Quantum Physics Berkeley Physics Course Volume 4 PDF
Quantum Physics Berkeley Physics Course Volume 4 PDF
Quantum Physics Berkeley Physics Course Volume 4 PDF
r Temperature: zz 12,000ffK
Eyvind H. Wichmann
Professor of Physics
,
University of California Berkeley
quantum physics
Copyright © 1967, 1971 by Education Develop¬
ment Center, Inc. (successor by merger to
Educational Services Incorporated). All rights
reserved. Printed in the United States of
America. This book, or parts thereof, may not
be reproduced in any form without the written
permission of Education Development Center,
tnc.r Newton> Massachusetts.
04861
1234567890 MWMW 79876543210
Preface to the Berkeley Physics Course
The initial work on this course led Alan M. Fortis to devise a new
elementary physics laboratory, now known as the Berkeley Physics
Laboratory. Because the course emphasizes the principles of
physics, some teachers may feel that it does not deal sufficiently
with experimental physics. The laboratory is rich in important ex¬
periments, and is designed to balance the whole course.
In recent years several new college physics courses have been
planned and developed in the United States. The idea of making a
new course has come to many physicists, aware of developments in
science and engineering, and of the increasing emphasis on science
in elementary schools and in high schools. Our own course was
conceived in a conversation between Philip Morrison, now of the
* Of these Volumes I, II, III, and V have appeared previously. The changes in our
common preface for the present volume reflect some organizational changes under
which this volume was prepared
V
f?i Preface to the Berkeley Physics Course
Eyvind H. Wichmann
October, 1967
Berkeley, California
Acknowledgments
IX
Notes for Teaching and Study
The material is presented in nine chapters. Each chapter is divided
into many short, consecutively numbered sections, each corre¬
sponding roughly to one idea or one step in the train of thought
Equations, figures, and tables within the text are numbered by the
numbers of the sections within which they occur or to which they
refer. Specific references to particular topics in the text are given
in footnotes. General references are given at the end of the chap¬
ters. Tables of physical data are given in the Appendix, as well as
on the insides of the covers of the book.’ Problems for individual
study are given at the end of the chapters. The serious student
should do a very substantial fraction of these problems.
My references are to original papers, other textbooks, and ele¬
mentary review articles of the kind which can be found in Scientific
American, To my student-reader I want to say the following. You
will get a distorted picture of physics if you confine your reading to
textbooks alone, A textbook provides a framework for orderly and
systematic studies, but it cannot possibly mirror the richness and
variety of the intellectual effort in physics. This book, for instance,
is very deficient in its description of experimental procedures. To
encourage you to begin to become acquainted with the literature,
I have included references to papers which report on original re¬
search. I certainly do not expect you to read more than a small
fraction of these papers, but when you encounter a subject which
you find particularly interesting I urge you to go to the library to
look up the original sources. You will probably find other papers
which also interest you, and soon you will become a habitual reader.
Do not try to read papers for which you clearly do not possess the
necessary background. There are many papers, especially on ex¬
periments, which you can read with your present preparation, and
you should select among these. Your instructor can give you
further advice on where to look. The elementary review articles
0 (Note added in proof.) The completed manuscript was submitted to the publisher
at the end of 1967, and the book accordingly does not contain references to more recent
work. It may be stated, however, that nothing has happened in the meantime which
would significantly affect the contents of this book.
xi
xii Notes for Teaching and Study
Chapter 1 Introduction 1
Appendix 405
A General Physical Constants 406
B The Most Stable Elementary Particles 408
C The Chemical Elements 410
Index 413
Chapter 1
Introduction
Sections I to 7 The Scope of Quantum Physics 2
Problems 39
Chapter 1 Introduction
body” as a whole, but we do not try to discuss the motions of all the
elementary constituents of the body. This is a characteristic
feature of classical theories of physics as applied to macroscopic
systems; the finer details of the behavior of the system are ignored
and no attempt is made to consider all aspects of the situation.
In this sense the laws of classical physics are approximate laws of
nature. We should regard them as limiting forms of the more basic
and comprehensive laws of quantum physics.
The classical theories are, in other words, phenomenological
theories. A phenomenological theory attempts to describe and
summarize experimental facts within some limited domain of
physics. It is not intended to describe everything in physics, but
if it is a good phenomenological theory it does describe everything
within the limited domain very accurately. The philosophically
minded reader may want to remark that ultimately every physical
theory is “phenomenological,” and that the difference between a
basic theory and a phenomenological theory is only a question of
degree. As physicists we recognize, however, a clear difference
between the two kinds of theories. Our basic laws of nature are
distinguished by their great generality; we are not aware of any
exceptions to what they state. We regard them as true and exact
and universally valid until there is clear experimental evidence to
the contrary. In contrast to this, the laws contained in a phenom¬
enological theory are recognized not to be of universal validity;
we know that they are valid (i.e., sufficiently accurate) only in some
Albert Einstein. Born 1879 in Ulm, Germany; died
limited domain of physics, and that outside this domain the 1955. Studied at the Institute of Technology (ETH)
phenomenological theory may be completely meaningless. in Zurich. Switzerland. After receiving his diploma in
1900 he held a position as a patent examiner in the
Swiss Patent Office in Bern. During this time he wrote
4 We should, of course, not be contemptuous of phenomeno¬
three famous papers, all of which appeared in the
logical theories. They serve the very useful purpose of summa¬ 1905 issue of Annaien der Physik, dealing with the
rizing our practical knowledge of the various domains of physics. photo electric effect. Brownian motion, and special
There are many instances in physics in which we do believe that relativity. Subsequently he held positions in Bern,
Zurich, and Prague, and as director of the Kaiser Wil¬
we have available a basic theory, but where the complexity of the
helm Institute in Berlin. In 1933 he became a mem¬
phenomena prevents us from making accurate predictions based ber of the Institute for Advanced Study in Princeton,
on “first principles.” In such a case wre try a simplified phenomeno¬ INIJ., settling permanently in the United States. He
logical theory which is partly based directly on the experimental received the Nobel prize in 1921.
Einstein is generally regarded as the most outstand¬
facts, and partly based on some general features of the basic theory.
ing physicist of this century, and as among the greatest
We let, in other words, “the physical systems do some of our theo¬ scientists of all time. He possessed to an extraordi¬
retical work.” There are, furthermore, many instances in physics nary degree the ability to grasp the essence of physical
where the basic theory is missing. Any phenomenological theory phenomena, and no short summary could do jus¬
tice to his numerous, always profound contribu¬
which we can construct (based on some simple model) is then useful
tions on the fundamental problems of physics. His
as a steppingstone in the search for a more comprehensive theory. theory of General Relativity stands out as one of the
When we try to understand an unfamiliar physical phenomenon most remarkable intellectual creations of all time.
it is clearly rational to try the simplest thing first, i,e., to try a theory, (Photograph by courtesy of Physics Today.)
4 Introduction Sec. 1.5
Proton
- •
Proton
This is not the only thing which can happen in a proton-proton
collision: the protons may disappear and a number of entirely new
particles, known as K-mesons and hyperons, may appear instead*
Similarly it can happen, in a violent collision of two electrons,
1 that the final reaction products consist of three electrons and one
positron* (The positron is an elementary particle similar to the
electron, except that it carries the opposite charge*) On the other
Proton hand, if an electron and a positron collide with each other it can
happen that these two particles disappear (we say that they are
Fig, 12A Schematic representation of the creation
annihilated) and we are left with only electromagnetic radiation
of two pi mesons in a high-energy collision of two
protons. One pion carries the charge +e, and the in the form of gamma rays.
other the charge — e, where e is the magnitude of the
electronic charge. The total charge is thus conserved 13 An interesting example of a creation process is the creation
in this event.
of an electron-positron pair when a gamma ray passes through the
Since the two protons remain after the collision and
two new particles appear it is strikingly obvious that electric field in an atom* Material particles can thus be created
naive models of the kind shown in Figs. 9A and UA from electromagnetic radiation* Figure 13A, which is a doud-
cannot apply here: the event cannot be thought of as chamber photograph of so-called cascade showers, “shows” many
a “rearrangement of the elementary constituents (?)
instances of this phenomenon. The explanation (see also Figs*
of the two protons."
13B and C) for what is seen is as follows. If an energetic charged
particle, say an electron or a positron, passes through one of the
horizontal lead plates seen in the photograph it may be very slightly
deflected in the field of one of the atoms in the plate* Such a de¬
flection constitutes accelerated motion, and consequently electro¬
magnetic radiation in the form of an energetic gamma ray is
emitted* (The particle may, of course, be deflected by several
atoms in a single plate, in which case several gamma quanta will
be emitted*) The gamma rays arising in this manner then create
electron-positron pairs in the fields of the atoms which they en¬
counter when they traverse the plates* These charged particles
in turn give rise to more gamma rays as they are deflected in the
plates, and the new gamma rays give rise to new pairs, and so on*
A single energetic charged particle, or a single gamma ray, can thus
give rise to a cascade of gamma rays, electrons and positrons* The
charged particles leave visible tracks in the cloud chamber; these
are the tracks we see in Fig, 13A* The gamma rays are not visible
in the figure*
The cascade shower in the right part of the photograph appears
to have been initiated by a gamma ray, incident from above. The
energy of this gamma ray was probably about 20 BeV* The shower
Fig. 13A Cloud chamber photograph showing cascade showers. Most of the visible tracks are due to electrons and positrons,
which generally move toward bottom of picture. The particle entering at the top right and penetrating three plates before
stopping in the fourth may be a pion. See text for further comments. (Courtesy of Professor W. B, Fretter, Berkeley.)
10 Introduction Sec. 1.14
18 Let us ask the following: What are the forces which keep an
electron together? What fraction of the mass of an electron is of
an infrirtsic nature and what fraction is due to the energy of the
electrostatic field of the electron? To try to deal with these ques¬
tions we assume a not unreasonable model according to which the
electron is a small uniformly charged sphere of radius r. The dif¬
ferent parts of this sphere repel each other electrostatically, and
there must therefore be some other kind of force which keeps the
sphere together. What is the nature of this force?
In Volume II of this series \ we have learned how to compute the
total energy “residing” in an electrostatic field: we integrate
(l/8w)E2 over all space, where E is the local electric field For our
model we obtain the expression W = %(e2/r) for the electrostatic
energy,§ where e is the electronic charge. (The coefficient in front
of the expression e2/r depends on the details of the model: for a
uniformly charged sphere it happens to have the value What is
important here is not the value of this coefficient but the propor¬
tionality of W to the expression e2/r* That W depends in this way
on e and r is immediately obvious on dimensional grounds.) We
can now write the mass of the electron in the form m = me + m*,
where mP = W/c2 is the electromagnetic contribution and m*
is the “intrinsic” part. The problem is: How large is m^? Could
it perhaps be that m = mP, in which case the entire mass would be
of electromagnetic origin? If we make this assumption we can
t It is not only the beginning physics student who has such prejudices, the senior
physicist has them too. Since rigidity of mind appears to increase with age it is
plausible that the senior physicist actually suffers more from his "classical prejudices”
than the beginning student does.
| Berkeley Physics Course, Vol. II, Electricity and Magnetism* Chap. 2, p. 51.
§ This holds for the egs-system of units. In the MKS system we have
Sec. 1J9 Introduction 13
compute the radius r and we find r = 1*7 X 10^13 cm. There are
many experimental facts which suggest that the electron must be
very “small/* and it is therefore comforting that we did obtain
something small. Note that we cannot make t much smaller, unless
we wish to contemplate the possibility that mi is negative.
Since the electron is supposed to be elementary it might appear
particularly tempting to try a model with r - 0, in which case the
electron would be a “point-particle** with no extent and no struc¬
ture. This, however, would lead to an infinite electromagnetic
self-energy W* and to a negative infinite intrinsic mass mi , which
hardly makes sense. (This circumstance, which raises an in¬
surmountable obstacle in the way of the mathematically simple
and attractive model of a point-electron, is referred to in the lit¬
erature as “the difficulty of the infinite self-energy of the elec¬
tron,”)
t A dynamical variable is any variable which characterizes the state of the system;
for example, a position coordinate, a component of momentum, a component of
angular momentum, a component of velocity, the total energy, etc.
Sec, 1,21 Introduction 15
i,APaA
Here Aq is the root-mean-square error in q and Ap is the root-
mean-square error in p, and the inequality thus asserts that the two
variables q and p cannot be known more accurately than that the
product of the “uncertainties” of the two variables is of the order
of Planck's constant.
We note immediately that because of the smallness of Planck’s
constant, ht the uncertainty relation is of no importance in macro-
physics; other sources of errors in q and p always mask the funda¬
mental uncertainty expressed in the inequality (22a). The relation
(22a) therefore does not in any sense contradict our empirical
knowledge of macrophysics even though it does contradict our
classical theories about macroscopic systems.
they make no sense, how then can the relation make any sense?
The answer is as follows. In the quantum-mechanical description
of the behavior of a particle we can introduce certain mathematical
objects q and p which in many respects correspond to the classical
position and momentum variables. These objects, however, are
not identical with the classical variables. The relation (22a) tells us
that if we try to interpret the quantum-mechanical objects q and p
as “position* and “momentum," and thus inteipret the motion in
classical terms, then there is a fundamental limitation on the ac¬
curacy with which “position” and "momentum” can be known.
In other words, the relation tells us that if we try to introduce
classical variables, and try to interpret the motion classically, then
the precision with which these variables can be specified is limited.
tFor some reason the billiard ball has oome to play the role of the prototype of a
classical particle in textbooks on Quantum mechanics. The author, of course, conforms
to this tradition, It may amuse the reader to know that the author has never played
billiards and has never held a billiard ball in his hand His knowledge of the alleged
properties of billiard balls is, therefore, book knowledge, derived from texts on quantum
mechanics.
Sec. 1.28 Introduction 19
occupied the physicists of the time, but we single them out because
they illustrate the dilemma of classical physics in a particularly
clear-cut way.
The reader should realize that our discussion is extremely de¬
ficient as a historical account: we could not possibly hope to do
justice to the very interesting development of quantum physics in a
few pages. We are looking at the situation at the beginning of this
century in retrospect, and it is then easy to see that these three
problems were key problems. However, if we examine the publica¬
tions for the year 1900 in Annalen der Physik (which was one of
the leading journals in physics at the time) we find that the majority
of physicists were concerned with very different things. The ability
to distinguish the truly significant from the insignificant is a rare
ability indeed (at any time), and we have every reason to admire the
remarkable insight and imagination of the early pioneers of quantum
physics.
e F
A#p NqMp
and this constant is therefore not independent of the constants
mentioned earlier*
We should also note that given precise values of e/m and e/Mp
we can find a precise value for
Afp _ e/m
(28d)
m e/Mp
PV = iV0£kiQ (33a)
where P is the pressure, V the volume of the container, and £kin
Fig* 33A The relation PV = JAofiyn can be easily
the average kinetic energy per (monatomic) molecule. understood. Consider a container of volume V in
The absolute temperature is so defined that, within this model, which there are \:n molecules. Let us first assume
it simply expresses the average kinetic energy by £kin = where that all the molecules move to the right, with velocity r.
the constant of proportionality k is known as Boltzmann's constant. The number of molecules colliding with a unit area of
We can therefore write (33a) in the form the wall per unit time is Each molecule
transfers an amount of momentum 2mc to the wall.
FV=Sr0kT = RT (33b) The pressure J># is equal to the total momentum trans-
fer per unit area per unit time, and hence we have
where the constant fl = \rok is the universal gas constant. It is />' = 2muB(ArV'V) = 4 ftmOVo/V).
an experimental fact that this law holds accurately for all sufficiently in reality the direction of motion is random, and the
true pressure P is related to the pressure PT computed
rare gases: i.e., for any real gas the law holds better the less dense
above by P = which leads to Eq. (33a). (We can
the gas is. We can exploit this fact to construct a gas thermometer understand the factor £ if we imagine that the mole*
calibrated to show absolute temperature, cules move in six standard directions: in both direc¬
tions along three perpendicular axes. Only one-sixth
34 The universal gas constant has the value of the molecules will then contribute to the pressure
on the right wall.)
R = N0k =s 8.3 X 107 erg (QK)-i (mol)-i
= L99 cal (°K)~1 (mol)-1 (34®)
It is a macroscopic constant which can be readily measured on
the basis of the relation (33b),
Boltzmann's constant, k = fi/A'o, is the gas constant per mole¬
cule* It can be found provided No is known, and it has the value
a curve which goes to zero for very long, and for very short wave¬
lengths; in general the curve will have a single maximum at some
wavelength \max> which depends on the temperature* The loca¬
tion of this maximum, and the total amount of radiation emitted,
is very roughly the same for all material surfaces. Instead of study¬
ing the radiation from a material surface one may observe the radia¬
tion emerging from a small hole in the wall of a dosed material
surface kept at a fixed temperature. In this kind of measurement
we thus have an enclosure, or "oven,” made of any suitable refrac¬
tory material, with a small hole in the side (i,e,, a hole small com¬
pared with the linear dimensions of the cavity). We direct our
instruments at the hole and thereby measure the radiant energy
emerging from the interior of the enclosure. It was found in such
measurements that
(i) A graph (see Fig* 35A), of the intensity of radiation from the
hole versus wavelength is a smooth curve falling to zero for long,
0 1 2 3 4 5 6 as well as for short wavelengths, and with a maximum at a wave¬
x 10 “4 cm length Amiut which depends on the temperature T of the walls in
Wavelength a very simple manner, namely
WT = C0 = 0.2898 cm *K (35a)
Fig. 3SA Graphs showing power emitted by a black-
body radiator per unit area per unit wavelength inter¬ (ii) The spectral distribution of the emitted radiation, i*e., the
val f tor four different temperatures. The total power
shape of the curve mentioned in (i), is independent of the shape
emitted is proportional to the areas under the curves;
it is proportional to the fourth power of the absolute of the cavity, and independent of the material of which the walls
temperature. Note how the location of the maximum are made. The constant Co in equation (35a) which expresses
depends on the temperature; the precise relationship Hfens displacement law, is thus a universal constant which de¬
is expressed by WienTs law. scribes a remarkable property of cavities in general
(iii) The intensity of the radiation emerging from the hole is
always larger, at every wavelength, than the corresponding intensity
of emission from a material surface kept at the same temperature
as the walls of the cavity; the order of magnitude of the intensity,
however, is the same.
^ X kT = X, = (37a)
c c
where c is the velocity of light, k is Boltzmann's constant, and X%
is a new constant* Since the left-hand side of (37a) has the phys¬
ical dimension [time] X [energy] = [action] the constant Xi is a
quantity of action* How can we derive a theoretical expression for
Xj? How are we to produce a quantity having the physical dimen¬
sion of action from the constants of nature available? This is cer¬
tainly a dilemma because it is hard to see how the constants m,
Mr and e could possibly enter into the expression for Xi* The
physical situation seems very clear-cut; the radiant energy inside
the cavity is in thermal equilibrium with the walls* The radiation
emitted from the cavity is, however, independent of the size and
shape of the cavity and also independent of the material in the
walls; how then could constants like m and e, which refer to prop¬
erties of the walls, be relevant? Our suspicion that Xi cannot be
derived from the remaining constants seems quite justified and, as
a matter of fact, the relation (37a) cannot be understood on the
26 Introduction Sec. L38
fFor latet tests of Planck's law* see H. Rubens and G, Michel, "Prufung der
Flanckschen Strahhmgsformel,” Physikalische 7>eitschTift 22,569 (1921),
28 Introduction Sec. 1.41
Fig. 44B A graph taken from Millikan's paper [R. A. v will thus be a straight line, as shown in Fig. 44B taken from
Millikan, Physical Review 7, 355 (1916)] showing the Millikan's paper. From the slope of this line we can find the
linear relation between the critical retarding potential constant h/e> and from its intercept with the Vo-axis we can find
Vo and the frequency of the light, for a photosensitive
surface of sodium. As we see, Millikan presented his
the material constant W/e.
computation of Planck's constant on the basis of his This is a conceptually simple and clear-cut experiment, but to
curve on the graph. (Courtesy of The Physical Ra- obtain accurate and reproducible results considerable care is re¬
view.) quired.
45 Let us consider the relation (44a) numerically. With h = 6,63
X 10"27 erg sec = 6.63 X 10-34 joule sec, and e = 1.60 X 10_ie
coul, we obtain h/e = 4.14 x 10“15 volt sec. For visible light the
wavelength lies in the range 4000 A to 7000 A, where 1 A(ngstr6m)
= 10“s cm.t This corresponds to the frequency range (4.3 to 7.5)
X 1014 sec-1. Blue light corresponds to a frequency of about
7 x 1014 sec-1, and for this case we obtain {h/e)v ^ 2.8 volts. For
light in the visible or near-ultraviolet region the retarding potential
will thus be of the order of magnitude of one volt, It is an experi¬
mental fact that the material constant W/e is also typically of this
f It may be mentioned here that Einstein did not use the term photon for the electro-
magnetic quanta In his paper; that term was introduced much later.
32 Introduction Sec. 1.47
The fact that we can obtain a new solution from a given solution
through a scaling of the kind described is really an extension of
Kepler's third law. Applied to the particular case of a single eleo-
tron moving around a fixed nucleus, our argument tells us that for
two elliptical orbits of the same eccentricity the ratio of the squares
of the periods is proportional to the ratio of the cubes of the semi¬
major axes.
Since we can give any value we please to q we actually have a
whole family of solutions, and there is no reason why we should
prefer any particular one of these; there is, in other words, no prin¬
ciple which tells us why a particular “size" for the atom should be
preferred. One might, of course, argue that the actual size of an
atom is determined by “accident" but such an argument would
hardly stand up. How is it possible that this “accident" always
leads to the same size of orbit for a given species of atoms? Why
don't we have a continuous distribution in size for, say, the hydro¬
gen atom?
those solutions of the equations of motion for which the total angular
momentum of the atom is a definite multiple of h are realized in
nature. If we accept this principle we must give up our scaling
argument since under the transformations described by Eq. (49a)
the angular momentum will be scaled by the factor q, which now is
not allowed. This means that there will be preferred solutions, and
hence we now have a principle available for the determination of
the size of an atom.
In 1913 Niels Bohr presented a theory of the hydrogen atom along
these lines, f In the simplest version of this theory a single electron
moves in a circular orbit of radius do around a proton. The orbit is
determined by the equation of motion
m (51a)
J = mvao = — (51b)
27T
1) It will be assumed in this book that the reader has some familiarity with the most
basic facts of quantum physics, at the level at which these topics are discussed in a high
school physics text such as Pfiysfcs. by the Physical Science Study Committee (D. C.
Heath and Company* Boston, 1980). (Part IV* in particular.)
In the cases in which the above assumption does not correspond to the facts some
supplementary reading is called for. Any library will have a variety of semi-popular
accounts of "atomic physics," some of which are had and some of which are good.
Taken with a grain of salt such a book might serve the purpose. Articles in magazines
such as Scientific American can be very useful, and are jfrongty recommended. The
reading of such articles will probably whet the reader's appetite and lead to further
self-study and reading, Whenecer the reader & background permits it he should try to
reed original papers, but highly technical, or mathematically complicated accounts are
best avoided at this stage.
2) The reader may be interested in reading selected portions of some text on quantum
physics in which a more complete account of experiments is given than in this book.
Among the many such texts we mention the following:
a) E. Grimsehl and R> Tomaschek: A Textbook of Physics, vol V, Physics of the
Atom (Blackie and Son Limited* London, 1945).
b) G> P. Harnwett and J. J. Livingood: Experimental Atomic Physics (McGraw-Hill
Book Company* New York* 1933).
3) The following books are historical surveys of the development of modem physics:
a) M. Jammer: The Conceptual Development of Quantum Mechanics (McGraw-Hill
Book Company, New York* 1968). A magnificent piece of work, which, however,
requires a substantial knowledge of quantum mechanics for a full appreciation. The
beginning of the book* which deals with the early history, can be read with a modest
background. The many carefully compiled references to original papers is a valuable
feature.
b) E, Whittaker: A History of the Theories of Aether and Electricity, vols. I and II
(Harper Torchbooks* Harper and Brothers* New York 1960), The second volume dit-
cusses the development of quantum mechanics. These books (as well as Jammer's)
discuss the interesting false leads as well: the theories once taken seriously but now
forgotten.
Problems Introduction 39
Problems
1 (a) First consider, and then describe very briefly, the kind of reasoning
and the kinds of measurements that have led to definite assignments of atomie
and molecular weights.
(b) In 1815 William Frout suggested that all elements might be combina¬
tions of hydrogen, which would thus be the primordial material from which
everything else is made. What might have ted him to such a hypothesis,
and why was his suggestion rejected during the nineteenth century?
(a) Show that the above law of disintegration results if we assume that
each atom disintegrates independently of the other atoms, and also assume
that the probability that an atom* which has survived until the time tt will
disintegrate dining the time interval (f, t + Af) is independent of
(b) In the decay of radium atoms an alpha particle is emitted. If this
alpha particle strikes a zinc sulfide screen, a flash of light (called a scintil¬
lation) will mark the point of impact. It is thus possible to count directly
the number of alpha particles emitted per second from I gram of radium,
and this number was determined by Hess and Lawson as 3.72 X 1010* The
atomic weight of radium is 226. Use these data to find the half-life of
radium. (Measurements with radioactive substances have been used to give
independent estimates of Avogadro’s number. In the above problem the
procedure is reversed, and we instead determine the half-life of radium.)
3 The moving parts of a wrist watch are pretty “small.” Making reason¬
able estimates of the magnitudes of the physical parameters which charac¬
terize a ‘"typical" wrist watch, show, however, on the basis of the general
criterion given in Sec, 20, that quantum mechanics is totally irrelevant to
the art of watchmaking.
II to IS Energy 49
Molecular Physics gg
of Nature 78
Problems 85
Chapter 2 Magnitudes of Physical Quantities
in Quantum Physics
1 One of the aims of this chapter is to give the reader a feeling for
the orders of magnitude of various physical quantities in the realm
of quantum physics* Many of the important physical quantities,
such as the electronic charge, the electronic mass, Planck’s constant,
etc., have numerical magnitudes when expressed in our familiar
macroscopic units which are inconvenient and untransparent be¬
cause they are so very small. It is difficult to grasp directly what it
means that Planck’s constant has the value h = 6*6 x 10*"27 erg
sec. It is therefore important that we study in detail how these
various constants occur in physics, and what their numerical values
actually mean* t
Every domain of physics has what we may call natural units for
the physical quantities involved, which means that when we express
any physical quantity in terms of these natural units then the nu¬
merical values are reasonable in the sense that their significance can
be readily grasped* The numerical values may vary over the range
10-6 to 106, but we will not encounter numbers such as 10~27.
Our familiar macroscopic units (in the MKS system) are particularly
suited for our everyday experiences with physical phenomena, and
they are based on readily available macroscopic standards. We
note that they are really “human units"'; units such as the meter and
the kilogram, and the second unmistakably refer to human char¬
acteristics. The so-called “scientific,” or cgs system of units is more
appropriate for small animals, such as cockroaches. We shall try to
free our discussion from the arbitrary standards of the human sys¬
tem, or the cockroach system, and try to identify the natural units
in the various domains of quantum physics.
standards are arbitrary and “accidental.” This does not mean that TABLE 2A Some Physical Constants
the list is unimportant. Once we have selected our macroscopic
Planck*s constant:
standards we naturally want to relate the basic parameters of
h = = (6.62559 ± 0.00015)
quantum physics to these, and that is the purpose of the list.
X 10-27 erg sec
We have also quoted the estimated errors for the constants to # = h/2* - (1.05449 ± 0.00003)
give the reader a feeling for how accurately they are known at pres¬ X 10“27 erg sec
ent. At the level of this book the reader wiD almost never have Velocity of light:
c = (2.997925 ± 0.000001) x 10* u cm sec-1
occasion to carry out any computations to greater accuracy than
Electronic charge:
can be obtained with the slide rule, which accuracy is about 0.2 per
e = (4.80298 ± 0.00006) X 10-«> esu
cent per multiplication or division. The reader should also learn to
= (1.60210 ± 0.00002) X 10-lflcoid
make simple estimates which may vary in accuracy from 10 per cent E/ecfron mass:
to merely an estimate of an order of magnitude. On the inside of m - (9.10908 ± 0.00013) x 10“^ gm
the front cover of this book the reader will find a table of very rough
Proton mass:
values of the most important constants, and this table should prop¬ Mp = (1.67252 ± 0.00003) X I0~24gm
erly be committed to memory. Other, more detailed tables of
Avogadro's number:
physical data can be found in the Appendix. ATo = (6.02252 dr 0.00009) X 1023{moi)-i
Boltzmann's constant:
3 The definition of Avogadro's number requires discussion. k - (1.38054 ± 0.00006) x 10-™ erg (^K)-1
When chemists tabulated atomic weights in the past they employed
a scale in which naturally occurring oxygen was assigned by defini¬
tion the atomic weight 16 exactly. The atomic weight of hydrogen,
for instance, was thus defined as
The word “atom” is embellished with quotation marks because Element z Atomic weight
the weight of the “atom” always refers to the element as it occurs
H 1 1.00797
in nature. The atomic weights, as defined in Eq, (3a), are deter¬
He 2 4.0026
mined by the chemist through careful tveighing operations; he wilt Li 3 6.939
for instance, determine the amount, in grams, of naturally occurring Be 4 9.0122
hydrogen which will combine with 16 grams of naturally occurring B 5 10.811
oxygen to form water with nothing left over. The resulting number C 6 12.01115
N 7 14.0067
divided by two is the atomic weight of hydrogen.
0 3 15.9994
The atomic weights determined by the chemists in this manner F 9 18.9984
are called atomic weights on the chemical scale* Many of the ele¬ Ne 10 20.183
ments have atomic weights which are close to integers, but there are Na 11 22.9893
also notable exceptions; the atomic weight of chlorine, for instance, Mg 12 24.312
Al 13 26.9815
is 35.5.
Si 14 28.086
R 15 30.9738
4 As the reader knows, the mass of an atom is mainly concentrated S 16 32.064
in the nucleus. Nuclei are built of protons and neutrons. The Cl 17 35.453
A 18 39.948
number of protons plus the number of neutrons is known as the
mass number of the nucleus. This integer is commonly denoted t See Table C in Appendix for a complete listing.
46 Magnitudes of Physical Quantities Sec. 2.5
TABLE 4A Naturally Occurring Isotopes by A* The number of protons is called the atomic number of the
of Selected Light Elements nucleus. It is denoted by Z, and the nuclear charge is thus eZ,
where e is the elementary charge. The chemical properties of an
Iso¬ Natural
Ele¬ Atomic atom are determined almost exclusively by the nuclear charge, and
z tope
mass
abundance
Z is thus a characteristic of the chemical element. It is found that
ment
A percent
there are many instances of families of nuclei with the same charge,
H 1 1 1,007825 99,985 but with different mass numbers, and these different nuclei are
2 2.01410 0.015 referred to as different isotopes of the element. Isotopes differ in
He 2 3 3.01603 , 0.00013 the number of neutrons. The mass of the proton is almost equal
4 4.00260 100 to the mass of the neutron, and the masses of all nuclei are very
Li 3 6 6.01513 7.42 closely proportional to the integral mass number A. The explana¬
7 7.01601 92.58 tion for the occurrence of markedly non-integral atomic weights is
Be 4 9 i 9,01219 100 that many naturally occurring chemical elements are mixtures of
B 5 10 10.01294 19.6 two or more different isotopes, in which case the “atomic weight**
11 11.00931 80.4 of the element as measured by the chemists is an average of the
C 6 12 12.000000 98.89
more fundamental atomic weights of the different isotopes, t It is
13 13.00335 Ml an experimental fact that the relative abundances of the different
N 7 14 14.00307 99.63
isotopes occurring in an isotopic mixture of an element are very
15 15.00011 0.37 nearly the same all over the surface of the earth. Furthermore the
0 8 16 15.99491 99.759
different isotopes have, for all practical purposes, identical chemical
17 ! 16.99914 0.037 properties and are therefore almost impossible to separate from
13 17.99916 0.204 each other by “chemical * means. If this were not the case, the
F 9 19 18.99840 100 chemists* tables of atomic weights would be worthless.
■ p 4
t That a chemical element can consist of different isotopes was firmly established
by J. J. Thomson. [J. J. Thomson, “Rays of Positive Electricity," Proceedings of th&
Royal Society (London, Series A) SO, 1 (1913).]
Sec* 2.6 Magnitudes of Physical Quantities 47
w = 2at (8a)
It follows that
X_
ft
2tt
Energy
m = [mass] = [time]-1
n
me = [momentum] (13a)
The reader should check that the physical dimensions (with the
conventions of the cgs system) are correctly stated* All these quan¬
tities “hang together1* through the constants h and c. Based on
the above relations an energy can be associated with a mass or a
frequency or an inverse length, and the magnitude of the energy
can be expressed in terms of the magnitudes of the associated
quantities*
= (2.41804 ± 0.00002)
(energy)
X 1014 (cycles/sec)/(eV) (14b)
(wave number)
= (8.06573 ± 0.00008) x 10^ (cm-1)/(eV) (14c)
(energy)
The table on the inside of the back cover of this book is partly
based on the above conversion factors. Each horizontal row shows
a set ,of corresponding quantities associated with the quantity in the
first column. The second and the third columns give the energy E
in eV and ergs. The seventh column gives the corresponding mass
E/c2, in atomic mass units, amu; the eighth column the correspond¬
ing frequency E/h, in cycles/sec; and the ninth column gives the
wave number E/(hc)} in cm^1. The tenth column gives the asso¬
ciated wavelength (hc)/E, in Angstrom-units, and this is the only
quantity in the table which is not directly proportional to E.
52 Magnitudes of Physical Quantities Sec* 2.15
I keal = 1000 cal, 1 cal - 4.186 joule = 4,186 X 107 erg (15a)
= Xo = 23,050 (cal)/(eV)
In the table on the inside of the back cover the fourth and fifth
columns give bulk energies in erg/mol and cal/mol.
The sixth column in the table on the inside of the back cover
gives the equivalent temperature, in Kelvin degrees.
tfl ”7 =7 2 1—108
■2
-
DK cm
magneton separated by 1 A
Precession frequency of proton in 1000 gauss
Magnitudes Characteristic of
Atomic and Molecular Physics
m 1 (20a)
1836
(20b)
r
_ e2f{Hfmc) _ e2
a (me'2} He
The reader should note, by inspection of Eq* (21a), that the mass
of the electron does not enter into the expression for a. Conse¬
quently a is the coupling constant which describes the coupling of
any elementary particle carrying the elementary charge e to the
electromagnetic field*
In Table 21A we list some important quantities which we can
form from m, f£, c, and et and we give the names under which these
quantities are known.
The velocity is thus 137 times smaller than the natural unit,
which is the velocity of light c. This is the a posteriori justification
for our nonrelativistic discussion of this problem.
The kinetic energy, and the potential energy, Epot, are
given by
Arnold Sommerfeld. Born 1868, in Kbnigsberg, Ger¬
Ekin = imv2 = £m(ac)2 = R*> (24b) many (now Kaliningrad, USSR); died 1951. For many
years professor of physics at the University of Munich.
Epot =—— = £ — £jdn = —2R„ = -2Ekia (24c) Sommerfeld made important contributions in the
Qq development of quantum physics, and in particular in
early atomic theory. He refined Bohr’s theory in two
On the basis of these considerations we can say that the hydro- ways: to include elliptical orbits, and to take special
relativity into account. His relativistic theory of the
hydrogen atom introduced the fine structure constant
f The subscript oc in 7?* and refers to the model in which the proton is infinitely into physics. (Photograph by courtesy of Professor
heavy and stays fixed. The actual ionization energy is slightly smaller. L B. Loeb, Berkeley.)
60 Magnitudes of Physical Quantities Sea 2,25
and
_ e2 _ e4m
■ia2ntc2 (25b)
2 " 2<io “ 2
the arguments seem to work out particularly well in all the cases in
which either the experimental results or the results of a more com¬
plete theory are known.
In defense of what has been done we say the following: (i) we
wish to form a picture of the orders of magnitude in atomic and
molecular physics. Instead of just telling the reader that the ioniza¬
tion energy of hydrogen is 13.6 eV we should try to relate this 13.6
eV to expressions formed from the basic constants. It is nice to
know that 13.6 eV equals a2 me2/2, and it is nice to know that
0.53 A equals (1/a) (ft/mc). Our discussion of quantum electro¬
Energy Wave number Wavelength
dynamics and its relevance to the hydrogen atom gives us at least eV cm'1 A
some understanding of how it all might hang together. The author
would certainly not have presented these ideas if they did not have
H 40000 2500
their counterparts in the precise theory. Our “derivations” are
therefore at least useful as mnemotechnic devices.
(ii) The Bohr theory is admittedly wrong. On the other hand the h- 35000
I— 3000
reader undoubtedly knows that it was successful in some instances, r 4-0
although it faded badly in other cases. In a vague sense the theory
\- 30000
therefore has some elements of truth in it. It introduces Planck’s f- 3300 Fraunhofer
Uqqi
constant into physics and thereby introduces a relation between >--* Ultraviolet
position and momentum which does not occur at all in a purely _/ * <Ga)
\- 23000 — 'if (Ca)
~h (h) Violet
classical theory: something like rp — #. We can take the point of
—-G (Ca}
view that our derivation based on the Bohr theory was really in Blue
- 4500
essence an experimentation with a relation of this kind: rp
Later we will play with this relation in a different way, and we will h 2.5 I— 20000
1' - y (H) j 1
— 5000
. Ch (Mg)
discuss a method of estimating the size and ionization energy of E‘ <Fe)
, Green
Maximum
the hydrogen atom on the basis of the uncertainty relation. At the — 5500
visibility
Yellow
same time we will gain a much better understanding of why the — 6000
-D <N»)
2.0 Orange
hydrogen atom does not collapse. -
— 10 6500 — t; <H)
(hi) The dimensional argument in Sec. 25 would be much more Bed
(27a)
and a moment’s reflection shows us that the discussion for the case
of the hydrogen atom still applies, provided we replace the fine
structure constant or by aZ, In other words, the first electron will
he bound at an energy
e, = -Z*Rk = — Z2 (13.6 eV) (27b)
ZttHc
zz 1000 GO (29a)
a2mc2 a
about 0.75 A. These values are quite typical for molecules in TABLE 3€A Characteristics of Somewhat
general; molecular binding energies are of the order of 1-10 eV, Randomly Selected Diatomic Molecules
and the intemuclear separations are of the order of one Angstrom
Distance
unit, i.e., of the order of 10-8 cm. Dissociation
between
The same “mechanism” which holds a molecule together also Molecule Energy
holds solids together, and the typical separation between two Nuclei
eV
neighboring atoms in a solid is also of the order of 1 A. A
There are some marked exceptions among the very light nuclei,
and there is also a slight systematic decrease of the average binding
energy as the mass number A increases, as we see in Fig* 33a,
t The neutron was discovered by Chadwick in 1932, [J, Chadwick, The Existence
of a Neutron," Proceedings of the Royal Society (London), sex. A, 136, 692 (1932),]
Sec. 2.34 Magnitudes of Physical Quantities 67
34 The reader should note that the mass values listed in most
tables of “nuclear” masses actually refer to the masses of the cor¬
responding neutral atoms* If Atf(A,Z) is the mass of a nucleus, and
M(A,Z) the mass of the corresponding atom, we have
fF< W, Aston, "Isotopes and Atomic Weights,7' Nature 105, 617 (1920). Also
F. W. Aston, Mass Spectra and Isotopes (Edward Arnold and Company, London. 1942).
68 Magnitudes of Physical Quantities Sec. 2.35
Some nuclei are stable, whereas other nuclei are unstable and
decay through the emission of particles or gamma rays. The com¬
monly occurring nuclei are either absolutely stable or else have
very long lifetimes; if this were not so they would have decayed
at some early stage in the history of the earth and would no longer
be present. Nuclei formed in nuclear reactions may have very
short lifetimes, of the order of a small fraction of a second. When
the lifetime is very short we often speak of an excited state of a
nucleus, especially if the decay takes place through the emission
of a gamma ray, in which case A and Z remain unchanged.
At present about 900 nuclei are known, of which about 280 are
stable. If we plot these nuclei on a (ZA)-plane, then the points
representing the individual nuclei tend to cluster along a smooth
curve, in accordance with what was said before. (See Fig. 35a.)
The more removed a nucleus is from the "central curve" the more
unstable it tends to be.
(The unit fermi = 10~13 cm, named in honor of Enrico Fermi, is Fig. 34C Mass spectrum recorded with the appara¬
often employed as a unit of length in elementary particle physics.) tus shown in Figs. 34A-B, for xenon extracted from a
stony meteorite. Graph is taken from J. H. Reynolds,
Since the volume of the nucleus is proportional to r3, and hence,
"Determination of the age of the elements/' Physical
by formula (36a), to the nucleon number A, we conclude that the Review Letters 4,8 (I960). The short horizontal bars
density of nuclear matter in the different nuclei is approximately show the isotopic abundances for terrestial samples
constant. of xenon. As we can see, the meteorite sample is
The sizes of the nuclei, as summarized by the fonnula (36a), richer in the isotope Xelzs. Note that the graph is
drawn in two different vertical scales. (Courtesy of
have been determined from a variety of experiments.! The most Physical Review Letters,)
straightforward method consists in the measurement of the effec¬
tive cross-sectional area which a nucleus presents to a beam of
very-high-energy particles in a scattering experiment.
37 Let us now try to say something about the nature of the forces
which hold a nucleus together. All our experimental evidence
says that
(i) The nuclear force is not electromagnetic in nature; compared
with the electromagnetic forces the nuclear forces are much
stronger.
(ii) Hie nuclear force is of short range; when the separation
.
Fig 35A The stable and almost stable nudei. All
known nudei having half-fives greater than 5 x 10™
years are shown. This somewhat arbitrary lower
limit on half-life was chosen because it is about ten
times the estimated age of the solar system, and the
included nuclei are thus long-lived even on a geological
time scale. In this graph (consisting of two parts)
the ordinate is the neutron number (A — Z) and the
abscissa is the atomic number Z, It is at once appar¬
ent that the nuclei cluster about a smooth curve. For
the light nudei the number of protons is about equal
to the number of neutrons, but as the atomic number
increases the neutron number increases faster.
The steplike appearance of the pattern of nuclei
derives from the fact that the stability of a nucleus
depends on whether the number of protons and the
number of neutrons are even or odd, as follows: even-
even nudei are most stable, even-odd (and odd-even)
nudei are less stable, and odd-odd nuclei are most
unstable. The reader should study the graph care¬
fully to see how these rules manifest themselves.
There are very few odd-odd nudei on the graph. For
some neutron numbers, and some proton numbers
there are “gaps” corresponding to the absence of
stable nuclei. Note that these always occur when the
neutron or proton number is odd.
Sec. 2.38 Magnitudes of Physical Quantities 71
provided that the separation r is larger than 10“13 cm,t The con¬
stant b is a measure of the range of the force; it has the value
b = 1-4 X 10-13 cm. The constant C expresses the strength of
the force. The nature of the force at distances smaller than
10-13 cm is much more complicated and very poorly understood
at present.
We emphasize that the potential function l/(r) does not describe
the interaction between two nucleons precisely, but it does repre¬
sent the most important feature of this interaction which is that
the potential falls off exponentially with distance,
Let us see what this really means. At the distance r = b we have
U(b) = C/e. (This constant is roughly of the order of 10 MeV,)
At the distance r = 10b — 1,4 X 10~12 cm the potential equals
U(10b) = (0.1C) exp ( —10) 5 X 10_6C. At the distance
r — 100b = 1,4 X 10“11 cm the potential equals U(100b) —
(0*01C) X exp ( — 100) ^ 10*45C. We can oonclude from this
numerical exercise that when the separation between the two
nudeons exceeds 10~n cm, then the nuclear force is utterly negli¬
gible, For all practical purposes there is no nuclear force beyond
the distance quoted.
The reader should think carefully about this. At first sight the
expression (38a) might resemble the Coulomb potential. However,
the exponential factor makes all the difference in the world. Our
numerical exercise was intended to impress the reader with this
fact.
The specific nuclear force between nuclei in molecules and solids
is thus nonexistent for all practical purposes, and this is the kind
of situation in which the electromagnetic forces have a chance to
play a dominant role. At smaU distances, r — 10~13 cm, the specific
nuclear force is considerably stronger than the electromagnetic
forces, and the latter are relegated to play a secondary role. That
this is the case is immediately obvious from the fact that nuclei
exist. The electrostatic forces of repulsion try to disperse the
charged partides in a nucleus, but the nuclear forces try to keep
them together, and the nuclear forces win; they are stronger,
t In Chap. 9 we shall present a theoretical explanation for this form of the potential
function.
Sec♦ 2,40 Magnitudes of Physical Quantities 73
M*Gfr* M 2G
= 8.1 X 10~37 (40a)
eVr2 e2
R^ = ^ct2mc2 (45a)
as we have written it. It was rather written as
_ e*m
R TO (45b)
~ 2W
and for this reason a was not called the “gross structure constant,”
which would be more appropriate. The expression (45a) must be
regarded as a "better” expression for as it gives us a much better
insight into the nature of atoms. As we have explained, a is the
fundamental coupling constant between the electromagnetic field
and the elementary charge. Atoms are "loosely bound structures,”
with “slowly” moving electrons, because ot is small compared to
unity. For this reason a nonrelativistic theory leads to a good
approximation. The relativistic corrections are of the order of
(vfc)2t hence of order a2.
8 =
H Mp 1836
= 5 902 xl0—3,
(MpC2)
= tFs
Mpc1
= 135 X 10-3
t We are here stating a reasonable inference. What King actually did was to estab¬
lish that the hydrogen molecule and the helium atom aTe neutral to the stated accuracy,
[J. G. King, “Search for a small charge carried hy molecules," Physteal Review Letters
5, 562 (1900)0
Sec. 2.50 Magnitudes of Physical Quantities 81
52 For the meter there exist, or rather have existed, two stand¬
ards. In the old standard the meter is the distance between two
notches on a certain metal rod kept in Paris; we shall denote this
meter by (m)P, the “Paris meter.” The new standard is “atomic”
in nature, and the corresponding meter, which we denote (m)fl,
the “atomic meter,” is defined as a certain multiple of the wave¬
length of a definite orange line in the spectrum of krypton, the
multiple being n2 = 1,650,763*73, by international agreement*
The wavelength of the orange krypton line is something which
we can compute in principle (but not in practice), and we can write
this wavelength in the form
\ = can-* ^ (52a)
w
where c2 is a constant which depends only very weakly on a and /?*
In a first approximation it is a purely numerical constant, and if
we should master the mathematics of atomic physics we could find
this number*
The atomic meter can thus be written
*o = c3a (53b)
Even if we cannot compute the quantities ci, c2, 03, and c4 accu¬
rately in practice, we know that they are, to a first approximation,
just purely numerical parameters, independent of a and j8. If we
could really compute these numbers it would mean that we could
compute the value of the velocity of light, in units of (m)fl/(sec)a.
Our theoretical expressions for the macroscopic standards enable
us to deal with the following question: What would the world be
like if our constants of nature were slightly different? This means:
what would the world look like if the two empirical constants a
and were slightly different? This is an interesting question be-
Sec* 2*57 Magnitudes of Physical Quantities 85
*
References for Further Study g 10-4 Red blood cell
CA
Problems
Atom
1 In 1903, P. Curie and Laborde studied the heat emission of radium*
They found that 1 gram of pure radium (we now know that the isotope
86 Chap. 2 Problems
asRa226 is involved) emits about 100 cal/hr. From this and from the known
half-life compute the approximate energy in MeV with which the emitted
alpha-particles emerge. In the experiment of Curie and Laborde these par-
tides were captured within the source and the calorimeter, and their kinetic
energy therefore was converted to heat energy. (The half-life is 1622 years.)
2 (a) The radium nucleus has a posttfoe mass defect, but nevertheless
it is unstable and decays. How is this possible? Is not the necessary and
sufficient condition for stability that the mass defect be positive? Explain
in detail.
(fc) The radium isotope referred to above is asRa226; this is the isotope
discovered by P. and M. Curie. It decays through the emission of alpha-
particles, which are nothing but helium nuclei, 2He4.
We might be led to believe that only stable nuclei or long-lived isotopes
occur naturally since any short-lived isotopes would have decayed in geo¬
logical times. Now, measured against the age of the earth the half-life of
1622 years is not particularly large; it is, rather, quite small How do you
then explain the natural occurrence of radium?
4 The sun radiates energy from its surface at the rate of 3.86 X I02fl
watts. Before the development of nuclear physics it was something of a
problem to explain where this huge amount of energy comes from. Let us
try to make some simple estimates.
The sun is believed to be at least 4 billion years old. The mass of the sun
is 1.98 x lO^o kg.
(a) What fraction of the sun's mass has to be converted into radiant energy
per year in order to account for the power radiated? You will find that this
number is quite consistent with the idea that the sun has not changed much
during its lifetime, i.e., during the last 4 billion years.
(b) Rule out chemical reactions as the source of energy.
(c) Do you know of any nuclear process which might take place in the
interior of the sun, and which might provide us with the explanation of
where the energy comes from? Consult some introductory astronomy book,
Problems Magnitudes of Physical Quantities 87
and convince yourself through some simple estimates that your explanation
is plausible, or at least not in violent contradiction with the facts.
5 We have said that the density of nuclear matter, i.e,, the density of the
“substance” inside a nucleus, is roughly the same for all nuclei* Give this
density in macroscopic units, i,e*, in gm/'crn3.
6 (a) With reference to the discussion in Sec. 17, estimate the mean
energy and the mean velocity of a nitrogen molecule in nitrogen gas held
at room temperature. The nitrogen molecule consists of two nitrogen atoms,
(Give energy in eV,)
(b) Under atmospheric pressure, and at room temperature, 1 mole of
nitrogen gas (or any gas) occupies a volume of 22.4 liters. Estimate the
number of collisions that a nitrogen molecule will undergo per second, assum¬
ing that a nitrogen molecule is of “typical molecular size." Compare this
collision frequency with a typical optical frequency.
8 Singly ionized helium, i.e,, a helium atom with one electron removed
is, like the hydrogen atom, a system consisting of a single electron moving
around a nucleus. We may therefore expect that the spectral lines emitted
by singly ionized helium are entirely analogous to the spectral lines emitted
by the hydrogen atom* The two systems are, however, not identical; the
helium nucleus carries two elementary charges whereas the hydrogen nucleus
88 Chap 2♦
Froblems
(proton) carries only one. In view of what has been said in this chapter it
should be possible to find the consequences for the spectrum of the increased
central charge in singly ionized helium as compared to hydrogen, and it
should, therefore, also be possible to predict the wavelength of any line
emitted by singly ionized helium, given the wavelength of the corresponding
line in hydrogen. In other words, it is possible, without a detailed theory of
atomic structure, to find the ratios of the corresponding wavelengths.
One of the visible hydrogen lines has the wavelength 6562.99 A. What is
file wavelength of the corresponding line emitted by singly ionized helium?
Does ttus line fie in the visible region?
We may here assume that both nuclei are infinitely heavy. This example
teaches us that primitive dimensional arguments, like the one given in Sec. 27,
can sometimes be employed to make precise, quantitative predictions.
Let us furthermore assume that the number of neutrons equals the number
of protons, i.e.( that A = 2Z. From (i) we then obtain an expression for the
electrostatic energy per nucleon, namely
(b) A small chunk of U233 metal will not explode spontaneously, whereas
a large chunk will. How do you explain this?
(c) To study the origin of the energy released in fission consider, on the
90 Chap. 2 Problems
basis of the relation (i) in Prob. 10, the electrostatic energy of the nudens
(say U235) before fission and the total electrostatic eneigy of the fragments.
Obviously some of the electrostatic energy will be released* Estimate this
energy and compare with the value 200 MeV per fission.
12 The mass of two deuterium nuclei is larger than the mass of the alpha-
particle (= the nucleus 2He4). (See Table 4A for atomic masses,)
(a) Compute the energy released if 1 gm of deuterium undergoes fusion to
form helium* and compare this with the energy released in fission,
(b) Why is it that a container filled with deuterium does not explode
spontaneously?
and then remain constant at their new numerical values. We shall assume
that the numbers u and w are small, say of the order of 1 percent; otherwise
the change in the world order may be too drastic. This natural catastrophe
will certainly be noticed, and for some time (after recovering from the initial
shock) physicists will be busy with the re measurement of their sacred con¬
stants. Let us denote quantities after the catastrophe by primes.
(a) Find (mW(m)i.
(b) What are the new values of the mass of the electron and the mass of
the proton? [In (gm)p.]
(c) What is the new value of the velocity of light, c\ in units of (m)o/(sec)^?
(d) What is the new value li' of Planck’s constant?
(e) What is the new value of the electronic charge in electrostatic units,
and what is the new value in coulombs?
(/) What will be the density of copper [in (gm)p/ (cm3 )o] after the
catastrophe?
Energy Levels
Sections I to 13 Term Schemes 94
Problems 137
Chapter 3 Energy Levels
Term Schemes
4742.5 —4728.6 A.
1 The fact that each chemical element is associated with a unique
Wave¬ Ele¬ Intensities Wave¬ Ele¬
length ment Arc R length ment optical spectrum is one of the striking aspects of nature. This
*742.589 Mo 10 4737.642 Sc I feature of nature is furthermore very general: not only do atoms
4742.549 Er 3w — — 4737.626 U
4742.5 bh 5c 5 —
Me 4737.561 Pt I have characteristic spectra, but so do molecules and nuclei as well.
4742.481 Sm 3 — *+ 4737.350 Cr
4742.392 Md 4 — - 4737.282 Ce These objects emit and absorb electromagnetic radiation at certain
4742.333 \J W 3 — 4737.1 bh C
4742.325 Pr 7 — — 4737.05 Tl U definite frequencies, which range from the radio frequency region
4742.266 Th 41 2 — 4736,965 Zr (for molecules) up to the region of very short wavelength X-rays,
4742.25 So I — [600] Rd 4736.958 Sm
4742+227 5m 2 — — 4736.945 Er or gamma rays (for nuclei). Historically the optical spectra of the
4742+110 Tf I 15 1 _ 4736+9 bh 2
4742.04 Ho 10 3 Ex 4736,79 Dy elements were discovered first, by G. R* Kirchhoff and R+ Bunsen
4741.997 Er 3w — — 4736.782 Ca
4741.937 Ge 11 - 50 - 4736.780 Fe in the middle of the nineteenth century, whereas the radio-fre¬
4741.922 Sr I 30 — JSn 4736.688 Pr
quency spectra of molecules, and the gamma ray spectra of nuclei,
4741.78 Cd n 3 Vs 4736637 Me
4741.775 Eu low — — 4736.608 Eu were discovered much later, during this century*
4741.726 Sm II 80 — — 4736.6 Rt
4741,71 o u — [20] FI 4736.491 Cl We interpret the spectra in terms of energy levels of atoms,
4741.539 Dy 3 2 — 4736.490 S'
4741.533 Fe 1 12 1 S 4736.30 T molecules and nuclei. Through our study of spectra we become
4741.620 W 12 2 - 4736,203 r
4741+503 Pr 30 — - 4736.151 i aware of an extremely important property of composite systems,
4741+404 Yt I 2 3 — 4736.116
4741.396 Er 20 — - 4736.069 which is this: with each such system there is associated a set of
4741.282 U 1 2 — 4736,062 energy levels, or stationary states, characteristic of the system.
4741+269 Ru 4 — — 4735+94
4741.10 Tm 3 - Me 4735.93 We find these levels in “small” systems, such as atoms, molecules
4741.018 Sc I 1W 60 h - 4735.848
4741.005 Pr 6 - — 4735.847 and nuclei, in which cases the levels manifest themselves very
4740.97 Sc u - [600] Bi 4735.77 directly in the spectra which we observe. We also find these levels
4740.928 Dy 3 2 — 4735,76
4740.68
4740,514
Cl I
ct>
—
3
[10]
3
Ks
-
4735.66
■ 4735,4?
in “large systems,” such as solids, liquids and gases. At first it might
4740.524 Eu 500 2 — 4735.4: not occur to us that there is a relationship between the emission
4740.517 Th 20 15 — 4735.3
4740.5 bh Zr 8 — L 4735.3 and absorption of gamma rays by a nucleus, and the vibrations of
4740+40 Cl H ,- [150] Ks 4735.1
4740.359 Me 5 5 — a quartz crystal in some electronic device, but there is*
4740.331 Ru 7 —
concerning energy levels more generally, and therefore we shall '■ if-
:
also discuss nuclei, although the properties of nuclei became known
only much later.
- . f ^ ■. f ■ r . . . •
We then have
hv = fiio = Eu — Ei (5a)
t As we explained in See. 8, Chap. 2, both v and the associated quantity = 2nr are
called "frequency.” Similarly both h and ft = h/2m are called “Planck’s constant”
In the fallowing we will mostly use w and because the author likes them better than
v and h.
Several portions of the spectrum of iron photographed
on the same glass plate. Wavelengths written on the
plate are in Angstrdms. The purpose of this partic¬
ular photograph was not to measure the wavelengths
’■% of iron, but rather to use these well-known wave¬
lengths to calibrate the quartz prism spectrograph.
(Photograph by courtesy of Professor S. P. Davrs,
Berkeley.)
98 Energy Levels Sea 3.6
(£3 - Eg)
“10 “32
n
5790 (visible) light, unless the atoms are excited by some other (external)
6770
M61 (7,7) means.
4916 (9,2)
8 If we study the emission spectrum of a gas of the atoms, excited
for instance by an electrical discharge, we might observe all the
spectral lines indicated. If an atom, originally in its ground state,
4858 (7,1 4348 (2,5)
collides with an energetic electron the electron may transfer part
4109 (9,7)
of its energy to the atom. This causes the atom to jump to one of
4078 (7,5
4047 (7,7) the higher states from which it subsequently decays to a lower
8900 (9,8) level with the emission of light* It is self-evident that this process
cannot take place unless the electron has sufficient energy to raise
the atom to one of the excited states. If the energy of the electron
8663)
3665} (8,8) is less than (£h — Eo) then the electron can only undergo an elastic
86601
collision with the atom. If the energy is higher, inelastic collisions
no leading to light emission become possible.
<i 5
<
O
<1
o There is an obvious experimental test of this picture, and indeed
of the general ideas underlying our postulates of Sec* 5* We simply
Fig* 8A The spectrum of the mercury atom when vary the energy of the electrons used in exciting the atoms, and as
excited by electron collisions, for two different elec- the energy increases new emission lines ought to appear. Fig. 8A
tron energies. This photograph is taken from G. shows some results in such an experiment, for a gas of mercury
Hertz, N0ber die Anregung von Spektrallinien durch atoms* As we see the appearance of the emission spectrum changes
Elektronenstoss, I/1 Zeitschr/ft fur Pfiysik 22t 18
(1924).
When the electron energy is increased from 8,7 eV
(left spectrum) to 9.7 eV (right spectrum) a whole set
of new lines appears, of which there is no trace in the cm eV
left spectrum. The numbers within parentheses in
the figure show the electron energies at which the 79661
lines first appear, and the numbers without paren¬ 78404
77064
theses are the wavelengths in Angstroms. (Courtesy
74405
of Springer Vertag.)
71431
71396
71336
71333
flpooccxaD'Tj ©<
SO'fHdit'-L.
co^f'TtL0 to ic CO CO CO
Fig* ISA A mechanical model of an atom, which is 15 The observed facts can be understood very easily in terms of
helpful in understanding resonance fluorescence. another model* Let us regard the atom as a mechanical system in
If this contraption is excited by a kick (say in a colli¬ which the electrons are bound to the nucleus by springs. Such a
sion with an electron), it will oscillate, and since the
system will have a number of resonant frequencies, one of which is
electrons are charged, electromagnetic radiation wilt
be emitted at the resonant frequencies of the system* the frequency In the ground state of the atom this system is at
The motion is necessarily damped, because the sys¬ rest, but an incident electromagnetic wave will excite oscillations in
tem loses energy through radiation. the system* As a result the oscillating elections will radiate an
Under the influence of an incident electromagnetic electromagnetic wave of the same frequency as the incident wave.
wave, the atom will undergo forced oscillations at the
The amplitude of the oscillations will be larger the closer we are to
frequency of the incident wave, and hence will emit
radiation at the same frequency. This is the phe¬ the resonant frequency too and the effectiveness of the atom as a
nomenon of resonance fluorescence. scatterer will clearly be largest when the incident frequency is
Sec. 3.16 Energy Levels 105
emitted by the atom? Do they also have finite widths? The an¬
swer is yes. The width of an emission line is the same as the width
of the corresponding absorption line. (We must mention here that
the lines of optical spectra, such as we observe them in practice,
are broadened because of several different effects. We are here
concerned with the width of a spectral line emitted, or absorbed,
by an isolated atom, originally at rest with respect to the observer.
This width is an intrinsic property of the atom. Let us forget
temporarily about all other causes of broadening: we will discuss
these causes later in this chapter.)
What does it mean that an emission line has a finite width? It
literally means what it says: if we photograph the line with a spec¬
trograph of extremely high resolution, we find that the line has a
finite width. The frequency of the emitted light is not precisely
too, but we also find frequencies in the immediate neighborhood
of ido-
\
energy of oscillation has diminished to 1/e times its original value.
(The time required for this is the “mean-life of the oscillatory
state.”) Let the time interval between two successive outswings
to the right be one second.
Suppose now that someone asks us for the frequency of the
pendulum. Without much reflection we would answer that the
frequency is one per second. This is certainly a reasonable answer,
but strictly speaking it is wrong: by “frequency” we understand
the repetition rate of a periodic phenomenon. The motion of the
pendulum is, however, only approximately periodic since the
amplitude of oscillation does diminish as time goes on. The fre¬
quency of a damped harmonic motion is not precisely defined,
although it may, for all practical purposes, be very wefl defined
indeed. Fig* 19A An exponential damped oscillatory proo
An atom emitting radiation is in some respects analogous to a ess, showing the amplitude as a function of time.
Since the process is not strictly periodic in time it is
damped pendulum. The emission process does not go on forever,
wrong to say that the frequency of the oscillation is
and this must mean that the “oscillation inside the atom” is a coo, because the concept of frequency refers to a
damped oscillation. There is not, therefore, a precisely defined periodic phenomenon. If the damping is not too large
frequency, because the oscillatory phenomenon is not strictly it is fair to say that the frequency is approximate/y
periodic. The electromagnetic radiation emitted by that “some¬ It is intuitively clear that the smaller the damping,
Le., the smaller the decrease in amplitude for two
thing which is oscillating inside the atom” is thus not monochro¬
successive maxima, the better is the frequency de¬
matic. The emitted line has a finite width. fined.
emitted wave must be of the same form (20a). The intensity l(t)
of the emitted radiation is proportional to the absolute square of
the amplitude:
1 2
S(«) proportional to
(w — wo) + i/2t
or
Fig. 21A The Universal Resonance Curve. It de¬
scribes the response of any linear (or approximately
(1/2t)2
S(w) = S(<d0) (2 Id) linear) system to a sinusoidally varying external force
(d0)2 + (1/2t)2 in the neighborhood of a resonant frequency provided
no other resonant frequency is close by.
where S(«o) is the amount of scattering “at resonance,” i.e., when (Two bell shaped curves play a particularly impor¬
CO — LO0h
tant role in physics: the resonance curve and the
A schematic plot of S(co) versus co is shown in Fig, 21A. gaussian curve. As usually drawn they may look very
similar, It must be remembered, however, that the
gaussian falls off very rapidly outside the central
22 The function S(tj) expresses “the intensity of response" of the
region, whereas the resonance curve has a long
system under an external perturbation at the frequency co* This "tail.")
kind of resonant response is a very general phenomenon in quan¬
tum physics, and it is by no means restricted to the interaction of
l&ght with atoms* We find the same resonant response when we
study the scattering of material particles, such as protons, of a
well-defined energy, from a nucleus, or the scattering of pions from
a proton. One might well say that a quasi-stable energy level of
a quantum-mechanical system “exists” in precisely the sense that
the system exhibits a resonant response, as given by equation (2Id),
at the appropriate frequency*
In nuclear physics the resonance formula (2 Id) is known as the
Breit-Wigner one-level resonance formula, after G* Breit and
E. P. Wigner*
“ = “° ± ~kr (23a)
(23b)
T
110 Energy Levels Sec* 3.24
This agrees with our conjecture, in Sec. 20, about the relation¬
ship between the uncertainty in the frequency and the mean-life
of the excited state,
Since we can define the width of the (excited) energy level by
AE = fi Aw, we immediately derive from (23b) the very important
relation
AE = - (23c)
T
Further Discussion of Levets and Term Schemes Fig. 27A Portion of a tabfe in a paper by J, Sugar,
MDescription and Analysis of the Third Spectrum of
Cerium (Ce III)/1 Journal of The Optical Society of
27 Let us now look at a number of typical term schemes. They America 55, 33 (1965); exhibit taken from page 44.
have been constructed on the basis of actual measurements, inter¬ First column shows wavelength in air of observed lines
preted within the framework of quantum mechanics. We should of the doubly ionized cerium atom. Second shows
relative intensity of line. Third shows energy of pho¬
look at them with proper respect: each diagram, or rather the asso¬
ton, expressed as a wave number. The fourth column
ciated table of wavelengths, is the fruit of a considerable amount shows spectral terms involved, with energies ex¬
of human labor. pressed as wave numbers,
112 Energy Levels Sec. 3.27
Fig. 28A Term scheme for the neutral lithium atom, We have drawn our term schemes in the form in which the
The slanted lines represent observed electric dipole reader will find them in the literature. The drawing of such
transitions. The numbers on these lines are the scheme^ and the labeling of the different energy levels, is governed
wavelengths in Angstrdms. For other details, see the
by a number of conventions of long standing. For greater realism
explanation in the text. Based on a figure in W.
Grotrian, Graphische Darsfef/ung der Spektren von we wanted to adhere to these conventions, even if we cannot here
Atomen. , , r vol. M, p. 15 (Verlag von Julius Springer, explain every detail of the drawings. The reader may want to
Berlin, 1928). object that we should not show anything in the diagrams that we
Sec. 3.28 Energy Levels 113
31 The theorem which we have stated is, however, not the whole
story of the selection rules which operate for the lithium atom. In
atomic physics there is an additional, approximate, selection rule
governing electric dipole transitions, and it can be stated as follows;
in an electric dipole transition the orbital angular momentum of the
electrons must change by precisely one unit, or
Al = If — k — — 1 or +1 (31a)
columns in Fig. 28A are labeled, are in fact code letters for the
orbital angular momentum, as follows: 'V' means 1 = 0; <fp*’ means
1=1; means 1 = 2 and “f* means 1 = 3* The selection rule
which we mentioned in Sec. 28 is equivalent to the selection rule
(31a).
It is not always possible to make a definite assignment of orbital
angular momentum quantum number to an energy level in an atom,
although it so happens that it can be done without ambiguity for
an alkali atom, such as lithium. The reason for this is that whereas
the total angular momentum is a constant of motion, neither the
orbital angular momentum, nor the spin angular momentum is.
The levels, in other words, do not, in general, have a definite value 90
of L This is the sense in which the rule (31a) is only approximately Na I—Continued
|
valid. As we have said, it is a good rule for alkali atoms (and for Config. De&ig. J j Level Inters'*] i1
hydrogen).
V »F°
{ %X } ssjm i
32 Consider again Fig. 28A* What about /, and the selection eft 4U 33403. i
{ }
rule (30b)? This rule does not quite show up in Fig. 28A because rp 7p »P* 33540. 4Q
X a 74
we have exhibited a simplified form of the term scheme* We should W 33541.14
actually have drawn the p-, d-f and /-columns double* The sub¬ a# 8s *S x 3S9G& 35
7d 3920a 962
scripts f and £ on the column-labels s, p, d and / indicate the IX 39200. 963 -OL001
Sjj X 39338. 54
a 47
mentum is due to the electron spin)* For all other values of l, IX 39899. 01
j can have the values / = l + £ and / = £ — ■£■* (For other atoms 0s *S X 39574. 51
2X
the rules are different)* The level 2p is thus actually double, but the ed 8cf *D
{ } 39729. 00
separation in energy between the two levels of the doublet is quite TF* [S9734. O]
{ }
small, and within the accuracy of the figure, the levels coincide.
Op *PD X 39794. S3 a 47
Fig. 32A shows the term scheme of the sodium atom. Sodium is 39795.00
also an alkali atom, and it is quite obvious that its term scheme is, 10s 10s *8 X 39983. 0
sible for the characteristic yellow light of a sodium lamp are the
transitions from the 3pi/2- and 3p3/2-levels to the ground state
Fig. 32B A portion of a ta ble of energy levels in the
3$i/2- “The yellow sodium line” is in fact a doublet* neutral sodium atom. The energies (fourth column)
The reader should ponder the term scheme in Fig* 32A and are expressed in wave numbers, cm-1, measured
convince himself that the transitions shown obey the selection rules from the ground state. The column labeled / gives
(30b) and (31a) for ; and respectively. the angular momentum of the state.
Exhibit taken from C. E. Moore, Atomic Energy
Levefs, vol, lF p. 90 (Circular of the National Bureau
33 The energy levels of the helium atom, shown in Fig* 33A form, of Standards 467, U.S. Government Printing Office,
as it were, two almost completely independent systems: the singlet Washington, 1949).
118 Energy Levels Sec. 3.34
Fig. 33A Term scheme for the neutral helium atom. system and the triplet system. The observed spectral lines arise
Note the remarkable separation between the singlet from transitions within these systems: from singlet levels to singlet
and the triplet systems ol levels. In the triplet states
the electron spins are parallel, and in the singlet states
levels, and from triplet levels to triplet levels.
anti-parallel. There is an obvious correspondence The helium atom is a two-electron atom. In the singlet levels
between the triplet and singlet Jevels, except that the the two electron spins are oppositely directed, whereas in the
singlet ground state has no analog among the triplet triplet levels the two electron spins are parallel,
states. This circumstance is a consequence of the
The letters S, P, D, F, . . , , designate the total orbital angular
Pauli exclusion principle: two electrons whose spins
point in the same direction cannot both occupy the momentum of the electrons. The left superscript I or 3 designate
lowest level. There is no such restriction if the spins the multiplicity (singlet or triplet). For the singlet levels the total
are oppositely directed. angular momentum equals the orbital angular momentum. For the
triplet levels the total angular momentum / can assume the values
/ = f — 1, Z, Z -|- I, with the provision that we always have ; ^ 0.
In the triplet system the S-levels are single, and the remaining levels
triple. The singlet levels are, of course, single.
S 1/2
^5/2 ^S/2,7/2
0
8F 5000
- 1
10,000 H
eV
15,000
- 2
20,000 H
- 3
25,000 —
30,000
-4-
35,000—
cm'1
-5 40,000 H
6^3/2
45,000—
- 6h
1/2 50,000 J
Fig* 34A Term scheme for the neutral thallium
atom. The numbers on the slanted lines give the
wavelengths in Angstroms of observed transitions.
(“After Grotrian.)
120 Energy Levels Sec, 3.35
Fig*35A Term schemes of zinct cadmium and mer¬ thallium atom, Fig, 34A, An atom in the state 72S±/2 can decay
cury, shown together to illustrate the fact that chemi¬ either to the state 62IV2? or to the ground state 62Pi/2. The atom
cally similar elements have similar term schemes.
has a choice which way to make the “jump,” There are other
This figure is taken from W. Grotrlan, Graphiscfte
DarsteMung cfer Speitfren von Afomen und Jonen . , . , examples of this feature in the thaDium term scheme, as well as in
voL II, Sfrukfur der Materie, Band VIII, p. 131 (Verlag some of the other term schemes shown in this chapter* (The reader
von Julius Springer, Berlin, 1928). (Courtesy of should hunt for these examples,) If an excited state can decay in
Spr/nger Vertag>) several different ways, each mode of decay occurs with a definite
probability. This probability is known as the branching ratio for
the decay mode in question. That the branching ratios are intrinsic
properties of the excited state, i.e., insensitive to how the excited
state was reached, is an experimental fact.
1 2
H He
1,0080 4.008
3 4 5 6 7 8 9 10
Li Be B c N 0 F Ne
6.940 9.013 10.82 12,011 14.008 16.000 19,00 20.183
11 12 13 14 15 18 17 18
Na Mg Al Si P s Cl At
22.991 24,32 26.98 28,09 30,975 32,000 35,457 39.944
19 20 21 22 23 24 25 26 27 23 29 30 31 32 33 34 35 36
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
39,100 40.08 44.96 47.90 50.95 52.01 54,94 55.85 58.94 58.71 63.54 65.38 69,72 72.60 74.91 78,96 79.916 83,80
38 39 40 41 42 43 44 45 46 47 48 49 50 mm 52 53 54
Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Te I Xe
37.63 38.92 91,22 92.91 95.95 101,1 102.91 106.4 107.880 112.41 114.82 118.70 1S1.70 127.61 126,91 131.30
55 56 57-71 72 73 75 76 77 78 79 80 81 82 83 84 85 86
Cs Ba La Hf Ta Be Os Pt Au Hg Tl Pb Ri Po At Bn
Ir
132,91 Series 178.50
137,36 180.95 186.22 190,2 192.2 195.09 197.0 200.61 204.39 207.21 208.99
87 KW 89-103 <104) (105) (106) (107) BRgn
Fr Ac ■1
Series
Lanthanide
57 58 59 80 ei 62 63 64 65 66 67 68 69 70 71
series La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
138.92 140.13 140.92 144.27 150.35 152,0 157,20 156,93 162.51 164.94 107.27 103.94 173,04 174.99
Actinide
89 90 91 92 93 94 95 96 97 98 99 100 101 102 103
series Ac Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lw
227,04 232.05 231.05 238.04
elements have similar term schemes* Figure 35A shows an example Fig* 35B The periodic chart of the elements. The
of this* The reason for this is that the optical spectrum and the atomic number Z is given above the chemical symbol,
chemical properties of an element are both determined by the and the atomic weight (for reasonably stable elements)
below.
electronic configuration in the atom, and in particular by the con¬
Note the lanthanide series (rare earth series) con¬
figuration of the outermost electrons. sisting of 15 chemically very similar elements. All
The remarkable periodic table of the chemical elements, shown these atoms have the same configuration of electrons
in Fig, 3SB, can be understood in terms of the shell structure of in the outermost shell. The series arises because
atoms* In this rectangular table the elements are arranged in a Inner shells which were "bypassed" are being filled
as we progress in the series. On the basis of this
certain way, in order of increasing atomic number Z, and with
picture Bohr predicted that the element of atomic
elements of similar chemical properties standing in the same col¬ number 72, hafnium, at that time undiscovered, would
umn, The number of electrons in an atom equals Z, and as we be chemically similar to zirconium rather than to the
progress in the table, in the direction of increasing Z, the “shells” rare earths. Hafnium was indeed later found in a
zirconium mineral, which was a striking triumph for
are filled with electrons in a regular manner. The chemical prop¬
the theory.
erties depend on how the shells are filled. For instance, noble gases The so-called actinide elements form an analogous
will occur in the table when certain shells are completely filled series.
122 Energy Levels Sea 3.36
atomic spectra and energy levels, and this would be a bit too much
I < in an introductory course. To whet the reader’s appetite we show,
in Fig* 35C, a portion of a table of electronic configurations of
1
atoms*
He 2
Li 3
Be 4 36 When the periodic table was first proposed by D* I. Mendele-
B 5 jeff in 1869 neither electrons nor nuclei were known* Mendelejeff
c 6 therefore did not arrange the elements according to the charge Z,
N 7 but rather in order of increasing atomic weight. Fortunately this
0 8 gives the right order, with very few exceptions* The sequence
f; 9
argon-potassium is such an exception: argon has a larger atomic
Ne! 10
weight than potassium although the chemical properties of these
Na 11
elements (argon is a noble gas and potassium an alkali metal) estab¬
Mg 12
lish without a trace of doubt that argon must come first. From
A1 13
Si 14 the standpoint of chemistry the order of the elements in the table
P 15 is quite clear, and on this basis it is thus possible to assign an atomic
S 16 number Z to each element.
Cl 17 We should mention here that Mendelejeff had the foresight to
A 13 leave some empty spaces in his table, to accommodate elements not
K 19 yet discovered.!
Ca 20
Sc 21 37 The realization that the atomic number actually measures the
Ti 22 nuclear charge, and hence equals the number of electrons, was an
important step forward in atomic theory* The work of H, G* J.
Kg, 35C The shell structure of light atoms. The Moseley around 1913 was particularly important in settling this
main shells, designated by the letters Kf L, M, Nt,. , , question, He systematically measured the wavelengths of X-rays
are divided into subshells as shown. The different from a large number of elements and was able to show that the
periods are indicated by the thin horizontal lines.
wavelengths of analogous lines (in different elements) depend on
Completed noble gas configurations are shown in
gray. For the first three periods the shells are sue- the atomic number in a very simple way.f Let us discuss this
cessively filled in a pleasingly regular manner, but question briefly*
beginning with potassium an outer shell is being filled
before an inner shell has been completed. This phe¬ f For an account of Mendelejeff $ work, and the history of the periodic table, see
nomenon also occurs later in the periodic table. It is The World of the Atom, Vol. 1, edited by H. A. Boorse and L. Motz (Basic Books, Inc.,
well understood theoretically. New York, 1966).
An $'$ubshell can accommodate 2 electrons, a p- J H. G. J, Moseley, "The High-Frequency Spectra of the Elements,” Philosophical
shell 6P and a d'Shell 10 electrons, Magazine 26> 1024 (1913), and 27, 703 (1914),
Sea 3.38 Energy Levels 123
Bk = (37a)
in Fig, 38A shows the nuclear energy levels of the boron isotope
gB11, as they have been determined experimentally.
In this figure we have assigned the energy zero to the ground
state. The total angular momentum of the ground state is j = f.
Levels which are particularly broad have been drawn hatched,
and the hatching is an approximate measure of the widths.
The dissociation limit of this nucleus is at 8.667 MeV: above
this energy the nucleus can dissociate into an alpha-particle and
the lithium isotope 3L17. This mode of dissociation is indicated at
the right of the main term scheme. Above an energy of about 11
MeV the boron nucleus can dissociate in two different ways: either
into a neutron and the boron isotope gB10, or else into a proton
V
and the beryllium isotope 4Be10. These modes of dissociation
n are likewise indicated at the right of the level scheme for the
isotope 5B11.
Note, however, that the isotope gB11 has a system of energy
levels above the dissociation energy 8.667 MeVt Below this energy
the nucleus can only emit gamma rays, but above it the nucleus
can also emit material particles. (The observed gamma ray tran¬
sitions in 5B11 are indicated by vertical lines.)
10.61
10.32 As this example shows, we have to be a bit careful about inter¬
IIHMHlDHt
tni)iitnm»tiinwi preting the “continuum.” Levels can very well exist above the
□ 19 9.28t3/a+)
—8,92= dissociation limit. The dissociation energy is merely an energy at
8.57- which the system can dissociate into two material particles. Below
7,99 3Li7 + a 8.607
this limit the system can still "dissociate/7 but only into a photon
7.30
IIII If] I (3/JD and one material particle. If we wish to treat photons on the same
ML
6.76 footing as material particles we can conclude that the levels above
the dissociation limit (which are often called “virtual levels”) are
5,03 not different in principle from the levels below the dissociation
4.46 (5/2-)
limit: all levels above the ground state are unstable. Actually even
the ground state may be unstable: consider the ground state of a
radioactive nucleus. In our example in Fig. 38A the ground state
2.14 is stable: the isotope gB11 occurs in the boron found in nature.
39 Two nuclei are said to form a pair of mirror nuclei if one can
be obtained from the other by changing all protons into neutrons,
B 11 J = 3'a~
and vice versa.
As we said in Sec. 37, Chapter 2, the sfrong interactions, which
Fig, 38A Term scheme showing energy levels of the are the dominant interactions in nuclear physics, are believed to be
boron nucleus 5B11. This figure is a simplified version invariant under this change. The proton-proton force is the same
of a graph which appeared in F. Ajzenberg and T.
as the neutron-neutron force. If this belief is correct* and if there
,
Lauritsen, IJEnergy levels of light nuclei," Reviews of
Modern Physics 27 77 (1955). It is recommended
that the reader look at the original.
were no other interactions but the strong interactions, then the
level systems in two mirror nuclei must be identical.
Sec. 3AQ Energy Levels 125
In Fig. 39A and Fig. 39B we show the experimentally found 7.47“ — 7.10
energy levels of two pairs of mirror nuclei. As we see, it is possible 6.50- 0.35
to establish a correspondence between the levels of the pairs.
The energies of the corresponding levels are not* however, identi¬
--4.«5
cal, as the figures show. The reason for this is that electromagnetic
forces are also present, and the electromagnetic forces are not MeV MeV
invariant under neutron-proton interchange.
40 The term scheme in Fig. 40A explains why the alpha particles
emitted by a radioactive nucleus do not always emerge with a single ,477-.430
0-;-=-0
well-defined energy. The figure shows the alpha decay of the 3 Li7 4 Be'
bismuth isotope gaBi212 to the isotope siTl208 of thallium. The
decay takes place from the ground state of the parent nucleus to Fig. 39A The lithium and beryllium isotopes of
one of several excited states, or to the ground state, of the daughter mass number 7 form a pair of mirror nuclei: if the
neutrons in the lithium nucleus are changed into pro¬
nucleus. The term scheme is drawn so that the ground state of
tons, and vice versa, we obtain the beryllium nucleus.
the parent nucleus lies 6.2 MeV above the ground state of the Mirror nuclei have similar but not identical, level sys¬
daughter: this energy is the maximum kinetic energy with which tems. The difference is an effect of electromagnetism.
the alpha particle can be emitted. It is clear that if the decay takes
place to an excited state of the daughter, then the alpha particle
will emerge with a smaller energy. For the level system shown in
the figure, the alpha particle may be emitted with one of five well-
defined different energies* The tilted lines show these decays.
The numbers within parentheses are the branching ratios for the
different decay modes.
If the daughter nucleus is left in one of the excited states it
will emit gamma rays, indicated by the vertical lines, and eventually
reach the ground state.
8,93 ^ /8M
For many other alpha-active nuclei the decay always takes place 8.57- -8,44
to the ground state of the daughter, because no suitable excited 7,99 — ^8.12
states are available. The alpha particles will then emerge with a 7.30 — - 7.33
6,81--; - 6,77
single well-defined energy, and there will be no gamma-rays asso¬
6.78 / ^ 0.40
ciated with the alpha-decay.
5,03
4.77
4.48
41 By beta-disintegration we understand a process in which a 4.23
nucleus emits an electron or a positron. The simplest process of MeV MeV
this kind is the beta-decay of a neutron, which is a phenomenon
well established experimentally. The mean-life of the free neutron 2.14-
1,85
is 16 minutes. Since the neutron-proton mass difference is
— mp) = 1.3 MeV, we might draw a term scheme like the one
shown in Fig. 41A. The oblique line indicates the transition. If
only an electron were emitted, it would always be emitted with the
same energy (about 1.3 MeV), just as is the case in alpha-decay. Fig* 39B The boron and carbon isotopes of mass
Experimentally it is found that the electron, in fact, can be emitted number 11 form another pair of mirror nuclei,
126 Energy Levels Sec* 3.42
with any energy between the rest energy 0*5 MeV and the available
energy L3 MeV,
The explanation for this is that another particle, in this case the
massless anti-neutrino, is also emitted, and the available energy is
shared between the electron and the anti-neutrino. The reaction
formulas for beta-decay thus read
2TCo60 _► MNi“* + e- + v
of Spectral Lines
-Ekin = £M(u0l2 + Dq22 + Uq32) = ffeT (44a) ftg*42A Term scheme showing beta-gamma cas¬
cade emitted by the cobalt isotope btCo00. This Iso¬
where M ^ AMP is the mass of the atom, (Mp is the mass oi a tope first beta-decays to an excited state of the nickel
proton.) The mean velocities in the three different coordinate isotope seNi™ which lies 2.4 MeV above the ground
directions are clearly equal, and we obtain state, The maximum k/net/c energy of the electron
is 0,3 MeV. The excited state of the nickel isotope
subsequently decays through the emission of two
t)0 — Dos (44b)
gamma rays in rapid succession.
128 Energy Levels Sec. 3.45
(Au)c — — (45a)
Tc
which says that the mean number of atoms found inside a cylinder
of radius 2r swept out by an atom in the time rc shall be of order
unity.
Sec. 3A6 Energy Levels 129
One mole of any gas contains ATo = 6 X 1023 molecules (in our
case the molecules are atoms). At a temperature of 273° K, and
at a pressure of 1 atm* 1 mole occupies a volume of 22.4 liters. In
other words, at this temperature and pressure the number of atoms
per unit volume is given by
Np
no = = 2.7 X 101& atoms/cm3 (45c)
(22.4 liters)
The number of atoms per unit volume at any other pressure F,
and temperature T, is then given by
n — no (45d)
^- = $kT (45e)
2 2
where M = AMP is the mass of the atom. Combining all the above
equations (45) we finally obtain
(Au)« ~ j- ~
(2 X 10- sec-) X (j^) X y^(^F) <«>
48 The answer to the first question is: "Because the fine structure
constant a is so small.” Let us try to see what this means*
First of all, we recall the conclusion which we reached in Secs*
29 and 39 in Chapter 2 that the wavelength of the emitted electro¬
magnetic radiation is in general large compared with the size of
the atom or nucleus which emits the radiation. This circumstance
has important physical consequences, and it also simplifies the
mathematical discussion of radiation phenomena* Let us first sup¬
pose that an atom, or nucleus, in an excited state acts like an oscil¬
lating electric dipole* Let o> be the frequency of oscillation; this
is also the frequency of the emitted light. Let a denote the size
of the object* Since the thing that oscillates is one or more ele¬
mentary charges, we can assume that the electric dipole moment
is of the order of magnitude ea. That the object is small compared
with the wavelength is expressed by the condition
a<d
< 1 (48a)
W = (48b)
3?"* W
This formula gives the power emitted. Since we know that our
atom (or nucleus) will only emit a single photon, we are interested
in the time r it takes for the object to emit an amount of energy Tico.
This time is given by
1 con (48d)
r cf
We interpret t as the mean-life of the excited state: this is the
time it takes for the? excited state to decay through the emission
of a photon. Let- us consider the dimensionless quantity
a\cj
This quantity is proportional to the number of oscillations which
the system has time to perform during the time t, before it decays.
Clearly the excited state is the more stable the larger is the quantity
cdT. As we see <or is large for two reasons: it is proportional to the
‘Targe” quantity 1/a = 137, and it is proportional to the inverse
of the square of the quantity (cmo/c), and as we have said (ato/c) is
in general small.
and we now understand why excited states which can only decay
electromagnetically live long compared with the inverse of the
frequency of the light emitted.
* (c ~ f) w (50a)
where C(0) is a function of 9 which is proportional to the electric
dipole moment. Its precise form need not concern us here*
If both dipoles are now present, as in the figure, then the elec¬
tric fields due to the two dipoles almost cancel, but not quite,
because the distance from P to dipole 1 is ^ (r + § cos 9% whereas
the distance from P to dipole 2 is ^ (r — f cos 9): consequently
the field due to dipole 1 differs in phase from the field due to
dipole 2. The electric field, E2> is therefore given by
Fig* 50A Schematic picture of electric quadrupole
source. The arrows represent two electric dipoles,
Eo =
m exp X
oscillating with the same frequency w. They are of
equal magnitude, but oppositely directed. The elec¬
tric, as well as the magnetic, dipole moment of this
configuration vanishes, but the electric quadrupole
moment does not. if a is small compared with the
X
(iao> cos 9 — iaos cos ff'
j - exp (—-
2c
(50b)
wavelength X the rate at which energy is radiated from
the system is smaller by a factor (a/X)2 than the rate 51 We shall now make use of our assumption (48a) that (aw/c) is
from a single dipole. very small compared to unity: this assumption is clearly valid for
Sec. 3.52 Energy Levels 133
£&d\
(cos 6) E1 (51a)
c /
where Ei is given by (50a). The electric field £2 produced by the
electric quadrupole shown in Fig. 50A is therefore everywhere
smaller by at least a factor (au/c) than the electric field E\ pro¬
duced by a single one of the dipoles "making up the quadrupole.”
Since the radiation rate is proportional to the square of the electric
field we can conclude that the typical rate of electric quadrupole
radiation is smaller than the typical rate of electric dipole radiation
by a factor (ao>/c)2, The corresponding lifetimes are then related by
?E2 (51b)
1) Energy levels of atoms* molecules and nuclei are, of course, discussed in very many
tests on these subjects. Among these we mention the following fairly elementary ones:
a) G, Herzberg: Atomic Spefrtrti and Atomic Structure (Dover Publications, New
York, 1944).
b) H, White: Introduction to Atomic Spectra (McGraw-Hill Book Co,, New York,
1934).
c) G. Hersberg: Molecular Spectra and Molecular Structure: I, Spectra of Diatomic
Molecules (D. van Nostrand Co., New York, 1953).
d) D. Holliday: Introductory Nuclear Physics (John Wiley and Sons, Inc., New York,
1950),
e) E. Segr&: Nuclei and Particles (W. A, Benjamin, New York, 1964),
2) a) Term schemes of many atoms can be found in the book: W, Grotrian; Graph-
ische DarsteUung der Spektren von Atomen und lonen mit Eiro, Ztvei und Drei
Valenzelektronent vol II (Verlag von Julius Springer, Berlin, 1928),
b) For energy level diagrams of selected nuclei see: F. Ajzenberg and T, Lauritsen:
“Energy levels of light nuclei,” Ren. Mod, Phys, 27, 77 (1955)>
3) For shorter tables relating to spectra and energy levels we refer to:
a) Handbook of Chemistry and Physics (Chemical Rubber Publishing Co>).
b) American Institute of Physics Handbook (McGraw-Hill Book Co>, New York,
1957).
Froblems Energy Levels 137
4) There are several articles in the Scientific American which the reader can read
with profit at this point:
a) A- Bloom: ‘Optical Pumping," October I960, p, 72.
b) Hr Lyons: “Atomic Clocks," February 1957, p. 71.
c) G. E. Pake: “Magnetic Resonance,” August 1958, p. 58.
d) J, Pr Gordon: “The Maser," December 1958, p. 42.
e) A, L, Schawlow: “Advances in Optical Masers,” July 1963, p. 34.
f) Sr de Benedetti: “The MossbaueT Effect/' April I960, p. 72.
Problems
1 The following spectral lines were observed (early in this century) for
a certain atom:
figure- Can you explain these phenomena, and can you explain the physical
characteristic of the light sources in which the line can be expected to be of
the kind shown in the upper figure?
6 (a) From the experimental data presented in Fig. 37A, compute the
constant C in Eq, (37b).
(b) In the study of X-ray emission it is found that in order to make one of
the characteristic lines (of frequency u) appear, the energy E of the bombard¬
ing electrons must be quite a bit higher than tfw. For the Ka-lines, to which
Fig. 37A refers, the condition for the appearance of the lines is roughly
E > Why does not the line appear as soon as E >
7 Although the author cannot assume any responsibility for the harmful
mental images which might be formed if the reader studies Bohr*s planetary
model of the atom, he does not want to go as far as to outright forbid the
reader to consider this model, Bohr assumed that the electron moves in a
circular orbit in the hydrogen atom, and in such a way that the angular
momentum of the electron is a positive integral multiple of ft. It is a remark¬
able accident that this model gives the correct location of all the energy
levels to a very high accuracy. Since this model is of considerable historical
interest the reader may wish to follow in Bohr's footsteps, and construct the
term scheme, and identify the lines shown in Fig. IB. (The wavelengths
noted on the photographic plate are: 4861.3 A, 4340,5 A, 4101.7 A, 3970,1 A*
3889.1 A, and 3835.4 A.)
This latter bar is joined to some of the levels of B11 by lines with arrowheads.
This feature of the diagram also refers to some measurements. Discuss these
measurements, and explain what the arrows represent.
for the ease of the electric quadrupole shown in Fig. 50A, Here A and B are
constants. The intensity is independent of the azimuthal angle. This ex¬
ample indicates how the different kinds of multipole radiation can be distin¬
guished from each other by their characteristic intensity patterns.
Photons
Sections 1 to 17 The Photon as a Particle 142
Problems 175
Chapter 4 Photons
1 In this chapter and the next we shall explore both the particle
and wave aspects of such fundamental entities as the photon, the
electron, the proton, the neutron, and the other elementary particles
found in nature. We shall look at some basic experimental facts
and try to obtain a preliminary consistent picture of what is ob¬
served. In many instances, the outcome of a particular experiment
may suggest a new experiment: when this is the case, we shall try
to make a prediction and then study what has actually been ob¬
served. Our approach is one of experimentation with ideas, and we
should be careful not to commit ourselves too firmly to any specific
model yet: let us see how things work out.
f Berkeley Physics Course, Vol. I, Mechanics. The formula for the longitudinal
Doppler shift was derived in Chap, II, and the transformation law for energy and
momentum was derived in Chap. 12.
144 Photons Sec♦ 4.6
E = 4toP E* = (5c)
to eliminate E and Ef from (5b), and then eliminate w' from the
resulting equation by using (5a), we obtain
fibs — vp
VI - (v/W
or
p = ± (5e)
c
where the first term gives the energy density due to the incident
radiation, and the second term gives the energy density due to the
outflowing radiation. The flux and the radiation pressure are thus
related by
9 Let us now look upon this situation from the standpoint of the
photon picture. In this picture there is a flux of, say, N photons
per unit time through a unit area against the mirror. Each photon
carries an energy E = -fiw and a momentum p = ftw/c* After
colliding with the mirror each photon has its momentum reversed
(regard the mirror as being infinitely heavy since it remains at rest),
and each photon therefore transfers an amount of momentum 2p
to the mirror: in this picture the radiation pressure arises from the
bombardment of the mirror by the photons.
The radiation pressure P is equal to the amount of momentum
Mirror transferred per unit time to a unit area of the mirror, and we thus
have
Incident photons
P = 2Np = ^ (9a)
c
Reflected photons On the other hand the energy flux $ is simply given by
0 — Af-Hu (9b)
Fig. 9A Reflection of light from a mirror according and the energy density (since each photon travels with the velocity
to the particle picture. The radiation pressure arises of light) is given by
when the photons collide with the mirror, and have
their momenta reversed (for the case of normal inci- ^ = f jyrito (Qc)
dence). The relation between radiation pressure and c
energy density is the same as in the wave theory, (See . p
Pig gA,) If we combine the formulas (9a)-(9c) we recover the relations
Sec. 4.10 Photons 147
(8a)-(8c), which means that for the case considered the photon
picture is consistent with the wave picture*
Here we have taken into account the fact that the mirror may
have a (slightly) different velocity v' after the collision: the direction
of velocity will, however, remain unchanged* The reflected photon
will travel in the opposite direction, and hence the term — pf in
(10a)*
Let the frequency of the reflected photon be to' = E'/fi. We
may rewrite equations (10a) and (10b) in the form
1 - (10g)
1 - (lla)
1 - (lib)
with respect to the mirror, and the mirror itself moves with a veloo
ity v with respect to the stationary observer* Since o is small we Moving
Stationary image
can employ the non-relativistic law for addition of velocities, and
source
we conclude that the image of the light source appears to move L—f-r'
away from the observer with the velocity 2v. The frequency must
therefore be Doppler shifted, and the reflected frequency os' will To stationary 2v
be given (in the non-relativistic approximation) by os' = w(l — 2u/c), observer
which is in accordance with (10g),
Moving mirror
13 Let us next consider the intensity* In Vol. II of this seriesf
we have discussed the transformation laws for the electromagnetic Fig. 12A Light from a stationary source reflected
fields under Lorentz transformations* Let E and B be the ampli¬ from a moving mirror appears to come from a moving
tudes of the electric and magnetic fields of the wave in the rest source: the image moves with twice the velocity of the
frame of the source. The fields E and B are perpendicular to the mirror* The wave theory accord ingfy predicts that
the frequency of the reflected Fight will be Doppler
direction of propagation. We denote the corresponding amplitudes
shifted. (Imagine, for simplicity, that the figure shows
in the frame in which the source moves away from the observer a monochromatic candle.)
with the velocity vf by E' and Bf, For a plane linearly polarized
wave we in fact have E = B and Ef = B\ The transformation
laws then say that the primed and unprimed amplitudes are
related by
E = <13*>
c — vf
(13b)
c + vf
where <6 is the intensity in the rest frame of the source, and where
is the intensity in the frame in which the source moves away
from the observer with velocity v\ If we now write v' = 2vy and
expand the right side of (13b) in powers of u/c, assuming that this
quantity is small, we recover the expression (lib) in the linear
approximation.
We see that the particle picture leads to conclusions identical
with those which can be drawn from the wave picture, i.e.5 from
classical electromagnetic theory*
go? Since the mirror is moving, the radiation pressure will perform
work on the mirror: this work accounts for half the net flux* The
other half goes toward building up the electromagnetic field in the
space between the mirror and the plane of observation: since the
volume of this space increases steadily but the energy density re¬
mains constant, energy has to be supplied at a steady rate. In the
photon picture we would rather say that the number of photons in
transit between the mirror and the plane of observation increases
uniformly because the distance increases. The reader should carry
out the indicated very simple computations in detail to convince
Fig. 14A The intensity, i.e., the flux of energy per
himself that the energy flow is balanced.
unit area per unit time, of light reflected from a mirror
moving away from the source and observer is smaller
than the incident intensity. The radiation pressure 15 Let us next consider an example which teaches us the need
does work on the mirror, and the volume filled with for caution* An extremely monochromatic beam of light of fre¬
radiative energy increases. quency Wo (as may be obtained using a laser as a light source) is
Both the particle picture and the wave picture cor¬
rectly account for the energy balance.
incident perpendicularly on a mirror which vibrates with the fre¬
quency in the direction of the beam* We wish to find the
frequency of the reflected light.
On the basis of a naive particle picture, one might argue as fol¬
lows: if the photon happens to hit the mirror at an instant when
the velocity of the mirror is v, away from the source, then the
reflected photon will be of frequency w = Wo(l — 2u/c), in accord¬
ance with our earlier discussion* The photons arrive randomly at
the mirror, and therefore we will encounter a continuum of fre¬
quencies ranging from w<)(l — 2vo/c) to wq(1 -f 2vq/c) in the re¬
flected light: the spectral distribution of the initially almost
monochromatic light will be broadened. In the above formula t?o
is the maximum velocity of the mirror.
Since
Two
P = -; p' — p • p' = pp' cos 9 (18d)
c
we can solve (18c) for and we obtain
(18e)
W — 1 + (fco/mc2) (1 — cos 8)
Fig. 18A To illustrate the kinematics of Compton 19 If we introduce the wavelengths X = 27tc/w, Xf — 2mcf(d\ we
scatter!ngt in which a photon collides with an electron may write (18e) in the alternative form
originally at rest. The conservation laws of energy
and momentum imply a unique frequency aA and Xf = X + 27r{K/mc) (1 — cos 8} (19a)
momentum p\ of the scattered photon as a function
of the scattering angle 0. The quantity 2tt(iifnw) = h/mc is known as the Compton wave-
Sec. 4.20 Photons 153
f W, C. Rontgen, “Ober eine neue Art von Strahlen/’ Sitzungsberlchte Med. Phys.
Ges. Wurzburg, 1895, p. 137; 1896, p. 11. These papers have been translated; W, C.
Rontgen, "On a New Kind of Rays,” Science 3, 227 (1896); "A New Form of Radia¬
tion," Science 3, 726 (1896),
| C* G. Bar Ida, “Polarized Rontgen Radiation/7 Phil. Trans. Rot/ Sac. 204, 467
(1905). C. G. Barkla, “Polarisation in Secondary Rontgen Radiation/' Proc. Roy. Soc>
(London) 77, 247 (1906). (The latter paper reports on the double-scattering experi¬
ments.)
W. Friedrich, P. Knipping and M, von Laue, Annalen der Physik 41, 971 (1913).
156 Photons Sec. 4.24
Pi — Pf + Pn + P (momentum) (24a)
&i + Me2 — Ef + En + 'foo (energy) (24b)
Sea 4.25 Photons 157
where E* and E/ are the initial and final energies respectively of the
electron, and where En is the final energy of the nucleus.
These equations taken together give us four conservation equa¬
tions. There are, however, nine variables which characterize the
final situation, namely the nine components of the three vectors
pn and p. The detailed investigation of the permissible range of
these vectors is somewhat involved and we shall not attempt it.
One can show, that for any given direction, the photon can emerge
with an energy ranging from zero to some maximum* This maxi¬
mum actually occurs when both the electron and the nucleus
have the same velocity, say v, after the collision: that this must
be so is immediately obvious if we consider the problem in the
center of mass system. Let us rewrite the conservation equations
for the case when the final electron and nuclear velocities are
indeed equal:
(M + m)y
Pi “ P = y , ,' (24c)
V1 - (v/c)2
(M + m)c2
Ei + Me2 — cp ~ (24d)
VI - (t?/c)z
Multiplying the first equation by c, and subtracting the square of
the result from the square of the second equation gives us
_Ej — me2_
(24e)
1 + (Ei — piC cos 8 )/(Mc2)
where 8 is the angle between the emerging photon and the incident
electron. The above formula thus gives the maximum photon
energy at the angle 8, We note that is is approximately equal to
{Ei— me2), the kinetic energy of the incident electron, which in
turn is equal to eVo. The second term in the denominator in the
right side of (24e) is very small for X-ray tubes since the constant
Me2 ~~ 940 A MeV, for a nucleus of mass number A, is large com¬
pared with Ej, which may range from 1 keV to 100 keV.
^ 2*ne ch /at£ ^
Amin —-= —rr (25a)
o) eVo
e+ + e~ = ny
where the symbol y stands for a photon (a gamma quantum). Sup¬
pose that the electron and the positron are practically at rest when
the reaction takes place (in the laboratory frame), and suppose
furthermore that the reaction takes place iiji free space, far away
from all other particles.
First of all we note that there must be at least two gamma rays:
n ^ 2, since we cannot otherwise conserve energy and momentum.
(If the electron and positron are initially at rest the initial momen¬
tum is zero: if only one photon is emitted, the final momentum
would not be zero.) Let us therefore assume that two photons are
emitted. Since the initial momentum is zero it follows that the
final (total) momentum must also be zero, and the momenta of the
two photons are thus equal and opposite, which means that their
energies, and hence their frequencies are also equal. Let us denote
the frequency by u: the conservation of energy then implies that
= 2me2 or X = = — (27a)
to me
The wavelength of the emitted photons therefore equals the
Compton wavelength of the electron, h/mc — 0.0243 A: to this
wavelength corresponds the rest energy of the electron* me2 =
0.511 MeV.f
For the positrons which have been slowed down and captured
in the material we can assume that the above prediction is valid: the
presence of the other particles in the material may have an effect
but this effect should be small since the atomic binding energies are
very small compared to the rest energy of the electron.
We may therefore look for the two gamma rays formed in the
annihilation. They should emerge in opposite directions, and their
wavelengths should be the Compton wavelength of the electron.
It has been found experimentally that these predictions are cor¬
rect in every detail: annihilation into two gamma rays indeed takes
place-! In addition it has been found that annihilation into three
gamma rays also takes place.
f Note that the quantity h/mc = 0.00386 A is also often called the Compton wave¬
length.
\ See, for instance, O, Klemperer, "On the Annihilation Radiation of the Positron/'
Proceedings of the Cambridge Philosophical Society 30, 347 (1934).
160 Photons Sec♦ 4,29
f Very recently some experiments have been performed which seem to indicate that
the weak interactions axe not invariant under the exchange of particles for anti-particles.
This implies that whereas the strong and electromagnetic interactions might obey the
symmetry principle mentioned, the weak interactions do not. Since the strong and
electromagnetic interactions are the dominant interactions in the world, it is perhaps
fair to say that the symmetry principle is almost (but not quite) true.
J For the discovery of the antiproton* see O. Chamberlain, E. Segre, C. Wiegand and
T. Ypsilantis* “Observation of Antiprotons,” The Physical Review 100, 947 (1955).
Sec. 4.30 Photons 161
more, and sometimes less energy becomes available for light emis¬
sion. What is even harder to understand is why two totally dif¬
ferent atoms, say a sodium atom and a mercury atom, which emit
light of different frequencies, wna and <OHg> should emit wave trains
of total energy and Hg> respectively. From a classical
standpoint the appearance of the universal constant of propor¬
tionality ii is very mysterious.
If we think about the totality of all the many experimental facts
which we discussed in Chapter 3 it is very clear that the phenomena
cannot be understood on a classical basis. Let us, however, try to
forget temporarily what we already know about the emission and
absorption processes, and instead concentrate on the study of
“isolated” photons. We consider wave trains which have been
emitted from some source, and we study these wave trains with
a photocell in the experimental region.
QassicaDy, we can understand these facts easily: each wave train Screen
arriving at the mirror is split into two parts, with
Consider now what happens when a single wave train arrives at slit Beam
the mirror* On the classical model, we expect it to be split into I splitter
two parts, and in such a way that the energy carried by the trans¬
mitted part of the wave train is one-half the energy of the incident Light -~a
Photocell 2
wave train. Therefore, the photocell 2 should never click! source
This prediction* based on classical theory, is in flat contradiction
p
with experience. The light which goes through is still blue, of
i
frequency cj, and the register of cell 2 does click as long as
Hw > Emin, which shows that the energy of the transmitted light Photocell 1
comes in packets of Huy. What does happen when the mirror is
inserted is that the counting rate is only half its value in the absence Fig. 35A Schematic figure to Illustrate the discus¬
of the mirror. sion in Sec. 35. The light beam from the source is
divided into two parts by the beam splitter, which may
be a hall-silvered mirror. Are the individual photons
36 How convincing is the evidence that photons cannot be split split?
with the experimental arrangement shown in Fig. 35A, or with any
other arrangement of a similar kind? The evidence is exceedingly
good* and, in fact, photon-splitting experiments go on all the time.
Every optical device in which we have a photoceD, or a photo¬
graphic plate, can be regarded as an instrument in which we try
to split photons, but fail* The simplest observation of this kind is
the observation of the photoelectric effect at different distances r
from the light source. If an atom is like an antenna, then it should
emit light in the form of a spherical wave train. The intensity of
the emitted light is proportional to 1/r2, and on the classical pic¬
ture, the amount of energy carried by a single wave train through
a unit area at a distance r is proportional to 1/r2. Therefore, since
the photocell has some definite cross section, it would seem that
by merely placing the photocell at a sufficiently large distance, the
fraction of the energy in a wave train which can possibly be active
in the photocell can be made as small as we please, and for a fixed
retarding potential, the photoceD should cease entirely to register
as soon as the distance exceeds a certain limit. This is most cer¬
tainly not what we observe: all that happens is that the counting
rate of the photoceD diminishes as 1/r2. Perhaps the most striking
example is the observation of the photoelectric effect for light from
a very distant star. The wave train was emitted thousands of years
ago, and should have spread out over a large portion of space.
Only a minute fraction of the energy carried in the wave train
should be able to reach the photoceD through the telescope. Never¬
theless, the amount of energy given to the electron in the photoceD
is found to be just as if the source were a lamp close to the
photoceD.
166 Photons Sec, 4.37
A0 = e-*»* (40a)
iwa
A = /(r,0)e_w exp sin +
where
t G. I. Taylor, “Interference Fringes with Feeble Light,1' Proc. C ambridge Phil, See
15, 114 (1909).
170 Photons Sec. 4.45
After the photon has been emitted it spreads out like a spherical
shell. By the time the wave arrives at the detector the energy
carried by the wave has spread out over a very large region in spaee3
say within some spherical shell of radius one light year. How is it
then possible that the entire amount of this energy can suddenly
become concentrated within the photocell in case the cell does
register? It should take more than a year for the energy within the
“far side" of the shell to reach the cell, otherwise we violate the
principle that no signal can propagate faster than the velocity of
light.
The fallacy in this line of reasoning is to believe in the classical
of radiation expression for the energy density in terms of the electric and mag¬
netic fields. We must remember that the whole purpose of intro¬
Pig. 47A The atom, shown in the center, emitted ducing the concept of the electromagnetic field in physics is to
fight one year ago. The spherical shell of radiation
accordingly has a radius of one light year. It is about describe the interactions between charges* In Vol. II of this series
to reach the photocell at right. If the ceil registers, we learned that it is a convenient concept, and we also learned
the entire energy of the wave is suddenly concentrated that it is sometimes convenient to imagine (in typical macroscopic
in the cell. How is this possible? How can the energy situations) that energy is distributed in space with a density propor¬
from the far side of the shell reach the photocell in
tional to the square of the field amplitude. There was, however,
less time than two years?
The J<paradox’1 evaporates if we give up the das- no physical fact discussed in Vol, II which says that we must take
sical idea that the energy density is proportional to the this idea literally, and we now know that the classical expression
square of the field amplitude. According to quantum for energy density refers to the average energy density which we
mechanics the transfer of energy from the atom to the will observe for a very large number of photons, but it does not
cell is governed by a probabilistic law, and the square
describe the energy density associated with a single photon.
of the field amplitude must be interpreted as a prob¬
ability density. The real question is this: what laws govern the transfer of energy
from an atom in the source to an electron in the detector? This is
what we are studying, and we have now discovered some features
of these laws.
1) A set of reprints has been published by the American Institute of Physics, 335 East
45th Street, New York, N.Y., under the title: Quantum and Statistical Aspects of Light
As the title indicates, these papers deal with various properties of photons* and the
reader may find some of them interesting. A short survey of the literature is included.
2) We refer again to the books The World of the Atom, edited by H. A. Boorse and
L. Motz, Vols. I and II, (Basic Books, Inc., New York 1966), for translations and reprints
(with editorial comments) of many early papers relevant to the subject matter of this
chapter.
Problems Photons 175
3) The following articles in the Scientific American can properly be read at this stage
of our discussion;
a) G, E, Henry: "Radiation Pressure/' June 1957, p. 96.
b) W. H, Jordan: "Radiation From a Reactor/' Oct. 1951, p. 54 (discusses the
Cerenkov radiation).
c) G, Burbidge and F. Hoyle: "Anti-Matter/' April 1958, p. 34,
d) G. B, Collins: "Scintillation Counters/' Nov., 1953, p, 36.
Problems
compute (wo — w)/w for the 113 keV gamma ray emitted by the hafnium
isotope 72Hf177.
The formula above describes the recoil-effect In photon emission. As we
see, the emitted photon is always of smaller frequency (in the rest frame of
the emitter) than the frequency w0 which it would have for an infinite Mf,
For optical photons emitted by atoms the effect is extremely small,
3 On the basis of the data given in the graph in Fig. 23A, determine h/e
to the accuracy permitted by the accuracy of the graph. (The velocity of
light is regarded as known.)
of the gamma ray emitted from the excited Fe57 nucleus is 14,4 keV,
Knowing this, can you relate the velocity to the frequency in the two scales
shown on the graph?
ft Pu ft Po
absent absent horizontal absent
absent horizontal vertical absent
circular horizontal vertical circular
circular horizontal horizontal circular
circular horizontal vertical absent
A refinement of Fig. 39A. Polarization filters have
been placed as follows: in front of the source, Pv
In the table above “horizontal” refers to a filter which lets through only and in front of the upper and lower slits, and Po in
front of the observer.
light polarized in the horizontal direction, “vertical” refers to a filter which
This figure refers to Prob. 12 on this page. What
lets through only light polarized vertically, and “circular” refers to a filter kind of fringes will we see, for various choices of
which lets through only left-circularly polarized light. filters?
Chapter 5
Material Particles
Sections 1 to 15 The de Broglie Waves 180
Problems 217
Chapter 5 Material Particles
tSee, for instance, PSSC, “Physics" (D. C. Heath and Company Boston, 1965),
Part IV,
f The author believes that the historical order of discovery can be understood
theoretically, on the basis of the smallness of the fine structure constant a.
Sec* 5.3 Material Particles 181
particle, but nothing was known about its wave properties, That
the photon has some corpuscular properties was, however, known.
Let us play a make-believe game in which we ask whether a
material particle, say an electron, might possibly have some prop¬
erties of a wave. To find out we must turn to experiments, but
before we do that we should try some theoretical ideas first to see
what we might expect.
ner we would expect that this wave must be moving in the same
direction as the particle* We represent the wave by the complex
wave function
^(x,t) = A exp (ix - k — icof) (4a)
1 _ dk^ _ dto dv
v do) 1 dv dk
dk _ 1 dw
dv v dv
ftk = p
^(x',0 = A' exp (?xy * k' — i<oY) (6a) Louis Victor de Broglie. Born 1892 in Dieppe, France.
De Broglie first studied history, but later changed over
where A' is a constant amplitude which may or may not be equal to physics. He obtained his doctoral degree from the
to A, University of Paris in 1924. Since then he has held
Let us suppose that the primed frame is the rest frame of the positions at the Sorbonne, at the Henri Poincar£ Insti¬
tute, and at the University of Paris. He received the
particle. In this frame we thus have k' = 0, p' = 0, and Ef = me2,
Nobel prize in 1929.
Let us furthermore suppose that the relation (5b) holds in the rest The title of de Broglie's doctoral thesis was iJRe-
frame (but perhaps not in any other frame): under this assumption cherches sur la Th4orie des Quanta,’' and it contained
we have to' = me2/ft. the essence of his ideas about matter waves. (Photo¬
graph by courtesy of Physics Today.)
7 The phase of the wave is given in any one frame by the expres¬
sion (x • k — <of), and this quantity we assume to be an invariant:
if the phase has a certain value at the point x', at the time f, in the
primed frame, then the phase should have the same value at the
corresponding point x, at the corresponding time t, in the unprimed
frame, We defend this assumption by pointing to the periodic
nature of the wave. If the phases for two events in space-time
differ by an integral multiple of 2*it in one frame of reference, then
the phases of the same wave must differ by the same integral mul¬
tiple in every frame of reference. From this it follows that the
184 Material Particles Sec* 5.7
x * k — uf = — H — — f (7a)
f - (x • v)/c2
(7b)
Vi - (u/c)2
and if we insert this expression into (7a) we obtain
Since this relation must hold for every x and every t it follows that
w = (7d)
VI - («/c)2
k _ (7nv/K)
(7e)
>
C*
rH
V
1
me2 mv
(7f)
vT VI - Wcf
Taken together the equations (7d) to (7f) lead to
E (7g)
We accordingly recover the result (5e) and furthermore we see
that the equation (5b) which was introduced ad hoc in Sec* 5 indeed
must be generally true if it holds in the rest frame. This line of
reasoning thus shows us that the relations (7g) are consistent with
h __ 2?r
X
p k
X= (MV1-fr/c)2
\mc / (u/c)
shows us that X decreases as the velocity v increases* For a fixed
velocity v the wavelength X is inversely proportional to the mass m.
he (hc/E)
VE2 “ m2c4 Vi - K /£)2
X = £ (9b)
E = T + me2 (9c)
186 Material Particles Sec* 5.i0
X = = k - 1 - (9d)
\/T(T + 2me2) V§mT VI + T/(2mc2)
For a fixed rest mass m the wavelength A decreases when the
kinetic energy T is increased. For a fixed kinetic energy T the
wavelength A likewise decreases as m is increased.
In the limiting case when the velocity of the particle is very small
compared to c, the ratio T/mc2 becomes very small. Setting this
ratio equal to zero in Eq* (9d) we thus obtain an expression for the
wavelength A in the non-relativistic approximation
X gs h - - eg A (9e)
y2mT mV
particle with the smallest possible mass, namely the electron, and
furthermore we should try to keep the velocity as small as is feasible.
Since we want to consider the case when the velocity is very small
we can employ the nonrelativistie approximate expression (9e) for
the de Broglie wavelength. If we rewrite this expression specifically
for electrons, of mass m, and kinetic energy T, we obtain
x = = A“gs“,m (lla)
It was implicit in the theory that beams of electrons like beams of light
would exhibit the properties of waves, that scattered by an appropriate
grating they would exhibit diffraction, yet none of the chief theorists
mentioned this interesting corollary, The first to draw attention to it
was Elsasser, who pointed out in 1925 that a demonstration of diffraction
would establish the physical existence of electron waves. The setting of
the stage for the discovery of electron diffraction was now complete.
the first with electrons and the second with X-rays. The samples
were in both cases aggregates of small crystals of white tin. The
similarity in the pattern of circles is very striking. Even if we know
nothing about the detailed theory of diffraction of waves in lattices,
a glance at these two photographs will at once convince us that
X-rays and electrons are diffracted in the same way.
t D, P. Mitchell and P, N, Powers, "Bragg reflection of slow neutrons,” The Physical Fig. 16C Three-dimensional lattice. The edges of
Review 50, 486 (1936), See also E. O. Wollan and C. G. Shull, “Neutron diffraction the unit cell are again drawn heavier, The position
and associated studies," Nucleonics 3, 8 (1948), vector of any lattice point is some linear combination
| Sections 16-22 can be omitted in a first reading, but do not fail to look at the photo¬ with integral coefficients of the vectors ©i, ©2, and ©3.
graphs in Sec, 22, (These vectors are not necessarily perpendicular,)
192 Material Particles Sec. 5,18
x0 — niei (17c)
ei * (uj — up)
= mi (18d)
X
where mi is an integer. This we might have concluded imme¬
diately* The waves from any pair of atoms arrive in phase with
each other if and only if the waves from two neighboring atoms
arrive in phase with each other* and this is just what the condition
(18d) says.
Making use of the de Broglie relation we can rewrite (18d) in a
physically interesting form, as follows. Let p* be the incident
momentum, and let po be the momentum in the scattered beam*
We then have
_ Pi u0 _ po
X “ h * X ~ h
and the condition (18d) can be written
assume that the number is large and the scattered particles there¬
fore emerge in very well-defined directions, as given by (19a) and
(19b). These equations define a set of cones, one for each integer
mi. These integers are, of course, subject to the constraint
by the atoms in the surface layer, and the theory for the two-dimen¬
sional lattice therefore applies.
ei - qi = ht ©2 ’ qi = 0, e3 * qi = 0
ei - q2 = 0, e2 ■ q2 = h, e3 ■ q2 = 0 (21c)
ei * q3 = 0, e2 ■ q3 = 0, e3 * q3 = h
(22b)
where mu m2, and m3 are any integers, and where qi, q2l and q3
196 Material Particles Sec. 5,23
are the vectors discussed in the preceding section, for some par¬
ticular orientation of the crystal lattice* The above equations do
have solutions, and we see that the diffracted rays emerge along
a set of cones centered about the incident direction,
Fig, 14A shows schematically how a diffraction experiment based
on the above theory is performed. In X-ray work the sample is
often a small amount of a fine powder consisting of many small
microcrystals. This is the way the photograph 14C was obtained.
The lines on the film strip are the intersections of the cones [defined
by the conditions (22a) and (22b)] with the film strip.
We can easily understand that if the sample is too small, in the
sense that it does not contain sufficiently many crystals, then the
distribution of the diffracted rays along the cones will be very non-
uniform, We will not see continuous circles on the photographic
Figs, 22A-B The upper photograph shows electron plate, but rather individual dots. This effect is shown very beauti¬
diffraction rings obtained with the method illustrated
fully in the photographs 22A and 22C, These photographs, which
schematically in Fig. 14A. The sample consisted, as
for the photograph 14B, of small white tin crystals should be compared with the photograph I4B, show the diffraction
deposited on a thin film of silicon monoxide. The of 100 keV electrons by tin crystals. In this case the electron waves
lower photograph shows the appearance of the sample completely penetrate the small crystals. An electron microscope
in the electron microscope (8 mm corresponds to was used as a diffraction apparatus. Photographs 22B and 22D,
1000 A). The dark spots are the images of the crys¬ taken with the same electron microscope, show the appearance of
tals (their darkness depends on their orientation).
The lightest spots are pits in the SiO once occupied the sample.
by crystals which have disappeared in the preparation
of the sample. The average size of the crystals is
about 600 A.
In obtaining the diffraction photograph the electron There Is But One Planck's Constant
beam was limited to a comparatively small area of the
sample. In view of the theory in Sec, 22 we expect 23 The subheading above perhaps surprises the reader. Of
to see individual dots, as the photograph indeed course there is only one Planck's constant, by definition. What
shows, rather than well-developed rings,
deep conclusion does the author intend to draw from this trivial
(Photographs by courtesy of Dr. W. Hines and Pro¬
fessor W. Knight, Berkeley.) fact?
The entirely nontrivial point is that we do not need more than
one “Planck-type constant” in physics. Consider the de Broglie
relation written in the form
h = Ap (23a)
where p is the momentum of the particle, and A its de Broglie wave
length. Both p and A are independently measurable quantities,
and by measuring a pair of corresponding variables (p,A) we can
determine Planck's constant ft. It is a remarkable empirical fact
that we always get the same value for ft, irrespective of which kind
of particle we are observing, and that this is so is not a triviality.
The reader is perhaps not impressed by this. We could, after all,
derive this relation on the basis of some very simple ideas. Let us,
however, examine the premises of our derivation.
Sec. 5.24 Material Particles 197
E = p = #k (24a)
where E is the energy, p the momentum, to the frequency and k the
wave vector, and where # is a constant defined by
£0 = me2 = #w o (24b)
Figs. 22C-D These two photographs were obtained
in terms of the rest energy Eq and the “rest frequency” coq. in the same way as the photographs 22A-B. The
How did we conclude that the constant # is actually Plancks sample here consists of smaller crystals (average size
constant? By guessing. The relation E = holds for photons, about 200 A), and the diffraction pattern arises from
and it is tempting to guess that it will also hold for material particles. a much larger number of crystals. The rings are
better developed, although individual dots can be
But that is just the crucial point: Does the first relation (24a) really
seen. The photographs 22A and 22C should be com-
hold for all material particles? pared with the photograph 14B, in which individual
Therefore, what we really derived in Secs. 3-5 was that the con¬ dots can no longer be seen. The latter photograph
nection between energy, momentum, frequency and wave vector is was taken with an electron beam traversing a much
larger portion of the film. We accordingly expect a
E = Cto, p = Ck (24c) well-developed ring pattern since all orientations of
the crystals are well represented in the sample.
where C is a constant characteristic of the particle, and this constant The electron energy was 100 keV for the pictures
is defined, for instance, by 14B, 22A and 22C. This corresponds to a wavelength
of about 0.04 A.
(Photographs by courtesy of Dr. W. Hmes and Pro¬
(24d)
fessor W. Knight, Berkeley.)
If this were not the case the theory of the'elementary particles and
their interactions would have to be very different indeed.
How well is the hypothesis that C = H for every kind of particle
verified experimentally? Direct experiments, analogous to the
experiments of Davisson and Germer, or Thomson, have been
performed only for a few kinds of particles. Such experiments are
very readily interpreted as tests of the relation h = but they
are naturally of limited precision* They support our belief in the
universality of the relations (24a), but the real basis of our faith in
these relations is the general success of quantum mechanics. There
exists an enormous amount of experimental evidence which sup¬
ports the relations (24a) indirectly. The interpretation of this
evidence is not always as clear-cut and simple as in the case of elec¬
tron diffraction in crystals, but in its totality it is very convincing.
Our belief that the relations (24a) are exactly true is somewhat
analogous to our belief that the relation Eq = me2 is also exactly
true* The direct evidence for this latter relation is quite strong,
but the totality of the indirect evidence, concerning the general
validity of the ideas of special relativity, is what really convinces us*
There is not the slightest hint in our experimental material that the
relations (24a) or the relation Eq = me2, might be only approx¬
imately true* We assume that they are true exactly, and we regard
them as cornerstones of physical theory.
Let us recall our discussion in Sec* 12, Chap. 2* We argued that
because of the fundamental role played by the constants c and #
in relativistic quantum physics one might well select a system of
units in which ft = c — 1. Such a system of units would clearly
not make much sense if there were a different Planck-type con¬
stant C for every particle* Since we believe that there is only one
such constant it means that, for instance, mass, energy and fre¬
quency are always related in the same way, and we can regard
the words “mass," “energy," and “frequency” as different names
for the same thing*!
other. Let their momenta be pi, p£, ■ * ■ , pi, and let their energies
be Ei, , EJi When we say that the particles are initially
“well separated” from each other we mean that the particles initially
move in such a way that all interactions between the particles are
effectively absent at the early time. This idea makes sense if we
assume that the inter-particle forces tend rapidly to zero as the
distance between the particles increases. At first each particle
therefore moves as if the other particles were absent. As time
goes on the particles may converge in a “collision region,” and
the inter-particle forces come into play. An interaction takes place,
and in this process the particles are deflected Furthermore some
of the particles may be destroyed, and new particles may be
created.
If we wait for a sufficiently long time the particles involved in the
collision event will again be dispersed, and the interactions between
the particles will effectively cease for the simple reason that they
are no longer close together. At some late time each particle will
move as if the other particles were absent. Let the momenta of
the particles after the collision event be pL . .., pj', and let their
energies be E", £2 * * ■ « ? £"■
The conservation laws then read
if if
z*
T—l s—1 r=l 5=1
The total initial energy equals the total final energy, and the total
initial momentum equals the total final momentum. The condition
that the particles effectively do not interact with each other at the
“early time” or at the 'late time” is essential; because otherwise the
total energy would not be equal to the sum of the energies of the
individual particles. If the particles do interact with each other we
have to include an “interaction energy” in the expression for the
total energy.
The reader should note that the particles need not be elementary
particles; they can just as well be composite particles, like atoms or
nuclei. When we discuss collision events we mean by “particle”
any object which is reasonably stable so that it can be assigned a
momentum, an energy and a (rest) mass as soon as it is well sepa¬
rated from other similar objects. As an example we can consider
the collision between a neutral helium atom and an electron. Sup¬
pose that the helium atom is ionized in the collision. There are
then two initial particles, namely the electron and the neutral helium
atom. There are three final particles, namely two electrons and
200 Material Particles Sec. 5.27
one singly-charged helium ion, (This is* of course, not the only
possible outcome of the collision event. The helium atom might
lose both its electrons in the event* or it might lose none. Further¬
more the event can lead to the emission of one* or several, photons,)
K = V K'
/ j
(27a)
T—l 8=1 r—1 8=1
The sum of the initial frequencies equals the sum of the final
frequencies, and the sum of the initial wave vectors equals the sum
of the final wave vectors. These conservation laws are com¬
pletely equivalent to the conservation laws (26a), Each set of laws
implies the other. This is so because there is only one Planck's
constant.!
30 Now it may happen that the reader does not have any firm
opinions about the properties of a “classical wave,” and therefore
the statement that an electron is not a classical wave might appear
a bit colorless. What we have in mind here is that for a classical
wave the absolute square of the wave amplitude at a given instant
of time, and at a given point in space, represents a physical quantity
such as a charge density, or an energy density. This idea is analo¬
gous to the idea in classical electromagnetic theory that the squares
of the electric and magnetic fields represent energy densities.
Suppose, for instance, that the square of the wave amplitude is
proportional to the charge density. We could then compute the
flux of charge into one of the counters, and since the wave is
202 Material Particles Sec. 5,31
32 We have said that the wave is divided into two (or several)
“parts” in the diffraction experiment shown in Fig. 29A. The
reader may then ask: Can the wave traveling in the direction of
counter Cj be made to interfere with the wave traveling in the
direction of counter C4P If an electromagnetic wave is divided by
a half-silvered mirror, then the two “parts” can oertainly interfere
with each other, and we expect the same behavior from the de
Broglie waves. In other words, if we somehow deflect the wave
f This may not be true in practice, but, for the sake of argument, we can assume that
every incident electron goes into either counter Ci or counter C4.
Sec. 5*33 Material Particles 203
traveling in the direction of counter C4, and “mix” it with the wave
traveling in the direction of counter Ci, will we then see interfer¬
ence effects?
The answer is that we certainly expect to see interference effects*
On the other hand it must be admitted that it would be very diffi¬
cult in practice to perform this experiment exactly as described,
with electrons* Fortunately we do not have to do this experiment,
because the very fact that we can observe electron diffraction at all
with the crystal is conclusive proof of the reality of the interference
effects* Each atom in the crystal surface gives rise to a diffracted
wave when “illuminated” by the incident wave, and all these
diffracted waves combine to produce the overall interference
pattern which we observe with the crystal* What does it mean
that the waves diffracted by the individual atoms “combine?”
How do we describe the “combination?” We describe it by adding
the amplitudes of all the separate waves to obtain the total ampli¬
tude of the wave emerging from the crystal* The square of this
resultant amplitude is an intensity variable, to be understood
quantum mechanically, which describes the response of a detector*
= -w2^(x,t;p) (37c)
and similarly for the second derivatives with respect to the other
two space coordinates %2 and £3.
We thus obtain, taking into account the relation (37b),
r* = JL + _?L + _?!_
03Ci2 GX22 dXs2
in empty regions in space-time, Le*, far away from all other par¬
ticles* Similarly the homogeneous Maxwell's equations, i.e., with
the current density and charge density both equal to zero* describe
the propagation of electromagnetic waves only in regions free of
charges and currents, Le,? in regions free of other particles,
i£(x,t) = A' exp (ix ■ p' - ito'*) + A” exp (ix * p" - to"f) (39a)
where A' and A" are two arbitrary complex constants, also satisfies
the differential equation (37e). In other words,
= -to2^(x,*) (39b)
We shall not prove this theorem here, and we shall not really
depend on the theory of the Fourier integral in our discussion in
this book* In due time the reader will learn how to state this
theorem precisely, and how to prove it, in her calculus course* Our
aim here is to discuss the physical implications of the theorem, and
thereby provide the reader with a strong “physical” motivation for
learning about the Fourier integral* It is of central importance in
physics,
yp = + c2yp2 (47a)
facts, such as
We shall not list all the trivial theorems here, as the author is
confident that the reader's intuition will not lead him astray.
What is the virtue of introducing the concept of an abstract
complex vector space? The answer is that in our study of mathe¬
matical theories we encounter over and over again sets of elements
which, among any other properties which these sets may have,
have the particular property of satisfying all the axioms of an ab¬
stract complex vector space. When we encounter such a set we
need not list the properties of an abstract vector space anew but
we can simply say that the set is a complex vector space, and then
everyone who knows the axioms for a vector space immediately
knows quite a bit about the set.
1) For the history of the topics discussed in this chapter we again refer the reader
to the books mentioned at the end of Chap. 1. (Items 3 and 5.)
2) There exists a vast literature on the mathematical theory of linear partial differen¬
tial equations. The author does not expect the reader to go deeply into this theory
at this time, but he wants to mention one treatise which has played an important role
in physics, namely R* Courant and D* Hilbert; Methoden der mathematischen Physik,
vols. I and II (Verlag von Julius Springer, Berlin, 1931 and 1937.) This work has been
translated under the title; Methods of Mathematical Physics, vols. I and II (Interscience
Publishers, Inc*, New York, 1953 and 1962)*
The second volume discusses partial differential equations* Hie first discusses a
variety of subjects of interest in physics, Such as Fourier analysis, the theory of matrices
and vector spaces, the calculus of variations, and the theory of certain ordinary linear
differential equations which appear in many physical problems.
It so happened that important developments in mathematics, which Later were found
to be '‘made to order*' for quantum mechanics, took place about the time quantum
mechanics was discovered. David Hilbert at the University of Gottingen played a
central role in these developments, and the infinite-dimensional vector space in terms
of which quantum mechanics is formulated today is called Hilbert space, after him.
Hilbert did not originally develop his theory of linear spaces for physical applications,
but the discovery of quantum mechanics naturally stimulated mathematical inves¬
tigations of the questions raised by the physical applications. The period was one of
considerable interaction between mathematicians and physicists.
The theory of quantum mechanics from a mathematician's standpoint is presented
in J* von Neumann: Mathematische Grundlagen der Quantenmechanik. (Verlag von
Julius Springer, Berlin, 1932* Reprinted by Dover Publications, New York, 1943.)
An English translation has appeared under the tide: Mathematical Foundations of
Quantum Mechanics (Princeton University Press, 1955)*
3) Matrix mechanics is discussed in most advanced textbooks on quantum mechanics.
For an introductory account of quantum physics in which the algebraic approach is
discussed and used we refer to R* P* Feynman, R. B* Leighton, and M* Sands: The
Feynman Lectures on Physics, vol* III (Addison-Wesley Publishing Co*, Inc*, 1965)*
This is the final volume in a series of three books on basic physics. The presentation
in these books is magnificent, and the reader is strongly advised to become acquainted
with them,
4) We say very little about solid state physics in this volume. Among introductory
books on this subject we mention C. Kit tel: Introduction to Solid State Physics 3rd ed.
(John Wiley and Sons, Inc., New York, 1966)* Among many other topics the reader
will find discussions of crystal structure, diffraction theory, and the theory of phonons.
Concerning crystals the reader should note the long article on the subject in the
Encyclopaedia Britannica under the heading "crystallography.”
5) The reader may find the following articles in the Scientific American interesting:
a) K* K. Darrow: "The quantum theory,” March 1952, p, 47
b) K* K. Darrow; "Davisson and Germer,” May 1948, p. 50
c) E. Schrodinger: "What is matter?” Sept. 1953, p* 52
Problems Material Particles 217
Problems
helps in your search for good ideas; You know that your quantum mechanical Grating
theory must in this case yield a known result*
(c) According to classical dynamics the particle will not be reflected at
the interface, but only refracted. Light incident on the interface between
two different dielectrics is both reflected and refracted Express your opin¬
ions about what the situation ought to be in the case of a quantum mechan¬
ical particle, he., a real particle.
to tlie crystal surface, at a distance of 5 cm from the surface. This figure must
be drawn in the correct scale, and all diffracted rays must be shown.
Theory of Measurements
Problems 262
Chapter 6 The Uncertainty Principle and
the Theory of Measurements
Position poorly defined length and momentum are then related by A = 2 m/p, and we do
Momentum well defined not have to distinguish between wave vector and momentum.
We shall argue in terms of pictorial representations of waves,
(A)
and for this purpose we have plotted four particular wave trains of
finite length in Figs. 5ABCD. (The ^-coordinate is the abscissa in
these figures.) Now the reader should note that the wave function
$(x,0) is in general a complex-valued function, which fact creates
Position better defined problems when we want to represent it graphically. We can, how¬
Momentum less well defined
ever, plot the real part and the imaginary part of the function sepa¬
(B) rately, and the reader can interpret the Figs. 5ABCD as showing
either the real or imaginary parts of r//(x,0) ■
The graphs show "interrupted sine waves,” described by the
function sin (px) in the region in which the wave function does not
Position well defined
Momentum very poorly defined
vanish. The wave is, however, not really a pure sine wave, because
it is “cut off” at both ends. For this reason the wavelength (and
the momentum) are not precisely defined: these quantities can be
(C)
precisely defined only if the wave is a pure sine wave.
Looking at the figures 5ABCD we can see very clearly that the
better the position is defined, the more poorly is the momentum
Position very well defined defined. Let us denote the uncertainty in the position x by Ax.
Momentum very poorly defined As a rough measure of the uncertainty in position we may take the
length of the wave train: if the wave train consists of n full waves
(D) we have
Ax ~ nA = (5a)
Figs. SABCD To illustrate our discussion of the P
position momentum uncertainty relation. A well-de¬
fined position requires a short wave train. A well- where A is the wavelength. Now it is dear that the wavelength
defined momentum requires many well-developed must be better defined the larger the number of full oscillations in
sinusoidal cycles. The two requirements are in con¬ the wave train. As a rough measure of the fractional uncertainty
flict with each other. in the wave length we may take the quantity
1 AA _ Ap
n X ~~ p
Ax Ap 1 (5c)
6 The relation (5c) is the form the uncertainty relation takes for
the particular kinds of waves shown in Figs, 5ABCD. The general
uncertainty relation, which holds for all waves, is in the form of an
inequality* To convince the reader of this fact we show another
Fig. 6A For the above wave train the position is as
kind of wave in Fig. 6A* It is clear that for this wave the uncer¬ poorly defined as in Fig. 5A, The momentum is, how¬
tainty in the position is about the same as in Fig, 5A. The un¬ ever, a/so very poorly defined, and certainly much
certainty in the momentum (or wavelength) must, however, be more poorly defined than in Fig. 5A. The correct
much larger for the wave in Fig. 6A than for the wave in Fig, 5A* uncertainty relation must be an inequality: it is pos¬
sible to imagine wave trains for which the uncertain¬
The correct position-momentum uncertainty relation must there¬
ties in both momentum and position are arbitrarily
fore be of the form large.
Ax Ap > 1 (6a)
The reader will recognize that this is the uncertainty relation
which we discussed very briefly in Chap* 1.
The inequalities (7a) and (7b) are the uncertainty relations for
waves (= particles) in three-dimensional space*
9 These ideas can be given a precise form, and we can relate the
concentration of the function A(p) to the concentration of the func¬
tion ^(x,0). The result is an uncertainty relation; the precision
with which the position is defined is inversely related to the preci¬
sion with which the momentum is defined. Since we have pro¬
mised the reader not to depend on the theory of the Fourier integral
in this book we shall not present a rigorous derivation of the un¬
certainty relations.! The important thing for us is to understand
dp > 1 (10a)
With the cgs-system of units this criterion reads dp ^ and
this is the same criterion which we discussed in Secs. 20-26 of
Chap. 1.
Ap — ™ (11a)
Fig* 11A We attempt to produce a narrow beam of
electrons by limiting a broad beam incident from the If we assume that Ap is small compared to p we can restate (Ha)
left by the slit in the screen at left. The beam is in terms of the uncertainty A6 in the angle 9 (with respect to the
diffracted at the slit, and the uncertainty M in the incident direction) with which the electrons emerge* and we have
angle at which the electrons leave the slit is inversely
proportional to the width d of the slit. The size of the
spot on the screen at right is given by A* — d + L M.
P
Let Ax measure the size of the spot produced on the screen at
right. The magnitude of Ax is determined by two things: the size
of the opening in the screen at left, and the spreading of the wave
by diffraction at the slit. (See Fig. 11A.) We can therefore write
“tracks” of the electrons between the two screens are thus narrow
and well defined from a macroscopic point of view.
2m r
rp H (14a)
or, for definiteness, let us assume that
rp — (14b)
'3£\ _ Po _ _ o
(14d)
dp jp~Po m *
Solving for p0, and defining r0 = fi/po, we thus obtain
e2m *2
Po = ro = (14e)
IT’ e2m
and
Time poorly defined nucleon, confined in a nucleus within a sphere of roughly the radius
Frequency very well defined fo = 1.2 X 10^13 cm, The uncertainty relation then tells us that
the momentum must be at least of order p ~~ ftfro9 and hence the
(A)
kinetic energy of the nucleon must be of order
e^~M^ ~10Mev m
Time better defined Since the nucleon is bound in the nucleus the average of the
Frequency less well defined
potential energy, denoted (17), must be negative and larger in
(B) magnitude than the kinetic energy, and we may conclude that
Time well defined This estimate is very rough, but it does give an idea of the order
Frequency very poorly defined of magnitude involved.
Aw At > 1 (18c)
r Aw — 1 or r AE ~ ji (19a)
an amplitude such as the one shown in Fig, 18B} for which both the
time and the frequency are very poorly defined.
The relations (19a) are just the relations we derived in Secs, 20-
23, Chap. 3, by a seemingly different line of reasoning. If the
reader thinks more about this matter he will notice that the basic
ideas of our two derivations are not really so different. Our dis¬
cussion in Chap. 3 could be characterized as “Fourier analysis in
disguise.”
in the following that our beams are of such a low intensity that only
one particle is in transit at a time. In actual practice we would
not deliberately limit the intensity of the beam, but, on the con¬
trary, we would try in general to work with the highest intensity we
could achieve.
Nu (24b)
P12
N
where is the number of elementary experiments in the sequence
in which both counters 1 and 2 clicked.
(c) The counter 1 clicked an average number of p(l;2) times per
click of counter 2. This number is defined by
Sec. 6.25 Uncertainty Principle and Theory of Measurements 237
N12
P(l;2) (24c)
n2
where N2 is the number of times counter 2 clicked, and N12 the
number of times both counters 1 and 2 clicked.
pose that there are two independent lamps in the source, say a
sodium lamp which emits yellow photons and a mercury lamp
which emits blue photons. A photon in a particular single experi¬
ment can thus be either yellow or blue, and the color is one of the
variables characterizing the photon which we might determine in
the experiment. Suppose that we do this for a long sequence of
photons. We can then report that the probability is pi that the
photon in any particular experiment is blue, and that the prob¬
ability is P2 that the photon is yellow. We assume that the inten¬
sities of the two lamps are kept steady so that these probabilities
are reproducible: if we perform several runs of large numbers of
repeated experiments we always find the same probabilities pi and
P2 in each run.
Are we willing to say, under these circumstances, that the pho¬
tons are all prepared “in the same way” in the source? Whether
this is an appropriate mode of expression or not is not immediately
obvious. One might argue that our arrangement with the two
lamps introduces an element of chance into the preparation process
which could easily be avoided if we would carry out our observa¬
tions with only one lamp operating at a time. Perhaps we should
not say that the photons have all been prepared in the same way
unless we are assured that the photons are in some sense identical
to the highest possible degree?
The difficulty with such a position is that we would then have
to decide, for each kind of experiment, whether the particles are
“identically prepared to the highest possible degree” or not. This
2000 3000 4000 9000 6000 7000
is obviously not a trivial problem. Furthermore, the two-lamp
WAVELENGTH-ANGSTROMS
experiment is really just as respectable as the one-lamp experiment
Fig. 24C Graphs showing detection efficiency of the in the sense that the probabilities pi and P2, as well as any other
photomultiplier tube in Fig. 24B. Note the curve probabilities describing the responses of detectors, are stable and
labeled "quantum efficiency." It represents the reproducible* This, of course, is the essential thing for any experi¬
probability for the detection of a photon, as a function ment in which we determine counting rates and probabilities, and
of the wavelength. The maximum probability is about
unless the source is stable in just this sense the discussion in Sec. 25
25 percent, which is a very high efficiency for a photo*
tube. The graph is taken from the manufacturer's would be irrelevant and meaningless.
booklet describing the tube, (ti/ustrat/on by courtesy It is therefore more practical to regard the photons as being all
of Radio Corporation of America, Harrison, NJ.) prepared in the same way whenever the source can be kept steady
in such a manner that all the relevant probabilities are stable and
reproducible. This is the position which we shall adopt in the
following.
sity may be very weak). Let us consider two examples to illustrate Counter D
what we have in mind here.
Fig, 27A shows a semi-realistic electron diffraction experiment
in which the objective is to observe the diffraction pattern due to
the two slits in the screen S2. The electrons are emitted by the
filament F, and accelerated toward the screen Si, which is provided
with a slit* Let the electrons emerge through this slit with a mo¬
mentum p. We observe the two-slit diffraction pattern with the
help of the counter Dt placed at a very large distance from the
center of the second screen This counter can be moved along
the circular arc shown in the figure. We assume, for simplicity,
Fig. 27A To illustrate our discussion in Secs. 27-
that the distance from the counter to the slits is so large that we
30, of a two-slit electron-diffraction experiment, The
can regard the rays connecting the entrance slit of the counter with counting rate is observed as a function of the angle B
the two slits in S2 as parallel (This is not what the figure shows, when the counter and the entrance slit Sa are moved
because if we drew the figure correctly it would be hard to see the along the circular arc. If the separation between the
two slits in the screen S2. The essence of our discussion is, how¬ slits in S2 is large compared to the wavelength t and
if the source produces monoenergetic electrons, the
ever, not affected by whether the rays are parallel or not.)
counting rate will be an extremely rapidly varying
Let the separation between the two slits in S2 be 2a. The angular function of B. The diffraction pattern wifi not be
distribution 7(0,p) of the radiation detected by D can then be observable unless the angular resolution defined by
written, as we found in Sec. 40, Chap* 4 the counter slit arrangement is very good. If the elec¬
trons are not monoenergetic, as would be the case if
l(9yp) = 47O(0) cos2 (ap sin ff) (27a) the source is a simple filament the patterns for the
different energies overlap and the diffraction maxima
where 7o(0) is the angular distribution which we would observe might be smeared out to such an extent that they can
with a single slit.f no longer be seen*
m=ik)lZ’ *im=
cos (2opo sin 9) sin (2aq sin 9)
(29a)
2aq sin 9
accuracy, for the two frequencies. On the other hand, if this dif¬
ference is <n, i.e. 15(cor) — 5(w") | = then constructive inter¬
ference for the frequency of corresponds to destructive interference
for the frequency w", and vice versa. The system of fringes for
the two frequencies are complementary, and if they are super¬
imposed on each other with equal intensities no fringes will be
observed. This leads to a simple criterion for the visibility of the
fringes: the frequency spread Aw in the source must be such that
d Aw < rr (31c)
d<jt=fe)(t) <32a>
In Sec. 44, Chap* 3, we derived an expression for the fractional
Doppler broadening, namely
(is''vW^sjxl°5 (32c)
For T — 293°K (room temperature), and for X = 5000 A (visible
light), and for A — 100, we thus obtain d < 50 cm* This estimate
is in accordance with observations. The maximum path difference
for which interference fringes are seen is of the order of 1 meter
for "ordinary7’ light sources, such as gas discharge tubes (lasers
excepted).
where the equality sign applies if and only if all the numbers pft,
k = 1> 2, . . * , 2V, are equal, in which case their common value
equals Av(p;p)* In this particular case the particles in the beam all
have precisely the same momentum,
The quantity in the left side of (35c) measures the statistical
spread in the variable p* In general it will be greater than zero,
which we can express by saying that there is an uncertainty in the
momentum for the particular ensemble*
Av(D;p) = d (37c)
Suppose now that lamp 1 gives rise to a flux of fti photons per
unit time in the beam, and that lamp 2 gives rise to a flux of ri2
photons per unit time in the beam. The total flux in the beam is
thus (ni + n2) photons per unit time. In any single experiment
the photon is either “yellow” or “blue,” depending on whether it
came from lamp 1 or lamp 2, and we conclude that the probability
that we find a “yellow” photon in any single experiment is
ni
Fig.36B The beta spectrum of P3S. The graph
shows the relative number of emitted electrons as a
(»i + ns)
function of momentum. The momentum is expressed whereas the probability that we find a “blue” photon is
in terms of the quantity Bp (in units of gauss centi¬
meters), where p is the radius of curvature in the field n2
B. The maximum momentum at 7200 gauss cm (37e)
(n± + n2)
corresponds to the maximum kinetic energy 1.7 MeV.
The electrons can emerge with an energy ranging The numbers 0i and 02 satisfy the conditions
from zero to the upper limit because the total (kinetic)
energy released in the decay is shared (in a random 1 g 02 ^ 0, 01 + 02-1 (37f)
fashion) between the electron, the daughter nucleus,
and an anti-neutrino. as a consequence of the definitions (37d) and (37e). The conditions
Sec, 6,38 Uncertainty Principle and Theory of Measurements 247
d = + 02d2 (38a)
or
1 s 0* a 0, V ek = 1 (39a)
k-1
This means that the average of any physical variable Q for the
ensemble p is given by
n
We shall make the assumption that if pi, pa, P3, * * *, Pn is any set
of possible statistical ensembles, then every incoherent super¬
position of these ensembles is also a possible ensemble, This as¬
sumption is mathematical rather than physical, and we make it
because we want the set of aU statistical ensembles to have the
property that it is closed under incoherent superposition This
means that if the set contains any finite number of ensembles it also
contains all possible incoherent superpositions of these ensembles.
f=y jL * ^ <4od)
Amplitudes and Intensities
in S is unity. Let flu be the amplitude of the wave at the slit 1"
when the amplitude at the slit V is unity, but the amplitude at the
slit 2' is zero. Similarly, let B$i be the amplitude of the wave at
the slit l" when the amplitude at the slit V is zero, but the ampli¬
tude at the slit 2' is unity. Ci stands for the amplitude at the slit in
the screen D when the amplitude at the slit 1" is unity, but the
amplitude at the slit 2" is zero. The remaining amplitudes are
defined in an analogous manner. We can call these amplitudes
transfer amplitudes, because they describe the propagation of the
wave between the slits, from left to right. The dotted lines in Fig.
41A represent this propagation symbolically. A transfer amplitude
is associated with each dotted line, as we have explained.
The transfer amplitudes are complex numbers. Their absolute
squares define transfer probabilities as follows. Pi = | Ai 2 equals
the probability that a particle which has passed through the slit in
S is detected immediately behind the slit V. P& = A212 equals
the probability that a particle which has passed through the slit in
S also passes through the slit 2'. P12 = | Si2 [2 equals the prob¬
ability that a particle which has passed through the slit lr also passes
through the slit 2". In this case the slit 2' must be closed, to make
sure that the particle really passed through the slit V. The absolute
squares of the other transfer amplitudes have analogous interpreta¬
tions. Let us list all the transfer probabilities corresponding to the
eight amplitudes:
The reader should think carefully about how these transfer prob¬
abilities can be measured with counters, closing some of the slits
when required.
43 Suppose we now ask the question: with all the slits open, what
is the probability P that a particle which enters through the slit in S
will emerge through the slit in D?
Let us first give a thoughtless answer: since we know all the
transfer probabilities between the slits we can find P by compound¬
ing these probabilities according to the rules of probability theory.
The probability that the particle passes through the slit 1" should
thus be equal to the sum of the probability that it passes through
i" via the slit V and the probability that it passes through 1" via
Sec. 6.44 Uncertainty Principle and Theory of Measurements 251
the slit 2\ or, in other words equal to (FiPn + P2P21). This kind
of reasoning leads to the final wrong result
tJPpg§; (43a)
where
V= + BiaGiJARBSiCi + BM)
- AUBhCl + B?2C ?)A2{B21Ci + B33C2) ] (45c)
as the reader should convince himself.
If we like we can rewrite the expression for P(8) in the form
that a particle coming through the screen S' will also pass through
die slit in the screen D? Since every particle coming through the
slit in D must have passed through the screen S' it follows that F(0)
must be equal to the ratio of the probability P(0), given by Eq. (45a),
and the probability that a particle coming through the slit in S
passes through S'. This latter probability is clearly equal to
[|Ai|2 + | As | s]3 and we thus obtain
Similarly we can ask: for the situation shown in Fig, 46A, what is
the probability PI that a particle coming through the screen Sy will
also pass through the slit in DP We easily see that
(47d)
the counter may, or may not, click, and we do not know beforehand
what will happen with certainty: we can only state the probability
that the counter will click The reader may say: Well, that is only
because the ensemble is not pure, But what does the reader sug¬
gest that we should do in order to make it purer?
a pure state has the property that the outcome of every single
measurement is exactly predictable. If a given counter clicks in
one single experiment, it clicks in every subsequent experiment as
well. Each time an experiment is repeated the same things happen
which have happened before. For a pure state there is no statistical
spread in any physical variable.
Physicists have recognized for a long time, and long before the
development of quantum mechanics, that happenings in the macro¬
scopic world cannot in practice be predicted with unlimited pre¬
cision. Thermal noise and many other kinds of “disturbances”
over which we have no control are always present, and in macro¬
scopic situations these causes of uncertainty in the value of a phys¬
ical variable completely mask the characteristic quantum-mechan¬
ical uncertainty. The classical physicist's belief that pure states
are characterized by a complete absence of statistical spread in the
variables was never really critically tested in macroscopic situations,
and this explains why the belief could persist for so long.
light linearly polarized along the line which bisects the quadrant
bounded by the positive x- and y-axis (counter variable 045°) and
filters which let through light polarized perpendicular to this bi¬
sector (counter variable D1350).
For the ensemble p£, we find the following averages:
Av(Z>L;pL) = 1, Av(DapL) = 0 (58a)
1) It is proper that the reader supplement the theoretical studies ii> this chapter with
some reading on actual counters and related equipment.
a) Chapter 5, of D, Halliday; Introductory Nuclear Physics, (John Wiley and Sons,
Inc., .1950), is devoted to a discussion of the detection of charged particles and photons.
Various types of counters, and associated electronic equipment, are discussed,
b) The statistical analysis of counter data is discussed in the above reference. See
also, L. J. Rainwater and C, S. Wu: "Applications of Probability Theory to Nuclear
Particle Detection,” Nucfeomcff vol. 1, no. 2, p, 60, (1947) for a simple and clear dis¬
cussion,
c) G, D, Rochester and J. C. Wilson: Cloud Chamber Photographs of the Cosmic
Radiation (Academic Press, Inc.t New York, 1952), It is eminently worthwhile to look
at this book and its many interesting pictures,
d) An elementary discussion of the detection of particles is given in Chap. 3, in
D, H, Frisch and A, M. Thorndike: Elementary Particles (D, van Nostrand Company,
Inc., 1964),
e) For a collection of stereoscopic bubble chamber pictures, sec Intnxlurtion to the
De tertian of Nuclear Particles in a Bubble Chamber (Prepared at the Lawrence
Radiation Laboratory, The University of California, Berkeley.) (The Ealing Press, 1964,)
2) Note the following articles in the Scientific American magazine:
a) O. M. Bilaniuk, "Semiconductor Particle-Detectors,” Oct. 1962, p. 76.
b) G. B. Collins, "Scintillation Counters,” Nov. 1953, p. 36.
c) G. K. O’Neill, "The Spark Chamber,” Aug. 1962, p. 36.
d) H, Yagoda, "The Tracks of Nuclear Particles,” May 1956, p. 40.
e) D, A. Glaser, "The Bubble Chamber,” Feb, 1955, p. 46.
f) D, E. Yount, "The Streamer Chamber,” Oct. 1967, p. 38.
Problems
1 One of the favorite arguments of those who want to refute the un¬
certainty relation goes as follows. (See adjoining figure.) A monoenergetic
beam of electrons, of momentum p, is incident normally on the screen Si
from the left. This screen has a circular hole, of diameter a♦ At a distance d
Problems Theory of Measurements 263
from the screen Si we have another screen 82, which likewise has a circular
hole of diameter a. We assume that the two holes are lined up in the direc¬
tion of the incident beam. Some of the electrons which pass through the first Incident
hole might be deflected, but some of them will go on to pass through the beam r*-d-J
Si S2
second hole. Consider an electron which has passed through the second hole.
The uncertainty in its lateral position is of order Ax ^ a, The magnitude of
This figure refers to Prob, lf in which the author fal¬
its momentum is p, the same as in the incident beam, because the electrons laciously argues that the uncertainty relation can be
do not lose or gain energy in this experiment. Since we know that the elec¬ violated if we make the slits narrow and the separation
tron has passed through both holes the uncertainty in the direction of the d large. It then seems that the product of the un¬
momentum must be less than, or equal to, A6 — a/d, It follows that the certainty in lateral momentum with the uncertainty in
lateral position at the moment the particle passes
uncertainty in the lateral component of the momentum of the electron is of
through the second slit can be made as small as we
order Ap = (a/d )p. We thus have
like. What is wrong with this idea?
ap
fur the product of the uncertainties in lateral position and lateral momentum.
By making a small and d large we can make this product as small as we please,
and hence violate the uncertainty relation, which is one of the cornerstones
of quantum mechanics.
Can you demolish this argument? Make sure that you meet all counter¬
arguments to your arguments.
The above argument is one among many that have been advanced against
quauturn mechanics via the uncertainty relation. Now it should be clear
that there is never any danger that the uncertainty relation will be refuted by
this or any similar arguments, as long as the premises of toave mechanics are
acceptedJ because with these premises the uncertainty relation can be proved.
One might group the “refutations” of wave mechanics into two classes:
(a) Arguments in which the ideas of wave mechanics are really denied,
although this is not always stated explicitly,
(b) Arguments which are "muddled” but which are based on some of the
ideas of wave mechanics.
A careful conceptual analysis will clarify the nature of the “refutation.” An
outright denial of the principles of wave mechanics cannot, of course, be
contradicted on logical grounds, but one can always appeal to experimental
facts: the “refutation” carried to its logical conclusion might contradict one
of these facts. The arguments in the category (b) are simply faulty.
A dn
0.075
n dX
The velocity of the pulse of light let through the shutter could be measured
by another shutter placed at a certain distance from the first, and opened at
a slightly later time. What is the velocity by which the pulse propagates in
the carbon disulfide?
3 The author has a new idea for violating the uncertainty principle: this
time the time-frequency uncertainty relation. The arrangement is shown
very schematically in the adjoining figure. Almost monochromatic light is
incident through the slit at left which is fitted with an extremely fast shutter.
We shall not be concerned here with purely technical difficulties, and we
thus assume that the shutter can be opened for an arbitrarily short interval of
time, to admit a sharply defined pulse into the spectrograph indicated sym¬
bolically by the prism in the figure* The entering fight will, of course, no
longer be monochromatic, but will exhibit a spread in frequency as discussed
in Prob. 2. We can, however, fit the spectrograph with a suitable narrow exit
slit, shown at right in the figure, and thereby select a portion of the incident
light whose wavelength falls within an extremely narrow range. The fight
emerging through the exit slit can therefore be made monochromatic to an
arbitrarily high degree: the uncertainty in the frequency can be made as
Fast shutter
small as we like. On the other hand the duration of the pulse can be made
as short as we like with the help of the shutter The pulse emerging through
the exit slit can therefore be of arbitrarily short duration, and be of an arbi¬
trarily precise frequency, contrary to what the uncertainty relation says* Can
you find the fallacy in this argument?
4 With reference to our discussion in Sec, 29, suppose that the tempera¬
This figure refers to Prob, 3. The author again tries
ture of the filament is 1000° C, and suppose that the accelerating potential is
to violate the uncertainty relation. The prism sym¬
bolizes a spectrograph of very high resolution which 10 volts* Estimate the fractional precision in the momentum of the emerging
is used to select an extremely narrow frequency electrons, i,e*, estimate the quantity q/po- A crude estimate is enough.
range of the transmitted light* The incident light is Explain your ideas*
controlled by a fast shutter. The author erroneously
maintains that the light pulse emerging from the exit
slit can be arbitrarily well defined both in frequency 5 If we could produce electron beams of very low energy it would be
and time. What is wrong? possible to do macroscopic” electron-diffraction experiments. Suppose
that we try to produce a beam of well-defined momentum, with a mean
Problems Theory of Measurements 265
energy of 0.01 eV* Discuss the practical difficulties which we might en¬
counter in this attempt. It is clear that a hot filament with a single accelerat¬
ing electrode would not do, but there are perhaps other methods which you
might think of* If so, state your ideas, and discuss their technical feasibility*
& The adjoining figure shows a refinement of the double double-slit ex¬
s ^ a 9X
periment discussed in Secs* 41-43. Ideal polarization filters are placed (or not
^M
\
placed) in front of the slits and in front of the source %nd the detector. We
0
\ / P21
X
assume that the transfer amplitudes discussed in Secs* 41-43 are independent /\
B
of the state of polarization, and we assmne that the light source emits un¬
polarized light. Derive expressions, analogous to (43b), for the probability
Nil/ B\lk'
F'|—%—fF''
that a photon entering through the slit in S will pass through the slit in D, for
the various combinations of polarization filters listed below*
This figure refers to Prob, S. It illustrates a refine¬
F
L X Fi F'a F£' F? Frf ment of the double double-slit experiment shown in
Fig. 41 A. Ideal polarization filters can cover the
abs H V abs abs abs various slits. The problem is to determine the prob¬
LC H V abs abs abs ability that a photon entering through the slit in S
LC H V abs abs RC emerges through the slit in D for various combinations
LC H V RC LC H of filters. The numbers AWl and Cm are the
abs II abs abs H abs transfer amplitudes in the absence of the filters. We
assume that the transfer amplitudes do not depend on
In this table “abs” means that the filter is not there, H means a horizontal the state of polarization.
266 Chap. 6 Problems
of Schrodinger
Problems 307
Chapter 7 The Wave Mechanics
of Schrddinger
(V'- S) <1M>
and we then have
and hence
\M^)\*= |* (lot)
As we see from equation (lOf) the two wave functions and
\pB differ only in a complex factor of modulus unity and this factor
is independent of the state of motion of the particle, i.e*, it is in¬
dependent of p* The absolute squares of the two wave functions
are identical everywhere and at all times* To describe the prob¬
ability distribution of the particle we can employ the wave function
t/'s just as well as the “correct” de Broglie wave function ^ This
Ls precisely what is done in the Schrodinger theory, and as
given by (lOd), is thus taken to be the Schrodinger wave function
describing a free particle moving with the small momentum p. This
convention is purely a matter of convenience; why should we carry
along the factor exp (—itmc2/1i) in our calculations when it ulti¬
mately has “no physical effect?”
72 >Kx’f) (lla)
<o p p2 P
where « — k (12a)
k ~ 2m 2mH *
The phase velocity Vf of the de Broglie wave given (in the non-
relativistie approximation) in (10c) is, on the other hand,
me2 P (12b)
V 2m
The reader may be bothered by the fact that the two phase
velocities v/ and v-f are not equal, although the two kinds of waves
^ and are supposed to describe exactly the same physical situ¬
ation. However, there is no cause for alarm: the phase velocity is
not the same thing as the velocity of the particle, and it does not
correspond to anything directly observable. The group velocity o,
on the other hand, is given by
1 _ dk __ m
(12c)
v do* p
for the Schrodinger wave, and this velocity Ls indeed equal to the
velocity of the particle as it should be. We have already shown,
in Chap. 5, that the group velocity of a de Broglie wave is also
equal to the velocity of the particle, and the two kinds of waves
therefore do propagate with the same group velocities.
13 Let us now try to go one step further and consider the motion
of the particle in an external field of force derivable from a poten¬
tial. We shall denote the potential energy of the particle by V(x):
the potential is a function of position but not of time.
The reader may have some doubts about the idea of introducing
a potential in quantum mechanics to describe the forces acting on
a particle. The forces on a particle are, of course, due to the pres¬
ence of other particles, and consistency requires that these other
particles should also be described quantum mechanically. AH the
particles in a given physical situation should be described as waves,
and a fundamental theory of particle interactions must therefore
Sec. 7.14 The Wave Mechanics of Schrodinger 275
=£ (13a)
p*
E = (16b)
2m
It follows that each one of these plane waves satisfies the dif¬
ferential equation
V2 x[s(x,t) = (16c)
2m
Therefore, the Schrodinger wave corresponding to a particle of
energy E must satisfy the differential equation (16c) throughout
the region IIL
Consider now the wave in the region I. If we resolve the wave
in this region into plane waves of the form exp (ix * p/K) the mag¬
nitude of the momentum p will now be determined, in accordance
with (15a), by
Pi = Eton = E — VI
2m
(16d)
-~v
Zm
2 = {E~ Vn) #£,#) (iff)
where V(x) is the potential function which assumes the values V/,
V// and Vr/i — 0 in the three regions. It should l>c noted, however,
that we presented no arguments for what the “correct” differential
equation should be in the boundary-regions in which the potential
changes rapidly, and it is therefore not self-evident that the equation
(17a) must hold everywhere. In fact the author now wants to eon-
fess that he arranged the arguments leading to our equations, and
drew the picture in Fig* 15A in a manner deliberately designed to
lead the reader to believe that die equation (16e), for instance*
must be true* There is actually a flaw in our argument* As long
as the region II is very large compared to the wavelength of die
de Broglie wave in this region we can safely accept our conclusion
(I6e) as extremely plausible. The heal behavior of the wave in the
region should not depend on the potential elsewhere, and the rela¬
tion between wavelength and kinetic energy must then be in ac¬
cordance with our assumption. The situation is different, however,
if the region II is small compared to the wavelength, i,e., if the
potential V(x) varies considerably over a wavelength* In this case
it is not so clear what the spatial dependence of the wave function
should be, because the “wavelength” at the point x, as defined
through the de Broglie relation in terms of the kinetic energy
[E — V(ac)]s is a function of position.
It is therefore not self-evident that the equation (17a) is the
correct equation everywhere in space, and for every potential
function V(x)* Nevertheless we shall assume, following Schrti-
dinger, that the equation (17a) is correct* As an equation describ¬
ing the behavior of the Schrodinger waves it is at least a reasonable
equation, and we should give it a fair trial* We wanted to make it
clear, however, that our discussion is not a proof of the correctness
of equation (I7a) but rather a plausibility argument in its favor.
Actually one can do a bit better. One possible approach is to start
from quantum electrodynamics, in which case one can show that
the equation (17a), as applied to non-relativistic problems involving
atoms and molecules, arises as an approximation to the field theoiy
formulation. Another approach is to study systematically what the
possible wave equations are which allow a sensible physical inter¬
pretation, including the probability interpretation discussed in Sec.
8* We wish to preserve this interpretation of the wave function in
the case when the particle is subject to forces. One may then show
that the equation (17a) is, in a certain sense, the simplest wave
Sec. 7.18 The Wave Mechanics of Schrodinger 279
~ 2^ + ^ ^ ^ (18a)
Turning
21 Consider now the situation shown in Fig, 21 A, The dashed
point line indicates the total energy E, whereas the solid line represents
the potential function V(a;). We assume that as we go to the left in
Fig, 21A To illustrate discussion in Sec. 2L The the figure the potential tends to the constant value zero, whereas as
solid line represents the potential, and the heavy we go to the right it tends to the constant value Vq ]> Er The point
dashed line indicates the magnitude of the total erv r0, where the kinetic energy has the value zero, is known as the
ergy E. The point x0 at which the potential equals E
turning point According to classical mechanics a particle incident
is the classical turning point. According to quantum
mechanics there is a finite probability that the particle from the left would stop and turn around at this point. The region
will be found in the classically forbidden region. to the right of #q is inaccessible to the classical particle.
Sec, 7,22 The Wave Mechanics of Schrodinger 281
We should now solve the equation (20c) for the potential shown
in Fig* 21A* The solution <p(x) is some function of x which is con¬
tinuous and has a continuous first derivative* Without really solving
the equation explicitly we may guess that the wave function <p(x)
will not vanish to the right of x&, which, according to our probability
interpretation of the wave function, means that there is a certain
non-vanishing probability that we will find the particle to the right
of xq. Quantum mechanics therefore predicts that a particle can
penetrate into a region forbidden to it according to classical
mechanics*
(24a)
A + B = 1, ik(A — B) = -q (24e)
A _ (1 + *?A) B _ (1 - *Q/k)
1 -iyWo/E - 1
for x < 0 (25a)
1 +iy/%f&= T_
and
2e~w
for x > 0 (25b)
l + WVo/E-T
where
2mE _ /2m(Vo - E)
& ’ q ~y W (25c)
1-WVo/E-l
l + WVo/E - 1
and the two waves therefore have amplitudes of the same magni¬
tude, The absolute square of the amplitude of a wave must some¬
how be proportional to the “flux” of the particle, and we can con¬
clude that the wave function in (25a) describes the situation in
which a particle incident from the left is reflected back to the left
by the potential “hill,” This interpretation is in accordance with
our classical picture of what goes on.
The wave function in the region x > 0, as given by (25b), de¬
scribes the penetration of the Schrodinger wave into the region
forbidden to the classical particle. The amplitude of the penetrat¬
ing wave decreases exponentially as we go further into the forbidden
region, and at large distances from the barrier the amplitude is for
284 The Wave Mechanics of Schrodinger Sec. 7.26
find it explicitly is, however, often not an easy matter, but the
complications are only of a technicahmathematical nature. Even
without knowing the precise explicit solution we can often say
quite a lot about the nature of the solution, and hence make general
statements about the behavior of the physical system. On the basis
of our studies so far we can conclude that it is a general feature of
quantum mechanics that the Schrodinger wave can penetrate into
regions forbidden to the particle within classical mechanics,
2m(V0 - E)
<p(x) = B exp (-xq), q =
V- (32b)
Set?. 7.33 The Wave Mechanics of Schrodinger 289
2m(V0 - E)
2a
exp
Y &
(33b)
for the situation shown in Fig. 34B is very small whenever the
thickness of the slab of the optically rare material is large compared
to the wavelength of the incident radiation. As the thickness de¬
creases the transmission coefficient increases, and it reaches the
value one when the thickness is zero.
in T == /n 1 + in Tg -h in H- in -f- in T5 (35b)
hrs.2j[^v@MEI (36b)
Theory of Alpha-Radioactivity
f The justification for trying to find the explanation on the basis of the Schrodinger
theory is that the velocity of the alpha-particle outside the nuclear surface is "noil-
relativistic,” as the reader can estimate for himself. Remember that the energy of the
alpha-particle does not exceed 10 MeV,
294 The Wave Mechanics of Schrodinger Sec. 7.39
2e2Z'
Rc (39a)
E
R ^ roA1/3 - L2 X cm (39b)
and for ggRa226, for which A = 226, we thus obtain R = 7.3 fermi.
The picture is therefore qualitatively correct; the alpha-particle
indeed has to penetrate a potential barrier. Quantitatively the
picture is wrong: the barrier should have been drawn much thicker.
Our execution of the figure has been motivated by aesthetic con¬
siderations: the important features of the situation are nevertheless
reproduced qualitatively.
The inequality Rc R is true generally for the alpha-active
Sec. 7,40 The Wave Mechanics of Schrodinger 295
“ T“ - jC * Vi-* (40b)
The integral occurring in (40b) can be evaluated in closed form
easily enough. Since, however, the quantity xc = R/Rc is in
general a fairly “small” quantity, it will suffice for our purposes to
carry out an approximate evaluation in which we retain only the
first two terms in an expansion in We proceed as follows:
(40c)
The first term in the extreme right-hand side of (40c) can be
evaluated trivially if we make the substitution x = sin2 8, and we
get
f-2w (4fle)
and if we substitute this expression into (40b) taking into account
(39a) we obtain
<«*>
To
r (42a)
or
148
Log r = Log r0 + 32*5 (42b)
\/Ef MeV
To estimate tq we may assume, on the basis of our native model,
that the alpha-particle inside the nucleus moves with the same
velocity v as it will have after emission* We then have
2R
to v (42c)
v
148
Log (r/sec) ^ 53*5 (42d)
\jEfMeV
298 The Wave Mechanics of Schrodinger Sec. 7,43
Energy E in MeV
ti/2 = -1 In 2 = rm In 2
{ G. Gamow, “Zur Quantentheorie des Atoinkernes/’ Xeitschrift fur Physik 51, 204
{1928). See also G. Gamow, “Quantum theory of nuclear disintegration/' Nature 122,
805 (1928); R. W, Gurney and E, U, Condon, “Wave mechanics and radioactive
disintegration/' Nature 122, 439 (1928).
300 The Wave Mechanics of Sckrodinger Sec* 7A6
Tir1 2
' \. i 'lavs
lifetime, or else it is a member of a decay chain originating from
a long-lived element. Among heavy nuclei with long lifetimes we
note U238 with half-life 4.5 X 109 years; Th232 with half-life Ik fi.7 hr
1.4 X 1010 years, and U235 with half-life 7,13 X 108 years. The
most long-lived member of the (4n + l)-family is a neptunium ,.|C
isotope, Np237 with half-life 2.2 X 106 years. This is a short time X id' vr
measured on a geological time scale, and tfye (4n + 1)-family there¬
fore does not occur naturally.
A few of the naturally occurring light nuclei are also radioactive. 8.3 X to4 vr
Examples are the beta-active nuclei K40 with half-life 1,3 X 109
years and Rb87 with half-life 4,7 X 1010 years.
yr
47 The phenomenon of natural radioactivity makes it possible
to determine the ages of rocks, i.e., the time which has passed since
the rocks were last chemically transformed. The principle is ffggyliMBiir
3.8 da vs #
Au ~ Aro e
1.8X10 4s»x*
where No is the number of U23S atoms present originally, A is the
«j
decay rate of uranium, and T is the age of the sample. Since
No — A u + Npb we have 21 yr
XT _ (A7Pb + Aru)
Nlt
(47b) ft
5.0 dav.s
X0 10 20 30 40 50 60 70 80 90
Atomic number Z
we can find out how many uranium atoms have disintegrated since
the rock was formed* t
Through methods such as these it has been found that the oldesi
rocks in the crust of the earth are about 3 X 10e years old. This
is definitely a lower limit on the age of the earth, because the crusi
has undergone many chemical transformations in the past. Mete¬
orites have also been studied, and they have been found to be about
4*6 X 109 years old, How the meteorites were formed is not
known with any certainty, but there is good evidenoe that the)
f The first estimate of the age of the earth on the basis of radioactivity was made bj
Rutherford. See E. Rutherford, “The Mass and Velocity of the a particles expellee
from Radium and Actinium,” Philosophical Magfizine 12, 348 (1906). See pp* 388-
369, where Rutherford arrives at the estimate of 400 million years for the minerah
he studied.
Sec. 7A8 The Wave Mechanics of Schrddlnger 303
were formed (crystallized) at about the same time as the other solid TABLE 48B The Eight Most Common Elements
bodies in the solar systerrii The age of the earth as a body would in the Earthrs Crust
thus be about 4,6 X 10e years. It is furthermore possible to esti¬
Number of Atoms
mate, using radioactive “clocks,” the time that elapsed between
Element percent
the latest formation of the chemical elements in the meteorite and
its crystallization* According to one such estimate,! it appears Oxygen 62.6
that this time was about 0.35 X 10e years. This implies that the Silicon 21.2
final building of the chemical elements contained in planets and Aluminum 6.5
meteorites took place about 5 billion years ago. This is thus the Sodium 2.64
estimated age of our solar system. Calcium 1.94
iron 1.92
48 It is natural to speculate further. How old is the universe? Magnesium 1.84
How were the chemical elements formed? We cannot discuss here Potassium 1.42
the ideas which lead to estimates of the age of the universe* It is
believed that the age of the universe might be something like 10 This table shows the estimated composition of the
billion years, roughly of the same order of magnitude as the age ten outermost miles of the earth's crust together with
the oceans and the atmosphere. These eight ele¬
of the solar system.
ments make up nearly 99 percent of the mass of this
It is believed that the chemical elements were formed from domain. The weak gravitational field of the earth
hydrogen in nuclear reactions in the stars. Fig. 48A shows the cannot retain the light elements hydrogen and helium,
estimated abundances of the chemical elements in the solar system. which explains their low abundances compared to the
The dots which represent individual chemical elements are not "cosmic" data. The abundances for the heavier ele¬
ments in the earth can be expected to be similar to
experimental points in the sense that they would all derive from
the cosmic abundances, but geological processes on
measurements on a single "standard sample.” The dots represent the earth have led to a chemical segregation of the
estimates based on a large number of different kinds of measure¬ elements, and the data for the crust are not represen¬
ments, such as spectroscopic determinations of relative abundances tative for the earth as a whole,
in the solar atmosphere, relative abundance measurements in mete*
orites and estimates of the chemical composition of the earth s
crust. Notice that hydrogen is by far the most abundant element.
Notice also the peaks in the abundance curve corresponding to
particularly stable elements. There is a clearly visible systematic
trend in that elements of even atomic number are more abundant
than neighboring elements of odd atomic number, which reflects
the fact that nuclei which have an even number of protons and an
even number of neutrons tend to be more stable than other nuclei.
To explain this curve in all details and thereby trace out the early
history of the solar system is a fascinating problem. The main
features of the abundance curve are believed to be fairly well under¬
stood at this time.
The author has absolutely nothing to say about the question of
where the hydrogen came from originally*
Advanced Topic:
Normalisation of the Wave Function t
P(xi,x2) = N fJx
1
dx |ip(x,t)|2 (49a)
M*>*) = (49c)
where N is given by Eq, (49b). This wave function has the nice
property that
ft * I 12 = 0 (50b)
= - V(x)t*(x,t) (50d)
(50e)
2m<br\r S# dx /
We therefore obtain
306 The Wave Mechanics of Schrodinger Sec* 7*51
a/_r *= 4-00
(50f)
51 At this point the reader may be quite alarmed since our firm
conclusion that every physically meaningful wave function must be
square-integrable seems to cast doubts on our discussion of plane
monochromatic waves earlier in this chapter, It is clear that a wave
function of the form exp (ixpf'h — itp2/2m:H) is not squaro-integ-
rable and therefore cannot be normalized to unity. We are forced
to the conclusion that a wave of precisely defined momentum p,
which depends on the coordinate x only through the factor exp
(ixpffi), does not correspond to a physically realizable state of
motion of the particle. On the other hand we are not forbidden to
consider a wave which over a very large interval on the x-axis
depends on x through the factor exp (fecpM), provided that this
wave function does tend to zero as x tends to + oc or — oo. We can
therefore resolve our difficulty by agreeing that when we discuss
“waves of precisely defined momentum” we do not really mean
that the wave is everywhere of the form exp (ixpfff). We realize
that the wave function must tend to zero at infinity, but we assume
that the wave function is of this form in a very large interval of
the x-axis which includes the region in which we are primarily
interested. Our “monochromatic waves” are thus to be understood
as “almost monochromatic waves,” With this understanding we
can safely continue to talk about waves which depend on the co¬
ordinates through the factors exp (■ixp/&\ or exp (ix - p/#), as is done
in almost all texts on quantum mechanics. We can regard the non-
normalizable waves as limiting cases of normalizable waves, and if
we like we may call the former kind of wave functions improper
References for Further Study 307
wave functions* This term will also serve to placate the mathema¬
ticians whose sensibilities are rightly offended by the frequently
careless manner in which physicists talk about “plane waves" as if
these would be bona-fide Schrodinger wave functions.
Problems
1Consider the barrier shown in Fig. 28A for the case when E> VP.
(a) Consider first the case of a particle incident from the left* It, i.e., the
wave packet, will be partly reflected and partly transmitted by the step*
To discuss this case we desire a solution which in the right-hand region
describes a wave traveling to the right. Find this solution everywhere, and
308 Cfeap, 7 Problems
4 Let us fuss about a small detail: Is the Fig, 34B appropriately drawn?
Consider the relationship between the transmitted ray and the incident ray.
Perhaps the transmitted ray should have been drawn as a continuation of the
incident ray, and not as in the figure? To find out how this picture should
really be drawn we could perhaps perform some experiments. Suppose that
the thickness of the optically rare medium is of the order of a wavelength of
the light used. By an arrangement with slits we select an extremely narrow
pencil of light as the incident beam, and this pencil is represented by the
dotted line in the lower right part of the figure. We can then study the
transmitted pencil of light, and find out whether it actually follows the dotted
line in the upper left part of the figure. You do not really have to do this
experiment in the laboratory; you can do it as a mental experiment instead,
because there is nothing in this experiment which cannot be accurately
predicted within electromagnetic theory.
After a consideration of this experiment, state your opinions as to whether
Fig. 34B is correctly drawn.
Problems The Wave Mechanics of Schr&dinger 309
6 There are further interesting questions which we can ask about the
situation illustrated in the figure associated with the preceding problem.
For instance: Is the transparency of the barrier the same in both directions?
Theorem: The transmission coefficient when the particle is incident from
the left is the same as when the particle is incident from the right, provided
that the energies are the same in both cases.
Prove this theorem. Hintx Note that if <p(x), as discussed in the preceding
problem, is a solution of the Schrodinger equation, so is <p*(x), and so is every
linear combination of <p(x) and qp*(x). Consider a suitable linear combination
of <p(x) and qp*(x).
= m |*(*,*) (3b)
Inserting a wave function of this form into (3b) we obtain the time-
independent Schrodinger equation
where
>-v¥
and where A and B are constants. If we now first impose the con¬
<4d)
dition that the wave function must vanish at x = 0, we find that the
physically acceptable solution must be of the form
<p(x) = C sin (xfc) (4e)
5 We have thus found that for our particle in a box the Schro-
dinger equation (3b) has stationary solutions with a simple ex¬
ponential time-dependence, i,e., solutions of the form ^(xs#) =
<p(x) exp (— itEffi), only if the energy E assumes one of a discrete
set of values Elt E2j E3, ..., E^ *. . , given by
inside the interval (0,n), and zero outside. (That this wave function
is correctly normalized to unity we see trivially by integrating
[ i^(x,t) 2 = (2fa) sin2 {nwx/a) from 0 to a\ the result is L)
H-oo +QP
We have represented the energies En in the form of a term
scheme for our system in Fig. 4B, which shows the first six energy
levels. In Fig. 4C we have drawn the corresponding wave func¬
tions <p,t(x). These functions are, of course, equal to the functions
at the particular time t = 0.
See also Fig. 5A for a composite picture,
x=Q x—a
inside the interval (0,a)9 and Pn(x) = 0 outside this interval. As we
see, the probability density does not depend on the time for a
stationary solution. Fig. SA It is common practice in texts on quantum
Let us next consider a non-stationary solution. Since the Schro¬ mechanics to show figures such as the one above.
The three figures 4ABC have been collapsed into a
dinger equation (3b) is a linear differential equation the linear single figure. This is perhaps a deplorable practice,
combination of any two solutions gives a new solution. This new but since the author has never been misled by such
solution will satisfy the same boundary conditions ip(0,£) = pictures, he assumes that the reader will not be misled
ip(a9t) = 0, provided that the two original solutions satisfy these either.
The energy levels are indicated by the thin broken
lines. Each one of these lines also serves as the
f For a discussion of the normalization of the Schrd dinger wave function, see Sec, r-axis in a superimposed figure showing the corre*
49, Chap, 7. sponding wave function.
316 Theory of Stationary States Sec* 8*7
4>(x4) = yA (6b)
where we assume that n' =^= nff. We claim that this new solution
of the Schrodinger equation is normalized to unity (for all times £).
The probability density P(x,t) corresponding to the solution (6b) is
given by
, * /nVx\ . /n"m\
+ 2 an (—j sin (—j COS
fcVn" —
interactions between the electrons and the nucleus are taken into
account, but in which the radiation of electromagnetic waves from
the moving particles is ignored* We can nevertheless hope that the
Schrftdinger theory is a good approximation in atomic and molecular
physics. We can thus expect that a stationary state predicted by
the SchrQdinger equation would correspond to an almost stationary
state in the “true” theory, and that the “mean energy” of the
latter state would be very close to the precise energy predicted by
the Schrodinger equation*
17 Let us next consider the case when V_ > £ > V0. An exam¬
ple of this is the energy £3, indicated by the dashed line so labeled
in Fig. HA. In this case we have a region to the left and a region
to the right in which [£ — V(x)] < 0, and a region in the middle
in which [£ — V(x)] > 0. The two boundary points separating
these regions from each other are the classical fuming points: we
shall denote them by xi and x2.
To the left of xi the wave function must tend asymptotically to
the x-axis, and the behavior must be as indicated by the left seg¬ For the convenience of the reader we show Fig. 11A
once more. The potential function tends to the con’
ment in Fig. 12C (except for the sign of the wave function, which
slant values V+ and v_ as x tends to +00 or -00.
is immaterial). Unless the wave function behaves in this way it The dashed horizontal lines indicate four energies
would grow as x tends to — 00, and a steadily growing wave func- representative of the cases which can occur.
322 Theory of Stationary States Sec. 8*18
are shown in the left part of the figure. Note that the first wave E
function has one extremum (and no node), the second wave func¬ - 0.2B
tion has two extrema (and one node), and the fourth wave function*
corresponding to the highest discrete energy level, has four extrema - 0
(and three nodes). For a deeper potential well we would have — 0.2B
more bound states, and in the extreme case of an infinitely deep
- 0.503 — 0.43
well, which is the problem discussed in Sec. 4, we have an infinite
number of bound states. The term schemes in Figs. 4B and 19A -0.63
- 0.773
should be compared: the locations of the first four bound state -0.83
levels are similar, although not identical, in the two cases. — 0,943
-1.03
The reader may wish to try to solve the problem of finding the
bound states for the situation shown in Fig. 19A: it is not particu¬ Fig. 19A The case of a particle in a potential well
larly difficult. of depth 6. This figure is based on an example given
We have now learned that we can understand, on the basis of in R. B. Leighton, Princfp/es of Modern Physics, p. 154
the SchrOdinger theory* why a quantum-mechanical system will (McGraw-Hill Book Co., New York, 1959). The poten¬
have bound states, and why there will in general be a continuum tial well is shown at left, and the term scheme at right.
There are four bound states (four discrete energy
of possible energies above a certain limit. The beginning of the
levels). The corresponding eigenfunctions are shown
continuum is simply the energy above which the system can dis¬ at left, superimposed on the graph of the potential
sociate, which in our simple examples means that the particle can function. The continuum begins at the top of the
behave like a propagating wave packet far away from the 'central well, as indicated in gray on the term scheme.
region.”
we are closer to the lower limit for the third excited state in Fig.
19A. The formula (24e) therefore represents a compromise.
p = — V) (25b)
Comparing (25a) with (24a) we find
dx — — \/2m(E — V) dx (25d)
n
Let us now use (25d) as an approximate expression for the change
of phase with x in the case that V(x) is not a constant, This approxi¬
mation is better justified the more slowly the potential V(x) varies
with position. Within this approximation the total change of phase
Between the two turning points xi and X2 is given by
where n — 0, 1, 2, * * . .
The approximate method which we have just found for determin¬
ing the energy levels of a particle in a “potential valley/* like the
one shown in Fig* 23A> is known as the WKB-method. f In many
cases it gives fairly accurate results, and it is always useful when
we try to form a rough idea of the location of the levels* The
nature of the approximation is quite similar to the nature of the
approximation which we made in deriving the formula (36b) in
Chap. 7 for the transmission coefficient of a potential barrier: the
same kinds of integrals in fact appear in both cases.
It is interesting to note that our equation (25f) which we derived
within the framework of wave mechanics, is identical to the so-
called Bohr-Sommerfeld quantum condition of the old Bohr theory.
We thus achieve a certain understanding of why the Bohr theory
sometimes works quite well, and we can also see why the Bohr
theory might sometimes fail badly: the equation (25f) is not
rigorously true but is only an approximate relation*
V(*) = | x2 (27a)
/* -
(27b)
Wo = Vm
To carry out the quantization procedure described in Sec, 26 we
must first find the turning points. They are symmetrically placed
with respect to the origin, and we can write x± = — Xq, = x$r
where, in accordance with (26a), we have
stable for n > 0. One can show that the selection rule for electric
dipole transitions is that n changes by one unit. The quantum
emitted must accordingly be of the classical frequency coo for any
transition of this kind. This is also what we would predict on the
basis of classical theory.
Electron
EU = + 2^-4^ = ^(VS-4) (33a) —e
aV2 a a
This potential energy should be compared with the total potential
energy £p0t of two hydrogen atoms separated by a very large
distance from each other. This potential energy is given by
E'm = (33b)
r - r0
U(r) + U(r0) (35a)
( flo -y
This is a reasonable guess. For r = r<> the right side assumes the
correct value E7(ro). For |r — Tq\ = ao the potential is larger than
t/(ro) by the amount Rw. Since the size of a molecule is of order
oo» and since the binding energy is of order R.*, we would expect
the potential to behave roughly in this manner.
The right side in (35a) is the potential for a harmonic oscillator,
The “spring constant” K for this oscillator is given by
2RW a2mc2
(35b)
Qq2 do2
= (35c>
atomic physics we concluded that we can regard the quantity TABLE 35A Vibrational Frequencies of
Selected Diatomic Molecules
(35d)
Frequency Wave number
Molecule cycles/sec cm
as a “typical” frequency associated with optical transitions in an
atom or molecule; i.e., transitions in which the electronic con¬ c2 4,921 x 1013 1641.35
figuration changes. We can then write (35c) in the form n2 7.074 x 1013 2359.61
Os 4.374 x 1013 1580.36
NO 5.708 X 1013 1904.03
(35e) CO 6.506 X 1013 2170.21
IBr 0.805 x 10” 268.4
The quantity M is for all molecules of the order of a nuclear $2 2,176 x 1013 725.68
mass, whereas m is the electronic mass. The “typical” electronic
frequencies, (Ce> lie in the visible region of the electromagnetic
spectrum. As we see the “typical” vibrational frequencies of
molecules, are smaller by the factor They will thus be
found in the near infrared region, and this prediction is in accor¬
dance with observations*
two nuclei from the center of mass, as indicated in Fig. 36A, The
kinetic energy of this system is then given by
Mxt
T= W = *(m7Tk) * <“■> Ml + M2
where the dots indicate time derivatives. The potential energy of
the oscillator is given by (35a) as a function of f, and the kinetic
energy is given by (36a) as a function of r. The effective mass M
of this oscillator is then the coefficient of fi/% i.e.,
We can then write the kinetic energy of the molecule in the form 6 r1
Tt = £ (38e)
where we have eliminated the angular velocity <oa from the express D i
sion (38b) by the use of the relations (38d),
3
(39a)
Tr % 1
03r
a 27 0
Fig* 39A Term scheme showing the eight first rota¬
According to (38d) the angular momentum is given by J — 7<oa,
tional energy levels of a diatomic molecule (regarding
and since we assumed J — H it follows that wa — fi/7. The angular the molecule as a rigid dumbbell). According to Eq.
velocity and the characteristic rotation frequency defined (39c) the energy E* of the state of angular momen¬
by (39b) are thus of the same order of magnitude, as we would tum f is given by Ej = + l)t where B = fi2/{21)
expect on the basis of a classical model. is the rotationa/ constant of the molecule. The ver¬
tical arrows show electric dipole transitions in which /
The complete quantum-mechanical theory of the dumbbell
changes by one unit.
molecule leads to a very simple formula for the energy levels. Each
rotational state is characterized by a non-negative integral value
of the angular momentum quantum number j, and the energy of
the state is given by TABLE 39B The Rotational Constant Be for
Selected Diatomic Molecules
* _ id +1)*2
*** — - (39c) Be r
Molecule (Mc/sec) (A)
where / = 0, 1, 2, 3.Although we shall not derive the
formula in this book, the author felt it was worthwhile to quote it BrF 10700 1.76
KCI 3800 2.79
anyway.
KBr 2400 2,94
C12013 57900 M3
40 The separation between the nuclei in any molecule is of the OH 566000 0.97
order of the Bohr radius We thus estimate the moment of NO 51100 1,15
inertia as 7 ~ Mao2, and if we insert this expression for 7 into (39b)
The constant B (see Fig. 39A) is here expressed in
we obtain
terms of the corresponding frequency Be = B/k =
&/(8ffai), in megacycles per second. The third column
(40a) gives the internuclear distance t.
2Ma02
338 Theory of Stationary States Sec, 8.41
■ ■r.1
■ ■'
?h'.V
■hi
J i 41 The key idea in the complete explanation of the very com¬
Waveguide^ Gas inlet plicated optical band-spectra emitted by molecules is the idea that
■1 -11 * m 4 . VI
"
Scale
Spectrum IQ me
under low resolution
Theoreticof pottern
due to CJ quadruple done
Spectrum under
high resolution
Theoretical pottern
indudrng N r—r-4+ It
quadrupole effects X-XJ ii|i iL_f X X 4-4
; XX
LX%~XX ! XX
H-XX
, - XX ■
XX
V/2*V/2-ljj \ X sniXX-% % -J <~Xh~XX XX I
XX*XX~‘\ XVXX '4 %■*% X
X%
X V1 t XX
X X~X %-i
and ^ ^2' Ji [X %-%X -i X H
\XX-1>XX
e2Z
V2 <p(x) <p(x) = E<p(x) (42a)
2m
where x =
a
x = where a = (43a)
mcaZ he
and let us also introduce a new “energy parameter'* A by
E = (aZ)2mc2\ (43b)
n = oo OeV
tt5* 5 0.54 eV as well as the wave function/(y) defined by
n=4 0.85 eV
n bs 3 1.51 eV <p(x) = /(y) (43c)
Paschen
series
Brackett Rewriting the differential equation (42a) in terms of our new
series variables and parameters we obtain
n=2 3.39 eV
Balmer First excited
series state
- \^v2 f(y) --Ky) = M(y) (43d)
57
Lyman
dinger equation (42a). It is dimensionless in the sense that the
series physical constants m, e, h, c, and Z no longer appear in it. If we
can solve (43d) we can re-introduce the old variables by making use
of Eqs. (43a) to (43c) and the two equations (43d) and (42a) are
clearly completely equivalent.
<Pio(x) = (45b)
where ao — Ji/(mca).
The reader may wish to verify for himself that the wave function
<Pio(x) does satisfy the wave equation (42a), and that it is normalized
to unity, which means that the integral over all of space of the Fig* 45B Term scheme for a hydrogen-like atom.
square of the wave function equals unity. The levels have been arranged into columns corre¬
sponding to different values of the orbital angular
momentum quantum number l All electric dipole
46 Our discussion so far has been based on the assumption that
transitions between levels ot principal quantum num¬
the nucleus remains fixed at the origin. We can very easily general¬ ber four or less are shown. In such transitions l must
ize our discussion to the case when the nucleus also moves. Let the change by one unit. Note that the state 2s cannot
nuclear mass be M, and let the mass of the electron be m. The decay by an electric dipole transition: the level is
reduced mass /a of the nucleus-electron system is then given by metastable.
The above term scheme should be compared to the
alkali term schemes in Figs. 28A and 32A, Chap. 3.
t Compare vvith the discussion in Secs. 30-31 and 54, Chap. 3. There are many similarities.
342 Theory of Stationary States Sec* 8*47
^ H1+sr
in accordance with our discussion in Sec, 36,
<**
The problem of studying the motion of two particles in their
center of mass system, moving under the influence of a force
describable by a potential which depends only on the distance
between the two particles is completely equivalent to the problem
of studying the motion of a single (fictitious) particle, carrying the
reduced mass of the system. This particle moves in sl fixed potential
field of force, described by the original potential as a function of
the separation of the particles. In order to take the nuclear motion
into account we must thus replace the mass m everywhere in our
formulas by the reduced mass ju. The energy levels of the system
are given by
£*= (46b)
<46c)
Av (x2) = f
J—®
dxx2 |i//(x)|2 (49b)
(49c)
or
50 Let us think very carefully about what the above really means.
The probability interpretation of the Schrodinger wave function
farces us to define the average of the position variable x as in Eq*
(49a). The integral in the right-hand side of this equation thus
permits us to find the numerical value of the average of the
quantummechanical position variable x, given the wave function
which describes any particular state of the particle. But what is
the numerical value of the ^quantum-mechanical variable x itself ”?
The answer is that a quantum-mechanical variable does not have
a numerical value: it is defined only in terms of a procedure whereby
its average can be computed for any given wave function.
The position variable x is a particularly simple variable in the
Schrodinger theory, and for this variable the full implication of
the basic principle that quantum-mechanical variables are defined
through their averages (for all states) is not immediately obvious.
The symbol x also occurs as an independent variable in the wave
function, and therefore the definition (49a) might not strike us as
being particularly profound* Consider, however, the quantum-
mechanical momentum variable (which we shall denote by p). The
symbol p does not “occur” in the wave function, and in view of
this we may at first wonder whether a momentum variable “exists”
at all. To settle this question we shall define the quantum-
mechanical momentum variable p through a definite prescription
whereby the average of p can be computed for any given state.
The real problem is therefore whether we can define the average
momentum in a physically reasonable way,
*K*) = pW*)
and, since the wave function is normalized to unity, we have
p = (51d)
dx
P2 = -*2 § (52a)
where p = Av (p).
Sec. 8.53 Theory of Stationary States 347
Notice that the same argument which led us to define the average
momentum as in Eq* (51c) applies to the definition of the average
of p2, as in Eq* (52b)*
- iL- (54a)
2m 2m dx2
fl=l;+vw=-isiiP + vw
in accordance with our discussion in Sec, 10 of this chapter.
cl Av ^ = m Av ^ (56a)
= m (56c)
ot
two variables Q' and Q". We now have available a precise defini¬
tion of Ax, as given by Eq. (49e), and a precise definition of Ap, as
given by Eq. (52d). We could, without too much difficulty, prove
the precise uncertainty relation
AxAp^Z (58b)
i*e*, prove that the inequality (58b) holds for all wave functions,
and that furthermore there are some wave functions for which
(58b) holds as an equality* We shall not undertake this task here
since we already have a good qualitative understanding of why a
relation such as (58b) must hold, and that is sufficient for this course.
Problems
2 For the sake of argument the author makes the following assertion
(inspired by some attempts to “explain*' quantum mechanics in certain
popular writings). The probability density P(x) — |2 for a stationary
state represented by the wave function can be understood as repre¬
senting the thne-averagp of the probability density for a particle moving
classically in the potential, with the energy of the stationary state. In other
words: the particle moves classically, but if we average this motion over a
time which is large compared with the natural period of the motion, then we
are led to the probability density P{r). For a particle in three-dimensional
motion, for instance an electron in the hydrogen atom, we can give a similar
interpretation of the absolute square of the wave function representing
a stationary state. The particle moves classically, but our measuring instru¬
ments are too crude to follow the details of the motion, and we therefore
observe instead a probability distribution of the electron in the atom, and
this can be understood as arising from averaging the classical motion over
a large time.
The reader will note that this assertion, interpreted literally, can be im¬
mediately refuted, The author therefore backs away a bit. He asserts
instead that whereas this interpretation of the square of the wave function is
not strictly correct, it is nevertheless a very useful way of thinking about the
quantum-mechanical motion of a particle; it gives us real insight into what
goes on, provided it is understood in an approximate sense.
It is the task of the reader to utterly refute these ideas; both the naive
first assertion and the modified second assertion. In so doing the reader
should take into consideration the discussion in the beginning of this chapter,
as well as our discussion of “double-slit experiments” in Chaps. 4 and 5.
352 Chap. 8 Problems
and determine the correct exponent fc* Your expression should be carefully
compared with the expression (37a), which describes the isotope effect on the
vibrational spectrum* The dependence on the isotopic masses is different in
the two cases*
10 (a) For the potential U(r) shown in Fig. 30A we note that the spacing
between neighboring levels decreases as the quantum number n increases.
Explain in a qualitative way why this is so,
(b) Draw a parabola representing the potential for a strictly harmonic
oscillator* On the same graph draw two other potential curves, symmetric
with respect to the origin, representing two “almost harmonic" potentials,
and such that the radius of curvature of all three curves is the same at the
origin (= the minimum of the potential). These two curves are to be drawn
such that for the first one of these the spacing between neighboring levels
increases with the quantum number n, whereas for the second one the
spacing between neighboring levels decreases with n. It is not necessary tx>
find the levels explicitly, but you should explain why the two curves have
the stated properties.
12 The mean life of the 2p-state in hydrogen is 0*16 X I0_s sec* What
is the mean life of the 2p-state in singly ionized helium?
Problems Theory of Stationary States 355
15 Find the “Bohr radius" of (a) a muonic aluminum atom; (h) a muonic
lead atom, and compare these radii with the nuclear radii. It is interesting to
make this comparison, because if the “Bohr radius" turns out to be com¬
parable with the nuclear radius then we can clearly not regard the nucleus
as being a charged point without extension, which means that the energy
levels of the muonic atom cannot be given precisely by a formula like (46b).
It has been found, experimentally, that the level systems of heavy muonic
atoms deviate considerably from the prediction (46b) * Through a systematic
observation of these deviations it has been possible to draw definite con¬
clusions about the charge distribution in nuclei, and about the sire of nuclei.
Problems 401
Chapter 9 The Elementary Particles
and their Interactions
Of =■ — (2a)
n
where P is the probability that an A-particle incident perpendic¬
ularly on the layer undergoes some interaction with one of the
B-particles such that it is removed from the incident beam. In this
definition it is essential that the layer be sufficiently thin so that the
observed probability P will be small compared to unity. (We shall
elaborate on this point in Sec. 4.)
This equation must hold for all positive real numbers n\ and n2.
Its general solution is
lim^= C (4d)
»->o ri
A + B —» C + D (5a)
The reaction cross section oae-»cd for this process is then defined
by
0AB-*)D = (5b)
where Pab-*cd is the probability that the reaction (5a) takes place
when an A-particle is removed from the beam through an inter¬
action with a B-particle in the target. Let us assume that (5a) is
the only reaction, Le., the only inelastic process, which can take Fig*5A Graph showing anti proton-proton cross
sections as functions of the Kinetic energy of the anti¬
place. A particle can, however, also be removed from the beam by
proton. The three experimental points denoted by
an elastic scattering, in which the A- and the B- particles remain open circles were obtained in the experiment de¬
after the collision. We define the elastic cross section oe by scribed in Fig. 2A, The cross section for proton*
proton scattering are shown on the same graph for
ae = aT Pf? (5c) comparison. Note that the total anti proton-proton
cross section is about twice as large as the total
where Fe is the probability that a collision event which removes a
proton-proton cross section.
particle from the beam is elastic. The three cross sections are thus This graph is taken from R. Armenteros et al.h “Anti-
related by proton-Proton Cross Sections at 1.0, 1.25, and 2.0
BeV/1 The Physics/ Review 119, 2068 (1960). (Cour¬
aT — + <Jab-»cd (5d) tesy of The Physical Review.)
(12a)
Sec. 9.13 The Elementary Particles and their Interactions 367
(12c)
where k is the same constant which occurs in (12a) and (12b)*
Considering the expressions (12b) and (12c) we can say the fol¬
lowing* In a sequence of repeated scattering experiments (in
which the A-particles always have the same initial momentum p*)
the ratio of the number of scattered particles emerging within the
cone of solid angle dQ to the number of particles incident on the
unit disc is equal to
i£o(x,f) = ^o(x^) +
~2Axp (t1 + 2ip/] exp {ixp) - exp (-ixp)) exp (-iuf) (16a)
370 The Elementary Particles and their Interactions Sec. 9.17
(16c)
(17b)
^ (l^c)
i / r/2 (18d)
p \ (u — a>o) + iT/2
and
(r/2)*
(18e)
(<o - u0)2 + (r/2)2
372 The Elementary Particles and their Interactions Sec. 9.19
Am ~ JL (21a)
Fig. 24A Yield curve for the reaction Al27 + p are very long compared to the lifetime of the neutral pion, or in
Si28 4- y, from a paper by K. J. Brostrom, T. Hu us, fact long compared to the lifetime of the neutron. Some of the
and R. Tangen, “Gamma-Ray Yield Curve of Aluminum
Bombarded with Protons,” Physical Review 71, 661
excited states decay through the emission of material particles,
(1947), The ordinate is a measure of the cross sec¬ and some decay through the emission of photons* Is it fair to
tion for the reaction. The abscissa is the kinetic exclude the excited states if we admit the "ground state” of 88Ra226t
energy of the incident protons, in keV, in the lab frame which also decays through the emission of a particle? Furthermore:
of reference. The sharp peaks are resonances. They perhaps some of the hyperons should be regarded as “excited states”
reveal the existence of excited states in the silicon
of the nucleon? (The hyperons all decay into other particles, one,
nucleus produced in the reaction, (Courtesy of The
Physical Review,) and only one, of which is a nucleon.) We find it very difficult to
resist these pressures, and we accordingly admit the “excited
states,”
*
tiveness of an atom as a scattered of light, as a function of the N,3/2
frequency of the light, we find sharp maxima at the frequencies
corresponding to the energy differences between the excited states
and the ground state. This phenomenon is, however, not restricted
to the scattering of light: we also encounter it in the scattering of
material particles. Figure 24A shows an example* The ordinate
is a measure of the cross section, and the curve thus shows the
experimentally measured cross section, as a function of the energy,
for the absorption of protons be aluminum. The sharp peaks in the
cross section reveal the locations of the excited states in the silicon
nucleus produced in the reaction.
The width T of a resonance peak measures the uncertainty in the
energy of the corresponding excited state. As long as the reson¬
ances are very sharp the interpretation of the resonances as mani¬
festations of excited states is clear-cut* We have agreed that such
excited states are “particles.” Let us now look at Fig* 24B which Fig, 24B The two curves show the observed cross
shows the cross section for the scattering of pions on protons, as a section in the scattering of positive and negative pions
against protons. The ordinate is the total cross sec
function of the energy* The cross section for positive pions shows tion in millibarns, and the abscissa is the totaf energy
one pronounced peak, as well as a slight “bump” at a higher energy* of the pion and the proton in the center-of-mass sys¬
The cross section for negative pions shows three moderately well tem of reference. It is convenient to express the
defined peaks* Do all these peaks correspond to particles? The in¬ energy in this manner, because the location of a
clination of many physicists today is to say that they do. The prominent peak directly corresponds to the mass of
the “ particle" or resonant state.
masses of these “particles” (?) are simply the abscissas of the Note the large peaks at an energy of about 1.238
maxima* BeV, This energy corresponds to a pion kinetic energy
of about 195 MeV in the laboratory system of refer¬
25 The dilemma which we are facing is where to draw the line* ence in which the incident pion collides with a proton
We certainly do not want to say that every small “bump” in a at rest.
We have denoted these resonances by the symbol
curve showing a cross section as a function of energy corresponds to A^/2- The notation A(1238) also occurs frequently in
a particle, but on the other hand, any rule according to which a the literature.
resonance must be “sufficiently” narrow if we are to accept it as
defining a particle is somewhat arbitrary* In other words; if an
object is to be admitted into the set of particles then its lifetime
cannot be too small, but where do we draw the line?
Let us re-examine our aims* Perhaps nothing is really gained by
trying to define precisely what we mean by a particle in general.
Our attempts have led us to a class of objects with millions of mem¬
bers, containing, among other particles, such qualitatively distinct
objects as pions and protein molecules* According to common
English usage these objects can reasonably be called particles, but
we can hardly expect to learn anything profound about fundamental
interactions if we try to treat pions and protein molecules as
equals in our basic theory* Some of the particles are obviously
composite systems, and we should describe them as such in our
376 The Elementary Particles and their Interactions Sec. 9.26
TABLE 26A The Leptons TABLE 26B The Principal Meson Octet
muons 77
f»+ 105.7 neutral pion 134.98 0.89 X 10-m
ye+ e~
e-neutrino 0 0
^ e-antineutrino 0 0 V
^ charged ft- it0
493.8 1.23 x 10-8
0 0 „ & mesons
rr
Vp ^-neutrino K~
77± tt+
Vp jti-antineutrino 0 0
7T+ ft~
0.87 X 10-™1
The muons are unstable and decay according to: K°| neutral f*1 IT^TT0
^ f + p, (Of the neutrinos ond is presum¬ *" K-Mesons < 497.9
ably a ^-neutrino, and the other an e-neutrino,) The X°J l** 5.68 X 10-* IT0 77° 7T°
mean life of the muons is 2.20 x 10-fl sec, The 7T+ rr~
other particles are stable. The leptons all have spin ft H V
angular momentum ft e v
7 7
TABLE 26C The Principal Baryon Octet 71
5T°
P7T()
2+ 1189.47 0.81 x 10-“
fl7T+
sigma- theory: we should “explain” them in terms of the interactions of
20 hyperons 1192.56 < io-« Ay their more elementary constituents.
From a practical point of view we can think about a hierarchy of
2“ 1197.44 1.65 X 10-10 tw7” increasingly more elementary particles, Depending on the kind of
i physical phenomenon we wish to consider our notion of the “ele¬
1314.7 1 3.0 x 10-“ Ait* mentary constituents” of a composite system changes. It is com¬
cascade
mon English usage to say that a molecule is a bound state of atoms,
r_ particles 1321.2 Ait-
1,7 X 10-“ that an atom is a bound state of a nucleus and a number of electrons,
These particles all have spin angular momentum and that a nucleus is a bound state of protons and neutrons. The
and baryon number +1. There exists an anti'baryon proton, neutron and electron are, however, not in any obvious way
octet consisting of the anti-particles of the above par¬ bound states of anything else: they might well be among the ulti¬
ticles. The anti-particles have the same masses,
mate elementary particles. As such they are objects of particular
spins, and mean lives, but opposite charges and
baryon number. interest in basic theory.
Sec. 9*27 The Elementary Particles and their Interactions 377
we shall not use absolute stability as a condition for admissibility. Electric charge
The elementary particles are divided into four classes. The Fig* 27A Mass spectrum of the meson octet to
which the pions and JC-mesons belong. These par¬
photon is the sole member of the first of these. The other classes ticles ell have baryon number zero and spin angular
are the leptonst the mesons and the baryons (including the anti- momentum zero. The two neutral tf-mesons and
baryons)* Tables 26A-C list some of the properties of the leptons indicated by the double line in the term scheme,
and the most respectable mesons and baryons.t (See also Table B have the same mass to the accuracy of this drawing.
in the Appendix.) Particle-antiparticle pairs are symmetrically situated
with respect to the vertical line corresponding to zero
charge. The particles ir° and t? are their own anti¬
27 In Figs. 27A-B the mesons and baryons listed in Tables 26B-C particles. R® is the anti-particle of
are shown in diagrams which strongly resemble the term schemes
discussed in Chap. 3. Each particle is represented by a short
horizontal bar on a graph in which the ordinate is the rest mass
(in MeV) and the abscissa is the electric charge. (The center of
the bar indicates the charge of the particle.)
According to current ideas our particle diagrams should be
>to 2+
regarded as entirely analogous to term schemes for atoms. Each
EQ
diagram corresponds to a “multiplet” of closely related particles .3 A
which might in some sense be regarded as different states of the
“general” particle of the multiplet. 2
Fig. 27C shows the anti-baryon multiplet of eight particles which
are the anti-particles of the eight baryons in Fig. 27B. The anti¬ 1.0^
particles of the mesons shown in Fig. 27A are contained in the n
same diagrams; we say that the meson octet is self-conjugate. The
0.9^ 1-1-T"
negative pion is thus the anti-particle of the positive pion and the - 1 0 + 1
Electric charge
Fig. 27B Mass spectrum of the baryon octet to
t The nomenclature for elementary particles appears to have been invented to give which the proton (p) and neutron («) belong. These
a certain classical Greek flavor to the subject. Although the author's knowledge of the particles all have baryon number +1 and spin angular
classical languages is extremely limited, he nevertheless feels that he has very good momentum The diagram can be interpreted as a
reason to suspect that the linguistic principles which underlie the construction of the term scheme, showing the eight different states of
“Greek”-sounding terms are not entirely correct. the "general particle'f associated with this multiplet
378 The Elementary Particles and their Interactions Sec. 9.28
p n
~1 I T"
- 1 0 + 1
Electric charge
ir~ + p ~> K° + A0
(0) (+i) (+i) (0) (29a)
is allowed by the hypercharge conservation principle, and it is well
known to occur readily whenever negative pions of sufficient energy
collide with protons* (The numbers below the symbols for the
particles show their hypercharges*) The reaction Fig* 29A Diagram showing electric charge and
hypercharge of the particles in the meson octet whose
7T- + p tT0 + A0 mass spectrum appears in Fig. 27A, The total hyper¬
(29b) charge is conserved in any strong or electromagnetic
(0) ( + 1) (0) (0)
interaction. The total charge is conserved in aH inter¬
is forbidden by the hypercharge conservation principle* This im¬ actions.
The pattern becomes particularly pleasing to the
plies in particular that a lambda particle cannot be produced in a
mind if the particles are plotted on a hexagonal co
pion-proton collision unless there is sufficient energy available for ordinate net as above. The Eightfold Way theory of
the creation of a £-meson according to the reaction (29a). No symmetries predicts that the pattern should be as
example of the reaction (29b) has ever been observed The reaction shown for such a plot. In particular the theory pre¬
dicts two particles in the center of the diagram, in this
n + p A0 + V case the particles tr° and ij.
(29c) The abscissa expresses another commonly used
(+1) ( + 1) (0) (+1) quantum number called the third component of iso¬
is also forbidden* Tliat it actually does not occur in nature as a topic spin. This quantity (denoted by Ja) is also con¬
served in all strong and electromagnetic interactions.
strong interaction is well established experimentally*
The particles in the diagram all have baryon number
0 and spin angular momentum 0.
380 The Elementary Particles and their Interactions Sec. 9.30
A0 iTT~ + P
(29d)
(0) (0) ( + 1)
which violates hypercharge conservation. The weak interactions
are responsible for this decay, as indicated by the comparatively
slow rate at which it takes place. The explanation for the long life¬
time (a mean life of 10™10 sec is long on a “nuclear” time scale)
of the lambda particle is that the principles of conservation of
baryon number and hypercharge prevent it from decaying in any
other way than through the weak interactions*
Mann that the particle denoted f2“ in Figs. 27D and 29D exists.!
(three) quarks, then the binding energy must be very large com¬
pared to the mass of the nucleon. The nucleon would thus be a
very tightly bound system* and in this respect it would differ radi¬
cally from the bound systems which are familiar to us, namely
atoms, molecules, and nuclei. (The binding energy of an atom, a
molecule, or a nucleus is small compared to the mass of the system.)
It is therefore safe to say that whereas nucleons might one day be
found to be composite, they are certainly not composite in the
same sense as, for instance, the deuteron is composite.
35 Today the mechanical ether has been banished from the world
of physics, and the word “ether” itself, because of its “bad” con¬
notations, no longer occurs in textbooks on physics. We talk osten¬
tatiously about the “vacuum” instead, thereby indicating our lack
of interest in the medium in which the waves propagate. We no
longer ask what it is that "really oscillates” when we study electro¬
magnetic waves or de Broglie waves. All we wish to do is to for¬
mulate wave equations for these waves, through which we can
predict experimentally observable phenomena. As we have already
said, these wave equations must be non-linear if they are to describe
Elementary Particles and their Interactions Sec. 9.35
velocity larger than c> and it follows that the action of the force
cannot be instantaneous. If the position, or state of motion, of W/W////////J
one particle suddenly changes it must take some time before this
’©
J
change is perceived by the other particle, and the minimum time it
must take is the time it takes for a light signal to pass between the Schematic picture of a liquid hydrogen bubble cham¬
two particles. ber. The chamber is activated by a sudden decrease
It is not at all a trivial matter to formulate a relativistically invari¬ in the pressure of the hydrogen liquid. The temper¬
ature of the liquid is above the boiling point at the
ant theory of interacting classical particles. A profound change in reduced pressure, but boiling does not start imme¬
the non-relativistic idea of an instantaneous action at a distance is diately: the liquid stays for a short interval of time in
required, a superheated stage. The passage of a charged par¬
ticle through the liquid leads to local vaporization
along the path. A visible track consisting of very tiny
37 One possible way out of this dilemma is through the intro¬
gas bubbles is formed, and it is photographed with
duction of a (classical) field. Each particle is the source of a field the camera above the chamber The pressure is then
which can propagate in space, but never with a velocity greater raised again and all tracks disappear. The chamber
than c, and this field may then influence the motion of other par¬ is ready for the next exposure, (//lustration by cour¬
ticles, In a relativistic classical theory of this kind we are thus led tesy of Lawrence Radiation Laboratory, Berkeley.)
386 The Elementary Particles and their Interactions Sec, 9.38
Bubble chamber picture showing the production and p + — A0 + K9. The decay interactions are all
decays of a (neutral) lambda particle and a neutral weak. The negative muon emitted in the decay of
K-meson. The various tracks are identified in the the K° decays into an electron, a neutrino, and an
insert. Only the charged particles leave visible tracks, antineutrino. The last two particles are neutral and
and these are curved because the chamber is in a cannot be seen. (Photograph by courtesy of Lawrence
magnetic field. The strong production reaction is: Rad/arton Laboratory, Berkeley.)
-r
388 The Elementary Particles and their Interactions Sec. 9,40
f The fields are actually not "ordinary" complex-valued functions of position and
time. They are mathematical objects known as “operator-valued distributions.”
However, for our purposes we can think about them as ordinary functions (representing
“sound waves in the non-linear ether” )♦
390 The Elementary Particles and their Interactions Sec, 9.43
Bubble chamber photograph showing the production quently collides with a proton and annihilates into
and subsequent decay of a lambda-antilambda pair, pionsr of which four are charged and leave visible
The drawing in the upper right corner identifies the tracks.
tracks of the various particles. An incoming anti’ This photograph is shown in the middle of our dis¬
proton collides with a proton and produces the lambda- cussion of quantum fields to remind the reader that
antilambda pair. The latter particles do not produce one of the objectives of quantum field theory is to
visible tracks since they are neutral. The lambda give us a theoretical understanding of events such as
decays into a negative pion and a proton (via the weak those seen in the photograph. (Photograph by cour¬
interaction) and the antilambda decays into a positive tesy of Lawrence fiadiafion Laboratory, Berkefey:)
pion and an anti-proton. The anti-proton subse¬
392 The Elementary Particles and their Interactions Sec, 9.45
fore not been found possible to solve the field equations which have
been proposed and we cannot tell whether these equations are
really correct. Most likely they are not. There is actually an
infinite latitude in the selection of equations, and our only guiding
principle in the past has been the “principle of simplicity.” In
quantum electrodynamics we have been further guided, in a de¬
cisive way, by the classical analog of charged billiard balls inter¬
acting with the electromagnetic field,
48 The reader has very likely heard that these particles do exist,
and that they are none other than the pions* In the days of
Yukawa's work, however, no mesons were known and his sugges¬
tion that they existed was truly a prediction* He knew the two
394 The Elementary Particles and their Interactions Sec . 9.49
(50b)
V2 V(r) 1 dfrdVtf
(51e)
r2 dr l dr
1 d U dV(r)
m^2V(r) (52a)
r2 dr \ dr
where C* is a constant.
Our rejection of the second solution illustrates again an import
tant principle which we have encountered before: not every solu¬
tion of the wave equations of quantum mechanics has a physical
meaning. The wave functions which are physically meaningful
not only have to satisfy the wave equation but they must also satisfy
a number of boundary conditions, one of which is that the solution
must not increase indefinitely at infinity.
Bubble chamber photograph showing the annihilation track is hard to distinguish from the pion track, but
of a proton and an anti-proton into pions. The main the beginning of the positron track can be clearly seen.
event takes place in the middle of the field of view. The chamber is in a magnetic field perpendicular
The anti-proton is incident from below, and its path to the plane of the picture. The tracks of negative
is shown by the almost straight “dotted" track. Eight particles turn in the clockwise direction and the tracks
charged pions are produced in the annihilation. One of positive particles in the opposite direction. Slowly
of these, the one whose track is initially directed against moving particles leave dense tracks, whereas the
the direction of the incoming anti-proton, decays into tracks of very fast particles tend to have a “dotted"
a muon and a neutrino. The muon subsequently de¬ appearance. (Photograph by courtesy of Lawrence
cays into a positron and two neutrinos. The muon Radiation Laboratory, Berke/ey.)
400 The Elementary Particles and their Interactions Sec. 9.56
general type as the one given by the Yukawa potential, The fact
that we have employed a linear approximation therefore does not
invalidate our main conclusion which was that the range of the
force is inversely proportional to the mass of the particle exchanged.
Concluding Remarks
Problems
2 The total cross section for the interaction of a K+-meson with a proton
is about 15 mb when the kinetic energy of the K-meson (incident on a proton
at rest) is 400 MeV. What is the average number of interactions per cm of
path for a K-meson of this energy in liquid hydrogen (in a bubble chamber)?
The density of liquid hydrogen is 0.071 gm/cm3.
4 At a gamma ray energy of 100 keV the cross section for Compton
scattering was measured in one experiment as 0.49 bam. At this energy,
which is considerably smaller than the rest energy of an electron, a simple
402 Chap. 9 Problems
6 Using our simple theory in Secs. 17-18 for resonant scattering, esti¬
mate the cross section for the resonant absorption of gamma rays of energy
14.4 keV by the Fe57 nucleus. (This estimate is relevant to the experimental
results presented in Fig. I6A, Chap. 4.) Assuming that the absorbing iron
nuclei are in a foil 1 mil thick, what is the probability that a gamma ray passes
through the foil?
Note that our simple theory does not really apply to photons, among other
things because photons have spin angular momentum one. You can there-
The Elementary Particles and their Interactions 403
fore not expect to obtain a numerically correct value for the cross section.
The dependence of the maximum cross section on the wavelength is how¬
ever given correctly by our theory, and your estimate is therefore useful as
an order of magnitude estimate,
8 Consider the particles forming the baryon octet for which the mass
spectrum is shown in Fig, 2715 and the Eightfold Way symmetry diagram in
Fig, 29B, Of these particles one is stable. Among the remaining unstable
particles one decays through an electromagnetic interaction (it has a notably
shorter lifetime than the other particles) and the others decay through the
weak interaction. See whether you can account for these features of the
octet in tenns of the conservation laws for baryon number, charge, and hyper¬
charge which we have mentioned* To do this you should investigate all
possible decays you can think of into particles which have been mentioned
in the text, taking into account the experimentally determined masses of
these particles. For example: you might begin by asking whether the 2+
particle could decay into a K+-meson and something else. You will soon
discover that the possibilities are severely limited, and that there are not too
many cases to consider.
The problem is thus to show in detail that the conservation laws which
404 Chap, 9 Problems
we have discussed imply that none of the particles can decay through the
strong interaction, and that only one can decay through the electromagnetic
interaction,
9 The symmetry diagrams in Figs, 29A-D show the values for the various
particles of a quantity called the third component of isotopic spin (denoted
h)- We have mentioned that this quantity is also conserved in all strong
and electromagnetic interactions.
Investigate whether this conservation law implies anything more than the
other conservation laws which we have mentioned and which concern con¬
servation of charge, hypercharge, and baryon number*
at large distances from the scattering center. Show that for the special case
of spherically symmetric scattering, in which f{0) = / is independent of the
scattering angle 6, the wave function given in (a) is actually a solution of the
KJein-Gordon equation (except for the point x = 0) in empty space. It will
be helpful to consider our discussion in Secs, 51-52 in this connection,
(b) Show that for an arbitrary f[8) the expression in (a) is an approximate
solution of the Klein-Gordon equation. You should show that if this wave
function is substituted into the Klein-Gordon equation, then the equation
is satisfied except for an error term which tends to zero as 1/x2 when x tends
to infinity.
Appendix
For tables of units and conversion factorSj see inside back cover of book.
For a table of crude values of important physical constants, see inside front cover of book.
406 Appendix
- = 137.0388 ± 0.0006
a
* Most of the data in this table are taken from an article by E. R. Cohen and J. W. M. DuMond, “Our Knowledge of the Fundamental
Constants of Physics and Chemistry in 1965,” Reviews of Modem Physics 37, 537 (1965).
Appendix 407
Rydberg constant for infinite proton mass: fl«,= -q~- = = (109737.31 ± 0.01) cm-1
Bohr magneton: _ eH
^B 2me
— (9,27314 ± 0.00021) x 10-21 erggauss-
Important Decays t
Mass Mean life Partial Branching Q
Particle Spin MeV sec mode fraction MeV
r photon 1 0 stable stable
LEPTONS
e- neutrino 1 0 « 0.2 keV) stable stable
v* ji-neutrino 2 0 « 2 MeV)
e+ electron-positron 1 0.511006 stable stable
2
Important Decaysf
MESONS
0 139.58 2.608 x 10-8 pv 100% 34
charged pions
ev 1.24 X 10-* 139
pvy 1.24 x 10-4 34
1.0 X 10-8 4.08
The numbers within parentheses in the atomic mass column arc the mass numbers of the most stable isotopes of radioactive elements.
Appendix 411
The numbers within parentheses in the atomic mass column are the mass numbers of the most stable isotopes of radioactive elements.
Index
Barkla, C. G., 155 Broadening of spectra lines, 127 (Se£ also Line widths)
Bam, 362 de Broglie, VP, 181
Bames, V, E., 381 biography, 183
Barrier penetration, 288 theory of matter waves, 181
and alpha-radioactivity, 392 wavelength, 185-187
optical analog, 290 Brostrom, KP JP, 374
transmission coefficient, 292 Bubble chamber design, 385
Barrow, G. M., 350 Bubble chamber picture, charge exchange scattering, 162
Baryon number, 379 antisigma-zero, 384
Baiyons, 376-381, 408 lambda and jK-meson, 387
Bearden, J. A,, 158 lambda-antilambda pair, 391
Becquerel, H,, 39 proton-antiproton annihilation, 399
Bellicard, J* B., 365 Burbidgc, E* M., 307
de Benedetti, S., 137 Burbidge, C. R., 175, 307
Beta decay, 125, 300
of cobalt, 127 Cadmium atom, energy levels, 120
evolution of heat, 310 Calorie, 52
Beta particle, beta rays (see Electron- Beta decay) Cascade particle, 376
Beta spectroscopy, 246 Cascade shower, 8
Bethe, H+ AP, 401 Cerenkov counter, 176, 358
Bilaniuk, O. M., 262 Cerenkov radiation, 176
Binding energy, in atoms, 58, 62, 102, 123, 230 Cerium atom, portion of spectrum, 111
of deuteron, 103, 352 Cesium chloride crystal structure, S3
of molecules, 64, 330 Chadwick, J+, 66
of nuclei, 66,102, 232 Chain reaction, 89
Black-body radiation, 23 Chamberlain, (X, 160
Planck's radiation law, 26 Chambers, R. G*> 203
Bloom* A. L,, 137 Characteristic X-ray radiation, 156
Bohr, N., 35, 95 Charge, of electron, 19, 20,45, 57,406
biography, 35 of elementary particles, 376—381
correspondence principle, 349 fractional, 381
prediction of hafnium, 121 of nuclei, 66, 123
theory of atoms, 32, 58, 139 Charge conservation, 80, 378
theory of emission and absorption of light, 95 Charge exchange scattering, 162
Bohr magneton, 90 Charge independence of nuclear forces, 71, 124
Bohr radius, 35, 58, 407 Chemical elements, 46
Bohr-Sommerfeld quantum condition, 328 origin, 303
Bolef, D. I., 151 periodic chart, 121
Boltzmann, L., 6 shell structure, 122
Boltzmann’s constant, 23, 52,406 tables, 45, 46, 410
determination, 27 Chemical reactions, 6
Boorse, H. A., 39, 122, 174 bulk reaction energies, 65
Born, M., 215,272 Chew, G, FP, 401
biography, 272 Classical physics, 2
Bom*Oppenheimer approximation, 331 limiting case of quantum physics, 229,349
Boron nucleus, energy levels, 124 Classical radius of electron, 13, 34, 407
Boundary conditions for wave functions, 282, 284, 287, Cloud chamber picture, alpha emission, 293
304, 314, 318, 397 cascade shower, 9
Brackett series, 340 positron, 158
Branching ratio, 120, 125 Cohen, E. R,, 85,406
Breit-Wigner formula, 109, 371 Coherent superposition, ensembles, 249
Bremsstrahlung, 10, 156 light, 253, 262
Brillouin, L>, 328 Collins, G. B., 175, 262
Index 415
: 1,6021 1.0736 2.
1 I
x 10 13 XlO9 X
1 0303 0.3499
XlO5 x 10 171 XlO7 XlO
1 barn (b) =
I millibarn (r = iO~2T cm2
Energy equivalent of
atomic mass unit: (1 amu) X c2 = (9.31478 ± 0. 00005) x 108 eV
Activity of radioactive
sample: 1 curie = 3.7 x lO10 disintegrations per second