Introduction To Particle Physics
Introduction To Particle Physics
Particle Physics
Introduction to
Particle Physics
By
All rights for this book reserved. No part of this book may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any means,
electronic, mechanical, photocopying, recording or otherwise, without
the prior permission of the copyright owner.
Foreword xii
I Particle phenomenology 1
1 Particles and symmetries 3
1.1 Symmetries in particle physics . . . . . . . . . . . . . . . . 3
1.2 Symmetry groups and spin . . . . . . . . . . . . . . . . . . 4
1.3 Fermions and bosons . . . . . . . . . . . . . . . . . . . . . 6
1.4 Mirror reflection: parity . . . . . . . . . . . . . . . . . . . 8
1.5 Charge conjugation . . . . . . . . . . . . . . . . . . . . . . 10
1.6 CPT symmetry . . . . . . . . . . . . . . . . . . . . . . . . 10
1.7 Isospin and strangeness . . . . . . . . . . . . . . . . . . . . 11
3 Quark model 19
3.1 Coloured quarks . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2 Colour interaction, QCD . . . . . . . . . . . . . . . . . . . 21
3.3 Reminder: summing up spins . . . . . . . . . . . . . . . . . 21
3.4 Lightest mesons . . . . . . . . . . . . . . . . . . . . . . . . 22
3.5 Meson nonet (flavour SU(3)) . . . . . . . . . . . . . . . . . 22
3.6 Ground-state baryons . . . . . . . . . . . . . . . . . . . . . 25
3.7 Baryon multiplets . . . . . . . . . . . . . . . . . . . . . . . 26
3.8 Three families of fermions . . . . . . . . . . . . . . . . . . 26
v
vi Table of Contents
4 Dirac equation 31
4.1 Covariant formalism . . . . . . . . . . . . . . . . . . . . . 31
4.2 Gamma matrices . . . . . . . . . . . . . . . . . . . . . . . 32
4.3 Bilinear products of spinors . . . . . . . . . . . . . . . . . . 33
4.4 Free fermions . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.5 Lagrangians and equations of motion . . . . . . . . . . . . . 34
4.6 Conservation of fermion current . . . . . . . . . . . . . . . 35
4.7 Isospin algebra and conservation . . . . . . . . . . . . . . . 35
4.8 Nucleon as quark atom . . . . . . . . . . . . . . . . . . . . 37
5 Interactions 39
5.1 Three interactions of particle physics . . . . . . . . . . . . . 39
5.2 Electromagnetic interaction . . . . . . . . . . . . . . . . . . 40
5.2.1 Local U(1) invariance . . . . . . . . . . . . . . . . 40
5.2.2 Quantum electrodynamics (QED) . . . . . . . . . . 42
5.2.3 Current-current interaction . . . . . . . . . . . . . . 42
5.2.4 Photon . . . . . . . . . . . . . . . . . . . . . . . . 43
5.3 Mandelstam variables . . . . . . . . . . . . . . . . . . . . . 44
5.4 Strong interaction . . . . . . . . . . . . . . . . . . . . . . . 45
5.4.1 Colour charges . . . . . . . . . . . . . . . . . . . . 45
5.4.2 Nuclear forces . . . . . . . . . . . . . . . . . . . . 47
5.4.3 Local SU(3) invariance . . . . . . . . . . . . . . . . 48
5.4.4 Running coupling . . . . . . . . . . . . . . . . . . . 48
5.4.5 Gluons . . . . . . . . . . . . . . . . . . . . . . . . 50
5.5 Electroweak interaction . . . . . . . . . . . . . . . . . . . . 50
5.5.1 Spontaneous symmetry breaking . . . . . . . . . . . 51
5.5.2 BEH mechanism . . . . . . . . . . . . . . . . . . . 52
5.6 Basic bosons . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.7 Electroweak Lagrangian with interactions . . . . . . . . . . 54
II Experimental methodology 57
6 Accelerators 59
6.1 Magnets: bending and focusing . . . . . . . . . . . . . . . . 60
6.2 Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.3 Colliders . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.4 Flux and luminosity . . . . . . . . . . . . . . . . . . . . . . 62
6.5 Beam cooling . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.6 CERN’s accelerator complex in the LEP era . . . . . . . . . 63
Introduction to Particle Physics vii
7 Detectors, calorimetry 73
7.1 Event registration . . . . . . . . . . . . . . . . . . . . . . . 73
7.2 Energy loss in matter . . . . . . . . . . . . . . . . . . . . . 74
7.3 Particle identification . . . . . . . . . . . . . . . . . . . . . 77
7.4 Detector types . . . . . . . . . . . . . . . . . . . . . . . . . 78
7.4.1 Multiwire chambers . . . . . . . . . . . . . . . . . 79
7.4.2 Scintillation counters . . . . . . . . . . . . . . . . . 79
7.4.3 Shower detectors . . . . . . . . . . . . . . . . . . . 80
7.4.4 Cherenkov detectors . . . . . . . . . . . . . . . . . 80
7.4.5 Transition radiation detectors . . . . . . . . . . . . . 81
7.5 The CMS detector . . . . . . . . . . . . . . . . . . . . . . . 82
8 Event registration 85
8.1 LEP events . . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.2 Transverse momentum, pseudorapidity . . . . . . . . . . . . 87
8.3 Observation of the top quark . . . . . . . . . . . . . . . . . 89
8.4 Mysterious events . . . . . . . . . . . . . . . . . . . . . . . 89
9 Data analysis 91
9.1 Statistical concepts of particle physicists . . . . . . . . . . . 92
9.2 Basic concepts of statistical analysis . . . . . . . . . . . . . 93
9.3 Fitting parameters . . . . . . . . . . . . . . . . . . . . . . . 95
9.3.1 Goodness of fit . . . . . . . . . . . . . . . . . . . . 95
9.3.2 Confidence level . . . . . . . . . . . . . . . . . . . 95
9.4 Estimating (fitting) parameters . . . . . . . . . . . . . . . . 96
9.4.1 Arithmetic mean and standard deviation . . . . . . . 96
9.4.2 Linear fitting . . . . . . . . . . . . . . . . . . . . . 97
viii Table of Contents
13 Neutrinos 133
13.1 Weak currents . . . . . . . . . . . . . . . . . . . . . . . . . 133
13.2 Neutrino mass . . . . . . . . . . . . . . . . . . . . . . . . . 135
13.3 Early neutrino mysteries . . . . . . . . . . . . . . . . . . . 135
13.4 Neutrino oscillation . . . . . . . . . . . . . . . . . . . . . . 137
Introduction to Particle Physics ix
Intermezzo 209
23 Outlook 351
Bibliography 355
Index 36
xii Foreword
Foreword
One of the methods of studying Nature is to penetrate deeper and deeper in
the structure of matter ever increasing the spatial resolution, i.e., studying
smaller and smaller objects. In the history of natural sciences new and new
particles appeared which were thought to be elementary: the four atoms
(a-tom = not divisible) of Anaximenes and Democritus, the elements/atoms
of Dalton and Mendeleev, the atomic nucleus of Rutherford and the so-called
elementary particles of which the proton, the neutron, the electron and the
neutrino are the most well-known. Between 1930 and 1960 hundreds of such
particles were discovered, thus a new level of elementariness was needed
and the quark model appeared. We will see that, in fact, the proton and the
neutron are also composite particles although the electron stays elementary.
This development was crowned by the standard model (SM) in the late
sixties and it is still the uncontested global theory of matter, supported by all
available theoretical and experimental evidence.
In this textbook we summarize the present knowledge of particle physics
at an introductory level. Particle physics is a very broad subject including
many different sub-fields. While we mention many of these, a detailed
account on all is impossible. Our clear focus is on high energy collider
physics that is among the most widely pursued subfields where the threshold
of current research is high. With the Large Hadron Collider in operation new
results appear regularly. Our goal here is to keep the level introductory, yet
help students reach this high threshold making them acquainted with both
the experimental and theoretical minimum needed to comprehend current
research at colliders. Our treatise is detailed on established results of collider
physics while mostly marginal on current developments with the exception
of the discovery of the Higgs particle due to its utmost importance.
The first part (written by D. Horváth) is planned to be accessible for
advanced BSc or freshmen MSc students, while the second part (written
by Z. Trócsányi) on the theory is intended for advanced MSc or freshmen
PhD students in particle physics, with some attempt to go into the rather
complex mathematical formalism of the field. Our aim is to provide concise
but hopefully comprehensive account on the subject and also try to help
students in their decision whether to orient themselves towards experiment
or theory. We assume that the book can be covered during a full academic
year with about 10 hours of serious effort per week. Although the reader
may be confused on several occasions when not all details are given, the
theory is very precisely elaborated and its predictions beautifully agree with
the experimental observations. All present day experimental evidence is
Introduction to Particle Physics xiii
Particle phenomenology
1
Chapter 1
MOTTO:
Central to that theory is the concept of sponta-
neously broken gauge symmetry. According
to this concept, the fundamental equations of
physics have more symmetry than the actual
physical world does.
(Frank Wilczek)
x = a + b = x cos Θ + y sin Θ ,
y = y − c = y cos Θ − x sin Θ .
coordinate transformation:
x cos Θ · x + sin Θ · y cos Θ sin Θ x
= = · .
y − sin Θ · x + cos Θ · y − sin Θ cos Θ y
x x
This means that the vector is obtained by multiplying the vector
y y
with the matrix in front of it. An important property of these rotation
transformations is that they do not change the length of the vector pointing
to P (its absolute value) as
x 2 + y 2 = (x 2 + y 2 ) · cos2 Θ + sin2 Θ = x 2 + y 2 .
The condition that the length of the vector remains unchanged demands
that the complex transformation matrix be unitary:
∗ ∗
† U11 U21 U11 U12
U U= ∗ ∗ · =
U12 U22 U21 U22
2 + U2 ∗ U + U∗ U
U11 U11 12 21 22 1 0
∗ U + U∗ U
21
2 + U2 =
U12 11 22 21 U12 22
0 1
Rotations of this type have the following mathematical properties:
Sets with operation among its elements obeying these properties are
called groups. Spin is a three-dimensional quantity with the properties of
the rotation group and its mathematical description (representation) is called
the SU(2) group of special (determinant = 1) unitary complex 2 × 2 matrices.
SU(2) can be applied not only for spin, but for any physical quantity with
similar symmetry properties, like for example the isospin to be introduced
later.
When we increase the degrees of freedom we get higher symmetry groups
of similar properties. The next step, SU(3), which is also used in particle
physics, is the symmetry group of special unitary complex 3 × 3 matrices.
It has three possible eigenstates which can be interpreted as three corners
of a triangle with an SU(2) symmetry between any two of its corners (see
Chapter 3).
In case of complex quantities we can also decrease the degrees of freedom
of rotations, then we get the U(1) group of 1 × 1 unitary matrices, i.e. eiφ
complex phases. That is the symmetry group of the gauge transformations
of the electromagnetic interaction. The simplest manifestation of the gauge
symmetry of electromagnetism is that we can freely choose the zero point
of the electrostatic potential as demonstrated by the birds sitting on high-
voltage wires. The global U(1) symmetry of Maxwell’s equations leads
to the conservation of the electric charge. In the more general case, the
U(1) symmetry of the Dirac equation [Dirac, 1931], the general equation
describing the motion of a fermion, causes the conservation of the number
of fermions or fermion charge [Halzen and Martin, 1984].
Condensation no yes
possible asymmetry between particle and its antiparticle, and that is one of
the great mysteries of physics. Were there antimatter galaxies, they would
emit antiparticles and they would be encircled by a halo of annihilation when
meeting the particles emitted by neighbouring galaxies of ordinary matter,
but the astronomers do not see such phenomena anywhere.
Figure 1.2: Mirror reflection and parity change (after D. Kirkby, APS, 2003)
from right-handed to left-handed coordinates.
anti quarks, their ground state ( = 0) parity is P(qq) = P(q) · P(q) = −1.
By definition for the nucleons Pp = Pn = +1, thus the parities of quarks are
+1, and of anti-quarks –1. For the particles spin J and parity P are denoted
as J P , e.g. for charged pions π ± : 0− .
We shall see later that parity is not conserved by the weak interaction, it
is a broken symmetry. Stephen Weinberg calls such symmetries accidental
symmetries.
the electric current we call this particle current; in the above example the
incoming electron and positron constitute a lepton current.
In the simplest case of particle collision two such particle currents ex-
change a boson. This is made possible by the uncertainty principle of
Heisenberg, as it allows a violation of energy and momentum conservation
for very small time and space intervals: ΔE · Δt ≥ /2 and ΔpΔx ≥ /2,
where Δ indicates a very small change in the quantity behind it and E, p, t, x
the energy, momentum, time and space position. The very small value of the
reduced Planck constant ( 1.055 · 10−34 J·s) ensures that the conservation
laws are fulfilled in the macro-world. The boson mediating the interaction
can be real or virtual depending on whether or not it satisfies the on-shell
condition E 2 = m2 c4 + p2 c2 . Effects of virtual particles can be detected
experimentally: in the inelastic scattering of high energy electrons on each
other quark pairs could be produced when a virtual photon emitted by one
of the electrons is absorbed by a quark of a virtual quark-antiquark pair
produced momentarily by another photon emitted by the other electron.
CPT invariance is supported by ample experimental evidence. Its role
is so important in quantum field theory that according to some theorists it
is impossible to test experimentally: in the case of observing a small de-
viation one should rather suspect the violation of a conservation law than
CPT violation. In spite of this, there are considerable efforts to test it exper-
imentally. The most precise of those tests is the very small possible relative
mass difference between neutral kaon and anti-kaon which is less than 10−18 .
The European Particle Physics Laboratory, CERN has built the Antiproton
Decelerator facility in 1999 with the aim to test CPT invariance using pre-
cision spectroscopy of antihydrogen, the bound state of an antiproton and a
positron and also that of anti-protonic atoms where an electron is replaced
by an antiproton. The latter measures the mass and charge of the antiproton
(antimatter physics).
If CPT invariance is indeed a fundamental symmetry of nature, then
the violation of time reversal is equivalent to the violation of the combined
CP symmetry. We shall see later the weak interaction breaks not only P
(maximally), but also CP (a little). As a result, time invariance is also
violated by the weak interaction, in contrast to classical mechanics.
these have almost the same mass and apart from a charge effect the strong
interaction within the atomic nucleus affects them identically. Heisenberg
introduced the concept of the nucleon which has two eigenstates, the proton
and the neutron. This needed a new quantum number characterizing it; as
its symmetry properties are identical to those of the spin he called it the
isospin I from isotopic spin (isobaric would be more precise as isotopes
have different numbers of nucleons whereas isobaric nuclei have the same
numbers of nucleons). The nucleon has an isospin I = 12 , the proton is the
nucleon state with I3 = + 12 and the neutron3 is that of I3 = − 12 .
With the development of experimental methods many strongly interacting
particles, hadrons were observed and all had characteristic isospins, i.e. all
could be arranged in groups of particles of similar properties but different
charges according to their isospins. The nucleon has an isospin I = 12 and
two similar states, with I3 = ± 12 . The lightest hadron, the π-meson or pion
has I = 1 with three eigenstates (I3 = -1, 0 and +1) and three charge states
π + , π 0 and π − . The Δ hyperon has I = 32 :
3 1 1 3
Δ− (I3 = − ) , Δ0 (I3 = − ) , Δ+ (I3 = + ) , Δ++ (I3 = + ) .
2 2 2 2
A unit change of I3 involves a corresponding unit change in charge.
Then a third quantum number, strangeness S was discovered. Pairs of
particles were produced in collisions of energetic protons with probabilities
characteristic of the strong interaction and lived long enough that they must
have decayed via weak interactions. They were called V-particles as their
tracks curved in the magnetic field of the detectors in opposite directions.
To explain this Murray Gell-Mann, Abraham Pais and Kazuhiko Nishijima
introduced strangeness S as a new additive quantum number which is con-
served in strong interactions but not in weak reactions. For instance, the Σ−
hyperon (S = −1, I = 1) created in π − p→ K+ Σ− decays via Σ− → nπ − with
a lifetime of τ ∼ 10−10 s whereas the Δ+ hyperon (S = 0) decays in Δ+ →
nπ + with a lifetime of τ ∼ 10−23 s. It was postulated that only the weak
interaction can change the new quantum number.
Strangeness and isospin made together an SU(3) group which made it
possible to construct a unique frame for all known particles. In order to
explain this, Murray Gell-Mann and George Zweig suggested the quark
model of hadrons. Using three new elementary fermions, three quarks
(Table 1.2), all observed hadrons could be described. Isospin became the
quantum number of the two lightest quarks and because of the analogy to the
spin the I3 = + 12 state was named up quark with the sign u and the I3 = − 12
state down quark, d. The third quark’s quantum number is the strangeness,
so that is the strange (s) quark. The isospin and strangeness, characterizing
the various kinds of quarks are called flavour quantum numbers.
The quark model postulates that quarks can bind together only in two
ways: in quark-antiquark pairs, those are called mesons, and three-quark
states, those are the baryons. As the quarks have spin 12 , naively we expect
that mesons are bosons and baryons are fermions. Quarks have baryon
number 13 and fractional electric charges: in units of elementary charge e
the u quark has charge + 23 while the d and s quarks − 13 . This of course gives
the proper charges to the proton p = [uud] and the neutron n = [udd] or the
pions: π + = [ud], π 0 = √1 [uu + dd], π − = [ud]. Thus the third component
2
of the isospin is directly connected to the charge, as its unit increase means
replacing a d by a u quark, i.e. increasing the total charge by + 23 − (− 13 ) = 1.
quark J eq B I3 S Y = B+S
u 1
2 + 23 1
3
1
2 0 + 13
d 1
2 − 13 1
3 − 12 0 + 13
s 1
2 − 13 1
3 0 −1 − 23
Exercise 1.1
What invariance principles are violated by the weak, electromagnetic and
colour interactions?
Exercise 1.2
What are the analogies and differences between spin and isospin?
Exercise 1.3
How can the three-dimensional spin be characterised by two independent
quantities?
Exercise 1.4
What gauge symmetry facilitates the conservation of electric charge and
fermion number?
Chapter 2
What is measured in
experiment?
MOTTO:
I was brought up to look at the atom as a nice
hard fellow, red or grey in colour, according
to taste.
(Ernest Rutherford)
2π
W(i→ f ) = |Mi f | 2 ρ f
W 1 (2Sc + 1)(2Sd + 1) 2
σ= = |Mi f | 2 pf
na vi π4 vi v f
dp f Ec E d 1
= ≈ .
dE0 p f E0 vf
What is measured in experiment? 17
2.2 Resonance
A simple, structureless, unstable system will have an exponential decay law
|ψ(t)| 2 = |ψ(0)| 2 e−Γt (2.1)
with a lifetime τ = Γ−1 where Γ is the decay rate. The state function of the
system will have a time evolution ψ(t) ∼ ψ(0)e−iMt e−Γt/2 . Here we use the
natural units of particle physics . In SI units τ = /Γ 6.582 · 10−22 /Γ s
where the Γ energy is measured in MeV. The time development of the system
(assuming a plain wave initial state) is ψ(t) = ψ(0)e−iMt e−Γt/2 , and its energy
amplitude will be the Fourier transform of the time amplitude:
∫ ∞ ∫ ∞
1
χ(E) = ψ(t)eiEt dt = ei(E−M)t e−Γt/2 dt =
0 0 i(E − M) − Γ/2
The result of the Fourier transformation is a probability distribution,
which appears in the measurement as a Lorentz curve:
1
| χ(E)| 2 = (2.2)
(E − M)2 + Γ2 /4
This is the Breit-Wigner resonance (Fig. 17.2, left): an energy distribution
with a maximum at the M mass of the decaying particle and a Γ width
corresponding to its decay probability. New particles can be discovered by
observing the Breit-Wigner resonances of their decays, those of course have
to be at the same (so-called invariant) mass in the different decay channels.
One of the most important resonances in high energy physics is the Z peak
in electron-positron collisions (Fig. 17.2). Two colliding beam accelerators,
LEP at CERN and SLC at Stanford were especially built to study it, and they
provided an incredible wealth of information for testing and confirming the
standard model, the theory of particle physics.
We have shown above how the lifetime of a decaying particle is connected
to its energy spectrum. In low-energy physics exponential decay curves of
the type (2.1) can be directly measured by measuring the time difference
between the birth and decay of a system. In high energy physics the reactions
are usually much too fast for that, so the lifetimes have to be estimated by
measuring the width of the mass resonance of the decaying particle.
Exercise 2.1
Why can we use cross sections to characterize interactions in particle scat-
tering?
18 Chapter 2
Exercise 2.2
How can we discover a new particle by measuring energy in particle scat-
tering?
Exercise 2.3
How can one estimate the lifetime of a decaying particle by measuring the
energy of its decay products?
Chapter 3
Quark model
MOTTO:
Just because things get a little dingy at the
subatomic level doesn’t mean all bets are off.
(Murray Gell-Mann)
Figure 3.1:
Colour-SU(3) with its three eigenstates and SU(2)-type stepping operators
The quarks have colour charges whereas anti-quarks have anti-colour ones.
Colour solved all those problems: the new quantum number made each
quark state in the observed particles specific and so the Pauli principle was
fulfilled. Moreover, colour is considered to play the same role for the strong
quark–quark interaction as the electric charge in electrodynamics: it is the
strong charge. The existence of all observed particles is consistent with the
condition that only colourless (white) states are allowed in Nature. This
condition accounts for the non-existence of free quarks (quark confinement),
as only the combinations of colour+anti-colour (those bound states are the
mesons) or that of all three colours (baryons) can exist. This is an algebraic
explanation for quark confinement. The dynamical understanding is far more
complex.
Thus there is a good analogy between the strong colour charge of quarks
and colour vision in humans. The three eigenstates of the quark charges
correspond to the three basic colours and strong anti-charge is the analogy
of the complementary colour. The equilibrium mixture of the three charges
or the mixture of charge + anti-charge will be neutral, colourless in colour
vision.
In order to get free particles we have to construct colourless states.
According to the experimental observations colour makes the state functions
of the composite bosons, the mesons symmetric and of the baryons, as
required for fermions, antisymmetric. For the mesons it is of the form
√1 (RR + GG + BB) and for the baryons √1 (RGB − RBG + BRG − BGR +
3 6
GBR − GRB). Anti-baryons have the colour arrangement √1 (RGB − RBG +
6
Quark model 21
BRG − BGR + GBR − GRB). Thus when building the state function of a
mesonic or baryonic system, we just have to construct a symmetric (space
⊗ spin ⊗ flavour) function and let the colour part make it symmetric or
asymmetric as needed.
The quark model was very successful: it explained all observed hadrons
and correctly predicted the existence of new hadronic states. There are no
exceptions, like e.g., states with charges Q > 2 or isospin I > 3/2.
Triplet states:
|S = 1, M = +1 >= ↑↑ ⎫
⎪
⎬
⎪
|S = 1, M = 0 >= √1 (↑↓ + ↓↑) (3.1)
2 ⎪
⎪
|S = 1, M = −1 >= ↓↓ ⎭
Singlet state:
1
|S = 0, M = 0 >= √ (↑↓ − ↓↑) (3.2)
2
The conserved quantities and eigenvalues of angular momentum are:
J 2 |S, M >= S(S + 1)|S, M > and J3 |S, M >= M |S, M >.
The projections are increased and decreased with the stepping operators:
J± = J1 ± iJ2 such that J± | j, m >= j( j + 1) − m(m ± 1)| j, m ± 1 >.
2 (uu − dd) = π 0
1
1 0 0 135.0
2 (uu + dd) = η0
1
0 0 0 547.9
Table 3.1: The lightest mesons: the three pions, and the η0 .
Figure 3.2: Construction of the meson nonet of the three light quarks (3 ⊗ 3 =
8 ⊕ 1). A, B and C are the three possible symmetric combinations of uu, dd
and ss.
The quark flavour states of the meson nonet have to be combined with
spin states. If the spins of the quarks are anti-parallel they will make J P = 0−
pseudo-scalar mesons, parallel spins make J P = 1− vector mesons. Orbital
momentum L between the quarks will change the state parity P = −(−1) L .
The lightest, ground-state, neutral pseudo-scalar mesons are listed in
Table 3.2. States B and C have the same quantum numbers (I = 0), they will
mix and create two mass eigenstates: η1 , η8 ⇒ η, η and the latter can be
observed in experiment:
η = η8 cos Θ p − η1 sin Θ p
η = η8 sin Θ p + η1 cos Θ p
where Θ p ≈ 10◦ is the mixing angle. In particle physics all states that may
mix happen to mix with mixing angles to be determined experimentally. We
shall see more such cases later.
As the SU(3) group corresponds to 3 SU(2) subgroups, in the scheme
drawn in Fig. 3.2 in addition to the usual SU(2) isospin connected to the u↔d
exchange one can define similar symmetries for u↔s and d↔s; historically,
the u↔d SU(2) was called I−spin, the d↔s SU(2) U−spin and the u↔s
SU(2) V−spin, but these obsolete quantities are not used any more.
24 Chapter 3
Figure 3.3: Construction of the lightest hadrons using the first four quarks. Left:
pseudo-scalar (a) and vector (b) mesons ; right: baryons of octet (a) and decuplet (b)
structure [Beringer et al., 2012]). Note the nucleons at the back of the baryon octet.
Adding the fourth quark, charm to the set we get an SU(4) scheme.
In the left panel of Fig. 3.3 each horizontal plane corresponds to the Y
vs. I3 SU(3) plane of Fig. 3.2, the vertical axis shows the charm quark
content. In addition to the quark compositions also the historic names1 of
the mesons are noted. The u+d SU(2) is quite good symmetry as those
quarks have very little mass, of the order of 5 MeV. Flavour SU(3) for u+d+s
is already broken as the s quark is much heavier, around 100 MeV. The c
quark is even heavier, Mc 1275 MeV, so flavour-SU(4) cannot really be
used for quantitative predictions. We have to point out here that the mass
of our macroscopic world is predominantly due to the energy content of the
nucleons, the masses of the u and d quarks contribute very little.
There are quite a few meson nonets observed (Table 3.3). Generally,
when we speak of a particle, we always write down its mass to make the
situation unambiguous. For instance, the ground-state vector mesons (J P =
1− ) are ρ(770), K∗ (892), ω(782) and φ(1020) (in Table 3.3 average mass
values are listed).
1Enrico Fermi: Young man, if I could remember the names of these particles I
would have been a botanist.
Quark model 25
Table 3.3: Some of the observed meson nonets. ω ≈ √1 (uu − dd) and φ ≈ ss
2
flavour−SU(3) ⊗ spin−SU(2)
(qqq) 1
2 + 1
2 + 1
2 (L = 0)
3⊗3⊗3 2⊗2⊗2
(10 ⊕ 8 ⊕ 8 ⊕ 1) ⊗ (4 ⊕ 2 ⊕ 2)
TS MS MA TA TS MS MA
ν ν ν 0 + 12
e μ τ
Leptons
e L μ L τ L −1 − 12
u c t + 23 + 12
Quarks
d L s L b L − 13 − 12
Table 3.5: Leptons and quarks, the three families of basic fermions. T3 is
the third component of the weak isospin, the rest of the notation is explained
in the text step by step.
families decay quickly to lighter ones. For the quarks the eigenstates of
weak and strong interactions are different. The bound states of quarks, the
hadrons are produced in strong interaction. These particles are identified by
their masses, so if we speak of a given quark, we mean their mass eigen-
states. However, the weak isospin doublets should contain weak eigenstates,
mixtures of mass eigenstates, symbolized by the prime behind their symbols
in Table 3.5. It was shown by Makoto Kobayashi and Toshihide Maskawa
(discovery in 1973, Nobel Prize in 2008) that in order to explain CP-violation
(see Section 12.4) there have to be at least 3 families of quarks. It is sufficient
to mix one row of quarks, the lower ones are chosen.
According to our present knowledge Table 3.5 contains all of the basic
fermions of the standard model. The reader should not be frightened by the
extraordinary caution of the above statement. Theoretically, the standard
model was extended in many-many ways and all those extensions predicted
numerous hypothetical new basic particles [Collins et al., 1989]. Although
till now no real evidence was found against the standard model, there are
many possible extensions that do not contradict to the present experimental
data, and so one cannot a priori exclude their validity.
Exercise 3.1
How many different quarks are in the standard model? Explain in what
sense 3, 6, 12, 18 and 36 could be all correct answers.
Exercise 3.2
How could electromagnetic interaction explain the fractional charges of the
quarks?
Exercise 3.3
How can the bound state of 3 spin- 12 quarks have a spin 92 ?
Exercise 3.4
Is there any direct evidence for the 3 quark colours?
Exercise 3.5
Quark confinement is explained by gluon exchange. How could the 983 MeV
proton mass created by three quarks of 5–15 MeV masses?
Exercise 3.6
The existence of gluons is shown by detecting 3-jet events in lepton collisions.
Why must the 3 jets be in the same plane?
Quark model 29
Exercise 3.7
Compare and explain the signs of the terms in Eqs. (3.1, 3.2) and Table 3.1.
Hint: consider symmetries.
Exercise 3.8
How can the quark model explain the non-zero magnetic moment of the
neutron?
Chapter 4
Dirac equation
MOTTO:
When I was a young man, Dirac was my hero.
He made a breakthrough, a new method of
doing physics. He had the courage to simply
guess at the form of an equation, the equation
we now call the Dirac equation, and to try to
interpret it afterwards.
(Richard P. Feynman)
Table 4.1: Bilinear products of spinors, component numbers and the effect
of coordinate reflection.
to Lagrangian (4.2) one immediately obtains the Dirac equation for the
adjoint spinor:
i∂μ ψγ μ + mψ = 0 .
Thus the Dirac equation of the free fermion in its final form is
|I = 1, I3 = +1 >= pp ⎫
⎪
⎬
⎪
|I = 1, I3 = 0 >= √1 (pn + np) triplet states
2 ⎪
⎪
|I = 1, I3 = −1 >= nn ⎭
|I = 0, I3 = 0 >= √1 (pn − np) singlet state
2
Here I, I3 and I , I3 are the whole isospin and its 3rd component in the
initial and final states. Conservation laws dictate that I = I; I3 = I3 . We
know the measured values Iπ = 1 and Id = 0. The first reaction:
3
μp = μq = < p↑| μi σ3 |p↑ >
i=1
38 Chapter 4
1 0 1 0
where σ3 = , u= , d= .
0 −1 0 1
1 1
μp = [(μu − μu + μd ) + (−μu + μu + μd ) + 4(2μu − μd )] × 3 = (4μu − μd ) .
18 3
By switching u and d in the formula we get the magnetic moment of the
neutron, neutron = proton (u ↔d): μn = 13 (4μd − μu ) . Their ratio is
μn 4 − μu /μd 2
= − .
μp 4μu /μd − 1 3
Exercise 4.1
Considering the Dirac equation in what sense can we speak of anti-bosons?
Exercise 4.2
What property of the Dirac equation leads to the conservation of the lepton
and baryon numbers?
Exercise 4.3
2 †
Prove Eq. (4.4) using γ 0 = I, γ 0 = γ 0 and γ μ† = γ 0 γ μ γ 0 ).
Exercise 4.4
Explain the signs of the terms when summing up the isospins in Table 4.2.
Hint: consider symmetries.
Exercise 4.5
Why is it a rough approximation to consider the three quarks only for calcu-
lating the magnetic moment of the nucleon?
Chapter 5
Interactions
MOTTO:
When you’ve exhausted all possibilities, re-
member that you haven’t.
(Robert H. Schuller)
two quarks exchange colour so the mediating boson, the gluon (from glue)
should carry a colour and an anti-colour. This means 8 different gluons: the
3 × 3 = 9 combinations have one less degree of freedom as the combination
RR + GG + BB changes white to white and so does not carry any colour.
Pion decay provides an excellent example to compare the strengths of
the electromagnetic and weak interactions. The decay of the neutral pion to
two photons, π 0 →γγ, is a typical electromagnetic reaction with a lifetime
of 8 · 10−17 s. The charged pion can decay via weak interaction only to a
muon and its neutrino: π − →μ− + ν μ and its lifetime is 26 ns = 2.6 · 10−8 s, 8
orders of magnitude longer than that of its neutral brother. Please note that
in the above reaction a boson disappeared and a lepton was created together
with an anti-lepton: the fermion number is conserved, whereas the boson
number is not.
∂μ ∂ μ φ − ie(∂μ Aμ + Aμ ∂ μ )φ − e2 Aμ Aμ φ = ∂μ ∂ μ φ + V φ
where V = −ie(∂μ Aμ + Aμ ∂ μ ) − e2 Aμ Aμ is the scalar potential of the inter-
action derived from the Aμ vector potential.
The transition probability from an i initial to an f final state is propor-
tional to the overlap of their state functions in space-time,
∫ the matrix element,
when embracing the V potential: M f i = −i φ∗f V φi d4 x. As the coupling
α ≡ 4π
2
e 1 2 μ
137 is small, the e A Aμ term can be neglected in first approxima-
tion (first order perturbation calculation), and then V ≈ −ie(∂μ Aμ + Aμ ∂ μ ).
Assuming that the potential disappears at infinity, i.e. in the initial and final
states, ∫
∂μ (φ∗f Aμ φi )d4 x = [φ∗f Aμ φi ]+∞
(0,−∞) = 0,
pC pD
1 q 2
t
pA pB
where jμ = −ie[φ∗f (∂μ φi )−(∂μ φ∗f )φi ] is the electromagnetic current density.
fi
5.2.4 Photon
The scattering of point charges A + B → C + D (AB) involves a momentum
exchange q ≡ pD − pB = pA − pC and the matrix element ∫ (1) μdetermining
the reaction probability will have the form M (1)
fi = −i jμ A(2) d4 x. The
μ μ μ
potential is the solution of D’Alembert’s equation: 2 A(2) = j(2) with j(2) =
μ μ
−eNB ND (pD + pB )μ e−iqx , giving A(2) = − q12 j(2) . Thus we have for the
transition matrix element
∫
1 μ
M f i = −i jμ(1) (− 2 ) j(2) d4 x
q
g
and in the given case −iM = [ie(p A + pC )μ ](−i qμν2 )[ie(pB + pD )ν ] where
gμν is the metric tensor (4.1).
Thus the interaction is described in a symmetric current × current form.
It is mediated by the photon which is obviously virtual as it transfers energy
and momentum and so it has a finite mass, q2 > 0. That, however, is allowed
by Heisenberg’s uncertainty principle. The photon appears between the two
currents as 1/q2 : that corresponds to the photon propagator. It has finite
mass and lifetime and is denoted by a wavy line in the Feynman diagrams
(Fig. 5.1).
44 Chapter 5
s = (p A + pB )2 ,
t = (p A − pC )2 ,
u = (p A − pD )2 .
The use of these variables makes the handling of the Feynman diagrams
extremely simple as it facilitates the interrelation of different reactions via
diagram rotations, i.e. space-time reflections, called crossing. The scattering
reaction of Fig. 5.1 the original AB→CD reaction is called s channel, a
(s↔ − t) rotation gives the DB→C A t channel, and (s↔ − u) the u channel.
There are many sum rules for the Mandelstam variables, some of them are
tested in the Exercises.
As the simplest possible case, let us consider electron-positron collisions.
The momentum is p = (E, k) and its square is p2 ≡ m2 = E 2 − k 2 where
m is the electron mass. In the centre-of-mass system (like at the LEP
collider, CERN, 1989-2000) p A = (E, ki ), pB = (E, − ki ), pC = (E, k f ), and
pD = (E, − k f ). Then s = ECM 2 (E
CM = 2E is the total collision energy),
t = −2 k 2 (1 − cos θ) and u = −2 k 2 (1 + cos θ), where θ is the scattering angle
between the incoming and outgoing particles.
Without going into details, let us list a few examples for diagram rotation,
or crossing. The square of invariant amplitudes averaged for initial and
summed over final state spins for the e− e− →e− e− Møller scattering is
e4 s2 + u2 s2 + t 2 2s2
|M| 2 = + +
2 t2 u2 tu
where the terms describe the forward and the backward scattering, and
their interference. This interference term has a negative sign from direct
computation, but there is a subtle point in computing the square of the
amplitude: as there are identical fermions in the final state and under particle
interchange fermionic systems are odd, we have to square the difference of
the two contributions shown in the left panel of Fig. 5.2, which turns the sign
of the interference term positive. If we want to obtain the squared amplitude
for the e+ e− → e+ e− Bhabha scattering via an s↔ − u crossing, we should
use the sum of the two diagrams, i.e. the Møller formula with minus sign for
the interference term as this time the fermions are different in the final state.
Interactions 45
Thus, we find
e4 s2 + u2 u2 + t 2 2u2
|M| 2 = + + .
2 t2 s2 ts
Another example is the connection between the γe→ γe Compton scat-
tering and positron annihilation. From the invariant amplitude of the former,
e4 u s
|M | 2 = − −
2 s u
(there is no interference term in this case), one obtains that of annihilation
(shown in Fig. 5.3):
e4 u t
|M | 2 = +
2 t u
by an s↔ − t crossing.
Another interesting case is electron scattering on muon, described with
a photon exchange (shown in Fig. 5.4 left). Its invariant amplitude is
e4 s 2 + u2
|M| 2 = ,
2 t2
which gives after an s↔ −t crossing the amplitude for the creation of a muon
pair in an electron-positron annihilation. At high energies the lepton masses
can be neglected as compared to their energies and we obtain
e4 u2 + t 2 e4 8k 4 (1 + cos2 θ) e4
|M | 2 = = = (1 + cos2 θ)
2 s2 2 (4k 2 )2 4
for the invariant squared amplitude of the process e− e+ →μ− μ+ . Its differ-
α2
dΩ CM = 4s (1 + cos θ)
ential cross section in the centre-of-mass system is dσ 2
where α is the fine structure constant. Integrating over the polar and azimuth
angles, we find its total cross section σ(e+ e− →μ+ μ− ) = 4πα
2
3s in very good
agreement with the experimental result. √
In experimental papers very frequently s is used to denote the centre-
of-mass energy of particle collisions.
e− e− e− e−
e− e−
A C A D
e− e−
+
e+ e+
B D B C
e− e− e− e− +
e e+
Figure 5.2: From Møller scattering on the left we get via s↔ − u crossing
(diagram rotation) Bhabha scattering
e−
γ
γ γ e−
p k p
γ γ k k’ k
γ
k k’ p−k’
p+k
p+k
γ
p−k’ k’
e+ p’
e− p p’ e− e− p p’ e− p’ k’ γ e+
e− e−
e− μ−
μ− μ− e+ μ+
(this time the summation over the repeating indices is over the 8 generators).
Here, similarly to the QED case, αa (x) are real space-time functions. The
SU(3) gauge field is introduced via the covariant derivation
Dμ = ∂μ + igs Ta G μa
with a coupling gs . Note that the first two terms are exact analogies of the
QED field. Similarly to QED the QCD Lagrangian is
1 a μν
LQCD = q j (iγ μ ∂μ − m)q j − gs (q j γ μ Ta q j )G μa − G μν Ga .
4
The QCD Lagrangian in its expanded form (Fig. 17.1), in addition to a linear
term similar to that in the QED case, has terms of higher order interactions.
QED QCD
Elementary fermions leptons quarks
Charge electric colour
Gauge boson photon (γ) 8 gluons (g)
no charge have colours
Coupling α(Q2 = 0) = 137
1
αs (Q2 = m2Z ) = 0.12
Q2 dependence weak strong
Free particles leptons hadrons
Calculation precision < 10−8 5 − 20%
the two particles can have parallel or opposite spins; the former state, with
spin S = 1 called ortho-positronium decays by emitting three photons be-
cause of parity conservation and has a lifetime of 140 ns, whereas the other
one, the S = 0 para-positronium, decays to two photons with a lifetime of
125 ps. A high-energy hadron jet contains dozens of particles and so cannot
be produced by other interactions, just hadronization.
5.4.5 Gluons
Gluons are produced most frequently in hadron-hadron collisions. In the
proton-proton collisions of the Large Hadron Collider that is the most fre-
quent phenomenon. Higgs bosons are produced most frequently in gluon-
gluon collisions (gluon fusion). Gluon production is manifested best in 3-jet
events (Fig. 17.25); at the production of quark pairs in electron-positron
collisions, the quarks can interact with each other emitting a gluon and we
see an event of three jets in a common plane. The third jet must be from a
boson because of the conservation of fermion charge and that boson must be
coloured to be able to form a third jet.
The properties of gluons were established via studying such 3-jet events
in electron-positron collisions where two of the jets were identified as b
quarks (due to its longer lifetime and increased lepton emission), so that the
third jet had to be a gluon. The unit spin and zero mass of the gluons were
this way confirmed experimentally.
n→p + W− →p + e− + ν e ,
• One had to introduce the masses of elementary particles, those for the
weak bosons and the elementary fermion, the quarks and leptons. As
gauge invariance prohibited that, it had to be somehow violated.
where (λ > 0) is a real constant. If μ2 > 0 this is a scalar field with a finite
mass; however, if μ2 < 0 its stable vacuum state is not at φ = 0, but at a
different vacuum expectation value.
The Mexican hat illustrates the BEH potential in two dimensions: it has
a perfect axial symmetry (Fig. 5.5) and that is not violated by putting a ball
on the top of the potential at point (0,0). However, the symmetry will be
spontaneously violated when the ball will roll down. In the valley it can
move without spending energy, so we can put one axis of our coordinate
system on the position of the ball. The new system will be characterized
by one parameter, the distance of the minimum from zero, the vacuum
expectation value (generally abbreviated as vev). Then we parametrize the
BEH field around the displaced minimum and write it in the form φ(x) =
1 0
.
2 v + h(x)
52 Chapter 5
The BEH field introduces four new degrees of freedom in the system,
three of them will give masses to the weak bosons and the fourth one the
Higgs boson, a heavy particle with all quantum numbers zero. From the
known couplings one can calculate the masses of the weak bosons and the
vacuum expectation value of the BEH field, v 246 GeV.
The BEH mechanism was independently published in 1964 by Peter
Higgs and two research groups, but at that time it was received by a great
deal of scepticism. It actually introduces an artificial new field, a force field
with no source, which fills vacuum and creates a new scalar particle with all
zero quantum numbers except its mass which is not predicted by the theory.
It is very hard to detect such a particle. It was clarified in the early seventies
only that the BEH mechanism, in addition to the production of masses for
the weak bosons, solves a whole bunch of other problems: facilitates to
introduce the fermion masses and creates the badly needed scalar boson to
eliminate some mathematical difficulties. As the theory and experiment of
particle physics developed, the belief of physicists increased that the BEH
mechanism should be correct: the predicted neutral currents were observed,
the properties of the weak bosons determined, and all experimental data
seemed to agree with the calculations of the standard model. Finally, 48
years after its prediction, the Higgs boson was observed at the LHC in 2012.
The first two terms are similar to those of QED: the first one is the field
energy density, the second term stands for the interaction of the fermion,
/ = γ μ Dμ where D denotes the covariant derivation. The third term is
D
a possible gauge-invariant coupling between the fermions and the φ BEH
field that leads to fermion masses, the term (Dμ φ)2 generates the masses
of the gauge bosons and the last term is the self-interacting BEH-potential.
The fermion interaction terms need to be complemented by their Hermitian
conjugates. This equation is nicely explained in a generally comprehensible
way in [Woithe et al., 2017].
Interactions 55
Exercise 5.1
What is common in the three interactions described by the Standard Model?
Why is gravity different from them?
Exercise 5.2
To what extent does the magic number 3 connect the 3 fermion families with
the 3 quark colours and 1/3 quark charges?
Exercise 5.3
Conservation of the electric charge and colour charge is connected to which
Noether theorem? What are the two physical quantities whose conservation
is violated by the weak interaction?
Exercise 5.4
Why do quark states mix in ground state? Why is the number of families
limited to 3 in the standard model?
Exercise 5.5
Why does the number of quarks equal that of the leptons?
Exercise 5.6
Explain in what sense it is correct to consider for the numbers of different
quarks 3, 6, 9, 12, 18 and 36.
Exercise 5.7
Explain in what sense it is correct to consider in the standard model for
the numbers of different elementary particles (i.e. of the basic fermions and
bosons) 17, 21, 44, 52, 61 and 109.
Exercise 5.8
How did we determine the number of fermion families? Why cannot quarks
heavier than the top quark exist within the standard model?
Exercise 5.9
Why are the masses of the charged and neutral weak bosons different?
Part II
Experimental methodology
57
Chapter 6
Accelerators
MOTTO:
I sometimes think about the tower at Pisa as the
first particle accelerator, a (nearly) vertical linear
accelerator that Galileo used in his studies.
One needs higher and higher energies to get deeper and deeper in the
structure of matter. Cosmic rays can have very high energies, but at extremely
low intensity and with no choice of parameters. At the times of Columbus
the front line of discovery in Europe was the shore of the Atlantic Ocean:
he used ships to reach India and found America. Nowadays on the front line
of physics we see the Higgs boson and the quark-gluon plasma, the state of
matter right after the Big Bang. We use particle accelerators, now the Large
Hadron Collider (LHC) of CERN to study them, and in the 21st century after
many decades of hard work we see them both. Of course, the devil is in the
details, one needs more and more precise measurements as that is the only
way to find deviations from the predictions of the standard model, i.e. new
physics.
The word accelerator is a bit confusing in contemporary particle physics
as all machines produce particles with velocities extremely close to c, the
59
60 Chapter 6
ure 17.4 shows the cross section of a dipole magnet of the Large Hadron
Collider. The LHC keeps two proton beams in orbit in opposite directions,
and so it has two storage rings with two sets of dipole magnets.
6.2 Acceleration
At low energies sufficient particle acceleration was achieved using electro-
static fields. The largest electrostatic accelerators are the Van de Graaff
generators which can provide about 35 MeV energies. The contemporary
accelerators use resonance acceleration in microwave cavities rather than
electrostatic fields: E = −∂ A/∂t. The modern accelerator cavities can pro-
vide as high as 45 MeV/m acceleration. Figure 17.4 shows such a resonator
cavity built for the TESLA project at DESY, Hamburg.
The first cyclotron was built by E. Lawrence in Berkeley in 1929. He
spent 20 $ on it and was awarded the Nobel Prize for it in 1939. It is
a circular machine (Fig. 6.2) consisting of two half-cylinders (D’s): the
particles circulate in orbits of increasing radii as gaining energy in the gaps
between the D’s where the direction of the field changes in phase with the
circulation of the particles. In the synchrotron the radius of the particles is
fixed and the magnetic fields are tuned to keep the particles in the same orbit.
At CERN all circular machines are synchrotrons.
The linear accelerator (nicknamed linac) developed parallel with the
circular ones. It was first built by Rolf Widerøe in 1928; now usually the
first stage of any accelerator system is a linear one. The accelerator cavity is
in principle similar in the circular and linear accelerators, but in the circular
62 Chapter 6
ones the same cavities are used again and again in every turn of the beam,
whereas in the linear accelerators they are passed only once.
6.3 Colliders
In an ordinary accelerator the beam hits a fixed target. For the collision of
identical particles the centre-of-mass energy is ECM 2 = s ≡ (p1 + p2 )2 =
√
2MX Eb + 2MX , i.e. ECM ∼ Eb . With increasing beam energy ECM , the
2
useful energy of the collision that can produce new particles grows with the
square root of the beam energy. If both particles are accelerated, the energy
released in the collision is the sum of the beam energies in the CM system.
Another advantage of the colliders with identical particles is the centre-
of-mass system in the laboratory. This means not only an easier interpretation
of the data, but also the possibility to encircle the interaction point by a com-
pletely symmetric detector system with which one can detect all produced
particles.
∫
Year E(e+ e− ), GeV Ldt/4, pb−1 main study
1989–94 91 140 properties of Z
1995 130–136 5
1996 161–172 20 properties of W±
1997 184 60 W+ W− , ZZ production
1998 189 190 W+ W− , ZZ production
1999 192–202 220 search for the Higgs boson
2000 204–209 220 search for the Higgs boson
proton and heavy ion acceleration for the Super Proton Synchrotron (SPS),
electron and positron acceleration for LEP and antiproton deceleration for
LEAR, the Low Energy Antiproton Ring. Let us study these phases.
4π ep β γ
2 2 4
ΔE = − (6.1)
3 r
6.6.2 Protons
The protons started from a duo-plasmotron ion source where the electrons
were shaken off from hydrogen molecules, then accelerated in a proton linac
up to 50 MeV. They were injected into the PS Booster where they got to
1 GeV and injected into the PS. The PS provided:
The PS Booster also provided 1 GeV protons for the ISOLDE radioactive
beam facility of CERN to prepare radioactive atomic beams.
66 Chapter 6
6.6.4 Antiprotons
25 GeV protons from PS were injected into an iridium target to produce
proton-antiproton pairs; antiprotons of momentum 3.75 GeV/c were col-
lected in the AA-AC (antiproton accumulator - antiproton collector) double
ring to be cooled and stored until the SPS in its SppS collider regime or later
LEAR, the Low Energy Antiproton Ring received them. The PS decelerated
the antiprotons to 600 MeV/c before ejecting them into LEAR that further
accelerated or decelerated them for the LEAR experiments.
Protons in LHC
The LHC detectors are quite different from each other. ATLAS and
CMS can handle 15 times higher luminosities than LHCb, and 1500 times
higher than ALICE. Depending on the circumstances the LHC could deliver
stable collisions for several days with one fill, although they were frequently
interrupted by glitches in the electric network. The accelerator has many
information pages, for instance we could learn that at the time of taking
Fig. 17.9 the two proton beams met at the collision points at an angle of
170 μrad different from the head-on 180 degrees, their overlap was 25 mm
along the beam direction and 20 × 16 μm in the x − y plane orthogonal to
the beam. The proton bunches followed each other by 25 ns and each ring
contained 2220 proton bunches. In ATLAS and CMS there were 2208 bunch
crossings, in LHCb 2036, and in ALICE 1940.
LHC was started with protons in 2009 and went through an incredible
development. During 2010, its first full data taking year, the luminosity
went through a gradual increase of 8 orders of magnitude in 8 months, it
delivered 46 pb−1 integrated luminosity of proton-proton collisions at 7 TeV
energy. In 2011 it worked at 7 TeV and provided 4.5 fb−1 . In 2012 the
integrated luminosity was 23 fb−1 at 8 TeV. That was the time of observation
of the Higgs boson. After a long shutdown in 2015, the LHC restarted
with 6.5 + 6.5 GeV p-p collisions: it provided 4 fb−1 in 2015, 40 fb−1 in
2016 and 51 fb−1 integrated luminosity in 2017. Of course, there are always
inefficiencies, things sometimes break down and the inner detector part of the
detector is always delayed a bit behind the announcement of “Stable beams”.
Nevertheless, the experiments did very well, with beam usage efficiencies
well above 90%.
Lead ions are prepared and stored in LEIR, the Low Energy Ion Ring that
was built using the elements of LEAR (Fig. 17.6). The lead ions are then
injected to PSB, PS, SPS and LHC for acceleration. At the end of 2010 and
2011 LHC worked a few weeks with Pb-Pb collisions, and after its 2012
proton run, it collided Pb ions with protons. The beam energy of the LHC in
its heavy ion regime was 2.76 TeV / nucleon, so for reference LHC was also
running for a short while colliding protons at 2.76 TeV. For the same reason,
after the 13 TeV proton runs similarly reduced collision energies were also
studied: 5.02 and 8.16 TeV/nucleon p-Pb collisions in 2016, and 5.02 TeV
p-p collisions in 2017. Heavy ion physics will be treated in Chapter 15.
70 Chapter 6
6.8.2 Neutrinos
SPS has a special facility that provides a neutrino beam to study neutrino
oscillations: CERN Neutrinos to Gran Sasso (CNGS). The Italian Laboratori
Nazionali del Gran Sasso (LNGS) is the largest underground laboratory on
Earth devoted to particle and nuclear physics. It is located at 120 km from
Rome near a highway tunnel through the Gran Sasso mountains, under 1.4 km
of rock. It hosts 15 experiments of 900 scientists from 29 countries. We shall
describe the neutrino beam in Chapter 13 devoted to neutrino experiments.
6.8.3 Antiprotons
CPT invariance, the equivalence of matter and antimatter is deeply embedded
in particle physics. However, there are no antimatter galaxies in the Universe,
although during the Big Bang particles and antiparticles should have been
produced in identical quantities. There are also CPT violating extensions of
the standard model. All this demands the precise check of CPT invariance.
LEAR, the Low Energy Antiproton Ring was built mostly for meson spec-
troscopy and it was stopped when the major LEAR experiments finished data
taking. LEAR was converted into LEIR, but the CPT experiments asked
for a low-energy antiproton source. Thus CERN built the Antiproton Decel-
erator facility using the space and elements of the dismounted Antiproton
Accumulator / Collector rings (Fig. 17.10).
The Antiproton Decelerator (AD) works the following way. An extracted
beam from the Proton Synchrotron of 1.5 × 1013 protons at p = 26 GeV/c
momentum is produced in 5 pulses with a total duration of 500 ns. These
protons hit an iridium target where they produce proton-antiproton pairs.
Antiprotons of 3.57 GeV/c momentum are collected and focused using the
magnetic horn technique developed by Simon van der Meer (Nobel Prize,
1984). The antiprotons are injected into the AD ring where they are decel-
erated to 100 MeV/c in four steps, in the first two steps with stochastic and
then electron cooling (Fig. 6.3).
The AD delivers 3 × 107 antiprotons at 100 MeV/c momentum to several
(seven in 2017) experiments, which trap them in electromagnetic fields and
using slow positrons make antihydrogen (pe+ ) atoms. ALPHA and ATRAP
prepares spectroscopy on trapped antihydrogen, ASACUSA and BASE com-
pares the properties (mass, charge and magnetic moment) of protons and
antiprotons at high precision, AEGIS and GBAR tries to measure the gravi-
tational mass of antihydrogen, and ACE studies the effects of antiprotons on
living tissue.
Accelerators 71
Exercise 6.1
What is the energy loss per turn of 100 GeV electrons and 7 TeV protons in
the LEP/LHC tunnel of 27 km circumference due to synchrotron radiation?
Exercise 6.2
What is the energy limit one can reach with protons in a ring with a circum-
ference of 27 km filled with 8 T magnets?
Exercise 6.3
Why is it impossible to keep the LHC detectors ready to take data during
beam acceleration and adjustment?
Exercise 6.4
LHC was designed for 7 + 7 TeV collisions, but so far that was impossible
to reach. According to Fig. 17.9 the 6.5 TeV beam energy could be reached
72 Chapter 6
very fast. What could be the reason why 7 TeV beams were not delivered in
2016-17?
Exercise 6.5
How do you get the 2.76 TeV / nucleon energy for the Pb + Pb run of LHC?
Hint: To what collision energy does it correpond for protons and lead ions?
Chapter 7
Detectors, calorimetry
MOTTO:
Experimenter: A physicist who does experi-
ments. Theorist: A physicist who does not do
experiments.
(Leon M. Lederman)
Fast particles lose their energy in collisions with the atoms of the medium.
At lower energies, around 1 MeV Coulomb scattering dominates. Let us
consider a particle of mass M and charge z × e moving at velocity v = cβ
along the x axis in a medium of atomic number Z and atomic weight A. The
forces along its trajectory compensate
∫ each other, there will be no exchange
2
of momentum along x, px = Fx (t)dt ≈ 0. It applies a force Fy = x 2ze+b2 on
an atomic electron (assumed to be at rest) located at longitudinal coordinate
x in a cylinder at a radius b (Fig. 7.1) and thickness db. The momentum loss
Detectors, calorimetry 75
Figure 7.1: Derivation of the Bethe equation: a particle of mass M and charge
ze moving at velocity v interacts with the atomic electrons in a cylinder of
radius b of a medium
∫ ∫
ze2 b dx ze2 b dx 2ze2
√ = =
x 2 + b2 x 2 + b2 v v (x 2 + b2 )3/2 vb
The energy transfer from the incoming particle to a single electron is
p2e 2z2 e4
= .
2me me v 2 b2
We have nZ(2πbdbdx) electrons in unit length of the cylinder, so in order
to calculate the average energy loss per unit path length we have to integrate
over all impact parameters b between its minimal and maximal values:
∫ 2
dT bmax
2 ze2 4π(ze2 )2 nZ bmax
− = nZ 2πb db = ln
dx bmin me vb me v 2 bmin
The minimum distance cannot be shorter than the de Broglie wavelength,
bmin ≈ mhe v , whereas the maximal distance is limited by the ionization
γhcβ
potential of the medium, ie. the minimal transferable energy: bmax ≈ I .
76 Chapter 7
Figure 7.2: Energy loss of a positive muon in copper. Up to 500 GeV the
Bethe-Bloch mechanism works well for muons. Below the minimal ioniza-
tion region, 100 MeV/c, for different energy regions different approximative
calculation methods are used. Above the Bethe-Bloch region radiative losses
dominate for muons
Bragg’s rule can be used for calculating the radiation length as well, but it
overestimates the average ionization potential for molecules as the electrons
have tighter bonds there than in the atoms.
Electrons lose energy via radiation and photons via bremsstrahlung1 and
pair creation. As a result, both produce electron-photon showers, and thus
electrons and photons behave very similarly at high energies [Leo, 1987].
Det. type space res., rms (μm) time res. dead time
Photo-emulsion 1
Bubble chamber 10 − 150 1 ms 50 ms
Streamer chamber 300 2 μs 100 ms
√
Prop. chamber 50 − 300 (d/ 12) 2 ns 200 ns
Drift chamber 50 − 300 2 ns 100 ns
Scintillator 100 ps/n 10 ns
LAr drift chamber 170 − 450 200 ns 2 μs
Microstrip ch. 30 − 40 < 10 ns
Resistive plate ch. ≤ 10 1 − 2 ns
Silicon strip d/(3 − 7) readout readout
Silicon pixel 2 readout readout
Table 7.1: Detector types used in high-energy physics. The time resolution
of the various scintillators depends on the material and size of the detectors.
For the silicon detectors the read-out speed defines the timing
the trajectory of a charged particle in a magnetic field one can determine its
charge and momentum. A mixed beam of different particles with a given
momentum can be separated by their time of flight between two detectors.
Different particles lose their energy in collisions with the atoms of a medium
at different rates approximately at dE/dx ∼ z 2 /β2 where x is the thickness
of matter, z is the charge of the particle in units of the electron charge and
β = v/c is its relative velocity. Cherenkov radiation is also characteristic of
a particle velocity: when a particle flies in a medium faster than the light, it
emits Cherenkov light at an angle θ c = arccos nβ 1
where n is the refraction
index of the medium.
Table 7.2: Basic properties of various scintillators: name, light yield relative
to NaI(Tl), decay time of the signal, maximum of wavelength distribution,
density, stopping power and radiation length.
where d1 is the thickness of the foil and Ω is the solid angle. This is a radiation
in the X–ray region, which is strongly forward peaked with a cone angle of
θ c ≈ 1/γ. The total energy radiated for ω p1 ω p2 is Et r = αz 2 γω p1 /3.
82 Chapter 7
Exercise 7.1
What are the advantages and disadvantages of drift chambers against ordi-
nary multiwire chambers and scintillation counters?
Exercise 7.2
Arrange the following particles according to their penetration depths in
matter: muons, pions, protons, electrons, photons, neutrinos.
Detectors, calorimetry 83
Exercise 7.3
Why does the muon has so much longer free flight path in matter than the
pion that has almost the same mass?
Exercise 7.4
Arrange the following detectors in order of detection speed: scintillator,
bubble chamber, emulsion, multiwire chamber, Cherenkov detector, drift
chamber.
Exercise 7.5
Arrange the following detectors in order of energy resolution: scintillator,
bubble chamber, multiwire chamber, Cherenkov detector, drift chamber.
Chapter 8
Event registration
MOTTO:
For those who want some proof that physicists
are human, the proof is in the idiocy of all the
different units which they use for measuring
energy.
(Richard P. Feynman)
into account weighted by their cross sections. Of course one has to check
first how well the simulation reproduces the known processes. At the end
that has to be accounted for in the final uncertainty of the result.
Let us view a few typical event pictures recorded at LEP, Tevatron and
LHC.
φ azimuthal and θ polar angles. Note that at beam crossing, where the
collisions occur, the beams actually intersect each other at a very small angle
different from zero (so the beams indeed cross each other), the transverse
momentum corresponding to this can be neglected. While the transverse
momentum and azimuthal angle are invariant under Lorentz boost along the
z axis, the polar angle is not so. Thus instead of the polar angle we can use
rapidity that is invariant under Lorentz-boost:
E + pL
y = 12 ln (8.1)
E − pL
where E and pL are the energy and longitudinal momentum of the particle.
In high-energy collisions the mass of the detected particles (mostly leptons,
pions and kaons) is negligibly small as compared to their energy and their
rapidity is simplified to pseudorapidity, η = − ln tan θ2 (Fig. 8.1), which is
much easier to calculate as for that one does not need to determine the energy
of the particle. Furthermore, the kinematics of the events are such that most
of the secondary particles fly out at small angles and pseudorapidity gives a
more uniform scattering distribution than the polar angle.
Figure 8.1 shows the dependence of pseudorapidity on the polar angle.
In the CMS experiment pseudorapidities |η| < 2.1 − 2.5 (the exact value
depends on the subdetector) belong to the barrel region, outside that is the
forward region.
At hadron colliders the number of events is frequently plotted on an η − φ
picture which corresponds to the plane that you get if you cut the cylindrical
part of the detector at a given radius along the z axis and fold it out. The
angular distance of two tracks is expressed at hadron colliders in the rather
peculiar angle-like quantity
Exercise 8.1
What is the expected number of events of a signal with the cross section of
2.5 pb at a detection efficiency of 20% if the collected luminosity is 10 fb−1 ?
Exercise 8.2
Group the following data whether they must or may be written as part of
the registered events: amplitude and timing of detector hits, temperatures
of detector elements, temperatures of parts of the electronics, temperature
of the counting room, atmospheric pressure, field strengths of the detector
magnet, mains voltage, comments of the physicists on shift.
Exercise 8.3
How do multi-jet events like those in Fig. 17.22 prove the existence of quarks
and gluons?
Exercise 8.4
Why do transverse momentum and pseudorapidity play more important roles
at hadron collisions than at leptonic ones?
Exercise 8.5
In what conditions is pseudorapidity invariant under Lorentz boost?
Exercise 8.6
Why do we expect the t quark to decay predominantly to a b quark?
Chapter 9
Data analysis
MOTTO:
If you are receptive and humble, mathematics
will lead you by the hand.
(Paul A. M. Dirac)
In high energy physics data analysis is very critical for several reasons.
• The huge collaborations (CMS and ATLAS each have more than 3000
participating researchers) can afford to let several groups analyse the
same data, and that makes a strong competition for the best analysis
that can make it to the paper at the end. The competing groups are the
best testers for each others’ work quality.
91
92 Chapter 9
with the parameter vector p. The aim of the analysis is to determine whether
or not the model function describes the data well and to estimate its parameter
values. When the values of p are determined on the basis of the data (via
fitting the parameters of the model function to the data) we can also determine
94 Chapter 9
where x and x both denote the estimations of the expectation values of
x. The diagonal elements are the variances (the squares of the statistical
uncertainties) and the off-diagonal ones the covariances:
−1 ≤ C( p̃i, p̃k ) ≤ +1
so that its values fall in the range [−1, +1].
The correlation between two parameters is +1 Figure 9.1: Uncor-
if they are proportional, and –1 if they are inversely related and correlated
proportional to each other. High correlation means coordinates for the
that the model function is poor, or we have chosen same point.
the wrong parametrization. For instance in Fig. 9.1
the same point P can be described in the uncorrelated (x, y) and the highly
correlated (x , y ) coordinate systems. If P changes in the x direction, its y
coordinate can remain the same, but if P changes in the x direction, its y
coordinate must also change.
If a secondary variable z = z( p) is calculated as a function of p, its
uncertainty according to the law of error propagation (here the misleading
name error has stayed on) is
m
∂z ∂z
σz2 = Δpi Δpk . (9.1)
i,k=1
∂pi ∂pk
have an estimation of the parameters, the question is how well the function
fits the data. There are several ways to check that, the most common being
k
the χ−square test. The χ−square function is defined as χk2 = i=1 Xi2 where
Xi are independent variables following a Gaussian distribution
1 X2
P(X) = √ exp(− )
2π 2
with a mean value Xi = 0 and variance σx2i = 1. χk2 follows the Γ distribution,
thus it has a mean value of χk2 = k and variance σ 2 ( χk2 ) = 2k where k is its
degree of freedom.
When fitting a spectrum by, e.g., the least-squares fitting method to a
set of n independent data and m parameters, the χk2 will have a degree of
freedom k = n − m − 1:
n
1
χn−m−1
2
= [y − f (xi ; p1 · · · pm )]2 .
2 i
σ
i=1 i
χ2
The relative or reduced χk2 is defined as χr2 = n−m−1n−m−1
with a mean χr2 = 1.
χr2 1 means a too good fit, that is too little information in the data for the
given model, whereas χr2 1 means a bad fit: a model function of poorly
fitting parameters.
1 (x − μ)2
P(x) = √ exp(− )
2πσ 2 2σ 2
n
S(a) = wi (yi − ỹ)2 = min , (9.2)
i=1
weighted average is
1
i σ 2 yi
!
1
ỹ = i1 ± 1 . (9.3)
i σ2 i σ2
i i
Data analysis 97
1
n
S
σSD = √ with S 2 = (yi − ỹ)2
n−1 n i=1
more reasonable error: σSD ≈ 26 6 = 2.1. If we are not sure about the
statistical nature of the measurements, we should use the standard deviation
for the estimation of the uncertainty.
m
∂ f (x; p)
m
f (x; p) = pk = pk Φk (x),
k=1
∂pk k=1
98 Chapter 9
∂S( p)
i.e. the system of equations ∂pk = 0; (k = 1, ...m). Using the notation
fi ≡ f (xi ; p) we get
m n
1 ∂ fi ∂ fi 1
n
∂ fi
pl = yi ,
l=1
σ
i=1 i
2 ∂p k ∂pl σ
i=1 i
2 ∂p k
m
equivalent to l=1 pl Mkl = bk . We can solve this if the matrix Mkl is
symmetric and positive definite:
Mp = b ⇒ p = M−1 b ≡ Q b.
It can be easily shown that the covariance matrix of p will be just the inverse
of the Mkl matrix.
For fitting a straight line f (x; p) = p1 x + p2 to points {xi ; yi ± σi }i=1
n :
⎡ xi2 xi ⎤ ( x i yi )
⎢ 2 ⎥ σ2
M = ⎢⎢ xiiσ
1i ⎥⎥ ; b = yi i
σ
2
.
⎢ σ σ 2 ⎥ σi2
⎣ i ⎦
2
i
p˜1
The solution is = M −1 b with the covariance matrix Q = M −1 .
p˜2
Data analysis 99
but in all other methods one has to find a local minimum or maximum.
In particle experiments the MINUIT program [James and Roos, 1975] has
been the most popular for fitting for almost half a century.
The Monte Carlo method can be used if we have just a vague idea of
the location of the minimum: one makes random probes in the given region
and tries to find a minimal value of the function. It is extremely slow and
inefficient.
The simplex method is very popular. The value of the function S( p) to
be minimized is calculated in m + 1 points in the m-dimensional space of the
parameters p making a simplex of m + 1 edges R1 . . . Rm+1 . Take the worst
point (where e.g. S( p) is maximal in the case of minimization) and send it to
a point in the m-dimensional space on opposite side of centroid calculated
from remaining m points (Fig. 9.5). The speed of convergence will depend
on the distance you send the point away from the centroid: the usual value is
100 Chapter 9
once or twice the original distance. This method is somewhat slow, but does
not require calculating derivatives.
In particle physics the most popular minimiza-
tion methods are the gradient ones. The gradient
of S( p) in the point of p = p0 is
T
∂S ∂S( p) ∂S( p)
g= = ,...
∂p ∂p1 ∂pm p= p0
9.5 Uncertainties
The uncertainty of the estimated pa-
rameter value is usually obtained by
changing the χ2 of the fit by unity
at the two sides of the minimum.
This way one can get asymmetric
uncertainty values. A quite popu-
lar method to gain error estimation
of a single parameter is the MINOS
method (Fig. 9.6). It changes the
value of the given parameter and op-
timizes the rest of the parameters of
the fitting so that χnew
2 2 + 1. Figure 9.6: Estimating MINOS uncer-
= χmin
This will account for part of the cor- tainties by changing the parameter in
relations as well. question while optimizing the rest so
All uncertainties of the estima- that the χ of the fit would increase by
2
tion which do not come from the unity [James and Roos, 1975].
number of detected events are con-
sidered to be systematic. They come from various sources, like Monte Carlo
statistics and inputs, experimental calibration factors, detection efficiencies
etc, with the common name nuisance parameters. They can be estimated
Data analysis 101
n0 0 1 2 3
. nmax (90%) 2.30 3.89 5.32 6.68
nmax (95%) 3.00 4.74 6.30 7.75
For instance, there was a case when a fitted mass-square was negative:
π + →μ+ νμ ⇒ m(νμ )2 = (−0.016 ± 0.023)MeV2,
which is of course non-physical. The distribution was renormalized and the
limit m(νμ ) < 0.17 MeV (90%CL) accepted.
Data analysis 103
(a) Eliminate those events where a Z-boson was emitted together with a
photon (radiative return to Z). If an isolated
√ photon is found, remove
it from the total centre-of-mass
√ energy s and apply a cut on the
remaining energy s > 170 GeV.
√
(b) Cut on the total visible energy of the event by Evis / s > 0.7.
(c) Choose 4-jet events by cutting on the variable y34 of the Durham jet-
determining algorithm, which is the threshold track distance where
the event becomes 4-jet-like from 3-jet-like.
104 Chapter 9
(d) Select those events where all jets have at least two charged tracks.
One can see that the simulation describes the data very well. We tried to
apply cuts with maximum efficiency, i.e. eliminate much of the background
and little of the signal. Of course the total number of events went down as
we applied the cuts consecutively from (a) to (d).
The next step was a likelihood optimization with the basic aim to reduce
the main background of our searched signal e+ e− →H+ H− →cs cs. The only
difference between W± and H± of the same mass is spin as the W is a
J = 1 vector boson whereas the charged Higgs bosons are scalar with J = 0.
That means we should try to find variables to select between them like
angular distributions and jet structure. Using the simulations we created
reference histograms for several such variables for signal and background
and determined the probability of signal likeness for each event. The analysis
was mass-dependent, i.e. for each assumed mass of the hypothetical charged
Higgs boson a different optimization was made with different signal but the
same background. As Fig. 17.24 shows the signal was obviously rejected
for M(H± ) = 60 GeV, but that cannot be told of a signal at 75 GeV. Actually,
after collecting all information we obtained M(H± ) > 75.5 GeV at a 95%
confidence level. We got close to the mass of W± , but could not get over it.
Exercise 9.1
What are the main sources of the statistical and systematic uncertainties?
Exercise 9.2
What is the use of the covariance matrix in the interpretation of the ana-
lyzed data? How can it be used to estimate the statistical and systematic
uncertainties of derived physical parameters?
Exercise 9.3
What is the meaning of χ2 ? What shall we do if it is too high or too low?
Exercise 9.4
Deduce Eq. 9.3 for the weighted average from Eq. 9.2 using the error prop-
agation law, Eq. 9.1.
Exercise 9.5
How can one include former results in the analysis of new data?
Data analysis 105
Exercise 9.6
How can one optimize the event selection for data analysis?
Part III
Basic experiments
107
Chapter 10
MOTTO:
If I could remember the names of all these
particles, I’d be a botanist.
(Enrico Fermi)
+ −
Γ(V0 → ) ∝ eq2 |ψ(0)| 2 /MV2 .
Two vector mesons, ρ(770) and ω(782) have very similar masses, but
due to their different quark charges ρ(770) should decay to electrons with 9
times higher probability than ω(782) (Table 10.1): Γe (ρ) : Γe (ω) 9 : 1.
In spite of the very rough approximation when we neglected their different
masses and state functions, the experimental observation is quite close to it,
[Γe (ρ) : Γe (ω)]exp = 7.02(11)
0.60(2) 12. Of course, if the u and d quarks had
different charges, the ratio would be quite different.
quarks of the nucleons (Fig. 10.2). The reaction rates will be proportional
to the charges of the annihilating quarks:
π − = (ud) ⇒ σ ∝ 18Q2u = 18 · 4
9
π + = (ud) ⇒ σ ∝ 18Q2d = 18 · 1
9
confirmed by experiment.
10.4 Colours
The quark model postulates that quarks come with three colour charges and
colour makes the state function of baryon systems antisymmetric against
baryon exchange. Of course, the three colours make possible the existence
of such particles like the Δ++ = (u↑u↑u↑): J = 3/2; I3 = 3/2 hyperon.
Also, the J = 3/2 ground state needs at least three quarks, and no impossible
baryon state (like one having I > 3/2 isospin) was discovered. The simplest
solution is a new SU(3) quantum number, the three colours.
The standard model assumes that only 3-quark and quark + anti-quark
states exist. The requirement of colourlessness of the free hadrons could also
be fulfilled by e.g. bound states of 2 quarks + 2 anti-quarks (tetra-quark),
of 4 quarks + 1 antiquark (penta-quark), or 3 + 3 quarks and/or anti-quarks
(hexa-quark or dibaryon). Previously, all sightings of such states were
subsequently refuted, but recently the LHCb Collaboration reported the
observation of such states.
The existence of three colours were proven, e.g. by measuring the total
cross section of hadron production as compared to that of muon pair emission.
114 Chapter 10
Exercise 10.1
How can we discover a new particle by measuring energy in particle scat-
tering?
Experimental tests of the quark model 115
Exercise 10.2
What is the significance of having 3 kinds of light neutrinos for the rest of
fermions? How is the number of lepton families connected to those of the
quarks?
Exercise 10.3
How could the total Z-width be determined by studying the resonances of
individual decay channels in electron-positron scattering?
Exercise 10.4
Why does the observation of hadron jets prove the existence of coloured
particles? Why can 3 jets originate from gluons only?
Exercise 10.5
How was the idea of fractional charges of the quarks proven experimentally?
And that of the three colours?
Exercise 10.6
Using the quark compositions of the lightest vector mesons, deduce the cross
section ratio of electromagnetic decay widths Γe (ρ) : Γe (ω) 9 : 1 (Hint:
sum up the charges in Table 10.1).
Chapter 11
MOTTO:
I cannot believe God is a weak left-hander.
(Wolfgang Pauli)
60 Co → 60 Ni∗ +e− +ν e
J=5 J=4 + 12 + 12
−→ −→ → →
Figure 11.2: Parity violation in pion decay to muon and neutrino: only
process B is realized in Nature
Figure 11.3: Positive pion decay produces right-polarized positive muons and those
emit positrons in the direction of polarization
forced to be left polarized, the emitted positive muons are also left polarized,
and they can be used to study local magnetic fields in solid, liquid or gaseous
matter just by detecting the direction of the positrons emitted at muon decay.
The muons first slow down via Coulomb scattering (not losing their polar-
ization) and then get trapped somewhere in the sample. In a magnetic field
the slow muons decay with a lifetime of τμ = 2.2μs while precessing with a
frequency ω proportional to the magnetic field strength B (Fig. 11.3):
qB qB
ω = gμ S≈
mc mc
where q = ±e is the charge and m is the mass of the muon. For several
decades μSR was used to study the behaviour and measure the field strength
of local magnetic fields in solid state physics and chemistry, even meson
factories were built to facilitate this method. Today it is much less cultivated.
g−2
The anomalous part of the magnetic moment, a = 2 is due to all kinds
of corrections. The standard model predicts it to be
However, if there is physics beyond the standard model with new particles,
they could also contribute to the corrections, that is why the anomalous
magnetic moment of the muon is one of those important quantities which
could make it possible to discover new physics even at low energies. Thus
it is continuously probed at higher and higher precision. If we find a devi-
ation from the standard model prediction, it can be due to new physics or
incomplete calculation, and thus useful in any case. Let us add that precise
measurements are always useful in physics, many discoveries were made just
by trying to make a measurement more precise. Two bright examples are
cosmic background radiation and CP violation: both discoveries changed
our view of Nature and earned Nobel Prizes for the researchers.
eB eB
ωs = g + (1 − γ) ,
2mc mcγ
One can see from Eq. (11.3) that for muons having a magic momentum
(hence speed) for which a − γ21−1 = 0 the electric field effect is completely
Parity violation, pions and muons 123
Figure 11.5: The storage ring of the muon g-2 experiment at Brookhaven
National Laboratory, 1999-2006
eliminated from the measured frequency. This experiment was first made
at CERN in 1979 by J. Bailey et al. in a homogeneous magnetic field
of strength B = 1.5 T with the following parameters: pμ = 3.094 GeV/c
magic momentum; γμ = 29.37; γτμ = 64.4μs. Using the method of magic
momentum increased the precision of the measurement by another factor of
3 to aμ = 1165924(85) × 10−9 .
The last such experiment was made at the AGS accelerator of Brookhaven
National Laboratory using the same magic momentum method. The stor-
age ring (Fig. 11.5) with 24 detectors collecting the electrons from μ−
and the positrons from μ+ decay was operated in 1999–2002, and mea-
sured separately the anomalous magnetic moments for μ+ and μ− . The
results agreed with each other and their average [Patrignani et al., 2016] was
(116592089 ± 54stat ± 33syst) × 10−11 , two orders of magnitude better than
the CERN result. There is a 3.5σ (σ is the experimental uncertainty) differ-
ence between theory and experiment. This could mean some new physics,
but may also be due to the incomplete calculation of hadronic contributions,
so the muon (g −2) value is worth further improvement. The experiment will
continue at Fermilab and for that the storage ring was transported from New
York to Chicago. As it was much too large for trains, airplanes or trucks,
it was shipped via the Atlantic Ocean, the Mexican Bay and the Mississippi
river, see the home page at http://muon-g-2.fnal.gov/bigmove/. The ring is
124 Chapter 11
in operation already, we can expect the first results in the near future.
Exercise 11.1
Why is the polarization of the muon definite when the muon is emitted in
pion decay?
Exercise 11.2
Derive Eq. (11.2).
Exercise 11.3
Show how the muon storage ring facilitates direct measurement of the anoma-
lous magnetic moment of the muon.
Exercise 11.4
How could new physics phenomena beyond the standard model be discovered
via measuring (g − 2) of the muon?
Exercise 11.5
How do we measure the polarization of the muon?
Exercise 11.6
Why does the muon have deeper penetration in matter than both the much
lighter electron or the much heavier proton and neutron?
Chapter 12
MOTTO:
Alice laughed. “There’s no use trying,” she
said: “one can’t believe impossible things.”
“I daresay you haven’t had much practice,”
said the Queen. “When I was your age, I al-
ways did it for half-an-hour a day. Why, some-
times I’ve believed as many as six impossible
things before breakfast.”
12.1 Kaons
As discussed in Chapter 1, CPT invariance is one of the most fundamen-
tal symmetries in physics. As for elementary particle interactions the time
reversal invariance is also quite basic and generally established by experi-
mental evidence, it was generally assumed that CP invariance also has to be
valid for all microscopic processes, even when parity symmetry has proved
to be completely violated by the weak interaction. Fig. 12.1 shows the four
possible CP-variations for pion decay: only two of them will be allowed,
both C and P have to be reflected in order to get from the case where the
neutrino is left-polarized to the right-polarized anti-neutrino.
125
126 Chapter 12
Figure 12.1: Pion decay: out of the four possible CP-variations only two
will be allowed, both C and P have to be reflected.
Strangeness I3 = + 12 I3 = − 12
S = +1 K+ = us K0 = ds
0
S = −1 K = ds K− = us
Figure 12.2: The lightest K-mesons. Left: kaon formation in pion-nucleon colli-
sions. Right: the strong eigenstates of kaons.
Kaons are the second lightest mesons after the pions (Fig. 12.2). They
contain s quarks which carry the strangeness S flavour quantum number.
Pseudo-scalar mesons, their spin-parity quantum number is J P = 0− . We
met them already in Chapter 11: Lee and Yang discovered parity violation
when noted that K+ , the positive kaon appeared as two different particles
in the experiment. Among the kaons, in addition to K± we have another
particle-antiparticle pair: K0 and K0 .
The kaons are created via colour interaction in flavour eigenstates and decay
via weak interaction in weak eigenstates. Assuming that the weak interaction
conserves CP parity, one can look for weak decays in the forms of CP
0
eigenstates. K0 and K , however, are not those: CP |K0 >= |K0 > and
CP |K0 >= |K0 >. We can construct CP eigenstates:
K10 1 1
0
= √ (K0 ± K0 ) = √ (ds ± ds)
K2 2 2
1 −Γ1 t
IK0 (t) = 12 (a1 + a2 ) · (a1∗ + a2∗ ) = (e + e−Γ2 t + Re(a1 a2∗ ))
4
where
Γ1 t Γ2 t Γ1 t Γ2 t
Re(a1 a2∗ ) = exp − −im1 t − +im2 t +exp − +im1 t − −im2 t
2 2 2 2
Γ1 + Γ2
= exp − t [ei(m2 −m1 )t + e−i(m2 −m1 )t ] .
2
Using eix = cos(x) + i sin(x), the expression in the brackets can be written
as cos(Δmt),
1 −Γ1 t −Γ2 t Γ1 + Γ2
IK0 (t) = e +e + 2 exp − t cos Δmt ,
4 2
0
|m(K0 ) − m(K )|
< 10−18
m(mean)
Another interesting feature of kaon physics is the regeneration process
(Fig. 12.4). After a long flight the neutral kaon beam consists mostly of
long-lived K20 = √1 (ds − ds). As shown by Pais and Piccioni in 1955, if it
2
is injected into matter, the d quark is just scattered on nucleons, whereas its
antiparticle component can annihilate and so K0 is partially reconfigured.
Kaons and CP violation 129
12.4 CP violation
Around 1960 everybody believed in the conserved CP symmetry, but af-
ter the discovery of parity violation, CP invariance called for experimen-
tal testing as well. J.H. Christenson, James W. Cronin, Val L. Fitch and
R. Turlay published the observation of the CP-violating decay K02 →ππ in
1964, (Cronin and Fitch were awarded the Nobel Prize in 1980). Their
method was based on the different kinematics of 2- and 3-pion decay modes
of the neutral kaon: a 3-pion decay produces pions more-or-less uncorrelated
in angle, but for the 2-pion decay the kinematic conditions are determined.
The experiment fixed the angle and energy of the detected pions and gave
45 ± 9 π + π − pairs above background (note the 5σ excess!). This means
130 Chapter 12
that the long-lived neutral kaon is not a clean CP-odd eigenstate, it has a
CP-even component as well. This mixing can be written as KS = K1 + K2
and K L = K2 + K1 where KS is the short-lived and K L is the long-lived
neutral kaon. Here | | = (2.28 ± 0.02) · 10−3 so — as opposed to parity
violation — CP invariance is just a little bit violated.
CP invariance would mean |e+ νe π − > = |e− ν e π + >, but CP violation
should violate that as well. We expect that in the decay of the long-lived
neutral kaon,
* + −
e νe π
KL →
e− νe π +
N(K L →π 0 π 0 ) N(KS →π + π − )
R= .
N(KS →π 0 π 0 ) N(K L →π + π − )
N(K L →π 0 π 0 ) N(KS →π + π − )
= 0.99098 ± 0.00101 ± 0.00126
N(KS →π 0 π 0 ) N(K L →π + π − )
with the first uncertainty being statistical, the second one systematic. Clearly,
the experiment has demonstrated the direct CP-violation.
Exercise 12.1
In what sense can we speak of anti-kaons? How are antiparticles introduced
in particle physics?
Exercise 12.2
Show that K01 and K02 are CP eigenstates and that is why they can only decay
to 2 and 3 pions.
Exercise 12.3
How can energy and momentum conservation simultaneously fulfilled when
the kaon changes its mass in flight?
Exercise 12.4
What is the mechanism of kaon regeneration? How does the short-lived kaon
become long-lived again?
Exercise 12.5
What is the most precise test of CPT invariance and why?
Chapter 13
Neutrinos
MOTTO:
I have done a terrible thing, I have postulated
a particle that cannot be detected.
(Wolfgang Pauli)
Figure 13.1: Weak interaction processes. From left to right: neutron beta-
decay at nucleon level, depicted in a crossing symmetric diagram, and at
quark level (charged currents); antineutrino scattering on an electron (neutral
current). The arrows follow the CPT convention
are large, built under water, in deep mines, or tunnels under earth, and one
is sunk in the ice of the Antarctica.
It was quite a breakthrough when in 1987 the neutrinos from supernova
SN1987A were detected by the existing large neutrino experiments, the
Irvine–Michigan–Brookhaven (IMB) detector in the USA, Kamiokande in
Japan and Baksan in Russia, although the whole neutrino burst lasted 13
seconds only.
which practically means one detected neutrino interaction a day per 1030
atoms, i.e. per 10–100 tons of material. The Homestake experiment of
Davis collected and analysed radioactive 37 Ar atoms from the reaction
νe +37 Cl→37 Ar + e− .
The expected rate from the standard solar model was 8.2 ± 1.8 SNU, but they
found only 2.56 ± 0.23 SNU. As all subsequent measurements confirmed
both this result and the validity of the model, this huge discrepancy remained
mysterious for a long time.
A second mystery is connected to the disappearance of the atmospheric
muon neutrinos. As cosmic particles (mostly protons and helium nuclei) hit
the atmosphere they produce pions which decay to muons and the muons to
electrons :
π+ −→ μ+ νμ π− −→ μ− ν μ
| and |
−→ e+ ν μ νe −→ e− ν e νμ
Neutrinos 137
These processes should yield twice as many muon neutrinos than electron
ones, but the measurements gave much less than the expected factor of two.
As the neutrinos are neutral point-like fermions, they could be Majorana
particles which are their own antiparticles with the opposite polarization:
νL = ν R . In that case, however, the neutrino-less double beta-decay should
be possible as shown in Fig. 13.3. Many experiments were and are looking
for such a process, and so far upper limits were estimated only for its half-life,
typically about 1021 years (on a 90 % confidence level).
where the first uncertainty is statistical, the second one systematic. Thus it
was confirmed that only 40% of the electron neutrinos of the Sun predicted by
the standard solar model arrive to Earth. The Sun neutrinos were identified
by type and energy only and their origin reconstructed by the Cherenkov
cones nicely showed the shape of the Sun (Fig. 17.29).
Part of the atmospheric muon neutrinos also disappeared, SKK measured
the following yield:
(Nμ /Ne )data
= 0.688 ± 0.016 ± 0.050
(Nμ /Ne )MC
where, as usual, the first uncertainty is statistical, the second one systematic.
However, SKK could also separate muons and electrons, and sort them by
energy and direction, which allowed to make an amazing discovery: the
electron neutrinos of cosmic origin (i.e. of higher energies than those from
the Sun) are in order, whereas the higher energy muon neutrinos from the
atmosphere at the opposite side of Earth partially disappear (Fig. 17.30).
The ratio (flux up)/(flux down)
N(−1.0 < cos θ < −0.2)
= 0.54 ± 0.04
N(0.2 < cos θ < 1.0)
The conclusion from the SKK results was that (1) low-energy electron
neutrinos oscillate into other types on the Sun-Earth distance and (2) high-
energy muon neutrinos at the distance of Earth’s diameter oscillate into
τ neutrinos as they do not appear in the electron neutrino spectra. The
measured flux ratio was consistent with 1.3 × 10−3 eV2 ≤ ΔMatm 2 ≤ 3.0 ×
−3 2
10 eV . Note that these measurements cannot determine the neutrino
masses, just the differences in their mass squares.
ν e + p→e+ + n
with energies between 1.8 and 8 MeV. Energy conservation prohibits the
reactions involving muons or tau-leptons, and the oscillation is detected by
the disappearance of the electron anti-neutrinos.
For three families the survival probability of ν e is [An et al., 2017]
Psur ≈ 1 − cos4 Θ13 sin2 (2Θ12 ) sin2 Δ21 − sin2 (2Θ13 ) sin2 Δee ,
L[m]
Δi j ≈ 1.267 Δmi2j [eV2 ]
Eν [MeV]
with the mass-squared difference measured in eV, the flight distance in meters
and the neutrino energy in MeV. The effective neutrino disappearance phase
Δee (not to be mixed with Δi j ) is an empirical quantity independent of the
neutrino mass hierarchy. The reactor experiments help to determine the
two quantities Θ13 and Δee . From the latter an empirical mass-squared
difference Δmee 2 can be deduced, characterizing the disappearance of reactor
anti-neutrinos.
The Daya Bay experiment (Fig. 17.32) is built near Hong Kong in China.
It consists of eight ν detectors in three groups in the vicinity of six nuclear
reactors. Each detector contains 20 tons of liquid scintillator. It is aimed
at determining the mixing of neutrinos of the first and third families. The
results published in 2017 [An et al., 2017] were the following:
• For the ν1 ↔ν3 mixing they measured a mixing angle of
|Δmee
2
| = 2.50 ± 0.06 stat ± 0.06 syst) × 10−3 eV2 .
muons (they get captured in atomic orbits and then absorbed by the nuclei),
whereas the positive particles decay: π + →μ+ + νμ , μ+ →e+ + ν μ + νe . These
reactions produce three kinds of neutrinos, but not electron antineutrinos; if
they appear, it is a sign of ν μ →ν e oscillation.
LSND saw a 4σ excess of ν e above expected background using the
ν e p→e+ n reaction. Considering that we do not see this oscillation with
the atmospheric neutrinos, the difference has to be connected to the different
energy of neutrinos (in Los Alamos it was up to 52.8 MeV) and their different
flight distances (30 m between target and detector). The conclusion of the
collaboration was that Mν > 0.4 eV.
The LSND result contradicts the other experiments. If it is valid then
a fourth, sterile neutrino must exist. As usual, no new observation is ac-
cepted in particle physics before another experiment confirms it (although the
LSND result did not reach the required 5σ level). A new collaboration was
formed and it built the MiniBooNE (Booster Neutrino Experiment) detector
at Fermilab (Fig. 13.4) in 2005 to confirm or refute the LSND result.
Fig. 13.4 presents the beam line of the MiniBooNE experiment at Fer-
milab. It reproduces the conditions of LSND at an order of magnitude
higher energy and distance. LSND was located at L1 = 30 m from the
target and received E1 = 30 MeV neutrinos on average. The MiniBooNE
detector is located at L2 = 500 m from the production target which pro-
vides E2 = 500 MeV neutrinos. Thus the ratios determining the oscillation
frequency were the same, E1 /L1 = E2 /L2 .
In 2007 the MiniBooNE Collaboration stated that they did not see the
antineutrinos of LSND. In 2009 they announced that they started seeing
something and in 2012 they published a 3.8σ excess, almost exactly the
144 Chapter 13
• What is really the mass of the neutrinos? The oscillation gives the
difference of mass-squares only.
We can conclude that neutrinos present a real enigma: they are bright
signal posts showing the holes and deficiencies in our standard model in
spite of all its success in describing the experimental data, and also a way to
extend our knowledge.
Exercise 13.1
Why does the existence of neutrino oscillations point toward the existence
of new physics beyond the standard model whereas that of kaon oscillation
does not?
Exercise 13.2
How can energy and momentum conservation simultaneously fulfilled while
the neutrino changes its mass during flight?
Exercise 13.3
What is the difference between mass and flavour eigenstates for neutrinos?
Exercise 13.4
How was the number of light neutrinos measured?
Exercise 13.5
How does the existence of neutrino oscillations prove the non-zero mass of
neutrinos?
Exercise 13.6
What is the difference between left and right polarised neutrinos and an-
tineutrinos?
Chapter 14
Higgs boson
MOTTO:
It was in 1972 . . . that my life as a boson really
began.
The standard model seems to describe all experimental data so well that
any observed deviation from it is looked upon with deep suspicion. For
almost 40 years, more and more precise new data have been acquired at the
particle accelerators and all seem to agree very well with the predictions of
the standard model. Hundreds of experiments are summarized in Fig. 17.33
according to the LEP Electroweak Working Group [Schael et al., 2013]. It
shows the 2012 situation of the analysis of electroweak data: all experimen-
tal data and theoretical estimates agree within the experimental uncertain-
ties. The only parameter which deviates at more than 2σ uncertainty is the
forward-backward asymmetry of the decay of the Z boson to two b quarks.
Since the observation of the top quark in 1995 the Higgs boson was
the only missing elementary particle of the standard model. Its observa-
tion in 2012 put the dot on i, proving the mechanism of mass creation via
spontaneous symmetry breaking.
147
148 Chapter 14
14.1.1 Methodology
When searching for a new particle, what we usually try to observe is a
resonance. As described in Chapter 2, for a particle with lifetime τ = Γ−1
and decay rate Γ the event rate against the energy corresponding to the
invariant mass of the decay products is
1
| χ(E)| 2 = ,
(E − M)2 + Γ2 /4
i.e. a Lorentz curve. It shows a peak at the invariant mass M of the decaying
system with a full width Γ at half maximum. We claim the discovery of a
new particle if we see a resonance with a confidence level at least 5σ at the
same invariant mass of the particle in all expected decay channels, by all
related experiments.
The search involves several consecutive steps as outlined already in
Chapter 9.
• Compose a complete standard model background using Monte Carlo
simulation taking into account all types of possible events normalized
to their cross sections.
• Compose Higgs-boson signals, simulations of all possible production
and decay processes at all possible Higgs-boson masses.
• Put all these through the detector simulation to get events analogous
to the expected measured ones.
• Optimize the event selection via reducing the B background and en-
hancing the S signal via maximizing some figure of merit.
• Check the background, i.e. the description of data by the simulation for
the given luminosity: the simulation should reproduce the observed
background distributions in all details. For instance, you can check
the background of the decay of a neutral particle to charged leptons
by selecting lepton pairs of identical charges.
Neutrinos 149
• Check the signal: does it agree with the expectation by the theoretical
model?
Once we are happy with the simulations and the event selection, we
must choose a test statistic. That could be any kind of probability variable
characteristic of the given phenomenon: probabilities for having background
only, signal or combinations. One of the favourite is the Q likelihood ratio of
signal + background over background: Q = L s+b /L b . What most frequently
plotted is
Nch ⎡
⎤
⎢ nk
sk (mH )Sk (x jk ; mH ) ⎥⎥
−2 ln Q(mH ) = 2 ⎢
⎢sk (mH ) − ln 1 +
bk Bk (x jk ) ⎥
k=1 ⎢⎣ j=1 ⎥
⎦
where the variables are the following:
• CLb , the signal confidence level assuming background only, i.e. the
complete absence of the signal, or
projections were not very promising for a discovery of the SM Higgs boson
(the effect seen by ALEPH only was far too large, much higher than the
prediction of the standard model), and the contractors for building the LHC
were already prepared to start.
had pointed toward a light Higgs boson with a mass around 100 GeV. As
LEP excluded the Higgs boson below 114 GeV the LHC experiments had
to be prepared for detecting the Higgs boson in the most complicated mass
region, around 120 GeV, with several competing decay channels (Fig. 17.34).
It is well seen that below 110 GeV the bb̄ decay, above 160 GeV the WW
dominates, that is why those mass regions could be excluded earlier. The-
oretical calculations have shown that the best channels to observe a light
Higgs boson at the LHC should be the two-photon, H→γγ and four-lepton,
H→ZZ∗ → + − + − channels. The hadronic decay channels are hampered
by the very high hadron background. However, these signal channels have
very low branching ratios, for H→γγ it is BR = 2.27 × 10−3 and the sig-
nal/background ratio is small, S/B 1, and for H → ZZ∗ → + − + − (as
usual, = e, μ, i.e. in the experiment only the electron and the muon are
considered leptons, because the taus can decay to hadrons as well) it is even
lower, BR = 1.24 × 10−4 , but with a much cleaner signal, S/B > 1. In 2012
the LHC luminosity was already so high that every bunch crossing (event in
every 50 ns) contained 10–20 p-p collisions leading to copious hadron pro-
duction. Both large experiments, CMS and ATLAS designed their tracking
systems and electromagnetic calorimeters with these two processes in mind.
The electromagnetic calorimeter of CMS consists of 75,848 PbWO4 single
crystal scintillators, whereas that of ATLAS is a sampling calorimeter based
on liquid argon shower detectors.
By the beginning of 2012, when all 2011 data were analysed, the possible
mass of the SM Higgs boson was already confined to the region of 114 <
MH < 127 GeV by CMS with very similar results from ATLAS. In that
region 2– 3 σ excesses were found at 125 GeV in the two main decay
channels, H →γγ and H → ZZ. It seemed more and more probable that the
Higgs boson would be observed at the LHC in 2012. It was even decided
by the CERN administration to extend the data taking scheduled for 2012
before the long shutdown for accelerator development (for increasing its p-p
collision energy from 8 to 13 TeV and the event rate from 20 to 40 MHz) if
necessary for the discovery.
On July 4th, at the beginning of the large annual high-energy physics
congress in Melbourne, the spokespersons of ATLAS and CMS gave talks
from CERN (in internet connection to the whole word, including, of course,
the main auditorium of the Australian conference) on Higgs search. They
announced that at the LHC collision energies 7 and 8 TeV, in the two most
significant decay channels H → γγ and H → ZZ → + − + − , at an in-
variant mass of m 125 GeV a new boson is seen by both experiments at
a convincing statistical significance of 5σ confidence level, with properties
Neutrinos 153
corresponding to those of the standard model Higgs boson. The fact that
the new particle could decay to two photons or Z bosons confined its spin
to an even integer, i.e. a boson of S = 0 or S = 2. Of course, as the data
analysis was optimized to find the SM Higgs, it was very unlikely to find
something very different. Nevertheless, the two experiments emphasized
that it has to be studied, whether or not its spin is really zero with a +
parity (the pseudo-scalar mesons have spin 0 with negative parity), and that
its decay probabilities to various final states follow the predictions of the
standard model. After reanalysing their data the Tevatron experiments, CDF
and D0 also found an excess at this mass (after the LHC started the Tevatron
accelerator of Fermilab was stopped).
14.3.2 Observations
On 31 July the two experiments submitted papers of the discovery to Physics
Letters B, they were published 14 August. Both papers, [Aad et al., 2012]
and [Chatrchyan et al., 2012b], were 15 pages long followed by 16 pages
long lists of close to 3000 authors, and both were dedicated to the memory
of those participants who could not live to see the result of the more than
two decades of construction work. Their measured mass distributions, after
analysing about a quarter of the data to be collected in 2012, are shown in
154 Chapter 14
Figs. 17.41 and 17.42. The mass distributions of the di-photon and 4-lepton
decay channels were quite similar for the two experiments. They showed a
significant peak at the same invariant mass of about 125 GeV. In all cases the
signal strengths agreed within uncertainties to the predictions of the standard
model.
What was really convincing of the observation was the distribution of
the p-values of the events selected in the various decay channels of the
hypothetical Higgs boson. It was a joke of statistics that in July 2012 adding
together two decay channels, H → γγ and H → 4 gave the same 5σ
significance for both ATLAS and CMS, whereas adding to it the results for
other channels increased the significance to 6σ for ATLAS and left it at 5σ
for CMS (Fig. 17.42).
The LHC experiments studied the cross sections of the processes con-
nected to the new particle. Fig. 17.45 shows the signal strengths of production
and decay in various possible channels of the Higgs-like boson measured by
CMS [Chatrchyan et al., 2013] as compared to those predicted by the stan-
dard model for the Higgs boson with a mass of 125 GeV. The amplitudes of
Neutrinos 155
all observed signals are in agreement with the expectations of the standard
model. ATLAS got similar results, of course.
Thus what we found is most likely the standard model Higgs boson. On
one hand this is a great success of particle physics. On the other hand this
is somewhat of a disappointment as the SM has theoretical shortcomings
which need new physics to resolve. Just to list a few of them: it cannot unite
the interactions at large energies, cannot include gravitation, cannot account
for the dark matter of the Universe or for the dominance of matter against
antimatter, and cannot explain neutrino masses and oscillations. There are
many extensions of the theory, that can resolve some of those problems, most
popular among them is supersymmetry, which among other things predicts 5
kinds of Higgs bosons, but none of its predicted phenomena could be found
yet experimentally. The observables of the Higgs boson should be sensitive
to some of the features of new physics and these studies will be the main job
of the LHC experiments in the future.
Right after the confirmation of the discovery, in 2013 François Englert
and Peter Higgs were awarded the Nobel Prize in physics.
It is very interesting that the 125 GeV mass of the Higgs boson seems to
be exciting for theoreticians. There was even a special workshop in Madrid
organized to discuss this mass in 2013. The reason is that mH = 125 GeV is
at the border line of the stability of electroweak vacuum on the plane of top
mass against Higgs mass.
If the BEH vacuum has one minimum only, our vacuum is stable, if it has
more, then it can be metastable. If our vacuum is metastable and we are
not in the deepest minimum, then the universe exist in a false vacuum, and
the world can shift into the deeper one by a quantum mechanical tunnelling
effect. In some sense it could be the end of our world as we know it. At the
new BEH minimum all the particles would have different masses and also
the strength of the weak interaction would change. Of course, this assumes
that the standard model is valid up to extreme high energies. As Turner and
Wilczek stated in 1982 [Turner and Wilczek, 1982], even if our vacuum is
metastable, its expected lifetime could be much longer than the age of the
Universe, hence appear stable to us.
These estimations were refined in time and as the mass of the Higgs
boson became definite, it became clear that our world is located somewhere
at the edge of stability. Fig. 17.46 shows such an estimation, presently most
precise in the standard model. The location point of the measured masses of
the Higgs boson and of the t quark fall in a narrow metastable region.
up to the Planck scale. If below that energy some new physics starts, the
picture is invalid.
Thus we do not have to be afraid that our Universe shifts into another
vacuum. At the same time we see another excellent feature of the BEH
mechanism: it could explain the cosmic inflation without assuming another
force field. Apparently we get more and more proof for Lederman’s joke
that the Higgs boson and the BEH mechanism remind him of the deus-ex-
machina in the Greek-Roman dramas, that puts everything in order.
Exercise 14.1
Why do we need symmetry breaking in order to create masses? Why does
the existence of the Higgs boson prove the validity of the symmetry breaking
mechanism?
Exercise 14.2
Why do we need such overly sophisticated statistical methods like −2 ln Q or
the p-value in order to exclude or discover new physics?
Exercise 14.3
Why do hadron colliders have higher chance to discover new physics in spite
of the more difficult data analysis requirements than the electron-positron
colliders with their much cleaner events?
Exercise 14.4
What reactions helped to discover the Higgs boson at the LHC? Why those
decay channels were used which have the lowest yields?
Exercise 14.5
How can we tell that the Higgs-like signal observed at the LHC belongs to
the Higgs boson predicted by the standard model?
Chapter 15
MOTTO:
“It seems very pretty, ... but it’s RATHER
hard to understand! ... Somehow it seems to
fill my head with ideas — only I don’t exactly
know what they are!”
Figure 15.1: Lead-lead collision event at LHC energy of 5.02 TeV per
nucleon pair recorded by the time projection chamber of the ALICE detector
in November 2015. In this collision 1582 positively-charged (darker tracks)
and 1579 negatively-charged (lighter tracks) particles are produced; about
80 % of them are pions
15.2 Hydrodynamics
As mentioned above, due to the low internal viscosity of the quark gluon
plasma its partons move almost free. If their interaction were stronger,
the system could not possibly remember its structure before the collisions,
but it does: hydrodynamics converts spatial anisotropies into momentum
anisotropy. Another characteristic feature is the elliptic flow. The central-
ity (related to the overlap) of the nuclei in nucleus–nucleus collisions (see
Fig. 15.2) is rarely perfect. If the newly formed quark matter were gas-like
162 Chapter 15
then it should expand in all directions with the same momentum. How-
ever, it seems to behave as an almost perfect liquid and expands elliptically
(Fig. 17.47). This azimuthal anisotropy in particle production can be quanti-
tatively characterized by the elliptic flow, the coefficient v2 of the second term
in the Fourier transform of the angular distribution of measured (charged)
particles:
dN N ∞
= 1+2 v n cos (n(φ − ψ n )) . (15.1)
dφ 2π n=1
The observables are the following: φ is the angle in the transverse plane,
ψn are the event plane angles (where the nth harmonic component has its
maximal track multiplicity) and N is the average number of particles per
event. They can be measured for different particle species as functions of
rapidity, centrality and transverse momentum. v2 is the elliptic flow reflect-
ing two-parton correlations in the quark gluon plasma. Higher correlation
coefficients can also be measured and calculated. Useful quantities are also
the event-by-event values of these data, not only the averages.
Elliptic flow is one of the most important measured quantities in heavy
Heavy ion physics 163
15.3 Experiments
For heavy ion experiments the main problem is to follow and identify the
thousands of emitted particles (Fig. 15.1). The detectors of heavy ion physics
look very similar to those in high-energy particle physics: a large magnet, a
vertex detector closest to the beam pipe, an inner tracker, usually a chamber
encircled by electromagnetic and hadronic calorimeters, and muon detectors
outside. The experiment specialized on heavy ion physics at LHC, ALICE
(A Large Ion Collider Experiment) sacrificed data acquisition speed for
precision: it has the largest existing time projection chamber (TPC, see
Chapter 7) around its beam pipe, which helps to follow and identify thousands
of particles, but cannot handle the amount of luminosity ATLAS and CMS
accept with their semiconductor tracker systems. ALICE was placed in
and around the huge magnet of the former LEP experiment L3. Fig. 15.2
shows the correlation between centrality and particle multiplicity measured
by ALICE at 2.76 TeV/NN [Aamodt et al., 2010].
The measuring apparatus of the fixed-target NA49 experiment, now con-
tinued as NA61 (NA stands for North Area of SPS and the number is a
sequence number of accepted experimental proposals) also based on TPC’s.
It mapped the transition from ordinary nuclear matter to quark-gluon plasma
by comparing the following reactions from the point of view of particle
yields:
experiments were the first to demonstrate the transition from nuclear matter
to quark matter by detecting jet quenching in central Au+Au collisions as
compared to proton-proton, proton-Au, deuteron-Au events and peripheral
Au+Au collisions. In the case of peripheral collisions most of the nucleons
will avoid collisions and just shoot along the beam pipe. In the case of
central collisions there is a large activity in the transverse directions, which
is easily measured by the cylindrically built detector systems. Every heavy
ion experiment has a zero degree calorimeter, a detector close to the beam
pipe, in order to check the centrality of the collisions by detecting the forward-
scattered particles and also for measuring the total number of collisions, the
luminosity.
Exercise 15.1
Why pions dominate the hadron jets?
Exercise 15.2
What is the importance of the transverse momentum of jets and collision
centrality in heavy ion collisions?
Exercise 15.3
What are the characteristic features of the transition of nuclear matter into
quark-gluon plasma?
Exercise 15.4
What are the advantages and disadvantages of using time projection cham-
bers for studying heavy ion collisions?
Exercise 15.5
Deduce how one gets 2.76 GeV/NN collision energy at colliding 3.5 TeV Pb
beams. At what beam energies one gets 2.76 GeV/NN for Pb-Pb collisions
in the case of a fixed Pb target?
Chapter 16
Practical applications
MOTTO:
Prediction is very difficult, especially about
the future.
(Niels Bohr)
16.1 Informatics
The first and probably most important example is the World Wide Web. It
was invented and first developed by Tim Berners-Lee and his group at CERN
in 1989-90. As presented in the CERN museum, his boss in 1989 wrote on
the original proposal Vague but exciting. Tim Berners-Lee wrote in 1991:
“The WWW project was started to allow high energy physicists to share data,
news, and documentation. We are very interested in spreading the web to
other areas, and having gateway servers for other data.” First, of course, it
was used at HEP laboratories like Stanford, Fermilab and at the associated
universities, but in 1992 CERN announced that anybody can use the Web
free of royalty and the platform-independent Mosaic browser was opened
for the public: that began such an information explosion in the world that
seems to surpass even Gutenberg’s printing. Since Mosaic many browsers
have been developed and a lot of features added. Of course, with time the
Web became commercial as more and more companies use it for trading, but
it retains its completely free access established by CERN. Thus the free use
of the internet browsers is due to particle physics research. We mean free in
the monetary sense. Internet providers recognized quickly that information
is a new source of wealth and they collect information from their users
At the start of designing the LHC it was clear that it would become impossible
quickly to store and analyse all its data at CERN, so CERN established its
Worldwide LHC Computing Grid, WLCG. The experimental data at LHC
are analysed in several steps, in a distributed manner. The WLCG consists
of tiers. The centre, called Tier-0, is located at CERN and in the Wigner
data centre in Budapest. It receives the primary data from the LHC. After
a preliminary event reconstruction the data are distributed to the Tier-1
data storage centres, for instance these stations of CMS are located in Lyon,
Barcelona, Oxford, Bologna, Karlsruhe, Fermilab (near Chicago) and Taipei.
The actual data analysis is done at about a hundred Tier-2 computer centres
which get their data from the Tier-1 storage stations. This system was
developed by CERN for the LHC, but it helps other applications as well. For
a few years CERN helped to establish MammoGrid, a virtual organization
for diagnosing breast cancer.
Practical applications 169
16.2 Radiation
Some of the discoveries in particle physics were immediately used in practice,
mostly in medicine. Wilhelm Röntgen discovered the X-rays in 1895 and
used it right away to photograph the bones in his wife’s hand. He refused to
patent the X-rays as he realized its enormous significance for the society. In
1901 Röntgen was awarded the very first Nobel Prize in Physics, which he
donated to his university.
Henri Becquerel discovered natural radioactivity in 1896, which was
followed the separation of radium by Pierre and Marie Curie in 1898, the
three of them were awarded the 1903 Nobel Prize in Physics. As early as in
1902 radioactive irradiation was already used in Stockholm for local tumour
therapy.
16.3 Accelerators
There are nearly 40 000 particle accelerators in operation worldwide, about
half of them employed for biomedical uses and the other half in industry
(for surface treatment and to produce microchips). These days less than 200
accelerators are used for particle and nuclear research.
The first cyclotron was created in Berkeley by Ernest Lawrence in 1930,
it was as small as his palm. This was followed by larger and larger machines
(Fig. 17.51) and Lawrence was awarded a Nobel Prize in physics in 1939.
His brother, John Lawrence, a physician, also worked at Berkeley and in
1936 together they produced phosphorus-32 and used it as the first artificial
radioactive isotope for treating leukaemia. This is an excellent example of
how an interdisciplinary environment helps both science and applications.
170 Chapter 16
16.5.1 Teletherapy
The main problem of teletherapy is that the irradiation is not selective, it
destroys tissue not only in the tumour, but also that around it. In the case of
X-rays several beams are used from different directions with collators (In-
tensity Modulation Radiation Therapy, IMRT): using computer simulations
the crossing beams are shaped to minimize the dose in the healthy organs at
risk. Fig. 17.54 shows a simulation of a 3-dimensional conformal radiation
therapy, in which the profile of each radiation beam is shaped to fit the profile
of the target using a multi-leaf collimator and a variable number of beam
shots. Note that the radiation dose is most commonly measured in absorbed
radiation energy per unit mass: 1 Gy = 1 J/kg.
Another modern way of applying conformal radiation therapy is to-
motherapy: the accelerator rotates around the patient with changing colli-
mators. The gamma knife uses hundreds of 60 Co sources for local operation
172 Chapter 16
of brain tumours or arteries. The cyber knife is similar, but uses X-rays from
an electron linac (Fig. 17.54).
There is a way to prevent loosing cancerous tissue during operations,
which can cause metastasis, achieved by electron irradiation of the tumour
area during operation. Usually electrons of 3–9 MeV energy are applied for
a couple of minutes delivering a dose of about 10 Gy.
16.6 Conclusion
We see that particle physics offers useful methods to many other fields of
science, especially to informatics and medicine. For the latter physicists,
engineers and medical doctors should closely work together. Thus particle
Coloured figures 173
physics brings not only new knowledge about Nature, but also useful new
techniques and technologies.
Exercise 16.1
Why was the World Wide Web developed for particle physics?
Exercise 16.2
What is the advantage of hadrons to photons (X- or γ-rays) for cancer
therapy?
Exercise 16.3
Why should we prefer to use protons and carbon ions for hadron therapy to
other nuclei?
Chapter 17
Coloured figures
MOTTO:
Colour in certain places has the great value
of making the outlines and structural planes
seem more energetic.
(Antoni Gaudí)
175
176 Chapter 17
Figure 17.1:
Graphic picture of the Lagrangian of quantum chromodynamics (QCD)
Figure 17.3: Left: Bending of charged particles in a dipolar magnetic field. Right:
a dipole magnet, with two (blue) quadrupoles in the back
Coloured figures 177
Figure 17.4: Left: Cross section of a dipole magnet of the Large Hadron Collider.
Right: Microwave resonator to accelerate electrons for the TESLA project at DESY,
Hamburg
Figure 17.5: The accelerator complex of CERN in the LEP era. LEAR: Low
Energy Antiproton Ring, EPA: Electron-Positron Accumulator, PS: Proton
Synchrotron, PSB: Proton Synchrotron Booster, AAC: Antiproton Accu-
mulator Collector, SPS: Super Proton Synchrotron, LEP: Large Electron-
Positron collider
178 Chapter 17
Figure 17.6: The Low Energy Ion Ring (LEIR) at CERN. The Pb ions are collected
and stored in the ring while continuously cooled. The pipes of the stochastic cooling
are above the ring. The electron cooling is set up on left: the two guns are for
introducing and removing the electron cloud. The bending magnets are painted
orange, the quadrupole doublets and triplets blue. LEIR uses the elements of the
former Low Energy Antiproton Ring (LEAR) functioning in 1982–1996
Figure 17.7: The accelerator complex of CERN in the LHC era, after 2008
Coloured figures 179
Figure 17.8: Left: LHC dipoles waiting for installation. They are 15 m long,
weigh 35 tons, maintain up to B = 8 T magnetic field at T = 1.9 K. Right: The
radio-frequency cavities accelerating the particles in the LHC ring
Figure 17.9: Normal LHC operation at 13 TeV p-p collision energy . The upper
plot shows the energy (black line corresponding to the right scale) and the intensities
of the two beams (red and blue lines). On the bottom the background radiations of
the two lines around the four experiments. ATLAS and CMS needed the maximal
possible luminosities, whereas ALICE and LHCb much less
180 Chapter 17
Figure 17.10: The Antiproton Decelerator of CERN: its building and main
experimental area
Figure 17.12: The toroidal magnet of the ATLAS detector before the detection
elements and the solenoid magnet were installed in it
Coloured figures 181
Figure 17.13: Left: Working principle of a time projection chamber. Right: The
time projection chamber of the ALICE detector at the LHC
CMS Detector
SILICON TRACKER
Pixels (100 x 150 μm2)
~1m2 ~66M channels
Microstrips (80-180μm)
Pixels ~200m2 ~9.6M channels
CRYSTAL ELECTROMAGNETIC
Tracker CALORIMETER (ECAL)
~76k scintillating PbWO4 crystals
ECAL
HCAL
Solenoid PRESHOWER
Steel Yoke Silicon strips
Muons ~16m2 ~137k channels
SUPERCONDUCTING
SOLENOID
Niobium-titanium coil
carrying ~18000 A FORWARD
CALORIMETER
Steel + quartz fibres
~2k channels
HADRON CALORIMETER (HCAL)
Total weight : 14000 tonnes Brass + plastic scintillator MUON CHAMBERS
Overall diameter : 15.0 m ~7k channels Barrel: 250 Drift Tube & 480 Resistive Plate Chambers
Overall length : 28.7 m Endcaps: 468 Cathode Strip & 432 Resistive Plate Chambers
Magnetic field : 3.8 T
Figure 17.14: Structure and main parts of the CMS detector at the LHC
182 Chapter 17
Figure 17.15: Preparation of the PbWO4 single crystal scintillators for CMS.
The electromagnetic calorimeter of CMS consists of 75,848 such crystals
read by avalanche photo-diodes
Figure 17.16: Scintillation tiles of CMS: the wavelength-shifting fibres collect and
carry the light to hybrid photo-diodes to be converted to electronic signals
Coloured figures 183
Figure 17.17: The forward part of the hadron calorimeter of CMS: quartz
fibres in steel to collect Cherenkov light from secondary electrons
Figure 17.18: Inserting the central ring with the superconducting solenoid into the
CMS detector in the experimental cave of LHC underground
184 Chapter 17
Figure 17.22: Hadron jets recorded by the OPAL experiment. The yellow
blocks show the energy deposits in the electromagnetic calorimeter, the purple ones
those in the hadron calorimeter. Left: two-jet event, Z decay to a quark pair,
e+ e− →Z→qq; middle: three-jet event, Z decay to a quark pair when one of them
emits a gluon: e+ e− →Z→qqg; right: WW production and decay to quark pairs,
e+ e− →W+ W− →qqqq where the quarks have different flavours to conserve charge
186 Chapter 17
Figure 17.24: Search for charged Higgs-bosons at LEP. Left: Preselection cuts of
the OPAL search in the 4-jet channel [Horvath, 2003]. Right: Likelihood analysis in
the 4-jet channel for two assumed H± masses [Abbiendi et al., 2012]. Note how well
the data (black dots) are described by the background simulation. The signal curves
are filled yellow (left) and blue (right). The cut values are denoted by arrows
Coloured figures 187
Figure 17.26: Some of the major neutrino experiments. At the end of 2017 there
were 37 such detector systems under earth, water, ice, and near nuclear power plants
188 Chapter 17
Figure 17.27: KATRIN, the KArlsruhe TRItium Neutrino experiment. The spec-
trometer lets one electron of the 1010 tritium decays per second through into the
detector
Figure 17.31: The SNO detector in the Sudbury Neutrino Observatory (1999-2003).
Left: a 3-dimensional drawing of the apparatus, right: the acrylic ball
Figure 17.32: The Daya Bay experiment near Hong Kong: eight ν detectors in
three halls in the vicinity of six nuclear reactors
Coloured figures 191
Figure 17.33: Parameters of the standard model [Schael et al., 2013] as determined
by the experiment (2nd column) with uncertainties (3rd column), its prediction or fit
by the standard model (4th column) and a bar plot showing the difference between
theory and experiment divided by the experimental uncertainty. The agreement is
purely statistical as the difference is in only one case more than 2 uncertainties
192 Chapter 17
Figure 17.34: The various decay channels of the Higgs boson according to the
standard model. At lower masses bb, at higher ones W+ W dominates. Note how
small is the contribution of the important γγ channel
Figure 17.36: Exclusion of the Higgs boson at LEP [Barate et al., 2003]. The
test statistic, −2 ln Q shows a significant signal (well above the standard model
expectation) for ALEPH and nothing for the other three LEP experiments at equivalent
statistical and experimental conditions
Figure 17.37: Exclusion of the Higgs boson at LEP [Barate et al., 2003]. Spaghetti
diagrams of 17 Higgs-like events detected by the four LEP experiments: signal
weights against the simulated Higgs mass. The ALEPH events crowd around
115 GeV, whereas for the other three experiments there are less of them with a
rather random mass distribution
194 Chapter 17
gluon fusion
vector boson
fusion
Figure 17.38: Production of the SM Higgs boson in p-p collisions at LHC. Gluon
fusion (upper right) has the highest cross section, vector boson fusion (lower right)
comes next
Figure 17.41: First observation of the Higgs-like boson in invariant mass distribu-
tions at around 125 GeV by ATLAS [Aad et al., 2012]. Left: H→γγ, the raw events
and also the sum of event weights according to their signal-likelihood. Top right:
H→` + ` − ` + ` − . Bottom right: the local p-values averaged for all channels. The total
p values have shown more excess events than the prediction of the standard model
(dotted line in the lower right plot)
Figure 17.42: First observation of the Higgs-like boson in invariant mass distri-
butions at around 125 GeV by CMS [Chatrchyan et al., 2012b]. Left: H→γγ, the
raw events and also the sum of event weights according to their signal-likelihood.
Middle: H→` + ` − ` + ` − . Right: the local p-values for the individual channels and
their averages. CMS saw somewhat less events than the prediction of the standard
model (dotted line in the right plot)
Coloured figures 197
Figure 17.43: Invariant mass spectra obtained by ATLAS [Aaboud et al., 2017]
and CMS [Sirunyan et al., 2017] with 13 TeV p-p collisions for the H→` + ` − ` + ` −
reaction
Figure 17.44: Invariant mass spectra obtained by ATLAS [ATLAS, 2017] and CMS
[Sirunyan et al., 2017] with 13 TeV p-p collisions for the H → γγ reaction
198 Chapter 17
Figure 17.45: Relative signal strengths in the most significant decay channels
normalized to the predictions of the standard model as measured by ATLAS and
CMS [Aad et al., 2016] for MH = 125 GeV. Left: for various production channels
against the invariant mass of the new boson. Middle: for the various decay channels.
Right: 68% contours for various decay channels when produced in weak vs. strong
processes. All results agree within statistics with each other and with the expectations
of the standard model
Figure 17.46: Stable, unstable and metastable regions of the BEH vacuum on the
plane of the masses of the Higgs boson and the t quark according to the standard model
[Branchina et al., 2014]. The experimental point of the two masses at m H ' 125
and mt ' 173 GeV are encircled by 1σ, 2σ and 3σ contours
Figure 17.47: Simulated collision of heavy ions. In the central part the
temperature is high enough to de-confine coloured quarks and gluons
Coloured figures 199
Figure 17.48: Jet quenching observed at RHIC. Left: The π 0 yield observed by
PHENIX from central Au+Au collisions is suppressed as compared to the d+Au
collisions of minimum bias. Right: The 2-jet events in STAR are suppressed on the
side opposite to the highest energy track for central Au+Au collisions, whereas they
are free to come out in d+Au collisions
Figure 17.50: Jet–jet asymmetry in peripheral (left) and central (right) p-p and Pb-
Pb collisions at the LHC observed by the CMS experiment [Chatrchyan et al., 2012a]
√
at s N N = 2.76 TeV per nucleon pair. The simulation shows the expected asymmetry
with no nuclear medium
Figure 17.52: Left: computer tomography at work. Right: PET pictures of healthy
and cocaine-addict brains
Coloured figures 201
Figure 17.53: Brachytherapy on prostate cancer: about 100 radioactive pins are
placed in the tumour, while checking the placement with an ultrasound probe
Figure 17.54: Left: conform dose distribution with IMRT: the tumour gets 70 Gy,
the metaphase region 50 Gy and the spinal cord less than 25 Gy dose. Right: Cyber
knife, electron linac on a robot arm with very precise, multiple orientations and depth
penetration
202 Chapter 17
Figure 17.55: Left: a charged particle when penetrating matter loses most of its
energy at the end of its flight path (Bragg curve). Right: Gantry for proton therapy
at a hospital
0.6
R4(ycut)
0.2 0.1
LO LO
NLO NLO
NLO+NLL NLO+NLL
NLO+NLL+K 0.0 NLO+NLL+K
0.0
4(ycut)
0.1 0.5 < x < 2 ALEPH & HERWIG 0.1 0.5 < x < 2 HERWIG
0.0 0.0
-0.1 -0.1
-0.2 -0.2
-3.0 -2.5 -2.0 -1.5 -1.0 -3.0 -2.5 -2.0 -1.5 -1.0
log10(ycut) log10(ycut)
e−(k ")
e−(k) θ
q
p(P )
X
Figure 17.63: Evolution of valence quark, sea quark and gluon distributions
Coloured figures 207
Intermezzo
MOTTO:
What is it that breathes fire into the equations
and makes a universe for them to describe?
. . . Why does the universe go to all the bother
of existing?
(Stephen Hawking)
The first three parts of this book discussed the most important results
of experimental particle physics during the past fifty (or in some cases
more) years. As a consequence we know with certainty that the physical
manifestations of the three fundamental particle interactions–the weak, the
electromagnetic, and the strong forces–below the TeV energy scale can be
described by a model based on local quantum field theory that we call the
standard model of particle interactions. There are three almost identical
copies of particle families, containing two quarks and two leptons each (see
Table 3.5), the only difference being the mass of the particles. In the standard
model the neutral leptons are massless, but we already have experimental
evidence (the observation of neutrino oscillations) that they have masses even
210 Intermezzo
if we have not been able to measure them directly. While the experimental
evidence for the existence of only three families is strong, it remains a mystery
why there are exactly three families and how neutrinos get masses.
The mathematical model that describes the interaction is a gauge field
theory based on the SU(3)c ⊗ SU(2)L ⊗ U(1)Y symmetry where “c” refers
to colour degrees of freedom, Y stands for hypercharge and “L” abbreviates
left-handed. The gauge fields that emerge due to the gauge symmetries me-
diate the forces. These are the massless mediators of the electromagnetic and
strong forces, the photon and the gluon, furthermore the massive mediators
of the weak force, the W and Z bosons. In addition there is also a scalar
boson, the elementary excitation of the BEH field, called Higgs particle that
emerge in the spontaneous symmetry breaking of the BEH field together with
the longitudinal components of the W and Z bosons. There is continuously
increasing evidence that the scalar particle observed at the LHC has the exact
properties as predicted in the standard model. The interaction of the BEH
field with the fermions explains the masses of the fermions, although the
reason for the large hierarchy of the fermion masses is unknown. Accord-
ing to the standard model the interactions between the Higgs particle and
the fermions are proportional to the masses of the fermions, but complete
experimental verification will require new accelerators. The same is true for
the precise experimental exploration of the Higgs potential. We also lack
the reason for the particular choice for the gauge interactions by Nature and
we do not know if there is only one scalar particle.
We mentioned that apart of the continuous symmetries of gauge interac-
tions and space-time, there are the discrete symmetries of charge conjugation,
parity and time reversal, too. We discussed at length the experimental proof
for the chiral characteristics of the electroweak interactions, which means
that weak force violate the parity symmetry maximally. We also argued that
the combined symmetry of CPT is expected to be a fundamental symmetry
of Nature and mentioned the strong constraints on the experimental searches
for CPT violation. As a result, the naive expectation that particle interac-
tions respect time reversal symmetry would imply combined CP symmetry,
which is however weakly violated. While this violation can be explained
in the standard model, there is also another symmetry violation — called
baryon-anti-baryon asymmetry — that the standard model cannot explain.
There are two more big mysteries at the fundamental level in Nature.
The energy density in the Universe comes from several main sources. Some
of these are well known, like electromagnetic radiation (which turns out to
provide a small fraction, not reaching 1 % of the total energy density) and
baryonic matter that has also a quite small contribution of about 5 %. The
Introduction to Particle Physics 211
rest of the energy density is in unknown forms, called dark matter and dark
energy. While these results emerge in cosmological observations, it might
be that the explanations will come from particle physics, but clearly beyond
the standard model, therefore these subjects are omitted from this book.
We devote the last part of our book to the introduction to the quantum
field theoretical basis of the standard model of elementary particle interac-
tions, needed to understand high energy collision events. We explore what
we can learn from studying the theory of the standard model, to what extent
the experimental observations mentioned above can be understood theoreti-
cally, and what calls for a new theory beyond the standard model. In order to
facilitate the comprehension of the mathematical details, most of the compu-
tations are performed step by step, or in some cases left to the exercises. As
prerequisite familiarity with quantum electrodynamics is assumed, although
its basics are summarized.
Our primary goal is to describe collisions at the LHC, i.e. proton-proton
scatterings at high momentum transfer. As the initial state contains parti-
cles composed of coloured constituents of the proton–quarks and gluons–,
quantum chromodynamics (QCD) is the key to the mathematical descrip-
tion. QCD is a predictive theory due to its two main features: (i) asymptotic
freedom that allows for a perturbative description and (ii) the factorization
theorem that separates the short and long-distance physics in a universal
(process independent) way. Thus we put significant emphasis to develop
these two concepts in detail and exhibit the theoretical uncertainties that
are inherent in this theoretical framework, which is crucial to estimate the
theoretical systematics of experimental results correctly. In closing the book
we discuss the basics of the theory of electroweak interactions.
Part IV
Standard model of
elementary particles
213
Chapter 18
MOTTO:
We need something new. We can’t predict
what that will be or when we will find it be-
cause if we knew that, we would have found it
already!
(Stephen Hawking)
N
2 −1
H= a Ta , a ∈ R.
a=1
The constants f abc are the structure constants of the Lie algebra. The
structure constants are completely antisymmetric. For SU(2) the generators
satisfy the well-known angular momentum algebra
+ a b,
J , J = i abc J c .
In this case the structure constants f abc = abc are the Levi-Civita symbols
(if any two of its indices are equal, the symbol is zero, otherwise abc =
(−1) p 123 , with 123 = 1 where p is the number of interchanges of indices
necessary to unscramble a, b and c into the order 1, 2, 3). In the SM the
SU(3)c symmetry is unbroken, this means that SU(3) is a symmetry of the
Lagrangian and also of the vacuum, whereas the SU(2)L ⊗ U(1)Y symmetry
is spontaneously broken to U(1). First we shall discuss the unbroken part
and turn to the broken part in the last chapter.
The gauge theory based on local SU(3)c invariance is called quantum
chromodynamics (QCD). There are many similarities and some important
differences between QED and QCD. In order to see these clearly, we recall
the basics of QED.
n
- .
Lf = f¯j i∂/ − m j f j
j=1
The physical meaning of this question is the following: can we choose the
phase of the fermion field at any x μ at will, without changing observable
quantities? The answer is yes! To make the Lagrangian invariant under
local phase transformations, called gauge transformations, we have to take
the following steps:
i
Aμ (x) → Aμ (x) = Aμ (x) + [∂μ U (x)]U −1 (x) ,
e
U(x) = exp [i θ (x)] .
218 Chapter 18
then
Dμ [A] f j (x) → Dμ [A] f j(x) = U j (x)Dμ [A] f j (x) ,
so Dμ [A] f j (x) transforms the same way as the fermion field f j (x),
hence Dμ [A] is called covariant derivative.
3. The new field requires a gauge-invariant kinetic term:
1
L A = − Fμν [A] F μν [A] Fμν [A] = ∂μ Aν − ∂ν Aμ .
4
L A is a Lorentz scalar and therefore Lorentz-invariant. Since F μν [A]
is gauge invariant (the proof is left as exercise), L A is also gauge
invariant: L A = L A .
The Feynman rules can be read off the action
∫
S = i d4 x(L f + L A) ≡ S0 + SI ,
where ∫ ∫
S0 = i 4
d xL0 SI = i d4 xL I .
L0 contains all terms that are bilinear in the fields, L I contains the rest (called
interactions). The photon propagator Δγμν is the inverse of the bilinear term
in Aμ [Ryder, 1996]. In momentum space we have the condition
+ , ρ
Δγμν (p) · i p2 gνρ − pν pρ = δμ ,
equivalent form (action remains the same) such that Δγμν exists, which
is called gauge fixing. For example, the covariant gauges are defined by
requiring ∂μ Aμ (x) = 0 for any space-time point x μ . Adding
1 - .2
LGF = − ∂μ Aμ , λ ∈ R,
2λ
to L, the action S remains the same. The term LGF can be seen as a
constraint, taken into account in the Lagrangian by a term with Lagrange
multiplier (like in classical mechanics). The bilinear term becomes in this
case
2 νρ 1 ν ρ
i p g − 1− p p ,
λ
with inverse
i pμ pν
Δγμν (p) = − 2 gμν − (1 − λ) 2 .
p p
Of course, physical results must be independent of λ. It is customary
to choose λ = 1 (called Feynman gauge). In covariant gauges unphysical
degrees of freedom (longitudinal and time-like polarizations) also propagate,
which can be avoided by choosing axial (physical) gauges, defined with an
arbitrary, but fixed vector nμ (different from pμ ):
1 - μ .2
LGF = − n Aμ ,
2λ
which leads to
- 2 .
i pμ nν + nμ pν n + λ p2 pμ pν
Δγμν (p, n) = − 2 gμν − + .
p p·n (p · n)2
Since p2 = m2 = 0, we have:
with
pμ nν + nμ pν
dμν (p, n) = −gμν + = μ(λ) (p)ν(λ) (p)∗ , (18.2.3)
p·n λ=1,2
Feynman rules
In the Feynman gauge we have the following Feynman rules:
p
gμν
photon propagator: μ ν = Δγμν (p) = −i p2
p / +m j
p
electron propagator: = Δ j (p) = i p 2 −m2j
μ
μ
vertex: = Γγ f = −ie j eγ μ
j f¯j
Exercise 18.1
Show that in QED the covariant derivative transforms the same way as the
field itself, namely if f (x) → U(x) f (x), then Dμ f (x) → U(x)Dμ f (x) where
Dμ = ∂μ + i e Aμ .
Exercise 18.2
Show that in QED
+ ,
Dμ, Dν = i e Fμν
where Dμ = ∂μ + i e Aμ .
Exercise 18.3
Show that the generators of a special unitary group are traceless and her-
mitian.
Gauge theories in the standard model 221
Exercise 18.4
The generators of the SU(2) group in the fundamental representation are the
Pauli matrices divided by two:
σa 0 1 0 −i 1 0
tf =
a
, σ1 = , σ2 = , σ3 =
2 1 0 i 0 0 −1
The adjoint representation of a group is defined as
(t Ab )ac = i f abc
Tr[tRa tRb ] = δ ab TR
• compute this constant from the explicit form of the fundamental (TF )
and the adjoint (TA) representation.
The quadratic Casimir factor C2 (R) of a representation R is defined by
C2 (R)1 = tRa tRa
a
where ECM is the total incoming energy, Q μ is the total incoming four-
momentum (in the centre-of-mass system ECM = Q2 ), and
n 1n
dd p j
μ
dφn = (2π)d δ d Q μ − pj δ+ p2
j − m 2
j (18.3.5)
j=1 j=1
(2π)d−1
222 Chapter 18
the cross section for the scattering of two incoming particles pa and pb into
n final-state particles obtained as the sum over all possible squared diagrams
and over all possible cuts of these diagrams:
1 1
1
G G
σ(pa pb → p1, . . . , pn ) = .
2s S G,cuts n
The Feynman rules for the cut diagrams are the usual ones with the following
additional rules:
√
1. The sign of explicit factors of i = −1 and directions of fermion
arrows and those of all momenta are reversed in G̃ as compared to G.
• /p j + m j if j is a fermion,
• /p j − m j if j is a anti-fermion,
• dμν (p j ) if j is a (massless) gauge boson.
- .
• /p ± m j 2πδ+ p2j − m2j if j is a fermion/anti-fermion,
• dμν (p j ) 2π δ+ p2j if j is a (massless) gauge boson.
Example: e+ e− → μ+ μ−
We consider the reaction e+ e− → μ+ μ− as a very simple application of the
Cutkosky rules. At leading-order (LO) accuracy there is only one Feynman
diagram, shown in Fig. 18.1, that contributes to the amplitude at lowest order
to this reaction.
In order to describe the kinematics of the
reaction we use the Mandelstam variables e− pe− μ−
(see Section 5.3)
p μ−
γ∗
- .2
s = (pe− + pe+ )2 = pμ− + pμ+ ,
- .2 - .2 q
t = pe − − pμ − = pe + − pμ + , pe+ p μ+
- .2 - .2 e+ μ+
u = pe − − pμ + = pe + − pμ − .
(18.3.6) Figure 18.1: Feynman dia-
gram for e+ e− → μ+ μ− at LO
Furthermore we express the coupling as
e2 = 4π α. The cross section is then given
by
p e+
pμ−
+ − 1 1 + −
σ(e e → μ μ ) = .
2s S p e−
pμ+
Using the Feynman rules we obtain the squared matrix element from this
diagram as
(4πα)2 2 3
|M2 | 2 = 2
Tr /pe+ − me γ α /pe− + me γ β
spin
s
2 3
×Tr /pμ− + mμ γα /pμ+ − mμ γβ
224 Chapter 18
Evaluating the traces1 we obtain for the squared matrix element, summed
over spins
( )
(u − me2 − mμ2 )2 + (t − me2 − mμ2 )2 mμ2 + me2
|M2 | = 8 (4πα)
2 2
+2 .
spin
s2 s
(18.3.7)
We leave the integration over the phase space as an exercise. As mμ 200me ,
we usually neglect the mass of the electron.
Exercise 18.5
Derive the result in Eqn. (18.3.7) and integrate it over the two-particle phase
space dφ2 .
Exercise 18.6
Use Mathematica and the Package Tracer.m (or FORM) to compute the
following traces:
Tr /p2 γν ( /p1 − k/ 1 )γ μ /p1 γμ ( /p1 − k/ 1 )γν
Tr γ μ1 γ μ2 γ μ3 γ μ4 γ μ5 γ μ6 γ μ7 γ μ8 γ μ9 γ μ10
× γμ1 γμ2 γμ3 γμ4 γμ5 γμ6 γμ7 γμ8 γμ9 γμ10
nf
Nc
- .
Lq = q̄ kf i∂/ − m f δkl qlf .
f =1 k,l=1
1The Mathematica package Tracer.m is useful for computing the traces. (see
http://library.wolfram.com/infocenter/MathSource/2987/)
Gauge theories in the standard model 225
f 1 2 3 4 5 6
qf u d s c b t
mf 2.12 MeV 4.69 MeV 92.5 MeV 1.27 GeV 4.2 GeV 172.6 GeV
Table 18.1: The six quark flavours. Their baryon number is B = 1/3.
For light flavours the masses correspond to running MS quark masses (see
definition in section 19.4) at 2 GeV, for heavy flavours c and b the mass
values of Ref. [Bazavov et al., 2018] and for t quark it is the pole mass.
If we apply a transformation q k → q k = Ukl ql , with
⎧
⎪ c −1
2 ⎫
⎪
⎨ N
⎪ ⎬
⎪
Ukl = exp i t θ
a a
≡ exp {it · θ} kl
⎪
⎪ a=1 ⎪
⎪
⎩ ⎭ kl
where θ a ∈ R, then the Lagrangian again remains invariant under the global
SU(Nc ) transformation: Lq (q) = Lq (q ), therefore the colour charges are
conserved. The (t a )kl are Nc × Nc matrices, which constitute the fundamental
representation of the generators T a (called colour-charge operators). Both
satisfy Eq. (18.1.1). For SU(3) the matrices t a are the Gell-Mann matrices.
Can we make Lq (q) invariant under local SU(Nc ) transformations? The
answer is again yes, we can and the steps are similar as in the case of QED:
i
t · Aμ → t · Aμ = U (x) t · Aμ U −1 (x) + (∂μ U (x))U −1 (x) ,
gs
1 a
L A = − Fμν [A] F aμν [A] ,
4
226 Chapter 18
a given by
with the non-abelian field strength Fμν
a
Fμν [A] = ∂μ Aνa − ∂ν Aμa − gs f abc Aμb Aνc ,
789:
a
≡ Aμ × Aν
The necessity and form of the ghost-field Lagrangian can most easily be
derived from the path integral formalism. For a physical motivation we refer
to [Ellis, 1988].
In summary, the complete Lagrangian of QCD is
In Eq. (18.4.8)
Nc
Lq (q f , m f ) = q̄ kf (iγμ Dμ [A] − m f )kl qlf ,
k,l=1
with
Nc2 −1
(Dμ )[A]ab = ∂μ + igs Aμc T c
c=1 ab
and the kinetic term for the gluon field, the gauge field of QCD:
Nc −1 2
1
Lg (A) = − F a [A]F a μν [A] .
4 a=1 μν
The ghost fields are absent if we use physical gauges, which will be our
choice (defined precisely later).
It is clear that there is an unprecedented large number of degrees of
freedom we have to sum over:
1. spin and space-time as in any field theory, not exhibited above,
2. flavour and colour, specific to QCD.
As a result, computations in QCD are rather cumbersome. During the last
25 years a lot of effort was invested to find simpler ways of computing QCD
cross sections and to automate the computations.
Vertices: a, μ
μ, a
quark-gluon: j i
= Γgq q̄ = −igs (t a )i j γ μ
228 Chapter 18
a, α
p
q
a, b, c
three-gluon:
b, β
r
c, γ
≡ Γα, β, γ (p, q, r) =
a,b,c,d
four-gluon: ≡ Γα, β, γ, δ
c, γ d, δ
a, μ
μ, a
ghost-gluon: j
≡ Γgη η̄ = −igs (F a )ab pμ (not needed in
i
physical gauges)
c, γ
We can check that a single four-gluon vertex can be written as a sum of three
diagrams (see exercise). This way the summation over colour and Lorentz
indices factorize completely, which helps automation and makes possible
for us to concentrate on the colour algebra independently of the rest of the
Feynman rules.
Gauge theories in the standard model 229
Exercise 18.7
Show that in SU(N) gauge theories
+ ,
Dμ, Dν = igFμν
a a
T where a
Fμν = ∂μ Aνa − ∂ν Aμa − g f abc Aμb Aνc .
Exercise 18.8
a transforms according to the adjoint representation of SU(N):
Show that Fμν
a
Fμν → Fμν
a
− f abc θ b Fμν
c
.
Exercise 18.9
Show that the four-gluon vertex can be written as a sum of three graphs with
the help of the fake field such that in each graph the colour and Lorentz
indices are factorized.
- .
Tr t a t b ≡ a b TR a b = TR δ
ab . The usual choice is
- .
(F a )bc = F b ca = (F c )ab = −i fabc =
b c
- . . -
where F a with a = 1, . . . , (Nc2 − 1) are Nc2 − 1 × Nc2 − 1 matrices that
again satisfy the commutation relation (18.1.1). The graphical notation in
the adjoint representation is not unique. For the matrix (F a )bc we assume
an arrow pointing from index c to b, opposite to which we read the indices of
(F a ). On the structure constants the indices are not distinguished, therefore
arrows do not appear. However, these are completely antisymmetric in their
indices, so the ordering matters. By convention, in the graphical represen-
tation the ordering of the indices is counterclockwise. The representation
matrices are invariant under SU(N) - transformations.
.
The sums a tiaj t jk a and Tr F a F b have two free indices in the fun-
j j CF CA a
k i k i a b b .
a b b a a b
.
Gauge theories in the standard model 231
Let us first multiply this commutator with another colour charge operator.
Next, sum over the fermion index and then multiply with δik . As a result, we
obtain the resolution of the three-gluon vertex as products of colour charges:
or graphically
TR 1
Nc
We now show some examples of how one can compute the colour algebra
structure of a QCD amplitude, in particular we will also find explicit values
for CF and C A in the exercises. Taking the trace of the identity in the
fundamental and in the adjoint representation we obtain
= Nc , = Nc2 − 1 .
Then, using the expressions for the fermion and gluon propagator corrections,
we immediately find
- .
= CF Nc , = CA Nc2 − 1 .
These graphical rules make the evaluation of colour algebra easy. Never-
theless, nowadays computer algebra codes make computation of colour sums
an automated procedure. For instance, you can type In[1]:= Import["http:
//www.feyncalc.org/install.m"] in a Mathematica session to explore
some code for that purpose.
Exercise 18.10
Consider the process q q̄ → ggg. Compute the colour structures that appear
in the squared matrix element.
Exercise 18.11
Insert a quark loop on a single gluon loop and try to find the value of CF .
Compute CA in a similar way. Hint: use the resolution of the three-gluon
vertex.
Exercise 18.12
Determine the colour factors A,B,C in
A
,
B ,
C .
properties of strong interactions (C, P and T violating strong decays are not
observed).
We have already discussed exact symmetry in colour space: gauge in-
variance. In addition to the classical Lagrangian of Eq. (18.4.8), there exists
an additional gauge invariant dimension-four operator, the Θ-term:
The only difference is that the quark q transforms under the fundamental,
while the gluino λ under the adjoint representation of the gauge group.
An important approximate symmetry of the classical Lagrangian is re-
lated to the quark mass-matrix. Let us introduce the quark flavour triplet
u q
1
ψ = d = q2 ,
s q3
with each component being a four-component Dirac spinor , and the combi-
nations
1- .
P± = 1 ± γ5 , γ5 = iγ0 γ1 γ2 γ3 . (18.7.11)
2
The latter are projections:
P+ P− = P− P+ = 0 , P±2 = P± , P+ + P− = 1 .
a b
exp(iα) exp(−iβ) exp i αa T a exp −i βb T b
a b
∈ UV (1) ⊗ SUL (Nf ) ⊗ UA (1) ⊗ SUR (Nf )
where the matrices T a represent the generators of the group (Nf × Nf
matrices). The transformations (exp (i a αa T a ) , exp (i a αa T a )), acting
as ψ → exp (i a αa T 1) ψ, form
a
- a vector . subgroup
- SUV.(Nf ). How-
ever, the transformations (exp i β T b , exp −i β T b ), acting as
- . b b b b
ψ → exp i b βb T b γ5 ψ, do not form an axial-vector subgroup because
the algebra is not closed,
[T a γ5, T b γ5 ] = i f abc T c 1 (γ52 = 1 γ5 ).
c
Gauge theories in the standard model 237
Exercise 18.13
Show that the classical action of QCD in Eq. (18.4.8) is invariant under scale
transformations (18.7.10). Is the classical Lagrangian also scale invariant?
Exercise 18.14
Show that the classical Lagrangian of QCD in Eq. (18.4.8) is invariant under
the charge conjugation defined by C(ψ) = −iγ2 ψ ∗ .
n 2 3 4 5 6 7 8 9
nd 4 25 220 2485 34300 559405 10525900 224449225
where the sum runs over (n − 1)! non-cyclic permutations of the indices,
denoted by the prime on {. . . }. The usual normalization used in this repre-
sentation is TR = 1. The different traces are orthogonal only at leading order
5Similar, but more complex decomposition can be given also for loop amplitudes.
Gauge theories in the standard model 239
in Nc :
c −1
2
N ∗
1
Tr (t a1 . . . t an ) Tr t b1 . . . t bn = Ncn−2 Nc2 − 1 δ {a } {b } + O
ai =1
Nc2
We introduce the following formalism from Ref. [Mangano and Parke, 1991]:
• ψ (p) is a massless Dirac spinor that satisfies the Dirac equation,
/pψ (p) = 0, and the on-shell condition, p = 0 .
2
|p ± ≡ ψ± (p) , p ±| ≡ ψ± (p)
1
2 |p± q±| = (1 ± γ5 ) γ μ q±| γμ |p±
2
1
⇒ |p± p±| = (1 ± γ5 ) /p and /p = |p+ p+| + |p− p−| .
2
Furthermore,
p±| γμ |k±
μ± (p, k) = ± √ (18.9.12)
2 k ∓ |p±
Moreover,
∗
μ± (p, k) = μ∓ (p, k) , (18.9.13)
μ+ (p, k) ν− (p, k) + μ− (p, k) ν+ (p, k) = dμν (p, k) , (18.9.15)
- .
h (pi, k) · h p j , k = 0 , with h = +, − , (18.9.16)
- .
h (pi, k) · −h p j , pi = 0 , with h = +, − , (18.9.17)
- . =
/h pi, p j phj = 0 , with h = +, − , (18.9.18)
> - .
phj /−h pi, p j = 0 , with h = +, − . (18.9.19)
μ gs √
Γgq q̄ = i √ γ μ the factor of 2 is due to TR = 1,
2
gs
Γαβγ (p, q, r) = i √ Vαβγ (p, q, r) all momenta incoming, (18.9.20)
2
g2 - .
Γαβγδ = i s 2gαγ gβδ − gαδ gβγ − gαβ gγδ .
2
q
p4 p2
e+ μ+
- .2
where si j = pi + p j and in the third line we used momentum conservation
(4/ = −1/ − 2/ − 5/) to rewrite [14] 45 as
; < ; < ; <
[14] 45 = 1 + 4/ 5+ = 1 + −1/ − 2/ − 5/ 5+ = 1 + −2/ 5+
= − [12] 25 . (18.9.22)
The other helicity amplitudes can be found from Eq. (18.9.21), using discrete
symmetry transformations of parity and charge conjugation. Parity transfor-
mation reverses all helicities of the helicity amplitude. It is implemented by
the complex conjugation operation which substitutes i j ↔ [ ji]. Charge
conjugation changes anti-fermions into fermions and vice versa, which in
the present case amounts to interchanging indices 4 with 5 and/or 1 with 2.
Thus,
a4 (+, −, −, +) = a4 (+, −, +, −)| 4↔5 (from charge conjugation)
2 24 2
= −i (18.9.23)
12 54
a4 (−, +, −, +) = a4 (+, −, +, −)| i j ↔[ji] (from parity transformation)
2 [25]2
= −i (18.9.24)
[12] [45]
2 [24]2
a4 (−, +, +, −) = −i . (18.9.25)
[12] [54]
Computing the square of the amplitude, summed over helicities, we obtain
2 + s2
4 s25 24 t 2 + u2
|A| 2 = e4 2 = 8 (4πα)2 ,
helicity
s12 s45 s2
Exercise 18.15
Consider highly relativistic fermions in QED. Show that the electromagnetic
interaction conserves helicity. Is helicity conserved also for axial currents?
Exercise 18.16
Show that for colour ordered amplitudes m(1, . . . , n) ≡ m(p1, 1 ; . . . ; pn, n )
the following statements hold:
Exercise 18.17
Show the following properties of spinor products:
• i j [ j i ] = 2k i · k j
• i| γ μ | j] = [ j | γ μ | i
• Fierz identity: [ i| γ μ | j k | γμ | l] = 2 [ i l ] k j
Exercise 18.18
Consider the matrix element μ∗ (k 1 )ν∗ (k 2 )M μν (e+ e− → γ(k1 )γ(k 2 )) for the
QED process e+ e− → γ(k 1 )γ(k2 ) in leading order of perturbation theory.
Verify the Ward identity
μ
k1 M μν = 0 .
Exercise 18.19
What changes in QCD? We consider the matrix element M (q q̄ → g(k1 )g(k 2 )).
• Verify that
Exercise 18.20
Nevertheless, in QCD the Ward-identity holds thanks to additional (unphys-
ical) fields called ghosts. Square the above amplitude both in covariant and
in physical gauge. Do the results match? Why?
Exercise 18.21
Compute the cross section for e+ e− → q q̄ in leading order in perturbation
theory.
Exercise 18.22
Consider the matrix element M (q q̄ → g(k1 )g(k 2 )) for the QCD process
q1 q¯2 → g(k 1 )g(k2 ) in leading order in perturbation theory.
• Determine graphically the colour structures and the corresponding
colour subamplitudes m(q1, q2, g1, g2 ).
• Decompose the latter into contributions of fixed helicities.
Exercise 18.23
Compute the remaining colour-ordered subamplitudes from the previous
exercise. Use the spinor helicity formalism.
Chapter 19
Electron-positron
annihilation into hadrons
MOTTO:
“Well, in OUR country,” said Alice, still pant-
ing a little, “you’d generally get to somewhere
else — if you ran very fast for a long time, as
we’ve been doing.”
“A slow sort of country!” said the Queen.
“Now, HERE, you see, it takes all the running
YOU can do, to keep in the same place. If you
want to get somewhere else, you must run at
least twice as fast as that!”
1The sixth flavour, the t quark is so heavy that it cannot contribute at CM energies
attained in e+ e− experiments so far.
2There is even a more serious reason that we shall point out at the end of this
chapter.
Electron-positron annihilation into hadrons 249
The loop expansion in terms of the bare coupling gs(0) ≡ 4παs(0) , which is
the coupling that appears in the classical Lagrangian, is as follows:
q2 ( )
α(0) = αs(0) (1) =
(0) (0) 2
|A m = s A m + A m + O (αs ) , (19.2.2)
4π 4π
and .
i − ∫ 1 m2 −
= Γ () p2
dx x − x (1 − x) − iε (19.2.3)
(4π)2− 0 p2
∫ μ
μ dd
I2 = + ,
(2π)d ( − p)2 − m2 2
i − ∫ 1 m2 −
μ
= Γ () p 2
p dx x 2 − x (1 − x) − iε x (19.2.4)
(4π)2− 0 p
where the gamma (or factorial) function Γ () has the property
Γ () 1
Γ () = = Γ (1 + ) .
It is important to distinguish the deviation from four dimensions, represented
by , from the Feynman prescription in the propagator, represented by ε.
Finally we obtain
2 - . 3
iΣ/ (p, m) ≡ i m Σ1 p2, m + /p − m Σ2 p2, m
where
αs 1 4π μ2
Σ1 p , m =
2
Γ (1 + ) (−3C F ) + O 0 ,
4π p2
αs 1 4π μ2
Σ2 p2, m = Γ (1 + ) (+CF ) + O 0 .
4π p2
that are related to the bare ones by simple multiplication, hence the name
renormalization (R):
√
q(0) = Zq q(R) , A(0) = Z A A(R) , η(0) = Zη η(R) ,
m(0) = Zm m(R) , μ gs(0) = μR Zg gs(R) , λ(0) = Zλ λ(R) .
where we simplified slightly the notation, denoting αs(R) with αs . The Z (n, j)
coefficients ( j > 0) can be determined order by order in perturbation theory
from the requirement that the renormalized Lagrangian,
L (R) q(R), A(R), η(R), m(R), g (R), λ(R) =
= L q(0), A(0), η(0), m(0), g (0), λ(0)
− LCT q(R), A(R), η(R), m(R), g (R), λ(R)
(the counter-term for gauge fixing is not included). In principle, the renor-
malization factors of the various terms in the Lagrangian have to be computed
independently for each term. However, one can derive relations among those,
called Slavnov-Taylor identities, which reflect the symmetry of the original
Lagrangian. These relations have been already taken into account, when we
expressed the vertex renormalization factors as combinations of field renor-
malization factors and Zg . We have also omitted the superscript (R) for the
sake of brevity. The constants Z (n,0) are not determined by the requirement
of ultraviolet renormalizability. The simplest choice is Z (n,0) = 0, which de-
fines the minimal subtraction (MS) scheme. In QCD the modified minimal
subtraction (MS) scheme is used, which is defined by
at one loop, with γE being the Euler constant, emerging in the expansion
Γ (1 + ) = 1 + γE − O 2 .
where
αs 1
c = Γ (1 + ) (4π)
4π
thus
The results in Eq. (19.2.8) are valid for any value of the gauge parameter λR .
When computing scattering amplitudes in massless QCD at one-loop
accuracy, the renormalization amounts to the simple substitution
( - . )
αs μ2R β0
(0) 2 MS
αs μ S → αs μR μR 1 − 2 2
+ O αs 2
(19.2.9)
4π
Exercise 19.1
Compute the contribution to the beta function from the fermion loop:
1. Write down carefully the amplitude and compute the trace assuming
n f number of massless quarks..
3. Obtain I2 from
∫
dd l 1
I2 (p, m) = + ,
(2π) (l − p) − m2 l 2
d 2
i − ∫ 1 m2 −
= Γ () p2
dx x − x (1 − x) − iε
(4π)2− 0 p2
(19.2.12)
The contribution to β0 is the coefficient of the 1/ pole without the coupling
factor.
Electron-positron annihilation into hadrons 255
μ2
and defining
∂αs ∂αs (μ2 )
β (αs ) = μ2 =− ,
∂ μ αs(0) fixed
2 ∂t
we obtain the partial differential equation
∂ ∂ - .
− + β (αs ) R et , αs = 0. (19.3.14)
∂t ∂αs
- .
We can solve this equation by defining αs Q2 , the running coupling as
∫ αs (Q2 )
dx
t= . (19.3.15)
αs (μ )
2 β (x)
256 Chapter 19
- .
This
- 2gives
. αs Q2 implicitly for a known, arbitrarily fixed number αs ≡
αs μ . The derivative with respect to the variable t of (19.3.15) gives
- .
1 ∂αs Q2
1 = - - 2. . ,
β αs Q ∂t
which implies - .
∂αs Q2
β αs Q =
2
.
∂t
The derivative of (19.3.15) with respect to αs gives
- .
1 ∂αs Q2 1 ∂αs
0 = - - 2. . − ,
β αs Q ∂αs β(αs ) ∂αs
from which it follows that
- . - - ..
∂αs Q2 β αs Q2
= .
∂αs β(αs )
- - ..
It is now easy to prove that R 1, αs Q2 solves (Eq. (19.3.14)):
- .
∂ ∂R ∂αs Q2 ∂R
− R 1, αs Q 2
=− - 2. = −β αs Q2 - .
∂t ∂αs Q ∂t ∂αs Q2
and
- .
∂ ∂αs Q2 ∂R
β(αs ) R 1, αs Q2 = β(αs ) - .
∂αs ∂αs ∂αs Q2
∂R
= β αs Q2 - ..
∂ αs Q2
- .
Thus if we
- know
. αs Q2 , we automatically
- - 2. . know the Q2 -dependence of
R. To get αs Q , we need the β αs Q function:
2
∂αs
Q2 = β(αs ) . (19.3.16)
∂ Q2 αs(0) fixed
Our discussion so far is valid generally, with the single condition that R
depends on Q2 only and our computations were non-perturbative.
We can obtain the β function in perturbation theory. In the MS scheme
μ2 αs(0) = μ2
R Zg αs . Taking the derivative μ R dμ2 on both sides we obtain
2 2 d
R
( )
d
0 = αs + αs μ2R + β (αs, ) Zg2
dμ2R
Electron-positron annihilation into hadrons 257
and thus
μ2R d Zg2
β (αs, ) = −αs − αs .
Zg2 dμ2R
μ2R d Zg2
Let us define f (αs ) = Zg2 dμR2
. By definition, in the MS scheme Zg depends
- 2.
on μR only through αs μR , so
∂ Zg2 ∂ Zg2 ∂ Zg2
Zg2 f (αs ) = β (αs, ) = −αs − αs f (αs ) .
∂αs ∂αs ∂αs
Since is arbitrary, the equality must hold at each power of . The terms
independent of are
∞ α n−1 2 Z (n,1)
s g
1 · f (αs ) = − αs n
n=1
4π 4π
- 2. - 2.
- 2 .gives αs Q as a function of αs μ if both are small. The coupling
which
αs μ is a number to be extracted from data. We observe that the running
coupling tends to zero at asymptotically large energies,
Q2 → ∞ 1
αs Q2 −→ → 0. (19.3.20)
b0 t
This behaviour is called asymptotic freedom. The sign of b0 plays a crucial
role in establishing whether or not a theory is asymptotically free. If it is
positive, so the β function is negative, then the use of perturbation theory
is justified: the higher Q2 , the smaller the coupling. The coefficient b0
is easiest to compute in background-field gauge where only two diagrams
contribute:
Measuring αs (μ2 )
- . - .
We know αs Q2 if αs μ2 is known. Therefore, we have to measure αs
at some scale μ. The perturbative solution of the renormalization group
Electron-positron annihilation into hadrons 259
Exercise 19.2
The running of the strong coupling is given by Eq. (19.3.16). The pertur-
bative expansion of the QCD beta function is given by Eq. (19.3.18) with
b0, b1 ≥ 0. Determine (i) the expression for the coupling in leading order
(b0 0, b1 = 0) and the corresponding scale Λ0 (see below) (ii) the ex-
pression for the coupling in next-to-leading order (b0 0, b1 0) and the
corresponding scale Λ1 (see below).
Suggested steps:
1
αs (μ) =
K ln( Λμ 2 )
2
where K is a constant.
4. This time the solution cannot be solved for αs analytically. One can
nevertheless find an approximate solution by expanding αs in log Λμ 2 .
2
1
The constant K is not equal to the one in the first part of this exercise.
1 1
αs = ln(c2 +b0 αs )
μ2
K ln 1 + c1 μ2
Λ21 ln
Λ2
1
where γm is called the mass anomalous dimension and the minus sign before
γm is a convention. In perturbation theory we can write the mass anomalous
dimension as
γm (αs ) = c0 αs 1 + c1 αs + O αs2 , (19.4.22)
Electron-positron annihilation into hadrons 261
∂m
Q2 = −γ m (αs ) m Q2 , (19.4.25)
∂ Q2
from which - . ∫ Q2 2
m Q2 dq
ln - 2. = − γ m αs q2 .
m μ μ 2 q 2
γm (αs ) c0 2 3 bc0
− = ⇒ m Q2 = m αs Q2 0 ,
β (αs ) b0 αs
- . + - . , − c0
where we introduced the abbreviation m = m μ2 αs μ2 b0 . At NLO
the solution becomes
2 3 bc0 c0
m Q = m αs Q
2 2 0
1+ (c1 − b1 ) αs Q − αs μ
2 2
+ O αs2 .
b0
262 Chapter 19
2)
In terms of the running coupling and mass R 1, αs (Q2 ), m(Q Q is a solu-
- .
tion of Eq. (19.4.24), proven similarly as R 1, αs (Q2 ) being the solution of
Eq. (19.3.13). Expanding around m(Q2 ) = 0, we obtain
2 ∞ n 2
m(Q2 ) Q 1 m(Q2 ) Q
R 1, αs (Q2 ), = R 2 , αs, 0 + R(n) 2 , αs, 0 .
Q μ n=1
n! Q μ
(19.4.27)
We see from Eq. (19.4.27) that derivative terms are suppressed by factors of
m( Q 2 )
1/Q n at large Q2 . From the dependence of R on Q we can conclude
that the effect of mass is suppressed at high Q2 by its physical dimension
and also by its anomalous dimension, which justifies the assumption about
negligible quark masses for Q2 m2 , i.e. for the first three, four or five
flavours depending on the characteristic energy scale of the collision. The
condition Q2 mt2 is not generally fulfilled for the t quark in colliders at
present.
Exercise 19.3
Find the definition of γm (αS ) in (19.4.21). How does one determine γm
using Zm (μ)? Compute the value of c0 in the perturbative expansion of the
mass anomalous dimension (in Eq. (19.4.22)).
m( Q 2 )
2. If we want to avoid large logarithms of Q , we should consider
physical observables (that is physically measurable quantities) that
Electron-positron annihilation into hadrons 263
with
SMS = (4π) exp(−γE ) ,
which amounts to a simple shift in the amplitude:
= =
(0) (0)
M m = C(μR, μ, q; ) A m q∈N (19.5.29)
=
(1)
M m = C(μR, μ, q; )
- . ( 2 −1 = q β =
)
αs μ2R μR (1) 0 (0)
× SMS A m − A m
4π μ2 2
(19.5.30)
- . q
2
μ2 q
2
− q2
where C(μR, μ, q; ) = 4παs μ2R μ2
R
SMS . The ampli-
(i)
tudes A m are the terms in the expansion of the unrenormalized ampli-
tude, whereas the formal perturbative
= =expansion of the renormalized
(0) (1)
amplitude is |M m = M m + M m + . . .. The factors μ2R /μ2 re-
move the dependence on the dimensional regularization scale μ and
introduce the dependence on the renormalization scale μR .
=
(1)
4. The renormalized theory is ultraviolet finite. We shall see that M m
is divergent also in the infrared. We can use dimensional regularization
264 Chapter 19
- .
≡ A4(1) 1h1 , 2h2 , 4h4 , 5h5 .
4παs μ2
= CF e2 eq Iρσ (p1, p2 )
2s12
; h μ ρ α σ −h < ; h −h <
× 1 1 γ γ γ γ γμ 2 2 5 5 γα 4 4 ,
Electron-positron annihilation into hadrons 265
where
ρσ ρ
I ρσ (p1, p2 ) = I3 (p1, p2 ) − (p1 + p2 )σ I3 (p1, p2 )
∫ ρ σ
ρσ dd
I3 (p1, p2 ) = =
(2π)d 2( − p1 )2 ( − p1 − p2 )2
ρ ρ - ρ ρ.
a0 g ρσ s12 + a11 p1 p1σ + a22 p2 p2σ + a12 p1 p2σ + p1σ p2
∫ ρ
ρ dd ρ ρ
I3 (p1, p2 ) = = a1 p1 + a2 p2 .
(2π) d 2( − p1 ) ( − p1 − p2 )2
2
(19.6.32)
∫
dd 1
I2 (p) =
(2π) d 2( − p)2
i −
= −p2 Γ () B (1 − , 1 − ) (19.6.33)
(4π)2−
i 1 Γ2 (1 − ) Γ (1 + ) 2 −
= −p ,
(4π)2− Γ (2 − 2)
∫
dd 1
I3 (p1, p2 ) =
(2π) d 2(
− p1 ) ( − p1 − p2 )2
2
i1 Γ2 (1 − ) Γ (1 + )
=− (−s12 )−1− (19.6.34)
(4π)2− 2 Γ (1 − 2)
1 1 − 2
= I2 (p1 + p2 ) ,
s12
where we used h1 = −h2 . Therefore, we need to compute a0, a1, a2 and a12
only. From Eqs. (19.6.35) and (19.6.36) we find
∫
1 dd ( − p1 )2 − ( − p1 − p2 )2 + s12
a1 =
s12 (2π) d 2(
− p1 )2 ( − p1 − p2 )2
⎡ ⎤
1 ⎢⎢ ⎥
⎥
= ⎢ I2 (p1 + p2 ) − I2 (p1 ) +s12 I3 (p1, p2 )⎥
s12 ⎢ 789: ⎥
⎢ ⎥
⎣ =0 ⎦
1 1
= − 1 I2 (p1 + p2 ) , (19.6.43)
s12
∫ 2 − ( − p )2
1 dd 1
a2 =
s12 (2π)d 2 ( − p1 )2 ( − p1 − p2 )2
⎡ ⎤
1 ⎢⎢ ⎥
⎥
= ⎢ I2 (p2 ) −I2 (p1 + p2 )⎥
s12 ⎢789: ⎥
⎢ ⎥
⎣ =0 ⎦
1
= − I2 (p1 + p2 ) . (19.6.44)
s12
From Eqs. (19.6.37) and (19.6.38) we find
where the sums run over all coloured external legs. Note that
2 2
μ 1 μ 1 1 π2
= + iπ − Θ (sik ) + O()
−sik 2 |sik | 2 2
where the second term on the right hand side is related to the +i prescription
in the Feynman propagator. The flavour constant γi are γq = 32 CF and
γg = 12 β0 . T i denotes the colour charge associated with the emission of a
gluon from parton i,
T i = Tic |c ⇒
c1, . . . , ci, . . . cn, c |T i | b1, . . . , bi, . . . , bn = δc1 b1 . . . Tcci bi . . . δcn bn
a = F a and T a = t a are the representation matrices in the adjoint
where Tcb cb kl kl
and fundamental representations. The colour algebra is as follows:
Ti · Tk = Tk · Ti if i k and T 2i = Ci (CA or CF ) .
3The I n operator is known also with finite quark masses and also for the two-loop
amplitude.
270 Chapter 19
n
for any a and consequently i=1 T i |M n = 0. We denote the general colour
connected squared matrix element as
2
M n | T i · T k |M n ≡ M n (i,k) .
Γ2 (1− )Γ(1+ )
In our example q = 0 , n = 2 , s12 = s > 0 , Γ(1−2 ) = 1
Γ(1− ) +
- 3.
O so
2
(4π) μ 1 3
I , μ ; p1, p2
2
= −2CF − CF . (19.6.47)
Γ (1 − ) −s12 2
> = > =
= M4(0) I 2 ()M4(0) − 8CF M4(0) M4(0) .
Electron-positron annihilation into hadrons 271
= 2
(0) ; (0) (0) <
.
Two diagrams contribute to M5 : M5 M5 =
We compute them first using the helicity formalism and considering the
μ μ μ μ μ
crossing symmetric channel 0 → p1 + p2 + p3 + p4 + p5 (the new label ‘3’
refers to the gluon). Both diagrams contribute to the only colour structure
tic13i2 and thus a single helicity amplitude has the form
A5 1h1 , 2h2 , 3h3 , 4h4 , 5h5 = e2 eq tic13i2 a5 (h1, h2, h3, h4, h5 ) ,
i 2
a5 (+, −, +, −, +) = − [31] 12 [15] 42 + [31] 32 [35] 42
s45 23 s13
24 2
=i 12 [15] + 32 [35]
23 31 45 [45]
789:
24[45]
24 2
= 2i
23 31 45
(19.6.51)
25 2
a5 (+, −, +, +, −) = −2i .
23 31 45
[14] 2
= −2i
[23] [31] [45]
and
[15]2
a5 (+, −, −, −, +) = 2i . (19.6.53)
[23] [31] [45]
Electron-positron annihilation into hadrons 273
The amplitudes in the other four cases (for h1 = − and h2 = +) are obtained
from parity inversion
a5 (h1, h2, h3, h4, h5 ) = a5 (−h1, −h2, −h3, −h4, −h5 ) | i j ↔[ji]
and give a factor of 2 in the squared matrix element, which is then given by
2 2 + s2 + s2 + s2
s14
(0) 24 15 25
A5 = e4 eq2 Nc CF 8 .
hel.& col.
s s s
23 13 45
Q μ Qν
H μν = −H1 (s) g μν + H2 (s) .
s
L μν describes the leptonic part and is equal to (we take into account the
negative sign due to the fermion loop on the left hand side)
274 Chapter 19
1 e2
−L μν = μ ν= Tr (γ μ k/ 1 γν k/ 2 )
4 4
- μ μ .
= e2 k 1 k 2ν + k1ν k 2 − k1 · k2 g μν .
gμν L μν = e2 (d − 2) k 1 · k2 = e2 (1 − ) s = 4πα s (1 − ) ,
1 1
dim σtot = dim = dim 2 L μν ,
s s
so that dim H1 (s) = 0 and thus H1 (s) must be a constant. Hμν is also U(1)
(of QED) gauge invariant and so
g μν Hμν
H1 (s) = H2 (s) ≡ H(s) = − .
d−1
There is one cut diagram that contributes to the hadron tensor Hμν (s) at LO
in perturbative QCD. It is formally the same as for the lepton tensor, only
the fermion lines refer to quarks of momenta p1 and p2 . Thus,
∫
1 - μ μ .
−H μν = dφ2 (p1, p2 ; Q) e2 eq Nc 4 p1 pν2 + pν1 p2 − p1 · p2 g μν ,
2s q
(19.6.54)
where dφ2 is the two-particle phase-space measure (see Eq. (18.3.5) for
n = 2) and summation
√ extends over quark flavours whose masses are much
smaller than s. From Eq. (19.6.54) we find
e2 q eq2 Nc ∫
(0)
−H (s) = − 2s (2 − d) dφ2
2 (d − 1) s
∫
d−2
= Nc e2 eq2 dφ2 (19.6.55)
q
d−1
α
= Nc eq (1 + O ()) .
2
3 q
Electron-positron annihilation into hadrons 275
Then
1 4πα2
σtot = lim L μν 2 − gμν H (0) (s) = Nc eq2 ,
→0 s 3s q
The virtual correction emerges from cuts that lead to the loop×tree diagrams:
We already know that only the vertex correction is non-zero, which gives
(cf. Eqs. (19.6.47) and (19.6.48))
CF αs 1 4π μ2 2 3
HV (s) = − 2 − − 8 + π 2 + O () H (0) (s) .
2π Γ (1 − ) s
The real correction comes from cuts that lead to the tree×tree diagrams:
where
∫ ∫
g μν dφ3 d1 (1, 2, 3)μν = dφ3 μ μ
∫
1
= − 4παs μ e2 2
Nc CF dφ3 (p1, p2, p3 ; q)
eq2
q
2s
1 2 3
× 2 Tr γμ − /p2 γ μ /p1 + /p3 γα /p1 γ α /p1 + /p3
s13
276 Chapter 19
(the negative sign is due to the fermion loop). We can evaluate the trace
using
γμ − /p2 γ μ = (d − 2) /p2 , γα /p1 γ α = − (d − 2) /p1,
which leads to
+ ,
Tr [. . .] = − (d − 2)2 (p1 + p3 )α (p1 + p3 )β Tr /p2 γα /p1 γβ
- .
= − (d − 2)2 (p1 + p3 )α (p1 + p3 )β 4 p1α p2β + p1β p2α − p1 · p2 gαβ
= − (d − 2)2 [2s13 (s12 + s23 ) − 2s13 s12 ] = −8 (1 − )2 s13 s23
and thus
∫
1
dφ3 g μν d1 (1, 2, 3)μν = 4παs μ2 e2 eq2 Nc CF
q
2s
789:
the other diagrams have the same prefactor
∫
s23
× dφ3 (p1, p2, p3 ; q) 8 (1 − )2 .
s13
Similarly ∫ ∫
g μν dφ3 d2 (1, 2, 3)μν = dφ3 μ μ
∫
−1
∝ dφ3 (p1, p2, p3 ; q)
s13 s23
⎡ ⎤
⎢ ⎥
⎢ ⎥⎥
⎢
× Tr ⎢⎢ γμ − /p2 γν − /p2 − /p3 γ μ /p1 γν /p1 + /p3 ⎥⎥ .
⎢ 789: ⎥
⎢ ⎥
⎢(4−d)[ p/ γν ( p/ + p/ )]−2( p/ + p/ )γν p/ ⎥
⎣ 2 2 3 2 3 2 ⎦
ν
Using the equality γν γα γβ γ = 4 gαβ − (4 − d) γα γβ , we find for the trace
* 2 3
Tr [. . .] = (4 − d) 2 (s12 + s13 ) Tr /p2 /p1 + /p3
2 3
− (4 − d) Tr /p2 /p2 + /p3 /p1 /p1 + /p3
* 2 3
− 2 2s12 Tr /p2 + /p3 /p1 + /p3
2 3
− (4 − d) Tr /p2 + /p3 /p2 /p1 /p1 + /p3
Electron-positron annihilation into hadrons 277
2
= 2 (4 − d) (s12 + s13 ) 2 (s12 + s23 ) + s23 s13
3
− (s12 + s13 ) (s12 + s23 ) + (s12 + s13 + s23 ) s12
2 3
− (4 − d)2 s23 s13 − s12 (s12 + s13 + s23 ) + (s12 + s23 ) (s12 + s13 )
− 4s12 2 (s12 + s13 + s23 )
= −8s12 s + 4 [2s12 s + 2s13 s23 ] − 4 2 2s13 s23
= −8 (1 − ) [s s12 − s13 s23 ] ,
where s12 + s13 + s23 = s. Collecting the contributions from the four cut
diagrams, we finally have
∫ 2
1
−HR (s) = g μν d1 (1, 2, 3) + d1 (2, 1, 3)
d−1
3
+ d2 (1, 2, 3) + d2 (2, 1, 3)
μν
∫
1
= 4παs μ2 e2 eq2 Nc CF dφ3 (p1, p2, p3 ; q)
q
2s
d−2 s23 s13 s s12
×4 (1 − ) + +2 − 2
d−1 s13 s23 s13 s23
where the three-body phase space can be rewritten as
d−3
dφ3 (p1, p2, p3 ; q) = (2π)3−2d 2−1−d q2 dd−2 Ω dd−3 Ω
d−4
× (y12 y13 y23 ) 2 dy12 dy13 dy23
× δ (1 − y12 − y13 − y23 ) [dy123 δ (1 − y123 )] ,
s
with scaled invariants yi j = si j and y123 = y12 + y13 + y23 = 1. dd−1 Ω is the
measure of the hyper-surface element in d dimensions
∫ d
2π 2
dd−1 Ω = Ωd = .
Γ d2
Figure 19.1: Region of integration for the real correction
Then
CF αs 1 4π μ2
HR (s) = H (0) (s)
2π Γ (1 − ) s
∫
× dy12 dy13 dy23 δ (1 − y12 − y13 − y23 ) (y12 y13 y23 )− (19.6.57)
y23 y13 y12
× (1 − ) + +2 − .
y13 y23 y13 y23
The region of integration is shown in Fig. 19.1. The integrand is singular
at y13 = 0 or y23 = 0. The singularities are regularized by the (y13 y23 )−
factor in the phase space measure.
The phase-space integral in HR (s) (equation (19.6.57)) can be computed
easily (see exercise) and it is equal to
2 3 19
+ + − π 2 + O () . (19.6.58)
2 2
We find that the sum of the real and virtual corrections is finite,
HR (s) + HV (s) CF αs 3 αs
= + O () = + O () . (19.6.59)
(0)
H (s) 2π 2 π
Electron-positron annihilation into hadrons 279
Thus we see that σtot , a totally inclusive quantity is infrared finite (i.e.
infrared safe) at one loop.
The natural question at this point is whether there exist other (less inclu-
sive) infrared-safe quantities?
Exercise 19.4
In the lectures we have seen that the 3-point Green’s function I (3) (p1, p2 )
can be expressed–up to a "fairly simple" rational polynomial–in terms of
the two-point function I (2) (p1 + p2 ). Why was this relation so simple?
In this exercise we will derive this relation I (3) (p1, p2 ) = P(d)I (2) (p1, p2 )
(P(d) is a rational polynomial in d) from so called integration-by-parts
(IBP) identities. We do this without computing any integrals explicitly. In
dimensional regularization the following identities hold (generalization of
Gauss’ law):
∫
∂ μ
dd k
k I (k, p, m, . . .) = 0 ,
(2π) ∂k μ d
∫ d
d k ∂ μ
p I (k, p, m, . . .) = 0 .
(2π)d ∂k μ
μ
(p12 = p1 + p2, p21 = p22 = 0, gμ = d) in order to reduce the three-
point function to two-point functions; the propagators of the two-point
functions might acquire higher exponents.
4This method of computing integrals has the advantage that it is easily automated.
The systems of equations can however become large.
280 Chapter 19
• Find two equations that reduce the (two) two-point functions with one
squared propagator each down to the so-called "master integral"
∫ d
(2) d k 1
I (p1 + p2 ) = .
(2π) k (k − p12 )2
d 2
• Check the relation using the integrals evaluated explicitly in the main
text.
Exercise 19.5
Show that the d-dimensional three-particle phase space for q → p1 + p2 + p3
can be expressed in terms of the Lorentz-invariants si j = (pi + p j )2 as
Hints:
• The d-dimensional volume measure in spherical coordinates is given
by
dd+1 p = E dE d d pE = E d E E d−1 dd Ω
dd Ω = (sin θ 1 )d−1 dθ 1 dd−1 Ω .
• Show that
1 s12 s13 s23
sin2 θ 1 = ,
4 q2 E12 E22
where θ 1 is the angle between p1 and p2 .
Exercise 19.6
s
Let yi j = qi2j . Using the previous exercise, compute the real correction to
the process e+ e− → q q̄:
CF αS 1 4π μ2
HR (s) = H (0) (s)
2π Γ(1 − ) s
∫
× dy12 dy13 dy23 δ (1 − y12 − y13 − y23 ) (y12 y13 y23 )−
y23 y13 y12
× (1 − ) + +2 − .
y13 y23 y13 y23
Electron-positron annihilation into hadrons 281
Hint: Transform the triangular integration region into the unit square
and evaluate the B (Euler β) functions.
d3 p s 1 1
= Es dEs dcos θ dφ Es dEs dθ 2 dφ .
2Es 2 4
Therefore in the cross section we find logarithmic singularities both in the
dθ 2
soft and the collinear limits: dE s
Es or θ 2 . These are called the infrared
singular limits. In dimensional regularization the logarithmic singularities
appear as poles: ∫
−1− 1
dy13 y13 =− .
Thus, the singular behaviour arises at kinematically degenerate phase space
configurations, which at the NLO accuracy means that one cannot distinguish
the following configurations: (i) a single hard parton, (ii) the single parton
splitting into two nearly collinear partons, (iii) the single parton emitting a
soft gluon (on-shell gluon with very small energy). Then an answer to the
question posed at the end of the previous section is given by the Kinoshita-
Lee-Nauenberg (KLN) theorem:
In massless, renormalized field theory in four dimensions, transition rates
282 Chapter 19
are infrared safe if summation over kinematically degenerate initial and final
states is carried out.
For the e+ e− → hadrons process, the initial state is free of infrared
singularities that can emerge only off coloured particles (partons). Typical
infrared-safe quantities are (i) event shape variables and (ii) jet cross sections.
like thrust even NNNLL is known), but the discussion of this technique is
beyond the scope of these lectures.
Exercise 19.7
The thrust variable T is used to characterize final states in e+ e− →hadrons.
For m particles in the final state it is defined by Eq. (19.7.60). The momenta
are taken in the centre of mass system.
Exercise 19.8
Compute the thrust distribution at LO accuracy. Hints:
• Write the squared matrix elements for three
√ partons in the final state
using scaled energy variables xi = 2Ei / s = 1 − y jk (i, j, k = 1, 2, 3,
with i j k).
• Convince yourself that the value of thrust is T = max{x1, x2, x3 }.
• Find the region of integration in terms of x1 and x2 .
• Find the region where T = x1 , x2 and x3 and integrate the fully
differential cross section with the δ(T − T(x1, x2 )) function.
predictions where NLO and NLL are matched. The curve at NLO accuracy
gives a good description of the data measured by the ALEPH collaboration
[Barate et al., 1998], but only for ycut > 0.01. As αs ln2 (100) = 2.5, for
smaller values of ycut resummation is indispensable. However, the resummed
prediction is not expected to give a good description at large ycut because
in the resummation formula only the collinear approximation of the matrix
element is used. Matching the two predictions gives a remarkably good
description of the data over the whole phase space.
At hadron colliders the k T -algorithm needs modifications. First, instead
of energy, the boost-invariant measure of hardness, transverse momentum
R2
is used to define the distance between tracks, di j = min(p2T i, p2T j ) Ri2j where
Ri2j = (yi − y j )2 + (φi − φ j )2 (distance in rapidity–azimuthal-angle plane),
R is a small positive real number, and we need a distance from the beam
diB = p2T i , too. Also, the algorithm needs modification because either
di j or diB can be the smallest distance. If a di j is the smallest value,
then i and j are merged, while if the smallest is a diB , then momentum pi
becomes a jet momentum and is removed from the tracks to be clustered. We
then call jet candidates with transverse momentum larger than a predefined
value, pT i > ER resolved jets. The merging rule changes as well. In the
usual merging we add transverse momenta, pT (i j) = pT i + pT j and we add
rapidities y and azimuthal angles φ weighted, y(i j) = (wi yi + w j y j )/(wi + w j )
and φ(i j) = (wi φi + w j φ j )/(wi + w j ). The weight can be wi = pT i , p2T i , ET i ,
or ET2 i . Such a merging is boost invariant along the direction of the beam.
The parameter R plays a similar role as dcut in electron-positron annihilation
or the cone radius R in the cone algorithms: the smaller R, the narrower the
jet. Popular values in the experimental analyses are R = 0.4 − 0.7.
The iterative k T -algorithm is infrared safe and resummation of large log-
arithmic contributions of the form αsn ln2n and αsn ln2n−1 is possible, which
is a clear advantage from the theoretical point of view. The logarithms are
those of 1/R and/or Q/ER , Q being the hard process scale. By taking ER
sufficiently large in hadron-hadron collisions, we avoid such leading con-
tributions from initial-state showering and the underlying event (collisions
not belonging to the hard parton-parton scattering), so these terms are deter-
mined by the time-like showering of final-state partons (when the virtuality
of the parent parton is always positive). Particles within angular separation
R tend to combine and particles separated by larger distance than R from all
other particles become jets. The algorithm assigns a clustering sequence to
particles within jets, so we can look at jet substructure.
Nevertheless, at the TEVATRON experiments the kT -algorithm did not
286 Chapter 19
become a standard for several reasons. The jets have irregular, often weird
shapes as seen on Fig. 17.59 (left) because soft particles tend to cluster
first (even arbitrary soft particles can form jets). As a result there is a
non-linear dependence on soft particles, energy calibration and estimating
acceptance corrections are more difficult. The underlying event correction
depends on the area of the jet (in η − φ plane). It was also very expen-
sive computationally, so experimenters had a clear preference of cone algo-
rithms. The breakthrough occurred with Refs. [Cacciari and Salam, 2006,
Cacciari et al., 2008] where variants of the kT -algorithm and an improved,
fast implementation was introduced. The distance formula was modified to
R2
di j = min(p2nT i, pT j ) R 2 (n = −1, 0, 1). Infrared safety is independent of n,
2n ij
μ μ
second case the two collinear momenta pi and p j are replaced by a single
momentum pμ denoting the collinear direction. The cross section is the phase
space integral of the squared matrix element weighted by the jet function,
for instance ∫
1 (n)
σ= dφn |M n | 2 Jm (p1, . . . pn ) .
2s
In our case of electron-positron annihilation into partons, for an arbitrary jet
function J (n) we have the leading-order term
∫
(0) 2 d−2
H (s, J) = Nc e 2
eq dφ2 (p1, p2 ; q) J (2) (p1, p2 ) ,
q
d − 1
For the total cross section J (n) = 1 and cancellation of infrared poles takes
place as in Eq. (19.6.59). For thrust T (2) (p1, p2 ) = 1 (by momentum conser-
vation the two outgoing quarks are back-to-back, this direction defines the
thrust axis event-by event), and
Exercise 19.9
Generalize the requirements for the jet function for a computation at NLO
accuracy as given in Eqs. (19.7.62)–(19.7.64) to NNLO accuracy.
Mm ∝ M mTis gs ū(pi, si )γ μ i
pi sis
μ
ps → 0 1 pi
Tis gs ū (pi, si ) γ μ /pi = Tis gs ū (pi, si ) ,
sis pi · ps
where Tis is the colour charge operator of the soft gluon s, gs is the strong
coupling, ū(pi, si ) is the Dirac adjoint spinor for the outgoing quark labelled
with i and sis = (pi + ps )2 = 2pi · ps . In taking the limit, we used the
μ
anti-commutation relation (18.2.2) to write γ μ /pi = − /pi γ μ + 2pi and the
μ
p
Dirac equation of the massless bi-spinor, ū(pi ) /pi = 0. The factor pi ·i ps is
the “square root” of the eikonal factor Sik (s) = si2ss si kk s .
The emission of a soft gluon off an external gluon (in light-cone gauge)
is given by
μ, s
ps
Mm
λ, b
ν, a
∝ M m ε μ (ps, n) 1
si s d λλ (pi + ps, n)
pi
ν
× Γνμλ (−pi , −ps, pi + ps ) ε (pi , n)
asb
We use d λλ (pi + ps, n) (pi + ps )λ = 0 and εν (pi, n) pi ν = 0, thus
1 λλ ν ps → 0
d (pi + ps, n) Γνμλ
asb
(−pi , −ps, pi + ps ) ε (pi , n)
sis
piμ λλ piμ
Tbs gs d (pi, n) gλν εν (pi, n) = Tbs gs (−ε λ (pi, n)).
pi · ps pi · p s
where cs is the colour index of the soft gluon s, Ŝs is an operator which takes
the soft limit and keeps the leading singular term in 1/λ, and J μ (s) is given
by
m
pk μ
J μ (s) = T ks .
k=1
pk · p s
The soft gluon can be emitted from any of the external legs, therefore the sum
in the previous formula runs over all external partons. A soft quark leads
to an integrable singularity because the fermion propagator is less singular
than that of the gluon. Colour conservation implies that the current J μ (s) is
conserved,
m
μ
ps J μ (s) |M m = T ks |M m = + + = 0.
k=1
> =
(0) (0)
Then the soft limit of M m M m is as follows:
∝ + + ... .
(19.8.65)
The gauge terms give zero contribution on on-shell matrix elements due to
gauge invariance.
290 Chapter 19
μ μ
2
k iT nμ μ μ
2
krT nμ
pi = zi pμ + kiT − , pr = zr pμ + krT −
zi 2 p · n zr 2 p · n
μ μ
where k iT + krT = 0 and zi + zr = 1. The momentum pμ is the collinear
direction and
fir → fi + fr
q → q + g
g → q + q̄
g → g + g
We compute explicitly the first case and leave the second and third ones as
exercise. For the case of a quark splitting into a quark and a gluon we have
pr
/pi + /pr μ ν /p + /pr
= CF gs2 μ2 γ /pi γ dμν (pr , n) i
pi sir sir
/ /r
p + p / / n/ + n/ /pi /pr /pi + /pr
p p
= CF 4παs μ2 i −γ μ /pi γμ + r i .
sir pr · n sir
Using (19.9.66)
−γ μ /pi γμ = (d − 2) /pi , /pi /pi = pi 1 , /pi /pr /pi = sir /pi − pi /pr = sir /pi ,
2 2
we find
/pi + /pr −γ μ /pi γμ /pi + /pr = (d − 2) sir /pr ,
/pr /pi n/ = − /pi /pr n/ + sir n/ = /pi n/ /pr − 2 /pi pr ·n + sir n/ =
= −/n /pi /pr + 2pi ·n /pr − 2 /pi pr ·n + sir n/ .
Electron-positron annihilation into hadrons 291
Then
/pi + /pr /pr /pi n/ + n/ /pi /pr /pi + /pr =
= 2 /pi + /pr pi ·n /pr − pr ·n /pi + pi · pr n/ /pi + /pr
2
= 2 pi ·n sir /pi − pr ·n sir /pr
3
+pi · pr 2 (pi + pr )·n /pi + /pr − (pi + pr )2 n/
+ ,
= sir 4pi ·n /pi + 2pi ·n /pr + 2pr ·n /pi − sir n/ .
Substituting these results and then the Sudakov parametrization of the mo-
menta into Eq. (19.9.66) we obtain
( )
z 2
p i ! pr 1 i
CF 4παs μ2 2 (1 − ) zr + 4 + 2zi + 2zi + O (k T ) /p
sir zr
1 zi
= CF 8παs μ2 2 + (1 − ) zr /p
sir zr
789:
1+z 2
i
1−z i − (1−zi )
Similarly to the soft case we can define an operator Ĉir which takes the
collinear limit and keeps the leading singular (O(1/kT2 )) terms:
1 > (0) =
(0) 2 (0) (0)
Ĉir M m+1 = 8παs μ2 M m (p, . . .) P̂qg (zi, zr , k T ; ) M m (p, . . .) .
sir
(19.9.67)
The kernel P̂qg , called Altarelli-Parisi splitting kernel for the process q →
q + g. It is diagonal in the spin-state of the parent (splitting) parton:
zi
s| P̂qg |s = CF 2 + (1 − ) zr δss . (19.9.68)
zr
Similar calculations give the splitting kernels for the gluon splitting pro-
cesses, which display azimuthal correlations of the parent parton
( μ
)
(0) μν
k T kTν
μ| P̂q q̄ (zi, zr , kT ; ) |ν = TR −g + 4zi zr 2 , (19.9.69)
kT
(0)
μ| P̂gg (zi, zr , kT ; ) |ν =
( μ
)
μν zi zr k T k Tν
= 2CA −g + − 2 (1 − ) zi zr 2 . (19.9.70)
zr zi kT
292 Chapter 19
The soft and collinear limits overlap when the soft gluon is also collinear to
its parent parton:
2 2z j 2
(0) (0)
Ĉ jr Ŝr M m+1 (pr , . . .) = −8παs μ2 M m(j,k) (. . .)
s
kj jr r
z
2 z j (0) 2
= 8παs μ2 T 2j M m . (19.9.71)
s jr zr
The notation for the splitting kernels in these lectures is different from the
usual notation in the literature. Usually, P̂i(0) j (z, k T ; ) denotes the splitting
kernel for the process fi (p) → f j (zp) + fk ((1 − z)p), which does not lead
to confusion for 1 → 2 splittings because the momentum fraction of parton
j determines that of parton k as their sum has to be one, z j + zk = 1.
For splittings involving more partons, it is more appropriate to introduce
as many momentum fractions zi as the number of offspring partons, with
the constraint i zi = 1, and use the flavour indices to denote the offspring
partons in the order of the momentum fractions in the argument. For 1 → 2
(0)
splittings this means the use of P̂ir (zi, zr , kT ; ) for the splitting process
fk (p) → fi (zi p) + fr (zr p). The flavour of the parent parton fk is determined
uniquely by the flavour summation rules, q + g = q, q + q̄ = g + g = g.
These flavour summation rules are unique also for 1 → 3 splittings.
Exercise 19.10
Compute the Altarelli-Parisi-splitting kernel P̂qg (z) for the process q → qg
from the collinear limit of the matrix element for the process e+ e− → q q̄g:
M3 (e+ e− → q q̄g)2 = 8παs μ2 M2 (e+ e− → q q̄)2
1 y23 y13 y12
× (1 − ) + +2 − .
s y13 y23 y13 y23
Exercise 19.11
The Altarelli-Parisi splitting kernel P̂q q̄ (z) for the process g → qi q̄r is
Electron-positron annihilation into hadrons 293
1
8παs μ2 M n(0) (p, . . .) P̂q(0)q̄ (z, k T )M n(0) (p, . . .)
sir
1
= 8παs μ2 M n(0) (p, . . .) μμ P̂q(0)q̄ (z, kT ) νν M n(0) (p, . . .)
sir
Compute μ P̂q(0)q̄ (z, kT ) ν in leading order in k T .
Exercise 19.12
Derive the flavour summation rules for 1 → 3 splittings.
Exercise 19.13
Compute the soft limit of Eq. (19.9.67) and the collinear limit of Eq. (19.8.65).
σNLO = σ LO + σ NLO ,
where σ LO is the integral of the fully differential Born cross section over the
available phase space defined by the jet function, while σ NLO is the sum of
the real and virtual corrections:
∫
σ LO = dσ B Jm ({p}m ) ,
m
∫ ∫
σ NLO = dσ R Jm ({p}m+1 ) + dσV Jm ({p}m ) .
m+1 m
Both contributions to σ NLO are divergent in four dimensions, but their sum
is finite for infrared-safe jet functions.
The factorization of the squared matrix elements in the soft and collinear
limits allows for a method independent of the processes and observables
to regularize the real corrections in their singular limits. The essence of
294 Chapter 19
The definition of the approximate cross section is not unique and the best
choice may depend on further requirements, which we do not discuss here.
We also skip the precise definition of the momenta p̃μ which are obtained
by mapping the (m + 1)-particle phase space onto an m-particle phase space
times a one-particle phase space. A widely used general subtraction scheme
that can be used also for processes including massive partons with smooth
massless limits is presented in Ref. [Catani et al., 2002] where these defini-
tions are given explicitly. This method uses the factorization of the squared
Electron-positron annihilation into hadrons 295
matrix element in the soft and collinear limits. The challenge posed by
the overlapping singularity in the soft-collinear limit is solved by a smooth
interpolation between these singular regions.
The factorization properties of Eqs. (19.8.65) and (19.9.67) play other
very important roles in pQCD. The numerical implementation of the squared
matrix element is a process prone to errors. Testing the factorization in the
kinematically degenerate phase space regions serves a good check of the
implementation. The computation is even more difficult for the virtual
corrections. Similar factorization holds for those, which more recently has
been utilized to find the virtual corrections (“collinear bootstrap”). The
factorized form of the squared matrix element can be used in resumming
logarithmically enhanced terms at all orders, or in devising a parton shower
algorithm for modelling events (see Sect. 21.3). The splitting kernels that
appear in the collinear factorization have a role in the evolution equations of
the parton distribution functions (see Sect. 20.7).
The state of the art in making precision predictions assaults on the one
hand the full automation of computations at NLO, and on the other the realm
of next-to-next-to-leading order (NNLO) corrections. The automation of
computing jet cross sections at NLO accuracy has been accomplished and
several programs are available with the aim to facilitate automated solutions
for computing jet cross sections at NLO accuracy:
• aMC@NLO (http://amcatnlo.web.cern.ch),
• BlackHat/Sherpa (https://blackhat.hepforge.org),
• FeynArts/FormCalc/LoopTools (http://www.feynarts.de),
• GoSam (https://gosam.hepforge.org),
• HELAC-NLO (http://helac-phegas.web.cern.ch),
• MadGolem
(http://www.thphys.uni-heidelberg.de/˜ lopez/madgolem-corner.html).
Deeply inelastic
lepton-proton scattering
MOTTO:
If you want to learn about nature, to appre-
ciate nature, it is necessary to understand the
language that she speaks in.
(Richard P. Feynman)
20.1 Kinematics
Perturbative QCD stems from the parton model that was developed to un-
derstand deeply inelastic lepton-hadron scattering (DIS). The purpose of
those experiments was to study the structure of the proton by measuring
the kinematics of the scattered lepton. In Fig. 17.61 (left) we show a real
event in the H1 experiment at the HERA collider. The value of Q2 , which
is the modulus squared of the momentum transfer between the lepton and
the proton is 21475 GeV2 >> 1 GeV2 , signifying that the scattering is well
in the deeply inelastic region. The parton model interpretation of the event
is shown in Fig. 17.61 (right): the lepton is scattered by an angle θ due to
the exchange of a virtual photon with one of the constituents of the proton
297
298 Chapter 20
(a parton).
The measurement is inclusive from the point of view of hadrons (X
means any number of hadrons that are not observed separately), thus the
process can be described in pQCD.
The DIS kinematics is described by the following variables (using the
notation of Fig. 17.61 (right)):
The summation runs over all possible (unobserved) hadronic final states. We
factorize the phase space and the squared matrix element into two parts, one
for the lepton (denoted by L) and one for the hadrons (denoted H):
d3 k 1 e4
dφ = dφ L dφ H , dφ L = , |M | 2 = 4 L μν Hμν .
(2π)3 2E 4 spin Q
∫
1 1
dσ̂ = dφ2 |M | 2 ,
2ŝ 4 spin
1 eq2 e4 ŝ2 - .
|M | 2 = 4 L μν Q μν = 2eq2 e4 4 1 + (1 − y)2 .
4 spin Q Q
Deeply inelastic lepton-proton scattering 301
Also Q2 = 2p · q = 2ξP · q, so p 2 = Q2 (ξ/x − 1). Then the two-particle
phase space is
d3 k d4 p - .
dφ2 = 2πδ+ p 2 (2π)4 δ4 (k + p − k − p)
(2π) 2Ek (2π)
3 4
(20.3.3)
dϕ E x
= dE d cos θ 2 δ(ξ − x) ,
2π 4π Q
√
2yx
or using E = ŝ (1 − y) and cos θ = 1 − , we obtain dφ2 =
2 ξ(1 − y)
dϕ y ŝ
dy dx δ(ξ − x) . The differential cross section in x and y
(4π)2 Q2
d2 σ̂ 4πα2 + ,1
= 2
1 + (1 − y)2 eq2 δ(ξ − x) . (20.3.4)
dx dy Q 2
Comparing Eqs. (20.2.2) and (20.3.4), we find the parton model predic-
tions
Exercise 20.1
Compute the contribution to the DIS cross section in Eq. (20.2.2) with the
exchange of a transversely polarized photon. Hint: Use Eq. (18.2.3) for the
numerator in the propagator of the transversely polarized photon and the
Callan-Gross relation in Eq. (20.3.5). Can you identify the result with any
of the terms in Eq. (20.2.2)? What is the source of the remainder?
302 Chapter 20
The combination of the measurements and sum rules gives separate infor-
mation on the quark distributions in the proton fq (x). The result of such
measurements performed by the NMC collaboration [Arneodo et al., 1997]
is shown in Fig. 17.62 (left) together with a fit to the data by the CTEQ
collaboration [Pumplin et al., 2002]. The parton distributions deduced from
the fit are shown in Fig. 17.62 (right).
We can infer the proton momentum from the measurements. The surpris-
ing result is that quarks give only about half of the momentum of the proton,
∫1
q 0 dx x fq (x) 0.5. By now we know that the other half is carried by
gluons, but clearly the naïve parton model is not sufficient to interpret the
gluon distribution in the proton. With our experience in perturbative QCD
we try to compute radiative corrections to the quark process to see if that
helps to find the role of the gluon distribution.
Exercise 20.2
It is not feasible to use a neutron target experimentally. Instead deuteron
is used which is the bound state of a proton and a neutron. The corre-
p p
sponding structure function is F2d (x) = 12 (F2 (x) + F2n (x)), with F2 and F2n
given in Eqs. (20.4.7) and (20.4.8), respectively. Which combination of the
structure function on proton and deuteron targets gives the u- and d-quark
distributions?
αs dE dθ αs dz dk T2
∝ σh+g σh 2CF = σh CF .
p
σh θ
zp π E θ π 1 − z k T2
Integrating over z up to one and over k T we find soft and collinear divergences.
In studying perturbative QCD we found that these infrared singularities in
the final state cancel against infrared divergences in the virtual correction for
infrared safe quantities:
p αs dz dk T2
σh
∝ σh+V −σh CF .
π 1 − z kT2
If there is a coloured parton in the initial state, then the splitting may
occur before the hard scattering and the momentum of the parton that enters
the hard process is reduced to zpμ , so
p zp σh αs dE dθ
θ ∝ σh+g (p) σh (zp)2CF
(1 − z)p
π E θ
αs dz dk T2
= σh (zp) CF .
π 1 − z kT2
Integrating over z up to one and over k T we again find soft and collinear
divergences. The corresponding poles multiply σh (zp), while in the vir-
tual correction the poles multiply σh (p), irrespective whether the infrared
divergence is in the initial or final state:
p p αs dz dk T2
σh
∝ σh+V −σh (p) CF .
π 1 − z k T2
Deeply inelastic lepton-proton scattering 305
The sum of the real and virtual corrections then contains an uncancelled
singularity,
∫ Q2 ∫
αs dkT2 1
dz
σh+g + σh+V CF [σh (zp) − σh (p)] ,
π m2g k T20 1−z
789: 789:
infinite if mg =0 finite
i.e. it is zero in the gluon channel because the virtual photon does not interact
with the gluon directly. At one order higher in αs we find (the diagrams below
show the contributions to each term: one diagram at LO, one loop diagram
and two real radiation diagrams at NLO)
d2 σ̂ αs Q2 q
F̂2,q (x) = = eq x δ(1 − x) +
2
P (x) ln 2 + C2 (x) ,
qq
dx dQ2 F2 4π mg
(20.5.10)
and (only one real radiation diagram with gluon splitting in the initial state
contributes)
d2 σ̂ αs Q2 g
F̂2,g (x) = = eq x 0 +
2
P (x) ln 2 + C2 (x) ,
gq
dx dQ2 F2 q
4π mq
(20.5.11)
function. The space-like splitting function P ji (x) is obtained from the time-
like splitting kernel P̂ jk in four dimensions through the following steps: (i)
crossing the kernel for time-like splitting into space-like splitting:
1
P̂(0), i k (x, kT ) = −(−1)2s(i)+2s(j) x P̂(0)
¯
jk x T
, k ; = 0 (20.5.12)
relabelling superscripts P̂(0), ik (x, kT ) → P j i (x) and (iv) adding the con-
¯ ¯ ¯
tribution from the one-loop diagram (see exercise). The flavours satisfy the
flavour summation rule fi = f j + fk . We see that at NLO the prediction
for F̂2 does not contain ultraviolet divergences. The final state infrared di-
vergences also cancel, but an un-cancelled singularity remains in the initial
state, regularized with a small mass (mg or mq ) here.
The hard scattering function is not measurable, only the structure function
is physical:
(
F2,q (x, Q ) = x
2
eqi fq(0)
2
i
(x)
i
∫ )
αs 1
dξ (0) qq x Q2 q x
+ f (ξ) P ln 2 + C2 .
2π 0 ξ qi ξ mg ξ
Exercise 20.3
g
Compute the coefficient C2 (x) in Eq. (20.5.11).
where the summation runs over the light quark flavours.1 The long distance
physics is factored into the parton distribution functions that depend on
the factorization scale μF . The short distance physics is factored into the
hard scattering cross section that depends on both the factorization and the
renormalization scales. Both scales are arbitrary, unphysical scales. The
term K i j defines the factorization scheme. It is not unique, finite terms can
be shifted between the short and long distance parts, but it is important that it
must be chosen the same in all computations (the MS scheme is the standard
where K i j = 0).
Defining the convolution in x-space,
∫ 1 ∫ 1
( f ⊗ g)(x) ≡ dy dz f (y) g(z)δ(x − yz) ,
0 0
Exercise 20.4
The regularization of the splitting function at x = 1 is achieved by the
+-prescription defined by
∫ 1 ∫ 1
- .
dx[g(x)]+ f (x) = dxg(x) f (x) − f (1)
0 0
for any smooth test function f (x). The contribution of the loop corrections
has the same kinematics as the LO one, so it has to be proportional to
δ(1 − x). Thus the complete regularized splitting function has the form
1 + x2
P (x) = CF
qq
+ Kδ(1 − x) .
(1 − x)+
We can obtain the parton distribution for a ‘quark in a quark’ from Eq. (20.6.13)
by the substitution fq(0) (x) → δ(1 − x)
αs qq μ2
fq (x, μF ) = δ(1 − x) + P (x) ln F2 .
2π mg
hand side of the renormalization group equation is not exactly zero, but may
contain terms that are higher order in perturbation theory. Only infinite order
is expected to give exact independence of the scales. The renormalization
group equation gives the missing piece of information needed to make the
theory predictive.
To write the renormalization group equation, we introduce Mellin trans-
∫1
forms defined by f (N) ≡ 0 dx x N −1 f (x), which turns a convolution into a
real product:
∫ 1 ∫ 1 ∫ 1
N −1
dx x dy dz δ(x − yz) f (y) g(z) =
0 0 0
∫ 1 ∫ 1
= dy dz (yz) N −1 f (y) g(z) = f (N)g(N) .
0 0
So F2,q (N, Q2 ) = x i eq2 i fi (N, μF )F̂2,i (N, μR, t) is independent of μF , ex-
pressed as
dF2 - . - .
μF = 0 = O αsn+1 in PT at O αsn .
dμF
Let us explore the consequences of this renormalization group equa-
tion. For simplicity, we assume the existence of only one quark flavour,
F2,q (N, Q2 ) = x eq2 i fq (N, μF ) F̂2,i (N, μR, t). Then the renormalization group
equation reads
d fq dF̂2,q
F̂2,q (N, t) (N, μF ) + fq (N, μF ) (N, t) = 0 .
dμF dμF
d ln fq d ln F̂2,q
μF (N, μF ) = −μF (N, t) ≡ −γqq (N) , (20.7.15)
dμF dμF
where γqq (N) is called the anomalous dimension because it acts as a factor
−γ (N )
μF q q in the dimensionless function fq (N, μF ). Taking the Mellin moment
of Eq. (20.6.13) and then its derivative with respect to μF , we obtain that the
anomalous dimension is
d ln fq αs (μR ) qq - .
γqq (N) = −μF (N, μF ) = − P (N) + O αs2 , (20.7.16)
dμF π
thus it is the Mellin transform of the splitting function, which can be com-
puted in PT. Equation (20.7.15) implies that the scale dependence of the
310 Chapter 20
parton distribution functions can be predicted in PT. This together with the
universality of parton distribution functions makes perturbative QCD pre-
dictive: we can measure the parton distribution functions in one process at
a certain scale and then use it in another process at another scale to make
predictions.
How shall we choose the renormalization and factorization scales? If we
want to avoid large logarithms that spoil the convergence of the perturbative
series, the scales should be chosen near the characteristic physical scale of
the process Q, e.g. μ2R = μ2F = Q2 . Then the renormalization group equation
becomes
2 d ln fq 1 - 2.
Q (N, Q ) = − γqq N, αs Q ,
2
(20.7.17)
dQ2 2
which is the Mellin transform of
- .
2 d fq αs Q2 qq - 2.
Q (x, Q 2
) = (P ⊗ fq Q (x) . (20.7.18)
dQ2 2π
Our discussion was highly simplified by considering only one quark flavour
and neglecting the mixing of partons. If we make the full computation we
obtain the celebrated formula
- .
2 d fi αs Q2 i j - 2.
Q (x, Q 2
) = P ⊗ f j Q (x) , (20.7.19)
dQ2 2π j
Exercise 20.5
Compute the anomalous dimension γqq (x) using Eq. (20.7.16).
312 Chapter 20
Hadron collisions
MOTTO:
I cannot define the real problem, therefore I
suspect there’s no real problem, but I’m not
sure there’s no real problem.
(Richard P. Feynman)
Before collision the colliding partons may emit other partons nearly
collinear with the beam. These collinear emissions in the initial state give
rise to divergences that can be factored into the renormalized parton dis-
tribution functions. After collision few energetic partons appear that may
emit less energetic partons and each develops showers of partons. Emissions
into almost the same direction as the original parton occur with enhanced
probability (due to the collinear divergence) as well as emissions of soft glu-
ons. This is represented in Figure 17.64 by red quark and gluon lines. Both
factorization and parton showering can be described from first principles
based upon known physics of QCD and are universal, meaning that these are
independent of the process and observable. We have seen how factorization
works, but have not discussed how parton showers are modelled with shower
Monte Carlo (SMC) programs [Sjostrand et al., 2006, Corcella et al., 2001].
We mentioned marginally how the large logarithms emerging in the final
state splittings can be resummed. Such resummations give improved pre-
diction for the cross section (as seen in Fig. 17.60), but does not simulate
events.
Parton showers still only give a description of events in terms of quarks
and gluons, whereas detectors detect only hadrons. We do not know how
to compute hadronization, the transition from quarks and gluons to hadrons,
from first principles. Yet the idea of local parton-hadron duality (see,
e.g. Ref. [Dokshitzer et al., 1991]) provides some sort of theoretical un-
derstanding. It states that
after accounting for all gluon and quark production down to scales
ΛQCD , the transition from partons to hadrons is essentially local in
phase space, i.e. there is no rearrangement of energy and momentum.
Thus the directions and momenta of hadrons will be closely related to those
of the partons and the hadron multiplicity will reflect the parton multiplicity,
too. This is illustrated by the green lines with dots.
In addition to the energetic partons in the initial state, there are also low-
energy ones that may collide, which is energy and process dependent. This
low-energy physics is described in models of underlying event, which are
also part of modern shower Monte Carlo programs. The underlying event
produces low-energy partons. Also at the end of the shower low-energy
partons emerge. As QCD confines partons, these partons turn into hadrons
before detection, a process called hadronization. We do not have a theory
of hadronization based on first principles. Instead, shower Monte Carlo pro-
grams include models that describe hadronization in a process independent
way. These models contain parameters that are fixed experimentally.
Hadron collisions 317
21.4 Conclusions
In the last three chapters we discussed the theoretical basis of interpreting
the results of high-energy collider experiments. We discussed how pertur-
bative QCD can be made predictive and also the main uncertainties in the
predictions. We used the following key ingredients in this tour:
(i) gauge invariance that allows us to write down the Lagrangian and
which predicts many important features of the theory;
(ii) renormalization that cancels ultraviolet divergences systematically
order-by-order in perturbation theory and introduces a dimension-
ful scale into even the scaleless Lagrangian of massless QCD, leading
to scaling violations of one-scale observables that would be scale
independent in the classical theory;
(iii) asymptotic freedom at high energies emerging from the quantum struc-
ture of the theory and the non-abelian nature of the gauge group;
(iv) need for infrared safety, emerging from asymptotic freedom, to en-
sure that the infrared divergences, associated with unresolved parton
emission, cancel between real and virtual contributions, allowing the
perturbative calculation of jet cross sections, without a detailed under-
standing of the mechanism by which partons become jets;
(v) factorization that makes possible to use perturbative QCD to calculate
the interactions of hadrons, since all the non-perturbative physics gets
factorized, into parton distribution functions;
(vi) evolution and universality of parton distribution functions that allows
us to extract those measuring cross sections in one process, like DIS,
and then used to predict the cross sections for any other process. Again,
this factorization introduces a scale dependence into the parton model
so that the structure functions of DIS, and other one-scale observables
become scale dependent.
These features make perturbative QCD predictive, without forcing us to
solve the theory at all possible scales: unknown or not calculated high and
low-energy effects can be renormalized, factorized and cancelled away.
In the last chapter we study the sector of the standard model with bro-
ken symmetry. The photon appears here, which is massless like the gluon.
However, the photon does not carry colour charge and does not have self in-
teraction, it only mediates the electromagnetic interaction between fermions.
318 Chapter 21
MOTTO:
And as they drifted up, their minds sang with
the ecstatic knowledge that either what they
were doing was completely and utterly and
totally impossible or that physics had a lot of
catching up to do.
G = SU(2)L ⊗ U(1)Y .
Here L stands for left (and later R for right) polarized particle currents and
Y is the hypercharge. The field content in one family (we consider the first
319
320 Chapter 22
one here) is
u
ψ1 = ψ2 = uR , ψ3 = dR quarks
d L
νe
ψ1 = ψ2 = 0 , ψ3 = eR− leptons
e− L
Here τa represent the Pauli matrices, while α = (α1, α2, α3 ) and β are real
numbers. The number y j denotes the eigenvalue of the U(1) generator Y /2,
called weak hypercharge, for field ψ j (the factor 12 is included to maintain
the traditional Y = B + S, baryon number + strangeness definition for the
first three quarks). The Lagrangian for one family (family replication is
implicitly understood) is
3
L= / (j) ψ j (x),
iψ̄ j (x) D (22.1.1)
j=1
(j)
with Dμ = ∂μ + δ j1 ig T · W μ + ig y j Bμ . The Lagrangian in Eq. (22.1.1) is
invariant under the gauge transformation of the ψ j fields, provided the four
gauge fields introduced in the covariant derivative transform according to
the rules
G 1
Bμ → Bμ (x) = Bμ (x) − ∂μ β(x) (22.1.2)
g
G i + ,
T · W μ (x) → T · W μ (x) = U(x) T · W μ (x) U † (x) + ∂μ U(x) U † (x)
g
(22.1.3)
Electroweak sector of the standard model 321
where U(x) = exp [iT · α (x)]. The gauge invariant kinetic term for these
vector fields is
1 1
L B,W = − Bμν Bμν − W μν · W μν,
4 4
with Bμν = ∂μ Bν − ∂ν Bμ ≡ ∂[μ Bν] and W μν = ∂[μ W ν] − g W μ × W ν . Bμν
is invariant under G transformations, while T · W μν transforms covariantly:
G
T · W μν → U(x) T · W μν U † (x).
In quantum field theory the mass of a particle is the position of the pole
in the inverse two-point function in momentum space.1 Gauge symmetry
forbids such mass terms for gauge bosons. In case of the Dirac Lagrangian
the fermion propagator Δ j (q) = i q2 /+iε depends only on the momentum q,
q
i.e. does not contain mass. Thus, the mass of the fermion is zero. An extra
term proportional to iψ̄ψ would lead to a pole-position different from zero.
However, fermion masses must also be absent because
where
1 Wμ3 Wμ1 − iWμ2
T · Wμ = .
2 Wμ + iWμ2
1 −Wμ3
The off-diagonal terms lead to charged-current interactions
g
LCC = − √ Wμ† [ūγ μ (1 − γ5 ) d + ν̄γ μ (1 − γ5 ) e] + h.c. ,
2 2
with the stepping operators
1 1
Wμ = √ Wμ1 + iWμ2 , Wμ† = √ Wμ1 − iWμ2 ,
2 2
1This is the pole mass definition that differs from the renormalized mass discussed
earlier. The two definitions are related in perturbation theory.
322 Chapter 22
and the letters u, d, e, ν stand for the spinor of the relevant particle. The
diagonal terms lead to neutral-current interactions
LNC = − / 3 + g yj B
ψ̄ j g T 3W / ψj . (22.1.4)
j
We would like to identify the fields Wμ3 and Bμ with the observed fields Zμ
and Aμ (the latter denoting the electromagnetic field), but both Wμ3 and Bμ
are massless, therefore an arbitrary combination of them is possible:
3
Wμ cos θ W sin θ W Zμ
= .
Bμ − sin θ W cos θ W Aμ
The angle θ W is called Weinberg (or weak mixing) angle. Equation (22.1.4)
becomes then
LNC = − / g T 3 sin θ W + g y j cos θ W
ψ̄ j A
j
+ Z/ gT 3 cos θ W − g y j sin θ W ψ j
(Z)
≡ LQED + LNC .
The term with Aμ should give QED and for this we require
Y
g sin θ W = g cos θ W = e and Q= + T3 (22.1.5)
2
where Q is the electromagnetic charge operator, with
Qu/ν 0
Q1 = Q2 = Qu/ν Q3 = Qd/e .
0 Qd/e
experiments and for this reason we exclude it from the standard model. We
finally obtain
μ (Z) −e μ
LQED = −e Aμ JE M , LNC = Z μ JZ ,
sin θ W cos θ W
μ
JE M = Q j ψ̄ j γ μ ψ j ,
j
μ μ μ
JZ = T3 − sin2 θ W Q j ψ̄ j γ μ ψ j ≡ J3 − sin2 θ W JE M .
j
Exercise 22.1
The part of the standard model Lagrangian that determines the interaction
between fermions and electroweak gauge bosons is given by
1 - .
L= ψ̄L γ μ i∂μ + g τ · Wμ + g yBμ ψL + ψ̄R γ μ i∂μ + g yBμ ψR .
ψ
2 ψ
L R
The fields Wμ and Bμ denote the gauge fields of the weak isospin and the
hypercharges. The photon field Aμ and the Z boson field Zμ can be written
as linear combinations of these fields:
Aμ = Bμ cos θ W + Wμ3 sin θ W
Zμ = −Bμ sin θ W + Wμ3 cos θ W .
The relations of the various couplings are given by Eq. (22.1.5)
e = g sin θ W = g cos θ W .
Show that one obtains the QED Lagrangian by substituting the photon and
the Z boson fields for the W 3 and the B fields in the Lagrangian density
above.
324 Chapter 22
(b) has a degenerate set of states with minimal energy, which transform
under G as the members of a given multiplet.
The system arbitrarily selects one of these states as the ground state of
itself, we say in these cases that the symmetry is spontaneously broken. The
simplest example in field theory is the G = U(1) Brout-Englert-Higgs (BEH)
model. We consider the spontaneous symmetry breaking (SSB) of such a
model and move later on to the case of standard model.
Let us assume a Lagrangian L of a complex
scalar field φ which is symmetric under global U(1)
transformations
L = ∂μ φ∗ ∂ μ φ − V (φ)
where
V (φ) = μ2 φ∗ φ + λ (φ∗ φ)2 Figure 22.1: The Mex-
is the potential. L is invariant under global ican hat potential
φ → φ = e−i eθ φ
From the last equation we can read off that the mass of φ2 vanishes, mφ2 2 = 0,
which is a realization of Goldstone’s general theorem:
If a Lagrangian L is invariant under a continuous symmetry group
G but the vacuum is only invariant under a subgroup H ⊂ G, then
there exist as many massless spin-0 particles (Goldstone bosons) as
broken generators.
In the U(1) case there is one broken generator. The corresponding Goldstone
boson is φ2 that describes excitations around the flat direction in the potential.
We are more interested in the case when the symmetry is local. Then the
gauge-invariant Lagrangian is
∗ 1
L = Dμ φ Dμ φ − V(φ) − Fμν F μν,
4
with Dμ [A] = ∂μ − i eAμ . We parameterize φ(x) around the ground state as
1 ξ (x)
φ (x) = √ [v + h(x)] ei v .
2
326 Chapter 22
2
1
Aμ (x) = Aμ (x) − ∂μ ξ(x).
ev
In terms of the transformed fields (we drop the prime to ease the notation)
the Lagrangian becomes
1- . 1 1 1
L= ∂μ h (∂ μ h) + e2 v2 A2 + e2 v h (x) A2 + e2 h2 A2 − Fμν F μν
2 2 2 4
v λ
− V √ − λ v2 h2 (x) − λ v h3 (x) − h4 (x) ,
2 4
from which we conclude
√ the following spectrum: a scalar Higgs particle h
with mass mh = 2λ v2 , cubic and quartic self interactions, and a massive
U(1) field Aμ with mass m A = e v. The massive vector field has three
components Ai (i = 1, 2, 3) and as A2 = A20 − 3i=1 A2i and A0 is unphysical
due to gauge fixing, the mass terms for the Ai field components have the
correct (negative) sign in the Lagrangian. The excitations of this gauge
field interact with the Higgs particle (third and fourth terms). We could
identify the U(1) symmetry with that of QED, but then the vacuum would
be electrically charged, which contradicts to our observations.
This ground state is electrically neutral, and so it does not couple to the
photon, which consequently remains massless as shown below. Eliminating
the unitary phase (in other words “choosing unitary gauge”), the scalar
kinetic term becomes
2
1 - . μ - . 0
∂μ h (∂ h) + ig T · W μ + ig yφ Bμ
2 v+h
1 2
2
MW Wμ†W μ + M Zμ Z μ
2 Z
with
vg vg MW
MW = and MZ = = . (22.3.7)
2 2 cos θW cos θW
Thus we find that SSB predicts the masses of the vector bosons (note that the
Aμ field of QED is massless), with a non-trivial relation between MW and
MZ :
MW g
= cos θW = .
MZ g + g 2
2
The vector boson propagator is that of a massive vector field (with three
components propagating),
328 Chapter 22
p α pβ
g αβ − p
β M2
α = −i p2 −M 2 +iM V
, where V = W ±, Z.
MV V V ΓV
GF
HI = √ J μ Jμ, (22.3.8)
2
with sin θ C 0.22 (θ C is called Cabibbo’s angle as this mixing of the quark
fields was introduced by Nicola Cabibbo).
In the low energy limit, the propagator of the massive gauge bosons
becomes
p
g αβ
α β +i 2 ,
MV |p| MV MV
and from the electroweak theory we can recover the V − A model of weak
interactions (recall Table 4.1) with
GF g2 1
√ = = . (22.3.9)
2 8 M W
2 2 v2
where the fine structure α is not to be confused with the index α above. Using
the Feynman rules of the electroweak theory, from the tree-level diagrams
of Fig. 22.2,
Electroweak sector of the standard model 329
ν ν
Z W
q q q
Figure 22.2: Neutral and charged current neutrino scattering at tree level
we can deduce the following predictions for neutral and charged current
neutrino scattering:
8π g 4
σ (ν q → ν q) ≡ σNC (s) ∝ s a 2
q + a v
q q + v q ,
2
3 MZ4 cos4 θ W
8π g 4
σ (ν q → l q ) ≡ σCC (s) ∝ 4
s.
4 MW
(22.3.10)
(aq and vq denote the axial-vector and vector couplings of quark q). The
ratio of the two cross sections depends only on sin θ W (at LO accuracy):
4
σNC MW 4 2
= aq + aq vq + vq2 .
σCC MZ cos θ W 3
789:
f (sin θW )
Fitting this prediction to the measured value of the ratio one finds sin2 θ W =
0.231. Then with α(MZ ) 1/128, obtained from the measured value
α( 0) 1/137 by running the coupling, we can predict the values of the
gauge boson masses (at lowest order in perturbation theory),
SM prediction at LO measured
MW c2
GeV 80.23 80.385
MZ c2
GeV 91.49 91.188
Higgs sector:
1 1 M2 M2
Lh = ∂μ h ∂ μ h − Mh2 h2 − h h3 − h2 h4
2 2 2v 8v
+ −μ 1 2 μ h (x) h2 (x)
+ MW Wμ W
2
+ MZ Z μ Z 1+2 + .
2 v v2
√
where Mh = −2μ2 = 2λ v. The Higgs boson was observed at the
LHC by the ATLAS and CMS experiments. Its mass was measured to be
approximately 125.1 GeV/c2 , so λ 0.13. The Lagrangian also contains
cubic and quartic Higgs self interactions as well as vector boson-Higgs
interactions, proportional to MV2 . In order to prove that the standard model
Higgs boson was found the experiments have measured these couplings and
showed that these are consistent with the standard model predictions.
Exercise 22.2
In addition to explaining the mass of vector bosons, there is another puzzle
that the existence of the Higgs boson can explain. In the standard model
scattering of the electroweak vector bosons W + and W − is possible via
exchange of a photon or a Z boson both in the s and t-channels, as well as
through four-point interaction. Compute the cross section for longitudinally
polarized vector bosons assuming these processes at LO accuracy using the
Feynman rules in Sect. 22.7. What can you observe√ as the total centre-of-
mass energy of the colliding vector bosons grow, s → ∞?
With SSB, the colliding vector bosons can also interact through an off-
shell Higgs boson. Compute the contribution of these diagrams to the cross
section. How does the cross section now behave at large s?
Exercise 22.3
Compute the decay width of muon decay, μ− → e− ν¯e νμ (i) in the four-
fermion interaction theory of Fermi, (see Eq. (22.3.8)), (ii) in the standard
model (the necessary Feynman rules can be found in Sect. 22.7).
Exercise 22.4
Compute the cross sections in Eq. (22.3.10)
Electroweak sector of the standard model 331
dc = d cos θ C + s sin θ C
instead of d quark only as seen in Eq. (22.1.1). However, if we use the field
(Z)
dc in LNC instead of d, we have
(Z) + μ ,
LNC ∝ cos θ C sin θ C Zμ dγ¯ (vd − ad γ5 ) s + s̄γ μ (vd − ad γ5 ) d
¯ μ (vd − ad γ5 ) d + · · · ,
+ cos2 θ C Zμ dγ
which gives flavour-changing neutral currents (FCNC) in the first two terms
with similar strength as Z d d¯ itself (third term), in contradiction to experi-
mental facts. For instance, K L = [ds̄] decays via neutral current into μ μ̄,
while K + = [us̄] decays via charged current into νμ μ̄ according to the fol-
lowing diagrams:
d μ− u νμ
Z0 W+
and
s̄ μ+ s̄ μ+
Γ (K L → μ+ μ− )
- . = 2.8 · 10−9,
Γ K + → μ+ νμ
which means that writing the Lagrangian in terms of the rotated fields is
equivalent to writing it using the original ones. In the latter case flavour
changing neutral currents do not appear, so these must be absent also when
we write the Lagrangian in terms of the rotated field. This prediction of
the charm, based on the absence of FCNC (in 1970) was confirmed by the
discovery of the particle J/ψ, a cc̄ bound state (in 1974). The problem
comes back at one loop if the c quark is too heavy, mc 1 GeV. The two
diagrams
μ− d μ−
d
W− W−
c νμ
u W+
νμ W+
μ+ μ+
s̄ s̄
cancel because both are proportional to cos θ C sin θ C , but with different
signs. This cancellation is incomplete if the second diagram is suppressed by
2 , which happens when mc 1 GeV, suggesting
1
mc in the propagator p2 −m
c
that the mass of the c quark should be around 1 GeV.
so
- . i - .
δ ψ̄L · φ = −i T · + i T · − i yL + ψ̄L · φ
2
1 - .
= −i yL − ψ̄L · φ
2
where yL = y1 = y3 + 12 (from Eq. (22.1.6)). Thus,
- . - .
δ ψ̄L · φ = −i y3 ψ̄L · φ +- . ,
⇒ δ ψ̄L · φ dR = 0.
δ dR = i y3 dR
It is left as an exercise to check the gauge invariance of the second term,
proportional to cu . In a unitary gauge
1 0
φ (x) = √ ,
2 v + h (x)
from which
1 + ,
LY = √ (v + h(x)) cd d¯L dR + cu ūL uR + ce ēL eR + h.c.
2
We see that there are mass terms with mi = − c√i v where i = d, u, e:
2
h(x) + ,
LY = − 1 + ¯ + mu ūu + me ēe .
md dd
v
The standard model predicts that the coupling to the Higgs boson of the
fermions is proportional to the fermion masses. This prediction awaits
experimental confirmation. The first direct measurements of Higgs-fermion
couplings with large uncertainties are appearing at the time of writing this
book.
Exercise 22.5
Let us define a conjugate field φ̃ to the Higgs field φ by φ̃ = iτ2 φ∗ , where τ2
(0) ∗
φ
is the second Pauli matrix, so φ̃ = . Check that the infinitesimal
−φ(+) ∗
- .
transformation of φ̃ under SU(2) L × U(1)Y is δ φ̃ = i T · − 2i φ̃. Prove
that the
+- second , in Eq. (22.5.11) is invariant under SU(2) L × U(1)Y ,
. term
i.e. δ ψ̄L · φ̃ u R = 0.
334 Chapter 22
where the fermion doublets are weak eigenstates and the couplings c(i) jk
are
arbitrary elements of 3 × 3 matrices (in flavour space). After spontaneous
symmetry breaking the Lagrangian becomes
h (x) ? ¯
LY = − 1 + dL M d dR + ūL M u uR + l¯L M l lR + h.c. ,
v
where d , u and l are vectors in flavour space. The mass matrices are given
by
- . v - . v - . v
Md i j = −ci(d)
j √ , Mu i j = −ci(u)
j √ , Ml i j = −ci(l)j √ .
2 2 2
Their diagonalization determines the mass eigenstates. If none of the fermion
masses are zero (actually they are positive), M is invertible. Furthermore,
the flavour space is finite dimensional. Then M can be decomposed as
M = H · U
where U U † = 1, H = H † and H is positive definite. (A general n × n
complex matrix has 2n2 independent real components, while both a unitary
and a hermitian matrix has n2 independent real components.) The hermitian
matrix H can thus be diagonalized (its diagonal form is denoted by M):
M = S† M S U ,
with S S † = 1 and M has positive diagonal entries:
M d = diag (md, ms, mb ) ,
M u = diag (mu, mc, mt ) ,
- .
M l = diag me, mμ, mτ .
Electroweak sector of the standard model 335
the CKM matrix into the definition of the quark wave functions, which is
reflected by the usual parametrization of the CKM matrix:
V Vus Vub
ud
V = Vcd Vcs Vcb
Vt d Vts Vtb
1 0 0 c 0 s13 eiδ c s12 0
13 12
= 0 c23 s23 0 1 0 −s12 c12 0
0 −s23 c23 −s13 e
iδ 0 c13 0 0 1
(22.6.13)
|V | =
0.97446 ± 0.00010 0.22452 ± 0.00044 (3.65 ± 0.12) · 10−3
+0.00010
0.22438 ± 0.00044 0.97359−0.00011 (42.14 ± 0.76) · 10−3 .
+0.24 −3 (41.33 ± 0.74) · 10−3
(8.96−0.33 ) · 10 0.999105 ± 0.000032
(22.6.14)
Exercise 22.6
Prove that the CKM-matrix is real if there are only two families.
Electroweak sector of the standard model 337
1 1 2 2
MH g
MW = gv , MZ = g + g 2 v , λ= , tan θ W = ,
2 2 2v 2 g
1
v2 = √ ,
2GF
√ 4παem
g = 2 2 MZ G F 1 + 1 − √
2 2
,
2MZ2 GF
2
√ 4παem
g = 2 2 MZ G F 1 − 1 − √
2
2MZ2 GF
Feynman rules
We list the Feynman rules of electroweak interactions in unitary gauge, using
the notation of Sect. 18.4:4
• Cubic gauge field interactions V1,α V2,β V3,γ (with all-incoming kine-
matics, pμ + q μ + r μ = 0): Γα, β, γ (p, q, r) = ieCVα, β, γ (p, q, r)
where Vα, β, γ (p, q, r) = (p − q)γ gαβ + (q − r)α gβγ + (r − p)β gαγ is
the same function as in QCD, while C depends on the type of the
gauge boson participating in the interaction:
V1V2V3 C
γW +W − 1
cos θ W
ZW +W −
sin θ W
V1V2V3V4 C
1
W +W +W −W −
(sin
θW ) 2
2
cos θ W
W +W − Z Z −
sin θ W
cos θ W
W +W − γZ −
sin θ W
W +W − γγ −1
4We focus on the interaction vertices only as the propagators of the various fields
were already presented earlier, those of the photon and charged leptons in the section
on QED, those of the massless scalar and quarks among the QCD Feynman rules
and those of the massive gauge bosons in Sect. 22.2. Finally, the denominator in the
propagator of a massive scalar is the same as for the massive gauge boson, while its
numerator coincides with that for the massless scalar. Some authors use opposite
sign convention for the massive gauge bosons as we do (e.g. Ref. [Denner, 1993]).
As a result, their vertices have opposite signs if odd number of massive gauge bosons
are involved, which does not influence values of observable quantities.
Electroweak sector of the standard model 339
where
sin θ W T f3 − sin θ W
2 e
f
g +f = − ef , g −f = .
cos θ W sin θ W cos θ W
1 m f ,i
C = −δi j .
2 sin θ W MW
The relations between the vector, axial-vector couplings of the Z boson and
the g ±f couplings read as
vf 1 − +
T f3 − 2(sin θ W )2 e f
= gf + gf = ,
2 sin θ W cos θ W 2 2 sin θ W cos θ W
af 1 T f3
= g −f − g +f = .
2 sin θ W cos θ W 2 2 sin θ W cos θ W
fermions, then five phases can be absorbed into the state vectors of charged
leptons, as for quarks, so only one phase remains. In such cases the usual
parametrization of the PMNS matrix is the same as for the CKM matrix in
(Eq. (22.6.13)), U = V, with c f m = cos θ f m , s f m = sin θ f m and δ 0, is
α21 α31
neutrinos (ν ≡ ν̄), U = V P
the CP-violating phase. In case of Majorana
α n−1
where P = diag 1, ei 2 , ei 2 , . . . , ei 2 , (n is the number of Majorana
neutrinos that can be different from 3). We do not know the physical
origin of the PMNS matrix. According to experimental observations charged
leptons do not mix, but neutrinos may mix. If they do so, then the matrix
Sν = U PMNS is an analogue of the CKM matrix. In charged currents flavour
eigenstates participate (those are created and annihilated). In high-energy
scattering experiments of particles the masses of neutrinos are negligible.
Our measurements are by far not sufficiently precise to measure any effect
of neutrino masses.
Neutrinos are stable particles due to their small masses and weak inter-
actions with other particles. Thus, they exist in asymptotically free states.
Moreover, according to observations and depending on the question we raise,
we can use a macroscopic, point-like particle approach in describing their
motion. For instance, we can measure the speed of electrons directly using
time-of-flight measurements and thus their masses if we know their ener-
gies. Therefore, it depends on the question which description is the most
convenient: (i) classical, (ii) quantum mechanical, (iii) or quantum field
theoretical. As a matter of fact, neutrinos show classical, quantum and
relativistic features, similarly to electrons.
A characteristic quantum phenomenon is the double-slit interference
with particles. For example, even single photons or electrons produce the
characteristic interference lines if we let sufficiently many pass (one by one)
through the two slits. These “characteristic” interference lines are those
that are also produced by a plane wave of light. Such light is the classical
manifestation of an ensemble of plane wave photons.
Similar quantum mechanical phenomenon is the oscillation of neutrinos,
which is interference in flavour space. If neutrinos have different masses,
then their phases change differently as they travel the distance L between
the source and the detector. As a result flavour interference occurs, called
neutrino oscillation. We do not measure L, the time T needed to travel
the distance L, the energy E and momentum p of each individual neutrino,
but we measure their average values in a neutrino beam. Thus, we use
the plane wave approximation as for classical (beam of) light, |νm ( x, t) =
342 Chapter 22
e−iφm (x,t) |νm (0, 0) where the phase is φm ( x, t) = (Em t − pm · x)/.6 We
assume that the neutrinos move along the x-axis inertially and we use the
abbreviation | pm | ≡ pm . According to the dispersion relation Em 2 = p2 c 2 +
m
mm c . If m1 m2 , then the neutrinos travelling in time T a distance L will
2 4
which means that E/T = E /T . As a result, in all inertial reference frames
T/L = E/(pc2 ).
In real experiments it is easier to measure L (instead of T) so we write
the phase difference as
L ΔE 2 E Δp2 L Δ(E 2 − pc2 ) Δm12
2 c4
Δφ12 = − = = L.
2 E pc2 p 2c2 p 2c2 p
(22.8.18)
Such a phase difference can lead to a change of neutrino flavour. The
probability of the change ν f → ν f is
2
∗ −iφ
P(ν f → ν f ) = |ν f |ν f (t)| = U f mU f n e n νm |νn
2
m,n
2 (22.8.19)
= U f mU ∗f m e−iφm .
m
In order to compute the sum, let us first consider the simplified case of two
6In this section–as opposed to the rest of the book–we use SI instead of natural
units so and c appears explicitly.
Electroweak sector of the standard model 343
summands (Δφ = φ1 − φ2 ):
(u1 eiφ1 + u2 eiφ2 )(u1∗ e−iφ1 + u2∗ e−iφ2 ) = |u1 | 2 + |u2 | 2 + u1 u2∗ eiΔφ + u1∗ u2 e−iΔφ
= |u1 | 2 + |u2 | 2 + 2Re(u1 u2∗ ) cos(Δφ) − 2Im(u1 u2∗ ) sin(Δφ)
∗ Δφ
= |u1 + u2 | − 4Re(u1 u2 ) sin −
2 2
+ 2Im(u1 u2∗ ) sin(−Δφ) .
2
(22.8.20)
For Dirac neutrinos, if the complex phase in the PMNS matrix vanishes,
δ = 0 (no CP-violation in the neutrino sector), then the PMNS matrix is real
and the last term vanishes in Eq. (22.8.21). In this case the only source of
neutrino oscillation is the second term, in which the square of the sinus is
periodic in π. Therefore, it is useful to rewrite its argument as
Δφmn L
= π (r)
2 Lmn
where
(r) 2hc2 p 2hcE pc
Lmn = = 2 c4 E
Δmmn c
2 4 Δmmn
is the characteristic length of oscillation. In oscillation experiments one can
usually measure the energy of the neutrinos and they travel with the speed
of light in vacuum. (Their energy is more than 1 MeV, while their masses
are smaller than 1 eV.) So it is more convenient to use the second form, in
which pc/E 1 so it is natural to drop it.
(r)
We obtain another form if we substitute Lmn expressed as a function of
the neutrino energy,
Δφmn π Δmmn2 c4
= L.
2 2 hcE
344 Chapter 22
Δφmn π Δmmn
2 c4 Δmmn
2 c4
GeV L
= L 1.27 ,
2 2 hcE eV2 E km
while the oscillation length is
Δφmn Δmmn
2
= L.
2 4E or p
Neutrinos can show interference in flavour space because they remain
coherent over long distances due to their weak interactions. Thus, the plane-
wave approximation can be applied. The experiments are performed with
neutrino beams. Oscillation experiments confirm our formula for neutrino
oscillation (see Fig. 22.3) within the uncertainty of the data. Neutrino oscil-
lation was observed in the following experiments:7
7The 2015 Nobel prize in physics was awarded to Profs. Takaaki Kajita and
Arthur B. McDonald for the discovery of neutrino oscillations.
Electroweak sector of the standard model 345
All these experimental observations prove that neutrinos have masses and the
lepton flavour number is not conserved, which calls clearly for explanation
beyond the standard model.
In the near future the most important tasks in the neutrino sector from
the experimental side is to measure the mass of the neutrinos and the pa-
rameters of the PMNS mixing matrix. At present we have (sometimes
contradictory) experimental information only for the sum of the neutrino
masses. From cosmological observations the current conservative upper
bound is 3i=1 mi ≤ 0.23 eV. From oscillation experiments we know two
squared mass differences, δm2 and Δm2 , with the latter being about thirty
times larger than the former. We conclude that there are two mass eigenstates
of approximately equal masses (one of which can even vanish), while the
third one has much different mass. However, we do not know if the very
different mass is much bigger than the other two, called normal hierarchy
(NH), or much smaller than the other two, called inverse hierarchy (IH).
The further parameters of neutrino oscillation are the mixing angles θ i j
and the complex phase δ of the PMNS matrix. In the oscillation formula
sin2 θ i j appears, therefore those are measured directly.
The PMNS matrix, obtained from a global fit of the parameters to data,
(at 99 % confidence level) [Esteban et al., 2017] is the following:
Compare these numbers with the precision of the parameters of the CKM
matrix in (Eq. (22.6.14)). From the experimental point of view the main
difficulty is to improve the precision of the parameters. From the theoretical
point of view the challenge is to understand the origin of the PMNS matrix
(or in popular terms “to understand the origin of neutrino masses”) and why
it is so different from the CKM matrix. We do not know the answer to these
theoretical questions, but we can state with certainty that in order to answer
them, we have to go beyond the standard model of particle interactions.
Γμ
q
W
iγμ ∂ μ ψ = mψ ⇒ −γμ ∂ μ ψ = i mψ
∂ μ Jμ5 = 2i mψ̄γ5 ψ ≡ 2m J5 ,
Electroweak sector of the standard model 347
Γμ
O(e5 ) ,
Γ5μ
· · · + O(e7 ) .
The last diagram8 which contributes to Γμ5 violates the conservation of the
axial current. We define
where p1 p2
γκ γλ
Sκλμ (p1, p2 ) = − p1 + p2
q = p1 + p2
γμ γ5
axial: q μ Tκλμ = 0 ,
vector: p1κ Tκλμ = 0 ,
vector: p2κ Tκλμ = 0.
Using
2 3
q μ γμ γ5 = / + /p2 − / − /p1 γ5 = / + /p2 γ5 + γ5 / − /p1
and /p /p = p2 1, we find
∫ ⎡ ⎤
⎢ / − /p1 γκ /γλ γκ /γλ / + /p2 ⎥
μ d4 ⎢ ⎥
q Sκλμ (p1, p2 ) = − Tr ⎢γ5 + γ5 ⎥.
(2π) 4 ⎢ ( − p1 )2 2
( + p2 )2 2 ⎥
⎢ ⎥
⎣ ⎦
We now shift by p1 : → + p1 in the first term and use γλ γ5 = −γ5 γλ
to obtain
⎡ ⎤
∫ ⎢ γλ /γκ / + /p1 γκ /γλ / + /p2 ⎥
d4 ⎢ ⎥
q μ Sκλμ (p1, p2 ) = Tr ⎢γ5 − γ5 ⎥
(2π)4 ⎢⎢ 2 ( + p )2
1
2 ( + p )2 ⎥
2 ⎥
⎣ ⎦
= −q μ Sλκμ .
Thus
q μ Tκλμ (p1, p2 ) = 0
provided the shift is allowed. However, the integral is divergent, so the shift
may be forbidden. For instance, with a cut-off regularization the momentum
shift leads to a finite term which is different from zero. Use of dimensional
regularization looks better as then the momentum shift is allowed. Neverthe-
less, the ambiguity remains also in this case, but disguised in a different form:
in d 4 dimensions γ5 is ambiguous (γ5 is intrinsically 4-dimensional). The
’t Hooft-Veltman prescription says for γ5 = iγ0 γ1 γ2 γ3 in d 4 dimensions
/ 0
γ μ , γ5 = 0 for μ = 0, 1, 2, 3
+ ,
γ μ , γ5 = 0 for other values of μ.
we find
∫ ( )
μ dd / − /p1 / / + /p2
q Tκλμ (p1 + p2 ) ∝ 2 Tr 2γ5 /̂ γκ γλ ,
(2π)d ( − p1 )2 2
( + p2 )2
Electroweak sector of the standard model 349
which has 5 terms. The only one that does not vanish, after integrations
gives
∫ ˆ2
β dd
4 (−4i) αβκλ p1α p2 =
(2π) d 2(
− p1 )2 ( + p2 )2
β d − 4 (2)
= −16i αβκλ p1α p2 I (p1 + p2 )
d
1
= 2 αβκλ p1κ pλ2 + O().
2π
Dimensional regularization respects the vector Ward identities, but the axial-
vector Ward identity is violated at one loop. There is no regularization
which respects both vector and axial-vector Ward identities simultaneously.
We prefer to keep the vector Ward identities, which is an expression of
insisting on gauge invariance. The axial-vector Ward identity is then violated,
which
- . axial anomaly. It is supported by the observed decay rate
is called
Γ π 0 → γγ = (7.7 ± 0.6) eV. The chiral symmetry forbids π 0 → γγ at tree
level, but the anomaly triangle diagrams predict
N 2 α 2 m2
c π
Γ π 0 → γγ = = 7.73 eV,
3 64π 3 fπ
T a1 T a2 T c + T a2 T a1 T c = {T a1 , T a2 } T c
for a given fermion in the loop (T c is the generator at the vertex with γμ γ5 ).
The generators are T a = τ2a or ya with anti-commutation relations
/ 0 / 0
τi, τj = 2δi j , yi, y j = 2yi y j .
Then
Tr ({T a1 , T a2 } τc ) ∝ Tr (τc ) = 0
350 Chapter 22
τa
for both T a = 2 and ya , while
Tr ({T a1 , T a2 } yc ) ∝ δ a1 a2 Q c,
or
Tr ({y a1 , y a2 } yc ) ∝ Q c (Tr (yc ) = Tr (Q c ) ∝ Q c )
Based on these considerations the total contribution
in a family is propor-
tional to Q e + Qν + Nc (Qu + Qd ) = −1 + 0 + 3 23 − 13 = 0 for left-handed
members. For right-handed members Qν = 0 is absent from the sum. Con-
sequently, the axial anomaly cancels in the standard model. Any theory
beyond the standard model should also have this cancellation if we want a
renormalizable model.
Chapter 23
Outlook
In the theoretical part of this book we emphasized that QCD radiative cor-
rections play an essential role in understanding the structure of the theory,
interpreting data and finding signatures of new physics that cannot be ex-
plained by the standard model. The theory of radiative corrections in the
electroweak part of the standard model is equally important though the ef-
fects are usually less pronounced. As a result, electroweak corrections have
also been computed to many processes that play important role in particle
phenomenology. These computations are technically challenging and the re-
sulting corrections are usually much smaller than those of QCD, simply due
to the large difference in the size of the couplings. Nevertheless, in certain
kinematic regions as well as for interpreting the non-hadronic final states at
the LEP experiments, taking such corrections into account is crucial.
As the LHC collides hadrons, the strong interaction has a role in every
single interaction. As a thumb rule, we can state that the effect of first
order electroweak corrections is similar to that of the second order QCD
corrections. Thus in a computation at NNLO QCD accuracy (O(αs2 ) relative
correction) the first order electroweak corrections (O(α) relative correction)
should also be taken into account. Such computations are hampered by the
mixed QCD and electroweak corrections at O(αs α) relative accuracy.
We put more emphasis on the computation of QCD corrections because
their role of quantum corrections is qualitatively different in QCD than in
the electroweak theory. In general QCD predictions at LO accuracy provide
only an order of magnitude estimate due to the strong dependence on the
unphysical scales. The NLO corrections are crucial in order to decrease the
dependence on these scales. In order to estimate the accuracy of the scale
351
352 Chapter 23
the proton is much smaller than that of the electron in the ordinary hydrogen
atom. Thus it is more sensitive to the electric charge distribution inside the
proton. The radius of the proton measured in muonic atoms differs from the
radius measured in ordinary hydrogen.
Between writing and publishing this book several anomalies have been
observed by the LHCb experiment at the LHC. These concern a fundamental
assumption of the standard model that is known as flavour universality in
the lepton sector. The second and third particle families assumed to be exact
copies of the first one with only one exception, namely the particle masses.
Therefore the standard model predicts that processes with identical initial
state and with only difference in the final state in the type of leptons, such
as B0 → K ∗0 μ+ μ− and B0 → K ∗0 e+ e− ) should occur with approximately
equal probability. Quantitatively it means that the standard model predicts
B(B0 → K ∗0 μ+ μ− )
RK ∗0 = 1 (23.0.1)
B(B0 → K ∗0 e+ e− )
for the ratio of the branching fractions (the exact value depends on the value of
the invariant mass squared q2 of the charged lepton pair). Modern particle
detectors can identify the leptons in the final state with high accuracy so
such ratios can be measured fairly well. The results published by the LHCb
collaboration [Aaij et al., 2017]
⎧ +0.11
⎨ 0.66−0.07 (stat) ± 0.03(syst) for 0.045 < q < 1.1 GeV /c
2 4
⎪
⎪
2
RK ∗0 =
⎪
⎪ 0.69+0.11 (stat) ± 0.05(syst) for 1.1 < q2 < 6.0 GeV2 /c4
⎩ −0.07
(23.0.2)
seem to suggest violation of the principle of flavour universality. The ob-
servations currently show deviations from the standard model predictions at
about three sigma and await confirmation by other experiments. If deviations
with more than five sigma significance are confirmed, these will constitute
discoveries that will require non-trivial extension of the standard model,
similarly to the impact of the discovery of parity violation on the extension
of the Fermi theory.
In recent years particle physics has started to employ new techniques
inspired by developments in information technology. The particle detector
prepares a three-dimensional image of the final state measuring the direction,
energy and momentum of particle flow and the bend of particle tracks by
magnetic field. The purpose of data analysis is to find out the hard scattering
process from this information. This is a typical pattern recognition problem.
Engineering of pattern recognition has undergone tremendous development
354 Chapter 23
There is a theory which states that if ever anyone discovers exactly what the
Universe is for and why it is here, it will instantly disappear and be replaced by
something even more bizarre and inexplicable.
There is another theory which states that this has already happened.
(Douglas Adams)
Why God Particle? Two reasons. One, the publisher wouldn’t let us call it the
Goddamn Particle, though that might be a more appropriate title, given its villainous
nature and the expense it is causing. The title ended up offending two groups: (1)
those who believe in god and (2) those who do not.
(Leon Lederman)
We need something new. We can’t predict what that will be or when we will find
it because if we knew that, we would have found it already!
(Stephen Hawking)
“Would you tell me, please, which way I ought to go from here?”
“That depends a good deal on where you want to get to,” said the Cat.
“I don’t much care where–” said Alice.
“Then it doesn’t matter which way you go,” said the Cat
“–so long as I get SOMEWHERE,” Alice added as an explanation.
“Oh, you’re sure to do that,” said the Cat, “if you only walk long enough.”
(Lewis Carroll)
Bibliography
[Aaboud et al., 2017] Aaboud, M. et al. (2017). Measurement of the Higgs boson
√
coupling properties in the H → Z Z ∗ → 4 decay channel at s = 13 TeV with
the ATLAS detector.
[Aad et al., 2010] Aad, G. et al. (2010). Observation of a centrality-dependent dijet
√
asymmetry in lead-lead collisions at s N N = 2.77 TeV with the ATLAS detector
at the LHC. Phys. Rev. Lett., 105:252303.
[Aad et al., 2012] Aad, G. et al. (2012). Observation of a new particle in the search
for the Standard Model Higgs boson with the ATLAS detector at the LHC.
Phys.Lett., B716:1–29.
[Aad et al., 2015] Aad, G. et al. (2015). Combined measurement of the Higgs boson
√
mass in pp collisions at s = 7 and 8 TeV with the ATLAS and CMS experiments.
Phys. Rev. Lett., 114:191803.
[Aad et al., 2016] Aad, G. et al. (2016). Measurements of the Higgs boson produc-
tion and decay rates and constraints on its couplings from a combined ATLAS and
√
CMS analysis of the LHC pp collision data at s = 7 and 8 TeV. JHEP, 08:045.
[Aaij et al., 2017] Aaij, R. et al. (2017). Test of lepton universality with B0 →
K ∗0 + − decays. JHEP, 08:055.
[Aamodt et al., 2010] Aamodt, K. et al. (2010). Elliptic flow of charged particles in
Pb-Pb collisions at 2.76 TeV. Phys. Rev. Lett., 105:252302.
[Abbiendi et al., 2012] Abbiendi, G. et al. (2012). Search for charged Higgs bosons
√
in e+ e− collisions at s = 189 − 209 GeV. Eur. Phys. J., C72:2076.
[Abbott, 1981] Abbott, L. (1981). The background field method beyond one loop.
Nucl.Phys., B185:189.
[Adams et al., 2003] Adams, J. et al. (2003). Evidence from d + Au measurements
for final state suppression of high p(T) hadrons in Au+Au collisions at RHIC.
Phys. Rev. Lett., 91:072304.
[Adler et al., 2003] Adler, S. S. et al. (2003). Suppressed π 0 production at large
transverse momentum in central Au+ Au collisions at S(NN)**1/2 = 200 GeV.
Phys. Rev. Lett., 91:072301.
355
356 Bibliography
[Cacciari and Salam, 2006] Cacciari, M. and Salam, G. P. (2006). Dispelling the
N 3 myth for the k t jet-finder. Phys.Lett., B641:57–61.
[Cacciari et al., 2008] Cacciari, M., Salam, G. P., and Soyez, G. (2008). The anti-k(t)
jet clustering algorithm. JHEP, 0804:063.
[Catani et al., 2002] Catani, S., Dittmaier, S., Seymour, M. H., and Trocsanyi, Z.
(2002). The dipole formalism for next-to-leading order QCD calculations with
massive partons. Nucl.Phys., B627:189–265.
[Catani et al., 1991] Catani, S., Dokshitzer, Y. L., Olsson, M., Turnock, G., and
Webber, B. (1991). New clustering algorithm for multi - jet cross-sections in e+
e- annihilation. Phys.Lett., B269:432–438.
[Chatrchyan et al., 2012a] Chatrchyan, S. et al. (2012a). Jet momentum dependence
√
of jet quenching in PbPb collisions at s N N = 2.76 TeV. Phys. Lett., B712:176–
197.
[Chatrchyan et al., 2012b] Chatrchyan, S. et al. (2012b). Observation of a new boson
at a mass of 125 GeV with the CMS experiment at the LHC. Phys.Lett., B716:30–
61.
[Chatrchyan et al., 2013] Chatrchyan, S. et al. (2013). Observation of a new boson
√
with mass near 125 GeV in pp collisions at s = 7 and 8 TeV. JHEP, 1306:081.
[Cohen et al., 1998] Cohen, A. G., De Rujula, A., and Glashow, S. (1998). A matter
– antimatter universe? Astrophys.J., 495:539–549.
[Collins et al., 1989] Collins, P., Martin, A. D., and Squires, E. (1989). Particle
physics and cosmology.
[Corcella et al., 2001] Corcella, G., Knowles, I., Marchesini, G., Moretti, S., Oda-
giri, K., et al. (2001). HERWIG 6: An event generator for hadron emission
reactions with interfering gluons (including supersymmetric processes). JHEP,
0101:010.
[Cousins and Highland, 1992] Cousins, R. D. and Highland, V. L. (1992). In-
corporating systematic uncertainties into an upper limit. Nucl.Instrum.Meth.,
A320:331–335.
[Cowan et al., 2011] Cowan, G., Cranmer, K., Gross, E., and Vitells, O. (2011).
Asymptotic formulae for likelihood-based tests of new physics. Eur. Phys. J.,
C71:1554. [Erratum: Eur. Phys. J.C73,2501(2013)].
[Denner, 1993] Denner, A. (1993). Techniques for calculation of electroweak ra-
diative corrections at the one loop level and results for W physics at LEP-200.
Fortsch. Phys., 41:307–420.
[Dirac, 1931] Dirac, P. A. (1931). Quantized singularities in the electromagnetic
field. Proc.Roy.Soc.Lond., A133:60–72.
[Dokshitzer, 1977] Dokshitzer, Y. L. (1977). Calculation of the structure functions
for deep inelastic scattering and e+ e- annihilation by perturbation theory in
quantum chromodynamics. Sov.Phys.JETP, 46:641–653.
358 Bibliography
[Dokshitzer et al., 1991] Dokshitzer, Y. L., Khoze, V. A., Mueller, A. H., and Troyan,
S. I. (1991). Basics of perturbative QCD.
[Durr et al., 2008] Durr, S., Fodor, Z., Frison, J., Hoelbling, C., Hoffmann, R., et al.
(2008). Ab-Initio determination of light hadron masses. Science, 322:1224–1227.
[Ellis, 1988] Ellis, R. K. (1988). An introduction to the QCD parton model.
[Esteban et al., 2017] Esteban, I., Gonzalez-Garcia, M. C., Maltoni, M., Martinez-
Soler, I., and Schwetz, T. (2017). Updated fit to three neutrino mixing: Exploring
the accelerator-reactor complementarity. JHEP, 01:087.
[Gehrmann-De Ridder et al., 2007] Gehrmann-De Ridder, A., Gehrmann, T.,
Glover, E., and Heinrich, G. (2007). NNLO corrections to event shapes in e+
e- annihilation. JHEP, 0712:094.
[Glashow et al., 1970] Glashow, S., Iliopoulos, J., and Maiani, L. (1970). Weak
interactions with lepton-hadron symmetry. Phys.Rev., D2:1285–1292.
[Gribov and Lipatov, 1972] Gribov, V. and Lipatov, L. (1972). Deep inelastic e p
scattering in perturbation theory. Sov.J.Nucl.Phys., 15:438–450.
[Halzen and Martin, 1984] Halzen, F. and Martin, A. D. (1984). Quarks and leptons:
An introductory course in modern particle physics.
[Horvath, 2003] Horvath, D. (2003). Search for charged Higgs bosons with the
OPAL detector at LEP. Nucl. Phys., A721:C453–C456.
[James and Roos, 1975] James, F. and Roos, M. (1975). Minuit: A system for func-
tion minimization and analysis of the parameter errors and correlations. Com-
put.Phys.Commun., 10:343–367.
[Klein and Roodman, 2005] Klein, J. and Roodman, A. (2005). Blind analysis in
nuclear and particle physics. Ann.Rev.Nucl.Part.Sci., 55:141–163.
[Lee and Wick, 1974] Lee, T. D. and Wick, G. C. (1974). Vacuum stability and
vacuum excitation in a spin 0 field theory. Phys. Rev., D9:2291–2316.
[Leo, 1987] Leo, W. (1987). Techniques for nuclear and particle physics experi-
ments: a how to approach.
[Mangano and Parke, 1991] Mangano, M. L. and Parke, S. J. (1991). Multiparton
amplitudes in gauge theories. Phys.Rept., 200:301–367.
[Martin, 1997] Martin, S. P. (1997). A supersymmetry primer. pages 1–98. [Adv.
Ser. Direct. High Energy Phys.18,1(1998)].
[Miller et al., 1972] Miller, G., Bloom, E. D., Buschhorn, G., Coward, D., DeStae-
bler, H., et al. (1972). Inelastic electron-proton scattering at large momentum
transfers. Phys.Rev., D5:528.
[Mohr et al., 2016] Mohr, P. J., Newell, D. B., and Taylor, B. N. (2016). CODATA
recommended values of the fundamental physical constants: 2014. Rev. Mod.
Phys., 88(3):035009.
Bibliography 359
[Nagy and Trocsanyi, 1999] Nagy, Z. and Trocsanyi, Z. (1999). Multijet rates in
e+ e- annihilation: Perturbation theory versus LEP data. Nucl.Phys.Proc.Suppl.,
74:44–48.
[Patrignani et al., 2016] Patrignani, C. et al. (2016). Review of particle physics.
Chin. Phys., C40(10):100001.
[Perkins, 1982] Perkins, D. (1982). Introduction to high energy physics.
[Prosper and Lyons, 2011] Prosper, H. B. and Lyons, L. (2011). Proceedings of
the PHYSTAT 2011 Workshop on statistical issues related to discovery claims in
search experiments and unfolding, CERN, Geneva, Switzerland, 17-20 January
2011.
[Pumplin et al., 2002] Pumplin, J., Stump, D., Huston, J., Lai, H., Nadolsky, P. M.,
et al. (2002). New generation of parton distributions with uncertainties from
global QCD analysis. JHEP, 0207:012.
[Ryder, 1996] Ryder, L. (1996). Quantum field theory.
[Schael et al., 2013] Schael, S. et al. (2013). Electroweak measurements in electron-
positron collisions at W-boson-pair energies at LEP. Phys.Rept., 532:119–244.
[Shaposhnikov and Wetterich, 2010] Shaposhnikov, M. and Wetterich, C. (2010).
Asymptotic safety of gravity and the Higgs boson mass. Phys. Lett., B683:196–
200.
[Sirunyan et al., 2017] Sirunyan, A. M. et al. (2017). Measurements of properties
of the Higgs boson decaying into the four-lepton final state in pp collisions at
√
s = 13 TeV. JHEP, 11:047.
[Sjostrand et al., 2006] Sjostrand, T., Mrenna, S., and Skands, P. Z. (2006). PYTHIA
6.4 physics and manual. JHEP, 0605:026.
[Skands, 2012] Skands, P. (2012). Introduction to QCD. arXiv.org, 1207.2389:1–85.
[Turner and Wilczek, 1982] Turner, M. S. and Wilczek, F. (1982). Might our vacuum
be metastable? Nature, 298:633–634.
[Weinzierl, 2009] Weinzierl, S. (2009). Event shapes and jet rates in electron-
positron annihilation at NNLO. JHEP, 0906:041.
[Woithe et al., 2017] Woithe, J., Wiener, G. J., and Van der Veken, F. F. (2017).
Let’s have a coffee with the Standard Model of particle physics! Phys. Educ.,
52(3):034001.
Index
Bold-face page numbers indicate sections dedicated to the given entry, whereas
page numbers in normal type point to textual references. In order to maintain an
alphabetic order, Greek symbols and numbers in entries are spelled out in Latin (e.g.
“τ-lepton” is listed as “Tau-lepton” or “3-jet events” as “Three-jet events”).