Essential Quantum Physics
Essential Quantum Physics
Essential Quantum Physics
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:15 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:15 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
Essential Quantum Physics
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:15 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:15 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
Essential Quantum Physics
Peter Landshoff
University of Cambridge
Allen Metherell
University of Central Florida
Gareth Rees
University of Cambridge
CAMBRIDGE
UNIVERSITY PRESS
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:15 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore,
Sao Paulo, Delhi, Dubai, Tokyo
Published in the United States of America by Cambridge University Press, New York
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:15 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
Contents
1 Preliminaries 1
Atoms; Photons; Wave nature of matter; Problems
3 Special solutions 19
Particle in a box; The one-dimensional square well; The lin-
ear harmonic oscillator; The tunnel effect; The delta-function
potential; The WKB approximation; Alpha decay; Problems
Revision quiz 62
vn
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:26 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
Vlll
8 Spin 82
Two kinds of angular momentum; Spin | ; The electro-
magnetic interaction; The Zeeman effect; Spin precession;
Problems
12 Transistors 141
Impurities; n- and p-type semiconductors; Impurities and
crystal colour; Semiconductor junction; The diode; The junc-
tion transistor; Two simple circuits; Problems
Appendices
A Power-series solutions 158
B The delta function and Fourier transforms 162
C Orbital-angular-momentum operators 173
D Electrodynamics 175
E Bloch waves 178
Index 201
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:26 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
Preface
IX
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:38 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.001
Cambridge Books Online © Cambridge University Press, 2014
Constants of quantum physics
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:09:48 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.002
Cambridge Books Online © Cambridge University Press, 2014
1
Preliminaries
Atoms
An atom consists of a positively charged nucleus, together with a number
of negatively charged electrons. Inside the nucleus there are protons, each
of which carries positive charge e, and neutrons, which have no charge.
So the charge on the nucleus is Ze, where Z, the atomic number, is the
number of protons. The charge on each electron is —e, so that when the
atom has Z electrons it is electrically neutral. If some of the electrons are
stripped off, the atom then has net positive charge; it has been ionised.
The electrons are held in the atom by the electrostatic attraction between
each electron and the nucleus. There is also an attraction because of the
gravitational force, but this is about 10~40 times less strong, and so may
be neglected. The protons and neutrons are held together in the nucleus
by a different type of force, the nuclear force. The nuclear force is much
stronger than the electrical force, and its attraction more than counteracts
the electrostatic repulsion between pairs of protons. The nuclear force
does not affect electrons. It is a very short-range force, so that it keeps
the neutrons and protons very close together; the diameter of a nucleus is
of the order of 10~15 m. By contrast, the diameter of the whole atom is
about 10~10 m, so that for many purposes one can think of the nucleus
as a point charge. The mass of the proton or neutron is some 2000 times
that of the electron, so nearly all the mass of the atom is in the nucleus.
It is natural to think of the electrons as being in orbit round the nucleus,
figure 1.1, just as the planets are in orbit round the sun. The electrostatic
force that keeps the electrons in their orbits is an inverse-square-law force,
just as is the gravitational force that keeps the planets in orbit, so that the
two systems would seem to obey precisely similar equations. However,
there is a serious difficulty. When a particle moves in a curved orbit its
velocity vector is continuously changing: the particle is being accelerated
towards the centre of its orbit. According to classical electrodynamics,
when a charged particle is accelerated it inevitably radiates energy (this
is the basic principle of radio transmission). So according to classical
physics the electron would continuously lose energy and its orbit would
form a spiral which would gradually collapse into the nucleus.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:02 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.003
Cambridge Books Online © Cambridge University Press, 2014
1 Preliminaries
Figure 1.1. Classical picture of negatively charged electrons in orbit round the posi-
tively charged nucleus of an atom.
The reason that this does not happen is that very small systems, such as
atoms, do not obey classical mechanics. To describe an atom one has to
use quantum mechanics. In quantum mechanics, as opposed to classical
mechanics, one cannot arbitrarily choose a value for the energy of the
orbiting particle and then find an orbit corresponding to that energy; only
certain discrete values of the energy are allowed. When the electron is in
its lowest allowed energy level, it cannot radiate any more energy, and so
the total collapse of the atom is not possible.
One can also use quantum mechanics to describe the solar system. Just
as for the electrons, the allowed energy levels of the planets are discrete.
If a planet in orbit is given an impulse, its energy is allowed to change
only to that of one of the other allowed discrete levels. However, the
separation between these levels is so small that this is not a very real
restriction, and classical mechanics is perfectly adequate to describe the
system. The effects of quantum mechanics are generally only important
for submicroscopic systems.
The chemistry of an atom is determined by the charge on its nucleus.
Thus atoms whose nuclei differ only in the number of neutrons that they
contain have similar chemical properties; they are said to be isotopes
of the same element. For example, the atom of the common form of
hydrogen contains just a single proton, that is, Z = 1; but hydrogen has
a stable isotope, called deuterium, whose nucleus consists of one proton
and one neutron. Atoms can be bound together to form molecules (see
chapter 6), and different isotopes of the same element do this in the same
way. Ordinary water H2O consists of molecules containing two hydrogen
nuclei and one oxygen nucleus, while 'heavy' water D2O has deuterium
nuclei instead of the ordinary hydrogen. The chemical properties of heavy
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:02 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.003
Cambridge Books Online © Cambridge University Press, 2014
Photons
water are exactly the same as those of ordinary water, but there are some
differences in its physical properties. In particular, it is denser because of
the extra neutrons.
Photons
In a metal, the atoms are effectively anchored to fixed sites by the electro-
static forces due to all the other atoms. The outermost orbital electrons of
the atoms are almost free, and move through the metal when an electric
field is applied (see chapter 11). If the metal is bombarded with light,
some of the electrons can actually escape from the surface of the metal
and can be detected as an electric current. This is the photoelectric ef-
fect. The number of electrons that escape in a given time rises with the
intensity of the beam of light, but the energy with which they escape does
not depend on the beam intensity. Rather it depends on the colour or
frequency v of the light. The kinetic energy T with which the electrons
escape is found to obey the equation
hu = T + W. (1.1)
h = 6.626 x 10" 34 J s,
E = hv. (1.2)
The electron is ejected from the metal when one of the photons collides
with it and is absorbed by it, so giving up all its energy to the electron.
The number of electrons ejected rises as the intensity of the light is in-
creased because there are then more incident photons, and so there is a
greater chance of a photon being absorbed.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:02 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.003
Cambridge Books Online © Cambridge University Press, 2014
4 1 Preliminaries
Photons move with the speed of light, so their kinematics must be de-
scribed by the laws of special relativity. The energy of a particle whose
speed is v and whose rest mass is m is
E = mc2/{l-v2/c2)ll\ (1.3a)
so that when v = c the energy can be finite only if m = 0; that is, photons
have zero mass. In terms of the momentum p of the particle, (1.3a) reads
E = c(m2c2+p2)1'2, (1.36)
hv = E2 -Ei.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:02 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.003
Cambridge Books Online © Cambridge University Press, 2014
Wave nature of matter
Slits Screen
Figure 1.2. The double-slit experiment. There is darkness at points on the screen such
that the path difference between rays that pass through the two slits is (n -f ^
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:02 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.003
Cambridge Books Online © Cambridge University Press, 2014
6 1 Preliminaries
A - h/p. (1.5)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:02 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.003
Cambridge Books Online © Cambridge University Press, 2014
Wave nature of matter
Problems
1.1 A radio transmitter operates on a wavelength of 100 m at a power of
1 kW. How many photons does it emit per second?
1.2 Using energy-momentum conservation, show that an electron that is
not in a bound state cannot absorb a photon.
1.3 A particle has mass 1 kg. How long does it take to move through a
distance of 1 m if its de Broglie wavelength, (1.5), is comparable with
the wavelength of visible light? What is the corresponding answer if
the particle is an electron?
1.4 A photon of momentum p, and therefore of wavelength /i/p, scatters
on an electron that is initially at rest. Using relativistic kinemat-
ics, deduce from the conservation of energy and momentum that as
the result of the scattering the wavelength of the photon changes by
(h/mc)(l — cos#), where 9 is the angle through which it scatters and
m is its rest mass. (This scattering process is known as Compton
scattering, and the quantity h/mc is the Compton wavelength of the
electron.)
1.5 Associated with the electron there is an antiparticle, the positron,
which has equal mass and equal, but opposite, charge.
A positron impinges on an electron which is at rest. They annihilate
into two photons. Show that the sum of the wavelengths of the two
photons is Ao(l — cos#), where 6 is the angle between their directions
of motion and XQ is the Compton wavelength of the electron.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:02 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.003
Cambridge Books Online © Cambridge University Press, 2014
2
The Schrodinger equation
E=p2/2m. (2.1)
and the wave vector fc, whose direction is in the direction of wave propa-
gation and whose magnitude is
k = 2TT/A.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
Wave functions and operators
p = hk
E = hu. (2.2)
For a free particle, which is not interacting with any other particle or with
a potential, p and E are constant. Hence we expect such a particle to be
described by a wave for which k and u are constant:
(-iftV)tf = p * . (2.46)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
10 2 The Schrodinger equation
qV. (2.5)
(2.7a)
f Although H is an operator, we shall follow the usual convention and not write the
operator symbol " over it.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
Example: the one-dimensional potential well 11
or
HV = ihdV/dt. (2.8b)
This is the time-dependent Schrodinger equation.
This is the equation that is valid in all circumstances whether or not the
particle is free and whether or not it is known to be in a state of definite
energy. In the particular case where the particle is known to be in a state
of definite energy E, the time-independent equation (2.7) is valid also.
When both equations hold,
We repeat that a solution of the form (2.9) is applicable only when the
energy takes a definite value. For reasons that will be explained below, a
solution of this type is known as a stationary-state solution. Comparing
(2.9) with (2.3), we see that in the special case of a free particle
if>(r)=Neik'r (2.10a)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
12 2 The Schrodinger equation
0 0<x<a ( 2 n )
oo otherwise.
and its general solution for ^ is a linear combination of smkx and coskx,
where E — Ti2k2/2m.\ As we shall see at the end of this chapter, ip(x)
has to be continuous; in particular, it has to be continuous at x = 0 and
x — a. Hence, since it vanishes for x < 0 and for x > a, we must impose
the conditions tp(O) = ip(a) — 0. The condition at x = 0 picks out the
sin&x solution for -0, and that at x — a imposes a restriction on the
allowed values of k:
k = nn/a n = 1,2,....
f Throughout we shall use primes to denote differentiation with respect to a space vari-
able and dots to denote differentiation with respect to a time variable.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
Probability interpretation and normalisation 13
We could also allow negative integers n, but this would not give any
additional wave functions; the allowed wave functions are
where the Cn are normalising constants (we explain below how these are
chosen). The corresponding allowed energy values are
That is, the allowed energy values for the bound system are discrete.
(2.14)
/ •
Here we have used the symbol / d 3 r as a shorthand for JJJ dx dy dz, and
the integration is supposed to extend over all space. So (2.14) corresponds
to saying that there is unit probability that a measurement of the position
of the particle finds that the particle is actually somewhere.
If we have found a solution # to the Schrodinger equation that is not
correctly normalised so as to satisfy (2.14), we can usually get one that
is by simply multiplying by a suitable constant. An exception is the
special case of the plane wave (2.3), for which the normalising integral
(2.14) diverges for all values of the multiplying constant N. The reason
for this is that strictly the plane wave does not correspond to a physical
situation. It has |\I>|2 = \N\2, independent of r, so that it gives equal
probability of finding the particle anywhere throughout space. In practice,
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
14 2 The Schrodinger equation
one knows that the particle is confined within some volume £)o> f° r example
within a given building, and that therefore the wave function vanishes
outside fio- No problems arise for finite f2o> &s then the normalisation
integral converges, and by dividing ^ by a suitable finite constant we can
make it converge to 1, as is required in (2.14). The plane-wave solution
(2.3) is best thought of as a mathematical limit, where QQ —>• oo, of a
physically allowed solution. Apart from the normalisation problem, which
can in fact be circumvented by methods that we shall not discuss here, the
mathematics of the f2o —> oo situation is rather simpler than that of the
proper wave functions, and if the volume QQ is large (for example when
fi0 represents the volume of a typical experimental apparatus, which is
almost always very large when measured on a quantum-mechanical scale),
the numerical results are changed very little by taking the limit QQ —> oo.
Suppose that the particle is confined to a finite volume £2o> so that its
wave function * vanishes outside that volume. Then if \l/(r, t) is correctly
normalised at a given time i, it remains so. For, if we integrate over any
volume fi,
d
* Jn Jn
= Jn[ d r q3
3
(*** + * * * ) . (2.15a)
To evaluate this, we use the Schrodinger equation and its complex conju-
gate:
[(-h2/2m)V2
[(-h2/2m)V2 +
Thus
—
at 2miJa '
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
Probability interpretation and normalisation 15
Figure 2.1. The integration volume for (2.15), and its bounding surface S.
=- [ [dS-j
Js
where
j = (ft/2rai)(\I>*V\I/ - * V * * ) . (2.15c)
This relation is valid for any volume Q. For the normalisation integral
(2.14), we have to integrate over all space. However, since we are assuming
that \I> vanishes outside some volume f2O5 it is sufficient to integrate over
any volume that contains fi0- Because the surface S is then outside the
volume $~20? ^ , and therefore j , vanishes all over it, and so the integral
(2.15b) vanishes.
What we have said so far about normalisation applies to any wave function
\I/(r, t). When the wave function corresponds to a state of definite energy,
f The divergence theorem states that for any suitably well-behaved vector U
= f dS u.
Js
Here the vector dS is a vector whose magnitude is equal to an element dS of
the area of the surface S bounding the volume Q, and its direction is that of the
outward normal to S. See, for example, H Jeffreys and B S Jeffreys, Methods of
Mathematical Physics, (Cambridge University Press)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
16 2 The Schrodinger equation
,*)| 2 = h/>(r)|2
so that the probability density at each point r is independent of time.
This is the reason that we call a state of definite energy a stationary state.
Beams of particles
So far, we have considered a single particle, such as an electron. Exper-
iments often involve beams of particles, for example a beam of electrons
from an accelerator fired at an atom. If the beam is not too intense, that
is, the particle density is not too high, each electron interacts with the
atom independently of the presence of the other electrons, and interactions
between the electrons are negligible. This means that we may describe the
problem in terms of the same Schrodinger equation as describes a single
electron moving in the potential produced by the atom.
Suppose that when there is just one electron the potential produced by the
atom results in a wave function \&(r, t). In the case of an electron beam,
the same wave function describes the density of particles at the point r at
time t. However, it is convenient then to normalise \I> not by the integral
(2.14) appropriate for a single-particle problem, but instead such that the
integral Jd 3 r|\I/(r, i)\2 is the total number of particles at time t. Then
(2.17)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
Continuity conditions 17
j A ( . (2.18)
2rai
In most cases, the potential and the rate at which the beam delivers par-
ticles to the target vary very slowly, if at all, on the atomic timescale.
Hence there is effectively a stationary-state situation, with the wave func-
tion factorising as in (2.9) and p and j independent of t.
Continuity conditions
The Schrodinger equation (2.8) is supposed to be valid at all points r.
This means that if the potential V is well behaved, the wave function and
its space derivative must both be continuous functions of r. For if this
were not so, the second derivative V 2 ^ would be infinite at the points of
discontinuity, and this cannot be because there cannot be just one term
in the Schodinger equation that has an infinity. The infinity would have
to be balanced by another infinite term and there is no such term unless
V has an infinity.
The physical consequences of this are not surprising: both p(r,t) and
j(r,t) have no discontinuous changes as r varies. That is, no particles
are being created or destroyed at any point r.
Problems
2.1 How are the stationary-state solutions of the Schrodinger equation
changed if a constant is added to the potential V(r)? Show that
this change has no observable consequences. (This means that, as in
classical mechanics, the point at which V vanishes may be chosen ar-
bitrarily. By convention, V is often defined so as to vanish at infinity.)
2.2 For each of the following, estimate the difference between the speeds
of the particle when it is in the first excited state and when it is in the
lowest-energy state: (i) an electron confined by an infinite square-well
potential whose width is roughly equal to the radius of an atom (about
2Q-io m j . ^ a t e n n i s b a n confined by an infinite square-well potential
whose width is equal to the width of a tennis court. What do your
answers tell you about the mechanics of microscopic and macroscopic
systems?
2.3 Determine the constant Cn in the wave function (2.13a) for the infinite
square well, so as to satisfy the normalisation condition (2.14). Show
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
18 2 The Schrodinger equation
Show that as n —> oo these average values approach the values that
are obtained from classical mechanics.
2.4 Show that for a one-dimensional system in a stationary state the par-
ticle flux (see (2.18)) is independent of both t and x. Does this result
have a simple generalisation to the three-dimensional case?
2.5 A system consists of two particles, so that its stationary-state wave
functions VKri? r2) are functions of the coordinates of each. What is
the obvious probability interpretation of ip?
Write down the time-independent Schrodinger equation. In the general
case, the potential will be a function F(ri, r2). What is the structure of
the function V in the special case where the particles interact only with
an external field of force, and not with each other? Show that in this
case the stationary-state wave functions have the structure ^(ri, r2) =
)V;2(^2)- What equations do the functions ^i and V>2 obey?
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:12 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.004
Cambridge Books Online © Cambridge University Press, 2014
3
Special solutions
Particle in a box
Consider the stationary states of a particle confined within the rectangular
box-shaped volume
Suppose that within the box the potential is zero and outside it is infi-
nite. Within the box, the particle moves freely and the time-independent
Schrodinger equation is
2ra (ITS
~t~ dy2 +dzin
\dx2 + ITS 2
x = 0, y = 0, z = 0,
x = a, y = 6, z = c.
The first three conditions pick out the sin functions, rather than the cos
functions. The last three conditions impose restrictions on the allowed
values of &i, &2 and k3:
. qn rn sir
ki = — , k2 = — , k3 = —,
a b c
19
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
20 3 Special solutions
where q, r and s are integers. Thus the allowed wave functions are
1/2
sm
Vqrs ( r ) = — - — sin —^- sin , (3.4)
v }
\abcj a b c
and the allowed values of the energy are
hn (q2 r2 s2
^ 2m \ a2 b2 cz
Notice that all these allowed values of Eqrs are positive. In (3.4) we
have included the normalisation factor (8/abc)1/2, so as to satisfy the
normalisation condition
the integration being over the volume of the box, or equivalently, since
i\) — 0 outside the box, over all space.
The significant feature of these results is that the allowed values Eqrs
of the energy are discrete, because of the condition that 9, r and 5 are
integers. Notice that we may as well confine them to being non-negative
integers, since changing the sign of any one of them merely changes the
sign of the wave function, a change that has no physical significance.
These are the stationary-state solutions. The corresponding time-vary ing
wave functions are, from (2.9),
c
qrs®qrs{rj), (3.7)
q,r,s=l
where the cqrs are constants, we obtain more general wave functions \I>,
which can be verified to satisfy the time-dependent Schrodinger equation
(2.8) and also have built into them the correct boundary condition that
they vanish on the faces of the box. Imagine we know that we have
introduced the particle into the box in such a way that its wave function
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
The one-dimensional square well 21
(a) (b)
Figure 3.1. (a) The one-dimensional square-well potential. (6) The Coulomb potential
produced by a charged atomic nucleus.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
22 3 Special solutions
The solution to the first equation is a linear combination of elkx and e~lkx,
where E = h2k2/2m. However, for a bound state the particle should be
more or less confined to values of x in the vicinity of the potential, and so
we should impose the boundary condition \ij)\ -> 0 as x -+ ±oo. Hence we
reject the case E > 0, since when k is real the solutions do not satisfy this
requirement. To satisfy the boundary condition, we must have E < 0. If
we write, with a and (3 > 0,
where A and B are constants. (We have rejected a term eax for x > | 6 ,
and a term e~aa; for x < —16, so as to make |^| —> 0 as x —> ±oo). The
solution to the second equation in (3.9) is
= Cpcos |/36 |
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
The linear harmonic oscillator 23
(6) Odd-parity wave functions, which are antisymmetric about the origin,
il>(x) = ~^{-x). For these, A = -B and D = 0, and
The reason for the existence of these two classes of wave function is that
the potential is symmetric about x = 0.
If we combine either (3.12a) or (3.126) with (3.10), we obtain a discrete
set of allowed values for the energy E. Each of these lies in the range
0 > E > - [ / ; that is, the bound states correspond to energies lying
within the well. The equations (3.10) and (3.12) can only be solved by
numerical methods, but we can get some information about the solutions
by graphical methods. If we eliminate a between (3.10) and (3.12), we
find that either
We have sketched the two sides of these equations, against /?, in figure 3.2.
In each case the two curves intersect in just a finite number of pointsf,
so resulting in a finite set of allowed values for E. The number of such
allowed values depends on the magnitudes of U and 6. One can show
that if U < 0, there are no bound-state solutions. This is not unexpected:
the potential is then repulsive instead of attractive, and so has no bound
states.
f Because we have squared equations (3.12) we have to take care to accept only those
points of intersection for which tan ^fib or — cot \fib is positive.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
24 3 Special solutions
central role in classical physics, and the same is true in quantum physics.
So the mathematics that follows, with suitable adaptations, applies also,
for example, to the quantisation of the electromagnetic field, since accord-
ing to classical electrodynamics the electromagnetic field in a light wave
oscillates harmonically at each point of space. Thus, just as we shall find
that in quantum mechanics the energy of a simple linear oscillator with the
potential (3.14) can only take certain discrete values, so also the energy
in an electromagnetic field appears in discrete quanta. These quanta are
what we know as photons.
The time-independent Schrodinger equation for the potential (3.14) is
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
The tunnel effect 25
tl>n(x)=e-m»x2'MHn(x), (3.17)
x,t), (3.18)
n=0
where the constants cn may take any values. For arbitrary choices of
these constants, the probability density at each point x oscillates in time
with period 2TT/U;, though the oscillation is not simple harmonic. Notice
that although each \I/n is a stationary-state wave function and so satisfies
the time-independent Schrodinger equation, this is not true of the wave
function \I> that we get by superposing them on each other. This wave
function satisfies only the time-dependent Schrodinger equation; it does
not correspond to a stationary state.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
26 3 Special solutions
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
The tunnel effect 27
, (3.24)
where T and R are constants. The complete wave function is then (3.24)
in x > a, (3.23) in 0 < x < a and
in x < 0.
Applying the boundary conditions at x = 0 and x — a now gives us
A + R = 5 + C,
a
= «(Be Ka - Ce~Ka). (3.26)
T = ^hA ( 3 27)
eika[e«a(k + IK)2 + e~^a(n + ik)2)' K }
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
28 3 Special solutions
In a stationary state, where the particle density does not vary with time,
this must be independent of rr, so that no particles disappear. (This
particular simplicity of the consequences of the particle-conservation re-
quirement is peculiar to one-dimensional problems; see problem 2.4.) In
x < 0, (3.25) gives
j = (hk/m)(\A\2-\R\2), (3.30a)
and in x > a (3.24) gives
j = (hk/m)\T\2. (3.306)
Putting (3.30a) and (3.306) equal to each other, we retrieve (3.28). Notice
that, according to (3.24), in x > a the particle density is \ip\2 = \T\2.
The factor hk/m in (3.306) is the speed of the particles (see the relation
between wave number and momentum in (2.2)), so that (3.306) says that,
as we expect, the particle flux is equal to the particle density times the
speed. The form (3.30a), applicable in x < 0, is more complicated because
some of the particles are moving in one direction, the rest in the other. In
a more general situation, where the particles do not have a single speed,
the form of j does not have such a simple interpretation.
(For a description of the delta function, see appendix B.) To find the sta-
tionary states, we have to solve the time-independent Schrodinger equation
(3.32)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
The WKB approximation 29
This equation is to be valid for all x. Prom it, we conclude that i/;(x)
is continuous even at x = 0. For suppose that ij)(x) were discontinuous
at x = 0. Then V'O^) would contain a delta-function part (or a worse
singularity if the discontinuity of i\) is infinite) and so ip" would contain the
derivative of S(x) (or worse). This cannot be, because such a singularity
in \j)" is not balanced by a similar singularity in any other term of the
Schrodinger equation (3.32). However, ip'(x) must be discontinuous at
x = 0, so that ij)"{x) contains a delta-function singularity that balances
that of the second term of the Schrodinger equation.
Take the wave function in x < 0 to be an incoming plane wave together
with a reflected wave, just as in (3.25):
%l>(x) = Aeikx + Re~ikx. (3.33a)
Let the transmitted wave be as in (3.24):
i/j(x)=Teikx x>0. (3.336)
In order that the wave function be continuous at x — 0,
A + R = T. (3.34)
f
To obtain the correct discontinuity of ip (x) at x = 0, integrate the
Schrodinger equation (3.32) from x — — e to x = +e and then let e —> 0.
The various terms of the Schrodinger equation give
A /
J -£-e
p£
ip'(x) at x just greater than zero minus its value at x just less than zero.
Hence
(-h2/2m) disc V' + A^(0) = 0 (3.35)
and, from the wave function (3.33),
{-n2/2m)[ikT - ik(A - R)] + XT = 0.
With (3.34), we thus find that
R = m\A/(ih2k - m\),
T = ih2kA/{ih2k - mX). (3.36)
It is straightforward to check that the particle-conservation condition
(3.28) is satisfied.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
30 3 Special solutions
K2 = 2m{U-E)/h2. (3.37)
Then the intensity transmission coefficient \T/A\2 (the ratio of the trans-
mited intensity to the incident intensity) is approximately
(Jf^J>2"a (3.38a)
and, if Ka is large enough,
Because our approximations have led to a result for ln|T/A| 2 that is lin-
ear in Ka, if we follow the first potential barrier by another one for which
the corresponding parameters are K! and a', then ln|T/A| 2 will become
modified by the addition to it of —2n'a'. A smooth, arbitrarily shaped
potential barrier V(x) can be treated as a series of adjacent rectangular
barriers, each of infinitesimal width Ax, as shown in figure 3.3. In the re-
gion xi < x < x2 we may use (3.37) and (3.38b) to write the contribution
to ln|T/^4|2 from one of these infinitesimal barriers:
dx[2m(V(x)-E)/h2]^\ (3.396)
This is the WKB formula, which was derived more rigorously by Wentzel,
Kramers and Brillouin.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
Alpha decay 31
V(x)
\
\
\ \
Alpha decay
We use the WKB approximation to derive a very simple model of the
a-decay of a radioactive nucleus. In this decay, the nucleus emits an a-
particle, which is composed of two neutrons and two protons, so that its
mass M is about 4 times that of the proton and its charge is 2e. The a-
particle is held within the nucleus by a strong attractive force, the nuclear
force. This force has finite range, so that if the a-particle escapes from
it by tunnelling, it is subjected outside the nucleus only to the weaker
Coulomb force.
Consider first the wave function of an a-particle bound within the nucleus
by the nuclear force; we include the effect of the Coulomb force later.
Model the potential corresponding to the nuclear force by a square well,
and simplify further by pretending that the potential is one-dimensional
and that the a-particle moves only in the ^-direction. So the nuclear-force
potential is
' -U 0 < x < i?,
V(x) = (3.40a)
oo otherwise.
A stationary-state wave function corresponding to positive energy E that
satisfies the Schrodinger equation for this potential in the well 0 < x < R
and vanishes at x = 0 and x = R, as is required when the potential is
infinite outside the well, is
with
E + U = h2K2/2M such that sin«i? = (3.416)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
32 3 Special solutions
V(x)
E
0
-U
Figure 3.4. One-dimensional model potential for a-decay.
We now include the Coulomb force, so that for x > R the potential be-
comes
2Ze2
V(x) = (3.406)
where Ze is the charge of the 'daughter' nucleus that remains after the
a-particle has escaped. We leave the potential unaltered for x < i?, so
that it is as shown in figure 3.4.
With classical mechanics, the a-particle cannot escape if its energy E is
less than Vb, the maximum height V(R) of the barrier, but in quantum
mechanics it can tunnel through the barrier. Because the potential is no
longer infinite for x > i?, we no longer require the wave function to vanish
at x — R, and so the wave function (3.41) will become modified. The
coefficients of e1KX and e~1KX will no longer have equal magnitude, and
they will even become time-dependent so that we should use the time-
dependent Schrodinger equation. However, for energies E much less than
Vb the tunnelling probability is small and the wave function in 0 < x < R
is changed only a little by the addition of the Coulomb tail to the potential,
so it is a reasonable approximation to keep the form (3.41); that is, we
can assume that A varies very slowly with t. For very large positive x the
potential is extremely small, so that the wave function there is to a good
approximation Telkx with h2k2/2M = E, as in (3.24). The current j then
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
Alpha decay 33
rR dx\iJ>(x)\ 2
= 2S\A\2R, (3.43)
=s ,
Jo
where S is the cross-sectional area of the nucleus in our one-dimensional
model. As we have said, A varies very slowly with t\ in fact P obeys the
continuity condition (2.17):
dP P
-* - "7 <
with
/M\2 i 7,P2 / M\W
(3.456)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
34 3 Special solutions
Problems
3.1 Sketch the wave functions for the one-dimensional square-well potential
(3.8).
It is found that the average distance of the particle from the centre of
the well can be rather greater than the width of the well. How is this
possible, in terms of your sketches?
3.2 A particle of mass m is bound by a one-dimensional potential well of
depth U and width b. Show that the condition that there is just one
bound state is that
Show that each energy level is reduced by e2£2 /2mu2. What are the
new wave functions? [Hint: replace x by a new variable.]
3.4 By consulting appendix A, calculate the n = 0 and n = 2 wave func-
tions of the linear harmonic oscillator (leave them unnormalised).
Verify that they are orthogonal, that is,
£ dxipo(x)ijj2{x) = 0.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
Alpha decay 35
energy of each particle. Show that in case (i) the particles are nearly
all reflected, while in case (ii) they are nearly all transmitted.
Notice how different these results are from those which obtain in the
corresponding situations in classical mechanics.
3.6 Find the energy level of the bound state for the potential V(x) =
—fiS(x). Verify that this agrees with that for the square-well potential
(3.8) in the limit U -» oo, with Ub fixed and equal to JJL.
3.7 A potential V(x) is such that V(x) = V(—x) and each bound-state
energy level corresponds to only one independent wave function. Show
that each of these wave functions has either even parity or odd parity.
[Hint: the first step in the proof is to show that if ip(x) is a solution of
the Schrodinger equation, so also is ij){—#), with the same value of E.]
3.8 The nuclei IQQFm and 928U have approximately the same radius and
both are a-particle emitters. The a-particle energies are 7.4 MeV and
4.2 MeV, respectively. The half-life of ^oFm is \ hour; assuming that
the probability of the formation of an a-particle is the same, what is
the half-life of jf
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:23 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.005
Cambridge Books Online © Cambridge University Press, 2014
4
The superposition principle
Linear operators
In the examples of chapter 3, we took linear combinations of solutions to
the Schrodinger equation and so constructed further solutions: see (3.7)
and (3.18). The reason that this works is that ihd/dt and the Hamiltonian
operator if, which appear in the Schrodinger equation, are linear, and,
furthermore, the boundary and continuity conditions on the wave function
^ are linear.
By linear boundary or continuity conditions, we mean that if ^ i and \I/2
are two functions that satisfy them, then so does the function (ci^i +
c2\I/2), where C\ and c2 are any complex numbers. This is evidently the
case for wave functions, which we normally require to vanish outside some
region of space, or at infinity.
By a linear operator Q, we mean that
(4.1)
36
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
Wave packets 37
are complete. This means that any allowable # ( r , i ) , that is, any wave
function that satisfies the time-dependent Schrodinger equation and the
appropriate boundary conditions for the system, can be written as a sum
of stationary-state wave functions
Wave packets
As an example, consider a free particle moving in the ^-direction. The
stationary states are the plane waves
- [dpN{p)eipx/h-ip2t/2mh. (4.4)
Here N may take a different value for each value of p: it is now a function
of p. The resulting wave function is not a stationary state; if we wish
to localise the particle in space, it does not have definite momentum or
energy. Rather, (4.4) expresses the wave function as a superposition of
components having different values of momentum p and energy p2/2m.
Setting t = 0, we have
= [dpN(p)eipx'h. (4.5)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
38 4 The superposition principle
That is, N(p) is the Fourier transform of $(x,0) (see appendix B). By
inverting the transform, we may express N(p) in terms of \I>(:r,0):
The wave function (4.7) is called a wave packet: the probability density
|^(x,0)| 2 is localised around the point x = x$. If we prepare the state a
large number of times and each time measure the position of the particle
at t = 0, because |^| 2 is the probability density the average value will be
Ax is called the width of the wave packet. With (4.8) we find that, at
t = 0, Ax = \a.
If we insert (4.7) into (4.6), we obtain an integral whose evaluation is
discussed in appendix B. It gives
2 2 2
N(p) =
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
EhrenfesVs theorem 39
This has the same functional form as #(£,0), a property that is peculiar
to the Gaussian. One can check that J dp\(/)(p)\2 = 1, where
d>(p) = [l/(2nh)1/2]N(p).
AxAp= \U.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
40 4 The superposition principle
Ehrenfest's theorem
Having obtained N(p) in (4.10), we may now use (4.4) to calculate *(x, t)
for t / 0 (see problem 4.2). The result is that ^(x,t) remains Gaussian
in shape at subsequent times t. The average position (x) of the particle
moves with velocity po/m, and the width Aa: of the wave packet increases
with increasing time. The reason that the wave packet spreads out as t
increases is that the different momentum components of the wave function
correspond to different velocities p/m.
There is a theorem, due to Ehrenfest, that states that whatever the shape
of the wave packet, and for a general potential V,
and
(d/dt)(p) = <-W>, (4.146)
where
= /*d3rW|«f
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
Hermitian operators 41
Hermitian operators
We have said that it is assumed in quantum mechanics that to each ob-
servable Q there corresponds a linear operator Q. We now introduce an
additional assumption about this operator: that it is Hermitian.
Let ^ i and ^2 be any two of the allowed wave functions for the system
under discussion. We write
(4.15)
In this definition the integration is, as usual, over all space. The operator
Q is said to be Hermitian if, for all choices of \&i and #25
*, (4.16a)
that is, if
fd3r^l(Q^2) = /d3r(Q#i)*#2. (4.166)
where
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
42 Jf. The superposition principle
Take the complex conjugate of (4.17) and use (4.16a) for the case \&i =
To show this, since by definition Q^i = qi^i and Q\&2 = #2^2, we write
the chain of equations
q2 (
Here we have used (4.16a) and the fact that qi is real. So if q\ ^ q2->
we have the required result. If qi = q2, we cannot conclude that \I>i
and \&2 are orthogonal. However, in this case any linear combination
(ci^i +02^2) is also an eigenfunction of Q, with the same eigenvalue, and
we can choose two linear combinations that are orthogonal (see problem
4.7). If we normalise each eigenfunction, so that it satisfies ( ^ , ^ ) = 1,
we then have an orthonormal set:
(4.18)
#r. (4.19)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
Operators and observables 43
To verify this, use the orthonormality condition (4.18). From this condi-
tion we find also that
(*.,*) = c , . (4.21)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
44 4 The superposition principle
(4-23)
That is, the probability of getting some result for the measurement is 1.
Let us now recapitulate our basic assumptions about observables:
If we somehow prepare the same state ^ many times, and on each occasion
measure Q at the same relative instant in time, the average answer will be
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
Commutators 45
Commutators
Suppose that we measure an observable Q and then another observable
R. The first measurement will throw the system into an eigenstate of
Q. In general, this will not be an eigenstate of R, so that the second
measurement will disturb the system again. Thus, by measuring i?, we
lose the information we have about Q.
This is another manifestation of the uncertainty principle, which we have
already encountered for the particular case of the observables r and p.
In general, one cannot know the exact values of two observables Q and R
simultaneously, because generally their eigenstates are not the same.
However, some pairs of observables may simultaneously be measured ex-
actly. It can be shown that two operators Q and R have a complete set of
common eigenstates if, and only if, their commutator
vanishes. That is, if we operate on any wave function first with Q and
then with R, the result is the same as when we apply the operators in the
reverse order. We then say that the operators Q and R commute. So the
condition that we may simultaneously know the values of two observables
is that the corresponding operators commute.
The generalisation of the uncertainty principle (4.13) for Ax Apx to an
arbitrary pair of observables can be shown to be (see problem 4.15)
(4.27)
On the right of this inequality there appears the modulus of the expectation
value of the commutator of the corresponding operators, in the state for
which the measurement is being made. The commutator is, of course,
itself an operator. For the particular case of components of position and
momentum, it turns out that the operator is a simple number:
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
46 4 The superposition principle
Thus the two sides of the relation (4.28) are equal when applied to any
wave function #, which is the same thing as saying that (4.28) is a true
identity. Because the commutator (4.28) vanishes when i ^ j , we can
measure simultaneously the exact values of a given component of r and a
different component of p. When we take the same components of r and p,
the commutator (4.28) is ift, and so the general uncertainty relation (4.27)
then agrees with (4.13): we cannot measure the exact values together.
The theory of measurement in quantum mechanics is a deep, and even
controversial, subject; our discussion has necessarily been brief.
Problems
4.1 Calculate (p) and Ap for the Gaussian wave packet (4.7) at t = 0 by
taking expectation values of the appropriate operators.
4.2 A Gaussian wave packet is given at t = 0 by (4.7). Calculate 3/(z,£)
from (4.10) and (4.4) far enough to verify that
(x) — x0 + pot/m,
(Ax) = \a2 + h2t2/m2a2.
2
4.3 Show that, for any one-dimensional wave function, the values of (p) and
(p2) calculated from the momentum-space probability density |^(p)|2
are equal to the values obtained from the expectation values of the cor-
responding operators using the coordinate-space wave function tp(x).
4.4 Prove Ehrenfest's theorem, equations (4.14), for the one-dimensional
case where the wave function depends on x and t only.
4.5 Show that both (x) and (p) are zero for a one-dimensional harmonic os-
cillator and use the uncertainty principle to deduce that the minimum
energy must be greater than, or equal to, \hw.
4.6 A dart of mass 1 kg is dropped from a height of 1 m, the intention being
to hit a given point on the ground below. The uncertainty principle
imposes a limitation on the accuracy that can be achieved; show that
this is only of the order of a tenth of a nuclear diameter. [Neglect
the uncertainties in position and momentum in the vertical direction;
these produce second-order effects on the accuracy.]
4.7 The wave functions ipi(x) and -02(x) are normalised, and satisfy
(/0ij'02) = c. Find a normalised linear combination of them that is
orthogonal to ipi.
4.8 Show that the momentum operator — i W is Hermitian for a bounded
system.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
Commutators 47
4.11 Show that, for any linear operators A,B and (7,
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
48 4 The superposition principle
(Q + i\R){Q-i\R)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:33 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.006
Cambridge Books Online © Cambridge University Press, 2014
5
The hydrogen atom
[Q,H] = 0. (5.1)
49
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
50 5 The hydrogen atom
where 6 and (j) are polar and azimuthal angles of spherical polar coordi-
nates (figure 5.1). In spherical polar coordinates,!
r.2
V
1 (5.4)
=-
so that the part of V 2 that involves the angles 9 and </> is just —L2/r2.
If we now define the polar axis 9 = 0 of the spherical polar coordinate
system to lie along the ^-direction, it can be shown from the definition
(5.2) of L that its z-component takes the simple form
Lz = —i d/d4>. (5.5)
f For the calculation of V in spherical polar coordinates, see any textbook on vector
calculus, for example H Jeffreys and B S Jeffreys, Methods of Mathematical Physics,
(Cambridge University Press).
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
Orbital angular momentum 51
According to (5.3), L2 does not involve <j> explicitly, only d/d(/). This
means that L2 commutes with Lz, that is, [L2,LZ] = 0, so that the two
operators have simultaneous eigenfunctions. It is usual to denote these by
Y/m(0,</>), where the labels I and m are defined in terms of the eigenvalues
of L2 and Lz as follows:
That is, m is the eigenvalue of Lz but, for a reason that will become ap-
parent shortly (see (5.9) below), the eigenvalue of L2 is written as 1(1 + 1).
The first equation in (5.6), together with (5.5), tells us that Y/m varies
with <j> simply as eim<f>. That is,
1 d
sin0d0
and then one of the two solutions of the differential equation (5.8) is free
from infinities.
Strictly, a given eigenvalue 1(1 + 1) for L2 corresponds to the square of
the angular momentum being equal to U2l(l + 1). However, the usual
terminology is to call the value assigned to / the orbital angular momentum
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
52 5 The hydrogen atom
of the state. The value of m is called the magnetic quantum number, the
reason for this is that effects resulting from the quantisation of the z-
component of the angular momentum are observed when atoms radiate
photons in a magnetic field which points in the ^-direction.
Notice that [L^Lj] ^ 0 for i =fi j . Hence two different components of L
do not have simultaneous eigenfunctions, and we cannot simultaneously
assign definite values to the corresponding observables. However, there
is nothing special about the z-direction. Let La be the component of
L in any other direction. Then it is easy to show that La commutes
with L2. This can be done by direct calculation or by realising that any
three orthogonal directions can be chosen for the coordinate axes, so that
this new direction could equally well have been called the z-direction and
the result follows because we already know that Lz commutes with L2.
Instead of the simultaneous eigenfunctions Yim of Lz and Z 2 , one can
work with the analogous simultaneous eigenfunctions of La and Z 2 , and
the possible eigenvalues of La are m', where m' = —/,—/ + 1 , . . . , / — 1 or
I
This means that it is possible to know simultaneously the exact values of
both the orbital angular momentum I and of any one of its components.
It is not possible, however, to have certain knowledge simultaneously of
more than one component of the angular momentum. (The only exception
is when there is no orbital angular momentum, that is, / = 0. Then all
components of the orbital angular momentum are zero.) Suppose that
a measurement of the square of the orbital angular momentum gives the
value h2l(l + 1), and that a measurement of the ^-component gives the
value mh, with m an integer between —/ and +1. If we now measure some
other component, the value of / will remain unchanged but the system
will have been disturbed in such a way that the information about Lz is
lost. Instead, we will have determined a value m' for this new component,
where m! also is some integer between —/ and +/. If we know the value of
I and the original value of m, we can predict the probability of obtaining
any of the possible values ra', but we shall not describe this calculation
here.
There is in quantum mechanics another type of angular momentum, apart
from orbital angular momentum. This is an intrinsic angular momentum,
not associated with the motion of a particle, and it is given the name
spin. In quantum theory even a point particle may have spin. It turns
out that in the case of spin angular momentum the restriction on the
quantum number / is a little less severe: in addition to possibly being
integral, it may also be half-integral. In particular, for the case of electrons
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
Spherically symmetric potentials 53
and protons the spin angular momentum has / = | ; we say that these
particles have spin | , and this is always true for all electrons and all
protons. The corresponding possible values for the magnetic quantum
number associated with the spin are m = ± | . (Photons, on the other
hand, have spin 1.) In addition to their spin, the particles may carry
orbital angular momentum, as we have already described. We discuss
spin further in chapter 8.
[H, L2] = 0,
[H,Lz] = 0 (5.10)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
54 5 The hydrogen atom
where Y\m is defined by (5.6) and R(r) is some function that remains to
be found. Inserting this structure (5.12) into (5.11), and replacing L2 by
its eigenvalue /(/ + 1), we find that R(r) satisfies the differential equation
where we have taken the proton to be at rest at the origin. We insert this
potential into (5.13) and look for a solution for the electron wave function
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
The hydrogen atom 55
E = -ftV/2Af. (5.15)
We show in appendix A that this has solutions such that R(r) —> 0 at
infinity and the behaviour at r = 0 is suitable if, and only if, Me2/AKSQKTI
is a positive integer n, and / is restricted to integers from 0 to (n — 1).
That is,
Me 4 1
E
& ^ " = 1,2,3,...,
Z = 0,l,2,...,(n-l). (5.17)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
56 5 T/ie hydrogen atom
where
. (5.18)
It happens that ao is the radius of the ground-state orbit of the electron
in Bohr's model; it is known as the Bohr radius. In the proper quantum
theory the expectation value of the radius of the atom in its ground state
i S
o /
Many-electron a t o m s
Consider now an atom of atomic number Z > 1, so that the charge on
the nucleus is Ze and there are Z electrons when the atom is electrically
neutral. Each electron interacts with the nucleus through a potential
—Ze2/A-KE^r so that if we neglect the interactions between the electrons
the possible electron energy levels are obtained from (5.17) by making the
replacement e2 —> Ze2. The neglect of the electron-electron interaction
turns out to be a fair approximation if Z is not too large.
As we have explained, the electron has spin | . A basic principle of quan-
tum mechanics, the Pauli exclusion principle, states that no two identical
particles of spin | may occupy the same quantum-mechanical state. Cor-
responding to each wave function R(r)Yim(6', </>) there are two different
electron states, because the ^-component of the electron spin can be ei-
ther + | or — | . Hence in the ground state of the atom, two electrons
occupy the lowest-energy level (5.17), that is, n = 1, Z = 0. In the next
level, n = 2, there can be two electrons with / = 0, and six with / = 1,
because then m can take any of the three values - 1 , 0 , + 1 . We fill in the
levels in this way, starting from the lowest, until we have allocated all the
electrons, and so obtain the configuration of the ground state of the atom,
(For an atom with very many electrons, the discussion is more compli-
cated, because the energy of the interactions between pairs of electrons
cannot then be neglected.)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
Two-body systems 57
Two-body systems
For the hydrogen atom at rest, we assumed that it is a good approxima-
tion to regard the proton as being at rest. We then wrote the Schrodinger
equation for the wave function ip{r) for the electron. If we do not make
this approximation, which reduces the problem to the single-body case,
we must introduce a wave function ip{ri,r2) that depends on the coordi-
nate ri of the electron and also on the coordinate r 2 of the proton (see
problem 2.5).
Generally, for a system of two interacting particles the potential will be a
function of the separation between the particles. Hence the Hamiltonian
operator takes the form
so that r is the relative separation of the two particles and JR is the position
of their centre of mass. In terms of these variables, (5.19) may be reduced
with a little algebra to the form
where
M = M1+M2, n = M1M2/{M1 + M2). (5.22)
The quantity \x is known as the reduced mass of the system.
Because V does not depend on R, the time-independent Schrodinger equa-
tion
H<ip{rur2) =Et
has solutions of the form
where
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
58 5 The hydrogen atom
The second equation in (5.23) says that the centre of mass R of the system
has a motion equivalent to that of a single particle of mass M moving
freely, in zero potential. If the system is at rest, E1 = 0. The first
equation states that the relative motion of the two particles is the same as
the motion of a single particle of mass \i moving in the potential V. This
result is precisely that obtained in classical mechanics.
Hence in expression (5.17) for the energy levels of the hydrogen atom, the
mass M of the electron should be replaced by the reduced mass \i of the
system. However, because the mass of the electron is much smaller (by a
factor of about 2000) than that of the proton, /z « M.
When the result (5.17), modified in this way, is compared with experi-
ment, there is good agreement. As we explained in chapter 1, the com-
parison is made by examining the emission or absorption spectrum of the
atom. The main discrepancy between theory and experiment may be ac-
counted for by including in the Hamiltonian magnetic interaction terms
in addition to the electrostatic Coulomb potential. The most important
of these terms arises because the electron has both orbital and spin an-
gular momentum. It can be shown from relativistic quantum mechanics
that a charged particle of spin | , such as the electron, necessarily carries
an intrinsic magnetic dipole moment, and the magnitude of this dipole
moment can be calculated with high accuracy. The orbital motion of the
electron generates a magnetic field; in rather crude terms this is because a
charged particle moving in a closed orbit is equivalent to a current loop.
The energy of interaction between this magnetic field and the electron's
dipole moment must be included in the Hamiltonian. As we explain in
chapter 8, it leads to a small correction to the energy levels (5.17). For
a given value of n, the states with different values of / are then not ex-
actly degenerate. There are other effects of a relativistic origin, and when
all corrections have been made the calculations agree with experiment to
better than one part in 106.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
The deuteron 59
The deuteron
The nucleus of the isotope of hydrogen known as deuterium is called the
deuteron. It is a bound state of a neutron and a proton. As their masses
M are almost equal, the reduced mass of the system is /i « | M .
According to (5.23), the relative motion of the proton and the neutron is
described by the Schrodinger equation
Suppose that V depends only on \r\ — r, and treat this as in (5.11), where
now L is associated with the relative orbital angular momentum of the two
particles. Replace L2 by its eigenvalue, as in (5.13). We explained that
we expect to find that the ground state of the system, with the tightest
binding, corresponds to / = 0, so that the repulsive centrifugal potential
is absent.
As a simple model for the potential, take the spherical well
—U for r < a
^0 for r > a.
Then for / = 0
where
E = -h2k2/2^ E + U = h2K2/2fi. (5.27)
\ie:k" f Vr
? >a (5.28)
v
for r < a, '
where A and B are constants. Here we have imposed the constraint that
R(r) —> 0 as r -» oo, so that we indeed have a bound state. We have
also required that R(r) does not diverge at the origin: if it possessed a
singularity of the form r " 1 , it would not, in fact, satisfy the Schrodinger
equation at the point r = 0. At r = a, we want both R(r) and Rf(r) to
be continuous. This gives the equation
k = -K cot Ka (5.29)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
60 5 The hydrogen atom
Problems
5.1 How does the function Y/m (#,</>) change if the coordinate axes are
rotated through an angle a about the z-direction?
5.2 The quantity e2/Airsohc is called the fine-structure constant. Verify
that it is dimensionless; it is approximately equal to 1/137. Calculate
also the Bohr radius, given in (5.18).
5.3 Calculate the energy levels of the hydrogen atom according to Bohr's
theory as follows. Assume that the electron is in circular orbit and cal-
culate its orbital angular momentum TiL in classical mechanics. Then
require L to be an integer. Verify that this leads to the same energy
levels as the correct procedure based on Schrodinger's equation.
Calculate the velocity of the electron in the lowest-energy Bohr orbit
and determine whether the use of non-relativistic kinematics is justi-
fied.
5.4 Find the most probable distance from the nucleus of an electron in the
ground state (5.18) of the hydrogen atom.
5.5 A hydrogen atom has electron wave function
iP{r,O,(f)) = Cre~Kr ^
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
The deuteron 61
where
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:10:44 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.007
Cambridge Books Online © Cambridge University Press, 2014
Revision quiz
This quiz is designed to check that you have followed and understood the
basic features of quantum theory.
1 What are the diameters of an atom and of its nucleus?
2 What do we learn from the photoelectric effect?
3 What is the wavelength associated with an electron, and how do we
know this?
4 What are the operators corresponding to the momentum and energy of
a particle?
5 Write down the time-dependent Schrodinger equation. How and under
what circumstances can one derive the time-independent Schrodinger
equation from it?
6 What is the probability interpretation of the wave function?
7 What is the expression for the particle flux j in terms of the wave
function?
8 What are the continuity conditions that are usually satisfied by the
wave function, and how are they derived?
Why are the conditions different for the problem of a particle in a box
and for a particle moving in a delta-function potential? What are the
conditions in these cases?
9 What is the tunnel effect?
10 How does one use stationary-state solutions to express general solutions
of the time-dependent Schrodinger equation? What mathematical as-
sumption is implied?
11 What is the Heisenberg uncertainty principle?
Under what conditions can one simultaneously know the values of two
observables?
12 What is meant by a Hermitian operator? What properties do its eigen-
values and eigenfunctions have r
13 What are the basic assumptions about observables?
14 What is meant by the expectation value of an observable, and how is
it calculated?
15 Why is the orbital-angular-momentum operator useful in problems
where the potential is spherically symmetric?
16 Describe the energy levels of the hydrogen atom. What are their de-
generacies?
17 What is the Pauli exclusion principle?
18 How does one deal with a two-body system?
62
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:11:11 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.021
Cambridge Books Online © Cambridge University Press, 2014
6
The hydrogen molecule
' (6.1)
63
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:11:40 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.008
Cambridge Books Online © Cambridge University Press, 2014
64 6 The hydrogen molecule
In the other state, the wave function is <f>(r — r 2 ), where r2 is the coordinate
vector of the other proton. Notice that these wave functions are real.
Now reduce R to a value such that the Coulomb field of one proton is
beginning to be felt in the vicinity of the other. In classical physics the
configuration of the system would be little altered, because it would re-
quire energy from outside to tear the electron from one proton and attach
it to the other. But in quantum mechanics the tunnel effect does allow
this transition to occur without any external energy being supplied. This
means that </>(r — r\) and <fi(r — r2) are no longer stationary-state wave
functions. The system eventually settles down into a rather different sta-
tionary state, which we now discuss.
The stationary-state wave functions are the eigenfunctions of the Hamil-
tonian
2
P2 P2 P2
Lv
2m r v
4TT£O \r —
2
h (6 2)
where m is the mass of the electron. Here, the first term is the kinetic-
energy operator for the electron — remember that we are neglecting the
kinetic energies of the protons. We write the stationary-state wave function
for the ground state of the molecule as a superposition of <f>(r — r±) and
(p(r- r 2 ):
(6.3a)
where
fa = 0 ( r - n) 1 = 1,2. (6.36)
Then |ci| 2 is the probability of finding that the electron is in the con-
figuration with wave function </>i, that is, in orbit round the first proton.
Similarly, |c 2 | 2 is the probability offindingthat it is in orbit round the sec-
ond proton. The expansion (6.3) represents an approximation; it should
contain also terms corresponding to configurations in which the electron
is bound to one of the protons with a wave function associated with one of
the excited states of the hydrogen atom. However, these terms are impor-
tant only when R is taken to be so small that the extra Coulomb energy
of repulsion between the protons is readily able to excite the electron into
one of these states.
The wave function ip(r) has to satisfy the time-independent Schrodinger
equation H^ — EI/J. Insert in this equation the expansion (6.3) for %j).
Then pre-multiply the equation by <f>i and integrate over all space, and
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:11:40 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.008
Cambridge Books Online © Cambridge University Press, 2014
The ionised hydrogen molecule 65
n-E) = c2(K12E-H12),
E)
where
Hij = (faHfa) ij = 1 or 2,
/ TIT J?}2 ( J{ fT TT \ 2
n 12
(i) E = E+= , withci=c2,
In both these solutions, |ci| 2 = |C212, that is, there is equal probability of
finding the electron in orbit round either proton. This result shows how
dramatic are the consequences of the tunnel effect; the situation is very
different from that in classical mechanics. Crudely speaking, the electron
continually tunnels back and forth from the potential well surrounding one
proton to that surrounding the other.
We must now attempt a rough calculation of Hu and Hi2. The first two
terms of the Hamiltonian H in (6.2) are nothing but the hydrogen-atom
Hamiltonian of which cf>i is the ground-state eigenfunction, so
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:11:40 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.008
Cambridge Books Online © Cambridge University Press, 2014
66 6 The hydrogen molecule
throughout that part of the integration from which the integral receives
most of its contribution. Hence we may use the expansion
When we insert this expansion into the integral, the first term integrates
to give —e2/AneoR- The second term, being odd in each of the three
coordinates of r', integrates to give zero. Hence for large i?, (6.7) reduces
to
i f n « E o. (6.10a)
This approximation is correct up to and including terms of order ao/i? 2 ,
where the factor CLQ appears here because (pi is appreciable only in the
region where r' « oo- That is, the correction is of order OQ
On the other hand, when R —> 0 the integral (6.8) evidently remains
finite. The last term in the Hamiltonian (6.7) diverges as R -> 0, and so
dominates for small R. Hence for small R we have
(6.106)
H12 = Ai2^o - ~A / d r- r+
4T£ J \ \r-n
r \
(6.12)
This integral does not diverge at r' = 0, because the volume element d 3 r '
contains a factor r'2 when it is written in spherical polar coordinates. Nev-
ertheless, relatively small values of r' are most important in the integral,
and so when R is large
Hence for large R we may neglect the last term in (6.11) compared with
the others, and
(6.13a)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:11:40 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.008
Cambridge Books Online © Cambridge University Press, 2014
Other molecules 67
Finally, from its definition in (6.5), #12 is small for large i?, because when
(pi is large 02 is nearly zero, and vice versa. If we use this, and put (6.10)
and (6.13) together, the energies E± in (6.6) become
£0 =F V(R) R large,
E± (6.14)
R small.
Other molecules
A similar type of covalent binding provides most of the binding in the neu-
tral hydrogen molecule, in most organic molecules, and in some inorganic
molecules. (However, most inorganic molecules are bound mainly by a
different mechanism, known as ionic binding.) When only one electron
is involved, the covalent bond can be strong only between similar atoms:
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:11:40 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.008
Cambridge Books Online © Cambridge University Press, 2014
68 6 The hydrogen molecule
Problems
6.1 As a one-dimensional model for the ionised hydrogen molecule, sup-
pose that the potential near an atom is a delta function, so that for
the molecule
Show that the wave function of the electron has the form
,, v (A{eKX±e~KX) 0<x<R
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:11:40 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.008
Cambridge Books Online © Cambridge University Press, 2014
Other molecules 69
6.4 Three atomic nuclei are fixed at the corners of an equilateral triangle
and an electron is introduced into the system. If this electron were
bound to the atom i in isolation, it would have wave function <^.
Obtain expressions analogous to (6.6) for the energy levels.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:11:40 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.008
Cambridge Books Online © Cambridge University Press, 2014
7
Introduction to perturbation theory
and that they form a complete set, as discussed in chapter 4. This allows
the eigenfunctions ^ of H to be expanded as linear combinations of the
_ y&, (7.3)
3
70
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
Time-independent perturbation theory 71
= 1
' (7-4)
To simplify this equation, use the definition HQ(/)J = Ejfy of the eigen-
functions (f)j. Also, neglect terms that are quadratic in small quantities;
that is, omit those terms where either Hi or AEi is multiplied by cy with
i ^ j , and replace ca by 1. Then
That is, the shift AEi in the level Ei resulting from the addition of the
perturbation Hi to the original Hamiltonian HQ, is just the expectation
value of Hi calculated from the original wave function fa.
Usually this is as far as we need go in the calculation: the level shift
is the quantity of most interest. However, if we wish also to calculate the
perturbed wave function ^ , we instead pre-multiply (7.6) by </>£(r), with
k ^ i, and again integrate over all space. We then find that
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
72 7 Introduction to perturbation theory
Example
If we want to apply perturbation theory to calculate the eigenvalues of a
Hamiltonian H, the first step is to write it in the form (7.1). This division
of H into two parts Ho and Hi may be a purely mathematical procedure;
the only requirements are that we can find the eigenstates of HQ, and that
Hi is small in the sense that we have described. Usually, however, the way
in which we choose to divide H will have an obvious physical significance.
As a simple example, consider again the problem of a particle in a rect-
angular box, as described at the beginning of chapter 3. Suppose that
the particle does not now move freely within the box, but rather that it
carries a charge e and is acted on by a uniform electric field £. Take the
field to be parallel to the rr-axis, so that the potential takes the form
V = -e£x. (7.9)
. / . / 8 \ . qnx . my . snz ,_ .
4>qrs(r) = -T- I s i n - — s i n — - sin , (7.10)
\abcj a b c
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
Time-dependent perturbation theory 73
In this simple problem, the perturbation has turned out to be the same
for every level.
H = H0 + H1(t). (7.13)
where the <j>j(r) are the same eigenfunctions as in (7.3). Because H varies
with t, it does not have stationary-state wave functions; we must solve the
time-dependent Schrodinger equation
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
74 7 Introduction to perturbation theory
When the perturbation Hi (t) begins to switch on, \I>; will start to become
different from $;. We use the expansion
r,t). (7.16)
Notice that the coefficients a^ are functions of t, because the $j are not
stationary-state wave functions for the perturbed system. At t — — oo,
an = 1, and all the other a^ are zero. Inserting this expansion into the
Schrodinger equation (7.15), and using the equation Ho$j(r,t) = ih$j
which defines $ j , we find that
y^i^-. (7.17)
ihdik =
Notice that in (7.17) and (7.18) we have had to be careful about the order
in which we have written the wave functions $ relative to Hi, because H\
may contain the derivative operator d/dr. However, we do not allow Hi
to contain the operator d/dt, so that H does not operate on the coefficient
functions a^(t); these simply multiply H.
So far, we have made no approximations: (7.18) is exactly equivalent
to the time-dependent Schrodinger equation. We now suppose that Hi
is small. This means that we expect the coefficient functions a^-(t) to
depart rather little from their original values, at least if the time t is not
too large. We shall suppose that the value of t is such that au(t) is still
close to 1, and that aij(t) for i ^ j is still small. We then retain in (7.18)
only those terms that are first order in small quantities, which means that
on the right-hand side we keep only the term in the sum for which j — i
and in it we replace an by 1:
ihaik(t) = ($*,#!$;) = (&,tfi&) e ' ^ - W * . (7.19)
Integrating with respect to t, we have
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
Transition probability 75
Transition probability
We have supposed that initially, at t = —00, the Hamiltonian is just i?o,
and that the system is then in the stationary state $i of Ho- Suppose that
after some time T the perturbation Hi(t) vanishes, so that the Hamiltonian
is again Ho. If we now measure the energy of the system, the system will
be thrown into a stationary state of Ho and will remain in this state. The
probability that the measurement results in the value Ek, and so throws
the system into the stationary state <!>£, is, according to the discussion
given in chapter 4 and to (7.16), just |a^| 2 . Provided that this probability
is small, so that our approximations are valid, we may calculate it from
(7.20):
m 2
\«ik(t)\2 = To
(7.21)
This, then, is the probability that our measurement finds that the per-
turbation Hi has induced a transition from the state $; to the state $fc.
Notice that a^ involves the term (fa, Hi fa). Frequently one can show
without detailed calculation that this term vanishes for all but a few final
states fa, regardless of the value of t'. That is, without detailed calcula-
tion one can find those final states fa that are readily accessible from the
initial state fa. Those final states fa for which (7.21) vanishes may still
be possible final states, but with a transition probability that is non-zero
only at second or higher order in the perturbation H\. Such transitions
are known as forbidden transitions; usually (though not always) their
corresponding transition probabilities are very small.
Example
As an example, consider again the problem of a charged particle in a box,
and let the perturbation again be caused by a uniform electric field in the
^-direction, as given in (7.9). Suppose that this perturbation is switched
on at time t = 0, and then switched off again at time T. We assume that
for t < 0 the system is in the stationary state (f)qrs. The probability that a
measurement at some time t > T will find that there has been a transition
to the stationary state 4>QRS is
2
2
(1/h ) I dt'(<f>QRS, -
Jo
(7.22)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
76 7 Introduction to perturbation theory
where
I(UJ) = / dt'eiuJt'
Jo
u = (EQRS-Eqrs)/h. (7.23)
{(/>QRS,x<f>qr8)= / dxdydz(j)QRSx(j)qrs
Jbox
0 Q - q even (Q ^ q)
X | 8Q/[2(Q2 2)2] Q _ q
(7.24)
A plot of this function against UJ is shown in figure 7.1. The height of the
central peak is T 2 , while that of the peaks next to it on either side is some
twenty times less. Hence the transitions that occur most readily are those
to levels for which UJ lies within the central peak. This peak is centred on
UJ — 0 and it has half-width 2TT/T. SO the levels to which transitions are
most likely are those whose energy EQRS differs from the original energy
Eqrs by an amount huj of order h/T.
The shorter the duration T of the perturbation, the greater will be the
spread in energy of this band of preferred final levels. This somewhat
surprising result is peculiar to quantum mechanics, and is a manifestation
of the energy uncertainty principle. We explained in chapter 4 that making
a measurement on a system is liable to induce a perturbation, and, in
fact, this principle also governs the accuracy with which it is possible to
measure the energy of a system: it can be shown that if a limited time
T is available for making the measurement, then the energy cannot be
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
Sudden change in the Hamiltonian 77
-2TT/T 2TT/T
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
78 7 Introduction to perturbation theory
Note that here the prime does not denote differentiation, and the functions
$'• may be completely different from the $j. We do not need to assume
that the difference between HQ and H'o is small. For t < to, the wave
function \I/(r,t) of the system is, by assumption, equal to $*. For t > to
we expand it as a linear combination of the <fr ^:
($i{r,t) t<t0}
* ( r , t ) = < Ebj&jfat) t>t0. (7.27)
I 3
For t > to the <J>^- are stationary-state wave functions of the Hamiltonian,
and therefore the coefficients bj are constant. The wave function (7.27)
will automatically satisfy the Schrodinger equation
HV = ihd$/dt (7.28)
for both t < to and t > t0. At t = t 0? * must be continuous. For if it were
not, d^jdt would have a singularity proportional to S(t — to) at t = to
(see appendix B), and the Schrodinger equation (7.28) would not then be
satisfied at t = t0 since H merely has a finite discontinuity there. The
continuity condition at t — t0 gives
^ ( r , t o ) = *t(r,«o). (7.29)
3
Multiplying this equation by $'fc* (r, t 0 ), integrating over all space and using
the orthonormality conditions in (7.26), we obtain
(7.30)
Hence the probability that a measurement of the energy at some time
t > to yields the value E'k is
M 2 = l(&,*)| 2 . (7.31)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
Example: decay of tritium 79
(The neutrino is a particle that has zero charge and also zero mass, so
that like the photon it moves with velocity c.) Consider this decay in the
frame of reference in which the neutron is at rest. The reason that the
decay is possible is that the rest mass mn of the neutron is greater than the
sum of the rest masses of the proton and the electron, so that the initial-
state energy mnc2 is sufficient both to create the final-state particles and
to give them some kinetic energy. However, when the neutron is bound
in a nucleus, the decay may not be possible. Consider, for example, the
deuteron. If its neutron were to decay, the nucleus would have to break
up, because there is no bound state of two protons. However, this break-
up is not allowed by energy conservation: the binding of the deuteron is
strong enough to ensure that the rest mass of the deuteron is less than the
sum of the rest masses of two protons and an electron.
In the case of tritium, energy conservation does allow one of its neutrons
to decay as in (7.32). This is because there is a bound state of two protons
and a neutron, namely the isotope 3 He of helium. The decay is
Further, the nucleus 3 He is bound rather more tightly than the original
tritium nucleus, so that quite a lot of kinetic energy is available in the
final state. The electron is so energetic that it rapidly escapes, rather
than being trapped in an orbit corresponding to one of the atomic states
of the helium atom.
Suppose now that, before its decay, the tritium nucleus was accompanied
by an orbiting electron in the atomic ground state, so that the tritium
atom was neutral. When the decay (7.33) occurs, the charge on the nu-
cleus suddenly changes from e to 2e, so that the stationary atomic states
correspond to new electron wave functions. Because the 3 He has only one
orbiting electron, it is not neutral but positively ionised. If we measure
the energy of the orbiting electron after the decay of the nucleus, we do not
necessarily find that the orbiting electron is in the new ground state. The
probability that we do find this to be so is calculated from (7.31) by using
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
80 7 Introduction to perturbation theory
the original ground-state wave function for fc and the new ground-state
wave function for (f)'k.
According to (5.18) and the subsequent work in chapter 5,
^ ( ) a o ) - 3 / 2 e - 2 r / a 0 ) (7.34)
Problems
7.1 A hydrogen atom is placed in a weak electric field 5. Show that to
first order in £ the ground-state energy is unchanged.
7.2 Calculate the first-order perturbation to the ground-state energy of a
linear harmonic oscillator due to an additional potential ex2. Verify
that your result agrees with the exact answer to first order in e.
7.3 A quantum system is capable of existing in only two independent
states 0i and </>2, which are eigenstates of the Hamiltonian Ho with
eigenvalues E\ and E2, respectively. The system is modified by the
addition of a term H± to the Hamiltonian. Find expressions for the
new eigenvalues E[ and E'2\ (i) by exact calculation, (ii) in lowest-
order perturbation theory. Verify that the perturbation-theory result
is an approximation to the exact result.
7.4 A particle of mass m and charge e is contained within a cubical box
of side a. Initially the particle is in the stationary state of energy
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
Example: decay of tritium 81
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:03 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.009
Cambridge Books Online © Cambridge University Press, 2014
8
Spin
82
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
Spin \ 83
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
84 8 Spin
Spin \
The special case of spin \ is particularly important, because it applies
to the principal constituents of matter: the electron, proton and neutron.
For spin 5, and only for spin | , we represent the algebra in a very simple
particular concrete form. First, it is useful to write for this case
5 = \a. (8.6)
The operator a is known as the Pauli spin matrix. Since we are describing
the case where S2 has eigenvalue 5(5 + 1), a2 has the single eigenvalue
3. Since we only use the operator a when it is applied to states of spin | ,
whenever we use it the operator 5 2 takes the value 3, and so effectively
a2 = 3. (8.7)
a2z = 1. (8.8a)
There is nothing special about the ^-direction: we may choose the z-axis
to be in any direction without changing the physics. So the square of the
component of a in any direction must be 1; in particular
a\ = 1, a2y = 1. (8.86)
To derive further relations, we now use the algebra satisfied by the com-
ponents of a, which we obtain from (8.2) by replacing L with \a\
Multiply the last of these relations by crx, multiplying each term first on
the left and then on the right:
We use (8.86) to set <J\ — \ and then add these relations, so obtaining
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
The electromagnetic interaction 85
While everything involving spin \ can be calculated by using the six simple
relations (8.8), it is often useful to cast them into a more concrete form.
We can do this by associating a 2 x 2 matrix with each component of a:
/on x
{ 8 U a )
'
This set of matrices is called a "representation" of the components of a.
Although we have written = signs in (8.11a), more properly we should
say "is represented by". Each matrix is a Hermitian matrix, because
the operator it represents is a Hermitian operator. The representation
(8.11a) is not unique: we could have used larger matrices. It is not even
unique if we restrict ourselves to 2 x 2 matrices. It has the feature that
the matrix that represents GZ is chosen to be diagonal. This means that
its eigenstates have a particularly simple representation. We usually label
these eigenstates by t and | . They are represented by
!)(i)-(i)- (S!)(?)~(!
so that u^ is the eigenstate of GZ with eigenvalue +1, and u± is its eigen-
state with eigenvalue —1.
To obtain a general spin-| stationary state, we take a linear combination
of u^ and u^. The coefficients that multiply them will usually be functions
of r. So the state has a representation of the form
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
86 8 Spin
_ dp__8H
(8 5)
dt - dp' d* " dr'
V(r), (8.16)
r = p/M, p = - VV,
A, B = VAA. (8.17)
In order to verify this, one shows that Hamilton's equations (8.15), to-
gether with (8.17), now give
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
The Zeeman effect 87
We do this in appendix D. In (8.19), the last two terms are just the correct
Lorentz force for the action of the electromagnetic field on the charged
electron.
In quantum mechanics, the Hamiltonian (8.18) becomes an operator, be-
cause p must be replaced by the operator p = —i/iV. The vector potential
A is not fully determined from the definitions (8.17), and it is convenient
to impose on it the further gauge condition V • A = 0 (see appendix D).
This has the consequence that, when we multiply out the term (p + eA)2
in (8.16), the order in which we write p and A in their dot product does
not matter. For if \I> is any wave function,
H = HQ + i?i,
Hi = (e/M)p - A-e<l> = {e/M)A • p - ecf). (8.20)
A = \rf\B. (8.21a)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
88 8 Spin
This has to be added to the original energy levels (5.17), which were
independent of m, so placing the atom in a magnetic field removes some
of the degeneracy.
This affects the spectrum of photon frequencies seen when the atom makes
transitions between different levels. For example, consider a transition
from the first excited state of the hydrogen, n = 2, to its ground state,
n = 1. Here n is the principal quantum number: see (5.17). The ground
state is necessarily / = 0 and so m = 0; it is not affected by the external
magnetic field. The n = 2 level can be I = 0, which again is unaffected;
but it can also be / = 1, for which m can take any of the values 1,0 or
— 1. The m = ±1 states have their energies shifted by ±(e/2M)B, so that
the n — 2 level is split into three levels that are close together (because
in practice B is "small" in the sense we have already explained). So we
expect that the n = 2 to n = 1 transition should give three lines in the
spectrum whose frequencies are close to one another.
This is not what is observed experimentally. The discrepancy is explained
by the fact that the electron has spin, and therefore it can have an intrinsic
magnetic dipole moment. A spinless particle cannot have an intrinsic
magnetic dipole moment, because there is no direction intrinsic to the
particle along which it can point. However, for a particle that has spin
vector 5, there can be an intrinsic magnetic dipole moment pointing in
the direction of S. Its magnitude may be found from experiment, but in
the case of the electron it may also be calculated from a relativistic wave
equation for the electron which was derived by Dirac. The electron's
intrinsic magnetic dipole moment vector is found to be — (eh/M)S, so
that in the presence of the external magnetic field B pointing in the z-
direction there is an additional term
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
Spin precession 89
Spin precession
In (8.13) we introduced a two-component spinor u(r) to describe a station-
ary state of a spin-1 particle. It satisfies the time-independent Schrodinger
equation
Hu(r) = Eu{r). (8.23a)
The Hamiltonian H is represented by a 2 x 2 matrix; for example it might
contain a term involving az as in (8.22). If we are interested instead in a
state that is not stationary, we must use the time-dependent Schrodinger
equation and allow the spinor to depend on t as well as r;
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
90 8 Spin
where
u) = 2hvB, (8.256)
with /i the magnetic moment of the proton. The effect of adding a spin-
dependent term to the Hamiltonian is to make the spin state of the proton
change with t. We look for a solution to the time-dependent Schrodinger
equation (8.236) of the form
(8.26)
Substitute this into (8.236) and use (8.246) and the representation (8.11a)
for ax:
0 -±7i(
fa 0 JU2(t)J -""dt\u2{t)
or
= ik
diUl^
ih^-u2(t). (8.276)
dt
If we introduce
= U1(t)+U2(t),
= Ul(t)-u2(t). (8.28a)
Their solution is
We have said that the spin state at t = 0 is such that t*i(0) = 1 and
« 2 (0) - 0, so that Uo = Vo = 1. So, from (8.26), (8.28a) and (8.28c),
«(r,0 = • ( , .
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
Problems 91
What we have found is that, if initially we know that the spin points
in the +z direction, at later times there is non-zero probability that a
measurement will find that it points in the — z-direction. The expectation
values of the components of the spin vector 5 vary as if, in classical terms,
the spin vector steadily rotates in the yz plane with angular frequency u:
this is called spin precession.
Because the magnetic moment of the proton points along its spin vector,
when its spin precesses the direction of its magnetic moment vector also
oscillates. Just as in classical physics, this causes the emission of electro-
magnetic radiation with angular frequency u. Conversely, if electromag-
netic radiation of angular frequency u is shone on the precessing proton,
there is a relatively high probability that it is absorbed. An important
application is to nuclear magnetic resonance imaging, used in medicine
and chemistry. A specimen is subjected to a magnetic field, so that the
spins of the protons within it precess. Electromagnetic radiation is shone
on the specimen and measuring the frequencies that are absorbed by dif-
ferent parts of it gives information about the magnetic moments of the
nuclei concerned and hence about its structure.
Problems
8.1 Derive the commutation relations (8.2).
8.2 What are the eigenvalues of ax? With the representation (8.11a) find
representations for its corresponding eigenstates.
8.3 With the definitions S± = (Sx ± iSy), show that the spinors given in
(8.116) obey S-U^ = u± and S+u± = u^. Evaluate also S-U± and
S
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
92 8 Spin
8.4 The operators S^1) and S^ are the spin vectors of two electrons and
the components of S^ commute with those of S^2\ The total spin
vector is the operator S — S^ + S(2\ What are the eigenvalues of
Szl What does (8.5) give as the eigenvalues of S2 ?
It is clear that the two-particle spinors u^ and u^ are both eigen-
spinors of Sz with eigenvalue 0; however, they are not eigenspinors
of S2. Show that the superpositions x^ = ^(uti ^ ^It) a r e e ig e n "
2
spinors of S , and find the eigenvalues. [Hint: use the results of the
previous question.]
8.5 A particle of spin | and magnetic moment | TIJAG is at rest with its spin
pointing along a magnetic field B. A small additional magnetic field B'
is switched on, such that the angle between the directions of B and B'
is 45°. Use perturbation theory to calculate the new stationary-state
energy correct to order B' and compare with the exact answer.
8.6 An electron is in a state in which the y-component of its spin is known
to be + | . The x-component of its spin is now measured; calculate the
probabilities of obtaining each of the possible results. What can be
said about a further measurent of the y-component of its spin?
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:37 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.010
Cambridge Books Online © Cambridge University Press, 2014
9
Masers and lasers
Radiative transitions
When a system changes its energy as the result of the emission or the
absorption of a photon, it is said to undergo radiative transition. There
are three kinds of radiative transition. In the presence of an electromag-
netic field the system can absorb a photon, so that its energy is raised to
a higher level. If the system is initially in a state other than the ground
state, it can emit a photon and so shed energy. This can happen with-
out the presence of any external electromagnetic field, in which case a
spontaneous emission is said to occur. On the other hand, if the excited
system is placed in an electromagnetic field that varies in time with the
appropriate frequency, the probability that it will emit a photon can be
greatly enhanced. In this case the emission process is known as stimu-
lated emission. Stimulated emission is the basis of maser and laser action,
which we describe in this chapter.
First, we construct a simple model that allows us to investigate how the
internal structure of a quantum system changes as the result of a radia-
tive transition. Although the model is very simple, it will allow us to
understand the main features of the proper, more exact treatment. We
shall deal mainly with absorption and stimulated emission, the two types
of radiative transition where there is an external electromagnetic field.
93
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
94 9 Masers and lasers
electromagnetic field.
We shall take this field to be the plane wave given by
The second of these equations also makes A satisfy the gauge condition
V • A = 0. We shall choose the directions of the coordinate axes such
that the direction of propagation, parallel to the wave vector K, is in the
^-direction, and such that Ao is in the ^-direction. Then, from (8.20)
and (9.1),
We suppose that the field is weak enough to allow us to use the time-
dependent perturbation theory developed in the last chapter. Imagine
that the field is switched on at time t = 0 and is switched off again at
some time T. We suppose that the electron was initially in the eigenstate
<!>i of Ho, and calculate the probability that a measurement, after the field
has been switched off, will find that a transition has occurred to some
other eigenstate $ / of Ho. This transition probability (see (7.20) and
(7.21)) is the squared modulus of
10
}, (9.5)
JO
where
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
Resonant absorption and stimulated emission 95
If Ef > Ei, the transition must have occurred through the absorption of
energy from the electromagnetic field, and the more complete treatment in
which the field is quantised shows that, in fact, the electron has absorbed
a single photon from the field. If Ef < E^ the electron has emitted a
single photon. In either case, we see that the transition probability is
proportional to the square of the intensity of the field. In the case of
absorption, it is easy to understand intuitively that the probability should
increase with the field intensity, but the result is perhaps surprising for
the case of emission: the presence of the field stimulates the electron to
emit a photon.
The integral (9.5) gives
aif = - ^ ^ - . (9.7)
a Ufi — UJ ujfi + UJ
Suppose that the angular frequency u of the external field is chosen such
that hu is close to the difference between the energies E{ and Ef of the
initial and final states. Then the denominator of one of the two terms in
(9.7) is small, and so u) has been tuned to give a resonance in the transition
probability. The term with the small denominator then dominates over the
other term, which may be neglected in comparison with it. In the case
of absorption, where Ef > E^ so that cj/i > 0, the first term becomes
dominant and the transition probability is
K / l
h2 {ufi-u)* • (9 8)
'
In the case of emission, where Ufi < 0, the second term in (9.7) leads to
an analogous expression.
We have already encountered a function very similar to (9.8): see (7.25)
and figure 7.1. Unless u is within a distance of about 2TT/T from a;/*, the
transition probability is very small. In practice, Ufi is often in the optical-
frequency range, so that when T is equal to one second, the transition
occurs readily only if u differs from Ufi by less than one part in 1014.
(However, this estimate ignores the fact that excited levels of atoms have a
finite lifetime and so, because of the energy uncertainty principle which we
described in chapter 7, their energy is not precisely determined. Allowing
for this typically increases the frequency spread to about one part in 105.)
Let us consider what happens when u;, the frequency of the incident
radiation, is equal to CJ/J, the frequency corresponding to the energy
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
96 9 Masers and lasers
difference between the initial and final states. Then (9.8) becomes
\aif\2 = lu^^T2/^2. Recall that \aif\2 is the probability that a transi-
tion has taken place by time T. In macroscopic systems, the transition
rate is, however, observed to be constant, that is \a>if\2 is linear in T.
We can reconcile these facts if we consider not a transition to a single
final state / , but rather to a narrow range of states. The probability P
of a transition from state i into any state whose energy is in the range
{Ef,Ef + AEf) is given by
Here g(Ef) is the density of states at energy Ef, that is the number of
states in the range (Ef,Ef + AEf) is g(Ef)AEf. The integration in (9.9)
is over the energy interval AEf. However, as we have already remarked,
for large T the transition probability (9.8) is extremely small unless Ufi
is almost equal to a;, so that it is an excellent approximation to replace
g(Ef) and | ^ | 2 with their values at E = (Ei + TIUJ) and to extend the
integration to ±oo. Making the change of variable a = |(cj/i — o;)T, (9.9)
can then be rewritten as
(9.10)
(9.12a)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
The ammonia molecule 97
Now, —ex is the rr-component of the electric dipole moment of the atom.
For this reason, the transitions corresponding to the approximation we
have made, that e±lKz « 1, are known as electric dipole transitions. If
we had not replaced the exponential in (9.6) by unity, we should have
obtained additional terms corresponding to electric quadrupole and higher
multipole transitions, and also terms corresponding to magnetic dipole
and multipole moments. The magnetic terms may be understood from
the fact that, crudely speaking, the electron in orbit round the nucleus is
equivalent to an electric current loop, which gives the atom a magnetic
moment that interacts with the electromagnetic field.
The transition probabilities are greatest for the electric dipole transitions.
However, for many pairs of levels E{ and Ej the quantity (</>/,— exfc)
vanishes: compare the result (7.24) for the case of a particle in a box. This
leads to certain selection rules: radiative transitions readily occur only
between certain pairs of atomic levels. It may be shown that spontaneoues
emission is subject to exactly the same selection rules. In consequence,
the corresponding lines in the emission or absorption spectrum of the atom
are much more pronounced than those that are associated with transitions
that can occur only via a magnetic or higher electric multipole mechanism.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
98 9 Masers and lasers
(a) (b)
Figure 9.1. Two configurations of the ammonia molecule (neither is a stationary state).
S is the spin vector and /X is the electric dipole moment.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
The ammonia maser 99
where H^ = (
As in the discussion of the hydrogen molecule in chapter 6, the wave func-
tions (j>i and (fi2 do not describe stationary states. Their significance is
merely that they represent configurations for which (9.15) applies. Be-
cause of the tunnel effect, transitions occur between these states, and as
in chapter 6 the stationary states are
V>± = ( 0 i ± f c W 2 (9.16)
(see (6.3) and (6.6); the l / \ / 2 is just a normalisation factor). The cor-
responding energies E± are to be calculated from a very complicated
four teen-particle Hamiltonian H and, unlike in chapter 6, the wave func-
tions fa and <f>2 are functions of many coordinates. The calculation is not
possible in practice.
Without making a detailed calculation, it may however be shown that the
ground state of the molecule corresponds to the wave function ?/>+. It is
found experimentally that the level separation
E_ - £ + = hu0 (9.17)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
100 9 Masers and lasers
The separation
The separation is achieved by passing the gas through a static electric field
S. This field interacts with the electric dipole moment /JL of the molecule.
This dipole moment exists because the molecule is not spherically sym-
metrical. In fact, the electrons of the molecule tend to be found closer
to the nitrogen nucleus than to the plane containing the hydrogen nuclei,
so that the two states fa and fa have equal and opposite dipole moment,
normal to the plane of the hydrogen nuclei, as shown in figure 9.1. Thus,
if /i 0 is ^ e expectation value of the dipole moment in the state fa,
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
The ammonia maser 101
due to the applied field £ can be neglected. Hence not only the dipole
moment /io can be taken to be unchanged by £, but also H12 and #21,
which, as in (6.13), depend mainly on the overlap if 12 = (</>i, ^2) between
the wave functions <f>i and fa.
In practice, K\2 < 1, so that in the expressions (6.6) for E± we may
neglect Ku compared with 1. According to (9.17), we define the difference
between E+ and E- to be huo, so that (6.6) gives
If we now use the values (9.18) for Hn and #22 in equations (6.4), elim-
inating c\ and C2 from them we find that in the presence of the electric
field £ the energies E± are changed to
(9.20)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
102 9 Masers and lasers
Output
Input
Ammonia
jet Separator
Collimator Resonant
cavity
where
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
Population inversion
the conducting walls of the cavity. When u is tuned in this way, the
emission occurs fairly readily, and the photon that is emitted adds to
the intensity of the electromagnetic field (9.22), so that the probability of
further molecules undergoing stimulated emission is then enhanced. Each
transition further reinforces the field and so increases the probability of
yet more transitions occurring, so that the field builds up rapidly. That
is, the original electromagnetic field (9.22) rapidly becomes amplified.
The standing wave (9.22) corresponds to a collection of photons that
bounce back and forth between the two conducting planes. The momen-
tum vector of each photon is normal to the planes, and when a photon hits
one of the planes its momentum is reversed. What our simple discussion
does not show is that each extra photon that results from the stimulated
emission actually has the same wave vector, and therefore from (2.2) the
same momentum, as the other photons that are already present in the
cavity. It therefore exactly reinforces the existing field. To show this,
one must consider in more detail the equivalence between electromagnetic
fields and systems of photons, which is beyond the scope of this book.
The speed with which the ammonia gas passes through the cavity deter-
mines the time that the molecules spend in the resonant system. This is
an important factor in the design of the maser. The reason for this is
that once a molecule has undergone stimulated emission and so made a
downward transition to the ground state ?/>+, until it has left the cavity it
is now liable to absorb radiation and so make a transition back to the state
V>-. If this is allowed to occur, it decreases the intensity of the field in the
cavity instead of amplifying it. This problem is avoided by ensuring that
the molecules which have dropped into the ground state escape through
the exit hole before an appreciable probability of reabsorption occurs.
Population inversion
Because a molecule in the ground state -0+ is liable to make a transition
that absorbs energy from the field instead of adding to its intensity, the
initial separation of the gas into two components is a vital stage in the
operation of the maser. If an unseparated, equal mixture of the two com-
ponents ?/>+ and ip- enters the cavity, the net effect is zero: there are
as many upward transitions that absorb photons as there are downward
transitions that emit them.
Even a small quantity of gas contains a very large number of molecules
— about 1025 kg" 1 . The molecules continually collide with one another
and so exchange energy. It is completely impracticable to try to follow the
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
104 9 Masers and lasers
effect of the collisions on the state of each individual molecule. Instead, one
resorts to statistical methods, and it turns out that this is quite sufficient
to give an accurate description of the macroscopic properties of the gas.
There is little interest in determining at any instant the state of a given
molecule; what matters is the configuration of the system as a whole. That
is, one needs to know how many molecules at any instant may be expected
to be in a given quantum-mechanical state, rather than which particular
molecules are in that state.
The molecules continually collide with each other and so change their
states. The collisions become more frequent and more violent as the tem-
perature of the gas increases, since by definition the average kinetic energy
of the molecules is proportional to the temperature (measured on the ab-
solute, or Kelvin, scale). A remarkable result of statistical physics is that
it predicts the average distribution of the states of the molecules without
knowing anything in detail about what happens in the collisions. The
demonstration of this may be found in any book on statistical physics;
here we just quote the results.
Consider any system of particles that is in dynamical equilibrium, so
that in particular no heat flows into or out of the system. Suppose that
each particle may be found in any one of a number of different states.
Let these have energies £?», i — 1,2,..., with corresponding degeneracies
gi (so that the level E{ is associated with gi independent states). Then
according to classical statistical mechanics, the average number U{ of par-
ticles expected to be in the ith level at any given instant is given by the
Maxwell-Boltzmann distribution:
m = Cgie~*i/KBl. (9.24)
C =N V>e-*'/*Br . (9.25)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
The laser 105
The laser
The principles underlying the operation of the laser are much the same as
those of the maser, the basic difference being that the photons produced
correspond to the optical range of frequencies. In lasers, and nowadays
also in masers, the material that emits the photons is kept stationary
within the radiation cavity, instead of being squirted through the cavity. If
the amplifying material is solid, the parallel mirrors that form the radiation
cavity may be the end faces of the material itself. A resonant cavity
of this form was used in the first successful laser, whose operation was
announced in 1960. This laser had as its active medium a cylinder of
synthetic ruby, which is an aluminium oxide crystal containing a small
percentage of chromium. The end faces of the ruby cylinder were rendered
accurately parallel to each other and then both were silvered, one fully
and the other partially. Population inversion was achieved by a process
known as optical pumping, which uses a helical flashtube wrapped around
the ruby cylinder. The flashtube emits white light, that is, light with a
wide range of frequencies.
The active ingredient of the ruby is the chromium, whose (simplified)
energy-level diagram is shown in figure 9.3. When the flashtube is fired,
the ruby crystal is flooded with photons, most of which dissipate their en-
ergy as heat. However, some in the blue and green ranges of the spectrum
are absorbed by the chromium and cause transitions from the ground state
to one of the two energy bands shown in the figure. As we have explained,
the relative probabilities of transitions between different levels of a sys-
tem are governed by selection rules. In the case of the energy bands it
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
106 9 Masers and lasers
Energy
Non-radiative
transitions
turns out that the preferred decay is to a metastable state, also indicated
in figure 9.3. (This decay is a non-radiative transition: the energy is not
given up by the emission of a photon, but rather it appears in the form of
phonons, which are quanta of vibration of the other atoms in the crystal
lattice about their mean positions.) So if the flashtube is fired with high
power, a large number of chromium ions are driven into the metastable
state: a large population inversion occurs.
The metastable state decays to the ground state by spontaneous emission
of photons. (Its lifetime is rather long on the atomic scale: a few millisec-
onds, compared with the 10~8 s typical of spontaneous transitions when
dipole transitions are allowed.) Most of the photons emitted by sponta-
neous emission pass out of the crystal and are lost; but just a few move in
the right direction to be reflected back and forth between the two silvered
end faces. In this way, a standing-wave electromagnetic field of the form
(9.22) begins to be set up. This field stimulates the emission of further
photons, so that the intensity of the field (9.22) builds up very rapidly as
the metastable state depopulates. The partially silvered end of the crystal
allows the leakage of an intense pulse of laser light. This is the output of
the pulsed ruby laser.
Laser technology has developed very rapidly since 1960, and there are
now many types of laser, for example the semiconductor lasers described
in chapter 12.
The light emitted by a laser has two important properties. First, it can
be very intense: the power per unit area of cross-section in a laser beam
can be at least 109 times that which may be obtained from a conventional
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
Holography 107
Figure 9.4. Light reaching a point P directly from a coherent source S and reflected
from a point object O.
light source. Secondly, laser light is not only highly monochromatic, but
also highly coherent. In a conventional light source, the atoms decay in
an uncorrelated fashion and so the phases of the electromagnetic wave
packets associated with the different photons are random. In a laser, the
stimulated emission is precisely in phase with the electromagnetic field
that causes it. There are some variations in phase of the field as a whole,
mainly caused by fluctuations in the distance between the mirrors resulting
from thermal and mechanical disturbances, but these are extremely small.
In a laser beam, the phase difference between the light at points along the
beam separated by as much as 100 km is likely to be unchanging with
time.
Holography
The high intensity of laser light has obvious applications, such as to cutting
and welding in engineering or medicine. It also makes possible the study
of fundamental phenomena, such as non-linear optics, arising from the
changes in the properties of materials when they are subjected to very
strong electromagnetic fields. An important application of the high degree
of coherence of laser light is to holography. We digress briefly from our
main subject of quantum physics in order to describe this technique of
three-dimensional photography.
Imagine first a coherent, monochromatic point source of light S at a dis-
tance R from a point object O (figure 9.4). Light that comes from S in the
direction of O will be scattered by O, and hence O behaves effectively as a
second point source. The two sources S and O will be mutually coherent;
that is, the phase difference between the light from them will be constant
in time at any point P in the space surrounding them. Hence, an inter-
ference pattern is established: if the two components of the light reaching
P are exactly out of phase, then there is darkness at P, while if they are
in phase, then P is a point of maximum brightness. The condition that
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
108 9 Masers and lasers
(a)
Photographic
plate
(b)
T
Figure 9.5. (a) Interference fringes corresponding to figure 9.4, with a photographic
plate on which a hologram is recorded. (6) Reconstruction of a virtual image O1 and
a real image O", using a light source Sf.
where
= 0,±l,±2,.
Here K is the wave number of the light (so that its wavelength is 2TT/K),
v\ and r 2 are the distances of P from S and O as in figure 9.4, and a is a
possible phase change induced by the reflection of the light from S on the
object O. Similarly, the condition that there be constructive interference
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
Holography 109
Mirror
Photographic
Laser plate
Mirror
at P is
K(rx - R - r2) + a = 2ra'7r, (9.266)
where
The points P that satisfy either of the two conditions (9.26) lie on a sys-
tem of hyperboloids of revolution having O and S as foci, as is shown
in figure 9.5(a). In this figure we show also the cross-section of a pho-
tographic plate that has been placed so as to intersect the interference
pattern. This plate records a hologram, which is a thin section of the
interference pattern. Development of the plate causes silver grains to be
deposited in the emulsion, with a high density at points P correspond-
ing to constructive interference, and low density in between at points of
destructive interference.
To view the holographic image, one illuminates the hologram with a source
S", having the same position relative to it as the original source S and emit-
ting light of the same colour. Unlike 5, the source S' does not have to be
coherent. The light from S' that falls on the hologram is diffracted by the
light and dark fringes recorded on it. Diffraction occurs in two directions,
one such that the rays emerging on the other side of the hologram appear
to have originated from the point O' corresponding to the position of the
original object O, and the other such that the rays converge on the point
O" that is the reflection of O' in the photographic plate (see figure 9.56).
Thus there is a virtual image at O' and a real image at O". To view the
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
110 9 Masers and lasers
virtual image, one looks at the plate from behind but focuses the eye on O'
because this is the point from which the rays appear to have diverged; this
is why holography produces a three-dimensional effect. The real image is
made apparent by placing a screen behind the plate, passing through the
point O", or by placing the eye beyond the point O" and focussing it on
that point (see figure 9.5(6)).
In practice, because a suitably coherent point source of light S is not, in
fact, available, a laser beam is used to produce the hologram. The beam is
spread out by a lens and split into two parts. One part is reflected directly
onto the photographic plate, and the other is reflected onto the plate by
the object. The set-up needed to produce a hologram of an extended
three-dimensional object is shown in figure 9.6. The interference fringes
now have a much more complicated geometry, but the principles involved
are the same as for a point object.
Problems
9.1 The wave number K corresponds to the radiation associated with
transitions between the two lowest energy levels of the hydrogen atom
(see (5.17)). Verify that the ground-state wave function (5.18) is small
for all values of r that do not satisfy Kr <C 1.
9.2 Assuming that the states fa and fa of the ammonia molecule are eigen-
states of the electric-dipole-moment operator, show that the stationary
states ip± have zero dipole moment.
9.3 For a black body, the power emitted per unit area in the wavelength
interval from A to A + dA is
(2nhc2/\s)[d\/(ehc'Xk*T - 1)].
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
Holography 111
(NA/L3){kBT/2<irm)1/2 e-
m
where the Fermi energy ( is independent of E{. What form does this
distribution take at T = 0, and why?
Notice that when E{ » £ the two distributions are approximately the
same.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:49 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.011
Cambridge Books Online © Cambridge University Press, 2014
10
Band structure of crystals
Electrons in crystals
A crystal consists of a collection of atoms arranged in a regular array,
the spacing between atoms being of the same order of magnitude as the
dimensions of the atoms. Each atom is more or less anchored to one point,
called its site in the lattice, by the electrostatic forces produced by all the
other atoms. We shall not find it necessary here to discuss the details of
how this comes about; nor shall we consider the various patterns in which
the atoms can be arranged. It will be sufficient to remember the essential
feature that the structure of the crystal is periodic in space.
We have seen in chapter 5 that the energy of an electron bound to an atom
is restricted to certain discrete values. Imagine that we can assemble a
crystal of identical atoms whose spacing L can be altered at will. If L is
large enough, the motion of an electron in one of the atoms will be affected
to a negligible extent by the electrons and nuclei of the other atoms. Each
atom then behaves as if it were isolated, with its electrons in discrete
bound states. In figure 10.1 (a) we have drawn a schematic diagram of
the potential V(r) in which an electron moves in this situation. Suppose
that the spacing L is now reduced (figure 10.1(6)). The potential V(r)
in the neighbourhood of a given atom is now affected by the presence
of the nuclei and electrons of the other atoms, particularly those that are
closest. An exact calculation of this situation is a complicated many-body
problem and is not tractable. To a reasonable approximation, however,
the motion of a given individual electron in a crystal can be treated as a
single-particle problem, with the potential determined by the electric field
produced by all the other particles in the crystal.
We shall show in this chapter that, as L is reduced, each original discrete
atomic energy level spreads out into a band of closely spaced levels (fig-
ure 10.2). These bands are separated by energy gaps that are forbidden
to the electrons. The band structure is all-important in determining the
properties of electrons in crystals.
In a crystal, an electron is not bound to any one particular atom, but
rather to the crystal as a whole. If it is not to escape from the crystal,
112
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
Band structure 113
V(r)
(a)
V(r)
(b)
Figure 10.1. Schematic diagram of the potential in a crystal: (a) when the atoms are
widely separated; (b) when they are closer together.
its energy must be negative (where the zero of energy is defined to be the
potential of the free space outside the crystal). If the crystal temperature
is high enough, an electron can gain sufficient energy by collisions with
the vibrating nuclei to escape from the crystal; this is called thermionic
emission. However, we shall confine the discussion to the electrons that
remain bound in the crystal. An electron in one of the low-lying bands,
such that its energy is less than F max (see figure 10.1(6)), passes from
one atom to another by tunnelling through the potential barriers between
the nuclei. We shall find that there are also energy bands between Vmax
and 0; electrons in these bands are readily transmitted over the potential
barriers.
To explore the band structure, we consider mainly a one-dimensional
model. This provides a good guide to electron behaviour in real crys-
tals. Certain aspects, however, require a discussion of electron motion in
more than one dimension; this is necessary, for example, in order to under-
stand why divalent metals such as calcium and magnesium are reasonably
good electrical conductors.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
114 10 Band structure of crystals
E
A
Figure 10.2. The band structure that develops when the atomic separation L is de-
creased.
Band structure
We do not need to know in detail how the potential varies with position
in the crystal. The most important feature is its periodicity. To have
a readily solvable problem, we consider an electron in a one-dimensional
periodic square-wave potential (figure 10.3). This model is sufficiently
realistic for many purposes. For convenience, we define the zero of the
potential to coincide with the top of the wells, rather than with the value
of the potential in free space outside the crystal.
Consider first an electron with energy E such that —U<E<0. The
Schrodinger equation is
in the wells,
between the wells, (10.1)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
Band structure 115
V(x)
-U
or
„ _ r a2i> between the wells, (10.2)
in the wells,
where
a2 = -2mE/hz, (32 = (2m/hz){E + U).
The procedure for solving (10.2) is suggested by a theorem due to Bloch,
which is derived in appendix E. One looks for solutions of the form
(10.3)
where k is real and Uk{x) is periodic, with the same period L as the
period of the crystal lattice. Bloch's theorem states that solutions of the
type (10.3) form a complete set, at least in the limit where the total length
of the crystal is taken to be infinite. Evidently (10.3) represents a wave of
wavelength 2?r/A;, whose amplitude Uk{x) varies across each atomic site,
but is the same from one atom to the next. The electron belongs not to
one atom, but has equal chance of being found in the vicinity of any of
them.
The solution to each of the two differential equations (10.2) consists of
simple exponentials. We write
Aei({3-k)x
a - L < x < 0, (10.5)
•{ Ce(a-ik)x
0 < x < a,
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
116 10 Band structure of crystals
At both edges of each well we require if) and rj)1 to be continuous, or equiv-
alently u and u' to be continuous. The periodicity of Uk{x) guarantees
that if these conditions are satisfied at the edges of just one well, then
they are satisfied for every well. This, of course, is the reason for looking
for a solution of the Bloch type (10.3). Imposing the conditions at x = 0
on the two functions in (10.5) we have
A + B =C + D,
i{p - k)A - i(/3 + k)B = {a - ik)C - (a + ik)D. (10.7)
where
b = L-a. (10.9)
Equations (10.7) and (10.8) are four homogeneous equations in four un-
knowns, A, B, C and D. In order that they have a non-trivial solution the
determinant of their coefficients must vanish. When this 4 x 4 determinant
is expanded it results in the condition
Although we have derived this equation for the case where E lies in the
range -U < E < 0, it is valid also when E > 0; then a 2 , defined in (10.2),
is negative and in (10.10) we write a = \al and so obtain
According to (10.2), when a,L and U are fixed, the quantities a (or a1)
and p are functions of E, and so the left-hand sides of (10.10) and (10.11)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
Number of levels in a band 117
Gap Gap
and in figure 10.4 we plot the function f(E) against E for a typical set of
values of a,L and U. Because k is real, |cosA;I/| < 1. Thus those values
of E for which \f(E)\ > 1 are not accessible; that is, the allowed values of
E fall into bands determined by the condition |/(i£)| < 1, with forbidden
energy gaps between the bands.
Notice that, according to figure 10.4, the bands corresponding to the
lowest-lying levels are narrowest: the levels associated with the electrons
that are most tightly bound to the atoms are affected least by the presence
of the other atoms. Notice also that the band structure persists even for
positive values of E, that is, above the top of the potential wells. The
energy gaps decrease in width with increasing E, but they can still be of
appreciable width even when E is so large that the electron nearly has
enough energy to escape from the crystal into the surrounding free space.
As we shall describe below, when the level structure is calculated in more
than one dimension, an alternative possibility is that neighbouring energy
bands overlap, and this is found to be the case in certain types of crystal.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
118 10 Band structure of crystals
k = 2nn/NL, n = 0, ± 1 , ± 2 , . . . . (10.13)
Furthermore, we find from (10.3), (10.5), (10.6), (10.10) and (10.11) that
both if) and the corresponding energy E are unchanged if k is either in-
creased or decreased by an integral multiple of 2n/L. Thus we may,
without losing any of the states, confine k to any chosen interval of length
2n/L, for example
-n/L <k< n/L. (10.14)
As k takes each of the allowed values (10.13) that fall within this interval,
we pass through every one of the allowed levels in each band. Thus each
band has just N allowed levels, where N is the total number of atoms in
the crystal.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
Band overlap 119
Figure 10.5. One-dimensional model for a crystal having two atoms per unit cell.
This result is not unexpected. When the atoms are very far apart each
band consists of just a single level, the energy level of the solitary atom.
However, the single level corresponds to N different states, because the
electron can be attached to any one of the N different atoms. As the
atoms are brought together, the AT-fold degenerate levels spread out into
bands. This is a generalisation of the two-atom problem that we discussed
in chapter 6.
The property that the number of levels in each band is equal to the number
of atoms in the crystal is special to the simplest crystal structure. In many
cases the crystal lattice consists not of a regularly spaced array of single
atoms, but rather of an array of unit cells that each consist of more than
one atom. The model of figure 10.3 represents a crystal in which the unit
cell consists of a single atom; in figure 10.5 we show a one-dimensional
potential that can be used as a model for a crystal in which the unit cell
consists of two atoms. Generally, the number of levels in each energy band
is equal to the number of unit cells in the crystal.
Band overlap
Equations such as (10.10) and (10.11), which relate k and i£, are called
dispersion relations. We consider now the solution to these equations,
which expresses E as a function of A;. It is not possible to obtain an
explicit expression for this solution; it is necessary to use either graphical
or numerical methods. We discuss the general features of the solution.
A given value of k corresponds to many values of E. The number of
values for which E < 0, corresponding to (10.10), depends on the depth
and width of the atomic potential wells, but, for E > 0, (10.11) always
gives an infinite number of E values for a given k. Because (10.10) and
(10.11) involve k only through cos kL the solutions for E are symmetric
about the origin in k. According to figure 10.4, the edges of the allowed
energy bands occur when cos kL = ± 1 , and if the range of definition of
k is chosen to be (10.14) this corresponds to either k = 0 or k = ±n/L.
Prom these considerations, we deduce that the dispersion curves have the
form drawn in figure 10.6.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
120 10 Band structure of crystals
k
-n/L n/L
In the one-dimensional model, gaps always occur between the bands, re-
gardless of how high up in energy the bands are. However, in the real
world of three dimensions the bands may overlap. In three dimensions the
Bloch wave function (10.3) becomes
tl>(r)=eik'ruk(r), (10.15)
so that A; is a vector. The dispersion curve becomes a three-dimensional
surface in four-dimensional (£7, k) space. The properties of the crystal
may be different in different directions, so that if we draw sections of this
surface corresponding to different directions for k the resulting curves
need not be the same. Each such section will have an appearance similar
to figure 10.6, with gaps between the bands, but the positions of these
gaps and their widths may vary with the direction chosen for k. As this
direction is varied, it may be that the lowest position of the top of a given
gap between two bands is lower in energy than the highest position of the
bottom of that gap (figure 10.7). If this happens, the bands overlap and
there is no gap between them: it is possible to get from one to the other by
continuously varying k. This occurs in divalent metals and semimetals,
and has important consequences.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
Simple consequences of band structure 121
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
122 10 Band structure of crystals
have said that the maximum number of electrons in any one band is 2N.
Hence at the absolute zero of temperature, T = 0, there are three basically
different ways in which the bands can be populated, distinguished by
whether the number Zm of electrons per unit cell is even or odd, and if it
is even by whether or not there is band overlap. We consider these cases
in turn.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
Simple consequences of band structure 123
Conduction
band
Valence
band
Figure 10.8. Excitation of electrons from the top of the valence band of a semiconductor
into the bottom of the conduction band.
Metals
The atoms of the noble metals (copper, silver etc.) and of the alkali
metals (sodium, potassium etc.) have an odd number of electrons and one
atom per unit cell. Hence their valence band is completely filled and their
conduction band half full. The high concentration of conduction-band
electrons ensures that metals are good electrical conductors.
For insulators and semiconductors, the magnitude A of the energy gap be-
tween the valence and conduction bands has an important influence also
on the colour of the pure crystal. Consider, for example, diamond, which
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
124 10 Band structure of crystals
is one of the crystal forms of carbon. This has A = 5.6 eV. A photon
in the visible frequency range has energy in the range 1.7 to 3.5 eV, less
than A, so that an electron in the valence band cannot absorb such a pho-
ton. There are not enough conduction-band electrons to give appreciable
absorption, so all visible light passes through the crystal. Diamond is
therefore colourless. Sulphur, on the other hand, has A = 2.4 eV, so that
its valence-band electrons can absorb photons of frequency u) such that
TIUJ > 2.4 eV and be knocked into the conduction band. Thus the blue
end of the spectrum is absorbed, and the crystal is yellow in appearance.
These remarks apply to pure crystals; as we indicate in chapter 12, the
introduction of impurities can have an important effect.
Problems
10.1 In the free-electron model of a metal the potential is taken to be
constant, independent of position in the crystal. What form do the
Bloch waves take in the one-dimensional version of the model? Sketch
the dispersion curves, analogous to those of figure 10.6. (A more
realistic model is the nearly-free-electron model, which starts from the
free-electron model and then uses perturbation theory to introduce a
small additional potential around each atomic site.)
10.2 A crystal cube has N atoms arrayed in a simple cubic lattice and
its volume V is large. How does (10.14) generalise to this three-
dimensional case? How many allowed fc-vectors are there per unit
volume of fe-space?
If there are Z electrons per atom, what is the largest value of \k\ for
the states occupied at zero temperature?
10.3 (i) As a model for a one-dimensional crystal, take the potential as
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
Simple consequences of band structure 125
and the binding is tight. In the crystal, the atoms are not very close
together. The wave function
clearly satisfies the Bloch condition (E.10) and one can show that it
approximately describes a stationary state. By considering the expec-
tation value of the Hamiltonian if, show that
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:12:58 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.012
Cambridge Books Online © Cambridge University Press, 2014
11
Electron motion in crystals
Electron velocity
In the previous chapter we showed that in a perfectly regular crystal the
stationary-state electron wave functions are the Bloch waves
*k(r,t)=uk(r)Sk-r-E>t'h\ (11.1)
This expression is the same as (10.15), except that we have added the
factor e~lEkt/h so as to write the time-dependent wave function. The
values allowed for the components of the vector k are discrete, but if the
crystal is large enough these discrete values are so close together that for
many purposes k may be treated as a continuous variable. As we know
from the discussion in chapter 4, apart from questions of degeneracy the
wave functions corresponding to different values of k are orthogonal, and
we may choose the normalisation
= 6kk,. (11.2a)
Here, the integration is over the volume of the crystal and dkk> is zero
unless the vectors k and k' are the same, in which case it is equal to 1.
From (11.1) and (11.2a),
= <W- (11.26)
with
126
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
Electron velocity 127
where
The Bloch wave function (11.1) is not an eigenfunction of the electron mo-
mentum operator p = —i/iV. However, we may calculate the expectation
value of the momentum:
We now pre-multiply this equation by u*k and apply the integration / d3r.
The first term on the left-hand side becomes
(uk,h(k)uk>) = Ek(uk,uk>),
where we have again used (11.4). This vanishes because of (11.26). The
first term on the right-hand side of (11.6) similarly gives zero contribution.
The other terms in (11.6) give
where again (11.26) has been used to simplify the right-hand side. Because
the direction of 6k is arbitrary, (11.7) remains true if we cancel the factor
5k on each side. Prom the definition of the operator h(k) in (11.4),
or
v = (l/h)(dEk/dk), (11.96)
where v = (p)/m is the expectation value of the velocity of the electron.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
128 11 Electron motion in crystals
Figure 11.1. Plots of Ek, dEk/dk and d2Ek/dk2 against k for the three lowest bands
of figure 10.6.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
Motion in an external electric field 129
,t), (11.10)
where the function N(k) determines the shape and spread of the wave
packet, and /d 3 fc means integration over each of the components of the
vector k. Because each Bloch wave (11.1) satisfies the time-dependent
Schrodinger equation (11.3), so also will any superposition (11.10). How-
ever, (11.10) is not a stationary state.
We shall suppose that N(k) is sharply peaked at some value of k and
that it falls rapidly to zero outside the immediate vicinity of this value.
This means that each component of the wave packet corresponds to ap-
proximately the same value v = dEk/dk of the expectation value of the
electron velocity. We explained in chapter 4 that a narrow spread in k
corresponds to a wide spread in r. If we had, instead, chosen a wave
packet whose spread in r was less than the dimension of a unit cell of the
crystal, the expectation value of the electron velocity would have varied
as the wave packet moved across each unit cell. This variation may be
traced back to the variation over the unit cell of the factor uk(r) in the
Bloch wave function (11.1). However, for a wave packet that spans many
unit cells, the variation of Uk{r) is averaged out and the time variation of
v is negligible.
That is the situation in the absence of an external electric field. In prob-
lem 4.4 we proved Ehrenfest's theorem (4.14), that a wave packet moves
like a classical particle. Hence in an electric field £ the vector k about
which N(k) is peaked changes with time in such a way that the rate of
change of the electron energy is equal to the rate of work done by the field:
k = -(e/h)£. (11.116)
This result holds provided that the necessary empty electron states exist
so as to allow k to change.
When £ is constant, k increases uniformly with time in the direction
opposite to 5. At whatever point of the dispersion curve k starts, it
moves in this direction until it reaches the limit of the range of definition
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
130 11 Electron motion in crystals
-n/L n/L
Figure 11.2. Motion of the representative point in the presence of an electric field £.
When the point reaches G, it reappears at Gf.
Electric current
In a full band, all the allowed representative points along the correspond-
ing dispersion curve are occupied and the exclusion principle therefore
forbids the distribution of these points to change when a field is applied.
For each electron with velocity v there is another with velocity — v, and
so the total current due to all the electrons in a full band vanishes.
But suppose that the conduction band is half full and is separated from
the band above it by a gap. Let the dispersion curve for the conduction
band have the shape drawn in figure 11.2; if, instead, it is the other way
up like the middle band in figure 11.1, the discussion must be modified
accordingly. The representative points for the electrons, which are initially
in the lower half of the band, will be strung out along the (v, k) curve like
a chain of beads. When there is no electric field the points are distributed
symmetrically about k = 0, as in figure 11.3(a), so that the total current
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
Electric current 131
\v
.r\
I
(a) (b)
i
1 1
r\ 1
1
j 1
k
1 % 1
1 1
Figure 11.3. The velocity of conduction band electrons (a) when there is no electric
field; (6), (c) and (d) with an electric field 8, at successive times.
is zero. When a field is applied towards the left, the representative points
begin to move to the right, as shown in figure 11.3(6). The vector sum
of the velocities of the collection of electrons is now non-zero, and so a
net electric current flows in the direction of the electric field (remember
that the electrons have negative charge). When the electrons reach the
right-most edge of the range of definition of k they reappear at the left,
figure 11.3(c), until in figure 11.3(rf) the sum of the velocities has changed
sign. An electric current nowflowsin the reverse direction. Subsequently,
it changes direction again, and so on.
This rather surprising result, that the direction of the electric current
oscillates back and forth with time, does not apply if the conduction band
overlaps the next higher band, and if all the higher bands similarly overlap
each other. Then the electric current increases steadily, remaining in the
direction of the applied field.
In either case, the steady state described by Ohm's law is not found; the
crystal has no electrical resistance. This is because the discussion has
assumed that the crystal lattice is perfectly regular. In practice this is
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
132 11 Electron motion in crystals
never exactly true. Even at the absolute zero of temperature the nuclei
are not completely anchored at their lattice sites, but oscillate with zero-
point energy (see the discussion of the harmonic oscillator in chapter 3).
In addition, the exact periodicity of the lattice is disturbed, in nearly all
crystals, by the presence of imperfections such as geometrical dislocations
and concentrations of impurity atoms.
Imperfections and impurities apart, the main source of resistance in a
solid is the thermal vibration of the atoms about their lattice sites.
Because the displaced atoms exert forces on one another, their vibrations
are correlated. This correlation allows an analysis in terms of lattice
waves; these travel through the crystal with speeds of the order of the
speed of sound. Just as the energy of an electromagnetic wave is quantised,
so is the energy of a lattice wave; the quantum of energy is called a phonon.
Phonons behave in many respects like particles, as do photons.
As the electrons move through the crystal, they repeatedly collide with
phonons. As the temperature T increases, the collisions become more
frequent and more violent. When an electric field is applied to the crys-
tal, the representative points illustrated in figure 11.3(a) begin to move
according to (11.11), but this motion is opposed by the phonon scatter-
ing. The scattering with phonons tends to prevent the electrons from
increasing their energy and so from rising into the uppermost parts of the
conduction band. A steady state is rapidly reached: the distribution of
representative points becomes static, displaced asymmetrically from its
original position as in figure 11.3(6). A steady electric current then flows.
dt dt [h dk ) ~h Ok* k- (1L12)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
Effective mass and holes 133
with
1 1 d2Ek
(11.14)
m* h2 dk2
The relation (11.13) is just Newton's law for the motion of the electron,
but with an effective mass m*.
We have plotted d2Ek/dk2 for the one-dimensional model in figure 11.1;
the effective mass m* is proportional to (d2Ek/dk2)~1. For each of the
bands, m* is negative near the top of the band. This means that the
acceleration of an electron near the top of a band is opposite to the applied
force. Suppose that a particular band is nearly full, with a few levels near
the top of the band unfilled. This occurs in a semiconductor, where at
zero temperature the valence band is filled and the conduction band is
empty, but at non-zero temperature a few of the electrons from the top
of the valence band are thermally excited into the conduction band. It
is convenient to regard the unfilled states as fictitious particles, called
holes. If the levels were filled with electrons, these would have negative
charge and negative effective mass. The holes correspond to an absence
of these electrons and so they have positive charge and positive mass.
The dynamics of the charge carriers in the nearly filled band may then be
described either in terms of the large number of electrons in the band or
in terms of the small number of holes.
We illustrate this by considering the case in which a single electron is
missing from an otherwise full band. When an electric field is applied,
the total electric current contributed by the band is
u (11.15a)
where the summation extends over all the states i except for the unfilled
state j . However, when the band is full there is no net electric current,
because for each state corresponding to velocity v there is another corre-
sponding to velocity — v. Hence
0= -
where the summation extends over the whole band, and so (11.15a) may
also be written as
J = evj. (11.156)
This is just the current that would be contributed by a solitary positively
charged carrier with velocity Vj equal to that which an electron occupying
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
134 11 Electron motion in crystals
the level j would have had. Similarly, the rate of change of momentum of
all the electrons in the band is
dP x—s *dv{ /i 1 1^? \
(1L16a)
X>'dT
However, when the band is filled, the exclusion principle forbids its overall
quantum state from changing, so that then the rate of change of its total
momentum vanishes:
Thermal excitation
In a metal or semimetal at zero temperature, all states in the conduction
band with energies up to some maximum value (o are filled, and the rest
are empty. At temperatures T > 0 the distribution of occupied states
does not end abruptly at £o- As a result of thermal excitation, electrons
are found with energies greater than (o; the distribution tails off smoothly
beyond energy (Q. In a semiconductor or insulator at T = 0, there are
no electrons in the conduction band and no holes in the valence band.
At T > 0 some electrons are excited into the conduction band and leave
behind holes in the valence band.
To study the effects of temperature it is rarely necessary to consider the
details of the collision processes that cause electron excitation. In a solid
in thermal equilibrium, the electrons are distributed in the available en-
ergy levels in accordance with the predictions of statistical mechanics.
Although the electrons continually change levels as a result of collisions,
statistical mechanics gives their expected average distribution among the
levels independently of the details of the collisions. Because the effects of
the exclusion principle are important, the classical Maxwell-Boltzmann
distribution (9.24) is not applicable; instead, the appropriate distribution
is the Fermi-Dirac distribution. If the level E{ has degeneracy Qi (with
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
Thermal excitation 135
Ui =
(^
Here, the parameter ( is called the Fermi energy or, more correctly, the
chemical potential. It is independent of energy, but it varies with the tem-
perature T. It is determined in terms of the total number N of electrons
by the implicit equation
This usually has the effect that £ varies as log &BT, SO that it varies
rather slowly with T. In practice, the variation between T = 0 and room
temperature is almost negligible.
The Fermi-Dirac distribution (11.17) is derived in any textbook on sta-
tistical mechanics. It has the property that at zero temperature
9i Ei < Co,
where C = e^^kBT. So for energies for above the Fermi energy we again
have the classical Maxwell-Boltzmann distribution (9.24). This is because
the average number n; of electrons in each level is then very small, so that
the exclusion principle has little effect.
In any bound quantum system of finite size the allowed values of Ei are
discrete, but for a large system they are usually so close together that it
is an excellent approximation to treat Ei as a continuous variable. Let
p(E) &E be the number of states in the energy range E to E + dE. The ex-
pected average number of electrons in this range is then, as the continuum
limit of (11.17),
p(E)dE
i. (11.21)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
136 11 Electron motion in crystals
Figure 11.4. The Fermi-Dirac distribution at temperatures T > 0, for the case in
which all the levels have the same degeneracy g.
The average number density n(E) given by (11.21) is the product of p(E)
and a factor that has the shape drawn in figure 11.4 (with g set equal to 1).
In the energy bands, p(E) is a function that depends on the nature of the
crystal, whereas in the gaps p(E) = 0. Unless the temperature T is very
high, the Fermi level £ for a metal or semimetal lies in the conduction
band, while for an insulator or a semiconductor it lies somewhere in
the energy gap between the valence band and the conduction band. We
demonstrate this last result. In an insulator or an intrinsic semiconductor,
there are just enough electrons to fill all the levels up to the top of the
valence band:
N
-L dEp(E),
where Ev is the energy at the top of the valence band. However, from
(11.23)
(11.22) evaluated at T = 0,
/•Co
JV= / dEp(E). (11.24)
J — OO
That is, at T = 0 all the levels up to the Fermi energy ( 0 are filled, and
all those above it are empty. For (11.23) and (11.24) to be consistent,
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
Pair excitation in intrinsic semiconductors 137
p(E) must vanish for values of E between Ev and (o? so that ( 0 must be
in the energy gap. As T is increased ( changes slowly, but remains in the
energy gap until T becomes much greater than room temperature.
On the other hand, the total number of electrons in the valence band and
in lower bands is
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
138 11 Electron motion in crystals
P(E)
Figure 11.5. Model in which the state-density function p(E) has the same shape at the
top of the valence band as at the bottom of the conduction band.
-L dE-
p(E)
(11.27)
p(E)=p(Ev (11.28)
= Ev (11.29)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
Pair excitation in intrinsic semiconductors 139
Inserting (11.29) into (11.25) and evaluating it for the case k&T < A, so
that the exponential in the denominator dominates, we have
-L Ev+A
(11.30)
Since most of the contribution to this integral arises from values of E near
the lower end of the integration, we see that approximately
Problems
11.1 Consider the free-electron model for a crystal (problem 10.1), defining
the constant potential to be zero inside the crystal.
p(E) = {NL/Kh)(m/2E)ll2.
(ii) In the case of a simple square lattice of area A, show that
p(E) = (Am/nh2).
(iii) In the case of a simple cubic lattice of volume V, show that
p(E) =
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
140 11 Electron motion in crystals
travel freely in the film. How does the density of states p(E) vary with
El
11.3 In the free-electron model of a metal, the electrons behave as a set
of independent particles obeying the Fermi-Dirac distribution. An
electron having momentum component normal to the metal surface
greater than a critical value (2mVb)1/2 can escape from the surface;
this is called thermionic emission. In practice Vb ^> ( + k&T. If a
collecting electrode is placed near the metal surface and is maintained
at a sufficiently positive potential relative to the metal, it will collect
all the electrons that are emitted. Show that the current per unit area
emkBT c(C
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:17 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.013
Cambridge Books Online © Cambridge University Press, 2014
12
Transistors
Impurities
We have so far described the properties of intrinsic semiconductors, in
which the crystal lattice is perfectly regular. However, the electrical prop-
erties of semiconductors, unlike metals or semimetals, are drastically af-
fected by the addition of small traces of impurity atoms. Observable
effects occur with impurity concentrations as low as a few parts in 108,
and increasing the impurity concentration to one part in 105 can increase
the conductivity by as much as a factor of 103 at room temperature and
1012 at liquid-helium temperatures. The semiconductor is said to be doped
with impurity atoms.
In order to study the effect of doping, we first consider a crystal consisting
of a periodic array of one type of atom, except that just one of the atoms
has been replaced by an atom of a different type. As a one-dimensional
model of this situation, we take the same infinite chain of square wells as
in chapter 10, but with one of the wells having a different depth, Ui say
(figure 12.1).
We recall that for the perfectly regular crystal, the stationary state solu-
tions are Bloch waves of the form (10.3). These have the property that
when the required continuity conditions on the wave function are satisfied
at the two edges of one of the potential wells, they are automatically sat-
isfied also at the edges of all the other wells. For this reason, we might
guess that in the case of the potential of figure 12.1, the stationary-state
wave function is again composed of Bloch waves in the parts of the crystal
where the lattice is perfectly regular, but that it has a different form in
the neighbourhood of the impurity.
The Bloch wave elkxuk(x), with k real, represents a wave travelling from
left to right. If such a wave is incident on the impurity, part of it is re-
flected, and the remaining part is transmitted across the impurity (com-
pare (3.24) and (3.25)). Hence the total wave function to the left of the
impurity atom is
141
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
142 12 Transistors
*— L
u
1
Figure 12.1. One-dimensional model for a crystal with a single impurity atom,
Teikxuk(x). (12.2)
Here, R and T are constants. Inside the potential well that represents the
impurity atom, the solution to the time-independent Schrodinger equation
is
JLJ e "T u e yiz.o)
where
(12.5)
where the functions vi(x) and V2(x) are both periodic. Because of the
exponential factors, the wave function goes to zero as x -» ±oo, so the
electron is trapped in the neighbourhood of the impurity.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
n- and p-type semiconductors 143
By making vi(x) and v2(x) periodic, we again ensure that when we satisfy
the required continuity conditions at the edges of any single well on either
side of the impurity, they are automatically satisfied at the edges of the
well corresponding to all the host atoms. For a given K, the functions v\
and V2 are determined in a way precisely similar to the determination of
Uk(x) in chapter 10. Each contains a multiplicative constant that remains
to be fixed and that determines its normalisation. Since the complete
potential is symmetric about x = 0, the solutions fall into two classes. One
class has positive parity and the other negative parity, ip(x) = ±if;(—x).
For definiteness, consider the positive-parity solutions (the negative-parity
ones may be discussed similarly). For these v\{x) = +V2(—x), so that the
pair of functions v\ and v2 now contains one multiplicative constant that
remains to be fixed, A say. Inside the impurity well, the positive-parity
solution is
B(ehx + e''nx), (12.6)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
144 12 Transistors
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
Semiconductor junction 145
Lattice site
carries charge
Si (a)
Fifth valence
electron
Conduction band
Impurity
level
(b)
Valence band
Figure 12.2. (o) A silicon crystal doped with the donor impurity phosphorus, with (b)
the impurity level just below the conduction band.
of the impurity levels is such that red light is absorbed. The crystal now
appears blue; it is sapphire.
Semiconductor junction
Consider two pieces of a given semiconductor, one doped with donor atoms
and the other with acceptor atoms. Suppose that each piece has a face
that is accurately plane, and imagine bringing the two pieces into contact
at their plane faces. This forms a pn-junction. (In practice, rather than
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
146 12 Transistors
The diode
We now explain how the pn-junction acts as a diode rectifier: when a bat-
tery is connected such that the p-side is at a higher electrostatic potential
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
The diode 147
Conduction
band
(a)
Valence o o o o
band ° °o°o
P-type n-type
\
\
(b)
\
\
<t>0
(c)
p-type n-type
Figure 12.3. The pn-junction: (a) the band structure at the instant the two pieces are
brought together and (b) in dynamical equilibrium. The difference between (a) and
(6) is caused by the electrostatic potential <j> drawn in (c).
than the n-side, current flows readily, but only a limited current flows
when the polarity of the battery is reversed.
Suppose first that no external electrostatic potential difference is applied.
In chapter 11 we explained that in an intrinsic semiconductor the Fermi
energy ( lies somewhere in the energy gap. Its position is shifted when
the crystal is doped, and the Fermi energies of each of the two pieces, p
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
148 12 Transistors
and n, are further shifted when the pieces are pressed together to form the
junction. However, statistical mechanics tells us that any system of elec-
trons in dynamical equilibrium has a unique Fermi energy. So dynamical
equilibrium for the pn-j unction is reached when the Fermi energies ( of
the p-part and n-part coincide. We do not need to know the precise final
value of the common Fermi energy, but in practice it lies somewhere in
the overlap between the energy gaps on the two sides of the junction, as
drawn in figure 12.3(6).
According to the Fermi-Dirac distribution (11.21), the number of electrons
on the n-side that are above the bottom E% of the conduction band on the
p-side is
where pn(E) is the density of states on the n-side. Some of these electrons
will be moving in the right direction to cross over to the p-side, and some
of these will actually reach the p-side conduction band before collisions
reduce their energy to below E%. Hence there is a diffusion current of
electrons from the n-side to the p-side (that is, an electric current from
the p-side to the n-side):
/•OO
Jo oc / dEfr(E)e-(E-Mk»T. (12.8)
JEI
Similarly, some of the electrons in the conduction band on the p-side dif-
fuse into the n-side. In order that there be dynamical equilibrium, the
corresponding current must be equal and opposite to Jo- Exactly similar
considerations apply to the hole carriers.
Suppose now that an external electrostatic potential difference V is ap-
plied across the junction, by connecting the two sides of the crystal to a
battery. We assume that all this potential difference appears across the
junction region; if this is not a good approximation, the Ohmic potential
drop IR in the rest of the crystal may be allowed for once the current I
is known. With the extra driving potential V, the crystal is no longer in
dynamical equilibrium; there is no longer a single Fermi energy through-
out the crystal. However, on each side far from the junction there is a
situation close to equilibrium, not much affected by the additional current
flow caused by V. This is provided that V is not too large, so that the
drift velocity superimposed on the electron motion, corresponding to the
additional current, is small. Then outside the junction region each piece
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
The diode 149
EP+eV-
EP+eV
n (a)
I I
(b)
p n 1-
V
Figure 12.4. (a) The band structure of figure 12.3(6) when there is an additional
potential difference V whose sign is indicated in (b).
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
150 12 Transistors
-*» V
Figure 12.5. Plot of current / against potential difference V for a semiconductor diode.
The signs of V and / are defined in figure 12.4(6).
have sufficient energy to pass over to the p-type conduction band, and so
they carry a diffusion current proportional to
dEpn (12.9)
Assuming that the density of states pn(E) varies only slowly with E, this
is approximately
Hence the total electric current that passes through the crystal from the
n-side to the p-side is
(12.10)
This is the contribution from electron carriers; the hole carriers contribute
similarly.
/ is plotted against V in figure 12.5. When V is positive, / < /Q, but for
negative V, the current increases exponentially with the magnitude of V
(until the approximations made in the derivation are no longer valid).
The semiconductor diode can be made to operate as a laser (see chap-
ter 9). The first semiconductor laser was made in 1962, using stimulated
emissions associated with transitions from the conduction to the valence
band as the source of laser radiation. The pn-junction is forward-biased
(the p-type side is at a higher electrostatic potential than the n-type side),
providing the necessary pumping, and the end-faces of the diode, perpen-
dicular to the plane of the junction, are polished to give high reflectivity.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
The diode 151
Conduction
band
(a)
Valence
band
n-type p-type n-type
R
(b)
4-'
Emitter Base Collector
Conduction
band
(c)
Valence
band
Figure 12.6. The npn-junction transistor (a) with no external potential, (c) with
external potentials applied as in (6).
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
152 12 Transistors
Semiconductor lasers can be made very small — the size of a pin head
— and have wide application, for example in compact disc players and in
fibre-optic communications.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
The junction transistor 153
Gate—^
Insulator
— n
c ' ^ Drain
Source
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
154 12 Transistors
Gate /Insulator
Conduction
band
Figure 12.8. Lowering of conduction and valence bands near the surface of the p-type
material when a positive potential is applied to the gate.
ure 12.8), and the Fermi level £ appears in the conduction band. According
to the Fermi-Dirac distribution, there is then a substantial population of
electrons in the conduction band in this region, even though in the bulk of
the p-type material holes are the majority carriers. This inversion layer
at the surface allows electrons to flow from the source across to the drain.
The larger the gate voltage, the greater the electron concentration in the
inversion layer and the greater the current.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
Two simple circuits 155
Diode
t -M-
o
t
Input Output
o
L
Earth (a)
Diode
T -H- T t
Input Output
J T
Earth (b)
Hence, the input is the sum of two high-frequency components, and so the
capacitor would short circuit the whole signal if the rectifier were not there.
The signal passed by the rectifier is F(t) = cosptcosut6{cosptcosut),
where 9{x) = 1 for x > 0 and 0(x) = 0 for x < 0. If this function
is represented as a Fourier integral, it is found to have a low-frequency
component corresponding to the audio signal. It is not easy to calculate
the Fourier transform, but a function that has much the same shape as
F(t) is G(t) = cos2 \pt cos2 \ut. Like F(t), G(t) is non-negative, and
both F(t) and G(t) oscillate in a similar fashion. However,
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
156 12 Transistors
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
Two simple circuits 157
Problems
12.1 In a one-dimensional crystal an impurity atom replaces the atom at
the origin. Investigate the energy-level structure, taking as a model
for the potential
12.2 The tunnel diode is a junction between p- and n-type materials which
are both very heavily doped, so that the impurity levels are spread out
into bands. In the n-type material the Fermi level lies in the conduction
band, and in the p-type material it lies in the valence band. The
transition region at the junction is made so narrow that tunnelling
readily takes place across the interface.
Sketch the energy bands near the interface (a) when no external po-
tential is applied, (6) when the n-type material is connected to a small
negative potential V relative to the n-type, and (c) when it is connected
to a fairly large negative potential.
Sketch the graph of the current I through the junction against the
potential V. (The correct curve has a region of negative 'resistance'
dl/dV.)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:29 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.014
Cambridge Books Online © Cambridge University Press, 2014
Appendix A
Power-series solutions
We are interested in solutions having the property I/J —>• 0 as x -> dboo, so
we begin by examining the differential equation (A.I) at very large values
of x. For any given eigenvalue £?, when x is sufficiently large E is much
less than ^muj2x2 and so the term — Ety in the equation is then much less
important than the other two terms. This suggests that it may be useful
to write
tl>(x)=e~mu'x2'2hH(x), (A.2)
because the exponential factor has the property that for large x
The exponential e+mujx l2h has the same property, but if we used this
instead, H(x) would have to obey very much more stringent conditions
at x — ±oo in order to satisfy the required boundary conditions for -0-
Substitute (A.2) into (A.I):
Y,asXs. (A.4)
s=0
158
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:39 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.015
Cambridge Books Online © Cambridge University Press, 2014
The hydrogen atom 159
We substitute this into the left-hand side of the differential equation (A.3)
and equate to zero the coefficient of xs:
2m \ ( ) /
[ \
For any choice of ao and ai, this relation determines as for all s > 1. There
are two parameters ao and a\ that can be chosen arbitrarily, because the
differential equation (A.3) is second-order.
When s is large, (A.5) gives
2mcj
(A.6)
(22 fi S
This means that, for any fixed value of #, the ratio of successive terms in
the series (A.4) tends to zero as s -> 00, and therefore the series converges.
Consider now the series expansion of the function AemuJX /h xB, where A
and B are constants. Its coefficients also have the property (A.6) for large
s. If we make the plausible assumption, which can be justified, that the
form of H(x) at large x is controlled by the high powers in the series, we
conclude that for large x
Inserting this behaviour into expression (A.2) for -0, we see that, whatever
may be the values of A and J5, we do not have the desired behaviour for
ip at infinity.
The only way to avoid this is to arrange that the series expansion of H(x)
terminates at some value of s, so that H(x) is a polynomial. From (A.5),
we see that this will happen if and only if:
either a\ = 0, so that all odd powers of x are absent, and for some even
value 5 = 2n, E— \TIUJ = 2nTujo, so that the series terminates at 5 = 2n,
or ao = 0, so that all even powers of x are absent, and for some odd
value s = 2n + 1, E - ±hu = (2n + l)hw, so that the series terminates
at s = 2n + 1.
Together, these two possibilities correspond to the results quoted in (3.16)
and (3.17).
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:39 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.015
Cambridge Books Online © Cambridge University Press, 2014
160 Power-series solutions
5=0
and substitute this expansion into the left-hand side of the differential
equation (A.9). The lowest resulting power of r that we find is ra~2, and
on equating the coefficient of this to zero we find
ao[a(a-l)~ 1(1 + 1)] = 0. (A.ll)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:39 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.015
Cambridge Books Online © Cambridge University Press, 2014
The hydrogen atom
For any choice of ao, this determines all the other coefficients as. When s
is large,
as+i/as ~ 2K/S. (A.13)
f(r)~Ae2KrrB,
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:39 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.015
Cambridge Books Online © Cambridge University Press, 2014
Appendix B
The delta function and Fourier transforms
a:) = l. (B.2)
When plotted against #, as in figure B.I, the function A£(x) has a peak
at the origin. This peak has height 1/ey/n and width of order e (exactly
how we define the width does not matter), so that if e is allowed to become
very small the peak becomes very tall and very narrow. Outside the peak,
the function becomes extremely small. This means that if we define an
integral
poo
Af] = / dxA£(x)f(x), (B.3)
J -c
then for a wide class of functions f(x) the value of the integral when e
is very small depends almost entirely on the values that f(x) takes very
162
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
The delta function 163
dxS(x)f(x)=f(0) (B.6)
for a suitable class of 'test functions' f(x). The test functions are ordi-
nary functions. Different types of generalised function or distribution are
defined by using different classes of test function, for example tempered
distributions are defined using functions f(x) that, together with all their
derivatives, are bounded by some fixed power of x at x = ±00.
By making a simple change of variable in (B.6), one obtains
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
164 The delta function and Fourier transforms
dxS(x) — 1. (B.8)
J —<
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
The delta function 165
1—
Because of the explicit definition (B.I) of Ae(x), the first term on the
right-hand side vanishes unless f(x) explodes violently at infinity. So, by
letting e —> 0, we arrive at the definition of 8'{x)\
pCO pOO
We have used (B.6) for the case where f(x) is replaced by f'(x). Higher
derivatives of S(x) may be defined similarly. Because the rth derivative
of A£(x) changes sign r times very near the origin when e is very small,
successive derivatives of 5(x) are more and more singular at the origin.
Consider now the indefinite integral of Ae(x),
This is plotted against x in figure B.3. As e -> 0, the step in this function
becomes progressively steeper, until finally the function changes abruptly
from 0 to 1 at the origin. In fact
dy6(y)=6(x), (B.12)
where
6(x) =
0 x<0.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
166 The delta function and Fourier transforms
and so, applying the usual rule for differentiation of a product of functions,
In the last term, we obtain initially [g(x) — f(x)]5(x — a), but we may
replace x by a inside the square bracket because the delta function vanishes
unless x — a.
Fourier transforms
In our discussion of the delta function, we used a particular function
Ae(ai), defined by the Gaussian function (B.I), which has a peak that
becomes progressively taller and narrower as e —> 0. There is nothing
special about this particular choice of A£(x); almost any function having
this property will be just as good, provided it is chosen so as to satisfy
the integral condition (B.2) for each value of e. Another example is the
function
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
Fourier transforms 167
f(x)
/ d k e ikx+ek
+
J-
1
IX + £ IX — 6
= 2TTA£(X). (B.17)
(B.19)
The integral here does not converge in the usual sense. It is understood
that, as in (B.18), a damping factor e~£^ is to be included in the inte-
grand so as to make the integral converge, and after the integral has been
calculated e is allowed to tend to zero.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
168 The delta function and Fourier transforms
[ (B.20)
V27T J-
for some suitable function <j>(y). If (j)(y) -> 0 sufficiently rapidly as y ->
±00, this integral converges in the usual sense. Otherwise it is necessary
to include in the integrand a damping factor e~£^ and let e —> 0 after the
integral has been calculated. If <j>(y) is bounded by some power of y as
y -> ±00, this procedure will give a convergent result. We now calculate
1
4
Again, a damping factor e~e\h\ may be needed in order to make this
integral converge. We insert the representation (B.20) for $(k) and obtain
dxeikx(f)(x)
(B.22)
lkx
(f)(x) = -4= I dke-' $(k).
This is the Fourier inversion theorem: if one of the relations holds, so does
the other. We say that the functions $ and <j> are Fourier transforms of
each other.
It may be that both the integrals in (B.22) converge without needing any
additional damping factors. This is the case of classical Fourier trans-
forms. Most of the textbooks consider only classical Fourier transforms;
in this case it may be shown that both $(k) and (p(x) are ordinary func-
tions. The possibility of including damping factors allows the very pow-
erful extension of Fourier transform theory to include the possibility of $
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
Fourier transforms
1 f°
\/2W-oo V2^
^2^ J_
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
170 The delta function and Fourier transforms
dye-y la = a v ^
(B.23c)
where 6(x) is the step function defined in (B.12). For, putting in the
necessary damping factor,
1 f°° ., i / 1 \
lim dfce
. -/*= / ^ = TT7Z ' ( B - 24 )
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
Fourier transforms 171
/ \
/ \
/ \
/ \
I /
\ ^Pole /
\ /
\
ie tells us that it approaches the real axis, along which lies the contour of
integration, from below rather than from above. See figure B.5. Suppose
now that x < 0. Then the exponential factor in the integral (B.24) vanishes
as k —> oo in the upper half of the complex k plane sufficiently rapidly
that the integral taken over an infinite semicircle in the upper half-plane
would be zero. Hence we may add this contour to the original integration
contour along the real axis, without changing the value of the integral.
The integration contour is now closed and, because the pole is in the
lower half-plane, the integral vanishes by Cauchy's theorem. When x > 0,
we instead add on to the original integration an integration over an infinite
semicircle in the lower half-plane. The integral is then evaluated by taking
the residue at the pole, which now lies within the contour.
(v) Let </>i(x) and fafe) be the Fourier transforms of and
respectively. Then
4= f dk> (B.23e)
J_
dk2 5(k - h - k2)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
172 The delta function and Fourier transforms
/ dx<f)*(x)<t>{x)
J — oo
/•o
= /
J—
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:13:52 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.016
Cambridge Books Online © Cambridge University Press, 2014
Appendix C
Orbital-angular-momentum operators
y = psin</>,
z = pcot9, (C.3)
where p = r sin#.
To calculate Lz in spherical polar coordinates, we need
^ + +
dx dx dp dx 89 dx dcf)
and the similar expression for d/dy. Here the partial derivative 8/dp
is calculated keeping 9 and (j) constant, 8/86 is calculated with p and <£
constant, and d/dcj) is calculated with p and # constant. On the other
hand, each partial derivative with respect to x is calculated with y and z
constant. Prom (C.3),
dx = cos (f)dp — ps'mcf) d(/>,
dy = sin <f> dp + p cos <f> d</>,
dz = cot 9 dp - p cosec2(9 d0. (C.5)
Solving these three simultaneous equations for dp, d<j> and d0 we have
dp = cos (/> dx + sin (/> dy,
pdcj) — — sin 0 dx + cos 0 dy,
pd9 = sin 0 cos 0(cos 0 dx + sin 0 dy) — sin2 0 dz. (C.6)
173
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:04 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.017
Cambridge Books Online © Cambridge University Press, 2014
174 Orbital-angular-momentum operators
We insert these in (C.4), and calculate d/dy similarly. Using these expres-
sions, with x and y from (C.3), we calculate Lz from (C.2). The result is
given in (5.5). The other two components of L may be calculated in the
same way, so that altogether
Lz = -id/d<f>. (C.8)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:04 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.017
Cambridge Books Online © Cambridge University Press, 2014
Appendix D
Electrodynamics
The electric field £ and the magnetic field B obey four equations known
as Maxwell's equations. Maxwell wrote down these equations before either
special relativity theory or quantum mechanics were known, but neither
of these two developments led to any need to change the equations. They
are
V • £ = 0,
V • B = 0,
V A£ = -B,
c2 = 1/eoMo. (D-2)
A general theorem of vector calculus allows one to deduce from the second
of Maxwell's equation that, at least locally, B may be expressed in terms
of a vector potential A:
B = VAA. (D.3)
By calculating V • B , it is straightforward to verify that the second equa-
tion of (D.I) is indeed satisfied if B is expressed in this way, but the
converse result is also valid.
If we insert (D.3) into the third of Maxwell's equations, we find that
V A (£ + A) = 0. (D.4)
(D.5)
175
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:13 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.018
Cambridge Books Online © Cambridge University Press, 2014
176 Electrodynamics
then the fields B and £, given by (D.3) and (D.5), remain unaltered. The
fields B and S are physically measurable quantities, but the functions A
and (j) are not directly measurable. Hence the transformation (D.6), which
is called a gauge transformation, can have no physical consequences. This
freedom allows one to impose on (p and A certain constraints known
as gauge conditions, and one example of a possible gauge condition is
V • A = 0, which is the condition introduced in chapter 8.
In (8.18) we wrote the Hamiltonian for an electron of charge — e in inter-
action with an electromagnetic field:
Here, the operator d/dr is usually written as V, and d/dp denotes the
corresponding differentiation with respect to p. The differentiation d/dr
is carried out with p kept constant, and r is kept constant for d/dp. If
we apply Hamilton's equations to the Hamiltonian (D.7), we obtain
r = (p + eA)/m,
d e
at VV m {[(p + eA) • V)A + (p + eA) A (V A A)} + eV</>.
(D.9)
To write the second of these equations, we have used the vector identity
Vu2 = 2(u • V)u + 2u A (V A u).
The first equation in (D.9) gives
(d/dt)p = mr-e(d/dt)A
= mr -eA- e(r • V)A. (D.10)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:13 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.018
Cambridge Books Online © Cambridge University Press, 2014
Electrodynamics 177
Here, we have used the fact that A generally varies with both t and r,
so that its total derivative dA/dt is not simply equal to dA/dt = A,
but rather dA/dt — A + (r • V)A. We reduce the right-hand side of the
second equation of (D.9) using the first equation of (D.9) and also (D.3)
and (D.5), and so obtain the equation of motion
mr = - W - eS - er A B. (D.ll)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:13 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.018
Cambridge Books Online © Cambridge University Press, 2014
Appendix E
Bloch waves
We define the translation operator D such that for any function f(x)
where we have used (E.I) and (E.2). Hence, for any function f(x),
[D,H] = 0. (E.4)
). (E.6)
178
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:32 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.019
Cambridge Books Online © Cambridge University Press, 2014
Block waves 179
The properties (E.I) and (E.6) are valid for all pairs of points (x,x + L)
that lie within the crystal. Suppose that the crystal extends from x = —oo
to x = +00, so that (E.6) implies that
/•OO 1 /*O
Together, (E.7) and (E.8) give |c| 2 = 1, so that c = eia with a real.
Without loss of generality, we may set a = kL:
c = eifcL, (E.9)
where k is real. According to (E.6) and (E.7), each time that we displace
the coordinate x through distance L the wave function is multiplied by
elkL. So for any positive or negative integer m
xl>(x)=eikxuk(x), (E.ll)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:32 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.019
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems
Chapter 1
1.1 5 x 1029 photons per second.
1.2 Work in the frame in which the electron is initially at rest, so that its
energy is me2. Let the momentum of the photon be p; according to
(1.4), its energy then is cp. So if the photon is absorbed, the electron
afterwards has momentum p and energy E = me2 + cp. This means
that E2 > m2c4 + c2p2, in violation of (1.36).
1.3 The wavelength of visible light is of the order of 500 nm, so for a mass
of 1 kg the answer is about 3 x 1020 years, and for an electron about
10"3 s.
1.4 Suppose that after the scattering the photon has momentum p'. Then
according to (1.4) its energy has changed from cp to cp1. Applying
(1.36) to the electron after the scattering gives
Squaring both sides, using p • p' = pp' cos 0, and dividing through by
2mc3ppf/h gives the answer.
1.5 The calculation is similar to problem 1.4.
Chapter 2
2.1 According to (2.76), adding a constant VQ to V(r) is equivalent to
instead subtracting the same constant from E. Hence, according to
(2.9), the stationary-state solutions become multiplied by eiVot/h. This
leaves unchanged the probability density |\I/|2 for all r and all t.
2.2 The allowed values of \mv2 are given by (2.136) so the allowed values
of v are nhn/ma. We need the difference between the values of v for
n = 2 and n = 1. For the electron, this is about 3 x 106 m s" 1 , and for
the tennis ball it is about 10" 33 m s" 1 . In the case of the tennis ball,
the answer is negligibly small: the velocity is effectively a continuous
variable and quantum mechanics need not be used to describe the
motion.
180
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems i
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
182 Hints for the problems
2rrii
and E = Ex + E2.
Chapter 3
3.1 According to (3.11), within the well i/> oscillates sinusoidally, while
outside the well it goes exponentially to zero. At the two edges of the
well, both ij) and dip/dx are continuous. From (3.10), the lowest levels
correspond to the smallest allowed values for ft and so, from (3.116),
their wave functions have the smallest number of oscillations.
If the values of U and b are such that in a given state the value of a
is very small, the exponential fall-off of the wave function beyond the
edges of the well is very slow, and then there is a good chance that the
particle will be found outside the well.
3.4 Prom (3.17), the wave functions are Hermite polynomials Hn(x) multi-
plied by an exponential factor. Prom appendix A, HQ(X) is a constant
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems 183
to show that in case (i) \T\2 « : \A\2, and in case (ii) \T\2 « \A\2.
Notice that (3.27) is valid also when U is negative, as in (i); then n is
imaginary.
3.6 For x 7^ 0, the Schrodinger equation is i/>" = a2tp, where
E = -h2a2/2m. We choose the solutions Ae~ax in x > 0 and Beax
in # < 0, so that ip —> 0 as x —> ±oo as required for a bound state.
This requirement also picks out the case E < 0, so that a is real,
rather than E > 0. At x = 0, I/J must be continuous, so that A — B.
The other condition at x — 0 is (3.35), with A replaced by —/i. This
gives a = m/i/h2, and so E = —m/j,2/2h2. When [/ —>> oo, (3.10) gives
/? -> oo such that [/ - h2f32/2m. Hence if C/6 = /i, \fib - mn/h2f3.
Thus in (3.12a) the right-hand side becomes mfi/h2, but in (3.126) it
diverges and gives no finite value for a.
3.7 The first step is achieved simply by replacing x by —x throughout
the Schrodinger equation (always valid for any equation!) and using
V(x) = V(—x). Because each value of E is supposed to correspond to
only one independent wave function, it must be that ^{—x) = Cfi/j(x))
for some constant C. Replacing x by —x in this equation, we see that
ip(x) = Cip(-x). Combining these two equations gives C2 = 1, so
C = ±l.
3.8 The mass of the a-particle is about 4 times that of the proton, that is
M « 3750 MeV/c 2 , and e2/he0 = 4TTC/137. SO (3.456) gives for nuclei
of the same radius r oc E~ 1 / 2 exp(4.0Z J E~ 1 / 2 ), where the a-particle
E is in MeV. This gives r\j/rFm w 5 x 1013 and hence T\J W 3 x 109
years. The experimental value is 4.5 x 109 years, so the simple model
gives a surprisingly accurate result.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
184 Hints for the problems
Chapter 4
4.1 /»OO
= / dx \&*(—ihfy1).
J—oo
J — oo
f
Integrating the first term by parts twice, and using the fact that for
a wave packet the wave function ^ vanishes at x = ±00, gives the
result. The second part is done similarly.
4.5 (x) and (p) are zero because of the symmetry about the origin.
(Ap)2/2m+±muj2(Ax)2.
So, from the uncertainty relation (4.13),
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems 185
ft2 f 3 h2 f
= -— d rV-(rV$) + - - /
2m J 2m J
The divergence theorem transforms the first integral to an integral
over a surface that encloses the system so that \I> vanishes over
this surface; this integral therefore vanishes. The second integral is
(h2/2m) J d 3 r | V#| 2 , which is positive. It follows that in each bound
state
fa = (fa -
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
186 Hints for the problems
The first of these has eigenvalue +1, so the system is left in this state
after the first measurement. At a later time t the wave function is
lElt/h
e~lElt/h
= (</>!e~' elE2t/h)/V2.
I/J2 2 e-'lE2t/h
++ I/J
Cos[(E1-E2)t/h}}.
Chapter 5
5.1 Under the rotation, 0-^6 and <f> -» <j> — a. So, according to (5.7), YJm
is multiplied by e~ imQ .
5.2 e2/47re0r2 is a force, so dim (e2/47re0) = ML 3 T~ 2 . The momentum
operator is (-iftV), so dim (h) = ML 2 T" 1 .
a0 = (47rsohc/e2)(h/mc) « 5 x 10~ n m.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems 187
5.3 The force towards the centre of the orbit is mv2/a = e2/4neoa2. So
TiL = mva = (mae2/A-KSQ)1!2, = rih say. The corresponding energy
is
\mv2 - e2/4ne0a = -e2/8n£0a = me4/32TT2
as in (5.17). When n = 1, the velocity is e2 /A-KSQU = c/137, so that
v <ti c and non-relativistic kinematics are justified.
5.4 When the wave function is spherically symmetric, the probability
that the distance is between r and r + dr is P(r) dr, with P(r) =
Anr2 \ip(r)\2. The most probable distance is where P(r) is largest,
which is at r = 2CLQ.
5.5 Applying L2 in (5.3) to ip multiplies it by 2, so that according to (5.6)
1 = 1. Also (5.6) shows that m = - 1 . (5.13) and (5.14) yield
\Mr 2M Aneor-EJ
«r r>a,
Kr
-eKr) r<a.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
188 Hints for the problems
m dr2
and ip2(R) = 0 (see (5.23)). However, the first of these equations
is obtained from equation, (3.15), for the linear harmonic oscillator
by making the changes m —> \m and u —> y/2cu. So the energy
eigenvalues are E = (n + \)THJO\/2.
5.9 Equations (5.27) and (5.29) give
1
cot Ka = — — = —
K
Chapter 6
6.1 For x 7^ ±i?, the Schrodinger equation is if)" = K2ip, where
E — — TI2K2/2m. The potential is symmetric about x = 0 and so
(see problem 3.6) the wave functions have definite parity: either
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems
Chapter 7
7.1 Take the z-axis in the direction of £. In (7.1), take Ho as the Hamil-
tonian before the field is introduced and Hi = e£z. With the ground-
state wave function (6.1), the energy shift (7.7) is
AE=(na30)-1 f
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
190 Hints for the problems
- Ei)fa
To this equation, apply in turn the operations Jd3r<f>l and /d 3 r(/> 2 ,
using the orthonormality of fa and fa (see (4.18)). Eliminating c\ and
c2 then gives
but not to either of the other two states of energy 3h27r2/ma2. The
transition probability is calculated from (7.22), with
t'-t'2/T2
-/-,
(see (B.236)).
7.5 The normalised wave functions are ipo(x) = (a/7r)1/4e~2aa;2 and
•0i (a;) = (27r) 1 /2( a / 7r )3/4 ;re -iax 2 ) w h e r e a _ muJ/fl T h e required
probability is
1°
J —c
f
2 2 2
= (2a2v2\2/h2) dtte -a v t +iut
J —a
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems
The integral is
(see (B.236)).
7.6 The probability is given by (7.31), with fa = (2/a)1/2 sm{nx/a) and
0'fc = (l/a) 1 / 2 sin(7rx/2a). The integration is from 0 to a, and the
answer is 32/9TT2.
Chapter 8
8.1
[Lx,Ly] = -^[VPz ~ ZPy, ZPx ~ Wz]
a
Z]px+X[z,pz]py}
8.2 Solve
-o C
C
1 0) \u2
giving U2 = cui and u\ = CU2- So the eigenvalues of ax are c = ±1, the
same as those of a z . The corresponding eigenstates are represented
by the normalised spinors
0
.3 With 5 = | a , the matrix representations (8.11a) give
s+ s
-{o oj' --{i o
and when these matrices multiply the spinors given in (8.11b) the
results given for S-u^ and S+u± follow. Also S-u^ = S+u^- = 0.
.4 The possible results of measuring Sz and Sz are t t , J 4 ? t i a n d
It, for which Sz has eigenvalues +1,-1,0 and 0 respectively. If the
eigenvalues of S2 are s(s + 1), then according to (8.5) the allowed
values of s axe | | — | | and ( | + | ) .
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
192 Hints for the problems
= 0, etc
••(!
HUB'
l
2 2 2
-\\& \B + B'\ = -±nhB ((1 + B'/BV2) + (B'I By/2) ) ,
where the factor l/\/2 normalises the spinor to unity. The possible
results of the measurement are the eigenvalues of |cr x , namely ± | .
The corresponding spinors u± obey ayu± =± u± and are
The two probabilities are calculated from (4.22) and are each equal
to | . A further measurement of the y-component of the spin will sim-
ilarly give one of the two results ± | , each with probability | .
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems 193
(Of course, usually the probabilities will not be equal; we have ob-
tained this result because we have chosen directions that are orthogo-
nal to each other.)
Chapter 9
9.1 Equation (5.17) gives Tiu = 3me4/327r2h3SQ. For a photon, K = u/c.
So, with (5.18), Ka0 = 3e2/8ne0hc = (3/2)(l/137). Hence if it is not
the case that Kr <^ 1, then r/ao » 1 and the exponential in the wave
function (5.18) is very small.
9.2 The stationary states ip± are not eigenstates of the electric-dipole-
moment operator /i, but we may calculate the expectation of JJL in
these states. Because the eigenstates </>i and </>2 correspond to different
eigenvalues ±/i 0 , they are orthogonal. Hence, using (9.16),
nqrs = Ce-E«rs/kBT,
(The fact that the states in this problem are degenerate makes no
difference here; the degeneracy enters only if we calculate the number
of particles in a level rather than in a state, as in (9.24).) The average
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
194 Hints for the problems
fraction of particles that are in the state is nqrs/N, which is also the
probability that a given particle is in the state.
The velocity of the particle is v = p/m = hk/m = (hir/mL)(q,r,s).
As L -> oo, the allowed values of EQRS become arbitrarily close, and
r r i 2 i"1
= N UmL/h7r)3(27rkBT/m)3/2j .
In this limit, the number of allowed sets of values of (g, r, s) correspond-
ing to the volume element d 3 i; of velocity space ~ {mL/fnr)3d3v. If
d3v is infinitesimal, so that all the corresponding states have essen-
tially the same energy, the average number of particles in these states
is C(mL/h7r)3d3ve~^mv2/kBT. Integrate over vy and vz to obtain the
number having vx in the desired interval.
Notice that the final answer is in fact valid for a large box of any
shape.
9.6 Take the x-direction normal to the membrane, and use the result of the
previous problem. A particle whose rr-component of velocity is u will
reach the wall containing the membrane during the time interval At if
it is within a distance uAt of the wall, and a fraction A/L2 of such
particles will hit the membrane region of the wall. Hence the number
of particles in time At that pass through the membrane is
N f du(m/2nkBT)^2e'mu^2kBT(uAt/L)(A/L2).
JuO
Chapter 10
10.1 The Bloch waves here have Uk(x) in (10.3) equal to a multiple of
e2lrnx/L, where r is any integer, and V + E = h2(k + 2rn/L)2/2m.
The dispersion curves are shown in the figure below. In this model,
there are no gaps between the bands, but these appear when the
model is modified so as to become the nearly-free-electron model.
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems 195
r=-3
r=-2
10.3 (i) For 0 < x < L, the potential is zero and so I/J and Uk have the
forms shown in the second equations of (10.4) and (10.5), respectively.
For — L < x < 0, Uk is obtained from the second equation of (10.5)
by replacing x by (x + L). Impose the conditions that, at x = 0, if) is
continuous and disc xp' = —Uip (0) - see (3.35). This gives
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
196 Hints for the problems
(ii) In 0 < x < L, take rp and uk as before, and in L < x < 2L take
similar forms but with different coefficients. Impose the conditions
on ip and disc rp' at x = L. Because V(x) now has period 2L, so
does uk(x). Use this to obtain uk(x) in -L < x < 0 from uk(x) in
L < x < 2L, and impose the conditions on ip and disc ipf at x = 0.
Eliminate the four coefficients from the four conditions.
10.4
where
and where
n)L
f dx 0*(:r - nL)0(a; - ml),
m,n
n)L
fdx<l>*(x-nL)H<l>(x-mL).
The first term in (?/>, -H"^) contributes £Jo to £", and the second con-
tributes
[dx(/)*(x-nL)V0(x-rL)<f)(x-mL)
Because the binding is tight and the atoms are not very close together,
the main contribution to the denominator comes from those terms
having m — n, so that the denominator is approximately equal to TV,
the number of atoms. Likewise, the main contribution to the numerator
comes from those terms having r = n = m ± 1. There are N pairs of
such terms and together they give the desired result.
10.5 In the case of periodic boundary conditions, the allowed values of
k are given by (10.13), and are used to form Bloch waves ipk(x) =
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems 197
Chapter 11
11.1 (i) Prom (10.13), there is one allowed wave vector in the interval
2ir/NL of fc-space. Taking into account the two spin states, there are
therefore NL/n states per unit interval of &-space. In the free-electron
model, E = h2k2/2m and dE/dk = h2k/m, and so
p(E) = (NL/7r)/{h2k/m) = (NL/nh){m/2E)1/2.
(ii) This time there is one allowed wave vector in the area 4TT2 /A of
fc-space, so that there are Aj2ix2 states per unit area. For |fc| between
k and (k + dk) there are therefore (yl/27r2)27r/jdfc states having all the
same energy. So p(E) = (Ak/n)/(h2k/m).
(iii) There are V/4TT3 states per unit volume. For |fc| between k and
(k+dk) there are (V/A7r3)A7tk2dk states of the same energy. So p(E) =
(Vk2/7r2)/(h2k/m).
11.2
p(E)/(Am/nJi2)
J I I II
E/(nznz/2ma2)
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
198 Hints for the problems
11.3 From part (iii) of problem 11.1, for unit volume of the metal
p(E) = (2m3E/n4h6)1/2. The Fermi-Dirac distribution (11.21) gives
the number of electrons n(E) in the unit volume per unit E. Since
E = p2/2m, the number per unit p is obtained by multiplying this by
dE/dp = p/m and is
11.4 Near the valence band, p(E) = NS(E + A). The number of electrons
in the valence band is obtained by inserting this into the Fermi-Dirac
distribution (11.21) and integrating around E = — A:
" --N-U.
e-(A+C)/kBT
AVEdE e-(E-O/kBT
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Hints for the problems 199
Chapter 12
12.1 The function Uk{x) determined in problem 10.3 may be used to con-
struct the solution given in (12.1) and (12.2). The constants R and T
are determined by making I/J continuous at x = 0, and disc if)' = —U21P
there. The same level structure obtains as in problem 10.3.
Consider now the trapped-electron levels. In 0 < x < L, where
the potential vanishes, we have if)(x) — Ceax + De~ax, where E =
—h2a2/2m. Writing rj> as in (12.4), and making Vi(x) periodic, we find
that in L < x < 2L we have i/>{x) = Cea^x~L^KX a L KX
+De- ^~ ^ .
Imposing the conditions at x — L, that ip is continuous and disc ipf =
—C/iV>, gives two equations. With the even-parity solution, one of the
two corresponding conditions is satisfied automatically. The other,
together with the two previous equations, gives on eliminating C, D
and ekL
(C/1C/2 + \Ul ~ 2a 2 ) t a n h a L = 2aUx.
There is no negative-parity solution.
12.2
Conduction
band
Valence
band
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
200 Hints for the problems
Normal
diode *V
Tunnel
diode
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:46 GMT 2014.
http://dx.doi.org/10.1017/CBO9781139171243.020
Cambridge Books Online © Cambridge University Press, 2014
Index
201
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:56 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
202 Index
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:56 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
Index 203
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:56 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014
204 Index
Downloaded from Cambridge Books Online by IP 150.244.221.208 on Thu Feb 13 17:14:56 GMT 2014.
http://ebooks.cambridge.org/ebook.jsf?bid=CBO9781139171243
Cambridge Books Online © Cambridge University Press, 2014