)

We assume the nucleus Z,A

Z+ \,A.

to be stationary,

and ignore the

recoil

of the nucleus

THE NUCLEUS

628

The quantity pf
the quantity vfi

is

the energy density of final states of the system,

That

and

the matrix element of the interaction potential V, the

is

perturbation which causes the transitions, taken between the


final states.

[Ch. 16

initial

and

is,

vmj v *V Vi dx

(16-56')

As we know nothing about this interaction potential, we are at liberty


make the simplest possible assumption concerning its form. Following
Fermi, we assume that V is a universal constant g, which is called the
beta decay coupling constant. Then
to

/<

gj VtVt dr

gL- 3

iK

Jy>Z + i,Ae~

'"e~

iKv ' t

Vz,A dr

(16-57)

There

will be contributions to the integral only from values of r of the


order of the nuclear radii, since for larger values both y>z+i.A anc* Vz.a

But for electron emission processes, except those of


show that Ke and Kv are
both always small enough that K r < 0.1 and K v r < 0.1, for r of the
order of or smaller than the nuclear radii. To a fairly good approximation,
we may therefore set both exponentials in equation (16-57) equal to unity
and obtain
obviously vanish.

unusually high energy, a quick calculation will

% - gL~ ] Vz+i,aV>z,a dr
3

or
v fi

~gir 3M

(16-58)

where the quantity

M=]y>z+

i,

(16-58')

A yz ,AdT

Considering (16-56), we see that,


independent of the electron and neutrino momenta, these
quantities enter into the transition rate, or into the shape of the electron
is

called the nuclear matrix element.

since

is

spectrum, only through the density of final states p f


As the transitions are assumed to lead to final states of the system in
.

which the stationary nucleus Z + I, A is in a single state (usually the


ground state), pf contains only the density of states of the electron and
the neutrino. In fact, p f is essentially the product of the densities of states
of these particles. According to our discussion of section 3, Chapter 15,
both particles do have finite densities of states because of the restrictions
imposed by the periodic boundary conditions used to normalize the
eigenfunctions (16-54) and (16-55). As in that discussion, we may immediately take

over the results (12-37) of the calculation of the density

Sec.

1]

BETA DECAY

629

Te

of states for an electron of kinetic energy


present notation, this

is

N(Tt)dTe

=m

in a

Let us write (16-59) in terms of the electron

box of length L. In our

l'

(16-59)

momentum pe by
,

using the

equation

=T
^
2m

(16-60)

Differentiation yields

P^Es

dT,

Multiplying this by the square root of (16-60),


p2 dp

we

obtain

=T?dT.

2*m A

which allows us to write (16-59) as

N(T )dT
t

^^

(16-61)

can be shown that (16-61), which involves only the momentum pe ,


relativistically correct expression, even though (16-59) comes from
the non-relativistic Schroedinger equation and (16-60), which is nonrelativistic.t
Consequently, equation (16-61) may be applied to the
It

is

relativistic electrons,

transitions

we

and also

to the relativistic neutrinos, emitted in the

are dealing with. For the latter particles the equation reads
3
= l pUP

^^

N(Tv )dTy

2tt

Now

(16-62)

!!

/j

equation (16-51) shows that in a given transition T, is determined


Te In fact, by differentiating that equation we obtain

by the value of

dTv
Therefore

= -dT

(16-63)

suffices to specify the final states

number of final states,


Te and Te + dTe is

which

in

of the transition.

this electron kinetic

energy

lies

The

between

d*N
t

The reason

is

= -N(T ) dT N(T ) dT = V

that the periodic

boundary condition

L6 tf

dP*P* dPe

4 6
47T h
restrictions

which lead

to the

of states are essentially restrictions on the allowed values of the de Broglie


wavelength or, through the relativistically as well as classically valid expression /> = h/X,
finite density

are restrictions

on the allowed values of the momentum.

THE NUCLEUS

630

[Ch. 16

fact that dTv and dT, are intrinof the opposite sign. The number of these states per unit electron
kinetic energy is

where the minus sign compensates for the


sically

d*N

LpUp vPUPe

To

simplify this

energy

is

the

we note

same as the

that for the zero rest

mass neutrino the

kinetic

total relativistic energy and, as in (3-10),

cp Y
Differentiating,

(16-64)

InW dZ

dT,

=T

(16-65)

we obtain
c dp,

= dT = -dT
v

(16-66)

where we have used equation (16-63).


have

Substituting this in (16-64),

d*N_LPtf

d%-^^c dP

we

(16 ~ 67)

'

The quantity /f may be eliminated by using equations (16-65) and

(16-51),

which give
__v

K1 e

I e)

Thus equation (16-67) may be written


max ~ T
- f ,
_ L8(r
^p!dp

d2N

4ttW

dTe

(16-68)

Taking (16-68) as the density of final states for which the electron
is between p, and
pe + dpe and using (16-58), we have an
expression for R(pe) dp e the rate of transitions into these states. This is

momentum

R(P.) d Pe

^^ (rr - T
x

Ztt

The function

n c

fp\ dp,

(16-69)

momentum

spectrum of the electrons emitted in


appear in a convenient form since most of
the experiments yield the momentum spectrum directly. They predict
that a plot of the quantity [R(p^lpl\A versus (Tf Te ), or simply
versus T should yield a straight line. Figure (16-41) shows such a Kurie
plot for the simplest of all electron emission processes,
R(p,)

the transitions.

is

the

These

results

V
1

the decay of the neutron "w into

H1 +

+v
a proton H
e~

(16-70)
plus an electron e~ and a

The neutron decays this way because [M01


+0.78 Mev. The lifetime t of the decay is about 1 100 sec.

neutrino

v.

M^c2 =

Sec.

BETA DECAY

1]

631

These data provide a fairly typical example of the kind of agreement


obtained between the simple Fermi theory and experiment for the beta
decay of low nuclei. The deviations at the low energy end of the spectrum

are often found to be a result of experimental difficulties such as selfabsorption of the low energy electrons in the source. However, for high
nuclei there are definite deviations between experiment and the simple

theory at the low energy end of the spectrum which are more pronounced,

02

0.1

0.3

0.4

Te
Figure 16-41.

0.5

0.6

0.7

0.8" - 0.9

1.0

(Mev)

Kurie plot for the beta decay of the neutron.

From

J.

M. Robson,

Phys. Rev., 83, 349(1951).

and are of the opposite


deviations for high

compared to those of figure (16-41). These


from the use of plane waves (equation

sense,

nuclei arise

16-54) for the electron eigenfunctions. This

is

an excellent approximation

for the neutrino eigenfunctions because neutrinos interact so weakly that

they really do behave like free particles.

Coulomb
to make %

strongly with the

field

interaction tends

larger than

of the

But electrons interact rather


final nucleus.

This attractive
a plane wave in the region near

and thus increases the transition rate. As the Coulomb


more important for small Te the effect is to enhance the low
energy part of the spectrum. The simple Fermi theory also applies directly
to positron emission, but the Coulomb effect has a different sign because
the interaction is repulsive. This makes the positron eigenfunction smaller
than a plane wave near the nucleus and leads to a depletion of the low
the nucleus

attraction

is

energy part of the spectrum.


are included in the theory,

When

its

corrections for these

Coulomb

effects

predictions are usually in

with accurate experiments, even for

good agreement
electrons or positrons of quite low

energy.

Of

course, the question of the shape of the spectrum does not arise in

THE NUCLEUS

632
electron capture.

the nucleus

state,

away

carries

In this process there are only two particles in the final

Z I, A

and the neutrino. The

essentially all the available energy

Although only a neutrino

so massive.

[Ch. 16

is

always

latter particle

since the nucleus is

emitted in the primary process

of electron capture, the process can be observed experimentally by detecting the X-rays which are subsequently emitted in the filling of the hole left
in the electronic system of the atom.

By

integrating (16-69) over the entire spectrum, the transition rate into

all final states

may be

As

calculated.

in equation (13-118), this total

probability per second that the initial nucleus will decay


reciprocal of

its

lifetime t.

is

equal to the

Thus

^sr^-^*t

To

2tt

convert the integrand completely to a function of pe

(T

where

mc2

max

TJ2

(i6-7i >

Jo

<r

we

write

^ ^ + ^^

_
= [T +
= (fir" " Eef
max

the rest mass energy and Ee is the total relativistic energy of


and where we have used equations (1-22') and (1-24). But,

is

the electron,

according to (1-25),

E e = Vc2^ +

mV

So
i
t

^ f^Wr" + V - Vc^ +
2irn<r Jo

rnV/p! dA

Integrating this gives

;"S^

(16-72)

F(a)

where
F(a)

= -Ja - Aa3 + 3^a5 + jVl +


-

a 2 sinh" 1

Vl +

a2
(16-72')

with

= p**jmc

a.

The function

F(<x)

has the behavior


-

F()~

ifsoc

7
,

<x

and an intermediate behavior between these

>5
< 0.5
limits.

(16-72*)

Sec.

BETA DECAY

1]

When

the

Coulomb

An

function F(a, Z).

633

corrections are included, F(a)

graphs and tables can be found in several


f(a,

Z)

~ F(a).

For

is replaced by a
Z) cannot be given, but
references. For small Z,

analytic expression for F(oc,

Z=

emission, and F(x,Z)

100, F(a,

~ 0.1 F(a) in

Z)

100F(a) in the case of electron

the case of positron emission,

when

a has a value corresponding to the typical available beta decay energy


E = 1 Mev.
Equation (16-72) shows that the lifetime t of a beta decaying nucleus
depends strbngly on />ax and therefore strongly on the available energy E.
,

The observed lifetimes for the naturally occurring beta unstable nuclei
range from values of the order of 1 sec for an available energy of several

Mev to values of the order of 10' sec for available energies of several
hundredths of an Mev. (Note, however, that the lifetime of an alpha
decaying nucleus is an even stronger function of the energy available in
the alpha decay.)

Equation (16-72) also shows that the quantity

F(a z>~^Tl7^r;
>

g'Tn^c*

M*M

(16-73)

depends only on a collection of universal constants, and on the value of


the nuclear matrix element

M =JVzi,aVz.a dr
where we

now

(16-73')

indicate that the theory applies equally well to electron or

The product Ft is inversely proportional to M*M and


when it takes on its maximum possible value M*M = 1. This
occurs when the eigenfunction y> zi.A is identical with the eigenfunction

positron emission.
is

smallest

Vz.a because, in these circumstances, the normalization condition for


eigenfunctions requires that (16-73') yield
1.
If the eigenfunctions
1, and this quantity becomes
Vzi.a and Vz.a ar e not identical,

M=

M*M <

smaller as the eigenfunctions

become less similar. In fact, M, and therefore

M*M, are exactly zero if y>zi,A and Vz.a are so dissimilar as to correspond
to different values of nuclear spin

i,

or to different nuclear

parities.

The

because the eigenfunctions y> z\.A an d Vz.a-> like


any other eigenfunctions, are orthogonal with respect to all their quantum
numbers, including the quantum number i. The second restriction arises
first restriction arises

because, if the parities of these two eigenfunctions differ, the entire


integrand will be of odd parity.t and the contribution to the integral
(16-73')
t That

from each point


is,

(even)(odd)

\x, y, z) will

(odd),

be canceled by the contribution

and (odd)(even)

(odd).

THE NUCLEUS

634

from the point

(-

x, y, ). These two

[Ch. 16

restrictions constitute the

Fermi

selection rules:

M=
The nuclear
on the

that give conditions


nuclei Z,

element

and

M to vanish

parity

(16-74)

must not change.

relative spins

and

parities of the states of the

1,A which, if violated, cause the nuclear matrix


Transitions which satisfy the selection

identically.

rules are called allowed transitions.

'
!

Neutrons Protons

Neutrons

Figure 16-42.

A schematic

Protons

representation of the ground states of a typical set of mirror

nuclei.

It

immediately follows from the shell model that beta transitions between

the ground states of any set of mirror nuclei are always allowed transitions.

As mentioned

in section 3, the two nuclei of a set of mirror nuclei consist


of an odd number of protons and the same odd number of neutrons, plus
an extra proton for one of the nuclei and an extra neutron for the other.

The neutron and proton

eigenstates filled

by the nucleons of a

of mirror nuclei are indicated in figure (16-42).


nuclear spin

is

equal to the total angular

nucleon and, since

this

nucleon

is

in the

momentum

same

typical set

In the shell model, the

of the single odd

eigenstate in each nucleus

set, the spins of these two nuclei will be the same. Also, in the shell
model the complete eigenfunctions yzi,A and Wz.a can De written as
products of the eigenfunctions for each of the A independently moving
nucleons, and the nuclear parity will be the product of the parities of each
of these eigenfunctions. Since there are two nucleons in each eigenstate,
except for the eigenstate containing the single odd nucleon, and since the
product of the parities of two eigenfunctions of the same parity is always

of the

even,f the nuclear parity


single

odd nucleon. As

nucleus of the

same.
t That

set, it

is

equal to the parity of the eigenfunction of the

this

nucleon

is

in the

same

eigenstate in each

follows that the parities of these two nuclei are the

These arguments lead us not only to the conclusion that the


is,

(even)(even)

(even),

and (odd)(odd)

= (even).

Sec.

BETA DECAY

1]

complete eigenfunctions
similar

enough

635
ip z1

A and y z<A

for a set of mirror nuclei are

to satisfy the selection rules (16-74), but also to the

much

stronger conclusion that these two complete eigenfunctions should be

They differ only in that the eigenfunction for one


odd nucleon) is a proton eigenfunction for one of the

essentially identical.

nucleon (the single


nuclei and a neutron eigenfunction for the other and, since these eigenfunctions are for the same eigenstate, they are identical except for

Coulomb

M*M ~

1. Thus
f zi,A V>z,a>
we conclude that for all beta transitions between the ground states of sets of
mirror nuclei the product Ft should assume its smallest possible value.
This conclusion is verified by the experiments, which show that for all these
35 sec, and that
transitions the Ft values are clustered around Ft ~ 10
the Ft values for these transitions are the smallest which are ever observed.
By using these experimental values of Ft in (16-73), and making the

effects.

But we have seen

M*M =

assumption that

tnen

that, if

for these cases,

constant can be evaluated. The result

gil.4x

the beta decay coupling

is

10- 49 erg-cm3

(16-75)

Of course, there are many allowed transitions which do not take place
between the ground states of mirror nuclei, since there are many other
Although
possibilities of satisfying the selection rules (16-74).
is greater than zero for all these transitions, it is less than unity because

M*M

y z1 A and y Z-A lead to cancellations which reduce


the value of the integral (16-73'). With increasing A, these differences
the differences between

should become greater because the nuclei become larger and more complicated.

This

is

seen in the experiments since the measured

these allowed transitions increase with increasing A.


55
10
sec.
for allowed transitions are around Ft

The

Ft values for
Ft values

largest

If either part of the selection rules

However, for such forbidden

become

transitions

infinitely large, as predicted

only by about a factor of 103

not exact because

it is

(16-74)
it is

is

M*M = 0.

violated,

observed that Ft does not

by equation (16-73), but increases

This simply

reflects the fact that

based upon several approximations.

(16-73)

One of

is

these

e~ iK '* equal to unity in (16-57). A correction to this approximation is found by using the more accurate expression
of (16-57) obtained by expanding e~' K '* in a power series and taking the
first two terms. This gives for the beta decay interaction potential matrix
is

the setting of e~' Ke-r e~' K,,

'

element

gL~
J

V>zi,aVz,a dT

- K

Vzi,A r V>z,A d i

<

(16-76)

'J

Situations can obviously arise in which the

first

term vanishes but the

THE NUCLEUS

636

second does not. As

J ip%i,A*y>z,A </t

<K

of normal energy, this second term makes

Ft values

transitions as long as the

it

<

10

-1

[Ch. 16

for beta decays

possible to have forbidden

are larger than those for allowed

by a factor >(10 -1 )' 2 = 10 2 A similar possibility arises from


_1
a term of order v/c < 10
which should actually appear in the beta decay
transitions

interaction potential matrix element because of relativistic effects.

The

situation

further complicated by the experimental observation

is

that there are certain beta transitions that violate the

(16-74) but nevertheless have


that are allowed

by these

Ft values

assumed, but

of a

is

V in

(16-56')

is

not a constant, as Fermi

more complicated type

interaction potential matrix element

which gives

show that

Theoretical investigations

Fermi selection rules


any of the transitions

This implies that in these transitions the

rules.

beta decay interaction potential


initially

as small as for

that leads to

an

different selection rules.

there are several types of matrix

elements which should be possible, and that certain of them lead to the

Fermi selection

rules, while others lead to the

Ai

0,

(but not

Gamow -Teller selection

->

rules:

0)

(16-77)

The nuclear

By

parity

must not change.

evaluating the consequences of the various possible types of matrix

elements, and comparing these consequences with experiment,

it is

found

that the beta decay interaction potential matrix element actually existing
in nature contains a mixture of

approximately equal amounts of the types

leading to the Fermi selection rules and the types leading to the

Gamow-

These comparisons are not based on the shape


of the beta decay spectrum, because this is always dominated by the same
density of states factor for all types of matrix elements. We shall not go
into the theory underlying these comparisons because these points are
really intelligible only in the language of Dirac's relativistic theory of
quantum mechanics. Similar motivation leads us to avoid making a
Teller selection rules.

between neutrinos and anti-neutrinos.


However, we shall consider one recently discovered property of the beta
decay interaction potential matrix element that is extremely interesting
and not difficult to describe. It is found that this matrix element is not a
distinction

scalar but instead

is

at least partly a pseudo-scalar.

when

The

distinction

is

that

performed that is,


when the coordinates of each point (x, y, z) are changed to (%, y, z)
whereas a pseudo-scalar changes sign in this operation. An example
of a scalar is the Coulomb interaction potential, or any other interaction
potential that can be written V(r). As the argument r is the magnitude of
a scalar does not change

the parity operation

is

a vector

r,

V(r) does not

change in the parity operation of

reflecting r

BETA DECAY

Sec. II]

637

through the origin, even though r -> r in this operation. Since the
matrix element of V(r) is a certain average of this quantity, the matrix
element is a scalar. In fact, with the exception of the beta decay interaction
potential matrix element, the matrix elements of

all

the

known

interaction

of classical, atomic, and nuclear physics are scalars. For


instance, consider the spin-orbit interaction potential matrix element,
potentials

L. We investigate what happens to S L


by writing L = r X P and realizing that in this
operation both r and p change sign.f Thus L does not change sign. To
bring out the distinction between their behavior in the parity operation,
vectors which change sign, such as r and p, are called polar vectors, and
vectors which do not change sign, such as L, are called axial vectors.

which

is

proportional to

in the parity operation

All angular

we

momentum

vectors, including S, are axial vectors.

see that the spin-orbit interaction potential matrix element

Therefore
is

a scalar

does not change sign in the parity operation. An example


of a pseudo-scalar is the quantity P A, where P is a polar vector and A
is an axial vector.
This is a pseudo-scalar since it changes sign in the
parity operation because P changes sign and A does not. The beta decay
because S

interaction potential matrix element contains quantities of this type.

the parity operation

is

equivalent to leaving the point

As

and
the operation amounts to
(x, y, z)

fixed

through the origin,


changing from a right-handed to a left-handed coordinate system and,
as at least part of the beta decay interaction potential matrix element
changes sign in this operation, the description of the beta decay processes
which this matrix element produces will depend on whether a right-handed
or a left-handed coordinate system is used. When this was first discovered,
it was extremely surprising because it had been a fundamental assumption
reflecting the coordinate axes

that the description of all physical processes must be independent of


whether a right-handed or left-handed coordinate system is used.
This assumption was proved incorrect for beta decay in experiments
performed by Wu and collaborators (1957), at the suggestion of Lee and
Yang (1956). In these experiments, which we describe first in a right-

handed coordinate system, the spin angular momentum vectors I of


electron emitting 27 Co 60 nuclei were aligned with respect to an external
magnetic field by using the hyperfine structure interaction at temperatures
so low that the thermal equilibrium energy kTis comparable to the hyperfine splitting energy. In these circumstances, a majority of atoms will be
in the lowest energy state, and in this state I is essentially aligned with
respect to the applied magnetic field. The angle between the direction of
I and the direction of the momentum vector p was then measured for
tp
other

changes sign since

m dxjdt -* m d{-x)jdt =

momentum components.

-mdxjdt, and

similarly for the

THE NUCLEUS

638

[Ch. 16,

each emitted electron, and the distribution of these angles was plotted.
It was found that this distribution is asymmetric, with about a 30 percent
preference for the emission of electrons whose

momentum

vectors p are

from the spin vector I. So in the average


emission process p I < 0. But p I is a pseudo-scalar since p is a polar
vector and I is an axial vector. Thus in any left-handed coordinate
in the opposite hemisphere

system the sign of


process p

>

0,

this

pseudo-scalar will change, in the average emission

and the description of

depends

this process therefore

on whether a right-handed or a left-handed coordinate system

is

used.

must be a
were purely

This proves the beta decay interaction potential matrix element


pseudo-scalar, or at least contain a pseudo-scalar part.
scalar, the electron distribution

in the average emission process

If

it

necessarily be symmetric so that

would

0,

and the description of

this

process would be independent of whether a right-handed or a left-handed


coordinate system is used, even though the pseudo-scalar p I changes sign,

because

+0.

Since the beta decay interaction potential matrix element is not a


z)
it is not true that for the interaction potential V(x, y,

scalar,

V(x, y,

Thus the eigenfunctions

z).

for the entire system

do not

necessarily

have a definite parity (cf. footnote on page 465), and it is not necessary
that the entire system conserve parity. Experiments indicate that it does
not. The parity of the initial state consisting of the nucleus Z, A is not
the

same

as the parity of the final state consisting of the nucleus

Z \, A

plus the emitted neutrino and electron or positron.

We close this

section with brief descriptions of

ments involving beta decay and


In the period since 1950 a
Allen, Sherwin, and others

on the

recoil velocity of the nucleus

As

the nucleus

is

remaining in
very massive

and neutrino, the recoil velocity is low enough


measured
by measuring electronically the time (several
can be

compared
it

recent experi-

number of measurements have been made by

the final state of a beta decay process.

that

some other

neutrinos that are of particular interest.

to the electron

microseconds) required for the nucleus to travel a given distance (several


centimeters) in a given direction through a field free vacuum. This
measurement determines the momentum vector of the recoiling nucleus.

By

momentum vector of the emitted electron, it is


momentum can be conserved only if a neutrino
process. Furthermore, the required neutrino momentum

also measuring the

possible to

show

that linear

emitted in the
agrees within the experimental accuracy (10 percent) with the neutrino
energy required to conserve energy, if the neutrino rest mass is zero as
is

assumed by Pauli.
Even after these

recoil

a strong feeling in

some quarters

experiments had been performed, there was still


that the neutrino is not a particle but

BETA DECAY

Sec. II]

only a

639

myth cleverly designed to provide for the conservation of energy,


momentum, and linear momentum. There was a certain merit in

angular
this

point of view as long as the neutrino remained undetectable. However,


Cowan and Reines in 1956 succeeded in detecting the

experiments by

neutrino, and thereby gave

tremely

it

the status of any other particle.

->

V + e+

the capture of a neutrino v by a proton

positron e +

In these ex-

experiments they observed the reaction

difficult

This

(16-78)

leading to a neutron n 1 and a

the inverse of the reaction

is

V + e+ - W +

(16-79)

which is an alternative form of the neutron decay process (16-70). A point


which will be of interest shortly is that the reaction (16-79) should have
10 3 sec^ as the reaction (16-70) because the
about the same lifetime, t
matrix elements and densities of states governing the lifetimes contain
essentially the same factors, albeit in a different order. The CowanReines reaction (16-78) took place on the protons contained in a very large
tank of hydrogenous scintillator exposed to the enormous flux of neutrinos
emitted from the fission induced beta decays in a nuclear reactor, and the
positrons were detected by the scintillations they produced in the same
tank. Elaborate methods were used to minimize background scintillations.
This was necessary because less than one reaction per minute was expected,
despite the magnitude of the neutrino flux and the size of the hydrogenous

target, since the interaction of neutrinos with

To show

we make

this,

Mev

incident

upon a proton, and assume

interaction between these

two

particles

their relative separation is

maximum

~10~ 10

that there

lifetime

~103

justified

cm. As their relative velocity

lifetime for the reaction (16-78)

sec for

its

a possibility of

This will occur

when

is

is

~10 10 cm-sec -1 the


20
is ~10~
sec. Now
,

about the same as the


this assumption is

inverse reaction (16-79);

by the principle of microscopic

equation (13-84).

is

comparable to the de Broglie wavelength of

time during which there can be overlap

assume that the

extremely weak.

whenever they are close enough to

allow appreciable overlap of their wave functions.


the neutrino

is

Consider a neutrino of the typical energy

reaction (16-78) as follows.


1

matter

a rough estimate of the cross section for the

reversibility

mentioned

after

Then

the probability that the reaction (16-78) will


10~ 23 and the cross section
occur in the time ~10~ 20 sec is ~10~~ 20 /10 3
10~ 20 cm 2 for
a for this interaction is the cross section <~(10~ 10 cm) 2

with overlapping wave functions times the probability just


10~ 43 cm 2 An accurate
evaluated. This gives <r~ 10~ 20 cm 2 x 10~ 23
collisions

THE NUCLEUS

640

on the method we used

calculation, based

[Ch. 16

in evaluating the lifetimes of

the processes leading to neutrino emission, gives

(16-80)
rk*C

For a positron of momentum p e and velocity v e corresponding to the


1 Mev, (16-80) predicts an interaction cross section
10~ 44 cm 2 The measurements of Cowan and Reines agree with this
a

typical kinetic energy

value, within the large experimental uncertainty.

Note

that equation (16-80)

shows

clearly that the extreme smallness of

the neutrino interaction cross section

is a result of the extreme smallness


10~ 49 erg-cm 3 The cross section
of the beta decay coupling constant g
small
that
a
neutrino
is so
emitted in a nuclear reaction in the interior of

the sun can traverse

its entire bulk with little probability of being absorbed.


Beta decay interactions are very weak compared to nuclear interactions.
A numerical comparison can be made by dividing the beta decaycoupling

constant, which has the dimensions of energy times volume, by the

volume

of a typical nucleus. This gives a characteristic energy ~10~ 49 ergcm3 /^ x 10~ 13) 3 cm 3 -~ 10~ 12 ergs <~ 1 ev, which is smaller by a factor of

~10~ 7

than the energy

~10 Mev

that characterizes nuclear interactions.

Since the square of the beta decay coupling constant enters into measurable
quantities, such as transition rates

and cross

sections,

it

is

appropriate

to say that beta decay interactions are weaker than nuclear interactions by a

factor of <~10~ 14 The reader will recall that the comparison of figure
(16-12) indicates that electromagnetic interactions are only somewhat
.

weaker than nuclear interactions;


will

show

in section 16 a quantitative

actions by a factor of'

TO -3

This conclusion, plus the conclusion drawn

in the first footnote of section 3,

shows that gravitational

weaker than nuclear interactions by a factor o/"~10~ 38

12.

comparison

that electromagnetic interactions are weaker than nuclear inter-

Gamma

interactions are

Decay and Internal Conversion

The nucleus

a system of nucleons

is

bound to a certain region of space.


quantum mechanics, it should

Since this system seems to obey the laws of

have a

set of

quantized energy states.

we have considered
we turn to the

Heretofore

only the lowest energy ground state, but in this section


higher energy excited states.
excited states has

gamma

rays,

state of

energy

come from

Much

of what has been learned about the

a study of the high energy quanta, called

which are emitted in transitions between some initial excited


E and a final state of energy Et < E These spontaneous
t

GAMMA DECAY AND INTERNAL CONVERSION

Sec. 12]

gamma

641

decay transitions are completely analogous to the optical transi-

between the quantized energy states of an atom. In


both cases, the transitions take place through the interaction of the charge
moments and magnetic moments of the system with electromagnetic
fields, and energy is conserved because the energy of the emitted quantum
tions that occur

is

hv

=E -E
t

(16-81)

There are a number of processes through which the nucleus can be


its higher energy states.
For instance, it can happen that
an alpha decay or beta decay process leaves the final nucleus in an excited
state instead of its ground state. The possibilities are indicated schemaexcited to one of

diagram of figure (16-43).


decay is, of course, the available beta decay
energy E, equal to the total energy carried away by the neutrino and the
electron or positron. Usually only a few states of the final nucleus are
tically,

for beta decay, in the energy level

The energy change

in the beta

excited because beta decays to states of certain spins and parities are
forbidden by the beta decay selection rules, and because the strong
dependence of the beta decay transition rate on E makes the transition
rate negligible to states higher

than several

In excitation following alpha decay,

Mev above

the

ground

state.

the even stronger dependence of

on the alpha decay energy usually limits


somewhat less than 1 Mev
above the ground state.f States of energies up to 6 or 8 Mev above the
ground state can be excited when this binding energy is liberated by the
capture of a low energy nucleon in a nucleus. States of even higher
the alpha decay transition rate

the excitation of the final nucleus to states

energy can be excited by the capture of a higher energy nucleon, processes


in which a high energy nucleon is inelastically scattered from a nucleus,

and by a number of other processes involving nuclear reactions.


After the excitation of a nucleus, it can decay by the emission of gamma
rays.

If

nucleon

the excitation energy


in the nucleus,

it

is less than the binding energy of the last


must decay by gamma emission, or by the related

process of internal conversion that will be explained later in this section.


t

Note

that,

when alpha decay between two

nuclei leads to excited states of the final

nucleus, the alpha particle energy spectrum will contain

of lower energy.
of the

initial

It is

nucleus,

of higher energy.

a. fine structure of components


also possible for the alpha decay to originate from excited states

and

The

in this

case the spectrum will contain long range components

similar situations which occur in beta decay result in several

values of the available beta decay energy E, and electron or positron energy spectra
which are a superposition of spectra corresponding to each value of E. Note, then, that
information about the energies of the excited states can be obtained directly from a study

of the energies of the particles emitted in alpha or beta decay. For these processes,
and particularly the latter, a study of the lifetimes can also give information about the
spins and parities of the excited states.

THE NUCLEUS

642

[Ch. 16

For excitation energies higher than the binding energy of the last nucleon*
decay by nucleon emission is possible and, for reasons to be explained in
the next section, with increasing excitation energy nucleon emission
rapidly becomes the dominant decay mode. As a consequence, most

gamma

rays are emitted with energies less than about 6 or 8 Mev. In fact,

their energies are generally

much

less

than

this

because the de-excitation

often proceeds in a cascade through a succession of low energy states,

instead of by the emission of a single

gamma

ray which carries

away

all

Excited states

'. ',

Zl,A
The

Figure 16-43.

the excitation energy.


is

set

'

Ground

state

excitation of a nucleus as a result of beta decay.

The lower

limit for the energy of these

gamma

rays

by the minimum energy difference between the low-lying nuclear

energy

states.

It is

of the order of 10 3 ev.

Most of the methods for measuring the energies of gamma rays, or


quanta in the energy range 10 3 ev to 6 or 8 Mev and higher, involve
transferring their energy content to electrons by means of the photoelectric
and Compton processes described in Chapter 14, or to electrons and
positrons by means of the pair production process described in the same
chapter. The energies of these charged particles can then be measured by
standard techniques. In perhaps the most widely used method, the energy
is transferred to electrons of the scintillating crystal Nal, and the electron
energies are measured in the same crystal by measuring the intensity
of the light pulse which it emits, due to scintillations produced by the
electrons. For hv < 0.3 Mev, it is most probable that all the energy of a
gamma ray will be transferred to a single electron of the 1 atoms by the
photoelectric process. For 0.3 Mev et hv < 7 Mev, the Compton effect
is the most probable, and only a fraction of the gamma ray energy is
transferred to an electron in a single collision. In this range, it is necessary
to use a large Nal crystal in order to absorb all the energy. Forhv > 7 Mev,
the pair production process dominates, and the gamma ray energy is

GAMMA DECAY AND INTERNAL CONVERSION

Sec. 12]

643

divided between the total energy of an electron and a positron. At even


larger hv, the ranges of these particles are too long to allow their energies
to be

measured

an Nal

in

called a pair spectrometer


foil

of some high

crystal of convenient size and, instead, a device


is

In this device the pair is produced in a


and the energies of the emitted electron and

used.

Z material,

momenta in a magnetic field.


The methods just described have gamma ray energy resolutions of several
positron are measured by analyzing their

percent at best, but other methods of different types and much better
resolutions are also available. The one with the best resolution involves

gamma rays from a crystal lattice of known lattice spacing d.


Although the gamma ray wavelengths X = cjv are as much as several
orders of magnitude smaller than d, diffraction can be used by making
the distances between the source, crystal lattice, and detector very large.
This gives a large linear displacement even though the angular displacement
diffracting the

~ Xjd

However, as all the solid angles involved are very small,


the diffraction method has low efficiency and can be used only in very
favorable circumstances. The other high resolution methods also suffer
6

is

small.

from low efficiency.


The measured energy spectrum of gamma rays, emitted in transitions
between the energy states of a nucleus, is used to determine the energies
of these states the connection being made by the same trial and error

procedures that were used in the very early days of the study of atomic
spectra and energy states. Of course, the energies of the nuclear states
give important information about nuclear structure.
Another source of important information is the lifetimes of these states.
If the lifetime t

>

10~ 10 sec,

it

can be measured by electronically timing

the delay between the instant the state was excited (e.g., by detecting the
beta decay which led to the excitation) and the instant it decayed (by
detecting the emitted gamma ray). In extreme cases, lifetimes as long as
t
10 8 sec are found. All gamma ray transitions with directly measurable

lifetimes are called isomeric transitions to distinguish

them from

the

more

which have lifetimes t ^ 10~ 10 sec that are not directly


-16
sec, it can be measured indirectly by
measurable. However, if t < 10
measuring the natural width of the gamma ray spectral line emitted in the
transition, and then employing the relation (13-125). This technique can
be used for nuclear transitions only with the aid of an ingenious method,
called resonance absorption, that employs absorption in a target, contypical transitions,

taining

some

species of nuclei, as a spectrometer of extremely high resolu-

tion to measure the width of the spectral lines emitted

containing nuclei of the same species.


10~ 17 sec have been observed.

Gamma

decay

by a source,

lifetimes as short

asT~

The tremendous

variation which

is

found

in the lifetimes for

gamma

THE NUCLEUS

644

decay

due, in part, to the dependence of t on the energy of the emitted

is

gamma

[Ch. 16

ray.

But an equally important part of this variation


and parities of the initial and

influence of the relative spins

is

due

to the

final nuclear

states.
Therefore analysis of the experimentally measured values of
t gives information about the spins and parities of the states. These

quantities are just as important in characterizing the states as their energies.

The analysis is based on the theory of gamma decay transitions in nuclei.


As the theory is basically the same as the theory of optical transitions in
atoms which we treated in section 12, Chapter 13, it would be appropriate
for the reader to review that section.

In our treatment of optical transitions in atoms

we considered only

the electric dipole transitions, which arise


electric

from periodic oscillations in the


dipole moment of the charge distribution, and we found that the

electric dipole transition rate for

spontaneous emission

is

proportional

to the "square of the matrix element of the interaction potential for the

moment,"f taken between the states involved in the transition.


completely rigorous treatment would also consider the electric quadru-

associated

pole, electric octupole,% etc., transitions arising


in the associated
dipole,

from periodic

moments of the charge distribution's well as

magnetic quadrupole, magnetic octupole,%

from periodic

oscillations in the associated

tion (or of the magnetic

transitions arising

moments of the current distribu-

moments produced by

would be found that the

etc.,

oscillations

the magnetic

electron spin). In each case

transition rate for spontaneous emission

by a
proportional to the square of the matrix element of the
interaction potential for that moment. We did not give the completely rigorit

particular moment

is

ous treatment because

The

it

is

really

not necessary for optical transitions.


magnetic optical transitions is

justification for neglecting all the

that the current distribution

is

pvjc,

of the moving charges, and this

by the factor

moments

v/c,

which

is

where

is

the characteristic velocity

smaller than the charge distribution p


typically ~10~ 3
As a result, the magnetic
is

moments by this
(The magnetic moments produced by electron spin are of the same
order as those produced by the current distribution.) Thus the square
of the matrix element of the interaction potential, and therefore the
are

all

smaller than the corresponding electric

factor.

We

use this somewhat loose terminology as an abbreviation. The transition rate


t
for spontaneous emission is proportional to the transition rate for absorption, and
(13-77) shows that the latter quantity is actually proportional to a sum of terms each of
which is the matrix element of the interaction (perturbation) potential for a component
of the electric dipole moment times its own complex conjugate.
J Except for a special process involving the entire atom, there are no electric mono-

pole transitions because nuclear charge conservation requires the associated


to be constant.

moment

is

zero.

moment

There are no magnetic monopole transitions because the associated

GAMMA DECAY AND INTERNAL CONVERSION

Sec. 12]

transition rate,

smaller for each of the magnetic

is

moment by

corresponding electric

The
etc.,

the factor (v/c) 2

moments than

~ 10~

is

that the matrix element of the interaction

moment

of the charge distribution

moment by

smaller than the matrix element for the preceding

2-nRjl,

for the

justification for neglecting electric quadrupole, electric octupole,

optical transitions

potential for each successively higher


is

645

where

is

the radius of the charge distribution and X

length of the radiation.

This factor

is

is

the factor

the wave-

~10~3 The matrix


successively higher moment

also typically

element of the interaction potential for each


of the current distribution decreases by the same factor. The factor arises
because the interaction potential of a dipole
field involves

products of the

field

moment

in a spatially varying

times distances of the order of R, the

interaction potential of a quadrupole

moment

involves products of

first

R2

the

spatial derivatives of the field times distances of the order of

interaction potential of an octupole

moment

involves products of second

spatial derivatives of the field times distances of the order of


t

Equation (13-65) shows that

potential,

and

it

electric potential

is

this is true for the electric dipole

y, z)

and the charge distribution be

S<7j.

R3

moment

easy to verify for the higher electric moments.

be V(x,

etc.t But,
interaction

Let the external

where

(x h

yH

z,)

are

the coordinates of charge q s and where the sum is taken over all the charges. Then the
total energy of the distribution is S^Ktx,, Vi, z,). Make a Taylor expansion of V(x, y, z)
,

about the origin

(0, 0, 0)

V(x,y,z)

of the charge distribution. This gives

[V

"l/a K\
2

+
+

Sx dy/

Use

this to evaluate the total energy.

l/a K\
2

xy
v

().*

d2

2\

ft?

\ dx dz/

\ By dzf

It is

where we employ the relations Ex = dVjdx,Ev = dVISy,E z = ft V/dz to introduce


the components of the electric field. The first square bracket is the electric monopole
interaction potential, and the sum it contains is the total charge; the second square
bracket is the electric dipole interaction potential, and the three sums it contains are

THE NUCLEUS

646

[Ch. 16

which is a sinusoidal function of position with wavelength X,


such as sin (Inx/X), each spatial derivative brings in a factor of lirjX.

in a field

Thus the interaction potential of each successively higher moment contains


an extra factor of 2-ttR/A. The square of the matrix element, and thus the
transition rate,

smaller for electric quadrupole transitions than for


10~ 6 and the transition
by the factor (IttRJI) 2

is

electric dipole transitions

moments

rates for the higher

are smaller by successively higher even

powers

of this factor. All optical transitions in atoms, other than electric dipole
transitions, have very small transition rates, or very long lifetimes. These
transitions

can be neglected because in the long lifetimes the atom

will

almost always be de-excited by collisions, or some other non-radiative


process, before it gets a chance to radiate.

These simplifications do not obtain for gamma decay transitions in


-4
10~ 2 and (2ttR/X) 2
For one thing, (v/c) 2
10
in the typical
case A = 200 and hv = 1 Mev. Therefore the factor suppressing the rates
for magnetic transitions is not so small, and neither is the factor suppressing the rates for transitions associated with the higher moments. Even more

nuclei.

the components of the electric dipole moment the third square bracket is the electric
quadrupole interaction potential, and the six sums it contains are the components of
the electric quadrupole moment. This proves our point.
It is also interesting to develop the relation between the components of the electric
quadrupole moment and the quantity Q defined by equation (16-38). In the case to
which that definition applies, there is an axis of symmetry along the z axis, each of the
2
last three components vanishes, and ~5Lq s x = YfliV^Then the electric quadrupole
;

interaction potential

is

/3 2 k\1

(d 2 V\

l/^W

where we have changed back to the external electric potential


potential, Laplace's theorem of electrostatics demands
d2

d2

d2

dx

So

By

V.

But, for an external

dz 2

the electric quadrupole interaction potential can be written

where

KSM^-H-^"
Q=

2>,[2z?

2x 2 ]

J^q&z

-x - y
2

2
]

= 2*p*j -

(*?

y*

*?)]

Equation (16-38) is obtained by


in units of proton charges.

letting the

sum go

to an integral

and expressing

the q s

GAMMA DECAY AND INTERNAL CONVERSION

Sec. 12]

important

no

the fact that there are

is

collisions

between nuclei, or other

non-radiative processes (except internal conversion)

if

the excitation energy

Thus all states with excitation


than the nucleon binding energy must eventually decay by

than the nucleon binding energy.

is less

energy

less

gamma

ray emission (or internal conversion), even

is

647

if

the lifetime for decay

extremely long.

Although

we

a task, which

it is

shall avoid,

even to write expressions for

the matrix elements of the higher electric

and magnetic moments, the real


difficulty in evaluating these matrix elements arises from the lack of
information concerning the nuclear eigenfunctions which they contain.
This is the same difficulty that arises in evaluating the matrix elements for
beta decay transitions. However, if some nuclear model is assumed, the
eigenfunctions for that model can be found, the matrix elements can be
evaluated, and predictions for the gamma decay transition rates for spontaneous emission can be obtained. The accuracy of these predictions will,
of course, be only as good as the accuracy of the model. Using the shell
model with a nucleus of radius R, Blatt and Weisskopf have obtained the
following expressions for approximate values of the transition rates S
for spontaneous emission in electric and magnetic transitions.
For electric transitions:

^ TJfiL(W"(f
W W
L[l'3-5---(2L+l)]\L

For magnetic

_ 82)

(1(i

3/

transitions:

+
s^ L[l-3-5---(2L+
./r-....
i^j(r'(r'^
+
1.9(L

1)

. r

where

L =

l)]

for dipole transitions,

3 for octupole transitions, etc.,

=
R =
E =
S

with

the energy of a

gives values of t

several values of hv,

l.2A 1A

gamma

10~ 13

cm

gamma

\L

L=

2 for quadrupole transitions,

and where

10 21 sec" 1

10- 13
197

cm
Mev

ray for which 2ttR JX

1/5, in seconds, calculated

and for

with

A =

R =

200.

ray transition indicated,


value of L, as

1.

Table (16-3)

from these equations for

cm corresponding to R =
can decay only by the single

10~ 13

If a state

is equal to this t. Note that


and magnetic transitions with any
well as the relative magnitudes of S for electric

the relative magnitudes of

common

its

lifetime

for electric

648

THE NUCLEUS

[Ch. 16

L and the relative magnitudes of S for


magnetic transitions with various values of L, agree with the implications
of the previous paragraph. Also note that for each value of L the dependence of S on the energy hv of the emitted gamma ray agrees with the
transitions with various values of

implications of that paragraph since X

dependence before

cjv.

We

have seen

energy

this

in the case of electric dipole transitions in

atoms

(the

v 3 factor in equation 13-99).

TABLE
The Reciprocal

Transition

10

Mev

Elec. dipole

Mag. dipole

4 x lO" 19
4 x 10" 17

Elec. quadrupole

Mag. quadrupole

Elec. octupole

Mag. octupole

Elec. sixteenpole

Mag.

sixteenpole

(16-3)

Model Transition Rates for

of the Shell

x 10 -16
x 10" 14

x
x
x
x

Mev

0.1

6 x 10" 18
6 x 10" 14
1

lO" 13
10" 11

10~ 10
10~ 8

Gamma

Mev

0.01

4 x 10~ 13
4 x 10" u

x 10" 11
x 10" 9
x 10 -6

x
x
x
x
x
x

1
1
1

x 10~ 4
x 10" 1
x 10 1

1
1
1

Decay

Mev

10" 6

4 x 10" 10
4 x 10~ 8
2 x 10~ 2

10" 4

2 x 10

10 1

10 3

10 8

x 10 10
x 10"

10 10

x 10 19

x 10 8

Since the electric dipole transition has the largest transition rate for

a given value of

hv, the question

immediately

Why

arises:

do not

all

gamma
is

ray decays proceed by an electric dipole transition? The answer


that in a gamma decay between initial and final states of given spin

quantum numbers
sitions are allowed

For

and if and given parities, only certain types of


by the following selection rules.

i(

tran-

electric transitions:
\i

A/

<L< +
i{

(but not

it

-+

i,

0)

(16-83)

The nuclear parity must change if L is


odd and must not change if L is even.
For magnetic
\it

transitions:

Ai

^L^ +L

(but not L

i{

= > =
i,

0)

(16-83')

The nuclear parity must change


even and must not change if L

The gamma decay

will

if

is

odd.

is

always proceed by the allowed transition having


Because of the strong L dependence of S, this

the largest transition rate.

means a

transition with

L=

Ai.

electric transition if the initial

If this value of

and

final states are

is

odd,

it

will

be an

of the opposite parity,

GAMMA DECAY AND INTERNAL CONVERSION

Sec. 12]

and a magnetic
value of

same

is

parity,

transition

even,

it

these states are of the

if

be an electric transition

will

and a magnetic transition

if

same

649

parity.

If this

these states are of the

they are of the opposite parity.

if

always the case, the selection rules for the various transitions arise
from the form of the matrix elements of the interaction potentials which
induce these transitions. The part of the rules relating the possible values

As

of

is

to the values of

and

i(

i,

that all these matrix elements

tum

in the transition,

momentum

angular
subject

it is

shown

can be understood most easily if we realize


as to conserve angular momen-

must be such

and that the

gamma

ray

must therefore carry away

In advanced treatments of this

as well as energy.

magnetic transitions associated

that, in the electric or

with a certain value of L, the quantum number specifying the magnitude


of the angular momentum carried by the gamma ray is L, and also that
ray cannot carry zero angular momentum. Granting these
statements, and writing the first part of the selection rules as

gamma

L=
we

see,

\h

i \,
f

\h

if\

1.

( but

this part

by comparison with (13-40), that

not

'i

simply represents the

requirements of conservation of angular momentum.


To understand the second part of the selection rules,

we must actually
potentials.
interaction
of
the
some
write
We do this for the simplest casesthe electric and magnetic dipole interaction potential matrix elements, ignoring the motion of the CM of the

down

the matrix elements for

These

nucleus.

are, respectively,

E-j"v/*2iWlT

H
The parameter a which

jy*

is

I {bjL, +

+e

(16-84)

(16-84')

c,S,M dr

for protons

and

for neutrons, relates the

of the nucleons to their electric dipole moments the parameters


bj and c which are the orbital and spin g factors for protons and neutrons,
relate the orbital and spin angular momenta L 3 and S,- of the nucleons to
positions

r,

their

magnetic dipole moments

the sums, taken over

all

the nucleons of

the nucleus, give the total nuclear electric and magnetic dipole moments.f
The matrix elements of these dipole interaction potentials initially contain
inside the integral signs the dot product of the

moments with

the electric

and magnetic fields E and H. However, both fields have a negligible


spatial dependence over the nucleus as far as the dipole moments are
t

The

effect

of the

matrix elements.

CM

motion

is

to increase the contribution of neutrons to these

650

THE NUCLEUS

concerned, and can be factored out.


section, r 3

is

[Ch. 16

Now, as we mentioned in the previous


S are axial vectors. By inspecting

a polar vector and L, and

Coulomb's law and Ampere's law, the reader can immediately verify that
E is a polar vector and H is an axial vector. Thus the electric dipole
interaction potential matrix element

is a dot product of two polar vectors,


the magnetic dipole interaction potential matrix element is a dot product
of two axial vectors, and so we see that both are scalars, in agreement with

a statement

made

in the previous section.

vanishes identically, unless


the polar vector o 3 r 3

y> t

and

y> f

We

also see that (16-84)


are of the opposite parity, because

of odd parity and,

if the entire integrand is of odd


from each point (x, y, z) will be
canceled by the contribution from the point ( x, y, ). We further see
that (16-84') vanishes identically, unless ip and rp are of the same parity,
t
as the axial vector 6,L, + c,S, is of even parity. Both of these conis

parity, the contribution to the integral

clusions are in agreement with the second part of the selection rules for
L
imagine evaluating the dot product in the electric dipole
1.

Now

interaction potential matrix element (16-84).


integral containing

This gives the

sum of an

x j in the integrand, an integral containing y


}

integrand, and an integral containing

in the integrand.!

in the

As we know,

the terms x t , y t z } are of odd parity. The electric quadrupole interaction potential matrix element consists of a similar set of integrals, each
all

containing a term such as xj, x{y t yj, etc., in the integrand. | As each of
is of even parity, it follows that all the integrals vanish iden,

these terms

%p t and y> f are of the same parity.


This is in agreement
with the selection rule quoted in (16-83) for L
2.
For L
3 the
integrals contain terms of odd parity such as xf, x-y-z^, etc., and this leads

tically unless

to the selection rules quoted in (16-83) for

L=

3.

Similar things happen

in the integrals that comprise the magnetic transition matrix elements,

and these lead

to the selection rules

quoted in (16-83') for the higher

values of L.

In

many gamma decay transitions,

and hv

is

along with the


cally to

particularly those in

which L

is

large

small, several groups of monoenergetic electrons are emitted

gamma rays. The energies E of the groups are found empirigamma rays by the equation

be related to the energy hv of the

E= hv - W

(16-85)

where the value of Wiox the most prominent group


energy of a

K shell electron

of the

gamma

is equal to the binding


decaying atom, and the values

t In this form the interaction potential matrix element (16-84) times its own complex
conjugate can easily be identified with the expression (13-77) that we derived for the
electric dipole transition rate
% Cf.

of atoms.
footnote on page 645.

GAMMA DECAY AND INTERNAL CONVERSION

Sec. 12]

651

of W-'for the other groups are equal to the binding energies of the electrons
in the L, M, N, etc., shells. This makes it quite evident that the various
groups of electrons are emitted from the various shells of the atom, so
that the energy of one of these electrons upon leaving the atom equals the
energy of the gamma rays less its binding energy. The similarity between

(16-85) and the photoelectric effect equation (3-8) suggested to


coverers, Ellis

and Meitner

(1921), that the electrons are ejected

its dis-

through

gamma ray delivers


energy to an atomic electron by the photoelectric effect before
leaving the atom.
a process of internal conversion, in which an emitted
all

its

However, calculations of Taylor and Mott (1933) show that the probainternal conversion by the photoelectric effect is very small
compared to the probability for internal conversion by the direct transfer
of energy from the nucleus to the electron through the Coulomb interaction between the two. In this process the nucleus is de-excited and the
electron is ejected. Except for the differences in the interaction potential
and in the initial and final eigenfunctions for the system, the calculations
are quite similar to those involved in beta decay theory. In recent years
Rose has used this picture of direct energy transfer to make a number of
bility for

calculations of the ratio of the probability for emitting an electron

given shell to the probability for emitting a


called the interna! conversion coefficient a.

The

gamma

ray.

from a

This ratio

is

calculations of a should be

very accurate because the factors involving the


eigenfunctions cancel

known Coulomb

accurately

known

initial and final nuclear


each other, leaving only factors involving the

interaction

eigenfunctions for the

electron.

initial

potential

bound

and the accurately


and the final free

electron

Figure (16-44) shows a 7C the value of a for the emission of a AT


,

shell electron, as calculated for the 40 Zr

atom. Because the probability


of internal conversion increases as the value at the nucleus of the bound
electron eigenfunction increases, a /c rapidly becomes larger as
becomes

For the same reason, at a given Z and hv, ol k is usually larger than
y.
L or a u or a Y etc. Furthermore, at a given Z and hv, the quantity
yK jy. L depends quite sensitively on the character of the gamma ray
transition, that is, on the value of L and on whether the transition is
electric or magnetic. Thus accurate measurements of a. ja.
K L which are
relatively easy to make, provide a very good method of determining the
larger.
.

character of a

gamma

ray transition and, therefore, the relative spins and

parities of the nuclear states involved in the transition.

compete with gamma ray emission in the


The processes are independent,
and the probabilities for the two processes are additive. Thus the total
rate S for transitions between the initial and final nuclear states is the sum
Internal conversion does not

sense that one process inhibits the other.

THE NUCLEUS

652

[Ch. 16

10>

10

ia

io

10

g io~

10

-3
i

10"

*
10"
0.4

0.2

0.1

Nuclear transition energy (Mev)

Figure

16-44.

The K

shell

internal

conversion coefficients.

broken curves, magnetic transitions. From


McGraw-Hill Book Co., New York, 1955.

transitions;

of the transition rate

Sic

SY

for

gamma
That

for internal conversion.

Solid

R. D. Evans,

curves,

electric

The Atomic Nucleus,

ray emission plus the transition rate

is,

Sy

Sic

(16-86)

This can be written in terms of the internal conversion coefficients since

sic
where

<x

is

( k a-

ai

<*-m

-)S y

If the initial state

S,(l

can decay only to a single

==
S

final state, as is usually true

shows that the

lifetime r

initial state is

Thus

for isomeric transitions, equation (13-118)

As

a<S y

the totai internal conversion coefficient.

of the

S 7 (l

(16-87)

()

can be accurately calculated, the experimental values of t can be

compared with the values predicted by the Sy of (16-82) and

(16-82').

Sec. 12]

GAMMA DECAY AND INTERNAL CONVERSION

653

Figure (16-45) shows such a comparison for a group of isomeric transitions


that have been identified as magnetic sixteenpole (A/
4, parity change).

The dots are experimental and the curve is theoretical. For these transitions, we see that the predictions of (16-82) and (16-82'), which are based
on the shell model, are reasonably accurate. But inspection of figure
(16-23) will demonstrate that in the shell model all A/ = 4, parity change

100

200

Energy (kev)

Figure 16-45.

The

lifetimes for a

M. Goldhaber and A. Sunyar, Phys.

500

1000

group of magnetic sixteenpole transitions.


906 (1951).

From

Rev., 83,

transitions are between states quite near those filled at the

magic numbers,!
where the model should be accurate. For transitions of
a different character, the shell model predictions are not so accurate.
However, the shell model does explain one striking feature of gamma

and

this is exactly

ray transitions from nuclear states of excitation energy of several Mev or


less. In transitions from these states, which are the states excited by beta
decay, electric dipole transitions are almost never observed. The selection

show that electric dipole transitions require A/ = 1, parity change.


But figure (16-23) shows that there is a particularly large energy difference
between any two states for which the transitions satisfy this condition.
rules

The

parity of a shell

example of a A<
and 3p,A
.

model

4, parity

state is even if / is even [cf. equation (13-112)]. An


change transition would be the transition between 5g%

THE NUCLEUS

654

[Ch. 16

A^
CO"!

^ 3 =*X
c

as

"?o

^ o
O 3

E<
I =

I-

LU?
!5<
txo

E X
suo)OJd )0 JsquinN

PROPERTIES OF EXCITED STATES

Sec. 13]

This energy

65S

almost always greater than several Mev, and so it is enerwhich can decay by an electric dipole

is

getically impossible for a state

transition to be excited

energy

is

by beta decay. However, when enough excitation


and they have

available, electric dipole transitions are found,

transition rates

which are not

in disagreement with theory.

Properties of Excited States

13.

In addition to the information about the excited states of nuclei obtained


from the study of internal conversion, gamma ray emission, and alpha
and beta decay to excited states, much information is obtained from the
study of nuclear reactions. Consider the general case (16-16), which is
written

a
If the

A^B + b

energy of the bombarding particles a

distribution of the product particles b


is

usually found to consist of a

is

is

held constant, and the energy

measured,

this

energy distribution

number of discrete groups. Figure (16-46)

shows an example the energy spectrum of protons produced in the


reaction 1 H 2 + 13 A1 27 -- 13 A1 28 + 1 H 1 for a bombarding deuteron energy
of 2.1 Mev. (A nucleus of the hydrogen isotope 1 H 2 is called a deuteron.)
The labeled groups can be identified as originating from impurities in the
target foil, but all others are proton groups from reactions on the target
nucleus 13 A1 27 When a proton is emitted in the highest energy group, the
residual nucleus 13 A1 28 is left in its ground state; when a proton is emitted
in a lower energy group, the 13 A128 is left in one of its excited states. Thus
the spectrum of excited states in the residual nucleus is immediately
obtained from the energy spectrum of the product particles [after using
.

equation (16-20) to

This

is

shown

make

small corrections for the recoil of this nucleus].

in figure (16-47).

Excited states can also be located by measuring the cross section for
reactions between a

bombarding

and a target nucleus, while


With low A nuclei and
low bombarding energies, the cross section is usually found to exhibit
resonances, or certain energies at which the probability of reactions taking
place is unusually large. As an example, figure (16-48) shows the relative
cross section for producing reactions by the bombardment of 13 A1 27
particle

varying the energy of the bombarding particle.

with

W.

Resonances occur at energies that excite the system consisting of the


bombarding particle plus the target nucleus, which is called the compound
nucleus, to one of its excited states. This is because at these energies the

THE NUCLEUS

656

[Ch. 16

4~
i

>

0)

c
LlJ

2-

Figure 16-47.

The energy

and A. Sperduto, Phys.

levels of

the nucleus

13

AI 28

From H.A. Enge, W. W. Buechner,

Rev., 88, 961 (1952).

eigenfunction for the bombarding particle has a particularly large value


in the region of the nucleus,

and thus the cross section

is

greatly increased.

These resonances are analogous to the resonances discussed in section


6, Chapter 15, and the excited states of the compound nucleus are analogous
to the virtual states mentioned there. This is a very interesting situation.
Since the

compound

nucleus

is

obviously energetically unstable to the

some other particle, it might


bombarding
excited
states
would be observed if
well-defined
no
unbound, its allowed energies form a continuum. We shall

re-emission of the

particle, or

be expected that
a particle

is

can actually exist


though the total excitation

see in the next section that well-defined excited states

for sufficiently long times to be observable, even

PROPERTIES OF EXCITED STATES

Sec. 13]

657

15.5

0.500

0.600

uy U uu

0.700

0.800

0.900

1.000

Proton energy (Mev)

Figure 16-48.

Resonances

in

1.100

1.200

1.300

1.400

the cross section for producing reactions by the bombardK. J. Brostrom, T. Huus, and R. Tangen, Phys.

ment of aluminum with protons. From


71,661 (1947).

Rev.,

energy

is

particle,

the

great

enough

to liberate a particle if it

because the energy

compound

nucleus,

and

is

usually divided

is all

concentrated on that

among many

particles of

also because reflection at the nuclear surface

inhibits particle emission.f

The data of

figure (16-48) locate states of the compound nucleus


a range of excitation energies starting at 0.5 Mev above the
11.59-Mev binding energy of a proton to that nucleus. Lower energy
protons cannot be used because they have negligible probability of getting
through the Coulomb barrier of the target nucleus. However, in reactions
14

Si 28 in

t These very sharp many particle resonances are definitely not the same as the very
broad single particle resonances, mentioned in section 3, that are observed in the cross
sections of nuclei for several Mev nucleons. The distinction will be made apparent in
the next section. In atoms the broad single particle resonances are just the Ramsauer

resonances in the cross section for scattering electrons, but the sharp many
have never been observed.
This is due, presumably, to difficulties in obtaining adequate electron energy resolution
(~10- 6 ev?). However, sharp many particle resonances probably do exist because the
c?-o^"potential for the interaction of a neutral atom and a free electron is qualitatively
similar to that in the nuclear case, in that reflection from the "surface" should inhibit
effect

particle resonances in the cross section for capturing electrons

electron emission.

THE NUCLEUS

658

compound

induced by neutrons states of the


the

way down

[Ch. 16

nucleus can be located

all

to excitation energies equal to the binding energy of the

neutron to that nucleus, which is typically 6 to 8 Mev.


By measuring the full width at half maximum T of the resonances
observed in these experiments, the lifetimes of the states may be obtained

from the

relation

(16-88)

HIT

which applies to the widths and


lifetimes of states excited by the resonant absorption of quanta (including
gamma rays). Its applicability to states excited by the resonant absorption
of particles has been justified by Breit and Wigner. This will be discussed
This

is

same

the

relation as (13-125),

in the next section.

proton or neutron experiments using bombarding energies


than several Mev, only states of the compound nucleus with spin

For
less

either

quantum numbers very

close to the spin

nucleus will be excited.

The reason

particle is so small that

momentum (KR

1)

it

is

quantum number of the target


bombarding

that the energy of the

must interact in a

and, therefore, angular

state

of low orbital angular

momentum can be conserved

the spin change is low. In higher energy nuclear reactions, more


can be excited. Furthermore, experimental and theoretical studies
of the angular distributions of the product particles can sometimes give
information about the spins and parities of the states of the residual

only

if

states

nucleus that are

excited after the emission of these particles.

left

some feeling for the procedures used in studying the


and parities of the excited states, let us consider
studies. Briefly put, there seems to be little rhyme or

Now that we have

energies, lifetimes, spins,

the results of these

reason in the

results.

That

is,

few of the regularities, which are so


an atom, are found in the excited

characteristic of the excited states of


states of
1.

a nucleus. There are only three well-established exceptions:

In the

first

few excited

states

of nuclei, particularly those with

N or

Z near magic numbers, certain shell model characteristics are often found.
An

example, which

we

discussed in the previous section,

is

the absence of

between these states.


2. In the first few excited states of even N, even Z nuclei, particularly
and Z far from magic numbers, certain characteristics of the
those with
rotational motion predicted by the collective model have been found.
For a rigid body rotating about a point at the center of its axis of symelectric dipole transitions

metry, the eigenvalues are

E =

21

i(i

1),

0, 2, 4, 6,

(16-89)

PROPERTIES OF EXCITED STATES

Sec. 13]

659

where /is the moment of inertia about the axis of rotation, and the parities
of all the eigenfunctions are even.j Every even N, even Z nucleus has
an i = 0, even parity ground state. In many of them the first excited
= 2, even parity, and the second excited state is i = 4, even
state is
;'

parity.

some

In

cases the third excited state

more, the energies of these


able values of
3.

even parity. Further-

6,

equation (16-89) for reason-

/.

In certain sets of light nuclei having

few excited

is

states agree with

After corrections are

made

of the different nuclei of a

approximately equal.

common

values of A, the

found to have corresponding spins and

states are

for the different

set,

This

is

Coulomb

first

parities.

repulsion energies

the energies of the corresponding states are


particularly true

if the' set

consists of

two

mirror nuclei.

Beyond

first few excited states, even these modest beginnings of


no longer observed, and the energies, lifetimes, spins, and
seem to be essentially random. But this should not be surprising

the

regularity are
parities

the nucleus

is

much more complicated than

theory of the atom, the net potential V(r)

mining the behavior of the electrons, and


perturbations. In the shell
V(r),

but the other

because

is

it

the atom.

the

is

all

in deter-

the other effects are small

model of the nucleus there

effects are

In the Hartree

dominant factor
is

also a net potential

not small perturbations.

Furthermore,

only a model instead of a theory, there are additional

and probably a number of others) that the


and these effects are also not small perturbations.

effects (the collective effects,


shell

model

leaves out,

Apparently in a real nucleus there are a number of different effects of


comparable importance that govern the properties of the excited states
and, as a result, it is not possible to give a description of the individual
properties of every excited state, except

However,

it is

by

them

listing

in a table.

possible to give a simple description of certain average

With respect to their energies, this is done


by specifying the average energy spacing D between adjacent states. Some
properties of the excited states.

values of

are given in table (16-4).

from one case

to the next,

average spacings must be taken


indicate the general trends,

Because of the large fluctuations

and also because of large

cum grano

uncertainties, these

Nevertheless, they do

salis.

and these trends are easy

to understand.

In the shell model, at very low excitation energies only one of the nucleons carries the total excitation energy of the nucleus, and so the spacing
Chapter 10. The / = 1, 3, 5,
states are absent because
odd, whereas only even parity states are possible since the
symmetrical system to which equation (16-89) applies is transformed into itself in the
parity operation; that is, the system is completely symmetrical with respect to reflection
t Cf. results of exercise 6,

in these states the parity

in the origin

is

and must therefore be described by eigenfunctions of even

parity.

THE NUCLEUS

660

TABLE

[Ch. 16

(16-4)

Average Spacings between Excited States of Nuclei


Excitation

energy

~10 6
~103
~102

~0 Mev
~6 to 8 Mev
~16 Mev
of the excited states
the shell model.

is

~ 20

~ 200

ev

~10 5

ev

~10 ev

ev

ev

the relatively large spacing of the low lying states of

In the collective model, the low lying excited states are

mainly rotational, and


excitation energies

their spacing

no known model

is

At higher

also relatively large.

adequate, and the nucleus

is

best thought of as a complicated system of

many

particles all of

is

perhaps

which can

ways in the excitation of the system. As there are a


number of divisions of excitation energy between the many particles
lead to approximately the same total excitation energy, it is evident

participate in various
large

that

that the spacing between the higher excited states of the system will be
small.

It is

also evident that this spacing will decrease with increasing

and with increasing A.


approximate description of the energy dependence of the average
spacings that are observed can be obtained by using the Fermi gas model

total excitation energy E,

An

Simple thermodynamic arguments, starting

for the highly excited nucleus.

from the

relation

E = aikTf
between the

energy

total excitation

of a Fermi gas and

T, lead to the following expression for the density

Po

(16-90)

e^VE

its

temperature

of excited states p

l/D:

(16-91)

The quantity a = tj^jV lAEt where jV is the number of Fermi particles in


the system and Ef is its Fermi energy (12-38). Treating the neutron gas and
the proton gas separately, and evaluating yT and Ef in terms of N, Z, and
the nuclear volume as in equation (16-35), we find that
,

where a

The

is

measured

in

~ A/10

(16-91')

-1

Mev

lifetimes of the excited states are usually expressed in terms of

As we have indicated, the relation (16-88) is valid for all


Thus any excited state with lifetime t has a full width at
half maximum Y, where
their widths V.

excited states.

HIT

(16-92)

PROPERTIES OF EXCITED STATES

Sec. 13]

661

We

have seen that for excitation energies up to several Mev the excited
can decay only by gamma ray emission and internal conversion,
and that the lifetimes range from about 10~ 15 sec to as long as 10 8 sec.
states

Since h
to

less
is

~ 10~
10~ 23

15

ev.

means that Y ranges from about T


lO" 1 ev
For excitation energies somewhat above this range, but

ev-sec, this

>

than the binding energy of a nucleon, the evidence indicates that T


E or A. It varies from r <~ 1 ev near the

not strongly dependent on

neutron binding energy in nuclei with


in nuclei with

neutron binding energy

in this energy range there are very

20,

<

~ 200.

many

to

T~

10"" 1

ev near the

This comes about because

lower energy states to which

the excited state can decay, so there will always be states available

and

whose

such as to allow decay of the excited state by electric


dipole transitions with lifetimes of the order of 10~ 15 sec.
spins

parities are

As the excitation energy goes above the binding energy of the last
nucleon in the nucleus, typically 6 to 8 Mev, decay by the emission of a
nucleon becomes energetically possible. In some nuclei, proton emission
becomes energetically possible first, and
possible

but in

first,

all,

in others

neutron emission becomes

except those of very small Z, the

Coulomb

barrier

proton emission until the excitation energy is quite


high. When nucleon emission is energetically possible, the total rate for
transitions from the excited state will be larger than that due to transitions
involving gamma ray emission and internal conversion alone. This
effectively inhibits

decreases the lifetime and increases the width.


inversely proportional to lifetimes

and

In fact, since widths are

lifetimes are inversely proportional

to transition rates, widths are proportional to transition rates

wnte

= ry + r B

because transition rates are additive.

width due to

all

and T n means

gamma

(16-93)

In this expression,

Y y means

the

ray and internal conversion transitions,

the width due to neutron emission (or nucleon emission if

proton emission
It is

possible

and we can

is

important).

easy to see what the general form of the energy dependence of the

neutron width

T n must

be.

Imagine using Golden Rule No. 2 to calculate


initial state consisting of the nucleus in an

the rate for transitions

from an

excited state of energy

toa

final state consisting

of the nucleus in

its

ground state plus a nucleon moving off with kinetic energy (E Eb ),


where Eh is its binding energy. The matrix element for the transition will
be a slowly varying function of E, and the energy dependence of the transition rate comes in mainly through the density of states for the nucleon.

From
to

equation (12-37), we see that this density of states


)'A
Therefore we expect
b

(E E

is

proportional

Tn

x(E- E

)'

(16-94)

THE NUCLEUS

662
at least for a restricted energy range.

This

[Ch. 16

the energy dependence

is

When r > T y
and these measurements simply involve finding the widths

exhibited by the average results of the measurements.

then

r,

of the resonances in data such as those given in figure (16-48).

Ty

oc constant,

compared

to

nuclei with

~
~

Tn

while

T y when (E

~ 20,

this

E)

(E

cc

Yn

Vi
,

will

Since

always become large

becomes larger than some value. For


occurs for (E Eb )
1 ev, and at this value
b)

10" 1 ev at
T = T y
1 ev.
For nuclei with A
200, F n = T y
3
(E - Eb)
10 ev.
When (E Eb) is not too large, there is an interesting empirical relation
between the average neutron widths r of the excited states of a nucleus
and the average spacings D between these states. It is

~ CK D

(16-95)

where

K =p
2

The quantity p 2 is
mass. For almost

the

2 \h

= V2M(E-E )lh
b

momentum

all nuclei,

C=

of the emitted nucleon, and

the constant
0.5

10" 13

is

is its

within a factor of 5 of

cm

(16-95')

T n vary by many orders of magnitude.


from (16-91) and (16-94), we see that TJD increases
with increasing excitation energy. At the point where TJD = 1, the
excited states begin to overlap and form a continuum. As an indication of
the general trend, it might be said that for nuclei with A <~ 20 the continuum
begins at (E Eb )
10 Mev, and for nuclei with A
200 it begins at
10" 1 Mev.
(E ~ Eb)
although the values of

From

(16-95), or

14.

Nuclear Reactions

Nuclear reactions provide


of excited

states.

much

useful information about the properties

In turn, the properties of excited states provide the key

to the understanding of nuclear reactions. In this section

we

shall use this

point of view to give a simplified version of the present theoretical description of nuclear reactions.

and Wigner

This description

is

due to Bohr (1936), Breit


and many others.

(1936), Weisskopf (ca. 1950), Butler (ca. 1953)

Before the discovery (1936) of very sharp resonances in the cross sections
by low energy neutrons, it had been assumed that

for reactions induced

nuclear reactions involve a direct transition from an

of the bombarding particle

initial state

and the target nucleus to a

consisting

final state consisting

NUCLEAR REACTIONS

Sec. 14]

663

of the product particle and the residual nucleus.

However, the discovery

of resonances replaced this direct interaction theory by the compound


nucleus theory. Bohr proposed the latter theory because he realized that
in reactions exhibiting resonances there

compound

long lifetime, which he called the

relatively
is

must be an intermediate
nucleus.

state of

This state

necessary because such reactions are essentially measurements of the

incident neutron energy, since the cross sections are very rapid functions

of energy. Therefore the uncertainty principle demands that the duration


hjT, where the resonance widths
of these measurements be at least t

The compound nucleus theory was

are their energy resolutions.

number of years

quite

was thought that it could explain


Recently, however, it has been found that some
all nuclear reactions.
reactions actually are most easily explained on the basis of a direct interand

successful,

for a

it

action theory.

We

shall describe first the

to distinguish

compound

nucleus

separated since

nucleus

rjD >

is

(a),

nucleus theory.

E)

(E

is

excited into a region in which

TJD <

(b),

(E

b ) is

large

It is

convenient

small enough that the


its

states are well

enough that the compound

which its states strongly overlap since


both cases the initial step in the nuclear reaction
This is the step described by the optical model.

excited into a region in

is
1

compound

between two cases:

However,

in

essentially the same.


Consider a beam of particles bombarding a target of nuclei. According
to the optical model, a particle striking a nucleus will either be scattered

is

beam or removed from that beam by absorption. If it


absorbed it will initiate a nuclear reaction, but the optical model cannot
tell us anything about the details of the reaction because it only describes
the particle moving through the absorptive nucleus in a single particle
state of the complex potential that represents the initial particle-nucleus
interaction. Because the potential is fairly absorptive for all positive
values of (E ),t this single particle state will have a very short lifetime,
or very large width. In fact, since it immediately follows from (15-106)
out of the incident

is

that, if the

W,

imaginary part of the complex potential has a constant value

the width of the single particle state

(16-10) shows that


the lifetime
ev-sec/10 7 ev

The very

2W ~

10

Mev

of the single particle


10~ 22 sec.

many

consisting of the

2W, and

state

is

(E

typically

since figure

- E ~ 10 Mev,
t = hjT ~ 10^ 15
b)

particle states that correspond, within the limits of

bombarding

E ) for the
E of figure (16-10).

Here we use (E

equals the

short lifetime single particle state rapidly decays into the

the uncertainty principle, to the

longer lifetime

It

is

for the typical case

same

total excitation energy

of the system
This is the

particle plus the target nucleus.


external kinetic energy of the

bombarding nucleon.

THE NUCLEUS

664

system we

compound

call the

These many particle

section.

One

many particle states


we were discussing in the

nucleus, and the

precisely the higher excited states that

particle state.

[Ch. 16

states are certainly different

difference

is

that in the cases

we

from the

are
last

single

are discussing the

measured in Mev while the widths of


measured in ev. Another is that the single
particle state and the many particle states do not describe the same physical
situation, even though they correspond to the same total excitation
energy. In the single particle state all the excitation energy is carried by a
single particle, but in any of the many particle states it is distributed over
a large number of particles. The excitation of the many particle states by
the single particle state takes place through a process in which energy is
transferred from one to the other in multiple collisions of the bombarding

width of the single particle


the

many

state is

particle states are

particle with the particles of the nucleus.!

Of

course, the

many

particle states

of the

eventually decay with an average lifetime r

take

~ 10 ev as a typical many

compound nucleus

hjY <~ 10~16

particle state width.

will

where we

sec,

We know that,

in

must be by nucleon emission since, except at small


(E Eh), r > T y for many particle states. But it is perhaps not apparent
how this can happen. The point is that in the many particle states the total
excitation energy E is distributed among many nucleons, and in order for
any nucleon to escape from the nucleus the nucleon must have an excitation energy at least equal to its binding energy Eb Thus a large fraction
of the excitation energy must somehow become concentrated again on a
single nucleon if it is to escape. Some idea of the way this comes about,
and also some idea of the process in which the many particle states are
initially excited, can be obtained from the following argument.
Let *F be the wave function for the compound nucleus, and y> k be the
general, this decay

eigenfunctions for the

Assume

many

particle states that are excited in that nucleus.

that the spacing between

particle states

is

all

adjacent eigenvalues

Ek of these many

a constant equal to their average spacing D, so that

Ek = E +

kD, where E is a constant and k = 1, 2, 3, ...


=
wave functions Tj. for the many particle states are
k
-*>c:Dt f
ip e-iEo l> e
linear
of
the
Next
write
as
a
combination
k
t

Y=2
k

Then the

tp

^=
k

e-^l* J

a kWk e

~ iWtin

lE" tin

That

is,

k e-

(16-96)

where the summation is taken over the many particle states that are excited.
In this expression t =
is the time at which they are initially excited, and
t This gives rise to

a simple relation between the sharp

the broad single particle resonances


strength,

TJD,

of the

many

the

many particle

resonances and

single particle resonances define the average

particle resonances.

NUCLEAR REACTIONS

Sec. 14]

= A ke'

ak

l4k

constants

same

<f>

zero.

where the

665

real quantities

Ak

are their amplitudes

are phase factors that allow the

Now

and the real


have the

in all the ^Yk to

note that the summation appears to be a perfectly


t, with repetition period

periodic function of

T=2TThjD

(16-97)

At t = 0, *F actually describes a single particle state, since at that time the


compound nucleus is just being formed and all the excitation energy is
carried by the bombarding particle. If the summation is perfectly periodic
with period T, then

(and also IT, 3T,

again describe a single particle state at

will

In fact,

etc.)

it

will describe exactly the

same

= T

single

was initially formed, and all the total excitation energy


be concentrated on the single particle. This particle will then escape
the nucleus, providing it is not reflected by the change in potential at
particle state that

will

the surface of the nucleus.

Let us consider case

which the quantity (E

Here
b ) is small.
For a proton, the
Coulomb barrier makes the probability of escape negligible. For a neutron,
the probability of escape, P, can be estimated from equation (8-46). It is
in

(a),

reflection at the nuclear surface

is

very important.

P = 4A 1 A 2 /(A 1 + K2f

(16-98)

where

K = pjh ~ V2MVJH
x

and

K = P% \h = VlM(E - E )lh
2

The quantity p1
the nucleus.

the

is

We

momentum

of the neutron of mass

equate this to the

momentum

it

M when inside

would have, approxi-

when moving in an optical model potential with real part of depth


The quantity p 2 is the momentum of the neutron after escaping.

mately,

(E

Equation
b ) as shown.
based on the approximation that the potential at the surface of
the nucleus changes rapidly in a distance of the order of 1 de Broglie
wavelength. This approximation is not bad for small (E Eb ). FurtherIt is

related to the external kinetic energy

(16-98)

is

more, with small (E

E K >K
b ),

and

so

P ~ 4Kx K2 lKf = 4KJK!

Now

(16-99)

the probability per second that a neutron will be emitted

nucleus

is

equal to

T" 1

number of times per second

the

that

from the

T repeats

its

by the probability P that any one


escape. Thus the lifetime for neutron

original single particle form, multiplied

of these times the neutron


emission

will

is

1/r-ip

T/P

(16-100)

THE NUCLEUS

666

hjr

H4K D
_ h4K

__

and the neutron width r


(16-99) this

Tn

is

HPjT.

From

[Ch. 16

(16-97) and

is
2

2K 2 D

Ink

*i

ttK

or

2h

Tn

;J1 K 2 D

(16-101)

where

K = ^2M( - ,)/*
V = 50 Mev that is indicated
2

Taking the value

2h/W2MV =

the coefficient

(16-101)

is

in excellent

0.39

probability
repeats

its

is

small

by the

optical model,

cm, and we see that equation


Note from equation

agreement with (16-95).

(16-100) that the lifetime t of the


the repetition period T,

10" 13

when (E

many particle

b ) is

states is

much

longer than

small, simply because the escape

the wave function Y for the compound nucleus

original single particle

form many times before the

particle is

emitted.

not true for case (b), in which (E Eb ) is large, because in these


1. To handle this case we must
circumstances the escape probability P
The quantities A k which are
argument.
correct a weak point in the
are not constants, but on
equation
(16-96),
the amplitudes of the a k in
This

is

the average decrease in time according to the lawf

A k (t) = A k(0)e~ mT*

(16-102)

where rk = hjT k is the lifetime of the many particle state of width Y k


This can be ignored for the purpose of evaluating the repetition period T
T and the A k (t) are very nearly constant over
in case (a), since all the T
any repetition period. But it cannot be ignored in case (b), since the
T does not obtain. Now, if all the rk had exactly the same
condition rk
tl2r
this term would
value t, and all the A k (t) had the same t dependence e~
outside
as a term
on
the
and
appear
(16-96)
factor out of the summation
leave the
would
nucleus.
This
compound
the
the
decay
of
describing
not
all have
r
do
But
the
of
t.
function
periodic
perfectly
a
summation
k
exactly the same values. Instead they have a spread of values, and thus
.

fc

>

>

not a perfectly periodic function of t. Even so, it will be


if the spread in the rk is not too large.
for the compound nucleus should, within a cerThen the wave function
tain time, return to a form that has an approximately single particle

the

summation

is

an approximately periodic function

character,
t

and the

particle

is

emitted on the

first

occasion on which this

the nucleus in the many particle state k


simply the usual exponential decay law.

As the probability of finding

equation (16-102)

is

is

a?a k

Al(t),

NUCLEAR REACTIONS

Sec. 14]

667

However, the particular approximately single particle wave


returns should have no necessary connection with the
exact form of. the initial single particle wave function. This will certainly
happens.

function to which

be true

if a. large

nucleus,

and

number of overlapping

if the

states are excited in the

width of each of these

states is

compound

not strongly correlated

In these circumstances, the decay of the


be independent of the exact details of its formation.
This fundamental result of the theory of compound nucleus reactions

with the widths of the others.

compound nucleus

will

also holds for case (a).


fact that the values of

period, so

Thus

actually change slightly during one repetition

T deviates slightly from being a perfectly periodic function of

W returns repeatedly to a single particle wave function that

different

from the previous

number of
all

This can be thought of as a consequence of the

A k (t)

single particle

wave function and,

cycles that occur before the particle

memory

of the exact details of the

is

in the large

emitted,

is finally

initial single particle

t.

slightly

wave

W loses

function.

In both cases, this result is only strengthened by deviations from the


assumption that the eigenvalues Ek of the many particle states are evenly
spaced. But, in any circumstances, the decay of the compound nucleus

must be independent of the exact details of its formation for case (a).
The reason is that energy conservation within the limits of the uncertainty
principle demands that only many particle states of excitation energy
t = hjY n when the
within T of the energy E be excited by the time t
compound nucleus has existed long enough to decay.f For case (a),
FJD < 1, so there can only be one many particle state within these
limits. Therefore, in case (a) the decay of the compound nucleus must
necessarily depend only on the properties of that many particle state and
cannot depend on the details of how it was excited.
After this somewhat esoteric description of compound nucleus reactions,
it might be well to recapitulate in the kind of language used in the original

papers of Bohr.

When

a particle enters a nucleus,

collides with the

it

particles of the nucleus and very rapidly gives up its kinetic and binding
energy to those particles. The bombarding particle is thereby captured, and
an excited compound nucleus is formed in a state, somewhat reminiscent
of a boiling liquid, in which the excitation energy is distributed at random
among all its particles. Each particle then has an excitation energy con-

siderably less than

its

binding energy. Occasionally,

statistical fluctuations

cause enough energy to be concentrated on a single particle to make it


If the circumstances are such that
energetically possible to escape.
f Of course, at an earlier time many particle states whose excitation energies fall
within broader limits can be excited, as we have assumed. For example, when only

10

-22

sec has elapsed (the lifetime of the initial single particle state),

over a range of

10 Mev can be excited.

Cf. section

5,

Chapter

many particle states

9.

THE NUCLEUS

668

reflection at the nuclear surface is important, this

of times before a particle

compound
lifetime

that

it

Thus the

finally emitted.

must happen a number

But, in any circumstances, the

nucleus has a relatively long lifetime and during this long


forgets everything about

must remember

it

is

[Ch. 16

how

it

was

originally

decay process

formed (except

momentum,

to conserve energy, angular

etc.).

independent of the exact details of the formation process. This means, for example, that the ratio of the probability
that the
that

it

final

is

compound nucleus will decay by nucleon emission to the probability

will

decay by gamma ray emission and internal conversion is equal


r n /T y (since transition rates are proportional to widths),

to the ratio

where T n and T y are evaluated for the state or states excited in the compound nucleus, and where this is true independent of exactly how these
states were excited. If in a certain reaction a particular compound nucleus
is formed with a particular excitation energy, and in a different reaction
(different bombarding particle and different target nucleus) the same

compound nucleus is formed at the same excitation energy, both compound


nuclei should decay in the same way. At small values of (E E), T can
y
be comparable to r, and there can be an appreciable probability that
the compound nucleus decays by gamma ray emission and internal con-

At higher values of (E Eb ),
F > F y and the compound nucleus always decays by particle emission.
If enough excitation energy is available, the compound nucleus can decay
either by neutron emission or by proton emission. For quite high excitation
energies, more than one particle can be emitted.
Figure (16-49) shows the measured cross sections for the set of reactions
version to form a completely stable nucleus.
,

+ 01
+ 01 + 01
29
Cu62 + 01 + H
30

1 _|_ 29

Cu

63

_+

30

Zn 63
Zn62

and

also for the set

He

28

Ni

60

so

Zn 63

oi

30

62

01

_|_

01

+ R

+
_+ Zn +
29
Cu62 +

01
1

bombarding energies. The abscissa is the kinetic energy of


bombarding proton for the first set of reactions, and the kinetic energy
of the bombarding alpha particle in the second set of reactions; the
relative displacement of the two abscissas is equal to the difference in Et
for the two bombarding particles. The ordinate is the cross section for
the various reactions, and each reaction is labeled according to a common
at quite high

the

notation.

compound

This figure provides a test of the prediction that the decay of a


nucleus is independent of its formation, in case (b), where

NUCLEAR REACTIONS

Sec. 14]

(E

b)

is

large

669

enough that the compound nucleus

region in which there are presumably

many

widths are not strongly correlated, because for both

compound

nucleus

^Zn64 and
,

12

sets there is the

same

13

17

21

25

16

20

24

28

32

Energy of alpha particle (Mev)

Figure 16-49.

excited into a

because the abscissas have been adjusted

Energy of protons (Mev)

is

overlapping states whose

36

40

Cross sections for a set of reactions which test the predictions of the
theory. From S. N. Ghoshal, Phys. Rev., 80, 939 (1950).

compound nucleus
to equalize

its

We see that the prediction is well verified

excitation energy.

same relative probabilities for the


modes of the compound nucleus.
It has not been possible to make such a test for case (a), where (E Eb)
is small enough that the states of the compound nucleus are well separated
and only a single state is excited and such a test is hardly necessary. But
other predictions of the compound nucleus theory can be tested for this
case. For instance, Breit and Wigner have shown that the energy dependence of the cross section for absorbing an S-wave neutron (K2 R < 1) to

because for both

sets there are the

various decay

THE NUCLEUS

670

[Ch. 16

form a compound nucleus excited to a state of energy E and width


r = r + r which subsequently decays by gamma ray emission and
internal conversion, is essentially identical with the energy dependence
(13-124) of the probability for absorbing a quantum to form an excited
state of an atom, which subsequently decays by emitting another quantum.
They found
,

o(E)
KJ

^-

K\(E

-Ef+

(r/2)

(16-103)

where

X = V2M( - E
2

The term

ir\K\

S-wave neutron

is

the

(cf.

possible cross section for absorbing an

equation 15-98 with

The term YJ?\\(E


contains a factor

maximum

- Ef +

Tn

b )lfi

rj

0,

r\

rj

r\

(r/2) 2 ] gives the resonance

1).

shape;

it

its

which, being proportional to the probability for

emitting a neutron, must also be proportional to the probability for


absorbing one (cf. equation 13-84 and the associated discussion of microscopic reversibility), and

it

also contains a factor

The term r /r
compound nucleus will decay by gamma

normalization.
excited

that

comes from

is the ratio of the probability that the

conversion to the probability that

it

will

ray emission and internal

decay by any process.

The

energy dependence of (16-103) has been accurately verified in many


different experiments. Granting its validity, it may be used to obtain
values of T, F n and T y from measurements of a{E). Breit and Wigner also
found that equation (16-88), which is identical with (13-125), applies to
the resonant absorption of particles in any situation covered by case (a).
Although the predictions of the compound nucleus theory seem to be
verified for reactions that proceed through a compound nucleus process,
it has been found in recent years that there are certain situations in which
,

reactions proceed through a direct interaction process.

Figure (16-50)

These are energy


distributions of protons inelastically scattered (scattered with energy loss)
50
Sn nuclei
at five different angles in the LAB system from a target of
bombarded by a beam of protons of external kinetic energy 31 Mev.
The abscissa is the external kinetic energy of the inelastically scattered
protons, and the ordinate is the cross section for scattering a proton into a
presents

some data

illustrating

one of these

situations.

Thus the data also give the angular


Note that in the energy
section
does not depend very
cross
the
and
20
Mev
range between 10
decreasing
angle. On both
with
rapidly
increases
energy
but
on
strongly
points, this is exactly the opposite of what would be expected from the
compound nucleus theory. The theory predicts that in the long-lived
unit energy

and

solid angle range.

distribution of the inelastically scattered protons.

NUCLEAR REACTIONS

Sec. 14]

compound

nucleus essentially

all

671

sense of direction

would be

lost,

and that

the angular distribution of the protons which are finally boiled off should

be isotropic.
high

It also

Coulomb

Z=

is their external kinetic

E=

(31

N(e)

<

Mev, and N(e)

0,

10

compound

energy and

(16-90) with

51

above the 10-Mev

nucleus with a strongly

elkT
Maxwellian-like energy distribution N(e) cc ee~

energy dependent

where

predicts that the protons boil off

barrier of the

+ E ) Mev =
b

kT is

38 Mev. That
L7Mev

oc e-

obtained from equation


10

is,

the theory predicts

Mev

<e<

31

Mev.

6i

>h

135

10

15

g(Mev)
Figure 16-50.

20

25

30

35

Energy and angular distributions for a reaction which proceeds through


From R. M. Eisberg and G. Igo, Phys. Rev., 93, 1039 (1954).

a direct interaction process.

is no Coulomb barrier, and their predicted energy


N(e) cc ce -/i-7Mev over the ent i re ran g e f i n t ^ c com .
pound nucleus theory almost all the emitted nucleons would be neutrons,
since only a few protons in the tail of the Maxwellian with
10 Mev

For neutrons, there

distribution

is

>

can escape, and it would be expected that considerably less than 1 percent
of the reactions induced by the incident 31 -Mev protons should lead to
the emission of inelastically scattered protons. But, on this third point,
the data also contradict the predictions, because the magnitude of the
cross section indicates that about

1 5 percent of the reactions lead to the


emission of inelastically scattered protons.
This does not mean that the compound nucleus theory is wrong.

Instead,

it

and high

means

that inelastic proton scattering reactions

on intermediate

Z nuclei proceed through some process, other than the compound

nucleus process, which has a cross section for these reactions so much larger
than the small compound nucleus cross section that the compound

THE NUCLEUS

672

nucleus effects are completely hidden.

The process

that

[Ch. 16

involved

is

is

thought to consist of a direct interaction in which a bombarding proton


collides with a nucleon in the diffuse rim of the target nucleus, and either
the

bombarding proton or the struck nucleon

(or both) immediately

escapes.

Consider a 31-Mev proton incident upon a 50 Sn nucleus. As the de


Broglie wavelength of the proton is several times smaller than the nuclear
radius, it is legitimate to think of the collision as being well enough localized that a distinction can be
sions with the diffuse rim.

made between head-on

Now the

optical

collisions and collimodel shows that attenuation

Mev are also several times smaller than the


nuclear radius. Thus, in any head-on collision with the nucleus, the bomlengths in nuclear matter at 31

barding proton will soon collide with the nucleons of the nucleus, these
nucleons will in turn collide with other nucleons, and so the kinetic energy
of the bombarding nucleon will rapidly be shared with

all the nucleons of


nucleus process that will
eventually lead to the emission of low energy neutrons. However, in a

the nucleus. This

is

the

first

step in the

compound

rim of the nucleus, the bombarding proton could


nucleon in such a manner that either it or the struck nucleon
(or both) would immediately escape, carrying away a large fraction of the
collision with the diffuse
collide with a

As the optical model shows that the thickness of the diffuse


rim in the 50 Sn nucleus is about 10 percent of its radius, its projected area
is about 20 percent of the projected area of the entire nucleus.
This
kinetic energy.

indicates that the cross section for this direct interaction process could be

of the order of magnitude observed experimentally.

The other

character-

of this process are such that they could also be in agreement with
experiment. The process would lead to a slowly varying energy distribution
istics

of high energy inelastically scattered protons, and to an angular distribuwhich would increase rapidly with decreasing angle

tion of these particles


(if it

were not for the

initial

momentum

of the struck nucleon, and refrac-

tion in the nuclear potential, conservation of momentum

and energy would

require the angular distribution to be concentrated entirely in the forward


90 of the LAB system).

A number

of semi-quantum mechanical calculations based on this direct

shown that it is capable of a very adequate fit


proton scattering reaction, and to many similar
reactions. Butler (ca. 1953) and others have made completely quantum
mechanical calculations for this process for reactions in which it is possible
to measure the angular distribution of a single group of inelastically scattered
protons that leave the residual nucleus excited in a particular low lying
interaction process have
to the

31-Mev

excited state.

inelastic

The

calculations are basically the

mation calculations for

same

elastic scattering studied in

as the

Born approxi-

Chapter

15,

but the

Sec. 14]

NUCLEAR REACTIONS

details are

somewhat more complicated

if

673

the scattered particle imparts angular

nucleus. Instead of giving

them

in the

very much more complicated


momentum as well as energy to the

form of a long equation, the

results

of these calculations for a particular case are shown as the solid line in
figure (16-51). The points are experimental data. This is the angular
distribution for the inelastic scattering of alpha particles, of external
kinetic energy 31

Mev, from a

target of

12

Mg24

The alpha

particles

50

.4?

40

30

20

10

V^V
40

20"

Scattering angle

60

80

(CM)

Experimental and theoretical angular distributions for a direct inter-

Figure 16-51.

action leaving the residual nucleus in a single excited state.

From

S.

T. Butler, Phys. Rev.,

106,272(1957).

measured are the highest energy


leave the residual nucleus in

excited state at 1.37

in addition to the general increase in the

angle, there are very


elastic scattering

these cases.

The

pronounced

angular distribution with decreasing

oscillations that are reminiscent of the

angular distributions. This

is

typical of

from the

what

is

found

in

locations of the oscillations in the theoretical results turn

out to be a very sensitive function of the orbital angular


ferred

and they
Mev. Note that,

inelastically scattered group,

its first

momentum

inelastically scattered particle to the nucleus,

trans-

and thus

provide a powerful tool for studying the spins and parities of the excited
states of the nuclei involved.

There may be direct interaction processes other than those involving


nucleon-nucleon collisions in the rim of the nucleus. In fact, recent evidence indicates that in some cases the highest energy inelastically scattered

THE NUCLEUS

674
particles

come from a

[Ch. 16

process in which the incident particle collides with

the rim of the nucleus but, instead of interacting with a single nucleon,
the particle interacts coherently with the entire nucleus

of

low lying rotational

The

and puts

it

into one

immediately moves
having suffered only a small energy loss and angular deflection.
its

states.

particle

Direct interactions seem to play an important role in

which a

fairly

high energy product particle

reactions where the product particle

is

all

emitted;

is

different

off,

reactions in
this

includes

from the bombarding

Important examples are reactions where a deuteron 1 H 2 is the


bombarding particle, and a proton 1 H 1 or a neutron n 1 is the product
particle. These reactions apparently go through a process in which one
particle.

of the nucleons of the deuteron is captured, while the other continues on.
Recently there have been attempts to use the formalism of the compound
nucleus theory to describe direct interactions. This

is done by dispensing
with the condition that the widths of the overlapping states excited in the
intermediate system are not strongly correlated. By adjusting these cor-

which is of course permissible since the widths are not directly


measurable when the states overlap, it is possible to make the intermediate
system do almost anything from decaying immediately in a manner
dependent upon its formation (direct interaction) to decaying after a long
relations,

time in a manner independent of

its

formation (compound nucleus).

With regard

to direct interactions, this might be criticized as

complicated

way of

describing a simple process.

However,

an overly

has the
advantage of providing a single unified theory of all nuclear reactions,
including those which are compound nucleus processes but are in the

awkward

15.

range, lying between our cases (a) and (b), where

Fn

it

~ D.

Nuclear Forces

In the preceding sections

we have concentrated on

describing the

properties of the system of nucleons called the nucleus and have largely

avoided discussing the properties of the forces acting between the nucleons
of the system, even though these properties are clearly the origin of the
properties of the system.

Nevertheless,

we have

learned something about

nuclear forces.

Nuclear forces must be attractive, at least in regard to their net


on the average, because otherwise nuclei would not exist.
2. Nuclear forces must be of short range because the spatial distribution
of the nuclear interaction potential exerted by a nucleus on a passing
nucleon extends beyond the spatial distribution of the nucleons in the
nucleus by a distance which is only of the order of 10~ 13 cm.
1.

effect

NUCLEAR FORCES

Sec. 15]

675

3. Nuclear forces between two neutrons in particular quantum states


must be nearly the same as nuclear forces between two protons in the
same quantum states, not counting the Coulomb repulsion of the protons.
This follows from the fact that the total ground state binding energies of

each

member

of a set of mirror nuclei are equal, after correction

for the different total

Coulomb

is

made

repulsion energies of the two members,

and from the fact that for one member the number of neutron-neutron
interactions between nucleons in various quantum states is equal to the
number of proton-proton interactions between nucleons in those quantum
states for the other member, and vice versa (cf. figure 16-42). As the

number of neutron-proton interactions in various quantum states is the


same for both members of a set, mirror nuclei do not allow a comparison
between neutron-proton interactions and neutron-neutron or protonproton interactions.
4. Nuclear forces cannot be such that
attractive interaction potential between all
states. This follows from the fact that
energy A is approximately proportional

there

is

nearly the same purely

pairs of nucleons in all

quantum

the total ground state binding


to

for all nuclei with

A^A,

and from the following argument due to Heisenberg (1933). Consider


some nucleus and fix attention on a particular nucleon in its interior, say
the z'th nucleon. Now assume that there is a purely attractive interaction
potential, of range a, acting between this nucleon and all the other nucleons

Then the average potential energy V of that nucleon is


number of other nucleons found within a sphere of
centered on that nucleon, and this energy is negative. That is,

of the nucleus.

proportional to the
radius

a,

where p

is

the density of nucleons.

oc

-p

(16-104)

If this were the only effect, the radius

of the nucleus would collapse to the range a of the interaction potential,

since this

would make

t,

as well as the average potential energies of all

the other nucleons, as negative as possible and, therefore, lead to the

and the most stable arrangement. But there


which tends to oppose such a collapse. It is the increase in
the kinetic energy of the nucleons demanded by the uncertainty principle
if the volume in which they are confined decreases.
For a nucleus of
radius R, the uncertainty Ap in any component of the linear momentum
of the /th nucleon must be at least
largest total binding energy
is

an

effect

APi ~ h/R
The magnitude of the

linear

momentum
Pi

vector

~ &Pi

is

THE NUCLEUS

676

and the uncertainty

because

arises

average kinetic energy of the

its

direction

nucleon

fth

unknown.

is

[Ch. 16

Thus

the

is

Ti-pZjlMKiAptfKllR 2
But
3
p oc ljK

so|

From

we

(16-104) and (16-105)

binding energy of that nucleon

+ P^

oc

(16-105)

find that the p dependence of the total

is

AE = -E = -(T + V ) = -ap +
i

where
j8

is

the negative total energy of the ith nucleon,

are constants.

The

total binding

A = 2

pP
and where a and

energy of the nucleus

A,

= -yp V3 +

is

dp

(16-106)

=i

where y and <5 are other constants. If this equation is valid, the nuclear
radius R would collapse to the range a of the potential because it would
lead to a large p and maximize A.E.% In these circumstances, all the
nucleons of the nucleus are within the range of each other's interaction
potential, and the average potential energy of the /th nucleon is proportional to the total number of nucleons in the nucleus. That is,
JAcc

As

the

same

is

true for

energy of the nucleus

-A

any other nucleon, the

total average potential

is

V=
i

K <* ~ A
2
=

(16-107)

But, as the average kinetic energy of any of the nucleons

depend on A, the

total average kinetic energy

T= 2
With a collapsed nucleus,
true for large A. Thus

T, oc

V is much

would not

is

+A

(16-108)

larger than T,

A.E= (T + K)~ -Foe

and

+A

this is particularly

(16-109)

t Note that equations (12-38) and (12-39) show that the same result is obtained for a
Fermi gas in which the exclusion principle is explicitly taken into account.
t The second term of equation (16-106) must be larger than the first term, as A
must be positive. Therefore A increases with increasing p since the second term varies
faster

than the

first

term.

NUCLEAR FORCES

Sec. 15]

We

677

see that the total binding energy

is proportional to A 2 despite the


of the uncertainty principle, under the assumption that there is the
same purely attractive interaction potential between all pairs of nucleons
,

effect

in a nucleus.

is

the

a.

But this
same radius R =

not true, and neither is it true that all nuclei have


Therefore the assumption is false. It must be

that the interaction potentials are different between pairs


of nucleons in
different quantum states, being repulsive in some states,
or it must be that
there is some kind of impenetrable repulsive core in the interaction

poten-

tials,

or else

it

must be that there

is

some combination of both

these effects.

One of these alternatives is absolutely necessary in order to keep the


nucleons far enough apart that at any instant one nucleon interacts only
with a limited number of other nucleons. Then the binding energy of
each
nucleon will be independent of the total number of nucleons in the nucleus
and the total binding energy of the nucleus will be proportional to A,
as observed. This requirement is described by saying that
the nuclear
forces

must lead

to saturation.

Most of the information about nuclear

forces that can be obtained from


complex as a saturated nucleus {A
4) is contained

>

the study of systems as

in the four items listed above.

More

detailed information

is obtained by
studying simpler systems where the effects of the nuclear forces can be
analyzed without the complications that characterize a saturated nucleus.

The

simplest of these

is

the deuterium nucleus

or deuteron, in its
This system consists of a neutron and a proton bound
together by the interaction potential for the neutron-proton force.
Its

ground

state.

observed properties are

1H2

A =

2.014732
2.22

amu

Mev

*'=!

(16-110)

= +0.8574^
=
Q +2.74 x 10"27 cm

p iM

The deuteron has a small quadrupole moment Q, which means


its

probability density

immediately shows

that

not completely spherically symmetrical. This


that the neutron-proton interaction potential is
is

complicated because the ground state would be a spherically symmetrical


S statet if the interaction potential could be written V(r), where r is the
magnitude of the vector from one nucleon to the other. [We have seen
that this

is

true in the case of the

atom and

its

Coulomb

potential

The symbols, and the quantum numbers they represent, are taken over directly
from the theory of the electronic structure of atoms, except that i is often used for; as
t

in the shell

model.

THE NUCLEUS

678

V(r) oc
V(r)

by

(10-27)

not

It is

\\r.

difficult to

prove

it

for

any

[Ch. 16

attractive potential

showing that the lowest eigenvalue of the Schroedinger equation


is

always found for

0.]

Thus

not simply of the form V(r).


the existence of a quadrupole moment is
potential

is

V(x, y, z)

V(r)

V'{r) {3(S a

the neutron-proton interaction

form which

r)(S 2

r)/r

is

consistent with

Sx S2 }

(16-1

1)

The
first term is the dominant spherically symmetrical potential.
second term gives rise to small departures from spherical symmetry because
its value depends on the angles between the vector r and the nucleon spin

The

angular

momentum

vectors Sj

and S 2

the quantity

so that the average of the entire second term over

S x S 2 is

subtracted

all

directions of r

vanishes.!

conventional to call the second term the tensor potential.

It is

nomenclature
just like the

is

first

somewhat unfortunate
term.

as

it is

This

not a tensor but a scalar,

Neither term changes sign in the parity operation,

so both are scalars and

V(x, y, z)

V(x, y,

z).

spherically symmetrical term mixes a state of higher

The small nonvalue in with the

dominant S state to form the ground state of the deuteron. But, since
V(x, y, z) = V(x, y, z) and the ground state is not degenerate, the
eigenfunction for the system must have a definite parity (cf. footnote on
page 465), and the state which is mixed in must have the same parity as
the S state. As the parities of these states are even if / is even and odd if
/ is

The

odd

(cf.

state

equation 13-112), this state can be a


is

the only one of these which

is

state,

state, etc.

possible because for

all

the

1
Thus the
too large to lead to the spin quantum number i
ground state is assumed to be a mixture of S and D states. In order to
3
3
Calculations of probobtain i
1, these must be the S1 and
1 states.

others

/ is

and corresponding values of quadrupole moments, show


that the observed quadrupole moment is obtained with the mixture
ability densities,

96% 3 S1

4%W

(16-112)

where these percentages give the probability of finding the deuteron

in

either of the states.

confirmed by comparison with the observed magnetic moment


3
If the ground state of the deuteron were a pure S1 state, then

This
Hi

is

/"<,

(H-i)v

(f*i)n

(16-113)

t This term has the same directional dependence as the magnetic interaction potential
for two magnetic moments. However, it must not be thought of as due to magnetic
interactions between the magnetic moments associated with the spins, because the

strength of those interactions is several orders of magnitude smaller than what


required to explain the quadrupole moment.

is

NUCLEAR FORCES

Sec. 15]

679

components of intrinsic magnetic


from their spins. There
would be no orbital contribution from the motion of the charged proton
since / = 0, and the z components of the intrinsic magnetic moments
would add since the two nucleons are in the triplet state with their spins
"parallel."
Evaluating (fa) v and ((*() in equation (16-113) gives
H iz = +2.7896/^ - 1.9103^ = +0.8793^, which is 2.6 percent
greater than what is observed.! The general agreement confirms the
where

(/i t

moments of

and (fa) n are the

the proton and neutron that arise

Figure 16-52.

An

attractive square well potential.

conclusion that the ground state

is

almost a pure

S1

state

and the

dis-

crepancy confirms the conclusion that it does have a bit of some other
state mixed in. When the magnetic moment is calculated for the mixture
(16-113), agreement with the observed value

better than the accuracy

is

of the observation. The fact that the neutron-proton interaction potential


is

not completely spherically symmetrical

is

interesting but, to a first

can be ignored at low energies because the asymmetry


is small. In most of our discussion we shall ignore it and, therefore, take
the ground state of the deuteron to be a pure 3 51 state.
approximation,

it

Consider, then, the spherically symmetrical interaction potential V(r),

and assume that

it is

in the

as in equation (16-114)

form of a square well of radius


and figure (16-52).

-V

V{r)
0,

We

would

like to establish

it

will turn

<R

(16-114)

AE =

VQ

and the measured

As might be
we shall not be able to evaluate separately
single known quantity Ais, and we shall certainly
2.22 Mev.

out that

both R and V from the


not be able to evaluate the shape of the potential well from
t

The

signs give the relative orientations of the magnetic

momentum

vectors.

depth

r>R

a relation between R,

deuteron ground state binding energy


guessed,

R and

moment and

this single
spin angular

THE NUCLEUS

680

Therefore any shape

quantity.

we might

is

[Ch. 16

equally appropriate and, to simplify the

The time independent


Schroedinger equation that we must solve for this problem is (10-27),

calculation,

with

since

we

as well take the square well.

are considering

r dr\
where

/x is

the reduced

dr

mass

11

with

M the nucleon mass.

an S

state.

It is

which

is

(15-76) for

make

convenient to

^h [ -

(16-115)

u(r),

K(r(r)

Equation (16-116)

0.

the substitution

rR{r)

This leads to the differential equation for

dr

is,

MM _M
M + M~ 2

u{r)

two nucleons. That

for the system of

_
~

It is

we

the one dimensional square well equation

=
is

(16-116)
identical in

solved in Chapter

form with
8, and we

could adapt that solution to our present purposes by redefining the terms.

But

it is

just as easy to solve

^42 +

it

Write

again.

Kfu

it

<R

0,

as

(16-117)

dr

where
K\

j2/i(V

AE)/H

(16-117')

and

K\u

0,

dr 2

>R

(16-118)

where
K\\

The general

y/2/tAEIh

solution to equation (16-117)

(16-118')

is

r<R

= A sin K\r + B cos K\r,

(16-119)

Since the eigenfunction R(r) never diverges, equation (16-115) shows that

at r

0.

This requires setting


u

The general

B=

= A sin K\r,

solution to equation (16-118)

in (16-119).

= Ce~ K " r + De K " r

<R

Then
(16-120)

is

r>R

(16-121)

NUCLEAR FORCES

Sec. 15]

To

681

>

prevent this from diverging as r

we must

co,

D=

set

0.

Then

equation (16-121) becomes

Ce~ K " r

r>R

(16-122)

Now we must equate u and dujdr for both (16-120) and (16-122) at r =

R.

This gives

KR =

sin

Ce~ K " R

(16-123)

and

AK\

cos

K\R

= -CK

n e-

K" B

(16-124)

Dividing equation (16-124) by (16-123) gives

or

K, cot

~ A) cot

is

essentially (8-82),

method used

= -X

NWo ~ A)

This

K^R

and

it

(16-125)

VgMg

can be solved by the same graphical

The

solution only establishes a relation


between those values of V and R that will lead to the observed value of
AE. However, we shall soon see that there is evidence from scattering
experiments that can be used to evaluate R separately, so that V can be
AE. This being the case, it is a good
determined. It is found that V
for that equation.

>

approximation to neglect AZs in the term (V

AE) and

write equation

(16-125') as

AE
h

h
h

or
cot

The

first

N v

root (corresponding to the ground state)

is

J 2t*K R = Z
h

so

V R2 ~

R=

2
TT h*l8/j,

2
tt

/4M

(16-126)

x 10^ 13 cm, a value which is indicated by the scattering


experiments, we obtain V ~ 25 Mev and confirm the consistency of our
If

we

take

approximation.

From

this

information concerning the

potential that

fits

V R2

value for the interaction

the ground state of the deuteron,

it is

not

difficult to

THE NUCLEUS

682

[Ch. 16

The calculation predicts that


and this is verified by experiment.!
the assumed square well potential V(r) with its

calculate the locations of the excited states.

there are

no bound

excited states,

Figure (16-53) indicates

bound state, and also indicates the associated eigenfunction. There


no bound excited states of the neutron-proton system because the
interaction potential is only barely strong enough to bind the ground state.
For the same reason, the ground state eigenfunction has a long exponential
tail extending into the region r > R, and there is a probability higher than
50 percent that the separation between the two nucleons of a deuteron is
single

are

greater than the range of their interaction potential.

Figure

16-53.

The assumed

potential and the single

bound eigenfunction for the

deuteron.

<

We turn now from the bound state' (E 0) of the neutron-proton


system to the unbound states (E
0) of that system. They are investigated
experimentally by measuring the cross section for scattering neutrons

>

from protons, using techniques similar to those used for scattering nucleons
from nuclei, and are analyzed theoretically by the procedures of Chapter
15. As we saw in equations (15-73) and (15-74) of that chapter, the
Furthermore,
scattering cross section is specified by the phase shifts d
equation (15-78) showed us that the <5j will have appreciable values only
for values of / through / = /max where
.

max

~ KR

J2/J.E

(16-127)

-13
cm, and
2 x 10
Taking the radius of the interaction potential as R
10 Mev.
evaluating the total energy E which gives /max = 1, we find E
Thus, at energies small compared to this value, only d should have an

has an infinite
t Compare this with the H atom. This simplest of atomic systems
number of bound excited states whose properties provide almost all the information
that was needed in developing the theory of atomic systems. One of the reasons why
the development of the theory of nuclear systems
little

information because

it

is difficult is

has no bound excited states.

that the deuteron provides

NUCLEAR FORCES

Sec. 15]

683

appreciable value, and isotropic S-wave scattering should be observed.!


As isotropic angular distributions actually are observed up to energies of

about 5 Mev,

it is

certainly safe in treating scattering in this energy range

and (15-81) that apply to S-wave

to use the equations (15-80)

The

latter

scattering.

equation reads

= Lpo
K

(16 _ 128)

where

K=

(16-128')

yJljIEjh

assuming the square well potential (16-1 14), we

Still

may immediately take

over any useful results from the calculation for the identical potential
(15-82). However, we shall not just write down the expression (15-91)
for

(5

because

it

in terms of

gives d

terms of the known quantity A.E.


An approximate expression of

and we prefer an expression in

this type is easy to

value

is

if

we

realize

not going to

total energy

goes from the small negative

the small positive values

we

are considering here since, for

change very much as the

AE to

obtain

<R

that the form of the eigenfunction in the region r

<

V , and the form of the eigenfunction in the region


these values, ||
Thus we use
of the well is determined almost entirely by its depth V

all

equation (16-120) inside the potential well. That


u

~ A sin K\r,

is,

<R

(16-129)

Outside the potential well we use equation (15-88), which

= F sin (Kr +
u and du\dr at r = R gives
u

Matching

sin

Kr~
{

sin

),

(KR

is

r> R

(16-130)

(16-131)

<5)

and

AKi

cos K\r

Dividing (16-132) by (16-131),

X, cot

~ FK cos (KR +

(16-132)

we obtain

K R ~ K cot (KR +
{

But equation (16-125) shows that the


where

K =
n

<5

left side

c5

(16-133)

of (16-133) equals K^,

j2ftbEIH

(16-134)

Therefore

K cot (KR +
t It is

-K||

and that all energies are measured in that system.


E and the LAB energy is E = JElab-

(16-135)

CM system,
The relation between the CM energy

to be understood that in this section all statements refer to the

THE NUCLEUS

684

we

Since

<5

are going to use the S-wave scattering equations that are valid

KR <

only for

in (16-135)

1,

and write the equation

we may

little

work with trigonometric

(16-136)

drop

as well

KR in comparison with

as

K cot
A

[Ch. 16

~ -X

<5

(16-136)

identities will convince the reader that

identical with

is

Evaluating (16-128) from (16-137), and using (16-128') and (16-134), we


obtain an expression for the low energy neutron-proton scattering cross
section,
4t7

K*

4tt

Ki\

2fxElh

2ttH

2/iAEIh

E + A

(16-138)

E and the known deuteron binding energy


Mev. This expression was first obtained by Wigner in 1933.
It is in fairly good agreement with experiment for E from about 1 Mev up
to 5 Mev. However, the expression considerably underestimates the observed values of a when E is small compared to 1 Mev. For example,
(16-138) predicts that a approaches the constant value a
2nh 2 lfj, AE
24
2
2.4 x lfT cm = 2.4 bn, for E
O.j But the experimental value of
a for E
1 ev is a = 20.4 bn.
Although (16-138) was obtained from

in terms of the scattering energy

AE =

2.22

certain approximations, calculations in

which these approximations are not


can hardly be off by as much as a factor of 2. Thus the
factor of 10 discrepancy between the predicted and measured values of a
shows that something is definitely wrong with the theory.
used show that

Wigner suggested that the point

In 1935

wrong

is

it

at

which the theory might be

the assumption that the interaction potential used to

fit

the

deuteron binding energy can also be used to explain the scattering of a


free neutron by a proton. In the deuteron, the neutron and proton must
always interact in the 3 S 1 state because the spin quantum number is i = 1
free neutron and proton can interact in the 1 S state as well as
in the 3 St state. In fact, for an unpolarized beam of neutrons incident

whereas a

upon an unpolarized

target of free protons, J of the scatterings will take


state where the spins are "antiparallel," and f of
the scatterings will take place in Jhe triplet 3 S1 state where the spins are

place in the singlet


"parallel."

The

to 3 ratio occurs because there are

triplet states (cf. section 4,


t

The barn

10- 24

cm 2

(bn)

is

Chapter

12).

a unit of cross section

singlet state

and

If the target contains free protons,

commonly used

in nuclear physics

bn

NUCLEAR FORCES

Sec. 15]

685

all the others, and the observed


an average of the singlet scattering cross
section a s and the triplet scattering cross section a t weighted according
to the relative probabilities J and f That is,

each scattering

be incoherent with

will

scattering cross section a will be

The
The

= to + to

(16-139)

what was

triplet scattering cross section is

really evaluated in (16-138).

have the same general form as


have a different value if the neutron-

singlet scattering cross section will

(16-138), but the quantity AE will


proton interaction potential in the singlet state is different from what
in the triplet state. Allowing for this possibility, Wigner wrote

4(

fX

+ A

+
4(

s)

+ AE

it is

(16-140)
t ).

AE = 2.22 Mev. By equating this to the observed value a = 20.4


E ~ ev, AE may be determined. The result is AE ~ 0.06 Mev.
this is very different from A
it is apparent that the interaction

where

bn
As

at

potential

is

different in the singlet

and

triplet states.

corresponds to the spins "antiparallel" while the


the spins "parallel,"

we

This

is

corresponds to

see that the interaction potential of a neutron

and proton depends on the


nucleons.

Since the former

latter

relative orientation of the spins

of the two

described by saying that the interaction potential

is

spin

dependent.

The experimental value of a

because, for

at

E~

ev does not determine the sign of

<< |A|, the expression

~ 2rrh
ft

E+

|A|

can be shown to be as valid if A < 0, as (16-138) is if A > 0. Thephase


shift changes sign if AE changes sign, but the scattering cross section
(16-128) is independent of the sign of the phase shift. Thus the singlet
state could be either a stable bound state with binding energy AES > 0,
or it could be an unstable virtual state with binding energy AES <
(cf.
section 6, Chapter 15). The observation that there are no stable systems
containing two neutrons or two protons, for both of which the exclusion
principle

demands

that the lowest energy state be the singlet state

suggests that the singlet state of the system containing a neutron

proton

is

also unstable, or virtual.

The

and a

fact that this state actually is

proved in measurements of the scattering of low energy neutrons


from a target of molecular hydrogen at very low temperatures. The
virtual

is

THE NUCLEUS

686

[Ch. 16

hydrogen molecule can either have the two protons "parallel" (orthohydrogen) or "anti-parallel" (para-hydrogen). The latter type has the
lowest energy state, and so at very low temperatures essentially all the
molecules will eventually be converted to that type by collisions and other
processes. For neutrons of de Broglie wavelength comparable to the
internuclear spacing of the molecule, the total wave scattered from the
molecule will be a coherent combination of the waves scattered from

20.0

10.0

5.0

2.0-

1.0

0.5
,-5

10

10"

10"

10

_/!

10"

10"

10

CM (Mev)
Figure 16-54.

Experimental cross sections and theoretical

neutrons from free protons

in

fits

for the scattering of

the S-wave energy range.

two protons. Since in one of these scattering processes the neutron


and proton are in the singlet state while in the other they are in the triplet
the

state, the total scattering cross section is sensitive to the relative signs

the singlet and triplet phase shifts.

An

analysis

by

of

Teller (1937) of the

measurements, which were made at his suggestion, shows that these


phase shifts are of the opposite sign and, therefore, that the singlet state
of the neutron-proton system is virtual.
The points of figure (16-54) shows some measured values of the cross
section for scattering neutrons from free protons in the S-wave energy

To within their accuracy, these data can be fitted perfectly by an


equation basically the same as (16-140), but including corrections for the
approximations we made in its derivation. This is indicated by the solid
range.

These corrections make the exact equation more dependent on the


characteristics of the neutron-proton interaction potential than (16-140)
and thus allow the radius and depth of both the triplet and singlet interaction
line.

NUCLEAR FORCES

Sec. 15]

potentials to be determined

687

by

fitting

the data.f

(It is

also necessary to

bring in the data obtained in the scattering of neutrons from low temperature molecular hydrogen.)

The values so determined

R
VQ
Rs
Vn

'

=
=
=
=

2.05

x 10"13 cm

33.9

Mev

2.9

11.7

Note that the interaction potential


the singlet potential

are

(16-141)

x 10- 13 cm

Mev
certainly is spin dependent,

and that

considerably weaker than the triplet potential for

is

the neutron-proton interaction.

Let us turn

now

particles with s

to the proton-proton interaction.


\,

For these

identical

the exclusion principle plays an important part in

ruling out certain states because

it

demands that the complete eigenfunction

for the system of two particles be antisymmetric with respect to an exchange

of their labels. Thus the spin part of the eigenfunction must be antisymmetric if the spatial part is symmetric, and vice versa (cf. section 4,

Chapter

12).

The symmetry of

realized that, for this system of

the

same

is

is

easily evaluated if

it is

exchange of their labels

particles,

CM of the two particles.

at the

the eigenfunction
is

two

is

as changing the signs of the coordinates because the origin of

coordinates
parity

the spatial part

is

symmetric

odd, and the parity

if its

is

parity

even

if

is

is

Therefore the spatial part of

even and antisymmetric


even and odd

if

/ is

if its

odd

(cf.

states correspond to symmetric


equation 13-112). Thus the S, D, G,
antisymmetric
triplet spin parts. The opposite
require
the
spatial parts and
possible
states.
The
states in which two protons
true
for
the
P,F,
H,
is
.

states only.
can interact are, then, the 1 5', a P, 1 D S F, 1 G, 3 H,
In scattering experiments at energies somewhat less than 10 Mev,
KR is small enough that the two protons interact only in the 1 S state,
which is the 1 S state. The exclusion of the 35x state simplifies the analysis
of the proton-proton scattering experiments, and this certainly helps
because the analysis is complicated by the existence of a Coulomb as well
as a nuclear interaction between the two protons. Because of the Coulomb
interaction, the overall angular distribution is not isotropic in the
.

CM

system, even though

S-wave scattering as far as the nuclear interaction


is concerned. Figure (16-55) shows the differential scattering cross sections
for proton-proton scattering measured at several different low energies,
and also shows theoretical fits to these measurements. Consideration of
figure (15-2) shows that do/dCl must be symmetric about 90 because, if
it is

t For example, the radius R appears separately in the term which was dropped between
equations (16^135) and (16-136). Cf. exercise 20 at the end of this chapter.

THE NUCLEUS

688

one proton

[Ch. 16

must be scattered at
and it is sometimes used as an
experimental check. At angles smaller than about 30, dajdD. rises very
rapidly. In this region it is following the rapid angular dependence of
Coulomb interaction scattering (cf. equation 15-94) which dominates
the nuclear interaction scattering. Near 90, the S-wave nuclear interaction
180

6.

is

scattered at the angle

This

is

6,

the other one

verified experimentally,

0.18N\eM

0.17

0-16

0.15

0.14

0.13

0.12

EcM =

2.101 Mev

0.11

0.10

20

30

50

40

60

70

80

90

"cm

Experimental differential cross sections and theoretical fits for protonin the S-wave energy range.
From G. Breit and R. L. Gluckstern,
Annual Review of Nuclear Science, Vol. 2, Annual Reviews, Stanford, 1953.
Figure 16-55.

proton scattering

scattering dominates

and dajdQ.

is

region in which the nuclear and


tively.

almost isotropic. In between, there

Coulomb

is

interactions interfere destruc-

This interference allows a determination of the sign as well as the

magnitude of the nuclear interaction phase shift. It is found that, for


energies up to about 5 Mev, the phase shift can be fit with a square well
proton-proton interaction potential with radius and depth exactly the same
as R s and V of (16-141), the radius and depth that are found for the
1
S state neutron-proton interaction potential. Thus at low energies the
neutron-proton and proton-proton interaction potentials appear to be the
same in the 1 S state, which is the only state in which a comparison can be
made. Also recall that mirror nuclei indicate that the neutron-neutron

NUCLEAR FORCES

Sec. 15]

and proton-proton interaction

689

same

potentials are the

in the

low energy

range characterizing the relative energies of two nucleons in a nucleus.

From all of this it is concluded that at low energies the interaction potentials
between

all

pairs of nucleons are the same.

however, that

this

It

mean

does not necessarily

should be remembered,

that the shapes of the inter-

action potentials are the same, but only that the phase shifts they produce
are the same.

None
15

of the low energy measurements can give detailed

E CM = 45

Mev

L
/

10

7^

-X

"^x

*-jf

x^^^

.X""

<

b
C

20

40

60

80 100 120 140 160 180

Figure 16-56.
Experimental differential cross sections and theoretical fits for neutronproton scattering at a high energy. From R. Christian and E. W. Hart, Phys. Rev., 77>

441 (1950).

information about the shapes of the interaction potentials since, when


KR 1, the shapes are not important in determining the scattering cross

<

sections.

more about the interaction potentials it is necessary


measure and analyze nucleon-nucleon scattering at high energies,
where KR > 1. According to statements made in the previous chapter,
it would be expected that such measurements should be able to give
information about the shapes of the interaction potentials. Since, for
KR > 1, partial waves with several / values will be involved, the measurements should also be able to tell whether or not the interaction potentials
have an explicit dependence on /.
Figure (16-56) shows the neutron-proton differential scattering cross
section measured at E = 45 Mev, and a theoretical fit to these data.
At other energies from about 15 to 150 Mev, dajdQ. has much the same
In order to learn

to

THE NUCLEUS

690

[Ch. 16

appearance except that

its magnitude has an approximately l/E behavior.


form of this da/dQ indicates that there is an / dependence in the neutron-proton interaction potential, as can be seen from the
following argument. If there were no / dependence the potential could be
written as V(r), and its high energy differential scattering cross section
could be evaluated, at least qualitatively, by the Born approximation.
But, in section 4, Chapter 15, we saw that for KR^> 1 it seems to be a
general characteristic of all potentials V{r) that dajdQ is a highly asymmetric

Even the

qualitative

function of

6,

much

with

larger values for

0~O

than for d

180.

symmetry about 90 that the dajdQ of


figure (16-56) evidently possesses, and it shows that the interaction
potential cannot be written as V(r). Serber (1948) has proposed an /
dependent interaction potential which is capable of fitting the experimental
data for the neutron-proton interaction, at least up to energies of about
This Serber potential can be written
1 50 Mev.
This

is

in strong contrast to the

F (r) =

L"

+C-

1 )']

F(r)

Since

V (r)
t

vanishes for

all

odd

/,

it

(16-142)

gives zero phase shift for all these

which has symmetry about 90,


because only even / terms will remain in the general phase shift expression
(15-73) for dajdQ, and because all the Legendre polynomials P^cosd)
for even / have symmetry about 90 (cf. 15-53). Note that the Serber
potential helps the nuclear forces lead to saturation by removing the

values.

This, in turn, gives a da/dD,

nucleon-nucleon interaction potential in half of the possible quantum


states. However, it does not do enough. For nuclear forces actually to
lead to saturation,

it

is

necessary that there actually be repulsions in

certain circumstances.

The study of high energy proton-proton


the repulsions arise.

scattering indicates

where

Figure (16-57) shows dajdQ. for proton-proton

measured at E = 170 Mev. In the energy range from about 50


Mev, dajdQ. has almost the same form a long flat region where
10~ 3 bn per unit solid angle, and an abrupt rise at small
dajdQ.
angles to follow the Coulomb interaction dajdQ that dominates at these
angles. When these measurements were first performed, it was thought
that, because dajdQ is isotropic in the angular region where nuclear
interaction scattering dominates, this was somehow a case of pure S-wave
scattering even though KR > 1. However, the maximum possible value
of dajdQ for S-wave scattering of two protons isf
scattering

to 200

~4x

dajdQ

(2//Q

2
1. The factor of 2
t Except for the factor of 2, this is equation (15-80) for sin <5
because there are two identical particles in proton-proton scattering.

arises

NUCLEAR FORCES

Sec. 15]

At

E=

170

Mev

this gives the

unit solid angle, which


that several partial

together in such a

691

is less

upper

limit da/dD.

than what

is

x 10"3 bn per
Thus it must be
and happen to mix

waves contribute to the scattering


as to yield an isotropic da/dD,.

way

E CM =

Figure 16-57.

~ 2.5

observed.

30

The measured

60

170 Mev

90

differential cross section for

proton-proton scattering at

a high energy.

Jastrow (1951) has shown that this implies the existence of a strong
repulsive core in the interaction potential,

form of the type indicated


led to this conclusion

is

Figure 16-58.

i.e.,

in figure (16-58).

that the potential has a

The

detailed analysis

which

complicated, but the essential points can be seen

potential with a strong repulsive core.

from a simple argument. Consider proton-proton scattering at E = 170


Mev, and assume that, as for neutron-proton scattering, the interaction is
described by the Serber potential. Since
at this energy, only
S-, P-, and D-wave phase shifts can be appreciable.
But the P-wave

KR~2

phase

shift is

zero for the Serber potential, so the differential scattering

THE NUCLEUS

692
cross section can be obtained

zero except S and

=\
X

=
K

dil

As

2 .|

{e'
{e

=
K
dO.

<5

iS

id

(sin
l

sin d

sin d

5e

5e~
id

25 sin

10 sin S sin

{sin
l

from (15-73) by

(cos 8)

id

sin

(5

sin

^(cos

P2 (cos

P2 (cos

(5

phase

setting all

shifts

this gives

1,

[Ch. 16

0)}

0)}

0)

d^ "'^ + e-^-'^PJcos
5

25 sin

2
t5

2 Pl(cos

0)}

0)

(16-143)
P (cos 0)}
+ 20 sin sin cos
Now P (cos 0) = (3 cos - l)/2, so P2(cos 0) = P (cos 180) = and
P (cos 90) = |. Thus the second term of (16-143) is larger at 0 and
<5

<3

<5

(<5

2)

1,

180 than at 90 and, if both 5 and <5 2 are positive, the third term will
be positive at 0 and 180 and negative at 90. This will make da/dQ. much

smaller at 90 than at 0 and 180,


experimentally.

opposite signs.
to

the low energy data, both d

fit

Chapter
<5

will

15).

and

certainly not isotropic as observed

To achieve isotropy it is necessary for d and 6 2 to have


Now, for a purely attractive potential of the strength used
and

<5

will

be positive

(cf.

section

5,

But, for a potential of the form indicated in figure (16-58),

become negative

at energies high

compared

to the depth of the

attractive region since the effect of the attractive region

becomes

negligible

and only the repulsive core, which produces a negative phase shift, is
important. The repulsive core is, however, not important in determining
l
the phase shift <5 2 until the energy becomes extremely high because the r
dependence of the eigenfunctions, for

r - 0,

makes

the

2 partial wave

smaller in the region of the repulsive core than in the attractive region
1 70 Mev, and similar energies, d 2 will be positive. By fitting the high
at E

energy proton-proton scattering data, it is found that the radius of the


-13
cm. The attractive region can then be
repulsive core is about 0.5 x 10
adjusted to give agreement with the low energy data. If its strength is
made somewhat greater than that of the simple attractive square well

by the parameters (16-141), the correct positive value


can be achieved at low energies because the positive phase shift
produced by the attractive region becomes larger than the negative phase
shift produced by the repulsive core at these energies.
potential described

for S

As

is

1
states are possible. Thereproton-proton scattering, only the 1 S and
and one phase shift for 1 = 2.
only one phase shift for / =

this is

fore there

NUCLEAR FORCES

Sec. 15]

693

Recent work of a number of investigators shows that by using a repulsive


core in a spin dependent potential, which
in the sense that

is

similar to the Serber potential

weak for odd

/, an interaction potential can be found


which gives a good fit to both neutron-proton and proton-proton data at
all energies up to more than 100 Mev, providing the tensor potential and a
spin-orbit potential are included. The spin-orbit potential is proportional
to S L and is attractive if this quantity is positive and repulsive if it is
negative.f One reason for its inclusion is to help fit experiments which
find that the spin angular momentum vectors of two nucleons are partially
polarized by the process of scattering from each other. We avoid quoting

it is

the parameters for the very complicated interaction potential that fits
all the data, because these parameters have not been determined uniquely.

In fact,

felt by many that it will not be possible to find a unique


and that it is better to express what is known about nuclear

it is

potential,

forces in terms of sets of phase shifts describing the nucleon-nucleon


interactions in various quantum states at various energies. There are

even ambiguities concerning the values of these phase shifts.


However, there is, at present, little doubt concerning the existence of
repulsive cores (at least in the singlet states,

which are the ones most


important in the proton-proton interaction). For one thing, repulsive
cores obviously produce nuclear forces that lead to saturation, although it

was

would not lead to saturation at the right


~0.5 x 10-13 cm of the repulsive cores is conmean center-to-center spacing ~ 1.2 x 10~ 13 cm

originally thought that they

density because the radius


siderably smaller than the

of the nucleons in a saturated nucleus. Recent calculations indicate,


though, that repulsive cores should actually produce nuclear forces that
lead to saturation at densities and binding energies close to those observed
in nuclei.

These calculations are based on a theory of nuclear structure due to


Brueckner (ca. 1953) and others. The theory starts from the nucleonnucleon interaction potential described in the next to last paragraph,

and then goes through a

self-consistent procedure similar to that of the


Hartree theory of the atom. However, for the nucleus the procedure is
very much more difficult than for the atom, because there is no single

dominant interaction analogous to the electron-nucleus interaction of the


atom, and because the nuclear forces which must be treated in the theory
are so complicated. Although the theory is still in its infancy, some results
have been obtained, and they are quite encouraging. In addition to finding
nuclear densities and binding energies which are close to what is observed,
t

Note that the

probably related.

two nucleons is similar to the spinand a nucleus (cf. section 8). The two effects are

spin-orbit interaction between

orbit interaction between a nucleon

THE NUCLEUS

694

[Ch. 16

the real and imaginary parts of


good values are found for V and
Furthermore,
the theory gives some feeling
the optical model potential.
instance, the self-consistent
nucleus.
For
actually
happening
in
a
for what is
procedure indicates that the nucleons interact primarily two at a time, and
not three or more at a time. This is the justification for assuming that the
nucleon-nucleon potential, obtained from analysis of experiments involving
the interaction of two free nucleons, applies to the interactions between
fairly

the nucleons of a nucleus.

It is

action potential which are the


structure are:

the

(a),

also found that the features of the inter-

most important

there are repulsive cores;

in determining nuclear

(b), the attractive

regions of

S state interaction potentials are just barely strong enough to overcome

the repulsive cores and give a net binding for the neutron-proton system;

and

Of

(c),

the entire

state interaction potentials vanish or are very

course, the exclusion principle

is

also extremely important.

weak.t
Despite

the uncertainties in the present state of knowledge concerning nuclear

been made in the theory of nuclear structure because


which are apparently the most important
are fortunately those which are the least uncertain.
forces, progress has

the features of nuclear forces

16.

Mesons

In the preceding section

we presented

a description of certain features

of nuclear forces that are observed in experiment. Although theory was


used in the description, it was used only to correlate the experimental
observations, and not to explain them. There is, however, a theory that
attempts to explain how the properties of nuclear forces are based on

more fundamental

attributes of nature.

originated with the

work of Yukawa

This

is

the

meson

theory,

which

in 1935.

Let us imagine two nucleons in interaction.

Yukawa proposed

that

there is associated with this system a "field" which describes the interaction,

and that the actual mechanism of the interaction involves the exchange
of a "quantum" of the field between the two nucleons (i.e., in interacting,
one nucleon emits a "quantum" and the other subsequently absorbs it).
This is pictured schematically in figure (16-59). In such an interaction
the momentum carried by the "quantum" is transferred from one nucleon
to the other, and this is the origin of the force acting between the
nucleons.
In

making

of the

this proposal,

Coulomb

Yukawa was

guided by analogy with the case


two charged particles.

interaction between a system of

t States with / values higher than those of the P state are not very important at the
energies characterizing the motion of nucleons in nuclei.

MESONS

Sec. 16]

695

According to the well-developed and highly successful theory of quantum


electrodynamics (cf. footnote on page 462), associated with this system
there is an electromagnetic field which describes the interaction, and the
interaction actually arises from the exchange of a quantum of that field
between the two charged particles. The quanta of the electromagnetic
field are the zero rest mass entities which throughout this book we have
called simply quanta.

Quantum
interaction

electrodynamics shows that the long range of the

is

in the interaction,

Figure 16-59.

that quanta

Coulomb

a result of the zero rest mass of the quanta that are exchanged

had

and that the range would not be long

if it

had happened

schematic representation of a meson exchange process.

finite rest

Yukawa adapted the theory to the case of

mass.

a system of nucleons, interacting with a short range nuclear force, by

assuming that the "quanta" of the "field" describing the interaction have
mass. These "quanta" are called mesons, and the "field" is
called the meson field. We shall not attempt to go beyond the first steps of a
quantitative development of the meson theory because that would rapidly
carry us into some difficult mathematics. However, there are certain
features of the theory which are not at all difficult to describe.

finite rest

For

instance,

it is

easy to estimate the

lead to the observed range

meson

rest

mass m^ required to

Consider the process


represented in figure (16-59). In this process one nucleon emits a meson
of rest mass w, the meson travels a distance of the order of R, and finally
the meson is absorbed by the other nucleon. While the meson is in flight,
of nuclear forces.

the conservation of total energy

is violated because the total energy of the


system equals two nucleon rest mass energies in the initial and final
state, and equals two nucleon rest mass energies plus at least one meson

mass energy in the intermediate state. But the uncertainty principle


shows that violation of energy conservation by an amount
rest

AE ~
is

not impossible

if it

mc*"

does not happen for a time longer than At, where

At A

~h

THE NUCLEUS

696

[Ch. 16

because such a violation would be immeasurable. The velocity of the


meson can be no greater than c, so its flight time At is at least

At

~ Rjc

These three equations give

m^c 2

hc/R

/j/At

m,

~ hjRc

'

or

Setting

R~2 X

10~ 13 cm, this predicts

tron rest mass. It

is

worth while to

Figure 16-60.

way.

meson of

rest

mass

)T

(16-144)

m^

~ 200m, where m

restate the

The Yukawa

argument

is

the elec-

in the following

potential.

~ hjRc leads to an interaction of range

by a distance much larger than


R, there could be no interaction as its flight time would be so long that
the uncertainty principle would allow an accurate enough measurement
r~"R because,

if

the nucleons are separated

of total energy to detect the violation of energy conservation.


A detailed treatment, using relativistic quantum mechanics, leads to
a detailed prediction of the interaction potential arising from

exchange. This

is

the

meson

Yukawa potential,
V(r)

-rlR

(16-145)

where

R = hjm w c
which

is

plotted in figure (16-60).

leading to this potential.

not

It is

relativistic

(16-145')
difficult to

wave equation

obtained by writing (1-25),

2 =

y+m

?c

where

Vl

p!

follow the steps

for the

mesons

is

Sec. 16]

MESONS

making the

substitutions (7-77),

697

Px^ in
dx

Py~^

d
*
in

Pz~+

d
in

dy
t,

dz

->

in

d
dt

and then multiplying the

resulting operator equation into the equality

y>

rp

This gives

-**?? = -c 2 n

v* w

or

+ my v

or

V-^ =
c

which

is

m\ c
-$r<P

dPtp __

J_

06-146)

~dr~~~li

called the Klein-Gordon equation.

For m^

it

reduces to the

wave equation
T72

V*W
^

= 12 <P"fc

for the quanta of the electromagnetic


the

This has a

field.

static solution

of

form

= e /r,
2

f
as can easily be verified

by

ip(r).

For m^

>

substitution, using the relation

for

3f

dr \

dr'

the Klein-Gordon equation has a static solution

of the form
y>

= - q- e~ rlB

>

where
-R

h/m n c

as can also easily be verified by substitution.

wave equation

for zero rest

Since the solution to the

mass quanta gives the Coulomb interaction

THE NUCLEUS

698

[Ch. 16

potential for the electromagnetic field, the solution for non-zero rest

assumed to be the interaction potential for the meson

quanta

is

that

the

is,

Yukawa

field,

potential (16-145).

The constant q2 of (16-145) determines the


Although

mass

this potential is

strength of the potential.

obviously not capable of explaining the existence

of the short range repulsive core in the nucleon-nucleon interaction,

Or~-

Intermediate

Initial

Figure 16-61.

it

Final

The exchange

of a

n~ meson.

can explain the attractive region outside the core. Fitting it to this region
2
In terms of the dimensionless constant q 2 jhc, it is
fixes the value of q
.

\hc~

q
In the original form of the

mesons carry the

electric

meson

charge

(16-147)

15

theory,

or +e.

it was assumed
The theory then

that all
predicts

that a typical interaction process between a proton and a neutron

lead to the situation pictured schematically in figure (16-61).

would
In this

p represents a proton, and -n~ represents a


meson of charge e. We see that, as a result of the interaction, the
neutron is changed into a proton and the proton is changed into a neutron.
figure n represents a neutron,
7T~

_3i~o

^0
Final

Initial

Figure 16-62.

small angle scattering.

This might seem like an unreasonable prediction, but there is very direct
experimental evidence that it really can happen. Consider the high

energy neutron-proton differential scattering cross section curve of figure


(16-56). We have interpreted its symmetry about 90 as evidence for the
Serber

potential

(equation

16-142)

would lead primarily


(16-62). However, we could

because

an ordinary interaction

potential V(r)

to scattering at 8

figure

also interpret the

as evidence for the conclusion that there

is

0, as indicated in

symmetry about 90

a probability of \ that in the

MESONS

Sec. 16]

699

scattering the neutron is changed into a proton, and vice versa, by the
exchange of a tt~~ meson emitted from the neutron, or by the exchange of a
7T+ meson of charge +e emitted from the proton. Figure (16-63) indicates
such a scattering for a case in which a n~ meson is exchanged. An inter-

~<Z)

>
Intermediate

Initial

A small

Figure 16-63.

action potential, that

Final

angle scattering with the exchange of a it- meson.

would lead to

scatterings in

which there was a

probability of \ that such processes occur, can be written formally as

[1+*]
where

V(r)

(16-148)

an exchange operator that exchanges the labels of the two


This potential always comes into formulas for the cross section

is

particles.

through the matrix element


-

y>*

so

it is

V(r)ip

of interest to consider the quantity

[1

f]
V(r)y> t

iV(r)V

+ iVWVi

(16-149)

where we have specified the eigenfunction by its quantum number /. In the


paragraph following (16-141), we found that exchanging the labels of the
particles is the same as changing the signs of the coordinates. In this
operation,

yi

odd, because

will
its

be unchanged if / is even, and will change sign if / is


is even in the first case and odd in the second.

parity

Therefore
Py> t

ip t ,

p Vi = Yi>

even

lodd

or

FVi=(-l)'v.

THE NUCLEUS

700

[Ch. 16

and equation (16-149) can be written

U v(r)fi =

mr)xpi

i F(r)(

_ 1)V; =

[1

+^- 1 >'J V (r) Wl

We see, then, that (16-148) is completely equivalent to the Serber potential


(16-142),

and

that the

meson theory provides

a qualitative explanation of

the origin of that potential.

The high energy neutron-proton

differential scattering cross section

shows, however, that the original assumption that

cannot be correct.

If

it

(16-63) would take place with a probability of

all

mesons are charged

were, processes of the type indicated in figure

Initial

1,

instead of with the

Intermediate

Figure 16-64.

Final

The exchange

of a

w meson.

probability of \ that

is required to fit the data. Thus there must also be


uncharged mesons which allow interactions of the type pictured schematically in figure (16-64), where 77
represents an uncharged n meson.
Furthermore, the v mesons are necessary in order to allow nuclear
interactions between two nucleons of the same charge.f

Meson theory also provides a qualitative explanation of how the


neutron can have an intrinsic magnetic moment even though its net
charge is zero. According to the theory, a neutron can spontaneously emit
_
a 7T meson and become a proton, if the proton reabsorbs the meson within
a time At
h/m^c 2 and becomes a neutron again. As a result of such

processes, there are internal currents associated with the neutron,

and
moment. The theory
predicts that protons also spontaneously emit and reabsorb mesons. Thus
meson theory pictures nucleons as being surrounded by a "cloud" of
mesons, which have a transient existence and are bound to the nucleon
"core" by the requirements of energy conservation.
At the time of Yukawa's proposal, there were no known particles of
rest mass between the electron rest mass m and the proton rest mass
1836am. The n + and it~ mesons were first observed by Powell and collaborators in 1947 as a component of the cosmic radiation. Shortly after,
these currents produce the intrinsic neutron magnetic

t At the time of Yukawa's original proposal


only between neutrons and protons.

it

was thought that nuclear

forces acted

MESONS

Sec. 16]

701

mesons were produced artificially at a large particle accelerator in


between high energy (~300 Mev) nucleons and nucleons in a
target. Presumably the cosmic radiation mesons are also initially produced
in high energy collisions. Measurements show that the tt+ and ir mesons

these

collisions

have the same

rest

mass,

m,
This

is

274m

(16-150)

The

enough to Yukawa's prediction (16-144).

certainly close

mesons were first observed by Moyer and collaborators in 1950 as


products of high energy collisions. Their rest mass is found to be

77

m^o

264m

(16-151)

free n mesons, which are observed in these


from the clouds surrounding the colliding

According to the theory, the

experiments, are liberated


nucleons by the energy made available in the collision.
that these are the

same

as the

mesons discussed

It is

in the

assumed, then,

meson

theory.

Of

course, their discovery constituted a striking verification of the qualitative


validity of the theory.

Unfortunately, there has been

little

opportunity to

test the quantitative

validity of the theory. The problem is that the usual calculational techniques cannot be applied to meson theory because the dimensionless
constant (16-147) is so large. This constant, q*\hc <= 15, plays a part in

meson theory which


constant, a

e2 jhc

is

analogous to that played by the fine structure

1/137, in

quantum

electrodynamics.

It is

possible

very accurate calculations in the latter theory by the use of perturbation techniques, because it turns out that the perturbations associated
with electromagnetic interactions are always proportional to powers of
to

make

etjhc

and are therefore very small.t Perturbation techniques cannot be

used in meson theory because the perturbations are proportional to powers


of q 2 lhc and are therefore very large. Quantitative calculations in meson
theory await the development of more powerful techniques.
Mesons are unstable. The n+ and it mesons decay spontaneously in
free space according to the

scheme
77*

-^/X*

+V

(16-152)

ith lifetime

tt =2.54

x 10" 8

sec

(16-153)

The symbol /j,+ represents a /*+ meson of charge +e; (x represents a


pr meson of charge e; and v represents a neutrino. The a mesons
perturbation (11-51)
t As a particular example, note that the spin-orbit
(e^/ic) 2 .
tional to a 2

is

propor-

702

THE NUCLEUS

[Ch. 16

decay according to the scheme

sy +
N

y>

*%

+e-+y,

(16 _ 154)

\,

with lifetime

V>~l0 _15 sec

(16-155)

represents a gamma ray; e+ represents a positron; and


er represents an electron. The fi+ and fjr mesons also decay. Their decay

The symbol y
scheme

is

M ->e

+v +

and

(16-156)

their lifetime is

x 10" 6

2.22

sec

(16-157)

There are no uncharged


mesons. All values quoted for lifetimes are
values in frames of reference in which the mesons are at rest. In a frame
in which a meson is moving, its lifetime appears to be lengthened because
fj,

of the relativistic time dilation effect described by equation (1-7). Measurements of the observed values of t and T , made in both rest frames
and moving frames, give direct experimental verification of time dilation.
The fi mesons have no part in Yukawa's meson theory of nuclear forces,
although this was not appreciated until some time after their discovery in
1936 by Anderson and Neddermeyer. These investigators found the
particles as components of the cosmic radiation and showed that their
rest mass is intermediate between the rest mass m of an electron and the
rest mass 1836m of a proton. We now know that they are probably
produced in the cosmic radiation from the decay of it mesons. But, in
1936, it mesons had not been discovered, and it was naturally assumed
that the fi mesons were Yukawa's mesons. An ever increasing accumulation of evidence showed, however, that the interaction of
mesons with
matter is very weak. For instance, the fi mesons of the cosmic radiation
can penetrate great thicknesses of solid matter with little attenuation since
they can be detected in deep mines. This being the case,
mesons can
hardly be the particles responsible for the strong interactions of nuclear
forces, despite the fact that their rest mass
ft

/j,

/j,

m
is

quite close to the value predicted

207m

(16-158)

by Yukawa.

This situation was the source of considerable confusion in the ten years
before the discovery of the tt mesons. But, after their discovery, it was

immediately assumed that the

tt

mesons are Yukawa's mesons

early evidence indicated that their interaction with matter

was confirmed by

later

is

since the

strong. This

measurements which showed that the cross section

Ch.

BIBLIOGRAPHY

16]

for interaction of a

it

703

meson with a nucleus

close to

is

its

maximum

possible value, the projected geometrical cross sectional area itR*

equation 15-103). The interaction

meson

interacts with a nucleus,

splitting the nucleus into several

able energy.

It is

now known

is

also particularly violent;

most of

that

mass energy goes into

rest

its

fragments which

n mesons

(cf.

when a n

fly

apart with consider-

are closely associated with

nucleons and interact with the strong nuclear interaction, while fi mesons
are closely associated with electrons and interact with the very weak beta
decay interaction. One of the pieces of evidence for the association between

mesons and electrons is that they both have spin quantum number
i, antisymmetric symmetry character, and magnetic moments corresponding to the spin g factor g s = 2. The n mesons have i = 0, symmetric symmetry character, and no magnetic moments.
Within the past few years, more than a dozen other unstable particles
have been discovered. These particles fall into two classes the K mesons
with rest masses around 1000m, and the hyperons with rest masses around
fi
i

2500/w.

At

present, not too

much

is

known about

However, some of them seem to have strong


contribute to nuclear forces.

interactions,

their properties.

and so probably

In fact, the relation (16-144) makes

it

tempting to associate those that interact strongly with the short range
repulsive core of the nucleon-nucleon potential, because the relation shows
that the range of the force is inversely proportional to the

mass of the meson.


probably also true that part of the effects
leading to the core arise from the simultaneous exchange of several tt
mesons. We shall not even attempt to list the
mesons and hyperons that
are presently known, since such a list would undoubtedly soon be obselete.
For up to date information concerning this rapidly developing field, and
the important part which it surely will play in the fundamentals of modern
This

may

be true, but

it is

physics, the reader

must consult current research

literature

and review

articles.

BIBLIOGRAPHY
Bethe, H. A., and P. Morrison, Elementary Nuclear Theory, John Wiley and
Sons, New York, 1956.

M., and V. F. Weisskopf, Theoretical Nuclear Physics, John Wiley and


York, 1952.
Elton, L. R. B., Introductory Nuclear Theory, Isaac Pitman and Sons, London,
Blatt, J.

Sons,

New

1959.

Evans, R. D., The Atomic Nucleus, McGraw-Hill Book Co., New York, 1955.
Halliday, D., Introductory Nuclear Physics, John Wiley and Sons, New York,
1955.

THE NUCLEUS

704

[Ch. 16

EXERCISES
1.

Show

that the

minimum

from a thin hydrogenous


2. Evaluate, in

Coulomb

energy quantum that will eject a 4.5-Mev proton


is E c 50 Mev, as stated in section 2.
y

layer

Mev, the energy of gravitational attraction, and the energy of


two protons separated by a distance of 5 x 10- 13 cm.

repulsion, for

3. Prove equation (16-3). Hint: Use Gauss' law, which


Chapter 13.

4. Verify

is

stated in section 2,

equations (16-11), (16-11'), and (16-11"). Also show that (16-11')

reduces to (16-12).
5.

Use the appropriate

classical expressions to derive (16-20).

Use (16-20) to calculate the energy in the LAB frame of the protons emitted
in the forward direction, and at 30 from the forward direction, in the reaction
2
He 4 + 7 N 14 8 17 +
induced by 30-Mev alpha particles. For this reaction
Q = -1.18 Mev.
6.

Using the data of table (16-1), evaluate, in Mev, the energy released in
two
nuclei to form a 2 He 4 nucleus. Also evaluate, in Mev,
the height of the Coulomb repulsion barrier which must be overcome before there
is a large probability that the two nuclei can get close enough together for fusion
to take place, assuming that the radius of 1 H 2 is 2 x 10-13 cm.
8. Evaluate the terms of the semi-empirical mass formula for the 92 U M8
7.

the fusion of

Use the results to calculate the predicted values of total binding energy
and average binding energy per nucleon AE/A, and compare with table
(16-1). Calculate the fraction which each of the terms /i through
5 contributes
to AE.
nucleus.

AE

9. Look up the ground state spin quantum numbers of all the odd A
and compare with the predictions of the shell model and figure (16-23).

nuclei,

Assuming that 71 Lu173 is an ellipsoid of revolution, find the lengths of


the semiminor and semimajor axes that lead to the value of quadrupole moment
given in figure (16-24), and also to a volume consistent with normal nuclear
10.

densities.
1

is

1.

Show

V, 1 H

1
,

H2

or

He3

but that

it is

9 2rj 238

unstable to the

2
He 4 Explain why this is so. Hint: Take the masses of the emitted
from table (16-1) because the semi-empirical mass formula is not

emission of
particles

mass formula correctly predicts that

that the semi-empirical

stable to the emission of


.

accurate for very small A.

An

unstable nucleus whose probability per second of decay is A is being


artificial means in a particle accelerator at a constant rate of /
nuclei per second. If the production process commences at / = 0, calculate
the number of these nuclei which will be present at f = f by setting up and
12.

produced by

solving an equation similar to one of the equations (16-46).


13.

Use the semi-empirical mass formula to estimate how the data of

(16-25) would extrapolate to

A =

190, 180, 170, 160.

figure

Ch.

705

EXERCISES

16]

Use the semi-empirical mass formula

14.

similar to (16-36) for

A =

77.

to evaluate points

on a

figure

From it predict the energies available for electron

and positron decay. Then look up the data and compare.


15. Use the shell model to predict the spins and parities of the ground states
of several sets of odd A nuclei in the region N = 50 to N = 82. Then predict
which of the beta transitions between these ground states will satisfy the Fermi
selection rules (16-74), and which will satisfy the Gamow-Teller selection rules
(16-77).

Evaluate the attenuation length A for neutrinos passing through matter


2
gm/cm 3 using the typical interaction cross section a = 10" cm

16.

of density 10

predicted by (16-80).

Taking the predictions of the

17.

shell

model and

figure (16-23) for the spin

of the ground and first excited states of odd A


nuclei in the regions of the magic numbers, where the model should be valid,
use the selection rules (16-83) and (16-83') to predict the character (L value and

quantum numbers and

whether

electric

parities

or magnetic) of the

gamma

decay transitions from the first


Also determine the predictions
nuclei in the regions between magic

excited state to the ground state in these nuclei.

of the collective model for even N, even Z


numbers, where the model should be valid, by consulting section

13.

29
Cu63 and 25-Mev
18. Compare the value of the a{p,pri) cross section for
protons, quoted in figure (16-49), with the projected geometrical area for a
1
13
2
Discuss the implications of this
nucleus of radius R = 1.25 -* x 10~ cm
.

comparison in terms of the optical model. Also calculate the probability that
a 25-Mev proton would make such a reaction in traversing a foil 0.001 inch
29
thick and containing the normal mixture of Cu isotopes.
19.

R =

Solve equation (16-125') graphically, but without approximation, for


2 x 10~ 13 cm. Compare with the predictions of (16-126).

20. Derive the low energy neutron-proton scattering cross section formula
without making the approximation of using the bound state eigenfunction for
r < R in (16-129), and without dropping the KR in (16-135). Show that this
13
R) instead of (16-138). Taking/? = 2 x lO"
gives a = [4t7/(A: 2 + Ap(l +

cm, evaluate the error in (16-138) at


21.

section

is

dajdQ. in the

E=

ev.

ev the isotropic neutron-proton scattering cross


frame, use equation (15-17) to evaluate
20.4 bn in the

Given the

fact that at

=
LAB

CM

frame. Plot the results.

22. A tt meson comes to rest in the LAB frame and then decays into a n meson
and a neutrino according to the scheme (16-152). Show that the kinetic energy
of the n meson in the LAB frame is (m^ m^^llm^, where w and m^ are
the -n and ji meson rest masses, respectively. Evaluate this kinetic energy, in
Mev, from the rest mass values quoted in (16-150) and (16-158).
OT

Ind ex

Absorption, of electromagnetic radiation, 453-7, 505-12


Einstein's coefficient for,

Alpha decay,

459

and magic numbers, 609


and semi-empirical mass formula,
609

rate,

physical picture of, 461-2


transition rate for,

457

of nucleons by nuclei, 577-8

with several optical electrons, 425-8

of particles, 555

Alpha

discovery and nature


89-91
emission through barrier, 236-9,
610-2
Alpha particle scattering, 91-2
and atomic charge, 106-7
and discovery of nuclei, 99-100
and nuclear radii, 107-8
by Rutherford atom, 100-6
by Thomson atom, 92-8
Ampere's law, 339
Angular momentum, 313-4

by absorbing sphere, 556-7, 573


in complex potential, 557-8
in Schroedinger theory, 554-8
Absorption spectra, 113, 121
Actinide elements, 141, 411
Age of earth, 615
Alkali atoms, 418-25

418-24
421-4
quantum numbers for, 421
selection rules for, 424-5
spectra of, 112, 424-5
fine

structure

of,

in,

addition of, in JJ coupling, 442


in LS coupling, 430-2

Alkali elements, 413

in Paschen-Bach
grand total, 450
operators, 314-5

Allowed transitions, 634


Alpha decay, 89-90, 236-9, 609-16
barrier penetration

probability,

238,

610-2
energy, 609-10

447

momentum
in,

structure

641

in,

effect,

orbital, 323; see also Orbital angular

336-9, 449; see also Spin angu-

spin,

equilibrium decay
fine

particles,

of,

Absorptivity, 47

energy levels

potential, 237, 611

radioactive series and, 613-6, 620-1

614-5

lar
total,

Geiger-Nuttall law for, 612-3

momentum

346-9, 429-30, 442

total orbital, 429; see also Orbital

lifetime, 613

angular

long range components in, 641


motion of alpha particle before, 612
nuclear radii and, 239, 612

707

momentum

total spin, 428; see also Spin

momentum
Annihilation, 516-7

angular

708

INDEX

Anomalous Zeeman

effect,

447

Atomic theory,

Antiparticle, 517

Antiproton, 517
Antisymmetric eigenfunctions, 363-73
and exchange force, 373
and exclusion principle, 366-7
for interacting identical particles,

368
for non-interacting identical particles,

363-4, 367

370
369
triplet, 370
Associated Laguerre functions, 302,
304-5
Associated Legendre functions, 301,
304-5
Atomic bombs, 619
Atomic energy states, 124-8
Atomic form factor, 500
Atomic mass formula, 590-2
Atomic mass unit, 582
energy equivalent of, 584
Atomic model, of Rutherford, see
Rutherford's model of atoms
of Thomson, see Thomson's model of
atoms
Atomic number (Z), 87, 560
Atomic radii, in Bohr theory, 116, 133
in Hartree theory, 403
macroscopic estimate, 78-9
in Thomas-Fermi theory, 396
Atomic reactor, 619
Atomic spectra, absorption, 113, 121
of alkali atom, 112, 424-5
singlet,

and

spin,

emission, 113
structure of, see Fine structure
of hydrogen, see One-electron atom
hyperfine splitting of, 357-8, 449
of JJ coupling atom, 443
fine

lines

of

111

of,

LS coupling atom, 440-1

measurement

of,

110-1

of one-electron atom, 111, 119, 425


optical,

417

series of,

112

series limit of, 112

spark,

120-1

Zeeman, 447

of

Bohr,

see

Bohr's

theory of atoms
of Hartree, see Hartree's theory of

atoms
of

Schroedinger,

see

One-electron

atom
of

Thomas and Fermi,

see ThomasFermi theory of atoms


Atomic weight (A), 78, 560
Atoms, charge of, 87-8, 106-7, 488-9
charge density of, 395-6, 400-1
energy levels of, see Energy levels,
of atoms
excitation of, 119, 416-8
mass of, 78, 585
stability of,

108-9, 114-5, 152, 187,

414, 467

Attenuation, exponential law for, 511-2


of nucleons, 577-8
of X-rays, 493-512

Attenuation coefficient, 517


Attenuation length, for nucleons, 577-8
for X-rays, 512
Avogadro's number, 78
determination of, 481
Axial vectors, 637

Balmer series, 113, 120


Barn (bn), 684
Barrier penetration, and alpha particle
emission, 236-9, 610-2
for barrier potential, 233-9
by electromagnetic radiation, 235-6
for step potential, 226-7
and uncertainty principle, 239
and particle-wave duality, 235, 239
Barrier penetration probability, for bar-

233-4
610-2
in three dimensions, 610-1
Barrier potentials, 231-9
rier potential,

for nuclear potential,

^alpha particle emission and, 236-9

boundary conditions
classical

behavior

232
231

for,

of,

eigenfunctions for, 232

energy levels

Ramsauer

for,

232
234-5

effect and,

transmission coefficients for, 233-4


transmission resonances for, 234-5

709

INDEX

Bohr's theory of atoms, radii of orbits,

Bessel functions, spherical, 538

compared

Beta decay, 89-90, 620-40


coupling constant, 628, 635, 640

Coulomb

effects in,

308
Sommerfeld's

631

density of final states for, 628-30

selection rules for,

for,

in,

in,

relation to

636

638-40

638

neutrinos and, 626

Born's interpretation of wave functions,

172-5, 294-5

Bose probability distribution, 389-90

Boundary conditions, for eigenfunctions,

412-3
585-6

of one-electron atom, 300-2

590

and fission, 587


and fusion, 587-8
and saturation of nuclear
675-7
Black body radiation, 47-9
Planck theory of, 63-6

of simple harmonic oscillator po-

forces,

E>

compared

to

quantum mechanics, 313


of, compared to quantum mechanics, 303
and magnetic moments, 327-9

energy levels

postulates of, 114-5, 123, 128-30


postulates of, de Broglie's interpreta-

151-2
predictions of, 115-24
tion,

<

V, 223-4

K, 228

K,

periodic,

and spontaneous emission, 458, 460


Bohr magneton, 329
Bohr's microscope, 157-8
Bohr's theory of atoms, 1 10-37
of,

E>

of square well potential,

Black body spectrum, 48-9

momentum

258-9, 261

tential,

of step potential,

Rayleigh-Jeans theory of, 51-7, 62-3

angular

178, 300

of barrier potential, 232

373,

of, 586,

Boltzmann's constant, 61
Bombarding particle, 563
Born approximation, 524-30
applicability of, 534, 554
for Coulomb potential, 554
for Gaussian potential, 530-2
for relativistic electrons,

Binding energies, of electrons in atoms,

constancy

probability dis-

388-90

for nuclear direct interactions,

636-8
pseudo-scalar matrix element for,
636-8
shell model and, 634-5
spectra, 624
Beta particles, 89-90, 599
parity non-conservation in,

of nuclei,

quantum

tributions,

672-3
569
for square well potential, 532-3

origin of, 621

128,

458-60

derivation of, 57-61

matrix element, 628, 633-8


neutrino recoil

distribution,

applications of, 121,

630

neutrino induced reactions

modification

61

in,

632-3

lifetime,

relativistic

Boltzmann probability

624-5
Fermi selection rules for, 633-4
Fermi theory of, 627-33
Ft values for, 633-6
Kurie plots

quantum mechan-

131-4, 355-6

of,

energy, 621-4

energy conservation

Gamow-Teller

to

ics,

E< K

241

250

.set'

Periodic boundary condi-

tions

for Thomas-Fermi
Box normalization, 220

theory,

394-5

applications of, 252, 270, 525-6, 629

Brackett series,

113,

120

Bragg's law, 479

Breit-Wigner formula, 670


Bremsstrahlen, 489; see also X-ray con-

tinuum spectra
Broadening of spectral

lines,

see Line

widths

Brueckner theory of nuclear structure,


693-4
Bucherer's experiment, 37, 75-6

710

INDEX

Cascades in gamma decay, 642


Cathode rays, 71
Cathode ray tube, 71

Compound

nucleus reactions, independence of formation and decay in,


667

Center of mass, 122

random width assumption

separation of time independent

Schroedinger equation for,

295-8
Center of mass

(CM

frame, 39, 519


relation to laboratory frame, 519-22
Centrifugal potential, 404-5
)

Chain reactions, 618-9


Charge, of atoms, 87-8,

106-7,

406,

488-9
of electrons, 74

measurement

of,

73-5

of nuclei, 106, 560

Charge

density, of atoms, 339-400


of nuclei, 570-2

Charge independence of nuclear forces,


675, 688-9
Classical limit of quantum theory, 206-9
and correspondence principle, 135-6
and identical particles, 364-5
and orbital angular momentum, 324
and simple harmonic oscillators, 68,
264-5
and spin angular momentum, 337-8
and square well potentials, 248-50
and time dependent perturbation
theory, 289-90
Classical physics,

Cloud chamber, 513


Collective model of nuclei, 605-8
and electric quadrupole moments,
606-8
and low lying energy levels, 658-9
and magnetic dipole moments, 606
and liquid drop model, 605
and shell model, 605
Collision broadening of spectral lines,

470

Compound

667

of,

uncertainty principle and, 663

nucleus, 655

and resonances, 655-7

Compound

nucleus reactions, 662-71


angular and energy distributions of,

670-1
Breit-Wigner formula for, 670
excitation and decay of, 664-7
experimental evidence for, 668-70

Complex
Complex
Compton

conjugate, 172
potentials, 557-8,
scattering, 81-5,

572-80
502-5

cross section for, 503-5

from protons, 564


quantum theory of, 82-5, 503

Compton wavelength, 85
Conduction electrons, 382, 386-7
Configuration, 409
Conservation, of charge, 584, 644
of energy, in beta decay, 624-5
and limitations of uncertainty principle, 289, 667, 695-6, 700
in relativity theory, 35-6
and nuclear reactions, 582-5
of probability, 175; see also Normalization, of

wave functions and of

eigenfunctions

and complex potentials, 557-8


Conservation equation, 174-5
Continuity of eigenfunctions and their
derivatives, 178, 321

Continuum of energy
128,

levels,

atomic,

184

nuclear, 662

Contraction of length, see Lorentz contraction

Correspondence

135-6;

principle,

also Classical limit of

see

quantum

theory

Cosmic

radiation, 513

Coulomb
Coulomb

cross section, 554


potential, 295; see also

electron

Coulomb

One-

atom

scattering,

100-6

angular distribution, 106


cross section, 554
Coulomb's law, 42
and quantum electrodynamics, 694-5,
696-8

Coulomb
Coupling

repulsion energy of nuclei, 566


constant,

635, 640

beta

decay,

628,

INDEX

711

Coupling constant, comparison between


magnitudes of, 640
Coulomb field, 701

meson

field,

Degeneracy, and

698, 701

Coupling of angular momenta,


LS, 428-41
orbital, 426-7

of perturbations,

identical

particles,

363
J],

441-3

Sommerfeld theory, 133

in

Degenerate

perturbation

theory,

Perturbation theory

see

degen-

for

restrictions

of exclusion principle,
435-9, 442
spin, 426

erate case
Density of energy levels for free par-

spin and orbital, 428

Density, of atoms, 79

strong

field,

447

ticle,

527-8

of nuclei, 560, 572, 580

Cross product, 313


Cross section, 495, 504-5, 522
absorbing sphere, 557

Born approximation formulas


529-30
Breit-Wigner, 670
CM and LAB, 522-4
Coulomb, 554
differential, 498, 504
Gaussian potential, 531

of final states, in beta decay, 628-30


in

in
for,

nucleon emission, 661


time dependent perturbation

theory, 290
Deuterium, see Deuteron
Deuteron, 677-82

eigenfunction for, 682


electric

quadrupole moment

of,

677-9

excited states of, 682, 685

ground

state of,

678

geometrical interpretation of, 496


hard sphere, 548

potential for,

Klein-Nishina, 503-4

time independent Schroedinger equa-

phase

shift

formulas

for, 544,

556

510
proton-nucleus, 574
proton-proton, 688-91
square well, 532, 551, 684

S-wave, 549

Thomson, 495, 498


X-ray, 509-10
Current loops, 327-8

Daughter nucleus, 609, 613


Davisson-Germer experiment, 147-9
De Broglie's postulate, 139-41
experimental confirmation of, 146-51
to

Schroedinger's theory,

165-6

Differential cross section, 498, 522; see

also Cross section


Differential operators,

for angular

201-5

momentum, 314-5

and associated dynamical quantities,


203
and development of Schroedinger
equations, 293-4
and eigenvalue equations, 318-22
and relation to Schroedinger's theory,
204, 321

and

relativistic

Diffraction,
tion,

of

wave equations, 697

electromagnetic

140,

148,

149-50,

radia-

160-2,

467-80
of electrons, 147-50, 567-8

relation to uncertainty principle, 153

Broglie wavelength, 146

typical values of, 146-7, 150


Degeneracy, 278
in atoms wjth several optical electrons,

680
685

tion for,

virtual state of,

photoelectric, 507,

relation

679, 687

properties of, 677

neutron-proton, 685-6, 689


pair production, 508-9, 510

De

effect

278-9
and exchange of

4^5-6

of helium atoms, 150


of protons, 573-5

of quantum mechanical waves, 533,


548-9, 557
Diffraction

gratings,

467-80

110,

148,

149,

712

INDEX

Dilation of time, see

Time

Eigenfunctions, linear combinations of,

dilation

Dirac theory, of quantum mechanics,


338
and fine structure, 356-7

and pair production, 513-7


of vacuum, 514-6

189-90, 191, 269, 272-3, 276-7,


538
for LS coupling, 432-3
mathematical properties of, 184-92
for multi-electron atoms, 397-8

Direct interactions in nuclear reactions,

for non-interacting identical particles,

662-3, 670-4
angular and energy distributions for,
670-1, 673

normalization of, see Normalization,

363-4, 367-8
of eigenfunctions

Born approximation and, 672-3


excitation and decay of, 672

for one-electron atom, see One-elec-

experimental evidence for, 670-1,

orthogonality of, see Orthogonality of

tron-atom eigenfunctions

672-3

eigenfunctions
for simple harmonic oscillator poten-

Discrete energy levels, 183


finite widths of, 470
Doppler broadening of spectral
469

tial,

lines,

Dot product, 331

for

Doublets, of energy levels, see Fine

square

424
Dynamical quantities, 202
and associated differential operators,
203
fluctuations in, 318-21
lines,

well

potential,

E< V

<

224, 225

243, 247

E> V

structure

of spectral

263-4

370-1
for spin angular momentum, 370
singlet,

250-1

for step potential,

E> V

229

for string, 196


symmetric, 363-4, 368-9
triplet, 370-1

Eigenstates, 187

and emission of radiation by atoms,


461-2
Eigenvalue equations, 318-22
for angular momentum, 323, 538
for energy, 320-1
for linear momentum, 321-2
Eigenvalues, 185, 320; see also Energy

Earth, age of, 615


Effective charge, 401,

Ehrenfest's

theorem,

488
206-9

Eigenf unctions, 185, 320


antisymmetric, 363-4, 367-8
barrier

for

potential,

232

beta decay, 627

in

conditions of acceptability, 178, 300,

levels

Einstein's coefficients,

321

459

Einstein's photoelectric theory,

for deuteron, 682

Einstein's relativity postulates,


finiteness of,

178, 321

for free particle, standing wave,

217

three dimensional, 526, 540


traveling wave, 213

and

gamma

decay transitions, 647


375-8

for helium atom,

for incident and scattered waves, 535


for infinite square well potential, 253,

254
for

interacting

368-9

identical

particles,

79-81
15-6

Einstein's mass-energy relation, 34


of, 585-6
and nuclear binding energy, 585-6
and nuclear reactions, 582-5
Einstein's quantum postulate, 80
and theory of radiating atoms, 455,
462
Einstein's spontaneous emission theory,
458-60

experimental verification

Electric

dipole

460-1

matrix elements, 455,

INDEX

713

Electric

dipole

transitions,

454,

467,

Energy

644
Electric quadrupole moment, 606, 646
and collective model, 606-8
of deuteron, 677-9
of nuclei, 451-2, 607
and shell model, 608
Electric quadrupole transitions, 468,
644
Electromagnetic

levels,

of atoms,

for hydrogen, see One-electron

atom
in JJ coupling,

LS

in

441-2

coupling, 433-4, 439-40

for

multi-electron atoms, 403-5,


408-9, 41 1-2, 415, 425-8, 433-4,
439-40, 441-2

absorption

for one-electron atoms,

453-7, 461-2
classical theory of, 42-6
dual nature of, 85-6, 139, 161-2
energy density of, 43
energy flux of, 45
induced emission of, 457, 461-2
rate of energy emission, 46
spontaneous emission of, 453, 460,

303-4, 308-9, 403-5

radiation,

of,

461-2
Electromagnetic wave equation, 697
Electron, capture in beta decay, 623-4
diffraction,

Zeeman
of

of infinite square well potential, 255


of nuclei, average spacing, 659-60

model, 658
662
density of, 660
excitation of, 641
in Fermi gas model, 595
measurement of, 655-8; see also
Beta decay; gamma decay
in shell model, 601-2, 658-9
widths of, 660-2
in collective

continuum

by nuclei, 567-70

Electron charge, 74

73-5

Electron mass, 74

measurement

of,

73-5

Electron-proton mass ratio, 124


Electron spin, see Spin angular momen-

446-7

particle,

of free electron, 128, 513-7


of free particle, 312

emission in beta decay, 623-4

of,

splitting of,

134,

of Fermi gas, 383

147-50, 567-8

measurement

bound

117,

184
continuum, 128, 184
discrete, 183

discovery of, 70-3


scattering

helium,

for

374, 376

of,

of rigid rotator, 138, 325

of

tum

simple

harmonic

oscillator,

64,

131, 262

Electron volt (ev), 125

of step potential, 222


of square well potential,

Emission spectra, 1 1
Emission theories, 15

<V

244-5

E > V

Emissivity, 46

240, 250
unbound particle, 184
X-ray, 485-6
(

e/m

ratio for electrons, 73

Energy, equipartition
relativistic kinetic,

rest mass,

of,

61

33

Equilibrium decay, 614-5


Equipartition of energy, 61

34

total relativistic, 34
Energy conservation,
tions, 582-5

of

Ether, 7
in

nuclear reac-

in relativity theory, 35-6


and uncertainty principle limits, 289,
667, 695-6, 700
Energy level diagram, 64
Energy levels, of atoms, for alkali
atoms, 418-24

Ether drag, 14
Ether frame, 8

Even-odd effects in nuclei, 589, 591


and Fermi gas model, 596-9
Exchange, of labels for identical particles, 362-3
of mesons, 694-5
and repulsive core, 703

714

INDEX

Exchange, of mesons, and Serber potential,

698-700

Exchange degeneracy, 363


helium atom, 375
Exchange force, 372-3
in helium atom, 377-8, 380
Exchange integral, 380
and nuclear Coulomb energy, 566
Exchange operator, 699
Excitation, of atoms, 119, 416-8
in

of nuclei, 641

Excited

states,

119;

see

also

Energy

Fermi
Fermi
and
and
and
and

gas, exclusion principle and,

depth of nuclear potential, 595-6

596-9
595-6
semi-empirical mass formula,
596-8
and shell model, 595, 599
Fermi probability distribution, 388-90
Fermi selection rules, 633-4
Field, electromagnetic, 694-5
meson, 694-5
even-odd

effects,

optical model,

Fine structure, 131, 334-5, 338-9,

levels

Exclusion principle, 366-8

and angular momentum of holes, 485


and antisymmetric eigenfunctions, 367
and capacity of subshells, 408
and conduction electrons, 386-7
and experimental evidence, 381
and Dirac theory of pair production,
514
and Fermi gas, 384, 386-7, 595
and Fermi probability distribution,
390
and independent particle motion in
nucleus, 595
and JJ coupling, 442
and LS coupling, 435-9
and properties of atoms, 414
and proton-proton scattering, 687
and theory of nuclear structure, 694
Expectation values, 201-5
general formula, 205
order of terms in integrals, 204, 211
relation to perturbation theory, 272
Exponential attenuation, 511-2
Exponential decay, 468-9
of compound nucleus states, 666

355-7
atoms, 421-4

in alkali
in

JJ coupling, 441

in LS coupling, 433-4, 441


Fine structure constant, 134
and quantum electrodynamics, 701

and spin-orbit energy, 352


Finitenesse of eigenfunctions and their

Fermi beta decay theory, 627-33


Fermi energy, 385, 387
of conduction electrons, 386
of nuclei, 596
Fermi gas, 381-9
applications of, 382
conduction electrons and, 386
energy distribution of, 383, 388
energy level density of, 383
of,

383

178, 321

derivatives,

Fission of nuclei, 616-20

chain reactions

in,

energy diagram

618-9

for, 616,

energy emitted

618

587

in,

fragment distribution, 619-20


fragments, 616
induced, 617-8
neutron emission
origin of,

in,

618

616

spontaneous, 617
Fluctuations

in

dynamical

quantities,

318-21
Flux, classical, 174-5
incident, 504-5,

522

probability, 175; see also Probability


flux

scattered, 505,

energy levels

384

gas model of nuclei, 594-9

Forbidden

Form

522

transitions,

factor,

635

500

Fourier analysis, 154, 158, 191, 200


Franck-Hertz experiment, 124-8
Free particle, 212-20
classical behavior of, 212
eigenfunctions, standing wave, 217
three dimensional, 526
traveling wave, 213
energy levels, 128, 213

INDEX

715

Free particle,

normalization

wave

of

Groups of wave

functions for, 215, 219

functions,

142-5,

153-4, 159-60, 163, 215-6

probability density, standing wave,

Group

velocity, 142, 144, 166

218
traveling wave, 213

Halogen elements, 413

probability flux, standing wave, 218


traveling wave, 214

Hamiltonian, 321

time independent Schroedinger equa-

see

tion,

wave

212

in

functions, standing wave, 217

traveling wave, 213

Ft

values,

Hartree's

633-6

of atoms,

theory
also

ments, 400, 501


theory of

Heisenberg's

Gamma

cascades, 642

and magnetic
644-8

energy of

gamma

rays

transitions,

642
650-3

in,

internal conversion and,

isomeric transitions, 643


lifetimes,

652-3

matrix elements, 649-50

measurement of energy, 642-3


measurement of lifetime, 643

rays,

Gamow-Teller

splitting, 357-8, 449


Hyperfine structure, 449-52

and

647-8

89-90, 559-60
selection rules,

Hermite equation, 259


solution of, 259-62
Hermite polynomials, 263
Holes, and pair production, 515-7
properties of, 437, 516
quantum numbers for, 485
selection rules for, 486
subshells with, 437
X-ray spectra and, 484-5
Hydrogen, see One-electron atom
Hyperfine

selection rules, 648


transition rates in shell model,

Gamma

quadrupole interaction,

electric

451-2
and magnetic dipole interaction, 448-

636

Gaussian potential, 531


Gauss' law, 393
Geiger counter, 563-4

51

Hyperons, 703
Identical

Geiger-Nuttall law, 612-13

General theory of relativity, 38


g factor, Lande, 446
meson, 703
nuclear, 450
orbital,

quantum me-

chanics, 164-5, 337

decay, 640-55

electric

329

particles

ability,

and

and beta decay, 627-30


and Born approximation, 527-9
and widths of nuclear energy levels,
661-2

Independent particle motion in nucleus,

593-5
360-1
Induced emission, 457
Indistinguishability,

459

physical picture of, 461-2


transition rate for,

Induced

fission,

Inertial frame,
Infinite

457
617-8

square well potential, 251-6

applications of, 252


boundary conditions

Gravitational forces compared to nuclear forces, 566

eigenfunctions, 253, 254

Ground

energy

119

indistinguish-

360-1

Einstein's coefficient for,

336-7, 357
Golden Rule No. 2, 291
spin,

state,

atoms,

Hartree theory

and nuclear shell model, 600


and theory of nuclear structure, 693
and verification by X-ray measure-

Fusion of nuclei, 587-8


Galilean transformation, 5
and mechanics, 3-9

396-405;

Multi-electron

levels,

255

for,

252-3

716

INDEX

Infinite

square well potential, normalization of eigenfunctions for, 254

parity of eigenfunctions for,

253-4

time independent Schroedinger equation,

253

zero point energy


Interaction

255-6

of,

strengths,

Legendre functions, associated, 301,


304-5
Legendre polynomials, 537
Lifetime of quantum states, 469
and widths of energy levels, 472
Limiting velocity in relativity,

comparison

and range of nuclear

tween, 640
Interference of eigenfunctions and

wave

17,

27,

29, 38, 44

be-

Line, spectral,

forces,

696

111

153-4,

of stability, 618
Linear combinations, see Superposition

Internal conversion, coefficients, 651-2

of eigenfunctions, 189-90, 191, 269,


272-3, 276-7, 538

functions,

151-2,

148,

365
Interferometer,

of wave functions

energetics of, 650-1

gamma

of wave functions, 186, 192, 285

decay and, 650-3

Interval rule of Lande, 434, 441

Linearity of Schroedinger equation,

Ionization energy, 128, 413

166-7, 185-6
Line widths, 469-72

Isomeric transitions, 643


Isotopes, 581

abundance

of, 582,

588-9

hyperfine structure and, 449


JJ coupling, in atoms, 441-3
in nuclei,

603

Kirchhoff's law, 47
Klein-Gordon equation, 697

Klein-Nishina cross sections, 503-5

K mesons,
Kurie

703
630

plots,

Laboratory (LAB) frame, 39, 519


relation to center of mass frame,

519-22
Laguerre functions, associated, 302,

433-4, 441
of orbital angular momenta, 429
fine structure splitting,

304-5

Lamb

shift,

and collision broadening, 470


and Doppler broadening, 469
and natural broadening, 469-70
and uncertainty principle, 469-70,
472
Liquid drop model of nuclei, 589-91
and collective model, 605-8
and fission, 617
and semi-empirical mass formula,
590-1
and shell model, 605
Lorentz contraction, 14, 18, 21-2
Lorentz transformation, 26
derivation of, 24-7
LS coupling, 428-41
energy levels, 433-4, 439-40, 441
and exclusion principle, 435-9

357

quantum numbers, 429-33, 435

Lande g factor, 446


Lande interval rule, in atoms, 434, 441

selection rules,

in nuclei, 600
Lanthanide elements, 411, 414
Laplacian operator, in rectangular co-

of spin angular momenta, 428


of total angular momenta, 429

spectra,

Lyman

440-1

440-1

series,

113, 120

ordinates, 294
in spherical coordinates,

298

Larmor frequency, 332


Larmor precession, 331-2
of net magnetic moment, 444-5
Last nucleon in nucleus, 592-3

Magic numbers, 592-4


and alpha decay energy, 609
and electric quadrupole moments, 608
and fission fragment distribution,
620

INDEX

717

Magnetic dipole moment, of deuteron,


677-9
of electron, 336-7
of filled subshell, 438-9
net, 443
of neutron, 679
and meson theory, 700
in non-uniform magnetic field, 333-4
orbital, of electron, 327-9
of positron, 516
and precessional motions, 331-2,
444-5
of proton, 679
quantization, 329, 336
spin, of electron, 336-9
Magnetic dipole moments of nuclei,
335-6, 357-8, 449

and collective model, 606


and electrons in nuclei, 562
and shell model, 603-4
Magnetic dipole transitions, 468
Magneton, Bohr, 329
nuclear, 450
Mass, of deuteron, 677
of electron, 74
measurement of, 73-5
of n meson, 702
of v meson, 701
of neutrino, 626
of neutron, 585
of nuclei, 585
measurement of, 580-5
of positron, 516
of proton, 585
relativistic,
rest,

Mass
Mass

32

32

deficiency of nuclei, 585-6

density of nuclei, 572


Mass-energy relation, 35-6
and nuclear reactions, 565
Mass spectrometry, 580-2
Mass unit, atomic, 582
energy equivalent of, 584
Matrix element, 273, 286
for beta decay, 628, 633-8
electric dipole, 455, 460-1
for

gamma

decay, 649-50

nuclear, see Beta decay

Maxwell's equations, 7

Mechanics,

classical,

3-4

30-7
Mesons, 694-703
charge of m, 701
relativistic,

charge of *-, 698, 700


decay of /*, 702
decay of w, 701-2
discovery of

/t,

discovery of

tt,

702
700-1

hyperons, 703
interaction of m, 702
interaction of v, 702-3

K mesons,

703

lifetime of

/i,

lifetime of

tt,

702
701-2

and time dilation, 702


magnetic dipole moment of fi, 703
magnetic dipole moment of ir, 703
mass of /i, 702
mass of tt, 696, 701
and range of nuclear forces, 695-6
spin of ,u, 703
spin of tt, 703
lifetime

symmetry character of /i, 703


symmetry character of *, 703
Meson theory, 696-700
and role of perturbation theory, 701
Michelson-Morley experiment, 9-13
Microscopic reversibility, 458
Mirror nuclei, 565
beta decay of, 634-5
binding

energies

and nuclear

radii,

565-6
Models, 88, 559
and theories, 604-5

Modern

physics,

Multi-electron atoms, in Hartree theory,

396-405; see also Alkali atoms;

Atoms, with several optical

elec-

trons; Periodic table of elements


effective

charge

for,

401

eigenfunctions for, 397-8


net potentials for, 397,

probability

density

400-1

for,

397-401,

404-5

quantum numbers

for,

self-consistency in,
shells

in,

402-3

398

397

718

INDEX

Multi-electron atoms, in Hartree theory,

411-

subshells in, 403-5, 408-9,


2,

415

time independent Schroedinger


equation for, 396
in

gas model of nuclei

Thomas-Fermi theory, 393-6


boundary conditions for, 394-5
net potential for, 394-5
probability density for, 394-5, 401
self-consistency in, 393

Multiple scattering, 92-8

angular distribution for, 98


Multiplet, 434

Nal

scintillation detector, 537


Natural broadening of spectral

lines,

469-70
Natural width of spectral

Net

470

lines,

potential, in Hartree theory, 397,

400-1
in multi-electron
in

atoms, 392

Thomas-Fermi theory, 395

Neutrino, beta decay and emission

626-33
meson decay and emission
induced reactions, 638-40
properties of, 626
recoil in beta decay, 638
Neutron, 562
beta decay of, 630-1
discovery of, 562-5
magnetic dipole

moment

of,

of,

of,

701-2

679

and meson theory, 700


mass of, 585
spin angular

momentum

symmetry character

of,

of,

562

562

Newton's laws, 4
for angular

momentum, 331

187-8,

of eigenfunctions,

190

examples, 254, 367-8, 525-6


of probability density, 175

wave functions,
219-20
Nuclear bombs, 619
of

187,

190-1, 215,

Nuclear collective model, see Collective

model of nuclei

Nuclear fission, see Fission of nuclei


Nuclear forces, 674-94
angular momentum dependence of,
690
attractive character of, 556, 674
charge independence of, 675, 688-9
and deuteron, 677-82
and eigenfunction for deuteron, 682
and meson theory, 694-703
and neutron-molecular hydrogen
scattering, 685-6
and neutron-proton scattering, high
energy, 689-90
low energy, 682-7
and nuclear structure, 693-4
and potential for deuteron, 679, 687
and proton-proton scattering, high
energy, 690-3
low energy, 687-8
range of, 579, 674, 687, 692
compared to other forces, 579
and meson theory, 696, 697
repulsive cores in, 691-3
and meson theory, 703
and saturation of, 693-4
saturation of, 675-7, 690, 693

and Serber potential, 690, 691-2,


693, 699-700
spin dependence of, 685, 693
and spin-orbit potential, 693
and stability of nuclei, 566
strength of,

compared

to other

forces, 556, 579-80,

Noble gas elements, 412-3


Normalization,

Nuclear Coulomb repulsion energy,


590-1
Nuclear even-odd effects, 589, 591
and Fermi gas model, 596-9
Nuclear Fermi gas model, see Fermi

640
and tensor potential, 678-9, 693
and virtual state of neutron-proton
system, 685-6
Nuclear fusion, see Fusion of nuclei
Nuclear liquid drop model, see Liquid
drop model of nuclei
Nuclear magic numbers, see Magic
numbers
Nuclear magneton, 450

INDEX

719

Nuclear mass formula, 590-2


and alpha decay energy, 609
and beta decay energy, 621-2
Nuclear matrix element, 628, 633-6
Nuclear model of atoms, see Rutherford's model of atoms
Nuclear optical model, see Optical
model of nuclei
Nuclear potential, 572-80; see also Shell

model of nuclei
composition of total real part, 57980
Nuclear radii, for charge distribution,
565-72
comparison with potential
tion radii, 578-9
electron
of,

scattering

Nuclei, composition of, 561-5


electric

quadrupole moments

Electric quadrupole

energy levels
of nuclei

Energy

see

of,

measurements

567-72

levels,

magnetic dipole moments of, see


Magnetic dipole moments of
nuclei

mass of, 585


mass deficiency of, 585-6
mass density of, 560, 572, 580
number of protons and neutrons

in,

565
spin of, see Spin angular

distribu-

of, see

moments

momentum,

of nuclei
stability of,

587-8, 609, 615-6, 622


of, 561

symmetry character
Nucleons, 565

mirror nuclei binding energy measurements of, 565-6

Numerical

integration

equations,

of differential
180-2, 268, 276-7

values of, 571


for potential distribution,

572-80

alpha particle scattering and decay

measurements, 100, 107-8, 239,


612
comparison with charge distribution radii, 578-9
neutron cross section measurements
of, 572-3
proton scattering measurements of,
573-9
values of, 575-6
nucleus,

see

Compound

nucleus reactions
direct, see Direct interactions in nuclear reactions

history of, 563-5

136-7

angular momenta,
expectation values, 317
precessional motion of, 323-4
quantization of, 323

One-electron-atom

eigenfunctions,

be-

cause, 308-9, 404-5


normalization of, 307

orthogonality of, 307

466

quantum numbers

for,

302-4, 337,

349-50
One-electron-atom energy levels, degeneracy of, 303-4, 350-1

notation for, 563


Q value of, 583

Nuclear reactors, 619


Nuclear shell model, see Shell model
of nuclei

Nuclear structure, theory

of,

693-4

Nuclei, binding energies of, see Binding energies, of nuclei


of,

of,

One-electron-atom

parity of,

energy balance in, 582-5


energy level measurements and, 655-8

charge density

critique

One-electron atom, 295-325, 350-8


and role in theory of multi-electron
atoms, 402-3

havior near origin, 309


degeneracy of, 303-4, 350-1

Nuclear reactions, 662-74

compound

Old quantum theory, 136

570-2

cause, 308-9, 404-5


with fine structure, 355-7
in old quantum theory, 117, 134
in Schroedinger theory, 303

One-electron-atom probability densities,


angular dependence of, 309-13
behavior near origin, 309
measurability of, 311-2

INDEX

720
One-electron-atom probability densities,
radial
relation to

306-9
angular momenta, 312-3

dependence

of,

One-electron-atom time independent


Schroedinger

equation,

com-

295

plete,

464-6
and atomic transition selection rules,
467
and beta decay selection rules, 634,
636
of deuteron ground state, 678

Parity, 247,

of eigenfunctions, for infinite square

for relative motion, 298

295-8

253-4

well potential,

separation of complete equation,

for one-electron atom,

separation of relative equation, 298-

for simple

harmonic

466

oscillator,

263

for spherically symmetrical poten-

9
solution of relative equation,

299-

tial,

466
247-8

for square well potential,

302

Optically active (optical) electron, 418

and form of potential, 248


and gamma decay selection
648, 650

Optical model of nuclei, 572-80

non-conservation in beta decay, 636-

Operators, see Differential operators


Optical excitation of atoms, 417, 484

and Fermi gas model, 594-6


and initial stage of nuclear reaction,
663
and shell model, 594
values of parameters, 575-7
Orbital angular

of

momentum, 323

filled subshell,

438

of nuclear energy levels, 653, 658-9

measurement
decay,
origin of,

quantization of, 114, 129-30, 323

429
in atoms with several optical electrons, 426-7
quantization of, 429
Orbital g factor, 329
Orbital magnetic dipole moment, 327-

Partial

Ordinary integral, 380

magmoment, 330-1
magnetic dipole moment, 445

Orientational potential energy, of


netic dipole

of net

Orthogonality of eigenfunctions, 188-90


in

continuum, 189

when

degenerate, 189-90

decay

253-4
wave

equations,

248,

554-6
556-

analysis, 534-44,

for absorbing sphere scattering,

total,

quantization of, 329

658; see also Beta

465

Schroedinger

atom

precessional motion of, 323-4

of,

Gamma

and perturbation theory, 276.


and solution of time independent

of one-electron atom, see One-electron

rules,

applicability of, 542, 549,

554

cross section formulas for, 544, 556


for hard sphere scattering,

546-9

for neutron-proton scattering, 682-4

691-2
549-52
Partial waves, 537-8, 555
orbital angular momentum of, 538
and plane waves, 540-2
Particle-wave duality, 85-6
and barrier penetration, 235, 239
and uncertainty principle, 160-2,
226-7
Paschen-Bach effect, 447-8
for proton-proton scattering,
for S-wave scattering,

Pauli principle, see Exclusion principle

Periodic boundary conditions, for elecPairing energy of nuclei, 598

Pair production, 508-9, 513-7


cross section for, 508-9, 510

Parent nucleus, 609, 613

tromagnetic radiation, 52
in relativistic range,

for

629

wave functions and eigenfunctions, 220, 525-6

INDEX

721

Periodic table of elements, 406


actinide elements, 411,

Phase

414

elements, 413
halogen elements, 413
Hartree theory interpretation

and cross

alkali

of,

407-

interpretation of,

for

noble gas elements, 412-3

for

rare earth elements, 411, 414

group elements, 413-4


transuranic elements, 414
Perturbation, 269
and effect on degeneracies, 278-9
Perturbation theory, and meson theory,
701
and quantum electrodynamics, 701
transition

Perturbation theory for degenerate case,

278-83
and atoms with several optical electrons, 425-8
classical example of, 283-5
formulas for, 279-81
and helium atom, 373-81
and JJ coupling, 442
and LS coupling, 428-9, 432-4
and relativistic corrections to Schroedinger's equation,

353-4

simplified treatment of, 282-3, 350-1

and spin-orbit energy, 350-3


and Zeeman effect, 445-6
Perturbation theory for non-degenerate

268-73
277-8
formulas for, 272
case,

of,

use in degenerate case, 282-3

and helium atom, 373-6


and V-bottom potential, 274-8
Perturbation theory for time dependent

285-91; see also Golden

Rule No. 2
and absorption of electromagnetic
radiation, 453-7
and Einstein's quantum postulate, 455
formulas

286
and Golden Rule No. 2, 286-91
and induced emission of electromagnetic radiation, 457
Pfund series, 113, 120
for,

556

sections, 544,

553

of,

for hard sphere scattering, 547

lanthanide elements, 411, 414

case,

542, 555-6

energy dependence

14

accuracy

shifts,

for absorbing sphere scattering, 557

544-6

>

KR, 547
Ramsauer effect, 552
/

551-2

for resonance,

for S-wave scattering,

Photoelectric

cell,

551

76

Photoelectric effect, classical theory of,

78-9
cross sections for, 507-8, 510
early experiments on, 76-8, 81

quantum theory
X-ray

of,

line spectra

79-81, 507
and, 507

Photoelectrons, 77
Pilot waves,

140; see also

Wave

func-

tions

Planck's black body spectrum theory,

63-6
Planck's constant, 66

and uncertainty principle, 158


and X-ray continuum spectra, 482
Planck's quantization postulate, 64
comments on, 66-8, 128-31, 262-3
Polarization, of scattered nucleons, 602,

693
of scattered X-rays, 476

Polar vectors, 637


Positronium, 517
Positrons,

513-7

annihilation of, 516-7

Dirac theory of, 514-7


discovery of, 513

emission of, in beta decay, 623-4


holes and, 516
properties of, 516
Potential,

alpha

particle-nucleus,

see

Alpha decay
barrier, see Barrier potentials

constant, see Free particle

Coulomb,

see

Coulomb
atom

potential;

One-electron

deuteron, see Nuclear forces

Gaussian, see Gaussian potential

INDEX

722
Potential,

infinite

square well, see In-

nucleon-nucleon, see Nuclear forces


nucleon-nucleus, see Optical model of
nuclei; Shell

model of nuclei

harmonic

oscillator

square well, see Square well potential

step, see

Step potential

V-bottom, see V-bottom potential

Yukawa,
Potentials,

Yukawa

see

and

potential

general treatment,

coordinate

dependent,

separation of center of mass

of energy in Schroedinger theory,

of grand total angular


of nuclear spin angular

Lar-

294-

for atom emitting radiation, 461-2


for free particle, standing wave, 218
traveling wave, 213

of total orbital angular

and Wilson-Sommerfeld

for multi-electron atom, 397, 401,

404-5

momentum,

rules,

Quantum, of electromagnetic

atom
atom, 401
atom, 306-9

in multi-electron

well

oscillator,

potential,

264

<

248
251

for step potential,

E< V

two non-interacting
ticles, 363-5

128-

field,

80,

694-5
of meson field, 694-5
Quantum defects, 420
Quantum electrodynamics, 462
and origin of Coulomb interaction,

307

harmonic

347,

32

for one-electron atom, see One-elec-

momentum,

429-30

of total spin angular momentum, 429


and uncertainty principle, 157-8

wave function, 186-7


eigenstate wave function, 187

square

moment, 329

429

for general

in one-electron

114,

of simple harmonic oscillator, 64


of total angular

for simple

momentum,

of orbital magnetic

Probability density, 171-2, 172-5,

momentum,

323

Probability conservation, 175

for

momentum,

449
of orbital angular

mor, 331-2, 444-5


and uncertainty principle, 324, 430

E>V

117,

450

general treatment, 176, 294

tron

124-8

of atomic energy states,

178-84

Precession of angular momenta,

for

moment

and

285

radial,

Quadrupole moment, see Electric quad-

mo-

time dependent, general treatment,

for

quantum mechanics, 160

Quantization, 64; see also Energy levels

298

tion,

Probability in

Product particle, 563


Proper length, 22
Proper time, 21
Proton, 560
Pseudo-scalars, 636
and beta decay, 636-8

rupole

298-9
and spatial quantization, 324-5
relative

178

forces,

radial dependent,

for incident

simple harmonic oscillator, see Simple

214
and scattered waves, 536
for step potential, E < V
225
E > V 229
traveling wave,

Net potential

net, see

Probability flux, 175


for free particle, standing wave, 218

square well potential

finite

226

identical par-

694-5, 696-8
and perturbation theory, 701
and spin g factor, 357
and spontaneous emission of radiation, 458, 462
Quantum numbers, 64, 185
assignment of, 448-9

INDEX

723

Quantum numbers,

for atoms, with //

coupling, 442

LS

with

coupling, 429-33, 435

with one electron, 302^1


with one optical electron, 421

with several optical electrons, 425


for nuclear shell model, 600
spectroscopic notation, for

408

486
422-3

for n,
for

/,

s,

and energy balance in nuclear reactions, 565, 582-5


experimental verification

585

of,

30-7

Relativistic mechanics,

momen-

Relativistic transformation, of

tum and energy, 36-7

for X-ray levels, 485

Quantum

mass-energy relation, and


binding energy of atomic electrons, 421, 580
and binding energy of nuclei, 585-6

Relativistic

64
lifetime of, 469,. 472
Quantum statistical mechanics, 390
states,

of position and time, 26


of velocity, 29
Relativistic

wave equation, 697

Relativity theory, general theory, 38

Quantum

experimental verification

special theory,

theory, 41
value of nuclear reaction, 583

of,

37-8

3-38

Relativity theory applications, see Rela-

Radial probability density, 307


in multi-electron atoms, 401
in one-electron atoms, 306-9

tivistic

Compton

mass-energy relation
83

effect,

corrections

to

Born approximation,

Radiation by accelerated charges, 42-6


Radii, of atoms, see Atomic radii
of nuclei, see Nuclear radii

corrections to Schroedinger equation,

Radioactive decay, 89; see also Alpha


decay; Beta decay; Gamma decay

non-radiation of uniformly moving


charge, 43

Radioactive series, 613-6, 620-1


branching in, 621

pair production, 513-7

equilibrium

Ramsauer

Random

569

353-4

Sommerfeld theory of

614-5

in,

234-5, 552, 657

effect,

walk, 97

Range, of charged

563

compared to other forces, 579


and meson theory, 696, 697
Rare earth elements, 411, 414
theory

radiation, 52-7,

of black

and meson theory, 703


and saturation, 693-4
Residual

body

62-3

spectral distribution of, 62

Recursion relation, 260, 268


Reduced mass, 123, 519
in

time independent Schroedinger


equation for
motion, 297

CM

Reflection coefficient, for step potential,

E < V 225
V 229-31
,

33-4
34-5

Relativistic energy,
Relativistic mass,
Relativistic

mass-energy relation, 34-5

Coulomb

interactions,

426-8

Residual nucleus, 563


Resonances, 551-2

many

particle vs. single particle,

657, 664
in nuclear reactions,
in nuclei,
in

655-7

577

square well potentials, 251

for transmission through barrier potentials,

E>

precession, 340-4, 359

Repulsive cores in nuclear forces, 691

of nuclear forces, 579, 674, 687, 692

Rayleigh-Jeans

momentum, 338

spin angular

Thomas
particles,

fine structure,

133

and

234-5
552-3

virtual states,

Rest mass, 32
Rest mass energy, 34
Rigid rotator, energy levels

of, 138,

325

724

INDEX

Russell-Saunders coupling, see

LS

Selection rules, for alkali spectra, 424


for beta spectra, 634, 636

coupling

Rutherford scattering experiments, 912, 98-9, 106-8; see also Cou-

lomb

scattering

Rutherford's model of atoms, 99-100

and Bohr's theory, 114


predictions of, 100-6
stability of,

108-9, 114-5, 152, 187,

461
verification of,

106-7

Rydberg constant, for hydrogen, 112


for infinitely heavy nucleus, 119, 123
for nucleus of mass M, 123
Rydberg series, for alkalis, 112, 420,
424

Saturation of nuclear forces,

675-7,

690, 693-4

in

693-4

Thomas-Fermi theory of atom,


393

and alpha decay energy, 609


and beta decay energy, 621-2

and meson theory, 698-700

of alpha particles by nuclei, 91-108


of electrons, by atoms, 400

by nuclei, 567-70
of neutrons, by molecular hydrogen,

685-6

atomic, 112
Balmer, 113, 120
Brackett, 113, 120

Series,

Lyman,

113, 120

Pfund, 113, 120

by protons, 682-7, 689-90


and nuclear physics, 518
of protons, by nuclei, 572-80
by protons, 687-8, 690-3
of X-rays, 493-505
Schmidt lines, 604
Schroedinger's equation, 170, 294
and classical wave equation, 193
complex character of, 170
development of, 165-70, 293-4
and differential operators, 204
linearity of, 166-7, 185-6
and Newton's laws, 206-9
independent,

see

Time

Rydberg, 112, 420, 424


Series limit,

postulates of, 165-6, 321

counter,

Nal,

573

112

Series solution of differential equations,

259-60, 301-2

Shadow

scattering, 548-9,

557
599-604
and beta decay, 634-5
and collective model, 605-8
and electric quadrupole moments,

Shell

inde-

pendent Schroedinger equation


Schroedinger's theory of quantum mechanics, and de Broglie's postulate, 165-6
and energy quantization, 178-84

ZnS, 92

Hartree theory of atom, 397

in nuclear structure theory,

Separation constant, 176


Serber potential, 690, 691-2, 693

636
Scattering, 555

Scalars,

Scintillation

in

Semi-empirical mass formula, 590-2

for hydrogen, 112

time

example of evaluation of, 463-7


for gamma spectra, 648
initial and final eigenfunctions and,
463
for JJ coupling spectra, 443
for LS coupling spectra, 440-1
in old quantum theory, 134, 135-6
for X-ray spectra, 486
for Zeeman spectra, 447
Self-consistency, 392

model of

nuclei,

608
and Fermi gas model, 594-5
and gamma decay, 647
and Hartree theory, 600

and JJ coupling, 603


and liquid drop model, 605
and magnetic dipole moments, 6034
and optical model, 594
and spin angular momentum, 603
and spin-orbit interactions, 600-2
theoretical basis of, 593-5

INDEX

725

Shells, in multi-electron atoms,

402-3

602

in nuclei,

Spectroscopic notation, for

307-8

in one-electron atoms,

Spin angular

Shielding, see Effective charge

Simple harmonic

oscillator,

boundary conditions

for,

classical behavior of, 68,

of,

258-9
256

438
meson, 703
of T meson, 703
of neutrino, 626
of neutron, 562
of nuclei, and beta decay selection
rules, 634, 636
and electrons in nuclei, 562
and gamma decay selection rules,
648, 650
measurement of, 451, 658; see

theory, 64, 131

parity of eigenfunctions for, 263


probability densities for, 264

quantization of, in old


theory, 64, 130-1

quantum

selection rules for, 136


time independent Schroedinger equation for,

also Beta decay;

370-1
and exchange force, 372-3

quantum numbers

for,

372
684

Single-valuedness of eigenfunctions and


their derivatives,

Slater determinant,

178, 300, 321

theory of

and fine structure, 131, 338


and quantum mechanics, 355-7
Spatial quantization, 324
experimental verification

of,

335

and form of potential, 324-5


Spectra, analysis of, 448-9
atomic, see Atomic spectra
black body, 48-9

624-33
417
simple harmonic oscillator, 136
X-ray, continuum, 482-3, 489-92
line, 483-9
beta,

optical,

111
widths of, see Line widths
Spectroscopic notation, for /, 408
origin of,
for n,

486

424

quantum numbers

for, for s

X,

and symmetry character, 369


428
in atoms with several optical electrons, 426
quantization of, 371, 429
Spin dependence of nuclear forces, 685,
693
total, 371,

atoms, 131-4

Spectral line,

decay

336

step potential, 231

relativistic

Gamma

and shell model, 603


and symmetry character, 369
values of, 451
of positron, 516
of proton, 562
quantization of, for s = I, 336

367

Solid angle, 497

Sommerfeld's

filled subshell,

ij.

quantization of, 449

Singlet eigenfunctions,

Smoothed

of

of

257-8

zero point energy of, 262


Simultaneity, 16-8

Singlet state of deuteron,

of deuteron,

eigenfunctions for, 300, 337


of electron, 336-9

262

quantum

momentum,

256-66

eigenfunctions for, 263-4


in old

422-3

677

applications of, 256

energy levels

s,

Spherical Bessel functions, 538

Spin g factor, 336-7, 357


Spin-orbit interaction, in atoms, 339-46,

351-3
for alkali

atoms, 421-4

formulas for, 345, 353


for // coupling, 442

LS coupling, 433-4
magnitude of, 345
in nuclei, 600-3
and JJ coupling, 603
magnitude of, 602
and nuclear properties, 600-4
and nucleon-nucleon scattering,
693
and nucleon-nucleus scattering, 602
for

726

INDEX

Spontaneous emission, by atoms, 453,

Step potential, probability

E< V

458-60
459

Einstein's coefficient for,

E > Vu

dipole interactions, 468, 644-

physical picture of, 461-2


transition rate for, 460, 463
by nuclei, see Gamma decay
Spontaneous fission, 617
Square well potential, 239-51
applications of, 240-1
boundary conditions for, E < V Q
241
E > V 250
classical behavior of, 240
eigenfunctions, E
V 243, 247
E > V 250-1
energy levels, E
V 244-5

E> V

wave
,

240, 250

<

247-8
248

Stokes' law, 74

Strength function, 664


Subshells, 405
in alkali atoms,

419-20
405

ordering according to energy, 408-9,


spherical

symmetry of

filled,

398-9,

412

E> V

240,

144,

251

148,

143,

167; see also Linear

combinations

of atoms,

114-5,

108-9,

152, 187, 414, 467


of nuclei, 587-8, 609, 615-6, 622-3
Standing waves in cavity, 52-7

number of allowed frequencies

for,

57

Surface tension energy of nuclei, 590

S-wave

and

187; see also Eigen-

states

46

Stefan's law,

barrier penetration of,

E< V

boundary conditions,
223-4
E > V 228

E< V

E > Vn

E <C V

222

E<V

224-5

229

227

Symmetric eigenfunctions, 363-4,


368-9

E< V

222

226-

363-4, 368
and spin angular momentum, 369

Symmetry character of

227

levels,

E> V

682-7

proton-proton scattering, 687-8

eigenfunctions,

classical behavior,
,

neutron-proton scattering,

for non-interacting identical particles,

applications of, 221

E> V

549-52

proton-proton scattering

for interacting identical particles, 369

Step potential, 221-31

eigenfunctions,

scattering,

limits to

cross section, 690


in
in

Stationary state,

energy

224

Superposition of wave functions,

251

transmission coefficient,
Stability,

E< V

411-2, 415

probability density,
,

function,

and electron spin, 335


on hydrogen atoms, 335

tential

Stern-Gerlach experiment, 334-5

in multi-electron atoms,

parity of eigenfunctions for,

E>

E> V

229-3

see Infinite square well po-

infinite,

transmission coefficient,

()

<

E V 225
229-3
time independent Schroedinger equaV 222
tion, E
E > V 227
,

<

<

225

<

E< K

229

reflection coefficient,

E> K

density,

226

probability flux,

justification for treating only electric

for alpha particles, 368

and behavior of atoms, 368, 414


for electrons, 367
and exclusion principle, 367
for m mesons, 703
for tv mesons, 703
for neutrons, 368
for protons, 368

INDEX

727

Symmetry character of
and
369

spin

eigenfunctions,

momentum,

angular

Total relativistic energy, 34


Transformation, of cross sections,
to

Synchronization of clocks, 16-7, 23-4,

27

Galilean, 5

Lorentz, 26

of

Target nucleus, 563

relativistic,

Thermal emission of electrons, 124-5


Thermal radiation, 42
and quantum theory, 41
Thomas-Fermi theory of atoms, 393-6

Thomas frequency, 343


Thomas precession, 340-4, 359
Thomson scattering, 493-8, 505
cross sections, 495, 498, 505
Thomson's model of atoms, 88
and alpha particle scattering, 92-8

Time dependent perturbation

19,

20-1

176-7, 294

acceptable solutions of, 178, 300


classical

37

521
29
Transition group elements, 413-4
classical,

relativistic,

Transition probability, 290


Transition rate, 291
for absorption, atomic,

453-7

for induced emission,

atomic, 457

and old quantum theory, 137, 291


for spontaneous emission, atomic,

453, 458-60, 463


nuclear, 647-8

Transmission

coefficient,

potential,

for barrier

233-4

E>

Time independent Schroedinger equa-

and

energy, classical,

for square well potential, 240, 251


for step potential,
V 229-31

for mesons, 38, 702


tion,

of velocity,

theory,

see Perturbation theory for time


dependent case
dilation,

momentum and
36

Tensor potential, 678-9, 693

Time

CM

LAB, 523-4

wave equation, 194-5,

235

Transmission probability of nucleons


through nuclear surface, 665
Transmission resonances, for barrier
potential,
in

Ramsauer

234-5

552
Transuranic elements, 414

for deuteron, 680

and energy quantization, 178-84


for Fermi gas, 382
for free particle, 212
for infinite square well potential, 253
for multi-electron atom, 396
for one-electron

atom, see Oneelectron-atom time independent Schroedinger equation


real character of, 177

effect,

Triplet eigenfunctions, 370-1

and exchange

force,

quantum numbers
Triplets, of

energy

372-3

for,

372

levels, see

Fine

structure

of spectral

lines,

424

Triplet state, absence of, in


state of helium,

ground

381

of deuteron, 684

for simple harmonic oscillator, 257-8

by numerical integration,
180-2, 268, 276-7

solution,

by

method, 259-60, 301-2


for step potential, E
Va 222
E > Kn 227
series

<

Total angular
30,

momentum, 346-9, 429-

442

Total cross section, 510, 556; see also

Cross section
Total internal reflection, 235-6

Ultraviolet catastrophe, 63

Uncertainty principle, 155, 156, 159


and barrier penetration, 226-7, 239
and compound nucleus reactions, 663
and de Broglie's postulate, 153-5

and energy
tions,

definition in nuclear reac-

667

and indistinguishability, 360


and line widths, 469-70, 472
and meson rest mass, 695-6

INDEX

728
Uncertainty

wave

principle,

and

duality, 160-2,

particle-

Wave

321-2

physical basis of, 157-8

and quantization of angular momen-

general, 170-5
linear combinations of, 186, 192, 285,

tum, 323

and quantization of electromagnetic


radiation, 157-8
and saturation of nuclear forces,
675-7
and spontaneous emission and absorption of mesons, 700
and spreading of wave groups, 15960
and time dependent perturbation
theory, 288-9
and zero point energy, 255-6
Uniform translation, 4
Unsold's theorem, 312

288
mathematical properties of, 141-6,
184-92
normalization of, 187, 190-1, 206
V 224
for step potential, E

<

274-8
eigenfunctions, 276-7
energy levels, 276-7
potential,

Vector diagrams for addition of angular


momenta, 348-9, 43
Velocity of light,

Velocity transformation, classical, 521

27-9

relativistic,

Virtual level, 553


Virtual state, 553

of deuteron, 685-6
of neutron-proton system, 685-6

Wave

equation, electromagnetic, 235,

697

quantum mechanical, see Schroedinger's equation


relativistic,

697

for string, 193

Wave

functions, 165, 185; see also Pilot

waves
complex character
for

of,

for free particle,

170

664
standing wave, 217

compound nucleus

reactions,

traveling wave, 213

groups

of,

163,

142-5, 153-4, 159-60,

215-6

for identical particles, 364

interference of, 148, 151-2, 153-4,

167

superposition of, 143, 144, 148, 167;


see also Linear combinations, of

wave functions
time dependence
Wave number, 112

of,

191-2

Wave-particle duality, see Particle-wave


duality

Waves

Wave
V-bottom

functions, interpretation of, 205,

206, 214, 217, 294, 295, 318-21,

226-7

in string,

192-200

velocity, 141,

142, 144

Widths, of nuclear energy levels, 660-2


and Breit-Wigner formula, 670

measurement of, 658; see also


Alpha decay; Beta decay;

Gamma

decay

and predictions of compound nucleus theory, 666


and random width assumption of
compound nucleus theory, 667
of quantum states, 470-2
and lifetimes, 472
and uncertainty principle, 470, 472
of spectral lines, see Line widths
Wien's law, 50-1
Weizsacker's atomic mass formula, see
Semi-empirical mass formula
Wilson-Sommerfeld quantization rules,
128-9
applications of, 129-31
de Broglie's interpretation

of,

152

X-ray attenuation, 493


attenuation coefficient, 512
attenuation length, 512
cross section for, 509-11

exponential law for, 511-2


X-ray energy levels, 485-6
X-ray excitation of atoms, 417, 483r 4
X-ray continuum spectra, classical
theory of, 490-1

INDEX

729

continuum
489-92

X-ray

origin

spectra,

of,

X-ray

scattering,

Hartree theory and,

400, 501

Planck's constant and, 482

incoherent, 498-9

properties of, 482-3

Klein-Nishina cross sections for,

quantum theory

503-5
measurement of atomic structure by,

X-ray

line

491-2
and Bohr theory,

of,

spectra,

488
and determination of atomic charge,
488-9
and holes, 484-5
and Moseley experiments, 487-9
origin of, 483-7

400, 501

quantum theory of, 501-4


Thomson, 493-8, 505

Thomson

cross sections for, 495, 498,

505
total cross sections for,

510

photoelectric effect and, 507

Yukawa

properties of, 483

quantum numbers

for,

selection rules for, 486


X-ray pair production, 508-9, 513-7

cross sections for, 508-9, 510

X-ray photoelectric
cross

sections

effect,

for,

505-8

507-8, 510

X-rays, diffraction of, 476-80

discovery of, 475-6

measurement of

spectra, 476,

480-1

properties of, 476

X-ray scattering, 493-505

493-501
coherent, 476-80, 499-501

classical theory of,

coherent cross sections for, 499-501

Compton,

8 1-5,

502-5

potential,

696-8

485

Zeeman

effect, 332, 443-7


anomalous, 447
energy level splitting in, 446-7
Lande g factor and, 446
net magnetic moment in, 444-5
norma], 447
selection rules, 447
spectra, 447-8
Zero point energy, 255
and electrons in nuclei, 561-2

for simple harmonic oscillator, 262

and spontaneous emission of radiation, 462


and uncertainty principle, 255-6
ZnS scintillation detector, 92

You might also like