Elmer E. Anderson - Modern Physics and Quantum Mechanics-W. B. Saunders Company (1971)
Elmer E. Anderson - Modern Physics and Quantum Mechanics-W. B. Saunders Company (1971)
Elmer E. Anderson - Modern Physics and Quantum Mechanics-W. B. Saunders Company (1971)
PHYSIOS
AND
QUANTUM
MEOHANIOS:
ELMER E. ANDERSON, Ph.D.
Chairman, D"port*ent of Physics,
Clnrkson College of Technobgy
Potsdarn, New York
l2 Dyott Street
London, WCIA lDB
1835 Yonge Street
Toronto 7. Ontario
e D7l by W. B. Saunders Company. Copyright under the International Copyright Union. All
rights reseived. This book is protected by copyright. No part of it may be reproduced, stored in a
relrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying,
recording. or otherwise, without written permission from the publisher. Made in the United States
of America. Press of W. B. Saunders Company. Library of Congress catalog card number
77-145554.
Print No.:
ffiffiffiffiffiffiffi
the hvdrogen atom. The complexities resulting from spin and its interactions
u-ith orbital angular momenta are treated in Chapter B. The phenomenon
of exchange degeneracy resulting from the symmetry properties of identical
particles is also discussed in this chapter. Chapter 9 is an important chapter.
Here some approximate methods are introduced which enable the reader to
solrre large classes of real problems by means of the solutions obtained for the
ideal systems in earlier chapters. In Chapter 10, approximate methods are
applied to some specific atomic problems, and Chapter 11 is an introduction
to quantum mechanical scattering theory.
The book contains over 250 problems in addition to the numerical ex-
amples which are worked at appropriate points in the text. The importance
of solving a large number of problems in order to achieve a mastery of the
material cannot be overemphasized.
It is a pleasure to acknowledge the help of the following people and to
thank them for their indispensable roles in the preparation of this book: my
students over the past few years for their enthusiasm and inspiration; Agatha
Hollister and Amelia Anderson for typing and preparing the manuscript;
Rita Arajs for making all of the drawings; my colleagues, Professors Fred
Otter and Sigurds Arajs, as well as Professors Henri Amar of Temple University,
Paul H. Cutler of the Pennsylvania State University, Thomas B. W. Kirk of
Harvard University, John Reading of Northeastern University, John G.
Teasdale of San Diego State College, and an anonymous reviewer, for their
most valuable critical comments; to John J. Hanley and others of the staff
of the W. B. Saunders Co. for their assistance and encouragement; last of all,
to my mother and to my family-Amy, Kenneth, Mark, Scott, Ruth, and
Carl-to whom I dedicate this book.
Potsdam, Ir{ew York E. E. A.
wffiffiffiffiffiffiffi
Chopter 2
APPENDICES A To C 65
Chopter 3
,
no
THE CONCEPT OF THE NUCIEAR ATOM lc
l. The Atomic Models of Thomson and Rutherford . tc
2. Classical Scattering Cross Sections . lc
3. Scattering Cross Sections in the C and L Frames .80
4. Bohr's Theory of Atomic Spectra ,83
5. The Franck-Hertz Experiment . . .88
6. X-ray Spectra and the Bohr Theory .89
7. Nuclear Structure and Spectroscopy o9
8. Fundamental Forces and Exchange Particles .99
9. Angular Momenta and Magnetic Vloments l0r
10. Tlre Larmor Theorem and the Normal Zeeman Effect . . . 104
I l. Spatial Quantization . . l. . . 107
ll0
ll0
Chopter 4
The theory of relativity is concerned with the way in which observers who
are in a state of relative motion describe physical phenomena. The idea of
an absolute state of rest or motion has long been abandoned, since an observer
"at rest" in an earth-bound laboratory is sharing the motion of the earth
about its axis, the earth about the sun, the solar system through the Milky
Wuy, and so on. It is also common knowledge that one can perform simple
experiments with bouncing balls, oscillating springs, or swinging pendula in a
laboratory fixed on the earth or in a smoothly running truck moving at
constant velocity and obtain identical results in both sets of experiments.*
We describe this fact by saying that the laws of motion are coaatiamt, that is
they retain the same form when expressed in the coordinates of either frame of
reference.
Consider two frames having relative translat'ional velocity V in their
common r-directions. It is evident that the coordinates are related by
t(':x-Vt
)' :)
nl-n
/-
tt:t,
* This is true only for experiments that are insensitive to the rotation of the earth.
THE SPECIAL THEORY OF RELATIVITY
zz'
Figure l-l Two inertial frames of reference.
Equation 1.1 says that Newton's laws have the same form in both coordinate
frames, that is, they are covariant. Note that the coordinates and velocities
are different in the two frames but each observer always knows how to obtain
the other's values. Frames in which Newton's laws are covariant are called
inertial frames. Inertial frames are equiualent in the sense that there is no
mechanical experiment which can distinguish whether either frame is at rest
or in uniform motion; hence, tltere is no preferredframe. This is known as the
Galilean or classical principle of relatiuiQ and the coordinate transformation given
above is called a Galilean transformation. Strictly speaking, the earth is not an*
inertial frame because of its rotation and its orbital motion, but it can often
be treated as an inertial frame without serious error.
As a consequence of the principle of relativity, observers in different
inertial frames would all discover the same set of mechanical laws, namely,
\ewton's laws of motion. By way of illustration, consider the following thought
experiment. An observer, riding on a flat car which is moving at constant
velocity past an observer on the ground, fires a ball vertically upward from a
small cannon and notes its maximum height as well as the total time of flight
fior it to return to the muzzle of the cannon. The observer on the ground also
ELECTROMAGNETIC THEORY, GALILEAN TRANSFORMATION
measures the maximum height and the time of flight. Although the two
observers will not agree on the shape of the trajectoiy of the ball, they wilt
agree on the height, the time, the calculated muzzle velocity, and the value
of g, the gravitational acceleration.
PROBI.EM ,jliilii:ilia1jLii,:ii.
'-'
Show that the equation V .i : p is not covariant under
a Galilean transformation. That is, assume the covariance
of the Lorentz force, F : q(& + I o x h\ and show that
this leads to a contradiction. Hint: require p' : pi
otherwise, a simple measurement of charge would distinguish
one frame from another.
True
Positiorlgp-
l7
I
t Apptrrent
Position
!u A,t
months later, the telescope should be tilted the same amount in the opposite
direction. See Figure 1-2. An important conclusion can be drawn from the
fact that aberration of the correct amount is experimentally observed, namely,
that the earth mlaes through the ether. The ether, then, was assumed fixed with
respect to the stars.
b. A number of experiments were devised to measure the earth's motion
through the ether, that is, to detect an ether wind with respect to the earth.
The most famous of these was the Michelson-Morley experiments which will
be described below. This experiment was repeated many times but always
with a null result, that is, no ether wind was detected. Another famous but
unsuccessful attempt to detect an ether wind was the experiment of Trouton
and Noble. a They sought to detect a torque on a charged capacitor due to its
motion through the ether.
c. All attempts to explain the absence of an ether wind while retaining
the ether concept were also refuted. It could not be argued that the earth
carries the ether along with it since this would contradict the observed aber-
ration of starlight. Some proposed that the velocity of light adds vectorially
to the velocity of the source. This theory (historically called the emission
theory) would easily account for the null result of the Michelson-Morley
experiment, but it contradicts a known fact about wave motion, namely,
that the velocity of a wave depends only upon the properties of the medium.
Perhaps the most convincing evidence for rejecting the emission theory comes
from the study of the periods of binary stars as proposed by De Sitter. s The
relative velocity (to an earth observer) of a distdnt star performing circular
motion will vary sinusoidally with time. Therefore, the measured Doppler
shift* of the light emitted by such a star would show the same sinusoidal
3 A. A. Michelson, Amer. Journ. of Science (3),22t 20 (lBBl) ; A. A. Michelson and E. W. Morley,
ibid.34r 333 (lBB7). These results were confirmed by R. J. Kennedy, Proc. Nat. Acad.l2r 62l (1926)
and K. K. Illingworth, Phys. Reu.3Or 692 (1927).
4 F. T. Trouton and H. R. Noble, Phil. Trans. Ro1. Soc., A2O2r l65 (1903) ; Proc. Ro2. Soc. (London),
72, t32 (1903).
5 W. De Sitter, Proc. Amsterdam Acad.161 395 (1913).
* The mathematical details of the Doppler shift, which will be discussed in section 11 of this
chapter, are not required for the present argument.
ELECTROMAGNETTC THEORY, GAL|LEAN TRANSFORMATTON
variation if the speed of light were constant. On the other hand, if the emission
theory were true, the speed of the emitted light would be greater as the star
approached and less as the star receded. In this event the measured Doppler
shifts would no longer show a sinusoidal variation but the curve would be
distorted as in Figure 1-3(a). An actual plot of the Doppler shifts measured
for the components of the binary star Castor C is shown in Figure l-3(b).d
The absence of distortion in the curves is strong evidence lor the validity of
the constancy of the speed of light. Many such binaries have been studied
and there is no indication of distortion other than that due to known eccentric-
ities in the orbits. More recent experiments which refute the emission theory
of light are measurements of the velocity of light in the form of gamma rays
emitted from positron annihilation? and from the decay of zo mesons.s
A third attempt to explain the null result of the Michelson-Morley
experiment was the Lorentz-FitzGerald contraction theory. It was postulated
that a length is contracted by a factor (l - r'lt\ l, where a is the component
of the velocity parallel to the length and c is the velocity of light. This con-
traction is just the right amount to explain the null Michelson-Morley result,
although it was not derived from first principles. The Lorentz-FitzGerald
.n
60
a.
+
r20
(,
c)
vE o
Prirnrrrv
tio
0.4 0.8
Time-days
b
Figure l-3 The measured Doppler shifts due to the orbital velocity of astar performing circular
motion would appear as in (a) if the speed of light were ro increase as the star approached and to de-
crease as the star receded from the observer. Shown in (b) are the relative velocities of the components
of the binary star Castor C as determined from the Doppler shifts of their spectra. The absence
of distortion is strong evidence for the constancy of the speed of light. (The data are from A. H. Joy
and R. F. Sanford, Astrophysical lournol 64,250 (1926).)
contraction was extremely difficutt to refute,e and its physical significance did
not become clear until after the work of Einstein'
t_
/r- * lz : ?tt, :21, (r t ,-,4c (I + pr),
,^ - r62)-1
b2*-
c _aTfia:-p -az - c
where 0 : ulc. Since the velocity of the earth in its orbit is about 30 kilometers
per second,'F 10-a and terms high.. than pz are neglected in the expansion.
^For -
the puit, that is perpendiculir to the relative motion the situation is
analogous to the problim-of rowing a boat across a river having a current u.
lrlt :1/rz rz, and the time for a round trip transverse to the current is
-
1,
------+
_ ] Cornpensator
Half-silvered surface
Source
To eyepiece
Figure l-4 Michelson interferometer with arm f, parallel to the earth's motion.
The phase difference between the two beams as they enter the eyepiece is
Ad, : QrrlD.(pathdifference) : (2nl)').c(tr-tr):(2n11).12(/,
- fr) +
0'(2/, - {t)1. By rotating the interferometer through 90" the roles of the two
beams are interchanged and the phase difference for this case is
L6, :a 2n
l2(/,- /,) + p,(2/,- ()1.
Thus, the effect of rotating the instrument through 90'is to enable the observer
to detect a total phase difference of A{, -l L6r. This corresponds to a fringe
shift of
2"
t
(Ad' f Ad,) :#,t, * /r). (1.2)
For lt
- fz :10 meters and ,i ^,5000 A, Equation 1.2 predicts a shift of
about 0.4 fringe. Such a shift could be easily detected, as Michelson's instru-
ment could resolve 0.01 of a fringe at that wavelength.
The Michelson-Morley experiment was repeated many times, at all hours
of the day and in all portions of the earth's orbit, yet no fringe shift has ever
been observed. The null result requires us either to accept the non-physical
Lorentz-FitzGerald contraction or to reject the ether theory, since the ether
cannot be fixed with respect to the earth as well as the distant stars.
THE SPECIAL THEORY OF RELATIYITY
PROBLEM 1"2
This value was adopted by the International Union of Geodesy and Geophysics
in 1957.14
Before deriving the new relativistic transformation, let us look briefly at
the concept of time. According to classical physics, t : t', and it was tacitly
assumed that clocks in different inertial frames could be easily synchronized.
The concept of simultaneity was not questioned. Einstein, however, realized
that the process of synchronization requires the transmission of signals at a finite
velocity which can be at most the velocity of light in free space. Furthermore,
simultaneity is a meaningful concept only in a frame in which light sources,
clocks and observers are all at rest. By way of illustration, suppose observer O
has measured two equal distances tr1 and x, from his reference position and
installs identical flash lamps ,S, and ,S, as shown below:
OfO -'+v
Stlt- xr X2 -45,
o
He can fire these flash lamps simultaneously by means of matched cables or
radio operated relays, and he will assert that the flashes reach him simultane-
ously. However, to a second observer O', who is in a state of relative motion
parallel to the line joinins Sr and Sr, the flashes will not appear to be simultane-
ous because of the finite distance through which he moves during the transit
time of the light waves. Thus, if the flashes occur just as O'passes O, then O'
will claim that S, flashed earlier than Sr.
Since simultaneity cannot be defined independently of spatial coordinates
for observers who are in relative motion, their clocks are not synchronized and
their measured times are different. As a consequence of this, taken together
with the constancy of the observed speed of light in either frame, we conclude
that the statement t' : / is no longer valid.
s'-C(x -ut))
)' :) l (1.3)
z'-7, )
where C --- I as u ' 0, and where u is parallel to the common direction of the
x and "r' axes. Since the transformed time must, in general, depend upon both
IO THE SPECIAL THEORY OF RELATIYITY
However, by Einstein's postulates, the primed observer also must see a spherical
wave emanating at speed c from iis origin as its center. If he were not to see a
spherical wave but, say, an ellipsoidal wavefront, he would have a means of
determining which frame were moving with velocity u. The equation of the
wave front in the primed system is,
C2t'2 - X'2 *)'2 * Z',
Then we have,
,c2 +y + 22 - c2t2: x'' *)'2 i z'2 - czt',L (1.5)
6. RELATIVISTIC KINEMATICS.
We define an event as a physical occurrence that is localized in space and
time. That is, to an observer who is at rest in the frame of the event, the time
RELATIYISTIC KINEMATICS II
interval of the event can be measured by means of a single clock and the position
of the event can be determined by means of a meter stick. We call the time
interval in the rest frame the proper time interual, rr.
If the event consists of the measurement of the length of an object in its
rest frame, the coordinates of the two ends of the object can be measured,
simultaneously by an observer in that frame. We call the length of the object
measured in this fashion its proper length, Lo. To an observei in any ineitial
frame other than that of the event, a measurement of either a length or a time
interval will require the use of both a meter stick and at least oni clock. We
will now look in detail at the way lengths and time intervals transform under
a Lorentz transformation.
a. Preservation of lengths perpendicular to the direction of relative motion.
Consider the following thought experiment. The lengths of two meter sticks
are known to be exactly the same in a frame where they are both at rest.
If they are now placed in relative motion in a direction perpendicular to their
lengths, as in Figure 1-5, what do observers O and O' conclude about their
lengths ? O can install synchronized flash cameras at each end of his meter
stick which he can trigger just as O' passes him. O' will agree that the flash
Pictures were simultaneous since he is at all times equidistant from the two
flash lamps. Since both observers agree that the flashes were simultaneous,
!h.y must agree with the result shown on the photographs. That is, the res.rli
is an absolute one, not a relative one. Now the experiment can be repeated
with the flash cameras in the primed frame and once again both observers will
agree that the flashes were simultaneous. The same result must be obtained
in both sets of experiments. If, for example, the stick in the unprimed frame
were found to be shorter in the first set of pictures, then the stiik in the un-
primed frame must be found to be shorter in the second set of pictures. How-
ever, such a result would contradict the principle of relativity since in that
case motion to the right would stretch a stick and motion to the left would shrink
it. Therefore, we conclude that the lengths of the two sticks must remain
unchanged. This result is consistent with the symmetrical conditions of the
experiments.
b. Contraction of lengths parallel to the direction of relative motion. Let
the two meter sticks of Figure 1-5 now be placed in relative motion parallel
to their common .r-axes and the photographs repeated as before. The first set
w: (u) (b)
Figure l-6 Length measurements parallel to the line of motion. ln (a) the photographs are
simultaneous to observer O in S; in (b) they are simultaneous to observer O'in S'.
of photographs, taken by O, would look like Figure 1-6(a), and the second,
taken by O', would look like Figure 1-6(b). In the first, O would say that the
"moving" stick was contracted. O'would explain the photographs by claiming
that the flashes fired by O were zol simultaneous. In fact, he would say that
the leading end of his meter stick was photographed before it reached the coin-
cidence position and the lagging end of his meter stick was photographed after
it had passed the position of coincidence. Thus, both observers accept the
photographs but since they do not agree that the flashes were simultaneous
they have different explanations for them. In the second experiment the roles
are reversed and now obsewer O' says that the meter stick in .S is contracted
by virtue of its "motion." We cannot argue, as we did in the previous thought
experiment, that the sticks should appear equal in length since there is n0 agree-
ment on simultaneit2 in the present case. We must conclude that the stick that
is moving with respect to the observer is contracted.
This contraction may be demonstrated analytically as follows. Suppose
observer O' has a meter stick situated so that the end points are at xl and x'2
for any time l' in his frame. By the Lorentz transformation, an observer in ,S
will obtain the result
xi,:y(xr-Bctr)
xL:y(xr-|ttr),
from which we obtain
But xi - xl : Lo, the proper length, since the stick is at rest in S '. Further-
more, if r, and xL ate measured simultaneously in S, then tcz - x1 : Z, the
length of the stick as observed in .S. Thus, for l, : tz,
The same result can be obtained by using the inverse transformation, but
the algebra requires a few more steps. To illustrate this we write,
But the time transformations must be used here since the simultaneous measure-
ment of r, and x, in frame S will not appear simultaneous in frame S'. Then,
'^',i
bz : y(ctr
- fx')
from which
t
=YQh
- frr),
c(ti - ti) : yc(tz
- tr) - fy(r, - xr).
Putting \ : tz and elineinating (tL
- t{) from the above equations,
xz - xr: y(xl *'r)
- - fryr(r, - xt),
cti:TQtr-Fro)
ct'r:y(ctz-fro)
c(ti - t) : yc(tz - tr) - fy(ro - ro).
from a point to the right of observ er Oras shown in Figure l-7, since the mirror
is moving to the right with velocity u. The time at which the beam returns is
measured by observer ,B on a clock which had previously been synchronized
with the one used by observer O. If the elapsed time in the unprimed frame
is z, then from the figure we have,
T : TTo,
as in Equation 1.8. Observer O thus concludes that the clock used by O' runs
slow. Had O' photographed the clocks at O and at.B as he passed those points,
his photos would confirm the time interval reported by O. However, O' would
claim that the clocks used by O and B were not synchronized but that the
clock used by B was fast by the amount p2z. Therefore, according to O', the
time interval in the unprimed frame should be
PROEIEU 1.5
PROBIE'H
'{
If the galaxy has a diameter of 106 light-years, what will the
diameter appear to be to a cosmic ray particle traveling at
the relative speed 0 : 0.99 ? How long will the trip take
as measured by a clock riding with the particle?
PROBLEM
'.f
What synchronization error will an observer moving at
relative speed Bc detect in two stationary clocks separated
by r distance Z along the path of the motion ?
7. RELATIVISTIC DYNAMICS
By means of the Lorentz transformation given in Equations 1.6 we can
readily obtain expressions for the velocity components which each observer
will measure in his own frame. Thus, first taking the differentials,
dx' : y(dx - pc dt)
d)' :d)
dz' -dz
dt' : nr(r,
/ \-- -l*),
we define,
ul
dx' ur- fc
dt' t Urf
I ---
c
: d)' : u'
u!, (l.e)
dt'
,(-ry)
uL: dz'
uz
dt'
,F-ry)
Note that p, the relative velocity of the frames, is in the common r-direction
in Equations 1.9. The numerators of these expressions are just the velocity
-a
approach unity as P gets small. These equations can be used to obtain the
relative velocity of one body with respect to the other when two bodies are
moving with respect to the laboratory frame. For example, suppose two
electrons are fired in opposite directions, each with speed V with respect to
the laboratory. Let thg S frame be the rest frame of the electron moving in
the negative r-direction, and the ,S' frame be the rest frame of the laboratory.
Then, in the S frame the laboratory is moving to the right with velocity pc : V;
in the ,S ' frame, uL : Z is the velocity of the other electron. Since we seek z'
we must use the inverse transformation,
ua:
uL+pc V+V
, 7v-''
r -T--"Lp r |
(1.10)
cc"
If -, -
V is large, such as 0.9c, then u,: (1.80/1.81)c, which is very close to, but
still less than, the velocity of light in free space. An observer in the laboratory
will assert that the electrons are separating at a relative velocity of 1.80c.
This is not in conflict with the theory of relativity, since neither electron is
moving at a speed greater than c in any frame of reference. If the particles
in this example were photons fired by two lasers, Equation l.l0 tells us that
relative to each photon, the other would have a velocity of
2c
Ur:T:t,
which is the result required by Einstein's second postulate.
PROBLEM 1.9
PROBLEM I.IO
PROBLEM T.IT
PROBLEM T-T2
PROBLEM T.I1
that is,
dx dt
uor:Y:
dto dt dro
(1.r1)
:
2moun (2mrur)' : 2*[r{4 u* Provided u, 4c.
Then, letting m[ : m, we may write
ffi : Tt|to, (1. 12)
which is the expression for the relativistic change of mass. That is, a body
of rest mass mo has its mass increased by the factor 7 when it moves at a
S'
lv'
I
I ------ B"
I
,L\
I
\ \
Bc <-
./
---..
(a) (b)
Figure l-8 The collision of two identical volley balls: (a) as observed in S, (b) as observed in S'.
The y-axes of S and S'are colinear at the instant of impact.
RELATIYISTIC DYNAMICS t9
or
2mcz dm
- 2muz dm - 2m2u du : 0,
where the last expression holds for dd parallel to ri. From (1.13) and (1.14)
we have
dT : c2 dm,
Jo:o
f d,T : f* ,, d*.
J*o
Then,
T: (m
- mo)c, : (y - l)mocL, ( l.l5)
which says that the kinetic energy of a body is equal to the product of its mass
change times the square of the velocity of tight. This resuli can be compared
with the classical expression for the kinetic energy by writing
T_ ,/4lfrocz
- ntocz : moc'l(l - p')-+
- 1]
Thus, if B is small enough to permit discarding all terms higher than Br,
Equation 1.15 reduces to the classical result.
Still more information can be gleaned from Equation 1.15 by writing it
in the form,
E-mc2:T+ffi0c2, (1.16)
where we now call E the total energy and mocz the rest mass energy. Equations
l.l5 and 1.16 express the equivalence of mass and energy and provide the
basis for restating the conseryation laws of classical mechinics. N.ith"r mass
20 THE SPECIAL THEORY OF RELATIVITY
or
- (pt), I (mrc2)2.
E2 (1.18)
Further,
g +mocz)r: (pt)'+ (*ot\
and
(Pt)': Tz {2Tmrc,. (1.1e)
Equations 1.18 and l.19 provide useful relationships between the momentum
of a body and its energy.
It is worth pointing out that the kinetic energy of a body, according to
Equation 1.15, will increase rapidly as the velocity increases. In the limit as
u --+ ct 7 would become infinite and hence the kinetic energy would also be in-
finite. We conclude that c must represent a limiting velocity that cannot be
achieved by any particle that has a rest mass. Conversely, we also conclude
that photons and neutrinos must have zero rest mass since they are known
to travel with speed c. Another way of stating this is to say that only those
entities for which a rest frame exists can have a rest mass. Since an entity
traveling at the speed of light has no rest frame (its speed will be c to all ob-
servers), it can have no rest mass. A zero rest mass particle canrhowever, have
a momentum and kinetic energy. From Equation 1.18 we see that the
momentum and kinetic energy are p - Elc and T : Pc - E, respectively.
PROBLEM I.'6
8. FOUR.VECTOR NOTATION
In ordinary three-space the components of a vect or i in two coordinate
systems S and S' with a common origin are related by:
A| : r4r cos (r't, rr) * A, cos (.ri, xr) * ,43 cos (x't, xt)
AL : z{., cos (xL, xt) * Arcos (ri, *r) * r43 cos (rL, r") (1.20)
AL :,4, cos (xL, xt) * A, cos (.ri, xr) * ls cos (*'", r").
- rcos 0 +) sin
tc' B
(;):(;ll ;il0)
The cni are not all independent since",::
they are direction cosines and must
satisfy the relations
: f0, for i+f
citcit, 6or:
tr, for j : k
and (r.22)
det cri - l.
22 THE SPECIAL THEORY OF RELATIYITY
\i:/ \;,1
When the index is a Greek letter, a four dimensional space is implied; a
Roman letter will represent a three dimensional vector. With this convention
AX,, AXp : (As)z, as given above.
The matrix of the Lorentz transformation is given by
6:(r,,l :',l)(3
FOUR-VECTOR NOTATION 23
ct,:y(ct_px),
which was to be shown.
PROBLEM I-20
r-1 _
00
10 'r)
(,[, 01
00
24 THE SPECIAL THEORY OF RELATIYITY
U,:(t (1.26)
ffi)
and its transformation is given by Equation 1.25 as UL: Tp,u".
c. The momenfum four-vector. The four-momentum is formed simply
by multiplying the four-velocity by the rest mass. Then,
P:(l,W) (r.27)
W:(T)
The first three components reduce to the three-momentum components and
the fourth is the total energy multiplied by ilc.
1r..
I-'l ::tuu::i .,.....,,'.. .r1:::: :" :.
PROBLEM I.2T .,.1
examPle,
Ft: I ,dVrt:T
r, ,dP,
7t: T,FN'
and
FE: ,'#rrr: y':# : y'! F'a'
Thus, the first three components of the four-force are the three-vector force
components multiplied by y'.
F': (1.28)
(,1'!'')
e. The current density four-vector. If po is the charge density in a frame
where the charges are at rest, then a current density four-vector may be defined
as the product of po and the four-velocity. That is,
Ju: PN,r:
(l!';,h:(T;) (1.2e)
fi,)
Here we have defined p : y'po, illustrating the increase in charge density
due to the change in volume element. Then the product of local charge density
and local velocity gives the ordinary three-space current density, (i*, ju,i,).
f. The electromagnetic potential four-vector. Since a charge distribution
can be at rest in only one frame, it will be observed as a current accompanied
by.u magnetic field by all other observers. Therefore, we would not expect
the components of either electric or magnetic fields to become components
of the four-vector. Using the vector and scalar potentials defined by
Ar:h) (1.30)
0 g, 8*
-go C
0 g*
-9, -!c
a
6a
(l .31)
go
-g* 0
-,_ 8,
C
3mx
8. - 6ot 8"
c'c
PROBLEM 1.22
eir: gr +
: Et
8'u
g'L:y(Etao x6)'
CIt:
'(*':#)
where "parallel" and "perpendicular" refer to the line of
Q.g2)
relative motion.
PROBLEM I-23
rG See, for example, W. K. H. Panofsky and M. Phillips, Classical Electricit2 and Magnetism' Addison-
rA-esley Publishing Co., Inc., Reading, Mass. 1955.
RELATIVISTIC COLLISIONS
h. The proper time as a four-scalar. Using the symbol (Ar)2 for an interval
in three-space, we may express the length of an interval in four-space as
Now, when two events are observed by the same clock, Lr :0 and A, : Azo,
the proper time interval. Therefore, in any inertial frame the square of the
interval is
(At)' - -cz (A"o)'.
Since this is a scalar quantity in all frames, Azo is called a four-scalar. We
can show that the proper time is the minimum time interval by rearranging
the above expression as follows,
(A?'.
(Ar)r: (Azs)z +
when (A"o)' ) 0, Azo is real and,,, ;; ur r;*rt)ure'e interual.In this case one can
always find an inertial frame traveling at a velocity less than c such that the
two events occur at the same point in space but at different times. Two events
separated by u time-like interval can never be regarded as simultaneous by
any observer. When (A"o)' { 0, Azo is imaginar2 and is called a space-like interual.
In such a case one can always find an inertial frame traveling at a velocity
less than c such that the two events occur simultaneously but at different
points in space. No observer can eliminate the space interval between the
events regardless of his state of relative motion. The imaginary proper time
interval corresponds to the transit time for a tight pulse between the two
spatial points.
At this point it is worthwhile to summ arize the principal features of the
four-vector formulation. First, all four-vectors are Lorentz covariant, that is,
they transform like Equation 1.25. Second, the "length" of a four-vector is a
scalar that is invariant in any inertial frame, as, indeed, is any scalar product
of two four-vectors. Third, the first three components of a four-vector form a
vector in three-space, but not all three-vector components are components of a
four-vector (for example, the three-velocity). Fourth, unlike a three-vector,
a four-vector can have zero ma.gnitude without all of its components being zero.
9. RELATIVISTIC COLLISIONS
In the majority of experiments involving relativistic velocities, the ob-
server is at rest in the laboratory and the events under study are taking place
THE SPECIAL THEORY OF RELATIVITY
between particles which, for the most part, are moving at high velocities with
respect to the laboratory. The physical nature of these events, however,
depends upon the amount of energy available to do work in the zero-momentum
frame of the particles. (The zero-momentum frame is the center-of-mass frame
of classical mechanics.) If two particles approach each other and exchange
momentum and energy we call the event a scattering event. If the particles
are not the same in number or in kind after the collision we say that a reaction
has occurred. In either case, it is often convenient to do the theoretical analysis
in the zero-momentum frame and then transform the results to the laboratory
frame in order to check against experiments.
a. The zero-momentum frame (C-frame). We will designate the total
energy, total kinetic energy and total momentum in the zero-momentum frame
by E*, T*,and,P*, respectively. Of course, F* - 0, by definition. Consider
two identical particles of rest mass mo, one particle having momentum F, and
kinetic energy To and the other being at rest in the laboratory (Z-frame).
If we call p* the relative velocity of the C-frame, the Lorentz transformation
may be written,
0 0 i|*y*
r0
(il:( _'i ,
0
01 0
00 y*
(il
where Pt: Po, Pz: Ps :0, and Pa : iElc : (ilt)(E" + E) : (ilt)(To *
2moc2).
PI:T*(Pr*if*Pn)
Pt:P,
Pt:P"
Pf :/*(Pn-i|*Pr).
Substituting for P, and Pnit follows that:
p*_ r*(r"-ry-o
and
E*:y*(E-f*tP").
From the invariance of the length of the momentum four-vector we have,
E*z-Ez-Plcr.
From the first of these three equations we see that the velocity of the C-frame
is given by
From the third equation we can easily obtain an expression for the kinetic
energy of each particle in the C-frame in terms of the laboratory kinetic energy
of the incident particle, f. Thus,
E*2-(8"+mocz)z-Plcz
: (E: - (*ot\
PStr) ! 2Eomoc2 +
: (mocz), + 2Eomocz * (*ot\
:2mocz(Eo * mocz)
:2moc2(7, + 2mocz). (1.35)
Since the energy is equally divided between the two particles in the C-frame,
the energy of either particle is
Tt:+E* -ffiocz.
EXAMPLE
f* :H: o.BB,
d'
cosg: lW Eo I mocz
-VIE" + 3rnoc'z' (1.37)
V tf', * mocz)z
- (2msc2)2
In the classical limit the kinetic energy of the incident particle is small compared
to its rest mass, and Eo ffi0c2. Then cos 0 : {T12, and the angle between
-
the paths of the particles is 90'. This result holds even for asymmetric scattering
in the classical limit as long as the masses are equal. As the incident energy
increases so that Eo t ffi0c2, the scattering angle gets smaller and both particles
tend to go off in the forward direction.
c. Inelastic collisions. If kinetic energy is not conserved during a
collision we call that collision inelastic, and the kinetic energy that disappears
is converted to mass or potential energy. By way of illustration consider the
following example: A particle of rest mass mo moving with velocity ptc collides
inelastically with and sticks to a stationary particle of rest mass Mo. Find the
velocity of the composite particle. The momentum in the L-frame is
P :7tmo0{,
and the total energy in the Z-frame is
E:Tfltocz*Mocz.
RELAT|VISTIC COLLISIONS 3t
Then, from Equation 1.33, if we call B* the relative velocity of the C-frame,
D*,: pc Ttmo|{z
P' : : Ttmo|r
E r*ot, a 74t, r*, * 4'
and p*c is the velocity of the composite particle in the l,-frame. All of the
kinetic energy that existed in the C-frame before the collision has been converted
to mass, since there is no kinetic energy in the C-frame after the collision.
PROBLEM 1.24
PROBLEM 1.25
Q:T'a -T*'
where the prime denotes a value after the collision. Since the change in kinetic
energy is numerically equal to the change in rest mass energy in the C-frame
we can write,
For Q - 0, the collision is elastic; for I > 0, the rest mass decreases, the
kinetic energy in the C-frame increases, and the reaction is exoergic; for
Q < 0, the rest mass increases, the kinetic energy decreases and the reaction
is endoergic. Equation 1.38 is most useful when Q is expressed in terms of
the energy of the incident particle and its angular deflection by the collision.
Let us designate the incident particle as particle a with rest mass m and kinetic
energy To in the l-frame. The target particle D has a mass M and is initially
32 THE SPECIAL THEORY OF RELATIYITY
Po:Po +F;
P'o' : PZ + PL' - 2!'P'ocos o,
Eo-l-Eo:EL+E;
To * (M + m)cz : fL + m'cz * f{, + M'cz.
Since this equation gets quite complicated for relativistic velocities it will be
derived here for the classical case instead. It has been shownlT that the correct
relativistic expression can be obtained from the classical result by substituting
for each rest mass mo the value mr I (Tcl2C), where To is the value in the
Z-frame. Then, using the classical expression for the kinetic energy, Equation
1.39 becomes:
of,
Equation 1.40 is frequently used in nuclear physics to obtain the energy released
by u nuclear reaction or the threshold energy to induce a reaction, both of
which are given by the value of Q. Note that To, T',, and 0 are all measured
in the Z-frame.
assuming that all four particles are at rest in this frame. Before the collision,
the total energy in the C-frame was
2Moc2 + T*.
* Spin will be discussed in detail in section 9 of Chapter 3 and in Chapter B.
34 THE SPECIAL THEORY OF RELATIVITY
T* :2'
From Equation 1.34, the total energy in the Z-frame must be
PROBLEM 1.26
PROBLEM l-27 ,:' ',:.' -:'' ':.'' 'i:ir "i!,' ',:r' 'l
f The quantum theory of radiation will be developed in Chapter 2. However, the concept of the
photon as a quantum of energy of amount hy will be used freely here, since it should be familiar to the
reader from an earlier course.
THE RELATIVISTIC DOPPLER SHIFT 35
e+ * e- ---> n+ | n-?
Use 139.6 MeV for the rest mass energies of the pions.
(Ans. : 139.1 MeV.)
Bc
-->
Observer
E:h2,,,tr:$
/
Sotrlce
\ 1l_
Figure l.l0 Relative motion of observer and light source.
36 THE SPECIAL THEORY OF RELATIYITY
lower than the proper frequency. For the case of the moving observer we have
the transformation equation,
PL : rPu'
(Pn)':T(Pn-iFPr),
ort
and
+:r(+ -iPT*'u)
v
^., _vo(l -Bcos0)
- '\/L p,
-
As the observer in the S' frame moves toward the source from the left,
he light emitted at the angle 0 : n and he measures the frequency,
sees
,ot/l 1
',
u
\/-r-P
B
(Approaching observer)
As he passes the source he sees light emitted at the angle 0 : nl2 and he
measures,
yo
y' : (Transverse Doppler effect)
ffir.
As he recedes to the right, the observer in S' measures light emitted at the
angle 0 : 0 and he obtains,
, : ,o{T-p
vt' (Receding observer)
ffi.
Now suppose we consider the inverse transformation in which the observer
is regarded as stationary in the S frame and the source is moving. In matrix
form this is (see Problem 1-20),
Pu : f -1P;.
The equation for the fourth component is
Pe:y\i+ipP)
OI,
*f cos0')
- \/t_
.,_r,o(1
p,
Again we have the three cases:
ph.'a)
v':
vo(L
-
\/r- p, '
where B is always positive, i is the unit wave vector, and A is the unit vector
in the direction of the velocity of the primed frame.
PROBLEM I-29
PROBT.EM
'.30
Assume that the stars of a binary pair revolve about their
center of mass with a linear velocity of 6000 km/sec. If one
of the stars emits a spectral line of wavelength 6000 A, what
will an observer measure for this line when the emitting star
is receding, approaching, and moving transverse to the line
of sight if (a) the observer is in the rest frame of the center
of mass, and (b) the observer is moving with speed 0.6c with
respect to the center of mass frame of the binary ?
(Ans.: (a) 6120 A, 5BB0 A, SSSS A; (b) 12,2+A A, 3060 A,
50e4 A.)
18This is the experiment of H. E. Ives and G. R. Stilwell, J. Opt. Soc. Am.28r2l5 (1938) and 31,
369 (1941), but with an exaggerated aromic velocity.
THE SPEC|AL THEORY OF RELATIVITY
PROBT E/r4
'-3'
In its own rest frame, a no meson decays into two photons
of equal energy and equal but opposite momenta. If a n"
has a velocity ao in the laboratory, what are the highest
and lowest photon frequencies that can be produced when
it decays ?
Idx
tan?: cdt
':c <t.
*u,
Now consider a frame S' traveling at velocity pc relative to S in the x-
direction. The line r' - 0, which corresponds to ct : xl0 (from Equation 1.6),
becomes the time axis in the primed frame. The Iine ct' - 0, which corresponds
to ct : pxris the space axis in the primed frame. Hence, the angle oc between
Figure l'12 Calibration curves for the coordinate axes. The segments OP1,OP2, OQ, and OQ,
each represent unit interval along their respective axes.
the two space axes and between the two time axes is given by
tan a. : f,
and these axes can be drawn once B is known. Now it is necessary to obtain
the calibrations for the two sets of axes. For convenience define the new vari-
ables s : ct and s' : 6f,t . Then the Lorentz transformations are:
x'-y(x-fis) + fs')
x -- y(x'
I .41)
r'-y(s-Fx) s:y(s'+ 0x') (
Figure l-14 Graphical illustration of time dilation and the Lorentz contraction. The scale used
here is for y : ;.96.
SUMMARY 4t
I
Absolute Future
Ar, is time-like
a
,/p
Present Present
Ar,, is space-like
Present
Absolute Past
This section has been merely a brief introduction to some of the basic
ideas of the geometrical approach to the solution of problems in special
relativity. To some the geometrical approach is fresh and appealing; others
prefer the analytical methods stressed in this book. Those readers who wish
to pursue the intricacies of the geometrical method are referred to the recent
book by Shadowitz.rg
SU'I,IMARY q
The two postulates of the special theory of relativity are that (1) all
inertial frames are equivalent for the description of all physical laws, and (2)
that the speed of light in vacuum is the same for all observers. Since the
Lorentz transformation follows directly from these postulates, any physical
law which retains the same mathematical form under a Lorentz transformation
automatically satisfies the principle of relativity. Such a law is said to be
Lorentz couariant. Two important kinematical effects which derive from the
theory of relativity are the contraction of lengths parallel to the relative motion
and the dilation of time. An important dynamical effect is the increase of
mass with velocity, which leads to the statement of the equivalence of mass and
energy, E : mcz. The conservation of total energy, which includes rest mass
energy, now replaces the separate conservation theorems for mass and energy
in classical physics. The concept of a four-vector is developed from the treat-
ment of time as a fourth dimension on an equal footing with the three
dimensions of Euclidean geometry. The invariance of the "length" of a
four-vector is shown to be a useful property when solving physical problems.
Applications are made to particle collisions, particle creation and annihilation,
a.td the Doppler shift. The use of the space-time diagram to solve problems
graphically is briefly introduced.
REFERENCES
p. G. Bergmann, Introduction to the Theor2 of Relatiaitl. Prentice-Hall, Engle-
wood Cliffs, N.J., 1962.
A. Einstein, The Meanirg of Relatiuit2. Princeton University Press, Princeton,
N.J., 1946.
A. P. Fiench, Special Relatiuity. W. W. Norton, New York, 1968'
J. D. Jackson, classical Electrodlnamics. John wiley and sons, New York,
1962, Chap. ll, 12.
R. B. Leighton, Principles of Modern Physics. McGraw-Hill, New York, 1959,
Chap. l.
C. MOller, The Theoryt of Relatiuit2. Oxford University Press, New York, 1952.
W. K. H. Panofsky and M. Phillips, Classical Electricit2 and Magnetism. Addison-
Wesley, Reading, Mass., 1955, Chap.14-17'
R. Resnick, Introduction to special Relatiuit\. John wiley and Sons, New York,
r968.
W. G. V. Rosser, An Introduction to the Theoryt of Relatiuity. Butterworths,
London, 1964.
J. H. Smith, Introduction to Special Relatiaity. W. A. Benjamin, New York, 1965.
f. W. Van Name, Jr., Mod.ern Ph-ysics. Prentice-Hall, Englewood Cliffs, N.J.,
1962, Chap. 3.
F. K. Richtmyer, E. H. Kennard, and J. N. Cooper, Introduction to Modern
Physics. McGraw-Hill, New York, 1969, 6th edition'
A. Shadowitz, special Relati.uiti,. w. B. Saunders, Philadelphia, 1968.
CHAPTER 2
In this chapter we will review some of the significant events which led to
the realization that electromagnetic radiation is quantized, and further, that
these quanta-called photons-display both wave and particle characteristics.
What at first seems like a conflict between contradictory natures will be
resolved by adopting the point of view that a photon is a wave packet which
can participate in either wave-like or particle-like interactions.
I. BLACKBODY RADIATION
The quantum theory of radiation had its origin in the search for an
explanation of the spectral distribution of the radiant energy emitted by u
heated body. By the term spectral distribution, we refer to the relative amount
of energy associated with each wavelength interval of the emitted radiation.
It has been known for a long time that the color of a heated object changes
to a dull red at about I100"K and that the color of the visible light emitted
shifts toward the blue end of the spectrum as the temperature rises further.
Experimental curves of the distribution of energy with wavelength for a given
equilibrium temperature show the same general characteristics, regardless of
the material of the body. Hence, it is natural to define an ideal blackbodl,
which is a perfect absorber (and emitter) of radiation. Since it reflects no
light at all, it must appear perfectly black unless itis emittinglight in the visible
region of the spectrum. If a pulse of visible light, made up of a narrow band of
frequencies, were incident upon such a blackbody, all of the light would be
absorbed and would then be reradiated in all directions at greatly reduced
intensity and with a different spectral distribution. It turns out that a black-
body which is in thermal equilibrium with its surroundings has a constant
spectral distribution of radiated energy which is characteristic of all blackbodies
maintained at that same temperature. Moreover, the spectral distribution
curve for a real object can be predicted by multiplying the blackbody curve by
the absorptivity of the real body.
A study of the radiation from an ideal blackbody can be approximated
experimentally by observing the light emerging from a small hole in an iso-
thermal enclosure, such as a hollow block of carbon. If the hole is sufficiently
43
THE BEGINNINGS OF THE GIUANTUM THEORY
L
lr
E 100
Figure 2-l Spectral distribution
t< of blackbody radiation for several
a)
d ifferent temPeratu res.
d50
24
Wavelength, in microns
small,.the energy radiated through it will have a negligible effect upon the
equilibrium state in the cavity. For such a cavity, a typical plot of the spectral
distribution of energy for several absolute temperatures is shown in Figure 2-1.
The curves shown were obtained by Lummer and Pringsheiml in 1900,
although their general characteristics were known earlier. It should be noted
that (1) the short wavelength cutoff advance s toward the origin as the tempera-
ture increases, (2) raising the temperature increases the energy of all spectral
components, and (3) the peak of the curve shifts to shorter wavelengths as the
temperatuie increases. The shift of the peak of the curve was found to obey
the following empirical relationship, commonly called Wien's displacement law,z
where the symb oI )., refers to the value of the wavelength corresponding to
the peak of the curve. A thermodynamic expression also exists which relates
the total power radiated per unit area of a blackbody to its absolute tempera-
ture. This is known as the Stefan-Boltzmann law and is expressed mathe-
matically as
E(T) - oTa, (2.2)
where o : 5.6699 x l0-8 watts m-2 deg-a. Thus the total energy radiated in
a given time by u heated object is proportional to the fourth power of its
absolute temperature. The monochromatic emissive power, E()"r Z), is the
power radiated per unit area at a given wavelength. This is related to the total
power radiated per unit area by the integral,
1 O. Lummer and E. Pringsheim, Verhandlungen der Deutschen ph_rsikalischen Gesellschaft 1r 23, 215
(1899); 2, 163 (1900).
2 W.
Wien, Ann. Physik 58, 662 (1896).
THE RAYLEIGH.JEANS THEORY 45
EQ',7):# (2.3)
The quantities a and b are not derived but are simply curve-fitting parameters.
Although Wien's radiation law is consistent with both the Stefan-Boltzmann
law and the displacement law, it is quite unsatisfactory as a theory since it is
not derivable from a physical model. That is, there was no attempt to relate
the emitted radiation to physical processes taking place within the radiating
body.
PROBLEM 2.2
H:T+ ,r-PT
' - 2*- tkx',
ithas two such variables , p* and x (called degrees of freedom), and hence the
average total energy is simply fr 7. All that is necessary, then, to obtain the
spectral energy density in the cavity is to find n(r), the number of oscillators
per unit volume at each frequency a, and to multiply this number by kT. This
number is known as Jeans' number and is calculated in Appendix A of this
chapter. Its value is*
/\ : Bnyz /4\ : Bn
n\v) or n\/.) (2.+)
,"
^n
We then obtain the important Rayleigh-Jeans law,3 namely,
where I(1, T) is the energy per unit volume at wavelength 2 at the equilibrium
temperature ZoK. The relationship between I()", Z) and E(1,7) is given in
Problem 2-5.
A difficulty with Equation 2.5 immediately appears when we consider very
small wavelengths. Although the Rayleigh-Jeans law describes the experi-
mental curve quite well in the long wavelength region, it diverges (in the
mathematical sense of having no limit) as the wavelengths approach zero.
This failure was such a crushing blow to classical physics that it is historically
referred to as "the ultraviolet catastrophe." We can gain some appreciation
of the difficulty physicists faced at that time if we realize that the Rayleigh-
Jeans theory utilized the important equipartition theorem of classical physics,
and further, that it had no adjustable parameters. Its failure implied that
there was something fundamentally wrong with either the equipartition
theorem or the theory of electromagnetic radiation.
* Conversion between the parameters r., and ,i is easily done if one remembers lhat n(a) dv :
n(i.) d2.. -I\en
ldll c
n(u) : nQ,) ' : -;
lal "(i)
3 Lord Rayleigh, Phil. Mag.49, 539 (1900) and J. H. Jeans, Phil. Mag.l0' 9l (1905).
PLANCK'S OUANTUM THEORY OF RADIATION 47
(2) An oscillator cannot have an arbitrary energy but must occupy one
of a discrete set of energy states given by
en : nh'u'
where z is an integer or zero. It was assumed that the ground state corre-
sponded to the zero energy state. The value Planck gave for the constant of
proportionality, lt, was 6.55 x 10-27 erg-sec. He obtained this value by
fitting curves such as those shown in Figure 2-1 with his radiation law (which
will be derived below [see Equation 2.8]). However, the best values for i is
now believed to be
h : 6.626196 x 10-27 erg-sec.
Planck's constant is a universal constant which plays an important role in all
quantum phenomena.
The previous picture of a continuum of oscillator states is now replaced
by u discrete set of "quantized" states. Furthermore, the amount of energy
emitted or absorbed is also quantized, since each quantum must correspond to
the energy difference between two states of a given oscillator. Each quantum
of electromagnetic energy is called a photon. The absorption of a photon of
frequency r will raise the energy of an oscillator of frequency ? by an amount
given by ha; it will have no effect on an oscillator of frequency v' t' a. Emission
of a photon occurs when the oscillator energy drops to the next lower energy;
the frequency of the emitted light will correspond to the oscillator frequency.
In what state is an oscillator most likely to be found ? If nothing excites
it, it is most likely to be found in its lowest energy state or ground state. Hence,
at absolute zero one would expect to find all oscillators in the zero energy state
according to the above model. This would be true in classical mechanics and
can be assumed here without affecting our answer. But we will see later that
quantum mechanics predicts a so-called "zero-point motion" at absolute zero
instead of the complete cessation of all vibration. At higher temperatures
thermal agitation excites some oscillators to higher states so that some sort of
distribution of oscillators over all possible states will exist for each temperature.
The higher energy states are thus less likely to be populated, and as the energy
increases indefinitely the number of such oscillators becomes vanishingly small.
It can now be seen that this is going to eliminate the problem of the
ultraviolet catastrophe, because the latter arose as a result of the assumption
that oscillators of all bnergies were excited with equal probability; hence,
all energies contributed equally to the emitted radiation. From the quantum
hypothesis we see that a high energy oscillator can contribute more radiated
energy than a low energy oscillator only if it is excited. But its probability of
being excited is so low that the energy which appears at the high-frequency
end of the spectrum is much less than the classical Rayleigh-Jeans theory
predicted.
Mathematically, this can be shown by u calculation of the average energy
per oscillator, €, based on Planck's quantum hypothesis.
@
I N(n)e"
a : *- , where the sums are taken over the energy levels, €0: €1r €2 ' '
.
I
n:0
/v(n)
-
2
n:o
lr:ot-"'nlkr nhu
0* hue-hvlkr a 27rt-znvlter + 3hye-BhvlkT a
I a s-nnt*, + t-2hvlteT * fsnvl*T *
I Not-"onlo'
n:0
: hyx (r
+ + 4xB +
2x + 3x2
where * - t-hvltcT
+r 1-' + -gr -L
I
),
, ' (l - x)-z: hvx Itv
: nYi
(1 -
"1-r
T=x: w_r (2.7)
PROBLEM 2.3
PROBLEM 2-1
Using Jeans' number, Equation 2.4, and the average energy per oscillator
obtained in Equation 2.7,we write the Planck radiation law as
I(u, T) : n(r). €-
- ry (ehvttcr ,)-J
Planck's law in this form is in units of energy per unit volume per interval of
wavelength (or frequency). To express it in terms of radiated power per unit
area per interval of wavelength (or frequency), one need merely multiply by
the factor cl4.
PROBLEM 2.5
PROBLEM 2.6
Cv- : 3R,
*:
PROBLEM 2.8
P:tu;
where P is the pressure and u the energy density of isotropic
radiation. Assume that the ideal gas law holds in the cavity.
PROBLEM 2-9
lArr*Brrl(u)l .Nr,
and the transition rate for absorption is
BrrIQ') . ifr.
Now tlre equilibrium populations, frr and lr{r, are given by the Boltzmann
distribution function (Equation 2.6) as:
lrt
_::: Er\
N2
exp (- kr) 'p e): exp (#),
where 7 is the temperature corresponding to the equilibrium state. Then,
Bn : Brr,
52 THE BEGINNINGS OF THE OUANTUM THEORY
and
(2.10)
It is evident that this is the Planck radiation law (Equation 2.8) and that the
ratio of the Einstein coefficients is
A4 Bthvg
'
Bu: cs
(metastable)
Signal
E, (metastable)
Pump
hv Signal
Fast Decay
PROBLEM 2.' I
PROBLEIA 2.12
Much of the significance of laser (or maser) action lies in the fact that it
can provide an intense source of coherenl radiation. In fact, an incoherent
source acting as a pump can be converted to a coherent source at the pump
frequency if an appropriate pair of energy levels can be found in a physicat
system.
s4 THE BEGINNINGS OF THE QUANTUM THEORY
Although the term "laser" can be used for all such devices employing
photons, it is common practice to regard a device operating in the microwave
region as a maser and one operating at optical frequencies as a laser. However,
this distinction cannot be held too rigidly, since it is possible to pump at
microwave frequencies and to amplify at optical frequencies, or vice versa.
Since there are regions of the electromagnetiCspectrurn-wherein it is extremely
difficult either to obtain a monochromatic source or to amplify an existing
signal, it would be attractive to devise lasers which could operate at these
frequencies but which could be pumped at frequencies that are readily attain-
able (such as "white" light). This goal provides impetus for a great deal of
research on the energy levels of atomic, molecular and many-body systems.s
Light incident upon a metal surface can, under some conditions, eject
electrons from the surface. These electrons are called photoelectrons, not that
they differ from other electrons, but merely to identify their source. The
following facts must be explained by u satisfactory theory of the photoelectric
8 B. Lengyel, Introduction to Laser Physics. John Wiley and Sons, Inc. New York, 1966. See also,
Lasers and Light, Readings.from Scientifc American. W. H. Freeman and Co., San Francisco, 1969.
0 R. A. Millikan, Phil. Mag.19, 209 (1910) ; Phys. Reu. 32,349 (l9l l).
toJ.J. Thomson, Phil. Mag. tAt293 (1897).
56 THE BEGINNINGS OF THE OUANTUM THEORY
Light quanta
Figure 2-3 Apparatus for determining the photocurrent (switch in position. l) and the maximum
kinetic energy of the electrons produced by photoemission (switch in position 2).
t
I
b0
li
0)
0.)
!
!i
o
P
Saturation Current x
pi
2
0
Frequency --->
I u rlate voltage
Plate --
--+
-w
T.o*
Figure 2-4 Plot of photocurrent Figure 2-5 Maximum kinetic
for a given photon fre-
vs. plate voltage energy of photoelectrons vs. photon
quency. Tmax is determined from the frequency.
value of the reversed potential that cuts
off the photocurrent.
Figure 2-4. Figure 2-5 shows a plot of 7}a* ar. ? which permits the experi-
mental determination of both h and W in Einstein's equation. The first
experimental verification of this equation was made by Millikan in 1916.13
PROBLEM 2.13
PROBLEM 2.14
PROBLEI{I 2.15
PROBLEM 2.16
lo:p'cos0+pcos{
and
0:!'sin0-psin{,
lf M. C. Teich, M. Schroeer, and G. 6l I
15
J. J. Wolga, Ph1s. Reu. Letters 13, (1964).
R. L. Smith,
Ph1s. Reu.l'28r2225 (1962).
r'A. H. Compton, Ph2s. Reu.22r409 (1923).
58 THE BEGINNINGS OF THE AUANTUM THEORY
(a) (b)
from which we can eliminate d bV isolating the term in { on the left side of each
equation, squaring, and adding. Then,
Eo-E': T,
or
T
Po-P': c
(2. 13)
Squaring this,
PZ + P'" - 2Po!' : (Tr (2.r+)
h
I,
p2-+:2!op'(l
c -coso). (2.15)
Since the velocity of the particle after the collision could well be relativistic,
we must use the relativistic momentum, Equation 1.19,
b,
lo
:7 (T' + 2moczT).
Then
where Equation 2.13 has been used to obtain the last expression. Using
Equation 2.15 and Equation 2.16,
moc(Po
- p') : PoP' (1 - cos 0)
11 lrt cos 0)
Y - Po: n?oc' -
of,
).'-)": Ll 'Tltoc - cos 0). (2.r7)
Using the mass of the electron for ffi,, the quantity hf moc : 0.0243
angstrom and is now called the Compton wavelength. Notice that for 90"
scattering, Equation 2.17 predicts a new x-ray line just 0.0243 angstrom
longer than the primary line. Figure 2-6 shows this line as the large peak
measured by Compton. The presence of the primary peak at 90o might at
first seem surprising. Compton explained this by considering the scattering of
a photon from the atom as a whole. Thus, if one uses the mass of a whole
carbon atom instead of the electronic mass in Equation 2.17, the wavelength
shift will be reduced by a factor of 20,000 and amounts to roughly one millionth
of an angstrom. Therefore, the line scattered by an atom is for all practical
purposes unshifted.
Compton's work provided rather convincing evidence that a photon can
undergo particle-like collisions with both atoms and unbound electrons. Later
studies of the recoil electrons and their energies added further confirmation to
the predictions of the theory.u
PROBLEM 2.17
PROBLEM 2.'9
PROBLEM 2.20
PROBLEM 2.21
of the packet as a whole are transferred to another particle. In the latter case
the details of the phases of the constituent waves are unimportant. That both
the wave and particle aspects of photons are required for the description of
light is known historically as Bohr's principle of complementarit2, That is, the
wave and particle aspects complement each other. In order to acquire a better
understanding of the nature of a wave packet, let us first consider the super-
position of two plane waves.
In classical physics we frequently represent a plane wave traveling in
the positive x-direction by either the real or the imaginary part of one of the
following equivalent expressions
Here z4 is the wave amplitude, k :2nl), is the propagation constant, cr.r :2rv
is the angular frequency, and ?tr is the period of the harmonic oscillation. The
velocity of propagation of the wave front is the phase velocity, u : alk - Lu.
Suppose that two waves having slightly different frequencies and wave-
lengths are propagating together through a medium. For simplicity let us take
their amplitudes and initial phases to be equal. Then we may represent these
two waves by the expressions,
we may express the displacement resulting from the superposition of these two
waves by,
(*-)
)
Note thatk':k+dkf2,-,k, and 111':u1-dal2- @t since dk<k and
da K a. Then the'factor containing the sine is essentially the same function
as Y, and may be thought of as a "carrier wave" of phase velocity u : a'lk' -
,lk. The cosine factor has the effect of modulating the amplitude of the
carrier wave, and the modulation envelope moves at the so-called group
velocity given by , : daldk. The transmission of energy (that is, a signal)
must o...rr at the group velocity and not with the phase velocity. The reason
for this will be clearer after a discussion of the Fourier integral theorem in
Chapter 4, but it may be noted here that the transmission of a signal alwa_ys
involves modulation of one kind or another. An infinitely long wave train of a
single frequency can never be used to transmit information at its phase velocity.
Signaling always involves chopping (keying), amplitude modulation, frequency
52 THE BEGINNINGS OF THE OUANTUM THEORY
z"o.'g(*
)
Figure 2-8 One half-wavelength of the modulation envelope formed from the superposition
of two plane waves of nearly equal frequencies and equal amplitude. This pattern is repeated con-
tinuously throughout space.
@ - u(k) 'k.
du
zero, then the packet is so broad (its wave train is so long) that its emission
time is extremely large. Light radiated by atoms during electronic transitions
has a finite wave train which corresponds roughly to the lifetime of the state.
In the time domain these bursts, which are of the order of 10-8 seconds, result
in a spread of frequencies rather than in a single emitted frequency, and
produce what is called the natural linewidth of the spectral line. The longer
the lifetime, the fewer extraneous frequencies in the spectral line.
Denoting the uncertainty in the frequency (the frequency spread) by Ar,r
and the uncertainty in emission time by At, we can write,
^
A@ ,\, 2n r
N.*
.r
L'k* '--'
t*.*
Making use of the Planck expression for the energy of the photon and the
relativistic value for the photon momentum, the above relations suggest that
LE . Lt ,-,h
and (2.1e)
Lp*' L,x ,-, lt.
SU/I4IT4ARY
10 See Chapter II of W. Heisenberg, The Ph2sical Principles of the Quantum Theorlt. Dover Publi-
cations, Inc., New York.
t The photon also has unit spin, which is ignored in the present discussion.
CHAPTER 2 APPENDICES A to C
02u | 02u
0x2 7 at,'
where z is the displacement and c is the velocity of the wave. Since the variables
are independent, we assume a product solution of the form u : T(t) .
6@.
Substituting this into the wave equation we obtain the separated .q,ruiior,,' '
_ i(t)
'"ffi:fi)-
^16"@) -@2' (A1)
where -@2 is simply the separation constant. Solutions of the two equations
(Al) are
T:Csinicut+Dcosrcot
and (A2)
6:ysinkxfdcos,tr,
where k : alc :2r11.
An infinite medium can have any values of ro and ,tr, but a finite medium
such as a string fastened at each end has only certain allowed values. These
allowed values are called eigenualues, and the modes are known as standing
waves. If we impose the constraint thatu:0 for x:0 and x: L, then wi
find that d :0 and k : nnf L, where n is an integer. The general solution to
(Al ) becomes
u(x, t) : 7 sin
T t, sin a-l/ + D cos a-rt),
u(x, t) : Fnsin (k,x) cos (a;,t) + G*sin (fr,r) sin (a,r,t). (A3)
Although we will leave (A3) in its present form, one can show that it can be
written as the superposition of waves traveling to the right and to the left.
Also, the most general solution is one including a summation over alt possible
a)n. If we consider one frequency, orn, we note that
- nTrC
@n:knc:i:2run,
65
55 APPENDICES A TO C
of,
2L
fl--Tn'
C
If the frequency modes are closely spaced we can express the density of modes as
and constrain u(x,_y, z,t) to be zero for.r :0, L; I :0, L; z:0, L. Since
all of the variables are independent, the solution is a product function, the
first term of which is
Each set of integers (n*, flu, n") determines a point in lattice space and each such
point occupies unit volume of the lattice. (Each unit cube contains eight
points, but each lattice point is contained in eight such cubes.) Since there is a
one-to-one correspondence between the volume of the space, the number of
lattice points contained in that volume, and the number of allowed modes of
a given frequency, we can count modes by merely calculating the volume in
the lattice space. Therefore, the number of modes of frequency pn in the first
octant of a sphere of radius z is
nr-1.!n^,
iV:B'Trr":6
n
(?",r (,{6)
and the number of modes per unit volume in the same frequency interval is
dN 4no,
L":-u'"da'
In the case of transverse waves there are actually two times this number of
modes, one for each sense of polari zation. Finally, we obtain
Jeans' number,
which is
Bn
n(u) : Y2. (A7)
-c"
Stppose that a given volume is divided into z cells and that n distinguish-
able but otherwise identical particles are to be distributed among the z cells.
If each particle has the same a priori probability of occupying any unit volume
of the space, then intuition tells us that in the final distribution the number of
particles occupying each cell should be proportional to the volume of the cell.
Thus, a uniform distribution is the most probable distribution, and if the cell
volumes are the same each cell would be expected to containnf zparticles. We
will prove this conclusion formally before deriving the Max*ill-Bolt zmann
distribution function.
Let 9e be the fractional volume and no the number of particles occupying
the ith cell such that
2 go:l
i:l
and 2 ,o:
i:L
n.
Then the probability of a given distribution with n, particles in gt, ltz ir gr,
and so on, is
n!
W: nrlnrl ...n,!
(g')"(gr)"'''' (g")'" (81)
The first factor is the expression from combinatorial algebra for the number
of configurations corresponding to a given distribution (that is, the number
of ways it can be obtained) when neither the order in which particles enter the
cells nor the designation of specific particles is important- The remaining
factors give the a priori probability of obtaining a single configuration. Foi
example, consider the case of distributing 9 particles over 3 equal cells such
that n, : 2, ltz :3, and fts : 4. The 2 particles in the first cell can be put
there in 2!equivalent ways, ft2:3 can be obtained in 3!ways, and nr:g
can be obtained in 4 ! ways. Thus, the number of different configurations
corresponding to the distribution (2, 3, +) is given by
9!
ztgt+t
68 APPENDICES A TO C
But the a priori probability of putting any particle in a cell is l; for two
particles in the same cell it it (*)'; and so on for no particles. Therefore, the
probability of the distribution is
w:#^(l),(+),(+-)n,
which agrees with (Bl). Before maximizing (Bl) it is con'u'enient to take its
logarithm as follows:
F :f * 1'4, * 116r:0.
The conditions for the extremum are given by the five equations
AFAFAF^AF^AF
l:0. l:0. -- -- --
il-'' o7 -o
or-"' -o
aLr.-"' -n
olr-"'
wlrich permit us to obtain the five parameters xu)b zo, 1r, and )'r. The method
of Lagrange's multipliers provides a direct and symmetric approach to all such
problems involving large numbers of variables subject to any number_ of
ionstraints. To apply the method to the present problem we note that f :
ln (W), and since there is only one constraint we need only one multiplier, 2.
The constraint is
6:n flr :0'
3
dnn'tt" w)l+ ^y-0,
oftt
APPENDICESATOC 69
and we must solve the set of
z equations obtained by letting i run from I to z.
From the equation obtained by differentiating with t.rp.Jt to n, we obtain:
h/e\
\ntl
:).+I
or
3r : flro^*' (83)
Similar expressions are obtained from aII z equations, which, when added,
yield
(g' + 3z * ''' * g,) : (ry * nz + ''' + ,,)t^,''
I -r*'
-:u
n
Then (B3) becomes
ni
8t:;t
which says that the number of particles per cell is proportional to the size of the
cell. That is, the most probable distribution is the uniform distribution.
Now we will consider the same problem with an additional constraint
on the energy of the system. The cells may now be regarded as a discrete set
of energy states, er, Such that the total energy is
e : lnoe..
The i,D Lagrange equation is
,.06,, ^ a6,
*,rt^ (w)l * h6 + ^,#,: 0,
ttz:grg-(|ftt),e-xzez
Adding,
ft : s-()'r*t'(grt-^r', * gze-tzez * * g"e-xz€'1
_ ,-(),iL) 7,
Then,
,.: 9tft ,-i"'. (85)
rLi-
Z"
fli : floe-€tllcT,
Since the cells are infinitesimal, we can replace the summation over g, by
an integration over dr. That is,
r-
2 so - J d, : III
du, dun dt," : !* +nr' dr,
where u is the radius of a sph;. ," velocity'space. The total number of
molecules may be expressed using Equation B5 as
,: Zfti
-ii
: 2 Agor-^z'i : A l, go..O ( - ry)
-r,.*o
n '^Jo- ---r\(- ry\
: 4nAf 2 I
or. (cl)
E - 2'n'o
E:2,A*t*aaexp ?n n. (c2)
APPENDICESATOC 7I
rr : Io-
rr-',' or :
*.
r,:l**rr-or'
' Jo dx: -dlo
da -I 4Va3
17
fs :
J*
r"r-",' dx :- d*4!]
: *,
Ir : j* rnr-,* dx :
# :3r/;
rs:[o*xue-'r'dx:#:*
r,n: (-1)"
f)t,
Izn+t: (- ,)" rr.
f*
Then,
N:AffiT (c3)
and
E:*J(#f : nr' 3n
(c4)
Using the result that the energy of an ide al gas is ,!nkT, we have
gnkr'
or #:
1r: #
A:'(#J
72 APPENDICES A TO C
V p V Vrnr.
From (Ct) we then find that the number of molecules having speeds between
aandu*duis:
fto:+"r(Huzexp ?m, (cs)
PROBLEM 2.23
SUGGESTED
REFERENCES
A. Beiser, Perspectiues of Modern Ph2sics. McGraw-Hill Book Co., Inc., New
York, 1969.
C. H. Blanchard, C. R. Burnett, R. G. Stoner, and R. L. l\reber, Introduction to
Modern Ph2sics, 2nd ed. Prentice-Hall, Inc., Englewood Cliffs, New Jersey,
1969.
M. Born, Atomic Ph1,,sics, 6th ed. Hafner Publishing Co., New York, 1959.
R. M. Eisberg, Fundamentals of Modern Physics. John Wiley and Sons, Inc.,
New York, 1961.
Richard P. Feynman, Robert B. Leighton, and Matthew Sands, The Feltnman
Lectures on Physics. Addison-Wesley Publishing Co., Inc., Reading, Mass.,
1963.
R. B. Leighton, Principles of Modern Physics. McGraw-Hill Book Co., fnc., New
York, 1959.
F. K. Richtmyer, E. H. Kennard, and J. N. Cooper, Introduction to Modern
Ph2sics, 6th ed. McGraw-Hill Book Co., New York, 1.969.
Henry Semat, Introduction to Atomic and Nuclear Ph2sics, 4th ed. Holt, Rinehart
and Winston, New York, 1969.
Paul A. Tipler, Foundations of Modern Pfutsics. Worth Publishers, Inc., New
York, 1969.
F. W. Van Name, Jr., Modern Ph2sics, 2nd ed. Prentice-Hall, Inc., Englewood
Cliffs, N.J., 1962.
Hugh D. Young, Fundamentals of Optics and Modern Phltsics. McGraw-Hill Book
Co., New York, 1968.
CHAPTER 3
This chapter addresses the question of the structure of the atom, its
relationship to the more familiar elementary particles, the origin of electro-
magnetic spectra, and the source of the atomic magnetic moment. The Bohr
model, when combined with the Planck theory, leads to a simple but unified
quantum theory of matter and radiation.
The assumption that electrons are constituents of all atoms was a reasonable
inference from such experiments as the photoelectric effect in metals, the
ionization of gases in discharge tubes, x-ray bombardment, and so forth. A
difficulty arose, however, in attempting to design a stable atom which could
contain both electrons and the positive charges necessary to make the atom
electrically neutral. It was supposed that these positive charges were many
times heavier than electrons, and this was confirmed by the elm measurements
of positive ions by Thomson.l The hydrogen ion turned out to be 1836 times
as heavy as the electron if one assumed the same magnitude of charge for each.
Thomson proposed an atomic model in which electrons were embedded in a
massive matrix of positive charge filling a volume of roughly one atomic
diameter (which was known to be about I angstrom). This model has been
called the "plum-pudding" model and the "jellium" model, the former term
referring to the role of the electrons analogous to the raisins in a pudding, the
latter deriving from the fact that the electrons were permitted to vibrate in
order to account for the radiation spectrum discussed in the previous chapter.
After the identification of alpha particles by Rutherford and his co-
workers,z'3 a series of experiments was performed by Geiger and Marsdena-G in
which alpha particles of known energies were scattered from gold foils. The
deflection of an alpha particle caused by u Thomson atom can be estimated
by the following argument. Neglecting the binding energy of the electron to
the atom, the maximum momentum that can be transferred to an electron of
mass mo by an alpha particle of mass M and velocity u is 2mou. Then the
maximum angle of deflection of the alpha particle is approximately
since the momentum given to the electron represents the change in momentum
of the alpha particle. Similarly, one can estimate the deflection due to the
continuous charge within the atom by calculating the total impulse given the
alpha particle by the Coulomb force during its transit through the positive
jellium. Thus,
2ft' (lF)*,*
2R fz'2e dq 4Zez
muz Jo Rz MRuz'
-t
where R is the atomic radius (-10-t cm) and u'-'2 x lOe cm/sec. Substitut-
ing numerical values results in a A{,,ax - 10-a or 10-5 radian. Since the
probability of scattering from more than one electron within a single atom is
quite small, we will assume that A{-ax - l0-a radian per atom. Multiple
scattering from a layer of such atoms is completely random and obeys the laws
However, for angles much larger than Io, Equation 3.2 predicts an infinitesimal
probability. In particular, the fraction of particles scattered through 90"
would fs The earlier work of Geiger and Marsden8 showed that about
...,,s-s021
one alpha particle in 10a was scattered 90o or more. Though this might seem
like a small number, it is many orders of magnitude greater than any prediction
based on the Thomson model of the atom.
7 See footnote 5.
8 See footnote 4.
CLASSICAL SCATTERING CROSS SECTIONS 75
There are two principal methods of studying the forces between elementary
particles, namely, the study of the bound states of the particle system and a
statistical study of the scattering of a beam of particles of one type by u target
containing the other particles. The study of bound states is more limited
because we must be content with the bound systems provided by nature and
we must adapt our detectors to the energies required by those states. In the
case of the neutron-proton system (the deuteron) there is only one bound
state, so that there is not a great deal that one can learn. However, in a
scattering experiment a wide range of particle energies and scattering angles
may be studied, and the resolving power of the scattering apparatus is poten-
tially very great, because the de Broglie wavelength * of high-energy particles can
become quite small in comparison with the range of the forces being studied.
Furthermore, such quantum mechanical effects as spin-dependent forces can
be studied in scattering experiments by polarizing the particles of the incident
beam or target, or both.
In the present section we will discuss only classical scattering and will
consider the situation illustrated in Figure 3-1. The incident beam consists of
monoenergetic particles such that I particles pass through unit area normal to
Bettur
<lf
intensitY \ -
I -
-
Detector
Figure 3-l Schematic diagram of a scattering experiment.
the beam per second. That is, the beam intensity is .I particles per unit area per
second. Consider a ray of the beam that would miss the scattering center O
by the distance s if it were undeflected. We call s the impact parameter. Since
there is cylindrical symmetry about the scattering axis, we will call { the
azimuthal angle about that axis, measured from some arbitrary reference.
Then for a detector located at (0, il which detects dI\f (0, il particles per
second, we define the scattering cross section as
dN(o' 6)
d.o : (3.3)
I
The scattering cross section must have the dimensions of area since the product
I do gives the number of particles per second arriving at the detector.
Because of the cylindrical symmetry, the total number of particles
scattered into the conical wedge bounded by 0 and 0 + d0 is equal to the
number of particles incident on the washer of area2ns ds. That is,
dN(0) :2nls ds
oft
dI'{(O, 6) : Is ds d$, (3.4)
do:sdsd$ (3.5)
do
:-: sdsd$ sds (3.6)
da -sin 0d0d6 -sin 0d0'
For a specific interaction between particles one first obtains the relationship
between s and 0;
then dsld0 can be found and the differential cross section can
be calculated immediately. Experimentally, the number of particles counted
per second per unit solid angle at the detector is
dN(0, 6) _ ,ao
T -'ddr'
do I dN(O, +)
_: (3.7)
dolda
A comparison of theory and experiment can be made by noting the results
obtained for Equations 3.6 and 3.7.
CLASSICAL SCATTERING CROSS SECTIONS
o: In#on. (3.8)
PROBLEM 3.1
Tangent plane
Then,
r(o), : "+.
,,2
"z
(3.e)
"z!' : m(i
"o* - rdr). (3.10)
12 -
Letting r : llu,
0: SUottzt
drdu,
:h;;0: du
f -'rofr, (3.11)
and
i: -su, *ry, o
: -szfiu, #.
Substituting these expressions into Equation 3.10,
Define D, the collision diameter, as the closest distance of approach in the case
of a head-on collision. That is, D is the distance from the target at which all
of the kinetic energy of the projectile is converted to electrostatic potential
energy. Then,
2zZez
D_@,
and the equation of the trajectory becomes
d2u D
dfrr+u:-*' ' (3'12)
(g :L: B.
\dole:o r
Therefore our solution becomes
1D
u::sin0 +W(cos0-1). (3.13)
CLASSICAL SCATTERING CROSS SECTIONS 79
,:2/L:g'ot D o D o
(3'14)
2\ ,iro-/:Ttanr:Zcotr,
where o plays the role of 0 in Equation 3.3 and those which follow.
This provides the relationship which is needed to calculate the differential
scattering cross section in the center of mass system from Equation 3.6. Since
dsDO cscz
de: - ,,
then
do D cot O.. D O : D2/ .O\-1
(sin Q)-'4 csc2
de: , 2- , * \sina 7/ . (3.15)
This is the now famous Rutherford result. To find the fraction of particles
scattered at a given angle @ we use Equation 3.7 and write
dN@, D2 I
+) :t6
--f
-@do
stn4
U
per scattering center. Then for C scattering centers per unit area,
4{gD
I
:o:
16 +.2zsin
,O
@da
rrDzC sin (D dO
(3.16)
sln- sirrn
t f,
For gold, the constant C : plr{otlA:5.92t x 1022 atoms/cm2, where p is the
density in g/cm3, I is the thickness of the film in cm, A is the atomic weight, and
/[o ir Avogadro's number. For alpha particles having energies of the order of 5
MeV, D is about 2 x 10-11 cm. Thus the fraction of particles scatterdd into
the cone bounded by ,D and (D + dO, where
O : 90",
1S
ry91 :++dQ,_-,to-.dQ,
for a foil one micron thick.
Impact parameter
Scattering center
(a)
v(.
Figure 3-5 Scattering of mass m. from mass m2 as viewed in the laboratory frame (a) and in
t,he C-frame (b).
system with respect to the L-frame is $(m, + mz)a?, the reader should show that
for elastic scattering this energy is equal to T" - Tc. We then have the useful
relationship,
TcP (3.17)
mt
rL mr
for elastic scattering. Note that in the C-system each mass is scattered at
the same angle, 0r; thus, conservation of kinetic energy requires conservation
of velocities and each mass has the same kinetic energy after the collision that
it had before the collision. It is convenient to draw the following vector
triangle:
(rr.
- ar) sin 0, : u'tsin 0a
ut-4, f
cos o,
-+cosoc
llt2
r*cosd"
Using the fact that the total cross section is the same in each frame, we
write,
g, doLd+L:IIffi)",t' oc doc d+c,
ilffi),sin
,"1(fr)""in g1 d0L : r"l(f")"sin 0s d0g.
Then,
(do\ . sin 0" .d0"
(#)" _
- \aal" sin g, dot (3.1e)
From the second postulate note that we now have angular momentum (as
well as charge and energy) quantized in atomic systems. The third postulate
rejects the troublesome claim that an accelerated charge must radiate in
atomic systems, in spite of its validity in the macroscopic world. The fourth
postulate provides the link with Planck's theory of radiation, since the frequency
of the photon emitted or absorbed is given by the energy difference of the two
states divided by lt.
In order to appreciate the implications of the Bohr theory, let us consider
a one-electron atom of nuclear mass M, nuclear charge Ze, electronic mass m0
and electronic charge e. We will use the reduced mass pc given by
p: MmoM (3.23)
+mo'
11N. Bohr, Phil. Mag.26, I (1913).
84 THE CONCEPT OF THE NUCLEAR ATOM
in order to neglect any motion of the nucleus. Then, from the first postulate,
puz:ry, el+)
rr"
and from the second,
p,aT : nh. (3.25)
By combining Equations 3.24 and 3.25 we obtain the expression for the radius
of the nth orbit,
n2h2 nzao
Tn: (3.26)
@F-';.
The quantity,
h2
" r-0.53A,
ao: ffiooz
is the radius of the first orbit of hydrogen calculated for a fixed nucleus, that
is, for F : ltto. It is often simply called tlte Bohr radius. Note that Equation
3.26 specifies that the radii of the allowed orbits are proportional to the squares
of the integers.
Using Equation 3.24 we may write the kinetic energy of the electron as
'' :Z-!'
T: !uu' 2r
Since the potential energy of the electron in the Coulomb field of the nucleus is
,, -
-zt'r ,
where -
wo : To=to
^, I3.6 ev
2h2
is the magnitude of the ground state energy of hydrogen for a fixed nucleus.
Equation 3.27 indicates that the quantization of energy has arisen as a result
of assuming that the angular momentum is quantized. The frequency of a
photon emitted by u transition from the nth level to the kth level, where n ) k,
is given by
hy : .(:) : hcn - En - Ek : try #, - *)
Expressed in wave numbers, this becomes
- 2n2 p,Zzea
" hsc $-:,):R)(h-*), (3.28)
BOHR'S THEORY OF ATOMIC SPECTRA
where R7 is the Rydberg constant calculated from the reduced mass. For
hydrogen R) : 10968l cm-l, whereas the experimental value from optical
spectroscopy is R) : 109677.576 + 0.012 cm-r. The Rydberg constant
calculated for a fixed nucleus (that is, for p, : mo) is
The excellent agreement between the spectroscopic l,alues for the Rydberg
constant, the ionization potential, and the emission ipectrum of hydrogln wa"s
Energy
(eV)
Paschen Brackett
series series
Lyman -r3.53
series
4a,, l6q,
Radius, r
Figure 3-5 Bohr radii and allowed energies of hydrogen. The first few lines of several of the
spectral series of hydrogen are shown at the left.
86 THE CONCEPT OF THE NUCLEAR ATOM
a great triumph for the Bohr theory. However, it was immediately apparent
that the Bol-rr theory was good only for a one-electron atom, since it failed
badly in the cases of helium and lithium, the next simplest neutral atoms. But
for thd spectra ol singly ionized helium and doubly ionized lithium the Bohr
theory is again quite good. We call such atoms "hydrogen-like" or "hydro-
genic" since they are one-electron atoms but their nuclear masses and charges
differ from that of l-rydrogen. The spectra of the alkali metals can also be
calculated as if their single valence electron moves in an effective field of
reduced Z due to the noble gas core of electrons. Although the gross fcatures
of these spectra are quite similar to that of hydrogen, there are some serious
discrepancies due to the non-Coulomb nature of the electrostatic field and due
to spin-orbit coupling. *
Interesting applications of the Bohr theory can be made to the sliort lived
mesic atoms and to positronium. Mesic atoms are ordinary atoms which have
captured an orbital negative mu meson in lieu of one of the orbital electrons.
Alihough tl-rese atoms have short lifetimes, t the muon completes many millions
of orbits bcfore it decays or is captured by the nucleus. During its short lifetime,
it can make transitions to or from excited states by absorbing or emitting quanta
of radiation. The study of these transitions provides additional information
about tlre nucleus.12 Since the muon is about 207 times the mass of the electron,
the radius of a Bohr orbit for a bound muon is about 7616 the radius of the
corresponding orbit for an electron. Taking into account the difference
in nuclear charge, a muon bound to a heavy element will have orbital radii
"-,10*a times the hydrogen radii, or ^,10-12
centimeter. According to Wheelerl3
a muon in its ground state orbit spends nearly half of its lifetime within the
captor nucleus. Accordingly, we conclude that the nucleus is nearly transparent
to the muon, which is to say that the muon interacts weakly with nucleons.
PROEI-E/U 3-6
(u) Using the Bohr model find the energy of the photon
emitted when a mu meson makes a transition from the
first excited state to the ground state of mu-mesic
Pb2o8.
(b) The 2Pg-lS+ transition energy for mu-mesic Pb208 is
known to be 5.963 MeV. Suggest several reasons for
the discrepancy between this value and your answer
in (a).
PROBLEM 3.7
PROBTE/I4 3.8
PROBLEM 3-9
PROBLEM 3.IO
Gas inlet
ra
J. Franck and G. Hertz, Verhandl. deut' ph2sik. Ges.161 457 and 512 (1914)'
X.RAY SPECTRA AND THE BOHR THEORY 89
? 3oo
L
,i
': 200
L
Figure 3-8 Current vs. voltage in the Franck- i3
o loo
A')
510
Plrrte volttrge, \/olts
K"
a
5 25 kV
-4
!
L
ti
Figure 3-9 X-ray spectrum of molybdenum.
Kp The continuous spectrum is the bremsstrahlung
X 15 kv radiation.
hz
a)
t2
Wavelength, angstroms
15
F. K. Richtmyer, E. H. Kennard, and T. Lauritsen, op. cit.rp. 349.
90 THE CONCEPT OF THE NUCLEAR ATOM
photoelectric effect the photon frequency provides an upper limit on the change
in kinetic energy. Since the initial kinetic energy of an electron is determined
by the voltage on the x-ray tube, the short wavelength cutoff is given by
ltc 12,+00
^
Ac : (3.2e)
,V: V,
where Zis in volts and A is in angstroms. Thus, a 12 kilovolt x-ray tube has a
cut-off wavelength of about I angstrom. Not many photons having wave-
length )", are emitted, because the majority of electrons perform mechanical
work on the target when they stop and a large fraction of their kinetic energy
is converted into heat. Equation 3.29 provides an accurate methodr6 for
determining the ratio ef h, since a plot of r, vs. V is a straight line whose slope
is elh.
Superimposed upon the continuous spectrum are a few discrete peaks
whose wavelengths are characteristic of the target material. A few of these
peaks are also shown in the figure . The characteristic x-ray spectra of many
elements were studied by Mosel.y,tt who showed that the frequency of a given
line varies from element to element as the square of the atomic number. A
plot of some of Moseley's data is shown in Figure 3-10. This is precisely the
dependence predicted by the Bohr theory in Equation 3-28. Furthermore,
the Bohr theory accounts for the discreteness and the order of magnitude of
the photon energies of the characteristic spectrum by assuming that these
photons are emitted as a result of transitions involving inner electron shells.
If for example, a high energy electron knocks out an electron from the first
shell of a target atom, the vacancy may be filled by a transition from the second
shell, or the third, or higher shells. Since the electrons of the first shell have
been traditionally called K-electrons, the photon emitted by a transition from
the second shell to the .K shell is called Ko radiation, that due to a transition
from the third shell to the K shell is called Kuradiation, and so on. The second
9zo
X
--
N
16
Figure 3-10 Moseley's results for the effect of
irz atomic number on the frequency of a given x-ray line.
9s
f-
216202428323640
Atomic number
thgll is designated as the L shell, the third M, and so forth, so that Lo, LB,. . . ,
Mo, M*. . . radiation is defined in a similar fashion.
Since Equation 3.28 shows that the photon energy is proportional to Zz,
a transition from the second to the first shell for coppe, (Z : 29) would result
in the emission of a photon having an energy about 84l times the energy of the
longest line of the Lyman series of hydrogen. This approximate calculation
gives l.++ Afor the K* line of copper, wherlas the correci value is 1.54 A. The
agreement is not nearly so good for transitions involving shells of higher order
because of the shortcomings of the Bohr theory and the importanf screening
effect of the nuclear charge by electrons in the innermost shells.
A process which generally competes with x-ray emission is the internal
photoelectric effect or Auger efrtt. Here, the x-ray photon does not actually
appear, but an equivalent amount of kinetic energy is given to an outer
electron, which is in turn ejected from the atom.
PROBLEM 3.13
2*o
lo : t
rrr"
where wn is given in Equation 3-27.
(b) Show that in the limit as n gets very large, the Bohr
frequency relation reduces to the expression in (u).
This is an example of the correspondence principle, which
states that a quantum theory result should agree with
the equivalent classical solution in the limit of large
quantum numbers.
92 THE CONCEPT OF THE NUCLEAR ATOM
r: roA+, (3.30)
which expresses the radius of any nucleus in terms of its nucleon number and
t44
z
L
j
.J
/
Eao
=
L
264
CJ
z
^:.
48 =r/
32 7
l6
l6
Proton number, Z
Figure 3-l I Plot of N vs. Z for stable nuclei.
NUCLEAR STRUCTURE AND SPECTROSCOPY 93
the constantro. The accepted value of ru is 1.2 x l0-13 cm (or 1.2 F), which was
obtained from measurements of the nuclear charge distribution by means of
high-energy electron scattering.18 Other methods of estimating nuclear radii
are in good agreement with this 1g5uh.1s,2o Equation 3.30 leads to two
interesting conclusions. First, the nuclear force whicl-r holds the nucleus together
must be charge independent, that is, the n-n, p-p, n-p interactions must be
essentially the same since it is .z{, and not Z or //, which determines the nuclear
radius. Second, the nuclear density is apparently the same for all nuclei, since
it is given by the expression
A A
V 4rt
T,,U
Thus, nuclear matter is roughly (oolro)s:1012 times as dense as ordinary
matter.
The binding energ) of a nucleus can be expressed in terms of the difference
in rest mass energy between the constituent nucleons and the composite nucleus.
Thus, the binding energy is the mass defecl multiplied by c2. For example, the
binding energies of the deuteron and the alpha particle are as follows:
z,
j
O
O
L
q)
b0
(')
C.)
b0
7l
ll!*a for example, Robert Hofstadter, Annual Reuiews o.f Nuclear Science 7r231 (1957).
le L. R. B. Elton, Introductor2 Nuclear Theor2. W. B. Saunders Co., Philadetphia, 1966.
20 R. D. Evans,
The Atomic Nucleus. McGraw-Hill Book Co., New york, 1955.
94 THE CONCEPT OF THE NUCLEAR ATOM
for most nuclei except the lighter nuclei. The sharp increase in binding energy
for A : 4 accounts for the unusual stability of the alpha particle. The gradual
fall-off of binding energy per nucleon for the heavy nuclei is due to the effect of
the Coulomb repulsion, which is always present although it is dominated by
the nuclear potential. fhiq increase in Coulomb repulsion is partially com-
pensated by an increased nuclear attraction resulting from an increase in
the I,{IZ ratio as Z increases (see Figure 3-tl). The nearly constant nuclear
energy per particle contrasts sharply with the dependence of electronic binding
energy per electron on Z (and hence A) in atomic potentials.
The study of positron emission from mirror nuclei provides an estimate of the
Coulomb contribution to the binding energy as well as an approximate value
of the nuclear radius. Mirror nuclei are pairs of nuclei having odd ,4 such that
Z and i[ differ by one. For example, Oru-(Z : B, ff : 7) and N15(Z : 7,
1/ : B) are mirror nuclei, each containing 15 nucleons. Thus, the binding
energies of these nuclei should be the same except for the correction due to the
difference in the Coulomb interactions. Using the classical expression for
the Coulomb energy of a uniformly charged ball of charge Ze and radius r,
the difference in Coulomb energy between nuclei of atomic numbers Z and
Z-lis,
L,E :Yrt, - (z - t),1 : Y* {rt - t). (3.31)
PROBLEM 3-14
Find the mass defects of O15 and Nr5. Compare this differ-
ence in rest mass energy with the Coulomb energy calculated
from Equation 3.31.
where & is the range of the nuclear force, ^-,lQ-ra cm : I F. This expression
is not valid for r < 0.4 F, where a repulsive term dominates. Forr: D : I F,
Repulsive core
Coulomb repulsion
l
Figure 3-13 Schematic d iagram of the
nucleon- nucleon potential. Since the exact shape
of the well is not known, it is approximated by a
square well. For a neutron there is no Coulomb
repulsion. Yukawa potential
- -40 MeV
i t,i
V ,-.,, I to l0 MeV for a nucleon;- for r : ljb : l0 F, V ,-' l0 to 100 eV. A
schematic diagram of the nucleon-nucleon potential is shown in Figure 3-13.
Another property of the nuclear force, which is quite a departure from the
behavior of Coulomb and gravitational forces, is that of saturation That is,
a nucleon seems to interact with only a limited number of other nucleons,
analogous to nearest-neighbor interactions in solids or the saturation of
chemical bonds in ligand theory. It is this characteristic that is evoked to
explain the unusual stability of the alpha particle. There appears to be little
or no attraction between an alpha and another nucleon and hence there is no
stable nucleus of A : 5.
As a result of the discovery of natural and induced radioactivity, as well
as the research leading to the identification of the "mysterious emanations"
called q, f , and y rays, it became evident that the nucleus itself must have an
inner structure. From careful measurements of the energies of emitted particles
and y rays, schematic diagrams of the energy levels and decay schemes of many
nuclei'have been produced (see Figure 3-14). One striking difference between
Ground strrte
Grotrnd state
Figure 3-14 Schematic energy level structure for hypothetical nuclei. An a-emitter (a) or a
p-emitter (b) may decay directly to the ground state or to an excited state of the daughter nucleus.
95 THE CONCEPT OF THE NUCLEAR ATOM
optical and nuclear spectroscopy is associated with the energies of the photons
in each case. Recall that the energy of a photon in the visible region is of the
order of a few eV, while an x-ray photon is generally in the keV regime.
Since nuclear energies are of the order of MeV, 7 photons are frequently in
the MeV energy regime. However, there are no sharp demarcations between
these energy regimes and accordingly, "hard" x-rays may equally well be
called "soft" y rays. The atom emitting a photon must suffer a recoil in order
to conserve momentum as well as energy. For visible photons this recoil
energy is often negligible, but it can become significant for energetic gammas.
PROBIE/I4 3-15
E2
LE - wIr" (3.3 3)
The effect of such recoil on the light emitted from a source consisting of many
atoms is to broaden the spectral line associated with the transition. * Of
course, every spectral line has a natural width (there is no perfectly mono-
chromatic source), but the recoil broadening can be many times greater than
the natural line width. An important exception to this occurs in some crystals
when the emitting atom is bound to the lattice with sufficient energy to prevent
its recoil. In such cases the emitted line is very sharp since the mass in the
denominator of Equation 3.33 becomes the mass of the whole crystal and
AE ^,0. When recoilless emission occurs the linewidth is essentially its
natural width, and the photon may be absorbed by an unexcited nucleus of
the same species by the process of resonance absorption. This phenomenon is
known as the Miissbauer ffict after its discoverer.22 It has become an important
tool in many areas of physics, but particularly in solid state and nuclear
physics.2s
Brief mention should be made of the gamma emission process when recoil
does occur. Here we may distinguish three separate cases. (1) If the free-atom
Zero
p'i Phonon
Line
i^
Figure 3-15 A hypothetical gamma ray spec-
trum for an atom in a solid at low temperature.
The narrow zero-phonon line is the one normally
used for Miissbauer spectrometry. z
Plionon Wing
O
- 0.1 - 0.05 0
Energy Sliift - eV
recoil energy, Equation 3.33, is greater than the binding energy of the atom
in the solid (15 to 30 eV), the recoiling atom is dislodged from its lattice site.
(2) If the free-atom recoil energy is less than the binding energy but large
compared to the phonon* energy (-0.1 eV), the gamma-emitting atom will
remain in its site and dissipate its recoil energy to the lattice as heat. (3) If the
free-atom recoil energy is less than the phonon energy, then it is possible to
observe either recoil-free gamma emission or gamma emission accompanied
by the excitation of a discrete phonon of energy ha ^, 10-2 eV. This third
case is the effect discovered by Mijssbauer, and the gamma spectrum consists of
a sharp peak corresponding to recoil-free gamma emission (the zero phonon
peak) as well as a continuous spectrum on the low energy side (the phonon
wing) corresponding to gamma emission accompanied by phonon excitation
(see Figure 3-15). The reader should note that recoil-free gamma emission is
analogous to the elastic scattering of photons as in the production of the
unmodified line in the Compton effect (see section 7 of Chapter 2).
A frequently used source for Mossbauer studies is ,rCo57, which captures a
-K-electron and decays to an excited state of ,.Fe57. When Fe5? returns to its
ground state it emits a recoilless photon having a linewidth of about l0-1r
times its energy of 14.4 keV. Such a sharp line can easily excite an Fe5? atom
in its ground state by the process of resonance absorption. Thus, a sample of
Co57 provides a source of radiation of unbelievable spectral purity and un-
excited Fe57 provides an equally sharp detector. To get some appreciation fior
the sharpness of this line, note that a relative velocity of only a few hundredths
of a millimeter per second between source and detector will produce a Doppler
shift great enough to prevent the resonance absorption.
PROBLEM 3.17
Several nuclear models have been proposed to account for the observed
properties of nuclei. Unfortunately, no single model satisfactorily explains all
of tt experimental facts, so that more than one model must be invoked unless
one is" concerned only with a specific property of the nucleus.
Perhaps the most popular model is the shell model. In this model each
nucleon is regarded as moving in an orbital state under the influence of a
nuclear field *tri.h represents the average effect of all the other nucleons. The
allowed quantum states are found by using the j-j coupling scheme of combining
angular momenta, which will be discussed in section 5 of Chapter B. The
noieworthy successes of the shell model are its ability to account for nuclear
spins, the stability of the alpha particle and certain other nuclei (those associated
with the so-called "magic numbers"), and for many features of nuclear spectra.
The shell model not only ignores such problems as the saturation of the
nuclear force, the nearly constant nuclear density, and the binding energy per
nucleon, but it gives the wrong results for the scattering of neutrons from
nuclei. It turns out that neutrons interact so strongly with nuclei for certain
discrete energies (sometimes spaced only a few hundred eV apart) that they
are often trapped for a while before being ejected by the target nucleus. Such
trapping events are called "resonances."
Th; model proposed to account for these resonances is the liquid-drop
model. Here we regard the collective behavior of the nucleons as a many-body
system analogous to the lattice vibrations of a solid. The energy states of the
system are now the excitation energies of the collective system (analogous to
the normal modes), which one would expect to be a set of discrete but closely
spaced levels. Hence, a neutron having one of these energies will, in a collision
with the nucleus, immediately share its energy with the whole system. A
considerable time interval will elapse (compared to the transit time through a
distance equal to the nuclear diameter) before a neutron accumulates enough
energy to emerge as the "scattered" particle.
In addition to explaining the resonances, the liquid-drop model is useful
in visualizing the satuiation of nuclear forces and the phenomenon of fission.
It has also been used to develop a phenomenological formula known as the
semi-empirical mass formula.za
Another useful model of the nucleus is called the optical model. Here the
nucleus may be regarded as a cloudy crystal ball, since it is essentially opaque
to short wavelength neutrons (high energies) and it gets increasingly transparent
as the neutron wivelength increases (lower energies). Its cloudiness is a measure
2a See, for example, R. D. Evans, op. cit., p. 366, or L. R. B. Elton, op. cit., p. l1B.
FUNDAMENTAL FORCES AND EXCHANGE PARTICLES 99
of the absorption probability and its index of refraction is related to the deflec-
tion of scattered neutrons. The phenomena of absorption and resonances can
be accounted for by means of canceling and reinforcing traveling waves,
respectively.
(a) (b)
Figure 3.l6 (a) Emission and reabsorption of a virtual photon by an electron. (b) Creation
and annihilation of an electron-positron pair by a photon.
\
\
\
){
(a) (b) (c)
Figure 3-17 Feynman diagrams depicting (a) the interaction of two electrons, (b) the Compton
effecr, and (c) electron-positron annihilation. A dashed line represents a photon and a vertex indicates
the absorption or emission of a photon.
25J. Weber, Phys. Today2l,34 (1968) ; Ph2s. Reu. Letters 2l' 395 (1968).
26 H. Yukawa, Ph2s. Math. Soc. Japan 17r 48 (1935).
ANGULAR MOMENTA AND MAGNETIC MOMENTS l0t
where 2, is the Compton wavelength for the electron. Note the reciprocal
relationship between the mass of the exchange particle and the range of the
force. This explains why particles of zero rest mass are requireJ for the
inverse square forces which extend to infinity. Conversely, u fo.ce with a
finite range requires an exchange particle with a non-zero rest mass.
There is at least one additional force which is involved in the so-called
"weak interactions" of particle physics. This is the force that is responsible for
nuclear p'decay, muon decay, and other lepton decay schemes. The exchange
particle involved in this force has not yet been discovered.
it:i xiA,
where 7 it
i., absolute units of current and the caret signifies a unit vector. But
A: nrz andi: -ril2ncr, where e is in esu. Then
e: e;
Ft: - :-fvL?:---l
2c 2mc
xF:-2*,el (3.34)
Here m is the mass of the electron. The unit vector I is included to show that
the magnetic moment is directed antiparallel to the orbital angular momentum
because of the negative charge of the electron. The quantity ltn is called the
Boltr magneton, and its value is approximately
eh
Fn : : 9.27 x l0-2r erg/gauss.
%nc
102 THE CONCEPT OF THE N UCLEAR ATOM
The ratio of the magnetic moment to the orbital angular momentum is called
the classical gyromagnetic ratio (o. the magnetomechanical ratio) and is
expressed as
rt : li'l 2mc
0 :Fr
(3.36)
l7l: h
Ts: /'t e
(3.3 i )
SI mc
From the foregoing we note that the smallest unit of magnetic moment for the
electron is the Bohr magneton, whether one considers orbital or spin angular
momentum.
27 G. E. Uhlenbeck and S. A. Goudsmit, Naturwiss.13r 593 (1925).
28 P. A. M. Dirac, Proc. Rol,tal Sociefi,t (London) ll7, 610 (1928) and ll8' 351 (1928).
ANGULAR MOMENTA AND MAGNETIC MOMENTS t03
The intrinsic magnetic moment of the proton, which is also a spin- ] particle,
and the deuteron, which is a spin- l particle, were determined by Stern and
his co-workers from beam deflection experiments on hydrogen and deuterium
using the molecular species Hr, D., and HD.2e Later, more accurate values
were obtained for these nuclei by means of beam resonance and nuclear
magnetic resonance techniques. These methods have now been applied to
nearly all stable nuclei. The neutron magnetic moment was determined from
neutron scattering experiments in magnetized iron.s0
By analogy with the Bohr magnetorl we define the nuclear magneton as
L
en
FN : 2M"r:
5.050951 x l0-24 erg/gauss,
where the mass of the proton has replaced the electronic mass. The current
values of the proton, neutron and deuteron magnetic moments are:
Itrp : 2.792782p.'
Itrn: -1.9135Pr1-
Itra :0'8576Pr.''
Note that the deuteron magnetic moment is not the algebraic sum ofthe moment
due to the neutron and proton. This discrepancy is believed to be evidence for
a non-central contribution to the interaction between nucleons.
PROBLEM 3.19
2e For references and an excellent summary of the beam experiments, see N. F. Ramsey, Molecular
Beams, Oxford University Press, N.Y. (1956), p. 102 cf.
s0 L. W. Alvarez and F. Bloch, Phys. Reu.57,
lll (1940).
t04 THE CONCEPT OF THE NUCLEAR ATOM
df
dt
== i,xA:y,/xA.
ButldTl: / sin 0 d.4, so we may write the scalar equation
I sin 6.4!
dt
: ytl # sin 0.
ttJ.H.Van Vleck, The Theory d Electric and Magnetic Susceptibilities, Oxford University Press,
London, 1932, p. 22.
LARMOR THEOREM AND THE NORMAL ZEEMAN EFFECT t05
magnetic moment.
coordinate system. Thus, if the energy levels of an atom are known (with no
applied field) and then a magnetic field is turned on, a transformation to the
coordinate system rotating at the Larmor frequency will restore the original
field-free problem. It follows that the solution of a problem involving a static
magnetic field can be written approximately as the superposition of the no-
field solution and a rotation at the Larmor frequency. The exact expression
which shows the quadratic term in the field is considered in Problem 3-20.
Using the Planck relation, the energy associated with the Larmor frequency is
L,E: Iconh:'-2mc
*tlg -- +I Iua%, (3.3e)
where the signs refer to the sense of the rotation. This energy difference
should be recognized as the potential energy of a magnetic dipole whose
moment is one Rohr magneton, since the dipolar energy is given by
L,E : _i, .d
Hence, the plus sign (higher energy) in Equation 3.39 corresponds to anti-
parallel alignment (as in Figure 3-18), while the minus sign (lower energy)
indicates parallel alignment. If the energy of an electron having a moment
p-a is Eo with no applied magnetic field, then it can take on one of the energies
Eof FaCI in the magnetic field #,provided that one can neglect quadratic
terms in 0. It was found experimentally that in a collection of identical atomic
systems of the type we have described above, a magnetic field actually produces
a triplet of levels (called a Lorentz triplet) whose energies are{Eo, Eo * Ita%\.
The existence of these levels can be confirmed by studying the optical tran-
sitions to and from these states. This phenomenon is known as the normal
Zeeman tfttt.,, The reason for the appearance of three levels will become
apparent in the next section when we discuss spatial quantization' There we
will see that the three energies correspond to the three allowed projections of the
F:e@+L;"h1. c
3s See, for example,C. P. Slichter, Principles of Magnetic Resonance, Harper and Row, New York,
Ig63; also, M. Sparls, Ferromagnetic-Relaxation Theory, McGraw-Hill, New York, 1964.
SPATIAL OUANTIZATION 107
I t. SPAT|AL QUANTTZATTON
The classical model of the previous section is often useful for describing
certain aspects of resonance and relaxation experiments, but at the same timi
it can be misleading if taken too seriously. For example, the picture of a pre-
cessing angular momentum vector relaxing through a continuum of p^olut
angles conflicts with a fundamental concept of quantum mechanics which is
referred to as spatial quantization. Stated briefly, spatial quantization means
that the projection of an angular momentum vector along any single axis in
space can be only one of a discrete set of allowed values. At a given time there
can be only one axis of quantization; if it is altered, szy, by changing the
direction of an applied uniaxial magnetic field, quantized states ."irt-o.rly
along the new field direction. The close connection between space and angular
momentum will be elaborated further in section 3 of Chapter 7.
The experiment proposed by Stern and performed by Gerlach and Sternaa
provided the first conclusive evidence of spatial quantization in 1922. Silver
atoms were evaporated in an oven and collimated into a narrow beam which
was passed through an inhomogeneous magnetic field as shown in Figures
3-19 and 3'20. Although a uniform field produces no net force on a magnetic
dipole (only torque), a non-uniform field can deflect a dipole with I net
translational force. This can readily be seen by considering a magnetic fieId, g
which acts in the positive z-direction, and which is also designed to have a
gradient in the positive z-direction (FiS. 3-19). Assuming that all other
derivatives of the magnetic field are zero, we find that the translational force
on a dipole, oriented at an angle 0 with the z-axis, is
as ^
-l.rcos0-;-k,
oz
Fz: /rcos0 as
^ .
oz
Thus, a dipole aligned with the field is acted upon by an upward force of
magnitude p'(1CI102), while a dipole aligned antiparallel to the field is acted
uPon by u downward force of the same magnitude. For arbitrary orientations
the force could have any value between
-pQola) and + p,egaloz).
In the absence of spatial quantization, the beam of particles would contain
a continuous distribution of angular orientations of the precessing dipoles, and
the action of the non-uniform magnetic field would spread the narrow beam
into a band at the detector screen. However, instead of a continuous band,
Gerlach and Stern obtained two distinct lines whose breadth was due to the
spread in particle velocities rather than to a continuum of dipole orientations.
This can only be explained if the dipoles are permitted just two orienta-
tional states, one state having a component in the { z-direction and the other
3a O. Stern, Z. Pltsik 7 r 249 ( l92l ) ; W. Gerlach and O. Stern, Z. Ph2sik B, I l0 and 9, gag 0922) ;
Ann. Ph2sik 74,673 (1924).
t08 THE CONCEPT OF THE NUCLEAR ATOM
W, N
Collimator
( -t I Screen
Figure 3-20 Particle trajectories in the experiment of Gerlach and Stern. The deflections of
the beams are greatly exaggerated.
Detecting
screen
Beam of
particles
of
angular
momentum
)h
Screen
,h:tl:,,
Apparatus for verifying the quantum states of particles having one unit of angular
,.J","t:;.t-t,
happen if each of these three beams were, in turn, subjected to another appar-
atus identical with the first. In Figure 3-21 three such beams are shown
entering a second magnet with the same field orientation as the first. If only
one beam at a time is permitted to enter the second magnet, just one beam will
appear at the detecting screen. Hence, no further splitting will occur, regard-
Iess of which of the three beams is allowed to enter the second apparatus.
The second magnet produces an additional deflection of the upper and lower
beams from which one can verify that the magnetic moments of the particles
which comprise them correspond to + 1t,u and - Fa, respectively. The center
beam is undeflected and thus corresponds to zero projection of the magnetic
moment along the vertical axis. From this result we draw the following con-
clusions. The first magnet establishes an axis of quantization and sorts the
initial beam into three quantum states. The second magnet does not alter
these quantum states but serves to confirm that the sorting of the beam into
pure states has been preserved during the transit of the particles from one
magnet to the other. On the other hand, if one were to alter Figure 3-21 by
rotating the second magnet through an angle about the beam axis, the situ-
ation would be quite different. Each beam from the first magnet, which was
a pure state in the first magnetic field, now appears as a mixture of the three
quantum states in the new field orientation. Accordingly, if one beam at a
time from the first magnet is allowed to enter the second, each will split into
three components in the second apparatus.
PROBLEM 3-21
su/ylMARY .*,iiii:r
REFERENCES
American Institute of Physics, Nuclear Structure, Selected Reprints, 1965.
A. Beiser, Perspectiues of Modern Ph-ysics. McGrarv-Hill Book Co., Inc', New
York, 1969.
C. H. Blanchard, C. R. Burnett, R' G. Stoner, and R. L. Weber, Introduction
to Modern Ph\sics,2nd ed.Prentice-Hall, Inc. Englewood Cliffs, N.J., 1969.
M. Born, Atomic Ph2sics,6th ed. Hafner Publishing Co', New York, 1959.
R. M. Eisberg, Fundantentals of Modern Physics. John. Wiley and Sons, Inc.,
New York, 1961.
L. R. B. Elton, Introductory Nuclear Theor2. W. B. Saunders Co., Philadelphia,
1966.
Robley D. Evans, The Atomic Nucleus. McGraw-Hill Book co., Inc., New York,
1966.
Richard P. Feynman, Robert B. Leighton, and Matthew Sands, The Feynman
Lectures on Ph2sics. Addison-wesley Publishing co., Inc., Reading, Mass.,
I. INTRODUCTION
In this and the ensuing chapters we shall develop and use the formalism
of non-relativistic quantum mechanics as it is understood today. We begin
this chapter with a brief aside on the "old" quantum theory and its application
to the hydrogen atom. Although this material may be omitted without
hampering the reader's understanding of the current theory, it is included here
for the following reasons. It represented a forward leap in that it attempted
to incorporate quantum concepts into current theories. It correctly predicted
a large body of experimental results from a few simple rules. It is of consider-
able historical importance because it occupied many of the greatest minds, and
like the ether theory of pre-relativity days, it thus set the stage for the appear-
I
ance of the modern theory.
The new quantum mechanics appeared in two forms which were later
shown to be equivalent. One form is known as waae mecltanics and is generally
regarded as Schrcidinger's formulation,l although it is based heavily on the
work of de Broglie.z The other approach is called matrix mecltanics and it is
chiefly credited to Heisenberg, although Born and Jordan shared in its develop-
ment.8 Since wave mechanics is more easily grasped intuitively we shall begin
our study with the development of the de Broglie theory of particle waves and
the evolution of the Schrcidinger equation. All of Chapter 5 is devoted to
applications of the Schrodinger method to one-dimensional systems. The
essentials of matrix mechanics will be introduced in Chapter 6. In the re-
maining chapters the wave and the matrix formulations will be used freely
and interchangeably.
1E. Schrcidinger, Ann. Ph2sik (4)79,361 (1925); 79,489 (1925) ; 80,437 (1926); 8l' 109 (1926).
sL.de Broglie,Naturell2r540(1923); Thesis,Paris(1924); Ann.Ph2sique (10)2(1925).
3W. Heisenberg, Z. Physik33,B79 (1925); M. Born and P. Jordan, Z. Physik 34' B5B (1925);
M. Born, W. Heisenberg and P. Jordan, Z. Ph2sik 35r 557 (1926).
ilt
il2 THE DEVELOPMENT OF WAVE MECHANICS
mx +kx:0,
or
i +q'x -0,
where
k
692:-
m
Then,
x: xo sin cull
and
!* : mi -- mlJxo cos a-rl'
Equation 4.1 becomes
nh: or : *,'rfr cosz at dt
fn hq-
Ii
I .t
: maxil cos'odo
Jo
marxfi.
Therefore,
ua-
oflh
'rO- ,
@m7T
E:T+y:gmxzalkxz
: lmazxfi,
of,
En: nath : nltY., (+.2)
In both the classical theory and the old quantum theory, the ground
state energy of an oscillator is incorrectly given as zero. Howcver, th; level
spacings are correct in th.ese older theories. From the oscillator energy levels
obtained in Equation 4.2, there is no information about which transitions are
most likely to occur, or in fact, whether any are forbidden. Information of this
kind goes under the general heading of selection rules and is readily obtained in
the new quantum mechanics. However, in the old theory, selection rules were
inferred by comparing the system with the behavior of a classical system;
that is, by employing what is called Bohr's correspondence principle. Thus, since
a classical oscillator will emit only one frequency (and no harmonics) , if a
quantum mechanical oscillator is to correspond to the classical result in the
limit of large z, then we must have the selection rule Az : f l. We had
already assumed transitions between adjacent levels in our discussion of
Planck's oscillators in section 3 of Chapter 2.
If we treat a two-dimensional harmonic oscillator as two independent
one-dimensional oscillators in the x- and 2-directions, the energy levels are
En : nhat,
b
Figure Cl Bragg reflection of a photon I
by a crystal.
Atomic planes
6 W. Duane, Proc. Nat. Acad. Sci. 9n l58 (1923); A. H. Compton, ibid.91 359 (1923).
il4 THE DEVELOPMENT OF WAVE MECHANICS
amourrt of momentum in the z-direction equal to (2hl),) sin 0. But the momen-
tum must satisfy the quantum conditions given by Equation 4.3, so we have
nA:2dsin0, (+.+)
n!:2dsino.
ma
(4.5)
PROBLEM 4.2
PROBLE/I4 4.3
Although the Bohr theory was quite successful in predicting the spectrum
of hydrogen, there remained an unexplained fine structure or splitting of the
lines. This splitting amounts to about one part in 104 and cannot be seen in
spectrometers of low resolving power. Sommerfeld proposed that if elliptical
SOMMERFELD'S RELATIYISTIC THEORY il5
Zez
lti : pr p2
-,,
I16 THE DEVELOPMENT OF WAVE MECHANICS
.. {2 Zez
pr:Tr"- ,r.
Multiplying by i and integrating,
u!,ar:l(#-#) *,
or
iprr:-L*tt'-rF
2pr' ' r
-r r'' (4'7)
: /'L'
':!/f dt dP'
dP prz
(ldrY_ 1
,2p,Ze2 ,2pE
\*oBt 7- r', - f''
Introducing z : 1lr,
du Idr ,I o 2p,Ze2
u-,2pE
dP:-FaB:-V-u'r ,"'z /'
and
du
rd0: l2tL + 2t Zt' tt
l-i -t, - uz
lntegrating,
u :!r : rz:'
dz +.\ry!
'! f4 +ry.sin
' lz (B - oi, (4.8)
I 1*esin(p-pr) : a {a'z-62-.,n
It:1:ffi *+ n, sin(B - 0o),
where bfa: \/l - e'z. Comparing this with Equation 4.8 we immediately
SOMMERFELD'S RELATIVISTIC THEORY lt7
PROBLEM 1-1
Show that the time averages of 7" and Z satisfy the same
relation as that satisfied by circular Bohr orbits, namely,
E:+V:-f.
n oo : ktt (a.loa)
$
f
Q 1,, d4 : mh (4.10b)
J
Q pu d0 : neh (a.10c)
J
r
a, : n,lt
! f,
(4.10d)
The quantum number m is called the magnetic quantum number because of the
role it plays in distinguishing the energy levels of the atom in the presence of a
magnetic field. Since the axis about which P, is measured is neither unique
nor stationary, we must transform Equation 4.10c before it can be integrated.
I18 THE DEVELOPMENT OF WAVE MECHANICS
": ped;:podi+pedo.
Substituting this into Equation 4.10c,
: : (k -
f tn dP - lod6)
ltoh m)h,
of,
k -- ,u * m. (a.11c)
The integer ,t is called the azimutltal quantum number. It can take on the
values lr213r. . ., with zero excluded.
Sommerfeld's integration of Equation 4.10d will not be repeated here but
we will merely state his result, namely,
kh(-L-
\1/l e2
r)/ :n,h,
-
or
! :Y! k :: k'. (4.11d)
b- k
The quantity z in the last expression is called the total quantum number, since it
is defined as the sum of the radial and azimuthal quantum numbers, which
is to say that
tL:nr*ne*m.
Combining Equations +.9, +.11a, and 4.11d, we obtain the equalities
a nh Zez l- t,
b:7:- /lE'
from which the quantized orbits and energies are given by:
nzao
O:2,
.
h-_
ka knao
u--t
NL
and (4.r2)
LooZ'
un-
H
)
nZ
n:2
k:1
b : 2a,,
n:3
k: I
b: 3a,,
n:3
k:3
Figure 4-3 Sommerfeld's orbits for n : l, 2, and 3.
woZ'f . oz/z ll
En---;l'--\Z- *)l (4.13)
The quantity
e2
h.c
: 7.297 x l0-3 ^,Ir37
is called the fne structure constant since the term in Equation 4.13 in which a
appears, correctly accounted for the fine structure splitting of the lines of the
hydrogen spectrum.
Although the old quantum theory achieved many successes in atomic and
molecular spectroscopy, it was an incomplete theory in the sense that none
of its recipes for quantization were derived from first principles. Since it
THE DEVELOPMENT OF WAVE MECHANICS
The idea of associating both a wave and particle nature with the electron
was first proposed by de Broglie in his doctoral thesis in 1925.8 His-work was
motivatea ny the mystery of the Bohr orbits, which he attempted to explain
by fitting a standing wave around the circumference of each orbit. Thus,
de Broglie required that n). : 2rr, where 2 is the wavelength associated with
the nth orbit and r is its radius. Combining this with Equation 3.25 we immedi-
ately obtain the result that
hh
"1-_-_
- ml, P'
Assuming the existence of a natural symmetry in the properties of matter and
energy, he proposed that a material particle of total energy E and momentum I
must be accompanied by u phase wave, analogous to that ascribed to the
photon, whose wavelength is given by I : hlp and whose frequency is given
by the Planck formula, y : Elh. The Planck and de Broglie relations may be
expressed in the useful forms,
E:ha
and (+.r4)
!:hk'
where h. : hl2r, k :2rf )', and a-l :2nv.
The physical nature of such a particle wave was not clearly described by
de Broglie. Unlike a classical wave, the energy E of the particle wave is not
thought of as spread out over the extent of the wave, but is regarded as localized
with the particle. However, the. accompanying wave is essential in order to
account for the phenomena of interference and diffraction.
The concept of the de Broglie wavelength is one of the cornerstones of
modern quantum theory, and the simple relationship
1-- ph
holds for photons as well as for both relativistic and non-relativistic material
particles, provided that the appropriate expression for p is used. On the other
?
Albert Messiah, Quantum Mechanics, North-Holland Publishing Co., Amsterdam (1958),
Chapter l.
E
L. de Broglie, Ann. Ph2s. (Paris) 3,22 (1925).
THE DIFFRACTION OF PARTICLES t2t
hand, the de Broglie frequency has not been a very useftrl concept and it
comes into play only in the calculation of the phase velocity. Here a dis-
tinction appears between the relativistic and non-relativistic cases. If the rest
mass energy is included in the total energy E (as in de Broglie 's treatment),
then the phase velocity of the wave becomes
or'
t*) : cz. (4.15)
In obtaining Equation 4.15 we have expressed the relativistic momentum as
P : ymou, where , : Ft is the particle velocity and y : (l - 1lz\-tz. From
our knowledge of waves* we identify the particle velocity u with the group
velocity of the wave packet. Since special relativity requires that a be less
than c, we note that Equation 4.15 calls for phase velocities greater than c.
However, as no energy (that is, no signal or information) is transmitted at the
phase velocity, the fact that u > r constitutes no violation of the postulates of
special relativity.
In non-relativistic quantum mechanics the rest mass term is neglected
and the total energy E is merely the sum of the kinetic and potential energies.
Accordingly, the phase velocity of the wave associated with a non-relativistic
free particle (Z : 0) is,
*- @Ephka
k p 2m 2m-2'
that is, one-half of the particle velocity. Thus the phase velocity is not of any
physical significance. The group velocity of a particle wave, however, is given
by
da dE 1.
"-dk-kdk
expect from particles, the de Broglie wavelengths for a few special cases are
given in Table 4-1. Massive particles such as a bullet fired from a 0.22 rifle
or a fast baseball have de Broglie wavelengths of the order of l0-23 or 10-24
angstrom.
Since the wavelengths associated with 10 to 100 volt electrons correspond
to the atomic spacings of most crystalline solids, it was conjectured that
electrons of these energies ought to be diffracted by crystals in the same manner
as x-rays. This was confirmed experimentally in 1927 by the work of Davisson
and Germer,e in which they scattered low energy electrons from a nickel
crystal. Although electrons were scattered in all directions from the nickel
crystal, a distinct peak in intensity occurred at an angle which corresponded
to the first Bragg reinforcement for a wavelength of 1.65 A. The size of this
peak varied with the energy of the incident electrons and was found to reach
a maximum of intensity for 54-volt electrons whose de Broglie wavelength is
1.67 A! (See Figure 4-+.)
The Bragg equation, which gives the condition for intensity maxima in
the reflected beams, may be expressed as
where z is the order of the diffraction, d is the distance between the reflecting
planes, and the angles 0 and $ arc shown in Figure 4-5. For a mono-energetic
beam of particles at normal incidence, the first order maximum will be found
at an angle from the normal given by
20 :2 cos-'h.
44 eY 48 eV 5'1 eV 64 eV 68 eV
0o r: A
t.AS 0o r: t.ZZ A 0o I: 1.67 A 0' r: t.ss A 0' l,: t.+S A
Target
Figure4{ Angular plots of scattered intensity for low-energy electrons incident upon a nickel
single crystal. The energies and de Broglie wavelengths of the electrons are given for each plot.
(From C. Davisson and L. H. Germer,Phys. Rev.30,705 (1927). Used with permission.)
Incident beam
Reflected beam
Figure
s
G5 Bragg reflections of electron waves or x-rays for normal incidence on a crystal.
F :7, L
where the prime refers to the wavelength inside the crystal. For electron
energies that are small with respect to the rest-mass energy, the total energy
and the momentum are related as follows:
E-4*v. 2m
Then,
p:{2m(E-V)}+,
and the index of refraction becomes,
p:*:x:{t;n')', (4.r7)
where it has been assumed that V : 0 outside of the metal. If we take the
potential well of nickel to be V' ,-,
100 eV total energy we find that
-5 volts, then for incident electrons of
{iff}-: {r.05}} :
tt ,\-, r.02.
424 THE DEVELOPMENT OF WAVE MECHANICS
Figure 46 Diffraction of 50 kV
electrons from a disordered film of
CurAu alloy. The alloy film was
400 A thick. (Photograph kindly
furnished by Dr. L. H. Germer.)
Figure 48 Neutron Laue photograph of NaCl. (Photograph kindly furnished by Dr. E. O. Wollan.)
PROBLEM 1.6
using 2.15 A for the lattice constant of nickel, calculate
the angular positions of the first and second order maxima
for 100 eV electrons incident normal to the surface.
t26 THE DEVELOPMENT OF WAYE MECHANICS
Having established the fact that electrons have a wave nature, Davisson
and Germer continued their work by looking at higher order Bragg reflections
-tb
1"9 by studying grazing angles of incidence. They were able assign an
index of reflection slightly greater than one to nickel for 100-volt electrons.
_ Pressing the analogy with x-rays a step further, Thomson proposed that
electrons passing through a crystal should produce diffraction patteins similar
to Laue patterns. He subsequently observed these from foils of gold, aluminum
and other metals.lo Later, diffraction patterns were observed for other ele-
mentary particles (see Figure 4-B) as well as for atoms and molecules. *il
10
G. P. Thomson, Proc. Roy. Soc. A. ll7,600 (1927); 1l9p 651 (1928); 125,352 (1929); 133, I
(le3l).
* For an elementary review of this subject see Hugh D. Young, Fundamentals of Optics and Modern
Phltsics, McGraw-Hill Book Co., New York, 1968, Chapter 3.
Detecting
screen
_=_5
-'* ._1
--_->
--=+ 3_1,^
____-+
Incident
particles I
A
(a) (lr)
BA
|-t|l
(c) (d)
G9 (a) Double slit experiment with particles. (b) Distribution of particles recorded
Figure
on the screen due to diffraction from either slit A or B. (c) Distribution of particles recorded on the
screen due to diffraction with both slits A and B open. (d) Hypotheticil distribution of particles
recorded on the screen if wave effects are neglected.
refifiil
(d) Two slit electron Pattern
127
t28 THE DEVELOPMENT OF WAVE MECHANICS
'
of relative light intehsity versus position on the screen. The light intensity
at each point is interpreted as the square of a wave amplitude vector which
can be represented by the real or imaginary part of a function of the form,
q : )re(w-ail.
where the de Broglie relation has been incorporated in the last expression in
order to relate the frequency and wave number of the wave to the energy and
momentum of the particle. Although this complex wave function is not
directly observable (that is, measurable) its physical significance rests on the
assumption that the quantity
for the one-dimensional space along the r-axis. This integral must be finite
in order to represent a real particle. It is convenient to define the probabilit2
densit2 for the particle as
'
P(x,t) :
lY(x,t)12
{
(4.18)
t*t", t)t dr'
f F
I
f
I
where we then have I
i
dx : l. (4.1e)
I-:rr,q I
i
I
For physically acceptable wave functions it is'always possible to introduce an I
r
{.
appropriate factor in the wave function such that I*
i
i{!
J- t*{r,
t)lz dx :J]J.,', /)Y(r, t) dx - 1. (4.20)
I
I*
g
i-
,
f
I
THE WAVE FUNCTION FOR AN ELECTRON
'29
When Equation 4.20 is true, the wave function Y is said to be normalized. It is
evident from Equation 4.18 that when a wave function isnormalized the
probability density is simply the square of its absolute amplitude.
Furthermore, in order to account for interference effects we assume the
validity of the principle of superposition. When superposition is valid in optics
(that is, in the case of coherent light) we add the amplitudes at a point vectori-
ally and square the resultant amplitude to obtain the intensity at the point.
Thus,
I - (i, -l ir), : AL + A? + 2iL. i, - I, * I, I 2ir. ir,
where ,I represents the time average of the intensity over a full cycle. On the
other ha4d, when light is incoherent (the relative phases of the different
sources are washed out), the resultant intensity just goes as the sum
I-L|-Iz,
and the interference term 2i, .,4-, vanishes. Applying the principle of super-
position to the experiment of Figure 4-9, we write :
- Pz * pa * lYjYrl t lYsYel.
Here, the last two terms are the interference terms which depend upon the
relative phases of the two waves. The plus and minus signs correspond to
constructive and destructive interference. The phase factors in the wave
functions Y, and Vr play a role analogous to that of vector addition in the
above example from optics, thus indicating the importance of choosing complex
wave functions.
The device of representing an electron by u complex wave function can
be extended, of course, to other kinds of particles and even to atoms and mole-
cules. In order to be physically admissible, however, a wave function which
represents a particle must be finite, single-valued, and continuous. Furthermore,
it must vanish suitably as r --+ @ so that it can be normalized, as indicated
by Equation 4.20. The latter requirement is necessary for the validity of the
probabilistic interpretation of the modulus squared. The great significance of
the principle of superposition is that a superposition of physically acceptable
wave functions is itself acceptable for the representation of a real particle.
We will now study the Fourier integral theorem. Its use will enable us to
extend the mathematics of superposition from the simple case treated in section
8 of Chapter 2 to the formation of packets consisting of a continuum of fre-
quencies.
I3O THE DEYELOPMENT OF WAVE MECHANICS
PROBLEM 4.9
Y(") :lcosfi+Bsin!.
(u) Normalize this new function.
(b) Sketch the probability density as a function of x.
includes the end points of the period, then we add the assumption that the
value off (t) at each end pointis the arithmetic mean of its values at the right
and left end points. Then, in the interval, -(Tl2) < t < Tl2,
:
-f (t)
f + 2 o* cos o-nt + : bn sin ,.nt, (4.21)
where
an : + [:; ,,f
(t) cos ant' dt' ,
This may be written in a more convenient form* by means of the Euler identity:
f @ : -2-+ I:; , .f
(t'),n'n(t-t't ,,'l, (4.2s)
and note that we must eliminate the factor 1/7" before going to the limit.
Since z is restricted to integral values,
I Aar
(4.24)
-:-
T 2rr
"f
(t) ---*,,:." o' [1i,,-f
(t'1sn'*tt-t't 46"
if the integrals are absolutely convergent. The Fourier integrals are written
in a more nearly symmetric form by defining a function g(r) as follows:
(t) :
and
-f
#['*^ s(*),n"
(4.25)
g(r) :
#f_r, tr6'),-,,'i'
The functions/(t) and g(ar) are called Fourier transforms of one another.
Any reasonably well-behaved function /(l) can be represented by a super-
position of harmonic functions with continuously varying frequency and
weighting function g(ar). Conversely, g(r) can be represented by a super-
position of harmonic functions in time, each function multiplied by the
weighting factor ;f (t') .
The above expansion is in the time-frequenqt domain. It is evident that an
analogous expansion may be obtained in the position=wave-vector domain, or
what is conveniently called the coiirdinate-momentum domain. Physically, this
implies that our starting point was a spatially periodic function
,t@ + L) : rp(x),
,p(x) :
n:-@
where
',t,@):#l-- dkg(k)eck
I . (+.26)
6&):#[:_ dx'y(x')e-'*')
,t e) : ef y! norl)eifrr
l (4.27)
Since the delta function is zero everywhere but at the singular point, the only
contribution to an integral oecurs at that point. Then, using Equation 4.28
Thus the delta function may be thought of as a spike function having unit
area but a non-zero amplitude at only one point, the point of singularity, where
the amplitude becomeJ infinite. When integrated over all of space its effect
is to yiita the remaining factors of the integrand evaluated at the singular
point.
It is often convenient to place the origin at the singular point, in which
case the delta function may be written as
(:-4: sinax
d(") :
:g# t;*'dk: 1'-
2zr o-* \ ir /
lim
on* rrx
, (4.2e)
where a is positive and real. Let us examine the behavior of this function for
both small and large x. First, consider the limit as .r goes to zero:
Thus, d(0) : l:L*aln ---> oo, or the amplitude becomes infinite at the singularity.
For large lxl, sin axlrx oscillates with period 2rla, and it1 amplitude falls off
as |/lxl. But in the limit as a ---> o, the period gets infinitesimally narrolu so_
that the function approaches zero everywhere except for the infinite spike of
infinitesimal width ai the singularity. It now remains to show that the integral
of Equation 4.29 over all space is unity:
b ldE
r-',,,_tt-
dE
*- " - hdk -
dE :la,
and
E-L**r, (4.30)
* This is another example of the correspondence principle which states that quantum mechanics must
give the same result as classiial mechanics in the appropriate classical limit. For particles, the classical
limit occurs for very short de Broglie wavelengths.
t36 THE DEVELOPMENT OF WAVE MECHANICS
f(t\:ei'rt-
J\/)-
for -T <t <7.
-f (t) :0, for l/l > T.
The chopping (or keying) process will introduce many new frequencies in
varying amounts, given by the Fourier transform,
g(r) d(k)
I zr o I.
a;'r-1 ','*T
(a) (b)
Figure 4ll The spectral distribution resulting from chopping a pure, infinite sine wave in the
time domain (a) and in the spatial domain (b).
Multiplying by h and replacing the time interval T by At, we have the state-
ment that the minimum uncertainty product for this type of packet is
L,t.L,E:h.
Therefore, we may state Heisenberg's uncertainty principle for simultaneous
measurements of energy and time for such a packet as
L,t.L,E>h. (4.31)
+(k) : #,ll-rol,-ikn dx
:+ f" dr
! 2n J-" 'n'o'-o'*
l, sin (n'o
- k)o
v; fto - k)a
This function is also shown in Figure 4-lI, but here it is the wave vector (or
the momentum) that takes on a spread of values around fr,.* The breadth of
* Likewise, in single slit diffraction, we generally regard ks as a constant and allow its direction
to
take on a spread of values. This is equivalent to treating the slit as a source 9f Huygens wavelets.
t38 THE DEVELOPMENT OF WAVE MECHANICS
L,x.Lp,>h. (+.32)
For the packet shape shown in Figure 4-11, the minimum value of the
uncertainty product is fr at t :0. The packet shape that achieves the absolute
minimum value of hl2 is the Gaussian packet, which will be discussed later in
this section. (See Problem 4-2I.) However, as time progresses, any packet
will broaden in coordinate space so that for a given L,p, the uncertainty product
increases with time. (See Problem 4-16.) If we denote the time during which
the packet retains its form as To, then
mm
ro-E (Ar)',
where Ax is its initial breadth in coordinate space. The more massive the
particle, the more slowly its packet spreads with time. This is what we want
in order to satisfy the correspondence principle. Thus, for an electron which
is localized to I angstroffi, To n,,|Q-ro second, whereas for a .22 caliber bullet
localized to 0.1 mm, To
- l02a seconds. It is evident that the concept of a
classical trajectory, though meaningless for the electron, is quite appropriate for
a macroscopic particle.
Now suppose that we wish to represent a free particle as a superposition
of a continuum of plane wave states, namely,
Each value of $(k) serves as a weighting factor for the wave function it multi-
plies. That is, it tells how much of that particular wave function contributes
lo the packet at position x and time /. A plot of the values of $(k) verzus *
is known as the qpectral distribution of states, or simply as the spectrum of the
states. The plane wave states of Equation 4.33 may be represented by
Y*(', t) : Aen&v-@t\,
where ,4 is the normalization factor. Since infinite plane waves are not well-
behaved functions in the sense that they do not vanish at x : * @, special
proced.ures must be used in order to normalize them. One such procedure is
to use delta-function normalization; thus,
:
t- Yf,Yo dx : VPI*:'rue-ik'e dx IAPI:-enr&-k'l dx.
PARTICLES AS WAVE PACKETS I39
Comparing this expression with Equation 4.28, written for the delta function
in k,
6(k k')
1r- eie(tc-k')
- - 2n J--
I dX,
which is the normalized wave function for a packet containing a single mo-
mentum component hko. Note that the infinite plane wave ind thi delta
function are Fourier transforms of each other.
As a second example, consider a packet whose spectrum of ,t-values is the
Gaussian function,
f(k) : 4r-tcztzoz,
,p(x) : + f* dkr-r',zo2sika.
! 2rr J--
I I -)
t* zr/i- oj
r.
k
Figure 4-12 Gaussian packets. Each packet is the Fourier transform of the other. The breadth
of the packet is arbitrarily taken to be its standard deviation, as shown.
I4O THE DEVELOPMENT OF WAVE MECHANICS
:
,t,@)
f" I**on.,.o - # * [
ikx +
ry - ry)
=
#r-.,",,,f_dk
expl-
*&, - 2ikxoz - r,qf
: ir4')
fn'-."''' f*dk l- *, & +
exP
=#r_.,",,rf*e-u, du
- Aoe-'zo'lz'
PROBLEM 4.15
12
See J. L. Powell and B. Crasemann, Quantum Mechanics, Addison-Wesley Publ. Co., Inc.,
Reading, Mass. (1961), p.475.
THE scHnciorNGER wAvE EeuATtoN t4t
PROBLEM 4.'7
PROBLEM 1.18
,P(x) : 1Yt-k-agtzl2Kz.
PROBLEM 4.'9
"o --Pi
2*' (4.3+)
,h :
h2k2
,2* (4.35)
We have seen that such a free particle can be represented by an infinite plane
wave of the form
Y(r, t) : I o'r"-"' (4'36)
iZn
Note that a single differentiation of Equation 4.36 with respect to time gives
AV
(4.37)
;: -irouY,
a:Y : (4.38)
0x2 -k2Y.
From Equations 4.35, +.37, and 4.38 we see that we can write,
..
t6-
o'e
av h2 azY
(4.3e)
0t 2m 0x2'
., a ,-^,_ t)
., h, n, (+.41)
ih t),
i,Y(r, fiYzY(r,
,
ls E. Schrddinger, Ann. Ph1s.79r 361 (1926).
THE CONSERVATION OF PROBABILITY DENSTTY I43
a2 a2 a2
Trr*W--Azz'
If
the particle is not free but is acted upon by u force which can be ex-
pressed as the derivative of a scalar potential function, we merely add this
potential to the right hand side of Equation 4.3+. Thus,
E-f**''
and
ah.:H * r.
ih!Y(x. t\ : (
-Y I * v\v1r, t1 : trv(x, t).
dt^\^r"/-\ 2mOx2,'l'
The operator, ,ff : *
V, is called the Hamiltonian operator
-(h2l2m)(02l0xz)
since, analogous to the classical Hamiltonian function, it is the sum of the
kinetic and potential energy operators. In three dimensions the time-dependent
Schrodinger equation is,
where
.r -- - nv2 + v(i, t). (4.+3)
Equation 4.42 is a far more useful form than Equation 4.39 since all of the
important physical problems do contain forces. In fact, one can almost go so
far as to say that all of the interesting physics is contained in the potential term.
where Y(f, r) is the normalizedwave function associated with the particle. Irlow
suppose that as time progresses, p decreases for one volume element in space.
In order to be consistent with the fact that the total probability density over
all of space must be unity, we must then have p increasing in some other volume
144 THE DEYELOPMENT OF WAVE MECHANICS
+, Lp
dr :JF ],v. + Y* *,*) o"
Now, Y satisfies Equation 4.34,
trY : ih.!,v,
while V* satisfies the conjugate wave equation,
ffV* : _ih.
,Lrv..
The reason for the fact that the conjugate wave functions require a different
wave equation may be argued as follows. To claim that Y and V* satisfy the
same equation is to say that V : Y*.
To go a step further, if Y represents momentum transfer to the right, then
Y* transfers momentum to the left; Y:r€ - Y implies no momentum, that is,
no dynamical state. Then,
ry:
dt
!**: hffiv2+r)":*wY-in*,
th,
where the latter merely interchanges the roles of the functions f and g. Sub-
tracting,
v ' ["f(vg) - e(v/)] : "f (v'd - s(vy).
Letting/-Yandg -Y*,
v.[Y(vY*) - Y*(VY)l : Y(V2Y*) - \rt (VrV).
Then,
:
* [,0 a, I, - *v . ;v(vY*) - Y*(vY)] d".
Now we define the probabilit2 densitlt curcent,
Then
ff*v'.f:0. (4.+6)
Y(r, t):
' +ri&sa-a)tt.
{2,
Then the probability density current ill the x-direction associated with the
motion of this particle is
PROBLEM 4.20
where a and b are constants. The sum and the product of two linear operators
also form linear operators. Although the sum of two operators is commutative,
the product is, in general, not commutative.
We say that two quantities a and D commute under multiplication if
ab : ba; that is, if the order of multiplication is immaterial. This property
is taken for granted in ordinary algebra. However, you. have already seen in
vector algebia that the vector product Z x E + E x i, although the scalar
product n . E : E .i. Thus, the dot product is commutative, whereas the
cross product is not.
In operator algebra we must not assume commutivity unless it is proven
specifically for the operators in question. Consider the product of the operators
x and dldx operating on a function of x:
-
fr tra : xf '(x).
But,
!* rt f-l : -f (r) + xf ' (x).
Subtracting,
Therefore, the equivalent value for the bracketed quantity is - l. This bracket
is called the "commutator bracket" of r and df dx and is usually written in the
abbreviated form,
: -1.
[",
Evidently,
ld ,r] : *1.
ldr
Therefore, two operators commute if tlteir commutator bracket is zero.
Let us now consider a measurement of a position coordinate of a particle.
The probability of finding the particle in the volume element dr is, from
Equation +-lB,
z' t)Y-(r'''v' z' t)
!"-
p(x,), z, t) dr - j,V*
-Y!-|.*-'l'
(r,), z, t)Y(x,1t, z, t) dr'
.
The fact that p dr is called a probabilit2 implies that one would not necessarily
expect to obtain the same result for two successive measurements on the system.
We can define an aaerage ualue, or expectation ualue, of the position variable
by considering either a large number of measurements on the same system, or a
single measurement on each of a large number of identical systems. Thus, the
expectation value of, say, the variable x, is the weighted average,
(4.48)
It is also possible to obtain the expectation value (r) by means of the momentum
wave functions, provided that the appropriate operator is used. In order to
find the correct operator expression, let us insert Equation 4-26, the Fourier
transforms of Y* and Y, into Equation 4-+8, where we assume that Y is
normalized. Then,
(x) :
*fidxdkdk' 6* 1k)e-tk, xeik'u 6&')
xeik'' 6&).
+, fu I d.k 6* &),-o*.
Idk'
Performing a partial integration over k', the last integral becomes
where the first term is zero since the wave function vanishes at both limits.
i
Then,
:
IIor
dk,6*&) t
* +&,) 6(k - k'),
where Equation 4-28 has been used. Performing the ,t' integration,.
(p?)*-_ hz*,
otc"
E<-+ih!.
ot
A few of the commonly used operators are summarized in Table 4.2 for
convenience. il
Before closing this section, it is worth noting that the concept of the
expectation value enables us to make a more precise statement of the uncertainty
principle. The difference between a given measurement ofx and the expectation
value of r is called the deuiation of x. We define the uncertainty of r as the root-
mean-square deviation of r. Thus,
Ar: {GW
: \/ (x, - 2l4tc> + xrD
: \/ <tP> \x><,c> + (xY
: \/ <,cz) - <xy. (4.51)
OBSERVABLES, OPERATORS, AND EXPECTATION YALUES t49
Coordinate Momentum
Variable Representation Representation
a
x ,n
,n
q t
-o' un
a
P, - ift --dx P*
az
pi -ftz pi
-ox'
a o
E th= oh
dt u,
By means of Equation 4.51, we are now able to define the uncertainty product
given in Equation 4.32 asla
PROBLEM 4.2'
la See A. Messiah, Quantum Mechanics. North-Holland Publishing Co., Amsterdam, 1958, Vol. l,
p. 133.
rd lbid., p. 135.
I5O THE DEYELOPMENT OF WAYE MECHANICS
PROBLEM 4-22
lr' P*], l*, !of , lr, !'1, I !o !,], I P,, P'f '
Do the results just obtained depend upon whether the
coordinate or the momentum wave functions are used in the
calculation ?
Substituting this form of the wave function into Equation +.+2, we obtain
A
ihy(i) : x(t) ' trrP(i),
*x@
since .tr is independent of the time. This may be written in the separated
form,
ih a-
,ttl :
*y!!)
- E, (4.52)
m' }'x\t) 'Pe)
where the symb ol E has been used for the separation constant, since we will
soon see that it is readily identified as the total energy of the system. The
equation involving the time can be immediately integrated, and its solution is
x@ : Ce-QthtEt'. (4.53)
The spatial equation deriving from Equation 4.52 may be written in the
form
"/frpe) : Erp(f) (+.5+)
In this case the eigenvalue is -4 and the eigenfunction is sin 2x. If there is
a whole set of functions fn, there is an eigenvalue of Q associated with each
function. Thus Qf* : cn-fn. In the above example
-fn : sin nx
and
Cn: -n2.
Not all functions are suitable as eigenfunctions; for some functions there is
no opetator such that an eigenvalue exists. We say that an eigenfunction must
be 'iwell-behaved"; in particular, it must be single-valued ind the integral
of its modulus squared must be finite. In general, functions which are contiru-
ous, single-valued, and which are finite or vanish at infinity are suitable. It is
evident that the suitability of a function depends somewhat on the nature of
the operator, but generally the restrictions on eigenfunctions are the same as
those that we require for wave functions which describe particles.
One of the main activities of quantum mechanics is that of solving an
eigenvalue problem when only Q is known. That is, we want solutions of
Q.f : tcl where the problem is to find both c and f. As an example, consider
again the operator dzf dxz:
dT
_-LfT
,f
dxz - -L"J'
For positive c this has the solution-f : Aeri" 1 3s-{",. This function diverges
-> * oo, so it is not a suitable eigenfunction. However, if A : 0 and JtL :
as ,{
&s-\/i*, the function;{, is suitable for x > 0 and has the eigenvalue c. For x 1 0,
t52 THE DEYELOPMENT OF WAVE MECHANICS
P : Y* (i l)Y(1, r) : ''ph,(i)'pE,(i)e-uth)(En-En*It
'
+ V . ^S(i t) :
?
-h (8, - E|)rpfr^(i)rpr^(f) 0.
where the latter is a surface integral. For a sufficiently large bounding surface,
the probability flow through it must vanish, so we find that En : El , or En
is real. This important result tells us that when space and time are separable,
the time factor, Equation +.53, is a harmonic function and not an exponentially
rising or decaying function. Y(i, t) and yp*(i) differ, then, only by u time-
dependent phase factor. An immediate consequence of the reality of En is
that the probability density is independent of time, since
(E ) :,[*- Q, t)ih
*,*rr, t) dr
: i n)
e( t
E't
ih (- E -) v r,Q1 r-( t t h) E
^t dt r
!,vh ^tr) ;
: E,l,rh,(i)yp^(i) dr
: E*t,p d'
-tr
-
un.
.trrpr,(i) : Enrpn^(f),
d,.
t,vlv.
We will now also represent this integral by the symbol
where the form on the right implies that the operator is Hermitian. Since the
Hermitian property will not be discussed until Chapter 6, _we will restrict
ourselves to the form on the left for the time being. By way of illustration, the
expectation values given by Equations 4.48 and 4.49 now take the form:
(y rY)
(r):ffi and (!,):ffi
I (d I
p,+>
Although we will use Dirac's notation freely in the next chapter, its deeper
significince will not become evident until we take up Chapter 6.
SU/YIMARY :iri+:ii;::i':
The old quantum theory is briefly described and illustrated by the Sommer-
feld treatment of the hydrogen atom. Its demise was brought about by
Schrijdinger's wave mechanics, which is developed in -the rernainder of the
-Aft.t
chapter. a discussion of the de Broglie wavelength of a particle andparticle
didaction, a complex wave function is defined. Although the wave function
SUMMARY I55
length I, : h Tfloc
Compton wavelength divide d, by 2rr
h
time -+
ffioc' reciprocal of the de Broglie frequency
associated with electron rest mass
energy
In terms of the natural units and the coupling constant a, the relationships
that were previously found for the hydrogen atom may be expressed as follows:
ao:
hz 18,
First Bohr orbit:
ffloaz q.
-:-
ntoo4
Ground state energy: -Ws
:-- luzmocz
2h2
Thus, when expressed in natural units, we may regard the hydrogen atom
as being rather weakly bound, which is to say that thi electromag.r.ii. inter-
action is a weak interaction in the world of elementary particles. For instance,
suppose that the coupling constant o( were nearly unity- Then the Bohr radius
would be nearly the Compton wavelength, and the ground state of hydrogen
would be of the same order of magnituJe as the ,.ri-us energy of the elec-
tron. The density of matter then would be greater by the factor (137)3.
It is evident from the form of the orbital velocity that uz is the appropriate
correction for relativistic effects, which are always proportional to (iic1z.'
Some form of simplified units is always used in advanced courses,'although
it may not be the system described here. A frequent practice in quantulm
electrodynamics is to define h, : c : tnt: 1, in ordir to simplify all equations.
When the answer is obtained it is then multiplied by the-appropriite com-
bination of these universal constants which gives it the .otr.cf dimensions.
In a non-relativistic treatment of atomic systems, the "natural" units
turn out to be the so-called atomic units wherein the unit of length is the
Bohr radius ao, the unit of mass is mo, and energies are expressed in teims of ws.
SUGGESTED
REFERENCES
I. STEP POTENTIALS
We will begin with potentials that are constant in time and that are also
constant throughout prescribed regions of space. The simplest of these is the
step potential shown in Figure 5-1, in which the one dimensional space is
divided into two regions such thatV:0 for x 10,andV - Vofor x > 0.
Let us first consider the case when the total energy E of a particle of mass
m is greater than Zo. In the region where x I 0, which we will designate as
region I, the potential is zero and the time-independent Schrodinger equation
becomes
fftvt Yry:
2m dxz
Evr.
(a) (b)
t
Drift tubes II
x: 0 x --->
Figure 5-1. (a) A plot of energy vs position for a step potential. The dotted lines indicate
possible particle energies. (b) A physical approximation of a step potential. The dotted line indicates
a particle trajectory.
t58
STEP POTENTIALS I59
For a free particle in region I, the total energy of the particle is equal to its
kinetic energy, that is,
!? _,
2m-"'
or
!?:hzk?:2mE'
Thus, the Schrodinger equation for region I may be written as
drrl,,
-k?rp',
-:dxz
which has the general solution,
d'rp' : ,
k'rrltrr,
dxz -
and
grt:Qsih2n lDe-ikr, (5.2)
Thus far we have said nothing about the direction of the incident particle. It
turns out that the continuous states for free particles are doubQ degenerate, that is,
there are two solutions for each value of the total energy. These solutions
correspond physically to the fact that a particle of energy E may be traveling
in either direction, *x. For convenience we will adopt the initial condition
that the particle arrives from the left. Therefore, the coefficient A in Equation
5.1 represents the amplitude of the wave associated with the incident particle,
.B is the amplitude of a reflected wave at the potential discontinuity at x : 0,
C is the amplitude of the transmitted wave in region II, and D is the amplitude
of an incident wave from the right. Our initial condition requires that D : 0.
Therefore,
Vtt : Qtihsa (5.3)
Equations 5.1 and 5.3 will together constitute a satisfactory solution of the
problem only if they fulfill all of the requirements of good behavior which
quantum mechanics imposes upon allowed wave functions. It was shown in an
earlier section that plane wave solutions of this type can be normalized so as to
satisfy the integral-square criterion. Although we will not normalize Equation
5.1 it is important to know that it canbe normalized. Two other requirements
are the continuity of the wave function itself and the existence of the first
derivative everywhere. Both of these conditions will be met if we require
91 and ?rr to be equal in magnitude and to have the same slope where they
160 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
-^i:ll,:;;13,":,) (5 4)
A+B:C,
and from the second we obtain
c:\(A-
kr'
B\.
t
V:G4)
Using the expression for the probability current in the previous chapter, we
have
sro: bvr
m
S"": -bWt
m
(5.6)
hk'
sr. : m
pp
(5.i)
x:0
STEP POTENTIALS 16I
illustrates schematically the behavior of solutions for the step potential. There
will always be a reflected wave unless the discontinuity in the potential vanishes.
Furthermore, Equation 5.7 indicates that the reflection coefficient is independent
of whether the potential discontinuity is a step up or a step down. Here the
behavior is quite different from that of a classical particle.
PROBTE/I4 5-,
Now consider the case for a total particle energy less than Zo. Classically,
the particle would rebound from the step, but we will find that in quantum
mechanics there is a non-zero probability of finding the particle in region II.
The analysis of region I is identical with the first case and results in the solution
given by Equation 5.1. However, in region II we are now faced with the fact
that the momentum is imaginary since the total energy is less than the potential
energy. That is,
Kz : ikz
B kL - iK}
v k, a iK,
(5.e)
C 2kt
:-
v k, I iKr'
and the result that
R:1 and T:0. (5. ro)
Here the reflection coefficient of unity, which agrees with the classical result,
arises from the assumption that no absorption occurs in region II. Thus, the
152 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
PROBLEM 5.2
Consider the barrier shown in Figure 5-3, having height Vo and width
2a. Here there are three regions to be considered, but since the potential is
zero for x <
-o and for x ) a, these two regions have identical solutions.
For the case of E > Zo, following the same procedure as in the previous
section, we obtain the solutions:
Vttt:Feik", fotxla,
where hk.: t/Z*n and hkz: \/Wn.
t
"l
E<V.,
* This is confirmed experimentally by the coherent regeneration of short lifetime, neutral K mesons
inabsorbing material. See for example, Gunnar Kiillen, Elementar2 Particle Physics, Addison-Wesley
Publishing Co., Reading, Mass., 1964, p. 438.
THE FINITE POTENTIAL BARRIER t63
rl/
T
l
I
Applying the boundary conditions (Equation 5.4) to ,p, and rp11 at )c -- -a and
to tp' and rpr' at x : q., orte can calculate all of the coefficients in terms of the
coefficient A. Then, from the probability currents the transmission and
reflection coefficients are determined to be
T:
l*l:l,.l'
l?:
l*t:l#l'
For the case of E > Vo the transmission coefficient is an oscillating function
as shown in Figure 5-4. An eigenfunction for this case is shown in Figure 5-5.
Its functional form is given by
I
T:
cosz (2k,a) * i e,.Elrr"' (2k,a)
' - L +dLTJsi*(2k,a)
T-
r-T (5.11)
T_
cosh2 (2Kra) *i(bn,- ft)'i'r" (2Kra)
of,
T- (5.12)
r- W
4E(vo Eisinh2
(2K'a)
-
It is of interest to evaluate the transmission coefficient for the case of f
j
*l
T
I t'i
T:
D-7s ;1,kLay'
2m(vo-E)
'*#ulT
The same result is obtained by letting E ---> Vo in Equation 5.11.
Nature provides many examples of barrier penetration, although the
barriers in the real world are not rectangular in shape and they are generally
three dimensional. In spite of this, however, the idealized treatment given here
forms the basis for understanding all of the more sophisticated tunneling proc-
esses and frequently provides a reasonably good order-of-magnitude answer.
For example, alpha particle emission from heavy nuclei was first explainedl by
regarding it as a tunneling process through the nuclear barrier. Figure 5-6
shows a crude representation of a nuclear well which is about 40 MeV deep
for a heavy nucleus. The Coulomb barrier is roughly 16 MeV high in this
example and its breadth is of the order of 1 x 10-12 cm. A typical value for the
energy of an emitted alpha particle is 6 MeV, which is considerably below the
* At bombarding energies of about 0.7 eV, rare gas atoms are almost transparent to electrons.
This effect will be discussed in Chapter 10.
1 G. Gamow. Z. Playsik5lr204 (1928); R. W. Gurney and E. U. Condon, Ph1ts. Rea.33,127 (1929).
THE FINITE POTENTIAL BARRIER t55
E
(Mev) Coulomb repulsion
l6
-40
T--
1+ 4E(vovi Ei
sinh2 (2K'a)
-
: [1 + +*3 sinh2 (2Kra)]-1.
For 2Kra ) l, we may write,
T ,--, [sinh (2Kra)l-' ,-, lte'K'of- : l,s-4K2a
This means that an alpha particle incident on the barrier has about one chance
in 1025 of tunneling through the barrier. Although this is a very small proba-
bility, we will see that a 6 MeV alpha particle trapped in the above nucleus
bangs against the barrier about 1021 times per second (see Problem 5-15).
Thus a tunneling event can be expected to occur with a probability of 10-25 x
1021 : t6-e per second, or the decay time is about 3 hours. To illustrate how
t66 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
critically the mean decay time depends upon the strength of the barrier, suppose
that the barrier width were increased by only 20 percent, all other consiants
remaining the same. Then, T ,-' 10-30, and the mean decay time would be
,-,10e seconds or 100 years! Therefore, in spite of the crudeness of this model
it is able to accountfor tlte experimentalfact that the meun lifetirnes of atpha emitters aar)
-fro* 1 microsecond to l0ro Tears!
An elementary example of barrier penetration arises in optics in the case
of total internal reflection from a glass-air interface (see Figure 5-7). If a
second piece of glass is brought very near the reflecting surface, a weak trans-
mitted wave will appear in it. The intensity of the transmitted wave is strongly
dependent upon the thickness of the air barrier, and no transmission takei
place for barriers more than one wavelength thick.
There are numerous examples of the tunneling of electrons and quasi-
particles in solid state2 and molecular physics. However, there is little io be
gained by taking a detailed look at another example until we have developed
the mathematical techniques for treating more realistic potentials than the
rectangular potential. This will be done in Chapter 9.
PROBLEI4 5-3
Derive Equation 5. I l.
PROBTE/I4 5-5
2 C. B. Duke, ed,, Tunneling in Solids: Solid State Physics Supplement 10, Academic Press, New York,
1969.
THE SOUARE WELL IN ONE DIMENSION
PROBLEM 5.5
9r : AeK" , forx<-&
lprt : Ceikr' -f De-ikr*, fot -a < x < a
c :4 r,ik'a (5'I3)
D- '
c _ k, * i\ "_roo,o (5.r4)
D k2 - iKr-
-:-E
1
E
ttl
E>V,
AV
V\, Figure 5-9. Eigenfunctions for the square
E1
well potential.
/,\
E,]
\/iv
E1
,-l\..
E:O x:0
I
C : X.D.
tanu:JK,B
k2 -'- d.
ofr
q.2 +F':ry (5.17)
Thus, the symmetric solutions can be obtained by solving Equations 5.15 and
5.I7 graphically, and the antisymmetric solutions are obtained by solving
Equations 5.16 and 5.17. These solutions are sketched in Figure 5-10. Note
that the number of solutions depends upon the radius of the circle, Equation
5.17, which in turn depends upon the mass of the particle, the depth of the well
and the width of the well. The prodtct a% is a measure of the "strength" of
the well in terms of binding a particle. That is, the greater the depth and
breadth of a well, the greater the number of bound states and the greater the
probability of retaining a particle in the well. In particular, if
(rn),.ry<@*r),n,,
then there are n f I symmetric bound states. Similarly, if
then there are n antisymmetric bound states in the well. The total number
of bound solutions exceeds by one the largest integer containe din (alnh)\/rmv,
Note that there is always at least one bound symmetric state, regardless of how
shallow the well is.
In the limiting case of an infinitely deep well the energy eigenvalues are
readily obtained without performing a graphical solution. As Vo - oo, the
radius of the circle, Equation 5.17, becomes infinite and will intersect Equations
il
Irr2
2
I7O SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
.c io
'd-o\
uo
I-/ -::+-_ L_= t-
'tr
o
o $
o @ sfn
@
e(
!rt
o
(ltL
=- -t'- 5G'
IJ !L
crc
:i
_+
1
o o -o
o o @
+,
I o
*oo :h
'0€!
t VFt
_) t-
OJ
o
{
sl o
o
N
SvE
-.c
sO
(, fit
:o
et
o
cs
GtI
.e$
L
>\o
o o o bo -o
to !r
o
$
o
ts -,/ co
{rlJ
tu
,{
tr<
t--
9T
r( d
-o 6E
o{
o o
€ -c-
F
o o
@
1
cr=
'L- -
L>\
G'L
,d
-Od
o!
dlt
o' fit
6c
o an '-
o
@
c0
LF
o.g
&r,
bo c)
et
t-
q, (,
r_ 9+,
H9
s<
Hd
et
Q)6
o o JL
$ 9 (J(u
N si fit lt
o-g
O=
at1
__ r_ )
GI
:_=€_ ci
.; .90
24)
j
o
N
o
N
o
@
o
er!- b.E
'- t^
-dc
g.
_ _f - ?.e
L e-c
ioe':
o o
L i;
o @ o
N rl
THE SGIUARE WELL IN ONE DIMENSION t7t
,i
o-c
.9.c
_l
''
gT
!
-f
o o iri
q,
@l
ol
N
I io
T' O\
,!
En;
tr\
-1
I
cl-
*'l .tr tri
*
u __l +, tyi
o
a6
(ol
o
o
! Ef
cr. o-
9E
-t
--- €<
o
1 I
€i
ELr
-cl
o<===-
?.-/
o
N
F
rt
o T.:
>u
Qv,
frl
oc
cct
O1
a0)
--l -=
_-l
ll'-c
-t4 O
o
n o o
o
o o
o oo z-
O'J
c+
o
opo
ll ::_l -4- _l
F"i
{, T'
o o
o
o
o
38
G
o o o F<
=o
og
__J ...:I
---t '<
o
j Es
o
.r-- -=\
o
o
rf
o
o
N
df
a\
EL
o!n
r-LL .*,
@lr
CL
'tr Gt
I Oa
+, .=
..1
__==n
+,
8s
(,h
-1 6E
JO
q-c
o
rf
ol
{l o
N el @
(o f,+,
og
>o
Gt-
;6T'
c.
6l
--r .er
a3
cth
o o o (,r
o
N
(,
n
\i
@ z
1\l
.A t-
j E€?
3- o
-l .g0T't
E -.E
. --- o
@
o
N
o
€t
oL
t72 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
fn general, a : nnl2, where the lowest state is symmetric, the next antisym-
metric, and so on, alternately. Since u : k20.,
k2a:;,(ilh=7,
E^: nzrPhz
Bmaz
(5. rB)
This will be recognized as the answer to the problem of a particle in a box with
rigid walls as solved by the old quantum theory (Chapter 4, Problem 2).
Note that n : 0 does not correspond to a physical state since it would require
that Y : 0 everYwhere in the box.
The eigenfunctions for the infinite square well do not leak out of the walls
of the well as they do in the finite potential well (Fig. 5-9). Since each solution
must have zero amplitude at both walls, the eigenfunctions then become sine
or cosine functions.
Before closing this section it is important to point out that a rcal particle is
represented by a wave packet consisting of many * values instead of just a
single value of ,t as we have assumed in the foregoing analysis. Furthermore,
since the time dependence of each component differs from the others, a com-
plete analysis of the interaction of a packet with a barrier or a well is quite
complicated even in one dimension. Goldberg et al\ have obtained computer
solutions and computer-generated motion pictures of Gaussian packets
scattering from one dimensional barriers and wells. Two of their scattering
events are illustrated in Figures 5-11 and 5-12.
PROBLEM 5.7
PROBI.EM 5.8
Show that the energy of the lowest state in the infinite square
well is consistent with the uncertainty in momentum as
required by the uncertainty principle. Would the state
E : 0 satisfy the uncertainty principle ?
PROBLEM 5.9
(u) For the square well shown in Figure 5-8, find the
function tp11 defined in the region
-0. < x < a for the
caseofE>Vo.
(b) For what values of E > [ is the total probability with-
in the well a maximum ? A minimum ?
(") Is the maximum value of tL, p dx found in (b) greater
than what it would be in the absence of the well ?
Physically, this means that particles are "trapped" for
a short while in the well. Such events are called
resonances.
PROBLEM 5.IO
PROBLEM 5.12
Find the probability current density for the nth state of the
infinite square well.
PROBI.EM 5-'3
lnx 2.nx
1p: + srn
-a
{a
-COS- 2a \/a
-,-
(u) Normalize this wave function in the sPace -a < x I a.
(b) Sketch the probability density function, p(x).
(.) Include the appropriate time factor in each term of the
wave function and find the probability density function,
p(x, t).
(d) Using the normalizedwave function in (a), calculate the
expectation value of the kinetic energy.
(Ans. : (d) 17 Tznzl4\maz)
174 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
PROBLEM S.I4
PROBIE/I4 5-r5
If two identical square wells are isolated from each other and each has a
particle in its ground state, the energy of the system is just the sum of the two
ground state energies. Now, if the two wells are separated by only avery narrow
barrier they will interact and the total wave function is a linear combination
of the two single-well ground state functions as shown schematically in Figure
5-13. In (a) the total wave function is symmetric in the two wells, and in (b)
it is antisymmetric. The energies of these two states are not the same, and the
energy splitting increases as the width of the barrier decreases. This may be
understood qualitatively by means of the following argument. In the limit as
the barrier vanishes, the symmetric wave function corresponds to the ground
state of a well having twice the width of the original wells; hence the energy
decreases. As the wells coalesce, the antisymmetric function, on the other hand,
corresponds to the first excited state of the well of double width. Thus, the
energy decrease due to the increased width of the well is compensated by
excitation of the particle to the first state above the ground state. The energy
of the double-well system for these two cases is shown in Figure 5-I4, where r is
the width of the barrier.
The most important feature of this model is that the two wells, which have
the same energy levels at large separations, interact more strongly as they are
brought closer together, and the interaction splits each original level into two
separate energy levels.
An important generalization of this model was made by Kronig and
Pennet' in treating the electron energies in solids. Here the number of
(a) (b)
Figure 5-13. Total wave functions for two interacting square wetls, (a) symmetric case, (b)
antisymmetric case.
{ R. de L. Kronig and W. G. Penney, Proc. Ro1. Soc. (London) A l3O, 499 (1931).
MULTIPLE SOUARE WELLS t7s
t
E
Symmetric
interacting wells, l/, is extremely large so that each of the single-well levels is
split into y'/ Ievels spaced so close together that they form nearly continuous en-
ergy bands. Figure 5-15 shows a schematic square-well representation of the
potential seen by an electron due to a linear lattice of positive ions. Since the
potential function has the period of the lattice, that is, V(r + d) : V(x), we
expect our solution to have the same periodicity, provided that we avoid the
ends of the chain. This difficulty can be eliminated by imagining that the chain
is joined into a ring containing a large number of atoms (of the order of
Avogadro's number).
It was shown by Blochs that the solution for such a periodic lattice may be
written in the form of the product of a plane wave and a function having the
periodicity of the lattice. Thus,
,p(x) : eil'*u(x),
,P(x + d) : eo*rP(x).
The last result enables us to obtain the solution anywhere in the lattice once
it is known for a single well. If we denote the width of the well by a, the width
of the barrier by D and the height of the barrier by Vo, then we may define the
propagation constants,
6 F. Bloch, Z. Physik 52' 555 (1928). This was first proven by Floquet in 1883.
175 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
andforElVo,
+ 'tit srn
' h'a'
- (? ' stn k'b : 2nn
cos kra. cos krb cos
\ffi) ..n/ '
Sincg the right hand side of each of these equations is restricted to the range
of values between I and - l, solutions exist only for those energies for which the
magnitude of the left hand side does not exceed unity. As a consequence of
this restriction, the energieswhich correspond to "physically" realizableiolutions
form continuous bands which are separated by forbidden bands of energies for
which no solutions exist.
Although the details of the energy band structure of a solid require a
much more refined model than that used here, it is instructive to see that the
general quantum mechanical behavior of the electrons can be predicted by
just using the elementary concepts of wells and barriers devel,oped in the
previous sections.
ff:!" { $azmxz,
2m
Then the energy eigenstates of the oscillator will be given by the solutions of the
time-independent Schrodinger equation,
dy _dVdq _
dr-dqn- ^/:dV
"*dq
drrp: d 1&p1dqn4,
drrp
dr, aq\dr)n:
there results
:0,
# . (e - q')Y' (5'20)
, :;:
p2E
,h.
In order to find acceptable solutions of Equation 5.20, we will first con-
sider the form of the asymptotic solutions as q becomes infinite. Then it will be
assumed that a satisfactory solution of Equation 5.20 can be expressed as the
product of the asymptotic solution and a finite polynomial in q. That is,
V : Vo' H(q),
where H(q) is a power series in q which will later be terminated to the ap-
propriate polynomial. In order to arrive at the asymptotic solution, we note
that as 4 gets very large, Equation 5.20 may be approximated by
ry
all' - Qzqso
:0, (5.21)
d'rlto :
-
o
gzlso : t\o, where !)o frs-azlz'
#
I78 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
However, since go vanishes as Q * @, rpo satisfies both Equations 5.20 and 5.21
in the asymptotic limit. Our trial solution of Equation 5.20 may now be
written as
g- e-q'lrH(q),
H" (q)
- 2qH' (q) * (. - L)H(q) : 0, (5.22)
fn general,
(k + l)(r + 2)or*r* (. - I - 2k)a*: 0,
or
e - I -2k
ulc+z (5.23)
- - (ka116a27ax
Equation 5.23, which is called the recursion relation, shows how all of the co-
efficients can be determined from aoand a1. Thus, aoand arare the two arbitrary
constants required for the general solution of the second order differential
equation. The solution of Equation 5.22 may be written as the sum of two
series, one containing all odd powers of q and the other all even powers:
V - e-o't'H(q),
will diverge as q becomes infinite and
hence will be an unsatisfactory wave
function. This can be seen by examining the behavior of H(q) as & gets large.
From Equation 5.23,
at+z 2
o**k
Note that eq' : I + qz + *qn +. . . + 2kqzklk! + . . ., and the ratio of two
consecutive terms as ft gets large is
b**, 2
b* - k'
Thus, H(q) goes as e0', except for a constant factor, and
From this we conclude that the series solution of H(q) must be terminated to a pofu-
nomial in order to obtain a plrysicalfu acceptable wauefunction. The recursion relaiion,
Equation 5.23, tells us that whe n k : z, such that e : 2n * 1, one of the series
will terminate with a,,, since an*, and all higher coefficients will be zero. The
other series must be eliminated by setting ao equal to zero if z is odd, ot a1
equal to zero if z is even. As a consequence of the termination of the series we
obtain the energy eigenvalues from the condition
2E
::rh:2n * l' (5.25)
'
that is,
E : ah(n * il.
We may now label the energy states and the wave functions by means of the
index z which indicates the degree of the polynomial appearing in the solution.
That is,
,p"(q) : e-aztz17n7o1 and En : afi(n + t).
Note that the energy levels are equally spaced at intervals of o,h, as in the old
quantum theory. (See Figure 5-16.) However, the ground state is not zero, as
in the older theory, but is
Eo : $ah.
This energy is the so-called zero-point energy, which accounts for the fact that
7 P. M. Morse and H. Feshbach, Methods of Theoretical Physics. McGraw-Hill Book Co., Inc., New
York, 1953, Part II, p. 1640.
t80 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
v: j''mx'
E,:1dh
Figure 5-16. Energy levels of
*--:--r,.:f;,,h the harmonic oscillator.
Yr:|rn
+----
+rA
"r:
As 7 -t -2 2
d1 2.3
Then, from Equation 5.24,
Hr(q) : at(Q - &q").
H,(q) - 2q
Hr(q):4qz -2
Hr(q)-Bq'-l2q
Hn(il:16q4 -4$qz +12.
Note that H*(q) is odd or even as z is odd or even. In quantum mechanics the
term parit2 is used to describe "even-ness" or "odd-ness." That is, ^F/"(g) has
METHODS OF GENERATING THE HERMITE POLYNOMIALS I8I
even parity if z is even, and odd parity if z is odd. Since e-qztz is an even
function, we note that y*(q) has the same parity as H*(q) and n.
In order to obtain two additional prescriptions for constructing the
Hermite polynomials, let us first define the following function of two variables,
F(s, q) - exp lqn - (s - q)rf : exp [s(2q - r)] :|n,rq) .t;. (5.26)
The last expression is merely aTaylor expansion ofI (t, q) about s : 0. Thus,
the A"(q) are given by
A*(q)-eo,#(e_t,_ol,;]s:0(-1)",",#-(,_.,_o,,)],:o.
Ot,
A*(q): (-t)nraz{Ur". (5.27)
To show that the A"(q) are actually the Hermite polynomials, H,(q), we
proceed as follows: Taking the partial derivatives of Equation 5.26,
AF
(5.28)
*:2(q-s)F
!
dq
:2,, (5.2e)
H:2F
oq os
+2r{. os
(5.30)
a! _0F _r,n
a, - Tq -'v'',
and differentiating this sum with respect to q,
a2F
+2q{q (5.31)
@.Yr':2F
Eliminating the mixed derivatives from Equations 5.30 and 5.31,
AzF AF AF
(5.32)
@-'ouo*2s ar:u'
182 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
#:2,o-r, # (5.33)
? P'A"
aq' - ,o* * 2n A-) ''i : r.
"?u\
In order for this to hold for all J we must have
Att _
1Ln 2qA**2nAn:0,
where the primes signify differentiation with respect to the argument q. But
the latter expression is identical with the Hermite equation, Equation 5.22,
with e replaced by its proper value of 2n + 1. We therefore conclude that the
An are the Hermite polynomials. Replacin g A*with Hn, Equations 5.26 and 5.27
then provide us with the following two new definitions of the Hermite poly-
nomials:
The first of these, Equation 5.3+, says that if the generating function is
expanded in a power series in s, the coefficient of s" is the Hermite polynomial
of order z divided by nl. The second, Equation 5.35, enables the calculation of
the polynomial of any order by means of a sequence of differentiations.
PROBLEM 5.T6
Some useful identities may be easily derived which can provide a ready
means of obtaining Hermite polynomials of other orders if H"(q) is known. By
equating Equations 5.28 and 5.33 it follows that:
Of,
@
sn-' @ cn*L
2 H"(q) . (z 1)! -g
m:o
-
1 2 H"(q)
' n:o ;l -2q 2r"al ''n :0.
THE HARMONIC OSCILLATOR WAYE FUNCTIONS I83
hlr"*k) *, .
_ rqH"Jd;lql r, :o
.
ffi
fn order for this to hold for arbitrary s, we must have
In like manner, by equating the expressions for 0Fl0q from Equations 5.26 and
5.29, it follows that:
,P"(q) : Nne-o'/'H,(q).
Then we require that
t*-lr,{o)l'dq
: t.
I84 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
At this point we will use the Dirac notation for such integrals over all of space
(see section 13 of Chapter 4) :
(,p,(q) f*
l,p-(q)): J-a ,p\(q),p^(q) dq,
where it is understood that the complex conjugate of the quantity to the left of
the vertical bar is used in performing the integration. Then we may write the
normalization integral as
where it should be noted that the factor e-q' is understood in the last integral.
The condition n' : n may be imposed at alater time. The integral may be
evaluated in two ways. First, Equation 5.35 may be integrated by parts z
times, assuming n' > n, to obtain
(H*(q) I H*(q)) : (
- t)n'-n2n'n, ,l**ro(q) d.q: 0,
ffi e-o'
since e-q' and all of its derivatives vanishs for infinite q. For n' : flt the integral
becomes
Therefore, we note that the Hermite polynomials are ortltogonal with respect
to the weighting factor r-c', since the integral vanishes for n' * n. For n'-: n
the normalization factor is determined by the result,
lN,l'2"nlt/n : l.
The normalized wave functions for the linear harmonic oscillator are then
F (q,
') in!
G(q, r')
# n'!
8 R. Courant and D. Hilbert, Methods of Mathematical Phlsics. fnterscience Publishers, fnc.,
New
York, 1953, Vol. I, p.92.
THE HARMONIC OSCILLATOR WAVE FUNCTIONS I85
Then,
f*
,-u'du : t/n.
J--
Thus,
_r
:t/rlt*2ss'-T- , (2ss')" I
,(2ss,)r,
- ol -r"'l'
:G1ry
It is immediately obvious that there can be no solution unless n' - n, since the
two series expansions can be equal only when this condition is met. For n' : n
we have
Therefore,
(H^(q) I U"(q)) : 2"t/n n!, (5.3e)
:
*(,p*(q)lqy,*(q)),
186 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
where the wave functions in Equation 5.40 are used for the integral on the left
and those in Equation 5.38 are used for the integral on the right. Using the
latter and the definitions in Equation 5.26, *. *tit.:
:
Z,;. '# @-(q) I qH-,(q))
(,p,(q) I qy,*,(q))
: t/n(t I s')ezu'.
Then,
A,lrr"(q)lqH-,k)>;
';l: rn[('. 2szs' *T + "'
*ry + ) * (,, *2ss,z +T +.. .*r"'li'" . )]
By comparing series we see that the first series of the right member can be
equated to the left member only when nt - n
- I; the second series on the
right can be equated to the left only when n' - n -f l; for n' : n the integral
must vanish. The results are summarized below:
EXAMPLE
Calculate the expectation values (x), (!,), <T), <V) and (E) for the linear
harmonic oscillator. Utilizing the identities 5.36 and 5.37, we have
Substituting the values for (HnlqH**r) and, (HnlqH*-) from Equation 5.41,
we obtain
and
,po(q): (+)+r-oz1z
wnl
,p,(q) : (#Joe-qz,
,p,(q): (#f rrr -r)e-azrz
En:ath,(n++),
where 6sz - klp,, the force constant divided by the reduced mass of the two
atoms. A better fit to the real potential is obtained by using the so-called
Morse potential,
V(r) : VolL - r-a(o-totlz - Vo.
12
En:afi(n++)
+ +) -
-14 tzfn@ * +)J
The effect of the second term is to depress the levels as z increases. Hence, the
interval between successive levels diminishes instead of remaining constant as
in the harmonic oscillator. This is shown schematically in Figure 5-19.
Another model that sometimes proves useful in molecular problems is the
double oscillator with a barrier between the two oscillator wells (see Figure
5-20). Aparticle having energy E <Vocan tunnel from one oscillator to the
o See L. Pauling and E. B. Wilson, Introductionto Qwntum Meclwnics, McGraw-Hill Book Co., 1935,
p.271.
t90 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
1
E
other at a rate determined by the height and breadth of the barrier and the
total energy of the particle. This model can be used to describe the oscillation
of the nitrogen atom between two stable configurations of the ammonia mole-
cule. The geometrical arrangement of this molecrrle is that of a tetrahedron
whose basal plane contains the three hydrogen atoms at the vertices of an
equilateral triangle. The nitrogen atom is at the fourth vertex, either above or
below the basal plane (see Figure 5-21). The oscillation of the nitrogen atom
between these two equilibrium positions is a tunneling process whose qualitative
features are similar to those of the double square well of section 4. A rigorous
solution of the double oscillator shows that the energy splitting of the two
lowest levels, for example, isro
L,E : +r, lL g-vol€o ,
V zeo
10 Merzbacher, Quanturn Meclunics, John Wiley and Sons, New York, 1961, p. 64 cf.
THE HARMONIC OSCILLATOR: OPERATOR METHoD r9t
where Vo ) €0 : $ah, the ground state energy of either single oscillator. This
corresponds to a tunneling frequency of LElh. The ratio of the tunneling time
z to the period of either oscillator, ?", it
q: {nx'
p:+^:
h{ a'
* _+! : _+1jv: _io
{i a* t/i ar - aq Oq'
!': -y'
oq''
In terms of these operators, Equation 5.20 becomes
Since p and q are operators associated with r and y', (see Problem 4-22), we do
not expect them to commute. In particular,
If the commutator in 5.44 were zero, Equation 5.43 could be factored into
(q + ip)(q
- ip)tp : €v. However, because of the non-commutivity of p and,
q we can write:
(q+iP)(q-ip):q2+p2+r
(q - ip)(q + ip) : {rz + p, _ t.
Adding,
p' + q2 : tl? +;p)(q - ip) + (q - ip)(t +;p11.
If we define two new operators,
':+,(r +;p1
':aQ -i!),
a'1
then
P'+q':aaIaaIa,
and the Schrodinger equation, Equation 5.20, may be written as
Before actually solving the Schrodinger equation, we will show that we can
construct an infinite set of eigenfunctions if we can somehow find one solution
of any of Equations 5.45, 5.46, or 5.47. Operating on Equation 5.46 with at
from the left, we have
oscillator to the next higher energy state, at is called the raising operator. In like
manner we will show that a is a lowering operator for the same set of functions
such that the following relations will hold:
el el
-r-
2'2 ,- 2
€
t e3 el
aty) -r- -r- e*2
2'2 2'2
e5 e3
(oI)'rp
2'2
-_L- -r-
2'2 ei4
e 2n*l e 2n-l
(ot)"rp
2'
-_L-
2 2'
-l_-
2
et2n
el e3
e -2
,-, ,-,
e3 e5
e -4
a2yt
,-, ,-,
€ 2n'-l € 2n'+l
a''ll
,- ,- , -2n'
THE HARMONIC OSCILLATOR: OPERATOR METHOD t9S
lowering operator can be applied only a finite number of times because the
eigenvalue, el2
- (2o' + \12, cannot become negative. The physical reason
for this is that the energy states of the harmonic oscillator must always corre-
spond to positive energies. This, of course, was a well-known classical result
and could be accepted in quantum mechanics on the grounds of the corre-
spondence principle. However, we can readily show its validity in quantum
mechanics by examining the expectation value of the energy. Thus,
since there can be no lower state. Then, from Equation 5.47, we have
olorpo:0:(;-!r)r,
E-r- ,
2Eo
rh
or
Eo : trh,
the ground state energy of the oscillator.
We will now see how easily the ground state wave function is obtained by
this method. Inserting the expiicit foim of the lowering operator,
":#'(o* &)'
into Equation 5.51, we obtain the first order differential equation,
W
* rro:0.
196 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
This may be integrated immediately to obtain the ground state wave function,
111o: Coe-a'lz,
1P
z : r, : ?,
(or)r,t, o
?o,
vn:%^{or)"r0. (5.52)
It will be shown below that the normalization constant C,,, when divided by
(2")L, becomes the same factor as that indicated by I{, in the discussion pre-
ceding Equation 5.38. It should be pointed out that Equation 5.52 provides
still another method of obtaining the Hermite polynomials, since
H*(q) - saz/z(,
- &l
r*,,. (5.53)
Recall from the definitions of a and aI that a* : aI; that is, the operators
are complex conjugates. They are, in fact, Hermitian conjugates or adjoints of
one another. The full meaning of this property will not be evident until one
uses matrix representations, but its usefulness here lies in the fact that it per-
mits us "to move an operator past the vertical bar" as follows:12
Its validity can be readily deduced from the commutation relation, that is,
'(")=tr",{-M"i;f
;'.a'I6(6r)n-'l
: n(ot)"-l + (oI)"o .
o (or)n!t o : n (oI)n-'rl, o,
since ago :0. Then,
a'(aI)"rl,o: na(aI)"-'rpo :
- l) (ot)
n(n n-rrpo
kt*l !,*) :
*f"!(?o I vo) : *f^Pol't/n'
Therefore,
cn:(#J
and
Nn:(#J
The general form for the wave function for the nth state is then
't^k)
: (;1 e-a'zrz17n7o1
SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
I-4
V. ,/'
-5 -4 -3 -2 -l 0 I 2 3 4 5
(t ___+ (t: 1/ax)
Figure 5-22. Probability densities for a few harmonic oscillator states. The dotted curves show
the cf assical probabilities corresponding to thesame energies. (From Chalmers W. Sherwin, lntro-
duction to Quontum Mechanics, Holt, Rinehart and Winston, lnc., 1959. Used with permission.)
or
(5.56)
where
element dx is
p(x) dx :
,1/ trz - *z'
where ,4 is the amplitude of the motion.
(b) Show that the classical amplitudes corresponding to
the quantized oscillator energies are given by
A'n:(2n * r)
,
PROBLEM 5.21
Y(*, t) : 1
t) I Vr(r, t)1.
,rlYo@,
Find the expectation value of the energy.
li11!li,iii.Ii!ii-{
2OO SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
Ipn+L :T orr_,
or
aI,pn:
ff*,*,: (g+!'f ,-*,: \/n + r sn*I. (5.57)
aen:?oo(o')"r0.
cn
: nfr(ol)"-"1'o' (since avo : o)'
q:+r@+,ll
(5.5e)
e:hr,-^.J
Now,
(a) : (V*l atp,) : t/i(rp*l y,*-r) : 0
(oI) : (rp^l aIrp,, : 1, a l9,*l rp**r) : 0
* (ot)) : o
:72(@) 1
(q)
(!):m((at)-(a)) :o
(q') : *((aa) + (aaI) * (ata) -t @IaI))
: g(2n + l)
:n*t
(!2) :f f (aIaI: - @ta) - (aaI) * (aa))
:n*t'
Therefore,
(x): I<q>:o
!x
(!n):nt/i1py:O
(xz) : !r,, :la + +)
: Ir :2ah(n +
(T) +)
2*<pi>
ah ah
(E):Vk):i(pr> +(qr)) :toh.(n ++).
202 SOLUTIONS OF SOME ONE DIMENSIONAL SYSTEMS
The reader should compare these results with those obtained in section 7 using
the algebra of the Hermite polynomials.
PROBLEM 5.25
ff:q2+f':aaIaata,
then the algebraic results of this section hold, regardless of the exact nature of
these variables. An important application of this method is that of the creation-
annihilation operator representation of field theory. In that formulation aI
creates a particle state, a annihilates a state, and aI a counts or numbers the
states. The latter is called the number operator and its function is readily seen
as follows: otorlrn: arlirp*-r,: {n. \/;!)n: frIpn. Therefore, the eigen-
value of the operator at a is the integer which numbers the state of the eigen-
function upon which it operates. In physical problems involving the
interactions of several different kinds of particles, it is customary to define a set of
creation-annihilation operators for the states of each class of particle. Thus, in
solids the many-body properties are conveniently treated as the collective
normal-mode excitations of systems of harmonic oscillators. The quanta of
excitation are called phonons for the set of oscillators associated with lattice
vibrations. In a magnetic system the normal modes are called spinwaves
and the quantized excitations are known as magnlns. In a physical problem
involving the interaction of electromagnetic waves with a magnetic solid, one
might require pairs of creation-annihilation operators for photons, phonons,
and magnons,
SUGGESTED
REFERENCES
E. Merzbacher, Quanturn Mechanics. John Wiley and Sons, fnc., New York,
1961.
D. Park, Introduction to the Quantum Theory. McGraw-Hill Book Co., Inc., New
York, 1964.
L. Pauling, and E. B. Wilson, Introduction to Qtnntum Meclwni.cs. McGraw-Hill
Book Co., fnc., New York, 1935.
J. L. Powell, and B. Crasemann, Quantum Mechanics. Addison-Wesley Pub-
lishing Co., fnc., Reading, Mass., 1961.
F. K. Richtmyer, E. H. Kennard, and J. N. Cooper, Introduction to Modern
Ph2sics,6th ed. McGraw-Hill Book Co., New York, 1969.
D. S. Saxon, Elementaryt Quantum Meclanics. Holden-Day, fnc., San Francisco,
1968.
L. I. Schifl Quanturn Meclanicsr 3rd ed. McGraw-Hill
Book Co. Inc., New
York, 1969.
R. L. White, Basic Qwnturn Mechanics. McGraw-Hill Book Co., fnc., New
York, 1966.
K. Ziock, Basic Quntum Mechanics. John Wiley and Sons, Inc., New York,
1969.
GHAPTER 6
Wffiffi ffiffiffiffiffiffi
wweffiwwffi ffiffiffiffi&ffiffiffiffi
Postulate I
t)12 dx d1t dz - A,
,[,*,.
where A is a finite number. Then a wave function which is normalized to
unity may be constructed by writing the new function as Y' : 0l\/7)YQ, t).
(See Problems 4-8, 4-9, 4-10, and 4-15.) A quantum mechanical state is
defined by a specific set of values for the independent variables, such as
V(ir, tr) or +(il. Although neither Y(ir, t1) nor +(Er) is directly observable,
the square of its amplitude for a particular state is proportional to the probability
that the particle will be observed in that state. We defined the probability
density in Equation 4.18 as
It follows that the probability for finding the particle somewltere in space, that
is, the total probability, is unity. Thus,
Definition
Two non-zero wave functions are said to be orthogonal if their scalar
product is zero. Thus, the orthogonality condition is,
YIV, dr : 0.
I,
THE FORMAL STRUCTURE OF EUANTUM MECHANICS
Wave functions that are both normalized and orthogonal satisfy the combined
orthonormality condition,
f
I YIY, dr : 6ii, (6.2)
Jr
Postulate 2
The superposition principle is ualid for functions representing plrysicall2 admissible
states,
\r : 2tnYn, (6.3)
i
where the co are the expansion coefficients and the Y, are the known discrete
states.
Postulate 3
Tlte Schrtidinger equation describes the beltauior of a waue function in space and time.
"ryve, t) : th
*vQ, t), (6.4)
where ,tr is the Hamiltonian operator for the system. In the particular case
of a time-independent potential, Equation 6.4 is separable in space and time,
and Y(i t) may be written as the product of a spatial function and a time
function. The spatial part is an energy eigenfunction; that is, it is one of the
solutions of the energy eigenequation, Equation +.57,
where the eigenvalue En is the energy of t}re nth state. Since the energy E^
remains constant in time, the solutions are known as stationary states. The
THE POSTULATES OF QUANTUM MECHANICS 2O7
YQ, t) (6.i)
The sufficient condition for Equation 6.7 to be a valid expansion for a general
wave function Y is that the functions yp. form a complete, orthonormal set of
functions. This matter will be discussed further in section 3 of this chapter.
Postulate 4
Each d2namical uariable q can be directl2 associated witlt a linear, Hermitian
operator Q. Tlte onQ possible result of a measurement of tlte obseruable q is one of tlte
eigenualues of the operator Q.
Definition
The operator Q is linear if it commutes with constants and if it obeys the
distributive law. Specifically, if
Qk,D : tQ(v),
and if
Q(p, + lpr) : Q(rt) + Q(rtr),
Definition
An operator is Hermitian if it satisfies the following equality,
fr
IJr QrD*rp d, : JrI ,p* (QrD d",
Since a and b are arbitrary, and their relative phase is arbitrary, the only
way Equation 6.9 can be satisfied for all possible values of a and D is for each
term to be identically zero. This immediately gives Equation 6.8 and the
more general statement of the Hermitian property,
(6.10)
@r l,Qrp,) -- (Q'p' 1,P,>.
A Hermitian operator is frequently called a self-adjoint operator. This terminol-
ogy arises from the fact that an equation similar to Equation 6.10 can be
written for any operator Q, provided that the Hermitian adjoint* of Q, denoted
by Qt, is used when it operates on the other function. Thus,
to indicate that the operator, being Hermitian, can operate equally well on
the function to the left or to the right.
Now let us consider a measurement of the dynamical variable g which is
represented by a linear, Hermitian operator A. The fourth postulate says
that the measured value must correspond to one of the eigenvalues of Q, but
it does not tell us which one. In order to see what is implied here, let us suppose
that we wish to measure the energy of a hypothetical state which can be
represented by Equation 6.7, where the only non-zero expansion coefficients
are cL and cr, which are real numbers. That is,
The same result can be obtained by operating on Y with the energy operator
ih(7l0t) :
th :
*ty(r, r) ;nprvE,e) fir-ou,,,u * c{ps"@
*,r-ou*,1
: c1E1Yy * csEsYs. (6.14)
Equations 6.13 and 6.14 are evidently not eigenequations, since we cannot
gxpless the right hand side as the product of a scalar times the original function.
In this example the only accessible eigenvalues of ,ff are E. and4, ro we con-
clude that a measurement of the energy will yield either the value E, or E",
and not some intermediate or average value. The probability of measuring
the value E, is til^Q? c!) (or just cl if Y is normalized) and the probability o?
measurin1 EriscilQi^++ cil. The act of measurement, then, forces the sysiem
into one of its eigenstates, where it will remain unless it is disturbed. Once a
system is known to be in an eigenstate, the result of a repeated measurement
can be predicted with certainty.*
_ The previous discussion assumed, of course, that no time-dependent
fo-t::| act_ upon the system. Should the Hamiltonian be time-depetdent, an
additional c_omplication arises, since the Schrodinger equation is thin no longer
separable. We will assume that in such a case the wave function can still-be
expanded as in Equation 6.7 if we but add the assumption that the expansion
coefficients, co, are,also time-dependent (this will be dbne in Chapter 9j. The
1ct 9f measuring the energy will again yield one of the energy eiginvalues, say
Er, brlt in contrast with the previous case, the system will remain in
the jtk state. The time dependence in the coefficients will "o'to"g.r
cause some coeffi-
cients to grow at the expense of others, or even to oscillate. Therefore, successive
measurements of the energy will not, in general, yield the same eigenvalue.
Theorem
Tlte eigenualues of a Hermitian operator are real.
Pnoor. Given, Qytn : Qnvn; and Q : QI , For convenience let (tpn l rp,) :
l. 'Atro,
.(rtt*l Qrp"> : (pnl q*rp*) : gn(lpnlrp,) :' 4,.
The expectation value is
lQrprlr*):.(qnpnlrp,) : ql@*lrp^): qi.
g"i the lefi rru"a rid.r ur..q,rui
by the Hermitian property, so we have
8L : 8n,
ot In is real.
Theorem
where the fact that the eigenvalue is real for a Hermitian operator has been
used. The left hand integrals of each line are equal by the Hermitian property,
so by subtraction we obtain
o:(q,-er)@nlrp,).
Since Qr * gr, we must have (rpnlrp) : 0.
Postulate 5
(Y I 0Y)
(q) : (6.15)
(q) :
(cifl, * crY, I Q(r'Yt * crYr)) :7a*
tlqt, * c?q"
'
Note that this number turns out to be a weighted average of the accessible
eigenvalues of Q, although a single measurement must always correspond to
rHE POSTULATES OF OUANTUM MECHANICS 2I I
(q):fu
2 lttl' qt
(6.16)
lcnl'
2v)''
It has already been established that eigenvalues must be real-since they
correspond to physical measurements-and we have seen that the use of
Hermitian operators guarantees that this requirement is met. It is evident
also from Equation 6.16 that the expectation value must be real, but this can
be shown formally as follows.
Corollary
Tlte necessar.lt and suficient conditionfor a real expectation ualue is that tlte d2namical
uariable be represented b2 a Hermitian operator.
Pnoor. Given, (8) : (8)*, that is, the expectation value of I is real.
Then, (rp I Qil : (t, I QrD* : (Qrp l rD, which demonstrates that a is
Hermitian.
PROBLEM 6.1
PROBLEM 6.2
Pv : ltrl
Qrp : r{P,
Since p, and qL are simply numbers, they commute and the order of the
measurements is immaterial. Physically, this means that a system that is in
simultaneous eigenstates of two operators will be undisturbed by any sequence
of measurements of the observibles associated with these two operators.
Mathematically, we see that for this to occur the operators must commute;
that is,
lQ,Pl-QP-PQ-0. (6.17)
The bracket on the left is defined as the commutator bracket or simply the com-
of I and P. (See section l I of Chapter 4.)
mutator
Corollary
. .If tyo operators P and Q commute and either P or Q has non-degenerate eigenualues,
its eigerfunctions are also eigerlfunctions of the other operator.
Pnoor. Given [P, 0] : 0 and Py,n : !{p* where all po are distinct.
Then,
QPrpn : Qf rtpr : fn(Qrpr).
QPrpn:PQVI:P(Q1D.
Equating the right members,
P(Qrpr) : pn(Qrp).
: #@l e,r) :
!, <o>
<#lr,).(, ln"+*(, lr{)
:i@,plQ,p
-1 ,. (#) + !h(,pt etr,p),
2/4 THE FORMAL STRUCTURE OF OUANTUM MECHANICS
d n\ I _ *\di
lp ,\. (6.18)
dt<a):;6<lQ,.rl)
This tells us that the expectation value will be a constant of the motion if the
operator has no explicit time-dependence, and if it commutes with the
Hamiltonian.
(u) o!,<n.>
<#>
(b) *#,a>: (!,).
The statement that the variables in the classical equations of
motion can be replaced by quantum mechanical expectation
values is known as Ehrenfest's theorem.l
PROBLEM 5.7
PROBLEM 6-9
!,vIrrdr
: (v,.lv) : (ltzl vr)* (6.1e)
In a real vector space (or Euclidean space) the scalar product is real and
symmetric in the variables. In a complex vector space (or unitary space) the
scalar product is complex and possesses symmetry in the variables under
complex conjugation. The latter is known as Hermitian s2mmetr2.
To illustrate the linearity of the scalar product, let us form the product of
rp1 and (rp, * ,pr):
(rptl rp, * vs) : (Ttl rpr> + (rprl rpt).
Definition
Two vectors are orthogonal rf their scalar product is zero, provided that
neither vector is the null vector.
216 THE FORMAL STRUCTURE OF EUANTUM MECHANICS
Definition
A set of functions is linearl2 independent if the linear equation,
2 tort,o
- 0,
is satisfie d onl2 when all the cu's are zerl. For example, if the Euclidean vector
at+bj+c[-is 0 only when a:b:c:0t then t, j, and ft are linearly
independent. fn other words, no one of these quantities can be written as a
Iinear combination of the others.
Theorem
The representation of oo arbitrary uector on a giuen basis is unique.
o:g ko-r|)Qu
Since the basis veetors {tn arerr"."r;;:ndependent, this result requires that cu :
c'0. Therefore, the representation is unique.
Theorem
The inner product of a uector with itself is positiue definite.
/N
(Y lY) : 2^/\
t,,0,) : > cfci(0,r1,0,> : l.cfc, 601
\4'n{'n i / d,i i,i
: >lcilz.
Therefore, (V lY) > 0, where the equal sign holds only for the null vector.
* The subject of completeness is discussed in R. Courant and D. Hilbert, op. cit., Vol. I, p. 369.
LINEAR VECTOR SPACES 217
Definition
The norm of a vector is defined by
normY-/(YTY):{ffir.
A vector with unit norm is said to be normaliz.O. ;", non-zero vector can be
normalized and the result is unique except for an arbitrary phase factor.
Theorem
Two uectoru V, andYo satisflt the Schwartz inequalit2,
Pnoor. Let Y : Yo * bYo. Then, from the previous theorem, the inner
product of V with itself is positive definite. That is,
and
txl . trt >x.i,
cos(,P, n:#r.
By analogy we may define the "angle" between Yo and Y6 as
cos(Yo,Yu):ffi
2t8 THE FORMAL STRUCTURE OF OUANTUM MECHANICS
Definition
An eigenvalue q is n-fotd degenerate if there are n linearly independent
eigenfuncti,ons correqponding to this eigenvalue. The n functions are called
degenerate eigenfunctions.
Theorem
Any linear combination of degenerate, linearly independent eigerfunctions is also an
eigenfunction ltauing the same eigenaalue.
Pnoor. Given the set of z functions y,rall having the same eigenvalue q
associated with the operator Q. That is, Qrpo: QlPrfor i : L,2,...,n.nThe
most general linear combination of these z degenerate functions is V : Ztntpn.
Then,
QV : : : I : QY,
|r,Orn \rnort |r,rn
which proves the theorem.
Corollary
In the case of n-fold degenerac2,linear combinations of the n degenerate eigenfunctions
i
ma2 be found so as ti form linearll-independent, orthogonal eigenfunctions corresponding
to the same eigenualue.
condition. Thus,
@rl'Pt) : (hB
and
@rl rPr) -- azs-
Now take rits : y)s
- aBQr - ezs$, and require that
(0r 0) :0,
@, 0) :0,
and
@, 'Q) : l.
Let us test the orthogonality of ,0" with ,$, and $r:
I
*: r/r(tPt+2tPr*2tP")
is also an eigenfunction of H corresponding to the eigenvalue
Eo.
5. LINEAR TRANSFORMATIONS
Ya : QYo,
may now be thought of as a linear transformation in the vector space, such that
the vector Yo is transformed into the vector Y, by the action of the linear
operator a. The eigenequation, Equation 4.56, thus appears as a special
case in which the direction ofYo is not changed under the action of the operator
and V6 is simply Yo multiplied by u scalar, that is,
qX": QYo.
Two important mathematical operators which must be incorporated in the
formalism are the null operator and the identit2 operator. Their properties are
immediately evident from the following:
OYo : 0,
lYo : 5rr.
Definition
An operator Q is non-singular if the equation 0Y : 0 is satisfied only by
the null vector.
The Hermitian conjugate of each ket is a bra uector which is written as (al.
The bras comprise a space having the same dimensionality as the ket space
and which is dual to it. Since there is a bra for every ket, and vice versa, there
is a one-to-one correspondence between these two spaces. Operators act upon
bra vectors from the right, so that the equivalent form for the eigenbra equation
is,
(al QI : Io (al, (6.23)
(ol o).
@lb)'
which is not the same as the product of la) by lD). In agreement with Equation
6. r9,
(olb) : (b lr)*.
The expectation value of the operator Q then becomes, using Equation
6.22,
(allQla) : Io@la): qo,
We see that the vector (Dl is transformed to the direction of (al and
that it is
multiplied by a scalar factor which is numerically equal to the projectio.,
(bl along (al. Now let this same operator act upon thl ket oi
lc;:
nr
V@ : 2 t,,Lo.
i:L
.nr
which says that the projection operator summed laer arut complete basis is the identit2
operator.
Yo : QVo
becomes
NNN
Yo : 2 f*,Oo : Q : 2",Q,$,.
Pr"tLr
Taking the inner product with yto,
lrN
r) : 2", Q Q li> : 2 Qo$t
Z^P-(ul
Thus we see that the components of the transformed vector can be characterized
by the effect of the operator acting on the old components, where both sets
of components are defined on the same basis.
The coefficients Qo, can be written in a square array which we call the
"matrix of Q" or the matrix (Q). Then we can interpret the equation
MATRIX REPRESENTATIONS OF LINEAR OPERATORS 225
r:
fl(: :) 0
g
.n/
Let Y. yofo, where Tx: &lr). Then, analogous to Equation 6.26,
lc
: : ProZQtiei : r*0,,) o, :
4 (4
Tr, 2 Profn
1 4 ^o,o,.
Therefore, the operator equation R : PQ is equivalent to the matrix
equation,
(Rr,): (; PooQoo),
Yo : Q-tVu,
Definition
A matrix is non-singular if its determinant,
lQoi * 0.
225 THE FORMAL STRUCTURE OF EUANTUM MECHANICS
Definition
The determinant lQn,l muy be written as an expansion in its cofactors,
A : > (-I)n**Qr*lqnrl,
lQnrl: ior&
where the summation is over only one index. Here, lfrol is the minor of the
element Qn*. It is a scalar whose value is given by the (nf
- l) x (lf - 1)
determinant remaining after striking out the ith row and the ,ttr column of the
original determinant. The cofactor of the element Qn, ir (-l)o*, lqnl.
Definition
The cofactor matrix is the matrix of the cofactors of the elements of the
original matrix.
Cof (Q,,) : ((- l)r+r' lquil.
Definition
The transpose of a matrix is the matrix formed by interchanging the rows
and columns.
Definition
The adjoint of a matrix is the transpose of its cofactor matrix.
Definition
The inuerse of the matrix Q is determined by dividing the adjoint of Q
by the determinant of Q.
ee_t:e_te_1.
MATRIX REPRESENTATIONS OF LINEAR OPERATORS T27
Pnoor.
Theorem
Tlte inuerse of a product of non-singular matrices is equal to the product of the
inuerses in the reaerse order.
Pnoor'.
Let
(Anc;-' -o
Then,
(ABC)(ABC)-l : ABCD
1: ABCD
A-1 :A-IABCD:BCD
B-1A-1 :B-IBCD _CD
c_1B_1A._1 _C_ICD:D
Therefore,
(ABC)-I : C-rB-U.-I.
Definition
Determinant: L: lQoi
Transpose: (Qr) : (Qon)
Cofactor: Cof (Q) : ((-l)t+t 1roU,
Adjoint: Adj (Qn,) : Cof (Qo,)
fnverse: Q-l :i oo, O
Complex Conjugate: (Qno)* : (Qil)
Hermitian Adjoint: (Qn,)t -- (Qn,)* : (O,[)
THE FORMAL STRUCTURE OF OUANTUM MECHANICS
tfil ::3tat
(AB)t
Ba
(AB){€ : A*B*
(A'n;-t - B-1A-1
Properties of determinants
lAl : lAl
:
lA*l lAl*
lAtl :la*l :lAl*
We will now show that a Hermitian operator A is represented by u
Hermitian matrix. Using the notation of Chapter 5,
Q;t: $lqi>,
Qrj: (ajli): <ilail: Qr,
Qu : A qU>,
and
Qfi:<jlQli):Qt*
Therefore, (Q) : (Q)I .
It is evident that the elements along the principal diagonal of a Hermitian
matrix must all be real, that is,
(F|Q,D : (gfQ,pn)*,
MATRIX REPRESENTATIONS OF LINEAR OPERATORS 229
where ytnis a column vector and its transpose, Fnris a row vector. This expression
may be readily verified by starting with the right hand side:
Show that
PR'BLEM 6 ts
.':'T:j,#T,:-:"":T.'
fi\2 i2 sll)
(u) (b)
(; ::\
\-12
'l
' d
Ans.:
fi1) fi:l)
ft
l;_i -) U _i l)
d('l
PROBLEM EI6
'(_1
ffiffid,#P,
_i I 'l l)
Evaluate the detorminants of the four matrices in Problem
6-15.
(Ans. : 6, 41 5, 30.)
s#H
230 THE FORMAL STRUCTURE OF EUANTUM MECHANICS
PROBLEMs.T
;;:T:-
Ans.:
:H:ffiHffiilffi :Tff
(u) l& -+ +\ (b) l+ o -+\
l-;i il
(.) I -? -+\
l;-l )
+ (d) l+ o o\
('-; (: ;:)
-; )
PROBLEM 6.18
Q la) : )"la),
which is equivalent to writing,
.^/.^/
2Qn,n1\r: A2";Q,
of,
.^r
- 16J,Qr:0.
4",(Qno
The condition for non-trivial solutions of this set of linear equations is that
the determinant of the coefficients is zero. That is,
Qtt-1 Qt, Qt
Q^ Qrr-1 Q*
lQu-16n,1 : Q", Q* Qr"-I'l:o. (6.29)
THE MATRIX FORM OF THE EIGENVALUE PROBLEM 23I
This equation, which is called the secular equation or the characteristic equation,
is a polynomial of order Nin )". Its i/ roots are the eigenvalues of the matrix
a.
It is evident that the problem of obtaining these eigenvalues would be
greatly simplified if the matrix (Qn,) were itself diagonal. In that case,
Equation 6.29 would become
and the y'/ eigenvalues would be the diagonal elements Qrr.,Qrr,..., Qrr.
We will show in the next section that there is a transformation, called a unitar2
transformation, such that anlt Hermitian matrix ma2 be written in diagonal form. Such
transformations are analogous to the principal axes theorems of classical
mechanics and geometry, wherein cross products involving two different co-
ordinates can be eliminated by choosing an appropriate transformation of
coordinates. The off-diagonal terms of a matrix correspond to the cross terms
in the geometrical problem. For example, a surface in 3-space with its principal
axes along the coordinate axes may be described by the expression
a6
dxt -xn
of,
PROBLEM 5.19
Show that:
(u) The eigenvalues of a Hermitian matrix are real.
(b) The eigenvectors of a Hermitian matrix corresponding
to different eigenvalues are orthogonal.
We state without proof that Hermitian matrices have all of the properties
of Hermitian operators; that is, the operators and matrices are isomorphic
representations of the same abstract algebra.
THE FORMAL STRUCTURE OF EUANTUM MECHANICS
'0'n
: Lui,'4,,
i:L
where the basis {fr} ir a complete, orthonormal set of .l/ functions. Since we
require the new basis to be orthonormal, we have,
: 6,
@i1,0')
4,rr) : 6,i
/N
(> uh,O0
\ft
1,0,)
: 6nt
Zu,rrf,@r
un*u'f* : 6nt
7
(6.30)
Tuora|, -'6,r
.
Yi : P'Y;:P'Q'YL:R'YL,
under the unitary transformation which takes the basis {rir} into the new
basis {rii}. It is desirable that Equation 6.28 preserve its form. Furthermore,
*. r..d to know if the operatorS P', Q', and.R' differ from P, Q, and ^R.
Let us write Yu : QY" in terms of the old and new bases. First, consider
Y":
n/nrN
Yro : 2 oo,$n
i:Lii
CHANGE OF BASIS: UNITARY TRANSFORMATIONS 233
&k : 4 "iuf,6,*
: 4ilt"i.
(6.31)
fl:uf)
To obtain the inverse transformation multiply Equation 6.31 by u,* and sum
over *:
: 6rai
Fu,*n*: ; \u;*fi,fia.; >
where Equation 6.30 has been used.
Yr: a%
utY; : QUtY, (6.32)
Yi : UQU-lY;, (6.33)
since Ut : U-1. Thus we can preserve the form of Equation 6.28 if we define
Q^ Q* Qr" uZ
Qtt Q* Q* T u!
:) f
ul ul
ul u", 120
u!
(l u! 0 ).s
f
235 THE FORMAL STRUCTURE OF OUANTUM MECHANICS
If we consider the columns of the U matrix as separate vectors, note that the
above result is equivalent to the separate eigenequations:
We now have the prescription for obtaining the matrix of the transfor-
mation U:
First find the eigenvalues of the operator matri* Q.
For each eigenvalue 2,, obtain the eigenvector Qlo from Equation 6.35.
U is comprised of the Ql o as columns.
EXAMPLE
Given,
l1 0-2\
Q:
(; : i
Find the eige nvalues and eigenv ectors of Q. Construct U and U-l Show that
U-leU : A.
lr - ),
;'l : (-i)t(r -
0
I 0 - 4l :0.
I
-), 1)(4 - D
l-2
I
0 +-l
Then the eigenvalues ar e)":0, 0, 5. Taking lt : 0, we have from Equation
6.35, Q%r: 0, or
o -2\ /ui\
(i :;)l;;)-o
which yields the results, u!. :2t
2u!.
DIAGONALIZATION OF MATRICES 237
- Since az| is completely arbitrary, let it be zero. Then the simplest choice
for a normaliz ed Q/, is:
%'l:*(i)
Taking 1z :0, we have as above , uI : 2u3. Since e/, must be orthogonal
to Qly choose u3, : I and u?. : u3 : 0. Then,
%z:0 . 1
(-:;
\-, l) ffi:0
Then,
uI : -2u? r
uz : o'
and we obtain
%s:#( ;)
\ -zl
The transformation matrix is constructed by placing the %oalong the columns:
lz- o -r\
(
::, 'j) :+(i
'l i)
\/s {5/
and
IJ-l : U.
u.'QU:-(l
'l il( i I il(l 4 i)
:.( ,l
ilfi ; ,l)
:( ,l
)fi I il:fi Iil-^
Find the eigenvalues and eigenvectors for each of the
matrices given in Problem 6-15.
(Ans.: (a) I, 2, 3; (b) I, 2, 2; (.) 5, l, 1; (d) 2, 3, 5.
The eigenvectors are the columns of the matrices given in
Problem 6-21.)
PROBLEM 6.2'
I
-\/2
-:
I
t/z
':,j)
0
* It should be pointed out that we use the symbol U for the transformation matrix even for those
cases where Q is not Hermitian and U is not unitary. In general, the eigenvectors of Q will not be
orthogonal unless Q is Hermitian. Also, lJ-L + Ut unless U is unitary.
DIAGONALIZATION OF MATRICES 239
(.) ll
t/s \/,
I
--0
\/3
11
{)
t/g \/,
(d) lr 00\
(': o)
0' t/
Find the inverse U-r of each of these.
PROBLEM 6.22
PROBLEM 6.23
ir'l;i1::il::iltii,t iLJ
PROBLEM 5.24
il
PROBLEM 5.25
/v oilv\
r:l tt o I o l.
\-ify o y I
(u) Find the eigenvalues of f.
(b) Find the unitary matrix U for the transformation that
will diagonalize I.
(.) Verify by matrix multiplication that UUt : l.
(d) Verify that U-1IU is diagonal.
Ans.: (a) l, y(l + B).
"(L
:)
ffi.q.e.wP
matrix defined on the same basis vectors. The effect of applying a given
operator to a vector, then, can be determined by multiplying the matrix
representation of the operator times the column matrix of the vector components.
We have seen that the energy eigenfunctions of the harmonic oscillator
form a complete, orthonormal set, the only complication being that the set is
denumerably infinite. Thus, if we regard these eigenfunctions as our basis
vectors we will be using a space of infinite dimension. For example, let us
consider the energy matrix for the harmonic oscillator. This is, of course, the
matrix of the Hamiltonian operator using the oscillator wave functions as the
basis functions. Each matrix element is given by
Since both (il and 17) are eigenvectors of ,f , the effect of the operator is
merely to generate the eigenvalue Eti which, being a number, can be moved
outside of the integral. The integral is zero for all off-diagonal matrix elements
because of the orthogonality of the oscillator wave functions. If we label the
rows and columns in the order, Vo, Vt, 1pz, . . . , the matrix appears as follows:
(6.36)
Equation 6.36 illustrates the important fact that an operator has a diagonal
representation when its own eigenfunctions constitut6 the basis.
As a further example let us write the matrix of the operator x, whose
matrix elements were given in Equation 5.41. It is as follows:
0 IO 000
I o \/2 000
0 \/, 0 \/300
/\ I
o \/5
\/ t/20
(6.37)
0 002 o \/5
0 000
Notice that the matrix of x is not diagonal. This was to be expected since
the basis functions are not eigenfunctions of x.
THE FORMAL STRUCTURE OF OUANTUM MECHANICS
For problems involving pure states, the value of the matrix formulation
is not evident. A somewhat better example of its utility is that of Problem
5-22, in which an oscillator is in a superposition of two states. The time inde-
pendent wave function as well as the function which includes the time de-
pendence are given below:
I I o-ouorn
t/z \/r"
I l'-
---= r-'u,'tu
f2 {z
,p(x) : 0 Y(r, t) : 0
0 0
The expectation value of the energy in the time varying state is:
ah I
T 00 o-o,o',n
\/r"
\.- I
{2
,ouorro. +
'{2 ,iEltrh g
ryo \/r"
o-'ur,lu
,u o2 5ah, 0
l: ah o-inotlh
{2
rouoln r+
'{2 riErtrh, o > o, 2\/r"
3ah o-iELtrFL
2\/r"
0
:- ah ,+3ah :
I- a)h.
+
SUMMARY 243
Although the last-example is still quite elementary, one can easily visualize
extending the method to more complicated situations. The real power of the
matrix method will appear when we discuss perturbation theory in Chapter 9.
PROBLEM 6-26
PROBLEM 5.27
: +yo(r,
{ro v\ / ' .hY,(x,r) + hY,@,t)
Y(", t) /)
'
++Y"(x.
{to
o\'7 t).
SUMMARV
the accessible eigenvalues, where the weighting factor for each eigenvalue is
interpreted as the probability of obtaining that eigenvalue. After a measure-
ment of the variable q, the system is known to be in an eigenstate of the operator
a. If this is followed by a measurement of !, the act of measurement forces
the system into an eigenstate of P. If the operators P and Q commute, the
measurem ent of p will not disturb the eigenstate of Q, and, in fact, any sequence
of measurements of p and q will repeatedly yield the same two eigenvalues of
the operators P and Q. However, if P and Q do not commute, a measurement
ofp will destroy all previous knowledge concerning the state of Q, and vice
versa.
A quantum mechanical state, which was just regarded as simply a_wave
function, can be visualized as a vector in an abstract function space (Hilbert
space). Such a vector may be represented by u columnar matrix array of
numbers which characterize its projections on a given set of basis vectors.
In like manner, an operator Q may be represented by a matrix array composed
of the elements <,(tnl Q 1,,$,) calculated from a particular set of basis functions,
{{t,}. Just as a vector has an existence that is independent of the set of unit
vectors used to define its components, an operator transcends the particular
set of basis functions used to express its representation. The diagonal elements
of the matrix (i : j) are the expectation values of the operator Q, and the
ofldiagonal elements (i t' j) will later be identified as the transition probabilities.
Although ,Qo and {t, are orthogonal, the operator itself might connect the two
states. Physically, this means that the act of measurement produces an inter-
action which has a non-zero probability of causing a transition from one state
to the other.
A coordinate system in 3-space can be rotated so that an arbitrary vector
1 lies parallel to one basis vector, say fc. Then the vector has a particularly
simple form, namely,
t:
0
In like manner, a "rotation" in function space can be effected by applying a
unitary transformation to the matrix a . There is one particular transfor-
mation that will enable Q to be written in diagonal form such that its elements
are the eigenvalues of the oper ator Q and its basis functions are the eigen-
functions of q. An equally valid procedure would be to construct the new
basis by formittg upptbpriate linear combinations of the old basis functions
so as to make I diigo"ut. In general, however, it is easier to obtain the
eigenvalues directly by solving the secular equation._ -If !!. eigenfunctions
ari treeded, they can then be obtained easily by means of the diagonal matrix A.
SUGGESTED
REFERENCES
David Bohm, Qunnturn Theory. Prentice-Hall, Inc., N.Y., l95l'
R. H. Dicke and J. P. Wittke, Introduction to Quantum Mechanics. Addison-
Wesley Publishing Co., Inc., Reading, Mass., 1960.
SUGGESTED REFERENCES 245
h2
:
- 2*Y'rpt(r,), z) + V(r,), z)rln@,), z) Eyu@,1t, z).
In the special case of the free particle (V :0), the solutions are plane waves
of the form
Whenever the coordinates are independent, that is, they are not connected
by interactions, the Schrodinger equation is separable and its solutions may
be written as a product of one-dimensional solutions as in Equation 7 .1.
Furthermore, the total energy eigenvalue is equal to the sum of the eigenvalues
of the one-dimensional problems, that is,
E :ry2m: L
2m
fo:, + ki + kZ) - E* * Eo + E,.
246
RECTANGULAR COORDINATES IN THREE DIMENSIONS 247
f_h'a'
L 2m ox2 + v"(r)
r
)rt,n,@)
: E*rttn,(x).
Similar equations hold for the other coordinates with the result that
E-8,+Er+8,.
Another example is that of a particle in a box. For simplicity, let the box
be a cube of side I. Although there is no force acting upon the particle and
its total energy is positive, it does not have a continuum of available energy
states like a free particle. The presence of the rigid walls (infinite potential
barriers) forces the particle to occupy one of a discrete set of energy states.
We found these states for the one-dimensional case of an infinite square well
in section 3 of Chapter 5. From Equation 5.18,
n2rn2h,2
En, :
2mL2 '
and
,tt@) : t . nanx
Zsrn L
Then,
and
D
n2nzfP
Lnln,n, , (7.3)
- 2*L,
where
y72:nl+n?,+n2.
THE WAYE. EOUATION IN THREE DIMENSIONS
PROBLEM 7.1
PROBLEM 7-2
'
(n",., #frm)881.)
PROBLEM 7.3 'b-:-:i:1'srr;:i*rt}lt.i;i;llliii.i,,jlfillf:;i.:ffi;ig-Tr*ii;-?q:+i+iLf
t-:)
rc.- -:lrr
where ris the distance between the particles. The time-independent Schrcidinger
equation, Equation 4.54, becomes
l- h2 -,, hz 1
: EYu,
l- Trro? - frv3lv, - v(r)Ys (7.4)
where Vf operates only on the coordinates of the proton and Vr2 operates only
on the coordinates of the electron. A difficulty arises with this expression,
however, since the potential V(r) is a function of the coordinates of both particles.
This means that Equation 7.4 is not separable in the two sets of coordinates.
For this reason it is convenient to transform the equation to center-of-mass and
relative coordinates,
The position vector ,R for the center of mass of the two particles is given
by the relation,
Mlf, -i{t :mlR_.irl,
from which,
fi : M3! mrz. (7.5)
M +m
The kinetic energy of the center of mass is given by
T - +(M * m)R?,
while the kinetic energy of the two particles about the center of mass is
Mm
T' :'
M+m
Therefore, the Hamiltonian for the system becomes simply
F : pi : -ih,i na,
and is the reduced mass defined in Equation 3.23. Since the Schrodinger
p,
equation is now separable in these variables, we may write its solution in the
form of a product function,
Yz : IPr*' garr
For a conservative system the solution Vcm represents the continuum of free
particle states for the system as a whole, and the function yor, represents one
of the discrete eigenstates of the bound system. Thus the total energy eigen-
equation rr'ay be written as
Hereafter we will drop the label "Rel" and will write the equation to be solved
as
we let
l-
Y2:9 + ro g',
where
gE: (#.?*) (7,7)
and
g)z ra2
- ##(""'*,) + ,i"'0-
PROBLEM 7.4
v2 - (# .? *) . ; [# * (""' *) + # #,]
.'.
2pr' I
hz R(r) Y(0, 4)'
.
r29R(r) 6) grY(0,
R(r)
+ryw -v(,)): - -wT:r\' a
(7.8)
and
gzY(,, o) : -.LY(', o). (7.10)
Since we know the differential operators 92 and 9, Equations 7.9 and 7.10
can be solved by the standard techniques for ordinary differential equations
once the exact form of the potential is put into Equation 7 .9. However,
the reader should recognize Equation 7.10 as an eigenequation of the type that
we have encountered before. Therefore, we prefer to solve Equation 7.10 by
means of operator algebra in a manner similar to that used for the harmonic
oscillator. But first we must digress to discuss angular momentum and its
operators in three dimensions.
E:i x P,
having components given by
L, :fiu
Ln : Z!, tT;l (7.1 l )
L, : xpo
We use the same definitions for the components of the angular momentum
operators, where it is understood that po is the momentum operatot,,-ih(7ldxo).
As we discovered earlier (Problem 4-22), a coordinate operator and its
conjugate momentum operator do not commute. Hence we should not expect
the components of the angular momentum operator to commute with each
other.
PROBLEM 7.5
1 Although
t :; x p'b.hu'u.s like an ordinary vector (polar vector) under addition, it is actually
a pseudovector (axial vector). Its different behavior appears under a reflection ofall coordinates through
the origin: i u.rd 1urc reversed, but Z is not.
THE ANGULAR MOMENTUM OPERATORS 2!'3
00
:Ifi", z!,f - llp.
/v
x!,1 lzfo, zp,f + lzQ,
- x!"1
:il,lf u, zf * xgolz, !,f
-- -ihlP, ! ihxP, : ih(x!, - )P.) : ihL,.
In the second line above, the two indicated commutators are zero because all
of the factors in each bracket commute. In the third line, all commuting
factors have been taken outside of the commutator brackets. The commutation
relations for the other angular momentum components may be obtained easily
from cyclic permutations of x,1, z in the expression
Thus,
lL, L,f : (i.13)
and
lL,, L;
Asin the case of vectors, the square of the total angular momentum operator
may be defined by
L2:L'r*L'r+L'r.
PROBLEM 7.5
Ext:ihE.
If were an ordinary vector, E x
^d i would be identically
zero. Why is this not true here ?
PROBLEM 1.7
The commutation relations derived in Problems 7-6 and 7-7 have im-
portant consequences with regard to measurements of the total angular
THE WAVE EOUATION IN THREE DIMENSIONS
momentum and its components in a physical system. The fact that the operators
L, Ln, and L, do not commute means that a measurement of any one com-
ponent of the angular momentum introduces an uncertainty of the order of fi
in o.r. knowledge of any other component of angular momentum. Thus, in a
quantum system, a measurement of one component of the angular momentum
essentially nullifies what might have been learned previously about some other
component. In other words, a measurement of L, forces the system into an
eigenstate of Lr. This state cannot be an eigenstate of either Ln ot Lr. However,
the two operators Z, and 12 commute, so these operators can have simultaneous
eigenstates, and the knowledge of these states will be preserved throughout
any sequence of measurements of L* and L2. An attempt to measure L, or L"
will immediately destroy the eigenstate of L, but will not disturb the eigenstate
of L2.
PROBLEM 7.8
L* :;n(si' O
* * cot ocos rk)
/
Lo.: -ifr(cos d &_ cotosin r&)
_:
Lu
._a
-'O a6
PROBLEM 7.9
We see from the result of Problem 7 -9 that the angular part of the
Laplacian operator is identical with the square of the total angular momentum
operatorexceptfor the factor (-h}). Furthermore, if we rewrite Equation 7.10 as
it is evident that the eigenfunctions of the spatial operator 92 and the eigen-
functions of the angular momentum operator are the same set of functions.
These will be shown later to be the spherical harmonics. This close relationship
between geometry and momentum is real, and indeed, is summed up by the
statements that the conseruation of linear momentum in tlte absence of external forces
follows fio* the homogeneity of space and the conseruation of angular momentum in tlte
absence of external torques follows from tlte isotropy d space.
THE ANGULAR MOMENTUM OPERATORS 255
where the last expression is obtained from the first two terms of the Taylor
expansionof y(i + di) about ,t,e). Ifwe make useof the expression F : -ih.Y,
then we obtain
ri :ll: (t +iar.'l'"
: +;\;'
co 1 /; \n
F) (i.lB)
: ,Qlni*6 .
trTs-Ts,*-0,
or (7 .21)
l.f , Tsl - 0,
255 THE WAVE EOUATION IN THREE DIMENSIONS
which is the required result. Since the operator Ts depends explicitly on the
linear momentu^ F, the latter must also commute with the Hamiltonian and is
thus a conserved quantity.
Now let us consider an infinitesimal rotation through an angle dd by
regarding the displacement di to be given by di x i. Then the counterpart
of Equation 7.15 is:
As before, we can express a finite rotation about tlte same axis as the repeated
application of the infinitesimal rotation operator:
Ro :I," (t
\
* hu|
. tl'"
d--o
: i
'tt'Plf
Therefore, when the Hamiltonian has rotational symmetry about an
axis, the component of angular momentum along that axis commutes with the
Hamiltonian and is conserved. Further, if the Hamiltonian has spherical
symmetry the total angular momentum and its component along any single
axis are conserved.
PROBLEM 7.'O
with angular momenta and that the components of the angular momentum
do not commute with one another. Suppose we now consider two infinitesimal
rotations about the x- andlt-axes to see what effect the order of the operations
will have on the result. This is given by the commutator,
I
lRo,, Ruol
h2
lL,, Lul 6{" 66u (7.23)
ry0,
prouided the second-order infinitesimal 66" 66u is negligibly small. To the
extent that the latter is true, we can construct any finite rotation from a large
number of infinitesimal rotations. This difficulty does not occur in the case
of translations along different axes, since the coordinate operators commute
for both finite and infinitesimal translations.
As in the case of the harmonic oscillator, the algebraic method for obtaining
che eigenvalues of the angular momentum operators requires the use of raising
and lowering operators which we will denote by the symbols L* and L_,
respectively. They are defined as follows:
PROBLEM 7.1I
L2:L?+Li+rZ
: t(L*L_ + L_L*) * ,r) (7 .27)
: L+L- - hLu + Lz
: L-L+ + h.L, + L2
The advantages of using L* and L- and the expressions derived above will
become clear shortly.
The problem we have before us is that of solving Equation 7.14, which is
derived from Equation 7.10, for both A and Ir That is, for each allowed value
of A (eigenvalue) there is a function I (eigenfunction) that satisfies Equation
7.14. Our goal is to obtain the whole set of these functions. We also have
another important bit of information at our disposal: the operator Z2 commutes
with the operator for lne component of angular momentum, which we will
elect to be L,. This means that the functions Y arc also eigenfunctions of L,
(see section 2 of Chapter 6). It was stated in section I I of Chapter 31hat the
z component of the orbital angular momentum is an integer times h, say mh.
This will be derived in what follows below. Let us then write:
But we may use the commutation relation (Equation 7,26) to rewrite the left
member of Equation 7.30 as:
Equation 7.32 states that the quantity L+Y^ may be regarded as a new eigen-
function of Lu having a new eigenvalue (m + l)h. Now let us see if this new
EIGENVAIUES OF THE ANGULAR MOMENTUM OPERATORS 259
L,(LZ*Y*): (* +2)h(LiY^).
Also,
Lz(L\Y^) : L,*LrYm - fftz(L,"Y^).
Thus, once again we have obtained a new simultaneous eigenfunction of L"
and L2, where the eigenvalue of Lu has been raised another unit of fi but the
eigenvalue of 12 is unchanged. Generalizing these results, one may write
L,(I:*Y*) : (* {r)h(I\Y*)
Lz(L'*Y *) : Lh.' (LnY *).
PROBLEM 7.12
Since the angular momentum operators L,, L.r, L,, and L2 are Hermitian
operators with real expectation values, the squares of these expectation values
must be positiue numbers. Thus the operator (L2, + Li) must have a positive
eigenvalue and /\ > m2. This means that there is a limit on the number of
times that either the lowering or the raising operator may be applied to a
given function Y*, and thus the set of eigenfunctions must terminate with a
minimum and a maximum eigenvalu e of Lr. For convenience let It be the '
L*Y/' : 0
and (7.36)
L-Y -t, : 0
since, by definition, Y/' is the function associated with the top of the ladder
of states and. Y-t, is the bottom function. Letting z be the integer representing
the number of steps betweenl t' and Y-t', we have
This implies either that f, - lr or that l, (1, * l). The latter choice is
rejected, however, since it contradicts the assumption that both /, and /, are
positive (Equation 7.36). From Equations 7.37 and 7.38 we find that:
oft
/:/r -tZ:2 (i.3e)
and
A: t(/+r). (7.+o)
From Equation 7.39 we learn that the maximum value of m is I and the mini-
mum ,uui,r. of m is - f,, where / can be an integer 0r a half integer. For the
classical orbital angular momentum which we are considering here, only
integral values of f-are physically meaningful, so the integer z must be even.
Hoiever, the reader can probably guess that the existence of the half-integral
states for odd z will permit this same formalism to be used for treating half-
integral spin angular momentum. From Equation 7.40 we see that the eigen-
valu-e of the tqrrut. of the total angular momentum is /2 + /, where / is the
maximum obsirvable component of the total angular momentum. Thus, one
can never measure a component of angular momentum equal in magnitude
to the total angular -o*.ntn-. A measurement of, say, the z-component will
THE ANGULAR MOMENTUM EIGENFUNCTIONS 251
nl
(:3
FigureT-3 The allowed projections of the angular momentum for the cases of /: 1,2, and 3.
(*. icotA
fi)wro,6):
0.
I 0 ,, icot? A
f8:ikd6 (7.+2)
dP(0\
(7.43)
m:kcot1d?,
whose solutions may be readily written in the form
o(d) - rttcl
and
P(0) : (sin 0)k,
where the constants of integration have been neglected for the time being.
Then the product functron Y( is:
L'Yt' : /hY'
and
L'Y, -- /(/ + t)h,v,
to show that the separation constant k : /, and that the
product function is
YI : Q/s)$in o)teuo. (7.++)
operates. Thus,
(7.+7)
: ( _ L)scr-s eiv--,d
$n (# fil Gr^ u),,.
cos9: rl
dd 0n.
d0: -srn
Then,
Then,
YT?i : (- \*Cf PT?i'u*6. (7.50)
upon m which enters when we operate with I-. Our starting point is the re-
quirement that
Then,
<YilYh - 1.
f
L : lctlrJ,,liri" 0)teit67* . [(sin 0)/ea6f de
fn
:2n'lcll'Jo (sit o)2/sin o do.
This integral may be evaluated by means of the properties of beta and gamma
functions:
1+x:2t
L - x :2(1 _ l)
dx :2 dt,
and Equation 7.51 becomes:
: I)
2z/+tf.(( l- -l)l(/-'+
2r lcr' ' 1
't TQl+2)
:2r il 1
lcltz.2z/+r L .
Qf +t)t'
since l(/ + l) : ll Therefore
(7.52)
and
yt,::
'' 2/ll\lW4n gin,l)teit6. (7.53)
266 THE WAVE EQUATION IN THREE DIMENSIONS
yf : r cf t--r
r vm*L
/ (7.54)
h' C7*t'
where we use the fact that L* and L- are Hermitian conjugates. Applying the
commutation relations, the ket on the right becomes (see Equation 7 .27):
<yf*, l ytr*,,
r / : %l'. V,- - m - r)(/
l -/\- + m * 2)(Yf*'
t\ l vf*'>. u
lCy*rt
It follows that
Yr :(ir-;ffi m)l
"ol '
- m)l
rl-mvlr
l-- tt (7.s6)
--r
YT?r 6) (7.57)
ffiPT?t),'*d,
and
Y7?1,6) : #; - t)'
(7.58)
where rl : cos 0.
The above results were obtained for m positive. In order to obtain the
functions for negative m, it is most convenient to use the relations
and
Y7* (ri : (
- 1)- ( YT)* ?i . (7.60)
Of course, these expressions could have been obtained by following the pro-
cedure above, using Y;t as the starting point and operating with Z*. In this
case the counterparts of Equations 7.54 and 7.56 are:
Yr :iffiL*YT-" (7.61)
and
YT : (i)'.- J (2t)t(/
(/ - m)t Lf*Yr'. (7 .62)
m)l |
The angular momentum eigenfunctions, as given in Equations 7.56 and
7.58, are known as the spherical ltarmonics. They comprise a complete, ortho-
normal set of functions on the unit sphere. A few of the normalized spherical
268 THE WAV.E EOUATION IN THREE DIMENSIONS
vo- I
ro
- 2\/;
n:1/1
' 2rl.r .oro
r,*':Tl zrrl/Trin o'e+io
2n
vE ::^lt. (3 cos2 o - I)
^v?7
v+r : -- I
12 - r 2rl/E
z,
. sin 0 .cos 0 . erio
- :! +'Jit5
Y"L'
2n
. sin2 o . e+zio
4:l_
,3
+l;17'ftcos'o-3coso)
Y*':+l
rs -
-f ; I.lT sin0.(5cos2
dV
0_ l).e+to
rr
Since the ofldiagonal elements are characterized by either d' a { or m' t' m
(or both), the Kronecker deltas ensure that all off-diagonal elements are zero.
If we order the basisfunctionsasfollows, I?, Yl,Yl, Yrt, Y3, Yl, Y3,..., then
we can write the matrix of L2 as:
io
i1
i9
t1
lo
t/
t-
(L'.) : h' i6 i
(7.63)
t2
where all of the off-diagonal elements are zero. The boxes correspond to the
2l + 1 functions associated with the same eigenvalue of L2. That is, the one-
by-one submatrix corresponds to / : 0, the three-by-three to { : l, the
five-by-five to / : 2, and so forth. In like manner, the matrix representation
for the operator L, is
_1
i
: I
(L,) h (t.a+1
-1 cri
j
-1
Although the matrices shown in Equations 7.63 and 7.64 can be of infinite di-
mension, each set of 2/ + I functions associated with a given angular mo-
mentum forms a subspace of dimension 2/ + t. Furthermore, tlte 2/ + I
spherical ltarmonics associated with that f-ualue constitute a complete, orthonormal basis
in that subspace.
The operators L*, L-, L,, and Lo can also be represented by matrices,
although their representations are more difficult to obtain. One approach is
to begin with Equations 7.54 and 7.55 and to write:
L+L_Y\:hf@ L,YT_1
TO
:h'\/@Yr
or
L+YT:hf@Yry+r (7.66)
L|YT:h.f6Yr*t. (7 .67 )
(L*):hfi
(L*)-n\: l': ',l l\
and
:3:U'
lo o o o
l, o o o
(L-\:hl o \/6 o o
\o o fG
\o o o
o
2 il
where all other elements are zero. Then, from the definitions of Equation
THE ANGULAR MOMENTUM MATRICES 27t
0 u^/=y6 02
0 00 20
and
0-2 00
20 -t/a o
PROBLEM 7.18
PROBLEM T.'9
PROBLEM 7.20
(u) Find the matrix (t/) that will diagonalize (L,). (Do
this for the 3 x 3 matrix).
(b) What does this transformation do to (Ir) and (L,)?
(.) Show that the transformed matrices still satisfy
(L,)' + (L,)' + (L,)' :2h'(l). Physically, this is
equivalent to changing the axis of quantization from
the z-axis to the x-axis by, for example, rotating a
magnetic field from the z-direction to the x-direction.
272 THE WAYE EEUATION IN THREE DIMENSIONS
8. HYDROGENIC ATOMS ,
where the wave function has been labelled with the angular momentum
quantum numbers corresponding to those of the angular function. We already
know that the angular functions are the spherical harmonics given by Equation
7.57. It now remains to determine the radial functions. Rewriting Equation
7.9, we obtain:
R(r) : +,
which, upon differentiation with respect to r, has the derivatives:
g'(r) _ g(r)
R'(r) --
r 12
HYDROGENIC ATOMS 273
:-, I
=l-
+lE
IL
\
'-,
lc.l
\l I
b0
li
O
I
Coulomb potential
and
R"(r)--+-ryP+ry
Then Equation 7.69 becomes:
Since this cannot be solved closed form, we must employ either the
in
power series methodl or an analytical approach which utilizes a knowledge
fttn. asymptotiCs6lutions in the limits of very'small and very large r. Choosing
s See L. Pauling and E. B. Wilson, Introduction to Quantum Mechanics. McGraw-Hill Book Co., Inc',
New York, 1935, p. 12l.
274 THE WAVE EOUATION IN THREE DIMENSIONS
the latter approach, we note that for very large r, Equati on 7 .7 0 may be approxi-
mated by
s"(r) +ff s?) - o,
Since the plus sign leads to a function which diverges for large r, we cannot
use it for a physical system. Hence our only acceptable solution is
The oscillatory solutions which result from positive total energies (which were
excluded above) correspond, physically, to scattering states in the energy
continuum for free particles. Each continuum state is infinitely degenerate
since it can be assignid un infinite set of {-values. A detailed discussion of these
states will be deferred until the section on the quantum mechanical treatment
of scattering.
In the li*it of very small r, the dominant term in the bracket of Equation
7.70 is the llr2 term. Thus, in the asymptotic limit for small r, the solutions of
Equation 7.70 will be dominated by the solutions of
pZe' z (7.7+)
'o - h2)" aoA
att+t : alt
l+k*L-n
(k+L)(2/+k+2)'
In order to obtain a satisfactory wave function, the infinite power series must
be truncated to a polynomial so that it can be normalized. If the polynomial
is to terminate with the kth term, it is necessary that ar*, - 0. That is, we re-
quire that
n : / + k + l,
wherek:0,112,... Hence, z takes on integral values given by / * l,
/+2, /+3,.... It is customary to callntheprincipal or total quantum
number; it is allowed to take on integral values but not zero. For a given value
of n, the angular momentum quantum number / can have any one of the set
of values 0r Ir2, n - l.
The polynomial solutions of Equation 7.73 may be obtained by means of
series solutions in a' manner analogous to the treatment of the harmonic
oscillator in Chapter 5. .For a detailed treatment the reader is referred to one
275 THE WAVE EOUATION IN THREE DIMENSIONS
of the references listed at the end of this chapter. The polynomials are members
of a class of well-known functions called the associated Laguerre polynomialsra
To obtain the complete radial solution for the hydrogenic atom we must
retrace our steps through all of the substitutions and obtain
Io-'o
P"lLYf;(p)l' p' d P :ffffi
Normalization over r is obtained as follows:
rco 'zrz
dr : (fJI*t R*,(p)1, p, dp : W I,-l*^,(p)f,p, dp.
J. lR",(r)l
Then the normalized radial wave functions are given by:
Figure 7-6 shows the general behavior of some of the radial functions. A few
of the normalized radial functions are given below for convenience.
ft,,(,) :(ef
+ 7.*o (-#)
R,o(,) :61 ,[' -? fl*+ffi] exp (-#,)
ft,,(r) :(#,r + ';(, -L? exp
?a)
R,,(,) :(#: m (#.*o(- ?)
a See, for example, M. R. Spiegel, Mathemntical Handbook of Formulas and Tables. McGraw-Hill
Book Co., New York, 1968, p. 155.
HYDROGENIC ATOMS 277
Zrlas
Zrl as
l5-
Zr/ ao
Figure 7-6 Graphs of the radial wave functions R,,t|l for n : 1,2, and 3 and /:0,1,2.
(From Principlesof Modern Physics by R. B. Leighton. Copyright 1959 by McGraw-Hill BookCompany.
Used with permission of McGraw-Hill Book Company.)
The allowed energies are precisely the same as those obtained for circular
Bohr orbits and for Sommerfeld's non-relativistic elliptical orbits. Such good
agreement between classical and quantum mechanical calculations is fortuitous
in the case of the Coulomb potential, and should not be expected in general.
278 THE WAVE EOUATION IN THREE DIMENSIONS
Since the energy levels are independent of the quantum numbers / and m,
we can express the degeneracy of each level by the sum
n-L
>(2t+ 1) : I +3 +5 + "' + 2n -r :n2-
f:o
The (21 + I )-fold degeneracy of the m-values is characteristic of all spherically
symmetric potentials, while the z-fold degeneracy of the f-values occurs only
for a poteniial such as the Coulomb potential, which satisfies a specific invari-
ance iondition.s Historical usage has coined the unfortunate phrase "acci-
dental degeneracy" to describe the latter. The reader can anticipate some of
the results of later sections by speculating that any alteration of the physical
system which destroys the spherical symmetry or the pure Coulombic potential
will lead to the removal of some or all of the degeneracies'
5 The n-fold degeneracy will occur for any potential that is invariant under the group O(4). See
L. I. Schiff, Quantum Mechanics, 3rd. ed. McGraw-Hill Book Co., New York, 1969, p.237.
HYDROGENIC ATOMS 279
where n canbe any integer, I can take on the values 0rlr2r...r(n _i),
?,nd
*,:1n have any of the values /, I 1,...,0,
- given .'..',' f. 'ilthough
-1,
the radial and angular functions have been above, a few of the norm ali'i"d
spatial, hydrogenic wave functions are tabulated below to facilitate their use
in the problems:
Vrc- : ''-zrtas
Vlr,o
*Gf
Vzt - l?zoo :#c€)r (z 7).,-zr,za.
I /Z\z Zr
Vze o
a {Zn\aol ao
1Pze
I /Z\z';Zr ' sin o'e=io
m(;/ '-zrt,oo
Vst: lpsao
280 THE WAVE EAUATION IN THREE DIMENSIONS
?,-"''"'cos t;)
ege
rJ(, -
--?sro: *J; o
vsp :?arrr --
t;)
#rffi(t - 7o''-"'''oo
sin .'e+ao
ew: llszo:
#(tr # '
'-zrt,ao'
(3 cos2 0
- 1)
and
{2 y,*r,*, : -f(r)' r sin 0' erif : "f(r)(x I d.
HYDROGENIC ATOMS 28t
.,a.2s
tt""..
I ,'- -- -j-"''.
\
[4,. ..iri=-_...,_._ -l=.
...-....3P
\.,
.rt
,/
i \..
'tr
\
-t'.
s.rE6A212tt?t6$
cE6A212tr?!6
Figure 7-8 The radial probability distribution function lrRntlz for several values of the
quantum numbers n, /. (From E. U. Condon and G. H. Shortley, The Theory of Atomic Spectro,
Cambridge University Press, Cambridge, 1953. Used with permission.)
I
IPr, -- f (r) ' r t/z
(rpn , * ,ltnt,-t)
I
Voo : -f (r) 'l (rlrnr, - Vnt,-r)
i\/,
vro:-f (r)' z : gnto- (7.78)
THE WAVE EGIUATION IN THREE DIMENSIONS
Each of these wave functions resembles a figure B having rotational symmetry
about one of the coordinate axes. Such wave functions have applications in
crystal lattices and certain molecular structures when rectangular symmetry
exists.
PROBLEM 7.21
PROBLEM 7-22
Calculate the expectation value (r) for the g16s, the rprro
and the Vra20 states of hydrogen. Compare with the Bohr
theory results.
PROBLEM 7.23
PROBLEM 7.21
PROBLEM 7.25
Y(r,o,s):*,*+fia +# fia
Then, for the 2s state of hydrogen, ?(VV*) is real, and ,i : 0. It should be
evident that this will be true for all s-states. For the 2p states, however, we
write:
lpzp : Rrt(r)Yf Q, 6),
and
Vt, : Rrr(')Yt*(0, O)'
Then,
,pro(Ygto) : Rlrr(r) ' Rr, (r) lYil'z i
+ | rn,,t r)pYr fi rtne
+#tR,,(,) rYrf,rY,*)$.
Since the radial component has no imaginary part, it can make no contribution
to the current, as in the case of s-states. Let us now examine the d term. Its
angular factors may be written as the product YT : Pr(O)ei*o. Then,
rrm
YT
A
: Pt(0) 'P',t(O),
a0Yr*
which is also real. Therefore, the 6 term does not contribute to a current. It
remains to examine the fi te.-, which may be written in the form:
-tm
m-o ' lR,'(')l' ' lP'(0)l'.
Since this is an imaginary term it does contribute to a probability current
density in thef'ait.ciion, given by
a
Dd:
hm lRrr(r)Pr(0)l' a
*06-9'
io : _tSo'
of,
dtr[-" :4lR"(r.)Pt?)1'z
2moc srn 0
6
lRrr(r)P'(0)1'?
- -mltn d,,
sin 0
and
l@
M, : -mtta I lRr'(t) Pr(o)12 4m2 dr. (7.7s)
Jo
;,qrt1)1'lrii,;:4ii;:i
In Chapter 5 we noted, for both the square well and the harmonic
oscillator, that each eigenfunction can be either symmetric or antisymmetric in
the coordinates. In quantum mechanics the term parity is used to characterize
the symmetry properties of wave functions with respect to inversion (that is,
reflection through the origin), which is equivalent to changing the sign of each
coordinate. Thus, euen and odd parity refer to the symmetric and antisymmetric
cases, respectively.
It is convenient to define a parity operator II which changes the signs of
all coordinates. For example,
:
plus parity even : symmetry under inversion
parity
minus parity : odd parity : antisymmetry under inversion
Let us now examine the parity of the hydrogenic wave functions. The
parity operator transforms the variables as follows:
It follows that
frrpnr*:fIRntYT : RntfrYT : (-L)'rl,nr*,
which says that the parity of the hydrogenic functions is the same as the parity
of the spherical harmonics, namely, (-l)'. Using the same terminology, the
parity of a harmonic oscillator function it ( - I ) " and the parity of a square well
function is (-l)"+t.
When states of different parity have the same energy, the degeneracy is
called accidental degeneracy. This occurs for the different l-values in a hydro-
genic potential when all non-Coulombic interactions are neglected. * If these
interactions are included in the Hamiltonian, the f-degeneracy is removed and
all of the hydrogenic states become states of definite parity.
One might ask whether the parity of a state is a constant of the motion.
To determine the answer we examine the commutator,
* See footnote 5.
286 THE WAYE EEUATION IN THREE DIMENSIONS
Nobel prize was awarded to Lee and Yang in 1957 for their theoretical pre-
diction of this phenomenon.o
Frequent use of the concept of parity is made in calculating matrix
elements between states having definite parity. Since an integral of an odd
integrand over symmetric limits is zero, one immediately knows that all matrix
elements of an odd operator (antisymmetric in the coordinates) between states
of the same parity are zero. Likewise, matrix elements of a symmetric operator
between states of opposite parity are zero. This will be very useful in later
discussions of the selection rules governing transitions under the influence of
perturbations. Since the parity of an isolated system is conserved for all but
the weak nuclear interactions, it is often useful for obtaining information about
a state where the exact wave function is unknown.T
SU|'|/YIARY
: t(/ + r)\1
(r)
ryj * i{' _
nz Jl
I
(r') :
%l' * 1{' _r(r+r)-l\l
n2 l) -)
11\ Z
,h"'
11\
\,'/
/1\
m 22
z3
\d/ aflnsl(t++)(/+r)
,_1, _ gr, I
ur^h-t(t+L)\
11\ onr'(/+ +)(/ '+ r)(t + +)t(f
\rn / - +)
SUGGESTED
REFERENCES
David Bohm, Qtnntum Theorlt, Prentice-Hall, Inc., New York, 1951.
Sidney Borowitz, Fundamentals of Quantum Mechanics. W. A. Benjamin, Inc.,
'_ New York, 1967.
R. H. Dicke and J. P. Wittke, Introduction to Quantum Mechanics. Addison-
Wesley'Publishing Co., Inc., Reading, Mass., 1960.
P. T. Matthews, Introduction to Quantum Mechanics. McGraw-Hill Book Co.,
New York, 1968.
E. Merzbacher, Qtantum Mechanics. John Wiley and Sons, Inc., New York,
1961.
Albert Messiah, Qwntum Mechanics. North-Holland Publishing Co., Amster-
dam, 1958.
D. Park, Introduction to the Qrcntum Theory. McGraw-Hill Book Co., Inc., New
York, 1964.
L. Pauling and E. B. Wilson, Introduction to Qtnntum Mechanics. McGraw-Hill
,/ Book Co., Inc., New York, 1935.
Powell and B. Crasemann, Quantum Mechanics. Addison-Wesley Publish-
t,/J.t-. ing Co., Inc., Reading, Mass., 1961.
F. K. Richtmyer, E. H. Kennard, and J. N. Cooper, Introduction to Modern
Ph2sics,6th ed. McGraw-Hill Book Co., New York, 1969.
D. S. Saxon, Elementary Quantum Mechanics. Holden-Day, fnc., San Francisco,
1968.
Leonard I. Schifl Qtnntum Mechanics, Srd ed. McGraw-Hill Book Co., New
York, 1969.
Robert L. White, Basic Quantunt Mechanics. McGraw-Hill Book Co., New
York, 1966.
Klaus Ziock, Basic Quantum Mechanics. John Wiley & Sons, Inc., New York,
1969.
$ptsf;s dhnmilTtsffi 0F
JTHffi T,S ilAK MSil?T Hhf, TAu
A3{m tmHs*Tr*AL
PEKYffiffi&H$
We have freely alluded to the fact that the elementary particles possess an
intrinsic angular momentum which we caLI spin. Thus, we noted in Chapter 3
that a magnetic moment is associated with the spin of a nucleon, as well as
that of an electron, and that this moment can be accurately measured by beam
deflection or magnetic resonance experiments. It was from the number of
components in the beam experiments that the halflintegral value of $h for
the spin of the electron became apparent (see section ll of Chapter 3). The
existence of an electronic spin angular momentum of $h was postulated by
Uhlenbeck and Goudsmit in l92B in order to explain the fine structure splitting
of hydrogen.l It was soon observed that this assumption also explained the
"anomalous" Zeeman effect, the doublet structure of the spectra of the alkali
metals, and the anomalous magnetic moment reported by Compton,2 provided
that one also assumed the "anomalous" gyromagnetic ratio of roughly twice
the classical value.s
The Schrcidinger formulation of quantum mechanics is built, as we have
seen, on the classical Hamiltonian function. Thus, the Schrcidinger equation
does not include spin as an intrinsic variable, since spin is a relativistic phenom-
enon having no classical counterpart. For example, there are no additional
generalized coordinates that will permit us to write the spin in the classical
form,
^i x F,.:i
This conceptual disadvantage turns out to be an advantage, however, since the
uniqueness of the spin permits Sz to commute with all other dynamical vari-
ables. Hence, the spin is a constant of the motion and s is a good quantum
number. (S, may or may not be a constant of the motion, depending upon the
nature of the spin-dependent interactions, if any, that are included in the
Hamiltonian.)
Although it is an ad hoc assumption in the Schrodinger treatment of quan-
tum mechanics, the half-integral spin of the electron arises quite naturally out
of the relativistic theory of Dirac.a Thus, its validity as an additional coordi-
nate has been firmly established. Having this assurance we can, in many
instances, treat spin phenomenologically by including it in the Schrodinger
wave functions u"a i" the Hamiltonian in such a way as to account for the
physical effect being considered. The success of this ad hoc method is attested
ty the frequent use of a "spin Hamiltonian" in the Schrodinger equation for
solving problems.s
dl
m.
P. A. M. Dirac, Proc. Royal Societ2 (London) 117' 610 (1928) and 118' 351 (1928).
4
For examples in the fieli of magnetism see K. W. H. Stevens, "Spin Hamiltonians,"
5 in G. T.
Rado and H. S;hl, Magnetism, Vol I, Academic Press, New York, 1963'
SPIN IN THE SCHRODINGER FORMULATION 291
In order to include spin in the Schriidinger wave function we must use the
fact that the spin coordinates are independent of the cogrdinates of con-
figuration space. This enables us to write the total wave funCtion as a product
function,
Yroru, : Vu*(r, 0, O) . 1(spin) . ,-iEntltt : lRu . t-iEntlh\ lf, mt) ls, m,). (B.l)
Here the radial and time functions have been lumped together in the same
eigenket, since our main concern in this chapter will be with the last two kets,
representing the angular momentum and spin eigenfunctions.
For spin * *e know that there are two basis functions for the two-
dimensional spin space. If we adopt the convention that an ".rp" spin has a
positive projection on the axis of quantization (usually the z-axis), these
spin functions may be written in the following equivalent ways:
(8.2)
Since either of these functions may be used in Equation 8.1, we see that each
is doubl2 degenerate in spin. The spin wave functions, or
h2drogenic spatial state
spinors, given in Equation 8.2 are simultaneous eigenfunctions of the spin
operators 52 and S,. Thus, using ket notation:
: + r)h, li-, +) :
3h2 l1 1\
s, l+, *) 7 lEt 2/
+
'.(*
^t, l+, -*) : *(* +t)h, l+, -+) :7li'
3h2
-2/
1\
,s, l+, *) :
h
il+,,+>
ghzlt 0\
s2: z(o
r) t':;(;nlr 0\
-)
l0 l\ t0 0\
t*:n(o o/ s-:n(r ;/
s,:*(s*+s-) :X('r;) S,:irr--s*) :i(0, -;)
It is convenient to define the Pauli spin matrices:
6,:? r,:(l ;)
6v : i, t' : (o
/0 -i\ (B'3)
,)
o" :? ., : (l
-:)
PROBLEM 8'l ii.$fii*il$'4-\1l,'l.t
(b) Show that oron : iour 6ao, : io, and oror : io,
(.) Show that o,oro, : il.
(u)
(b)
Show that o| - o', - oli t.
Show that the commutation relations for the Pauli
THE SPIN.ORBIT INTERACTION
o X o :Zt,o
The total wave function given in Equation 8.1 for an electron possessing
both orbital momentum and spin assumes that there is no interaction between
L and $ that is, that [r, S] - 0. Hence, Equation 8.1 is an eigenfunction of
both L, and S,, which means that both ms arrd ms are good quantum numbers;
in other words, the projections of i and S are constants of the motion. In
reality, however, there is an interaction between f and ,S-called the spin-orbit
interaction-which removes the spin degeneracy, since it produces an energy
difference between the states for which the orbital and spin magnetic moments
are parallel and antiparallel. We will see later that this interaction can be
expressed in terms of the quantity t. S. Since t' S does not commute with
either i or ,S, Equation 8.1 is no longer the correct wave function, and lre, and
rts vte no longer independently quantized; hence, they cease to be good
quantum numbers.
An elementary way to picture the spin-orbit interaction is to regard the
stationary spin magnetic moment as interacting with the magnetic field
produced by the orbiting nucleus (see Problem 3-lB). In the rest frame of the
electron there is an electric field, A : Zeilrz, and a magnetic field
H":T
i xi
t
where f is directed from the nucleus toward the electron. Assuming that d is
the velocity of the electron in the rest frame of the nucleus, the current produced
by the nuclear motion is/ : -(Zelc)i in the rest frame of the electron. Then
!l- fr x 8.
2
Leu xr
E": - c
The spin moment of the electron precesses in this field at the Larmor
frequency (see section 10 of Chapter 3),
dr:TE": - 4o x&,
flloc"
(8.4)
Equations 8.4 and 8.5 hold in the rest frame of the electron. Since the rest
frame of the nucleus is of more importance experimentally, we must make a
transformation to that frame. This introduces a factor of $ which is called the
Thomas factor.6 This factor results from a relativistic effect known as the
Thomas precession Hence, an observer in the rest frame of the nucleus would
observe the electron to precess with an angular velocity of
Equations 8.6 and 8.7 can be put in a more general form if we restrict Z
to be any central potential with spherical symmetry. Then,
+AV+
F:-i+- or -eE,
and
-.i.. Z IAV+ -
DXd llaV- ,- I IAV--
--Tixi:--
e or ^-DXF: emor :-L.
er 0r dr
Using Equation 8.6,
+:
QL
I lav-_
- 2^rrrri A, "
and
L,E:- av-
.^l'S. (B.B)
2mficz r dr
Equation B.B illustrates the. explicit dependence of the spin-orbit energy on
the relative orientations of /] and,S. For spin : f the energy splitting is:
:
tAE;
ffii#
For the Coulomb potential this can be approximated by L,E : fr\mt. Ze2lrs.
Since the average value of 1/rB is?
z3
alns/(/++)(/+t) (for / + 0), (8.e)
LE,,
-m- (*)' .
# -,ro-a ev.
6 L. H. Thomas, Nature ll7,514 (1926).
? See Appendix A
of Chapter 7.
VECTOR MODEL FOR COMBINING ANGULAR MOMENTA
This is the same order of magnitude as the fine structure correction which goes
as a2 ^, l0-a (see section 2 of Chapter 4).
PROBTEM 8.5
PROBLEM 8.6
PROBLEM 8.7
(a)
(b)
Figure 8-2 two possible orientations of the spin and orbital moments for a single elec-
(a) The
tron' (b) spin'orbit coupling causes the vectors i and ito pr"."ss about
velocity. J'at the same angular
vector sum,
j:i +.9.
The vectors t an{S precess about the vector i at the same angular velocity
(see Figure 8'2). The conditions on angular momentum now apply
to J2 and J, instead of-quantum
to L,, sr, L, and s, separitely. That is, we ca.r d.nrrl
the eigenvectors U, m,) such that
possible values of m,. In the case of a single electron there are always just two
values of7, namely, j : f + +. For an s-electron the only possible value of7
is f, which has the two projections, mi : **. For a p-electron, j can have the
values,j : | * * : I or *; for a d-state, j :2 _[ * :8 or g.
The new angular momentum eigenfunctions to replace those of Equation
8.1 are not obtained so easily, since they must be the appropriate linear com-
binations of the old functions which diagonalize the matrices J2 and J,. We
know, in principle, how to determine these functions but it will suffice for the
time being merely to denote the eigenfunctions by their eigenvalues in ket
notation as follows:
J 1 t1 1\ 11 1\
z lZt 2/t lZt -2/
p 3
q
13 3\ 13 1\ 13
lEt E/t lE> 2/t lT> -2/t
1\ 13 3\
lEt -E /
1 11 1\ tt 1\ (8.10)
2 lZt 2/t l2t -2/
le 5\ 3\ 1\
l2) 2/) l5
l21 2/) l5
1\
l21 2/t l5
:)
z 12, -2,/,
15
lt, _3\ l_5 _ 5_\
-T /, lEt -E /
t3 3\ 13 1\ 13 1\ 13 s\
12>T/t lEt 2/t lT> -2/t lEt -E /
PROBI.EM 8.8
J+ : J, +iJ,
J2 - ti + ti + J2 : t(J*J_ + J_J+) * J"
i x i:ihi
[Jt, J,] : [J', Jr] : [J2, J,] : 0
[J*, J-] : 2hJ " (B.l l)
[J,, J*] : h,J
+
0l-il--o
I
li
io rl
l______-__i
J2:hz
lb
ib
r0
io
i+ oi
i 0 -l i
ir
i0
l0
Ju: h
VECTOR MODEL FOR COMBINING ANGULAR MOMENTA 299
0,
----l---------,
io ri
il
i0
ll
0i
t--------- i ----------_________,
t-i
io \/2 0i
i 0 y2i
lo lt
oi
io t/3 0 0
it- oz o
i o \/3
ioo
J,:
-r
I ---------___-____-__-_--
02 --------- |
i
0/6- i
o{6
02
0
o\E 0
o 2fr,
i0
ii
0i
it oi
,-------i------
i0 0
i {z o
i 0 y2
tt-
J-: h, i0
I
i2
I
I
o 0
i o {6
I
0
I
Go
I
I
i0 0
i0 0 o2
iU
i{s o 0
i02\/20
i'
300 SPIN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
0i
t9
o)E 0
\/i o fr.
o lz o
iri t-r t/t o oi
i v3 0 2 0
l-
i
il
i o 2 o ,4i
i-i
ii, o o rz oi
- l0
2 o on
i2 o rG oo
,0 v6 0 v6 0
r0 oG o2
i0 0 0 20
o -{z oi
Y2 rl
O -Y2l
o t5 oi
ih
JU
2
0--2 0 0 0i
ro-.nr6o oi
o.vG o -.G 0r
oo\Go _t!-i
0002 oi
i
If one wishes to account for the observed fine structure of the level splittings
in the spectra of one-electron atoms, there are two other effects which should be
considered. The relativistic correction is one of these, since its contribution is
INTERACTIONS IN THE ONE.ELECTRON PROBLEM 3OI
the same order of magnitude as the spin-orbit splitting. Although the Dirac
relativistic theory is the proper theory to use, it is possible to obtain the
relativistic corrections from the classical Hamiltonian by writing the kinetic
energy in the relativistic form,
where Equations l.16 and l.l8 have been used. Fory' 4 moc, the radical may
be expanded to give
lp'pn\
T - + t . ..) -
fu - W
mocz ttoc2
\t
o
p' pn
2*o - @''
ff x@-+z)
\2*o t ' I -L.
}mfrcz'
AEur:ry(r-&) (8.12)
Note that this would be in exact agreement with Equation 4.13 for k : / * z!.
The second effect, the nuclear hyperfine interaction, is several orders of
magnitude smaller than the spin-orbit and relativistic corrections. Although it
is often difficult to detect in atomic spectroscopy, the hyperfine interaction is of
considerable importance in electron and nuclear magnetic resonance experi-
ments. It was pointed out in section 9 of Chapter 3 that nucleons possess an
intrinsic spin of $fi. Consequently, nuclei have i net spin except for the special
cases in which the nucleon spins pair off to give a total spin of zero. The
interaction of the nuclear spin with the spin of the electron is known as the
hyperfine interaction. As in the spin-orbit case, the energy associated with
the hyperfine interaction may be expressed by the vector coupling model as
where .I is the nuclear spin and z4 is the hyperfine constant. Although the
nuclear spin is frequently larger than the spin and the orbital angular momen-
tum of the electron, it should be recalled that the nuclear magneton is smaller
than the Bohr magneton by a factor of about 2000. Hence, the hyperfine
energy is smaller than the spin-orbit energy by roughly a factor of 10-3.
302 SPIN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
As in the spin-orbit case, the kets lF, Mil may be expressed by linear corl-
binations of the kets I t, mt) ls, m,) lI, mt).
I
a
' + 'r*(2)1. (8.16)
?Si"
\/2 lx*Q) x-Q) x-(1)
The reader should note that V$i" does not change sign under exchange of the
two electrons, while ?*pi,, does change sign under particle exchange. ?.ft" is
said to be symmetric and tp*,r, antisymmetric under particle exchange.
The foregoing ideas may be extended to more general functions by
defining the two-electron wave functions,
Here the lette r a or D represents the set of quantum numbers (r, /, m, ffir)
which defines a specific one-particle state described by both spatial and spin
coordinates. The integer 1 or 2 refers to the set of spatial and spin coordinates
associated with a given electron. Then, the symmetric and antisymmetric
linear combinations of these functions are,
I
t[
+ ?rr)
{z 0|ou
and (B.lB)
I
ltz
\/2
0t,ou - ,-r')
We now define a particle exchange operator, Prr, which exchanges the co-
ordinates of the two electrons. Thus:
From the foregoing we see that although the functions yo6 and yuo are not
eigenfunctions of the particle exchange operator, Prr, the linear combinations
defined by ?u and yn are eigenfunctions. A symmetric eigenfunction of Pp has
the eigenvalue f 1, while an antisymmetric eigenfunction has the eigenvalue
-1. Furthermore, if the operator P* commutes with the two-particle Hamil-
tonian, then the exchange symmetry of each eigenfunction is constant for all
time.
Operations with the particle exchange operator should not be confused
with those of the parity operator II which was discussed in section 10 of
304 SPIN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
I
V.t (poo-tyno):0,
{z
I
11s :- (l,oo*1poo):12rp"".
f,
It follows that the Pauli exclusion principle will be automatically satisfied if we
require that only an antisymmetric total wave function be used to describe a
system of electrons. The antisymmetric wave function for two electrons
(Equation B.18) can bewritten conveniently as the 2 x 2 determinant:
I
IP,q.:wr. (8.20)
Note that the rule for expanding the determinant insures that the wave function
is antisymmetric; the fact that the determinant is zero if any two rows or
8 See, for example, J. C. Slater, Qtwntum Theor\ of Matter,2nd ed. McGraw-Hill Book Co., New
York, 1968, Chapter ll.
IDENT|CAL PARTTCLES, EXCHANGE, AND SYMMETRY
columns are equal guarantees that the Pauli principle is obeyed. An alter-
native definition which is equivalent to Equation 8.20 is,
:
V,t
#, t r..Pr,r,...,.'[v"(
\'PuQ)''''t',(N)1, (8.2 r )
Fermion states:
y'f-particle spatial function x antisymmetric
f 1tyr"*.tric
| \ff-putticle spin function
Tz :
I lurrtirymmetric i[-particle spatial function x symmetric
( \ff-putticle spin function
Boson states:
To illustrate the above with a specific example, let us separate the spatial
and spin coordinates 'of the two one-electron wave functions, rpo and ga, vs
follows:
ga : vnx+ and !)o : vn,x+,
I
?spatiar : (2) I vn' 0)'p
"(2)1,
ft,lrp"(L)rp "'
where the function having the plus sign is symmetric and the one having the
minus sign is antisymmetric in the exchange of the electrons. Likewise, we can
write the spin functions:
x,*Q)'x*(2)
x,-Q) ' x,-Q)
I
'x-Q) + x-0)'x*Q)1,
6lx"(\
where the last function containing the minus sign is antisymmetric and all
others are symmetric. Therefore, four antisymmetric product functions may
be written as follows:
x*Q)
I
-tr)' ) - (I) -(2)l (2) + r-(1) 'r.(2)rl
ftfr rp 2 rt, rp
*,( *,
\'j_';: rrr
rrr_r'
l1 'r.Or.l
frfv"tr)
',t'*,(2) + !,*,(L)'p"(2)l
&[r*(l)
' x-( 2) - u-(1)
(8.22)
PROBI.EM 8-9
PROBLEM 8.'O :i'r,:1r:1! r r::1i;ll;;;Fi:t i.l;iI 1l:l: Il -'ii i il lf:iii ifiE i;i:: l; i ri:l:f;::::*:;;:;;; '::;t I Xii
PROBLEM 8.1'
If the Hamiltonian for the two-electron system commutes with the particle
exchange operator, then all four of the functions of Equation 8.22 are energy
eigenfunctions. This will occur when the Hamiltonian itself is symmetric in
the coordinates of the two electrons. By way of illustration, consider the follow-
ing Hamiltonian for two non-interacting electrons: *
where Vtisany one of the four functions in Equation 8.22. Although these four
functions are degenerate when there are no spin-dependent forces or inter-
actions between the electrons, we will see that the motions of the two electrons
are correlated with their spin states. Thus, for the S : 0 state (singlet state),
when the spins are antiparallel, the spatial eigenfunction is symmetric and the
electrons can come very close together. In fact, the probability density for this
state has a maximum value when the electrons are so close together that their
coordinates are essentially the same. However, for the S : I state (triplet
* Although this is physically unrealizable, we will even neglect, for the time being, the Coulomb
interactions between the two electrons.
308 SPtN, ANGULAR MOMENTA, AND |DENT|CAL PART|CLES
/ l-
singlet
t/
/' Z/ (3-fold delenerate, .ttturc 8-3 The effect of the Coulomb re-
t)r' -.Triplet 3,:'#:"a:
tb.EEFarue degeneracv for two
En + En,
(4-fold degenerate)
-J
state), when the spins are parallel, the antisymmetric spatial function goes to
zero as the electrons approach each other. Thus, in the triplet state the motions
of the electrons are correlated so that they avoid each other as much as possible,
while in the singlet state they attract each other. That is, they move as if they
were under the influence of a force that is repulsive for parallel spins and
attractive for antiparallel spins. This force is called the exchange force since it
has its origin in the symmetry requirements on the wave function under the
exchange of indistinguishable particles. The exchange force is strictly a
quantum mechanical phenomenon with no classical analog.
In spite of the introduction of this new force, the exchange force, the
functions in Equation 8.22 are still degenerate when the Hamiltonian is
written as in Equation 8.23. Suppose we now add one term to the Hamil-
tonian, namely, the Coulomb repulsion between the two electrons, *e,lrrr.
Although this term does not depend upon the spin coordinates, it is evident
that it will partially remove the excha4ge degeneracy of Equation 8.22. Since
the electrons are closer together, on the average, in the singlet state than in the
triplet state, the Coulomb repulsion increases the energy of the singlet state
more than it increases the energ'y of the triplet state (see Figure 8-3). The
energy difference between the singlet and triplet states is called the exchange
enetgt; it is really an electrostatic energy, but its very existence depends upon
the fact that the exchange force causes the aligned spin state to have lower
energy. Accordingly, it is energetically more favorable for the spins of electrons
in the same orbital state to assume a parallel alignment. It is in this sense that
we say the exchange force aligns the spins. This topic will be discussed further
in Chapter 10 in connection with the helium atom and ferromagnetism.
PROBLEM 8.12
PROBtErvl 8-r3
IONIZATION POTENTIALS
li 4r i sa i oo
Figure 8-d lonization potential for the outermost electron of each element. (From Introduction
to Atoiic Spectra by Harvey E. White. Copyright 1934 by McGraw-Hill Book Company. Used with
permission.)
Thus far we have considered only the coupling of the spin and orbital
momentum of a singie electron by means of the spin-orbit interaction. A
question naturally arises as to how one should proceed in the case of two
electrons where there are four constituent momenta to be considered. One
model, known as the j-j coupting modll, assumes that the spin-orbit interaction
dominates the electrostatic interactions between the particles. Thus, we write
jt : ft + .St and i, : L, + & for each of the two particles, where the
quantum numbers jo are obtained as before. Then the total angular momen-
tum is obtained by combining J, and Jr. That is,
j:jr*j,
andT : ljt * jrl, l,r'. * j, - 11,.. .,Ij, - jlr.
We will illustrate y-7 coupling by applying it to iwo inequivalent* /_
electrons. For each electron we found above that7, : j2 : * or 3. Then the
possible ways of combining these are shown in Table B-1. In a weak magnetic
Number of States in
Spectral a Magnetic Field
Ir iz i Terms (Number of mo Values)
3
z ,a z, 2, l, o (9, 9)r,z,r,o 16
3 1 ,1
o, r!1r
2 2 2t 212,1 B
,1 +
3
l, 0 (*, t)r,o 4
1 9 Ir
at /1 lt\ B
2 2 \!t 2-)Z.L
10 rtut"t 36 states
* Inequivalent electrons occupy different atoms or different orbits of the same atom. Hence,
they automatically satisfy the Pauli principle.
312 SPIN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
field, each_ state of a given 7 will split into 2j + 1 states corresponding to the
allowed values of m,.
Although yy coupling is used extensivety for the description of the nuclear
states observed in ttn.l.ur spectroscopy, it is not appropriate for many atomic
systems because of the strong electrostatic (and other) interactions between the
two electrons. The model that has been most successful in accounting for
atomic spectra of all but the heavier atoms is known as the Russell-Saunders
coupling scheme.e In essence, this model simply assumes that the electrostatic
inteiaction (including exchange forces) between two electrons dominates the
spin-orbit interaction. In this case, the orbital momenta and the spins of the
tio electrons couple separately to form E - it * F, a1d S : Sl1 + Sr. Then
the total angulai momentum is given by J : i + S, as before. For two
inequivalent"p electrons we have: (:2, !, or 0 and s : I or 0. Fot each /
utd s thej values are | / * sl, ll +s - 11, . . .,1/ -sl and for eachT value
there are-(2j + 1) values of m,. The combinations are shown in Table B-2.
Once again-there are 36 states in a weak magnetic field, although of course,
their energies are not the same as those in the Yy coupling scheme.
Supp6se we now subject the two electrons to an intense magnetic field
such that p6I is much greater than the spin-orbit interaction as well as all
interactions between the electrons. This will uncouple the momenta, regardless
of whether Russell-Saunders or 7-7 coupling holdt it *.uk fields, so that -tr,
ir, ^fr, and.i, will precess independently about d. The good quantum num-
Uits *itt no longei tr.7 and m but will be msr, /rt/r, tTtrrr artd mrr. The total
number of states-that is, the possible combinations of these quantum num-
bers-will be (21, * l)(21, + 1)(2t' + t)(2s, + l) : ?6, as before'
'discussed
We have thus far only inequivalent electrons, whereas equi'
ualent electrons are of great physical importance in atomic spectroscopy. !inc9
equivalent electronr o-".rpythe same orbital state on the same atom, the Pauli
Number of States in
SPectral a Magnetic Field
i Terms (Number of m, Yalues)
, I 3,2, I tDr,r,, l5
2 0 2
,D, 5
I I 2, I,0 3D
r 0,1, 2
9
1 0 I ,P, 3
I I tS,. 3
0
0 0 0 tso I
l0 rt"t t 36 states
e H. N. Russell and F. A. Saunders, Astrophls, J. 61, 33 (1925). For a more detailed discussion,
see J. H. VanVleck, ih, Throry of Ehctrii and Mignetic Susceptibitities, Oxford University Press, Oxford,
1932.
ANGULAR MOMENTUM VECTOR COUPLING SCHEMES 3I3
I I 2
I _l
2 22 0 tD,
I 0 2
1
-2
I ll 0 tD,
I -1 I2 _1o 00 0 ,D,
0 -l 2
1 _12
-l -l 0 tD,
I ,D,
-l -l 1
, -2 -2 -2 0
0 ++2rl ,P,
-l ++ror "P,
0 -+-+ol-r ,P,
-1 -+-*-ro-l
_+ _1, _2 _l _l
,P,
-l "P,
0 -++ll0
+ + o -t l
"p,
tp,
-l
+ -+
+-+ooo -l -1
o tP,
-1
tPo
-l
_l 2 tSo
exclusion principle puts restrictions on the allowed values of m7 and m,. Let us
now re-examine the two coupling schemes just discussed and apply the ex-
clusion principle.
In the case of the Russell-Saunders coupling we find that for two equivalent
y' electrons we must exclude the following terms i ,Dr.r,r, lPr, and 3Sr. The
exclusion of the 3D terms is easily understood, since a D term requires that
/ :2, which means that ms, and, ms, must be the same ; then the Pauli
principle permits only a singlet spin state (opposite spins), or a tD term. Simi-
larly, the 3S term requires that ms, - rnh: 0 and that the spins be parallel,
which violates the Pauli principle. The exclusion of the lP term is not quite so
obvious. However, if a table is constructed such as Table B-3, which shows all
of the allowed states, every state is accounted for by the tDr, tPr,r,o, and 1,So
terms. There are no additional distinguishable states which can be attributed
to a rP term. In a weak magnetic field these five states split into 15 states. It is
easy to see why only l5 of the original 36 states for two inequivalent electrons
remain. Six of the 36 states violate the Pauli principle, and of the remaining 30
states only half are distinguishable.
When applying 7y coupling to two equivalent y' electrons, note first that
forJ, : jr:8 we must exclude the following cases:
,'-3 ,'-s ;-q.
Jt-Et J2-Vt J -.)
'jt:*, jz:*, j:I
jr:t, jz:t, j--1.
The first case is obvious since it would require parallel orbital moments and
parallel spins, whiCh would violate the Pauli principle. The third case would
require lt : lz : 0, as well as parallel spins. The middle case is not so
3t4 sPlN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
5 states 15 rt"t*
obvious, but as in the case of the lP term in the example from Russell-Saunders
coupling, it can be shown that no additional distinguishable states are available
to prorrlae for the 7 : I level when jt : j, : ,a. Also, the states given by
jr:8, jz: $ are indistinguishable from those given by.tt : t, jr: ,a. Th91,
for .q"i"alenl electrons, the distinguishable allowed states are shown in Table
B-+.
The above coupling with the empirical rules of
schemes, augmented
Hundrlo enable one to account for many features of the atomic spectra of
many-electron atoms without actually solving the many-body problem. A
quantum mechanical calculation is necessary, however, to obtain the energies
of the states and to verify Hund's rules.11 Hund's rules for the ground state of a
collection of equivalent electrons rrray be summatized as follows:
To see how these rules are applied, consider the ground state of samarium,
which has 6 electrons in the 4/shell and no other partially filled shell. Since the
4/ shell can hold 2(2 .3 1) 14 electrons, is less than half full in Sm,
+ : it
and all six spins are parallel. Thus, s :
3. A possible set of different m7 ualues
0; -!, -2, which combine to give tnl :3, corresponding to / : 3;
is 3, 2, 1,
that is, an F teim. By the third rule above, i -- / - r :-0, and the term
symboi is tFo. The superscript is the multiplicity,2s * 1, while the subscript
is the valffii*a
example, consider the ground state of iron, which has - 6
electrons in the 3d shell, the only partially filled shell. Since this shell can hold
2(2 .2 + l) : 10 electrons, it is more than half full. Then 5 electrons have
parallel spins and the sixth must be antiparallel, producing a total spin of 2.
ih" *ruuln.t of the 5 parallel electrons are 2, lr 0, -1, -2 which combine to
for example, G. Herzberg, Atomic Spectra and Atomic Structure, Dover Publications, Nbw York,
10 See,
1944,or H. n. Wnii", introductibn to Atomic Spictra, McGraw-Hill Book Co., New York, 1934.
it E. U. Condon and G. H, Shortley, The Theor2 of Atomib Spectra, Cambridge University Press,
Cambridge, 1953.
THE LANDE s FACTOR AND THE ZEEMAN EFFECT 315
zero; the sixth electron has a momentum of 2. Hence, { : 2, which corre-
sponds toaD term. Fromthethirdruleabove, j: I + s --+,andtheground
state term for Fe is 6Dn.
We will close this section by merely stating the selection rules for electric
dipole transitions which are appropriate for the two vector coupling models.
More will be said about selection rules in section 6 of Chapter 9. It is assumed
that only one electron at a time makes a transition.
Russell-Saunders coupling :
For the electron making the transition, L^l : tl, (that is, the parity
must change).
For the atom as a whole,
As:0
L,/:0, +l
Lj:0, *l (butatransition'from j:0 to7:git
forbidden)
Lm, : g, f I (but if Aj - 0, a transition from ffii : 0
to m, : 0 is forbidden)
j-j coupling:
For the electron making the transition, L^/': tl (parity must change),
and A7:0, tl.
For the atom as a whole,
We have previously obtained the result that the orbital and spin contri-
butions to the magnetic moment are given by (section 9 of Chapter 3),
+ke,*
lrr: -ntE: -g'Pu{@i
and
+g'oi
Itrs: - #_S: -g,prrlr1r + ji,
where gt : I and 3, : 2.002+ ,--',l. Now, when i and S are coupled we have
J:i+S
and
I
lr.. cos 0, s)
: Ferr
+
I
f,cz COS 0, L)
Figure &.5 The efflective magnetic moment of an atom is the comPonent of y' which lies along
the direction of the total angular momentum.
It is evident from the last two expressions that the total magnetic moment is not, in
general, colinear with the total angular momentum. This is illustrated in Figure 8-5.
Since E and,,S precess about i, it is apparent from the figure that y' also pre-
cesses about i. However, the efectiue magnetic moment, that is, the component
of p along J, maintains the constant value,
Fi: rt.i
lJl
: _Fat.J +zS'J
h tJt
puL'(E +S) +23'(i +S)
h tjt
palr+2Sz+3i.S
h tJt
irt:-
pa ljl ftj' + ,S' - i'\
h \W)
lpil-spu{ju+\. (8.27)
Note that for zero spin, Equation 8.26 reduces to the classical case of g
in a similar fashion 3 : Z It t: 0. - l;
Now it is possible to account for the so-called anomalous Zeeman effect
mentioned in Chapter 3. In a weak magnetic field, the angular momentum
will precess about I such that the projection of J along the"field direction will"/
be one of the allowed-values,-m,h. Thi correspondinginagnetic moment along
the field direction (taken to be the z-directio"; *itt lh.t i.,
Itrz : -9\amr
having a magnetic dipolar energy of
E: gmtpa%. (8.28)
The feature that distinguishes Equation 8.28 from the case of the normal
Zeeman effect is that here we see
that g is no longer unity, but depends upon
the quantum numbers /, s, and7. By way of illustiation lei us calcuiategfor an
electron in a p state and an s state:
orbitalstate / j g
p 134
LE5
p 112
L25
J 01r2
318 sPtN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
(4s, j) g:o g +o m;
J
I
t
(',;,;)
-tI
J
2
I
t
(t,rt,rt) _,L
('' l'l)
12345678910
Figure 8J Zeeman splittings for p and s states.
In a magnetic field 0, such that p,u98 is less than the spin-orbit energy, j and
miare good quantum numbers and the energies of the states split, as shown in
Table B-5. These splittings are shown schematically in Figure 8-6, and the
allowed spectral lines due to electric dipole transitions are shown in Figure B-7.
The lines corresponding to transitions in which m changes by +1 are labeled
o, and those corresponding to Lm: 0 are labeled z. The z lines are plane
polarized with the direction of polarization parallel to the field. The o lines
p 33+ ,
p 31L
'23
z23 5
p 3
z-23 _1 !
334
-5
p t-25
112
-,
1
p 823 3
p r12 1
2-23
J a1,
d6- I
1 _1 ,
.t oq -1
t
THE LANDE g FACTOR AND THE ZEEMAN EFFECT 3t9
-3 -lE,l
Energy (units
"rl u'a)
Figure 8-7 The spectrallines.resulting from the electronic dipole transitions shown in Figure 8-6.
lf the emitted light is observed perpendicular to the magnetic field, all of the above lines are seen
(transverse Zeeman effect). However, if the emitted light is observed parallel to the field only the
o lines are seen (longitudinal Zeeman effect).
are circularly polarized when observed parallel to the field, and linearly
polarized (perpendicular to the field) when observed at right angles to the field.
The selection rules for these transitions will be derived in the next chapter.
It is evident from the foregoing that the so-called "anomalous" Zeeman
effect is, in reality, what would normally be expected for an electron having
half-integral spin in a weak magnetic field. The "normal" or classical Zeeman
effect discussed in section l0 of Chapter 3 cannot occur for a single electron in a
weak field because of the spin term in Equation 8.26. However, in atoms in
which the spins are paired so that the total spin is zero, the g value for all
spectroscopic states is the classical value and only three spectral lines are ob-
served. The pair of lines whose energies are shifted by tpag are circularly
polarized, and the unshifted line is plane polarized.
fn a sufficiently strong magnetic field such that the spin-orbit splitting
is less than the magnetic splitting, the anomalous Zeeman effect converts to the
normal Zeeman effect. This phenomenon, known as the Paschen-Bach effect,
can be explained in the following manner. When the spin and the orbital
magnetic moments interact with the applied field more strongly than they
2.0 2.5
-t3 -, -L
ttsS _
(AE;
Figure 8-8 Energy level splittings in a magnetic field for the 2p states associated with j : 3/2 and
i :112. The zero-field fine structure splitting (AE)o is largely due to the spin-orbit interaction. The
weak-field Zeeman effect occurs for values of I below the point where the levels would first cross.
The Paschen-Bach effect occurs for very large fields where the splitting produces five levels that are
nearly equallyspaced. ln the latter case, three spectral lines are seen as in the normalZeeman effect.
320 SPIN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
interact with each other, E and,S are effectively uncoupled and they precess
independently about fr. The good quantum numbers are then nt1 vrtd flt*
insteid of 7 and mn which hold in weak fields. Hence, in strong fields the
projections of the magnetic moment on the field direction are proportional to
mt * 2m, rather than m,. This is illustrated graphically for / electrons in
Figure B-8, where the energy level splittings are shown as a function of applied
mignetic field strength. -Using the selection rules, L*, - 0, :tl and Lm, : 0,
only three spectral lines are observed, as in the normal Zeeman effect, fot
transitions from the 2p states to the 2s states in high magnetic fields.
PROBTEM 8.I5
PROBLEM 8.'6
(u) Show that in the vector coupling model one may define
an operator,
i.S:t(J'-Lz -s').
This operator will be used later to define the spin-orbit
operator (see Equation 8.8).
(b) Does this operator commute with J'rL', and 52?
PROBLEM 8.17
PROBtErtl 8.r8
(u) Consider a system for which ff,, L', Lu, S', and S,
form a set of mutually commuting operators. What
are the good quantum numbers for the system ? Write
a possible eigenfunction in symbolic ket form. With
these kets as a basis, which operators have diagonal
matrix representations ?
(b) Now assqme that a spin-orbit interaction is turned on
so that the Hamiltonian of the system becomes if -
tr o I n.d ' ^f. Evaluate the commutators, |,tr ,L,]
ls See American Insti.tute of Phlsics Handbook,2nd Ed. McGraw-Hill Book Co., New York, 1963,
page 5-222, Table 59-29.
EIGENFUNCTIONS OF COUPLED ANGULAR MOMENTA 321
5x5
ilx
9 x 9 matrix
I22 SPIN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
In order to diagon alize these matrices completely we must obtain the simul-
taneous eigenfunctions of J2 : (Jr * Jr)2 and Jr: Jy I Jzu. Our task,
then, will be to express the new eigenfunctions in terms of functions which we
already know, namely, ljrmr), the eigenkets of Jl and Jt", and ljrmr), the
eigenkets of J! and J2". Since the coordinates of J, and J2 are independent in
the old representation, the eigenkets may be written as the product functions,
These kets form a complete, orthonormal basis so that we may define a pro-
jection operator,
fnyffl2
' \IILL ) (Vtt'*''<i'*'t) - 1' (B'2e)
Let us designate the new eigenkets by jt jrj*), where I
lj,j,jm) : (8.30)
mym2
SinceT, and jrare fixed numbers for a specific calculation, the notation may be
simplified by omitting them from the kets. It should be remembered, however,
that the maximum values of m, and m, are j, and jr, respectively, and that the
maximum value of m is J, *Jr. Let us also include the requirement that m
always be equal to the sum ???1 I mz. Then we have
ljm) : (8.31)
mL+rmz:rm
The scalars, (mrmrl j*>, are merely coefficients that tell how much each old
eigenket lmrmr) contributes to the new eigenket ljm). These coefficients are
called the angular momentum coupling coeficients or the Clebsch-Gordon coffiients.
They are not difficult to evaluate for small values ofTr and j2, but the general
expression for them is quite complicated. Many of the Clebsch-Gordon
coefficients are available in tabular form.l3 We will illustrate how they may be
obtained by means of the angular momentum raising and lowering operators
for a few special cases.
' rs See E. U. Condon and G. H. Shortley, op, cit.; G. J. Ntjgh, A. H. Wapstra and R. van
Lieshout, Nuclear Spectroscop2 Tables,North-Holland Publishing Co., Amsterdam, 1969; M. Rotenberg,
R. Bivins, N. Metropolis, and J. K. Wooten, Jr., The 3-j and 6-j S2mbols, Technology Press, Cambridge,
Mass., 1959.
,j t.
'r \'.1
ll, 1), 11,0), 11, -1), 10, l), 10,0), 10, -l), l-1, l), l-1,0), l-I, -l).
(8.32)
The new kets, ljm), which are primed for convenience when 7 and m are
replaced by numbers, are:
r ; /,/,t,t
For ;l_ i
\,
j : 2: 12,2)', '2, l)', 12, 0)', '2, -L)', ',2, -2>'
j : l: ll, l)', ll, 0)', ll, -l)'
j : 0: 10, 0)'.
Note that state |12,2)' - ll, 1), since they both correspond to the case where
tn1 : t7t2: I and mL and mz are parallel. Likewise, lr2, -2)' - l-1, -l).
Thus, we can find all of the states for j : 2 by multiple applications of the
lowering operator on |12,2>' or by using the raising operator on 12, -2>'.
Choosing the former approach, we utilize Equation B.l I and proceed as follows:
\/Wlz,t)':/ml',t)
+/mll,o)
r>' : r) + ll, o)). ;;;
12,
+r00, 1,,r,
l ;r, :;;
:
fiV* -r Jr-)10, t) +
J-12,1)' 11, o))
@12,0),:#t/@t-r,t)
+/mlo,o) rtt rv,r Jt _n,
+/mro,o) l
+rZ6
Jz 'L fir Ji -.rrl
11,-l)l
12,0>' : I
(l-t, l) + 2 10, 0) + ll, -t)),
and in like manner,
-vb
-l)' :
I
12, .r/ru}, -1) + l-1, 0)),
Notice that all of the nine original kets have been used in forming the first
five primed kets. It turns out that all new kets having the same m value will be
linear combinations of the same group of old kets, each of which satisfies the
condition mr * rt2:7ys. We thus choose 11, l)'to be the linear combination
of 10, l) and ll, 0) which is orthogonal to 12, l)', namely,
12,2>' ll, 1) I
12, t>' 10, 1) U \/,
ll,0) tl\/,
12,0>' 11, -l) U \/6
10,0) 2l\/6
l-1, l) Ll\/6
12,
-l>' 10, -l) U\/2
l-1,0) \ \/,
12, -2>' l-1, -l) I
ll, 1)' 10, l) rl\/,
ll,0) -rt\/,
ll, 0)' ll, - 1) It\/,
10,0) 0
l-1, 1) -U \/,
11, - l)' 10, -1) tl\/2
l- t, 0) -rt\/,
10,0)' ll, -l) tl\/3
10,0) -t l\/i
l-1, l) tl\/g
EIGENFUNCTIONS OF COUPLED ANGULAR MOMENTA 325
t.*2b+d:0
a-c-0.
since these must be simultaneously satisfied we obtain,
a:-b:C.
Therefore,
0)' : I
10,
ilfll' -t) - l0' 0) + l_1, 1)).
PROBLEM 8-'9
Show that the nine kets obtained above are eigenkets of both
Jz - (Jr * Jr)z and Ju: Jru * Jr".
Write the six state functions ljm) for a single y' electron in
terms of the kets lmt, m,) - l/, m7) ls, m,). (Answers in
Appendix C of this Chapter.)
PROBLEM 8-2'
PROBLEM 8-23
The method of the previous section is readily applied to the task of ob-
taining the spin wave functions for two or more particles. Let us first consider
the case of two identical particles of spin $, where we will now label the spinors
of Equation 8.2 with a I or a 2 in order to distinguish the particles. Thus,
X*0) means that particle I has its spin up and 7-Q) means that particle I has
its spin down. In ket form these spinors are written as lrtrr, ffirr) - lsrfl?rr) lsrmrr),
so that we have for two identical spin f particles:
:
l1 1\
lZt 2/ r+Q) ' x*Q)
l+, -+) : x+(t) ' x-Q) (8.33)
I 1 1\ :
l-2t 2/ x,-(L)'x*Q)
l-+, -+) : )L(r) ' x-(2).
If we consider the total spin operators 52 - (S, * S2)2 and S, : Sr, * Sr,,
it is easy to show that the kets of Equation 8.33 are not eigenkets of 52, although
they are still eigenkets of S,.
PROBLEM 8.21
PROBLEM 8.25
Given that the kets of Equation B:33 are eigenkets of Sf, S!,
51" and S2,, show that they are not all eigenkets of 52 :
(S' * Sz)'.
Now let us find the spinors ls, m,) defined in the product spin space. This
is a 4-dimensional space containing a 3-dimensional subspace corresponding to
r :1, and a l-dimensional subspace, for s:0. The new kets are 11, 1)',
11,0)', ll, -1)', and 10,0)'. (Primes are included here although these kets
APPLICATION TO TWO-PARTICLE SPIN FUNCTIONS
The last ket, 10, 0)', must be orthogonal to ll, 0)'. Therefore, we have:
_ 11 1\
ll,1)' - l?t 2/
ll, -1)' _
-
I
f -Et
1
-2/
1\
10,0)' : l
+) - l+, -+)).
ft,(l-+,
The first three eigenkets in Equation 8.35 correspond to the triplet state (s : 1),
while the last is a singlet state (r : 0).
PROBLEM 8.25
For two spin-one particles there are nine product functions that can be formed
from the functions given in Equation 8.36. These are: )t*0) ' 1G(2), U*(1) '
Xo(2), )GQ)',GQ), Zo(1) 'X*(2), and so on. The appropriate lindar combin-
ations of these product functions which form simultaneous eigenfunctions of
Jz and J, have already been obtained in the previous section for the case of
jr:jr:I.
PROBLEM 8.27 s.i+,q,##
ji{:i;.fi"i8i,,$1s
(u) Write the expressions for the nine basis functions for
two spin-one particles which provide a diagonal
representation for J2 and Jr.
(b) Verify that these functions are the correct eigenfunctions
of J2 and J r.
ffi.t?,Ei8'g;tF,:l{,1,E3.'s.i$i8
emorecures
: ffi#illTrin)l*oto'. . .
[e(spatial) e(spin)]"r",t,oo, (8.37)
We already have seen that the two-electron wave function must be antisym-
metric in the exchange of the two electrons. Likewise, the two-proton wave
function must be antisymmetric in the exchange of the protons, since protons
are also Fermi particles. Then the possible proton functions have the same
form as the four electron functions given in Equation 8.22. Hence, ?molecule
has 16 differentforms, each of which is antisymmetric in the exchange of two
identical particles.
Let us now suppose that we are performing an experiment that would
detect a change in the spatial state of the protons. For example, we might
be looking at the infrared spectrum of the molecule. Since the interesting
electronic transitions are in the ultraviolet or visible spectral regions, we can
essentially ignore the electronic part of ?morecur" in the analysis of this experi-
ment. The principal contributions to pproron (spatial) are from the trans-
lational, rotational and vibrational states of the dumbbell-shaped molecule.
We will ignore the translational states since we are interested in changes of the
internal energy of the molecules. The vibrational states will be ignored for
APPLICATIONS TO SYSTEMS OF IDENTICAL PARTICLES 329
two reasons. First, they are symmetric in the exchange of the protons; second,
they require rather large excitation energies. * Our molecular wave function
now reduces to
?moleoule : ?protonr(rotation)' ?p.otoo.(spin),
which must be antisymmetric. The rotational states are the spherical harmonics
which are characterized by the angular momentum quantum number ;1,
where
?nrotoos(rotation) : Vt(l)''P,(2),
and
D
-iu + \h'
Lr:7,
z{ being the moment of inertia of the molecule. Since for a dumbbell molecule
the exchange of the two protons is equivalent to an inversion through the
origin, the symmetry of each state is given by the parity, (-l)r. This leads to
the important result (for hydrogen) that only odd aalues of j are allowed when the
proton spins are parallel and onll euen ualues of j are allowed when tlze proton spins are
antiparallel.
Hydrogen molecules which are symmetric in the nuclear spin are called
ortlto hldrogen, while those that are antisymmetric are called para lrydrogen.
Assuming the same statistical weight for all states and a uniform distribution
over the rotation states, there should be three times as many ortho molecules
as para molecules. This follows from the fact that there are three symmetric
spin functions which can be used with each odd rotation function, while only
one antisymmetric spin function can be used with each even rotation state.
This surprising prediction is confirmed experimentally for hydrogen at ordinary
temperatures. The infrared spectra consist of bands of lines corresponding to
odd-odd or even-even transitions, in keeping with the fact that transitions are
forbidden between states of different exchange symmetry (see Equation 8.24),
and each line associated with an odd-odd transition is roughly three times as
intense as each line in the even-even band.
At lower temperatures, however, when &Z becomes comparable with the
energy intervals between rotation levels, the relative populations of the rotation
levels are greatly altered. In fact, the system tends to go completely into the
para state-corresponding to zero rotational energ/ at about 20'K. This
ortho-para conversionf is an amazing verification of the role of symmetry and
statistics in quantum processes. Since the excited states of para hydrogen lie
lower than the corresponding ortho states, the heat capacity at low tempera-
tures is much greater for para hydrogen than for ortho hydrogen or for any
mixture of the two forms. So-called "ordinary" hydrogen is a mixture whose
concentration of para molecules can be as low as 25 per cent at room tempera-
ture and as high as 100 per cent at low temperatures.
* See section I of Chapter 5, particularly Problem 5-18.
f The conversion process is slow except when catalyzed by adsorption on carbon. The carbon
dissociates the metastable ortho molecule; when the atoms recombine at low temperatures they form
the stable para molecule.
330 sPtN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
PROBLEM 8.28
PROBLEM 8-29
Apply the particle exchange operator to the nine kets for two
spin-one particles (see Table 8-6 in section 10) and show
that the ratio of the number of ortho to para states is 2 :1.
PROELE'T4 8.30
PROBTEM 8.3I
SU'YIMARY
In this chapter we have seen how spin can be incorporated into the
Schrodinger formalism by postulating spin operators which act upon spin
functions in a manner analogous to the angular momentum operators. When
the spin and spatial coordinates are independent, the total wave function is a
product of the spin function and the solution to the spin-independent Schrodin-
ger equation. If any spin-dependent forces are to be included in the Hamil-
tonian, they must be expressed in terms of the spin operators. In particular,
the spin-orbit interaction is discussed and it is shown that it can be expressed in
terms of the operator scalar product t' S. The vector coupling models are
then introduced and applied to simple systems. The product space formed by
coupling angular momenta is described and a procedure given for obtaining
the eigenfunctions in the new space.
The concept of the exchange degeneracy of indistinguishable particles was
introduced. It was shown that a satisfactory wave function for two or more
identical particles must be either a symmetric or an antisymmetric linear
combination of the possible product states. This derives from the requirement
that the wave function be an eigenfunction of the particle exchange operator
and that its probability density remain unchanged under an exchange of
particles. The experimental connection between particle spin and the ex-
change symmetry of wave functions is summarized by the statements: a
collection of bosons (identical particles having zero or integral spin) is de-
scribed by u wave function that is symmetric under the exchange of any two
particles; a collection of fermions (identical particles having half-integral spin)
is described by a wave function that is antisymmetric under the exchange of any
332 sPtN, ANGULAR MOMENTA, AND IDENTICAL PARTICLES
two particles. Fermions tend to aaoid the same quantum state, while bosons
tend to condense into the sarne quantum state. Furthermore, the energy density
of a collection of fermions is greater than that for a collection of the same
number of bosons. It is in this sense that bosons attract one another and fer-
mions repel each other via an exchange force which has no classical analog.
These concepts are shown to account for the infrared spectra of diatomic homo-
nuclear molecules and for the electronic exchange energy which separates singlet
and triplet states in atomic spectra.
CHAPTER I APPENDICES A to C
n; !8*
Then the total number of ways of filling g, states of energy eo with n, fermions is
simply the number of combinations of g, identical states taken nn at a time,
gt!
ffit,
and
W-TTffi
This is the distribution given in Problem B-32.
For bosons the same combinatorial argument cannot be used, since a
state is no longer simply occupied or unoccupied but can contain any number
of particles. Using a somewhat different approach, the distribution of Problem
8-33 is obtained.ra
A more sophisticated method of deriving the distribution functions is to
use the grand partition functionls
Z_ 2 e-nto,
d
(A1)
Z-l*e-",
1r R. B. Leighton, Pinciples of Modern Ph2sics, McGraw-Hill Book Co., New York, 1959,
Chapter 10.
16 T. L. H:ill, An Introduction to
Statistical Tlurmodymmics, Addison-Wesley Publishing Co., Reading,
Mass., 1960, section 22-1.
333
334 APPENDICES A TO C
: - ln7,: e-d I
tlo:-kT*t" Z:-a iA2;
dAr anlnz- r + -': e'+l'
Defining the Fermi energy by .o .1r, this becomes
I
fi;:F' (A3)
.*p\ l,r )+r
For bosons the grand partition function becomes
I
Z- Ia s-o * e-'o * e-"' + "' : L-e*
Then,
n _*hZ:ft.:*:#
rti
-
-
-.
(A4)
'*P\ rr )-L
14/:ff-r- nr!(g,
',3n! .'.
- nn)!
PROBIEM 8-33: ]ri,! li;i;j:;rriill.;t4' .ijre'iJiir.i;tu.{'ii i'.i;i+.dur'9'li"ijlili3, :;',ri.xie,rll,r+.i*ill1
tr[-:nffi
_ _ l)! (_ gi
APPENDICES A TO C 335
Amelnc,an Institute of Ph\sic.s Handbookr 2nd ed., McGraw-Hill Book Co., New york, 1963,f"StiZi-i',,
p.7-14.
336 APPENDICES A TO C
The results of these problems will be ne eded for problems in later chapters.
PROBLEM 8.20
PROBLEM 8.22
(u) On the basis Il, +), ll, -+), 10, +), 10, -+), l-1, +),
- 1, -+),
l: _? ;
l
: ::\
' :;(l 'l
_4
s h2
l ,,'4'-:il
(b) On the basis ordered as in Problem 8-20 above,
,s :r(
;,, :,)
338 APPENDICES A TO C
PROBLEM 8.27
SUGGESTED
REFERENCES
David Bohm, Quantum Theorl. Prentice-Hall, Inc., New York 1951.
E. U. Condon and G. H. Shortley, The Theorlt of Atomic Spectra. Cambridge
University Press, Cambridge, 1953.
R. H. Dicke and J. P. Wittke, Introduction to Quantum Mechanics. Addison-
Wesley Publishing Co., Inc., 1960.
P. A. M. Dirac, The Principles of Quantunt Mechanics, 3rd ed. Oxford University
Press, London, 1947.
A. R. Edmonds, Angular Momentum in Quantum Mechanics. Princeton University
Press, Princeton, 1957.
R. M. Eisberg, Fundamentals of Modern Ph2sics. John Wiley and Sons, fnc.,
New York, 1961.
W. R. Hindmarsh, Atomic Spectra. Pergamon Press, New York, 1967.
R. B. Leighton, Principles of Modern Phltsics. McGraw-Hill Book Co., Inc., New
York, 1959.
E. Merzbacher, Quantum Mechanics. John Wiley and Sons, Inc., New York,
r961.
Albert Messiah, Quanturn Mechanics. North-Holland Publishing Co., Amster-
dam, 1958.
D. Park, Introduction to the Quantum Theor2. McGraw-Hill Book Co., Inc., New
York, 1964.
J. L. Powell and B. Crasemann, Qunntum Mechanics. Addison-Wesley Publish-
ing Co., Inc., Reading, Mass., 1961.
M. E. Rose, Elementarlt Theor2 of Angular Momentum. John wiley and sons,
Inc., New York, 1957.
D. S. Saxon, Elementarlt Quantum Mechanics. Holden-Day, Inc., San Francisco,
1968.
Leonard I. Schiff, Quantum Mechanics, 3rd ed. McGraw-Hill Book Co., New
York, 1969.
John C. Slater, Qunntum Theory of Matter. McGraw-Hill Book Co., New
York' 1968.
H. E. White, Introduction to Atomic Spectra. McGraw-Hill Book Co., New York,
1934.
Robert L. White, Basic Quantum Mechanics. McGraw-Hill Book Co., New
York, 1966.
CHAPTER 9
ewwffiffiKMM&wffiffiw
ffiffiffiw##ffiffi$ effiffi
&ffipkrffiewt&ffiffi
The easiest and most direct approach for solving a real physical system is
to regard it as a modification of one of the model systems whose solutions we
already know. In particular, we wish to express the Hamiltonian for the real
system, ff, as the sum of .f o, the idealized Hamiltonian whose solutions we
know, and an additional part Jt'ftt which contains the new interactions. If
it'o, is small comparedto ff0, the corrections to both the eigenfunctions and
I E. Schrddinger, Ann. Physik 80, 437 (1926).
339
APPROXIMATION METHODS AND APPLICATIONS
the eigenvalues resulting from Jt'o, will also be small. In such cases it is possible
to use the methods of perturbation theory.
We begin by writing the Hamiltonian for the real system as
where the rpjo) form an orthonormal set such that (i l.t) : dor. The equation we
wish to solve is
(tro * ldf ttt)rln : E,gn. (e.3)
It is assumed in this section that both the rpjo) and the q)n are discrete, non-
degenerate sets of time-independent functions. We further assume that there
is a one-to-one correspondence between the members of each set. Thus, rp,
goes into ?lo) us ,1 varies from I to 0, Vz + yft, etc. Stated mathematically,
In keeping with the spirit of these limits, let us represent the new eigenfunctions
and eigenvalues by the following power series in )";
:
tpn vlf) + tvll' + |ry,l?' + . . .
(e.5)
En : Ell' + lEf' + ffEf) +'''
It is assumed that successive terms of these series get smaller so that the series
converge. Our task will be to find these corrections to the eigenfunctions and
the eigenenergies to any desired order. Substituting Equation 9.5 into Equa-
tion 9.3, we obtain:
-- Elf'vf' + + nf',pt')
^(El!"tll'
+ 1, (Ef, rp'," + ELt' rp'"t' * E!i, rpli,) + . . .
Ot,
(trorpll' - El!'rpll') + i( i/f$\v(ot * troy,ll' - El,"rprct - E!f'rpl)')
+ |r(trorpllt * .ff<'trft) - El!,rpl"r, - Ef,rp*, - Ef,rpl,o,) +...- 0.
PERTURBATION THEORY FOR STATIONARY STATES 3{l
tr orpll) : El!'rp'f'
*tttt r<ot * ffi oy,*t : EllrPl,o' + Elf"pll' (e.6)
tr rrp'," +,ff"' rttll' : E!,1' rpl!' + Et' rpt' + n!," tplf'
:
The first of these is simply Equation 9.2, our starting point. The second may be
solved by expanding the first-order correction to the wave function in terms of
the unperturbed eigenfunctions. That is,
,Pll' :|o,ny,lo'.
The quantity (kl lf trt lz) should be recognized as the matrix element, "tr\\,
of the operator Jf o) with the unperturbed eigenfunctions as the basis. Then
we may write,
,tt') * a*o(Eft - Elf') : Elnil 6rn. (e.7)
In words, Equation 9.8 says: tltc frst order correction to the energ) of the nth eigen-
state is the diagonal matrix element corresponding to tlte nth row and tlte ntk column of the
matrix of the perturbation. Note that the unperturbed wave functions are used for
this calculation. For ft t' n, Equation 9.7 becomes
of,
,trf) + o,*(El,') - Eto)) - 0,
,trf)
Qntc:Mr. (e.e)
342 APPROXIMATION METHODS AND APPLICATIONS
Equation 9.9 tells us how to find the first order corrections to the wave function
for the nth state. Thus,
Each function that has a nun-zero matrix element with Vft contributes to the new
basis
waae function for the nth state. Hence, when the matrix of the perturbation has
non-vanishing, ofldiagonal elements, this means that the perturbation has
produced some interference or mixing of the original wave functions so that
they are no longer orthogonal. Note that the amount of this mixing depends
not only on the magnitude of the matrix element in the numerator, but also on
the "energy denominator." Thus, states that are widely separated in energy
are not expected to interfere to any large degree. States that are close together
will be expected to have a large amount of mixing. The denominator of
Equation 9.9 can never be zero, since we have excluded the possibility of any
degeneracy in this derivation.
To obtain the second-order corrections we will use the last expression
given in Equation 9.6,
!,llt : 2 b^ftt\ot,
i
,pflt : o*orp[o'.
7
Making these substitutions,
2b*oE'ftdr,
niii
* 2onotrL\'
b**El!'+ f amffl,|.' : bnoElft + on*Ell' + E'n"6on,
ofr
b,*(El!' - Ef') + > o*otrL\' - an*Efl' : Ellt6*n. (e.1 1)
PERTURBATION METHOD TO NON.DEGENERATE STATES 343
i
= 2 eno'trr')'
i+n
Then,
r_(2) s
un trr*t, <1 l,trtt)|,
tr:L,
(e.12)
- o?nEf' - E\ot - o?"EY' - E\ot '
H.:
where the results of the first-order theory expressed by Equations 9.8 and 9.9
have been used. The second-order correction to the energy has been obtained
using only the first-order wave function as expressed by the expansion coeffi-
cients arro.
To obtain the wave function to second order, we return to Equation 9.11
and seek solutions for k t' n. Then,
and,
where the expansion coefficients are given by Equations 9.9 and 9.13. For this
reason, perturbation theory is rarely used for wave function corrections beyond
the first order and for energy corrections beyond the second order. In principle,
however, higher order corrections can be obtained by proceeding to solve, in a
similar manner, the equations corresponding to higher powers of ,1 as in
Equation 9.6.
with energy eigenvalues given by Elf' : ah.(n + +). The unperturbed energy
eigenfunctions of this Hamiltonian are tabulated in section 7 of Chapter 5.
Now if we add.an anharmonic term of the form bxs to the potential, the
Hamiltonian becomes
itr : & * ikx, * bxs : tro * it2(t),
2m
This result follows immediately from parity arguments without performing the
integration. Since each oscillator state has definite parity, the square of any
state has even parity. For an odd perturbation, then, the integrand is odd and
the integral over symmetric limits will vanish.
When the first-order perturbation vanishes, it is necessary to proceed to
second order in order to find the energy correction. A glance at Equation 9.12
tells us that we are going to need the ofldiagonal matrix elements
It will be shown below that when (n i 'is the unperturbed ground state wave
function there are only two non-vanishing matrix elements, namely,
3b
ff["t :
t/Bu"
and
b
,ff[tt = 13
tVo.r'
' Then,
ll
83":- Busafi.
LT b2
Buzk'
b2
and
ll b2
Eo:irh- B"'k'
PROBLEM 9.1 :"*1. -: ; l"i;;. i; : : :. :" ll,"l;: f ,:l lliri; ;;:,:i :il1lt; i;,ii ;: :i i:1 J:ri
We will now use the raising and lowering operators developed in Chapter
5 to obtain the matrix of the perturbation. This will enable us to simply write
down the energy correction for any level without performing the integrations.
Recall that the unperturbed Hamiltonian for the harmonic oscillator can be
PERTURBATION METHOD TO NON.DEGENERATE STATES 345
q:#@ * oI),
from which we obtain
I
q" : + (ot)'+ a'aI I ata2 * a(al)z + (al1za I aata I aIaaIl.
2tilo"
(e.1i)
It is readily shown that:
{ap' I q'r) : fM6n,,n+s
+f@6n,*_s (9.18)
* 3(n + 1) fn + | 6n,,n+L * Snfi A*,,,_r.
PROBLEM 9-2
alln>:'y'n+llnll),
and
a ln) : \/; ln - I).
We can readily construct the matrix tlt\ from Equation 9.18, srnce
PROBLEM 9.3
PROBLEM 9.1
PROBLEM 9.5
PROBLEM 9.7
PROBTE/I4 9-8
llc
E, - ,Lwh *#,(' 2uzah)
PROBLEM 9.IO
(n"r.: EJ'):r(#)*)
PROBLEM 9.' I
/ ,^\
(e"*.: E[') - Eltt - 0; Et" : El" : -; e282 \)
PROBLEM 9-T2
/eE-,-\- \/rl,
(e"r. i en: ln) * #W-*l\/n + ln *L> -
r - r>1.)
E!," : -+
d"'tt
(3onz * 3on + 1 1).
it'$) : -w"
hn on
+ + 1 AZ
9:ix.Z |-":'.
and 8: -VO- cdt (e.21)
+
Since the potentials (D do not uniquely specify the fields / and 0, we
and, ,i
are at liberty to make an additional assumption relating them. For con-
venience, we assume that
i.A+1*:0,
c0t
(e.22)
which is known as the Lorentz condition. This choice arises quite naturally
during the simplification of Maxwell's equations.
where d is the velocity of the particle. Substituting Equation 9.21 into 9.23,
ln:Y*:m*+gA'
ALo
lu:6:mi+!Au
lu:Lun:m2+!Au'
where it is assumed that the components of i are not velocity dependent. Then
we can write
O-
F:md*zrq,'. (9.28)
o:!(
- m(o -r.r\
c'- l'
plus any mechanical potentials which may be present. Note that the kinetic
energy term for a particle in a magnetic field can be obtained by merely
substituting the quantity (p - qlc i) for p in the expression p2l2m.
In order to utilize the above Hamiltonian in quantum mechanics, we will
regard f and i as operators and will replace p with -ihV. It is therefore
useful to know under what conditions y' and ,Z commute. Since each com-
ponent behaves as follows,
To show that this satisfies our original definition of the vector potential,
Equation 9.2I, we operate as follows:
Vxi:-+Vx(;xA)
: -*ti(V.6 + (d.iii - dF.i) - fi.Vdl.
Since the divergenceof 6 is always zero, the first term vanishes. The second
term is(d.Vli : A. Since the divergence of iis 3, the third term becomes
34. The last term is zero because 6 j, u constant vector. Thus,
ixi:-+@-s4-d,
APPROXIMATION METHODS AND APPLICATIONS
Letting q be the electronic charge e, we see that the linear term in the magnetic
field is simply the Zeeman term. If the z-direction is taken along the field
direction, the magnetic dipole or Zeeman energy is just
L^E : mt\s%,
h A. (^d + zs),
and by adding the spin-orbit term, Equation B.B. The Zeeman energy then
becomes
L,E : (*, + 2m,)pry% : gmrpB%,
where the field 0 is taken to be in the z-direction. Since we must regard the
spin and orbital momenta as strongly coupled in weak magnetic fields, the
appropriate representation is lj, m,). Then, by first order perturbation theory,
the Zeeman energy corrections are
tr\!]*,,,,*, : r# U, m) J, I
j, mi) : €pahmr. (e.35)
These energies have already been calculated for p and s states of hydrogen, and
are given in Table B-5 in section 9 of Chapter B.
INTERACTIONS OF A CHARGED PARTICLE 353
The reader might well question the utility of the middle expression in
Equation 9.34 above. Its usefulness arises when the states u.. .*pr.rsed in the
lmt, m,) = l/, m) ls,.m,) representation. Thus, it would be used for the strong-
field Zeeman effect (Paschen-Bach effect) when the dipole energy is comparubTe
to or greater than the spin-orbit energy. It can also be used for the *.uk-fi.Id
Zeeman effect if the kets 17, m) are written in terms of the kets lms, m,). It will
no doubt be instructive to show this in detail for a few cases. Using the kets
for a/-electron given in Problem 8-20 we have:
:T.to(rt
ltaQ (1, L,
tril,g,t3+l * 25, ll, +) :2pn0,
trt,I,*,t:#'[3(0, +lL,+ 2s, 10, +) ++(1 , -+lL, *2s, lt, -+)]
: &png,
and so forth, in agreement with the values calculated from Equation g.35.
The spin-orbit energy correction for hydrogen can also be calculated
readily by means of perturbation theory. Using Equation B.B and the operator
given in Problem 8- 16, we can write
where 0Vl0r has been replaced by e2fr2. This form of the operator
^d. ,S was
chosen since it has a diagonal representation when the kets U, mi) are used as
basis vectors. Then, for the I states of hydrogen,
where : and tro li) : li). The vectors ) : are the eigen-
1t*n li)
1pn Elo' lf ,p:ot
cnn(En
- Ef') (e.38)
For convenience let us assume that we have only two degenerate states, the
ntk alnd /th'. Then the contributions to the energies of these states for i t' n, /
are negligible, and we may write
The necessary condition for these equations to have a solution is that the deter-
minant of their coefficients be zero. That is,
The meaning of what has been done in solving the degenerate case will
become apparent if one but compares Equation 9.40 with the diagonalization
procedure of Chapter 6. Thus, a matrix which has degenerate diagonal
elements and non-zero ofldiagonal elements in the representation of the
unperturbed eigenfunctions has a diagonal representation when expressed in
the new basis of the eigenfunctions of the perturbed Hamiltonian.
As an illustrative example, let us calculate the Stark splitting for the
n : 2level of the hydrogen atom by an electric field I in the z-direction. The
perturbation term of the Hamiltonian is zf$'t : eE z : e8r cos 0. There are
four degenerate wave functions corresponding to n : 2, !)zoo, Vzror grrr, and
Vzfi. Since each of these functions has definite parity and the perturbation is
odd, all diagonal matrix elements are zeroi that is, there is no first-order
correction to the energy and the degeneracy persists. Now let us obtain the
full matrix of the perturbation, whose elements are given by
Using parity arguments, one of the quantities / and f' must be odd and the
other must be even in order to have a non-zero matrix element. Furthermore,
we learned from the properties of the spherical harmonics that m' must equal
mfor a non-zero integral. Thus the only non-zero matrix elements are:
:
ffi|,**(' - f,)'-''"' o'1,'sin o cosz o do
: 3aoe8.
APPROXIMATION METHODS AND APPLICATIONS
3aoe8
-0,
-E\zt
and
EL" : !3aoe8.
The diagonalized matrix of the Hamiltonian is:
I 3aoe8 0
0 Bto) * Saoe8
(.tr) : (tr,) -t (./ftu1 :
("*' 0
0
0
0 4i) (s.+2)
Qlfrlprro - gno) has energy Eto' - 3aoef. This behavior can be described
by attributing a permanent electric dipole moment of magnitude Saoe to the
n : 2 state of hydrogen. It then follows that the three energy states correspond
to parallel, antiparallel, and transverse orientations of this dipole with respect
to the electric field.
PROBLEM 9.18
We will now consider'the case in which the perturbing part of the Hamil-
tonian is time-dependent; that is, when
Jry:ih,E,AY
where
tr # o : Elot V n and Ye : tpc-wntnlott -
and
#:j(0"-'ntf'"-)v,. (s.4+)
APPROXIMATION METHODS AND APPLICATIONS
Substituting Equations 9.43 and 9.44 into the Schrcidinger equation, we obtain
The first term of Equation 9.45 is trivially zeto since Y,, was defined to be an
eigenfunction of ff o. Multiplying the second term by Vf there results
and
a*(t) : - ei'oit), fo, k +i. (e.51)
H(I
The quantity lak(t)12 is the probability that the state Y* is occupied at time /.
Hence, it is a measure of the probability that the system will make a transition
from the initial state Vn to V* in time /. Squaring Equation 9.51, we have
Note that the probability oscillates at the angular frequency given by trro
but that its effect is appreciable only for values of ft near j, where the amplitude
reaches its maximum value. Recall that the derivation of Equation 9.52
assumes that the arremain small; this implies that ,trfr,t 4fr[0t
- Eo'.
For large / the amplitude aoQ) in Equation 9.52 resembles the delta
function, 6(E - E:o'), discussed in section 7 of Chapter 4. Assuming a quasi-
continuum of closely-spaced states, the area under the graph showin g lak(t)|,
versus E represents the total transition probability. Then, letting p@) be the
density of states,
p- p@)dE:12,tr;}'1, p(E)
!**bo{,)l, ,rt[_-#or
:- 2nt ,tr;t'l'
I
- p(E) , (e.53)
h
which says that the total probability of a transition out of the initial state Y, is
directly proportional to the time that the perturbation acts on the system. Both
p(E) and .ffft) are assumed to be slowly varying and are regarded as constants
for the integration of Equation 9.53. This is not a bad assumption, since the
integrand is a sharply-peaked function centered around Er10). The transition
rate from the initial state j be defined as,
^uy
, _dP _2n I tr!,'o'l' p(E)
rL- , (e.54)
dt- h
which is often referred to3 as Fermi's Golden Rule Number 2. It will be used later
in our discussion of scattering theory.
A few words should be said about energy conservation during the trans-
itions discussed in this section. In Equation 9.52, energy is conserved to
within the limits of the uncertainty principle expressed by LE . L,t ,-, h,
Although the system can oscillate between states of widely separated energies,
the greater this separation, the shorter the lifetime of the'excited state. How-
ever, as the resonance condition is approached, AE -* 0 and A, -' oo, implying
that a long-lived transition can occur. In this case the radiation field insures
that the total energy is conserved by creating or annihitating a photon. In
Equations 9.53 and 9.54 energy conservation is built into the formalism by
the assumption that all transitions occur between states which are closely
packed around the energy Ero'.
in solids and the study of resonance absorption in optical spectroscopy are both
easily described in the harmonic approximation.
Since either the electric vector or the magnetic vector (or both) of the
incident radiation rrray interact with an atom, let us omit the details of the
interaction for the time being and write the perturbation as
Substituting this into Equation 9.49 we obtain for the expansion coefficient
a*(t),
a*(t) : - tO *1,,,t (sik"a+att a sib*i-'tt1 dt
lr'
(e.57)
lah?)Iz --
#rl',r,"195"i,1,
where the minus signs correspond to absorption and the plus signs to emission.
Using the delta function approximation for large /, the transition rate to the
kth state from the jth state may be expressed as
2r
Ro: E I tr[",], 6(Eoi + E), (e.58)
Here, Fo : t8, + jE, + fc* the amplitude of the electric field and D : -0i
"is Then, from Equation 9.58, the rate of electric
is the electric dipole operator.
dipole transitions from the jth to the krh state is given by
8r:8n:8",
s2 -- 8tr : E2 : +8t,
since EB : Sl + Si + 92.
Furthermore, since the energy density of an
electromagnetic wave is given by
I(,) :: +77
@3>,
when the permeability is unity, the matrix element of Equation 9.60 may be
expressed as
where
lio,l, : (D,)r*, * (D,)rr, * (D,)?,,.
Then
R.
tttc
--Bn2I(a)
gh
.lDo,lrd(E[0,
- ptil + E). (9.62)
Recall that Bo,, the Einstein coefficient for induced emission or absorption
of photons, was defined in terms of the radiation energy density as (see section
4 of chaPter 2)'
R t : B*il(ot). (9.63)
By equating 9.62 and 9.63, we see that the Einstein coefficient may be expressed
in terms of the electric dipole matrix elements; that is,
It is interesting to note that the electric dipole model does not permit a direct
calculation of the Einstein coefficient for spontaneous emission, A*,, although
the latter can be obtained from the ratio given in section 4 of Chapter 2. In
order to treat spontaneous transitions rigorously, it is necessary to use a model
in which the electromagnetic field is quantized so that the physical system
consists of the quantum states of the field as well as those of the atom. The
absorption of a photon is then described by the simultaneous annihilation of a
photon state in the field and an upward transition in the atom. Emission is
iegarded as the creation of a photon state in the field, accompanied b.y u
downward transition in the atom. In the harmonic oscillator approximation,
the operators of Chapter 5 may be used. Although this topic belongs in a more
advanced course, the interested reader may wish to consult the excellent book
by Heitler.a
The selection rules for electric dipole transitions have been mentioned
previously, but it will be instructive to discuss them in the present context.
Note thai R* will be zero whenever Equation 9.61 is zero. Let us write the
hydrogenic wave functions in ket form vs !)nt* : ln) l/) lm). Then the matrix
.i.*.trtt of the components of D are:
Q"t D*IP) : (n't r tn) (/'t sin 0 th @' I cos { lrz)
lr)',',:'ir'r::-: rt:'*:'
Q"l Du h) : trrri'r'r'r".'r'u',',ti,"u!),,'1)*,'*'l sin 4lm)
A,/ : tI
(e.65)
Nm:0, +1.
a W. Heitler, The Quantum Theory of Radiation, 3rd ed., Oxford University Press, Oxford, 1958.
PERTURBATIONS THAT ARE HARMONIC IN TIME 363
i : ioer(i.i-,t1. (9.66)
A:Aoe-o"lL
1L I lrrv rJ + W+"'-l.
+UE'i) (e.67)
-1L0v L^ ' 2
Such an expansion is certainly valid for atomic radiation in the visible spectrum
' where)"f ao ,--' lgt. Keeping only the first term, the perturbation is
e*
' Ao cos at.
;F
Suppose that the wave is polarized in the r-direction. Then
F'io : !,Ao
and i
But
fn:m*:fflrr,*1,
from Equation 6. t l. Then,
.tolr : i2rh
.ff\','
ieA,
xtr,I j)l
l&l tror Ii) - (kl
: j2rhl@rkl
ieAo
x Ii) - <kl x l,tr i>l
:#(Ef,
ieA,
-Eo,)(klxlj)
:r#glexlj)
:'+ (D*)a,
energy density is
ir , : @'A3
t\a1
Brcr,
we write Equation 9.60 as
L-2
Rr,: i ,@ l(D*)nil, (e.6e)
For an unpolarized wave, the average of cos2 0 over solid angle yields a factor
of fi which makes Equation 9.69 consistent with Equation 9.62.
Returning to Equation 9.67, note that the next term of the expansion
introduces an additional factor of iE 'i which multiplies the quantity (ir'il.
These products are components of a second-rank tensor, some of which contain
orbitaf angular momentum while others contain charge times quadratic terms
in the coordinates.s The former correspond to magnetic dipole moments,
while the latter are called electric quadrupole moments.
Each higher power of iE. i introduces higher-ordered multipoles and a
corresponding new set of selection rules. We will not discuss them further
here except to remark that the appearance of a spectral line which is forbidden
in the dipole approximation merely means that other kinds of poles are playing
a role in the physical process. For example, the selection rules for electric
quadrupole radiation aret
,
Neglecting : 0
spin: L,/
Lm :0, +1
fncluding spin I Li - 0
Lm:0, +1
between these levels are induced by an oscillating field. The frequency of the
oscillating field is generally in the radio frequency band for nuclear spin
resonance, in the microwave band for electron spin resonance, and in the optical
band for electronic orbital transitions.
The magnetic term in the Hamiltonian is,
(do + 2d, co, ,r) 'T j : spu*roo rry# d, . icos a;r, (9.70)
where we have taken 9o to be in the z-direction. For a fixed value of 90,
we already know that the solutions for the Hamiltonian
are the kets lj,*,). Therefore, we will now regard Equation 9.71 as the
unperturbed Hamiltonian and the kets 17, m,) as the unperturbed wave functions.
It should be pointed out that Equation 9.70 assumes that g is a scalar
quantity. This is a good assumption for free or nearly free atoms, but is not
generally valid for ions or atoms in a crystalline lattice. In solids it frequently
turns out that the g factor is a tensor, and hence the perturbation must be
wrrtten as
jt'l:+d''3'i.
However, for simplicity we will regard g as a scalar in the present example.
Let us further assume that the oscillating field d, is in the x-direction. Then
the perturbation becomes
The transition rate from the jth to the kth state at resonance is given by Equation
9.58 as
where J* governs the absorption process and J- the emission. The rate of
absorption in the state I j, m) is
1T
2h
kpu%,),(j -*)(j +m|_I),
and the rate of induced emission is
fi
2h
kpue,)r(j + *)(j - m -l r).
APPROXIMATION METHODS AND APPLICATIONS
Of course, the number of transitions per second must take into account the
populations of the states as discussed in section 4 of Chapter 2.
PROBLEM 9-23
ihe string tension is slowly changed while a cirtain harmonic is bowed. The
harmonic used corresponds to a given oscillator state, and the change of pitch
is relate d to the change in the energy of the state. Another example is provided
by the elastic collisions of low-energy gas molecules. Since the electronic
motions are very rapid compared to the molecular motions, the distortions
produced during the collision process may be regarded as adiabatic and the
electronic wave functions are restored to their original form. On the other
hand, a high-speed molecule can induce a transition to an excited state, and
the collision process will no longer be elastic.
The analytical form for the adiabatic approximation may be obtained
by integrating Equation 9.+7. Since ,tr*' is now a function of the time, we
must integrate by parts:
at,: - +
# en'*,'
*,Ifir*;:llei'*t dt. (e.73)
The second term may be made arbitrarily small in the adiabatig approximation,
so that the first term gives the required result. Note that this result would be
identical with Equation 9.9 for a constant perturbation if the time factors were
included in the wave functions ,s'ed in section I of this chapter.
Now suppose that a system undergoes a sudden change at time / : 0
such that its Hamiltonian is given by trofor I < 0 and by tr - ffo * .ffrtt
forl)0,where
tr o li) : E,i li),
and
tr lq) : Er ltl)'
That is,, the orthonormal eigenfunctions of the original Hamiltonian are
designated by Roman letters, whereas the orthonormal eigenfunctions of the
ADIABATIC AND SUDDEN PERTURBATIONS 357
l ou lr> : 7
bo ln).
b,t:ZonOtli>. (e.7 4)
1,
Equation 9.74 would give the correct expansion coefficients for an instantaneous
perturbation. In a physical problem, the system will require a finite time A/
in order to make the transformation from the state characterized by ff o to
that described by tr. In such a case,
where the exponential has been replaced by the first two termd of its power
series expansion. Comparing Equations 9.74 and 9.75, it is evident that the
error introduced by the sudden approximation is proportional to At and to the
energy difference between the initial and final states.
As an example, consider the harmonic oscillator whose spring constant is
suddenly reduced to one-half its original value. We wish to know whether
this oscillator is likely to make a transition or whether it will remain in its
original state. For simplicity assume that the oscillator was initially in its
ground state, given by
rt:o : (!\
\7r/
r-*'r,,
where : l/f4n. The ground state wave function associated with the new
"
spring constant may be written by replacing a with
"112,
(Y-t
'' " : \T\/ 2l r-naztz{T,
ro-:n
'
APPROXIMATION METHODS AND APPLICATIONS
PROBLEM 9.24
PROBLEM 9.25
PROBLEM 9.27
PROBLEM 9.28
g' : >anli),
2lanl' Eo
<./f> _<tt'l (lqt) (e.76)
\rp I vl 2lo)'
i
Thus, Equation 9.76 provides an upper bound on the ground state of the system,
although not a very useful one. If Equation 9.76 can be minimized, however,
then the upper bound so obtained might well be a good approximation to the
actual ground state energy. The usefulness of the method lies in the fact that a
trial waue function V can be chosen such that it contains one or more variational
parameters for the minimization procedure. The effectiveness of the method
is dependent upon a judicious choice of the trial function. After the ground
370 APPROXIMATION METHODS AND APPLICATIONS
state energy and wave function are obtained, the method can be applied to the
first excited state by choosing a trial function that is orthogonal to the ground
state function.
For the purpose of illustrating the method, let us apply it to a problem
whose solution we already know, that of finding the ground state energy and
wave function for the harmonic oscillator. Choosing a trial function of the
form
yt : Ce-"', i
ff : -#fi* Lkx,,
Letting
!<*>-0,
oa
km
'-+h,
zt2-
-
,:T:I,,
\/k* d.
eo : (t)* ,-auztz,
\7rl
Eo: (ff)^in:+-+H:ry.
Because of our choice of the trial function, we obtained the exact solution in
this case.
THE JWKB SEMICLASSICAL APPROXIMATION 37t
PROBLEM 9.29
y, : Ce-f,',
PROBI.EM 9.30
hk:\/W.
A similar solution may often be used to approximate the real solution in physical
cases where the potential is not constant, provided that the potential is slowlt
uarlingin space. This approximation is quite useful in one dimensional problems
and in problems with radial symmetry. Although it has been credited to a
number of people, the method is usually called the WKB method or the JWKB
372 APPROXIMATION METHODS AND APPLICATIONS
metltod, after Jeffreys, Wentzel, Kramers and Brillouin6 who first applied the
mathematical techniques to physical problems.
By u "slowly varying potential" we mean a potential whose energy change
in a deBroglie wavelength is very much smaller than the kinetic energy of the
particle. Thus, the fractional change in kinetic energy per wavelength is much
less than one:
I I AVI
(e.7e)
(E - v)lax I
l2h ap
<1.
lp'a
This is analogous to an optical medium in which the index of refraction varies
continuously, but sufficiently slowly so that no reflection occurs.
Let us designate the slowly-varying potential by S(r) and require that
S(") : px when Z is constant. We will assume that S(r) may be represented
by an asymptotic expansion in powers of h as follows:7
Jeffreys, Proc. London Math. Soc. (2) 23, 428 (1923); G. Wentzel, Z. Physik 38' 5lB (1926);
6 H.
Combining like terms and setting the coeffi.cients of each power of fi separately
equal to zero, we obtain the following equations:
I
2*
(s6)' + V- E:0 (e.82)
,s6^ti
- ,S'; -0 (e.83)
where all terms beyond the quadratic terms in hhavebeen dropped. Rewriting
Equation 9.82 as
dso : t/Z*1n - V1 ar,
So:*lJ,, t/zm1n-v1
Pa
ax. (e.85)
fn a similar fashion,
sr:!r"1s6y
or
exp(fS') : l2m(E - V)l-+,
and
c
'Jz- ; # . t2m(E - v)t-B - { [ rr*rE - n]-t (#)' *
If V is a slowly-varying function of # so that Equation 9.79 is satisfied
and all higher derivatives of V are also small, then it usually is possible to
retain only the first two terms of the expansion given by Equation 9.80. Then
the wave function may be written as
and ysqv. If the region where the JWKB solutions break down is small
.tro.rgh, the potential can be represented by u straight line within the region.
Taking its slbpe to be that of the potential curve at x - 0, the equation of
the line is
v-E:(#),_"(x-a). \
Then the Schrcidinger equation in this region can be solved with the aid of
Bessel's functions of order $. The details of deriving the connection formulas
will not be given here,8 but for a simple barrier they may be summarized as
follows.
V(x)
k, : ,_l
Vem(n-v) kr:
I-
Vzm(v-P)
6a | fra
;-8"*o(- I:r,
*) o
ft",,ff,
-, dr
- t (e.8e)
h..o(,[0,
or)
= - +^'t(fo,
d, -) (e.e0)
t See E. C. Kemble, The Fundamental Principles of Quantum Mechanics, McGraw-Hill Book Co.,
NewYork, 1937, section 21; N. Frciman and P. O. Frcjman,JWKB Approximation,Contributiotu to the
Theory,North-Holland Publishing Co., Amsterdam, 1965; P. M. Morse and H. Feshbach, M*hods of
Theoieticat Physics, McGraw-Hill Book Co., New York, 1953, pp. 1092 tr
THE JWKB SEMICLASSICAL APPROXIMATION 375
.]'\
gm - ertio,u,,
where
t_
kt::Ot/Z*p-V).
For convenience later in applying the connection formulas, we will add a
phase factor to ?r'; thus,
In order to find Vr we must use the connection formulas for the barrier to the
right. First, let us rewrite Equation 9.92 in the following convenient form.
vr: -+[+.*o
!krt
(-l:r,ar) ,t"(fo, or -f
+ i2'"0 (fn , o,). ."'(fo , dr -;)] (e.e3)
n :f n'or
and
P:fo,dx -;.
Then Equation 9.93 becomes:
Ipr : - * i2e"cos P)
#r(Er-'sin P
A l- r
Le-')
l#,(e'+
SUGGESTED REFERENCES 377
A ,
T- A
lkr : ['"0(.ll lr, a*) + ie*P ( I"n,*)f-'
Gr@ * Le-")
(e.e4)
SU/T^IHARY
In spite of the fact that very few physical problems in the real world can
be solved exactly, methods for approximating complex systems have been
developed which are responsible for the widespread application and success
of quantum mechanics. The most important and simplest of these approximate
methods is perturbation theory, which can be used whenever the actual system
can be regarded as a slight modification of a system whose solutions are known.
Special techniques are required for the cases of degeneracy and time-depend-
ence of various sorts. These methods are illustrated by numerous examples.
When the physical system is not amenable to perturbation theory, a variatibnal
method can often be used to obtain an upper bound on the energies of the
lowest states. The success of this method hinges upon judicious choices of the
trial wave function and the number and kind of variation parameters. A semi-
classical method is also described and is applied to a problem in barrier
penetration.
SUGGESTED
REFERENCES
David Bohm, Quanturn Theory. Prentice-Hall, Inc., New York, 1951.
Sidney Borowitz, Fundamentals of Quantun Mechanics. W. A. Benjamin, fnc.,
New-York, 1967.
378 APPROXIMATION METHODS AND APPLICATIONS
&ffiffiffiw$ffiM&il"
&Pffifuffiffi&Yffiffiruffi
where wo : 13.6 eV, the ionization energy of hydrogen. Since the experi-
mental value for the ground state of helium is -78.62 eV, the reader can
readily see how large an error is produced by the effects which have been
neglected in Equation 10.1. In spite of the magnitude of the corrections
needed, it is an attractive idea to try a perturbation calculation, since the
interaction between the electrons is an obvious choice for the perturbing
potential. Pursuing this approach, the Hamiltonian becomes,
For spherically symmetric wave functions of the type used here, this
integration may be easily performed by regarding it as the electrostatic energy
of two overlapping spherical charge distributions. Thus, we first calculate the
potential at a point i, due to the infinite charge distribution defined by i, and
having charge density given by exp [ -(2Zrrlao)]. Since this potential is
constant for r, lrn but varies as l lrrfor rz ) rtz, we have the two integrals
*Although the simple product function is satisfactory for the present discussion, the reader
should note that the proper form is the symmetrized spatial function given on page 306.
THE GROUND STATE OF THE HELIUM ATOM 38I
,0,*t0)
r
: 5F"l .[o-,-,,vftnxz,dx,:W : *Zwo. (r0.7)
PROBLEM ti{it"iixl'lirrr+i"r
'0'l
Verify the integrations leading to Equation 10.7.
:*:i|(;l ''n*
r' ror r'r ) rz
1:l - /r-\t
rrz ,rA\r)rt3os o)' for rt I rzt
where 0 is the angle between i, and /r. Because of the orthonormality of the
spherical harmonics, performing the integration over Q, and O, before the
radial integrations will eliminate all of the terms containin g Px except for / : 0.
Then Equation 10.6 becomes:
etf,ro) :f
A[o*or,rl,-,,{['i !l'ar,rz,-,, +!),a***-,ll
5 ezZ
: :tZwo' (10.8)
B
'h
in agreement with Equation 10.7.
1 L. r. schifl Quantum Mechanics,3rd ed. McGraw-Hill Book co., N.y., 1969, p. 258.
ADDITIONAL APPLICATIONS
':i:.-:r
::i,Y#,.i)r+,1"ti6ij? ti;rirtilli .:ii'E:gli:;L:-"1;r
"r't1 ),i;iitii i.i, #;p;li6i.5lj*iiil6r
l; *, :
Ee,ou,,a : -2z2uo *1r*, : _
-74.8 eY, (10.e)
The second integral was just calculated above, and the first will be left as a
problem. The result is
(Eg.oo,,a) : 2(Zz - 4z)wo * *Zwo' (10. r0)
Show that
(01 trol0) : (2, - +4 .#r, (10.1 1)
THE LOWEST EXCITED STATES OF HELIUM
where l0) is given by Equation 10.4 and ,ffs is given by
Equation 10.1.
PROBLEM IO-1
PROE|-EM
'0-5
The first ionization energy of lithium can be calculated by
finding the difference in energy for the configurations 1s2
and 1s22s. Choose a product wave function for the ls22s
configuration, set up the integrals, and explain the procedure
to be followed. Do not evaluate the integrals.
PROBLEM 10.6
Find the first ionization energy for Li+. (This is the energy
difference between the ground states of Li+ and Li++.) The
experimental value is 5.56210.
PROBLEM IO.7
The lowest excited states of helium are those for which one electron
remains in the ls state and the other electron is raised to a 2s or 2p state.
Since the electrons are indistinguishable, for each pair of quantum states there
will be four possible spin wave functions as given in Equation 8.22. There
are also four two-electron spatial quantum states represented by n and n',
namely, Vnogzoo, T-oq'', Vnogzn, and tp.,s1p2..5. Neglecting all interactions,
then, there are 16 degenerate wave functions representing the first excited state
of helium. In reality, of course, we cannot turn off the Coulomb interactions,
so that the 16-fold degeneracy mentioned above does not physically exist. It
turns out that the four states associated with Vrcgz, are not raised as much as
384 ADDITIONAL APPLICATIONS
-
K,*. ,, 2'S,,
/
// ls 2s (l state) t
--
lsl I tS,,
fI state) (l state)
-78.62 Exp. ground state
/- -
/
I
/*,..,^
-108.8 E,tu'
Figure l0-l Schematic diagram of the lowest states of helium. The spin-orbit interaction has
been neglected since it is only eV. Degeneracies of leyels are given in parentheses.
-10-a
the twelve states associated with the yrrrprp wave functions. This is shown in
Figure 10-1. The remaining degeneracies are split further by spin-orbit
interactions and the correlation effects arising jointly from symmetry and the
Coulomb repulsion as discussed in sections 5 and 6 of Chapter B. fn order to
see this more clearly, let us look in detail at a perturbation treatment of the
four degenerate wave functions formed from !)*Vzr. We write g zs,
where the plus sign refers to the singlet state and the minus sign to the triplet
state. The ket ls, m,) can be any one of the three symmetric spin functions
when the minus sign is used; for the plus sign, the ket ls,m,) must be the one
antisymmetric spin function. It will not be necessary to express the spin
THE LOWEST EXCITED STATES OF HELIUM
EL" : <vlfiW>
ti,
-- rrr,( t),pr,(2) + r)pr,(2) lpt,(l)rpr,(2) + q)r,(l)vt,(2)) (s, nt,l t, *,)
?2,( I
*
- K +J, (10.13)
where
I
K - e2 (p*(1),pr,(2)l
fn l,pr,(1),pr,(2)),
(10. r4)
I
J: ez (,pt,(1)pr,(2)l t ) ).
fn lrltr,(2)y,r,(
The integral K is the direct or ordinary Coulomb integral and the integral J is
called the exchange Coulomb integral. Notice that when ,K is evaluated, the
integration over the coordinates of one of the electrons involves only one of the
functions !)u ot gz* On the other hand, when J is evaluated each configuration
integral involves both yy and gr, and, in fact, is a measure of the degree of
overlap of these functions.
The energy of the first excited state of helium is, then,
Ez:ELo'+K+J,
where the plus sign refers to the singlet state and the minus sign to the triplet
state. The calculation of the following energies will be left to the problems:
Note that the energies of both the singlet and triplet states are raised by the
direct Coulomb interaction, but that the exchange interaction raises the
singlet and lowers the triplet (see Figure l0-l).
3. THE HEISENBERG EXCHANGE INTERACTION
AND MAGNETISM
Heisenberg and Dirac have shownz that the phenomenological vector coupling
model provides a very simple spin Hamiltonian which gives the correct energy
splitting for this case. The Hamiltonian is
tr\"n: -2J&
.Sr, (10.17)
where ,31 and ,S, are the spin vectors for the two electrons. Since the total
spinisS:&*Sr,
,S, : (S, * Sr), : si + sr' + 23, .Sr,
and
Now, for the singlet state, where the spins are antiparallel,
(St 'Sr).tngtet : -?
and for the triplet state where the spins are parallel,
in certain atoms of the transition metal and rare earth series of the periodic
table. Moreover, it is now generally believed that this same exchange mecha-
nism is the origin of the intense internal field postulated by Weiss to account
for the parallel ordering of the spins in ferromagnets and the antiparallel
ordering in antiferromagnets (where the exchange integrcl J is negative).
A serious defect of Equation 10.17 is that its form is apt to lead one to
believe that the exchange energy has its origin in a spin-spin interaction (that
is, a dipole-dipole interaction). This certainly is not the case, and it was known
even at the time of Weiss that the dipolar forces were three or four orders of
magnitude too weak to account for spontaneous magnetic ordering. The reader
should bear in mind that the energy difference is due to an electrostatic interaction
which arises from the symmetry requirements of fermions. The fact that this
energy difference can be accounted for by means of a scalar product of spin
vectors should be regarded as fortuitous. In spite of this minor shortcoming,
the Heisenberg-Dirac-Van Vleck Hamiltonian, Equation 10.17, is of such great
importance that the overwhelming majority of papers in the field of magnetism
use it as a starting point.
In the discussion of perturbation theory it was pointed out that the smaller
the perturbation, the better the approximation. Therefore, since the Coulomb
repulsion between electrons is so large, we cannot hope to apply perturbation
theory alone to atoms with a large number of electrons-in spite of its fortuitous
success in the case of helium. On the other hand, the inclusion of the electron
interactions in the unperturbed part of the Hamiltonian would make it an
insoluble many-body problem.
A way out of this dilemma was suggested by Hartree,s who showed that
the spherically symmetric part of the electron interaction energy can be included
in the unperturbed Hamiltonian by means of a potential derived from the
average charge density of all electrons but one. That is, a single electron is
visualized as moving in an effective potential du'e to all of the other electrons
as well as the nucleus. The contribution to the potential seen by the ft'a
electron due to all the other electrons is calculated from the spatial average
of each unperturbed one-electron wave function ?pl.o), or
Then this potential is used along with the nuclear potential to solve Schrcidinger's
equation self-consistently. That is, the equation
is solved for each electron. Then the new wave functions ylt) are used in
Equation 10.18 to calculate an improved potential V;. The iterations are
continued until the potential ceases to change appreciably; that is, until it
becomes self-consistent.
fn recent years the self-consistent method has been used extensively
because it is readily
programmed for digital computation.a Some of the short-
comings of the model are that it neglects correlations and the Pauli principle,
and it retains only the spherically symmetric part of the Coulomb interactions.
An improvement due to Focks consists of including exchange by using properly
symmetrized product functions for the initial wave function.
where an additional factor of 2 has been included here because there are two
spin states associated with each spatial state. .Notice that the energy of the gas
increases as the number of electrons increases. The energy of a collection of Z
electrons is obtained as follows:
fEr
':)o "@)
dE:t#f'E+dE --ffi"',,
Ep : fi{z,, e)*. (10.21)
is simply the energy of the highest filled state. In the many-electron atom
which we are treating, we have Ep: -V(r), and from Equation 10.21,
P: L:
z em)* (10.22)
g*r1t ?vg11,,
Now, the self-consistency condition is that the potentials due to the electron
density in Equation 10.22, as well as that due to the nuclear charge, properly
reproduce the potential energy, -V(r). Consequently, the charge density t
-0 p,
and the electrostatic potential, -lV(r)lel, must satisfy Poisson's equation,
_.4n(-ep),
-1VrV:
e
Equations 10.22 and 10.23 may be solved simultaneouslyT for p and V, imposing
the boundary condition V(r) : -(Zezlr) as r--+0, and rV(r) -'0 as r --+ oo.
The result is that the "size" of a heavy atom is approximately
h2 ao
me'Ttr z+
(10.24)
PROBLEM IO.9
? L. I. Schifl Qtnntum Mechani,csr 3rd ed., McGraw-Hill Book co., New York, 1968, p. 430.
390 ADDITIONAL APPLICATIONS
PROBLEM IO.TO
where
ff crr : - h2 vt2ctut v
2*r*,
3fN --h, mp vk +t;
hz y12 ez ez
ffL- - 6"'t - i- ,*
nz hz ez ez
hfz- -
2*u'2 - ,z - ,*
t'
ffrr:
ftz
Here mnl2 is the reduced mass of the two protons and m" is the reduced mass
of the electron relative to a stationary proton. For simplicity we will neglect
the translational and rotational energy of the molecule and will regard the
distance between the protons, R, to be fixed. (W. are also ignoring the
exchange symmetry of the protons.) We expect both electrons to be in ls
statesin the ground state of the molecule. Then, taking into account the
exchange of the electrons, there are two possible approximate electronic wave
functions which are products of hydrogenic functions, namely,
where yo and Vb are both ls wave functions. Recall that in the helium atom,
where there is only one nuclerrs2 po6 and tpuo are degenerate eigenfunctions of
the same Hamiltonian. Here, however, you and y)ba are not degenerate, since
one is regarded as an eigenfunction of ,ff, and the other is an eigenfunction
of .ff r.
A variation method may be used to solve the problem by writing a trial
function that is a linear combination of the functions in Equation 10.26.
Thus we may write
V:Pao*A'Puo'
where / is the variation parameter. Thenre
(1 + A')B +2AD
Egroorro { (10.27)
| 4 trz +2AC )
where
B- (rl,,ul tr lrpou) : (rttu"l tr lrpor)
C: (rl,oulrpu")
-4
!o z
^,
a.
t"o
X
u
L
q)
t-1 -z
Figure lG3 Energy per atom for the symmetric and antisymmetric electron sPatial wave
functions for the hydrogen molecule.
WritingV:gm*rpoo,
Egroooa=
B+D (10.28)
f *"
Upon evaluation of the integrals B,C, and D, the difference between the energy
obtained from Equation 10.28 and the electronic energy of two separated hydro-
gen atoms (about 27 eY) should be the dissociation energy of the hydrogen
molecule. The experimental value for the latter is 4.72 eY.
Physically, one could have guessed at the proper linear combination of the
wave functions in Equation 10.26, since an antisymmetric spin function would
require a symmetric spatial function in order that the total wave function be
antisymmetric. That the symmetric spatial function has a lower energy than
its antisymmetric counterpart can be seen by reviewing the example of the
double wells in Chapter 5 (see section 4). The energies of these functions
versus nuclear separation are shown schematically in Figure 10-3. Since the
antisymmetric function can have no bound state, it is often called the anti-
bonding orbital, and the symmetric function is called the bonding orbital.
F- --i
b
Figure l0-{ Schematic cross sections of the spatial wave functions for the hydrogen molecule.
-afo*trrq*ffr%r, (10.2e)
where the summation is over all of the i electrons in the atom and where ffo
includes all of the contributions to the total energy which persist when the
magnetic field rt it turned off. Classically, the magnetic moment along the
field direction may be defined as
AE
(10.30)
t-z - ag'
where the energy E is
E - -rt'4.
The quantum mechanical equivalent of Equation 10.30, which is
(l't,)--:
/nn - - Ln
69 ,-)nn,
(10.31)
has been verified by Van Vleck.l0 The quantity (E),, is calculated from
perturbation theory as follows:
The perturbation ,ff, has been carried to second order, since all quadratic
terms in 0 should be considered if any are retained. Then Equation 10.31
becomes:
oz
: (nls,pmrln) -g>m@lxl +filn) *rU,nffi ln')|,
gpamt t /,nl
The middle term in Equation 10.33 is always negative and is the source of the
diamagnetec contribution to the magnetic moment; this term will be neglected
in what follows. Let us now obtain the paramagnetic susceptibility associated
with the induced moment of the last term of Equation 10.33. Assuming a
to
J. H. Van Vleck, op. ci.t., section 36 of Chapter VI.
394 ADDITIONAL APPLICATIONS
Boltzmann distribution* over the states z', the atomic susceptibility may be
expressed as
Qtu)nn I
r"-'"'r'
x:V: s>:@' (10'34)
,-,
tl_.tf, +s ry - 2sEfilt-'*'o'ro',
where only first order terms in CI are retained. Now the first term in the
bracket represents the net moment in zero field, which is zero. Then,
Z ,ur-',tkr
,-.,s >l% - ru*'fe-Ento)tkr,
and
r: tl%
-rrf,),-Bntotlrcr
(10.35)
In e-E*totlnr
x:
>lW + z
_Z,WH1r,-'""',r, (10.36)
le-8"(')lttr
If the separation of the states is much greater than kT, only the ground state
contribuies to the first term of Equation 10.36, and it becomes
J
Zl@i polmi>12
:- s,rrL 4*f s'\LJ(J +
skT
1)
(10.37)
KT kr 2J +t '
The second term in Equation 10.36 gives rise to the so-called Van Vleck
temperature-independent susceptibility.
Here there are two contributions, one due
towidely separated values of z and n', and the other due to widely separated J
values.lr Although these contributions are usually neglected, it is woith noting
that they arise when the Zeeman effect is carried to second order in perl
turbation theory.
REFERENCES
David Bohm, Qwntum Theorlt. Prentice-Hall, fnc., New York, 1951.
R. H. Dicke and J. P. wittke, Introduction to Quantum Mechani.cs. Addison-
Wesley Publishing Co., fnc., Reading, Mass., 1960.
H. Eyring, J. Walter, and G. E. Kimball, Quantum Chemistry. John Wiley and
Sons, Inc., New York, 1944.
E. Merzbacher,
Qtnnturn Meclwnics. John Wiley and Sons, fnc., New york,
t961.
Albert Messiah, Qwntum Mechanics. North-Holland Publishing Co., Amster-
dam, 1958.
L. Pauling and E. B. Wilson, Introduction to Quanturn Mechanics. McGraw-Hill
Book Co., New York, 1935.
D. Park, Introduction to the Qunnturn Theoryt. McGraw-Hill Book co., New york,
1964.
J. L. Powell and B. crasemann, Quanturn Mechanics. Addison-wesley publish-
ing Co., Inc., Reading, Mass., 1961.
D. s. Saxon, Elementary Quantum Mechanics. Holden-Day, Inc., San Francisco,
1968.
Leonard I. schifl Quntum Mechanics..3rd ed. McGraw-Hill Book co. New
York, 1969.
rr
J. H. Van Vleck, op. cit., section 55, Chapter IX.
CHAPTER IT
sffiewwffimffiffiffiffi wffi#ffiffiffiw
?sph : -f (0) +
Letting the particle beam be directed along the z-axis, the incident plane wave
is given by
Vin:Aeik'i-7titcz.
Hence the scattered wave is the superposition
beam is
I: lrl,rol,.nn
mm
:U,
fot ?,r' normalized, while the intensity of the beam scattered at an angle 0 is
/r( o) :
Iryl *: t -r (o) t''
-,Y,,'
Then the number of particles Scattered through the angle 0 into an element of
area da, normal to the beam, is
The factor-f (0) is called the scattering amplitude, since its square gives the differ-
ential scattering cross-section.
. Eqtation I 1.1 is a solution of the free particle Schrodinger equation. It
is convenient to write it in a different form, however, by expinding the plane
wave in terms of the Legendre polynomials. Thus,
eitcz
- Alt(rf + I)jr(kr) . Pr(cos 0), (11.3)
it(k )
: (- \'(Lr)##J'#, (r r.4)
and the first two are:
jo&r)
sin *r
j,(k):t1?fl-ry
(k kr )'
The asymptotic form of these functions for kr ) / is
js(kr) --*
,ir, (t, -+) (1r.5)
kr
1G. N. Watson, Theory of BesselFunctions, Macmillan, N.y., 1944, rev. ed., p. l2B.
SCATTERING THEORY
Thus, at large distances from the scattering center, Equation 11.3 may be
written as
Since Equation 11.6 is the free particle solution, it must also be the solution
for a non-zero potential at large distances.fro* the scatterer. We assume that in
the interaction region its form cannot change radically, but that the potential
must only alter the phase of the sinusoidal function. For the asymptotic form
of the scattered wave we then assume
l|sc
,i., (*, -+ *4)
- kr'
where d7 represents a phase shift of the (th partial wave. Rewriting Equation
1I.1,
oo
bxPl@os 0)
,i,'(r, -+*r)
t:0 kr
,i,' (r, - +)
: + l) ' Px@os o) +f(0)
t:
(1 I .7)
Zitel kr r
we find that
br : i/(2/ + l)ei6t (l 1.8)
and
-f (0)
:IAQI + t)darsin d7 . Pr(cos 0). (1 1.e)
Note that the scattering amplitude is zero if all of the phase shifts vanish. The
total scattering cross-section calculated from Equation 11.8 takes the simple
form
'.1
Show that Equation 11.5 is a solution of the Schrodinger
equation for a free particle.
PROBLEM I1.2
PROBLEM I1.3
/--
o :7Im t/(0)1. (11.1r)
This relationship is known as the optical theorem; its validity is far more general
than the present derivation implies.
In order to apply the partial-wave method of analysis to a scattering experi-
ment it is necessary to be able to calculate the phase shifts, il. In general,
the phase shifts will be positive for an attractive potential, which means that
the phase is advanced and the wave is pulled in toward the scatterer in the
interaction region. The peak of the outgoing wave is thus spatially behind
the free particle wave (see Figure l1-l). The reverse is true for a repulsive
potential. The scattering cross-section, however, does not depend upon the
sign of d7. The d, are obtained by matching the radial solutions of Equation
7.70 for V : 0 to the radial functions which include the effect of the scattering
potential, Equation 11.7. An explicit expression relating the phase shifts to
the scattering potential and the radial eigenfunctions can be obtained by means
provides a simple guide for determining how many values of / will be required
to represent the scattering. Thus, for low energy particles (small ft) or a
short-range potential (small ro) only the lowest / values are required. fn fact,
it often turns out that kro 11 so that / : 0 is the only angular momentum
state that need be considered. Such an event is called S-wave scatterin$, in
analogy with the terminology of spectroscopy. S-wave scattering is isotropic
(in the C-frame), since from Equation 11.9 we note that
-f (o)
: ! sin do'
'nuo
when all phase shifts vanish except for that associated with / : 0. The total
scattering cross section for S-wave scattering is
where
hk,--\/Wm.
fn a similar manner, the S-wave solution outside the well is
or (1 1.15)
tan(kra * do)
-kz& * do,
and then
do 1 k1a - k2a
!rr^"
or
sindo-aa(W-t).
From Equation 11.15, the total scattering cross section for S-waves is
d:4na',(w- tl (1r.16)
The scattering cross section becomes infinite here for kra : nl2, 3n12, . . .
where the tangent function becomes infinite. Since Equation I l. l6 was derived
from the assumption that do 4 kfi, the infinite cross section is inadmissible.
Returning to the exact expression, Equation 11.15, we see that sin do
k1a: nl2 andkra 41, and so the maximum cross sectionis - I when
4r
ures-7. (11.17)
fv2
PROBLEM I1.1
2. SCATTERING AS A PERTURBATION:
THE BORN APPROXIMATION
Time-dependent perturbation theory may be applied to the scattering
problem by treating the perturbation potential as being "turned on" when the
incident particle gets within its range, and "turned off"'again when the scattered
particle exceeds its range. The scattering process is thus regarded as a time-
dependent transition from an initial plane wave state lf) to a final plane wave
state l;f). The transition rate per incident particle is given by Fermi's Golden
Rule, Equation 9.54, as
R: Ir<ttvti>t, p(Et), (11.18)
where Z is the scattering potential and p(E) is the density of final states.
Since the final states here are free particle states, we may use the electron
gas density of states, Equation 10.20, omitting the factor of two for spin states,
p(E):mw:m (r1.1e)
We are interested in those states for which the vector E, ties within a small
solid angle centered around Er. Therefore, the density of states, Equation
'11.19, must be multiplied by the factor dQl4r. Then the transition rate per
unit solid angle centered around li, due to a particle flux of intensity
',ruhko
:-:-
Ls Lsm
ls
R -do
do - do'
-:l-
SCATTERING AS A PERTURBATION 403
R
do
-
do- /do -- \ffi)' *,. t<-n v ti>t, (r 1.20)
and
h: ffill,t-.nL',v(,)aEni d,f ,
-f (o): - #.[,ng1rn'Eo-E,t'r
4,. (1 1 .21)
This integral is known as the Born approximation. Since the volume normalization
factor Z-8 drops out, the integruiiot is independent of the size of the box.
Hence, contributions to the scattering cross section arise only where V(r) is
not regardless of the limits of integration on r. This approximation is
-small,
equivalent to assuming that only single scattering events u.. i*portant, and
thus the first-order term in V(r) is all that app.utrl.t Equation I 1.21. We will
see in the next section how multiple scattering events are treated.
Equation 11.21 may be simplified by means of the substitution
R:Eo-Et
^R is given by K :
Since fo and E, are the same length, the magnitude of
2,to_'.sin 012, where 0 is
the angle of scattering between Eo und, E, (see Figure
11-2). By defining an auxiliary set of polar ioordinates with the'polar"axis
directed along R,
subject, but we may make a few general remarks in the context of the assump-
tions used here. First, we have used perturbation theory, which presupposes
that the amplitude of the scattered wave is small compared to that of the in-
cident wave and that lv(r)l < E, the energy of the incident particles. Thus,
the energy of the incident particles must be large enough so that they may be still
represented by plane waves after the scattering occurs.
As an example, let us apply the Born approximation to the scattering of
electrons by u screened Coulomb potential,
vb\:
\'./ - -Zt'r-'/o
r " t
where a is the screening radius. From Equation Ll'.22,
- K "4]; cos
K2+0'
ful'-''"(-:sinKr
Thus, the scattering amplitude may be'written in the form
-f(o):w[,_.#l (r 1.23)
PROBLEM IT.5
PROBLEM I I-6
PROBtEtr4 tl-8
v(,) -- *vo"*o
[-(il'
using the Born approximation.
pQ):lrtt-
J
r')p(i')di', (11.25)
where "f d(i - i') di' : !. Suppose that a solution G(i,i') can be found for
the equation
(V, + kz)G(i,f') : -4tr6(i - i'). (11.26)
G(i,i'):+#, (11.31)
1.yi3a;;IS;13";;J;fjg;q;fiflgf*g3u;ggl*.fJ*;:;.ff$t.rif,+j:i*"_e"';11lS.jfrf:riJ.:i;#:fff:jij;y:,efFj
,ttcli-i'l,ittcr-i'.i'I
A/-
lf-i'l r '
and
._m
Vsc : eLkoz
- 6-
znhz
' '#l--' i'v(i,)ve,) di'. (11.33)
Scattering regiou
Figure I l-3 Schematic diagram of the scattering process for an extended scatterer.
SCATTERING THEORY
Comparing this with Equation 11.1, we see that the scattering amplitude is
given by
where the scattering angle, 0, is the angle between the incident beam and i.
A calculational difficulty still exists in that the total scattered wave function
appears in the integrand of Equation I 1 .3+. Note that if this total wave function
is replaced by only the frst term of Equation LL.32, namely, the incident plane
*urri, then Equation Ll.3+ is identical with the result for the Born approxi-
mation, Equation 11.21. An analytical justification for this substitution may
be sketched as follows. Assuming the convergence of iterative solutions, the
solution for g(i') may be written as
where pr?') : d(i') I p(i'), is the total charge density, that is, nuclear plus
electronic, at the point i'. Then Equation ll.2l becomes
-f(0):#, IIffidf,di
mZez f I ,rt.r:f.r'r:,!"
:-2m))er7)- di, di
V_i,l
:# ! o'{'')'rt''' di' ['#u "
The integral on the right is evaluated by taking f' :0 as the origin for the
integration over i and by including the inlegrating factor e-nr. Then,
"i
I'+di -:l
n "it.i PZt Ptt r.a
,[ J*,-",frirdrdodo
- tz'lim f*r-,,,'*'Joo
-'";5Jro a, lt ,ifl'r"or o 4\rr\
d(cos 0)
:+:T Kr dr
I-'-o'sin
: 4n-. K : 4n
t"{
T o,t + r, K,
It follows that
-f (o)
:ffi2mZe2 r .?
Jer{r,)riK.i'4i,
:WJ tat'l - p|)re&,'di'
:ffif'- Ip4)eikr di'l
:Ytr-r(,qr. (11.36)
4IO SCATTERING THEORY
Here F(K) is the atomic scattering factor, which is a measure of the amount of
shielding of the nuclear charge by the electrons. It is given by,
l+
F (K) : I p1)tiK i'7i' . (11.37)
J
Note that the atomic scattering factor is the Fourier transform of the electronic
density. For a spherically symmetric charge distribution, Equation 11.37 may
be integrated over the angle variables and simplified to
[n",.:
r(^K) : eXp t- (?l] ]
,*i.s,-&iS:.Ei#i1lSF.fd.+.SrS1E#9
The quantity hzKz in Equation 11.36 can be written in terms of the energy
of the incident electrons by means of the relationship derived from Figure LL-2,
hzKz : rrr 9
+hzkzrirr -- v,'eu 9LLL 9
BmEsirrr .
2 - 2.
Then,
f(o):frt 4E sinz ,
-F(K)1. (r 1.3e)
This gives the Rutherford result for large angle scattering, Equation 3.!5,
when th. .n.. gy E is much greater than the ionization energy so that the
shielding is ineffective. However, for very small scattering angles the shielding
THE EFFECTS OF EXCHANGE SYMMETRY AND SPIN 4il
do
-'"f(0)l'+l"fb-0)1"
dC-- |
_ (n 0)l'.
spatialfunction)
fO: l-fQ) -f -
(antis2mmetric (11.41)
su/vl/vlARv
Three approaches to scattering problems have been introduced in Chapter
I 1. In the partial wave method, the incident particle is represented by a
plane wave which is expanded in the Legendre polynomials and the spherical
Bessel functions. The scattering event is then treated as the superposition of
this incident wave and a spherical wave generated by the scattering center.
By comparing the forms of the scattered and incident waves at large distances
from the scatterer (the asymptotic limit), the effect of the scatterer can be
described in terms of phase shifts in the arguments of sine and cosine functions.
The calculation of these phase shifts constitutes a solution to the problem.
In the perturbation method, a scattering event is regarded as a transition from
one state in the continuum of positive energy states to another. Hence, the
mathematics used in calculating transition probabilities, including Fermi's
Golden Rule, can be applied directly. The Green's function method is equiv-
alent to using an infinite perturbation expansion. It can be applied to both
elastic and inelastic scattering, and multiple scattering events to any order can
be included. The Green's function is regarded as a propagator which carries
the particle from one scattering center to another. The Born approximation;
whiih is valid when the particle energy is large with respect to the interaction
energy, is derived and applied to simple potentials.
SUGGESTED
REFERENCES
David Bohm, Qwntum Theorlt. Prentice-Hal, fnc., New York, 1951.
R. H. Dicke and J. P. Wittke, Introduction to Qtnntum Mechanics. Addison-
Wesley Publishing Co., Inc., Reading, Mass., 1960.
E. Merzbacher, Quantum Mechanics. John Wiley and Sons, fnc., New York,
1961.
Albert Messiah, Qtnntum Mechanics. North-Holland Publishing Co., Amster-
dam, 1958.
D. Park, Introduction to the Qu.antum Theor2 McGraw-Hill Book Co., Inc., New
York, 1964.
J. L. Powell and B. Crasemann, Quantum Mechanics. Addison-Wesley Publish-
ing Co., Inc., Reading, Mass., 1961.
D. S. Sixon, Elementarlt Quantum Mechanics. Holden-Day, fnc., San Francisco,
1968.
Leonard I. Schifl Quantum Mechanics,3rd ed. McGraw-Hill Book Co., New
York, 1969.
ffipnm*sffi,$ffi
415
416 AUTHOR INDEX
Panofsky, W. K. H.,26,42,3M
Illingworth, K. K.,4 Park, D., 157, 203, 288, 338, 378, 395, 412
Ives, H. 8.,37 Parker, W.H.,47
Pauli, W.,289
Pauling, L.,157 , 189, 203, 273,288,378, E95
Penney, W. G., 174
Phillips, M., 26, 42,364
Jackson, J. D.,42, 864 Phipps, T. E., 108
Jeans, J. H., 46 Planck, M.,47
Jeffreys, H.,372 Powell, J. L., 140, 157, 203,245,288, 338, 378,
Jiinsson, C.,127 395,4r2
Jordan, P., lll Pringsheim ,8., 44
Josephson, B. D.,47
Joy, A. H.,5
Rado, G. T.,290
Rainwater, J.,86
Kiillen, G., 162 Ramsey, N. F., 103
Kemble, E. C.,374 Rayleigh, Lord, 46
Kennard, E. H.,42,72,89, ll0, 157,203,288 Resnick, R.,42
Kennedy, R. J., 4,6 Richtmyer, F. K.,42,72,89, ll0, 157,203,288
Kimball, G.8., 395 Rojansky, V.,157,245
Kjellman, J., 5 Rose, W. G.V.,42
Kramers, H. A.,372 Rosser, W. G. V., 42
Kronig, R. de L., 174 Rotenberg, M.,322
Royds, T.,73
Rusell, H. N.,312
Rutherford, 8., 7 3, 75, 80
Langenberg, D. N.,47
Lauritsen, C. C., 32, 89
Lawrence, E. O.,55
Sadeh, D.,5
Lederman, L. M.,286
Sanders, J. H., 8
Lee, T. D.,286
Sands, M., 72, ll0
Leighton, R. B., 42,72,110, 202, 245,277,279,
Sanford, R. F., 5
333, 339, 378
Saunders, F. A., 312
Lengyel,8.,54
Saxon, D. S., 157, 203, 288, 338, 378, 995,412
London, F., 390
Schey, H. M., 170,171,172
Lummer, O.,44
Schiff, L. 1., 157,203,245,278, 288, 338, 378,
381, 389, 391, 395, 4W,412
Schriidinger, E., lll, 142, 339
Schroeer, J. M.,57
Marsden, 8.,13,74 Schwartz, I. L., l70t 17 l, 172
Matthews, P. T., 202,245,288 Scott, G. D., l4
May, J. W.,124 Semat, H., 72, I l0
McDonald, D. F.,8 Shadorvitz, A.,4L,42
Merzbacher, 8., 157, 190, 203, 245, 264, 288, Shankland, R. S.,6
339, 378, 995,400,4t2 Sherwin, C. W., 198
Messiah, A., 120, 149, 157,245,288,338, 372, Shortley, G. H.,281, 314, 322,338
378,395,412 Slater, J. C., 804, 338, 388
Metropolis, N., 322 Slichter, C. P., 106
Michelson, A. A.,4 Smith, J.H.,42
Millikan, R. A., 54,56 Smith, R. L.,57
M6ller, C.,42 Snyder, C. W., 32
Morley, E. W.,4 Sommerfeld, A., I12, 176
Morse, H. G. J.,90 Sparks, M., 106
Miissbauer, R., 96 Spiegel, M. R.,276
Mulligan, J.F.,8 Stern, O., 107
AUTHOR INDEX 417
290
Stevens, K.W.H., Wapstra, A.H.,322
3
Stewart, A. B., Watson, G. N., 397
37
Stilwell, G. R., Weber, J., 100
72,ll0
Stoner, R. G., Weber, R. L., 72, 100, ll0
Suhl, H.,290 Weinrich, M.,286
Weisskopf, V. T., 14, 386
Wentzel, G.,372
Taylor, B. N.,47 \Vertheim, G. K.,96
Taylor, J.8., 108 Whecler, J. A., 86
Teich, M. C., bZ White, H. E., 310, gl4, g3g
Thomas, L. H., 294, Bgg White, R. L., lS7 ,209,248,299, Bgg
Thomson, G. p., 126 Wien, W.,44
Thomson, J. J.,54,73 Wilson, C. T. R., 59
Thorndike , E. M., 6 Wilson, E. B., 157, 189, 203, 27g,2gg, Z7g, ggb
Tipler, P. A.,72, ll0 Wilson, W., llz
Trouton, F. T.,4 fVittke, J. p., 1b7,202,244,2gg, B3g, g?g, 3g5,
4t2
Wolga, G. J., b7
Wollan, E. O., 125
Uhlenbeck, G. E., 102, 289 \ivooren, J. K., !r.,322
Wu, C. S.,286
-*i.&qffi
SUffiJffiCT ilNPffiK
419
420 SUBJECT INDEX
Matrix form of the eigenvalue problem, 230, Normalizat ion (C ont inue d)
23r of harmonic oscillator wave functions, 183-
Matrix mechanics, lll 188, 196, 197
application to the harmonic oscillator, 240- Nuclear, atom, 73-ll0
243 barrier,164-166
Maxwell velocity distribution ,70-72, 109 binding energy, 93
Maxwell's electromagnetic equations, 3, 8, 349 force, 80, 93, 94, 95, 100
Maxwell-Boltzmann distribution, 48, 51, 53, magnetic moment, 103
67 -70, 394 magneton, 103
Measurement, of compatible observables, 2I2- models,98, 99
2t4 liquid drop,98
Mesic atoms, 86 optical, 98
Meson, 100 shell, 98
mu, 16,86, 87 radius, 80, 92, 93, 94, ll0
pi,34, 35,38, 100, 101 spin, 104
Meson theory of nuclear force, 100, l0l structure and spectroscopy, 92-99
Metastable state, 52, 53 Nucleon, 92
Methods of generating Hermite polynomials, Nucleon number, 92
180-183 Null operator,220
Michelson-Morley experiment, 4-8 Null vector,2l5
Minimum uncertainty product, see Uncer- Number operator, 202
tainty product.
Minkowski diagram, 40, 4 I
force,24
Mirror nuclei,94 Observables, operators, and expectation values,
Mixed state, 109, 210 t46-I50,207
Molecular vibrations, 189-t9l Odd and even functions, see Symmetric and
Molecules, homonuclear antisymmetric.
nuclear spin eftects, 329, 330 Old quantum theory, l12-114, ll9
ortho and para forms, 329,330 One-electron atom, see Hydrogen atom and
rotational angular momentum, 329 hydrogenic atom.
rotational states, 329ff Operator, angular momentum, 252-257
Momentum, conservation of, angular, 254,256 anti-Hermitian, 2l I
linear,254,255 creation and annihilation, 202, 257
four-vector, 24 electric dipole, 361
operator, 148, 149 energy, 148
representation, 139 equation of motion,2l4
Morse potential, 189, 190 gradient in spherical coordinates, 283
Moseley's law,90 Hermitian, 154, 207 -2ll
M<issbauer eftect, 96, 97 idenriry, 220
Multiple square wells, 174, 17 5 inverses, 221
Multiplicity, 814 Laplacian in spherical coordinates, 251
Multipole radiation, 363ff linear, 207
Muon, 16,86,87 maBnetic moment,284
Muonium, 86 matrix representation of, 224-230
Mu-mesic lead,87 non-singular, 221
null,220
parity, 284ft
particle exchange, 303, 306, 331
Natural linewidth, see Linewidth. projection, 223, 224, 322
Natural units, 156,157 raising and lowering, angular momentum,
Neutrino, 305 257
Neutron, magnetic moment of, 103 harmonic oscillator, 193
number, 92 right and left inverses, 221
spin, 102, 305 rotation,256
statistics, 305 spin-orbit, 320,325
Non-singular matrix, 225 total angular momentum, 254
Non-singular operator, 221 translation, 255
Norm of a vector,2IT Operator solution of the harmonic oscillator,
Normalization, 129, 130 l9t-200
angular momentum eigenfunctions, 264-268 Optical model of nucleus, 98
box,247 Optical theorem in scattering theory, 399
delta function, 138, 139 Orbital angular momentum, 83, 84, l0l, 104,
426 SUBJECT INDEX