Martin - Nuclear and Particle Physics - An Introduction
Martin - Nuclear and Particle Physics - An Introduction
B. R. Martin
Department of Physics and Astronomy, University College London
Copyright # 2006 John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester,
West Sussex PO19 8SQ, England
All Rights Reserved. No part of this publication may be reproduced, stored in a retrieval system or
transmitted in any form or by any means, electronic, mechanical, photocopying, recording, scanning or
otherwise, except under the terms of the Copyright, Designs and Patents Act 1988 or under the terms of a
licence issued by the Copyright Licensing Agency Ltd, 90 Tottenham Court Road, London W1T 4LP, UK,
without the permission in writing of the Publisher. Requests to the Publisher should be addressed to the
Permissions Department, John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex
PO19 8SQ, England, or emailed to [email protected], or faxed to (+44) 1243 770620.
Designations used by companies to distinguish their products are often claimed as trademarks. All brand
names and product names used in this book are trade names, service marks, trademarks or registered
trademarks of their respective owners. The Publisher is not associated with any product or vendor
mentioned in this book.
This publication is designed to provide accurate and authoritative information in regard to the subject
matter covered. It is sold on the understanding that the Publisher is not engaged in rendering
professional services. If professional advice or other expert assistance is required, the services of a
competent professional should be sought.
John Wiley & Sons Inc., 111 River Street, Hoboken, NJ 07030, USA
Jossey-Bass, 989 Market Street, San Francisco, CA 94103-1741, USA
Wiley-VCH Verlag GmbH, Boschstr. 12, D-69469 Weinheim, Germany
John Wiley & Sons Australia Ltd, 42 McDougall Street, Milton, Queensland 4064, Australia
John Wiley & Sons (Asia) Pte Ltd, 2 Clementi Loop # 02-01, Jin Xing Distripark, Singapore 129809
John Wiley & Sons Canada Ltd, 22 Worcester Road, Etobicoke, Ontario, Canada M9W 1L1
Wiley also publishes its books in a variety of electronic formats. Some content that appears in print may not
be available in electronic books.
Library of Congress Cataloging-in-Publication Data
Martin, B. R. (Brian Robert)
Nuclear and particle physics/B. R. Martin.
p. cm.
ISBN-13: 978-0-470-01999-3 (HB)
ISBN-10: 0-470-01999-9 (HB)
ISBN-13: 978-0-470-02532-1 (pbk.)
ISBN-10: 0-470-02532-8 (pbk.)
1. Nuclear physics–Textbooks. 2. Particle physics–Textbooks. I. Title.
QC776.M34 2006
539.70 2–dc22
2005036437
British Library Cataloguing in Publication Data
A catalogue record for this book is available from the British Library
ISBN-13 978-0 470 01999 9 (HB) ISBN-10 0 470 01999 9 (HB)
978-0 470 02532 8 (PB) 0 470 02532 8 (PB)
Typeset in 10.5/12.5pt Times by Thomson Press (India) Limited, New Delhi
Printed and bound in Great Britain by Antony Rowe Ltd., Chippenham, Wiltshire
This book is printed on acid-free paper responsibly manufactured from sustainable forestry
in which at least two trees are planted for each one used for paper production.
To Claire
Contents
Preface xi
Notes xiii
1 Basic Concepts 1
1.1 History 1
1.1.1 The origins of nuclear physics 1
1.1.2 The emergence of particle physics: the standard model and hadrons 4
1.2 Relativity and antiparticles 7
1.3 Symmetries and conservation laws 9
1.3.1 Parity 10
1.3.2 Charge conjugation 12
1.4 Interactions and Feynman diagrams 13
1.4.1 Interactions 13
1.4.2 Feynman diagrams 15
1.5 Particle exchange: forces and potentials 17
1.5.1 Range of forces 17
1.5.2 The Yukawa potential 19
1.6 Observable quantities: cross sections and decay rates 20
1.6.1 Amplitudes 21
1.6.2 Cross-sections 23
1.6.3 Unstable states 27
1.7 Units: length, mass and energy 29
Problems 30
2 Nuclear Phenomenology 33
2.1 Mass spectroscopy and binding energies 33
2.2 Nuclear shapes and sizes 37
2.2.1 Charge distribution 37
2.2.2 Matter distribution 42
2.3 Nuclear instability 45
2.4 Radioactive decay 47
2.5 Semi-empirical mass formula: the liquid drop model 50
2.6 -decay phenomenology 55
2.6.1 Odd-mass nuclei 55
2.6.2 Even-mass nuclei 58
viii CONTENTS
2.7 Fission 59
2.8
-decays 62
2.9 Nuclear reactions 62
Problems 67
3 Particle Phenomenology 71
3.1 Leptons 71
3.1.1 Lepton multiplets and lepton numbers 71
3.1.2 Neutrinos 74
3.1.3 Neutrino mixing and oscillations 76
3.1.4 Neutrino masses 79
3.1.5 Universal lepton interactions – the number of neutrinos 84
3.2 Quarks 86
3.2.1 Evidence for quarks 86
3.2.2 Quark generations and quark numbers 89
3.3 Hadrons 92
3.3.1 Flavour independence and charge multiplets 92
3.3.2 Quark model spectroscopy 96
3.3.3 Hadron masses and magnetic moments 102
Problems 108
References 393
Bibliography 397
Index 401
Preface
Brian R. Martin
January 2006
Notes
References
References are referred to in the text in the form Ab95, where Ab is the start of the
first author’s surname and 1995 is the year of publication. A list of references with
full publication details is given at the end of the book.
Data
Data for particle physics may be obtained from the biannual publications of the
Particle Data Group (PDG) and the 2004 edition of the PDG definitive Review
of Particle Properties is given in Ei04. The PDG Review is also available at
http://pdg.lbl.gov and this site contains links to other sites where compilations of
particle data may be found. Nuclear physics data are available from a number of
sources. Examples are: the combined Isotopes Project of the Lawrence Berkeley
Laboratory, USA, and the Lund University Nuclear Data WWW Service, Sweden
(http://ie.lbl.gov/toi.html), the National Nuclear Data Center (NNDC) based
at Brookhaven National Laboratory, USA (http://www.nndc.bnl.gov), and
the Nuclear Data Centre of the Japan Atomic Energy Research Institute
(http://www.nndc.tokai.jaeri.go.jp). All three sites have extensive links to other
data compilations. It is important that students have some familiarity with these
data compilations.
Problems
Problems are provided for all chapters and appendices except Chapter 9 and
Appendices A and D. They are an integral part of the text. The problems are
mainly numerical and require values of physical constants that are given in a table
following these notes. A few also require input data that may be found in the
references given above. Solutions to all the problems are given in Appendix D.
xiv NOTES
Illustrations
Some illustrations in the text have been adapted from, or are loosely based on,
diagrams that have been published elsewhere. In a few cases they have been
reproduced exactly as previously published. In all cases this is stated in the
captions. I acknowledge, with thanks, permission to use such illustrations from the
relevant copyright holders.
Physical Constants
and Conversion Factors
1.1 History
Although this book will not follow a strictly historical development, to ‘set the
scene’ this first chapter will start with a brief review of the most important
discoveries that led to the separation of nuclear physics from atomic physics as a
subject in its own right and later work that in its turn led to the emergence of
particle physics from nuclear physics.1
Nuclear physics as a subject distinct from atomic physics could be said to date
from 1896, the year that Henri Becquerel observed that photographic plates were
being fogged by an unknown radiation emanating from uranium ores. He had
accidentally discovered radioactivity: the fact that some nuclei are unstable and
spontaneously decay. In the years that followed, the phenomenon was extensively
investigated, notably by the husband and wife team of Pierre and Marie Curie and
by Ernest Rutherford and his collaborators,2 and it was established that there were
three distinct types of radiation involved: these were named (by Rutherford) -, -
and
-rays. We know now that -rays are bound states of two protons and two
neutrons (we will see later that they are the nuclei of helium atoms), -rays are
electrons and
-rays are photons, the quanta of electromagnetic radiation, but the
historical names are still commonly used.
1
An interesting account of the early period, with descriptions of the personalities involved, is given in Se80.
An overview of the later period is given in Chapter 1 of Gr87.
2
The 1903 Nobel Prize in Physics was awarded jointly to Becquerel for his discovery and to Pierre and Marie
Curie for their subsequent research into radioactivity. Rutherford had to wait until 1908, when he was
awarded the Nobel Prize in Chemistry for his ‘investigations into the disintegration of the elements and the
chemistry of radioactive substances’.
3
J. J. Thomson received the 1906 Nobel Prize in Physics for his discovery. A year earlier, Philipp von Lenard
had received the Physics Prize for his work on cathode rays.
4
Why the charge on the proton should have exactly the same magnitude as that on the electron is a very long-
standing puzzle, the solution to which is suggested by some as yet unproven, but widely believed, theories of
particle physics that will be discussed briefly in Chapter 9.
HISTORY 3
units. At about the same time, the concept of isotopism (a name coined by Soddy)
was conceived. Isotopes are atoms whose nuclei have different masses, but the same
charge. Naturally occurring elements were postulated to consist of a mixture of
different isotopes, giving rise to the observed masses.5
The explanation of isotopes had to wait 20 years until a classic discovery by
Chadwick in 1932. His work followed earlier experiments by Irène Curie (the
daughter of Pierre and Marie Curie) and her husband Frédéric Joliot.6 They had
observed that neutral radiation was emitted when -particles bombarded beryllium
and later work had studied the energy of protons emitted when paraffin was
exposed to this neutral radiation. Chadwick refined and extended these experi-
ments and demonstrated that they implied the existence of an electrically neutral
particle of approximately the same mass as the proton. He had discovered the
neutron (n) and in so doing had produced almost the final ingredient for under-
standing nuclei.7
There remained the problem of reconciling the planetary model with the
observation of stable atoms. In classical physics, the electrons in the planetary
model would be constantly accelerating and would therefore lose energy by
radiation, leading to the collapse of the atom. This problem was solved by Bohr
in 1913. He applied the newly emerging quantum theory and the result was the
now well-known Bohr model of the atom. Refined modern versions of this model,
including relativistic effects described by the Dirac equation (the relativistic
analogue of the Schrödinger equation that applies to electrons), are capable of
explaining the phenomena of atomic physics. Later workers, including Heisenberg,
another of the founders of quantum theory,8 applied quantum mechanics to the
nucleus, now viewed as a collection of neutrons and protons, collectively called
nucleons. In this case, however, the force binding the nucleus is not the
electromagnetic force that holds electrons in their orbits, but is a short-range9
force whose magnitude is independent of the type of nucleon, proton or neutron
(i.e. charge-independent). This binding interaction is called the strong nuclear
force.
These ideas still form the essential framework of our understanding of the
nucleus today, where nuclei are bound states of nucleons held together by a strong
charge-independent short-range force. Nevertheless, there is still no single theory
that is capable of explaining all the data of nuclear physics and we shall see that
different models are used to interpret different classes of phenomena.
5
Frederick Soddy was awarded the 1921 Nobel Prize in Chemistry for his work on isotopes.
6
Irène Curie and Frédéric Joliot received the 1935 Nobel Prize in Chemistry for ‘synthesizing new
radioactive elements’.
7
James Chadwick received the 1935 Nobel Prize in Physics for his discovery of the neutron.
8
Werner Heisenberg received the 1932 Nobel Prize in Physics for his contributions to the creation of
quantum mechanics and the idea of isospin symmetry, which we will discuss in Chapter 3.
9
The concept of range will be discussed in more detail in Section 1.5.1, but for the present it may be taken as
the effective distance beyond which the force is insignificant.
4 CH1 BASIC CONCEPTS
By the early 1930s, the 19th century view of atoms as indivisible elementary
particles had been replaced and a larger group of physically smaller entities now
enjoyed this status: electrons, protons and neutrons. To these we must add two
electrically neutral particles: the photon () and the neutrino (). The photon was
postulated by Planck in 1900 to explain black-body radiation, where the classical
description of electromagnetic radiation led to results incompatible with experi-
ments.10 The neutrino was postulated by Fermi in 1930 to explain the apparent
non-conservation of energy observed in the decay products of some unstable nuclei
where -rays are emitted, the so-called -decays. Prior to Fermi’s suggestion,
-decay had been viewed as a parent nucleus decaying to a daughter nucleus and
an electron. As this would be a two-body decay, it would imply that the electron
would have a unique momentum, whereas experiments showed that the electron
actually had a momentum spectrum. Fermi’s hypothesis of a third particle (the
neutrino) in the final state solved this problem, as well as a problem with angular
momentum conservation, which was apparently also violated if the decay was two-
body. The -decay data implied that the neutrino mass was very small and was
compatible with the neutrino being massless.11 It took more than 25 years before
Fermi’s hypothesis was confirmed by Reines and Cowan in a classic experiment in
1956 that detected free neutrinos from -decay.12
The 1950s also saw technological developments that enabled high-energy
beams of particles to be produced in laboratories. As a consequence, a wide
range of controlled scattering experiments could be performed and the greater
use of computers meant that sophisticated analysis techniques could be devel-
oped to handle the huge quantities of data that were being produced. By the
1960s this had resulted in the discovery of a very large number of unstable
particles with very short lifetimes and there was an urgent need for a theory that
could make sense of all these states. This emerged in the mid 1960s in the form
of the so-called quark model, first suggested by Murray Gell-Mann and
independently and simultaneously by George Zweig, who postulated that the
new particles were bound states of three families of more fundamental physical
particles.
10
X-rays had already been observed by Röntgen in 1895 (for which he received the first Nobel Prize in
Physics in 1901) and -rays were seen by Villard in 1900, but it was Planck who first made the startling
suggestion that electromagnetic energy was quantized. For this he was awarded the 1918 Nobel Prize in
Physics. Many years later, he said that his hypothesis was an ‘act of desperation’ as he had exhausted all
other possibilities.
11
However, in Section 3.1.4 we will discuss recent evidence that neutrinos have very small, but non-zero,
masses.
12
A description of this experiment is given in Chapter 12 of Tr75. Frederick Reines shared the 1995 Nobel
Prize in Physics for his work in neutrino physics and particularly for the detection of the electron neutrino.
HISTORY 5
Gell-Mann called these quarks (q).13 Because no free quarks were detected
experimentally, there was initially considerable scepticism for this view. We now
know that there is a fundamental reason why quarks cannot be observed as free
particles (it will be discussed in Chapter 5), but at the time many physicists looked
upon quarks as a convenient mathematical description, rather than physical
particles.14 However, evidence for the existence of quarks as real particles came
in the 1960s from a series of experiments analogous to those of Rutherford and his
co-workers, where high-energy beams of electrons and neutrinos were scattered
from nucleons. (These experiments will also be discussed in Chapter 5.) Analysis
of the angular distributions of the scattered particles showed that the nucleons were
themselves bound states of three point-like charged entities, with properties
consistent with those hypothesized in the quark model. One of these properties
was unexpected and unusual: quarks have fractional electric charges, in practice
13 e and þ 23 e. This is essentially the picture today, where elementary particles
are now considered to be a small number of physical entities, including quarks, the
electron, neutrinos, the photon and a few others we shall meet, but no longer
nucleons.
The best theory of elementary particles we have at present is called, rather
prosaically, the standard model. This aims to explain all the phenomena of particle
physics, except those due to gravity, in terms of the properties and interactions of a
small number of elementary (or fundamental) particles, which are now defined as
being point-like, without internal structure or excited states. Particle physics thus
differs from nuclear physics in having a single theory to interpret its data.
An elementary particle is characterized by, amongst other things, its mass, its
electric charge and its spin. The latter is a permanent angular momentum
possessed by all particles in quantum theory, even when they are at rest. Spin
has no classical analogue and is not to be confused with the use of the same word
in classical physics, where it usually refers to the (orbital) angular momentum of
extended objects. The maximum value of the spin angular momentum about any
axis is shðh h=2Þ, where h is Planck’s constant and s is the spin quantum
number, or spin for short. It has a fixed value for particles of any given type (for
example s ¼ 12 for electrons) and general quantum mechanical principles restrict
the possible values of s to be 0, 12, 1, 32, . . .. Particles with half-integer spin are
called fermions and those with integer spin are called bosons. There are three
families of elementary particles in the standard model: two spin-12 families of
fermions called leptons and quarks; and one family of spin-1 bosons. In addition,
13
Gell-Mann received the 1969 Nobel Prize in Physics for ‘contributions and discoveries concerning the
classification of elementary particles and their interactions’. For the origin of the word ‘quark’, he cited the
now famous quotation ‘Three quarks for Muster Mark’ from James Joyce’s book Finnegans Wake. Zweig
had suggested the name ‘aces’, which with hindsight might have been more appropriate, as later experiments
revealed that there were four and not three families of quarks.
14
This was history repeating itself. In the early days of the atomic model many very distinguished scientists
were reluctant to accept that atoms existed, because they could not be ‘seen’ in a conventional sense.
6 CH1 BASIC CONCEPTS
at least one other spin-0 particle, called the Higgs boson, is postulated to explain
the origin of mass within the theory.15
The most familiar elementary particle is the electron, which we know is bound
in atoms by the electromagnetic interaction, one of the four forces of nature.16 One
test of the elementarity of the electron is the size of its magnetic moment. A
charged particle with spin necessarily has an intrinsic magnetic moment l. It can
be shown from the Dirac equation that a point-like spin-12 particle of charge q and
mass m has a magnetic moment l ¼ ðq=mÞS, where S is its spin vector, and hence l
has magnitude ¼ q h=2m. The magnetic moment of the electron very accurately
obeys this relation, confirming that electrons are elementary.
The electron is a member of the family of leptons. Another is the neutrino,
which was mentioned earlier as a decay product in -decays. Strictly speaking, this
particle should be called the electron neutrino, written e , because it is always
produced in association with an electron (the reason for this is discussed in
Section 3.1.1). The force responsible for -decay is an example of a second
fundamental force, the weak interaction. Finally, there is the third force, the
(fundamental) strong interaction, which, for example, binds quarks in nucleons.
The strong nuclear force mentioned in Section 1.1.1 is not the same as this
fundamental strong interaction, but is a consequence of it. The relation between
the two will be discussed in more detail later.
The standard model also specifies the origin of these three forces. In classical
physics the electromagnetic interaction is propagated by electromagnetic waves,
which are continuously emitted and absorbed. While this is an adequate descrip-
tion at long distances, at short distances the quantum nature of the interaction must
be taken into account. In quantum theory, the interaction is transmitted discon-
tinuously by the exchange of photons, which are members of the family of
fundamental spin-1 bosons of the standard model. Photons are referred to as the
gauge bosons, or ‘force carriers’, of the electromagnetic interaction. The use of the
word ‘gauge’ refers to the fact that the electromagnetic interaction possesses a
fundamental symmetry called gauge invariance. For example, Maxwell’s equa-
tions of classical electromagnetism are invariant under a specific phase transfor-
mation of the electromagnetic fields – the gauge transformation.17 This property is
common to all the three interactions of nature we will be discussing and has
profound consequences, but we will not need its details in this book. The weak and
strong interactions are also mediated by the exchange of spin-1 gauge bosons. For
the weak interaction these are the W þ , W and Z 0 bosons (again the superscripts
denote the electric charges) with masses about 80–90 times the mass of the proton.
15
In the theory without the Higgs boson, all elementary particles are predicted to have zero mass, in obvious
contradiction with experiment. A solution to this problem involving the Higgs boson will be discussed briefly
in Chapter 9.
16
Gravity is so weak that it can be neglected in nuclear and particle physics at presently accessible energies.
Because of this, we will often refer in practice to the three forces of nature.
17
See, for example, Appendix C.2 of Ma97.
RELATIVITY AND ANTIPARTICLES 7
For the strong interaction, the force carriers are called gluons. There are eight
gluons, all of which have zero mass and are electrically neutral.18
In addition to the elementary particles of the standard model, there are other
important particles we will be studying. These are the hadrons, the bound states of
quarks. Nucleons are examples of hadrons,19 but there are several hundred more,
not including nuclei, most of which are unstable and decay by one of the three
interactions. It was the abundance of these states that drove the search for a
simplifying theory that would give an explanation for their existence and led to the
quark model in the 1960s. The most common unstable example of a hadron is the
pion, which exists in three electrical charge states, written ðþ ; 0 ; Þ. Hadrons
are important because free quarks are unobservable in nature and so to deduce
their properties we are forced to study hadrons. An analogy would be if we had to
deduce the properties of nucleons by exclusively studying the properties of nuclei.
Since nucleons are bound states of quarks and nuclei are bound states of
nucleons, the properties of nuclei should, in principle, be deducible from the
properties of quarks and their interactions, i.e. from the standard model. In
practice, however, this is far beyond present calculational techniques and some-
times nuclear and particle physics are treated as two almost separate subjects.
However, there are many connections between them and in introductory treatments
it is still useful to present both subjects together.
The remaining sections of this chapter are devoted to introducing some of the
basic theoretical tools needed to describe the phenomena of both nuclear and
particle physics, starting with a key concept: antiparticles.
18
Note that the word ‘electric’ has been used when talking about charge. This is because the weak and strong
interactions also have associated ‘charges’ which determine the strengths of the interactions, just as the
electric charge determines the strength of the electromagnetic interaction. This will be discussed in more
detail in later chapters.
19
The magnetic moments of the proton and neutron do not obey the prediction of the Dirac equation and this
is evidence that nucleons have structure and are not elementary. The proton magnetic moment was first
measured by Otto Stern using a molecular beam method that he developed and for this he received the 1943
Nobel Prize in Physics.
8 CH1 BASIC CONCEPTS
classical radius of the proton, which is roughly 1015 m. This in turn requires
electron energies that are greater than 103 times the rest energy of the electron,
implying electron velocities very close to the speed of light. Hence any explanation
of the phenomena of elementary particle physics must take account of the
requirements of the theory of special relativity, in addition to those of quantum
theory. There are very few places in particle physics where a non-relativistic
treatment is adequate, whereas the need for a relativistic treatment is far less in
nuclear physics.
Constructing a quantum theory that is consistent with special relativity leads to
the conclusion that for every particle of nature, there must exist an associated
particle, called an antiparticle, with the same mass as the corresponding particle.
This important theoretical prediction was first made by Dirac and follows from the
solutions of the equation he first wrote down to describe relativistic electrons.20
The Dirac equation is of the form
@Cðx; tÞ
ih ¼ Hðx; p
^ÞCðx; tÞ; ð1:1Þ
@t
^ þ mc2 :
H ¼ ca p ð1:2Þ
The coefficients a and are determined by the requirement that the solutions of
Equation (1.1) are also solutions of the Klein–Gordon equation 21
@ 2 Cðx; tÞ
h2 ¼ h2 c2 r2 Cðx; tÞ þ m2 c4 Cðx; tÞ: ð1:3Þ
@t2
This leads to the conclusion that a and cannot be simple numbers; their simplest
forms are 4
4 matrices. Thus the solutions of the Dirac equation are four-
component wavefunctions (called spinors) with the form22
0 1
C1 ðx; tÞ
B C2 ðx; tÞ C
Cðx; tÞ ¼ B C
@ C3 ðx; tÞ A: ð1:4Þ
C4 ðx; tÞ
20
Paul Dirac shared the 1933 Nobel Prize in Physics with Erwin Schrödinger. The somewhat cryptic citation
stated ‘for the discovery of new productive forms of atomic theory’.
21
This is a relativistic equation, which is ‘derived’ by starting from the relativistic mass–energy relation
E2 ¼ p2 c2 þ m2 c4 and using the usual quantum mechanical operator substitutions, p ^ ¼ ihr and
E ¼ ih@=@t.
22
The details may be found in most quantum mechanics books, for example, pp. 475–477 of Sc68.
SYMMETRIES AND CONSERVATION LAWS 9
The interpretation of Equation (1.4) is that the four components describe the two
spin states of a negatively charged electron with positive energy and the two spin
states of a corresponding particle having the same mass but with negative energy.
Two spin states arise because in quantum mechanics the projection in any direction
of the spin vector of a spin-12 particle can only result in one of the two values 12,
called ‘spin up’ and ‘spin down’, respectively. The two energy solutions arise from
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the two solutions of the relativistic mass–energy relation E ¼ p2 c2 þ m2 c4 .
The latter states can be shown to behave in all respects as positively charged
electrons (called positrons), but with positive energy. The positron is referred to as
the antiparticle of the electron. The discovery of the positron by Anderson in 1933,
with all the predicted properties, was a spectacular verification of the Dirac
prediction.
Although Dirac originally made his prediction for electrons, the result is general
and is true whether the particle is an elementary particle or a hadron. If we denote
a particle by P, then the antiparticle is in general written with a bar over it, i.e. P .
23
For example, the antiparticle of the proton is the antiproton p, with negative
electric charge; and associated with every quark, q, is an antiquark, q. However,
for some very common particles the bar is usually omitted. Thus, for example,
in the case of the positron eþ , the superscript denoting the charge makes
explicit the fact that the antiparticle has the opposite electric charge to that of
its associated particle. Electric charge is just one example of a quantum number
(spin is another) that characterizes a particle, whether it is elementary or composite
(i.e. a hadron).
Many quantum numbers differ in sign for particle and antiparticle, and electric
charge is an example of this. We will meet others later. When brought together,
particle–antiparticle pairs, each of mass m, can annihilate, releasing their com-
bined rest energy 2mc2 as photons or other particles. Finally, we note that there is
symmetry between particles and antiparticles, and it is a convention to call the
electron the particle and the positron its antiparticle. This reflects the fact that the
normal matter contains electrons rather than positrons.
23
Carl Anderson shared the 1936 Nobel Prize in Physics for the discovery of the positron. The 1958 Prize
was awarded to Emilio Segrè and Owen Chamberlain for their discovery of the antiproton.
10 CH1 BASIC CONCEPTS
numbers, on their spins.24 Another very important invariance that we have briefly
mentioned is gauge invariance. This fundamental property of all three interactions
restricts the forms of the interactions in a profound way that initially is contra-
dicted by experiment. This is the prediction of zero masses for all elementary
particles, mentioned earlier. There are theoretical solutions to this problem whose
experimental verification (or otherwise) is probably the most eagerly awaited
result in particle physics today.
In nuclear and particle physics we need to consider additional symmetries of the
Hamiltonian and the conservation laws that follow and in the remainder of this
section we discuss two of the most important of these that we will need later –
parity and charge conjugation.
1.3.1 Parity
Parity was first introduced in the context of atomic physics by Eugene Wigner in
1927.25 It refers to the behaviour of a state under a spatial reflection, i.e. x ! x.
If we consider a single-particle state, represented for simplicity by a non-
relativistic wavefunction ðx; tÞ, then under the parity operator, P ^,
^ 2 ðx; tÞ ¼ PP
P ^ ðx; tÞ ¼ P2 ðx; tÞ; ð1:6Þ
then
^
P p ðx; tÞ ¼P p ðx; tÞ ¼P p ðx; tÞ ð1:8Þ
24
These points are explored in more detail in, for example, Chapter 4 of Ma97.
25
Eugene Wigner shared the 1963 Nobel Prize in Physics, principally for his work on symmetries.
SYMMETRIES AND CONSERVATION LAWS 11
a parity transformation on the position vectors of all particles in the system. Parity
is therefore conserved, by which we mean that the total parity quantum number
remains unchanged in the interaction. Compelling evidence for parity conservation
in the strong and electromagnetic interactions comes from the absence of
transitions between nuclear states and atomic states, respectively, that would
violate parity conservation. The evidence for non-conservation of parity in the
weak interaction will be discussed in detail in Chapter 6.
There is also a contribution to the total parity if the particle has an orbital
angular momentum l. In this case its wave function is a product of a radial part Rnl
and an angular part Ylm ð
; Þ:
where n and m are the principal and magnetic quantum numbers and Ylm ð
; Þ is a
spherical harmonic. It is straightforward to show from the relations between
Cartesian ðx; y; zÞ and spherical polar co-ordinates ðr;
; Þ, i.e.
r ! r; ! ; ! þ ; ð1:11Þ
Equation (1.12) may easily be verified directly for specific cases; for example,
for the first three spherical harmonics,
12
1
1
1 3 2 3 2
Y00 ¼ ; Y10 ¼ cos
; Y11 ¼ sin
ei : ð1:13Þ
4 4 8
Hence
^
P lmn ðxÞ ¼P lmn ðxÞ ¼ PðÞl lmn ðxÞ; ð1:14Þ
^ ja;
C ai ¼ Ca ja; a i; ð1:15aÞ
and
^ jb;
C bi ¼ j
b; b i; ð1:15bÞ
because interchanging the pions reverses their relative positions in the spatial
wavefunction. The same factor occurs for spin-12 fermion pairs f f , but in addition
there are two other factors. The first is ð1ÞSþ1 , where S is the total spin of the pair.
26
A phase factor could have been inserted in Equation (1.15b), but it is straightforward to show that the
relative phase of the two states b and b cannot be measured and so a phase introduced in this way would have
no physical consequences.
INTERACTIONS AND FEYNMAN DIAGRAMS 13
and
where "i ð#i Þ represents particle i having spin ‘up’ (‘down’) in the z-direction. A
second factor ð1Þ arises whenever fermions and antifermions are interchanged.
This has its origins in quantum field theory.27 Combining these factors, finally
we have
^ j f f ; J; L; Si ¼ ð1ÞLþS j f f ; J; L; Si;
C ð1:18Þ
for fermion–antifermion pairs having total, orbital and spin angular momentum
quantum numbers J, L and S, respectively.
1.4.1 Interactions
e þ n ! e þ p; ð1:19Þ
e þ p ! e þ p ð1:20Þ
27
See, for example, pp. 249–250 of Go86.
14 CH1 BASIC CONCEPTS
represents an electron and proton interacting to give the same particles in the final
state, but in general travelling in different directions. In such equations, conserved
quantum numbers must have the same total values in initial and final states.
Particles may be transferred from initial to final states and vice versa, when they
become antiparticles. Thus the process
þ p ! þ p; ð1:21aÞ
p þ p ! þ þ ; ð1:22Þ
which is obtained by taking the proton from the final state to an antiproton in the
initial state and the negatively charged pion in the initial state to a positively
charged pion in the final state.
The interactions in Equations (1.20) and (1.21a), in which the particles remain
unchanged, are examples of elastic scattering, in contrast to the reactions in
Equations (1.19) and (1.22), where the final-state particles differ from those in the
initial state. Collisions between a given pair of initial particles do not always lead
to the same final state, but can lead to different final states with different
probabilities. For example, the collision of a negatively charged pion and a proton
can give rise to elastic scattering (Equation (1.21a)) and a variety of other
reactions, such as
þ p ! n þ 0 and þ p ! p þ þ þ þ ; ð1:21bÞ
a þ A ! a þ A ; A ! A þ ; ð1:23Þ
n ! p þ e þ e ; ð1:24Þ
INTERACTIONS AND FEYNMAN DIAGRAMS 15
with a mean lifetime of about 900 s.28 The same notation can also be used in
nuclear physics. For example, many nuclei decay via the -decay reaction. Thus,
denoting a nucleus with Z protons and N nucleons as (Z, N ), we have
ðZ; NÞ ! ðZ 1; NÞ þ eþ þ e : ð1:25Þ
This is also a weak interaction. This reaction is effectively the decay of a proton
bound in a nucleus. Although a free proton cannot decay by the -decay
p ! n þ eþ þ e because it violates energy conservation (the final-state particles
have greater total mass than the proton), a proton bound in a nucleus can decay
because of its binding energy. This will be explained in Chapter 2.
The forces producing all the above interactions are due to the exchange of particles
and a convenient way of illustrating this is to use Feynman diagrams. There are
mathematical rules and techniques associated with these that enable them to be
used to calculate the quantum mechanical probabilities for given reactions to
occur, but in this book Feynman diagrams will only be used as a convenient very
useful pictorial description of reaction mechanisms.
We first illustrate them at the level of elementary particles for the case of
electromagnetic interactions, which arise from the emission and/or absorption of
photons. For example, the dominant interaction between two electrons is due to the
exchange of a single photon, which is emitted by one electron and absorbed by the
other. This mechanism, which gives rise to the familiar Coulomb interaction at
large distances, is illustrated in the Feynman diagram Figure 1.1(a).
In such diagrams, we will use the convention that particles in the initial state are
shown on the left and particles in the final state are shown on the right. (Some
authors take time to run along the y-axis.) Spin-12 fermions (such as the electron)
28
The reason that this decay involves an antineutrino rather than a neutrino will become clear in Chapter 3.
16 CH1 BASIC CONCEPTS
are drawn as solid lines and photons are drawn as wiggly lines. Arrowheads
pointing to the right indicate that the solid lines represent electrons. In the case of
photon exchange between two positrons, which is shown in Figure 1.1(b), the
arrowheads on the antiparticle (positron) lines are conventionally shown as
pointing to the left. In interpreting these diagrams, it is important to remember:
(1) that the direction of the arrows on fermion lines does not indicate the particle’s
direction of motion, but merely whether the fermions are particles or antiparticles,
and (2) that particles in the initial state are always to the left and particles in the
final state are always to the right.
A feature of the above diagrams is that they are constructed from combinations
of simple three-line vertices. This is characteristic of electromagnetic processes.
Each vertex has a line corresponding to a single photon being emitted or absorbed,
while one fermion line has the arrow pointing toward the vertex and the other away
from the vertex, guaranteeing charge conservation at the vertex, which is one of
the rules of Feynman diagrams.29 For example, a vertex like Figure 1.2 would
correspond to a process in which an electron emitted a photon and turned into a
positron. This would violate charge conservation and is therefore forbidden.
e–
e+
Feynman diagrams can also be used to describe the fundamental weak and
strong interactions. This is illustrated by Figure 1.3(a), which shows contributions
to the elastic weak scattering reaction e þ e ! e þ e due to the exchange of a
Z 0 , and by Figure 1.3(b), which shows the exchange of a gluon g (represented by a
coiled line) between two quarks, which is a strong interaction.
Feynman diagrams can also be drawn at the level of hadrons. As an illustration,
Figure 1.4 shows the exchange of a charged pion (shown as a dashed line) between
a proton and a neutron. We shall see later that this mechanism is a major
contribution to the strong nuclear force between a proton and a neutron.
We turn now to consider in more detail the relation between exchanged particles
and forces.
29
Compare Kirchhoff’s laws in electromagnetism.
PARTICLE EXCHANGE: FORCES AND POTENTIALS 17
Figure 1.3 (a) Contributions of (a) Z 0 exchange to the elastic weak scattering reaction
e þ e ! e þ e , and (b) gluon exchange contribution to the strong interaction q þ q ! q þ q
particle A, the lower vertex represents the virtual process (‘virtual’ because X does
not appear as a real particle in the final state),
and
where p ¼ jpj. The energy difference between the final and initial states is given by
E ¼ EX þ EA MA c2 ! 2pc; p!1
! MX c2 ; p!0 ð1:30Þ
and thus E MX c2 for all p, i.e. energy is not conserved. However, by the
Heisenberg uncertainty principle, such an energy violation is allowed, but only for
a time
h=E, so we immediately obtain
r R h=MX c ð1:31Þ
30
A resumé of relativistic kinematics is given in Appendix B.
PARTICLE EXCHANGE: FORCES AND POTENTIALS 19
as the maximum distance over which X can propagate before being absorbed by
particle B. This maximum distance is called the range of the interaction and this
was the sense of the word used in Section 1.1.1.
The electromagnetic interaction has an infinite range, because the exchanged
particle is a massless photon. In contrast, the weak interaction is associated with
the exchange of very heavy particles – the W and Z bosons. These lead to ranges
that from Equation (1.31) are of the order of RW;Z 2
1018 m. The funda-
mental strong interaction has infinite range because, like the photon, gluons have
zero mass. On the other hand, the strong nuclear force, as exemplified by Figure 1.4
for example, has a much shorter range of approximately (1–2)
1015 m. We will
comment briefly on the relation between these two different manifestations of the
strong interaction in Section 7.1.
In the limit that MA becomes large, we can regard B as being scattered by a static
potential of which A is the source. This potential will in general be spin dependent,
but its main features can be obtained by neglecting spin and considering X to be a
spin-0 boson, in which case it will obey the Klein–Gordon equation
@ 2 ðx; tÞ
h2 ¼ h2 c2 r2 ðx; tÞ þ MX2 c4 ðx; tÞ: ð1:32Þ
@t2
MX2 c4
r2 ðxÞ ¼ ðxÞ; ð1:33Þ
h2
where ðxÞ is interpreted as a static potential. For MX ¼ 0 this equation is the same
as that obeyed by the electrostatic potential, and for a point charge e interacting
with a point charge þe at the origin, the appropriate solution is the Coulomb potential
e2 1
VðrÞ ¼ eðrÞ ¼ ; ð1:34Þ
4"0 r
where r ¼ jxj and "0 is the dielectric constant. The corresponding solution in the
case where MX2 6¼ 0 is easily verified by substitution to be
g2 er=R
VðrÞ ¼ ; ð1:35Þ
4 r
where R is the range defined earlier and g, the so-called coupling constant, is a
parameter associated with each vertex of a Feynman diagram and represents the
20 CH1 BASIC CONCEPTS
basic strength of the interaction.31 For simplicity, we have assumed equal strengths
for the couplings of particle X to the particles A and B.
The form of VðrÞ in Equation (1.35) is called a Yukawa potential, after the physicist
who first introduced the idea of forces due to massive particle exchange in 1935.32
As MX ! 0, R ! 1 and the Coulomb potential is recovered from the Yukawa
potential, while for very large masses the interaction is approximately point-like
(zero range). It is conventional to introduce a dimensionless parameter X by
g2
X ¼ ; ð1:36Þ
4hc
that characterizes the strength of the interaction at short distances r R. For the
electromagnetic interaction this is the fine structure constant
31
Although we call g a (point) coupling constant, in general it will have a dependence on the momentum
carried by the exchanged particle. We ignore this in what follows.
32
For this insight, Hideki Yukawa received the 1949 Nobel Prize in Physics.
33
Like g, the coupling X will in general have a dependence on the momentum carried by particle X. In the
case of the electromagnetic interaction, this dependence is relatively weak.
OBSERVABLE QUANTITIES: CROSS SECTIONS AND DECAY RATES 21
derived from the quantum theory of the underlying interaction.34 We will not
pursue this in detail in this book, but rather will show in principle their relation to
observables, i.e. things that can be measured, concentrating on the cases of two-
body scattering reactions and decays of unstable states.
1.6.1 Amplitudes
The intermediate step is the amplitude f, the modulus squared of which is directly
related to the probability of the process occurring. It is also called the invariant
amplitude because, as we shall show, it is directly related to observable quantities
and these have to be the same in all inertial frames of reference. To get some
qualitative idea of the structure of f, we will use non-relativistic quantum
mechanics and assume that the coupling constant g2 is small compared with
4hc, so that the interaction is a small perturbation on the free particle solution,
which will be taken as plane waves.
If we expand the amplitude in a power series in g2 and keep just the first term
(i.e. lowest-order perturbation theory), then the amplitude for a particle in an initial
state with momentum qi to be scattered to a final state with momentum qf by a
potential VðxÞ is proportional to
ð
f ðq2 Þ ¼ d3 xVðxÞ exp½iq x=h; ð1:38Þ
and
where r ¼ jxj. For the Yukawa potential, the integral in Equation (1.38) gives
h2
g2
f ðq2 Þ ¼ : ð1:41Þ
q2 þ MX2 c2
34
In the case of the electromagnetic interaction, the theory is called Quantum Electrodynamics (QED) and is
spectacularly successful in explaining experimental results. Richard Feynman shared the 1965 Nobel Prize in
Physics with Sin-Itiro Tomonoga and Julian Schwinger for their work on formulating quantum electro-
dynamics. The Feynman Rules are discussed in an accessible way in Gr87.
35
See, for example, Chapter 11 of Ma92.
22 CH1 BASIC CONCEPTS
GF 4W
3
¼ ¼ 1:166
105 GeV2 : ð1:42Þ
ðhcÞ ðMW c2 Þ2
The numerical value is obtained from analyses of decay processes, including that
of the neutron and a heavier version of the electron called the muon, whose
properties will be discussed in Chapter 3.
All the above is for the exchange of a single particle. It is also possible to draw
more complicated Feynman diagrams that correspond to the exchange of more
than one particle. An example of such a diagram for elastic e e scattering, where
two photons are exchanged, is shown in Figure 1.6.
The number of vertices in any diagram is called the order n, and when the
amplitude associated p with
ffiffiffiffi n any given Feynman diagram is calculated, it always
contains a factor of ð Þ . Since the probability is proportional to the square of
the modulus of the amplitude, the former will contain a factor n. The probability
associated with the single-photon exchange diagrams of Figure 1.1 thus contain a
factor of 2 and the contribution from two-photon exchange is of the order of 4 .
OBSERVABLE QUANTITIES: CROSS SECTIONS AND DECAY RATES 23
As 1=137, the latter is usually very small compared with the contribution
from single-photon exchange.
This is a general feature of electromagnetic interactions: because the fine structure
constant is very small, in most cases only the lowest-order diagrams that contribute to
a given process need be taken into account, and more complicated higher-order
diagrams with more vertices can (to a good approximation) be ignored in many
applications.
1.6.2 Cross-sections
The next step is to relate the amplitude to measurables. For scattering reactions,
the appropriate observable is the cross-section. In a typical scattering experiment, a
beam of particles is allowed to hit a target and the rates of production of various
particles in the final state are counted.36 It is clear that the rates will be
proportional to: (a) the number N of particles in the target illuminated by the
beam, and (b) the rate per unit area at which beam particles cross a small surface
placed in the beam at rest with respect to the target and perpendicular to the beam
direction. The latter is called the flux and is given by
J ¼ nb vi ; ð1:43Þ
where nb is the number density of particles in the beam and vi their velocity in the
rest frame of the target. Hence the rate Wr at which a specific reaction r occurs in a
particular experiment can be written in the form
Wr ¼ JNr ; ð1:44aÞ
where nt is the number of target particles per unit volume and t is the thickness of
the target. If the target consists of an isotopic species of atomic mass MA (in atomic
mass units-defined in Section 1.7 below), then nt ¼ NA =MA , where is the density
of the target and NA is Avogadro’s constant. Thus, Equation (1.44b) may be written
36
The practical aspects of experiments are discussed in Chapter 4.
24 CH1 BASIC CONCEPTS
where ðtÞ is a measure of the amount of material in the target, expressed in units
of mass per unit area. The form of Equation (1.44c) is particularly useful for the
case of thin targets, commonly used in experiments. In the above, the product JN is
called the luminosity L, i.e.
L JN ð1:45Þ
and contains all the dependencies on the densities and geometries of the beam and
target. The cross-section is independent of these factors.
It can be seen from the above equations that the cross-section has the dimensions
of an area; the rate per target particle Jr at which the reaction occurs is equal to
the rate at which beam particles would hit a surface of area r , placed in the beam
at rest with respect to the target and perpendicular to the beam direction. Since the
area of such a surface is unchanged by a Lorentz transformation in the beam
direction, the cross-section is the same in all inertial frames of reference, i.e. it is a
Lorentz invariant.
The quantity r is better named the partial cross-section, because it is the cross-
section for a particular reaction r. The total cross-section is defined by
X
r : ð1:46Þ
r
dr ð
; Þ
dWr JN d; ð1:47Þ
d
where dWr is the measured rate for the particles to be emitted into an element of
solid angle d ¼ d cos
d in the direction ð
; Þ, as shown in Figure 1.7.
The total cross-section is obtained by integrating the partial cross-section over
all angles, i.e.
ð 2 ð1
dr ð
; Þ
r ¼ d d cos
: ð1:48Þ
0 1 d
The final step is to write these formulae in terms of the scattering amplitude f ðq2 Þ
appropriate for describing the scattering of a non-relativistic spinless particle from
a potential.
To do this it is convenient to consider a single beam particle interacting with a
single target particle and to confine the whole system in an arbitrary volume V
(which cancels in the final result). The incident flux is then given by
J ¼ nb vi ¼ vi =V ð1:49Þ
OBSERVABLE QUANTITIES: CROSS SECTIONS AND DECAY RATES 25
Figure 1.7 Geometry of the differential cross-section: a beam of particles is incident along the
z-axis and collides with a stationary target at the origin; the differential cross-section is
proportional to the rate for particles to be scattered into a small solid angle d in the direction
ð
; Þ
vi dr ð
; Þ
dWr ¼ d: ð1:50Þ
V d
In quantum mechanics, provided the interaction is not too strong, the transition
rate for any process is given in perturbation theory by the Born approximation37
ð 2
2 3
dWr ¼ dx f VðxÞ i ðEf Þ: ð1:51Þ
h
The term ðEf Þ is the density-of-states factor (see below) and we take the initial
and final state wavefunctions to be plane waves:
1 1
i ¼ pffiffiffiffi exp½iqi x=h; f ¼ pffiffiffiffi exp½iqf x=h; ð1:52Þ
V V
37
This equation is a form of the Second Golden Rule. It is discussed in Appendix A.
26 CH1 BASIC CONCEPTS
where the final momentum qf lies within a small solid angle d located in the
direction ð
; Þ (see Figure 1.7.). Then, by direct integration,
2
dWr ¼ j f ðq2 Þj2 ðEf Þ; ð1:53Þ
hV 2
where f ðq2 Þ is the scattering amplitude defined in Equation (1.38).
The density of states ðEf Þ that appears in Equation (1.51) is the number of
possible final states with energy lying between Ef and Ef þ dEf and is given by38
V q2f
ðEf Þ ¼ d: ð1:54Þ
ð2hÞ3 vf
If we use this and Equation (1.53) in Equation (1.50), we have
d 1 q2
¼ 2 4 f j f ðq2 Þj2 : ð1:55Þ
d 4 h vi vf
Although this result has been derived in the laboratory system, because we have
taken a massive target it is also valid in the centre-of-mass system. For a finite
mass target it would be necessary to make a Lorentz transformation on Equation
(1.55). The expression is also true for the general two-body relativistic scattering
process a þ b ! c þ d.
All the above is for spinless particles, so finally we have to generalize Equation
(1.55) to include the effects of spin. Suppose the initial-state particles a and b, have
spins sa and sb and the final-state particles c and d have spins sc and sd . The total
numbers of spin substates available to the initial and final states are gi and gf ,
respectively, given by
If the initial particles are unpolarized (which is the most common case in practice),
then we must average over all possible initial spin configurations (because each is
equally likely) and sum over the final configurations. Thus, Equation (1.55) becomes
d gf q2
¼ 2 4 f jMfi j2 ; ð1:57Þ
d 4 h vi vf
where
jMfi j2 j f ðq2 Þj2 ð1:58Þ
and the bar over the amplitude denotes a spin-average of the squared matrix
element.
38
The derivation is given in detail in Appendix A.
OBSERVABLE QUANTITIES: CROSS SECTIONS AND DECAY RATES 27
In the case of an unstable state, the observable of interest is its lifetime at rest , or
equivalently its natural decay width, given by ¼ h=, which is a measure of the
rate of the decay reaction. In general, an initial unstable state will decay to several
final states and in this case we define f as the partial width for channel f and
X
¼ f ð1:59Þ
f
Bf f = ð1:60Þ
where M is the mass of the decaying state and W is the invariant mass of the decay
products.39 The Breit–Wigner formula is shown in Figure 1.8 and is the same
formula that describes the widths of atomic and nuclear spectral lines. (The overall
factor depends on the spins of the particles involved.) It is a symmetrical bell-
shaped curve with a maximum at W ¼ M and a full width at half the maximum
height of the curve. It is proportional to the number of events with invariant mass W.
If an unstable state is produced in a scattering reaction, then the cross section for
that reaction will show an enhancement described by the same Breit–Wigner formula.
In this case we say we have produced a resonance state. In the vicinity of a resonance
of mass M, and width , the cross-section for the reaction i ! f has the form
i f
fi / ; ð1:62Þ
ðE Mc2 Þ2 þ 2 =4
where E is the total energy of the system. Again, the form of the overall constant
will depend on the spins of the particles involved. Thus, for example, if the
resonance particle has spin j and the spins of the initial particles are s1 and s2 , then
h2 2j þ 1 i f
fi ¼ : ð1:63Þ
qi ð2s1 þ 1Þð2s2 þ 1Þ ðE Mc2 Þ2 þ 2 =4
2
39
This form arises from a state that decays exponentially with time, although a proper proof of this is quite
lengthy. See, for example, Appendix B of Ma97.
28 CH1 BASIC CONCEPTS
In practice there will also be kinematical and angular momentum effects that will
distort this formula from its perfectly symmetric shape.
An example of resonance formation in p interactions is given in Figure 1.9,
which shows the p total cross-section in the centre-of-mass energy range
1.2–2.4 GeV. (The units used in the plots will become clear after the next section.)
Two enhancements can be seen that are of the approximate Breit–Wigner resonance
form and there are two other maxima at higher energies. In principle, the mass and
width of a resonance may be obtained by using a Breit–Wigner formula and varying
M and to fit the cross-section in the region of the enhancement. In practice, more
sophisticated methods are used that fit a wide range of data, including differential
cross-sections, simultaneously and also take account of non-resonant contributions
to the scattering. The widths obtained from such analyses are of the order of
100 MeV, with corresponding interaction times of order 1023 s, which is consistent
with the time taken for a relativistic pion to transit the dimension of a proton.
Resonances are also a feature of interactions in nuclear physics and we will return to
this in Section 2.9 when we discuss nuclear reaction mechanisms.
Problems
1.1 ‘Derive’ the Klein–Gordon equation using the information in Footnote 21 and verify
that Equation (1.35) is a static solution of the equation.
qffiffi
1.2 Verify that the spherical harmonic Y11 ¼ 38 sin
ei is an eigenfunction of parity
with eigenvalue P ¼ 1.
1.3 A proton and antiproton at rest in an S-state annihilate to produce 0 0 pairs. Show
that this reaction cannot be a strong interaction.
1.4 Suppose that an intrinsic C-parity factor is introduced into Equation (1.15b), which
then becomes C ^ jb; b i ¼ Cb j
b; b i. Show that the eigenvalue corresponding to any
^ is independent of Cb , so that Cb cannot be measured.
eigenstate of C
1.5 Consider the reaction d ! nn, where d is a spin-1 S-wave bound state of a proton
and a neutron called the deuteron and the initial pion is at rest. Deduce the intrinsic
parity of the negative pion.
1.6 Write down equations in symbol form that describe the following interactions:
(b) n ! p þ e þ e ;
PROBLEMS 31
(c) eþ e ! eþ e ;
g2 er=R
VðrÞ ¼ ;
4 r
where R ¼
h=mc is the range and r ¼ jxj.
1.11 A thin (density 1 mg cm2 ) target of 24 Mg (MA ¼ 24:3 atomic mass units) is
bombarded with a 10 nA beam of alpha particles. A detector subtending a solid
angle of 2
103 sr, records 20 protons/s. If the scattering is isotropic, determine
the cross-section for the 24 Mgð; pÞ reaction.
2
Nuclear Phenomenology
In this chapter we start to examine some of the things that can be learned from
experiments, beginning with basic facts about nuclei, including what can be deduced
about their shapes and sizes. Then we discuss the important topic of nuclear stability
and the phenomenology of the various ways that unstable nuclei decay to stable
states. Finally, we briefly review the classification of reactions in nuclear physics.
Before that we need to introduce some notation.
We will also refer to A as the nucleon number. The charge on the nucleus is
þZe, where e is the absolute value of the electric charge on the electron. Nuclei
with combinations of these three numbers are also called nuclides and are written
A
Y or AZ Y, where Y is the chemical symbol for the element. Some other common
nomenclature is:
The concept of isotopes was introduced in Chapter 1. For example, stable isotopes
of carbon are 12 C and 13 C, and the unstable isotope used in dating ancient objects
(see later in this chapter) is 14 C; all three have Z ¼ 6.
Just as in the case of electrons in atoms, the forces that bind the nucleons in
nuclei contribute to the total mass of an atom M(Z, A) and in terms of the masses of
the proton Mp and neutron Mn
and Mc2 is called the binding energy B. Binding energies may be calculated if
masses are measured accurately. One way of doing this is by using the techniques
of mass spectroscopy. The principle of the method is shown in Figure 2.1.
Figure 2.1 Schematic diagram of a mass spectrometer (adapted from Kr88 Copyright John
Wiley & Sons, Inc.)
uniform magnetic field B2 where it will be bent into a circular path of radius ,
given by
mv ¼ qB2 ð2:3Þ
and since q, B2 and v are fixed, particles with a fixed ratio q=m will bend in a path
with a unique radius. Hence isotopes may be separated and focused onto a detector
(e.g. a photographic plate). In the common case where B1 ¼ B2 ¼ B,
q E
¼ 2 : ð2:4Þ
m B
where we use tilded quantities to denote the energy, mass, etc. of A . Equating the
total energy before the collision
~ þ ma c2 þ m
Etot ðfinalÞ ¼ Ef þ E ~ c2 ð2:6bÞ
gives the following expression for the change in energy of the nucleus:
1
Practical details of mass spectroscopy may be found in, for example, Chapter 3 of Kr88.
36 CH2 NUCLEAR PHENOMENOLOGY
This formula can be used iteratively to deduce E and hence the mass of the excited
nucleus A , from measurements of the initial and final energy of the projectile by
~ ¼ mA on the right-hand side because E is small in comparison
initially setting m
with mA . One final point is that the energies in Equation (2.9) are measured in the
laboratory system, whereas the final energies (masses) will be needed in the centre-of-
mass system.2 The necessary transformation is easily found to be
A similar formula to Equation (2.9) may be derived for the general reaction
A(a,b)B:
ma mb 2 1=2
E ¼ Ei 1 Ef 1 þ þ ma mb Ei Ef cos þ Q; ð2:11Þ
mB mB mB
Figure 2.2 Binding energy per nucleon as a function of mass number A for stable and long-
lived nuclei
2
A discussion of these two systems is given in Appendix C.
NUCLEAR SHAPES AND SIZES 37
This shows that B=A peaks at a value of 8.7 MeV for a mass number of about 56
(close to iron) and thereafter falls very slowly. Excluding very light nuclei, the
binding energy per nucleon is between 7 and 9 MeV over a wide range of the
periodic table. In Section 2.5 we will discuss a model that provides an explanation
for the shape of this curve.
where E is the total initial energy of the projectile and is the angle through which
it is scattered. Note that Equation (2.12) is of order 2 because it corresponds to
the exchange of a single photon. Although Equation (2.12) has a limited range of
applicability, it is useful to discuss the general features of electron scattering.
Equation (2.12) actually describes the scattering of a spin-0 point-like projectile of
unit charge from a fixed point-like target with electric charge Ze, i.e. the charge
3
Robert Hofstader shared the 1961 Nobel Prize in Physics for his pioneering electron scattering experiments.
38 CH2 NUCLEAR PHENOMENOLOGY
where ¼ v=c and v is the velocity of the initial electron. At higher energies, the
recoil of the target needs to be taken into account and this introduces a factor E0 =E
on the right-hand side of Equation (2.13), where E0 is the final energy of the
electron. At higher energies we also need to take account of the interaction with
the magnetic moment of the target in addition to its charge. The final form for the
differential cross-section is
d d E0 2
¼ 1 þ 2
tan ; ð2:14Þ
dO spin 1 dO Mott E 2
2
where
q2
¼ ð2:15Þ
4M 2 c2
and M is the target mass. Because the energy loss of the electron to the recoiling
nucleus is no longer negligible, q, the previous momentum transfer, has been
replaced by the four-momentum transfer q, whose square is
4EE0
q2 ¼ ðp p0 Þ2 ¼ 2m2e c2 2ðEE0 c2 jpjjp0 j cos Þ 2 sin2 ð=2Þ; ð2:16Þ
c
4
To remove any confusion, in the non-relativistic case, which we use in the rest of this chapter, q is
interpreted to be q ¼ jqj 0 where q p p0 , as was used in Section 1.6.1. We will need the four-
momentum definition of q in Chapter 6.
NUCLEAR SHAPES AND SIZES 39
i.e. the Fourier transform of the charge distribution.5 In the case of a spherically
symmetric charge distribution, the angular integrations in Equations (2.17) may be
done using spherical polar coordinates to give
ð
1
qr
2 4h
Fðq Þ ¼ rðrÞsin dr; ð2:18Þ
Zeq h
0
where q ¼ jqj and ðrÞ is the radial charge distribution. The final form of the
experimental cross-section in this approximation is given by6
d d
¼
Fðq2 Þ
2 : ð2:19Þ
dO expt dO Mott
Publisher's Note:
Permission to reproduce this image
online was not granted by the
copyright holder. Readers are kindly
requested to refer to the printed version
of this chapter.
Figure 2.3 Elastic differential cross-sections as a function of the scattering angle for 450 MeV
electrons from 58 Ni and 758 MeV electrons from 48 Ca; the solid lines are fits as described in the
text (adapted from Si75 (58Ni data) and Be67 (48Ca data), Copyright American Physical Society)
5
Strictly this formula assumes that the recoil of the target nucleus is negligible and the interaction is
relatively weak, so that perturbation theory may be used.
6
If the magnetic interaction were included, another form factor would be necessary, as is the case in high-
energy electron scattering discussed in Chapter 6.
40 CH2 NUCLEAR PHENOMENOLOGY
rapid decrease in the cross-section with angle. These features are common to all
elastic data, although not all nuclei show so many minima as those shown.
The minima are due to the form factor and we can make this plausible by taking
the simple case where the nuclear charge distribution is represented by a hard
sphere such that
ðrÞ ¼ constant; r
a
ð2:20Þ
¼0 r>a
where b qa= h. Thus Fðq2 Þ will be zero at values of b for which b ¼ tanðbÞ. In
practice, as we will see below, ðrÞ is not a hard sphere, and although it is
approximately constant for much of the nuclear volume, it falls smoothly to zero at
the surface. Smoothing the edges of the radial charge distribution (2.20) modifies
the positions of the zeros, but does not alter the argument that the minima in the
cross-sections are due to the spatial distribution of the nucleus. Their actual
positions and depths result from a combination of the form factor and the form of
the point-like amplitude. We shall see below that the minima can tell us about the
size of the nucleus.
If one measures the cross-section for a fixed energy at various angles (and hence
various q2 ), the form factor can in principle be extracted using Equation (2.19) and
one might attempt to find the charge distribution from the inverse Fourier
transform
ð
Ze
f ðxÞ ¼ Fðq2 Þ eiqx=h d3 q: ð2:22Þ
ð2Þ3
However, q2 only has a finite range for a fixed initial electron energy and even within
this range the rapid fall in the cross-section means that in practice measurements
cannot be made over a sufficiently wide range of angles for the integral in Equation
(2.22) to be evaluated accurately. Thus, even within the approximations used,
reliable charge distributions cannot be found from Equation (2.22). Therefore
different strategies must be used to deduce the charge distribution. In one approach,
plausible – but very general – parameterized forms (for example a sum of Gaussians)
are chosen for the charge distribution and are used to modify the point-like
electromagnetic interaction. The resulting Schrödinger (or Dirac) equation is solved
numerically to produce an amplitude, and hence a cross-section, for electron–
nucleus scattering. The parameters of the charge distribution are then varied to give
a good fit of the experimental data. The solid curves in Figure 2.3 are obtained in
this way.
NUCLEAR SHAPES AND SIZES 41
Figure 2.4 Radial charge distributions ch of various nuclei, in units of e fm3 ; the thickness
of the curves near r ¼ 0 is a measure of the uncertaintity in ch (adapted from Fr83)
Some radial charge distributions for various nuclei obtained by these methods
are shown in Figure 2.4. They are well represented by the form
0ch
ch ðrÞ ¼ ; ð2:23Þ
1 þ eðraÞ=b
From this we can deduce that the charge density is approximately constant in the
nuclear interior and falls fairly rapidly to zero at the nuclear surface, as anticipated
above. The value of 0ch is in the range 0.06–0.08 for medium to heavy nuclei and
decreases slowly with increasing mass number.
42 CH2 NUCLEAR PHENOMENOLOGY
This can be found from the form factor as follows. Expanding Equation (2.17) for
Fðq2 Þ gives
ð X1
2 1 1 ijqjrcos n 3
Fðq Þ ¼ f ðxÞ d x ð2:26Þ
Ze n¼0
n! h
7
The constant comes from a fit to a range of data, e.g. the compilation for 55
A
209 given in Ba77.
NUCLEAR SHAPES AND SIZES 43
by A=Z. Then we find an almost identical nuclear density in the nuclear interior for
all nuclei, i.e. the decrease in 0ch with increasing A is compensated by the increase
in A=Z with increasing A. The interior nuclear density is given by
Likewise, the effective nuclear matter radius for medium and heavy nuclei is
These are important results that will be used extensively later in this chapter and
elsewhere in this book.
To probe the nuclear (i.e. matter) density of nuclei experimentally, a strongly
interacting particle, i.e. a hadron, has to be used as the projectile. At high energies,
where elastic scattering is only a small part of the total interaction, the nucleus
behaves more like an absorbing sphere. In this case, the incident particle of
momentum p will have an associated quantum mechanical wave of wavelength
¼ h=p and will suffer diffraction-like effects, as in optics. To the extent that we are
dealing at high energies purely with the nuclear strong interaction (i.e. neglecting the
Coulomb interaction), the nucleus can be represented by a black disk of radius R and
the differential cross-section will have a Fraunhofer-like diffraction form, i.e.
d J1 ðqRÞ 2
/ ; ð2:34Þ
dO qR
where qR pR for small and J1 is a first-order Bessel function. For large qR,
2
½J1 ðqRÞ2 sin2 qR ; ð2:35Þ
qR 4
Figure 2.5 Elastic differential cross-sections for 52 MeV deuterons on 54 Fe (adapted from Hi68,
copyright Elsevier, with permission)
nuclear case the imaginary part of a complex potential describing the interaction
takes account of all the inelastic reactions. It is an essential feature of the model
that the properties of nuclei are mainly determined by their size, as this implies
that the same potential can account for the interaction of particles of different
energies with different nuclei. Apart from the theoretical basis provided by
analogy with classical optics, the model is essentially phenomenological, in that
the values of the parameters of the optical potentials are found by optimizing the fit
to the experimental data. This type of semi-phenomenological approach is
common in both nuclear and particle physics.
In practice, the Schrödinger equation is solved using a parameterized complex
potential where the real part is a sum of the Coulomb potential (for charged
projectiles), an attractive nuclear potential and a spin-orbit potential, and the
imaginary part is assumed to cause the incoming wave of the projectile to be
attenuated within the nucleus, thereby allowing for inelastic effects. Originally,
mathematical forms like Equation (2.23) were used to parameterize the real and
imaginary parts of the potential, but subsequent work indicated substantial differ-
ences between the form factors of the real and imaginary parts of the potential and so
different forms are now used for the imaginary part. The free parameters of the total
potential are adjusted to fit the data.
The optical model has achieved its greatest success in the scattering of nucleons,
but analyses using data obtained from light nuclei targets are also possible. A wide
range of scattering data can be accounted for to a high degree of precision by
NUCLEAR INSTABILITY 45
Figure 2.6 Differential cross-sections (normalized to the Rutherford cross-section) for the
elastic scattering of 30.3 MeV protons, for a range of nuclei compared with optical model
calculations; the solid and dashed lines represent the results using two different potentials
(adapted from Sa67, copyright Elsevier, with permission)
the model and examples of this are shown in Figure 2.6. The corresponding
wavefunctions are extensively used to extract information on nuclear structure.
The conclusions are in accord with those above deduced indirectly from electron
data.
Figure 2.7 The distribution of stable nuclei: the squares are the stable and long-lived nuclei
occurring in nature; other known nuclei lie within the jagged lines and are unstable. [adapted
from Ch97.)
neutron into a proton; conversely, a nucleus with a large surplus of protons converts
protons to neutrons. These are examples of -decays, already mentioned. A related
process is where an atomic electron is captured by the nucleus and a proton is
thereby converted to a neutron within the nucleus. This is electron capture and like
-decay is a weak interaction. The electron is usually captured from the innermost
shell and the process competes with -decay in heavy nuclei because the radius of
this shell (the K-shell) is close to the nuclei radius. The presence of a third particle in
the decay process, the neutrino (as first suggested by Fermi), means that the emitted
electrons (or positrons) have a continuous energy spectrum. The derivation and
analysis of the electron momentum spectrum will be considered in Chapter 7 when
we discuss the theory of -decay.
The maximum of the curve of binding energy per nucleon is at approximately
the position of iron (Fe) and nickel (Ni), which are therefore the most stable
nuclides. In heavier nuclei, the binding energy is smaller because of the larger
Coulomb repulsion. For still heavier nuclear masses, nuclei can decay sponta-
neously into two or more lighter nuclei, provided the mass of the parent nucleus is
larger than the sum of the masses of the daughter nuclei.
RADIOACTIVE DECAY 47
Most such nuclei decay via two-body decays and the commonest case is when
one of the daughter nuclei is a 4 He nucleus (i.e. an -particle: 4 He 2p2n, with
A ¼ 4 ; Z ¼ N ¼ 2). The -particle is favoured in such decays because it is a very
stable, tightly bound structure. Because this is a two-body decay, the -particle has
a unique energy and the total energy released, the so-called Q-value, is given by:
where the subscripts refer to parent and daughter nuclei and the -particle, and E is
a kinetic energy.
The term fission is used to describe the rare cases where the two daughters have
similar masses. If the decay occurs without external action, it is called spontaneous
fission to distinguish it from induced fission, where some external stimulus is required
to initiate the decay. Spontaneous fission only occurs with a probability greater than that
for -emission for nuclei with Z 110. The reason for this is discussed in Section 2.7.
Finally, nuclei may decay by the emission of photons, with energies in the -ray
part of the electromagnetic spectrum (gamma emission). This occurs when an
excited nuclear state decays to a lower state and is a common way whereby excited
states lose energy. The lower energy state is often the ground state. A competing
process is internal conversion, where the nucleus de-excites by ejecting an electron
from a low-lying atomic orbit. Both are electromagnetic processes. Electromag-
netic decays will be discussed in more detail in Chapter 7.
where N(t) is the number of radioactive nuclei in the sample at time t. The activity is
measured in becquerels (Bq), which is one decay per second.8 The probability here
refers to the total probability, because
could be the sum of decay probabilities for a
number of distinct final states in the same way that the total decay width of an
unstable particle is the sum of its partial widths. Integrating Equation (2.37) gives
AðtÞ ¼
N0 expð
tÞ; ð2:38Þ
8
An older unit, the curie (1 Ci ¼ 3:7 1010 Bq) is also still in common use. A typical laboratory radioactive
source has an activity of a few tens of kBq, i.e. Ci.
48 CH2 NUCLEAR PHENOMENOLOGY
Thus
Ð
1
Ð t exp½
t dt
t dNðtÞ 0 1
Ð ¼ 1 ¼ : ð2:40Þ
dNðtÞ Ð
exp½
t dt
0
This is the quantity we simply called ‘the lifetime’ in Chapter 1. The mean
lifetime is always used in particle physics, but another measure more commonly
used in nuclear physics is the half-life t12 , defined as the time for the number of
nuclei to fall by one half. Thus t12 ¼ ln2=
¼
ln2. In this book, the term lifetime
will be used for the mean lifetime, both for radioactive nuclei and unstable
hadrons, unless explicitly stated otherwise.
A well-known use of the radioactive decay law is in dating ancient specimens
using the known properties of radioactive nuclei. For organic specimens, carbon is
usually used. Carbon-14 is a radioactive isotope of carbon that is produced by the
action of cosmic rays on nitrogen in the atmosphere.9 If the flux of cosmic rays
remains roughly constant over time, then the ratio of 14 C to the stable most abundant
isotope 12 C reaches an equilibrium value of about 1 : 1012 . Both isotopes will be
taken up by living organisms in this ratio, but when the organism dies there is no
further interaction with the environment and the ratio slowly changes with time as
the 14 C nuclei decay by -decay to 14 N with a lifetime of 8:27 103 years. Thus, if
the ratio of 14 C to 12 C is measured, the age of the specimen may be estimated.10 The
actual measurements can be made very accurately because modern mass spectro-
meters can directly measure very small differences in the concentrations of 14 C and
12
C using only milligrams of material. Nevertheless, in practice, corrections are
made to agree with independent calibrations if possible, using, for example, tree-
ring growth data, because cosmic ray activity is not strictly constant with time.
In many cases the products of radioactive decay are themselves radioactive and so
a decay chain results. Consider a decay chain A ! B ! C ! , with decay
constants
A ;
B ;
C etc.. The variation of species A with time is given by Equation
(2.38), i.e.
NA ðtÞ ¼ NA ð0Þexpð
A tÞ; ð2:41Þ
9
Cosmic rays are high-energy particles, mainly protons, that impinge on the Earth’s atmosphere from space.
The products of the secondary reactions they produce may be detected at the Earth’s surface. Victor Hess
shared the 1936 Nobel Prize in Physics for the discovery of cosmic radiation.
10
This method of using radioactive carbon to date ancient objects was devised by Willard Libby, for which he
received the 1960 Nobel Prize in Chemistry.
RADIOACTIVE DECAY 49
but the differential equation for NB ðtÞ will have an extra term in it to take account
of the production of species B from the decay of species A:
A
NB ðtÞ ¼ NA ð0Þ½expð
A tÞ expð
B tÞ: ð2:43Þ
B
A
Similar equations may be found for decay sequences with more than two stages.
Thus, for a three-stage sequence
expð
A tÞ
NC ðtÞ ¼
A
B NA ð0Þ
ð
B
A Þð
C
A Þ
expð
B tÞ expð
C tÞ
þ þ ð2:44Þ
ð
A
B Þð
C
B Þ ð
A
C Þð
B
C Þ
Figure 2.8 Time variation of the relative numbers of nuclei in the decay chain (2.45)
50 CH2 NUCLEAR PHENOMENOLOGY
This illustrates the general features that whereas NA ðtÞ for the initial species falls
monotonically with time and NC ðtÞ for the final stable species rises monotonically,
NB ðtÞ for an intermediate species rises to a maximum before falling. Note that at
any time the sum of the components is a constant, as expected.
In the following sections we consider the phenomenology of the various types of
radioactivity in more detail and in Chapter 7 we will return to discuss various
models and theories that provide an understanding of these phenomena.
The first of these is the mass of the constituent nucleons and electrons,
The remaining terms are various corrections, which we will write in the form ai
multiplied by a function of Z and A with ai > 0.
The most important correction is the volume term,
11
Latent heat is the average energy required to disperse the liquid drop into a gas and so is analogous to the
binding energy per nucleon.
SEMI-EMPIRICAL MASS FORMULA: THE LIQUID DROP MODEL 51
This arises from the fact that the strong nuclear force is short-range and each
nucleon therefore feels the effect of only the nucleons immediately surrounding it
(the force is said to be saturated), independent of the size of the nucleus. Recalling
the important result deduced in Section 2.2 that the nuclear radius is proportional
1
to A3 , this leads immediately to the binding energy being proportional to the
volume, or nuclear mass. The coefficient is negative, i.e. it increases the binding
energy, as expected.
The volume term overestimates the effect of the nuclear force because nucleons
at the surface are not surrounded by other nucleons. Thus the volume term has to
be corrected. This is done by the surface term
2
f2 ðZ; AÞ ¼ þa2 A3 ; ð2:49Þ
which is proportional to the surface area and decreases the binding energy. In the
classical model of a real liquid drop, this term would correspond to the surface
tension energy.
The Coulomb term accounts for the Coulomb energy of the charged nucleus, i.e.
the fact that the protons repel each other. If we have a uniform charge distribution
1
of radius proportional to A3 , then this term is
ZðZ 1Þ Z2
f3 ðZ; AÞ ¼ þa3 1 þa3 1 ; ð2:50Þ
A3 A3
where the approximation is sufficiently accurate for the large values of Z we will
be considering. A similar effect would be present for a charged drop of a classical
liquid.
The next term is the asymmetry term.
ðZ A=2Þ2
f4 ðZ; AÞ ¼ þa4 : ð2:51Þ
A
This accounts for the observed tendency for nuclei to have Z ¼ N. (There are no
stable nuclei with very large neutron or proton excesses – c.f. Figure 2.7.) This
term is purely quantum mechanical in origin and is due to the Pauli principle.
Part of the reason for the form (2.51) can be seen from the diagram of Figure 2.9,
which shows the energy levels of a nucleus near the highest filled level in the
approximation where all the energy levels are separated by the same energy .
Keeping A fixed and removing a proton from level 3 and adding a neutron to level
4, gives ðN ZÞ ¼ 2 and leads to an energy increase of . Repeating this for more
protons, we find that the transfer of ðN ZÞ=2 nucleons decreases the binding
energy by an amount ðN ZÞ2 =4. Although we have assumed is a constant,
in practice it decreases like A1 ; hence the final form of the asymmetry term.
If we start with an even number of nucleons and progressively fill states, then the
lowest energy will be when both Z and N are even. If, on the other hand, we have a
52 CH2 NUCLEAR PHENOMENOLOGY
Figure 2.9 Schematic diagram of nuclear energy levels near the highest filled levels
system where both Z and N are odd and the highest filled proton state is above the
highest filled neutron state, we can increase the binding energy by removing one
proton from the nucleus and adding one neutron. If the highest filled proton state is
below the highest filled neutron state, then we can produce the same effect by
removing a neutron and adding a proton. These observations are summarized in the
empirical pairing term, which maximizes the binding when both Z and N are even:
a1 ¼ av ; a2 ¼ as ; a3 ¼ ac ; a4 ¼ aa ; a5 ¼ ap : ð2:53Þ
The fit to the binding energy data for A > 20 using these coefficients in the SEMF
is shown in Figure 2.10. Overall the fit to the data is remarkably good for such a
simple formula, but is not exact of course. For example, there are a small number of
regions where the binding energy curves show enhancements that are not repro-
duced. (These enhancements are due to the existence of a ‘shell structure’ of
nucleons within the nucleus and will be discussed in Chapter 7.) Nevertheless, the
SEMF gives accurate values for the binding energies for some 200 stable and many
12
Note that some authors write the asymmetry term proportional to ðZ NÞ2 , which is equivalent to the form
used here, but their value for the coefficient aa will differ by a factor of four from the one in Equations (2.54).
SEMI-EMPIRICAL MASS FORMULA: THE LIQUID DROP MODEL 53
Figure 2.10 Fit to binding energy data (shown as solid circles) for odd-A and even-A nuclei
using the SEMF with the coefficients given in the text; the predictions are shown as open circles
and do not lie on smooth curves because A is not a function of Z
54 CH2 NUCLEAR PHENOMENOLOGY
more unstable nuclei. We will use it to analyse the stability of nuclei with respect to
-decay and fission. The discussion of -decay is deferred until Chapter 7.
Using the numerical values of Equation (2.54), the relative sizes of each of
the terms in the SEMF may be calculated and for the case of odd-A are shown in
Figure 2.11. In this diagram, the volume term is shown as positive and the other terms
are subtracted from it to give the final SEMF curve.
Figure 2.11 Contributions to the binding energy per nucleon as a function of mass number
for odd-A from each term in the SEMF; the surface, asymmetry and Coulomb terms have been
plotted so that they subtract from the volume term to give the total SEMF result in the lowest
curve
Finally, from its definition, one might expect the binding energy per nucleon to
be equivalent to the energy needed to remove a nucleon from the nucleus.
However, to remove a neutron from a nucleus corresponds to the process
A
ZY ! A1Z Y þ n ð2:55aÞ
A
ZY ! A1
Z1 X þ p; ð2:56aÞ
-DECAY PHENOMENOLOGY 55
Thus, Ep and En are only equal to the binding energy per nucleon in an average
sense. In practice, measurements show that Ep and En can differ substantially from
this average and from each other at certain values of (Z, A). We will see in
Chapter 7 that one reason for this is the existence of a shell structure for nucleons
within nuclei, similar to the shell structure of electrons in atoms, which is ignored
in the liquid drop model.
MðZ; AÞ ¼ A Z þ Z 2 þ 1 ; ð2:57Þ
A2
where
as
aa
¼ Mn av þ 1 þ
A 34
¼ aa þ ðMn Mp me Þ
aa ac ð2:58Þ
¼ þ 1
A A3
¼ ap
Odd-mass nuclei can arise from even-N, odd-Z, or even-Z, odd-N configurations
and in practice the number of nuclei that are stable against -decay are roughly
equally distributed between these two types. The example we take is the case of
56 CH2 NUCLEAR PHENOMENOLOGY
Figure 2.12 Mass parabola of the A ¼ 111 isobars: the circles are experimental data and the
curve is the prediction of the SEMF -- possible -decays are indicated by arrows
the A ¼ 111 isobars, which are shown in Figure 2.12. The circles show the
experimental data as mass excess values in atomic mass units, where
and the atomic mass unit (u) is defined as one twelfth of the mass of the neutral
atom 126 C.
The curve is the theoretical prediction from the SEMF using the numerical
values of the coefficients (2.54). The exact form of the curve depends on the
precise values of these coefficients. The minimum of the parabola corresponds to
the isobar 111
48 Cd with Z ¼ 48.
Isobars with more neutrons, such as 111 111 111
45 Rh, 46 Pd and 47 Ag, decay by converting
a neutron to a proton, i.e.
n ! p þ e þ e ; ð2:60Þ
so that
111
45 Rh ! 111
46 Pd þ e þ
e ð11 sÞ; ð2:61aÞ
111
46 Pd ! 111
47 Ag þ e þ
e ð22:3 minÞ ð2:61bÞ
and
111
47 Ag ! 111
48 Cd þ e þ
e ð7:45 daysÞ ð2:61cÞ
-DECAY PHENOMENOLOGY 57
Recall that we are referring here to atoms, so that the rest mass of the created
electron is automatically taken into account.
Isobars with proton excess decay via
p ! n þ eþ þ e ; ð2:63Þ
i.e. positron emission, which although not possible for a free proton, is possible in a
nucleus because of the binding energy. So for example, the nuclei
111 111 111
51 Sb; 50 Sn and 49 In could, in principle, decay by positron emission, which is
energetically possible if
MðZ; AÞ > MðZ 1; AÞ þ 2me ; ð2:64Þ
this takes account of the creation of a positron and the existence of an excess of
electrons in the parent atom.
It is also theoretically possible for this sequence of transitions to occur by
electron capture. This mainly occurs in heavy nuclei, where the electron orbits are
more compact. It is usually the electron in the innermost shell (i.e. the K-shell) that
is captured. Capture of such an electron gives rise to a ‘hole’ and causes electrons
from higher levels to cascade downwards and in so doing emit characteristic
X-rays. Electron capture is energetically allowed if
MðZ; AÞ > MðZ 1; AÞ þ "; ð2:65Þ
where " is the excitation energy of the atomic shell of the daughter nucleus. The
process competes with positron emission and in practice for the nuclei above this is
what happens. Thus, we have
e þ 111 111
51 Sb ! 50 Sn þ e ð75 sÞ; ð2:66aÞ
e þ 111 111
50 Sn ! 49 In þ e ð35:3 minÞ ð2:66bÞ
and
e þ 111 111
49 In ! 48 Cd þ e ð2:8 daysÞ; ð2:66cÞ
Even-mass nuclei can arise from even-N, even-Z, or odd-Z, odd-N configurations,
but for reasons that are explained below, nearly all even-mass nuclei that are stable
against -decay are of the even–even type, with only a handful of odd–odd types
known. Consider as an example the case of A ¼ 102 shown in Figure 2.13. (Recall
that the plot is of mass excess, which is a very small fraction of the total mass.)
Figure 2.13 Mass parabolas of the A ¼ 102 isobars: the circles are experimental data (open
circles are even--even nuclei and closed circles are odd--odd nuclei); the curves are the
prediction of the SEMF (upper curve is for odd--odd nuclei and lower curve for even--even nuclei)
and possible -decays are indicated by arrows
but this would involve a ‘double electron capture’ and would be heavily suppressed.
The reaction has never been observed. Thus there are two -stable isobars. This is a
common situation for A-even, although no two neighbouring isobars are known to be
stable. Odd–odd nuclei always have at least one more strongly bound even–even
neighbour nucleus in the isobaric spectrum. They are therefore unstable. The only
exceptions to this rule are a few very light nuclei.
The lifetime of a free neutron is about 887 s. The free proton is believed to be
stable and can only ‘decay’ within a nucleus by utilizing the binding energy.
Lifetimes of emitters vary enormously from milliseconds to 1016 years. They
FISSION 59
depend very sensitively on the Q-value for the decay and on the properties of the
nuclei involved, e.g. their spins.
2.7 Fission
Spontaneous fission has been defined as the process whereby a parent nucleus breaks
into two daughter nuclei of approximately equal masses without external action.
Precisely equal masses are very unlikely and in the most probable cases the daughter
nuclei have mass numbers that differ by about 45, with peaks around mass numbers
95 and 140. The reason for this is unknown. The binding energy curve shows that
spontaneous fission is energetically possible for nuclei with A > 100.13 An example is
238
92 U ! 145 90
57 La þ 35 Br þ 3n; ð2:69Þ
with a release of about 154 MeV of energy, which is carried off as kinetic energy of
the fission products. Heavy nuclei are neutron-rich and so necessarily produce
neutron-rich decay products, including free neutrons. The fission products are
themselves usually some way from the line of -stability and will decay by a series
of steps. For example, 145 145
57 La decays to the -stable 60 Nd by three stages, releasing a
further 8.5 MeV of energy, which in this case is carried off by the electrons and
neutrinos emitted in -decay. Although the probability of fission increases with
increasing A, it is still a very rare process. For example, in 238
92 U, the transition rate for
spontaneous fission is about 3 1024 s1 compared with about 5 1018 s1 for
-decay, a branching fraction of 6 107 . Spontaneous emission only becomes
dominant in very heavy elements with A 270, as we shall now show.
To understand spontaneous fission we can again use the liquid drop model. In the
SEMF we have assumed that the drop (i.e. the nucleus) is spherical, because this
minimizes the surface area. However, if the surface is perturbed for some reason from
spherical to prolate, the surface term in the SEMF will increase and the Coulomb term
will decrease (assuming the volume remains the same) and the relative sizes of these
two changes will determine whether the nucleus is stable against spontaneous fission.
For a fixed volume we can parametrize the deformation by the semi-major and
semi-minor axes of the ellipsoid a and b, respectively as shown in Figure 2.14. One
possible parametrization that preserves the volume is
1
a ¼ R ð1 þ "Þ; b ¼ R=ð1 þ "Þ2 ; ð2:70Þ
4 4
V ¼ R3 ¼ ab2 : ð2:71Þ
3 3
13
Fission in heavy nuclei was discovered by Otto Hahn, for which he received the 1944 Nobel Prize in
Chemistry.
60 CH2 NUCLEAR PHENOMENOLOGY
R
a
To find the new surface and Coulomb terms one has to find the expression for the
surface of the ellipsoid in terms of a and b and expand it in a power series in ". The
algebra is unimportant and the results are:
2 2
Es ¼ as A3 1 þ "2 þ . . . ð2:72aÞ
5
and
1 1
Ec ¼ ac Z 2 A3 1 "2 þ . . . : ð2:72bÞ
5
"2 2 1
E ¼ ðEs þ Ec Þ ðEs þ Ec ÞSEMF ¼ 2as A3 ac Z 2 A3 : ð2:73Þ
5
If E < 0, then the deformation is energetically favourable and fission can occur.
From Equation (2.73), this happens if
Z 2 2as
49; ð2:74Þ
A ac
where we have used experimental values for the coefficients as and ac given in
Equations (2.54). The inequality is satisfied for nuclei with Z > 116 and A 270.
Spontaneous fission is a potential barrier problem and this is shown in Figure 2.15.
The solid line corresponds to the shape of the potential in the parent nucleus. The
activation energy shown in Figure 2.15 determines the probability of spontaneous
fission. To fission, the nucleus could in principle tunnel through the barrier, but the
fragments are large and the probability for this to happen is extremely small.14 For
heavy nuclei the activation energy is about 6 MeV, but disappears for very heavy
nuclei. For such nuclei, the shape of the potential corresponds closer to the dashed
line and the slightest deformation will induce fission.
14
The special case of -decay will be discussed in Chapter 7. There we will show that the lifetime for such
decays is expected to have an exponential dependence on the height of the fission barrier and this is observed
qualitatively in fission data.
FISSION 61
Another possibility for fission is to supply the energy needed to overcome the
barrier by a flow of neutrons. Because of the absence of a Coulomb force, a neutron
can get very close to the nucleus and be captured by the strong nuclear attraction.
The parent nucleus may then be excited to a state above the fission barrier and
therefore split up. This process is an example of induced fission. Neutron capture by
a nucleus with an odd neutron number releases not just some binding energy, but also
a pairing energy. This small extra contribution makes a crucial difference to nuclear
fission properties. For example, very low-energy (‘thermal’) neutrons can induce
fission in 235 U, whereas only higher energy (‘fast’) neutrons induce fission in 238 U.
This is because 235 U is an even–odd nucleus and 238 U is even–even. Therefore, the
ground state of 235 U will lie higher (less tightly bound) in the potential well of its
fragments than that of 238 U. Hence to induce fission, a smaller energy will be needed
for 235 U than for 238 U. In principle, fission may be induced in 235 U using even zero-
energy neutrons.15
We consider this quantitatively as follows. The capture of a neutron by 235 U
changes an even–odd nucleus to a more tightly bound even–even (compound)
nucleus of 236 U and releases the binding energy of the last neutron. In 235 U this is
6.5 MeV. As the activation energy (the energy needed to induce fission) is about
5 MeV for 236 U, neutron capture releases sufficient energy to fission the nucleus. The
kinetic energy of the incident neutron is irrelevant and even zero-energy neutrons
can induce fission in 235 U. In contrast, neutron capture in 238 U changes it from an
even–even nucleus to an even–odd nucleus, i.e. changes a tightly bound nucleus to a
less tightly bound one. The energy released (the binding energy of the last neutron) is
about 4.8 MeV in 239 U and is less than the 6.5 MeV required for fission. For this
reason, fast neutrons with energy of at least the difference between these two
energies are required to fission 238 U.
15
Enrico Fermi was a pioneer in the field of induced fission and received the 1938 Nobel Prize in Physics for
‘demonstrations of the existence of new radioactive elements produced by neutron irradiation, and for his
related discovery of nuclear reactions brought about by slow neutrons’. Fermi’s citation could equally have
been about his experimental discoveries and theoretical work in a wide range of areas from nuclear and
particle physics to solid-state physics and astrophysics. He was probably the last ‘universal physicist’.
62 CH2 NUCLEAR PHENOMENOLOGY
2.8 c-Decays
When a heavy nucleus disintegrates by either - or -decay, or by fission, the daughter
nucleus is often left in an excited state. If this state is below the excitation energy for
fission, it will de-excite, usually by emitting a high-energy photon. The energy of these
photons is determined by the average energy level spacings in nuclei and ranges from a
few to several MeV. They are in the gamma ray () part of the electromagnetic
spectrum. Because -decay is an electromagnetic process, we would expect the
typical lifetime of an excited state to be 1016 s. In practice, lifetimes are very
sensitive to the amount of energy released in the decay and in the nuclear case other
factors are also very important, particularly the quantity of angular momentum carried
off by the photon. Typical lifetimes of nuclear levels are about 1012 s.
The role of angular momentum in -decays is crucial. If the initial (excited)
state has a total spin Si and the final nucleus has a total spin Sf , then the total
angular momentum J of the emitted photon is given by
J ¼ Si Sf ; ð2:75Þ
with
Si þ Sf J jSi Sf j; ð2:76Þ
mi ¼ M þ mf ; ð2:77Þ
where m are the corresponding magnetic quantum numbers. Both total angular
momentum and its magnetic quantum number are conserved in -decays.
-decays are further complicated because parity is conserved in these electro-
magnetic processes. Both the initial and final nuclear level will have an intrinsic
parity, as does the photon, and in addition there is a parity associated with the angular
momentum carried off by the photon, which is of the form ð1ÞJ , reflecting the
symmetry of the angular part of the wavefunction (see Equation (1.14)). We will not
pursue this further here, but defer a more detailed discussion until Chapter 7.
p þ 16 O ! d þ 15 O; ð2:78Þ
where we have used the notation Aða; bÞB for the general nuclear reaction
a þ A ! b þ B. This is an example of a pick-up reaction, because one or more
nucleons (in this case a neutron) is stripped off the target nucleus and carried away
by the projectile. The ‘inverse’ of this reaction is 16 Oðd; pÞ17 O. This is an example of
a stripping reaction, because one or more nucleons (in this case again a neutron) is
stripped off the projectile and transferred to the target nucleus.
The theoretical interpretation of direct reactions is based on the assumption that
the projectile experiences the average potential of the target nucleus. For example,
we have seen in the optical model of Section 2.2.2 how this approach can be used to
analyse differential cross sections for elastic scattering and be used to extract
information about nuclear shapes and sizes. It also leads to the prediction of
resonances of width typically of order 1 MeV separated by a few MeV, as observed
in cross-section as functions of centre-of-mass energy for nucleon scattering from
light nuclei. One way of viewing this is as a consequence of the reaction time for a
direct reaction, typically 1022 s , making use of the uncertainty relation between
energy and time, Et h.
A second important class of interactions is where the projectile becomes loosely
bound in the nucleus and shares its energy with all the nuclear constituents. This is
called a compound nucleus reaction. The time for the system to reach statistical
equilibrium depends on the nuclear species, the type of projectile and its energy, but
will always be much longer than the transit time and is typically several orders of
magnitude longer. An important feature of these reactions is that the properties of the
compound nucleus determine its subsequent behaviour and not the mechanism by
which it was formed. The compound nucleus is in an excited state and is inherently
unstable. Eventually, by a statistical fluctuation, one or more nucleons will acquire
sufficient energy to escape and the nucleus either emits particles or de-excites by
radiating gamma rays.
If the compound nucleus is created in a region of excitation where its energy
levels are well separated, the cross-section will exhibit well-defined resonances
64 CH2 NUCLEAR PHENOMENOLOGY
Figure 2.16 Energy-level diagram showing the excitation of a compound nucleus C and its
subsequent decay
Because the time for a compound nucleus to reach statistical equilibrium is much
longer than the transit time for a direct reaction, the cross-sections for a compound
nucleus process can show variations on much smaller energy scales than those for
direct reactions. The density of levels in the compound nucleus is high, and so a very
small change in the incident energy suffices to alter completely the intermediate
states, and hence the cross section. An example is shown in Figure 2.17, which gives
the total cross-section for neutron scattering from 12 C at neutron laboratory energies
of a few MeV. Peaks corresponding to resonance formation in 13 C are clearly
identified. Their widths vary from a few tens to a few hundreds of keV, consistent
with the characteristic times for compound nucleus formation and decay.
Figure 2.17 Total cross-section for n12 C interactions (adapted from Fo61. Copyright American
Physical Society.)
NUCLEAR REACTIONS 65
Figure 2.18 Direct and compound nucleus reactions in nuclear reactions initiated by protons
The general form of the yield NðEÞ of secondary particles at a fixed angle as a
function of the outgoing energy E, i.e. the number of particles with energy E
between E and E þ dE, is shown schematically in Figure 2.19 for the case of an
incident nucleon. At the upper end of the plot (which corresponds to low-incident
nucleon energies) there are a number of distinct peaks due to elastic, inelastic and
transfer reactions. Then as the excitation energy is reduced, the more closely-
spaced energy levels in the final nucleus are not fully resolved because of the
spread in energy of the incident beam and the uncertainty in the experimental
measurements of energy. At the lowest energies there is a broad continuum mainly
due to the decays of compound nuclei formed by the absorption of the projectile
nucleon by the target nucleus. The differential cross-sections for the two processes
will be very different. Direct reactions lead to a cross-section peaked in the
forward direction, falling rapidly with angle and with oscillations, as we have seen
66 CH2 NUCLEAR PHENOMENOLOGY
Figure 2.19 Typical spectrum of energies of the nucleons emitted at a fixed angle in inelastic
nucleon--nucleus reactions
in the case of elastic scattering in Section 2.2 (Figure 2.3). On the other hand, the
contribution from the compound nucleus at low energies where an isolated
compound nucleus is formed is fairly isotropic and symmetric about 90 .
Many medium- and large-A nuclei can capture low-energy ( ð10100Þ eV)
neutrons very readily. The neutron separation energy for the final nucleus is
6 MeV and thus capture leads to a compound nucleus with an excitation energy
above the ground state by this separation energy. Such excitation often occurs in a
region of high density of narrow states that show up as a rich resonance structure in
the corresponding neutron total cross-section. An example is shown in Figure 2.20.
The value of the cross-section at the resonance peaks can be many orders of
Figure 2.20 Total cross-section for neutron interactions with 238 U, showing many very narrow
resonances (with intrinsic widths of order 102 eV) corresponding to excited states of 239 U (from
Ga76, courtesy of Brookhaven National Laboratory)
PROBLEMS 67
magnitude greater than the geometrical cross-section based on the size of the
nucleus. This is because the cross-section is determined dominantly by the area
associated with the wavelength
of the projectile, i.e.
2 , which is very large
because
is large.
Once formed, the compound nucleus can decay to any final state consistent with
the relevant conservation laws. If this includes neutron emission, it will be the
preferred decay. However, for production by very slow (thermal) neutrons with
energies of the order of 0.02 eV, the available decay kinetic energy will reflect the
initial energy of the projectile, which is very small. Therefore, in these cases,
photon emission is often preferred. We shall see in Chapter 8 that the fact that
radiative decay is the dominant mode of decay of compound nuclei formed by
thermal neutrons is important in the use of nuclear fission to produce power in
nuclear reactors.
Problems
2.1 Electrons with momentum 330 MeV/c are elastically scattered through an angle of
10 by a nucleus of 56 Fe. If the charge distribution on the nucleus is assumed to be
that of a uniform hard sphere, and assuming the Born approximation is valid, by
what factor would you expect the Mott cross-section to be reduced?
2.2 Show explicitly that Equation (2.28) follows from Equation (2.26).
2.3 A beam of electrons with energies 250 MeV is scattered through an angle of 10 by a
heavy nucleus. It is found that the differential cross-section is 65 per cent of that
expected from scattering from a point nucleus. Estimate the root mean square radius
of the nucleus.
2.4 Find the form factor for a charge distribution ðrÞ ¼ 0 expðr=aÞ=r, where 0 and a
are constants.
2.5 A sample of 1 g of a radioactive isotope of atomic weight 208 decays via -emission
and 75 counts are recorded in a 24 h period. If the detector efficiency is 10 per cent,
estimate the mean life of the isotope.
2.6 A 1 g sample taken from an organic artefact is found to have a count rate of 2.1
counts per min, which are assumed to originate from the decay of 14 C with a mean
lifetime of 8270 years. If the abundance of 14 C in living matter is currently
1:2 1012 , what can you deduce about the approximate age of the artefact?
2.8 Natural lanthanum has an atomic weight of 138.91 and contains 0.09 per cent of the
isotope 138 138 138
57 La. This has two decay modes: 57 La ! 58 Ce þ e þ e ð -decayÞ and
138 138
57 La þ e ! 56 Ba þ e (electron capture), followed by the electromagnetic
decay of the excited state 138 138
56 Ba ! 56 Ba þ (radiative decay). There are
2
7:8 10 -particles emitted per s per kg of natural lanthanum and there are
50 photons emitted per 100 -particles. Estimate the mean lifetime of 138
57 La.
2.9 Use the SEMF to estimate the energy released in the spontaneous fission reaction
235
92 U ! 87 145
35 Br þ 57 La þ 3n:
2.10 The most stable nucleus with A ¼ 111 is 111 48 Cd (see Figure 2.12). By what
percentage would the fine structure constant have to change if the most stable
nucleus with A ¼ 111 were to be 11147 Ag? Assume that altering does not change
particle masses.
2.15 A radioisotope with decay constant
is produced at a constant rate P. Show that the
number of atoms at time t is NðtÞ ¼ P½1 expð
tÞ=
.
2.17 Consider the total cross-section data for the n238 U interaction shown in Figure 2.20.
There is a resonance R at the centre-of-mass neutron kinetic energy En ¼ 10 eV with
width ¼ 102 eV and the total cross-section there is max ¼ 9 103 b. Use this
information to find the partial widths n; for the decays R ! n þ 238 U and
R ! þ 238 U, if these are the only two significant decay modes. The spin of the
ground state of 238 U is zero.
3
Particle Phenomenology
In this chapter we shall look at some of the basic phenomena of particle physics –
the properties of leptons and quarks, and the bound states of the latter, the hadrons.
In later chapters we will discuss theories and models that attempt to explain these
and other particle data.
3.1 Leptons
We have seen that the spin-12 leptons are one of the three classes of elementary
particles in the standard model and we shall start with a discussion of their basic
properties. Then we shall look in more detail at the neutral leptons, the neutrinos
and, amongst other things, examine an interesting property they can exhibit, based
on simple quantum mechanics, if they have non-zero masses.
There are six known leptons and they occur in pairs, called generations, which we
write, for reasons that will become clear presently, as:
e
; ; : ð3:1Þ
e
Each generation comprises a charged lepton with electric charge e, and a
neutral neutrino. The three charged leptons ðe ; ; Þ are the familiar
electron, together with two heavier particles, the mu lepton (usually called the
muon, or just mu) and the tau lepton (usually called the tauon, or just tau). The
associated neutrinos are called the electron neutrino, mu neutrino, and tau
Ignoring gravity, the charged leptons interact only via electromagnetic and weak
forces, whereas for the neutrinos, only weak interactions have been observed.2
Because of this, neutrinos, which are all believed to have extremely small masses,
can be detected only with considerable difficulty.
The masses and lifetimes of the leptons are listed in Table 3.1. The electron and
the neutrinos are stable, for reasons that will become clear shortly. The muons
decay by the weak interaction processes
þ ! eþ þ e þ and ! e þ e þ ; ð3:3aÞ
Table 3.1 Properties of leptons: all have spin 12 and masses are given units of MeV/c2; the
antiparticles (not shown) have the same masses as their associated particles, but the electric
charges (Q) and lepton numbers (L‘ ; ‘ ¼ e ; ; ) are reversed in sign
with lifetimes ð2:19703 0:00004Þ 106 s. The tau also decays by the weak
interaction, but with a much shorter lifetime ð2:906 0:011Þ 1013 s. (This
illustrates what we have already seen in nuclear physics, that lifetimes depend
sensitively on the energy released in the decay, i.e. the Q-value.) Because it is
heavier than the muon, the tau has sufficient energy to decay to many different
final states, which can include both hadrons and leptons. However, about 35 per cent
1
Leon Lederman, Melvin Schwartz and Jack Steinberger shared the 1988 Nobel Prize in Physics for their use
of neutrino beams and the discovery of the muon neutrino. Martin Perl shared the 1995 Nobel Prize in
Physics for his pioneering work in lepton physics and in particular for the discovery of the tau lepton.
2
Although neutrinos have zero electric charge they could, in principle, have a charge distribution that would
give rise to a magnetic moment (like neutrons) and hence electromagnetic interactions. This would of course
be forbidden in the standard model because the neutrinos are defined to be point-like.
LEPTONS 73
of decays again lead to purely leptonic final states, via reactions which are very
similar to muon decay, for example:
þ ! þ þ þ and ! e þ e þ : ð3:3bÞ
Le
Nðe Þ Nðeþ Þ þ Nðe Þ Nð
e Þ; ð3:4Þ
where Nðe Þ is the number of electrons present, Nðeþ Þ is the number of positrons
present and so on. For single-particle states, Le ¼ 1 for e and e , Le ¼ 1 for eþ
and e , and Le ¼ 0 for all other particles. The muon and tauon numbers are defined
in a similar way and their values for all single particle states are summarized in
Table 3.1. They are zero for all particles other than leptons. For multiparticle
states, the lepton numbers of the individual particles are simply added. For
example, the final state in neutron -decay (i.e. n ! p þ e þ e ) has
Le ¼ Le ðpÞ þ Le ðe Þ þ Le ð
e Þ ¼ ð0Þ þ ð1Þ þ ð1Þ ¼ 0; ð3:5Þ
eþ þ e ! þ þ ð3:6Þ
In weak interactions more general possibilities are allowed, which still conserve
lepton numbers. For example, in the tau-decay process ! e þ e þ , a tau
converts to a tau neutrino and an electron is created together with an antineutrino,
rather than a positron. The dominant Feynman graph corresponding to this process
is shown in Figure 3.2.
Lepton number conservation, like electric charge conservation, plays an impor-
tant role in understanding reactions involving leptons. Observed reactions conserve
lepton numbers, while reactions that do not conserve lepton numbers are ‘forbidden’
and are not observed. For example, the neutrino scattering reaction
þ n ! þ p ð3:7Þ
which violates both Le and L conservation, is not. Another example which violates
both Le and L conservation is ! e þ . If this reaction were allowed, the
dominant decay of the muon would be electromagnetic and the muon lifetime would
be much shorter than its observed value. This is very strong evidence that even if
lepton numbers are not absolutely conserved, they are conserved to a high degree of
accuracy.
Finally, conservation laws explain the stability of the electron and the neutrinos.
The electron is stable because electric charge is conserved in all interactions and
the electron is the lightest charged particle. Hence decays to lighter particles that
satisfy all other conservation laws, like e ! e þ , are necessarily forbidden by
electric charge conservation. In the same way, lepton number conservation implies
that the lightest particles with non-zero values of the three lepton numbers – the
three neutrinos – are stable, whether they have zero masses or not. Of course, if
lepton numbers are not conserved, then the latter argument is invalid.
3.1.2 Neutrinos
nuclear -decays
ðZ ; NÞ ! ðZ þ 1; N 1Þ þ e þ e ð3:9Þ
and
ðZ; NÞ ! ðZ 1; N þ 1Þ þ eþ þ e ð3:10Þ
that were discussed in Chapter 2. The neutrinos and antineutrinos emitted in these
decays are not observed experimentally, but are inferred from energy and angular
momentum conservation. In the case of energy, if the antineutrino were not present
in the first of the reactions, the energy Ee of the emitted electron would be a unique
value equal to the difference in rest energies of the two nuclei, i.e.
where for simplicity we have neglected the extremely small kinetic energy of the
recoiling nucleus. However, if the antineutrino is present, the electron energy
would not be unique, but would lie in the range
depending on how much of the kinetic energy released in the decay is carried away
by the neutrino. Experimentally, the observed energies span the whole of the above
range and in principle a measurement of the energy of the electron near its
maximum value of Ee ¼ ðM me Þc2 determines the neutrino mass. The most
accurate results come from tritium decay and are compatible with zero mass for
electron antineutrinos. When experimental errors are taken into account, the
experimentally allowed range is
could not occur if the neutrinos had exactly zero mass. The consequences of
neutrinos having small masses have therefore to be taken seriously.
Because neutrinos only have weak interactions, they can only be detected with
extreme difficulty. For example, electron neutrinos and antineutrinos of sufficient
energy can in principle be detected by observing the inverse -decay processes
e þ n ! e þ p ð3:14Þ
and
e þ p ! eþ þ n : ð3:15Þ
However, the probability for these and other processes to occur is extremely small.
In particular, the neutrinos and antineutrinos emitted in -decays, with energies of
the order of 1 MeV, have mean free paths in matter of the order of 106 km.3
Nevertheless, if the neutrino flux is intense enough and the detector is large
enough, the reactions can be observed. In particular, uranium fission fragments are
neutron rich, and decay by electron emission to give an antineutrino flux that can
be of the order of 1017 m2 s1 or more in the vicinity of a nuclear reactor, which
derives its energy from the decay of nuclei. These antineutrinos will occasionally
interact with protons in a large detector, enabling examples of the inverse -decay
reaction to be observed. As mentioned in Chapter 1 (Footnote 12), electron
neutrinos were first detected in this way by Reines and Cowan in 1956, and
their interactions have been studied in considerable detail since.
The mu neutrino, , has been detected using the reaction þ n ! þ p and
other reactions. In this case, well-defined high-energy beams can be created in the
laboratory by exploiting the decay properties of pions, which are particles we have
mentioned briefly in Chapter 1 and which we will meet in more detail presently. The
probability of neutrinos interacting with matter increases rapidly with energy (this
will be demonstrated in Section 6.5.2) and, for large detectors, events initiated by
such beams are so copious that they have become an indispensable tool in studying
both the fundamental properties of weak interactions and the internal structure of the
proton. Finally, in 2000, a few examples of tau neutrinos were reported, so that
almost 70 years after Pauli first suggested the existence of a neutrino, all three types
had been directly detected.
Neutrinos are assumed to have zero mass in the standard model. However, as
mentioned above, data from the -decay of tritium are compatible with a non-zero
3
The mean free path is the distance a particle would have to travel in a medium for there to be a significant
probability of an interaction. A formal definition is given in Chapter 4.
LEPTONS 77
mass. A phenomenon that can occur if neutrinos have non-zero masses is neutrino
mixing. This arises if we assume that the observed neutrino states e ; and
which take part in weak interactions, i.e. the states that couple to electrons, muons
and tauons, respectively, are not eigenstates of mass, but instead are linear
combinations of three other states 1 ; 2 and 3 which do have definite masses
m1 ; m2 and m3 , i.e. are eigenstates of mass. For algebraic simplicity we will
consider the case of mixing between just two states, one of which we will assume
is and the other we will denote by x. Then, in order to preserve the orthonormality
of the states, we can write
and
where
are the usual oscillating energy factors associated with any quantum mechanical
stationary state.4 For t 6¼ 0, the linear combination (3.19) does not correspond to a
pure muon neutrino state, but can be written as a linear combination
AðtÞ ; p þ BðtÞjx ; pi; ð3:21Þ
The functions AðtÞ and BðtÞ are found by solving Equations (3.18) and (3.22) for
j1 ; pi and j2 ; pi, then substituting the results into (3.19) and comparing at with
(3.21). This gives,
and
and thus oscillates with time, while the probability of finding a muon neutrino is
reduced by a corresponding oscillating factor. Similar effects are predicted if
instead we start from electron or tau neutrinos. In each case the oscillations vanish
if the mixing angle is zero, or if the neutrinos have equal masses, and hence equal
energies, as can be seen explicitly from Equation (3.25). In particular, such
oscillations are not possible if the neutrinos both have zero masses.
Returning to Equation (3.25), since neutrino masses are very small,
E1;2 mi c2 ði ¼ 1; 2Þ and we can write
4
See, for example, Chapter 1 of Ma92.
LEPTONS 79
with
where ðm2 c4 Þ
m22 c4 m21 c4 . These formulae assume that the neutrinos are
propagating in a vacuum, whereas in real experiments they will be passing through
matter and the situation is more complicated than these simple results suggest.
This formalism can be extended to the general case of mixing between all three
neutrino species, but at the expense of additional free parameters.5
Attempts to establish neutrino oscillations rest on using the inverse -decay
reactions (3.14) and (3.15) to produce electrons and the analogous reactions for
muon neutrinos to produce muons, which are then detected. In addition, the time t
is determined by the distance of the neutrino detector from the source of the
neutrinos, since their energies are always much greater than their possible masses,
and they travel at approximately the speed of light. Hence, for example, if we start
with a source of muon neutrinos, the flux of muons observed in a detector should
vary with its distance from the source of the neutrinos, if appreciable oscillations
occur. In practice, oscillations at the few per cent level are very difficult to detect
for experimental reasons that we will not discuss here.
There are a number of different types of experiment that can explore neutrino
oscillations and hence neutrino masses. The first of these to produce definitive
evidence for oscillations was that of a Japanese group in 1998 using the giant
Super Kamiokande detector to study atmospheric neutrinos produced by the action
of cosmic rays.6 (Neutrinos of each generation are often referred to as having a
different flavour and so the observations are evidence for flavour oscillation.)
The Super Kamiokande detector is shown in Figure 3.3. (Detectors will be
discussed in detail in Chapter 4, so the description here will be brief.) It consists of
a stainless steel cylindrical tank of roughly 40 m diameter and 40 m height,
containing about 50 000 metric tons of very pure water. The detector is situated
deep underground in a mountain in Japan, at a depth equivalent to 2700 m of water.
This is to use the rocks above to shield the detector from cosmic ray muons. The
volume is separated into a large inner region, the walls of which are lined with
11 200 light-sensitive devices called photomultipliers (the physics of these will be
discussed in Chapter 4). These register the presence of electrons or muons
5
See, for example, the Review of Particle Properties published biannually by the Particle Data Group (2004
edition: Ei04). The PDG Review is also available at http://pdg.lbl.gov. This publication contains a wealth of
useful data about elementary particles and their interactions and we will refer to it in future simply as
PDG04.
6
Cosmic neutrinos were first detected (independently) by Raymond Davis Jr. and Masatoshi Koshiba, for
which they were jointly awarded the 2002 Nobel Prize in Physics.
80 CH3 PARTICLE PHENOMENOLOGY
Figure 3.3 A schematic diagram of the Super Kamiokande detector (adapted from an original
University of Hawaii, Manoa, illustration -- with permission)
indirectly by detecting the light (the so-called Čerenkov radiation – again, see
Chapter 4) emitted by relativistic charged particles (the electrons or muons) that
are created in, or pass through, the water. The outer region of water acts as a shield
against low-energy particles entering the detector from outside. An additional 1200
photomultipliers are located there to detect muons that enter or exit the detector.
When cosmic ray protons collide with atoms in the upper atmosphere they create
many pions, which in turn create neutrinos mainly by the decay sequences
! þ ; þ ! þ þ ð3:28Þ
and
! e þ e þ ; þ ! eþ þ e þ : ð3:29Þ
From this, one would naively expect to detect two muon neutrinos for every
electron neutrino. However, the ratio was observed to be about 1.3 to 1 on average,
suggesting that the muon neutrinos produced might be oscillating into other species.
Clear confirmation for this was found by exploiting the fact that the detector
could measure the direction of the detected neutrinos to study the azimuthal
dependence of the effect. Since the flux of cosmic rays that lead to neutrinos with
energies above about 1 GeV is isotropic, the production rate for neutrinos should
LEPTONS 81
be the same all around the Earth. In particular, one can compare the measured flux
from neutrinos produced in the atmosphere directly above the detector, which have
a short flight path before detection, with those incident from directly below, which
have travelled a long way through the Earth before detection, and so have had
plenty of time to oscillate (perhaps several cycles). Experimentally, it was found
that the yield of electron neutrinos from above and below were the same within
errors and consistent with the expectation for no oscillations. However, while
the yield of muon neutrinos from above accorded with the expectation for no
significant oscillations, the flux of muon neutrinos from below was a factor of
about two lower. This is clear evidence for muon neutrino oscillations.
In a later development of the experiment, the flux of muon neutrinos was
measured as a function of L=E by estimating L from the reconstructed neutrino
direction. Values of L range from 15 km to 13000 km. The results are shown in
Figure 3.4 in the form of the ratio of observed number of events to the theoretical
expectation if there were no oscillations. The data show clear evidence for a
deviation of this ratio from unity, particularly at large values of L=E.
Figure 3.4 Data from the Super Kamiokande detector showing evidence for neutrino
oscillations in atmospheric neutrinos (adapted from As04, copyright American Physical Society)
Other experiments also set limits on Pð ! e Þ and taking these into account
the most plausible hypothesis is that muon neutrinos are changing into tau neutrinos,
which for the neutrino energies concerned could not be detected by Super
Kamiokande. The data are consistent with this hypothesis and yield the values
1:9 103
ðm2 c4 Þ
3:0 103 ðeVÞ2 ; sin2 ð2Þ > 0:9 ð3:30Þ
82 CH3 PARTICLE PHENOMENOLOGY
e þ 37 Cl ! 37 Ar þ e ; ð3:31Þ
8
B ! 8 Be þ eþ þ e ; ð3:32Þ
where the neutrinos have an average energy 7 MeV. More recent experiments
have studied the same process using the reactions
to detect the neutrinos, where d is a deuteron. The first of these reactions clearly
can be initiated with electron neutrinos only, whereas the other two can be initiated
with neutrinos of any flavour. The measured flux of e from reaction (a) agrees
well with the standard solar model, but the ratio of the flux for e to that for x,
where x could be a combination of and , obtained by using data from all three
reactions, is less than unity. For example, the Sudbury Neutrino Observatory
(SNO) experiment finds a ratio of about 0.3. Thus there is a flux of neutrinos of a
type that did not come from the original decay process. The observations are
further clear evidence for flavour oscillation.
Although the neutrinos from (3.32) have been extensively studied, this decay
contributes only about 104 of the total solar neutrino flux. It is therefore important
7
See, for example, Chapter 4 of Ph94.
LEPTONS 83
to detect neutrinos from other reactions and in particular from the reaction
p þ p ! d þ eþ þ e ; ð3:34Þ
which is the primary reaction that produces the energy of the Sun and contributes
approximately 90 per cent of the solar neutrino flux. (It will be discussed in more
detail in Chapter 8.) The neutrinos in this reaction have average energies of
0:26 MeV and so cannot be detected by reaction (3.31). Instead, the reaction
e þ 71 Ga ! 71 Ge þ e ð3:35Þ
has been used, which has a threshold of 233 keV. (The experiments can also detect
neutrinos from the solar reaction e þ 7 Be ! 7 Li þ e .) Just as for the original
experiments of Davis et al., there are formidable problems in identifying the
radioactive products from this reaction, which produces only about 1 atom of 71 Ge
per day in a target of 30 tons of gallium. Nevertheless, results from these
experiments confirm the deficit of electron neutrinos and find between 60 and
70 per cent of the flux expected from the standard solar model without flavour
changing.
These solar neutrino results require that interactions with matter play a
significant role in flavour changing and imply, for example, that a substantial
fraction of a beam of e would change to antineutrinos of other flavours after
travelling a distance of the order of 100 km from its source. This prediction has
been tested by the KamLAND group in Japan. They have studied the e flux from
more than 60 reactors in Japan and South Korea after the neutrinos have travelled
distances of between 150 and 200 km. They found that the e flux was only about
60 per cent of that expected from the known characteristics of the reactors. A
simultaneous analysis of the data from this experiment and the solar neutrino data
yields the result:
7:6 105 ðm2 c4 Þ 8:8 105 ðeVÞ2 ; 0:32 tan2 ðÞ 0:48: ð3:36Þ
8
One possibility will be mentioned briefly in Chapter 9 as part of a discussion of the general question of how
masses arise in the standard model.
84 CH3 PARTICLE PHENOMENOLOGY
experiments – although producing compatible values for the mixing angle 40 –
yield wildly different values for the mass difference, as can be seen from Equations
(3.30) and (3.36). However, the analyses have been made in the framework of a two-
component mixing model, whereas there are of course three neutrinos. Thus it could
be, for example, that two of the neutrino states are separated by a small mass
difference given by Equation (3.36) and the third is separated from them by a
relatively large mass difference given by Equation (3.30). Progress will have to await
experiments currently being planned to detect oscillations directly using prepared
neutrino beams and which will make measurements at great distances from their
origin. These experiments are expected to produce data in the next few years and
should yield definitive values of the neutrino mass differences and the various
mixing angles involved.9
The consequences for lepton number conservation are unclear. In the simple
mixing model above, the total lepton number could still be conserved, but
individual lepton numbers would not. However, there are other theoretical
descriptions of neutrino oscillations and this is an open question. A definitive
answer would be to detect neutrinoless double -decay, such as
76
Ge ! 76 Se þ 2e ; ð3:37Þ
where the final state contains two electrons, but no antineutrinos. This could occur
if the neutrino emitted by the parent nucleus was internally absorbed by the
daughter nucleus (i.e. it never appears as a real particle) which is possible only if
e e . A very recent experiment claims to have detected this decay, but the result
is not universally accepted and at present ‘the jury is out’. Experiments planned for
the next few years should settle important questions about lepton number
conservation and the nature of neutrinos.
The three neutrinos have similar properties, but the three charged leptons are
strikingly different. For example: the mass of the muon is roughly 200 times
greater than that of the electron and consequently its magnetic moment is 200
times smaller; high-energy electrons are stopped by modest thicknesses of a
centimetre or so of lead, while muons are the most penetrating form of radiation
known, apart from neutrinos; and the tauon lifetime is many orders of magnitude
smaller than the muon lifetime, while the electron is stable. It is therefore a
remarkable fact that all experimental data are consistent with the assumption that
the interactions of the electron and its associated neutrino are identical to those of
the muon and its associated neutrino and of the tauon and its neutrino, provided the
9
For a review of these experiments see, for example, http://www.hep.anl.gov/ndk/hypertext/nuindustry.html.
LEPTONS 85
mass differences are taken into account. This property, called lepton universality,
can be verified with great precision, because we have a precise theory of
electromagnetic and weak interactions (to be discussed in Chapter 6), which
enables predictions to be made of the mass dependence of all observables.
For example, when we discuss experimental methods in Chapter 4, we will show
that the radiation length, which is a measure of how far a charged particle travels
through matter before losing a certain fraction of its energy by radiation, is
proportional to the squared mass of the radiating particle. Hence it is about 4 104
times greater for muons than for electrons, explaining their much greater
penetrating power in matter. As another example, we have seen that the rates
for weak -decays are extremely sensitive to the kinetic energy released in the
decay (recall the enormous variation in the lifetimes of nuclei decaying via -
decay). From dimensional arguments and the fact that they are weak interactions,
the rates for muon and tau leptonic decays are predicted to be proportional to the
fifth power of the relevant Q-values multiplied by G2F , the square of the Fermi
coupling.10 Thus, from universality, the ratio of the decay rates is given
approximately by
5
ð ! e þ e þ Þ Q
¼ 1:37 106 : ð3:38Þ
ð ! e þ e þ Þ Q
This is in excellent agreement with the experimental value of 1:35 106 (and is
even closer in a full calculation) and accounts very well for the huge difference
between the tau and muon lifetimes. The above are just some of the most striking
manifestations of the universality of lepton interactions.
A question that arises naturally is whether there are more generations of leptons,
with identical interactions, waiting to be discovered. This question has been
answered, under reasonable assumptions, by an experimental study of the decays
of the Z 0 boson. This particle, one of the two gauge bosons associated with the
weak interaction, has a mass of 91 GeV/c2. It decays, among other final states, to
neutrino pairs
Z 0 ! ‘ þ ‘ ð‘ ¼ e; ; Þ: ð3:39Þ
If we assume universal lepton interactions and neutrino masses which are small
compared with the mass of the Z 0 ,11 the decay rates to a given neutrino pair will all
be equal and thus
neutrinos
e þ þ þ ¼ N ; ð3:40Þ
10
The increase of the decay rate as the fifth power of Q is known as Sargent’s Rule.
11
More precisely, we assume m
MZ =2, so that the decays Z ! are not forbidden by energy
conservation.
86 CH3 PARTICLE PHENOMENOLOGY
where N is the number of neutrino species and is the decay rate to any given
pair of neutrinos. The measured total decay rate may then be written
total ¼ hadrons þ leptons þ neutrinos ; ð3:41Þ
where the first two terms on the right are the measured decay rates to hadrons and
charged leptons, respectively. Although the rate to neutrinos is not directly
measured, it can be calculated in the standard model and combining this with
experimental data for the other decay modes, a value of N may be found. The best
value using all available data is N ¼ 3:00 0:08, which is consistent with the
expectation for three neutrino species, but not four. The conclusion is that only
three generations (flavours) of leptons can exist, if we assume universal lepton
interactions and exclude very large neutrino masses.
Why there are just three generations of leptons remains a mystery, particularly
as the extra two generations seem to tell us nothing fundamental that cannot be
deduced from the interaction of the first generation.
3.2 Quarks
We turn now to the strongly interacting particles – the quarks and their bound states,
the hadrons. These also interact by the weak and electromagnetic interactions,
although such effects can often be neglected compared with the strong interactions.
To this extent we are entering the realm of ‘strong interaction physics’.
Several hundred hadrons (not including nuclei) have been observed since pions
were first produced in the laboratory in the early 1950s and all have zero or integer
electric charges: 0; 1; or 2 in units of e. They are all bound states of the
fundamental spin-12 quarks, whose electric charges are either þ 23 or 13, and/or
antiquarks, with charges 23 or þ 13. The quarks themselves have never been
directly observed as single, free particles and, as remarked earlier, this fact initially
made it difficult for quarks to be accepted as anything other than convenient
mathematical quantities for performing calculations. Only later, when the funda-
mental reason for this was realized (it will be discussed in Chapter 6), were quarks
universally accepted as physical entities. Nevertheless, there is compelling
experimental evidence for their existence. The evidence comes from three main
areas: hadron spectroscopy, lepton scattering and jet production.
Hadron spectroscopy
This is the study of the static properties of hadrons: their masses, lifetimes and
decay modes, and especially the values of their quantum numbers, including spin,
QUARKS 87
electric charge and several more that we define in Section 3.2.2. As mentioned in
Chapter 1, the existence and properties of quarks were first inferred from hadron
spectroscopy by Gell-Mann and independently by Zweig in 1964 and the close
correspondence between the experimentally observed hadrons and those predicted
by the quark model, which we will examine in more detail later, remains one of the
strongest reasons for our belief in the existence of quarks.
Lepton scattering
It was mentioned in earlier chapters that in the early 1960s experiments were first
performed where electrons were scattered from protons and neutrons. These
strongly suggested that nucleons were not elementary. By the late 1960s this work
had been extended to higher energies and with projectiles that included muons and
neutrinos. In much the same way as Rutherford deduced the existence of the nucleus
in atoms, high-energy lepton scattering, particularly at large momentum transfers,
revealed the existence of point-like entities within the nucleons, which we now
identify as quarks.
Jet production
High-energy collisions can cause the quarks within hadrons, or newly created
quark–antiquark pairs, to fly apart from each other with very high energies. Before
they can be observed, these quarks are converted into ‘jets’ of hadrons (a process
referred to as fragmentation) whose production rates and angular distributions
reflect those of the quarks from which they originated. They were first clearly
identified in experiments at the DESY laboratory in Hamburg in 1979, where
electrons and positrons were arranged to collide ‘head-on’ in a magnetic field. An
example of a ‘two-jet’ event is shown in Figure 3.5. The picture is a computer
reconstruction of an end view along the beam direction; the solid lines indicate the
reconstructed charged particle trajectories taking into account the known magnetic
field, which is also parallel to the beam direction; the dotted lines indicate the
reconstructed trajectories of neutral particles, which were detected outside this
device by other means.
The production rate and angular distribution of the observed jets closely matches
that of quarks produced in the reaction
eþ þ e ! q þ q; ð3:42Þ
by the mechanism of Figure 3.6. Such jets have now been observed in many
reactions, and are strong evidence for the existence of quarks within hadrons.
88 CH3 PARTICLE PHENOMENOLOGY
The failure to detect free quarks is not an experimental problem. Firstly, free
quarks would be easily distinguished from other particles by their fractional
charges and their resulting ionization properties.12 Secondly, electric charge
conservation implies that a fractionally charged particle cannot decay to a final
state composed entirely of particles with integer electric charges. Hence the
lightest fractionally charged particle, i.e. the lightest free quark, would be stable
and so presumably easy to observe. Finally, some of the quarks are not very
massive (see below) and because they interact by the strong interaction, one would
expect free quarks to be copiously produced in, for example, high-energy proton–
proton collisions. However, despite careful and exhaustive searches in ordinary
matter, in cosmic rays and in high-energy collision products, free quarks have
12
We will see in Chapter 4 that energy losses in matter due to ionization are proportional to the square of the
charge and thus would be ‘anomalously’ small for quarks.
QUARKS 89
never been observed. The conclusion – that quarks exist solely within hadrons and
not as isolated free particles – is called confinement. It is for this reason that we are
forced to study the properties of hadrons, the bound states of quarks.
The modern theory of strong interactions, called quantum chromodynamics
(QCD), which is discussed in Chapter 5, offers at least a qualitative account of
confinement, although much of the detail eludes us due to the difficulty of
performing accurate calculations. In what follows, we shall assume confinement
and use the properties of quarks to interpret the properties of hadrons.
Six distinct types, or flavours, of spin-12 quarks are now known to exist. Like the
leptons, they occur in pairs, or generations, denoted
u c t
; ; : ð3:43Þ
d s b
Each generation consists of a quark with charge þ 23 (u, c, or t) together with a quark
of charge 13 (d, s, or b), in units of e. They are called the down (d), up (u), strange
(s), charmed (c), bottom (b) and top (t) quarks. The quantum numbers associated
with the s, c, b and t quarks are called strangeness, charm, beauty and truth,
respectively. The antiquarks are denoted
d s b
; ; ð3:44Þ
u c t
Table 3.2 Properties of quarks: all have spin 12 and masses are given units of GeV/c2; the
antiparticles (not shown) have the same masses as their associated particles, but the electric
charges (Q) are reversed in sign (in the major decay modes, X denotes other particles)
Name Symbol Mass Q Lifetime (s) Major decays
Down d md 0:3 1=3
Up u mu md 2=3
Strange s ms 0:5 1=3 108 –1010 s!uþX
Charmed c mc 1:5 2=3 1012 –1013 c!s þX
c!d þX
Bottom b mb 4:5 1=3 1012 –1013 b!c þX
Top t mt ¼ 180 12 2=3 1025 t !b þX
90 CH3 PARTICLE PHENOMENOLOGY
bound states, together with models of quark binding.13 In this context they are also
referred to as constituent quark masses.
The stability of quarks in hadrons – like the stability of protons and neutrons in
atomic nuclei – is influenced by their interaction energies. However, for the s; c
and b quarks these effects are small enough for them to be assigned approximate
lifetimes of 108 –1010 s for the s quark and 1012 –1013 s for both the c and b
quarks. The top quark is much heavier than the other quarks and its lifetime is of
the order of 1025 s. This lifetime is so short that when top quarks are created they
decay too quickly to form observable hadrons. In contrast to the other quarks, our
knowledge of the top quark is based entirely on observations of its decay products.
When we talk about ‘the decay of quarks’ we always mean that the decay takes
place within a hadron, with the other bound quarks acting as ‘spectators’, i.e. not
taking part in the interaction. Thus, for example, in this picture neutron decay at
the quark level is given by the Feynman diagram of Figure 3.7 and no free quarks
are observed. Note that it is assumed that the exchanged particle interacts with
only one constituent quark in the nucleons. This is the essence of the spectator
model. (This is not dissimilar to the idea of a single nucleon decaying within a
radioactive nucleus.)
Figure 3.7 Spectator model quark Feynman diagram for the decay n ! pe e
Nf
Nð f Þ Nð f Þ ð f ¼ u; d; s; c; b; tÞ ð3:45Þ
where Nðf Þ is the number of quarks of flavour f present and Nðf Þ is the number of
antiquarks of flavour f present. For example, for single-particle states; Nc ¼ 1 for the
13
An analogy would be to deduce the mass of nucleons from the masses of nuclei via a model of the nucleus.
14
Again, these reactions and associated Feynman diagrams do not imply that free quarks are created.
Spectator quarks are implicitly present to form hadrons in the final state.
QUARKS 91
c quark; Nc ¼ 1 for the c antiquark; and Nc ¼ 0 for all other particles. Similar
results apply for the other quark numbers Nf , and for multi-particle states the quark
numbers of the individual particles are simply added. Thus a state containing the
particles u, u, d, has Nu ¼ 2, Nd ¼ 1 and Nf ¼ 0 for the other quark numbers with
f ¼ s, c, b, t.
In weak interactions, more general possibilities are allowed, and only the total
quark number
Nq
NðqÞ NðqÞ ð3:46Þ
is conserved, where NðqÞ and NðqÞ are the total number of quarks and antiquarks
present, irrespective of their flavour. This is illustrated by the decay modes of the
quarks themselves, some of which are listed in Table 3.2, which are all weak inter-
action processes, and we have seen it also in the decay of the neutron in Figure 3.7.
Another example is the main decay mode of the charmed quark, which is
c ! s þ u þ d; ð3:47Þ
Like the electric charge and the lepton numbers introduced in the last section, the
baryon number is conserved in all known interactions, and unlike the lepton
number, there are no experiments that suggest otherwise.15
15
However, there are theories beyond the standard model that predict baryon number non-conservation,
although there is no experimental evidence to support this prediction. These will be discussed briefly in
Chapter 9.
92 CH3 PARTICLE PHENOMENOLOGY
3.3 Hadrons
In principle, the properties of atoms and nuclei can be explained in terms of their
proton, neutron and electron constituents, although in practice many details are too
complicated to be accurately calculated. However, the properties of these con-
stituents can be determined without reference to atoms and nuclei by studying
them directly as free particles in the laboratory. In this sense atomic and nuclear
physics are no longer fundamental, although they are still very interesting and
important if we want to understand the world we live in.
In the case of hadrons, the situation is more complicated. Their properties are
explained in terms of a few fundamental quark constituents; but the properties of
the quarks themselves can only be studied experimentally by appropriate measure-
ments on hadrons. Whether we like it or not, studying quarks without hadrons is
not an option.
difference between the u and d quarks and also the electromagnetic interactions,
which is equivalent to setting all electric charges to zero, so that the forces acting on
the u and d quarks are exactly equal, then replacing the u quark by a d quark in the
proton would produce a ‘neutron’ which would be essentially identical to the proton.
Of course the symmetry is not exact because of the small mass difference between
the u and d quarks and because of the electromagnetic forces, and it is these that give
rise to the small differences in mass within multiplets.
Flavour independence of the strong forces between u and d quarks also leads
directly to the charge independence of nuclear forces, e.g. the equality of the force
between any pair of nucleons, provided the two particles are in the same spin state.
Subsumed in the idea of charge independence is the idea of charge symmetry, i.e.
the equality of the proton–proton and neutron–neutron forces, again provided the
two particles are in the same spin state. Evidence for the latter is found in studies
of nuclei with the same value of A, but the values of N and Z interchanged (mirror
nuclei). An example is shown in Figure 3.9. The two nuclei 115 B and 116 C have the
same number of np pairs, but 115 B has 10 pp pairs and 15 nn pairs, whereas 116 C has
15 pp pairs and 10 nn pairs. Thus, allowing for the Coulomb interaction, the
approximate equality of the level structures of these two nuclei, as seen in
Figure 3.9, means charge symmetry is approximately verified. To test charge
independence in a nuclear context we would have to look at the level structure in
three related nuclei such as 114 Be, 115 B and 116 C.
Here the test is not so clear-cut because an np pair is not subject to the
restrictions of the Pauli principle like pp and nn pairs and there is evidence (to be
discussed briefly in Chapter 7) that the np force is stronger in the S ¼ 1 state than
in the S ¼ 0 state. Nevertheless, the measured energy levels in such triplets of
nuclei support the idea of approximate charge independence of nuclear forces.
The symmetry between u and d quarks is called isospin symmetry and greatly
simplifies the interpretation of hadron physics. It is described by the same mathe-
matics as ordinary spin, hence the name. For example, the proton and neutron are
viewed as the ‘up’ and ‘down’ components of a single particle, the nucleon N, that
has an isospin quantum number I ¼ 12, with I3 values 12 and 12, assigned to the proton
and neutron, where I3 is analogous to the magnetic quantum number in the case of
ordinary spin. Likewise, the three pions þ ; and 0 are part of a triplet with
I ¼ 1 corresponding to I3 values 1, 0 and 1, respectively. In discussing the strong
interactions between pions and nucleons, it is then only necessary to consider the N
interaction with total isospin either 12 or 32.
As an example, we will consider some predictions for the hadronic resonance
ð1232Þ. The ð1232Þ has I ¼ 32 and four charge states þþ ; þ ; 0 and (see
Table 3.3) corresponding to I3 ¼ 32; 12; 12; 32, respectively. If we use the notation
jN; I; I3 i for a N state, then N; 32; 32 is the unique state þ p and may be written
N; 3; 3 ¼ ; 1; 1 N; 1; 1 : ð3:50Þ
2 2 2 2
94 CH3 PARTICLE PHENOMENOLOGY
Figure 3.9 Low-lying energy levels with spin-parity JP of the mirror nuclei 11
5B and 116 C. (data
from Aj90)
The other N states may then be obtained by applying quantum mechanical shift
(ladder) operators to Equation (3.50), as is done when constructing ordinary spin
states. This gives16 rffiffiffi rffiffiffi
N; 3; 1 ¼ 1þ n þ 20 p ð3:51Þ
2 2 3 3
and hence isospin invariance predicts
ðþ ! þ nÞ 1
¼ ; ð3:52Þ
ðþ ! 0 pÞ 2
which is in good agreement with experiment.
16
The reason for the minus sign and other details are given in, for example, Appendix D of Ma97.
HADRONS 95
rffiffiffi
p ¼ p1ffiffiffiN; 3; 1 2N; 1; 1 ð3:53aÞ
3 2 2 3 2 2
and
rffiffiffi
0
n ¼ 2N; 3; 1 þ p1ffiffiffiN; 1; 1 : ð3:53bÞ
3 2 2 3 2 2
1 2
Mð p ! pÞ ¼ M3 þ M1 ð3:54aÞ
3 3
and
pffiffiffi pffiffiffi
0 2 2
Mð p ! nÞ ¼ M3 M1 : ð3:54bÞ
3 3
At the ð1232Þ, the available energy is such that the total cross-section is
dominated by the elastic ( p ! p) and charge-exchange ( p ! 0 n) reac-
tions. In addition, because the ð1232Þ has I ¼ 32, M3 M1 , so
1
total ð pÞ ¼ ð p ! pÞ þ ð p ! 0 nÞ / jM3 j2 ð3:55aÞ
3
and
Thus, neglecting small kinematic corrections due to mass differences (phase space
corrections), isospin symmetry predicts
total ðþ pÞ
¼ 3: ð3:56Þ
total ð pÞ
Figure 3.10 shows the two total cross-sections at low energies. There are clear
peaks with Breit–Wigner forms at a mass of 1232 MeV corresponding to
the production of the ð1232Þ and the ratio of the peaks is in good agreement
with the prediction of Equation (3.56).
96 CH3 PARTICLE PHENOMENOLOGY
The observed hadrons are of three types. There are baryons and their antiparticles
antibaryons, which have half-integral spin, and mesons, which have integral spin. In
the quark model of hadrons the baryons are assumed to be bound states of three
quarks (3q), antibaryons are assumed to be bound states of three antiquarks (3q) and
mesons are assumed to be bound states of a quark and an antiquark (qq).17 The
17
In addition to these so-called ‘valence’ quarks there could also, in principle, be other constituent quarks present
in the form of a cloud of virtual quarks and antiquarks – the so-called ‘sea’ quarks – the origin of which we will
discuss in Chapter 5. In this chapter we consider only the valence quarks which determine the static properties of
hadrons. The masses of the constituent quarks could be quite different from those that appear in the fundamental
strong interaction Hamiltonian for quark–quark interactions via gluon exchange (i.e. QCD), because those
quarks are free of the dynamical effects experienced in hadrons. The latter are referred to as ‘current’ quarks.
HADRONS 97
baryons and antibaryons have baryon numbers 1 and 1 respectively, while the
mesons have baryon number 0. Hence the baryons and antibaryons can annihilate
each other in reactions which conserve baryon number to give mesons or, more
rarely, photons or lepton–antilepton pairs, in the final state.
The lightest known baryons are the proton and neutron, with the quark
compositions given in Section 3.3.1:
These particles have been familiar as constituents of atomic nuclei since the
1930s. The birth of particle physics as a new subject, distinct from atomic and
nuclear physics, dates from 1947, when hadrons other than the neutron and proton
were first detected. These were the pions, already mentioned, and the kaons,
discovered in cosmic rays by groups in Bristol and Manchester Universities, UK,
respectively.
The discovery of the pions was not totally unexpected, since Yukawa had
famously predicted their existence and their approximate masses in 1935, in order
to explain the observed range of nuclear forces (recall the discussion in Section
1.5.2). This consisted of finding what mass was needed in the Yukawa potential to
give the observed range of the strong nuclear force (which was poorly known at the
time). After some false signals, a particle with the right mass and suitable properties
was discovered – this was the pion. Here and in what follows we will give the hadron
masses in brackets in units of MeV/c2 and use a superscript to indicate the electric
charge in units of e. Thus the pions are ð140Þ; 0 ð135Þ. Pions are the lightest
known mesons and have the quark compositions
While the charged pions have a unique composition, the neutral pion is composed of
both uu and d d pairs in equal amounts. Pions are copiously produced in high-energy
collisions by strong interaction processes such as p þ p ! p þ n þ þ .
In contrast to the discovery of the pions, the discovery of the kaons was totally
unexpected, and they were almost immediately recognized as a completely new
form of matter, because they had supposedly ‘strange’ properties. Eventually, after
several years, it was realized that these properties were precisely what would be
expected if kaons had non-zero values of a hitherto unknown quantum number,
given the name strangeness, which was conserved in strong and electromagnetic
interactions, but not necessarily conserved in weak interaction. Particles with non-
zero strangeness were named strange particles, and with the advent of the quark
model in 1964, it was realized that strangeness S was, apart from a sign, the
strangeness quark number introduced earlier, i.e.
S ¼ Ns : ð3:59Þ
98 CH3 PARTICLE PHENOMENOLOGY
Kaons are the lightest strange mesons, with the quark compositions:
~
Nb
NðbÞ NðbÞ:
C
Nc
NðcÞ NðcÞ and B ð3:61Þ
The above examples illustrate just some of the many different combinations of
quarks that form baryons or mesons. These and some further examples are shown
in Table 3.3 and a complete listing is given in the PDG Tables.
Table 3.3 Some examples of baryons and mesons, with their major decay
modes; masses are in MeV/c2
Particle Mass Lifetime (s) Major decays
ðudÞ
þ
140 2:6 108
þ ( 100%)
0 ðuu; d dÞ 135 8:4 1017 ( 100%)
K þ ðusÞ 494 1:2 108 þ (64%)
þ 0 (21%)
K þ ðusÞ 892 1:3 1023 K þ 0 ; K 0 þ ( 100%)
D ðdcÞ 1869 1:1 1012 Several seen
B ðbuÞ 5278 1:6 1012 Several seen
pðuudÞ 938 Stable None
nðuddÞ 940 887 pe e (100%)
ðudsÞ 1116 2:6 1010 p (64%)
n0 (36%)
þþ ðuuuÞ 1232 0:6 1023 pþ (100%)
ðsssÞ 1672 0:8 1010 K (68%)
0 (24%)
þ
c ðudcÞ 2285 2:1 1013 Several seen
To proceed more systematically one could, for example, construct all the mesons
states of the form qq, where q can be any of the six quark flavours. Each of these is
labelled by its spin and its intrinsic parity P. The simplest such states would have
the spins of the two quarks antiparallel with no orbital angular momentum between
them and so have spin-parity J P ¼ 0. (Recall from Chapter 1 that quarks and
antiquarks have opposite parities.) If, for simplicity, we consider those states
composed of just u, d and s quarks, there will be nine such mesons and they have
quantum numbers which may be identified with the observed mesons ðK 0 ; K þ Þ,
0 ; K Þ, ð ; 0 Þ and two neutral particles, which are called and 0 . This
ðK
supermultiplet is shown Figure 3.11(a) as a plot of Y, the hypercharge, defined as
HADRONS 99
þ
Figure 3.11 The lowest-lying states with (a) JP ¼ 0 and (b) J ¼ 12 that are composed of u, d
and s quarks
Y
BþSþCþB ~ þ T, against I3 , the third component of isospin. This can be
extended to the lowest-lying qqq states and the lowest-lying supermultiplet
þ
consists of the eight J P ¼ 12 baryons shown in Figure 3.11(b).18
It is a remarkable fact that the states observed experimentally agree with those
predicted by the simple combinations qqq; qqq and qq and until very recently there
was no evidence for states corresponding to any other combinations. However, some
recent experiments have claimed evidence for the existence of a few states outside this
scheme, possibly ones involving five quarks, although other experiments have failed to
confirm this. Nevertheless, it is still a fact that hadron states are overwhelmingly
composed of the simplest quark combinations of the basic quark model. This was one
of the original pieces of evidence for the existence of quarks and remains one of the
strongest today.
The scheme may also be extended to more quark flavours, although the diagrams
become increasingly complex. For example, Figure 3.12 shows the predicted
þ
J P ¼ 32 baryon states formed from u, d, s and c quarks when all three quarks
have their spins aligned, but still with zero orbital angular momentum between them.
All the states in the bottom plane have been detected as well as many in the higher
planes and with the possible exception of the five-quark states mentioned previously,
no states have been found that are outside this scheme. The latest situation may be
found in the PDG Tables.
For many quark combinations there exist not one, but several states. For example,
the lowest-lying state of the ud system has spin-parity 0 and is the þ meson. It can
be regarded as the ‘ground state’ of the ud system. Here the spins of the quark
18
If you try to try to verify Figure 3.11, you will find that it is necessary to assume that the overall hadronic
wavefunctions ¼ space spin are symmetric under the exchange of identical quarks, i.e. opposite to the
symmetry required by the Pauli principle. This apparent contradiction will be resolved in Chapter 5.
100 CH3 PARTICLE PHENOMENOLOGY
þ
Figure 3.12 The J ¼ 32 baryon states composed of u, d, s and c quarks
constituents are anti-aligned to give a total spin S ¼ 0 and there is no orbital angular
momentum L between the two quarks, so that the total angular momentum, which
we identify as the spin of the hadron, is J ¼ L þ S ¼ 0. Other ‘excited’ states can
have different spin-parities depending on the different states of motion of the quarks
within the hadron.
An example is the K þ ð890Þ meson shown in Table 3.3 with J P ¼ 1 . In this state
the u and s quarks have their spins aligned so that S ¼ 1 and there is no orbital
angular momentum between them, i.e. L ¼ 0, so that the spin of the K þ is
J ¼ L þ S ¼ 1. This is a resonance and such states usually decay by the strong
interaction, with very short lifetimes, of order 1023 s. The mass distribution of their
decay products is described by the Breit–Wigner formula we met in Section 1.6.3.
The spin of a resonance may be found from an analysis of the angular distributions of
its decay products. This is because the distribution will be determined by the
wavefunction of the decaying particle and this will contain an angular part
proportional to a spherical harmonic labelled by the orbital angular momentum
between the decay products. Thus from a measurement of the angular distribution of
the decay products, the angular momentum may be found, and hence the spin of the
resonance. It is part of the triumph of the quark model that it successfully accounts
for the excited states of the various quark systems, as well as their ground states,
when the internal motion of the quarks is properly taken into account.
From experiments such as electron scattering we know that hadrons have typical
radii r of the order of 1 fm and hence associated time scales r/c of the order of
1023 s. The vast majority are highly unstable resonances, corresponding to excited
HADRONS 101
states of the various quark systems, and decay to lighter hadrons by the strong
interaction, with lifetimes of this order. The K þ ð890Þ ¼ us resonance, mentioned
above, is an example. It decays to K þ 0 and K 0 þ final states with a lifetime of
1:3 1023 s. The quark description of the process K þ ! K 0 þ þ , for example, is
From this we see that the final state contains the same quarks as the initial state,
plus an additional d d pair, so that the quark numbers Nu and Nd are separately
conserved. This is characteristic of strong and electromagnetic processes, which
are only allowed if each of the quark numbers Nu ; Nd ; Ns ; Nc and Nb is separately
conserved.
Since leptons and photons do not have strong interactions, hadrons can only decay
by the strong interaction if lighter states composed solely of other hadrons exist with
the same quantum numbers. While this is possible for the majority of hadrons, it is
not in general possible for the lightest state corresponding to any given quark
combination. These hadrons, which cannot decay by strong interactions, are long-
lived on a timescale of the order of 1023 s and are often called stable particles. It
is more accurate to call them long-lived particles, because except for the proton
they are not absolutely stable, but decay by either the electromagnetic or weak
interaction.
The proton is stable because it is the lightest particle with non-zero baryon
number and baryon number is conserved in all known interactions. A few of the
other long-lived hadrons decay by electromagnetic interactions to final states that
include photons. These decays, like the strong interaction, conserve all the
individual quark numbers. An example of this is the neutral pion, which has
Nu ¼ Nd ¼ Ns ¼ Nc ¼ Nb ¼ 0 and decays by the reaction
with a lifetime of 0:8 1016 s. However, most of the long-lived hadrons have non-
zero values for at least one of the quark numbers, and can only decay by the weak
interaction, in which quark numbers do not have to be conserved. For example, the
positive pion decays with a lifetime of 2:6 108 s by the reaction
þ ! þ þ ; ð3:64Þ
! p þ and n þ 0 ; ð3:65Þ
with a lifetime of 2:6 1010 s. The quark interpretations of these reactions are
ðudÞ ! þ þ ; ð3:66Þ
102 CH3 PARTICLE PHENOMENOLOGY
in which an s quark turns into a u quark and a ud pair is created, violating Nd and
Ns conservation.
We see from the above that the strong, electromagnetic or weak nature of a given
hadron decay can be determined by inspecting quark numbers. The resulting
lifetimes can then be summarized as follows. Strong decays lead to lifetimes that
are typically of the order of 1023 s. Electromagnetic decay rates are suppressed by
powers of the fine structure constant relative to strong decays, leading to observed
lifetimes in the range 10161021 s. Finally, weak decays give longer lifetimes,
which depend sensitively on the characteristic energy of the decay.
A useful measure of the decay energy is the Q-value, the kinetic energy released in
the decay of the particle at rest, which we metioned before in Section 2.3. In the weak
interactions of hadrons, Q-values of the order of 102 103 MeV are typical, leading
to lifetimes in the range 107 –1013 s, but there are some exceptions, notably
neutron decay, n ! p þ e þ e , for which
is unusually small, leading to a lifetime of about 103 s. Thus hadron decay lifetimes
are reasonably well understood and span some 27 orders of magnitude, from about
1024 s to about 103 s. The typical ranges corresponding to each interaction are
summarized in Table 3.4.
The quark model can make predictions for hadronic magnetic moments and masses
in a way that is analogous to the semi-empirical mass formula for nuclear masses, i.e.
the formulae have a theoretical basis, but contain parameters that have to be
determined from experiment. We start by examining the case of baryon magnetic
moments.
HADRONS 103
þ
These have been measured only for the 12 octet of states composed of u, d and s
quarks and so we will consider only these. In this supermultiplet, the quarks have
zero orbital angular momentum and so the hadron magnetic moments are just the
sums of contributions from the constituent quark magnetic moments, which we will
assume are of the Dirac form, i.e.
1 1
q
q; Sz ¼ ^z q; Sz ¼ ¼ eq eh=2mq ¼ eq M p =mq N ; ð3:69Þ
2 2
2Mp Mp Mp
u ¼ N ; d ¼ N and s ¼ N : ð3:70Þ
3mu 3md 3ms
Mp
¼ s ¼ N : ð3:71Þ
3ms
þ
For 12 baryons B with quark configuration aab, the aa pair is in the symmetric
spin-1 state with parallel spins (again this is to ensure that the predicted quantum
numbers of the supermultiplet agree with experiment) and magnetic moment 2a .
The ‘spin-up’ baryon state is given by
rffiffiffi
1 1 2 1 1
B; S ¼ ; Sz ¼ ¼ b; S ¼ ; S ¼ aa; S ¼ 1; S ¼ 1
2 2 3 2
z
2
z
rffiffiffi
1 1 1
b; S ¼ ; S ¼ aa; S ¼ 1; S ¼ 0 ð3:72Þ
3 2
z z
2
The first term corresponds to a state with magnetic moment 2a b , since the b
quark has Sz ¼ 12; the second term corresponds to a state with magnetic moment
b , since the aa pair has Sz ¼ 0 and does not contribute. Hence the magnetic
moment of B is given by
2 1 4 1
B ¼ ð2a b Þ þ b ¼ a b : ð3:73Þ
3 3 3 3
104 CH3 PARTICLE PHENOMENOLOGY
4 1 Mp
p ¼ u d ¼ N ; ð3:74Þ
3 3 m
where we have neglected the mass difference between the u and d quarks, as
suggested by isospin symmetry, and set mu md
m. The predictions for the
þ
magnetic moments of all the other members of the 12 octet may be found in a
similar way in terms of just two parameters, the masses m and ms . A best fit to the
measured magnetic moments (but not taking account of the errors on the data19)
yields the values m ¼ 0:344 GeV=c2 and ms ¼ 0:539 GeV=c2 . The predicted
moments are shown in Table 3.5. The agreement is good, but by no means perfect
and suggests that the assumption that baryons are pure three-quark states with zero
orbital angular momentum between them is not exact. For example, there could be
small admixtures of states with non-zero orbital angular momentum.
þ
Table 3.5 Magnetic moments of the 12 baryon octet as predicted by the
constituent quark model, compared with experiment in units of N , the
nuclear magneton; these have been obtained using m ¼ 0:344 GeV/c2 and
ms ¼ 0:539 GeV/c2 -- errors on the nucleon moments are of the order of 107
Particle Moment Prediction Experiment
4 1
p(938) 3u 3d 2.73 2.793
4 1
n(940) 3d 3u 1.82 1.913
(1116) s 0.58 0:613 0:004
þ (1189) 4 1
3u 3s 2.62 2:458 0:010
4 1
(1197) 3d 3s 1.02 1:160 0:025
4 1
0 ð1315Þ 3s 3u 1.38 1:250 0:014
4 1
ð1321Þ 3s 3d 0.47 0:651 0:003
We now turn to the prediction of hadron masses. The mass differences between
members of a given supermulitplet are conveniently separated into the small mass
differences between members of the same isospin multiplet and the much larger
mass differences between members of different isospin multiplets. The size of the
former suggests that they have their origin in electromagnetic effects, and if we
neglect them then a first approximation would be to assume that the mass
differences are due solely to differences in the constituent quark masses. If we
concentrate on hadrons with quark structures composed of u, d and s quarks, since
19
If we had fitted taking account of the errors, the fit would be dominated by the proton and neutron moments
because they have very small errors.
HADRONS 105
their masses are the best known from experiment, this assumption leads directly to
the relations
M M ¼ M M ¼ M MN ¼ ms mu;d ð3:75Þ
1þ
for the 2 baryon octet and
8 ei ej
E ¼ j ð0Þj2 Si Sj ; ð3:77Þ
3 mi mj
where ð0Þ is the wavefunction at the origin, rij ¼ 0. (When averaged over all
space, the interaction is zero except at the origin.) In atomic physics this is known
as the hyperfine interaction and causes very small splittings in atomic energy
levels. In the hadron case, the electric charges must be replaced by their strong
interaction equivalents with appropriate changes to the overall numerical factor.
The resulting interaction is called (for reasons that will be clear in Chapter 5) the
chromomagnetic interaction. As we cannot calculate the equivalent quark–quark
wavefunction, for the purposes of a phenomenological analysis we will write the
contribution to the hadron mass as
S1 S2
M / : ð3:78Þ
m1 m2
This of course assumes that j ð0Þj2 is the same for all states, which will not be
exactly true.
106 CH3 PARTICLE PHENOMENOLOGY
Consider first the case of mesons. By writing the total spin squared as
S2
ðS1 þ S2 Þ2 ¼ S21 þ S22 þ 2S1 S2 ; ð3:79Þ
we easily find the expect values of S1 S2 are 34h2 for the S ¼ 0 (pseudoscalar)
mesons and 14h2 for the S ¼ 1 (vector) mesons. Then the masses may be written
MðmesonÞ ¼ m1 þ m2 þ M; ð3:80Þ
where m1;2 are the masses of the constituent quarks and
3a 1 a 1
MðJ P ¼ 0 mesonÞ ¼ ; MðJ P ¼ 1 mesonÞ ¼ ð3:81Þ
4 m1 m2 4 m1 m2
and a is a constant to be found from experiment. The masses of the members of the
0 and 1 meson supermultiplets then follow from a knowledge of their quark
compositions. For example, the K-mesons have one u or d quark and one s quark and so
3a 1 1
MK ¼ m þ ms þ : ð3:82Þ
8 m2 m2s
Predictions for the masses of all the mesons are shown in Table 3.6, which also
gives the best fit to the measured masses (again ignoring the relative errors on the
latter) using these formulae. The predictions correspond to the values
Table 3.6 Meson masses (in Gev/c2) in the constituent quark model compared
with experimental values
Particle Mass Prediction Experiment
3a
2m 0.15 0.137
4m2
3a 1 1
K m þ ms þ 0.46 0.496
8 m2 m2s
2 4 a 1 2
m þ ms þ 0.57 0.549
3 3 4 m2 m2s
a
2m þ 2 0.77 0.770
4m
a
! 2m þ 2 0.77 0.782
4m
a 1 1
K m þ ms þ þ 0.87 0.892
8 m2 m2s
a
2ms þ 2 1.03 1.020
4ms
HADRONS 107
The comparison with the measured values is very reasonable, but omitted from the fit
is the 0 state where the fit is very poor indeed. Unlike the atomic case, the spin–spin
interaction in the strong interaction case leads to substantial corrections to the meson
masses.
The baryons are somewhat more complicated, because in this case we have three
pairs of spin–spin couplings to consider. In general the spin–spin contribution to
the mass is
X Si Sj
M / ; i; j ¼ 1; 3: ð3:84Þ
i<j
mi mj
þ
In the case of the 32 decuplet, all three quarks have their spins aligned and every
pair therefore combines to make spin-1. Thus for example,
Using this result, the mass of the ð1385Þ, for example, may be written
b 1 2
M ¼ 2m þ ms þ þ ; ð3:87Þ
4 m2 mms
and hence
Table 3.7 Baryon masses (in Gev/c2) in the constituent quark model compared
with experimental values
Particle Mass Prediction Experiment
3b
N 3m 2 0.89 0.939
4m
3b 1
2m þ ms 1.08 1.116
4 m2
b 1 4
2m þ ms þ 1.15 1.193
4 m2 mms
b 1 4
m þ 2ms þ 1.32 1.318
4 m2s mms
3b
3m þ 2 1.07 1.232
4m
b 1 2
2m þ ms þ þ 1.34 1.385
4 m mms
b 2 1
m þ 2ms þ þ 2 1.50 1.533
4 mms ms
3b
3ms þ 2 1.68 1.673
ms
Finally, setting mu ¼ md ¼ m and absorbing factors of h2 into the constant b, gives
Su Sd ðS1 S2 þ S1 S3 þ S2 S3 Su Sd Þ 3b
M ¼ 2m þ ms þ b 2
þ ¼ 2m þ ms 2 ;
m m ms 4m
ð3:91Þ
þ
where we have used Equation (3.89). The resulting formulae for all the 12 octet and
3þ
2 decuplet masses are shown in Table 3.7. Also shown are the predicted masses,
where for consistency we have used the same quark mass values obtained earlier in
fitting baryon magnetic moments, i.e. m ¼ 0:308 GeV=c2 and ms ¼ 0:482 GeV=c2 ,
and varied only the parameter b, giving a value 0.0225 ðGeV=c2 Þ3 . The fit is quite
reasonable and although better fits can be obtained by allowing the masses to vary
there is little justification for this, given the approximations of the analysis.
Overall, what we learn from the above is that the constituent quark model is
capable of giving a reasonably consistent account of hadron masses and magnetic
moments, at least for the low-lying states (the 0 is an exception), provided a few
parameters are allowed to be found from experiment.
Problems
3.1 Which of the following reactions are allowed and which are forbidden by the
conservation laws appropriate to weak interactions?
(a) þ p ! þ þ n;
(b) e þ p ! n þ e þ þ ;
(c) ! þ þ e þ e ;
(d) K þ ! 0 þ þ þ .
PROBLEMS 109
3.2 Draw the lowest-order Feynman diagram at the quark level for the following
decays:
(a) D ! K 0 þ ;
(b) ! p þ e þ e .
~ ) where
3.3 Consider the following combinations of quantum numbers (Q; B; S; C; B
Q ¼ electric charge, B ¼ baryon number, S ¼ strangeness, C ¼ charm and B ~¼
beauty:
Which of these possible states are compatible with the postulates of the quark
model?
3.4 Consider a scenario where overall hadronic wavefunctions consist of just spin and
space parts, i.e. ¼ space spin . What would be the resulting multiplet structure of
the lowest-lying baryon states composed of u; d and s quarks?
3.5 Draw Feynman diagrams at the quark level for the reactions:
(b) þ p ! K 0 þ 0 .
3.6 Find the parity P and charge conjugation C values for the ground-state ðJ ¼ 0Þ
meson and its first excited ðJ ¼ 1Þ state . Why does the charged pion have a
longer lifetime than the ? Explain also why the decay 0 ! þ has been
observed, but not the decay 0 ! 0 0.
! 0 þ
K þ ! þ þ 0
j
! 0 þ ; j and K 0 ! þ þ þ 0
j ! þ þ
! þ
Classify each process as strong, weak or electromagnetic and give your reasons.
110 CH3 PARTICLE PHENOMENOLOGY
3.10 Draw the lowest-order Feynman diagram for the decay K þ ! þ þ þ and
hence deduce the form of the overall effective coupling.
(a) p þ p ! þ þ ;
(b) p ! eþ þ ;
(c) 0 ! þ ;
(d) p þ p ! þ þ n þ K 0 þ þ ;
(e) ! þ ;
(f) þ ! p þ 0 .
3.13 Use the results of Section 3.3.1 to deduce a relation between the total cross-sections
for the reactions p ! K 0 0 ; p ! K þ and þ p ! K þ þ at a fixed energy.
3.14 At a certain energy ðþ nÞ ð pÞ, whereas ðK þ nÞ 6¼ ðK pÞ. Explain this.
4
Experimental Methods
4.1 Overview
To explore the structure of nuclei (nuclear physics) or hadrons (particle physics)
requires projectiles whose wavelengths are at least as small as the effective radii of
the nuclei or hadrons. This determines the minimum value of the momentum
p ¼ h= and hence the energy required. The majority of experiments are
conducted using beams of particles produced by machines called accelerators.
This has the great advantage that the projectiles are of a single type, and have
energies that may be controlled by the experimenter.2 For example, beams that are
essentially mono-energetic may be prepared, and can be used to study the energy
dependence of interactions. The beam, once established, is directed onto a target so
that interactions may be produced. In a fixed-target experiment the target is
stationary in the laboratory. Nuclear physics experiments are almost invariably of
this type, as are many experiments in particle physics.
In particle physics, high energies are also required to produce new and unstable
particles and this reveals a disadvantage of fixed-target experiments when large
1
See, for example, Fe86 and Kl86.
2
Nevertheless, important experiments are still performed without using accelerators, for example some of
those described in Chapter 3 on neutrino oscillations used cosmic rays and nuclear reactors. In fact cosmic
rays are still the source of the very highest-energy particles.
and using P2t ¼ m2t c2 etc., together with the general result
Pi Pj ¼ Ei Ej =c2 pi pj ; ð4:4Þ
we have
1=2
ECM ¼ m2b c4 þ m2t c4 þ 2mt c2 EL : ð4:5Þ
1
At high energies this increases only as ðEL Þ and so an increasingly smaller
2
fraction of the beam energy is available for particle production, most going to
impart kinetic energy to the target.
In a colliding-beam accelerator, two beams of particles travelling in almost
opposite directions are made to collide at a small or zero crossing angle. If for
simplicity we assume the particles in the two beams have the same mass and
laboratory energy EL and collide at zero crossing angle, then the total centre-
of-mass energy is
ECM ¼ 2EL : ð4:6Þ
This increases linearly with the energy of the accelerated particles, and hence is a
significant improvement on the fixed-target result. Colliding-beam experiments are
not, however, without their own disadvantages. The colliding particles have to be
stable, which limits the interactions that can be studied, and the collision rate in the
intersection region is smaller than that achieved in fixed-target experiments,
because the particle densities in the beams are low compared with a solid or
liquid target.
3
A brief summary of relativistic kinematics is given in Appendix B.
ACCELERATORS AND BEAMS 113
Finally, details of the particles produced in the collision (e.g. their momenta) are
deduced by observing their interactions with the material of detectors, which are
placed in the vicinity of the interaction region. A wide range of detectors is
available. Some have a very specific characteristic, others serve more than one
purpose. Modern experiments, particularly in particle physics, typically use several
types in a single experiment.
In this chapter we start by describing some of the different types of accelerator
that have been built, the beams that they can produce and also how beams of
neutral and unstable particles can be prepared. Then we discuss the ways in which
particles interact with matter, and finally review how these mechanisms are
exploited in the construction of a range of particle detectors.
4.2.1 DC accelerators
4
Sir John Cockcroft and Ernest Walton received the 1951 Nobel Prize in Physics for the development of their
accelerator and the subsequent nuclear physics experiments they did using it.
Figure 4.1 Principle of the tandem van de Graaff accelerator
ACCELERATORS AND BEAMS 115
In Figure 4.1, a high voltage source at I passes positive ions to a belt via a comb
arrangement at C. The belt is motor driven via the pulleys at P and the ions are
carried on the belt to a second pulley where they are collected by another comb
located within a metal vessel T. The charges are then transferred to the outer
surface of the vessel, which acts as an extended terminal. In this way a high
voltage is established on T. Singly-charged negative ions are injected from a
source and accelerated along a vacuum tube towards T. Within T there is a stripper
S (for example a thin carbon foil) that removes two or more electrons from the
projectiles to produce positive ions. The latter then continue to accelerate through
the second half of the accelerator increasing their energy still further and finally
may be bent and collimated to produce a beam of positive ions. This brief account
ignores many technical details. For example, an inert gas at high pressure is used to
minimize electrical breakdown by the high voltage. The highest energy van de
Graaff accelerator can achieve a potential of about 30–40 MeV for singly-charged
ions and greater if more than one electron is removed by the stripper. It has been a
mainstay of nuclear research.
4.2.2 AC accelerators
Linear accelerators
In a linear accelerator (or linac) for acclerating ions, particles pass through a series
of metal pipes called drift tubes, that are located in a vacuum vessel and connected
successively to alternate terminals of an r.f. oscillator, as shown in Figure 4.2.
Positive ions accelerated by the field move towards the first drift tube. If the
alternator can change its direction before the ions passes through that tube, then
they will be accelerated again on their way between the exit of the first and entry
to the second tube, and so on. Thus the particles will form bunches. Because
the particles are accelerating, their speed is increasing and hence the lengths
of the drift tubes have to increase to ensure continuous acceleration. To produce
a useful beam the particles must keep in phase with the r.f. field and remain
focused.
In the case of electrons, their velocity very rapidly approaches the speed of light.
In this case a variation of the linac method that is more efficient is used to
accelerate them. Bunches of particles pass through a straight evacuated waveguide
with a periodic array of gaps, similar to the ion accelerator. Radio frequency
oscillations in the gaps are used to establish a moving electromagnetic wave in the
structure, with a longitudinal component of the electric field moving in phase with
the particles. As long as this phase relationship can be maintained, the particles
will be continuously accelerated. Radio frequency power is pumped into the
waveguide at intervals to compensate for resistive losses and gives energy to the
electrons. The largest electron linac is at the SLAC laboratory in Stanford, USA,
and has a maximum energy of 50 GeV. It is over 3 km long.
An ingenious way of reducing the enormous lengths of high-energy linacs has
been developed at the Continuous Electron Beam Accelerator Facility (CEBAF) at
the Jefferson Laboratory in the USA. This utilizes the fact that above about
50 MeV, electron velocities are very close to the speed of light and thus electrons
of very different energies can be accelerated in the same drift tube. Instead of a
single long linac, the CEBAF machine consists of two much shorter linacs and the
beam from one is bent and passed through the other. This can be repeated for up to
four cycles. Even with the radiation losses inherent in bending the beams, very
intense beams can be produced with energies between 0.5 and 6.0 GeV. CEBAF is
proving to be an important machine in the energy region where nuclear physics
and particle physics descriptions overlap.
Cyclic accelerators
Cyclic accelerators used for low-energy nuclear physics experiments are of a type
called cyclotrons. They are also used to produce beams of particles for medical
applications, including proton beams for radiation therapy. Cyclotrons operate in a
somewhat different way to cyclic accelerators used in particle physics, which are
called synchrotrons. In a cyclotron,5 charged particles are constrained to move
in near-circular orbits by a magnetic field during the acceleration process. There
are several types of cyclotron; we will describe just one. This is illustrated
5
The cyclotron was invented by Ernest Lawrence, who received the 1939 Nobel Prize in Physics for this and
the experimental work he did using it.
ACCELERATORS AND BEAMS 117
Figure 4.3 Schematic diagram of a cyclotron (adapted from Kr88, copyright John Wiley &
Sons)
Figure 4.4 Cross-section of (a) a typical bending (dipole) magnet, and (b) a focusing
(quadrupole) magnet; the thin arrows indicate field directions; the thick arrows indicate the
force on a negative particle travelling into the paper
Figure 4.5 Magnitude of the electric field as a function of time at a fixed point in the rf cavity:
particle B is synchronous with the field and arrives at time tB ; particle A (C) is behind (ahead of)
B and receives an increase (decrease) in its rotational frequency -- thus particles oscillate about
the equilibrium orbit
beam, and a number of technical problems have to be overcome which are outside
the scope of this brief account.6
Both linear and cyclic accelerators can be divided into fixed-target and colliding-
beam machines. The latter are also known as colliders, or sometimes in the case of
cyclic machines, storage rings.7 In fixed-target machines, particles are accelerated
to the highest operating energy and then the beam is extracted from the machine
and directed onto a stationary target, which is usually a solid or liquid. Much
higher energies have been achieved for protons than electrons, because of the large
radiation losses inherent in electron machines mentioned earlier. The intensity of
the beam is such that large numbers of interactions can be produced, which can
either be studied in their own right or used to produce secondary beams.
The main disadvantage of fixed-target machines for particle physics has been
mentioned earlier: the need to achieve large centre-of-mass energies to produce
6
Very recently (2005) significant progress has been made on an ‘induction synchrotron’ in which a changing
magnetic field produces the electric field that accelerates the particles. This device has the potential to
overcome certain effects that limit the intensity achievable in conventional synchrotrons.
7
The use of the terms storage rings and colliders as synonymous is not strictly correct, because we will see
that the former can also describe a machine that stores a single beam for use on both internal and external
fixed targets.
120 CH4 EXPERIMENTAL METHODS
new particles. Almost all new machines for particle physics are therefore colliders,
although some fixed-target machines for specialized purposes are still constructed.
The largest collider currently under construction is the Large Hadron Collider
(LHC), which is being built at CERN, Geneva, Switzerland. This is a massive pp
accelerator of circumference 27 km, with each beam having an energy of 7 TeV.
A schematic diagram of the CERN site showing the LHC and some of its other
accelerators is shown in Figure 4.6. The acceleration process starts with a linac
Figure 4.6 A schematic diagram of the CERN site showing the LHC and some of its other
accelerators (CERN photo, reproduced with permission)
ACCELERATORS AND BEAMS 121
whose beam is boosted in energy in the Proton Synchrotron Booster (PSB) and
passed to the Proton Synchrotron (PS), a machine that is still the source of beams
for lower-energy experiments. The beam energy is increased still further in the
Super Proton Synchrotron (SPS) that also provides beams for a range of experi-
ments as well as the injection beams for the LHC itself. Four beam intersection
points are shown in the LHC and experiments (ALICE, CMS, LHC-b and ATLAS)
will be located at each of these. The extracted neutrino beam shown at the bottom
of the diagram is sent to the Gran Sasso laboratory 730 km away and is used,
amongst other things, for experiments on neutrino oscillations of the type
discussed in Chapter 3.
Another very large collider we should mention is the Relativistic Heavy Ion
Collider (RHIC), located at Brookhaven National Laboratory, USA. This unique
machine, which began operation in 2000 following 10 years of development and
construction, is the first collider in the world capable of accelerating heavy ions.
Like the LHC above, there are several stages, involving a linac, a tandem van de
Graaff and a synchrotron, before the ions are injected into the main machine.
There they form two counter-circulating beams controlled by two 4-km rings of
superconducting magnets and are accelerated to an energy of 100 GeV/nucleon.
Thus the total centre-of-mass energy is 200 GeV/nucleon. Collisions occur at six
intersection points, where major experiments can be sited. RHIC primarily
accelerates ions of gold and is used to study matter at extreme energy-densities,
where a new state of matter called a ‘quark–gluon plasma’ is predicted to occur.
We will return to this briefly in Chapter 9.
The performance of a collider is characterized by its luminosity, which was
defined in Chapter 1. The general formula for luminosity given there is shown in
Problem 1.10 to reduce in the case of a collider to the useful form
N1 N2
L¼n f; ð4:7Þ
A
where Ni ði ¼ 1,2) are the numbers of particles in the n colliding bunches, A is the
cross-sectional area of the beam and f is the frequency, i.e. f ¼ 1=T, where T is the
time taken for the particles to make one traversal of the ring.
An interesting proton synchrotron for nuclear physics studies is the COSY
facility located at the Research Centre Jülich, Germany. Low-energy protons are
pre-accelerated in a cyclotron, then cooled to reduce their transverse momentum
and injected into a synchrotron, where they are further accelerated to momenta in
the range 600–3700 MeV/c (corresponding to energies of 175–2880 MeV). The
protons can be stored in the ring for appreciable times and are available for
experiments not only in the usual way by extracting the beam, but also by using the
circulating beam to interact with a very thin internal target. Thus we have a
mixture of storage rings and fixed targets. The fact that the circulating beam may
make as many as 1010 traversals through the target compensates to some extent for
its low particle density.
122 CH4 EXPERIMENTAL METHODS
The particles used in accelerators must be stable and charged, but one is also
interested in the interaction of neutral particles, e.g. photons and neutrons, as well
as those of unstable particles, such as charged pions. Beams appropriate for
performing such experiments are produced in a number of ways.
We have seen that neutrons are the natural product of radioactive decays and we
will see in Chapter 8 that a large flux of neutrons is present in a nuclear reactor.
Typically these will have a spectrum concentrated at low energies of 1–2 MeV, but
extending as high as 5–6 MeV. Purpose-built reactors exist for research purposes,
such as the ILL reactor at the Institut Laue-Langevin, France. Another source of
neutrons is via the spallation process. The most important neutron spallation
source at present is ISIS located at the Rutherford Appleton Laboratory, UK.
Protons which have been accelerated in a linac to 70 MeV are injected into a
synchrotron that further accelerates them to 800 MeV, where they collide with a
heavy metal target of tantalum. The interaction drives out neutrons from the target
and provides an intense pulsed source. In each case, if beams of lower-energy
neutrons are required these are produced by slowing down faster neutrons in
moderators, which are materials with a large cross-section for elastic scattering,
but a small cross-section for absorption. In Chapter 8 we will see that moderators
are vital for the successful extraction of power from fission nuclear reactors.
Beams of unstable particles can be formed provided their constituents live long
enough to travel appreciable distances in the laboratory. One way of doing this is to
direct an extracted primary beam onto a heavy target. In the resulting interactions
with the target nuclei, many new particles are produced which may then be
analysed into secondary beams of well-defined momentum. Such beams will
ideally consist predominantly of particles of one type, but if this cannot
be achieved, then the wanted species may have to be identified by other means.
In addition, if these secondary beams are composed of unstable particles, they can
themselves be used to produce further beams formed from their decay products.
Two examples will illustrate how, in principle, such secondary particle beams can
be formed.
Consider firstly the construction of a K beam from a primary beam of protons.
By allowing the protons to interact with a heavy target, secondary particles will be
produced. Most of these will be pions, but a few will be kaons (that have to be
produced with a hyperon to conserve strangeness – this an example of so-called
associated production). A collimator can be used to select particles in a particular
direction, and the K component can subsequently be removed and focused into a
mono-energetic beam by selective use of electrostatic fields and bending and
focusing magnets.
The pion beam may also be used to produce a beam of neutrinos. For example,
the is unstable and as we have seen, one of its weak interaction decays modes is
! þ . So if the pions are passed down a long vacuum pipe, many will
PARTICLE INTERACTIONS WITH MATTER 123
decay in flight to give muons and anti-neutrinos, which will mostly travel in
essentially the same direction as the initial beam. The muons and any remaining
pions can then be removed by passing the beam through a very long absorber,
leaving the neutrinos. In this case the final neutrino beam will have a momentum
spectrum reflecting the initial momentum spectrum of the pions and, since
neutrinos are neutral, no further momentum selection using magnets is possible.
For hadrons, the most important short-range interactions with nuclei are due to the
strong nuclear force which, unlike the electromagnetic interaction, is as important
for neutral particles as for charged ones, because of the charge-independence of the
strong interaction. Both elastic scattering and inelastic reactions may result. At
high energies, many inelastic reactions are possible, most of them involving the
production of several particles in the final state.
Many hadronic cross-sections show considerable structure at low energies due to
the production of hadronic resonances, but at energies above about 3 GeV, total
cross-sections are usually slowly varying in the range 10–50 mb and are much
larger than the elastic cross-section. (The example of p scattering is shown
in Figure 4.7.) This is of the same order-of-magnitude as the ‘geometrical’
124 CH4 EXPERIMENTAL METHODS
Figure 4.7 Total and elastic cross-sections for p scattering as functions of the pion
laboratory momentum
This is called the collision length. An analogous quantity is the absorption length,
defined by
‘a ¼ 1=n
inel ; ð4:9Þ
Ionization energy losses are important for all charged particles, and for particles
other than electrons and positrons they dominate over radiation energy losses at all
but the highest attainable energies. The theory of such losses, which are due
dominantly to Coulomb scattering from the atomic electrons, was worked out by
Bethe, Bloch and others in the 1930s. The result is called the Bethe–Bloch
formula, and for spin-0 bosons with charge q (in units of e), mass M and velocity
v, it takes the approximate form (neglecting small corrections for highly relativistic
particles)
dE D q2 ne 2me c2 2 2 2
¼ ln ; ð4:11Þ
dx 2 I
h2
4
2
D¼ ¼ 5:1
1025 MeV cm2 ; ð4:12Þ
me
me is the electron mass, ¼ v=c and ¼ ð1 2 Þ1=2 . The other constants refer
to the properties of the medium: ne is the electron density; I is the mean ionization
potential of the atoms averaged over all electrons, which is given approximately by
I ¼ 10 Z eV for Z greater than 20. The corresponding formula for spin-12 particles
differs from this, but in practice the differences are small and may be neglected in
discussing the main features of ionization energy loses.
Examples of the behaviour of dE=dx for muons, pions and protons traversing a
range of materials is shown in Figure 4.8. It is common practice to absorb the
density of the medium by dividing by and expressing dE=dx in terms of an
126 CH4 EXPERIMENTAL METHODS
Figure 4.8 Ionization energy loss for muons, pions and protons on a variety of mate-
rials (reprinted from Ei04, copyright Elsevier, with permission)
equivalent thickness of gm cm2 – hence the units in Figure 4.8. As can be seen,
dE=dx falls rapidly as the velocity increases from zero because of the 1= 2 factor
in the Bethe–Bloch equation. All particles have a region of ‘minimum ionization’
for in the range 3–4. Beyond this, tends to unity, and the logarithmic factor in
the Bethe–Bloch formula gives a ‘relativistic rise’ in dE=dx.
The magnitude of the energy loss depends on the medium. The electron density
is given by ne ¼ NA Z=A, where NA is Avogadro’s number, and and A are the
mass density and atomic weight of the medium, so the mean energy loss is
proportional to the density of the medium. The remaining dependence on the
medium is relatively weak because Z=A 0:5 for all atoms except the very light
PARTICLE INTERACTIONS WITH MATTER 127
and the very heavy elements, and because the ionization energy I only enters the
Bethe–Bloch formula logarithmically. In the ‘minimum ionization’ region where
3–4, the minimum value of dE=dx can be calculated from Equation (4.11)
and for a particle with unit charge is given approximately by
dE Z
3:5 MeVg1 cm2 : ð4:13Þ
dx min A
Ionization losses are proportional to the squared charge of the particle, so that a
fractionally charged particle with 3 would have a much lower rate of energy
loss than the minimum energy loss of any integrally charged particle. This has
been used as a means of identifying possible free quarks, but without success.
From the knowledge of the rate of energy loss, we can calculate the attenuation
as a function of distance travelled in the medium. This is called the Bragg curve.
Most of the ionization loss occurs near the end of the path where the speed is
smallest and the curve has a pronounced peak (the Bragg peak) close to the end
point before falling rapidly to zero at the end of the particle’s path length.
The range R, i.e. the mean distance a particle travels before it comes to rest is
defined as
ð
xmax
R dxðÞ; ð4:14Þ
0
where F is a function of the initial velocity and we have used the relation
E ¼ Mc2 to show the dependence on the projectile mass M.
The range as given by Equation (4.15) is actually an average value because
scattering is a statistical process and there will therefore be a spread of values for
individual particles. The spread will be greater for light particles and smaller for
heavier particles such as
-particles. These properties have implications for the use
of radiation in therapeutic situations, where it may be necessary to deposit energy
within a small region at a specific depth of tissue, for example to precisely target a
cancer. The biological effects of radiation will be discussed in Chapter 8.
Because neutrons are uncharged, direct detection is not possible by ionization
methods. However, they can be detected via the action of the charged products of
induced direct nuclear reactions. Commonly used reactions are 6 Liðn;
Þ3 H,
10
Bðn;
Þ7 Li and 3 Heðn; pÞ3 H. All these reactions are exothermic and so are
very suitable for detecting neutrons with energies below about 20 MeV. Moreover,
as nuclear cross-sections tend to increase as v1 at low energies, detection
becomes more efficient the slower the neutron.
128 CH4 EXPERIMENTAL METHODS
When a charged particle traverses matter it can also lose energy by radiative
collisions, especially with nuclei. The electric field of a nucleus will accelerate and
decelerate the particles as they pass, causing them to radiate photons, and hence
lose energy. This process is called bremsstrahlung (literally ‘braking radiation’ in
German) and is a particularly important contribution to the energy loss for
electrons and positrons.
The dominant Feynman diagrams for electron bremsstrahlung in the field of a
nucleus, i.e.
e þ ðZ; AÞ ! e þ þ ðZ; AÞ; ð4:16Þ
are shown in Figure 4.9 and are of the order of Z 2
3 . The function of the nucleus
is to absorb the recoil energy and so ensure that energy and momentum are
simultaneously conserved (recall the discussion of Feynman diagrams in Chapter 1).
Figure 4.9 Dominant Feynman diagrams for the bremsstrahlung process e þ ðZ; AÞ ! e þ þ ðZ; AÞ
There are also contributions from bremsstrahlung in the fields of the atomic
electrons, each of the order of
3 . Since there are Z atomic electrons for each
nucleus, these give a total contribution of the order of Z
3 , which is small
compared with the contribution from the nucleus for all but the lightest elements.
A detailed calculation shows that for relativistic electrons with E
mc2 =
Z 1=3 ,
the average rate of energy loss is given by
dE=dx ¼ E=LR : ð4:17Þ
The constant LR is called the radiation length and is a function of Z and na , the
number density of atoms/cm3 in the medium. Integrating Equation (4.17) gives
E ¼ E0 expðx=LR Þ; ð4:18Þ
where E0 is the initial energy. It follows that the radiation length is the average
thickness of material that reduces the mean energy of an electron or positron by a
factor e. For example, the radiation length in lead is 0.566 cm.
From these results, we see that at high energies the radiation losses are
proportional to E=m2p for an arbitrary charged particle of mass mp . On the other
PARTICLE INTERACTIONS WITH MATTER 129
hand, the ionization energy losses are only weakly dependent on the projectile
mass and energy at very high energies. Consequently, radiation losses completely
dominate the energy losses for electrons and positrons at high enough energies, but
are much smaller than ionization losses for all particles other than electrons and
positrons at all but the highest energies.
Taking into account the above and the results of Section 4.3.2, we see that at low
energies, particles with the same kinetic energy but different masses can have
substantially different ranges. Thus, for example, an electron of 5 MeV has a range
that is several hundred times that of an
-particle of the same kinetic energy.
8
Arthur Compton shared the 1927 Nobel Prize in Physics for the discovery of the increase in wavelength that
occurs when photons with energies of around 0.5–3.5 MeV interact with electrons in a material – the original
Compton effect.
130 CH4 EXPERIMENTAL METHODS
Most gas detectors detect the ionization produced by the passage of a charged
particle through a gas, typically an inert one such as argon, either by collecting the
ionization products or induced charges onto electrodes, or (historically) by making
the ionization track visible in some form. The average energy needed to produce
an electron–ion pair is 30 10 eV, with a weak dependence on the gas used and
the energy of the incident particle. In practice, the output is a pulse at the anode
(which is amplified by electronic means), with the bulk of the signal being due to
the positive ions because of their longer drift distance. For a certain range of
applied voltages – the so-called ‘proportional region’ (see below) – these devices
are primarily used to provide accurate measurements of a particle’s position. As
position detectors, gas detectors largely replaced earlier detectors which used
visual techniques, such as cloud chambers, bubble chambers and stacks of
photographic emulsions, although the latter are still an ingredient in some neutrino
experiments. Although historically important, none of these visual devices are now
9
For more detailed discussions of particle detectors see, for example, Gr96 and the references in Footnote 1.
There are also useful reviews in Chapter 5 of Ho97 and Ei04.
132 CH4 EXPERIMENTAL METHODS
Figure 4.12 Gas amplification factor as a function of voltage V applied in a single-wire gas
detector, with a wire radius typically 20 mm, for a strongly ionizing particle (
) and a weakly
ionizing particle (electron)
in general use and they have been superceded by electronic detectors.10 In particle
physics experiments being planned at the new accelerators currently being built,
gas detectors themselves are being replaced by a new generation of solid-state
detectors based on silicon.
To understand the principles of gas detectors we refer to Figure 4.12, which
shows the number of ion pairs produced per incident charged particle (the gas
10
These early detector techniques produced many notable discoveries and their importance has been
recognized by the award of no less than five Nobel Prizes in Physics: a share of the 1927 Prize to Charles
Wilson for the invention and use of the cloud chamber; the 1948 Prize to Patrick Blackett for further
developments of the cloud chamber and discoveries made with it; the 1950 Prize to Cecil Powell for
development of the photographic emulsion technique and its use to discover pions; the 1960 Prize to Donald
Glaser for the invention of the bubble chamber; and the 1968 Prize to Luis Alvarez for developing the bubble
chamber and associated data analysis techniques resulting in the discovery of a large number of hadronic
resonances.
PARTICLE DETECTORS 133
amplification factor) as a function of the applied voltage V for two cases: a heavily
ionizing particle (e.g. an alpha particle – upper curve) and a lightly ionizing
particle (e.g. an electron – lower curve).
Ionization chamber
At low applied voltages, the output signal is very small because electron–ion pairs
recombine before reaching the electrodes, but as the voltage increases the number
of pairs increases to a saturation level representing complete collection. This is the
region of the ionization chamber. The simplest type of chamber is a parallel plate
condenser filled with an inert gas and having an electric field E ¼ V=d, where d is
the distance between the plates. In practice the gas mixture must contain at lease
one ‘quenching’ component that absorbs ultraviolet light and stops a plasma
forming and spreading throughout the gas.
Another arrangement is cylindrical with an inner anode of radius ra and an outer
cathode of radius rc , giving an electric field
V
EðrÞ ¼ ð4:24Þ
rlnðrc =ra Þ
at a radial distance r from the centre of the anode wire. The output signal is
proportional to the number of ions formed and hence the energy deposited by the
radiation, but is independent of the applied voltage. However, the signal is very
small compared with the noise of all but the slowest electronic circuits and
requires considerable amplification to be useful. Overall, the energy resolution and
the time resolution of the chamber are relatively poor and ionization chambers are
of very limited use in recording individual pulses. They are used, for example, as
beam monitors, where the particle flux is very large, and in medical environments
to calibrate radioactive sources.
As mentioned previously, neutrons cannot be directly detected by ionization
methods, but neutron flux measurements can be made with ionization chambers
(or proportional chambers – see below) filled with BF3 by utilizing the neutron
activation reactions of Section 4.3.1.
Proportional counters
If the voltage is increase beyond the region of operation of the ionization chamber,
we move into the proportional region. In this region, a cylindrical arrangement
as used in the ionization chamber will produce electric field strengths of the order
of 104 –105 V/cm near the wire and this is strong enough for electron–ion pairs
released in the primary ionization to gain sufficient energy to cause secondary
ionization. The rapid increase in amplification due to secondary ionization is called
a Townsend avalanche. The output signal at the anode is still proportional to the
134 CH4 EXPERIMENTAL METHODS
energy lost by the original particle. There are a number of different types of device
working in the proportional region and they are sometimes generically referred to
as track chambers or wire chambers.
The earliest detector using this idea was the proportional counter, which
consists of a cylindrical tube filled with gas (again a quenching component in
the gas is required) and maintained at a negative potential, and a fine central anode
wire at a positive potential. Again, neutrons can be detected indirectly by using the
direct nuclear reaction 3 Heðn; pÞ3 H mentioned in Section 4.3.2 in a proportional
chamber filled with a mixture of 3 He and krypton. Subsequently, the resolution of
proportional counters was greatly improved as a result of the discovery that if
many anode wires were arranged in a plane between a common pair of cathode
plates, each wire acts as an independent detector. This device is called a multiwire
proportional chamber (MWPC), and was introduced in 1968.11 An MWPC can
achieve spatial resolutions of 200 mm or less, and has a typical time resolution of
about 3 ns.
A schematic diagram of an MWPC is shown in Figure 4.13. The planes (a) have
anode wires into the page and those in plane (b) are at right angles. The wire
spacings are typically 2 mm. The cathodes are the faces of the chambers. A
positive voltage applied to the anode wires generates a field as shown in the upper
corner. A particle crossing the chamber ionizes the gas and the electrons drift along
the field lines to the anode wires. In this particular example, there would be signals
from one wire in the upper (a) chamber and two in the lower (a) chamber.
Figure 4.13 A group of three planes of an MWPC (from Po00 with kind permission of Springer
Science and Business Media)
11
The MWPC was invented by Georges Charpak and for this and other developments in particle detectors he
was awarded the 1992 Nobel Prize in Physics.
PARTICLE DETECTORS 135
Even better spatial resolutions are obtained in a related device called a drift
chamber, which has now largely replaced the MWPC as a general detector.12 This
uses the fact that the liberated electrons take time to drift from their point of
production to the anode. Thus the time delay between the passage of a charged
particle through the chamber and the creation of a pulse at the anode is related to
the distance between the particle trajectory and the anode wire. In practice,
additional wires are incorporated to provide a relatively constant electric field in
each cell in a direction transverse to normal incidence. A reference time has to be
defined, which, for example, could be done by allowing the particle to pass through
a scintillator positioned elsewhere in the experiment (scintillation counters are
discussed in Section 4.4.2). The electrons drift for a time and are then collected at
the anode, thus providing a signal that the particle has passed. If the drift time
can be measured accurately (to within a few ns) and if the drift velocity is known,
then spatial resolutions of 100–200 mm can easily be achieved, and specialized
detectors can reduce this still further.
Drift chambers are constructed in a variety of geometries to suit the nature of the
experiment, and arrangements where the wires are in planar, radial or cylindrical
configurations have all been used. The latter type is also called a ‘jet chamber’ and
a two-jet event in a jet chamber was shown in Figure 3.5 as evidence for the
existence of quarks.
One of the most advanced applications of proportional and drift chamber
principles is embodied in the time projection chamber (TPC) illustrated sche-
matically in Figure 4.14. This device consists of a cylindrical barrel, typically
Figure 4.14 Schematic diagram of a time projection chamber (TPC) (adapted from Kl86,
copyright Cambridge University Press)
12
In the new generation of colliders, drift chambers are largely being replaced by detectors based on silicon.
136 CH4 EXPERIMENTAL METHODS
Referring again to Figure 4.12, by increasing the external voltage still further one
moves into a region where the output signal ceases to be proportional to the
number of ion pairs produced and hence the incident energy. This is the region of
limited proportionality. In this region a type of gas detector called a streamer tube
operates, but this will not be discussed here. Eventually the process runs out of
control and we enter the Geiger–Müller region where the output signal is
independent of the energy lost by the incident particle. In this region a quenching
agent is not used. Detectors working in this region are called Geiger–Müller
counters. Physically they are similar to the simple cylindrical proportional counter
and are widely used as portable radiation monitors in the context of safety
regulations.
For completeness, we can mention that if the gas amplification factor is taken
beyond the Geiger–Müller region, the avalanche develops moving plasmas or
streamers. Recombination of ions then leads to visible light which can be made to
generate an electrical output. Eventually complete breakdown occurs and a spark
is emitted as the incident particle traverses the gas. Detectors in this region, called
streamer and spark chambers (these were of parallel plate construction, rather than
cylindrical), were widely used in the 1970s and 1980s and played an important role
in hadron physics, but are no longer in general use.
PARTICLE DETECTORS 137
For charged particles we have seen that energy losses occur due to excitation and
ionization of atomic electrons in the medium of the detector. In suitable materials,
called scintillators, a small fraction of the excitation energy re-emerges as visible
light (or sometimes in the UV region) during de-excitation. In a scintillation
counter this light passes down the scintillator and onto the face of a photodetector –
a device that converts a weak photon signal to a detectable electric impulse. An
important example of a photodetector is the photomultiplier tube, a schematic
diagram of which is shown in Figure 4.15.
Figure 4.15 Schematic diagram of the main elements of a photomultiplier tube (adapted from
Kr88, copyright John Wiley & Sons)
Electrons are emitted from the cathode of the photomultiplier by the photo-
electric effect and strike a series of focusing dynodes. These amplify the electrons
by secondary emission at each dynode and accelerate the particles to the next
stage. The final signal is extracted from the anode at the end of the tube. The
electronic pulse can be shorter than 10 ns if the scintillator has a short decay time.
The scintillation counter is thus an ideal timing device and it is widely used for
‘triggering’ other detectors, i.e. its signal is used to decide whether or not to
activate other parts of the detector, and whether to record information from the
event. Commonly used scintillators are inorganic single crystals (e.g. caesium
iodide) or organic liquids and plastics, and a modern complex detector in particle
physics may use several tons of detector in combination with thousands of
photomultiplier tubes.13 The robust and simple nature of the scintillation counter
13
For example, the Super Kamiokande experiment mentioned in Chapter 3, which detected neutrino
oscillations, although not using scintillation counters, has 13 000 photomultiplier tubes.
138 CH4 EXPERIMENTAL METHODS
has made it a mainstay of experimental nuclear and particle physics since the
earliest days of the subject.
Just as direct detection of neutrons is not possible by ionization methods, so the
same is true using scintillators. However, the
-particle and the 3 H nucleus from
the direct nuclear reaction 6 Liðn;
Þ3 H mentioned in Section 4.3.2 can produce
light in a LiI crystal scintillator and forms the basis for detecting neutrons with
energies up to about 20 MeV.
Solid-state detectors operate through the promotion of electrons from the valence
band of a solid to the conduction band as a result of the entry of the incident
particle into the solid. The resulting absence of an electron in the valence band
(a ‘hole’) behaves like a positron. Semiconductor detectors are essentially solid-
state ionization chambers with the electron–hole pairs playing the role of electron–
ion pairs in gas detectors. In the presence of an electric field, the electrons and
holes separate and collect at the electrodes, giving a signal proportional to the
energy loss of the incident charged particle. Most semiconductor detectors use
the principle of the junction diode. Since the band gap in some solids is as small as
1 eV and the energy loss required to produce a pair is only 3–4 eV on average
(cf. the 30 eV required in a gas detector), a very large number of electron–hole
pairs with only a small statistical fluctuation will be produced by a low-energy
particle. Solid-state detectors are therefore very useful in detecting low-energy
particles. Semiconductors (principally silicon or germanium) are used as a
compromise between materials that have residual conductivity sufficient to enable
conduction pulses due to single particles to be distinguished above background and
those in which the charges carriers are not rapidly trapped in impurities in the
material.
Such detectors have long been used in nuclear physics, where, for example, their
excellent energy resolution and linearity, plus their small size and consequent fast
response time, make them ideal detectors for -ray spectroscopy. Only recently
have thin planar detectors become important in particle physics, because of the
expense of covering large areas. Nevertheless, several square metres of semicon-
ductor detector are being planned for experiments at the LHC.
One example of a solid-state detector is a silicon microstrip detector, where
narrow strips of active detector are etched onto a thin slice of silicon, with gaps of
the order of l0 mm, to give a tiny analogue of an MWPC. Arrays of such strips
can then be used to form detectors with resolutions of the order of 5 mm. These
are often placed close to the interaction vertex in a colliding beam experiment,
with a view to studying events involving the decay of very short-lived particles.
Another example is the pixel detector. A single-plane strip detector only gives
position information in one dimension (orthogonal to the strip). A pixel detector
improves on this by giving information in two dimensions from a single plane.
PARTICLE DETECTORS 139
Methods of identifying particles are usually based on determining the mass of the
particle by a simultaneous measurement of its momentum together with some
other quantity. At low values of ¼ E=mc2 , measurements of the rate of energy
loss dE=dx can be used, while muons may be characterized by their unique
penetrating power in matter, as we have already seen. Here we concentrate on
methods based on measuring the velocity or energy, assuming always that the
momentum is known. We thus need to start with explaining how momenta are
measured.
Measurement of momentum
Time-of-flight
The simplest method, in principle, is to measure the time of flight between, for
example, two scintillation counters. If the distance between them is L, the time
difference for two particles of masses m1 and m2 travelling with velocities v1 and
v2 , is
L 1 1
t ¼ t2 t1 ¼ ; ð4:25Þ
c 1 2
m 2 ct
¼ : ð4:29Þ
m L
For example, using our previous values for L and t, Equation (4.29) shows that
the method is not useful for values of above about three, which corresponds to a
momentum of only about 3 GeV/c for nucleons. Of course this could be extended
by taking longer flight paths, but only at greater expense in instrumentation.
14
See Appendix B.
PARTICLE DETECTORS 141
C̆erenkov counters
The most important identification method for high-energy particles is based on the
erenkov effect. When a charged particle with velocity v traverses a dispersive
C
medium of refractive index n, excited atoms in the vicinity of the particle become
polarized, and if v is greater than the speed of light in the medium c=n, a part of the
excitation energy reappears as coherent radiation emitted at a characteristic angle
to the direction of motion. The necessary condition v > c=n implies n > 1 and by
considering how the waveform is produced15 it can be shown that cos ¼ 1=n for
the angle , where ¼ v=c as usual. A determination of is thus a direct
measurement of the velocity.16
erenkov radiation appears as a continuous spectrum and may be collected onto
C
a photosensitive detector. Its main limitation from the point of view of particle
detection is that very few photons are produced. The number of photons NðÞd
radiated per unit path length in a wavelength interval d can be shown to be
1 d 1 d
NðÞd ¼ 2
1 2 2 < 2
1 2 ð4:30Þ
n 2 n 2
and so vanishes rapidly as the refractive index approaches unity. The maximum
value occurs for ¼ 1, which for a particle with unit charge, corresponds to about
200 photons/cm in the visible region in water and glass. These numbers should be
compared with the 104 photons/cm emitted by a typical scintillator. Because the
yield is so small, appreciable lengths are needed to give enough photons, and gas
erenkov counters in particular can be several metres long.
C
erenkov counters are used in two different modes. The first is as a threshold
C
counter to detect the presence of particles whose velocities exceed some minimum
value. Suppose that two particles with values 1 and 2 at some given
momentum p are to be distinguished. If a medium can be found such that
1 n > 1 2 n, then particle 1 will produce C erenkov radiation but particle 2
will not. Clearly, to distinguish between highly relativistic particles with
1
also requires n 1, so that from Equation (4.30) very few photons are produced.
Nevertheless, common charged particles can be distinguished in this way up to at
least 30 GeV/c.
Another device is the so-called ring-image C erenkov detector which is a very
important device in both fixed-target machines and colliders. If we assume that the
particles are not all travelling parallel to a fixed axis, then the radiating medium
can be contained within two concentric spherical surfaces of radii R and 2R
centred on the target or interaction region where the particles are produced, as
illustrated in Figure 4.16. The outer surface is lined with a mirror, which focuses
15
This is Huygens’ construction in optics.
16 erenkov, Ilya Frank and Igor Tamm were
For the discovery and interpretation of this effect, Pavel C
awarded the 1958 Nobel Prize in Physics.
142 CH4 EXPERIMENTAL METHODS
Figure 4.16 Two particles P1 and P2 , produced from the target T, emit Cerenkov radiation on
traversing a medium contained between two spheres of radius R and 2R. The mirror M on the
outer sphere focuses the radiation into ring images at aa0 and bb0 on the inner detector sphere D.
The radii of the ring images depend on the angle of emission of the Cerenkov radiation and hence
on the velocities of the particles
erenkov radiation into a ring at the inner detector surface. The radius of this
the C
ring depends on the angle at which the C erenkov radiation is emitted, and hence
on the particle velocity v. It is determined by constructing an image of the ring
electronically. This was the technique used in the Super Kamiokande detector
discussed in Chapter 3 to detect relativistic electrons and muons produced by
neutrino interactions. In that experiment the radiating medium was pure water.
4.4.5 Calorimeters
Calorimeters are an important class of detector used for measuring the energy and
position of a particle by its total absorption and are widely used. They differ from
most other detectors in that (1) the nature of the particle is changed by the detector,
and (2) they can detect neutral as well as charged particles. A calorimeter may be a
homogeneous absorber/detector to detect photons and electrons. In early devices this
was often a block of lead glass, but is now more likely to be scintillator such as CsI.
Alternatively, it can be a sandwich construction with separate layers of absorber (e.g.
a metal such as lead) and detector (scintillator, MWPC etc.). The latter are also
PARTICLE DETECTORS 143
known as ‘sampling calorimeters’. During the absorption process, the particle will
interact with the material of the absorber, generating secondary particles which will
themselves generate further particles and so on, so that a cascade or shower,
develops. For this reason calorimeters are also called ‘shower counters’.
The shower is predominantly in the longitudinal direction due to momentum
conservation, but will be subject to some transverse spreading due both to multiple
Coulomb scattering and the transverse momentum of the produced particles.
Eventually all, or almost all, of the primary energy is deposited in the calorimeter,
and gives a signal in the detector part of the device.
There are several reasons why calorimeters are important, especially at high
energies:
the signal produced can be very fast, of the order of (10–100) ns, and is ideal for
making triggering decisions.
Although it is possible to build calorimeters that preferentially detect just one class
of particle (electrons and photons, or hadrons) it is also possible to design detectors
that serve both purposes. Since the characteristics of electromagnetic and hadronic
showers are somewhat different it is convenient to describe each separately. In
practice, it is common to have both types in one experiment with the hadron
calorimeter stacked behind the electromagnetic one.
Electromagnetic showers
When a high-energy electron or positron interacts with matter we have seen that
the dominant energy loss is due to bremsstrahlung, and for the photons produced
the dominant absorption process is pair production. Thus the initial electron will,
via these two processes, lead to a cascade of e pairs and photons, and this will
continue until the energies of the secondary electrons fall below the critical energy
EC where ionization losses equal those from bremsstrahlung. This energy is
roughly given by EC 600 MeV=Z.
Most of the correct qualitative features of shower development may be obtained
from the following very simple model. We assume:
each electron with E > EC travels one radiation length and then gives up half of
its energy to a bremsstrahlung photon;
144 CH4 EXPERIMENTAL METHODS
each photon with E > EC travels one radiation length and then creates an
electron–positron pair with each particle having half the energy of the photon;
electrons with E < EC cease to radiate and lose the rest of their energy by
collisions;
The main features of this simple model are observed experimentally, and in
particular the maximum shower depth increases only logarithmically with primary
energy. Because of this, the physical sizes of calorimeters need increase only
slowly with the maximum energies of the particles to be detected. The energy
resolution of a calorimeter, however, depends on statistical fluctuations, which are
neglected in this simple model, but for an electromagnetic calorimeter it is
1
typically E=E 0:05=E2, where E is measured in GeV.
Hadronic showers
Figure 4.18 The CDF detector at the pp collider at Fermilab, USA (Fermilab Graphic, reproduced
with permission)
The first is the pp Collider Detector at Fermilab (CDF), which is shown
schematically in Figure 4.18. The detection of the top quark and the measurement
of its mass were first made using this device. The dashed lines indicate some
particles produced in the collision. CDF is a large device, being approximately 8 m
wide and 26 m in overall length. The beams of protons and antiprotons enter from
each end through focusing quadrupole magnets and interact in the central
intersection region where there is a silicon vertex detector (1) to detect very
short-lived particles. The intersection point is surrounded by a 2000 tonne detector
system which, in addition to the vertex detector, consists of inner drift chambers
(2), electromagnetic calorimeters (4), hadron calorimeters (5) time-of-flight
detectors (not indicated) and further drift chambers (2) on the outside to detect
muons. The whole system is in a magnetic field with the solenoid coil shown at
(7) and steel shielding at (6). The rest of the detector consists of two symmetrical
sets of drift chambers (2) sandwiched between scintillation counters (3) and
magnetic toroids (8) to provide momentum measurements, primarily for muons.
The second example, shown schematically in Figure 4.19, is the ATLAS
detector currently under construction for use at the Large Hadron Collider
(LHC). It is hoped that this and other detectors at the LHC will be able to detect
the important Higgs boson, if it exists, and so help solve one the outstanding
current problems in particle physics – the origin of mass. The ATLAS detector is
even larger than the CDF detector and measures about 22 m high and 44 m long,
with an overall weight of approximately 7000 tonnes.
Figure 4.19 The ATLAS detector under construction for use at the pp collider LHC (also under construction) at CERN, Geneva, Switzerland (CERN photo,
reproduced with permission)
148 CH4 EXPERIMENTAL METHODS
Figure 4.20 The STAR detector at the RHIC accelerator at Brookhaven National Laboratory,
USA. (Courtesy of Brookhaven National Laboratory)
Finally, Figure 4.20 shows the STAR detector at the RHIC accelerator at
Brookhaven National Laboratory. This detects events resulting from the collisions
of heavy ions, typically those of fully-stripped gold nuclei, where the final state
may contain many thousands of particles. An example of an event is shown in
Figure 9.12.
Problems
4.1 At a collider, a 20 GeV electron beam collides with a 300 GeV proton beam at a
crossing angle of 10 . Evaluate the total centre-of-mass energy and calculate what
beam energy would be required in a fixed-target electron machine to achieve the
same total centre-of-mass energy.
4.2 What is the length L of the longest drift tube in a linac which operating at a
frequency of f ¼ 20 MHz is capable of accelerating 12 C ions to a maximum energy
of E ¼ 100 MeV?
4.3 Alpha particles are accelerated in a cyclotron operating with a magnetic field of
magnitude B ¼ 0:8 T. If the extracted beam has an energy of 12 MeV, calculate the
extraction radius and the orbital frequency of the beam (the so-called cyclotron
frequency).
PROBLEMS 149
4.4 Protons with momentum 50 GeV/c are deflected through a collimator slit 2 mm wide
by a bending magnet 1.5 m long which produces a field of 1.2 T. How far from the
magnet should the slit be placed so that it accepts particles with momenta in the
range 49–51 GeV/c?
4.5 Estimate the minimum length of a gas C erenkov counter that could be used in
threshold mode to distinguish between charged pions and charged kaons with
momentum 20 GeV/c. Assume that a minimum of 200 photons need to be radiated
to ensure a high probability of detection. Assume also that the radiation covers the
whole visible spectrum between 400 nm and 700 nm and neglect the variation with
wavelength of the refractive index of the gas.
4.7 What are the experimental signatures and with what detectors would one measure:
(a) the decay Z ! b
b, and (b) W ! e and W ! .
4.8 The reaction eþ e ! þ is studied using a collider with equal beam energies of
5 Gev. The differential cross-section is given by
d
2 h 2 c2
¼ 2
1 þ cos2
d 4Ecm
where Ecm is the total centre-of-mass energy and is the angle between the incoming
e and the outgoing . If the detector can only record an event if the þ pair
makes an angle of at least 30 relative to the beam line, what fraction of events will
be recorded? What is the total cross-section for this reaction in nanobarns? If the
reaction is recorded for 107 s at a luminosity of L ¼ 1031 cm2 s1 , how many events
are expected?
Suppose the detector is of cylindrical construction and at increasing radii from the
beam line there is a drift chamber, an electromagnetic calorimeter, a hadronic
calorimeter and finally muon chambers. If in a particular event the decays via
! þ þ and the þ decays to þ ! eþ þ þ e , what signals would
be observed in the various parts of the detector?
4.9 A charged particle with speed v moves in a medium of refractive index n. By considering
the wavefronts emitted at two different times, derive a relation for the angle of the
emitted C erenkov radiation relative to the particle’s direction in terms of ¼ v=c and n.
What is the maximum angle of emission and to what limit does it correspond?
If the momentum p of the particle is known from other detectors, show that the
mass squared x of the particle is given by x ¼ ðmc2 Þ2 ¼ p2 c2 ðn2 cos2 1Þ. If the
error on the momentum is negligible, show, by taking derivatives of this expression,
that for very relativistic particles, the standard error
x on x is approximately
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
x 2p2 c2 ðn2 1Þ
;
where
is the standard error on .
150 CH4 EXPERIMENTAL METHODS
4.10 Estimate the thickness of iron through which a beam of neutrinos with energy
300 GeV must travel if 1 in 109 of them is to interact. Assume that at high energies
the neutrino-nucleon total cross-section is given approximately by
1038 E cm2
where E is given in GeV. The density of iron is ¼ 7:9 g cm3 .
4.11 An electron with an initial energy of 2 GeV traverses 10 cm of water with a radiation
length of 36.1 cm. Calculate its final energy. How would the energy loss change if
the particle were a muon rather than an electron?
4.12 A beam of neutrons with kinetic energy 0.1 eV and intensity 106 s1 is incident
normally on a thin foil of 23592 U of effective thickness 10
1
kg m2 . The beam can
undergo (1) isotropic elastic scattering, with a cross section
el ¼ 3
102 b, (2)
radiative capture, with a cross-section
cap ¼ 102 b, or (3) it can fission a 235 92 U
nucleus, with a cross-section
f ¼ 3
102 b. Calculate the attenuation of the beam
and the flux of elastically-scattered particles 5 m from the foil.
4.13 A positron with laboratory energy 50 GeV interacts with the atomic electrons in a
lead target to produce þ pairs. If the cross-section for this process is given by
¼ 4
2 h2 c2 =3ðECM Þ2 , calculate the positron’s interaction length. The density of
lead is ¼ 1:14
107 kg m3 .
4.14 A liquid hydrogen target of volume 125 cm3 and density 0:071 g cm3 is bombarded
with a mono-energetic beam of negative pions with a flux 2
107 m2 s1 and the
reaction þ p ! 0 þ n observed by detecting the photons from the decay of the
0 . Calculate the number of photons emitted from the target per second if the cross-
section is 40 mb.
4.15 Assuming the Bethe–Bloch formula is valid for low energies, show that the rate of
ionization has a maximum (the Bragg peak) and find the kinetic energy of protons in
iron for which this maximum would occur.
4.16 A cylindrical proportional chamber has a central anode wire of radius 0.02 mm and
an outer cathode of radius 10 mm with a voltage of 500 V applied between them.
What is the electric field at the surface of the anode? If the threshold for ionization
by collision is 750 kV m1 and the mean free path of the particles being detected is
4
106 m, estimate the number of ion pairs produced per primary particle.
5
Quark Dynamics:
the Strong Interaction
5.1 Colour
We saw in Chapter 3 that the quark model account of the hadron spectrum is very
successful. However, it begs several questions. One is: why are the observed states
overwhelmingly of the form 3q, 3q and qq? Another arises from a particular
assumption that was implicitly made in Chapter 3. This is: if two quarks of the
same flavour uu, dd, ss . . . are in the same spatial state, they must also be in the
same spin state, with their spins parallel. This can be seen very easily by
considering the baryon state omega-minus that is shown in Table 3.3 and
Figure 3.12.1 From its decay products, it may be deduced that this state has
strangeness S ¼ 3 and spin J ¼ 32 and thus in the quark model it has the
composition ¼ sss, where all three quarks have their spins parallel and there
is no orbital angular momentum between them. This means that all three like-
quarks have the same space and spin states, i.e. the overall wavefunction must be
symmetric, which violates the fundamental requirement of the Pauli principle. The
latter states that a system of identical fermions has a wavefunction that is overall
antisymmetric under the interchange of any two particles, because identical
1
The discovery of the was a crucial step in gaining acceptance of the quark model of hadron
spectroscopy. The experiment is described in Chapter 15 of Tr75.
fermions cannot simultaneously be in the same quantum state. The three s quarks
in the therefore cannot be in the same state. So how do they differ?
The is an obvious example of the contradiction, but it turns out that in order
for the predictions of the quark model to agree with the observed spectrum of
hadron multiplets, it is necessary to assume that overall baryon wavefunctions are
symmetric under the interchange of like quarks.2 In order to resolve this contra-
diction, it is necessary to assume that a new degree of freedom exists for quarks,
but not leptons, which is somewhat whimsically called colour. The basic properties
of colour are as follows.
1. Any quark u, d, s, . . . can exist in three different colour states.3 We shall see
later that there is direct experimental evidence that just three such states exist,
which we denote r, g, b for ‘red’, ‘green’ and ‘blue’ respectively.
3. Only states with zero values for the colour charges are observable as free
particles; these are called colour singlets. This is the hypothesis of colour
confinement. It can be derived, at least approximately, from the theory of strong
interactions we shall describe.
Table 5.1 Values of the colour charges I3C and Y C for the
colour states of quarks and antiquarks
(a) Quarks (b) Antiquarks
I3C YC I3C YC
1 1
r 2 3 r 12 13
g 12 1
3
g 1
2 13
b 0 23
b 0 2
3
2
In Problem 3.4 it was shown explicitly that otherwise the wrong hadron spectrum is predicted.
3
Needless to say, nothing to do with ‘real’ colour!
4
This is one reason we were careful to use the qualifier ‘electric’ when talking about charge in the context of
electromagnetic interactions in earlier chapters.
QUANTUM CHROMODYNAMICS (QCD) 153
Returning to the quark model, it can be seen from Table 5.1 that a 3q state
can only have both I3C ¼ 0 and Y C ¼ 0 if we have one quark in an r state, one in a
g state and one in a b state. Hence in the , for example, all three s quarks
are necessarily in different colour states, and thus the Pauli principle can be
satisfied. Formally, we are assuming that the total wavefunction is the product of a
spatial part spatial ðxÞ and a spin part spin , as usual, but also a colour wavefunction
colour , i.e.
1
colour ¼ pffiffiffi ½R1 G2 B3 þ G1 B2 R3 þ B1 R2 G3 R1 B2 G3 B1 G2 R3 G1 R2 B3 ;
6
ð5:2Þ
where R, G and B represent quarks with colour red, green and blue, respectively.
One can also see from Table 5.1 part of the answer to the first question of this
section. Free quarks and fractionally charged combinations like qq and qqq are
forbidden by colour confinement, in accordance with experimental observation. On
the other hand, the combinations qq and 3q used in the simple quark model are
allowed. More unusual combinations like qqqq and qqqqq, which could give rise to
so-called ‘exotic’ mesons and baryons, respectively, are not in principle forbidden
by colour confinement and, as mentioned in Chapter 3, recent experiments may
possibly have provided some evidence for a small number of these, but this has yet
to be confirmed.
Gluons, the force carriers of the strong interaction, have zero electric charge,
like photons, but unlike photons, which couple to electric charge, gluons couple to
colour charges. This leads immediately to the flavour independence of strong
interactions discussed in Chapter 3; that is, the different quark flavours a ¼ u,
d, s, c, b and t have identical strong interactions. We now see that this is because
they are postulated to exist in the same three colour states r, g, b, with the
same possible values of the colour charges. Flavour independence has its most
striking consequences for u and d quarks, which have almost equal masses, where
it leads to the phenomenon of isospin symmetry. This results, among other
things, in the near equality of the masses of the proton and neutron, and charge
states within other multiplets such as pions and kaons, all of which we have
seen in Chapter 3 are confirmed by experiment. We will examine the con-
sequence of flavour independence for the bound states of the heavy quarks c
and b in Section 5.3.
Although QED and QCD both describe interactions, albeit of very different
strengths, that are mediated by massless spin-1 bosons which couple to conserved
charges, there is a crucial difference between them that profoundly effects
the characters of the resulting forces. While the photons which couple to the
electric charge are themselves electrically neutral, gluons have non-zero values of
the colour charges to which they couple. This is illustrated in Figure 5.1, which
shows a particular example of a quark–quark interaction by gluon exchange.
Figure 5.1 Example of quark--quark scattering by gluon exchange; in this diagram, the quark
flavours u and s are unchanged, but their colour states can change, as shown
In this diagram, the colour states of the two quarks are interchanged, and the
gluon has non-zero values of the colour quantum numbers, whose values follow
from colour charge conservation at the vertices, i.e.
1
I3C ðgÞ ¼ I3C ðrÞ I3C ðbÞ ¼ ð5:3Þ
2
and
Figure 5.2 The two lowest-order contributions to gluon--gluon scattering in QCD: (a) one-
gluon exchange, (b) contact interaction
Just as quarks can exist in three colour states, gluons can exist in eight colour
states, although we will not need the details of these. The first thing implied by the
non-zero values of the colour charges is that gluons, like quarks, are confined and
cannot be observed as free particles. The second is that since gluons couple to
particles with non-zero colour charges, and since gluons themselves have non-zero
colour charges, it follows that gluons couple to other gluons. The two types of
gluon self-coupling that occur in QCD are given in Figure 5.2, which shows the
two lowest-order contributions to gluon–gluon scattering.
The first is a gluon exchange process in analogy to gluon exchange in quark–
quark scattering, which we have encountered previously (see Figure 1.3), while the
second involves a so-called ‘zero range’ or ‘contact’ interaction. If the forces
resulting from these interactions were attractive and sufficiently strong, they could
in principle lead to bound states of two or more gluons. These would be a new type
of exotic state called glueballs. Although some experiments claimed to have
detected these, at present there is little compelling evidence that they exist.5
The gluon–gluon interactions have no analogue in QED (photons couple to
electrically charged particles and hence do not couple directly to other photons)
and it can be shown that they lead to properties of the strong interaction that differ
markedly from those of the electromagnetic interaction. These properties are
colour confinement, which we have discussed above, and a new property called
asymptotic freedom. The latter is the statement that the strong interaction gets
weaker at short distances; conversely, as the distance between the quarks increases,
the interaction gets stronger.6 In this strong interaction regime the situation is very
complicated, and it has not yet been possible to evaluate the theory precisely. We
therefore have to rely on results obtained by numerical simulations of the theory;
the approach is called lattice gauge theory. In these simulations, the theory is
5
A critical review is given in Ei04.
6
Asymptotic freedom was postulated in 1973 by David Gross, David Politzer and Frank Wilczek, who were
subsequently awarded the 2004 Nobel Prize in Physics.
156 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
7
Lattice calculations also support the view that gluon–gluon forces are strong enough to give rise to
glueballs.
8
The rather clumsy notation is because it was discovered independently by two groups, led by Burton Richer
and Samuel Ting. Richer’s group was studying the reactions eþ e ! hadrons and named it the particle.
Ting’s group discovered it in pBe reactions and called it the J. It is now known as the J= . Richer and Ting
shared the 1976 Nobel Prize in Physics for the discovery.
HEAVY QUARK BOUND STATES 157
Figure 5.3 Quark diagrams for (a) the decay of a charmonium state to a pair of charmed
mesons, and (b) an example of a decay to non-charmed mesons
MJ= < 2MD , where MD is the mass of the lightest meson having non-zero charm,
the Dð1870Þ. (These latter states had already been seen in neutrino experiments,
but not clearly identified.) The mass 2MD is referred to as the charm threshold.
Since the direct decay to charmed mesons is forbidden, the only hadronic decays
allowed must proceed via mechanisms such as that of Figure 5.3(b) and diagrams
like this where initial and final quark lines are disconnected are known to be
heavily suppressed.9
The explanation for this in QCD is that since both the decaying particle and
the three pions in the final state are colour singlets, they can only be connected by
the exchange of a combination of gluons that is also a colour singlet, i.e. not the
exchange of a single gluon. Moreover, the J= ð3097Þ is known to be produced in
eþ e annihilations via photon exchange, so it must have a charge conjugation
C ¼ 1. Thus the minimum number of gluons exchanged is three. This is
illustrated in Figure 5.4. In contrast, if M > 2MD then the decay may proceed
via the exchange of low-momentum gluons as usual.
Subsequently, higher-mass charmonium states also with J PC ¼ 1 , where
P ¼ ð1ÞLþ1 and C ¼ ð1ÞLþS , were discovered in eþ e reactions and states
with other J PC values were identified in their radiative decays. Thus the n ¼ 1; 1 S0
Figure 5.4 OZI-suppressed decay of a charmonium state below the DD threshold
9
This is known as the OZI Rule after Okubo, Zweig and Iizuka who first formulated it. Another example
.
where it acts is the suppression of the decay ! þ 0 compared with ! K K
158 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
ð3686Þ ! c ð2980Þ þ
and J= ð3097Þ ! c ð2980Þ þ
ð5:5Þ
ð3686Þ ! ci þ
ð5:6Þ
The latter themselves decay and from an analysis of their decay products they are
identified with the n ¼ 1 states 3 P0 , 3 P1 and 3 P2 . Some of these states lie below the
charm threshold and like the J= ð3097Þ are forbidden by energy conservation to
decay to final states with ‘open’ charm and thus have widths measured in keV.
Others lie above the charm threshold and therefore have ‘normal’ widths measure
in MeV. The present experimental situation for charmonium states with L 2 is
shown in Table 5.2.
of the 3 S1 ground states. There is a striking similarity in the levels of the two
systems, which suggests that the forces in the cc and bb are flavour independent, as
discussed in Chapter 3 and now seen to follow from the postulates of QCD. The
level structure is also very similar to that seen in positronium which suggests that,
as in positronium, there is a major contribution from a single-particle exchange
with the Coulomb-like form. In fact at short interquark distances r < 0:1 fm, the
interaction is dominated by one-gluon exchange that we can write as
a
VðrÞ ¼ ; ð5:7Þ
r
where a is proportional to the strong interaction analogue of the fine structure
constant in QED. Because of asymptotic freedom, the strength of the interaction
decreases with decreasing r, but for r < 0:1 fm this variation is slight and can in
many applications be neglected.10
In strong interactions we also have to take account of the fact that the quarks are
confined. The latter part of the potential cannot at present be calculated from QCD
and several forms are used in phenomenological applications. All reasonable forms
are found to give very similar results for the region of interest. If we choose a
linear form, then
VðrÞ b r: ð5:8Þ
This is an example of a confining potential, in that it does not die away with
increasing separation and the force between the quark and antiquark cannot be
neglected, even when they are very far apart. The full potential is thus
a
VðrÞ ¼ þ br: ð5:9Þ
r
If the form (5.9) is used in the Schrödinger equation for the cc and bb systems,
taking account of their different masses, it is found that a good fit to both sets of
energy levels can be obtained for the same values a
0:48 and b
0:18 GeV2 ,
which confirms the flavour independence of the strong interaction and is evidence
for QCD and the standard model.
10
The equivalent coupling in QED also varies with distance, but the variation is very small and can usually be
neglected.
THE STRONG COUPLING CONSTANT AND ASYMPTOTIC FREEDOM 161
Figure 5.6 The running coupling constant s corresponding to four flavours and a scale
parameter ¼ 0:2 0:1 GeV/c; the dashed, solid and dot-dashed curves correspond to
¼ 0:1 ; 0:2 and 0:3, respectively
162 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
variation with Q2 is small at large Q2 and over limited Q2 regions it can often be
neglected. In this large Q2 region, the coupling is sufficiently weak that calcula-
tions can be performed with reasonable accuracy by retaining only diagrams of
lowest and next-to-lowest order; and sometimes the short-range strong interaction
can be neglected to a first approximation, as we shall see.
Although there are other forces that increase with increasing separation (for
example, the force between two particles connected by a spring or elastic string),
the difference between those and the present case is that in the former cases
eventually something happens (for example, the string breaks) so that the particles
(or the ends of the string) become free. This does not happen with the strong force.
Instead, the energy stored in the colour field increases until it becomes sufficiently
large to create qq pairs and eventually combinations of these will appear as physical
hadrons. This latter process is called fragmentation and is rather poorly understood.
The behaviour of the strong interaction as a function of distance (or equivalently
momentum transfer) is so unlike the behaviour of other forces we are familiar with
(e.g. gravity and electromagnetism) that it is worth looking at why this is.
In QED, single electrons are considered to emit and reabsorb photons con-
tinually, as shown in Figure 5.7(a). Such a process is an example of a so-called
quantum fluctuation, i.e. one particle converting to two or more particles for a
finite time. This is allowed provided the time and the implied violation of energy
conservation are compatible with the uncertainty principle. Of course if another
electron is nearby, then it may absorb the photon and we have the usual one-photon
exchange scattering process of Figure 5.7(b).
Figure 5.7 (a) The simplest quantum fluctuation of an electron, and (b) the associated
exchange process
The emitted photon may itself be subject to quantum fluctuations, leading to more
complicated diagrams like those shown in Figure 5.8(a). Thus the initial electron
emits not only photons, but also indirectly electron–positron pairs. These are
referred to as a ‘sea’ of virtual electrons (cf. comments in Chapter 3 in the context
of the quark model). The equivalent contribution to elastic electron–electron
scattering is shown in Figure 5.8(b).
THE STRONG COUPLING CONSTANT AND ASYMPTOTIC FREEDOM 163
Figure 5.8 (a) A more complicated quantum fluctuation of the electron, and (b) the
associated exchange process
for
11
The name arises from the analogy of placing a charge in a dielectric medium. This aligns the particles of
the medium and produces a net polarization.
164 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
Figure 5.9 The two lowest-order vacuum polarization corrections to one-gluon exchange in
quark--quark scattering
discussed basic properties of quarks and gluons interactions earlier in this chapter.
They have been extensively studied in the reaction
eþ þ e ! hadrons ð5:18Þ
The fragmentation process that converts the quarks into hadrons is very
complicated, and the composition of the jets – i.e. the numbers and types of
particles in the jet and their momenta – varies from event to event. However, the
direction of a jet, defined by the total momentum vector
X
P¼ pi ; ð5:19Þ
i
where the sum extends over all the particles within the jet, closely reflects the
parent quark or antiquark direction. This is because the QCD interaction is
relatively weak at very short distances (asymptotic freedom), and the quark and
antiquark do not interact strongly until they are separated by a distance r of order
1 fm. At these relatively large distances, only comparatively small momenta can be
transferred, and hence the jets that subsequently develop point almost exactly in
the initial quark and antiquark directions. That is, the jet angular distribution
relative to the electron beam direction reflects the angular distributions of the
quark and antiquark in the basic reaction eþ þ e ! q þ q. The latter can easily
166 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
Events like these provided the first unambiguous evidence for gluons, because
the angular distributions of the jets are found to be in good agreement with the
theoretical expectation for spin-1 gluons, but are inconsistent with what would
be expected if, for example, the third jet originated from a particle of spin-0. The
ratio of three-jet to two-jet events can also be calculated, assuming that the third jet is
a gluon, because the probability that a quark or antiquark will emit a gluon is
determined by the strong coupling s , in the same way that the probability that an
electron or positron will emit a photon is determined by the fine structure constant .
This leads to a value of s and hence , the QCD scale parameter. The value obtained
is consistent with Equation (5.12) found from other determinations and lends further
support for the whole picture of quarks interacting via the exchange of gluons.
ðeþ e ! hadronsÞ
R ð5:20Þ
ðeþ e !
þ
Þ
is almost energy independent. The near constancy of this ratio follows from the
dominance of the two-step mechanism of Figure 5.10, with the total annihilation
rate being determined by that of the initial reaction eþ e ! q þ q. The value of
the ratio R then directly confirms the existence of three colour states, each with the
same electric charge, for each quark flavour.
To understand this, let us suppose that each quark flavour f ¼ u, d, s . . . exists
in NC colour states, so that NC ¼ 3 according to QCD, while NC ¼ 1 if the colour
degree of freedom does not exist. Since the different colour states all have the same
electric charge, they will all be produced equally readily by the mechanism of
Figure 5.10, and the rate for producing quark pairs of any given flavour f ¼ u, d,
s, . . . will be proportional to the number of colours NC . The cross-section is also
proportional to the squared charge of the produced pair (because this is a first-order
electromagnetic process), and since muon pairs are produced by an identical
mechanism, we obtain
When the small contribution from the three-jet events and other corrections of
order s are taken into account, this expression for R is modified to
R ¼ R0 ð1 þ s =Þ; ð5:23Þ
12
The cross-section for the production of muon pairs is essentially a purely electromagnetic one, except at
very high energies where the effect of the weak interaction may be seen. This will be discussed in Chapter 6.
13
There is no contribution from the top quark because it is too heavy to be produced, even at the high
energies we are considering.
168 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
Figure 5.12 Measured values of the cross-section ratio R and the theoretical prediction from
QCD for NC ¼ 3 colours; the dashed line shows the prediction without QCD corrections
14
The pioneering work on deep inelastic scattering done by Jerome Friedman, Henry Kendall and Richard
Taylor resulted in their receiving the 1990 Nobel Prize in Physics.
DEEP INELASTIC SCATTERING AND NUCLEON STRUCTURE 169
2M W 2 c2 þ Q2 M 2 c2 ð5:24Þ
x Q2 =2M: ð5:25Þ
Here, M is the proton mass, W is the invariant mass of the final-state hadrons and
Q2 is the squared energy–momentum transfer
Q2 ¼ ðE E0 Þ2 =c2 ðp p0 Þ2 : ð5:26Þ
The physical interpretation of x will be discussed below. In the rest frame of the
initial proton, reduces to
¼ E E0 ð5:27Þ
and so is the Lorentz-invariant generalization for the energy transferred from the
lepton to the proton.
In Chapter 2 we discussed several modifications to the formalism for describing
the structure of nuclei obtained from scattering experiments. Here we are dealing
with high-energy projectiles and so we will need to take all those corrections into
account. In particular, the magnetic interaction introduces a second form factor.
(cf Equation (2.14)). The two form factors, denoted W1 and W2 , are called structure
170 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
where is the lepton scattering angle. For values of W 2:5 GeV=c2 , the cross-
sections show considerable structure due to the excitation of nucleon resonances, but
above this mass they are smoothly varying. In the latter region, the values of the
structure functions can be extracted from the data by choosing suitable parameter-
izations and fitting the available data in an analogous way to the way charge
distributions of nuclei were deduced in Chapter 2.
Rather than W1 and W2 , it is usual to work with two related dimensionless
structure functions defined by
It is a remarkable fact that at fixed values of x the structure functions have only a
very weak dependence on Q2 . This behaviour is referred to as scaling and is
illustrated in Figure 5.14. As the Fourier transform of a spherically symmetric
point-like distribution is a constant, we conclude that the proton has a sub-structure
of point-like charge constituents.
The interpretation of scaling is simplest in a reference frame where the target
nucleon is moving with a very high velocity, so that the transverse momenta and
Figure 5.14 The structure function F2 of the proton as a function of x, for Q2 between 2 and
18 ðGeV=cÞ2 (reproduced from At82 with kind permission of Springer Science and Business
Media)
DEEP INELASTIC SCATTERING AND NUCLEON STRUCTURE 171
rest masses of its constituents may be neglected. The structure of the nucleon is
then given by the longitudinal momentum of its constituents. This approach was
first adopted by Feynman and Bjorken, who called the constituents partons. (We
now identify charged partons with quarks and neutral partons with gluons.) In
the parton model, deep inelastic scattering is visualized as shown in Figure 5.15.
The target nucleon is a stream of partons each with four-momentum xP, where
P ¼ ðp; pÞ is the four-momentum of the nucleon and p ¼ jpj, is its (very large)
three-momentum, so that the nucleon mass may be neglected.
Suppose now that one parton of mass m is scattered elastically by the exchanged
photon of four-momentum Q. Then
If x2 P2 ¼ x2 M 2 c4 Q2 , then
Q2 Q2
x¼ ¼ ; ð5:31Þ
2P Q 2M
where the invariant scalar product has been evaluated in the laboratory frame
in which the energy transfer is and the nucleon is at rest. This is our previous
definition Equation (5.25). Thus, the physical interpretation of x is the fractional
three-momentum of the parton in the reference frame where the nucleon has
a very high velocity. This is equivalent to having a parton of mass m stationary
in the laboratory system, with the elastic relation Q2 ¼ 2m. So provided
Q2 M 2 ,
Q2 m
x¼ ¼ ; ð5:32Þ
2M M
i.e. x may also be interpreted as the fraction of the nucleon mass carried by the
struck parton.
172 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
To identify the constituent partons with quarks we need to know their spins and
charges. For the spin, it can be shown that
and
The latter relation, known as the Callan–Gross relation, follows by comparing the
coefficients in the equation for the double differential cross-section Equation (5.28)
with that in Chapter 2 (Equation (2.14)). This gives
where ¼ Q2 =4m2 c2 and m is the mass of the target, in this case the mass of the
struck parton. Replacing W1 by F1 =Mc2 and W2 by F2 =, gives
F1 Q2
2
¼ 2 2 ð5:36Þ
Mc F2 4m c
and since now Q2 ¼ 2m, we have m ¼ Q2 =2 ¼ xM. Finally, using this mass in
Equation(5.36) yields the Callan–Gross relation. Figure 5.16 shows some results
for the ratio 2xF1 =F2 . It is clear that spin-12 is strongly favoured.
To deduce the parton charges is more complicated. We will assume that the
constituent partons are quarks and show that this is consistent with experimental
data. We start by defining qf ðxÞ to be the momentum distribution of a quark of
flavour f, i.e. qf ðxÞdx is the probability of finding in a nucleon a quark of flavour f,
with momentum fraction in the interval x to x þ dx. A given nucleon will consist of
a combination of valence quarks (i.e. those that give rise to the observed quantum
numbers in the quark model) and additional quark–antiquark pairs that are
continually produced and annihilated by the radiation of virtual gluons by the
quarks.15 (Recall the discussion of quantum fluctuations in electrodynamics in
Section 5.4.) Thus, in general, a structure function can be written as the sum of
contributions from quarks and antiquarks of all flavours. Also, from the cross-
section formula Equation (5.28), we would expect the structure functions to
involve the quark distributions weighted by the squares of the quark charges zf
(in units of e) for a given quark flavour f.
15
These are the ‘sea’ quarks referred to in the discussion of the static quark model in Chapter 3.
DEEP INELASTIC SCATTERING AND NUCLEON STRUCTURE 173
and
1 n n 4 n 1 n
F2‘n ðxÞ n n
¼ x ðd þ d Þ þ ðu þ u Þ þ ðs þ s Þ ; ð5:38bÞ
9 9 9
where, for example, un;p is the distribution of u quarks in the neutron and proton.
Using isospin symmetry, interchanging u and d quarks changes neutron to proton,
174 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
and
with similar relations for the antiquarks. Then if we work with a target nucleus
with equal numbers of protons and neutrons (an isoscalar target), its structure
function will have the approximate form (neglecting purely nuclear effects)
1 5 X 1
F2‘N ðxÞ ¼ ½F2‘p ðxÞ þ F2‘n ðxÞ ¼ x ½qðxÞ þ qðxÞ þ x½sðxÞ þ sðxÞ : ð5:40Þ
2 18 q¼d;u 9
The second term is small because s quarks are only present in the sea component at
the level of a few percent. Thus the mean squared value of the charges of the u and
5
d quarks is approximately 18 .
The final step is to extract information from deep inelastic scattering using
neutrinos and antineutrinos as projectiles. This is more complicated because, as we
shall see in Chapter 6, neutrinos and antineutrinos couple differently to the different
quarks and antiquarks and there is also a third form factor involved. Without proof,
we shall just quote the result:
X
F2N ðxÞ ¼ x ½qðxÞ þ qðxÞ : ð5:41Þ
q¼d;u
There is no electric charge factor outside the summation because, just as quarks form
strong interaction isospin multiplets with different electric charges, the leptons also
form weak isospin multiplets, but in this case the resulting weak charge is the same
for all quarks.16
From Equation (5.40) and (5.41), we expect
18 ‘N
F2N ðxÞ F ðxÞ: ð5:42Þ
5 2
The experimental data illustrated in Figure 5.17 show that F2‘N ðxÞ and F2N ðxÞ are
equal within errors except possibly at small values of x where antiquarks are more
important. Thus one can conclude that the partons do have charges 23 and 13, which
completes the evidence for identifying partons with quarks.
16
Weak isospin will be discussed briefly in Chapter 6.
DEEP INELASTIC SCATTERING AND NUCLEON STRUCTURE 175
Figure 5.17 Comparison of F2 ðxÞ from deep inelastic muon (data from Ar97) and neutrino
(data from Se97) scattering experiments; the data points are the average over a range of
Q2 > 2 ðGeV=cÞ2 and the error bars express the range of data values within the Q2 ranges
Combining data from different experiments, with electrons, muons, neutrinos and
antineutrinos as projectiles, enables individual quark/parton momentum distribu-
tions to be extracted from combinations of cross-sections. Some typical results at
Q2 ¼ 10 ðGeV=cÞ2 are shown in Figure 5.18 for the combinations
and
The difference
ðxÞ
Qv ðxÞ QðxÞ Q ð5:44Þ
can be identified as the distribution of the valence quarks of the quark model. It can
be seen that Qv is concentrated around x
0:2 and dominates except at small
values of x where the antiquarks q in the sea distribution are important.
The results of Figure 5.18 reveal an interesting and unexpected result concerning
gluons within the nucleon. If we integrate the momentum distributions for quarks
and antiquarks over all x we might expect to recover the total momentum of the
nucleon, whereas the curves of Figure 5.18 yield a value of approximately 0.5.
Thus it follows that about 50 per cent of the momentum is carried by gluons.
Although scaling is approximately correct, it is certainly not exact. In Fig-
ure 5.19 we show some deep inelastic scattering data plotted in more detail. The
176 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
deviations from scaling are due to QCD corrections to the simple quark model, i.e.
the quark in the proton that is struck by the exchanged photon can itself radiate
gluons. Again, without further details, the scaling violations are explained by QCD
using a value of the strong interaction parameter that is consistent with that
obtained from other sources (e.g. the three-jet events that we have discussed
above).17
Finally, it is worth noting that the nucleon structure functions and hence the
quark densities are found from lepton scattering experiments using a range of
different nuclear targets. We have seen that the average binding energy of nucleons
in heavy nuclei is of the order of 7–8 MeV per nucleon. As this energy is much
smaller than those used in deep inelastic scattering experiments, it might be
thought safe to ignore nuclear effects (except those due to the internal motion of
the nucleons – the Fermi momentum – that are typically about 200 MeV/c).
However, experiments have shown that the structure functions do in fact depend
17
Scaling violations are discussed in detail, but at a more advanced level than here in, for example Ha84.
PROBLEMS 177
Publisher's Note:
Permission to reproduce this image
online was not granted by the
copyright holder. Readers are kindly
requested to refer to the printed version
of this chapter.
Figure 5.19 A compilation of values of F2 measured in deep inelastic electron and muon
scattering from a deuterium target -- different symbols denote different experiments; for clarity,
the data at different values of x have been multiplied by the factors shown in brackets and the
solid line is a QCD fit with ¼ 0:2 GeV (adapted from Mo94, copyright the American Physical
Society)
slightly on the nuclear medium. Although the effects are very small and not
enough to alter the conclusions of this chapter, it is a reminder that there are still
things to be learnt about the role of nuclear matter and that this may hold
information on the nuclear force in terms of the fundamental quark–gluon
interaction.
Problems
5.1 The general combination of m quarks and n antiquarks qm q n , with baryon number
B > 0 has a colour wavefunction that may be written r g b
r g
b , where r
means that there are quarks in the r colour state, etc.. By imposing the condition of
colour confinement, show that m n ¼ 3p, where p is a non-negative integer and
hence show that states with the structure qq are not allowed.
178 CH5 QUARK DYNAMICS: THE STRONG INTERACTION
5.2 Draw the lowest-order Feynman diagrams for the following processes:
(a) the interaction of a quark and a gluon to produce a quark and a photon;
5.3 A pp collider with equal beam energies is used to produce a pair of top quarks. Draw
a Feynman diagram for this process that involves a single gluon. If the three quarks
of the proton (or antiproton) carry between them 50 per cent of the hadron total
energy–momentum, calculate the minimum beam momentum required to produce
the t t pair.
5.4 The lowest Feynman diagram for inelastic electron–proton scattering at high
energies
e ðE; pcÞ þ pðEP ; Pp cÞ ! e ðE0 ; p0 cÞ þ XðhadronsÞ
5.5 If hadron–hadron total cross-sections are assumed to be the sum of the cross-sections
between their constituent quarks, show that the quark model predicts the
relationship:
5.6 The 3
decay of positronium (the bound state of eþ e ) has a width that in QED is
predicted to be ð3
Þ ¼ 2ð2 9Þ6 me c2 =9, where is the fine structure constant.
If the hadronic decay of the cc bound state J=ð3100Þ proceeds via an analogous
PROBLEMS 179
mechanism, but involving three gluons, use the experimental hadronic width
ð3gÞ ¼ 80 keV to estimate the strong interaction coupling constant s . Use an
analogous assumption to estimate s from the radiative width ðgg
Þ ¼ 0:16 keV of
the bb bound state ð9460Þ.
5.7 Use Equations (5.38) and (5.39) to derive the Gottfried sum rule,
ð1 ð1
dx 1 2
½F2ep ðxÞ F2en ðxÞ ¼ þ ½ uðxÞ dðxÞ dx;
x 3 3
0 0
5.9 Common forms assumed for the momentum distributions of valence quarks in the
proton are:
If the valence quarks account for half the proton’s momentum, find the values of a
and b.
5.10 The cross-section ðud ! W þ Þ near the mass of the W þ is given by the Breit–
Wigner form
hcÞ2
ð 2 ud
¼ ;
3½4ðE MW c2 Þ2 þ 2
where ðMW ; Þ are the mass and total width of the W þ , ud is the partial width for
W þ ! ud , E is the total centre-of-mass energy of the ud pair and ¼ 2=E. Find the
maximum value of , i.e. max , given that the branching ratio for W þ ! ud is 1=3.
Use this result and the quark distributions of Question 5.9 to find an expression for
pffiffi cross-section ðp
the p ! W þ þ Þ in terms
pffiffi of the p
p total centre-of-mass energy
s and max and evaluate your result for s ¼ 1 TeV. (Use the narrow width, i.e.
delta function, approximation
!
W E2
ud ðEÞ ¼ max 1
MW c2 ðMW c2 Þ2
in integrals.)
6
Electroweak Interactions
Figure 6.1 Feynman diagram for the weak neutral current reaction þ N ! þ X
of the masses of the W and Z 0 bosons prior to the long-awaited detection of these
particles in 1983. In general, the theory is in agreement with all data on both weak
and electromagnetic interactions, which are now referred to collectively as the
electroweak interaction, in the same way that electric and magnetic interactions
are referred to collectively as electromagnetic interactions. Furthermore, the theory
predicts the existence of a new spin-0 boson, the so-called Higgs boson, which is
associated with the origin of particle masses within the model. This was mentioned
in passing in earlier chapters. Although a detailed discussion of the Higgs boson is
beyond the scope of this book, there is a brief discussion of the role of this very
important particle in Chapter 9.
The new unification only becomes manifest at high energies, and at low energies
weak and electromagnetic interactions can still be clearly separated. This follows
from the general form of the amplitude Equation (1.41):
g2 h2
Fðq2 Þ ¼ ; ð6:1Þ
jqj2 þ MX2 c2
where MX2 is the mass of the exchanged particle and g is the appropriate coupling.
For weak interactions, MX ¼ MW;Z 80 GeV=c2 and for the electromagnetic
interaction MX ¼ M ¼ 0. Thus, even with gweak
gem , the amplitudes for the
two interactions will only become of comparable size for jqj2
MX2 c2, i.e. at high
energies. We therefore start by considering the weak interaction at low energies
and deduce some of its general properties that are valid at all energies. Later we
will consider how unification arises and some of its consequences.
60
Co ! 60
Ni þ e þ e : ð6:2Þ
1
Two particles, called at that time and
, were observed to decay via the weak interaction to and
final states, respectively, which necessarily had different final-state parities. However, the and
had
properties, including the near equality of their masses, which strongly suggested that they were in fact the
same particle. Analysis of the ‘ –
puzzle’ suggested that parity was not conserved in the decays.
2
For their work on parity non-conservation, Chen Yang and Tsung-Dao Lee were awarded the 1957 Nobel
Prize in Physics.
3
This classic experiment is described in readable detail in Chapter 10 of Tr75.
184 CH6 ELECTROWEAK INTERACTIONS
Figure 6.2 Effect of a parity transformation on 60 Co decay: the thick arrows indicate the
direction of the spin of the 60 Co nucleus, while the thin arrows show the direction of the
electron’s momentum
that while P-violation and C-violation are large effects, there is a weaker combined
symmetry, called CP-invariance, which is almost exactly conserved. This has its
most striking consequences for the decays of neutral mesons, which are also
discussed below. We start by considering the P and C operators in more detail.
C-violation and P-violation are both conveniently illustrated by considering the
angular distributions of the electrons and positrons emitted in the decays
! e þ e þ ð6:3aÞ
and
þ ! eþ þ e þ ð6:3bÞ
of polarized muons. In the rest frame of the decaying particle these were found to
be of the form
1
ðcos
Þ ¼ 1 cos
; ð6:4Þ
2 3
where
is the angle between the muon spin direction and the direction of the
outgoing electron or positron, as shown in Figure 6.3(a). The quantities are
called the asymmetry parameters, and are the total decay rates, or equivalently
the inverse lifetimes, i.e.
ð
þ1
1
d cos
ðcos
Þ ¼ ; ð6:5Þ
1
e+−
µ+
− P π−θ µ+
−
e+
−
(a) (b)
Figure 6.3 Effect of a parity transformation on muon decays: the thick arrows indicate the
direction of the muon spin, while the thin arrows indicate the direction of the electron’s
momentum
þ ¼ ðC-invarianceÞ ð6:6Þ
and
þ ¼ ðC-invarianceÞ: ð6:7Þ
The parity transformation preserves the identity of the particles, but reverses
their momenta while leaving their spins unchanged. Its effect on muon decay is
shown in Figure 6.3, where we see that it changes the angle
to
, so that cos
ðcos
Þ ¼ ð cos
Þ ðP-invarianceÞ: ð6:8Þ
Substituting Equation (6.4), leads to the prediction that the asymmetry parameters
vanish,
¼ 0 ðP-invarianceÞ: ð6:9Þ
which shows that both C-invariance and P-invariance are violated. The violation is
said to be ‘maximal’, because the asymmetry parameters are defined to lie in the
range 1 1.
In view of these results, a question that arises is: why do the þ and have
the same lifetime if C-invariance is violated? The answer lies in the principle of
CP-conservation, which states that the weak interaction is invariant under the
combined operation CP even though both C and P are separately violated. The CP
operator transforms particles at rest to their corresponding antiparticles at rest, and
CP-invariance requires that these states should have identical properties. Thus, in
particular, the masses of particles and antiparticles are predicted to be the same.
Specifically, if we apply the CP operator to muon decays, the parity operator
changes
to
as before, while the C operator changes particles to anti-
particles. Hence CP-invariance alone implies that the condition obtained from
P-invariance is replaced by the weaker condition
þ ðcos
Þ ¼ ð cos
Þ: ð6:11Þ
þ ¼ ðCP-invarianceÞ; ð6:12Þ
þ ¼ ðCP-invarianceÞ; ð6:13Þ
antineutrinos first, assuming that they have zero mass for the purpose of this
discussion.
6.3.1 Neutrinos
Figure 6.4 Helicity states of a spin-12 particle: the long thin arrows represent the momenta of
the particles and the shorter thick arrows represent their spins
Figure 6.5 Effect of C, P and CP transformations; only the states shown in boxes are observed
in nature
where the spins of the nuclei are shown in brackets. The excited state of samarium
that is formed decays to the ground state by -emission
152
Sm ðJ ¼ 1Þ ! 152 SmðJ ¼ 0Þ þ ð6:15Þ
and it is these -rays which were detected in the experiment. In the first reaction
(Equation (6.14)), the electrons are captured from the K-shell and the initial
state has zero momentum, so that the neutrino and the 152 Sm nucleus recoil in
opposite directions. The experiment selected events in which the photon was
emitted in the direction of motion of the decaying 152 Sm nucleus, so that overall
the observed reaction was
where the three final-state particles were co-linear, and the neutrino and photon
emerged in opposite directions, as shown in Figure 6.6.
The helicity of the neutrino can then be deduced from the measured helicity of
the photon by applying angular momentum conservation about the event axis to
the overall reaction. In doing this, no orbital angular momentum is involved,
because the initial electron is captured from the atomic K-shell and the final-state
particles all move along the event axis. Hence the spin components of the neutrino
and photon, which can be 12 and 1 respectively, must add to give the spin
component of the initial electron, which can be 12. This gives two possible spin
configurations, as shown in Figures 6.6(a) and 6.6(b). In each case the photon and
neutrino have the same helicities. In the actual experiment, the polarization of the
photons was determined by studying their absorption in magnetized iron (which
depends on the polarization of the photon) and the results obtained were consistent
with the occurrence of left-handed neutrinos only, corresponding to Figure 6.6(a).
SPIN STRUCTURE OF THE WEAK INTERACTIONS 189
Figure 6.6 Possible helicities of the photon and neutrinos emitted in the reaction
e þ 152 EuðJ ¼ 0Þ ! 152 SmðJ ¼ 0Þ þ e þ for those events in which they are emitted in
opposite directions. Experiment selects configuration (a)
Later experiments have shown that only right-handed antineutrinos take part in
weak interactions.
To see the effect of the spin dependence in weak interactions involving particles
with mass, we will look at the decays of the pion and muon which are, of course,
examples of charged current reactions. The spin dependence is of a special form,
called a V–A interaction. This name is derived from the behaviour under a parity
transformation of the weak interaction analogue of the electromagnetic current.
The letter V denotes a proper vector, which is one whose direction is reversed by a
parity transformation (an example is momentum p). The familiar electric current,
to which photons couple, transforms as a proper vector under parity. Because
parity is not conserved in weak interactions, the corresponding weak current, to
which W -bosons couple, has in addition to a vector (V) component another
component whose direction is unchanged by a parity transformation. Such a
quantity is called an axial-vector (A) (an example of an axial-vector is orbital
angular momentum L ¼ r p). Since observables are related to the modulus
squared of amplitudes, either term would lead by itself to parity conservation.
Parity non-conservation is an interference effect between the two components.
Here we shall consider only the most important characteristic of this spin
dependence, which is that the results discussed above for neutrinos, hold for all
fermions in the ultra-relativistic limit. That is, in the limit that their velocities
approach that of light, only left-handed fermions L , e L etc. and right-handed
antifermions R , eþR etc. are emitted in charged current interactions. These
190 CH6 ELECTROWEAK INTERACTIONS
right-handed and left-handed particles are called chiral states and these are the
eigenstates that take part in weak interactions. In general, chiral states are linear
combinations of helicity states,4 with the contributions of the ‘forbidden’ heli-
þ
city states e
R , eL etc. suppressed by factors which are typically of the order of
2
ðmc2 =EÞ , where m is the appropriate fermion mass and E its energy. For massless
neutrinos this is always a good approximation and chiral states and helicity states
are identical. However, for particles with mass, it is only a good approximation for
large energies E. These spin properties can be verified most easily for the electrons
and muons emitted in weak decays, by directly measuring their spins. Here we
shall assume them to hold and use them to understand some interesting features of
pion and muon decays.
We start by considering the pion decay mode
þ ! ‘þ þ ‘ : ð‘ ¼ e; Þ ð6:17Þ
In the rest frame of the decaying pion, the charged lepton and the neutrino recoil in
opposite directions, and because the pion has zero spin, their spins must be
opposed to satisfy angular momentum conservation about the decay axis. Since the
neutrino (assumed to be zero mass) is left-handed, it follows that the charged
lepton must also be left-handed, as shown in Figure 6.7, in contradiction to the
expectations for a relativistic antilepton.
+ π+ ν
Figure 6.7 Helicities of the charged leptons in pion decays: the short arrows denote spin
vectors and the longer arrows denote momentum vectors
For the case of a positive muon this is unimportant, since it is easy to check
that it recoils non-relativistically and so both chirality states are allowed. However,
if a positron is emitted it recoils relativistically, implying that this mode is
suppressed by a factor that we can estimate from the above to be of the order
of ðme =m Þ2 105 . Thus the positron decay mode is predicted to be much rarer
than the muonic mode. This is indeed the case, and the measured ratio
ðþ ! eþ þ e Þ
¼ ð1:218 0:014Þ 104 ð6:18Þ
ðþ ! þ þ Þ
is in excellent agreement with a full calculation that takes into account both the
above suppression and the difference in the density-of-final states (i.e. the
difference in the Q-values) for the two reactions.
4
This is another example where linear combinations of states are the ones of physical interest; compare
neutrino mixing (Section 3.1.3).
SPIN STRUCTURE OF THE WEAK INTERACTIONS 191
and correspond to decays in which the neutrino and antineutrino are both emitted
in the direction opposite to the electron. This is illustrated in Figure 6.8 for the two
simplest cases in which the electron is emitted in the muon spin direction
(Figure 6.8(a)) and opposite to it (Figure 6.8(b)).
Figure 6.8 Muon decays in which electrons of the highest possible energy are emitted: (a) in
the muon spin direction, and (b) opposite to the muon spin direction
Since the neutrino and antineutrino have opposite helicities, the muon and
electron must have the same spin component along the event axis in order to
conserve angular momentum, implying the electron helicities shown in Figure 6.8.
When combined with the fact that the relativistic electrons emitted must be left-
handed, this implies that electrons cannot be emitted in the muon spin direction.
We thus see that the spin structure of the interaction automatically gives rise to a
forward–backward asymmetry in polarized muon decays. Of course not all the
electrons have the maximum energy and the actual asymmetry, averaged over all
electron energies, can only be calculated by using the full form of the V–A
5
This is in the rest frame of the decaying pion and assumes that the neutrino has zero mass. The degree of
polarization in the laboratory frame is a function of the muon momentum.
192 CH6 ELECTROWEAK INTERACTIONS
p þ p ! W þ þ X ;
p þ p ! W þ X þ ; and p þ p ! Z 0 þ X 0 ; ð6:20Þ
where X and X 0 are arbitrary hadronic states allowed by the conservation laws.
The beams of protons and antiprotons were supplied by a proton–antiproton
collider, which was specifically built for this purpose. At the time it had proton and
antiproton beams with maximum energies of 270 GeV each, giving a total centre-
of-mass energy of 540 GeV. Two independent experiments were mounted (called
6
See, for example, Chapter 12 of Ha84.
W AND Z 0 BOSONS 193
UA1 and UA2), both of which were examples of the ‘layered’ detector systems
that were discussed in Chapter 4.7 One of the main problems facing the
experimenters was that for each event in which a W or Z 0 is produced and
decays to leptons, there were more than 107 events in which hadrons alone are
produced and so the extraction of the signal required considerable care.
In contrast to the zero mass photons and gluons, the W and Z 0 bosons are both
very massive particles, with measured masses
while their lifetimes are about 3 1025 s. Their dominant decays lead to jets of
hadrons, but the leptonic decays
W þ ! ‘þ þ ‘ ; W ! ‘ þ ‘ ð6:22Þ
and
Z 0 ! ‘þ þ ‘ ; Z 0 ! ‘ þ l ; ð6:23Þ
7
Simon van der Meer lead the team that built the accelerator and Carlo Rubbia lead the UA1 experimental
team that subsequently discovered the bosons. They shared the 1984 Nobel Prize in Physics for their work.
8
A more detailed description of the UA1 experiment is given in, for example, Section 8.1 of Ma97.
194 CH6 ELECTROWEAK INTERACTIONS
At each vertex a boson is emitted or absorbed; while both fermion lines belong
to the same generation ‘ ¼ e, or , with one arrow pointing inwards and one
outwards to guarantee conservation of each lepton number Ne, N and N .
Finally, associated with each vertex is a dimensionless parameter with the same
value
at high energies for all three generations (because of lepton universality). This
constant is the weak analogue of the fine structure constant
1=137 in
electromagnetic interactions, with gW the weak analogue of the electronic charge
e in appropriate units.
We see from the above that, despite its name, the weak interaction has a similar
intrinsic strength to the electromagnetic interaction. Its apparent weakness in many
low-energy reactions, is solely a consequence of its short range, which arises
because the exchange bosons are heavy. From Equation (6.1) we see that the
scattering amplitude has a denominator that contains the squared mass of the
exchanged particle and so at energies where the de Broglie wavelengths ¼ h=p
of the particles are large compared with the range of the weak interaction, which is
an excellent approximation for all lepton and hadron decays, the range can be
neglected altogether. In this approximation the weak interaction becomes a point
or zero range interaction, whose effective interaction strength can be shown to be
eff ¼
W E =MW c2 2 ; MW c2 ;
E ð6:25Þ
where E is a typical energy scale for the process in question. (For example in muon
decay it would be the mass of the muon.) Thus we see that the interaction is both weak
and very energy dependent at ‘low energies’, but becomes comparable in strength
with the electromagnetic interaction at energies on the scale of the W-boson mass.
A typical semileptonic decay (i.e. one that involves both hadrons and leptons) is
that of the neutron, which at the quark level is
d ! u þ e þ e ; ð6:26Þ
WEAK INTERACTIONS OF HADRONS 195
as illustrated in Figure 6.10, where the other two quarks play the role of spectators.
Similarly, in the pion decay process
ðduÞ ! þ ð6:27Þ
have identical weak interactions. That is, one can obtain the basic W quark
vertices by making the replacements e ! u ; e ! d ; ! c ; ! s in
the basic W lepton vertices, leaving the coupling constant gW unchanged. The
resulting W quark vertices are shown in Figure 6.12.
Figure 6.12 The W quark vertices obtained from quark--lepton symmetry, without quark
mixing
Quark symmetry in the simple form stated above then implies that the
fundamental processes d þ u ! W and s þ c ! W occur with the same
couplings gW as the corresponding leptonic processes, i.e. in Figure (6.12) we
have gcs ¼ gud ¼ gW , while the processes s þ u ! W and d þ c ! W are
forbidden. This works quite well for many reactions, like the pion decay
! þ , but many decays that are forbidden in this simple scheme are
observed to occur, albeit at a rate which is suppressed relative to the ‘allowed’
decays. An example of this is the kaon decay K ! þ , which requires a
sþ u ! W vertex, which is not present in the above scheme.
All these suppressed decays can be successfully incorporated into the theory by
introducing quark mixing. According to this idea, the d and s quarks participate in
the weak interactions via the linear combinations
d 0 ¼ d cos
C þ s sin
C ð6:30aÞ
and
s0 ¼ d sin
C þ s cos
C ; ð6:30bÞ
WEAK INTERACTIONS OF HADRONS 197
This then generates new vertices previously forbidden. For example, the usW vertex
required for the decay K ! þ arises from the interpretation of the ud 0 W
vertex shown in Figure 6.13. In a similar way a new cdW vertex is also generated.
Figure 6.13 The ud0 W vertex and its interpretation in terms of udW and usW vertices
Quark mixing enables theory and experiment to be brought into good agreement
by choosing a value
C 13 for the Cabibbo angle. One then finds that the rates
for the previously ‘allowed’ decays occur at rates which are suppressed by a factor
cos2
C 0:95, while the previously ‘forbidden’ decays are now allowed, but with
rates which are suppressed by a factor sin2
C 0:05.
Historically, the most remarkable thing about these ideas is that they were
formulated before the discovery of the charmed quark. In 1971 seven fundamental
fermions were known: the four leptons e , e , and , and the three quarks u, d
and s. This led Glashow, Iliopolous and Maiani to propose the existence of a fourth
quark c to complete the lepton–quark symmetry and to solve problems associated
with neutral currents that we will discuss in Section 6.7. The existence of the
charmed quark was subsequently confirmed in 1974 with the discovery of the first
charmonium states (this is why their discovery was so important – see the
discussion in Section 5.3) and its measured weak couplings are consistent with
the predictions of lepton–quark symmetry and quark mixing.
We now know that there are six leptons
e
ð6:32Þ
e
9
This is yet another example of physical states being mixtures of other states.
198 CH6 ELECTROWEAK INTERACTIONS
When the third generation is taken into account, the mixing scheme becomes more
complicated, as we must allow for the possibility of mixing between all three
‘lower’ quarks d, s and b instead of just the first two and more parameters are
involved. In general the mixing can be written in the form
0 1 0 10 1
d0 Vud Vus Vub d
@ s0 A ¼ @ Vcd Vcs Vcb A@ s A; ð6:34Þ
b0 Vtd Vts Vtb b
10
The initials stand for Cabibbo, Kobayashi and Maskawa, the last two of whom extended the original
Cabibbo scheme to three generations of quarks.
11
A review is given Ei04.
WEAK INTERACTIONS OF HADRONS 199
Figure 6.14 Spin (thick arrows) and momentum (thin arrows) configurations for e e and e e
interactions: (a) e e before collision; (b) e e after scattering through 180 ; (c) e e before
collision; (d) e e after scattering through 180
both cases the total spin component along the z-axis is zero. This result is true for
all angles and the scattering is isotropic.
From this we can calculate the differential cross-section using the formulae of
Chapter 1. We will assume that the squared momentum transfer Q2 is such that
Q2max MW 2 2
c , so that the matrix element may be written [cf. Equation (6.1)]
4ðhcÞ3
W
GF ¼ ð6:36Þ
ðMW c2 Þ2
and
W ¼ g2 =4hc is the equivalent of the fine structure constant for charged
current weak interactions. Hence, using Equation (1.57) and recalling that the
velocities of both the neutrino and electron are equal to c,
d 1 G2 2
ðe e Þ ¼ 2 F 4 ECM : ð6:37Þ
d 4 ð
hcÞ
2
At high energies ECM is given by
2
ECM 2me c2 E ; ð6:38Þ
2me c2 G2F
tot ðe e Þ ¼ E ð6:39Þ
ðhcÞ4
200 CH6 ELECTROWEAK INTERACTIONS
and
MN c2 G2F Ev 1
tot ð
e NÞ ¼ H þ H ; ð6:42bÞ
ðhcÞ4 3
for scattering from an isoscalar nucleus, i.e. one with an equal number of neutrons
and protons, where MN is the mass of the nucleon. The quantities H and H are
given by
ð1 ð1
H
x½uðxÞ þ dðxÞdx and H
x½uðxÞ þ dðxÞdx; ð6:43Þ
0 0
12
This behaviour has arisen because of the approximation Equation (6.35). It cannot of course continue
indefinitely. At very high values of Q2 the full form of the propagator would have to be taken into account
and this would introduce an energy dependence in the denominator of Equation (6.39).
NEUTRAL MESON DECAYS 201
where uðxÞ etc. are the quark densities defined in Section 5.7 and the integral is
over the scaling variable x.
Setting y ¼ H=H, we have from Equations (6.42)
e NÞ 1 þ 3y
ð
R
¼ : ð6:44Þ
ðe NÞ 3þy
Some data for R are shown in Figure 6.15 from an experiment using muon–
neutrinos. These show that R is approximately constant, as predicted by
Equation (6.44), and has a value of about 0.51, which implies y 0:2, i.e.
antiquarks exist in the nucleon at the level of about 20 per cent. Other experiments
yield similar results in the range 15–20 per cent.
Figure 6.15 Neutrino and antineutrino total cross-sections (data from Se97)
states. Because most work has been done on the neutral kaons, we will mainly
discuss this system as an example. The equivalent formalisms for B- and D-decays
are similar. We start with a discussion of CP violation.
6.6.1 CP violation
Figure 6.16 Example of a process that can convert a K 0 state to a K0 state
so that
0 0
CPjK 0 ; 0i ¼ jK ; 0i; CPjK ; 0i ¼ jK 0 ; 0i: ð6:48Þ
0
Thus CP eigenstates K1;2 are
0 1 n 0
o
jK1;2 ; 0i ¼ pffiffiffi jK 0 ; 0i jK ; j0i ðCP ¼ 1Þ: ð6:49Þ
2
If CP is conserved, then K10 should decay entirely to states with CP ¼ 1 and K20
should decay entirely into states with CP ¼ 1. We examine the consequences of
this for decays leading to pions in the final state.
Consider the state 0 0 . Since the kaon has spin-0, by angular momentum
conservation the pion pair must have zero orbital angular momentum in the rest
frame of the decaying particle. Its parity is therefore given by [cf. Equation (1.14)]
C ¼ ðC0 Þ2 ¼ 1; ð6:51Þ
where C0 ¼ 1 is the C-parity of the neutral pion. Combining these results gives
CP ¼ 1. The same result holds for the þ final state.
The argument for three-pion final states þ 0 and 0 0 0 is more compli-
cated, because there are two orbital angular momenta to consider, If we denote by
L12 the orbital angular momentum of one pair (either þ or 0 0 ) in their
mutual centre-of-mass frame, and L3 is the orbital angular momentum of the third
pion about the centre-of-mass of the pair in the overall centre-of-mass frame, then
the total orbital angular momentum L
L12 þ L3 ¼ 0, since the decaying particle
has spin-0. This can only be satisfied if L12 ¼ L3 . This implies that the parity of the
final state is
C ¼ ðC0 Þ3 ¼ 1 ð6:53Þ
and combining these results gives CP ¼ 1 overall. The same result can be shown
to hold for the þ 0 final state.
The experimental position is that two neutral kaons are observed, called K 0 -short
and K 0 -long, denoted KS0 and KL0 , respectively. They have almost equal masses of
about 499 MeV/c2, but very different lifetimes and decay modes. The KS0 has a
204 CH6 ELECTROWEAK INTERACTIONS
lifetime of 0:89 1010 s and decays overwhelmingly to two pions; the longer-
lived KL0 has a lifetime of 0:52 107 s with a significant branching ratio to three
pions, but not to two. In view of the CP analysis above, this immediately suggests
the identification
However, in 1964 it was discovered that the KL0 also decayed to two pions13
KL0 ! þ þ ; ð6:55Þ
but with a very small branching ratio of the order of 103 . This result is clear
evidence of CP violation. This was confirmed in later experiments on the decay
K 0 ! 0 0 .
Because CP is not conserved, the physical states KS0 and KL0 need not correspond
to the CP-eigenstates K10 and K20 , but can contain small components of states with
the opposite CP, i.e. we may write
1
jKS0 ; 0i ¼ jK10 ; 0i "jK20 ; 0i ð6:56aÞ
ð1 þ j"j2 Þ1=2
and
1
jKL0 ; 0i ¼ 2 1=2
"jK10 ; 0 i þ jK20 ; 0 i ; ð6:56bÞ
ð1 þ j"j Þ
K 0 ! þ eþ þ e ð6:57aÞ
13
The experiment was led by James Cronin and Val Fitch. They received the 1980 Nobel Prize in Physics for
their discovery.
14
See, for example, Ei04.
NEUTRAL MESON DECAYS 205
and
0 ! þ þ e þ e :
K ð6:57bÞ
For example, if we start with a beam of K 0 particles, with initially equal amounts
of KS0 and KL0 , then after a time that is large compared with the KS0 lifetime, the KS0
component will have decayed leaving just the KL0 component, which itself will be
an equal admixture of K 0 and K 0 components. We would therefore expect to
observe identical numbers of electrons (N ) and positrons (N þ ) from the decays of
Equations (6.57). However, if KL0 is not an eigenstate of CP, then there will be an
asymmetry in these numbers, which will depend on the relative strengths of the K 0
and K 0 components in KL0 . The asymmetry is given by 2Re ", where " is the CP-
violating parameter defined in Equation (6.56).
Figure 6.17 shows data on the asymmetry ðN þ N Þ=ðN þ þ N Þ as a function
of proper time. After the initial oscillations there is seen to be an asymmetry whose
value is 2Re " 3:3 103 , which is consistent with the value of " obtained
from the modes. Thus CP-violation in K-decay occurs mainly, though not
entirely, by the mixing of the CP-eigenstates in the physical states rather than by
direct CP-violating decays, both of which are allowed in the CKM mixing scheme.
What do these results mean for the CKM mixing scheme? The CKM matrix is a
3 3 matrix and in general contains nine complex elements. However, the unitary
nature of the matrix implies that there are relations between the elements, such as
Vud Vub þ Vcd Vcb þ Vtd Vtb ¼ 0: ð6:58Þ
Figure 6.17 The charge asymmetry observed for K 0 ! eþ e and K0 ! þ e e as a function
of proper time, for a beam that is initially predominantly K 0 (adapted from Gj74, copyright
Elsevier, with permission)
206 CH6 ELECTROWEAK INTERACTIONS
Using these and exploiting the freedom to define the phases of the basic quark
states, the matrix may be parameterized by just four quantities. A number of
different parameterizations are used, but an approximate form that is commonly
used to discuss CP violation is
0 1
1 12 2 A3 ð iÞ
V ¼@ 1 12 2 A2 A þ Oð4 Þ; ð6:59Þ
3 2
A ð1 iÞ A 1
with parameters A; ; and . The quantity ¼ jVus j 0:22 plays the role of an
expansion parameter in this approximation and a non-zero value of would be
indicative of CP violation.
The parameter " in Equations (6.56) is just one CP-violating parameter that may
be measured in various K-decay modes. We will not pursue this further, but note
that by combining the values of the parameters with information on other elements
of the CKM matrix, a value of the CP-violating parameter may be deduced and
used to predict the size of CP-violating effects in other decays. There are very few
other places where such mixing effects can occur, but in principle they should be
0 and B0 B
possible in the D0 D 0 systems, which are analogues of the K-mesons, but
with a strange quark replaced by a charmed and bottom quark, respectively.
Mixing in the B0 B 0 states due to B0 B 0 oscillations has in fact been observed and
also very recently direct CP violation. The latter was established by comparing the
decay B0 ! K þ with the decay B 0 ! K þ. Moreover, the size of the effect is
much stronger than in neutral kaon decays and this is in agreement with the
predictions of the CKM mixing scheme.
There is still much to be done in studying CP violation. For example,
the cleanest measurement of the CP-violating parameter would be from the
decays B0 =B ! ðJ= ÞKS0 , where J= is the 3 S1 ground state of charmonium,
but the present limits on these decays are orders of magnitude from those required
to test the predictions. On the theoretical side, although the CKM mixing model
accounts for all CP-violating data to date, it fails by several orders of magnitude to
account for the observed matter–antimatter asymmetry observed in the universe
(which will be discussed in Chapter 9) and so there is probably a CP-violating
mechanism beyond the standard model awaiting to be discovered.
and m
and
are the mass and decay rate of the particle concerned. Here the first
exponential factor is the usual oscillating time factor eiEt associated with any
quantum mechanical stationary state, evaluated in the rest frame of the particle.
The second exponential factor reflects the fact that the particles decay, and it
ensures that the probability
2
1
pffiffiffi a
ðtÞ ¼ 1 e
t ða ¼ S; LÞ ð6:64Þ
2 2
in the form
0 ðtÞjK 0 ; pi ;
A0 ðtÞjK 0 ; pi þ A ð6:66Þ
208 CH6 ELECTROWEAK INTERACTIONS
where
1 0 ðtÞ ¼ 1 ½aS ðtÞ aL ðtÞ:
A0 ðtÞ ¼ ½aS ðtÞ þ aL ðtÞ and A ð6:67Þ
2 2
The intensities of the two components are then given by
1h i
IðK 0 Þ
jA0 ðtÞj2 ¼ eS t þ eL t þ 2eðS þL Þt=2 cosðmtÞ ð6:68aÞ
4
and
0 0 ðtÞj2 ¼ 1 h S t i
IðK Þ
jA e þ eL t 2eðS þL Þt=2 cosðmtÞ ð6:68bÞ
4
where m
jmS mL j and we have used Equation (6.63) explicitly to evaluate
the amplitudes.
0 0
The variation of the K intensity IðK Þ with time can be determined experi-
mentally by measuring the rate of production of hyperons (baryons with non-zero
strangeness) in strangeness-conserving strong interactions such as
0
K þ p ! þ þ 0
ð6:69Þ
! 0 þ þ
as a function of the distance from the K 0 source. The data are then fitted by
Equations (6.68) with m as a free parameter and the predictions are in good
agreement with the experiments for a mass difference
Figure 6.18 Higher order contribution to the reaction eþ ! eþ from the exchange of
two W-bosons
of such reactions before their discovery in 1973. This theory15 was proposed
mainly in order to solve problems associated with Feynman diagrams in which
more than one W boson was exchanged, like that shown in Figure 6.18, which
contributes to the reaction eþ ! eþ .
Such contributions are expected to be small because they are higher order in the
weak interaction and this appears to be confirmed by experimental data, which are
in good agreement with theoretical predictions that neglect them entirely. (For
example, in the experimentally accessible reaction eþ þ e ! þ þ .) How-
ever, when these higher-order contributions are explicitly calculated, they are
found to be proportional to divergent integrals, i.e. they are infinite. In the unified
theory, this problem is solved when diagrams involving the exchange of Z 0 bosons
and photons are taken into account. These also give infinite contributions, but
when all the diagrams of a given order are added together the divergences
cancel (!), giving a well-defined and finite contribution overall.16 This cancellation
is not accidental, but is a consequence of a fundamental symmetry relating the
weak and electromagnetic interactions. Here we will simply comment on some
phenomenological consequences of the theory.
15
The formulation of the theory is in terms of four massless vector bosons arranged as multiplets of ‘weak
isospin’ and ‘weak hypercharge’. Specifically, three states are a weak isospin triplet and the fourth is a weak
isospin singlet. The fact that they all have zero masses ensures that gauge invariance is satisfied. These fields
then interact with additional scalar fields associated with new postulated particles called Higgs bosons, which
we have mentioned elsewhere. This process, known as ‘spontaneous symmetry breaking’ generates the
observed masses of the W, Z and bosons, while still preserving gauge invariance. (For further details see,
for example, Section 8.4 of Pe00.) The originators of this theory, Sheldon Glasow, Abdus Salam and Steven
Weinberg, shared the 1979 Nobel Prize in Physics for their contributions to formulating the electroweak
theory and the prediction of weak neutral currents.
16
The first people to demonstrate that this occurred were Gerardus ‘t Hooft and Martinus Veltman. They
shared the 1999 Nobel Prize in Physics for this discovery.
210 CH6 ELECTROWEAK INTERACTIONS
The first is that to ensure the cancellation, the theory requires a relation between
the weak and electromagnetic couplings, called the unification condition. This is
e
pffiffiffi 1=2 ¼ gw sin
W ¼ gz cos
W ; ð6:71Þ
2 2"0
cos
W
MW =MZ ð0 <
< =2Þ ð6:72Þ
and gz is a coupling constant which characterizes the strength of the neutral current
vertices. The unification condition relates the strengths of the various interactions
to the W and Z masses, and historically was used to predict the latter from the
former before the W and Z 0 bosons were discovered.
Secondly, just as all the charged current interactions of leptons can be under-
stood in terms of the basic W -lepton vertices, in the same way all known neutral
current interactions can be accounted for in terms of basic Z 0 -lepton vertices
shown in Figures 6.19(a) and 16.19(b). The corresponding quark vertices can be
obtained from the lepton vertices by using lepton–quark symmetry and quark
mixing, in the same way that W -quark vertices are obtained from the W -lepton
vertices. Thus, making the replacements
e ! u; ! c; e ! d 0 ; ! s0 ð6:73Þ
e e Z 0 ; Z 0 ; e e Z 0 ; Z 0 ; ð6:74Þ
Figure 6.19 Z 0 and couplings to leptons and quarks in the unified electroweak theory, where
‘ ¼ e; and and a denotes a quark
NEUTRAL CURRENTS AND THE UNIFIED THEORY 211
Finally, we interpret the latter two of these using Equations (6.30). Thus, for
example,
d 0 d 0 Z 0 ¼ ðd cos
C þ s sin
C Þ ðd cos
C þ s sin
C ÞZ 0
¼ ddZ 0 cos2
C þ ssZ 0 sin2
C þ ðdsZ 0 þ sdZ 0 Þ sin
C cos
C ð6:76Þ
When all the terms in Expression (6.75) are evaluated, ones obtains a set of
vertices equivalent to
R L
APV
; ð6:78Þ
R þ L
value was found in agreement with other determinations, e.g. from deep inelastic
neutrino scattering. A later experiment confirmed the effect in polarized electron–
proton scattering.
A very recent experiment (2004) has measured APV for e e scattering. This
was done using electrons of about 50 GeV primary energy from the SLAC linear
accelerator in Stanford, USA, and scattering them from a liquid hydrogen target.
The experiment was able to distinguish final-state electrons scattered from the
atomic electrons from those scattered from protons. Taking account of all sources
of error, the measured value was APV ¼ ð175 40Þ 109 (note the exponent –
parts per billion) and the experiment also yielded a value of sin2
W consistent with
other determinations. These remarkable experiments provide unambiguous evi-
dence for parity violation in ‘electromagnetic’ processes at the level predicted by
theory and hence for the electroweak unification as specified in the standard
model.17
It should also in principle be possible to detect parity violating effects in atomic
physics, where the electromagnetic interactions of the electrons dominate. For
example, measurements have been made of the slight change in the polarization
angle of light passing through a vapour of metallic atoms. In this case the rotation
angle is extremely small (
107 rad), but very sensitive experiments can measure
the effect to an accuracy of
1 per cent. However, to predict the size of the effects
requires a detailed knowledge of the atomic theory of the atom and in all cases to
date the uncertainties on the predictions are such that a null effect cannot be ruled
out. Thus at present, atomic physics does not compete with particle physics
experiments in detecting parity-violating effects and measuring sin2
W , although
this could change in the future.
Problems
6.1 Define charged and neutral current reactions in weak interactions and give an
example of each in symbol form. How do they differ in respect of conservation of
the strangeness quantum number? Why does observation of the process
þ e ! þ e constitute unambiguous evidence for weak neutral currents,
whereas the observation of e þ e ! e þ e does not?
6.2 The reaction eþ e ! þ is studied using colliding beams each of energy 7 GeV
and at these energies the reaction is predominantly electromagnetic. Draw its lowest
order Feynman diagram. The differential cross-section is given by
d h 2 c2
2
¼ 2
1 þ cos2
;
d 4ECM
17
Incidentally, all these experiments are of the fixed-target type, showing that this type of experiment still has
a lot to offer.
214 CH6 ELECTROWEAK INTERACTIONS
F B
AFB ¼ ;
F þ B
where F ðB Þ is the total cross-section for scattering in the forward (backward)
hemisphere, i.e. 0 cos
1 ð1 cos
0Þ. Derive a relation between Cwk
and AFB .
6.3 Draw a Feynman diagram at the quark level for the decay ! p þ . If nature
were to double the weak coupling constant and decrease the mass of the W boson by
a factor of four, what would be the effect on the decay rate ð ! p þ Þ?
6.4 Neglecting the electron mass, the energy spectrum for the electrons emitted in muon
decay is give by
d! 2G2F ðm c2 Þ2 Ee2 4Ee
¼ 1 :
dEe ð2Þ3 ðhcÞ6 3m c2
What is the most probable energy for the electron? Draw a diagram showing the
orientation of the momenta of the three outgoing particles and their helicities for the
case when Ee m c2 =2. Show also the helicity of the muon. Integrate the energy
spectrum to obtain an expression for the total decay width of the muon. Hence
calculate the muon lifetime in seconds ðGF =ðhcÞ3 ¼ 1:166 105 GeV2 Þ:
6.5 Use lepton universality and lepton–quark symmetry to estimate the branching ratios
for (a) the decays b ! c þ e þ e (where the b and c quarks are bound in hadrons)
and (b) ! e þ e þ . Ignore final states that are Cabibbo-suppressed relative
to the lepton modes.
6.6 The couplings of the Z 0 to right-handed (R) and left-handed (L) fermions are given by
gR ðf Þ ¼ qf sin2
W ; gL ðf Þ ¼ 1=2 qf sin2
W ;
quarks; the negative sign is used for charged leptons and the q ¼ d; s; b quarks. If
the partial width for Z 0 ! f f is given by
GF MZ3 c6 2
f ¼ pffiffiffi 3
gR ðf Þ þ g2L ðf Þ ;
3 2ð hcÞ
calculate the partial widths to neutrinos and to q q pairs q and explain the
relation of q to the partial width to hadrons hadron .
The widths to hadrons and to charged leptons are measured to be
had ¼ ð1738 12Þ MeV and lep ¼ ð250 2Þ MeV, and the total width to all
final states is measured to be tot ¼ ð2490 7Þ MeV. Use these experimental results
and your calculated value for the decay width to neutrinos to show that there are
only three generations of neutrinos with masses M < MZ =2.
6.7 Explain, with the aid of Feynman diagrams, why the decay D0 ! K þ þ can
occur as a charged-current weak interaction at lowest order, but the decay
Dþ ! K 0 þ þ cannot.
6.8 Why is the decay rate of the charged pion much smaller than that of the neutral pion?
Draw Feynman diagrams to illustrate your answer.
6.9 Draw the lowest-order Feynman diagrams for the decays ! þ and
K ! þ . Use lepton–quark symmetry and the Cabibbo hypothesis with the
Cabibbo angle
C ¼ 12 to estimate the ratio
RateðK ! þ Þ
R
;
Rateð ! þ Þ
ð ! n þ e þ e Þ
R
ð ! þ e þ e Þ
and explain why the decay ðþ ! n þ eþ þ e Þ has never been seen.
6.11 One way of looking for the Higgs boson H is in the reaction eþ e ! Z 0 H. If this
reaction is studied at a centre-of-mass energy of 500 GeV in a collider operating for
107 s per year and the cross-section at this energy is 60 fb, what instantaneous
luminosity (in units of cm2 s1) would be needed to collect 2000 events in one year
if the detection efficiency is 10 per cent. For a Higgs boson with mass
MH < 120 GeV, the branching ratio for H ! b b is predicted to be 85 per cent.
Why will looking for b quarks help distinguish eþ e ! Z 0 H from the background
reaction eþ e ! Z 0 Z 0 ?
fictitious strangeness zero I ¼ 12 particle S0 in the initial state, find the prediction of
this rule for the ratio
ð ! þ Þ
R
:
ð0 ! þ 0 Þ
Assume that the state j0 ; S0 i is an equal mixture of states with I ¼ 0 and I ¼ 1.
6.13 The charged-current differential cross-sections for and scattering from a spin-12
target are given by generalizations of Equations (6.37) and (6.40) and may be written
2
where s ¼ ECM , y ¼ 12 ð1 cos
Þ and H is the integral of the quark density for the
target (cf. Equation (6.43)). The corresponding cross-sections for neutral current
scattering are
where the right- and left-hand couplings to u and d quarks are given by
1 2 2
gL ðuÞ ¼ sin2
W ; gR ðuÞ ¼ sin2
W ;
2 3 3
1 1 2 1
gL ðdÞ ¼ þ sin
W ; gR ðdÞ ¼ sin2
W :
2 3 3
Derive expressions for the ratios NC ðÞ=CC ðÞ and NC ð
Þ=CC ð
Þ in the case of
an isoscalar target consisting of valence u and d quarks only.
7
Models and Theories
of Nuclear Physics
Nuclei are held together by the strong nuclear force between nucleons, so we start
this chapter by looking at the form of this, which is more complicated than that
generated by simple one-particle exchange. Much of the phenomenological evi-
dence comes from low-energy nucleon–nucleon scattering experiments which we
will simply quote, but we will interpret the results in terms of the fundamental strong
interaction between quarks. The rest of the chapter is devoted to various models and
theories that are constructed to explain nuclear data in particular domains.
1
For reviews see, for example, Chapter 7 of Je90 and Chapter 14 of Ho97.
Figure 7.1 Idealized square well representation of the strong interaction nucleon--nucleon
potential. The distance R is the range of the nuclear force and R is the distance at which the
short-range repulsion becomes important. The depth V0 is approximately 40 MeV
interaction potential must be combined with the Coulomb potential in the case in
the case of protons.
A comparison of nn and pp scattering data (after allowing for the Coulomb
interaction) shows that the nuclear force is charge-symmetric (pp ¼ nn) and
almost charge-independent ðpp ¼ nn ¼ pnÞ.2 We have commented in Chapter 3
that there is also evidence for this from nuclear physics. Charge-symmetry is seen in
comparisons of the energy levels of mirror nuclei (see, for example, Figure 3.9) and
evidence for charge-independence comes from the energy levels of triplets of
related nuclei with the same A values. Nucleon–nucleon forces are, however,
spin-dependent. The force between a proton and neutron in an overall spin-1 state
(i.e. with spins parallel) is strong enough to support a weakly bound state (the
deuteron), whereas the potential corresponding to the spin-0 state (i.e. spins
antiparallel) has no bound states. Finally, nuclear forces saturate. This describes
that fact that a nucleon in a typical nucleus experiences attractive interactions only
with a limited number of the many other nucleons and is a consequence of the short-
range nature of the force. The evidence for this is the form of the nuclear binding
energy curve and was discussed in Chapter 2.
Ideally one would like to be able to interpret the nucleon–nucleon potential in
terms of the fundamental strong quark–quark interactions. It is not yet possible to
give a complete explanation along these lines, but it is possible to go some way in
this direction. If we draw an analogy with atomic and molecular structure, with
2
For a discussion of these data see, for example, the references in Footnote 1.
THE NUCLEON--NUCLEON POTENTIAL 219
quarks playing the role of electrons, then possibilities are: an ionic-type bond, a van
der Waals type of force, or a covalent bond.3 The first can be ruled out because the
confining forces are too strong to permit a quark to be ‘lent’ from one nucleon to
another and the second can also be ruled out because the resulting two-gluon
exchange is too weak. This leaves a covalent bond due to the sharing of single quarks
between the nucleons analogous to the covalent bond that binds the hydrogen
molecule. However, nucleons have to remain ‘colourless’ during this process and so
the shared quark from one nucleon has to have the same colour as the shared quark
from the other nucleon. The effect of this is to reduce the effective force (because
there are three possible colour states) and by itself it is unable to explain the depth
of the observed potential. In addition to the three (valence) quarks within the nucleon
there are also present quark–antiquark pairs due to vacuum fluctuations.4 Such pairs
can be colourless and so can also be shared between the nucleons. These quarks
actually play a greater role in generating the nuclear strong interaction than single
quarks. The lightest such diquarks will be pions and this exchange gives the largest
contribution to the attractive part of the nucleon–nucleon force (see, for example, the
Feynman diagram Figure 1.4).
In principle, the short-range repulsion could be due to the exchange of heavier
diquarks (i.e. mesons), possibly also in different overall spin states. Experiment
provides many suitable meson candidates, in agreement with the predictions of the
quark model, and each exchange would give rise to a specific contribution to the
overall nucleon–nucleon potential, by analogy with the Yukawa potential resulting
from the exchange of a spin-0 meson, as discussed in Chapter 1. It is indeed possible
to obtain excellent fits to nucleon–nucleon scattering data in a model with several
such exchanges.5 Thus this approach can yield a satisfactory potential model, but is
semi-phenomenological only, as it requires the couplings of each of the exchanged
particles to be found by fitting nucleon–nucleon scattering data. (The couplings that
result broadly agree with values found from other sources.) Boson-exchange models
therefore cannot give a fundamental explanation of the repulsion. The reason for
the repulsion at small separations in the quark model lies in the spin dependence
of the quark–quark strong interaction, which like the phenomenological nucleon–
nucleon interaction, is strongly spin-dependent. We have discussed this in the
context of calculating hadron masses in Section 3.3.3. When the two nucleons are
very close, the wavefunction is effectively that for a 6-quark system with zero
angular momentum between the quarks, i.e. a symmetric spatial wave function.
Since the colour wave function is antisymmetric, (recall the discussion of
Chapter 5), it follows that the spin wavefunction is symmetric. However, the
3
Recall from chemistry that in ionic bonding, electrons are permanently transferred between constituents to
form positive and negative ions that then bind by electrostatic attraction; in covalent bonding the constituents
share electrons; and the van der Waals force is generated by the attraction between temporary charges
induced on the constituents by virtue of slight movements of the electrons.
4
These are the ‘sea’ quarks mentioned in connection with the quark model in Chapter 3 and which are probed
in deep inelastic lepton scattering that was discussed in Chapter 6.
5
This approach is discussed in, for example, Chapter 3 of Co01 and also in the references given in Footnote1.
220 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
potential energy increases if all the quarks remain in the L ¼ 0 state with
spins aligned.6 The two-nucleon system will try to minimize its ‘chromomagnetic’
energy, but this will compete with the need to have a symmetric spin wavefunction.
The optimum configuration at small separations is when one pair of quarks is in
an L ¼ 1 state, although the excitation energy is comparable to the decrease in
chromomagnetic energy, so there will still be a net increase in energy at small
separations.
Some tantalizing clues exist about the role of the quark–gluon interaction in
nuclear interactions, such as the small nuclear effects in deep inelastic lepton
scattering mentioned in Chapter 5. There is also a considerable experimental
programme in existence (for example at CEBAF, the superconducting accelerator
facility at the Jefferson Laboratory, Virginia, USA, mentioned in Chapter 4) to learn
more about the nature of the strong nucleon–nucleon force in terms of the
fundamental quark–gluon strong interaction and further progress in this area may
well result in the next few years. Meanwhile, in the absence of a fundamental theory
to describe the nuclear force, specific models and theories are used to interpret the
phenomena in different areas of nuclear physics. In the remainder of this chapter we
will discuss a number of such approaches.
4V
nðpÞdp ¼ dn ¼ p2 dp; ð7:1Þ
ð2hÞ3
6
Compare the mass of the (1232) resonance, where all three quarks spins are aligned, to that of the lighter
nucleon, where one pair of quarks spins is anti-aligned to give a total spin of zero. This is discussed in detail
in Section 3.3.3.
FERMI GAS MODEL 221
Figure 7.2 Proton and neutron potentials and states in the Fermi gas model
VðpnF Þ3 VðppF Þ3
N¼ and Z ¼ ð7:2Þ
32 h3 32 h3
4 4
V ¼ R3 ¼ R30 A; ð7:3Þ
3 3
where experimentally R0 ¼ 1:21 fm, as we have seen from electron and hadron
scattering experiments discussed in Chapter 2. Assuming for the moment that the
depths of the neutron and proton wells are the same, we find for a nucleus with
Z ¼ N ¼ A=2, the Fermi momentum
h 9 1=3
pF ¼ pnF ¼ ppF ¼ 250 MeV=c: ð7:4Þ
R0 8
Thus the nucleons move freely within the nucleus with quite large momenta.
The Fermi energy is
p2F
EF ¼ 33 MeV: ð7:5Þ
2M
The difference between the top of the well and the Fermi level is constant
for most heavy nuclei and is just the average binding energy per nucleon
222 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
~ B=A ¼ 7–8 MeV. The depth of the potential and the Fermi energy are to a
B
good approximation independent of the mass number A:
~ 40 MeV:
V0 ¼ EF þ B ð7:6Þ
Heavy nuclei generally have a surplus of neutrons. Since the Fermi levels of the
protons and neutrons in a stable nucleus have to be equal (otherwise the nucleus
can become more stable by -decay) this implies that the depth of the potential
well for the neutron gas has to be deeper than for the proton gas, as shown in
Figure 7.2. Protons are therefore on average less tightly bound in nuclei than are
neutrons.
We can use the Fermi gas model to give a theoretical expression for some of the
dependence of the binding energy on the surplus of neutrons, as follows. First, we
define the average kinetic energy per nucleon as
ð p F ð p F
1
2 2
hEkin i Ekin p dp p dp : ð7:7Þ
0 0
3
Ekin ðN; ZÞ ¼ NhEn i þ ZhEp i ¼ ½NðpnF Þ2 þ ZðppF Þ2 ; ð7:9Þ
10 M
where again we have taken the radii of the proton and neutron wells to be equal.
This expression is for fixed A but varying N and has a minimum at N ¼ Z. Hence
the binding energy gets smaller for N 6¼ Z. If we set N ¼ ðA þ Þ=2,
Z ¼ ðA
Þ=2, where N
Z, and expand Equation (7.10) as a power series
in =A, we obtain
" #
3 h2 9 2=3 5 ðN
ZÞ2
Ekin ðN; ZÞ ¼ Aþ þ ...: ; ð7:11Þ
10 M R20 8 9 A
which gives the dependence on the neutron excess. The first term contributes to the
volume term in the semi-empirical mass formula (SEMF), while the second
SHELL MODEL 223
describes the correction that results from having N 6¼ Z. This is a contribution to the
asymmetry term we have met before in the SEMF and grows as the square of the
neutron excess. Evaluating this term from Equation (7.11) shows that its contribu-
tion to the asymmetry coefficient defined in Equation (2.51) is about 44 MeV=c2 ,
compared with the empirical value of about 93 MeV=c2 given in Equation (2.54). In
practice, to reproduce the actual term in the SEMF accurately we would have to take
into account the change in the potential energy for N 6¼ Z.
The binding energy of electrons in atoms is due primarily to the central Coulomb
potential. This is a complicated problem to solve in general because in a multi-
electron atom we have to take account of not only the Coulomb field of the nucleus,
but also the fields of all the other electrons. Analytic solutions are not usually
possible. However, many of the general features of the simplest case of hydrogen
carry over to more complicated cases, so it is worth recalling the former.
Atomic energy levels are characterized by a quantum number n = 1, 2, 3, 4, . . . :
called the principal quantum number. This is defined so that it determines the energy
of the system.7 For any n there are energy-degenerate levels with orbital angular
momentum quantum numbers given by
‘ ¼ 0; 1; 2; 3; . . . ; ðn
1Þ ð7:12Þ
(this restriction follows from the form of the Coulomb potential) and for any value
of ‘ there are ð2‘ þ 1Þ sub-states with different values of the projection of orbital
angular momentum along any chosen axis (the magnetic quantum number):
m‘ ¼
‘;
‘ þ 1; . . . ; 0; 1; . . . ; ‘
1; ‘: ð7:13Þ
Due to the rotational symmetry of the Coulomb potential, all such sub-states are
degenerate in energy. Furthermore, since electrons have spin-12, each of the above
7
In nuclear physics we are not dealing with the same simple Coulomb potential, so it would be better to call n
the radial node quantum number, as it still determines the form of the radial part of the wavefunction.
224 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
Again, both these states will have the same energy. So finally, any energy eigenstate
in the hydrogen atom is labelled by the quantum numbers ðn; ‘; m‘ ; ms Þ and for any n,
there will be nd degenerate energy states, where
X
n
1
nd ¼ 2 ð2‘ þ 1Þ ¼ 2n2 : ð7:15Þ
‘¼0
i.e. there is a strong pairing effect for closed shells. In these cases it can be shown
that the Pauli principle implies
L ¼ S ¼ 0 and J ¼ L þ S ¼ 0: ð7:17Þ
For any atom with a closed shell or a closed sub-shell structure, the electrons are
paired off and thus no valence electrons are available. Such atoms are therefore
chemically inert. It is straightforward to work out the atomic numbers at which this
occurs. These are
For example, the inert gas argon ArðZ ¼ 18Þ has closed shells corresponding to
n ¼ 1, 2 and closed sub-shells corresponding to n ¼ 3; ‘ ¼ 0; 1. These values of Z
are called the atomic magic numbers.
SHELL MODEL 225
Figure 7.3 Binding energy per nucleon for even values of A: the solid curve is the SEMF (from
Bo69)
In nuclear physics, there is also evidence for magic numbers, i.e. values of Z and N at
which the nuclear binding is particularly strong. This can been seen from the B=A
curves of Figure 2.10 where at certain values of N and Z the data lie above the SEMF
curve. This is also shown in Figure 7.3, where the inset shows the low-A region
magnified. (The figure only shows results for even values of the mass number A.)
The nuclear magic numbers are found from experiment to be
and correspond to one or more closed shells, plus eight nucleons filling the s and p
sub-shells of a nucleus with a particular value of n. Nuclei with both N and Z having
one of these values are called doubly magic, and have even greater stability. An
example is the helium nucleus, the -particle.
Shell structure is also suggested by a number of other phenomena. For example:
‘magic’ nuclei have many more stable isotopes than other nuclei; they have very
226 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
small electric dipole moments, which means they are almost spherical, the most
tightly bound shape; and neutron capture cross-sections show sharp drops compared
with neighbouring nuclei. However, to proceed further we need to know something
about the effective potential.
A simple Coulomb potential is clearly not appropriate and we need some form that
describes the effective potential of all the other nucleons. Since the strong nuclear
force is short-ranged we would expect the potential to follow the form of the density
distribution of nucleons in the nucleus. For medium and heavy nuclei, we have seen
in Chapter 2 that the Fermi distribution fits the data and the corresponding potential
is called the Woods–Saxon form
V0
Vcentral ðrÞ ¼ : ð7:20Þ
1 þ eðr
RÞ=a
However, although these potentials can be shown to offer an explanation for the
lowest magic numbers, they do not work for the higher ones. This is true of all purely
central potentials.
The crucial step in understanding the origin of the magic numbers was taken in
1949 by Mayer and Jensen who suggested that by analogy with atomic physics there
should also be a spin–orbit part, so that the total potential is
where L and S are the orbital and spin angular momentum operators for a single
nucleon and V‘s ðrÞ is an arbitrary function of the radial coordinate.8 This form for the
total potential is the same as that used in atomic physics except for the presence of
the function V‘s ðrÞ. Once we have coupling between L and S then m‘ and ms are no
longer ‘good’ quantum numbers and we have to work with eigenstates of the total
angular momentum vector J, defined by J ¼ L þ S. Squaring this, we have
J2 ¼ L2 þ S2 þ 2L S; ð7:22Þ
i.e.
1
L S ¼ ðJ2
L2
S2 Þ ð7:23Þ
2
and hence the expectation value of L S, which we write as h‘si, is
h2
‘=2 for j ¼ ‘ þ 12
h‘si ¼ ½ jð j þ 1Þ
‘ð‘ þ 1Þ
sðs þ 1Þ ¼ :
2
ð‘ þ 1Þ=2 for j ¼ ‘
12
ð7:24Þ
8
For their work on the shell structure of nuclei. Maria Goeppert-Mayer and J. Hans Jensen were awarded a
half share of the 1963 Nobel Prize in Physics. (They shared the prize with Wigner, mentioned in Chapter 1
for his development of the concept of parity.)
SHELL MODEL 227
(We are always dealing with a single nucleon, so that s ¼ 12.) The splitting between
the two levels is thus
2‘ þ 1 2
Els ¼ h hV‘s i: ð7:25Þ
2
Experimentally, it is found that V‘s ðrÞ is negative, which means that the state with
j ¼ ‘ þ 12 has a lower energy than the state with j ¼ ‘
12. This is the opposite of the
situation in atoms. Also, the splittings are substantial and increase linearly with ‘.
Hence for higher ‘, crossings between levels can occur. Namely, for large ‘, the
splitting of any two neighbouring degenerate levels can shift the j ¼ ‘
12 state of the
initial lower level to lie above the j ¼ ‘ þ 12 level of the previously higher level.
An example of the resulting splittings up to the 1G state is shown in Figure 7.4,
where the usual atomic spectroscopic notation has been used, i.e. levels are written
n‘j with S, P, D, F, G, . . . : used for ‘ ¼ 0, 1, 2, 3, 4, . . .. Magic numbers occur when
there are particularly large gaps between groups of levels. Note that there is no
restriction on the values of ‘ for a given n because, unlike in the atomic case, the
strong nuclear potential is not Coulomb-like.
The configuration of a real nuclide (which of course has both neutrons and
protons) describes the filling of its energy levels (sub-shells), for protons and for
neutrons, in order, with the notation ðn‘j Þk for each sub-shell, where k is the
occupancy of the given sub-shell. Sometimes, for brevity, the completely filled
Figure 7.4 Low-lying energy levels in a single-particle shell model using a Woods--Saxon
potential plus spin--orbit term; circled integers correspond to nuclear magic numbers
228 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
sub-shells are not listed, and if the highest sub-shell is nearly filled, k can be given as
a negative number, indicating how far from being filled that sub-shell is. Using the
ordering diagram above, and remembering that the maximum occupancy of each
sub-shell is 2j þ 1, we predict, for example, the configuration for 178 O to be:
and
Notice that all the proton sub-shells are filled, and that all the neutrons are in filled
sub-shells except for the last one, which is in a sub-shell on its own. Most of the
ground state properties of 178 O can therefore be found from just stating the neutron
configuration as ð1d52 Þ1 .
The nuclear shell model can be used to make predictions about the spins of ground
states. A filled sub-shell must have zero total angular momentum, because j is always
an integer-plus-a-half, so the occupancy of the sub-shell, 2j þ 1, is always an even
number. This means that in a filled sub-shell, for each nucleon of a given mj ð¼ jz Þ
there is another having the opposite mj . Thus the pair have a combined mj of zero and
so the complete sub-shell will also have zero mj . Since this is true whatever axis we
choose for z, the total angular momentum must also be zero. Since magic number
nuclides have closed sub-shells, such nuclides are predicted to have zero contribu-
tion to the nuclear spin from the neutrons or protons or both, whichever are magic
numbers. Hence magic-Z/magic-N nuclei are predicted to have zero nuclear spin.
This is indeed found to be the case experimentally.
In fact it is found that all even-Z/even-N nuclei have zero nuclear spin. We can
therefore make the hypothesis that for ground state nuclei, pairs of neutrons and
pairs of protons in a given sub-shell always couple to give a combined angular
momentum of zero, even when the sub-shell is not filled. This is called the pairing
hypothesis. We can now see why it is the last proton and/or last neutron that
determines the net nuclear spin, because these are the only ones that may not be
paired up. In odd-A nuclides there is only one unpaired nucleon, so we can predict
precisely what the nuclear spin will be by referring to the filling diagram. For even-
A/odd-Z/odd-N nuclides, however, we will have both an unpaired proton and an
unpaired neutron. We cannot then make a precise prediction about the net spin
because of the vectorial way that angular momenta combine; all we can say is that
the nuclear spin will lie in the range jjp
jn j to ðjp þ jn Þ.
Predictions can also be made about nuclear parities. First, recall the following
properties of parity: (1) parity is the transformation r !
r; (2) the wavefunction
SHELL MODEL 229
PYml ð ;
Þ ¼ ð
Þ‘ Yml ð ;
Þ; ð7:27Þ
(3) a single-particle state will also have an intrinsic parity, which for nucleons is
defined to be positive. Thus the parity of a single-particle nucleon state depends
exclusively on the orbital angular momentum quantum number with P ¼ ð
1Þ‘ .
The total parity of a multiparticle state is the product of the parities of the individual
particles. A pair of nucleons with the same ‘ will therefore always have a combined
parity of þ1. The pairing hypothesis then tells us that the total parity of a nucleus is
found from the product of the parities of the last proton and the last neutron. So we
can predict the parity of any nuclide, including the odd/odd ones, and these
predictions are in agreement with experiment.
Unless the nuclear spin is zero, we expect nuclei to have magnetic (dipole)
moments, since both the proton and the neutron have intrinsic magnetic moments,
and the proton is electrically charged, so it can produce a magnetic moment when it
has orbital motion. The shell model can make predictions about these moments.
Using a notation similar to that used in atomic physics, we can write the nuclear
magnetic moment as
¼ gj jN ; ð7:28Þ
where N is the nuclear magneton that was used in the discussion of hadron
magnetic moments in Section 3.3.3, gj is the Landé g-factor and j is the nuclear spin
quantum number. For brevity we can write simply ¼ gj j nuclear magnetons.
We will find that the shell model does not give very accurate predictions for
magnetic moments, even for the even–odd nuclei where there is only a single
unpaired nucleon in the ground state. We will therefore not consider at all the much
more problematic case of the odd–odd nuclei having an unpaired proton and an
unpaired neutron.
For the even–odd nuclei, we would expect all the paired nucleons to contribute
zero net magnetic moment, for the same reason that they do not contribute to the
nuclear spin. Predicting the nuclear magnetic moment is then a matter of finding
the correct way to combine the orbital and intrinsic components of magnetic moment
of the single unpaired nucleon. We need to combine the spin component of the
moment, gs s, with the orbital component, g‘ ‘ (where gs and g‘ are the g-factors for
spin and orbital angular momentum.) to give the total moment gj j. The general
formula for doing this is9
jðj þ 1Þ þ ‘ð‘ þ 1Þ
sðs þ 1Þ jðj þ 1Þ
‘ð‘ þ 1Þ þ sðs þ 1Þ
gj ¼ g‘ þ gs ; ð7:29Þ
2jðj þ 1Þ 2jðj þ 1Þ
9
See, for example, Section 6.6 of En66.
230 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
and
1 1
jgj ¼ g‘ j 1 þ
gs j for j ¼ ‘
12: ð7:30bÞ
2‘ þ 1 2‘ þ 1
Since g‘ ¼ 1 for a proton and 0 for a neutron, and gs is approximately þ5.6 for the
proton and
3.8 for the neutron, Equations (7.30) yield the results (where
gprotonðneutronÞ is the g-factor for nuclei with an odd proton(neutron))
1
jgproton ¼ ‘ þ 5:6 ¼ j þ 2:3 for j ¼ ‘ þ 12
2
1 1 2:3j
jgproton ¼ j 1 þ
5:6 j ¼j
for j ¼ ‘
12
2‘ þ 1 2‘ þ 1 jþ1
ð7:31Þ
1
jgneutron ¼
3:8 ¼
1:9 for j ¼ j ¼ ‘ þ 12
2
1 1:9j
jgneutron ¼ 3:8 j ¼ for j ¼ ‘
12:
2‘ þ 1 jþ1
Accurate values of magnetic dipole moments are available for a wide range of
nuclei and plots of a sample of measured values for a range of odd-Z and odd-N
nuclei across the whole periodic table are shown in Figure 7.5. It is seen that for a
given j, the measured moments usually lie somewhere between the j ¼ ‘
12 and the
j ¼ ‘ þ 12 values (the so-called Schmidt lines), but beyond that the model does not
predict the moments accurately. The only exceptions are a few low-A nuclei where
the numbers of nucleons are close to magic values.
Why should the shell model work so well when predicting nuclear spins and
parities, but be poor for magnetic moments? There are several likely problem areas,
including the possibility that protons and neutrons inside nuclei may have effective
intrinsic magnetic moments that are different to their free-particle values, because of
their very close proximity to one another.
In principle, the shell model’s energy level structure can be used to predict nuclear
excited states. This works quite well for the first one or two excited states when there
is only one possible configuration of the nucleus. However, for higher states the
SHELL MODEL 231
Figure 7.5 Magnetic moments for odd-N, even-Z nuclei (upper diagram) and odd-Z, even-N
(lower diagram) as functions of nuclear spin compared with the predictions of the single-particle
shell model (the Schmidt lines)
moving an unpaired nucleon to the next highest level, or moving a nucleon from the
sub-shell below the unpaired nucleon up one level to pair with it. Thus it is necessary
to consider levels just above and below the last nucleons (protons and neutrons).
As an example, consider the case of 178 O. Its ground-state configuration is given in
Equations (7.26). All the proton sub-shells are filled, and all the neutrons are in filled
sub-shells except for the last one, which is in a sub-shell on its own. There are three
possibilities to consider for the first excited state:
3. promote the 1d52 neutron to the next level, which is probably 2s12 (or the nearby
1d32 ), giving a configuration of ð1s12 Þ1 or ð1d32 Þ1.
Following the diagram of Figure 7.4, the third of these possibilities would
correspond to the smallest energy shift, so it should be favoured over the others.
The next excited state might involve moving the last neutron up a further level to 1d32 ,
or putting it back where it was and adopting configurations (1) or (2). Option (2) is
favoured over (1) because it keeps the excited neutron paired with another, which
should have a slightly lower energy than creating two unpaired protons. When
comparing these predictions with the observed excited levels it is found that the
expected excited states do exist, but not necessarily in precisely the order predicted.
The shell model has many limitations, most of which can be traced to its
fundamental assumption that the nucleons move independently of one another in
a spherically symmetric potential. The latter, for example, is only true for nuclei that
are close to having doubly-filled magnetic shells and predicts zero electric quadruple
moments, whereas in practice many nuclei are deformed and quadruple moments are
often substantial. We discuss this important observation in the next section.
classical charge distribution with charge density ðxÞ within a volume
, then the
first moment that can be non-zero is the electric quadrupole Q, defined by
ð
eQ ðxÞð3z2
r 2 Þd3 x; ð7:32Þ
where we have taken the axis of symmetry to be the z-axis. The analogous definition
in quantum theory is
ð
1X
Q¼ qi ð3z2i
r2 Þ d3 x; ð7:33Þ
e i
where is the nuclear wavefunction and the sum is over all relevant nucleons,
each with charge qi .10 The quadrupole moment is zero if j j2 is spherically
symmetric and so a non-zero value of Q would be indicative of a non-spherical
nuclear charge distribution.
If we consider a spheroidal distribution with semi-axes defined as in Figure 2.14,
then evaluation of Equation (7.32) leads to the result
2
Qintrinsic ¼ Zeða2
b2 Þ; ð7:34Þ
5
where Qintrinsic is the value of the quadrupole moment for a spheroid at rest and Ze
is its total charge. For small deformations,
6
Qintrinsic ZeR2 "; ð7:35Þ
5
where " is defined in Equation (2.70) and R is the nuclear radius. Thus, for a
prolate distribution ða > bÞ, Q > 0 and for an oblate distribution ða < bÞ, Q < 0,
as illustrated in Figure 7.6. The same results hold in the quantum case.
If the nucleus has a spin J and magnetic quantum number M, then Q will depend
on M because it depends on the shape and hence the orientation of the charge
distribution. The quadrupole moment is then defined as the value of Q for which M
has its maximum value projected along the z-axis. This may be evaluated from
Equation (7.33) in the single-particle shell model and without proof we state the
resulting prediction that for odd-A, odd-Z nuclei with a single proton having a total
angular moment j outside closed sub-shells, the value of Q is given by
ð2j
1Þ
Q
R2 : ð7:36Þ
2ðj þ 1Þ
10 PÐ
The electric dipole moment dz ¼ 1e qi zi d
vanishes because it will contain a sum of terms of the
i
form h i jzi j i i, all of which are zero by parity conservation.
234 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
Figure 7.6 Shapes of nuclei leading to (a) Q > 0 (prolate), and (b) Q < 0 (oblate)
Thus, Q ¼ 0 for j ¼ 12. For odd-A, odd-N nuclei with a single neutron outside
closed sub-shells Q is predicted to be zero because the neutron has zero electric
charge, as will all even-Z, odd-N nuclei because of the pairing effect.
Unlike magnetic dipole moments, electric quadrupole moments are not always well
measured and the quoted experimental errors are often far larger than the differences
between the values obtained in different experiments. Significant (and difficult to
apply) corrections also need to be made to the data to extract the quadrupole moment
and this is not always done. The compilation of electric dipole moment data shown in
Figure 7.7 is therefore representative. The solid lines are simply to guide the eye and
Figure 7.7 Some measured electric quadrupole moments for odd-A nuclei, normalized by dividing
by R2, the squared nuclear radius: grey circles denote odd-N nuclei and black circles odd-Z nuclei;
the solid lines have no theoretical significance and the arrows denote the position of closed shells
NON-SPHERICAL NUCLEI 235
have no theoretical significance. The arrows indicate the positions of major closed
shells. A change of sign of Q at these points is expected because a nucleus with one
proton less than a closed shell behaves like a closed-shell nucleus with a negatively
charged proton (a proton hole) and there is some evidence for this in the data.
Two features emerge from this diagram. Firstly, while odd-A, odd-Z nuclei with
only a few nucleons outside a closed shell do have moments of order
R2, in general
the measured moments are larger by factors of two to three and for some nuclei the
discrepancy can be as large as a factor of 10. Secondly, odd-A, odd-N nuclei also
have non-zero moments, contrary to expectations and, moreover, there is little
difference between these and the moments for odd-A, odd-Z nuclei, except that
the former tend to be somewhat smaller. These results strongly suggest that for some
nuclei it is not a good approximation to assume spherical symmetry and that these
nuclei must be considered to have non-spherical mass distributions.
The first attempt to explain the measured electric quadrupole moments in terms
of non-spherical nuclei was due to Rainwater. His approach can be understood
using the model we discussed in Chapter 2 when considering fission and used
above to derive the results of Equations (7.34) and (7.35). There the sphere was
deformed into an ellipsoid (see Figure 2.14) with axes parameterized in terms of a
small parameter " via Equation (2.70). The resulting change in the binding energy
EB was found to be
EB ¼
"2 ; ð7:37Þ
where
1 2 1
¼ ð2as A3
ac Z 2 A
3 Þ ð7:38Þ
5
and the coefficients as and ac are those of the SEMF with numerical values given
in Equation (2.54). Rainwater assumed that this expression only held for closed-
shell nuclei, but not for nuclei with nucleons in unfilled shells. In the latter cases he
showed that distortion gives rise to an additional term in EB that is linear in ", so
that the total change in binding energy is
EB ¼
"2
"; ð7:39Þ
where is a parameter that could be calculated from the Fermi energy of the
nucleus. This form has a minimum value 2 =4 where " ¼
=2. The ground
state would therefore be deformed and not spherical.
Finally, once the spin of the nucleus is taken into account in quantum theory, the
measured electric quadrupole moment for ground states is predicted to be
jð2 j
1Þ
Q¼ Qintrinsic : ð7:40Þ
ð j þ 1Þð2 j þ 1Þ
236 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
This model gives values for Q that are of the correct sign, but overestimates them by
typically a factor of two. Refined variants of the model are capable of bringing the
predictions into agreement with the data by making better estimates of the parameter .
The Rainwater model is equivalent to assuming an aspherical liquid drop and Aage
Bohr (the son of Neils Bohr) and Mottelson showed that many properties of heavy
nuclei could be ascribed to the surface motion of such a drop. However, the single-
particle shell model cannot be abandoned because it explains many general features
of nuclear structure. The problem was therefore to reconcile the shell model with the
liquid-drop model. The outcome is the collective model.11
This model views the nucleus as having a hard core of nucleons in filled shells, as
in the shell model, with outer valence nucleons that behave like the surface
molecules of a liquid drop. The motions of the latter introduce non-sphericity in
the core that in turn causes the quantum states of the valence nucleons to change
from the unperturbed states of the shell model. Such a nucleus can both rotate and
vibrate and these new degrees of freedom give rise to rotational and vibrational
energy levels. For example, the rotational levels are given by EJ ¼ JðJ þ 1Þh2 =2I,
where I is the moment of inertia and J is the spin of the nucleus. The predictions of
this simple model are quite good for small J, but overestimate the energies for larger
J. Vibrational modes are due predominantly to shape oscillations, where the nucleus
oscillates between prolate and oblate ellipsoids. Radial oscillations are much rarer
because nuclear matter is relatively incompressible. The energy levels are well
approximated by a simple harmonic oscillator potential with spacing E ¼ h!,
where ! is the oscillator frequency.
In practice, the energy levels of deformed nuclei are very complicated, because
there is often coupling between the various modes of excitation, but nevertheless
many predictions of the collective model are confirmed experimentally.12
11
For their development of the collective model, Aage Bohr, Ben Mottelson and Leo Rainwater shared the
1975 Nobel Prize in Physics.
12
The details are discussed, for example, in Section 2.3 of Je90 and Chapter 17 or Ho97.
SUMMARY OF NUCLEAR STRUCTURE MODELS 237
Liquid-drop model
This model assumes that all nuclei have similar mass densities, with binding
energies approximately proportional to their masses, just as in a classical
charged liquid drop. The model leads to the SEMF, which gives a good description
of the average masses and binding energies. It is largely classical, with some
quantum mechanical terms (the asymmetry and pairing terms) inserted in an
ad hoc way. Input from experiment is needed to determine the coefficients of the
SEMF.
Shell model
This is a fully quantum mechanical model that solves the Schrödinger equation with
a specific spherical nuclear potential. It makes the same assumptions as the Fermi
gas model about the potential, but with the addition of a strong spin–orbit term. It is
able to successfully predict nuclear magic numbers, spins and parities of ground-
state nuclei and the pairing term of the SEMF. It is less successful in predicting
magnetic moments.
Collective model
This is also a fully quantum mechanical model, but in this case the potential is
allowed to undergo deformations from the strictly spherical form used in the shell
model. The result is that the model can predict magnetic dipole and electric
quadrupole magnetic moments with some success. Additional modes of excitation,
both vibrational and rotational, are possible and are generally confirmed by
experiment.
238 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
It is clear from the above that there is at present no universal nuclear model. What
we currently have is a number of models and theories that have limited domains of
applicability and even within which they are not always able to explain all the
observations. For example, the shell model, while able to give a convincing account
of the spins and parities of the ground states of nuclei, is unable to predict the spins of
excited states with any real confidence. And of course the shell model has absolutely
nothing to say about whole areas of nuclear physics phenomena. Some attempt has
been made to combine features of different models, such as is done in the collec-
tive model, with some success. A more fundamental theory will require the full
apparatus of many-body theory applied to interacting nucleons and some progress
has been made in this direction for light nuclei, as we will mention in Chapter 9.
A theory based on interacting quarks is a more distant goal.
7.6 a-Decay
To discuss -decays, we could return to the semiempirical mass formula of Chapter
2 and by taking partial derivatives with respect to A and Z find the limits of -
stability, but the result is not very illuminating. To get a very rough idea of the
stability criteria, we can write the SEMF in terms of the binding energy B. Then
-decay is energetically allowed if
If we now make the approximation that the line of stability is Z ¼ N (the actual
line of stability deviates from this, see Figure 2.7), then there is only one independent
variable. If we take this to be A, then
dB
Bð2; 4Þ > BðZ; AÞ
BðZ
2; A
4Þ 4 ; ð7:42Þ
dA
28:3 4½B=A
7:7 10
3 A; ð7:44Þ
which is a straight line on the B=A versus A plot which cuts the plot at A 151.
Above this value of A, Equation (7.41) is satisfied by most nuclei and -decay
becomes energetically possible.
a-DECAY 239
Figure 7.8 Schematic diagram of the potential energy of an -particle as a function of its
distance r from the centre of the nucleus
Lifetimes of -emitters span an enormous range, and examples are known from
10 ns to 1017 years. The origin of this large spread lies in the quantum mechanical
phenomenon of tunelling. Individual protons and neutrons have binding energies in
nuclei of about 8 MeV, even in heavy nuclei (see Figure 2.2), and so cannot in
general escape. However, a bound group of nucleons can sometimes escape because
its binding energy increases the total energy available for the process. In practice,
the most significant decay process of this type is the emission of an -particle,
because unlike systems of two and three nucleons it is very strongly bound by
7 MeV/ nucleon. Figure 7.8 shows the potential energy of an -particle as a function
of r, its distance from the centre of the nucleus.
Beyond the range of the nuclear force, r > R, the -particle feels only the
Coulomb potential
2Zhc
VC ðrÞ ¼ ; ð7:45Þ
r
where we now use Z to be the atomic number of the daughter nucleus. Within the
range of the nuclear force, r < R, the strong nuclear potential prevails, with its
strength characterized by the depth of the well. Since the -particle can escape from
the nuclear potential, E > 0. It is this energy that is released in the decay. Unless E
is larger than the Coulomb barrier (in which case the decay would be so fast as to be
unobservable) the only way the -particle can escape is by quantum mechanical
tunelling through the barrier.
The probability T for transmission through a barrier of height V and thickness
r by a particle of mass m with energy E is given approximately by
T e
2r ; ð7:46Þ
240 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
where h ¼ ½2mjVC
E j1=2 . Using this result, we can model the Coulomb
barrier as a succession of thin barriers of varying height. The overall transmission
probability is then
T ¼ e
G ; ð7:47Þ
with ¼ v=c and v is the velocity of the emitted particle.13 This assumes that
the orbital angular momentum of the -particle is zero, i.e. we ignore
possible centrifugal barrier corrections.14 Since rC is the value of r where
E ¼ VC ðrC Þ,
rC ¼ 2Ze2 =4"0 E ð7:49Þ
and hence
where m is the reduced mass of the -particle and the daughter nucleus, i.e.
m ¼ m mD =ðm þ mD Þ m . Evaluating the integral in Equation (7.51) gives
1=2 " rffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
#
2mc2
1 R R R
G ¼ 4Z cos
1
: ð7:52Þ
E rC rC rC
Finally, since E is typically 5 MeV and the height of the barrier is typically
40 MeV, rC R and from (7.52), G 4Z=, where ¼ v =c and v is the
velocity of the alpha particle within the nucleus.
The probability per unit time of the -particle escaping from the nucleus
is proportional to the product of: (a) the probability wðÞ of finding the -particle in
the nucleus; (b) the frequency of collisions of the -particle with the barrier (this
13
These formulae are derived in Appendix A.
14
The existence of an angular momentum barrier will suppress the decay rate (i.e. increase the lifetime)
compared with a similar nucleus without such a barrier. Numerical estimates of the suppression factors,
which increase rapidly with angular momentum, have been calculated by Blatt and Weisskopf and are given
in their book B152.
a-DECAY 241
is v =2R); and (c) the transition probability. Thus, combining these factors, is
given by
v
¼ wðÞ e
G ð7:53Þ
2R
and since
Z Z
G/ / pffiffiffiffiffiffi ; ð7:54Þ
E
small differences in E have strong effects on the lifetime.
To examine this further we can take logarithms of Equation (7.53) to give
1
log10 t12 ¼ a þ bZE 2 ; ð7:55Þ
where t12 is the half-life. The quantity a depends on the probability wðÞ and so is a
function of the nucleus, whereas b is a constant that may be estimated from the above
equations to be about 1.7. Equation (7.55) is a form of a relation that was found
empirically by Geiger and Nuttall in 1911 long before its theoretical derivation in
1928. It is therefore called the Geiger-Nuttall relation. It predicts that for fixed Z,
12
the log of the half-life of -emitters varies linearly with E .
Figure 7.9 shows lifetime data for the isotopes of four nuclei. The very strong
variation with -particle energy is evident; changing E by a factor of about 2.5
changes the lifetime by 20 orders of magnitude. In all cases the agreement with the
Geiger–Nuttall relation is very reasonable and the slopes are compatible with the
Figure 7.9 Comparison of the Geiger--Nuttall relation with experimental data for some
-emitters
242 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
estimate for b above. Thus the simple barrier penetration model is capable of
explaining the very wide range of lifetimes of nuclei decaying by -emission.
7.7 b-Decay
In Chapter 2 we discussed in some detail the phenomenology of -decay using the
SEMF. In this section we return to these decays and examine their theoretical
interpretation.
The first successful theory of nuclear -decay was proposed in the 1930s by Fermi,
long before the W and Z bosons were known and the quark model formulated. He
therefore had to construct a theory based on very general principles, working by
analogy with the quantum theory of electromagnetic processes (QED), the only
successful theory known at the time for quantum particles.
The general equation for electron -decay is
A
ZX ! Zþ1A Y þ e
þ e : ð7:56Þ
where ! is the transition rate (probability per unit time), Mfi is the transition
amplitude (also called the matrix element because it is one element of a matrix
whose elements are all the possible transitions from the initial state i to different final
states f of the system) and nðEÞ is the density of states, i.e. the number of quantum
states available to the final system per unit interval of total energy. The density-
of-states factor can be calculated from purely kinematical factors, such as energies,
momenta, masses and spins where appropriate.15 The dynamics of the process is
contained in the matrix element.
15
This is done explicitly in Appendix A.
b-DECAY 243
The matrix element can in general be written in terms of five basic Lorentz
^:
invariant interaction operators, O
ð
Mfi ¼ f ðgO
^ Þi d3 x; ð7:58Þ
where f and i are total wavefunctions for the final and initial states, respectively,
g is a dimensionless coupling constant, and the integral is over three-dimensional
space The five categories are called scalar (S), pseudo-scalar (P), vector (V), axial-
vector (A), and tensor (T ); the names having their origin in the mathematical
transformation properties of the operators. (We have met the V and A forms
previously in Chapter 6 on the electroweak interaction.) The main difference
between them is the effect on the spin states of the parent and daughter nuclei.
When there are no spins involved, and at low energies, ðgO ^ Þ is simply the interaction
potential, i.e. that part of the potential that is responsible for the change of state of the
system.
Fermi guessed that O ^ would be of the vector type, since electromagnetism is a
vector interaction, i.e. it is transmitted by a spin-1 particle – the photon. (Decays of
the vector type are called Fermi transitions.) We have seen from the work of
Chapter 6 that we now know that the weak interaction violates parity conservation
and is correctly written as a mixture of both vector and axial-vector interactions (the
latter are called Gamow–Teller transitions in nuclear physics), but as long as we are
not concerned with the spins of the nuclei, this does not make much difference, and
we can think of the matrix element in terms of a classical weak interaction potential,
like the Yukawa potential. Applying a bit of modern insight, we can assume the
potential is of extremely short range (because of the large mass of the W boson), in
which case we have seen that we can approximate the interaction by a point-like
form and the matrix element then becomes simply a constant, which we write as
GF
Mfi ¼ ; ð7:59Þ
V
4ðhcÞ3 W
GF ¼ : ð7:60Þ
ðMW c2 Þ2
assumption is also used in -decay. Experimental results are consistent with the
theory under this assumption. However, the theory goes beyond nuclear -decay,
and can be applied to any process mediated by the W boson, provided the energy is
not too great. In Chapter 6, for example, we used the same ideas to discuss
neutrino scattering. The best process to determine the value of GF is one not
complicated by hadronic (nuclear) effects and muon decay is usually used. The
lifetime of the muon
is given to a very good approximation by (ignoring the
Cabbibo correction)
1 ðm c2 Þ5
¼ G2 ;
6 F
ð7:61Þ
1923 hð
hcÞ
from which we can deduce that the value of GF is about 90 eV fm3. It is usually
quoted in the form GF =ðhcÞ3 ¼ 1:166 10
5 GeV
2 .
We see from Equation (7.58) that the transition rate (i.e. -decay lifetime) depends
essentially on kinematical factors arising through the density-of-states factor, nðEÞ.
To simplify the evaluation of this factor, we consider the neutron and proton to be
‘heavy’, so that they have negligible kinetic energy, and all the energy released in the
decay process goes into creating the electron and neutrino and in giving them kinetic
energy. Thus we write
E ¼ Ee þ E ; ð7:62Þ
where Ee is the total (relativistic) energy of the electron, E is the total energy
of the neutrino, and E is the total energy released. (This equals ðmÞc2 , if m is
the neutron–proton mass difference, or the change in mass of the decaying
nucleus.)
The transition rate ! can be measured as a function of the electron momentum, so
we need to obtain an expression for the spectrum of -decay electrons. Thus we will
fix Ee and find the differential transition rate for decays where the electron has
energy in the range Ee to Ee þ dEe . From the Golden Rule, this is
2
d! ¼ jMj2 n ðE
Ee Þne ðEe ÞdEe ; ð7:63Þ
h
where ne and n are the density of states factors for the electron and neutrino,
respectively. These may be obtained from our previous result:
V
nðpe Þdpe ¼ 4p2e dpe ; ð7:64Þ
ð2hÞ3
b-DECAY 245
dp E
¼ 2; ð7:65Þ
dE pc
so that
4V
nðEe ÞdEe ¼ pe Ee dEe ; ð7:66Þ
ð2hÞ3 c2
with a similar expression for nðE Þ. Using these in Equation (7.57) and setting
M ¼ GF =V, gives
d! G2
¼ 3 F7 4 pe Ee p E ð7:67Þ
dEe 2 h c
where in general
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p c ¼ E2
m2 c4 ¼ ðE
Ee Þ2
m2 c4 : ð7:68Þ
d! dEe d! G2
¼ ¼ 3 F7 2 p2e p E : ð7:69Þ
dpe dpe dEe 2
hc
Figure 7.10 Predicted electron spectra: Z ¼ 0, without Fermi screening factor; , with Fermi
screening factor
above treatment is essentially correct: these are the so-called ‘allowed transitions’.
However, the nucleus may change its spin by more than 1 unit, and then the
simplified short-range potential approach to the matrix element is inadequate. The
decay rate in these cases is generally suppressed, although not completely
forbidden, despite these being traditionally known as the ‘forbidden transitions’.16
In the second case, the electric potential between the positive nucleus and a
positive -particle will cause a shift of the low end of its momentum spectrum to
the right, since it is propelled away by electrostatic repulsion. Conversely, the low
end of the negative -spectrum is shifted to the left (see Figure 7.10). The precise
form of these effects is difficult to calculate, and requires quantum mechanics, but
the results are published in tables of a factor FðZ; Ee Þ, called the Fermi screening
factor, to be applied to the basic -spectrum.
The usual way of experimentally testing the form of the electron momentum
spectrum given by the Fermi theory is by means of a Kurie plot. From
Equation (7.70), with the Fermi screening factor included, we have
16
For a discussion of forbidden transitions see, for example, Co01.
b-DECAY 247
Figure 7.11 Kurie plot for the -decay of 36 Cl (the y-axis is proportional to the function HðEe Þ
above)
where KðZ; pe Þ includes FðZ; Ee Þ and all the fixed constants in Equation (7.71). A
plot of the left-hand side of this equation – using the measured d!=dpe and pe ,
together with the calculated value of KðZ; pe Þ – against the electron energy Ee should
then give a straight line with an intercept of E. An example is shown in Figure 7.11.
If the neutrino mass is not exactly zero then it is straightforward to repeat the above
derivation and to show that the left-hand side of the Kurie plot is proportional to
1 1
fðE
Ee Þ½ðE
Ee Þ2
m2 c4 2 g2 : ð7:73Þ
This will produce a very small deviation from linearity extremely close to the end-
point of the spectrum and the straight line will curve near the end point and cut the
0
axis vertically at E0 ¼ E0
m c2 . In order to have the best conditions for measuring
the neutrino mass, it is necessary to use a nucleus where a non-zero mass would have
a maximum effect, i.e. the maximum energy release E ¼ E0 should only be a few
keV. Also at such low energies atomic effects have to be taken into account, so the
initial and final atomic states must be very well understood. The most suitable case is
the decay of tritium,
3
H ! 3 He þ e
þ e ; ð7:74Þ
where E0 ¼ 18:6 keV. The predicted Kurie plot very close to the end-point is
shown in Figure 7.12.
Since the counting rate near E0 is vanishingly small, the experiment is extremely
difficult. In practice, the above formula is fitted to data close to the end-point of the
spectrum and extrapolated to E0 . The best experiments are consistent with a zero
248 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
Figure 7.12 Expected Kurie plot for tritium decay very close to the end-point of the electron
energy spectrum for two cases: m ¼ 0 and m ¼ 5 eV=c2
neutrino mass, but when experimental and theoretical uncertainties are taken into
account, an upper limit of about 2–3 eV/c2 results.
Gamma emission is a form of electromagnetic radiation and like all such radiation is
caused by a changing electric field inducing a magnetic field. There are two
possibilities, called electric (E) radiation and magnetic (M) radiation. These
names derive from the semiclassical theory of radiation, in which the radiation
field arises because of the time variation of charge and current distributions. The
classification of the resulting radiation is based on the fact that total angular
momentum and parity are conserved in the overall reaction, the latter because it is
an electromagnetic process.
g-EMISSION AND INTERNAL CONVERSION 249
The photon carries away a total angular momentum, given by a quantum number
L17, which must include the fact that the photon is a spin-1 vector meson. The
minimum value is L ¼ 1. This is because a real photon has two possible polarization
states corresponding, for example, to Lz ¼ 1. Thus in the transition, there must be a
change of Lz for the emitting nucleus of 1 and this cannot happen if L ¼ 0. Hence,
if the spins of the initial and final nuclei states are denoted by Ji and Jf respectively,
the transition Ji ¼ 0 ! Jf ¼ 0 is strictly forbidden. In general, the photons are said
to have a multipolarity L and we talk about multipole radiation; transitions are called
dipole ðL ¼ 1Þ, quadrupole ðL ¼ 2Þ, octupole ðL ¼ 3Þ etc.. Thus, for example, M2
stands for magnetic quadrupole radiation. The allowed values of L are restricted by
the conservation equation relating the photon total angular momentum L and the
spins of the initial and final nuclei states, i.e.
Ji ¼ Jf þ L: ð7:75Þ
Ji þ Jf L jJi
Jf j: ð7:76Þ
17
As this is the total angular momentum, logically it would be better to employ the symbol J. However, as L
is invariably used in the literature, it will be used in what follows.
250 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
2 2 Yes 1 E1
þ þ
1 0 No 1 M1
2þ 0þ No 2 E2
3
1þ
2 2 Yes 1, 2 E1, M2
2þ 1þ No 1, 2, 3 M1, E2, M3
3
5þ
2 2 Yes 1, 2, 3, 4 E1, M2, E3, M4
restriction on its states of polarization that a real photon does. In practice, the energy
of the virtual photon can be transferred to an orbital atomic electron that can thereby
be ejected. This is the process of internal conversion. There is another possibility
whereby the virtual photon can create an internal eþ e
pair. This is referred to as
internal pair production.
In semi-classical radiation theory, the transition probability per unit time, i.e. the
emission rate, is given by18
E;M 1 8ðL þ 1Þ 1 E 2Lþ1 E;M
Tfi ðLÞ ¼ Bfi ðLÞ; ð7:77Þ
4"0 L½ð2L þ 1Þ!!2 h hc
where E is the photon energy, E and M refer to electric and magnetic radiation,
and for odd-n, n!! nðn
2Þðn
4Þ . . . 3:1. The function BE;Mfi ðLÞ is the so-called
reduced transition probability and contains all the nuclear information. It is
essentially the square of the matrix element of the appropriate operator causing
the transition producing photons with multipolarity L, taken between the initial and
final nuclear state wave functions. For electric transitions, B is measured in units of
e2 fm2L and for magnetic transitions in units of ðN =cÞ2 fm2L
2 where N is the
nuclear magneton.
To go further requires knowledge of the nuclear wave functions. An approxima-
tion due to Weisskopf is based on the single-particle shell model. This approach
assumes that the radiation results from the transition of a single proton from an initial
orbital state of the shell model to a final state of zero angular momentum. In this
model the general formulas reduce to
2
e2 3RL
BE ðLÞ ¼ ð7:78aÞ
4 L þ 3
18
See, for example, Chapter 16 of Ja75.
g-EMISSION AND INTERNAL CONVERSION 251
for magnetic radiation, where R is the nuclear radius and mp is the mass of the
proton. Finally, from the work in Chapter 2 on nuclear sizes, we can substitute
R ¼ R0 A1=3 , with R0 ¼ 1:21 fm, to give the final results:
2
e2 3
E
B ðLÞ ¼ ðR0 Þ2L A2L=3 ð7:79aÞ
4 L þ 3
and
2
10 eh 2 3
BM ðLÞ ¼ ðR0 Þ2L
2 Að2L
2Þ=3 : ð7:79bÞ
2mp c Lþ3
Figure 7.13 shows an example of the transition rates T E;M calculated from
Equation (7.77) using the approximations of Equations (7.79). Although these are
only approximate predictions, they do confirm what is observed experimentally: for
a given transition there is a very substantial decrease in decay rates with increasing
L, and electric transitions have decay rates about two orders of magnitude higher
than the corresponding magnetic transitions.
Finally, it is often useful to have simple formulas for radiative widths . These
follow from Equations (7.77), (7.78) and (7.79) and for the lowest multipole
transitions may be written
Figure 7.13 Transition rates using single-particle shell model formulas of Weisskopf as a
function of photon energy for a nucleus of mass number A ¼ 60
252 CH7 MODELS AND THEORIES OF NUCLEAR PHYSICS
Problems
7.1 Assume that in the shell model the nucleon energy levels are ordered as shown in
Figure 7.4. Write down the shell-model configuration of the nucleus 73 Li and hence
find its spin, parity and magnetic moment (in nuclear magnetons). Give the two most
likely configurations for the first excited state, assuming that only protons are
excited.
state has jP ¼ 12 . By reference to Figure 7.4, suggest two possible configurations for
the latter state.
7.4 What are the configurations of the ground states of the nuclei 93 33
41 Nb and 16 S and what
values are predicted in the single-particle shell model for their spins, parities and
magnetic dipole moments?
7.5 Show explicitly that a uniformly charged ellipsoid at rest with total charge Ze and
semi-axes defined in Figure 2.14, has a quadrupole moment Q ¼ 25Zeða2
b2 Þ.
þ þ
7.7 The decay 244 240
98 Cfð0 Þ ! 96 Cmð0 Þ þ has a Q-value of 7.329 MeVand a half-life of
19.4 mins. If the frequency and probability of forming -particles (see Equation
þ þ
(7.53)) for this decay are the same as those for the decay 228 224
90 Thð0 Þ ! 88 Rð0 Þ þ ,
228
estimate the half-life for the -decay of 90 Th, given that its Q-value is 5.520 MeV.
7.9 The reaction 34 Sðp, nÞ34 Cl has a threshold proton laboratory energy of 6.45 MeV.
Calculate non-relativistically the upper limit of the positron energy in the -decay of
34
Cl, given that the mass difference between the neutron and the hydrogen atom is
0.78 MeV.
PROBLEMS 253
7.10 To determine the mass of the electron neutrino from the -decay of tritium requires
measurements of the electron energy spectrum very close to the end-point where
there is a paucity of events (see Figure 7.12.). To see the nature of the problem,
estimate the fraction of electrons with energies within 10 eV of the end-point.
7.11 The electron energy spectra of -decays with very low-energy end-points E0 may be
approximated by d!=dE ¼ E1=2 ðE0
EÞ2. Show that in this case the mean energy is
1
3E0 .
7.13 Use the Weisskopf formulas of Equations (7.79) to calculate the radiative width
ðE3Þ expressed in a form analogous to Equations (7.80).
8
Applications of Nuclear Physics
Nuclear physics impinges on our everyday lives1 in a way that particle physics
does not, at least not yet. A minor example of this is radioactive dating of historical
artefacts which we discussed in Chapter 2. It is appropriate, therefore, to discuss
some of these applications. For reasons of space, we will consider just three
important areas: fission, fusion and biomedical applications, concentrating in the
latter on medical imaging and the therapeutic use of radiation.
8.1 Fission
Fission was discussed in Chapter 2 in the context of the semi-empirical mass
formula and among other things we showed that spontaneous fission only occurs
for very heavy nuclei. In this section we discuss fission in more detail in the
context of its use in the production of energy.
In Chapter 2 we saw that for a nucleus with A 240, the Coulomb barrier, which
inhibits spontaneous fission, is between 5 and 6 MeV. If a neutron with zero kinetic
energy enters a nucleus to form a compound nucleus, the latter will have an
excitation energy above its ground state equal to the neutron’s binding energy in
that state. For example, a zero-energy neutron entering a nucleus of 235 U forms a
state of 236 U with excitation energy of 6.5 MeV. This energy is well above the
fission barrier and the compound nucleus quickly undergoes fission, with decay
products similar to those found in the spontaneous decay of 236 U. To induce fission
1
This is literally true, because we shall see that the energy of the Sun has its origins in nuclear reactions.
in 238 U on the other hand, requires a neutron with kinetic energy of at least
1.2 MeV. The binding energy of the last neutron in 239 U is only 4.8 MeV and an
excitation energy of this size is below the fission threshold of 239 U.
The differences in the binding energies of the last neutron in even-A and odd-A
nuclei are given by the pairing term in the semi-empirical mass formula.
Examination of the value of this term (see Equation (2.52)) leads to the explana-
tion of why the odd-A nuclei
are ‘fissile’, i.e. fission may be induced by even zero-energy neutrons, whereas the
even-A (even-Z/even-N) nuclei
1. At energies below 0.1 eV, tot for 235 U is much larger than that for 238 U and the
fission fraction is large ( 84 per cent). (The other 16 per cent is mainly
radiative capture with the formation of an excited state of 236 U, plus one or
more photons.)
2. In the region between 0.1 eV and 1 keV, the cross-sections for both isotopes
show prominent peaks corresponding to neutron capture into resonances. The
widthsof these states are 0:1 eV and thus their lifetimes are of the order of
h f 1014 s. In the case of 235 U these compound nuclei lead to fission,
f
whereas in the case of 238 U, neutron capture leads predominantly to radiative
decay of the excited state.
3. Above 1 keV, the ratio fission =tot for 235 U is still significant, although smaller
than at very low energies. In both isotopes, tot is mainly due to contributions
from elastic scattering and inelastic excitation of the nucleus.
The fission fragments (which are not unique – several final states are possible)
carry away about 90 per cent of the energy of the primary fission reaction. The
accompanying neutrons in the primary fission process (referred to as prompt
FISSION 257
Figure 8.1 Total cross-section tot and fission cross-section fission as functions of energy for
neutrons incident on (a)235 U and (b)238 U (adapted from Ga76, Courtesy of Brookhaven National
Laboratory)
neutrons) carry away only about 2 per cent of the energy. For 235 U, the average
number of prompt neutrons per fisson is n 2:5, with the value depending a little
on the incident neutron energy and they have an average energy of about 2 MeV.
In addition to the neutrons produced in the primary fission, the decay products
will themselves decay by chains of -decays and some of the resulting nuclei will
258 CH8 APPLICATIONS OF NUCLEAR PHYSICS
themselves give off further neutrons. This delayed component constitutes about
13 per cent of the energy release in the fission of 235 U. Although the mean delay is
about 13 s, some components have very long lifetimes and may not decay until
many years later. One consequence of this is that a reactor will still produce heat
even after it has ceased to be used for power production and another is that the
delayed component may be emitted after the fuel has been used and removed from
the reactor, leading to the biological hazard of radioactive waste.2
We have seen in Chapter 2 that in each fission reaction a large amount of energy is
produced, which of course is what is needed for power production. However, just
as important is the fact that the fission decay products contain other neutrons. For
example, we have said that in the case of fission of 235 U, on average n ¼ 2:5
neutrons are produced. Since neutrons can induce fission, the potential exists for a
sustained chain reaction, although a number of conditions have to be fulfilled for
this to happen in practice. If we define
then for k ¼ 1 the process is said to be critical and a sustained reaction can occur.
This is the ideal situation for the operation of a power plant based on nuclear
fission. If k < 1, the process is said to be subcritical and the reaction will die out; if
k > 1, the process is supercritical and the energy will grow very rapidly, leading to
an uncontrollable explosion, i.e. a nuclear fission bomb.
Again we will focus on uranium as the fissile material and consider the length
and timescales for a chain reaction to occur. If we assume that the uranium is a
mixture of the two isotopes 235 U and 238 U with an average neutron total cross-
section tot , then the mean free path, i.e. the mean distance the neutron travels
between interactions (see Chapter 4), is given by
‘ ¼ 1=ð
nucl tot Þ; ð8:4Þ
where
nucl ¼ 4:8 1028 nuclei/m3 is the nuclei density of uranium metal. For
example, the average energy of a prompt neutron from fission is 2 MeV and at this
energy we can see from Figure 8.1 that tot 7 barns, so that ‘ 3 cm. A 2 MeV
neutron will travel this distance in about 1:5 109 s.
Consider first the case of the explosive release of energy in a nuclear bomb,
using the highly enriched isotope 235 U (for simplicity we will assume a sample of
2
We will return to the effect of radiation on living tissue later in this chapter, in Section 8.3.1.
FISSION 259
100 per cent 235 U). From Figure 8.1, we see that a neutron with energy of 2 MeV
has a probability of about 18 per cent to induce fission in an interaction with a 235 U
nucleus. Otherwise it will scatter and lose energy, so that the probability for a
further interaction will be somewhat increased (because the cross-section increases
with decreasing energy). If the probability of inducing fission in a collision is p, the
probability that a neutron has induced fission after n collisions is pð1 pÞn1 and
the mean number of collisions to induce fission will be
X
1
n ¼ npð1 pÞn1 ; ð8:5Þ
n¼1
provided the neutron does not escape outside the target. The value of n can be
estimated using the measured cross-sections
pffiffiffi and is about six. Thus the neutron will
move a linear (net) distance of 3 6 cm 7 cm in a time tp 108 s before
inducing a further fission and being replaced on average by 2.5 new neutrons with
average energy of 2 MeV.3
The above argument suggests that the critical mass of uranium 235 U that would
be necessary to produce a nuclear explosion is a sphere of radius about 7 cm.
However, not all neutrons will be available to induce fission. Some will escape
from the surface and some will undergo radiative capture. If the probability that a
newly created neutron induces fission is q, then each neutron will on average lead
to the creation of ðnq 1Þ additional neutrons in the time tp . If there are NðtÞ
neutrons present at time t, then at time t þ t there will be
Nðt þ tÞ ¼ NðtÞ 1 þ ðnq 1Þ ðt tp Þ ; ð8:6Þ
3
The square root appears because we are assuming that at each collision the direction changes randomly, i.e.
the neutron executes a random walk. Thus if the distance travelled in the ith collision is li , the displacement
P
n
vector d after n collisions will be d ¼ li and the net distance travelled d will be given by
i¼1
n X
X n
d2 ¼ ðli
lj Þ ¼ l21 þ l22 þ l23 þ
þ l2n þ 2ðl1
l2 þ l1
l3 þ
Þ:
i¼1 j¼1
When the average is taken over several collisions, the scalar products will cancel because the direction of
pffiffiffi
each step is random. Finally, setting li ¼ l, the mean distance travelled per collision, gives d ¼ l n.
260 CH8 APPLICATIONS OF NUCLEAR PHYSICS
The production of power in a controlled way for peaceful use is carried out in a
nuclear reactor and is just as complex as producing a bomb. There are several
distinct types of reactor available. We will discuss just one of these, the thermal
reactor, which uses uranium as the fuel and low-energy neutrons to establish a
chain reaction. The discussion will concentrate on the principles operating such a
reactor and not on practical details.
A schematic diagram of the main elements of a generic example of a thermal
reactor is shown in Figure 8.2. The most important part is the core, shown
schematically in Figure 8.3. This consists of fissile material (fuel elements),
control rods and the moderator. The roles of the control rods and the moderator
will be explained later. The most commonly used fuel is uranium and many
thermal reactors use natural uranium, even though it has only 0.7 per cent of 235 U.
Because of this, a neutron is much more likely to interact with a nucleus of 238 U.
However, a 2 MeV neutron from the primary fission has very little chance of
inducing fission in a nucleus of 238 U. Instead it is much more likely to scatter
inelastically, leaving the nucleus in an excited state and after a couple of such
collisions the energy of the neutron will be below the threshold of 1.2 MeV for
inducing fission in 238 U.
FISSION 261
Figure 8.2 Sketch of the main elements of a thermal reactor -- the components are not to scale
(after Li01, Copyright, John Wiley & Sons)
A neutron with its energy so reduced will have to find a nucleus of 235 U if it is
to induce fission, but its chances of doing this are very small unless its energy
has been reduced to very low energies below 0.1 eV, where the cross-section is
large (see Figure 8.1). Before that happens it is likely to have been captured into
one of the 238 U resonances with the emission of photons. Thus, to sustain a chain
reaction, either the fuel must be enriched with a greater fraction of 235 U (2–
3 per cent is common in some types of commercial reactor), or if natural uranium
is to be used, some method must be devised to overcome this problem.
This is where the moderator comes in. This surrounds the fuel elements and
its volume is much greater than that of the fuel elements themselves. Its main
purpose is to slow down fast neutrons produced in the fission process. Fast
neutrons will escape from the fuel rods into the moderator and are reduced to very
low energies by elastic collisions. In this way the absorption into resonances of
238
U is avoided. The moderator must therefore be a material with a negligible
cross-section for absorption and ideally should also be inexpensive. In practice,
heavy water (a form of water where the hydrogen atoms are replaced by atoms of
deuterium), or carbon (in the form of graphite), are the moderators of choice in
many thermal reactors using natural uranium. For enriched reactors, ordinary
water may be used.
Consider now the stability of the chain reaction. This is where the control rods
play their part. They are usually made of cadmium, which has a very high
absorption cross-section for neutrons. By mechanically manipulating the control
rods, i.e. by retracting or inserting them, the number of neutrons available to
induce fission can be regulated. This mechanism is the key to maintaining a
constant k value of unity and therefore a constant power output. However, safe
working of the reactor is not possible with prompt neutrons alone. To see this, we
return to Equation (8.9) and set nq 1 ¼ k 1, so that
NðtÞ ¼ Nð0Þ exp ðk 1Þt tp : ð8:10Þ
The value of tp is determined by the mean free path for neutron absorption and
unlike the case of pure 235 U we considered in Section 8.1.2, is given approximately
by tp 103 s. Thus, for example, if we take k ¼ 1:001, i.e. an increase of only
0.1 per cent, the reactor flux would increase by e60 1026 in only one minute.
Clearly a much smaller rate of increase has to be achieved for safe manipulation of
the control rods if a disaster is to be averted. This is where the delayed neutrons
play a crucial role.
In a nuclear weapon, the delayed neutrons are of no consequence, because the
explosion will have taken place long before they would have been emitted, but in a
power reactor they are vital for reactor safety. Taking account of delayed neutrons,
each fission leads to ½ðn þ nÞq 1 additional neutrons, where we have defined
n as the number of delayed neutrons per fission. In practice n 0:02. In the
steady-state operation, with constant energy output, the neutron density must
FISSION 263
remain constant (i.e. k ¼ 1 in Equation (8.3)). Thus q must satisfy the critical
condition
ðn þ nÞq 1 ¼ 0: ð8:11Þ
Equation (8.10) is now modified to have an additional term that depends on the
mean time td of the delayed neutrons, which is about 13 s. Provided
nðk 1Þ n, it is the latter term that dominates and (without proof) the modified
form of Equation (8.10) is given approximately by
nðk 1Þt
NðtÞ Nð0Þ exp : ð8:12Þ
½n nðk 1Þtd
Thus the timescale to manipulate the control rods is determined by that of the
delayed neutrons. For example, using n ¼ 2:5; n ¼ 0:02; k ¼ 1:001 and td ¼ 13 s
in Equation (8.12) gives an increase in reactor flux of less than a factor two in one
minute. Clearly, the precise increase is sensitive to the parameters chosen, but
factors of this size are manageable. The reactor design therefore ensures that
nq 1 < 0 always, so that the reactor can only become critical in the presence of
delayed neutrons.
This simple discussion has ignored many practical details that will modify the
real formulas used in reactor dynamics, such as the fact that the fuel and moderator
are not uniformly distributed throughout the core and that some of the fission
products themselves have appreciable cross-sections for neutron absorption and
will therefore reduce the flux of neutrons available to sustain the chain reaction.4
Returning to Figure 8.2, the core is surrounded by a coolant (often water), which
removes the heat generated in the core from the energy deposited by the fission
fragments. A thick concrete shield to prevent radiation leaks surrounds the entire
set-up. At start-up, the value of k is set slightly higher than unity and is kept at that
value until the desired power output is achieved and the operating temperature is
reached, after which the value of k is lowered by adjusting the control rods. It is
very important for safety reasons that dq=dT < 0, so that an increase in tempera-
ture T leads to a fall in reaction rate. The rest of the plant is conventional
engineering. Thus, the heated coolant gives up its heat in a heat exchanger and is
used to boil water and drive a steam turbine, which in turn produces electricity.
It is worth calculating the efficiency with which one can expect to produce
energy in a nuclear reactor. We can use the SEMF to calculate the energy released
during fission, by finding the binding energies of the two fission products and
comparing their sum to the binding energy of the decaying nucleus. For the fission
of a single 235 U nucleus this is 200 MeV or 3:2 1011 J. (As we have mentioned
above, about 90 per cent of this is in the form of ‘prompt’ energy.) We also know
4
More details of reaction dynamics are discussed in, for example, Section 10.3 of Li01. In Section 10.6 of
this reference there is also a discussion of several other types of commercial reactor.
264 CH8 APPLICATIONS OF NUCLEAR PHYSICS
Wf ¼ J N fission ; ð8:13Þ
where J is the flux, N is the number of nuclei undergoing fission and fission is the
fission cross-section. Consider, for example, a reactor containing 100 tonnes of
natural uranium, generating a neutron flux of 1013 cm2 s1 and with a fission
cross-section for 235 U of 580 b at the appropriate energy (see Figure 8.1). Since the
fraction of 235 U in natural uranium is 0.072 per cent, the number of 235 U nuclei
undergoing fission is given by
where A ¼ 238:03 is the mass number of natural uranium. The power generated is
thus
P ¼ Wf E; ð8:15Þ
where E ¼ 200 MeV is the total energy released per fission (see above). Evaluating
Equation (8.15) gives P 340 MW. In addition to causing fission, neutrons will be
absorbed by 235 U without causing fission. If the total absorption cross-section a is
680 b, then the number of 235 U nuclei that will be consumed per second will be
N Ja , i.e. 1:24 1019 s1 . Since we started with 1:82 1027 nuclei, the fuel will
be used at the rate of about 1.8 per cent per month.
We turn now to the fast breeder reactor mentioned above. In this reactor there
is no large volume of moderator and no large density of thermal neutrons is
established. In such a reactor, the proportion of fissile material is increased to
about 20 per cent and fast neutrons are used to induce fission. The fuel used is
239
Pu rather than 235 U, the plutonium being obtained by chemical separation from
the spent fuel rods of a thermal reactor. This is possible because some 238 U nuclei
in the latter will have captured neutrons to produce 239 U, which subsequently
FISSION 265
n þ 238 U ! 239 U ð23 minsÞ ! 239 Np ð2:4 daysÞ ! 239 Pu ð2:4 104 yearsÞ: ð8:16Þ
The mean number of neutrons produced in the fission of 239 Pu is 2.96, so this
nucleus is very suitable for use in a fast reactor. In practice, the core is a mixture of
20 per cent 239 Pu and 80 per cent 238 U surrounded by a blanket of more 238 U,
where more plutonium is made. The 238 U obtained from spent fuel rods in thermal
reactors is called depleted uranium. Such a reactor can produce more fissile 239 Pu
than it consumes, hence the name ‘breeder’. In principle such a reactor can
consume all the energy content of natural uranium, rather than the 1 per cent used
in thermal reactors, although in practice there are limits to its efficiency.
Whatever type of reactor is used, a major problem is the generation of
radioactive waste, including transuranic elements and long-lived fission fragments,
which in some cases may have to be stored safely for hundreds of years.5 Much
effort has been expended on this problem, but a totally satisfactory solution is still
not available. Short-lived waste with low activity (for example, consumables such
as protective clothing) is simply buried in the ground. One idea for long-lived
waste with high activity is to ‘glassify’ it into stable forms that can be stored
underground without risk of spillage, leakage into the water table, etc..
A particularly ingenious idea is to ‘defuse’ long-lived fission fragments by using
the resonance capture of neutrons to convert them to short-lived, or even stable,
nuclei. For example, 99 Tc (technetium), which concentrates in several organs of
the body and also in the blood, has a very long half-life. However, it has a large
resonant cross-section for neutron capture to a completely stable isotope 100 Ru
(ruthenium) and in principle this reaction could be used to ‘neutralize’ 99 Tc.
Needless to say, the problems to be overcome are far from trivial. First, the amount
of radioactive waste is very large, so one problem is to find a source of neutrons
capable of handling it. (Reactors themselves are one possible source!) Secondly,
the neutron energy has to be matched to the particular waste material, which
therefore would have to be separated and prepared before being bombarded by
the neutrons. All this would take energy and would increase the overall cost of
energy production by nuclear power, which is already more expensive than
conventional burning of fossil fuels. Nevertheless, there is considerable interest
in the principle of this method and proposals have been made to exploit it without
the attendant drawbacks above. We will return to this in Chapter 9, where we
discuss some of the outstanding problems in nuclear physics and their possible
future solutions. However, until such time as this, or some other, method is
realized in practice, the safe long-term disposal of radioactive waste remains a
serious unsolved problem.
5
In principle, there would be no such problem with fast breeder reactors, but in practice the ideal is not
realized.
266 CH8 APPLICATIONS OF NUCLEAR PHYSICS
8.2 Fusion
We have seen that the plot of binding energy per nucleon (Figure 2.2) has a
maximum at A 56 and slowly decreases for heavier nuclei. For lighter nuclei,
the decrease is much quicker, so that with the exception of magic nuclei, lighter
nuclei are less tightly bound than medium size nuclei. Thus, in principle, energy
could be produced by two light nuclei fusing to produce a heavier and more
tightly bound nucleus – the inverse process to fission. Just as for fission, the energy
released comes from the difference in the binding energies of the initial and final
states. This process is called nuclear fusion. Since light nuclei contain fewer
nucleons than heavier nuclei, the energy released per fusion is smaller than in
fission. However, as a potential source of power, this is more than balanced by the
far greater abundance of stable light nuclei in nature than very heavy nuclei. Thus
fusion offers enormous potential for power generation, if the huge practical
problems could be overcome.
1 ZZ 0 e2
VC ¼ ; ð8:17Þ
4"0 R þ R0
where Z and Z 0 are the atomic numbers of the two nuclei and R and R0 are their
effective radii. The quantity ðR þ R0 Þ is therefore classically the distance of closest
approach. Recalling, from the work on nuclear structure in Chapter 2, that for
1
medium and heavy nuclei R ¼ 1:2A3 fm, we have
e2 hc Z Z 0 Z Z0
VC ¼ 1=3 ¼ 1:198 MeV: ð8:18Þ
4"0 hc 1:2 A1=3 þ ðA0 Þ fm A1=3 þ ðA0 Þ1=3
Thus, for example, with A 8, VC 4:8 MeV and this energy has to be supplied
to overcome the Coulomb barrier.
This is a relatively small amount of energy to supply and it might be thought
that it could be achieved by simply colliding two accelerated beams of light nuclei,
but in practice nearly all the particles would be elastically scattered. The only
FUSION 267
practical way is to heat a confined mixture of the nuclei to supply enough thermal
energy to overcome the Coulomb barrier. The temperature necessary may
be estimated from the relation E ¼ kT, where kB is Boltzmann’s constant, given
by kB ¼ 8:6 105 eV K1. For an energy of 4.8 Mev, this implies a temperature
of 5:6 1010 K. This is well above the typical temperature of 108 K found in
stellar interiors.6 It is also the major hurdle to be overcome in achieving a
controlled fusion reaction in a reactor, as we shall see later.
Fusion actually occurs at a lower temperature than this estimate due to a
combination of two reasons. The first and most important is the phenomenon of
quantum tunnelling, which means that the full height of the Coulomb barrier does
not have to be overcome. In Chapter 7 we discussed a similar problem in the
context of -decay, and we can draw on that analysis here. The probability of
barrier penetration depends on a number of factors, but the most important is the
Gamow factor, which is a function of the relative velocities and the charges of the
reaction products. In particular, the probability is proportional to exp½GðEÞ,
where GðEÞ is apgeneralization
ffiffiffiffiffiffiffiffiffiffiffiffi of the Gamow factor of Chapter 7. This may be
written as G ¼ EG =E, where again, generalizing the equations in Chapter 7,
EG ¼ 2mc2 ðZ1 Z2 Þ2 : ð8:20Þ
Here, m is the reduced mass of the two fusing nuclei and they have electric charges
Z1 e and Z2 e. Thus the probability of barrier penetration increases as E increases.
Nevertheless, the probability of fusion is still extremely small. For example, if we
consider the fusion of two protons (which we will see below is an important
ingredient of the reactions that power the Sun), at a typical stellar temperature of
107 K, we find EG 490 keV and E 1 keV. Hence the probability of fusion is
proportional to exp½ðEG =EÞ1=2 expð22Þ 109:6 which is a very large
suppression factor and so the actual fusion rate is still extremely slow.
The other reason that fusion occurs at a lower temperature than expected is that
a collection of nuclei at a given mean temperature, whether in stars or elsewhere,
will have a Maxwellian distribution of energies about the mean and so there will be
some with energies substantially higher than the mean energy. Nevertheless, even a
stellar temperature of 108 K corresponds to an energy of only about 10 keV, so the
fraction of nuclei with energies of order 1 MeV in such a star would only be of the
order of expðE=kTÞ expð100Þ 1043 , a minute amount. We will return to
these questions in more detail in Section. 8.2.3.
The energy of the Sun comes from nuclear fusion reactions, foremost of which is
the so-called proton–proton cycle. This has more than one branch, but one of these,
6
Because of this, many scientists refused to accept that fusion occurred in stars when the suggestion was first
made. When challenged on this, Eddington’s reposte was simple: ‘. . . . go and find a hotter place’.
268 CH8 APPLICATIONS OF NUCLEAR PHYSICS
the PPI cycle, is dominant. This starts with the fusion of hydrogen nuclei to
produce deuterium via the weak interaction:
1
H þ 1 H ! 2 H þ eþ þ e þ 0:42 MeV: ð8:21Þ
The deuterium then fuses with more hydrogen to produce 3 He via the electro-
magnetic interaction:
1
H þ 2 H ! 3 He þ þ 5:49 MeV ð8:22Þ
and finally, two 3 He nuclei fuse to form 4 He via the nuclear strong interaction:
3
He þ 3 He ! 4 He þ 2ð1 HÞ þ 12:86 MeV: ð8:23Þ
The relatively large energy release in the last reaction is because 4 He is a doubly
magic nucleus and so is very tightly bound. The first of these reactions, being a
weak interaction, proceeds at an extremely slow rate and sets the scale for the long
lifetime of the Sun. Combining these equations, we have overall
Because the temperature of the Sun is 107 K, all its material is fully ionized.
Matter in this state is referred to as a plasma. The positrons produced above will
annihilate with electrons in the plasma to release a further 1.02 MeV of energy per
positron and so the total energy released is 26.72 MeV. However, of this each
neutrino will carry off 0.26 MeV on average, which is lost into space.7 Thus on
average, 6.55 MeV of electromagnetic energy is radiated from the Sun for every
proton consumed in the PPI chain.
The PPI chain is not the only fusion cycle contributing to the energy output of
the Sun, but it is the most important. Another interesting cycle is the carbon, or
CNO chain. Although this contributes only about 3 per cent of the energy output
of the Sun, it plays an important role in the evolution of other stellar objects. In the
presence of any of the nuclei 126 C, 136 C, 147 N or 157 N, hydrogen will catalyse burning
via the reactions
12
C þ 1 H ! 13 N þ þ 1:95 MeV
ð8:25Þ
13
N ! 13 C þ eþ þ e þ 1:20 MeV
13
C þ 1 H ! 14 N þ þ 7:55 MeV ð8:26Þ
14
N þ 1 H ! 15 O þ þ 7:34 MeV
ð8:27Þ
15
O ! 15 N þ eþ þ e þ 1:68 MeV
7
These are the main contributors to the neutrino flux observed at the surface of the Earth that was discussed
in Chapter 3.
FUSION 269
and
15
N þ 1 H ! 12 C þ 4 He þ 4:96 MeV ð8:28Þ
These and other fusion chains all produce electron neutrinos as final-state products
and using detailed models of the Sun, the flux of such neutrinos at the surface of
the Earth can be predicted.8 The actual count rate is far lower than the theoretical
expectation. This is the solar neutrino problem that we met in Section 3.1.4. The
solution to this problem is almost certainly neutrino oscillations, where some e
are converted to neutrinos of other flavours in their passage from the Sun to the
Earth. We saw in Chapter 3 that this is only possible if neutrinos have mass, so a
definitive measurement of neutrino masses would be an important piece of
evidence to finally resolve the solar neutrino problem. Such measurements should
be available in a few years.
The process whereby heavier elements (including the 12 C required in the
CNO cycle) are produced by fusion of lighter ones can continue beyond the
reactions above. For example, when the hydrogen content is depleted, at high
temperatures helium nuclei can fuse to form an equilibrium mixture with 8 Be via
the reaction
4
He þ 4 He Ð 8 Be ð8:30Þ
4
He þ 8 Be ! 12 C ð8:31Þ
to occur, where C is an excited state of carbon. A very small fraction of the latter
will decay to the ground state, so that overall we have9
8
The expectations are based on a detailed model of the Sun known as the standard solar model that we met in
Chapter 3.
9
The occurrence of this crucial reaction depends critically on the existence of a particular excited state of 12 C.
For a discussion of this and the details of the other reactions mentioned below see, for example, Section 4.3
of Ph94. Very recent experiments (2005) have found evidence for other nearby excited states that change the
accepted energy dependence (or equivalently the temperature dependence) of this reaction which will have
implications for theories of stellar evolution.
270 CH8 APPLICATIONS OF NUCLEAR PHYSICS
We have mentioned in Section 8.2.1 that quantum tunnelling and the Maxwellian
distribution of energies combine to enable fusion to occur at a lower temperature
than might at first be expected. The product of the increasing barrier penetration
factor with energy and the Maxwellian decreasing exponential actually means that
in practice fusion takes place over a rather narrow range of energies. To see this we
will consider the fusion between two types of nuclei, a and b, having number
densities na and nb (i.e. the number of particles per unit volume) and at a
temperature T. We assume that the temperature is high enough so that the nuclei
form a plasma, with uniform values of number densities and temperature. We also
assume that the velocities of the two nuclei are given by the Maxwell–Boltzmann
distribution, so that the probability of having two nuclei with a relative speed v in
the range v to v þ dv is
1=2
2 m 3=2 mv2 2
PðvÞ dv ¼ exp v dv; ð8:35Þ
kT 2kT
where m is the reduced mass of the pair. The fusion reaction rate per unit volume is
then
Rab ¼ na nb hab vi; ð8:36Þ
where ab is the fusion cross-section10 and the brackets denote an average, i.e.
ð
1
10
The product nA nB is the number of pairs of nuclei that can fuse. If the two nuclei are of the same type, with
nA ¼ nB ¼ n, then the product must be replaced by 12 nðn 1Þ 12 n2, because in quantum theory such nuclei
are indistinguishable.
FUSION 271
where the exponential follows from the previous discussion of quantum tunnelling
and SðEÞ contains the details of the nuclear physics. The term 1=E is conveniently
factored out because many nuclear cross-sections have this behaviour at low
energies. Using (8.35) and (8.38) in (8.37) gives, from (8.36):
1=2
3=2 1
ð "
1=2 #
8 1 E EG
Rab ¼ na nb SðEÞexp dE: ð8:39Þ
m kT kT E
0
Because the factor 1=E has been taken out of the expression for ðEÞ, the quantity
SðEÞ is slowly varying and the behaviour of the integrand is dominated by the
behaviour of the exponential term. The falling exponential of the Maxwellian
energy distribution combines with the rising exponential of the quantum tunnelling
effect to produce a maximum in the integrand at E ¼ E0 where
1=3
1
E0 ¼ EG ðkTÞ2 ð8:40Þ
4
and fusion takes place over a relatively narrow range of energies E0 E0 where
4 1=6
E0 ¼ EG ðkTÞ5=6 : ð8:41Þ
31=2 21=3
The importance of the temperature and the Gamow energy EG ¼ 2mc2 ðZa Zb Þ2
is clear. A schematic illustration of the interplay between these two effects is
shown in Figure 8.4.
As a real example, consider the pp reaction (Equation (8.21)), at a temperature
of 2 107 K. We have EG ¼ 493 keV and kT ¼ 1:7 keV, so that fusion is most
likely at E0 ¼ 7:2 keV and the half-width of the distribution is E0 =2 ¼ 4:1 keV.
The resulting function exp½E=kT ðEG =EÞ1=2 is shown in Figure 8.5.
In the approximation where SðEÞ is taken as a constant SðE0 Þ, the integral in
Equation (8.39) may be done and gives
1=2
8 2
hab vi SðE0 Þ 2 exp½ ; ð8:42Þ
9 3mEG
2=3
where ¼ 3 12 ðEG =kTÞ1=3 :
272 CH8 APPLICATIONS OF NUCLEAR PHYSICS
Figure 8.4 The right-hand dashed curve is proportional to the barrier penetration factor and
the left-hand dashed curve is proportional to the Maxwell distribution. The solid curve is the
combined effect and is proportional to the overall probability of fusion with a peak at E0 and a
width of E0
Figure 8.5 The exponential part of the integrand in Equation (8.39) for the case of pp fusion at
a temperature of 2 107 K
FUSION 273
with
1=3
Aa Ab 1 keV 1=3
¼ 18:8ðZa Zb Þ2=3 : ð8:44Þ
Aa þ Ab kT
The rate depends very strongly on both the temperature and the nuclear species
because of the factor 2 exp½ . This is illustrated in Figure 8.6 for the pp and
p12 C reactions, the initial reactions in the pp and CNO cycles.
Figure 8.6 The function 2 expðÞ of Equation (8.43) for the pp and p12 C reactions
and
2
1H þ 21 H ! 31 H þ p þ 4:03 MeV ð8:45bÞ
274 CH8 APPLICATIONS OF NUCLEAR PHYSICS
Figure 8.7 Values of the quantity h i for the d--t reaction of Equation (8.46) and the
combined d--d reactions of Equations (8.45) (adapted from Ke82 and reproduced by permission
of Annual Reviews)
suggest that deuterium might be a suitable fuel for a fusion reactor. Deuterium is
also present in huge quantities in sea water and is easy to separate at low cost.
An even better reaction in terms of energy output is deuterium–tritium fusion:
2
1H þ 31 H ! 42 He þ n þ 17:62 MeV: ð8:46Þ
The values of hvi for the d–t reaction of Equation (8.46) and the combined d–d
reactions of Equations (8.45) are shown in Figure 8.7. It can be seen that the
deuterium–tritium (d–t) reaction has the advantage over the deuterium–deuterium
(d–d) reaction of a much higher cross-section. The heat of the reaction is also
greater. The principal disadvantage is that tritium does not occur naturally (it has a
mean life of only 17.7 years) and has to be manufactured, which increases the
overall cost. From Figure 8.7 it can be seen that the rate for the d–t reaction peaks
at about E ¼ kT ¼ 3040 keV and a working energy where the cross-section is
still considered reasonable is about 20 keV, i.e. 3 108 K.
The effective energy produced by the fusion process will be reduced by the heat
radiated by the hot plasma. The mechanism for this is predominantly electron
bremmstrahlung. The power loss per unit volume due to this process is propor-
tional to T 1=2 Z 2 , where Z is the atomic number of the ionized atoms. Thus for a
plasma with given constituents and at a fixed ion density, there will be a minimum
FUSION 275
temperature below which the radiation losses will exceed the power produced
by fusion. For example, for the d–t reaction with an ion density 1021 m3,
kTmin 4 keV. It would be 10 times larger for the d–d reaction of Equation (8.45a)
because of the form of hvi (see Figure 8.7), which is another reason for using the
d–t reaction. In practice, the situation is worse than this because most of the
neutrons in Equation (8.46) will escape, so even at the theoretical ‘break-even’
temperature, external energy would have to be supplied to sustain the fusion
process. Only when the energy deposited in the plasma by the -particles exceeds
the radiation losses would the reaction be self-sustaining. This is referred to as the
‘ignition point’.
A numerical expression that embodies these ideas is the so-called Lawson
criterion, which provides a measure of how close to practicality is a particular
reactor design. We will assume a d–t reaction. To achieve a temperature T in a
deuterium–tritium plasma, there has to be an input of energy 4nd ð3kT=2Þ per unit
volume. Here nd is the number density of deuterium ions and the factor of 4 comes
about because nd is equal to the number density of tritium ions and the electron
density is twice this, giving 4nd particles per unit volume. The reaction rate in the
plasma is n2d hdt vi. If the plasma is confined for time tc , then per unit volume of
plasma,
where Q is the energy released in the fusion reaction. For a useful device, L > 1.
For example, If we assume k T ¼ 20 keV and use the experimental value
hdt vi 1022 m3 s1 , then the Lawson criterion may be written
Thus either a very high particle density or a long confinement time, or both, is
required.
At the temperatures required for fusion, any material container will vaporize and
so the central problem is how to contain the plasma for sufficiently long times for
the reaction to take place. The two main methods are magnetic confinement and
inertial confinement. Both techniques present enormous technical challenges. In
practice, most work has been done on magnetic confinement and so this method
will be discussed in more detail than the inertial confinement method.
In magnetic confinement, the plasma is confined by magnetic fields and heated
by electromagnetic fields. Firstly we recall the behaviour of a particle of charge q
in a uniform magnetic field B, taking the two extreme cases where the velocity v of
the particle is (a) at right angles to B and (b) parallel to B. In case (a) the particle
traverses a circular orbit of fixed radius (compare the principle of the cyclotron
discussed in Chapter 4) and in case (b) the path is a helix of fixed pitch along the
direction of the field (compare the motion of electrons in a time projection
276 CH8 APPLICATIONS OF NUCLEAR PHYSICS
chamber, also discussed in Chapter 4). Two techniques have been proposed to stop
particle losses: magnetic ‘mirrors’ and a geometry that would ensure a stable
indefinite circulation. In the former, it is arranged that the field in a region is
greater at the boundaries of the region than in the interior. Then as the particle
approaches the boundary, the force it experiences will develop a component that
points into the interior where the field is weaker. Thus the particle is trapped and
will oscillate between the interior and the boundaries.11 However, most practical
work has been done on case (b) and for that reason we will restrict our discussion
to this technique.
The simplest configuration is a toroidal field produced by passing a current
through a doughnut-shaped solenoid. In principle, charged particles in such a field
would circulate endlessly, following helical paths along the direction of the
magnetic field. In practice, the field would be weaker at the outer radius of the
torus and the non-uniformity of the field would produce instabilities in the orbits of
some particles and hence lead to particle loss. To prevent this a second field is
added called a poloidal field. This produces a current around the axis of the torus
and under the combined effect of both fields, charged particles in the plasma
execute helical orbits about the mean axis of the torus. Most practical realizations
of these ideas are devices called tokamaks, in which the poloidal field is generated
along the axis of the torus through the plasma itself.
One of the largest tokamaks in existence is the Joint European Torus (JET), which
is a European collaboration and sited at the Culham Laboratory in Berkshire, UK. A
schematic view of the arrangement of the fields in JET is shown in Figure 8.8(a).
This shows the external coils that generate the main toroidal field. The poloidal field
is generated by transformer action on the plasma. The primary windings of the
transformer are shown with the plasma itself forming the single-turn secondary.
The current induced in the plasma not only generates the poloidal field, but also
supplies several megawatts of resistive heating to the plasma. However, even this is
insufficient to ensure a sufficient temperature for fusion and additional energy is
input via other means, including rf sources.
In the inertial confinement method, small pellets of the deuterium–tritium ‘fuel’
mixture are bombarded with intense energy from several directions simultaneously
which might, for example, be supplied by pulsed lasers. As material is ejected from the
surface, other material interior to the surface is imploded, compressing the core of the
pellet to densities and temperatures where fusion can take place. The laser pulses are
extremely short, typically 107 109 s, which is many orders of magnitude shorter
than the times associated with the pulsed poloidal current in a tokamak (which could
be as long as 1s), but this is compensated for by much higher plasma densities.
Considerable progress has been made towards the goal of reaching the ignition
point. However, although appropriate values of nd , tc , and T have been obtained
11
The Van Allen radiation belts that occur at high altitudes consist of charged particles from space that have
become trapped by a magnetic mirror mechanism because the Earth’s magnetic field is stronger at the poles
than at the equator.
Figure 8.8 Schematic diagrams showing: (a) the main magnetic field components of the JET tokamak; (b) how these elements are incorporated into the JET
device (courtesy of EFDA--JET)
278 CH8 APPLICATIONS OF NUCLEAR PHYSICS
separately, to date no device has yet succeeded in achieving the Lawson criterion.
Tokamaks have reached the break-even point, but the best value of the Lawson
ratio that has been achieved is still about a factor of two too small. Much work
remains to be done on this important problem and in recognition of this at least one
major new tokamak machine is planned as a global collaboration. Even when the
ignition point is achieved, experience with fission power reactors suggests that it
will probably take decades of further technical development before fusion power
becomes a practical reality.
12
This has been known for a long time. For example, Hermann Muller was awarded the 1946 Nobel Prize in
Physiology and Medicine for his discovery that mutations can be induced by X-rays.
BIOMEDICAL APPLICATIONS 279
dD AðMBqÞ E ðMeVÞ
ðSv h1 Þ ð8:49aÞ
dt 6r2 ðm2 Þ
and for an internal source emitting radiation of energy ER, the effective dose rate
for an organ of mass M is
dD AER f
¼ ; ð8:49bÞ
dt M
13
For a discussion of Equations (8.49) and quantitative issues of acceptable doses for various sections of the
population and to different organs, see for example, Chapter 7 of Li01 and Chapter 11 of De99.
280 CH8 APPLICATIONS OF NUCLEAR PHYSICS
high, peaking near the end of its range and so the penetrating power is low. For
example, a 1 MeV -particle travels only a few tens of microns and is easily stopped
by skin. However, considerable damage can be caused to sensitive internal organs if
an -emitting isotope is ingested. An exception to the above is neutron radiation,
which being electrically neutral does not produce primary ionization. Its primary
interaction is via the nuclear strong force and it will mainly scatter from protons
contained in the high percentage of water present. The scattered protons will,
however, produce ionization as discussed above. The overall effect is that neutrons
are more penetrating than other heavy particles and at MeV energies can deposit
their energy to a depth of several centimetres. Electrons also lose energy by
interaction with electrons, but the rate of energy loss is smaller than for heavy
particles. Also, as they have small mass, they are subject to greater scatter and so their
paths are not straight lines. In addition, electrons can in principle lose energy by
bremsstrahlung, but this is not significant in the low Z materials that make up the
patient. The overall result is that electrons are more penetrating than heavy particles
and deposit their energy over a greater volume. Finally, photons lose energy via a
variety of processes (see Section 4.4.4), the relative importance of which depends on
the photon energy. Photons are very penetrating and deposition of their energy is not
localized.
In addition to the physical damage that may be caused by the primary ionization
process, there is also the potential for chemical damage, as mentioned above. This
comes about because most of the primary interactions result in the ionization of
simple molecules and the creation of neutral atoms and molecules with an
unpaired electron. The latter are called free radicals (much discussed
in advertising material for health supplements). These reactions occur on much
longer timescales of about 106 s. For example, ionization of a water molecule
produces a free electron and a positively charged molecule:
H2 O ! H2 Oþ þ e ð8:50aÞ
radiation
and the released electron is very likely to be captured by another water molecule
producing a negative ion:
e þ H2 O ! H2 O : ð8:50bÞ
Both ions are unstable and dissociate to create free radicals (denoted by black
circles):
H2 Oþ ! Hþ þ OH ð8:51aÞ
and
H2 O ! H þ OH : ð8:51bÞ
Free radicals are chemically very active, because there is a strong tendency for
their electrons to pair with one in another free radical. Thus the free radicals in
BIOMEDICAL APPLICATIONS 281
Equations (8.51) will interact with organic molecules (denoted generically by RH)
to produce organic free radicals:
RH þ OH ! R þ H2 O ð8:52aÞ
and
RH þ H ! R þ H2 : ð8:52bÞ
The latter may then induce chemical changes in critical biological structures (e.g.
chromosomes) some way from the site of the original radiation interaction that
produced them. Alternatively, the radiation may interact directly with the molecule
RH again releasing a free radical R :
Finally, if the irradiated material is rich in oxygen, yet another set of reactions is
possible:
R þ O2 ! RO2 ; ð8:54aÞ
followed by
with the release of another free radical. This is the oxygen effect mentioned above
that complicates the treatment of tumours.
Fortunately, for low-level radiation, living matter itself has the ability to repair
much of the damage caused by radiation and so low-level radiation does not lead to
permanent consequences. Indeed, if this were not so, then life may not have
evolved in the way it has, because we are all exposed to low levels of naturally-
occurring radiation throughout our lives (which may well have been far greater in
the distant past) and the modern use of radiation for a wide range of industrial and
medical purposes has undoubtedly increased that exposure. However, the repair
mechanism is not effective for high levels of exposure.
In the context of radiation therapy an important quantity is the linear energy
transfer (LET) which measures the energy deposited per unit distance over the
path of the radiation. Except for bremsstrahlung, LET is the same as dE=dx
discussed in Chapter 4. High-LET particles are heavy ions and -particles, which
lose their energy rapidly and have short ranges. LET values of the order of
100 keV=mm and ranges 0.1–1.0 mm are typical. Low-LET particles are electrons
and photons with LET values of the order of 1 keV=mm and ranges of the order of
1 cm. Much cancer therapy work uses low-LET particles. Treatment consists of
directing a beam at a cancer site from several directions to reduce the exposure
282 CH8 APPLICATIONS OF NUCLEAR PHYSICS
of healthy tissue, while maintaining the total dose to the tumour. Other techniques
include giving the dose in several stages so that the outer regions of the tumour,
which are relatively oxygen-rich, are successively destroyed as they become re-
oxygenated. Other treatments, particularly for localized cancers, involve the
introduction of a radionuclide either physically via a needle or by ingestion/
injection of a compound containing the radionuclide. Chemicals that preferentially
target specific organs or bones are commonly used.
Neutron therapy, as an example of a high-LET particle, is not widely used
because of the problem of producing a strongly collimated beam plus the difficulty
of ensuring that the energy is deposited primarily at the tumour site. Neutrons also
share with low-LET radiation the drawback that their attenuation in matter is
exponential. On the other hand, the rate of energy loss of protons and other
charged particles increases with penetration depth, culminating at a maximum, the
Bragg peak, close to the end of their range. In principle, this means that a greater
fraction of the energy would be deposited at the tumour site and less damage
would be caused along the path length to the site. There is also an increasing
interest in using heavy charged particles. For example, carbon ions at the
beginning of their path in tissue have a low rate of energy loss more like an
LET particle, but near the end of their range the local ionization increases
dramatically as it approaches the Bragg peak. Thus considerable energy can be
deposited at a precise depth without the danger of massive destruction of healthy
tissue en route to the target. Another potential advantage is that nuclear interac-
tions along the path length will convert a small fraction of the nuclei to radioactive
positron-emitting isotopes which could then be used to image the irradiated region
(using the PET technique described below) and thus monitor the effectiveness of
the treatment programme. Unfortunately, the use of protons and heavy particles
requires access to an accelerator and for this reason proton and heavy ion therapy
is not commonly used.
There are several techniques for producing images useful for diagnostic purposes
and in this section we will describe the principles of the main ones, but without
technical details.14
14
A readable account of medical imaging at the appropriate level is given in Chapter 7 of Li01 and a short
useful review of the whole field is He97.
BIOMEDICAL APPLICATIONS 283
front the patient and a detector (usually a special type of sensitive film) placed
immediately behind the patient. Because the radiation is absorbed according to an
exponential law, a measurement of the intensities just before and after the patient
yields information on the integrated mean free path (or equivalently the attenuation
coefficient 1=) of the photons in the body.
Thus, referring to Figure 8.9, we have for the ray shown, using Equation (4.17),
xð2
The full image reveals variations of this integral only in two dimensions and thus
contains no depth information. A three-dimensional effect comes from overlapping
shadows in the two-dimensional images and part of the skill of a radiologist is to
interpret these effects.
The most commonly used radiation is X-rays. The attenuation coefficient is
dependent on the material and is greater for elements with high Z than for elements
with low Z. Thus X-rays are good for imaging bone (which contains calcium with
Z ¼ 20), but far less useful for imaging soft tissue (which contains a high
proportion of water).
Images can also be obtained using an internal source of radiation. This is done
by the patient ingesting, or being injected with, a substance containing a radio-
active -emitting isotope. As photon detectors are very sensitive, the concentration
of the radioisotope can be very low and any risk to the patient is further minimized
by choosing an isotope with a short lifetime. If necessary, the radioisotope can be
combined in a compound that is known to be concentrated preferentially in a
specific organ if that is to be investigated, for example iodine in the thyroid. In
practice, more than 90 per cent of routine investigations use the first excited state
of 99
43 Tc as the radioisotope. This has a lifetime of about 6 hours and is easily
produced from the -decay of 99 42 Mo which has a lifetime of 67 hours. The
284 CH8 APPLICATIONS OF NUCLEAR PHYSICS
usefulness of this metastable state (written 99 Tcm ) is that it emits a single 140 keV
photon with negligible -decay modes, decaying to the very long-lived
(2 105 years) ground state.
Because the radiation is emitted in all directions, a different technique is used
to detect it. The patient is stationary and is scanned by a large-area detector consisting
of a collimated single-crystal scintillator, usually NaI, the output from which is
viewed by an array of photomultipliers (PMTs) via a light guide (see Section 4.4.2).
A schematic diagram of such a -camera is shown in Figure 8.10. The output from
the scintillator is received in several PMTs and the relative intensities of these signals
depend on the point of origin. The signals can be analysed to locate the point to
within a few millmetres. The collimator restricts the direction of photons that can
be detected and, combined with the information from the PMTs, the overall
spatial resolution is typically of the order of 10 mm, provided the region being
examined has an attenuation coefficient that differs by at least 10 per cent from its
surroundings.
Radioisotope investigations principally demonstrate function rather than anat-
omy, in contrast to X-ray investigations that show mainly anatomical features.
Thus better images of soft tissue, such as tumours, can be obtained than those
obtained using external X-rays, because the ability of the tumour to metabolize has
been exploited, but the exact location of the tumour with respect to the anatomy is
often lost or poorly defined.
Figure 8.11 shows part of a whole body skeletal image of a patient who had been
injected with a compound MDP which moves preferentially to sites of bone
cancer, labelled with the isotope 99 Tcm . The image clearly shows selective take-up
of the isotope in many tumours distributed throughout the body. (The concentra-
tion in the bladder is probably not significant.)
BIOMEDICAL APPLICATIONS 285
Figure 8.11 Part of a whole-body skeletal image obtained using 99 Tcm MDP (image courtesy of
Prof. R. J. Ott, Royal Marsden Hospital, London, UK)
286 CH8 APPLICATIONS OF NUCLEAR PHYSICS
Computed tomography
15
The CT system was devised independently by Godfrey Hounsfield and Allan Cormack who were jointly
awarded the 1979 Nobel Prize in Physiology and Medicine for their work.
BIOMEDICAL APPLICATIONS 287
Figure 8.13 (a) X-ray CT scan of the brain, and (b) SPECT brain scan using a 99 Tcm labelled
blood flow tracer, showing high perfusion in the tumour (indicated by arrows) (image courtesy of
Prof. R. J. Ott, Royal Marsden Hospital, London, UK)
288 CH8 APPLICATIONS OF NUCLEAR PHYSICS
some sense the ‘inverse’ of that in Figure 8.12. Thus the source is now within the
patient and the fixed ring of detectors is replaced by one or more -cameras
designed so that they can rotate in a circle about the patient. An example of an
image obtained using SPECT is shown in Figure 8.13(b).
For a number of technical reasons, including the fact that the emitted radiation is
isotropic, there are more stringent requirements on the -cameras and SPECT
images have a resolution of only about 10 mm. However, although not suitable for
accurate quantitative measurement of anatomy, they are of great use for clinical
diagnostic work involving function. For example, the technique is used to make
quantitative measurements of the functioning of an organ, i.e. clearance rates in
kidneys, lung volumes, etc..
Since radionuclide imaging provides functional and physiological information,
it would be highly desirable to be able to image the concentrations of elements
such as carbon, oxygen and nitrogen that are present in high abundances in
the body. The only radioisotopes of these elements that are suitable for imaging
are short-lived positron emitters: 11 C (half-life 20 min), 13 N (10 min) and 15 O
( 2 min). For these emitters, the radiation detected is the two -rays emitted when
the positron annihilates with an electron. This occurs within a few millimetres
from the point of production of the positron, whose initial energy is typically less
than 0.5 MeV. The photons each have energies equal to the rest mass of an
electron, i.e. 0.511 MeV and emerge ‘back-to-back’ to conserve momentum. This
technique is called positron emission tomography (PET) and was mentioned
earlier in connection with radiation treatment using heavy ions.
The arrangement of a PET scanner is shown in Figure 8.14. If the detectors D1
and D2 detect photons of the correct energy in coincidence, then the count rate is a
measure of the integral of the source activity within the patient along the line AB
passing through P. The ring of detectors defines a plane through the patient and the
Figure 8.15 Part of a whole-body PET scan showing uptake of the chemical FDG (labelled by
99 m
Tc ) in lung cancer (image courtesy of Prof. R. J. Ott, Royal Marsden Hospital, London, UK)
complete set of data from all combinations of detector pairs contains all
the information needed to generate the set of line integrals which can be
converted into a two-dimensional image of the source using standard CT image
reconstruction techniques. An example of an image using the PET technique is
shown in Figure 8.15.
This account of medical imaging has ignored many technical points. For
example, there are a number of corrections that have to be made to the raw
data, particularly in the SPECT technique, and the most useful radioisotopes used
in PET are produced in a cyclotron, so the scanner has to be near such a facility,
which considerably limits is use. The interested reader is referred to specialized
texts for further details.16
16
For a more detailed discussion see, for example, De99.
290 CH8 APPLICATIONS OF NUCLEAR PHYSICS
17
Felix Bloch and Edward Purcell were awarded the 1952 Nobel Prize in Physics for their discovery of
nuclear magnetic resonance (NMR) and subsequent researches. Although the term NMR is still used in
research environments, the term magnetic resonance imaging (MRI) is preferred in clinical environments to
prevent patients associating the technique with ‘harmful nuclear radiation’.
18
See, for example, De99, McR03 and Ho97a.
19
In general, the nuclear resonance frequency is defined by f ¼ jljB=j h, where j is the spin of the particle
involved and is its magnetic dipole moment.
BIOMEDICAL APPLICATIONS 291
Figure 8.16 (a) Precession of the magnetization M in the xy-plane under the action of a torque
M B resulting from an external field B; (b) motion viewed in a frame of reference ðx 0 ; y0 ; z0 Þ
rotating at the Larmor frequency about the z-axis the r.f. pulse Brf applied in the x 0 -direction
has rotated M so that it points in the y0 -direction
Crucially, the frequency of the external r.f. field must match the Larmor frequency
of the protons to be excited.
The induced signal will decay as equilibrium is restored. If B were uniform
throughout the selected region, all the protons would precess at the same frequency
and remain in phase. In that case the interaction of the proton spins with the
surrounding lattice, the so-called spin-lattice interactions, would cause M to relax
to its equilibrium state M0 parallel to B. Under reasonable assumptions the
radiated signal is proportional to the difference ðM0 MÞ and decreases expo-
nentially with a characteristic spin-lattice, or longitudinal, relaxation time T1 .
Typical spin-lattice relaxation times are of the order of a few hundred milliseconds
and are significantly different for different materials, such as muscle, fat and water.
However, because there are always small irregularities in the field due to local
atomic and nuclear effects, individual protons actually precess at slightly different
rates and the signal decays because the component of M orthogonal to B (i.e. in the
xy-plane) decreases as the individual moments loose phase coherence. This
decrease is characterized by a second time T2 , called the spin–spin, or transverse,
relaxation time. This is normally much shorter than T1 , but again varies with
material. Both relaxation times can be measured.
The above assumes that the external field B is perfectly uniform, but of course
the ideal is not realized in practice. The effects of macroscopic inhomogeneity in
the magnetic field can be eliminated by generating so-called spin echoes, which
may crudely be described as making two ‘orthogonal’ measurements such that the
unwanted effects cancel out exactly in the sum. Many MRI imaging sequences use
this technical device and again we refer the interested reader to the literature cited
in Footnote 18 for further details.
All the above assumes we are scanning the whole body. The original develop-
ment of the method as a medical diagnostic technique is due to the realization that
gradients in the static magnetic field could be used to encode the signal with
precise spatial information and be processed to generate two-dimensional images
corresponding to slices through the tissue of the organ being examined.20 The
patient is placed in the fixed field B pointing along the z-direction. A second static
field Bg parallel to z, but with a gradient in the z-direction is then applied so that
the total static field is a function of z. This means that the Larmor frequency (which
is proportional to the magnetic field) will vary as a function of z. Thus when the r.f.
field Brf is applied with a narrow band of frequencies about fr:f: , the only protons to
be resonantly excited will be those within a narrow slice of thickness dz at the
particular value of z corresponding to the narrow band of frequencies. The field
Brf is applied until the magnetization in the slice has been rotated through either
90 or 180 depending on what measurements are to be taken. Both Brf and Bg are
then turned off.
20
This discovery was first made by Paul Lauterbur and an analysis of the effect was first made by Sir Peter
Mansfield. They shared the 2003 Nobel Prize in Physiology and Medicine for their work is establishing MRI
as a medical diagnostic technique.
BIOMEDICAL APPLICATIONS 293
Figure 8.17 Two MRI scans of a brain (a) T1-weighted, and (b) T2-weighted showing a
frontal lobe tumour (images courtesy of the MRI Unit of the Royal Marsden NHS Foundation
Trust, London, UK)
294 CH8 APPLICATIONS OF NUCLEAR PHYSICS
Problems
8.1 The fission of 235 U is induced by a neutron and the fission fragments are 92 37 Rb and
140
55 Cs. Use the SEMF to calculate the energy released (in MeV) per fission. Ignore
the (negligible) contributions from the pairing term. The reaction is used to power a
100 MW nuclear reactor whose core is a sphere of radius 100 cm. If an average of
one neutron per fission escapes the core, what is the neutron flux at the outer surface
of the reactor in m2 s1 ? The core is surrounded by 1:3 m3 of ideal gas maintained
at a pressure of 1 105 Pa and a temperature of 298 K. All neutrons escaping the
reactor core pass through the gas. If the interaction cross-section between the
neutrons and the gas is 1 mb, calculate the rate of neutron interactions in the gas.
8.2 A neutron with non-relativistic laboratory speed v collides elastically with a nucleus
of mass M. If the scattering is isotropic, show that the average kinetic energy of the
neutron after the collision is
M 2 þ m2
Efinal ¼ Einitial ;
ðM þ mÞ2
8.3 A thermal fission reactor uses natural uranium. The energy released from fission is
200 MeV per atom of 235 U and the total power output is 500 MW. If all neutrons
captured by 238 U lead to the production of 239 Pu, calculate the rate of production of
plutonium in kg/year. The cross-sections at the relevant neutron energy are
capture ¼ 3 b and fission ¼ 600 b; and the relative abundance of 238 U to 235 U in
natural uranium is 138:1.
8.4 In a particular thermal reactor, each fission releases 200 MeV of energy with an
instantaneous power output 3 t1:2 , where t is measured in seconds. After burning
with a steady power output P0 ¼ 2 GW for a time T, the reactor is shut-down. Show
that the mean thermal power P from a fuel rod of the reactor after time t ð> 1 sÞ is
approximately
h i
PðtÞ ¼ 0:075P0 t0:2 ðT þ tÞ0:2
and, taking the mean age of the fuel rods to be 1 year, calculate the power output
after 6 months.
8.5 If the Sun were formed 4.6 billion years ago and initially consisted of 9 1056
hydrogen atoms and since then has been radiating energy via the PPI chain at a
detectable rate of 3:86 1026 W, how much longer will it be before the Sun’s supply
of hydrogen is exhausted (assuming that the nature of the Sun does not change)?
PROBLEMS 295
8.6 In the PPI cycle, helium nuclei are produced by the fusion of hydrogen nuclei and
6.55 MeV of electromagnetic energy is produced for every proton consumed. If the
electromagnetic radiation energy at the surface of the Earth is 8:4 J cm2 s1 and is
due predominantly to the PPI cycle, what is the expected flux of solar neutrinos at
the Earth in cm2 s1 ?
8.7 In a plasma of equal numbers of deuterium and tritium atoms (in practice, deuteron
and triton nuclei) at an energy kT ¼ 10 keV, the Lawson criterion is just satisfied for
a total of 5 s. Estimate the number density of deuterons.
8.8 A thermal power station operates using inertial confinement fusion. If the ‘fuel’
consists of 1 mg pellets of frozen deuterium–tritium mixture, how many would have
to be supplied per second to provide an output of 750 MW if the efficiency for
converting the material is 25 per cent?
8.9 In some extensions of the standard model (to be discussed in Chapter 9) the proton is
unstable and can decay, e.g. via p ! 0 þ eþ . If all the energy in such decays is
deposited in the body and assuming that an absorbed dose of 5 Gy per year is lethal
for humans, what limit does the existence of life place on the proton lifetime?
8.10 The main decay mode of 60 27 Co is the emission of two photons, one with energy
1.173 MeV and the other with 1.333 MeV. In an experiment, an operator stands 1 m
away from an open source of 40 KBq of 6027 Co for a total period of 18 h. Estimate the
approximate whole-body radiation dose received.
8.12 The flux of relativistic cosmic ray muons at the surface of the Earth is approximately
250 m2 s1 . Use Figure 4.8 to make a rough estimate of their rate of ionization
energy loss as they traverse living matter. Hence estimate in grays the annual human
body dose of radiation due to cosmic ray muons.
55
8.13 Calculate the nuclear magnetic resonance frequency for the nucleus 25 Mn in a field
of 2 T if its magnetic dipole moment is 3.46 N .
9
Outstanding Questions and Future
Prospects
In this chapter we shall describe a few of the outstanding questions in both nuclear
and particle physics and future prospects for their solution. The list is by no means
exhaustive (particularly for nuclear physics, which has a very wide range of
applications) and concentrates mainly on those areas touched on in earlier
chapters. The examples should be sufficient to show that nuclear and particle
physics remain exciting and vibrant subjects with many interesting phenomena
being discovered and questions awaiting explanations.
1
This process is called ‘spontaneous symmetry breaking’ and was mentioned in Chapter 6.
PARTICLE PHYSICS 299
eþ þ e ! H 0 þ Z 0 ; ð9:1Þ
Attempts were made to detect Higgs bosons by their decays to bb pairs, where
the quarks would be observed as jets containing short-lived hadrons with non-zero
beauty. The results were tantalizing. By the time LEP closed down in November
2000 to make way for another project, it had shown that no Higgs bosons existed
with a mass less than 113.5 GeV/c2; and some evidence had been obtained for the
existence of a Higgs boson with a mass of 115 GeV/c2. This is very close to the
upper limit of masses that were accessible by LEP, but because the Higgs boson
would have a width, its mass distribution would extend down to lower energies and
would give a signal. Unfortunately, while this signal was statistically likely to be a
genuine result rather than a statistical fluctuation, the latter cannot be completely
ruled out.
Future investigations will involve the use of new accelerators currently under
construction, particularly the LHC proton–proton collider mentioned in Chapter 4.
(One of the detectors at the LHC, ATLAS was shown in Figure 4.19.) This will
enable searches to be made for Higgs bosons with masses up to 1 TeV/c2 via
reactions of the type
p þ p ! H 0 þ X; ð9:2Þ
where X is any state allowed by the conservation laws. The mechanism for this
reaction is the weak interaction between the constituent quarks of the protons, an
example of which is shown in Figure 9.2, where the other quarks in the protons are
spectators, as usual.
300 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
Figure 9.2 An example of a process that can produce Higgs bosons in pp collisions
The reaction in Equation (9.2) will take place against a large background of
strong interaction processes and the method of detecting it will depend on the
actual Higgs boson mass. If MH > 2MW , then the Higgs boson can decay to a pair
of W-mesons or Z 0 -mesons, which themselves decay. For example, from the
leptonic decay of the Z 0 s, we could have overall the reaction
H 0 ! ‘þ þ ‘ þ ‘þ þ ‘ ; ð‘ ¼ e; Þ: ð9:3Þ
This would enable the mass range 200 GeV=c2 MH 500 GeV=c2 to be explored.
However, the branching ratios are such that only a few per cent of decays will have
such a distinctive signal and other decays modes will also have to be explored. For
lower masses such that MH < 2MW where these decays are energetically forbidden,
one might think of looking for decays to fermion–antifermion pairs. Because the
Higgs boson preferentially couples to heavy particles, the dominant decay of this
type will be H 0 ! b þ b with accompanying jets. This was the method used in the
LEP experiments referred to above. Unfortunately, it is very difficult to distinguish
these jets from those produced by other means. Rarer decay modes, but with more
distinctive signals, will have to be sought, such as H 0 ! þ , which in the
standard model has a branching ratio of about 103 .
All the above is based on the standard model with a single neutral Higgs boson,
but we will see in Section 9.1.3 that realistic extensions of the standard model
require several Higgs bosons, not all of which are electrically neutral. Experi-
mental investigations of the Higgs sector will undoubtedly play a central role in the
future of particle physics for many years to come.
Whether or not the Higgs boson exists is the most pressing unanswered question of
the standard model but, even if it is found with its predicted properties, this is not
PARTICLE PHYSICS 301
the end of the story, because one of the goals of particle theory is to have a single
universal theory that explains all the phenomena of the subject. Since we already
have a unified theory for the weak and electromagnetic interactions, the next
logical step is to try to include the strong interaction. Attempts to do this are called
grand unified theories (GUTs).
We have seen that unification of the weak and electromagnetic interactions does
not manifest itself until energies of the order of the W and Z masses. To get some
idea of the energy scale of a grand unified theory, we show in Figure 9.3 the
couplings2 pffiffiffi pffiffiffi
g 2 2gW ; g0 2 2gZ ð9:4Þ
Figure 9.3 Idealized behaviour of the strong and electroweak coupling as functions of the
squared energy--momentum transfer Q2 in a simple grand unified theory; gU is the unification
coupling
2
Recall that the electromagnetic coupling e is related to these couplings by the unification condition
Equation (6.71).
302 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
There are many potential grand unified theories, but the simplest incorporates
the known quarks and leptons into common families. For example, one way is to
put the three coloured d-quarks and the doublet ðeþ ; e Þ (strictly their right-handed
components) into a common family, i.e.
The fundamental vertex interactions allowed in this model are shown in Figure 9.4.
Figure 9.4 Fundamental vertices that can occur for the multiplet of particles in Equation (9.5)
In addition to the known QCD interaction in (a) and the electroweak interaction
in (b), there are two new interactions represented by (c) and (d) involving the
emission or absorption of two new gauge bosons X and Y with electric charges 43
and 13, respectively, and masses of the order of MX . In this theory all the
processes of Figure 9.4 are characterized by a single GUT coupling given by
g2U 1
U ; ð9:6Þ
4 42
which is found by extrapolating the known coupling of the standard model to the
energy MX c2.
This simple model has a number of attractive features. For example, it can be
shown that the sum of the electric charges of all the particles in a given multiplet is
zero. So, using the multiplet ðdr ; db ; dg ; eþ ; e Þ, it follows that
3qd þ e ¼ 0; ð9:7Þ
PARTICLE PHYSICS 303
where qd is the charge of the down quark. Thus qd ¼ e=3 and the fractional
charges of the quarks is seen to originate in the fact that they exist in three colour
states. By a similar argument, the up quark has charge qu ¼ 2e=3 and so with the
usual quark assignment p ¼ uud, the proton charge is given by
qp ¼ 2qu þ qd ¼ e: ð9:8Þ
Thus, we also have an explanation of the long-standing puzzle of why the proton
and positron have precisely the same electric charge.
GUTs make a number of predictions that can be tested at presently accessible
energies. For example, if the three curves of Figure 9.3 really did meet at a point, then
the three low-energy couplings of the standard model would be expressible in terms
of the two parameters U and MX . This could be used to predict one of the former, or
equivalently the weak mixing angle W . The result is sin2 W ¼ 0:214 0:004,
which is close to the measured value of 0:2313 0:0003, although not strictly
compatible with it. (This is true even if the effect of the Higgs boson is taken into
account when evaluating the evolution of the coupling constants.)
Figure 9.5 The three fundamental vertices predicted by the simplest GUT involving the gauge
bosons X and Y (these are in addition to those shown in Figure 9.4)
In addition to the interactions of the X and Y bosons shown in Figure 9.4, there
are a number of other possible vertices, which are shown in Figure 9.5. (There is
also another set where particles are changed to antiparticles.) A consequence of
these interactions and those of Figure 9.4(c) and (d) is the possibility of reactions
that conserve neither baryon nor lepton numbers. The most striking prediction of
this type is that the proton would be unstable, with decay modes such as
p !
0 þ eþ and p !
þ þ e . Examples of Feynman diagrams for these decays
are shown in Figure 9.6 and are constructed by combining the vertices of Figure 9.4
and 9.5. In all such processes, although lepton number L and baryon number B are
not conserved, the combination
X
R B L‘ ð‘ ¼ e; ;
Þ ð9:9Þ
‘
is conserved.
304 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
Figure 9.6 Examples of processes that contribute to the proton decay mode p !
0 þ eþ
Since the masses of the X and Y bosons are far larger than the quarks and
leptons, we can use the zero-range approximation to estimate the lifetime of proton
decay. In this approximation, and by analogy with the lifetime for the muon
Equation (7.62), we have for the proton lifetime
ðMX c2 Þ4
p : ð9:10Þ
g4U ðMp c2 Þ5
Proton decay via these modes has been looked for experimentally. The most
extensive search has been made using the Kamiokande detector described in
Chapter 4. To date no events have been observed and this enables a lower limit to
be put on the proton lifetime of about 1032 years, which rules out the simplest
version of a grand unified theory. However, there are other, more complicated,
versions that still cannot be completely ruled out by present experiments. Some of
these incorporate the idea of supersymmetry which is described below.
Finally, GUTs may offer an explanation for the very small neutrino masses
observed in the oscillation experiments discussed in Chapter 3. In Section 6.3 we
discussed the possibility that the neutrino was its own antiparticle (a so-called
Majorana neutrino). In GUTs the right-handed neutrino states are predicted to be
very massive (of order 1017 GeV/c2) and mix with the massless left-handed
neutrinos of the standard model to give physical neutrinos with masses
m
m2L =MX ; ð9:12Þ
where mL is the typical mass of a charged lepton or quark.
9.1.3 Supersymmetry
One of the problems with GUTs is that if there are new particles associated with
the unification energy scale, then they would have to be included as additional
PARTICLE PHYSICS 305
If the symmetry were exact then a particle and its superparticle would have equal
masses. This is clearly not the case or such states would have already been found. So
supersymmetry is at best an approximate symmetry of nature. Nevertheless, even in
an approximate symmetry, the couplings of the two states are equal and opposite,
thereby ensuring the required cancellation, providing their masses are not too large.
In practice, it is usually assumed in GUTs that incorporate supersymmetry that the
masses of the superparticles are of the same order as the masses of the W and Z
bosons. With the inclusion of superparticles, the evolution of the coupling constants
of the standard model as functions of Q2 changes slightly and when extrapolated
they meet very close to a single point. The unification mass is increased somewhat to
about 1016 GeV/c2, while the value of gU remains roughly constant. Thus the
predicted lifetime of the proton is increased to about 1032 – 1033 years, conveniently
beyond the ‘reach’ of current experiments. At the same time, the value of the weak
mixing angle is brought into almost exact agreement with the measured value.
Whether this is simply a coincidence or not is unclear.
306 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
~e ! e þ
e0 ; ð9:14Þ
giving overall
eþ þ e ! eþ þ e þ
e0 þ
e0 : ð9:15Þ
3
For a review of the current state of experimental searches for superparticles see, for example, Ei04.
PARTICLE PHYSICS 307
The problems here are formidable, not least of which is that the divergences
encountered in trying to quantize gravity are far more severe than those in either
QCD or the electroweak theory and there is at present no successful ‘stand-alone’
quantum theory of gravity analogous to the former two. The theories that have
been proposed that include gravity invariably replace the idea of point-
like elementary particles with tiny quantized strings as a device to reduce these
technical problems and are formulated in many more dimensions (usually 10 or
11) than we observe in nature. More recently, even strings have been superceded
by theories based on mathematical objects called membranes, or simply branes.
The problem with these theories, leaving aside their formidable mathematical
complexity, is that they apply at an energy scale where gravitational effects are
comparable to those of the gauge interactions, i.e. at energies defined by the so-
called Planck mass MP , which is given by
1=2
hc
MP ¼ ¼ 1:2 1019 GeV=c2 ; ð9:16Þ
G
Particle physics and astrophysics interact in an increasing number of areas and the
resulting field of particle astrophysics is a rapidly expanding one. The interactions
are particularly important in the field of cosmology where, for example, the
detection of neutrinos can provide unique cosmological information. Another
reason is because the conditions in the early Universe implied by standard
cosmological theories (the big bang model) can only be approached, however
remotely, in high-energy particle collisions. At the same time, these conditions
occurred at energies that are relevant to the grand unified and SUSY theories of
particle physics and so offer a possibility of testing the predictions of such
theories. This is important because, as mentioned above, it is difficult to see
other ways of testing such predictions. For reasons of space, we will discuss just
three examples of particle astrophysics. We will return to the question of
conditions in the early universe in Section 9.2.2.
4
This implies that strings have dimensions of order ‘P
h=MP c ¼ 1:6 1035 m.
308 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
Neutrino astrophysics
We have seen in Chapter 3 that cosmic rays and emissions from the Sun are
important sources of information about neutrinos and have led us to revise the view
that neutrinos are strictly massless, as is assumed in the standard model. At the
same time, there is considerable interest in studying ultra high-energy neutrinos as
a potential source of information about galactic and extra-galactic objects and
hence cosmology in general.
One of the first neutrino astrophysics experiments was the observation of
neutrinos from a supernova. Supernovas are very rare events where a star literally
explodes with a massive output of energy over a very short timescale measured in
seconds. The mechanism for this (briefly) is as follows. If a star has a mass greater
than about 11 solar masses, it can evolve through all stages of fusion, ending in a
core of iron surrounded by shells of lighter elements. Because energy cannot be
released by the thermonuclear fusion of iron, the core will start to contract under
gravity. Initially this is resisted by the pressure of the dense gas of degenerate
electrons in the core (electron degeneracy pressure), but as more of the outer core is
burned and more iron deposited in the core, the resulting rise in temperature makes
the electrons become increasingly relativistic. When the core mass reaches about
1.4 solar masses (the so-called Chandrasekhar limit), the electrons become ultra
relativistic and they can no longer support the core. At this point the star is on the
brink of a catastrophic collapse.
The physical reactions that lead to this are as follows. Firstly, photodisintegra-
tion of iron (and other nuclei) takes place,
which further heats the core and enables the photodisintegration of the helium
produced, i.e.
þ 4 He ! 2p þ 2n: ð9:18Þ
As the core continues to collapse, the energy of the electrons present increases to a
point where the weak interaction
e þ p ! n þ e ð9:19Þ
becomes possible and eventually the hadronic matter of the star is predominantly
neutrons. This stage is therefore called a neutron star. The collapse ceases when
the gravitational pressure is balanced by the neutron degeneracy pressure. At this
point the radius of the star is typically just a few kilometres. The termination of the
collapse is very sudden and as a result the core material produces a shock wave
that travels outwards through the collapsing outer material, leading to a supernova
(actually a so-called Type II supernova). Initially there is an intense burst of e
PARTICLE PHYSICS 309
with energies of a few MeV from the reaction of Equation (9.19). This lasts for a
few milliseconds because the core rapidly becomes opaque even to neutrinos and
after this the core material enters a phase where all its constituents (nucleons,
electrons, positrons and neutrinos) are in thermal equilibrium. In particular, all
flavours of neutrino are present via the reactions
Ð eþ e Ð ‘ ‘ ; ð‘ ¼ e; ; Þ ð9:20Þ
and these will eventually diffuse out of the collapsed core and escape. Neutrinos
of all flavours, with average energies of about 15 MeV, will be emitted in all
directions over a period of 0.1–10 s. Taken together, the neutrinos account for
about 99 per cent of the total energy released in a supernova. Despite this, the
output in the optical region is sufficient to produce a spectacular visual effect.
The first experiments to detect neutrinos from a supernova were an earlier
version of the Kamiokande experiment described in Chapter 3 and the IMB
collaboration, which also used a water C̆erenkov detector. Both had been
constructed to search for proton decay as predicted by GUTs, but by good fortune
both detectors were ‘live’ in 1987 at the time of a spectacular supernova explosion
(now named SN1987A) and both detected a small number of antineutrino events.
The data are shown in Figure 9.7. The Kamiokande experiment detected 12 e
events and the IMB experiment eight events, both over a time interval of
approximately 10 s and with energies in the range 0–40 MeV. These values are
consistent with the estimates for the neutrinos that would have been produce by the
reaction in Equation (9.20) and then diffused from the supernova after the initial
pulse.
Figure 9.7 Data for neutrinos from SN1987A detected in the Kamiokande and IMB experiments:
the threshold for detecting neutrinos in the experiments are 6 MeV (Kamiokande) and 20 MeV
(IMB) -- in each case the first neutrino detected is assigned the time zero
310 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
The data can be used to make an estimate of the neutrino mass as follows. The
time of arrival on Earth of a neutrino i is given by
L m2 c4
ti ¼ t0 þ 1þ ; ð9:21Þ
c 2Ei2
where t0 is the time of emission from the supernova and ðm; Ei Þ are the mass and
total energy of the neutrino. Thus
" #
L m2 c4 1 1
ðtÞij ti tj ¼ : ð9:22Þ
2c Ei2 Ej2
Using data for pairs of neutrinos, Equation (9.22) leads to the result
me 20 eV; ð9:23Þ
which, although larger than the value from tritium decay, is still a remarkable
measurement.
The neutrinos from SN1987A were of low energy, but there is also a great
interest in detecting ultra high-energy neutrinos. For example, it is known that
there exist point sources of -rays with energies in the TeV range, many of which
have their origin within so-called active galactic nuclei. It is an open question
whether this implies the existence of point sources of neutrinos with similar
energies. The neutrinos to be detected would be those travelling upwards through
the Earth, as the signal from downward travelling particles would be swamped
by neutrinos produced via pion decay in the atmosphere above the detector. Like
all weak interactions the intrinsic rate would be very low, especially so for such
high-energy events, but this is partially compensated by the fact that the –nucleon
cross-section increases with energy, as we showed in Chapter 6.
To detect neutrinos in the TeV energy range using the C̆erenkov effect in water
requires huge volumes, orders-of-magnitude larger than used in the Super-
Kamiokande detector. An ingenious solution to this problem is to use the vast
quantities of water available in liquid form in the oceans, or frozen in the form of
ice at the South Pole, and several experiments have been built, or are being built,
using these sources. The largest so far is the Antartic Muon and Neutrino Detector
Array (AMANDA) which is sited at the geographical South Pole. A schematic
diagram of this detector is shown in Figure 9.8.
The detector consists of strings of optical modules containing photomultiplier
tubes that convert the C̆erenkov radiation to electrical signals. The enlarged inset
in Figure 9.8 shows the details of an optical module. They are located in the ice at
great depths by using a novel hot-water boring device. The ice then refreezes
around them. In the first phase of the experiment in 1993/94 (AMANDA-A) four
detector strings were located at depths of between 800 and 1000 m. The ice at
PARTICLE PHYSICS 311
these depths is filled with air bubbles and so the detectors are not capable of
precision measurements, but they proved the validity of the technique. In the next
phase a few years later (AMANDA-B10), 10 more strings containing 320 optical
modules were located at depths between 1.5 and 2.0 km, where the properties of
ice are suitable for muon detection. Finally, the current version of the detector
(AMANDA-II) has an additional nine strings extending to a depth of 2.35 km. In
total there are 680 optical modules covering a cylindrical volume with a cross-
sectional diameter of 120 m.
The AMANDA detector has successfully detected atmospheric neutrinos and
has produced the most detailed map of the high-energy neutrino sky to date.
However, no source of continuous emission has yet been observed that would be a
candidate for a point source.
312 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
AMANDA can detect neutrinos with energies up to about 1015 eV, but an even
bigger detector, called IceCube, is under construction at the South Pole. This uses 80
strings each containing 60 optical modules regularly spaced over an area of 1 km2 at
depths of between 1.4 and 2.4 km (the volume covered by AMANDA is only
1.5 per cent of the volume to be covered by IceCube) and will be capable of detecting
neutrinos with energies as high as 1018 eV. IceCube is due for completion in 2010.
Dark matter
3H02
c ¼
1026 kg m3 5:1 ðGeV=c2 Þm3 ; ð9:24Þ
8
G
where G is the gravitational constant and we have used the best current value for
Hubble’s constant H0 to evaluate Equation (9.24). In the most popular version of
the model, called the inflationary big bang model, the relative density
=c ¼ 1: ð9:25Þ
¼ r þ m þ ; ð9:26Þ
5
For an accessible discussion of the big bang model and other matters discussed in this section see, for
example, Pe03.
PARTICLE PHYSICS 313
expectations range from 1–10 events per kg of detector per week. This is very
small compared with the event rate from naturally occurring radioactivity,
including that in the materials of the detectors themselves. The former is
minimized by working deep underground to shield the detector from cosmic
rays and in areas with geological structures where radioactive rocks are absent; and
the latter is minimized by building detectors of extreme purity. Finally, WIMP
recoils should exhibit a small seasonal time variation due to the motion of the
Earth around the Sun and the motion of the Sun within the galaxy. One experiment
claims to have seen this variation. Present experiments are at an early stage, but
some versions of SUSY theories with low-mass neutralinos can probably already
be ruled out.6
Matter–antimatter asymmetry
One of the most striking facts about the universe is the paucity of antimatter
compared with matter. There is ample evidence for this. For example, cosmic
rays are overwhelmingly composed of matter and what little antimatter is present
is compatible with its production in intergalactic collisions of matter with photons.
Neither do we see intense outbursts of electromagnetic radiation that would
accompany the annihilation of clouds of matter with similar clouds of antimatter.
The absence of antimatter is completely unexpected because, in the original
big bang, it would be natural to assume a total baryon number B ¼ 0.7 Then
during the period when kT was large compared with hadron energies, baryons
and antibaryons would be in equilibrium with photons via reversible reactions
such as
p þ p Ð þ ð9:27Þ
and this situation would continue until the temperature fell to a point where the
photons no longer had sufficient energy to produce pp pairs and the expansion had
proceeded to a point where the density of protons and antiprotons was such that
their mutual annihilation became increasingly unlikely. The critical temperature is
kT 20 MeV and at this point the ratios of baryons and antibaryons to photons
‘freezes’ to values that can be calculated to be
6
An up-to-date review of the status of dark matter searches is given in Pe03 and Ei04.
7
One could of course simply bypass the problem by arbitrarily assigning an initial non-zero baryon number
to the universe, but it would have to be exceedingly large to accommodate the observed asymmetry, as well
as being an unaesthetic solution.
NUCLEAR PHYSICS 315
with NB =NB
104 . The simple big bang model fails spectacularly.
The conditions whereby a baryon–antibaryon asymmetry could arise were first
stated by Sakharov. It is necessary to have: (a) an interaction that violates baryon
number conservation, (b) an interaction that violates charge conjugation, and (c) a
non-equilibrium situation must exist at some point to ‘seed’ the process. We have
seen in Chapter 6 that there is evidence that CP is violated in the decays of
some neutral mesons, but its source and size are not compatible with that required
for the observed baryon–antibaryon asymmetry and we must conclude that there
is another, as yet unknown, source of CP violation. Likewise a method for
generating a non-equilibrium situation is also unknown, although it may be that
the baryon-violating interactions of GUTs, which are necessary for condition (a),
may provide one. Clearly, matter–antimatter asymmetry remains a serious
unsolved problem.
In the standard model, the structure of nucleons is specified in terms of quarks and
gluons, but questions remain. One concerns the spin of the proton. This must be
formed from combining the spins and the relative orbital angular momenta of its
constituent quarks and gluons. Measuring these various contributions can be done
in deep inelastic scattering experiments of the type described in Chapter 5, but
using spin-polarized targets, sometimes with spin-polarized beams. Experiments to
date have shown the surprising result that the spins of all the quarks and antiquarks
together contribute only about 20–30 per cent to the total spin of the proton (the
so-called ‘proton spin crisis’). There is some information that the angular
momentum contributions of the quarks play an important role, but very little is
known about the contribution of the total angular momentum of the gluon. This is
an area where the type of experiment that can be pursued at the CEBAF accelerator
described in Section 4.2.2 will be vital in unravelling the details of each
contribution and thus further testing QCD.
Nucleons and mesons are the building blocks of nuclear matter, but there is no
guarantee that the properties of these particles in nuclei are identical to those
exhibited as free particles. According to QCD the properties of hadrons are
strongly influenced by the sea of quark–antiquark pairs and gluons that we have
seen in Chapter 5 are always present around confined quarks due to quantum
fluctuations. However, these influences could well be different in the case of
closely spaced nucleons in nuclear matter from those for a free nucleon. Indeed
there are theoretical predictions that the probability of finding a qq pair decreases
as the density of the surrounding nuclear matter increases. If such effects could be
established they would have a profound influence on our understanding of quark
confinement.
8
A comprehensive overview of the field as at 1999 is a report of the Board on Physics and Astronomy of the
National Research Council, USA: ‘‘Nuclear Physics: The Core of Matter, The Fuel of Stars’’, National
Academy Press, Washington, D.C. (1999) – NRC99. Other useful sources are the publications of the Nuclear
Physics European Collaboration Committee (NuPecc) and in particular its ‘‘Report on Impact, Applications,
Interactions of Nuclear Science’’ (2002) and the NuPecc Long-Range Plan 2004.
NUCLEAR PHYSICS 317
Figure 9.9 The ratios of the F2 structure function found from nuclear targets to that found
from deuterium, as a function of the scaling variable x (Carbon data from Ar95, calcium data from
Am95)
Figure 9.10 An energetic particle (typically several tens of MeV/u to GeV/u) is fragmented in
a nuclear reaction in a thin target, and radioactive reaction products are separated in-flight and
transported as a secondary beam to the experiment
The heaviest element made to date has Z ¼ 116 and was produced by fusion in the
reaction 48 248 292
20 Ca þ 96 Cm ! 116 Uuh þ 4n (the symbol Uuh is used as the element has
yet to be named). Strenuous efforts are being made to reach the predicted new
island of relative stability. This will require facilities to produce exotic short-lived
nuclear beams and there is much development work going on in this area. One
example of how such a beam can be formed is shown in Figure 9.10. The other
main method employs two independent accelerators: a high-power driver accel-
erator for production of the short-lived nuclei in a thick target that is directly
connected to an ion-source, and a second post-accelerator. Radioactive atoms
diffuse out of a hot target into an ion source where they are ionized for acceleration
in the post-accelerator.
Fewer than 300 stable nuclei occur naturally (see Figure 2.7) and outside the
stability region nuclei decay by the mechanisms discussed in Chapters 2 and 7. In
the uncharted regions there are many fundamental questions to be answered,
such as what are the limiting conditions under which nuclei can remain bound
and do new structures emerge near these limits? The answers to these questions
are important because theoretical descriptions of nuclei far from the line of
stability suggest that their structures are different from what has been seen in
stable nuclei. Nuclei far from stability also play an important role in astro-
physics, for example in understanding the processes in supernovae and how
elements are synthesized in stars. Another limiting region that is expected to
yield interesting information is that of angular momentum. Super-deformed
nuclei have been discovered with highly elongated shapes and very rapid
rotational motion. The states associated with these shapes are extremely stable.
Further investigation of these is expected to yield important information about
nuclear structure.
320 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
Figure 9.11 Stages in the formation of a quark--gluon plasma and subsequent hadron
emission: two heavy nuclei collide at high energies (a) and interact via the colour field (b); the
very high energy-density produced causes the quarks and gluons to deconfine and form a plasma
that can radiate photons and lepton pairs (c); finally, as the plasma cools, hadrons condense and
are emitted (d) (after NRC99, with permission of the National Academics Press)
NUCLEAR PHYSICS 321
Figure 9.12 View of a 200 GeV gold--gold interaction in the STAR detector at the RHIC
accelerator (Courtesy of Brookhaven National Laboratory)
those obtained from particle physics experiments. Other experiments plan to study
this effect in atomic francium, where the parity-mixing effect should be about 18
times larger. (The effect of an electric dipole moment of the electron is also
expected to be greatly enhanced in francium.) Unfortunately, francium is an
extremely rare element with no stable isotopes and so experiments will be carried
out with a small number of radioactive atoms collected in a magneto-optic trap.
In Section 8.3.1, we reviewed the use of radiation techniques for cancer therapy.
We also briefly mentioned that in principle heavier particles had advantages over
photons. For example, because of the form of the Bragg curve, protons deposit
more of their energy where they stop, not where they enter the body. Also their
depth of penetration can be precisely controlled so that they stop within the
tumour, thus allowing radiologists to increase the radiation dose to the tumour
while reducing the dose to healthy tissues.
This is illustrated in Figure 9.13, which compares the treatment plans (i.e.
simulations of the pattern of radiation that the patient would receive) for treating a
case of advanced pancreatic cancer. Figure. 9.13(a) shows an X-ray plan using
a ‘state-of-the-art’ nine-beam X-ray system. The amount of radiation received by
nearby organs and other critical areas (kidneys, liver and spinal chord) is seen to be
a substantial fraction of the dose received by the region of the cancer. This is
contrasted with the results of Figure 9.13(b), which is for treatment using a
single proton beam. Although there is some unwanted exposure at the input site
Figure 9.13 Treatment plans for a large pancreatic tumour: (a) using a nine-beam X-ray
system; (b) using a single proton beam. The diffuse grey areas in (a) indicate the spread of
energy deposition outside the region of the tumour (adapted from Zu00, copyright Elsevier, with
permission)
NUCLEAR PHYSICS 325
Nuclear fusion still holds the promise of unlimited power without the problem of
radioactive waste, but the road to realization of this goal is long and we are far
from the end. In Section 8.2 we introduced the Lawson criterion as a measure of
how close a design was to the ignition point, i.e. the point at which a fusion
reaction becomes self-sustaining. To date no device has yet succeeded in achieving
the Lawson criterion and much work remains to be done on this important
problem. In recognition of this, at least one major new tokamak machine (to be
built in France) is planned as a global collaboration, but even when the ignition
point is attained, based on experience with fission reactors, it could be many
decades before that achievement is translated into a practical power plant.
In the shorter term and assuming that renewable sources of energy are insufficient
to fulfil the world’s increasing energy needs, it does seem as if power plants based on
fission reactions are the only hope of replacing fossil fuels in the future. The
problems of reactor safety and the safe disposal of radioactive waste are therefore
paramount.
The waste from light-water reactors, the most common type of power reactor,
has two major components: the actinides, i.e. any of the series of radioactive
elements with atomic numbers between 89 and 103 (mainly uranium but also
smaller amounts of heavier elements, the transuranic elements like plutonium and
the minor actinides such as neptunium, americium and curium) and fission
products, which are medium-weight elements from fission processes in the nuclear
fuel. While it is generally agreed that radioactive nuclei with relatively short
lifetimes can be safely stored in deep geological disposal facilities, the same is not
true of waste with very long lifetimes, some of which are water-soluble and so
have the potential to contaminate ground water. An additional problem is the
disposal of material that could be used for nuclear weapons, i.e. 239 Pu and 235 U.
One option for handling waste with very long lifetimes, which was mentioned as a
theoretical possibility in Section 8.1.3, is to transmute it by neutron reactions into
NUCLEAR PHYSICS 327
9
The same man who shared the 1984 Nobel Prize in Physics for the discovery of the W and Z bosons.
328 CH9 OUTSTANDING QUESTIONS AND FUTURE PROSPECTS
The possible energy flow in a commercial system is shown in Figure 9.15. This
assumes a 1 GeV, 20 ma proton beam requiring about 20 MW of input power. The
latter is taken from the output of the reactor leaving a net electrical output of
580 MW, i.e. a gain factor of about 30.
The current situation on the energy amplifier is that a European collaboration
has shown that initial partitioning at the level of 95–99 per cent is possible
Figure 9.15 Possible energy flows in an energy amplifier system; the conversion efficiencies
are denoted by h
NUCLEAR PHYSICS 329
depending on the actinide species. They have also carried out a number of
successful reactor transmutation and spallation studies and the first full ADS
experiment (TRADE) is currently under construction. This consists of coupling a
cyclotron delivering a 140 MeV, 0.5–1.0 ma proton beam to an existing 1MW
water-cooled reactor sited in Italy and uses a spallation target of tantalum. The
operation date is planned for 2007/08. Additional work is being carried out in
Belgium on coupling a 350 MeV, 5 ma proton beam to a 100 MW subcritical
reactor (the Myrrha experiment) and has already shown that some long-lived
isotopes can be successfully incinerated. Although ADS has enormous potential,
there are still a great many problems to be overcome and questions to be answered.
The estimated time for completion of research and development work and
commencement of an industrial plant based on ADS could be as long as 50 years.
Appendix A
Some Results in Quantum Mechanics
and also a wave reflected at the barrier travelling from right to left of the form
eikx . Thus the total wavefunction in region I is
where A and B are complex constants. Within the barrier, region II ð0 < x < aÞ,
the solution of the Schrödinger equation is a decaying exponential plus an
exponential wave reflected from the boundary at x ¼ a, i.e. the total wavefunction
is
x
2 ðxÞ ¼ Ce þ Dex ; ðA:3Þ
Finally, in region III ðx > aÞ to the right of the barrier, there is only an outgoing
wave of the form
ikx
3 ðxÞ ¼ Fe ; ðA:5Þ
Figure A.1 Rectangular barrier with (a) wavefunction solutions, and (b) form of the incoming
and outgoing waves; (c) modelling an arbitrary smooth barrier as a series of rectangular barriers
T jF=Aj2 : ðA:6Þ
The values of F and A are found by imposing continuity of the wavefunction and
its first derivative, i.e. matching the values of these quantities at the two boundaries
x ¼ 0 and x ¼ a. The algebra may be found in any introductory book on quantum
mechanics.1 The result is
2
2keika
T ¼ : ðA:7Þ
2k coshðaÞ iðk Þ sinhðaÞ
2 2
The corresponding incident and transmitted waves are shown in Figure A.1(b)
(the reflected waves are not shown).
For large a, which corresponds to small penetrations, we can make the
replacement
1
sinhðaÞ coshðaÞ ea ðA:8Þ
2
and hence
2
4k
T e2a : ðA:9Þ
k þ 2
2
1
See, for example, Chapter 6 of Me61.
DENSITY OF STATES 333
The first factor is due to the reflection losses at the two boundaries x ¼ 0 and x ¼ a
and the decreasing exponential describes the amplitude decay within the barrier.
The first factor is slowly varying with energy and is usually neglected.
The result of Equation (A.9), ignoring the first factor, may be used to find the
transmission coefficient for an arbitrary smoothly-varying barrier by modelling it
as a series of thinPrectangular barriers. This is illustrated in Figure A.1(c). Thus by
replace 2a by 2 ðxÞx and taking the limit of small x, the summation goes
over to an integral, i.e.
ð 12
2m
2a ! 2 dx 2 ½VðxÞ E ðA:10Þ
h
and
" ð 12 #
2m
T exp 2 dx 2 ½VðxÞ E : ðA:11Þ
h
where C is a constant and the components of the wave number k ¼ ðkx ; ky ; kz Þ take
the values
nx ny nz
kx ¼ ; ky ¼ ; kz ¼ ; ðnx ; ny ; nz Þ ¼ 1; 2; 3:::: ðA:13Þ
2 2 2
The energy of the particle is given by
where k jkj ¼ p=h and p is the particle’s momentum. Negative values of the
integers do not lead to new states since they merely change the sign of the wave
function Equation (A.12) and phase factors have no physical significance.
The allowed values of k form a cubic lattice in the quadrant of ‘k-space’ where
all the values of ðnx ; ny ; nz Þ are positive. Since each state corresponds to
one combination of ðnx ; ny ; nz Þ, the number of allowed states is equal to the
number of lattice points. The spacing between the lattice points is ðL= Þ, so the
density of points per unit volume in k-space is ðL= Þ3 . The number of lattice
points nðk0 Þ with k less than some fixed value k0 is the number contained within a
volume that for large values of k0 may be well approximated by the quadrant of a
sphere of radius k0 , i.e.
14 3 L 3 V 4 k03
nðk0 Þ ¼ k0 ¼ : ðA:15Þ
83 ð2 Þ3 3
Hence the number of points with k in the range k0 < k < ðk0 þ dk0 Þ is
V
dnðk0 Þ ¼ 3
4 k02 dk0 : ðA:16Þ
ð2 Þ
Thus
ðk0 Þdk0 is the number of states with k between k0 and k0 þ dk0 , or
equivalently
4 V 2
ðpÞdp ¼ p dp ðA:18Þ
ð2 hÞ3
is the number of states with momentum between p and p þ dp. This is the form
used in Equation (7.1) when discussing the Fermi energy in the Fermi gas model.
Equation (A.18) can also be written in terms of energy using E ¼ p2 =2m, when it
becomes
4 V
ðEÞdE ¼ m p dE ðA:19Þ
hÞ3
ð2
and this was the form used in discussing -decay in Section 7.7.2.
Although the above derivation is for a particle confined in a box, the same
technique can be used for scattering problems. In this case we can consider a large
volume V ¼ L3 and impose ‘periodic’ boundary conditions
where
The density of lattice points in k-space now becomes ðL=2 Þ3 , but unlike the
standing wave case, permutations of signs in Equation (A.22) do produce new
states and the whole quadrant of lattice points has to be considered. Thus these two
effects ‘cancel out’ and we arrive at the same result for the density of states in
Equations (A.18) and (A.19). This approach was used in discussing the formal
definitions of cross sections in Chapter 1.
All the above is for spinless particles. If the particle has spin then the density of
states must be multiplied by the appropriate spin multiplicity factor, taking account
of the Pauli principle as necessary. Thus, for example, for spin-12 particles, with two
spin states, Equation (A.19) becomes
8 V
ðEÞdE ¼ mp dE: ðA:23Þ
ð2 hÞ3
2
We follow the derivation given in Chapter 9 of Ma92.
336 APPENDIX A: SOME RESULTS IN QUANTUM MECHANICS
dcf ðtÞ X
ih ¼ Vfn ðtÞei!fn t cn ðtÞ; ðA:26Þ
dt n
where the matrix element Vfn uf jVðtÞjun i and the angular frequency
!fn ðEf En Þ=h. If we assume initially ðt ¼ 0Þ that the system is in a state
jui i, then cn ð0Þ ¼
ni and the solutions for cf ðtÞ are found by substituting this result
into the right-hand side of Equation (A.22) giving, to first-order in V,
ðt
1
ci ðtÞ ¼ 1 þ Vii ðt0 Þdt0 ðA:27aÞ
ih
0
and
ðt
1 0
cf ðtÞ ¼ Vfi ðt0 Þei!fi ðt Þ dt0 ðf 6¼ iÞ: ðA:27bÞ
ih
0
2
For f 6¼ i, the quantity cf ðtÞ is the probability, in first-order perturbation theory,
that the system has made a transition from state i to state f.
The above is for a general time-dependent perturbation VðtÞ, but the results can
also be used to describe other situations, for example where the perturbation is
zero up to some time t0 and a constant thereafter. In this case, the integrals in
Equations (A.27) can be evaluated and, in particular, Equation (A.27b) gives, again
to first-order in V,
Vfi
cf ðtÞ ¼ 1 ei!fi t ðA:28Þ
h!fi
" #
2 4Vfi 2 sin2 ð12 !fi tÞ
Pfi ðtÞ ¼ cf ðtÞ ¼ : ðA:29Þ
h2 !2fi
The function in the square brackets in Equation (A.29) is shown in Figure A.2.
PERTURBATION THEORY AND THE SECOND GOLDEN RULE 337
" #
sin2 ð12 !fi tÞ
Figure A.2 The function
!2fi
For sufficiently large values of t, it has the form of a large central peak with
much smaller side oscillations. In this case Pfi is only appreciable if
h!fi ¼ Ef Ei 2 h=t ðA:30Þ
and then the square bracket can be replaced by a Dirac delta function3, i.e.
2 2
Pfi ðtÞ ¼ t Vfi
ðEf Ei Þ ðA:32Þ
h
3
The Dirac delta function was theÐ first so-called ‘generalized function’. It is defined by the two conditions:
þ1
(i)
ðx0 xÞ ¼ 0 if x 6¼ x0 and 0
Ð x2 (ii) 0 1 0
ðx xÞdx
0
¼ 1: It follows that if f ðxÞ is a function continuous in the
interval x1 < x < x2 , then x1 f ðx Þ
ðx xÞdx0 ¼ f ðxÞ if x1 < x < x2 or ¼ 0 if x < x1 or x > x2.
338 APPENDIX A: SOME RESULTS IN QUANTUM MECHANICS
The above assumes that the final state is discrete, but it is more common for the
final states to form a continuum defined by the density of states
ðEÞ derived in
Section A.2 above. In this case, since
ðEÞdE is the number of states with energy
between E and E þ dE, we can write the transition rate per unit time dTfi =dt to a
group of states f with energies in this range as
ð
dTfi dPfi ðtÞ 2 h 2 i
¼
ðEf ÞdEf ¼ Vfi
ðEf Þ ; ðA:34Þ
dt dt h Ef ¼Ei
where the integral has been evaluated using the properties of the delta function.
Equation (A.34) is called the Second Golden Rule (sometimes Fermi’s Second
Golden Rule, although strictly the result is not due to Fermi) and has been used
in several places in this book, for example in Chapter 7 when discussing nuclear
-decay.
Appendix B
Relativistic Kinematics
In particle physics, most scattering interactions take place between particles whose
speeds are comparable with the speed of light c. This is often true even in decays,
particularly if light particles are emitted. The requirements of special relativity
therefore cannot be ignored. In nuclear physics accurate predictions can also often
only be obtained if relativistic effects are taken into account. In this appendix we
review (usually without proof) some relativistic kinematical results and the use of
invariants to simplify calculations.
x0 ¼ x
y0 ¼ y
ðB:1Þ
z0 ¼ ðvÞðz vtÞ
t0 ¼ ðvÞðt vz=c2 Þ
1
where ðvÞ ¼ ð1 2 Þ2 is the Lorentz factor and v=c. From the definition of
velocity and using these transformations, the particle’s speed in S0 is related to its
speed in S by
uv
u0 ¼ ðB:2Þ
1 uv=c2
and hence
1
ðu0 Þ ½1 ðu0 =cÞ2 2 ¼ ðuÞðvÞð1 uv=c2 Þ: ðB:3Þ
a00 ¼ ða0 va3 =cÞ; a01 ¼ a1 ; a02 ¼ a2 ; a03 ¼ ða3 va0 =cÞ: ðB:4Þ
For example, the space-time four-vector is x ¼ ðct; xÞ and when used in Equations
(B.4) reproduces Equations (B.1). The scalar product of two four-vectors a and b is
defined as
ab a0 b0 a b ðB:5Þ
P mu; ðB:6Þ
where m is the rest mass and u is the four-velocity, whose components are
u ¼ ðvÞðc, vÞ, where v is the three-velocity and v jvj. In terms of the total
energy E (i.e. including the rest mass) and the three-momentum p,
and for P1 ¼ P2 ¼ P,
P2 ¼ E2 =c2 p2 : ðB:9Þ
E2 ¼ p2 c2 þ m2 c4 : ðB:10Þ
It follows that
The Lorentz transformations for energy and momentum follow from these
definitions and Equations (B.4). Thus, in S0 we have
and
and
where
X
N X
N
E¼ Ei and p¼ pi : ðB:13cÞ
i¼1 i¼1
In the general case where the relative velocity v of the two frames is in an arbitrary
direction, the transformations in Equations (B.12) become
0 vp 1
p ¼ p þ v E 2; E0 ¼ ðE v pÞ: ðB:14Þ
þ1 c
In the CMS, the three-momenta of the two particles a and b are equal and opposite,
so that the total momentum is zero,1 i.e.
1
Although ‘centre-of-mass’ system is the most frequently used name, some authors refer to this as the
‘centre-of-momentum’ system. Logically, a better name would be ‘zero-momentum’ frame.
342 APPENDIX B: RELATIVISTIC KINEMATICS
with
pa þ pb ¼ 0: ðB:16bÞ
In a colliding beam accelerator, these two views become mixed. The colliding
particles are both moving, but only if they have equal momenta and collide at zero
crossing angle is the system identical to the centre-of-mass system.
The four-vectors of the initial-state particles in the two systems may be written
(L ¼ laboratory, T ¼ target)
where
c2 pL EL þ mT c2 pL
v¼ ; ¼ pffiffi ; v ¼ pffiffi ðB:19Þ
EL þ mT c2 c2 s s
This result was used in Chapter 4 when discussing the relative merits of fixed-
target and colliding beam accelerators.
Substituting Equations (B.19) into Equations (B.18) gives
pL mT m2B c2 þ mT EL
p ¼ pffiffi ; Ea ¼ pffiffi ðB:22aÞ
s s
pL mT m2T c2 þ mT EL
p ¼ pffiffi ; Eb ¼ pffiffi : ðB:22bÞ
s s
FRAMES OF REFERENCE 343
Finally we state, without proof, the transformation of scattering angles for the
specific case of laboratory and centre-of-mass systems. Consider the general
scattering reaction
where L is the scattering angle in the laboratory system, i.e. the angle between the
beam direction and q. In the CMS,
where p0B and p0T are the CMS momenta of the beam and target, respectively. The
relation between the scattering angle C in this system and L is
1 q0 sin C
tan L ¼ ; ðB:26Þ
ðvÞ q0 cos C þ vE0 =c2
where
v ¼ pL c2 ½EL þ mT c2 1 ; ðB:28Þ
Substituting Equations (B.27), (B.28) and (B.29) into Equation (B.26) gives
2mT c 1=2 u sin C
tan L : ðB:30Þ
pL u cos C þ c
Thus, unless u c and cos C 1, the final-state particles will lie in a narrow
cone about the beam direction in the laboratory system. Similarly, when a
344 APPENDIX B: RELATIVISTIC KINEMATICS
high-energy particle decays, its decay products will emerge predominantly at small
angles to the initial beam direction.
B.3 Invariants
The transformations between laboratory and centre-of-mass systems for energy
and momentum have been worked out explicitly above, but a more efficient way is
to work with quantities that are invariants, i.e. have the same values in all inertial
frames. We have already met one of these: s the invariant mass squared, defined in
Equation (B.20). We will now find expressions for the energy and momentum in
terms of invariants for both the LS and the CMS.
First, in the LS, from Equations (B.15), we have
pB ¼ 0; EB ¼ mB c: ðB:31Þ
i.e.
and so
This function is invariant under all permutations of its arguments and in particular
Equation (B.35a) can be written in the form
h ih i12
c
pL ¼ s ðmT þ mB Þ2 s ðmT mB Þ2 : ðB:36Þ
2mT
PROBLEMS 345
The above formulae have many applications. For example, if we wish to produce
particles with a certain mass M, the minimum laboratory energy of the beam
particles is, from Equation (B.33),
M 2 c2 m2B c2 m2T c2
EL ðminÞ ¼ : ðB:39Þ
2mT
Hence the mass of the decaying particle is equal to the invariant mass of its decay
products. The latter can be measured if the particle is too short-lived for its mass to
be measured directly.
Problems
B.1 The Mandelstam variables s, t and u are defined for the reaction A þ B ! C þ D by
(b) In the case of elastic scattering show that t ¼ 2p2 ð1 cos Þ=c2 , where
p jpj, p is the centre-of-mass momentum of particle A and is its scattering
angle in the CMS.
B.2 A pion travelling with speed v jvj in the laboratory decays via ! þ
. If the
neutrino emerges at right angles to v, find an expression for the angle at which the
muon emerges.
px py pz
þ
A 0:488 0:018 2.109
B 0.255 0.050 0.486
B.6 Calculate the minimum laboratory energy Emin of the initial proton for the
production of antiprotons in a fixed-target experiment using the reaction
pp ! ppp p. If the protons are bound in nuclei, show that taking the internal motion
of the nucleons into account leads to a smaller minimum energy given by
0
Emin ð1 p=mP cÞEmin ;
B.7 A particle A decays at rest via A ! B þ C. Find the total energy of B in terms of the
three masses.
B.8 A meson M decays via M ! . Find an expression for the angle in the laboratory
between the two momentum vectors of the photons in terms of the photon energies
and the mass of M.
B.9 Pions and protons, both with momentum 2 GV/c, travel between two scintillation
counters distance L m apart. What is the minimum value of L necessary to
PROBLEMS 347
B.10 A photon is Compton scattered off a stationary electron through a scattering angle of
60 and its final energy is half its initial energy. Calculate the value of the initial
energy in MeV.
Appendix C
Rutherford Scattering
m vi ¼ m vf þ mt vt ðC:1Þ
and
m2t 2
m v2i ¼ m v2f þ v þ 2mt ðvf vt Þ ðC:3Þ
m t
1
For completeness one should also show that the observations cannot be due to scattering from the diffuse
positive charge present. This was done by the authors of the original experiment.
CLASSICAL PHYSICS 351
parameter). The derivation follows from considering the implications of linear and
angular momentum conservation.
Angular momentum conservation implies that
d
mvb ¼ mr 2 ; ðC:5Þ
dt
where v ¼ jvj. Since the scattering is symmetric about the y-axis, the component of
linear momentum in the y-direction is initially p ¼ mv sinð=2Þ and changes to
þmv sinð=2Þ after the interaction, i.e. the total change in momentum in the y-
direction is
p ¼ 2mv sinð=2Þ: ðC:6Þ
The change in momentum may also be calculated by integrating the impulse in the
y-direction due to the Coulomb force on the projectile. This gives
ð
þ1
zZe2
p ¼ cos dt; ðC:7Þ
4"0 r2
1
where we have taken t ¼ 0 to coincide with the origin of the x-axis. Using
Equation (C.5) to change variables, Equation (C.7) may be written
þ
ð
zZe2 1
2mv sinð=2Þ ¼ cos d; ðC:8Þ
4"0 bv
zZe2 1
b¼ cotð=2Þ; ðC:9Þ
8"0 Ekin
d d
dW ¼ J d
¼ 2J sin d ; ðC:11Þ
d
d
i.e.
d b db
¼ : ðC:12Þ
d
sin d
The right-hand side of Equation (C.12) may be evaluated from Equation (C.9) and
gives
2
d zZe2
¼ cosec4 ð=2Þ: ðC:13Þ
d
16 "0 Ekin
This is the final form of the Rutherford differential cross-section for non-
relativistic scattering.
d 1 p0 2 2
¼ 2 4 0 Mðq2 Þ ; ðC:14Þ
d
4 h vv
where v and p are the velocity and momentum respectively of the projectile (which
for convenience we take to have a unit negative charge) with v ¼ jvj, p ¼ jpj and
the primes refer to the final-state values. The matrix element is given by
ð
MðqÞ ¼ VðxÞeiqx=h dx; ðC:15Þ
ZðhcÞ
VðxÞ ¼ VC ðxÞ ¼ ; ðC:16Þ
r
where r ¼ jxj and Ze is the charge of the target nucleus. Inspection of the integral
in Equation (C.15) shows that it diverges at large r. However, in practice, charges
QUANTUM MECHANICS 353
To evaluate this, take q along the x-axis, so that in spherical polar coordinates
q x ¼ qr cos . The angular integration may then be done and yields
ð
1
4ðhcÞZh Lt
MC ðqÞ ¼ e
r sinðqr=hÞdr: ðC:18Þ
q
!0
0
The remaining integral may be done by parts (twice) and taking the limit
! 0
gives
4ðhcÞZh2
MC ðqÞ ¼ : ðC:19Þ
q2
d p0 2
¼ 4Z 2 2 ðhcÞ2 0 4 ; ðC:20Þ
d
vv q
which is the general form of the Rutherford differential cross-section. To see that
this is the same as Equation (C.13) in the non-relativistic limit, we may substitute
the non-relativistic approximations
2
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p2 ¼ p0 ¼ 2mEkin ; and v ¼ v0 ¼ 2Ekin =m; ðC:21Þ
d Z 2 2 ðhcÞ2
¼ 2 4 ; ðC:23Þ
d
4E sin ð=2Þ
Problems
C.1 Calculate the differential cross-section in mb/sr for the scattering of a 20 MeV
-particle through an angle 20 by a nucleus 209
83 Bi, stating any assumptions made.
Ignore spin and form factor effects.
C.2 Show that in Rutherford scattering at a fixed impact parameter b, the distance of
closest approach d to the nucleus is given by d ¼ b½1 þ cosec ð=2Þ=cosec ð=2Þ,
where is the scattering angle.
C.3 Find an expression for the impact parameter b in the case of small-angle Rutherford
scattering. A beam of protons with speed v ¼ 4 107 ms1 is incident normally on a
thin foil of 194 78 Pt, thickness 10
5
m (density ¼ 2:145 104 kg m3 ). Estimate the
proportion of protons that experience double scattering, where each scattering angle
is at least 5 .
Appendix D
Solutions to Problems
Chapter 1
1.1 Substituting the operators p ¼ i h@=@x and E ¼ i h@=@t into the mass–energy
relation E2 ¼ p2 c2 þ M 2 c4 and allowing the operators to act on the function
ðx, tÞ, leads immediately to the Klein–Gordon equation. To verify that the Yukawa
potential VðrÞ is a static solution of the equation, set VðrÞ ¼ ðxÞ, where r ¼ jxj,
and use
@2 2 @
r2 ¼ þ
@r2 r @r
1.3 Because the initial state is at rest, it has L ¼ 0 and thus its parity is
Pi ¼ Pp Pp ð1ÞL ¼ 1, where we have used the fact that the fermion–antifermion
pair has overall negative intrinsic parity. In the final state, the neutral pions are
identical bosons and so their wavefunction must be totally symmetric under their
interchange. This implies even orbital angular momentum L0 between them and
2 L0
hence Pf ¼ P ð1Þ ¼ 1 6¼ Pi . The reaction violates parity conservation and is thus
forbidden as a strong interaction.
1.4 Since ^ 2 ¼ 1, we must have C
C ^
^ 2 jb; b i ¼ Cb C b; b ¼ jb; b i, implying that
^ b; b ¼ Cb jb; b i with Cb Cb ¼ 1 independent of Cb . The result follows because
C
an eigenstate of C ^ must contain only particle–antiparticle pairs b b, leading to the
intrinsic parity factor Cb Cb ¼ 1, independent of Cb .
1.6 (a) e þ eþ ! e þ eþ ;
(b) p þ p ! p þ p þ 0 þ 0 ;
(c) p þ n ! þ 0 þ 0 ; þ þ þ .
1.7 (a) e þ ! e þ .
(b) n ! p þ e þ e .
CHAPTER 1 357
(c) eþ þ e ! eþ þ e .
(d)
þ
! eþ þ e .
ð
2 ð
1 ð
þ1
2 g2 er=R
2
f ðq Þ ¼ d dr r hÞ
d cos expðiqr cos =
4 r
0 0 1
ð
1 ð
1
h i
g2
h g2
h
¼ dre r=R
hÞþ1
½expðiqr cos = 1 ¼ drer=R eiqr=h eiqr=h
2iq 2iq
0 0
2 2
g
h
¼
q2 þ m2 c2
358 APPENDIX D: SOLUTIONS TO PROBLEMS
1.10 Let one of the beams (labelled by 1) refer to the ‘beam’ and let the other beam
(labelled by 2) refer to the ‘target’. Then in Equation (1.43), nb ¼ nN1 =2RA and
vi ¼ 2R=T, where R is the radius of the circular path. Thus the flux is
J ¼ nb vi ¼ nN1 f =A, where f is the frequency. Also N ¼ N2 , so finally the luminosity
is L ¼ JN ¼ nN1 N2 f =A.
1.11 From Equation (1.44c), ¼ WMA =Ið
tÞNA . Since the scattering is isotropic, the
total number of protons emitted from the target is W ¼ 20 ð4=2 103 Þ
¼ 1:25 105 s1 . I can be calculated from the current, noting that the -particles
carry two units of charge, and is I ¼ 3:13 1010 s1 . The density of the target is
t ¼ 1 mg cm2 ¼ 1032 kg fm2 . Putting everything together gives ¼ 161 mb.
Chapter 2
2.1 From Equation (2.21),
ðr
ðr 1
4
h
Fðq2 Þ ¼
r sin bðrÞdr 4 r2 dr ¼ 3½sin bðaÞ bðaÞ cos bðaÞb3 ;
q 0 0
1
from which q ¼ 2p sinð#=2Þ ¼ 57:5 MeV=c. Also, we know that a ¼ 1:21A3 fm and
so for A ¼ 56, a ¼ 4:63 fm and qa= h ¼ 1:35 radians. Finally, using this in the
integral, gives F ¼ 0:829 and hence the reduction is F 2 ¼ 0:69.
2.3 From Equation (2.28), r2 ¼ 6 h2 ½1 Fðq2 Þ=q2 , where q ¼ 2E sinð=2Þ: Thus,
pffiffiffiffiffiffiffiffi
q ¼ 43:6 MeV=c. Also, F 2 ¼ 0:65 and so hr 2 i ¼ 6:56 fm.
2.4 The charge distribution is spherical, so the angular integrations in the general result
of Equation (2.17) may be done, giving
21 32 1 31
ð ð
Fðq2 Þ ¼ 4
ðrÞ½sinðqr= hÞ4r2 dr54
ðrÞ4r2 dr5 :
hÞ=ðqr=
0 0
Ð
1
Substituting for
ðrÞ, setting x ¼ r=a and using x expðxÞ dx ¼ 1, gives, after
integrating by parts (twice), 0
1
ð qax
h 1
2
Fðq Þ ¼ ex sin dx ¼ :
qa
h h2
1 þ q2 a2 =
0
2.5 In 1 g of the isotope there are initially N0 ¼ 1 g=208 1:66 1024 g . Thus
N0 ¼ 2:9 1021 atoms. At time t there are NðtÞ ¼ N0 et= atoms, where is the
mean life of the isotope. Thus, provided t , the average decay rate is
N0 NðtÞ N0 75
¼ h1 :
t 0:1 24
2.6 The count rate is proportional to the number of 14 C atoms present in the sample.
If we assume that the abundance of 14 C has not changed with time, the artefact
was made from living material and is predominantly carbon, then at the time it
was made ðt ¼ 0Þ, 1 g would have contained 5 1022 carbon atoms of which
N0 ¼ 6 1010 would have been 14 C. Thus the average count rate would have been
N0 = ¼ 13:8 m1 . At time t, the number of 14 C atoms would be NðtÞ ¼ N0 expðt=Þ
and NðtÞ=N0 ¼ et= ¼ 2:1=13:8, from which t ¼ ln 6:57 ¼ 1:56 104 years. The
artefact is approximately 16 000 years old.
138
2.8 The total decay rate of both modes of 57 La is
1 1
2
ð1 þ 0:5Þ ð7:8 10 Þ kg s ¼ 1:17 103 kg1 s1 :
Also, since this isotope is only 0.09 per cent of natural lanthanum, the number
4
of 138 23
57 La atoms per kg is N ¼ ð9 10 Þ ð1000=138:91Þ ð6:022 10 Þ, i.e.
N ¼ 3:90 1021 kg1 . The rate of decays is dN=dt ¼ ! N, where ! is the
transition rate, and in terms of this the mean lifetime ¼ 1=!. Thus,
N 3:90 1021
¼ ¼ s ¼ 3:33 1018 s ¼ 1:06 1011 years:
dN=dt 1:17 103
2.9 The energy released is the increase in binding energy. Now from the SEMF,
Equations (2.46)–(2.52),
2.10 The most stable nucleus for fixed A has a Z-value given by Z ¼ =2
, where
from Equation (2.58), ¼ aa þ ðMn Mp me Þ and
¼ aa =A þ ac =ðAÞ1=3 .
Changing would not change aa , but would effect the Coulomb coefficient
because ac is proportional to . For A ¼ 111, using the value of aa from Equation
(2.54) gives ¼ 93:93 MeV=c2 and
¼ 0:839 þ 0:208 ac MeV=c2 . For Z ¼ 47,
ac ¼ 0:770 MeV=c2 . This is a change of about 10 per cent from the value given in
Equation (2.54) and so would have to change by the same percentage.
2.13 If there were N0 atoms of each isotope at the formation of the planet ðt ¼ 0Þ,
then after time t the numbers of atoms are N205 ðtÞ ¼ N0 expðt=205 Þ and
N204 ðtÞ ¼ N0 expðt=204 Þ, with
N205 ðtÞ 1 1 n205
¼ exp t ¼ ¼ 2 107 :
N204 ðtÞ 205 204 n204
and hence ½Mp þ Mð21; 46Þ ½Mn þ Mð22; 46Þ ¼ 1:59 MeV=c2 . We also need the
mass differences ½M þ Mð20; 43Þ ½Mn þ Mð22; 46Þ ¼ 0:07 MeV=c2 . We can
now draw the energy level diagram where the centre-of-mass energy of the
resonance is at (see Equation (2.10)) 2:76 ð45=47Þ ¼ 2:64 MeV.
35
2.16 The number of Cl atoms in 1 g of the natural chloride is
N ¼ 2 0:758 NA =molecular weight ¼ 7:04 1021 :
The activity AðtÞ ¼ N ¼ P 1 et Pt, since t 1. So
AðtÞ AðtÞt1=2
t¼ ¼ :
P ln2 F N
Substituting AðtÞ ¼ 3 105 Bq and using the other constants given, yields t ¼ 1:55
days.
2.17 At very low energies we may assume the scattering has ‘ ¼ 0 and so in Equation
(1.63) we have j ¼ 12; sn ¼ 12 and su ¼ 0. Thus,
h2 ðn n þ n
Þ 4
h2 n
max ¼ ¼ ;
q2n 2 =4 q2n
Therefore, n ¼ q2n max =4h2 ¼ 0:35 103 eV and
¼ n ¼ 9:65 103 eV.
Chapter 3
3.1 (a) Forbidden: violates L conservation, because L ð Þ ¼ 1, but L ð þ Þ ¼ 1.
(b) The quark compositions are: ¼ sud; p ¼ uud and since the dominant decay
of an s-quark is s ! u, we have
3.3 (a) This would be a baryon because B ¼ 1 and the quark composition would be ssb
which is allowed in the quark model.
(b) This would be a meson because B ¼ 0, but would have to have both an s- and a
b-quark. However, Qðs þ
bÞ ¼ 2=3, which is incompatible with the quark model
and anyway combinations of two antiquarks are not allowed. Thus this
combination is forbidden.
3.4 ‘Low-lying’ implies that the internal orbital angular momentum between the
quarks is zero. Hence the parity is P ¼ þ and space is symmetric. Since the
Pauli principle requires the overall wavefunction to be antisymmetric under
the interchange of any pair of like quarks, it follows that spin is antisymmetric.
Thus, any pair of like quarks must have antiparallel spins, i.e. be in a spin-0
state.
Consider all possible baryon states qqq, where q ¼ u; d; s. There are six
combinations with a single like pair: uud; uus; ddu; dds; ssu; ssd, with the spin
of (uu) etc. equal to zero. Adding the spin of the third quark leads to six states
þ
with J P ¼ 12 . In principle, there could be six combinations with all three quarks
the same – uuu; ddd; sss – but in practice these do not occur because it is
impossible to arrange all three spins in an antisymmetric way. Finally, there is one
combination where all three quarks are different: uds. Here there are no restric-
tions from the Pauli principle, so for example, the ud pair could have spin-0 or
þ
spin-1. Adding the spin of the s-quark leads to two states with J P ¼ 12 and 1 with
þ
J P ¼ 32 .
þ þ
Collecting the results, gives an octet of J P ¼ 12 states and a singlet J P ¼ 32 state.
This is not what is observed in nature. In Chapter 5 we will see what additional
assumptions have to be made to reproduce the observed spectrum.
364 APPENDIX D: SOLUTIONS TO PROBLEMS
3.5 (a)
(b)
3.6 The ground state mesons all have L ¼ 0 and S ¼ 0. Therefore they all have P ¼ 1.
Only in the case of the neutral pion is their constituent quark and antiquark also
particle and antiparticle. Thus C is only defined for the 0 and is C ¼ 1. For the
excited states, L ¼ 0 still and thus P ¼ 1 as for the ground states. However, the
total spin of the constituent quarks is S ¼ 1 and so for the
0 , the only state for which
C is defined, C ¼ 1.
For the excited states, by definition there is a lower mass configuration with the
same quark flavours. As the mass differences between the excited states and their
ground states is greater than the mass of a pion, they can all decay by the strong
interaction. In the case of the charged pions and kaons and the neutral kaon ground
states, there are no lower mass configurations with the same flavour structure and so
the only possibility is to decay via the weak interaction, with much longer lifetimes.
In the case of
0 decay, the initial state has a total angular momentum of 1 and
since the pions have zero spin, the final state must have L ¼ 1. While this is
possible for þ , for the case of 0 0 it violates the Pauli Principle and so is
forbidden.
BðY Þ ¼ 1. As charm and beauty for the initial state are both zero, these quantum
numbers are zero for the Y. The quark content is therefore dss. In the decay, the
strangeness of the is 1 and so strangeness is not conserved. This is therefore a
weak interaction and its lifetime will be in the range 107 –1013 s:
3.8 The quark composition is ¼ uds, then ðSu þ Sd Þ2 ¼S2u þ S2d þ 2Su Sd ¼ 2
h2 and
2
hence Su Sd ¼
h =4. Then, from the general formula given in Equation (3.84),
setting mu ¼ md ¼ m, we have
Su Sd Sd Ss þ Su Ss
M ¼ 2m þ ms þ b þ
m2 mms
Su Sd S1 S2 þ S1 S3 þ S2 S3 Su Sd
¼ 2m þ ms þ b þ
m2 mms
3.9 The initial reacton is strong because it conserves all individual quark numbers. The
decay is weak because strangeness changes by one unit and the same is true for
the decays of the
0 , K þ and K 0 . The decay of the þ is also weak because it
involves neutrinos and finally the decay of the 0 is electromagnetic because only
photons are involved.
The two vertices where the W-boson couples are weak interactions and have
pffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
strengths W . The remaining vertex is electromagnetic and has strength EM .
pffiffiffiffiffiffiffiffiffi
So the overall strength of the diagram is W EM .
3.12 Reactions (a), (d) and (f) conserve all quark numbers individually and hence are
strong interactions. Reaction (e) violates strangeness and is a weak interaction.
Reaction (c) conserves strangeness and involves photons and hence is an electro-
magnetic interaction. Reaction (b) violates both baryon number and electron lepton
number and is therefore forbidden.
and
Mðþ p ! þ K þ Þ ¼ M3 ;
where M1;3 are the amplitudes for scattering in a pure isospin state I ¼ 12; 32,
respectively. Thus,
3.14 Under charge symmetry, nðuddÞ Ð pðduuÞ and þ ðudÞ Ð ðd uÞ and since the
strong interaction is approximately charge symmetry, we would expect
ðþ nÞ ð pÞ at the same energy, with small violations due to electromagnetic
effects and quark mass differences. However, K þ ðusÞ and K ðs
uÞ are not charge
symmetric and so there is no reason why ðK þ nÞ and ðK pÞ should be equal.
Chapter 4
4.1 In an obvious notation,
2
ECM ¼ ðEe þ Ep Þ2 ðpe c þ pp cÞ2 ¼ ðEe2 p2e c2 Þ ðEp2 p2p c2 Þ þ 2Ee Ep 2pe pp c2
¼ m2e c4 þ m2p c4 þ 2Ee Ep 2pe pp c2
At the energies of the beams, masses may be neglected and so with p ¼ jpj,
2
ECM ¼ 2Ee Ep 2pe pp c2 cosð Þ ¼ 2Ee Ep ½1 cosð Þ;
where is the crossing angle. Using the values given, gives ECM ¼ 154 GeV. In a
2
fixed-target experiment, and again neglecting masses, ECM ¼ 2Ee Ep 2pe pp c2 ,
CHAPTER 4 367
1=2
where Ee ¼ EL ; Ep ¼ mp c2 ; pp ¼ 0. Thus, ECM ¼ 2mp c2 EL and for
ECM ¼ 154 GeV, this gives EL ¼ 1:26 104 GeV.
4.2 For constant acceleration, the ions must travel the length of the drift tube in half a
cycle of the rf field. Thus, L ¼ v=2 f , where v is the velocity of the ion. Since the
energy is far less than the rest mass of the ion, we can use non-relativistic kinematics
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
to find v, i.e. v ¼ c 200=ð12 931:5Þ ¼ 4:01 107 m s1 and finally L ¼ 1 m.
4.3 A particle with mass m, charge q and speed v moving in a plane perpendicular to a
constant magnetic field of magnitude B will traverse a circular path with radius of
curvature r ¼ mv=qB and hence the cyclotron frequency is f ¼ v=2r ¼ qB=2m.
At each traversal the particle will receive energy from the rf field, so if f is kept fixed,
r will increase (i.e. the trajectory will be a spiral). Thus if the final energy is E, the
pffiffiffiffiffiffiffiffiffi
extraction radius will be R ¼ 2mE=qB. To evaluate these expressions we use
q ¼ 2e ¼ 3:2 1019 C, together with B ¼ 0:8 T ¼ 0:45 1030 ðMeV=c2 Þs1 C1
and thus f ¼ 6:15 MHz and R ¼ 62:3 cm.
4.4 A particle with unit charge e and momentum p in the uniform magnetic field B of the
bending magnet will traverse a circular trajectory of radius R, given by p ¼ BR. If B
is in T, R in m and p in GeV/c, then p ¼ 0:3BR. Referring to the figure below, we
have L=R ¼ 0:3 LB=p and ¼ s=d ¼ 0:3BLp=p2 . Solving for d using the
data given, gives d ¼ 9:3 m.
erenkov condition is n 1. So, for the pion to give a signal, but not the
4.5 The C
kaon, we have n 1 K n. The momentum is given by p ¼ mv
where
1=2
¼ ð1 v2 =c2 Þ , so eliminating
gives ¼ v=c ¼ ð1 þ m2 c2 =p2 Þ1=2 . For
p ¼ 20 GeV=c, m ¼ 0:14 GeV=c2 and mK ¼ 0:49 GeV=c2 , ¼ 0:99997 and
K ¼ 0:99970, so the condition on the refractive index is 3 104 ðn 1Þ=n
3 105 . Using the largest value of n ¼ 1:0003, we have
1 1 1
N ¼ 2 1 2 2
n 1 2
as the number of photons radiated per metre, where 1 ¼ 400 nm and 2 ¼ 700 nm.
Numerically, N ¼ 26:5 photons/m and hence to obtain 200 photons requires a
detector of length 7.5 m. (You could also use
1 2 1
N ¼ 2 1 2 2
n 2
368 APPENDIX D: SOLUTIONS TO PROBLEMS
where is the mean of 1 and 2 , which would give 24.5 photons/m and a length of
8.2 m.)
4.6 Luminosity may be calculated from the formula for colliders, L ¼ n N1 N2 f =A,
where n is the number of bunches, N1 and N2 are the numbers of
particles in each bunch, A is the cross-sectional area of the beam and f is its
frequency. We have, n ¼ 12; N1 ¼ N2 ¼ 3 1011 ; A ¼ ð0:02 102 Þ cm2 and
f ¼ ð3 1010=8 105 Þ s1 , so finally L ¼ 6:44 1031 cm2 s1.
4.7 (a) The b quarks are not seen directly but, instead, they fragment (hadronize) to B-
hadrons, i.e. hadrons containing b quarks. So one characteristic is the presence of
hadrons with non-zero beauty quantum numbers. As these hadrons are unstable
and the dominant decay of b-quarks is to c-quarks, a second characteristic is the
presence of hadrons with non-zero values of the charm quantum number.
We need to observe the point where the eþ e collision occurred and the point
of origin of the decay products of the B-hadrons. The difference between these
two is due to the lifetime of the B-hadrons. As the difference will be very small,
precise position measurements are required. The daughter particles may be
detected using a silicon micro-vertex detector and an MWPC. In addition, any
electrons from the decays could be detected by an MWPC or an electromagnetic
calorimeter. The same is true for muons in the decay products, except they are
not readily detected in the calorimeter as they are very penetrating. However, if
one places an MWPC behind a hadron calorimeter then one can be fairly
confident that any particle detected is a muon, as everything else (except
neutrinos) will have been stopped in the calorimeter.
(b) In the electronic decay mode, the electron can be measured in both a MWPC
and an EM calorimeter. For high energies the better measurement is made in the
calorimeter. The neutrino does not interact unless there is a very large mass of
material (thousands of tons) and so its presence must be inferred by imposing
conservation of energy and momentum. In a colliding beam machine, the original
colliding particles have zero transverse momentum and a fixed energy. If one
adds up all the energy and momentum of all the final-state particles, then
any imbalance compared to the initial system can be attributed to the neutrino.
For the muonic mode, the muon can be measured in the MWPC but cannot be
measured well in the calorimeter because it only ionizes to a very small extent.
Since the muons only interact to a small extent they (along with neutrinos) are
generally the only particles that emerge from a hadronic calorimeter. So if one
registers a signal in a small MWPC placed behind a calorimeter then one can be
confident that the particle is a muon.
4.8 To be detected, the event must have 150 < < 30 , i.e. jcos j < 0:866. Setting
x ¼ cos , the fraction of events in this range is
ð
þ0:866 þ1:0
ð
d d þ0:866 þ1:0
f ¼ dx dx ¼ x þ x3 =3 0:866 x þ x3 =3 1:0 ¼ 0:812:
dx dx
0:866 1:0
CHAPTER 4 369
ð ð
2 ð
þ1 ð
þ1
d d h2 c2
2
¼ d ¼ d d cos ¼ 2 2
1 þ cos2 d cos :
d d 4Ecm
0 1 1
4.9 Referring to the figure below, the distance between two positions of the particle t
apart in time is v t. The wave fronts from these two positions have a difference in
their distance travelled of c t=n.
c t=n 1
cos ¼ ¼ :
v t n
The maximum value of corresponds to the minimum of cos and hence the
maximum of . This occurs as ! 1, when max ¼ cos1 ð1=nÞ. This value occurs
in the ultra-relativistic or massless limit.
1=2
The quantity may be expressed as ¼ pc=E ¼ pc½p2 c2 þ m2 c4 . Hence,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 p 2 c 2 þ m2 c 4
cos ¼ ;
n pc
4.10 The average distance between collisions of a neutrino and an iron nucleus is the
mean free path ¼ 1=n , where n
=mp c2 is the number of nucleons per cm3 .
Using the data given, n 4:7 1024 cm3 and 3 1036 cm2 , so that
7:1 1010 cm. Thus if 1 in 109 neutrinos is to interact, the thickness of iron
required is 71 cm.
4.11 Radiation energy losses are given by dE=dx ¼ E=LR, where LR is the radiation
length. This implies that E ¼ E0 expðx=LR Þ, where E0 is the initial energy. Using
E0 ¼ 2 GeV, LR ¼ 36:1 cm, x ¼ 10 cm, gives E ¼ 1:51 GeV. Radiation losses at
fixed E are proportional to m2 , where m is the mass of the projectile. Thus for
muons, they are negligible at this energy.
4.12 The total cross section is tot ¼ el þ cap þ f ¼ 4 102 b and the attenuation is
expðnxtot Þ where nx ¼ 101 NA =A ¼ 2:56 1023 m2 . Thus expðnxtot Þ ¼ 0:9898,
i.e 1.02 per cent of the incident particles interact and of these the fraction
that elastically scatter is given by the ratio of the cross-sections, i.e.
3 102 =4 102 ¼ 0:75 104 . Thus the intensity of elastically-scattered neu-
trons is 0:75 104 0:0102 106 ¼ 0:765 s1 and finally the flux at 5 m is
0:765=ð4 52 Þ ¼ 2:44 103 m2 s1 .
1
4.13 The total centre-of-mass energy is given by ECM ð2mc2 EL Þ2 ¼ 0:23 GeV and so
the cross-section is ¼ 1:64 1034 m2 . The interaction length is ‘ ¼ 1=n, where
n is the number density of electrons in the target. This is given by n ¼
NA Z=A,
where NA is Avogadro’s number and for lead,
¼ 1:14 107 kg m3 is the density,
Z ¼ 82 and A ¼ 208. Thus n ¼ 2:7 1033 m3 and ‘ ¼ 2:3 m.
4.14 The target contains n ¼ 1:07 1025 protons and so the total number of interactions
per second is N ¼ n flux tot ¼ ð1:07 1025 Þ ð2 107 Þ ð40 1031 Þ ¼
856 s1 . There are thus 856 photons/s produced from the target.
The latter has a maximum for v2 ¼ eI=2me. Thus for a proton in iron we can use
I ¼ 10Z eV ¼ 260 eV, so that Ep ¼ 12mp v2 ¼ mp Ie=4me ¼ 324 keV.
4.16 From Equation (4.24), EðrÞ ¼ V=r lnðrc =ra Þ and at the surface of the anode this is
0:5=ð20 106 Þ lnð500Þ ¼ 4023 kV m1 . Also, if Ethreshold ðrÞ ¼ 750 kV m1 , then
from Equation (4.24) r ¼ 0:107 mm and so the distance to the anode is 0:087 mm.
CHAPTER 5 371
This contains 22 mean free paths and so assuming each collision produces an ion
pair, the multiplication factor is 222 ¼ 4:2 106 ¼ 106:6 .
Chapter 5
5.1 We have m ¼ þ þ
> n ¼ þ þ
, where the inequality is because baryon
number B > 0. Using the values of the colour charges I3C and Y C from Table 5.1, the
colour charges for the state are:
By colour confinement, both these colour charges must be zero for observable
hadrons, which implies ¼ ¼
p and hence m n ¼ 3p, where
p is a non-negative integer. Thus the only combinations allowed by colour
confinement are of the form
ð3qÞp ðq
qÞn ðp; n 0Þ:
It follows that a state with the structure qq is not allowed, as no suitable values of p
and n can be found.
5.2 (a)
(b)
372 APPENDIX D: SOLUTIONS TO PROBLEMS
(c)
with
2
ECM pÞ2 ¼ x2 2m2 c4 þ 2E2 þ 2p2 c2 :
¼ x2 c2 ½PðpÞ þ Pð
Neglecting the masses of the proton and the antiproton at these energies, gives
2 2
W 2 c4 ¼ ½ðE E0 Þ þ EP ½ðp p0 Þ þ P c2 ¼ invariant mass of X:
CHAPTER 5 373
2M ¼ Q2 þ ðW 2 M 2 Þc2 Q2 :
From the definition of x, it follows that x 1. Finally, x > 0 because both Q2 and
2M are positive.
5.5 In the quark model, ¼ uds; p ¼ uud; K ¼ su; n ¼ udd and þ ¼ ud. From the
flavour independence of the strong interaction, we can set ðqqÞ ¼ ðudÞ ¼
ðsdÞ etc. and ðqqÞ ¼ ðudÞ ¼ ðs
uÞ etc.. Then ðpÞ ¼ ðppÞ ¼ 9ðqqÞ and
ðK nÞ ¼ ðþ pÞ ¼ 3ðqqÞ 3ðqqÞ. The result follows directly.
5.6 By analogy with the QED formula, we have ð3gÞ ¼ 2ð2 9Þ6s mc c2 =9, where
mc 1:5 GeV=c2 is the constituent mass of the c-quark. Evaluating this gives
s ¼ 0:31. In the case of the radiative decay, ðgg
Þ ¼ 2ð2 9Þ4s 2 mb c2 =9,
where mb 4:5 GeV=c2 is the constituent mass of the b-quark. Evaluating this gives
s ¼ 0:32. (These values are a little too large because in practice is replaced by
4
3s .)
ð1 ð1 ð1
dx 1 1
F2ep ðxÞ F2en ðxÞ ¼ ½uðxÞ þ
uðxÞ dx ½dðxÞ þ dðxÞdx:
x 3 3
0 0 0
However, summing over all contributions we must recover the quantum numbers of
the proton, i.e.
ð1 ð1
½uðxÞ
uðxÞ dx ¼ 2; ½dðxÞ dðxÞ dx ¼ 1:
0 0
374 APPENDIX D: SOLUTIONS TO PROBLEMS
Eliminating the integrals over u and d gives the Gottfried sum rule.
5.8 Substituting Equation (5.22) into Equation (5.23) and setting NC ¼ 3, gives
X
R ¼ 3ð1 þ s =Þ e2q ;
5.9 A proton has the valence quark content p ¼ uud. Thus from isospin invariance the u
quarks in the proton carry twice as much momentum as the d quarks, which implies
a ¼ 2b. In addition, we are told that
ð1 ð1
1
xFu ðxÞdx þ xFd ðxÞdx ¼ :
2
0 0
Using the form of the quark distributions with a ¼ 2b gives a ¼ 43 and b ¼ 23.
where we have used C-invariance to relate the distribution functions for protons and
antiprotons. In the narrow width approximation and using the quark distributions
from Question 5.9,
ð1 ð1 !
ð1 xu Þ3 ð1 xd Þ3 xu s
pp ðsÞ ¼ C 1 xd dxu dxd
xu xd ðMW c2 Þ2
0 0
ð1
3
ð1 xu Þ3 k
pp ðsÞ ¼ C 1 dxu ;
xu xu
k
CHAPTER 6 375
where k
ðMW c2 Þ=s and the lower limit is because k < xu < 1. The integral yields
8 W 2 3 11 2 11 3
pp ðsÞ ¼ max ð1 þ 9k þ 9k þ k ÞlnðkÞ 9k þ 9k þ k :
9 MW c2 3 3
pffiffi
Evaluating this for s ¼ 1 TeV gives k ¼ 0:0064 and pp ¼ 9:3 nb, which is about a
factor of two larger than experiment.
Chapter 6
6.1 A charged current weak interaction is one mediated by the exchange of charged W
boson. A possible example is n ! p þ e þ e . A neutral current weak interaction is
one mediated by a neutral Z 0 boson. An example is þ p ! þ p. Charged
current weak interactions do not conserve the strangeness quantum number, whereas
neutral current weak interactions do. For þ e ! þ e, the only Feynman
diagram that conserves both Le and L is:
which is a weak neutral current. However, for e þ e ! e þ e, there are two
diagrams:
Thus the reaction has both neutral and charged current components and is not
unambiguous evidence for weak neutral currents.
376 APPENDIX D: SOLUTIONS TO PROBLEMS
d d d 2 h2 c2
¼ þ ¼ 2
1 þ Cwk cos þ cos2 :
d d em d wk 4ECM
Then, using
ð1
F ¼ C 1 þ Cwk cos þ cos2 d cos
0
and
ð0
B ¼ C 1 þ Cwk cos þ cos2 d cos :
1
2 2 2
2
where C
2
h c =4ECM , gives
4 Cwk 4 Cwk
F ¼ C þ and B ¼ C
3 2 3 2
CHAPTER 6 377
and so
Cwk
AFB ¼ ; i:e: 8AFB ¼ 3Cwk :
2ð4=3Þ
The amplitude has two factors of the weak coupling gW and one W propagator
carrying a momentum q, i.e.
g2W g2
amplitude / 2
/ W2 ;
q2 c2 MW c 4 MW
When Ee m c2 =2, the electron has its maximum energy and the two neutrinos
must be recoiling in the opposite direction. Only left-handed particles (and right-
handed antiparticles) are produced in weak interactions. Since the masses of all
particles are neglected, states of definite handiness are also states of definite helicity,
so the orientations of the momenta and spins are therefore as shown:
378 APPENDIX D: SOLUTIONS TO PROBLEMS
c2 =2
m ð
2G2F ðm c2 Þ2 4Ee3 G2F ðm c2 Þ5
¼ Ee2 dE e ¼ :
ð2Þ3 ð
hcÞ6 3m c2 1923 ðhcÞ6
0
6.5 (a) In addition to the decay b ! c þ e þ e, there are two other leptonic decays
ð‘ ¼ ; Þ and by lepton universality they will all have equal decay rates.
There are also hadronic decays of the form b ! c þ X where QðXÞ ¼ 1.
Examining the allowed Wq q vertices using lepton–quark symmetry shows that
the only forms that X can have, if we ignore Cabibbo-suppressed modes, are
du and sc. Each of these hadronic decays has a probability three times that
of a leptonic decay because the quarks exist in three colour states. Thus,
there are effectively six hadronic channels and three leptonic ones. So finally,
BRðb ! c þ e þ e Þ ¼ 19.
(b) The argument is similar to that of (a) above. Thus, in addition to the decay
! e þ e þ , there is also the leptonic decay ! þ þ with
equal probability and the hadronic decays ! þ X. In principle, X ¼ d u
and sc, but the latter is not allowed because ms þ mc > m . So the only allowed
hadronic decay is ! d þ u þ with a relative probability of three because
of colour. So finally, BRð ! e þ e þ Þ ¼ 15. (The measured rate is 0.18,
but we have neglected kinematic corrections.)
6.6 For neutrinos, gR ðÞ ¼ 0; gL ðÞ ¼ 12. So, e ¼ ¼ ¼ 0 =4, where
GF MZ3 c6
0 ¼ pffiffiffi ¼ 668 MeV:
3 2ð hcÞ3
Thus the partial width for decay to neutrino pairs is ¼ 501 MeV. For quarks,
gR ðu; c; tÞ ¼ 16 and gL ðu; c; tÞ ¼ 13. Thus, u ¼ c ¼ 10
720 . Also, gR ðd;
P s; bÞ ¼ 12
1
5
and gL ðb; s; dÞ ¼ 12 . Thus, d ¼ s ¼ b ¼ 13
72 0 . Finally, q ¼ i , where
i ¼ u; c; d; s; b – no top quark because 2Mt > MZ . So, i
3 13 2 10 59
q ¼ þ 0 ¼ 0 ¼ 547 MeV:
72 72 72
pffiffiffiffiffiffiffi
In the case of the charged pion, there are two vertices of strength W , and there
will be a propagator
1 1
2 c2
2 2;
Q2 þ MW MW c
380 APPENDIX D: SOLUTIONS TO PROBLEMS
because the momentum transfer (squared) Q2 carried by the W is very small. Thus
the decay rate will be proportional to
pffiffiffiffiffiffiffipffiffiffiffiffiffiffi2
W W 2
2
¼ W
4
:
MW MW
pffiffiffiffiffiffiffiffi
In the case of the neutral pion, there are two vertices of strength em , but no
2
propagator. Thus the decay rate will be proportional to em and since em W , the
decay rate for the charged pion will be much smaller than that for the neutral decay,
i.e. the lifetime of the 0 will be much shorter.
Using lepton–quark symmetry and the Cabibbo hypothesis, the two hadron vertices
are given by gudW ¼ gW cos C and gusW ¼ gW sin C . So, if we ignore kinematic
differences and spin effects, we would expect the ratio of decay rates is given by
Rate ðK ! þ Þ g2usW
R¼ / ¼ tan2 C 0:05
Rate ð ! þ Þ g2udW
The measured ratio is actually about 1.3, which shows the importance of the
neglected effects. For example, the Q-value for the kaon decay is almost 20 times
that for pion decay.
6.10 To a first approximation the difference in the two decay rates is due to two effects.
First, ! n þ e þ e has jSj ¼ 1 and hence is proportional to sin2 C , where
C is the Cabbibo angle, whereas ! þ e þ e has jSj ¼ 0 and is propor-
tional to cos2 C . Secondly, the Q-values are different for the two reactions. Thus,
using Sargent’s Rule,
5
sin2 C Qn 5 257
R 0:053 ¼ 17:0:
cos2 C Q 81
6.11 The required number of events produced must be 20 000, taking account of the
detection efficiency. If the cross-section is 60 fb ¼ 6 1038 cm2 , then the inte-
grated luminosity required is 2 104 =6 1038 ¼ ð1=3Þ 1042 cm2 and hence
the instantaneous luminosity must be 3:3 1034 cm2 s1 .
The branching ratio for Z 0 ! bb is found from the partial widths to be 15 per cent.
Thus, if b quarks are detected, the much greater branching ratio for H ! b b will
help distinguish this decay from the background of Z 0 ! b b.
6.12 By ‘adding’ an I ¼ 12 particle to the initial state we can assume isospin invariance
holds. Consider
þ S0 ! þ . The final state is jI ¼ 1; I3 ¼ 1i and so is the
initial state because I3 ðS0 Þ ¼ 12. Thus the transition is pure I ¼ 1 and the rate is
jM1 j2 . For
0 þ S0 ! þ 0, the final state is again pure I ¼ 1 but with I3 ¼ 0.
However, the initial state is an equal mixture of I ¼ 0 and I ¼ 1, i.e.
0
S ¼ p1ffiffiffijI ¼ 1; I3 ¼ 0i p1ffiffiffijI ¼ 0; I3 ¼ 0i
2 2
and so the rate is 12jM1 j2 . Thus R ¼ 2: (The measured value is about 1.8.)
6.13 Integrating the differential cross-sections over y (from 0 to 1) gives for a spin-12 target
with a specific quark distribution
21 32 1 31
ð ð
NC ðÞ 4 2 2 2 54 dy5 ¼ g2 þ 1g2
¼ ½gL þ gR ð1 yÞ dy L
CC ðÞ 3 R
0 0
and
21 32 1 31
ð ð
NC ð
Þ 4 2
¼ ½gL ð1 yÞ2 þ g2R dy54 ð1 yÞ2 dy5 ¼ g2L þ 3g2R :
CC ð
Þ
0 0
For an isoscalar target, we must add the contributions for u and d quarks in equal
amounts, i.e.
NC ðÞ 1 1
ðisoscalarÞ ¼ g2L ðuÞ þ g2R ðuÞ þ g2L ðdÞ þ g2R ðdÞ
CC ðÞ 3 3
and
NC ð
Þ
ðisoscalarÞ ¼ g2L ðuÞ þ 3g2R ðuÞ þ g2L ðdÞ þ 3g2R ðdÞ:
CC ð
Þ
NC ðÞ 1 2 20 4 NC ð
Þ 1 20
¼ sin W þ sin W ; ¼ sin2 W þ sin4 W :
CC ðÞ 2 27 CC ð
Þ 2 9
382 APPENDIX D: SOLUTIONS TO PROBLEMS
Chapter 7
7.1 For the 73 Li nucleus, Z ¼ 3 and N ¼ 4. Hence the configuration is
2 1 2 2
protons: 1s1=2 1p3=2 ; neutrons: 1s1=2 1p3=2 :
By the pairing hypothesis, the two neutrons in the 1p3=2 sub-shell will have a total
orbital angular momentum and spin L ¼ S ¼ 0 and hence J ¼ 0. Therefore they will
not contribute to the overall nuclear spin, parity or magnetic moment. These will be
determined by the quantum numbers of the unpaired proton in the 1p3=2 sub-shell.
This has J ¼ 32 and ‘ ¼ 1, hence for the spin-parity we have J P ¼ 32 . The magnetic
moment is given by
1
¼ j gproton ¼ j þ 2:3 ðsince j ¼ ‘ þ Þ ¼ 1:5 þ 2:3
2
¼ 3:8 nuclear magnetons:
If only protons are excited, the two most likely excited states are:
2 1 2 2
protons: 1s1=2 1p1=2 ; neutrons: 1s1=2 1p3=2 ;
which corresponds to exciting a proton from the p3=2 sub-shell to the p1=2 sub-shell,
and
1 2 2 2
protons: 1s1=2 1p3=2 ; neutrons: 1s1=2 1p3=2 ;
which corresponds to exciting a proton from the s1=2 sub-shell to the p3=2 sub-shell.
7.2 A state with quantum number jð¼ ‘ 12Þ can contain a maximum number
Nj ¼ 2ð2j þ 1Þ nucleons. Therefore, if Nj ¼ 16 it follows that j ¼ 72 and ‘ ¼ 3 or
4. However, we know that the parity is odd and since P ¼ ð1Þ‘ , it follows that
‘ ¼ 3.
and the two d5=2 protons could combine to give jP ¼ 2þ , so that when this combines
with the single unpaired jP ¼ 32 proton the overall spin-parity is jP ¼ 12 . There are
many other possibilities.
7.4 For 93
41 Nb, Z ¼ 41 and N ¼ 52. From the filling diagram Figure 7.4, the configuration
is predicted to be:
þ
So ‘ ¼ 4, j ¼ 92 ) jP ¼ 92 (which agrees with experiment). The magnetic dipole
moment follows from the expression for jproton in Equations (7.31) with j ¼ ‘ þ 12, i.e.
¼ ð j þ 2:3Þ N ¼ 6:8 N . (The measured value is 6:17 N .)
For 33
16 S, Z ¼ 16 and N ¼ 17. From the filling diagram Figure 7.4, the configura-
tion is predicted to be:
with
¼ Ze=ð43b2 aÞ and the integral is through the volume of the spheroid
ðx2 þ y2 Þ=b2 þ z2 =a2 1. The integral can be transformed to one over the volume
of a sphere by the transformations x ¼ bx0 ; y ¼ by0 and z ¼ az0 . Then
ððð
3Z
Q¼ dx0 dy0 dz0 ð2a2 z02 b2 x02 b2 y02 Þ:
4
But
ððð ð1
02 0 0 1 02 0 4
x dx dy dz ði:e: z Þ ¼ r 4r02 dr0 ¼ ;
3 15
0
and similarly for the other integrals. Thus, by direct substitution, Q ¼ 25Zða2 b2 Þ.
7.6 From Question 7.5 we have Q ¼ 25Zeða2 b2 Þ and using Z ¼ 67 this gives
a2 b2 ¼ 13:1 fm2 . Also, from Equation (2.32) we have A ¼ 43ab2
, where
¼ 0:17 fm3 is the nuclear density. Thus, ab2 ¼ 231:7 fm3 . The solution of
these two equations gives a 6:85 fm and b 5:82 fm.
7.7 From Equation (7.53), t1=2 ¼ ln2= ¼ CR ln2 expðGÞ, where C is a constant
formed from the frequency and the probability of forming -particles in the nucleus.
384 APPENDIX D: SOLUTIONS TO PROBLEMS
Thus t1=2 ðThÞ ¼ t1=2 ðCfÞ exp½GðThÞ GðCfÞ. The Gamow factors may be calcu-
lated from the data given. Some intermediate quantities are: rC ¼ 45:96 fm (Th);
37.72 fm (Cf); R ¼ 9:268 fm (Th); 9.439 fm (Cf) (using R ¼ 1:21 ðA1=3 þ 41=3 Þ and
recalling that ðZ; AÞ refer to the daughter nucleus). These give G ¼ 66:5 (Th); 54.9 (Cf)
and t1=2 ðThÞ ¼ e11:6 t1=2 ðCfÞ ¼ 4:0 years. (The measured value is 1.9 years).
þ
7.8 The J P values of the 0 and the are both 12 (see Chapter 3), so the photon has
L ¼ 1 and as there is no change of parity the decay proceeds via an M1 transition.
þ
The 0 has J P ¼ 32 and again there is no parity change. Therefore both M1 and E2
multipoles could be involved, with M1 dominant (see Section 7.8.2). If we assume
that the reduced transition probabilities are equal in the two cases, then from
Equations (7.80), in an obvious notation,
3
E
ð0 Þ
ð0 Þ ¼ ð0 Þ;
E
ð0 Þ
i.e. ð0 Þ ¼ ð292=77Þ3 ð0:6 1023 Þ=0:0056 ¼ 5:8 1020 s (the measured
value is ð7:4 0:7Þ 1020 s).
7.10 From Equation (7.71) the electron energy spectrum may be written IðEÞ ¼ AE1=2
ðE0 EÞ2 , where E is the electron energy, E0 is the end-point, A is a constant and we
have neglected the Fermi screening correction and set the neutrino mass to be zero.
We need to calculate the fraction
2 32 31
ð0
E Eð0
6 74
F
4 IðEÞ dE5 IðEÞ dE5
E0 0
Ð
where is a small quantity. Using x1=2 ða xÞ2 dx ¼ 12a2 x2 23ax3 þ 14x4 1=2 , gives,
using E0 ¼ 18:6 103 eV and ¼ 10 eV, F ¼ 3:1 1010 .
is defined by
7.11 The mean energy E
2E 32E 31
ð0 ð0
4 E d!ðEÞ54 d!ðEÞ5 :
E
0 0
and
ð
2 3=2
E1=2 ðE0 EÞ2 dE ¼ E 35E02 42E0 E þ 15E2 :
105
¼ 1E0 , as required.
Substituting the limits gives E 3
From Figure 7.13, the dominant multipole for a fixed transition energy will be M1
for the 32 ! 52 and 32 ! 12 transitions and E2 for the 52 ! 12 transition. Thus
we need to calculate the rate for an M1 transition with E
¼ 178 keV. This can be
done using Equations (7.80) and gives 1=2 3:9 1012 s. The measured value is
3:5 1010 s, which confirms that the Weisskopf approximation is not very
accurate.
7.13 Set L ¼ 3 in Equation (7.78a), substitute the result into Equation (7.77) and use
¼ hT to give
ðE3Þ ¼ ð2:3 1014 ÞE
7 A2 eV, where E
is expressed in MeV.
Chapter 8
8.1 To balance the number of protons and neutrons, the fission reaction must be
n þ 235 92 140
92 U ! 37 Rb þ 55 Cs þ 4n;
i.e. four neutrons are produced. The energy released is the differences in binding
energies of the various nuclei, because the mass terms in the SEMF cancel out. We
have, in an obvious notation,
" #
2=3 ðZ NÞ2 Z2
ðAÞ ¼ 3; ðA Þ ¼ 9:26; ¼ 0:28; ¼ 485:0:
4A A1=3
The contribution from the pairing term is negligible (about 1 MeV). Using the
numerical values for the coefficients in the SEMF, the energy released per fission
EF ¼ 157:9 MeV.
The power of the nuclear reactor is P ¼ nEF ¼ 100 MW ¼ 6:25 1020 MeV s1 ,
where n is the number of fissions per second. Since one neutron escapes per fission and
386 APPENDIX D: SOLUTIONS TO PROBLEMS
contributes to the flux, the flux F is equal to the number of fissions per unit area per
second, i.e.
n P 6:25 1020 MeVs1
F¼ ¼ ¼ ¼ 3:15 1017 s1 m2 :
4r2 4r2 EF ð157:9 MeVÞ ð12:57 m2 Þ
8.2 The neutron speed in the CM system is v mv=ðM þ mÞ ¼ Mv=ðM þ mÞ and if the
scattering angle in the CM system is , then after the collision the neutron will have
a speed vðm þ M cos Þ=ðM þ mÞ in the original direction and Mv sin =ðM þ mÞ
perpendicular to this direction. Thus the kinetic energy is
2 32 31
ð1 ð1
4
Efinal ¼ E Eðcos Þ d cos 5 4 d cos 5 ¼ REinitial ;
1 1
where the reduction factor is R ¼ ðM 2 þ m2 Þ=ðM þ mÞ2 . For neutron scattering from
graphite, R 0:86 and after N collisions the energy will be reduced to
Efinal ¼ RN Einitial . The average initial energy of fission neutrons from 235 U is
2 MeV and to thermalize them their energy would have to be reduced to about
0.025 eV. Thus N lnðEfinal =Einitial Þ=lnð0:86Þ ¼ 116.
8.3 From Equation (1.44a), for the fission of 235 U, Wf ¼ JNð235Þf and the total power
output is P ¼ Wf Ef , where Ef is the energy released per fission. For the capture by
238
U, Wc ¼ JNð238Þc . Eliminating the flux J, gives
Nð238ÞC P
Wc ¼ :
Nð235Þf Ef
Using the data supplied, gives Wc ¼ 1:08 1019 atoms s1 135 kg year1 .
8.4 Consider fissions occurring sequentially separated by a small time interval t. The
instantaneous power is the sum of the power released from all the fissions up to that
time. If E is the energy released in each fission, then over the lifetime of the reactor,
i.e. up to time T, the power is given by P0 ¼ nE=T, where n is the total number of
fissions and t ¼ E=P0 .
CHAPTER 8 387
The power after some time t after the reactor has been shut down is
In this formula, the first term is the power released from the first fission and the last
term is the power released from the last fission before the reactor was shut down. To
sum this series, we convert it to an integral:
TP0ð=EF
X
n¼P 0 T=E
1:2
PðtÞ ¼ 3 ðT þ t nE=P0 Þ 3 ðT þ t nE=P0 Þ1:2 dn:
n¼0
0
ðt h i
P0
PðtÞ ¼ 3 u1:2 du ¼ 0:075P0 t0:2 ðT þ tÞ0:2 :
E
Tþt
Using T ¼ 1 year and t ¼ 0:5 year, gives a power output of approximately 1.1 MW
after 6 months.
8.6 A solar constant of 8:4 J cm2 s1 is equivalent to 5:25 1013 MeV cm2 s1 of
energy deposited. If this is due to the PPI reaction 4ð1 HÞ ! 4 He þ 2eþ þ 2e þ 2
,
then this rate of energy deposition corresponds to a flux of ð5:25 1013=2
6:55Þ 4 1012 neutrinos cm2 s1 .
8.7 For the Lawson criterion to be just satisfied, from Equation (8.46),
nd hdt itc Q
L¼ ¼ 1:
6kT
We have kT ¼ 10 keV and from Figure 8.7 we can estimate hdt i 1022 m3 s1 .
Also, from Equation (8.45), Q ¼ 17:6 MeV. So, finally, nd ¼ 6:8 1018 m3 .
388 APPENDIX D: SOLUTIONS TO PROBLEMS
8.8 The mass of a d–t pair is 5:03 u ¼ 4:69 109 eV=c2 ¼ 8:36 1024 g. The number
of d–t pairs in a 1 mg pellet is therefore 1:2 1020 . From Equation (8.45), each d–t
pair releases 17.6 MeV of energy. Thus, allowing for the efficiency of conversion,
each pellet releases 5:3 1026 eV. The output power is 750 MW ¼ 4:7 1027 eV=s.
Thus the number of pellets required is 8:9 9 s1 .
8.9 Assume a typical body mass of 70 kg, half of which is protons. This corresponds to
2:1 1028 protons and after 1 year the number that will have decayed is
2:1 1028 ½1 expð1=Þ, where is the lifetime of the proton in years. Each
proton will eventually deposit almost all of its rest energy, i.e. approximately 0.938
GeV, in the body. Thus in 1 year the total energy in Joules deposited per kg of body
mass would be 4:5 1016 ½1 expð1=Þ and this amount will be lethal if greater
than 5 Gy. Expanding the exponential gives the result that the existence of humans
implies > 0:9 1016 years.
8.10 The approximate rate of whole-body radiation absorbed is given by Equation (8.48a).
Substituting the data given, we have
dD AðMBqÞ E
ðMeVÞ ð40 103 Þ ð1:173 þ 1:333Þ
ð Sv h1 Þ ¼ ¼
dt 6r2 ðm2 Þ 6
¼ 1:67 102 Sv h1
8.11 If the initial intensity is I0 , then from Equation (4.18), the intensities after passing
through bone, Ib , and tissue, It , are
8.12 From Figure 4.8, the rate of ionization energy losses is only slowly varying for
momenta above about 1 GeV/c and given that living matter is mainly water and
hydrocarbons a reasonable estimate is 3 MeV g1 cm2 . Thus the energy deposited in
1 year is 2:37 109 MeV kg1 , which is 3:8 104 Gy.
8.13 In general, the nuclear magnetic resonance frequency is f ¼ jljB=jh. The numerical
input we use is:
Appendix B
B.1 (a) From the definitions of s, t and u, we have
ðs þ t þ uÞc2 ¼ ðp2A þ 2pA pB þ p2B Þ þ ðp2A 2pA pC þ p2C Þ þ ðp2A 2pA pD þ p2D Þ
and hence
X
ðs þ t þ uÞ ¼ mj 2 :
j¼A;B;C;D
E ¼
m c2 ; E ¼ cðm2 c2 þ p2 Þ1=2 ; E ¼ p c
and hence
2
m c2 p c ¼ c2 m2 c2 þ p2 : ð1Þ
p cos ¼ p ¼
m v; p sin ¼ pv ¼ E =c: ð2Þ
ðm2 m2 Þ
tan ¼ :
2
2 m2
390 APPENDIX D: SOLUTIONS TO PROBLEMS
E E
p pv ¼ p p ¼ m E ¼ m jp jc;
c2
because p ¼ 0 and E ¼ m c2 in the rest frame of the pion. However, jp j ¼ p
p
because the muon and neutrino emerge
h back-to-back.
i 1 Thus, p ¼ ðm2 m2 Þ c=2m ;
2
but p ¼
m v, from which v ¼ pc p2 þ m2 c2 . Finally, substituting for p gives
!
m2 m2
v¼ c:
m2 þ m2
B.5 If the four-momenta of the initial and final electrons are p ¼ ðE=c; qÞ and
p0 ¼ ðE0 =c; q0 Þ, respectively, the squared four-momentum transfer is defined by
where the invariant mass squared s is defined by s
ðp þ PÞ2 =c2 and P is the four-
momentum of the initial proton, i.e. P ¼ ðmp c; 0Þ. Thus,
invariant mass squared in the final state evaluated in the centre-of-mass frame has a
minimum value ð4mp cÞ2 when all four particles are stationary. Thus, Emin is given by
nucleons is 250 MeV=c (see Chapter 7), E mp c2 , while for the relativistic
incident protons pmin Emin =c, so using these gives
0
Emin 1 p=mp c Emin ¼ 4:8 GeV:
which on expanding gives EB ¼ m2A þ m2B m2C c2 =2mA .
B.8 If the four-momenta of the photons are pi ¼ ðEi =c; pi Þði ¼ 1; 2Þ, then the invariant
mass of M is given by M 2 c4 ¼ ðE1 þ E2 Þ2 ðp1 þ p2 Þc2 ¼ 2E1 E2 ð1 cos Þ, since
p1 p2 ¼ E1 E2 ð1 cos Þ=c2 for zero-mass photons. Thus, cos ¼ 1 M 2 c4 =2E1 E2 .
B.9 A particle with velocity v will take time t ¼ L=v to pass 1 between the two
counters. Relativistically,
1
p ¼ mv
with
¼ ð1 v2 =c2 Þ2 . Solving, gives
v ¼ cð1 þ m2 c2 =p2 Þ 2 and hence the difference in times-of-flight (assuming
m1 > m2 ) is
"
1
1 #
L m21 c2 2 m22 c2 2
t ¼ 1þ 2 1þ 2 :
c p p
ðE
; p
Þ þ e ðmc2 ; 0Þ !
ðE
0 ; p0
Þ þ e ðE; pÞ:
392 APPENDIX D: SOLUTIONS TO PROBLEMS
Appendix C
C.1 The assumptions are: ignore the recoil of the target nucleus because its mass is
much greater than the total energy of the projectile -particle; use non-relativistic
kinematics because the kinetic energy of the -particle is very much less that its rest
mass; assume the Rutherford formula (i.e. the Born approximation) is valid for
small-angle scattering. The relevant formula is then Equation (C.13) and it may
be evaluated using z ¼ 2; Z ¼ 83, Ekin ¼ 20 MeV and ¼ 20 . The result is
d=d ¼ 98:3 b=sr.
C.2 From Figure C.2, the distance of closest approach d is when x ¼ 0. For x < 0, the
sum of the kinetic and potential energies is Etot ¼ 12mv2 and the angular momentum
is mvb. At x ¼ 0, the total mechanical energy is 12mu2 þ Zze2 =4"0 d and the
angular momentum is mud, where u is the instantaneous velocity. From angular
momentum conservation, u ¼ vb=d and using this in the conservation of total
mechanical energy gives d 2 Kd b2 ¼ 0 where, using Equation (C.9),
K
2b=cotð=2Þ. The solution for d 0 is d ¼ b½1 þ cosecð=2Þ=cotð=2Þ.
C.3 The result for small-angle scattering follows directly from Equation (C.9) in the
limit ! 0. Evaluating b, we have, using the data given,
2
zZe2 e
hc 1
b¼ ¼ 2zZ ¼ 1:55 1013 m:
2
2"0 mv 4"0 hc mc ðv=cÞ2
2
Aj90 F. Ajzenberg-Selove (1990) Energy levels of light nuclei., Nucl. Phys., A506
1–158.
Am95 P. Amaudruz et al. (1995) A re-evaluation of the nuclear structure function ratios
for D, He, 6Li, C and Ca, Nucl. Phys., B441, 3–11.
Ar95 M. Arneodo et al. (1995) The structure function ratios F2C /F2D and F2Ca /F2D at small
x. Nucl. Phys., B441, 12–30.
Ar97 M. Arneodo et al. (1997) Measurements of the proton and deuteron structure
functions F2p and F2d and of the ratio L /T . Nucl. Phys., B483, 3–43.
As04 Y. Ashie. et al. (2004) Evidence for an oscillatory signature in atmospheric
neutrino oscillations. Phys. Rev. Letters, 93, 101801/1–5.
At82 W. B. Atwood (1982) Lectures on Lepton Nucleon Scattering and Quantum
Chromodynamics, Progress in Physics Vol. 4.
Ba77 R. C. Barrett and D. F. Jackson (1977) Nuclear Sizes and Structures, Clarendon
Press, Oxford, UK.
Be67 J. B. Bellicard et al. (1967) Scattering of 750-MeV electrons by calcium isotopes.
Phys. Rev. Letters, 19, 527–529.
Bl52 J. M. Blatt and V. F. Weisskopf (1952), Theoretical Nuclear Physics, John Wiley
and Sons, New York, USA.
Bo69 A. Bohr and B. R. Mottelson (1969) Nuclear Structure Vol. 1, W. A. Benjamin Inc.,
New York, USA.
Ch97 Chart of the Nuclides (1997) General Electric Company, Schenectady, New York,
USA.
Co01 W. N. Cottingham and D. A. Greenwood (2001) An Introduction to Nuclear
Physics, 2nd edn. Cambridge University Press, Cambridge, UK.
De99 P. P. Dendy and B. Heaton (1999) Physics for Diagnostic Radiology, 2nd edn.,
Institute of Physics Publishing, UK.
Ei04 S. Eidelman et al. (2004) Review of Particle Physics. Physics Letters B592, 1–
1109.
En66 H. A. Enge (1966) Introduction to Nuclear Physics, Addison-Wesley Publishing
Company.
Fe86 R. Fernlow (1986) Introduction to Experimental Particle Physics, Cambridge
University Press, Cambridge, UK.
Fo61 D. B. Fossan et al. (1961) Neutron total cross sections of Be, B10 , B, C, and O.
Phys. Rev., 123, 209–218.
Fr83 B. Frois (1983) Proc. Int. Conf. Nucl. Phys., Florence eds. P. Blasi and R. A. Ricci,
(Tipografia Compositori, Bolgona) Vol. 2, p. 221.
Pe03 D. H. Perkins (2003) Particle Astrophysics. Oxford University Press, Oxford, UK.
Ph94 A. C. Phillips (1994) The Physics of Stars. John Wiley and Sons. Ltd., Chichester,
UK.
Po99 B. Povh et al. (1999) Particles and Nuclei. An Introduction to the Physical
Concepts. Springer, Germany.
Sa67 G. R. Satchler (1967) Optical model for 30 MeV proton scattering. Nucl. Phys.,
A92, 273–305.
Sc68 L. I. Schiff (1968) Quantum Mechanics, 3rd edn.. McGraw-Hill, New York, USA.
Sc01 W. Scheider (2001) News Update to A Serious But Not Ponderous Book About
Nuclear Energy, Cavendish Press, Ann Arbour, USA.
Se80 E. Segrè (1980) From X-Rays to Quarks. W. H. Freeman and Company, USA.
Se97 W. G. Seligman et al. (1997) Improved determination of s from neutrino-nucleon
scattering. Phys. Rev. Letters, 79, 1213–1216.
Si75 I. Sick et al. (1975) Shell structure of the 58 Ni charge density. Phys. Rev. Letters,
35, 910–913.
Tr75 G. L. Trigg (1975) Landmark Experiments in 20th Century Physics. Crane, Russak
and Co. Ltd, New York, USA.
Zu00 A. Zurlo et al. (2000) The role of proton therapy in the treatment of large
irradiation volumes: a comparative planning study of pancreatic and biliary
tumors. Int. J. Radiat. Oncol. Biol. Phys, 48, 277–288.
Bibliography
Below are brief notes on a few books on nuclear and particle physics at the
appropriate level which I have found particularly useful. Other, more specialized,
texts are listed in the References section which follows.
1 Nuclear Physics
Two very readable concise texts at about the level of the present book although
covering more topics are: W. N. Cottingham and D. A. Greenwood, An Introduc-
tion to Nuclear Physics 2nd edn., Cambridge University Press, 2001, and N. A.
Jelley, Fundamentals of Nuclear Physics, Cambridge University Press, 1990. Both
deal with theoretical aspects only – there is nothing about experimental methods.
Both provide some problems for each chapter with either answers or brief hints on
solutions. Another good book at this level is: J. Lilley, Nuclear Physics –
Principles and Applications, John Wiley and Sons, 2001. This is in two parts.
The first covers the principles of nuclear physics, including experimental techni-
ques, and the second discusses an unusually wide range of applications, including
industrial and biomedical uses. An extensive range of problems is provided, with
detailed notes on their solutions.
Two good examples of comprehensive texts covering both theory and experi-
ment are: K. S. Krane, Introductory Nuclear Physics, John Wiley and Sons, 1988,
and P. E. Hodgson, E. Gadioli and E. Gadioli Erba, Introductory Nuclear Physics,
Oxford University Press, 1997. Both provide problems, but without solutions.
Finally there is the unique set of (hand written!) lecture notes by Fermi: E.
Fermi, Nuclear Physics, University of Chicago Press, 1950. Although old, these
are still well worth reading.
2 Particle Physics
There are several books covering particle physics at the appropriate level, For
obvious reasons, the one closest to the present book is: B. R. Martin and G. Shaw,
Particle Physics, 2nd edn., John Wiley and Sons, 1997, and some of the material
on particle physics in the present book has been developed from this previous
book. It covers both theory and experimental methods. Problems with full
solutions are provided for each chapter.
Some of the other texts available are now rather dated, but one that is not is: D.
H. Perkins, Introduction to High Energy Physics, 4th edn., Cambridge University
Press, 2000. This book is well-established and has changed substantially over the
years. It goes further than the present book in its use of relativistic calculations.
The latest edition (the fourth) has far less discussion of experimental methods
than earlier editions, but an expanded chapter on astroparticle physics. It is
therefore worth looking at the third edition also. Problems are provided, some
with answers.
Another older book, but still relevant, is: D. Griffiths, Introduction to Elemen-
tary Particle Physics, John Wiley and Sons, 1987. Griffiths’ book is written in an
unconventional conversational style, with interesting footnotes (and extensive
notes at the end of most chapters), giving further details and background. It is
exclusively theoretical – there is nothing on experimental techniques. It goes well
beyond the present text, as at least half of the book involves the detailed evaluation
of Feynman diagrams. A wealth of interesting problems is provided at the end of
each chapter, but no solutions are given.
A. Das and T. Ferbel, Introduction to Nuclear and Particle Physics, John Wiley
and Sons, 1994;
B. Povh, K. Rih, C. Scholz and F. Zetsche, Particles and Nucle, 2nd edn., Springer,
1995.
The first two books are concise readable introductions, although in Dunlap’s case
the particle physics part is very short – just 50 pages. This book is exclusively
about theory, whereas the book of Das and Ferbel also discusses experimental
NUCLEAR AND PARTICLE PHYSICS 399
methods. Both books provide problems, but neither supplies solutions. The book
by Williams is fairly comprehensive, although now a little old. The style is rather
discursive. There is a wealth of illustrations and many problems are given, with
answers to some of them supplied. (A full solutions manual is available as a
separate volume.) The book by Burcham and Jobes is also comprehensive and goes
further than the present text. There are many problems, all with solutions. Both of
the latter two books treat nuclear and particle physics as almost independent
subjects. The book by Povh et al. is closest in its coverage to the present book and
at a similar level, although experimental methods are only discussed in a brief
appendix. Some problems with solutions are provided for all chapters.
Index