Preview-9780192540270 A38255748
Preview-9780192540270 A38255748
The Oxford Master Series is designed for final year undergraduate and beginning graduate students in physics and
related disciplines. It has been driven by a perceived gap in the literature today. While basic undergraduate physics texts
often show little or no connection with the huge explosion of research over the last two decades, more advanced and
specialized texts tend to be rather daunting for students. In this series, all topics and their consequences are treated at a
simple level, while pointers to recent developments are provided at various stages. The emphasis is on clear physical
principles like symmetry, quantum mechanics, and electromagnetism which underlie the whole of physics. At the
same time, the subjects are related to real measurements and to the experimental techniques and devices currently
used by physicists in academe and industry. Books in this series are written as course books, and include ample
tutorial material, examples, illustrations, revision points, and problem sets. They can likewise be used as preparation
for students starting a doctorate in physics and related fields, or for recent graduates starting research in one of these
fields in industry.
Michael E. Peskin
SLAC National Accelerator Laboratory, Stanford University
1
3
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Michael E. Peskin 2019
The moral rights of the author have been asserted
First Edition published in 2019
Reprinted with corrections in 2023
Michael E. Peskin
Sunnyvale, CA
August, 2018
Figure permissions
Physics Institute of the University of Bonn. Figures 5.5 and 5.7 are re-
produced from the proceedings of the 1981 International Symposium on
Lepton and Photon Internations at High Energy and used by permission
of the institute.
Toshinori Mori. Figure 17.4 is used by permission of the author.
Aldo Serenelli. Figure 20.3 is used by permission of the author.
SLAC. Figures 6.7, 8.2, 9.1, 10.3, and 10.4 are reproduced from the
web site of the SLAC National Accelerator Laboratory and are used by
permission of the laboratory achivist.
Springer. Figures 10.9, 12.2, 12.3, 17.3, 17.6, and 20.1 are reproduced
from the European Journal of Physics C. Figures 8.3 and 15.5 are re-
produced from Zeitschrift für Physik C.
Taylor & Francis. Figures 9.4 and 11.2 are reproduced from An In-
troduction to Quantum Field Theory, by Michael E. Peskin and Daniel
V. Schroeder, c 1995, by permission of Westview Press, an imprint of
Perseus Books, LLC. This book was recently acquired by the Taylor &
Francis Group, a subsidiary of Informa UK Ltd. and is now published
under the imprint of CRC Press.
ZEUS Collaboration. Figure 9.6, from the ZEUS Collaboration web site,
is used by permission of the collaboration.
Contents
2 Symmetries of Space-Time 9
2.1 Relativistic particle kinematics 10
2.2 Natural units 12
2.3 A little theory of discrete groups 14
2.4 A little theory of continuous groups 18
2.5 Discrete space-time symmetries 22
Exercises 23
19 CP Violation 295
19.1 CP violation in the K 0 –K 0 system 295
19.2 Electric dipole moments 302
19.3 CP violation in the B 0 –B 0 system 303
Exercises 310
IV Epilogue 343
22 Epilogue 345
A Notation 355
References 367
Index 377
Part I
the nuclear force. The electron turned out to be only one of three appar-
ently pointlike particles with electric charge but no strong interactions.
All of these particles were observed to interact with one another through
a web of new, short-ranged interactions. Finally, as the 1960’s turned to
the 1970’s, the new interactions were sorted into two basic forces—called
the strong and the weak interaction—and simple mathematical expres-
sions for these forces were constructed. Today, physicists refer to these
expressions collectively as “the Standard Model of particle physics”.
Sometimes, authors or lecturers present the table of elementary par-
ticles of the Standard Model and imply that this is all there is to the
story. It is not. The way that the forces of nature act on the elementary
particles is beautiful and intricate. Often, the telling details of these
interactions show up through remarkable aspects of the data when we
examine elementary particle behavior experimentally.
These ideas elicit a related question: Of all the ways that nature could
be built, how do we know that the Standard Model is the correct one?
It seems hardly possible that we could pin down the exact nature of new
fundamental interactions beyond gravity and electromagnetism. All of
the phenomena associated with the new forces occur at distances smaller
than an atomic nucleus, and in a regime where both special relativity
It is important to remember that the and quantum mechanics play an essential role.
theory of particle physics must be stud- In this book, I will explain the answers to these questions. It turns
ied together with the understanding
of how experiments are done and how
out that the new forces have common properties and can be built up
their results are interpreted. from simple ingredients. The presence of these ingredients is revealed
by well-chosen experiments. The dynamics of the new interactions be-
comes more clear at higher energies. With the benefit of hindsight, we
can begin our study today by studying these dynamical ingredients in
their simplest form, working out the consequences of these laws, and
comparing the resulting formulae to data from high energy accelerator
experiments that illustrate the correctness of these formulae in a very
direct way.
Our quest for a fundamental theory of nature is far from complete.
In the final chapter of the book, I will discuss a number of issues about
fundamental forces for which we still have no understanding. It is also
possible, as we probe more deeply into the structure of nature, that we
will uncover new interactions that work at even smaller distances than
those currently explored. But, at least, one chapter of the story, open
since 1896, is now finished. I hope that, working through this book,
you will not only understand how to work with the underlying theories
describing the strong and weak interactions, but also that you will be
amazed at the wealth of evidence that supports the connection of these
theories to the real world.
Outline of the book. The book is organized into three Parts. Part I introduces the basic
materials that we will use to probe the nature of new forces at short
distances. Parts II and III use this as a foundation to build up the
Standard Model theories of the strong and weak interactions.
Part I Part I begins with basic theory that underlies the subject of particle
physics. Even before we attempt to write theories of the subnuclear
5
forces, we expect that those theories will obey the laws of quantum
mechanics and special relativity. I will provide some methods for using
these important principles to make predictions about the outcome of
elementary particle collisions.
In addition, I will describe the types of matter in the theories of
strong and weak interactions, the basic elementary particles that in-
teract through these forces. It turns out that there are two types of
matter particles that are elementary at the level of our current under-
standing. Of these, one type, the leptons, are seen in our experiments as particles of the Standard Model: the
individual particles. There are six known leptons. Three have electric leptons
charge: the electron (e), the muon (µ), and the tau lepton (⌧ ). The
other three are the neutrinos, particles that are electrically neutral and
extremely weakly interacting. Despite this, the evidence for neutrinos
as ordinary relativistic particles is very persuasive; I will discuss this in
Part III.
Matter particles of the other type, the quarks, are hidden from view.
Quarks appear as constituents of particles such as protons and neutrons particles of the Standard Model: the
that interact through the strong interaction. There are many known quarks
strongly interacting particles, collectively called hadrons. I will explain
the properties of the most prominent ones, and show that they are nat-
urally considered in families. On the other hand, no experiment has
ever seen an isolated quark. It is actually a prediction of the Standard
Model that quarks can never appear singly. This makes it especially
challenging to learn their properties. One piece of evidence that the
description of quarks in the Standard Model is correct is found from the
fact it gives a simple explanation for the quantum numbers of observed
hadrons and their assortment into families. I will discuss this also in
Part I. In the process, I will give names to the hadrons that appear
most often in experiments, so that we can discuss experimental methods
more concretely.
In a relativistic quantum theory, forces are also associated with parti-
cles that can be thought to transmit them. The Standard Model contains
four types of such particles. These are the photon, the carrier of the elec- particles of the Standard Model: the
tromagnetic interaction, the gluon, the carriers of the strong interaction, bosons
the W and Z bosons, the carriers of the weak interaction, and the Higgs
boson, which plays a more subtle role. You will have already encoun-
tered the photon in your study of quantum mechanics. I will introduce
the gluon in Part II and the W , Z, and Higgs bosons in Part III.
To understand experimental findings about elementary particles, we
will need to know at least the basics of how experiments on elementary
particles are done, and what sorts of quantities describing their proper-
ties are measureable. I will discuss this material also in Part I.
Part II begins with a discussion of the most important experiments Part II
that give insight into the underlying character of the strong interaction.
One might guess intuitively that the most convincing data on the strong
interaction comes from the study of collisions of hadrons with other
hadrons. That is incorrect. The experiments that were most crucial in
understanding the nature of strong interaction involved electron scatter-
6 Introduction
ing from protons and the annihilation of electrons and positrons at high
energy. This latter process has an initial state with no hadrons at all.
I will begin Part II with a discussion of the features of these processes
at high energy. Our analysis will introduce the concept of the current-
current interaction, which is an essential part of the physics of both the
strong and weak interactions. Starting from the surprisingly simple fea-
tures of these electromagnetic probes of strongly interacting particles,
we will develop a series of arguments that pass back and forth between
theory and experiment and eventually lead to a unique proposal for the
underlying theory of the strong interaction. We will then describe tests
of this theory of increasing sophistication, finally encompassing recent
results from experiments at the Large Hadron Collider.
The final chapter of Part II presents our current understanding of the
masses of quarks. At first sight, it might seems that it is straightforward
to measure the mass of a quark, although the fact that quarks cannot
be observed individually makes this more difficult. In fact, it turns
out that the question of the quark masses brings in a number of new,
subtle concepts. In particular, we will need to understand the idea of
spontaneous symmetry breaking, a concept that will also prove to be an
essential part of the theory of the weak interaction.
Part III Part III presents the description of the weak interaction. Here I
will begin from a proposal for the nature of the weak interaction that
again is based on a current-current interaction. I will present some
quite counterintuitive, and even startling, predictions of that proposal
and show that they are actually reproduced by experiment. From this
starting point, again in dialogue between theory and experiment, we
will build up the full theory. My discussion will include the precision
study of the carriers of the weak interaction, the W and Z bosons, and
the newest ingredients in this theory, the masses of neutrinos and the
properties of the Higgs boson.
In comparing theory and experiment, I will generally present deriva-
tions of the formulae that give the most important theoretical predic-
tions. To fully understand this book, it is important that you work
through these derivations rather than just skimming the mathematics.
For each one, take a large sheet of blank paper and carry out the cal-
culation yourself. The details of the calculation will reveal insights that
cannot by obtained from a purely qualitative discussion. You will find
that these insights accumulate as you go through the book, as successive
derivations cover similar ground from new perspectives.
Still, these derivations will be simplified with respect to a complete
treatment of the Standard Model of particle physics. Most of the pro-
cesses that I will consider will be studied in the limit of very high ener-
gies, where the mathematical analysis can be reduced as much as possible
and made more transparent. This will be sufficient to discover the equa-
tions of the Standard Model and understand their basic tests. But a full
discussion of the implications of the Standard Model would cover a more
complete list of reactions, including some whose theoretical analysis is
quite complex. A treatment of particle physics at that level is beyond
7
In multiplying matrices and vectors in this book, I will use the con-
vention that repeated indices are summed over. Then, for example, I
will write (2.3) as
p0µ = ⇤µ ⌫ p⌫ , (2.5)
In this book, unless it is explicitly in- omitting the explicit summation sign for the index ⌫. Lorentz trans-
dicated otherwise, repeated indices are formations leave invariant the Minkowski space vector product
summed over. This convention is one
of Einstein’s lesser, but still much ap-
preciated, innovations.
p · q = Ep E q p~ · ~q . (2.6)
To keep track of the minus sign in this product, I will make use of
raised and lowered Lorentz indices. Lorentz transformations preserve
2.1 Relativistic particle kinematics 11
p · q = pµ ⌘µ⌫ q ⌫ . (2.8)
Alternatively, let q with a lowered index be defined by I will use raised and lowered Lorentz
indices to keep track of the minus
qµ = ⌘µ⌫ q ⌫ = (Eq , ~q )µ . (2.9) sign in the Minkowski vector prod-
uct. Please pay attention to the po-
sition of indices—raised or lowered—
The invariant product of p and q is written throughout this book.
p · q = pµ qµ . (2.10)
I will define the mass of a particle as its rest-frame energy The mass of a particle is a Lorentz-
invariant quantity that characterizes
(mc2 ) ⌘ E0 . (2.13) that particle in any reference frame.
(mc2 )2 = p2 = E 2 p| 2 c 2
|~ (2.14)
pµ = (Ep , ~ pc
)µ or pµ = mc2 (1, ~ )µ , (2.15)
To illustrate these conventions, I will now work out some simple but
important exercises in relativistic kinematics. Imagine that a particle of
mass M , at rest, decays to two lighter particles, of masses m1 and m2 . In
the simplest case, both particles have zero mass: m1 = m2 = 0. Then,
energy-momentum conservation dictates that the two particle energies
are equal, with the value M c2 /2. Then, if the final particles move in the
3̂ direction, we can write their 4-vectors as
The next case, which will appear often in the experiments we will
consider, is that with m1 nonzero but m2 = 0. In the rest frame of the
original particle, the momenta of the two final particles will be equal
and opposite. With a little algebra, one can determine
These kinematic formulae will be used (for motion in the 3̂ direction), where
very often in this book.
M 2 + m21 2 M 2 m21 2
Ep = c , pc = c . (2.19)
2M 2M
It is easy to check that these formulae satisfy the constraints of to-
tal energy-momentum conservation and that pµ1 satisfies the mass-shell
constraint (2.14).
Finally, we might consider the general case of nonzero m1 and m2 .
Here, it takes a little more algebra to arrive at the final formulae
with
M 2 + m21 m22 M2 m21 + m22 2
E1 = c2 , E2 = c , (2.21)
2M 2M
and
c
p= ( (M, m1 , m2 ))1/2 , (2.22)
2M
where the kinematic function is defined by
These three sets of formulae apply equally well to reactions with two
particles in the initial state and two particles in the final state. It is
only necessary to replace M c2 with the center of mass energy ECM of
the reaction.
An electron with a momentum of the order of its rest energy has, ac-
cording to the Heisenberg uncertainty principle, a position uncertainty
h̄ 11
= 3.9 ⇥ 10 cm , (2.26)
me c
which I will equally well write as
1 1
= (0.51 MeV) . (2.27)
me
Natural units make it very intuitive to estimate energies, lengths, and
times in the regime of elementary particle physics. For example, the Natural units are useful for estimation.
lightest strongly interacting particle, the ⇡ meson, has a mass
factors for moving between distance, time, and energy units are given in
Appendix B.
One interesting quantity to put into natural units is the strength of
the electric charge of the electron or proton. In this book, the constant
e will be a positive quantity equal to the electric charge of the proton.
The electric charge of the electron will be ( e). More generally, I will
write the charge of a particle as Qe, with Q = +1 for a proton and
Q = 1 for an electron.
The Coulomb potential between charges 1 and 2 is given in standard
notation by
Q1 Q2 e2
V (r) = . (2.31)
4⇡✏0 r
I will use units for electromagnetism in which also
✏ 0 = µ0 = 1 . (2.32)
Q 1 Q2 e2 1
V (r) = . (2.33)
4⇡ r
Since r, in natural units, has the dimensions of (energy) 1 , the value of
the electric charge must have a form in which it is dimensionless. Indeed,
e2
↵⌘ (2.34)
4⇡✏0 h̄c
is a dimensionless number, called the fine structure constant, with the
value
↵ = 1 / 137.036 . (2.35)
The intrinsic strengths of the basic el- There are two remarkable things about this equation. First, it is sur-
ementary particle interactions are not prising that there is a dimensionless number ↵ that characterizes the
apparent from the size of their e↵ects—
or from their names. Here is a preview.
strength of the electromagnetic interaction. Second, that number is
small, signalling that the electromagnetic interaction is a weak inter-
Group theory plays an important role action. One of the goals of this book will be to determine whether the
in quantum mechanics, and this im- strong and weak subnuclear interactions can be characterized in the same
portance extends to the study of el-
ementary particle physics. You have
way, and whether these interactions—looking beyond their names—are
encountered group theory concepts in intrinsically strong or weak. I will discuss estimates of the strong and
your quantum mechanics course, but it weak interaction coupling strengths at appropriate points in the course.
is likely that those arguments did not It will turn out that the strong interaction is weak, at least when mea-
make explicit reference to group the-
ory. In particle physics, we lean much
sured under the correct conditions. It will also turn out that the weak
more heavily on group theory, and so interaction is also weak in dimensionless terms. It is weaker than the
it is best to discuss these concepts strong interactions, but not as weak as electromagnetism.
formally and give them their proper
names. Please, then, study Sections 2.3
and 2.4 carefully, especially if you are
uncomfortable with mathematical ab- 2.3 A little theory of discrete groups
straction. With careful reading, you
will see that the concepts I describe Group theory is a very important tool for elementary particle physics.
generalize physical arguments that are In this section and the next, I will review how group theory is used in
already familiar to you.
quantum mechanics, and I will discuss some properties of groups that we
2.3 A little theory of discrete groups 15
will meet in this book. To do this, I will mainly discuss systems that you
have already studied in your quantum mechanics course, giving a new
description of these systems using a more formal and abstract mathe-
matical language,. Please do not be put o↵by this. The mathematical
terms and concepts that I will introduce will generalize to and, I hope,
illuminate, many new systems that we will study in this book.
In quantum mechanics, we deal with groups on two levels. First,
there are abstract groups. In mathematics, a group is a set of elements
G = {a, b, . . .} with a multiplication law defined, so that ab is defined and
is an element of G. The multiplication law satisfies the three properties Here are the axioms that define a
group.
(1) Multiplication is associative: a(bc) = (ab)c.
(2) G contains an identity element 1 such that, for any element of G,
1a = a1 = a.
(3) For each a in G, there is an inverse element a 1 in G such that
aa 1 = a 1 a = 1.
Every symmetry of nature normally encountered in physics satisfies these
axioms and is described by an abstract group.
Next, we need to relate the abstract group to operators that act on
the space of states in a quantum mechanics problem. A group trans-
formation is a symmetry of a quantum-mechanical system if it leaves
the dynamics of that system invariant. This will be true if the trans-
formations commute with the Hamiltonian H. It is useful to state this
relationship more precisely.
In quantum mechanics, the quantum states are vectors in a Hilbert
space. Symmetries act to convert one of these states into another. A
symmetry transformation carries each state into another one in such
as way that the whole Hilbert space is mapped into itself, preserving
norms, that is, preserving quantum mechanical probabilities. Thus it
must act as a unitary transformation of the Hilbert space. A group G
is then described in a quantum mechanics problem as a set of unitary
transformations {U (a)}, one for each element of G, that obey the mul-
tiplication law of G. That is, if a, b, c are elements of G and ab = c, the
transformations in the set should satisfy U(a) U(b) = U (c). Such a set
of unitary transformations is called a unitary representation of G. The action of a group on the Hilbert
The group G will be a symmetry of the quantum mechanics problem space of states in quantum mechanics is
described through unitary representa-
if, for all elements a of G, the corresponding unitary transformations tions of the group, that is, through uni-
commute with the Hamiltonian, tary matrices with the same multipli-
cation law as the corresponding group
[ U (a), H ] = 0 . (2.36) elements. Thus, unitary group repre-
sentations will be used in many aspects
A direct consequence of (2.36) is that, if | i is an eigenstate of H with of the physics discussed in this book.
energy E, U (a) | i will also be an eigenstate of H with the same energy.
The inverse operation to a will be represented by the inverse transfor-
mation: U (a 1 ) = U(a) 1 = U (a)† .
Without reference to quantum mechanics, but just thinking about the
pure mathematics of groups, we can ask the following question: Given
a group G with elements {a}, can we find a set of finite-dimensional
16 Symmetries of Space-Time
unitary matrices Ua that obey the multiplication law of the group, that
is, such that Ua Ub = Uc ? This is a classical mathematics problem, and
mathematicians have classified the possible answers for all of the groups
commonly encountered in physics, and for more complex examples, in-
cluding one called the “Monster” group. These finite-dimensional uni-
tary representions of groups will appear in quantum mechanics problems
as tranformations among a finite number of eigenstates of the Hamilto-
nian that implement a symmetry of the problem.
The theory of group representations is a deep subject that can become
quite technically complex. It is certainly not necessary to master this
subject in order to study elementary particle physics. In this book, we
will make use of only the simplest examples. However, it is often good to
recognize that a physics problem of interest involves the representations
of a relevant symmetry group. This gives a starting point to analyze
the problem and connects the solution to those of other problems with
which we might be more familiar.
Here is a simple example: Consider the abstract group called Z2 that
contains two elements {1, 1} satisfying the multiplication law
1 · 1 = ( 1)( 1) = 1 , 1 · ( 1) = ( 1) · 1 = ( 1) . (2.37)
This is given by
✓ ◆ ✓ ◆
1 0 |⇡ + i
acting on . (2.41)
0 1 |⇡ i
U (~
↵) = exp[ i~ ~
↵ · J] (2.53)
where ↵~ gives the axis and angle of the rotation and J~ are the opera-
As in the previous example, the con-
tors of angular momentum. These operators satisfy the commutation servation law of angular momentum is
relation associated with the symmetry of invari-
[J i , J j ] = i✏ijk J k . (2.54) ance under rotations.
20 Symmetries of Space-Time
J i = 0, (2.56)
t0 = 1 (2.63)
This matrix commutes with all of the other ta . If we omit this element Definition of the group SU (n).
from the set of Hermitian matrices, we obtain a non-Abelian group of
matrices with n2 1 generators, the n ⇥ n Hermitian matrices with zero
trace. The group generated by these n2 1 matrices is called SU (n). It
is the group of n ⇥ n unitary matrices with determinant 1.
For n = 2, the Pauli sigma matrices (2.58) form a basis for the 2 ⇥ 2
traceless Hermitian matrices. Thus, SO(3) and SU (2) are names for the
same abstract group. (Mathematicians make a distinction between these
groups, but the di↵erence will not be relevant to the calculations done
in this book.) This abstract group describes rotations in three dimen-
sions, but it will also describe some internal symmetries of elemementary
particles that we will meet in the course of our discussion.
A continuous group of transformations generated by Hermitian matri-
ces, in the form (2.62), is called a Lie group. The commutation algebra
of the generators ta , This equation, which expresses the non-
[ta , tb ] = if abc tc (2.64) commuting nature of the generators of
a Lie group, contains the full informa-
is called the Lie algebra of the group. The constants f abc are called tion about the representations and the
the structure constants of the Lie algebra. It can be shown that we can geometry of the group.
always choose a basis for the ta such that the structure constants f abc are
completely antisymmetric in [abc]. These definitions straightforwardly
generalize the presentation that I have given of the rotation group in 3
dimension. In the case of the rotation group,
In the same way as for the rotation group, the Lie algebra of the gen-
erators determines the multiplication law of any two elements of the
group.
In this book, we will meet only special cases of Lie groups. The
particular groups U (1) = SO(2), SU (2) = SO(3), and SU (3) will
have important roles in our story. Still, the abstract properties of Lie
groups will be useful to us in understanding how to apply these groups
to particle physics. I will introduce some further formalism of Lie groups
when we will need it in Chapter 11.
Exercises
(2.1) Consider the decay of a particle of mass M , at both nonzero. With an appropriate choice of axes,
rest, into two particles with masses m1 and m2 , the momentum vectors of the final particles can be
24 Exercises
(@ 2 + m2 ) (x) = 0 . (3.5)
We can also see the invariance of (3.5) by examining the solutions of this
equation explicitly. These are
·~
iEt+i~px ip·x
(x) = e =e , (3.6)
E2 p|2 = p2 = m2 .
|~ (3.7)
E = ±Ep (3.10)
3.1 The Klein-Gordon equation 27
The principle that S is stationary with respect to all variations of the so-
lution x(t) yields the equation of motion of the system. Mathematically,
if x(t) ! x(t) + x(t), then we can write S in the form
Z
S[x(t), ẋ(t)] = dt x (t) E[x(t), ẋ(t), ẍ(t)] . (3.12)
L = @µ @µ m2 (3.15)
Putting this under the integral d4 x and integrating by parts in the first
term, we find
Z
S = d x (x) ( @ 2 m2 ) (x) .
4
(3.16)