Sei sulla pagina 1di 42

Elementary Particle Physics

David Paganin
School of Physics, Monash University, Victoria 3800, Australia
(Dated: August 22, 2007)
Welcome to the world of elementary particle physics! These lecture notes are based on the relevant
sections from the prescribed text: B.R.Martin and G.Shaw, Particle Physics (second edition), John
Wiley and Sons, Chichester (1999). Copies of the text are available for purchase from the Monash
University bookstore.

D. Charge conjugation

Contents

I. INTRODUCTION
A. Introduction
B. Antiparticles
1. Relativistic wave equations
C. Feynman diagrams
1. Virtual (basic) electromagnetic
processes
2. Real electromagnetic processes
3. Pair production and annihilation
D. Particle exchange
1. Range of forces
2. Yukawa potential
3. The scattering amplitude
E. Natural units
F. Amplitudes and cross sections
1. Rates and cross sections
2. Total cross sections
3. Dierential cross sections
G. Glossary of jargon
II. LEPTONS, QUARKS AND HADRONS
A. Leptons
1. Electron neutrinos
2. Further lepton generations
3. Lepton decays and universality
B. Strongly interacting particles
1. Quarks
2. Hadrons: General properties
3. Strange particles, charm and beauty

2
2
2
2
3
3
4
4
4
5
6
6
7
8
8
8
8
9
10
11
11
12
12
13
13
13
14

III. EXPERIMENTAL METHODS


A. Accelerators
1. Linear accelerators and synchrotrons
2. Fixed-target machines and colliders
B. Particle detectors
1. Time resolution: scintillators
2. Position measurement
3. Momentum measurement
4. Particle identification
5. Energy measurement
C. The Brookhaven experiment

15
15
15
16
16
16
17
17
18
19
19

IV. SPACE-TIME SYMMETRIES


A. Translational invariance
B. Rotational invariance
C. Parity

21
21
22
23

24

V. A CLOSER LOOK AT HADRONS


A. Revision
B. Isospin
C. Mass splittings of the baryons
D. Resonances
E. The quark-antiquark potential
F. The lightest hadrons
1. The light mesons
2. The light baryons
G. Colour

24
24
25
26
26
27
28
28
29
31

VI. QUANTUM CHROMODYNAMICS


A. Colour charges and confinement
B. Gluons as mediators of the strong force
C. Screening, anti-screening, asymptotic
freedom
D. Jets

31
32
33

VII. WEAK INTERACTIONS


A. Introduction
B. On the importance of the weak interaction
C. Charged-current weak interactions
1. Charged currents #1: the W -lepton
interactions
2. Charged currents #2: the W -quark
interactions
D. Neutral-current weak interactions
E. Electroweak unification: a very brief note
References

33
34
35
35
35
36
36
37
39
40
41

2
I.

INTRODUCTION
A.

Introduction

The project of particle physics, which entails a study


of the fundamental constituents of matter and their interactions, is incomplete in the sense that there exists
no grand unified theory which is able to unite the four
known forces of nature, namely:
the electromagnetic force;
the strong force;
the weak force;
the gravitational force.
Notwithstanding this, the currently dominant
paradigm of modern elementary particle physics is the
so-called standard model, the exposition of which will be
the central goal of this course.
The standard model attempts to explain the physics
of elementary particles in terms of three distinct basic
particles, which are treated in the theory as structureless
point particles. These are:
leptons;
quarks;
gauge bosons.
The gauge bosons are considered to be the force carriers of the theory, about which more will be said later.
The electron is an example of a lepton, this particle
often being bound to a nucleus by the electromagnetic
interaction. The neutrino is also a lepton, which is a light
particle produced in -decay, the force behind which is
the weak interaction.
Hadrons[1] such as protons and neutrons are considered to be composite particles in the standard model;
they are believed to be made of quarks bound together
by the strong interaction. Free quarks have never been
observed, a phenomenon known as confinement. The
strong force which binds quarks inside hadrons is also
believed to be responsible for the strong interaction between hadrons, such as the strong force between nucleons.
For most of this course, we will neglect the eects of
gravity in the physics of elementary particles.
We mentioned earlier that the gauge bosons are the
force carriers of the standard model. (a) For the electromagnetic interaction, the force between charged particles is considered to be due to the exchange of the familiar
spin-1 bosons called photons; we shall see later that the
long-range nature of this interaction is a consequence of
the fact that the rest mass of the photon is zero. (b) The
weak interaction is due to the exchange of very massive
spin-1 particles called W and Z bosons; the short range
nature of this force is a direct consequence of the massive

nature of W and Z. (c) The force carriers of the strong interaction are spin-1 bosons known as gluons. Despite the
fact that the gluons have zero rest mass, the associated
force is of a short range. (d) The gravitational force is
believed to be due to the existence of spin-2 rest-massless
bosons called gravitons.
B.

Antiparticles

For every charged particle, there exists an associated


antiparticle of the same mass but opposite charge. The
existence of antiparticles is a direct consequence of combining special relativity with quantum mechanics, as
shall be seen in our discussions on relativistic wave equations such as the Klein-Gordon and Dirac equations. We
shall also se that relativistic eects are important in considering many of the interactions between elementary
particles, as the velocities involved are very often comparable with that of light.
1.

Relativistic wave equations

The relativistic relation between energy E, momentum


p = |p| and mass[2] m is:
E 2 = p2 c2 + m2 c4

(1)

By making use of the usual Schrodinger correspondence rules:


p i, E it ,

(2)

this becomes the Klein-Gordon equation of relativistic


quantum mechanics:
2 t2 = 2 c2 2 + m2 c4

(3)

An important feature of this equation is the existence


of negative-energy solutions. To see this, note that for
every plane wave solution of the form:
(x, y, z, t) = N exp(i(p x Ep t)/)

(4)

with momentum p and positive energy:


E = Ep =

p2 c2 + m2 c4 mc2

(5)

there is also a solution:

(x,
y, z, t) = N exp(i(p x Ep t)/)

with momentum p and negative energy:

(6)

3
C.

E = Ep =

p2 c2 + m2 c4 mc2 .

(7)

Exercise 1: Derive equations (5) and (7).


From a mathematical point of view, the existence of
negative energy solutions to the Klein-Gordon equation
is a consequence of the fact that this equation is secondorder in time; to see this, note that equation (3) is invariant under t t, while the energy operator E = it
changes sign.
In an attempt to overcome this problem, Dirac looked
for a relativistic wave equation which was first order in
time, and which had the familiar Schrodinger form:
it (x, t) = H(x, p)(x, t)

(8)

for spin-1/2 particles such as the electron. This equation


is known as the Dirac equation. Now, you know from
your studies of non-relativistic quantum mechanics that
the wavefunction of a spin-1/2 particle is described by a
two-component wavefunction (Pauli spinor). You should
also know that a second-order dierential equation can
be reduced to a pair of first-order equation at the price
of doubling the number of unknowns (for example, the
dierential equation m
x = g mv = g, v = x ). It
should therefore come as no surprise that the reduction of
the second-order equation (3) to the first-order equation
(8) involves a doubling in the number of components of a
spin-1/2 wavefunction from 2 to 4, leading to the famous
Dirac spinor:

1 (x, t)
(x, t)
(x, t) = 2
3 (x, t)
4 (x, t)

Feynman diagrams

Feynman diagrams are a pictorial technique of great


power in elementary particle physics. Their introduction
is more properly motivated by a certain perturbationtheoretic argument based on the Dyson expansion; here,
we give an elementary motivation for the technique which
will suce for our purposes.

1.

Virtual (basic) electromagnetic processes

Earlier, we stated that the electromagnetic interaction


can be seen in terms of the exchange of massless spin1 gauge bosons called photons. The eight basic processes, whereby electrons or anti-electrons (i.e. positrons)
interact with photons, are sketched in Figure 1. These
diagrams are known as Feynman diagrams. Note that,
in the Figure, one arrow points towards each vertex,
and one arrow points away from each vertex, due to the
conservation of electric charge.

(9)

of relativistic quantum mechanics.


Now, what about the form of the Hamiltonian H in the
Dirac equation (8)? Since (x, t) is a four-component
wavefunction, this Hamiltonian must be a 4 4 matrix
operator. Also, since the equation is first order in time,
it must also be first order in the spatial variables x
(x, y, z) (by Lorentz invariance). Bearing all of this in
mind, we can see the motivation behind Diracs proposed
form for the Hamiltonian in (8), namely:
H = c p + mc2 ,

(10)

where (x , y , z ), and x , y , z , are 4 4 matrices whose form is not required for this course.
Despite the attempt of the Dirac equation, the problem
of negative-energy solutions remains. This is a feature of
relativistic quantum theories, and may be resolved, in
this context, by re-interpreting the negative-energy solutions as antiparticles. A rigorous justification of this
identification requires Quantum Field Theory, which is unfortunately! - beyond the scope of this course.

FIG. 1: Feynman diagrams for the eight basic processes


where electrons e and positrons e+ interact with photons.
Time runs from left to right.

4
2.

Real electromagnetic processes

Both momentum and angular momentum need to be


conserved at each vertex of a Feynman diagram. Now,
in free space, energy conservation is violated in all of the
basic processes shown in Figure 1.
Exercise 2: Demonstrate the violation of energy conservation in each of the basic electromagnetic processes
sketched in Figure 1. For each basic process, determine
whether energy conservation is violated in a positive or
a negative sense at the interaction vertex.
Because of the fact that they violate energy conservation, each of the processes sketched in Figure 1 is known
as a virtual process (as opposed to a real process, which
does conserve energy). To obtain a real process, two or
more virtual processes must be combined in such a way
that energy conservation is only violated for a time period
which is suciently short such that the energy-time uncertainty principle:
E

FIG. 3: Contribution to electron-electron scattering from twophoton exchange

stated earlier, that the contribution of multi-photon exchanges (such as that shown in Figure 3) is much smaller
than that due to single-photon exchanges.

(11)

is respected.

3.

Pair production and annihilation

Consider the following annihilation:


e+ + e p,

FIG. 2: Single photon exchange in electron-electron scattering

As an example of the application of these ideas, consider the pair of processes shown in Figure 2. For each
of the two Feynman diagrams, energy conservation is violated at the first vertex, with this violation being compensated for by energy nonconservation at the second
vertex, such that energy is conserved overall.
Now, multi-photon exchange, such as the two-photon
exchange in Figure 3, is also possible. However, the
contributions of such multi-photon exchanges is much
smaller than that of the single-photon exchange shown
in Figure 2.
Denote, by the order n of a process, the number of
vertices in a Feynman diagram of the said process. Thus,
single-photon exchange is of order n = 2, and m-photon
exchange is of order n = 2m. We state, without proof,
that the vertex of each basic process has a probability
of occurrence which is of the order of the fine-structure
constant 1/137 1. Thus, the probability of occurrence for single-photon exchange is O(2 ), and the probability of m-photon exchange is O(2m ). We conclude, as

(12)

where p 2 (i.e. at least two photons are produced)


because energy must be conserved overall. To produce p
photons we will need p vertices in the associated Feynman diagram. As an example, the case p = 2 is shown
in Figure 4; the resulting pair of diagrams are topologically equivalent because they can be continuously
deformed into one another. Topologically equivalent diagrams such as these are called dierent time orderings
and it is usual to draw only one time ordering, leaving
the others implied. Evidently, there are p! dierent time
orderings of reaction (12); for example, figure 5 gives one
of the 3! = 6 dierent time orderings for the reaction
e+ + e + + .
The process (12) is of order p (because there are p
vertices), with an associated probability of occurrence
that is O(p ). Hence, for example:
Rate(e+ e 3)
O(3 )
=
= O()
+

Rate(e e 2)
O(2 )
D.

(13)

Particle exchange

We have already stated that, in the standard model


of particle physics, the forces are associated with the exchange of particles. Here, we examine this in more detail, by looking at: (i) the relation between the range of
a force and the mass of the particle associated with that

5
strengths at each of the two vertices are both equal to g.
The lower vertex represents a virtual process, as is easily
seen by going to the rest frame of the incident particle (in
this rest frame, particle A has no kinetic energy before
the interaction; after the interaction, the total energy is
greater because (i) it is now moving and (ii) the particle
X has been emitted). Indeed, the energy dierence E
between the final and initial states is:

FIG. 4: Lowest-order contributions to e+ + e + . The


two diagrams are related to one another by dierent time
ordering.

FIG. 5: The process e+ + e 3. Only one of the 3! = 6


time orderings is shown; the remaining 5 time orderings are
implied.

force; (ii) the Yukawa potential; (iii) scattering amplitudes associated with a force that is well-approximated
by single-particle exchange.
1.

Range of forces

Consider Figure 6, which shows the elastic scattering


of particles A and B, which proceeds via the exchange of
a boson X of (rest) mass MX ; the (rest) masses of the
particles A and B are MA and MB , respectively.

FIG. 6: Contribution to the reaction A + B A + B from


the exchange of a gauge boson X.

As indicated in the Figure, we assume that coupling

E = EX + EA MA c2
(14)

2 c4 +
=
p2 c2 + MX
p2 c2 + MA2 c4 MA c2
2pc, p
MX c2 , p 0

Exercise 3: Prove the assertions made in equation


(14), i.e. please show that: (i) E 2pc as p ; (ii)
E MX c2 as p 0. Also, (iii) show that, for all p,
E MX c2 .
Bearing in mind the results of Exercise 3, we see that
the energy violation E at the lower vertex of Figure
6 obeys E MX c2 . Therefore, from the energy-time
uncertainty relation (11), we see that the energy violation can exist for a time /E /(MX c2 ); the
furthest distance r that the virtual particle X can travel
will equal the speed of light c multiplied by the maximum
permitted value of , and therefore the range R of the
force associated with the exchange of particle X is given
by:

Rc

MX

c2

.
MX c

(15)

Because of the presence of the rest mass MX of the


particle in the denominator, we see that the range of the
force decreases as the rest mass of the force particle
increases. Let us now examine three special cases of this
idea.
(i) The electromagnetic interaction, which is due to
photon exchange, has an infinite range because of the
zero rest mass of the photon; (ii) The fact that the graviton has zero rest mass is believed to imply that the gravitational force is infinite in range; (iii) The rest masses of
the W and Z particles associated with the weak interaction are very large (MW = 80.3 GeV/c2 and MZ = 91.2
GeV/c2 ), and so the weak interaction has a very short
range (from (15), we have RW = /(MW c) 2 103
fm, where 1 fm= 1015 m)[3]. Since the range of the
weak force is around a thousand times smaller that the
radius of a proton, it can often be approximated as a zerorange or point interaction, corresponding to the limit
MX , as sketched in Figure 7.
Exercise 4: Deduce Coulombs law from the photonexchange model for the force exerted by the electric field.

6
the scattering is caused by the interaction of the said particle with the potential (x); note that O is the origin of
coordinates for the position vector x, and the momentum transfer q is defined by q qf qi .

FIG. 7: As the rest mass MX of the particle X associated with the force between particles A and B, one approaches
a zero-range or point interaction.

2.

FIG. 8: Scattering by a potential (x)

Yukawa potential

With reference to the left panel of Figure 7, note that


in the limit where the mass of particle A becomes suciently large, we can regard the particle B as being scattered by a static potential, this static potential being due
to particle A. If we neglect spin, then we can work with
a complex wave-function (x, t) describing the spinless
bosons X associated with the force. Now, since relativistic eects are typically significant in the context of particle physics, one may postulate that the wave-function
(x, t) obeys the Klein-Gordon equation (c/f equation
(3)):

2 4
2 t2 (x, t) = 2 c2 2 + MX
c (x, t)

2 2
MX
c
(x)
2

(17)

Exercise 5: Show that:

(r) =

g 2 exp(r/R)
4
r

(18)

is a solution to equation (17), where R is given by


(15). Also, please show that, as MX 0, it reduces to
the Coulomb potential (up to multiplicative constants);
briefly interpret this result. Note that equation (18) is
known as the Yukawa potential, after Hideki Yukawa,
who derived this equation in 1935.

3.

M (q) =

(x) exp(iq x/)d3 x,

The scattering amplitude

Consider the scattering experiment which has been


sketched in Figure 8. Here, we see the scattering of a
particle, which has a momentum of qi before the scattering event and momentum qf after the scattering event;

(19)

an integral which has the form of a Fourier transform [4].


Exercise 6 (optional): Please prove equation (19).
Exercise 7: Show that, for scattering from the
Yukawa potential, the probability amplitude for scattering with momentum transfer q is given by:

(16)

Since we are interested in static potentials, let us drop


the time dependence from equation (16), leaving us with:

2 (x) =

We claim that, in lowest order perturbation theory, the


probability amplitude M (q) for the scattering process in
Fig. 8 is proportional to:

M (q) =

g 2 2
2 c2 .
|q|2 + MX

(20)

Now, remember that we introduced the concept of a


point interaction in Fig. 7; equipped with equation
(20), we are ready to explore this idea in a little more
detail. Suppose that the range R = /MX c of the force
in question is very small when compared to the de Broglie
wavelengths of all particles involved in a given interaction. Then:
R=

h
2
i =
=
MX c
pi
pi

MX c
pi
pi 2MX c,

(21)

where pi denotes the modulus of the momenta of the


particles involved in the interaction. Since the momentum transfer q is given by the appropriate sums and differences of the particle momenta, the said momentum
transfer is of a similar order of magnitude to the largest
of the pi , and so the final line of the previous equation
becomes:
2 2
|q|2 MX
c .

(22)

7
Hence, when R i , the |q|2 may be ignored in equation (20), leading to:

M (q)

g 2 2
2 c2 G.
MX

(23)

We see, therefore, that in the zero-range approximation, the resulting point interaction between A and B is
characterized by a single coupling constant G, independent of the momentum transfer. This concept is very
useful in the study of weak interactions, which are characterized by the Fermi coupling constant GF .

E.

Natural units

The SI system of units measures mass, length and time


in terms of kilograms, metres and seconds respectively;
while such a choice of units is a sensible one for the everyday life of a human, it is not very practical for elementary particle physics. Here, one often encounters the
so-called natural units, which are such that two of the
fundamental constants of nature have the dimensional
numerical value of unity:

= c = 1.

(24)

In natural units, energy is measured in terms of the


electron volt (eV), this being the kinetic energy acquired
by an electron which is accelerated through a potential
dierence of 1 volt.
In natural units, all quantities have the dimension of
some power of the energy, E. As an example, we consider
three such quantities below.
(i) Mass has the dimensions of energy, a fact which
is easily remembered by realizing that the natural-units
form of the famous Einstein relation E = mc2 is E = m,
implying that energy E and mass m have the same dimensions in natural units. (ii) Time has the dimensions of E 1 in natural units; to see this, note that the
natural-units form of the energy-time uncertainty principle is Et 1, implying that energy and time have
dimensions which are reciprocals of one another (since
the number 1 is dimensionless). (iii) Length also has the
dimension of E 1 in natural units, as is readily shown
(see the next exercise!).
Exercise 8: (a) Show that, in natural units, length
has the dimensions of E 1 . (b) Show that, for a quantity which has SI dimensions M p Lq T r (M mass, L
length, T time), the natural-units dimensions are
E pqr .
Unless stated otherwise, we will work in natural units
for the remainder of the course.

Quantity
p q r n
Action ()
1 2 -1 0
Velocity (c)
0 1 -1 0
Mass
1 0 0 1
Length
0 1 0 -1
Time
0 0 1 -1
Momentum
1 1 -1 1
Energy
1 2 -2 1
Fine structure constant () 0 0 0 0
Fermi coupling constant (GF ) 1 5 -2 -2
Table 1. Here, we see the MKS dimensions M p Lq T r
and the natural-units dimensions E n = E pqr of some
important physical quantities.
Exercise 9: Derive all entries listed in Table 1. For
entries that have already been derived in the text, give
an alternative derivation.
Having introduced the natural units, it is natural to
investigate the means of converting between SI and natural units. (a) It is easy to convert from SI to natural
units, because all one needs to do is set = c = 1 in all
equations! (b) Going the other way, i.e. from natural to
SI units, is a little more involved, and may be done using
the following steps:
Step 1: Use dimensional reasoning to restore the
factors of and c;
Step 2: Make use of the conversion factors:
= 6.582 1022 M eV.s,
c = 1.973 1013 M eV.m.

(25)

As example of such a conversion from natural to SI


units, consider the following natural-units formula:
=

8 2
,
3 m2e

(26)

where me is the electron mass, is the (dimensionless!)


fine-structure constant, and is a quantity known as a
cross section[5] which is known to have the SI dimensions
of an area. The SI version of this equation is:
=

8 2 a b
c ,
3 m2e

(27)

where a and b are to be determined. Now, from Table 1


we see that (27) implies:
L2 = M 2 (M L2 T 1 )a (LT 1 )b
= M 2+a L2a+b T ab ,

(28)

from which we see that:


a = 2, b = 2,

(29)

8
and so (27) has the following SI-units form:
=

8 2 2
.
3 m2e c2

2.

(30)

Exercise 10: Draw the topologically distinct lowestorder Feynman diagrams contributing to the following
processes : (a) + e + e ; (b) e+ + e e+ + e .
(c) Give examples of some higher-order contributions to
the same processes.
Exercise 11: To lowest order, the process e+ + e
+ is given by the Feynman diagrams of Figure 4.
Show that, for electrons and positrons that are almost at
rest, the distance between the two vertices is typically of
order m1
e in natural units, where me is the electron rest
mass. Convert this result to practical units and evaluate
it in fermis.
Exercise 12: Parapositronium is an unstable bound
state of an electron and a positron. Its lifetime is given
in natural units by = 2/(me 5 ), where me is the mass
of the electron and is the fine-structure constant. Restore the factors of and c by dimensional arguments
and evaluate in seconds.
F.

Rates and cross sections

Consider an idealized scattering experiment, which involves a mono-energetic beam of particles being fired at
a target of some form; the rates of production of the various particles can then be measured. The rate at which
a given reaction proceeds will be experiment dependent,
and proportional to:
the number N of particles in the target which are
illuminated by the beam;
the flux J of the incident beam.

Bearing this in mind, we write down the following formula for the rate Wr at which a given reaction r occurs
in a given experiment:
Wr = JN r ,

(31)

where r is known as the cross section for the reaction


r. Importantly, r is a characteristic of a given reaction and independent of the particular experiment used
to probe that reaction; compare this to Wr , J, N , which
are all dependent on the geometry and nature of a given
experiment.
It is easy to see from (31) that the cross section has
the dimensions of an area (prove this!). Note for future
reference that cross sections are often quoted in barns,
where:
1barn = 1b 1028 m2 .

The total cross section , as its name implies, is given


by:
r r ,

(33)

where we have summed over all reactions r which are


allowed by the relevant conservation laws. Similarly, the
total reaction rate W is given by (c/f (31)):
W r Wr = JN .

(34)

Now, to a good approximation for the vast majority


of particle-physics experiments, every reaction removes a
particle from the incoming beam, and so the total cross
section determines how much the beam intensity in diminished after having traversed a target of given density
and thickness.
3.

Dierential cross sections

We are now ready to introduce the concept of the differential cross section dr /d, via the following generalization of equation (31):

Amplitudes and cross sections


1.

Total cross sections

(32)

dWr = JN

dr (, )
d,
d

(35)

where dWr is the measured rate for a particle to be emitted in the direction (, ), into an element of solid angle
d = d(cos )d. The reaction cross section r is obtained from the dierential cross section dr (, )/d by
integrating over all solid angle:

r =

d
0

d(cos )

dr (, )
.
d

(36)

Evidently, the dierential cross section is a useful concept when one is considering the angular distribution of
the particles which are produced in a given scattering
experiment.
Although we have not explicitly indicated this in our
notation dr (, )/d, the dierential cross section will
have a dependence on both the energy and the spin state
(i.e. polarization) of the particles involved in the reaction r. Now, most particle-physics experiments use unpolarized beams and targets, together with polarizationinsensitive detectors which do not measure the spin states
of the product particles. The cross sections pertinent to
such experiments are known as unpolarized cross sections. Furthermore, such experiments typically have
cylindrical symmetry about the beam axis (i.e. the z
axis), and so the unpolarized cross section has no dependence on the azimuthal angle :

dr (, )
dr ()

.
d
d

(37)

One typically works with the following definition of an


unpolarized dierential cross section:
dr

d(cos )

d
0

dr ()
dr ()
= 2
,
d
d

(38)

with:

r =

G.

d(cos )
1

dr
.
d(cos )

(39)

Glossary of jargon

I thought it would be useful to give a glossary


of all of the major particle physics jargon in a single place! The following glossary of particle physics
has been taken from the website of the Particle
Data Group of Lawrence Berkeley National Laboratory, http://particleadventure.org/particleadventure/
index.html.
Annihilation A process in which a particle meets its
corresponding antiparticle and both disappear.
Antiparticle For most particle types there is another
particle type that has exactly the same mass but the
opposite value of all other charges (quantum numbers).
This is called the antiparticle. For example, the antiparticle of an electron is a particle of positive electric charge
called the positron. Most boson types also have antiparticles except for those that have zero value for all charges,
such as a photon or a composite boson made from a quark
and its corresponding antiquark. In these cases there is
no distinction between the particle and the antiparticle;
they are the same object.
Antiquark The antiparticle of a quark.
Baryon Definition 1: A hadron made from three
quarks[6]. The proton (up-up-down) and the neutron
(up-down-down) are both baryons. They may also contain additional quark-antiquark pairs. Definition 2: A
hadron which is a fermion (half-integral spin).
Boson A particle that has integer intrinsic angular
momentum (spin) measured in units of (spin =0, 1, 2,
...). All particles are either fermions or bosons.
Bottom quark The fifth flavor of quark (in order of
increasing mass), with electric charge of -1/3.
Charge A quantum number carried by a particle. Determines whether the particle can participate in an interaction process. A particle with electric charge has
electrical interactions; one with strong charge has strong
interactions, etc.
Charmed quark The fourth quark (in order of increasing mass), with electric charge +2/3.

Collider A collider is an accelerator in which two


beams travelling in opposite directions are steered together to provide high-energy collisions between the particles in one beam and those in the other.
Colour charge The quantum number that determines
participation in strong interactions. Quarks and gluons
carry nonzero colour charges.
Confinement The property of the strong interaction
that quarks or gluons are never found separately but only
inside color-neutral composite objects.
Conservation When a quantity is always the same
before and after a particle reaction, it is said to be conserved. Such quantities include electric charge, energy,
and momentum.
Dark matter Matter that is in space but is not visible
to us because it emits no radiation that we can observe.
The nature of the motion of stars around the centers
of their galaxies implies that about 90 percent of the
matter in a typical galaxy is dark. Physicists speculate
that there is also dark matter between the galaxies, but
this will be much harder to verify.
Down quark The down quark (d) is the second flavor
of quark (in order of increasing mass), with an electric
charge -1/3.
Electric charge The quantum number that determines participation in electromagnetic interactions.
Electromagnetic interaction The interaction due to
electric charge; this includes magnetic interactions.
Electron The least massive electrically-charged particle, hence absolutely stable. It is the most common
lepton, with electric charge -1.
Electroweak interaction In the Standard Model,
electromagnetic and weak interactions are related (unified); physicists use the term electroweak to encompass
both of them.
Fermion Any particle that has odd-half-integer (1/2,
3/2, ...) intrinsic angular momentum (spin), measured
in units of . All particles are either fermions or bosons.
Fermions obey a rule called the Pauli Exclusion Principle, which states that no two fermions can exist in the
same state at the same time. Many of the properties
of ordinary matter arise because of this rule. Electrons,
protons, and neutrons are all fermions, as are all the fundamental matter particles, both quarks and leptons.
Fixed-target experiment An experiment in which
the beam of particles from an accelerator is directed at a
stationary (or nearly stationary) target. The target may
be a solid, a tank containing liquid or gas, or a gas jet.
Flavour The name used for the dierent quarks types
(up, down, strange, charm, bottom, top) and for the
dierent lepton types (electron, muon, tau). For each
charged lepton flavor there is a corresponding neutrino
flavor. In other words, flavor is the quantum number
that distinguishes the dierent quark/lepton types. Each
flavor of quark and lepton has a dierent mass.
Fundamental particle A particle with no internal
substructure. In the Standard Model the quarks, leptons,
and gauge bosons (photons, gluons, W + and W bosons,

10
and the Z bosons) are fundamental. All other objects
are believed to be constructed from these basic building
blocks.
Gluon The carrier particle of the strong interactions.
Graviton The carrier particle of the gravitational interactions; not yet directly observed.
Hadron A particle made of strongly-interacting constituents (quarks and/or gluons). These include the meson and baryons.
Interaction A process in which a particle decays or it
responds to a force due to the presence of another particle
(as in a collision). The four fundamental interactions are
gravitational, electromagnetic, strong, and weak.
Lepton A fundamental fermion that does not participate in strong interactions. The electrically charged leptons are the electron (e ), the muon ( ), the tau ( ),
and their antiparticles. Electrically-neutral leptons are
called neutrinos (e , , ).
Mass The rest mass (m) of a particle is the mass defined by the energy of the isolated (free) particle at rest,
divided by the speed of light squared. When particle
physicists use the word mass, they always mean the
rest mass of the object in question.
Meson Definition 1: A hadron made from an even
number of quark-antiquark constituents. The basic
structure of most meson is one quark and one antiquark.
Definition 2: Hadrons which are bosons (integral spin
particles).
Muon The second flavor of charged leptons (in order
of increasing mass), with electric charge -1.
Neutral Having a net charge equal to zero. Unless
otherwise specified, it usually refers to electric charge.
Neutrino A lepton with no electric charge. Neutrinos
participate only in weak and gravitational interactions
and are therefore very dicult to detect. There are three
known types of neutrinos, all of which have very little
mass.
Neutron A baryon with electric charge zero; it is a
fermion with a basic structure of two down quarks and
one up quark (held together by gluons). The neutral
component of an atomic nucleus is made from neutrons.
Pauli exclusion principle The principle that no two
particles in the same quantum state may exist in the
same place at the same time. Particles that obey this
principle are called fermions; particles that do not are
called bosons.
Photon The force carrier particle of electromagnetic
interactions.
Pion he least massive type of meson, pions can have
electric charges of +1, -1, or 0.
Positron The antiparticle of the electron.
Proton The most common hadron, a baryon with electric charge +1 equal and opposite to that of the electron.
Protons have a basic structure of two up quarks and one
down quark (bound together by gluons). The nucleus of
a hydrogen atom is a proton.
Quark A fundamental fermion that has strong interactions. Quarks have electric charge of either +2/3 (up,

charm, top) or -1/3 (down, strange, bottom) in units


where the proton charge is 1.
Residual interaction An interaction between objects
that do not carry a charge but do contain constituents
that have that charge. Although some chemical substances involve electrically-charged ions, much of chemistry is due to residual electromagnetic interactions between electrically-neutral atoms. The residual strong
interaction between protons and neutrons, due to the
strong charges of their quark constituents, is responsible for the binding of the nucleus.
Spin A quantum particle property of intrinsic angular
momentum.
Stable Does not decay. A particle is stable if there
exist no processes in which a particle disappears and in
its place dierent particles appear.
Strange quark The third flavor of quark (in order of
increasing mass), with electric charge -1/3.
Strong force The interaction responsible for binding
quarks, antiquarks, and gluons to make hadrons. Residual strong interactions provide the nuclear binding force.
Tau The third flavor of charged lepton (in order of
increasing mass), with electric charge -1.
Top quark The sixth flavor of quark (in order of increasing mass), with electric charge 2/3. Its mass is much
greater than any other quark or lepton.
Up quark The least massive flavor of quark, with electric charge 2/3.
Virtual particle A particle that exists only for an
extremely brief instant in an intermediary process. Then
the Heisenberg Uncertainty Principle allows an apparent
violation of the conservation of energy. However, if one
sees only the initial decaying particle and the final decay
products, one observes that the energy is conserved.
W bosons The W + and W bosons are carrier particles of the weak interactions.
Weak interaction The interaction responsible for all
processes in which flavor changes, hence for the instability of heavy quarks and leptons, and particles that contain them. Weak interactions that do not change flavor
(or charge) have also been observed.
Z boson A carrier particle of the weak interactions.
It is involved in all weak processes that do not change
flavor.

II.

LEPTONS, QUARKS AND HADRONS

We mentioned, in section 1.A, that the standard model


attempts to explain the physics of elementary particles
in terms of three distinct basic particles, namely: leptons, quarks and gauge bosons. In this chapter, we concentrate on the leptons and quarks (which are treated as
structureless point particles), together with the hadrons
(such as protons and neutrons) which are treated as composites of quarks which are bound together by the strong
interaction.

11
Accordingly, this chapter is broken up into two major sections: section 2.A considers the leptons, while 2.B
considers the strongly-interacting particles (i.e. hadrons)
which are built from quarks.
A.

where N (e ) denotes the number of electrons, etc.


Exercise 13: By invoking the law of conservation
of the relevant lepton number, show that electrons and
positrons can only be created or annihilated in pairs, in
electromagnetic reactions.

Leptons
1.

Leptons are one of the three classes of basic particles


in the standard model, the other two being the quarks
and gauge bosons (see section 1.A). Leptons are spin-1/2
fermions which do not feel the strong interaction. We
know of the existence of six leptons, which are typically
grouped in three pairs (doublets) known as generations:

e
e

,
,
,

(40)

Electron neutrinos

The existence of the electron neutrino was first proposed by Wolfgang Pauli in 1930, in the context of trying
to understand -decay reactions such as:
(Z, A) (Z + 1, A) + e + e ,

(45)

where (Z, A) denotes the atomic and mass numbers of a


given nucleus. Considered at the level of nucleons rather
than atoms, reaction (45) becomes:

where:
n p + e + e .
e electron,
mu-lepton muon,
tau-lepton tauon.

(41)

The charged leptons in (41) all have the same charge


Q = e. Associated with each of the charged leptons of
(41) are the following three neutral leptons:
e electron neutrino,
mu neutrino,
tau neutrino.

(42)

All of the neutrinos have very small masses.


To complement the six leptons in equation (40), there
are six corresponding anti-leptons:

e+
e

+ +

,
,
.

(43)

The charged leptons feel both the weak and the electromagnetic forces, while the neutral leptons feel only the
weak force.
Each generation of leptons has, associated with it, a
quantum number which is conserved in all known interactions. These so-called lepton numbers are respectively
known as the electron number, the muon number
and the tauon number, and are given by the following
formulae:
Le N (e ) N (e+ ) + N (e ) N ( e )
L N ( ) N (+ ) + N ( ) N ( )
L N ( ) N ( + ) + N ( ) N ( )

(44)

(46)

While they were not experimentally observed in 1930,


the existence of the neutrino in the above reaction was
inferred from both the conservation of energy and angular
momentum[7]. Here, we consider the energy-conservation
part of this argument. If there were no anti-neutrino in
equation (46), then we would have a two-body decay in
which the energy Ee of the emitted electron would have
the unique value:
Ee = m = m(Z, A) M (Z + 1, A),

(47)

where we have neglected the recoil energy of the nucleus.


Now, this putative unique value of the electron energy
is not observed experimentally; rather, one sees that the
emitted electron in (46) has a range of allowed energy
values. This range of energy values is consistent with the
existence of the anti-neutrino on the right side of (46):
with the anti-neutrino present, we expect the energy Ee
of the emitted electron to lie in the following range of
values:
me Ee m m e .

(48)

This is indeed the case.


Since m is known, the mass m e of the electron-antineutrino can be determined by looking at the maximum
energy of the emitted electrons. This gives the following
upper limit for the mass of the anti-electron-neutrino:
m e < 3 105 me .

(49)

This mass is extremely small, and can be neglected


for the purposes of most calculations in particle physics.

12
Note, however, that neutrino masses may have important
consequences over cosmological distance and time scales.
We close this section by pointing out that electron neutrinos and anti-neutrinos can be detected by processes
such as the inverse decays:
e + n e + p
e + p e+ + n

(50)

but the reaction cross sections are extremely small, corresponding to a mean free path in lead of many light
years. Nevertheless, detectors have been built in which
such reactions may be observed.

3.

Lepton decays and universality

The lepton decays (51) and (52) are examples of the


weak interaction; as pointed out in section 1.D.3, such
interactions are often well-approximated by a point interaction with a characteristic strength given by the Fermi
coupling constant GF . For example, the muon decay
given by the second member of equation (51) is sketched
in the left panel of Fig. 9, while the inverse process,
namely inverse muon decay:
+ e + e

(53)

is given in the right panel of the same figure.


2.

Further lepton generations

So far, we have only considered the first of the lepton


generations appearing in equations (40) and (43). We
now have a brief look at the remaining generations.
Second lepton generation: Muons are very penetrating
particles with mass 105.7 MeV/c2 . They are produced in
the upper atmosphere in reactions involving cosmic rays;
the muons so produced decay on their journey down to
sea level, resulting in a sea-level muon flux of about 1
muon per cm2 per minute. Like the electrons, muons are
spin-1/2 fermions.
Third lepton generation: The tauon is heavier than the
muon, with a mass of 1777 MeV/c2 . It too is a spin-1/2
fermion.
The electromagnetic and weak interactions of muons
and tauons are identical to those of electrons, provided
that one takes into account their dierence in mass (we
return to this point in detail, soon).
The electron is a stable particle, because of the fact
that it is the lightest known charged particle. The
muon and tauon are unstable, with respective lifetimes
of 2.2 10 6 s and 2.9 10 13 s. (We will show, in the
next section, that the dierence in the muon and tauon
lifetimes is a consequence of their mass dierences.)
Both muons and tauons decay via the weak interaction.
For example, muons can decay via:
+ e+ + e +
e + e +

(51)

e + + e
+ + .

(52)

while tauons can decay via:

We close this section by introducing the concept of the


branching ratio. The branching ratio gives a measure
of the relative importance of various decay modes, and
is defined to be the fraction of all decays which lead to
a given final state. For example, 17.8 % of particles
decay via the first reaction in equation (52); this reaction
therefore has a branching ratio of 0.178.

FIG. 9: Feynman diagrams for muon decay e + e +


(left panel) and inverse muon decay + e + e (right
panel).

We can now make a comparison of the decay rates of


the muon and the tauon. To do this in a back of the
envelope way, we make two assumptions. (a) The first
assumption, already mentioned in section 2.A.2, is that
the electromagnetic and weak reactions of tauons and
muons are identical to those of electrons, provided that
their mass dierences are taken into account. Such an assumption is known as lepton universality. (b) The second
assumption makes use of the fact that me m m
in order to neglect the masses of the final-state leptons
in lepton decays such as are given in equations (51) and
(52).
Equipped with the above two assumptions, let us return to the muon decay which is sketched in the left panel
of Figure 9. Since we are ignoring the masses of all finalstate leptons, the only dimensioned constant available to
us are GF and m . With reference to Table 1, these constants have respective natural-units dimensions of E 2
and E. Now, the decay rate ( e + e + ) has
the dimensions of an inverse time, which is natural units
becomes the dimension of energy E. We conclude, based
on this purely dimensional reasoning, that the desired
formula for the muon decay rate has the form:
( e + e + ) = KG2F m5 ,

(54)

where K is a dimensionless proportionality constant


which is the same for both muons and tauons (by the
assumption of lepton universality).

13
We invoke lepton universality yet again to write down
the following expression for the rate at which tauon decay
takes place:
( e + e + ) = KG2F m5 .

(55)

Take the ratio of equations (54) and (55):


m5
( e + e + )
= 5.

( e + e + )
m

(56)

To proceed further, we need to make a link between


equation (56) and the lifetimes of the muon/tauon. To
do this, note that the lifetime L of a given lepton L is
related to the decay rate by:

L =

B(L e + e + L )
,
(L e + e + L )

(57)

(58)

The ratio of the branching ratios (!!!) in (58) is


known, by experiment, to have the value 0.178. Also,
m = 105.66 MeV/c2 and m = 1777 MeV/c2 . Thus
(58) yields:

= 1.3 107 ,

Quarks

u
d


c
t
,
,
.
s
b

(60)

The u,c,t quarks have a charge of 2/3, while the d,s,b


quarks have a charge of -1/3 (in units of the modulus
e of the electron charge). The letters u, c, t, d, s, b stand
for the deliberately-whimsical names up, charmed, top,
down, strange and bottom. In addition to the quarks, we
have the following antiquarks:

(59)

which is in excellent agreement with experiment (the


measured lifetimes are = 2.9 1013 s and = 2.2
106 s).
The accuracy of this calculation gives weight to the
theory of lepton universality. Indeed, all known experimental data is consistent with the assumption that the
interactions of the electron and its associated neutrino,
are identical with the interactions of the muon/tauon and
their associated neutrinos.

B.

1.

Six dierent flavours of quark are known to exist


(cf. the six leptons in equation (40)). The six quarks are
grouped into three pairs (generations):

where B is the branching ratio.


Exercise 14: Derive equation (57).
Substitute (55) into (56):
m5
B( e + e + )
=
B( e + e + )
m5
5
m

B( e + e + )
=
= 5
.

m
B( e + e + )

negligible when compared to the strong interactions (unless the strong interaction is forbidden in a given scenario).
Central to our discussion of the strong interaction will
be the quark model of hadrons, whereby the plethora
of hadrons can be understood as various bound states of
just three fundamental quarks (and their antiparticles,
the anti-quarks). These quarks have fractional charges
of 23 e or 13 e, a point which we shall soon examine in
more detail.
For many years, the quark model of hadrons encountered resistance, on account of the fact that free quarks
have never been observed in nature; this non-observation
of free quarks will not be considered until Chapter 7 of
this course. For now, however, we take the existence of
quarks as given, and show how the strongly-interacting
particles can be understood within the quark model.

d
u


s
b
,
,
,
c
t

(61)

which have the opposite charges to their quark counterparts.


Symbol Charge
Mass
d
-1/3 md 0.35
u
2/3 mu md
s
-1/3 ms 0.5
c
2/3
mc 1.5
b
-1/3 mb 4.5
t
2/3 mt 180
Table 2. Quark charges (in units of e) and quark
masses (in GeV/c2 ).

Strongly interacting particles


2.

We now turn our attention to the strongly interacting particles (i.e. the hadrons), which are built up from
quarks. Such strongly-interacting particles also interact
by the other known forces of nature (electromagnetic,
weak, gravitational) but these interactions are typically

Hadrons: General properties

There are over 200 dierent known types of hadrons,


all of which can be considered as quark bound states,
and all of which have an electric charge which is an integer multiple of e. Remarkably, all of these hadrons can

14
be understood in terms of just three dierent classes of
quark bound state [8]:
qqq (baryons)
qqq (antibaryons)
qq (mesons)

(62)

Tables 3 and 4 list a number of common baryons and


mesons; the third column of each of these tables lists the
quark constituents of a given hadron. These tables also
list the values of certain quantum numbers:
S = strangeness = (N (s) N (s))

T = truth = N (t) N (t)


where N (s) denotes the number of strange quarks, etc.,
to distinguish the beauty B
from the
and we write B
baryon number B (we will define the baryon number in
a moment). Note that T = 0 for all currently-known
hadrons.

Particle Mass Quark structure Q S C B


p
938
uud
1 0 0 0
n
940
udd
0 0 0 0

1116
uds
0 -1 0 0
c
2285
udc
1 0 1 0
Table 3. Some baryons, with their mass (in MeV/c2 ),
quark structure, charge Q, strangeness S, charm C and

beauty B.

Particle Mass Quark structure Q S C B


+

140
ud
1 0 0 0
K
494
su
-1 -1 0 0
D
1869
dc
-1 0 -1 0
Ds+
1969
cs
1 1 1 0
B
5279
bu
-1 0 0 -1
Y
9460
bb
0 0 0 0
Table 4. Some mesons, with their mass (in MeV/c2 ),
quark structure, charge Q, strangeness S, charm C and

beauty B.
The quark numbers N (u) N (u) and N (d) N (d)
do not have special names, since (see next exercise!)
B, where the baryon
they can be inferred from Q, S, C, B,
number B is defined to be:
(63)

with N (q) and N (q) denoting the total number of quarks


and anti-quarks, respectively.
Exercise 15: Show that:
1
+ T ),
B = (Nu + Nd S + C B
3

(65)

where Nu N (u) is the number of up quarks and Nd


N (d) is the number of down quarks. Hence show that
B.
Nu , Nd can both be inferred from Q, S, C, B,
B are conAll of the quantum numbers Q, S, C, B,
served in the electromagnetic and strong interactions. As
an example, consider the following strong interaction:
p + p p + n + + ,

(uud) + (uud) (uud) + (udd) + (ud).

= beauty = (N (b) N (b))


B

1
(N (q) N (q)),
3

2
1

(Nu + C + T ) (Nd S B),


3
3

(66)

the quark description of which is:

C = charm = N (c) N (c)

B=

Q=

(64)

(67)

While the above strong interaction clearly conserves


Nu and Nd (as do all strong interactions), the weak interaction does not necessarily conserve quark numbers
such as Nu and Nd . For example, the weak interaction:
n p + e + e

(68)

has the quark description:


(udd) (uud) + e + e ,

(69)

which conserves neither Nu nor Nd . Note, however, that


the weak interactions do conserve both B and Q.
T, B, Q may be used
Now, the quark numbers S, C, B,
to give some insight into the long lifetimes of certain
hadrons[9]. In this context, suppose that a given hadron
is the lightest hadron with a given set of quantum num T, B, Q. Since the strong interaction conbers S, C, B,
serves all such quantum numbers, the said lightest hadron
cannot decay via the strong interaction, and will be longlived (over timescales of the order of 1023 s).
3.

Strange particles, charm and beauty

Around the middle of last century, some new baryons


and mesons were discovered that were produced in strong
interactions, but decayed via the weak interaction. At
the time, this was puzzling, since there seemed to be no
valid reason why the said particles should not decay via
the strong interactions, with a lifetime of the order of
1023 s. Because of this bizarre behaviour, such particle
were dubbed strange particles. As an example of such
a strange particle, we cite the case of the baryon,
which has a mass of 1116 MeV/c2 and decays via the
reactions:
+ p, branching ratio = 0.64,
0 + n, branching ratio = 0.36,

(70)

15
with a lifetime of around 3 1010 s. This long lifetime
suggests that reactions (70) proceed via the weak rather
than the strong interaction. This suggestion, which is
correct, is perhaps most readily understood in terms of
the quark structure of the hadrons involved. For example, the upper member of equation (70) has the following
quark description:
(uds) (du) + (uud),

(71)

from which we see that a hadron of strangeness S = 1


is converted into a pair of S = 0 hadrons. Since neither
the quark number Nd nor the strangeness S is conserved,
this reaction can only proceed via the weak interaction.
Further, since the is the lightest hadron for its given
B, Q, S quantum numbers (B = 1, Q = 0, S = 1), it
can only decay by a weak interaction (which does not
conserve S), thus explaining its long liftime.
Two points of terminology: (a) A strange particle is
one which has a nonzero value for the strangeness quantum number S; (b) Baryons which contain quarks other
than u and d are called hyperons; the is the lightest of
the hyperons.
Until 1974, all known hadrons could be explained as
bound states of three dierent quarks, namely u,d,s. In
1974, however, a new particle named the J/ was discovered, which now understood to be a bound state of a
charmed quark and its anti-particle:
J/(3097) = cc.

(72)

Such states cc are known as charmonium. Since the


charm quantum number of charmonium is C = 0, such
states are said to have hidden charm. Conversely, particles with a nonzero value of the charm quantum number
C, are said to have naked charm. Particles with naked
charm have been detected, as have particles which contain the bottom quark. To date, the top quark has never
been found to form bound hadron states.
Exercise 16: Which of the following reactions are
allowed, and which are forbidden, by the conservation
laws appropriate to the weak interactions (you will need
to look at a table of particle properties):
+ p + + n
e + p e + + + p
+ + e + e
K + 0 + + +
Exercise 17: Classify the following processes into
strong, electromagnetic and weak reactions:
+ p + + + n
+ p + + n

+ n + p
0 e+ + e + e+ + e
p + p + + 0 +
+
D K + + +
0 + e + e
+ p K + p + p
Exercise 18: Six hadrons have the quantum
= (2, 1, 0, 1, 0), (0, 1, 2, 1, 0),
numbers (Q, B, S, C, B)
(0, 0, 1, 0, 1), (0, 1, 1, 0, 0), (0, 1, 1, 1, 0) and
(1, 1, 3, 0, 0). Identify the quark constituents of each
of these hadrons.
III.

EXPERIMENTAL METHODS

The experimental methods associated with elementary


particle physics are diverse, and often monumental in
scale. Because time is short, we will not be able to do
the subject justice. Rather, we will sketch the details behind some of the more important experimental methods
of particle physics.
Up until about the 1950s, cosmic rays were the only
source of suciently high-energy particles for elementary
particle physics experiments. However, nowadays, the
vast majority of particle-physics experiments are undertaken using beams produced by various classes of accelerator. The beams so obtained can then be directed
onto a target so that the interaction products may be
observed. The target may be fixed (e.g. liquid hydrogen) or it may be another beam (method of clashing
beams headlong into one another). The nature, energy
and trajectory of the particles so produced can then be
determined using the various classes of detector.
A.

Accelerators

Most accelerators use electromagnetic forces to accelerate charged particles. We give some important examples
below.
1.

Linear accelerators and synchrotrons

Accelerators may be divided into two classes - those


which accelerate particles along a straight line (linacs
= linear accelerators), and those which accelerate particles around a circle (synchrotrons).
As a simple example of a linac, consider the proton
linear accelerator sketched in Figure 10.
Here, protons from the source traverse a series of drift
tubes, all of which lie in an evacuated pipe; the tubes are

16

FIG. 10: Schematic of a proton linear accelerator.

connected to a radio-frequency (RF) source of alternating voltage as indicated. Protons are accelerated in the
following manner:
Suppose a proton is inside drift tube A; it feels no electric field when inside the tube, but upon emerging from
the tube it is accelerated towards B. Upon entering tube
B, it again enters a field-free region; while traversing B,
the RF changes the polarity of the fields, so that upon
emerging from B the particle is accelerated across the interface from B to C. This process continues, accelerating
the proton. Note that the tubes A,B,C are of increasing
length, due to the fact that the speed of the proton is
increasing as it moves along the accelerator. However,
tubes C,D, etc. are all of the same length, since by the
time the proton enters these tubes, its speed is essentially
that of light; by the axioms of Special Relativity, it can
go no faster.
Synchrotrons work on a similar principle, with the
charged particle being accelerated around a circle. A
magnetic field is used to constrain the particles to a circular path. As with the linac, acceleration is provided
by RF cavities. As the particle accelerates, the magnetic field must be increased, as must the radio frequency;
this must be done in an appropriately synchronized way,
hence the term synchrotron. In the course of its acceleration, a beam typically makes around 10,000 orbits
before attaining the desired maximum energy.

which means that much of the kinetic energy of the incident particles is wasted as kinetic energy of the product
particles; this energy is therefore unavailable for creating
rest-mass in ever more massive particles.
This disadvantage can be avoided by the method of
clashing beams, whereby particles are made to collide
in a head-on fashion. For example, electrons and antielectrons, or protons and anti-protons, can be made to
circulate in opposite directions in the same synchrotron,
and then made to collide head-on. Since the total momentum is zero in the laboratory frame of reference, all
energy is available for particle production, with little being wasted on the kinetic energy of the product particles.
The method of clashing particles has the disadvantages
that: (a) the colliding beams have to be composed of stable particles; (b) the interaction rates are typically lower
than for stationary targets, since liquids and solids are
much denser than the particles in either of the colliding
beams.
Exercise 19: Show that, for a fixed-target experiment, the total energy available for particle production
is, in un-natural units (!!):
ECM =

(73)

where mb is the mass of the beam particle, mt is the


mass of the target particle, and EL is the energy of the
beam particle in the laboratory frame.
Show that, for
high energies, this increases only as EL . Lastly, show
that, for a colliding-beam experiment, the total energy
available for particle production is:
ECM = 2EL .
B.

2.

m2b c4 + m2t c4 + 2mt c2 EL ,

(74)

Particle detectors

Fixed-target machines and colliders

We mentioned, in the opening paragraphs of this chapter, that accelerator experiments may have either fixed
or moving targets. We now examine this point in more
detail.
Fixed-target machines, as the name implies, direct the
beam onto a stationary target, which may be liquid or
solid. Because both liquid and solid targets are dense,
a large number of reaction products can be produced.
Such products may be studied in their own right, or used
to produce so-called secondary beams. Unlike the primary beams, secondary beams can consist of unstable
particles, such as charged pions. Other useful secondary
beams include uncharged particles such as photons or
neutrinos.
One significant problem with fixed-target machines regards their energy wastage. By momentum conservation, the product particles must have some kinetic energy,

Particle detection involves the localization of a given


particle in both space and time, together with the measurement of such relevant parameters as its energy and
momentum. Further, the nature/type of a given particle
must also be ascertained. No existing detector is able
to measure all of these quantities in an optimal manner.
Nowadays, therefore, one typically encounters complex
hybrid detectors which incorporate a number of dierent
classes of sub-detector, all of which are synchronized with
high-speed electronics and computing.
1.

Time resolution: scintillators

When a charged particle traverses a given medium, it


may lose energy by exciting or ionizing atoms in the said
medium. In certain media, known as scintillators, the
subsequent de-excitation of atoms in the medium leads

17
to the emission of visible light. This light can then be
directed (via light pipes) onto the cathode of a photomultiplier tube. This then releases electrons (via the
photoelectric eect), yielding an electronic pulse, the duration of which can be as short as 200 ps. Therefore, the
scintillator makes an excellent timing device, which can
then be used to trigger other detectors into action - i.e.,
based on the scintillator signal, one can decide whether
or not to engage other detectors in order to record more
about the present event.
Typical scintillator materials include sodium iodide
crystals, organic liquids such as toluene, and various plastics.

2.

Position measurement

There are many detectors able to measure the position/trajectory of an elementary particle. Here, we consider a few of the more important position-sensitive detectors.
3.2.2.1 Bubble chambers
While they are rarely used nowadays, bubble chambers were immensely important in the particle-physics
experiments undertaken from the 1950s through to the
1980s. A bubble chamber consists of a large vat of liquid,
typically hydrogen, which is held at a pressure which is
greater than the equilibrium vapour pressure at a given
temperature. Upon suddenly reducing the said pressure,
the liquid is now superheated, i.e. hotter than its boiling temperature at the new pressure. When elementary
particles traverse this superheated liquid, they produce a
trail of ions, at which bubbles nucleate and then expand.
Once these bubbles have expanded suciently, they can
be photographed with an array of cameras. Also, if the
whole chamber is placed in a magnetic field, then the curvature of the trajectories can be used to infer the particle
momenta. An example of a bubble chamber is shown in
Figure 11, while a sample image from a bubble chamber
is shown in Figure 12.
The slow repetition rate of bubble chambers has led
to their abandonment; they have been largely replaced
with the electronic detectors which we now proceed to
describe.
3.2.2.2 Proportional and drift chambers
The basic idea behind the so-called proportional counters is this: if an electric field permeates some region
filled with a given gas, then the electrons (produced by
the ionization events due to the passage of a charge particle) will migrate towards the anode. If the electric field is
suciently strong, then the said electrons will be accelerated so much that secondary ionization events will occur;
these can lead to tertiary ionization, etc.; one therefore
has an electron avalanche which is detected as an electronic pulse at the anode. Proportional counters are so
called because of the fact that the number of secondary
electrons produced is proportional to the number of primary ion pairs.

FIG. 11: The Big European Bubble Chamber (BEBC)


(Photo by CERN.)

The multi-wire proportional counter (MWPC) is


sketched in Figure 13. This detector works on the
same principle as the proportional counter which has just
been described, with each wire acting as an independent
detector. With accurate timing, the pulses registered by
both the anode wires and the cathode strips (see diagram) can yield spatial information at 0.05 mm resolution.
The MWPC has now been replaced by the drift chamber. The drift chamber is able to operate with many
fewer wires than the MWPC. The basic setup of the drift
chamber is sketched in Figure 14.
We once again see the anode and cathode wires used
to detect electron avalanches, but in addition there are
field shaping wires which cause the said avalanches
to drift towards the anode at speeds of the order of 40
m/ns. If a reference time is determined with the use
of fast timers such as a scintillator, then the drift time
can be used to accurately determine position, using far
fewer wires than would be needed for a MWPC of the
spatial resolving power. Very large drift chambers can
therefore be built, with volumes of 100 m3 or more.
Both MWPCs and drift chambers can be built in a variety of geometries: planar, radial, cylindrical etc. The
last-metioned type are called jet chambers, and example of which is shown in Figure 15. A sample of data
taken on a jet chamber is shown in Figure 16.

3.

Momentum measurement

As we have already mentioned, momentum may be


measured by looking at the curvature of a particles tra-

18

FIG. 13: Schematic of a multi-wire proportional chamber


(MWPC).

FIG. 12: Sample image taken on the Big European Bubble Chamber (Image from G.G.Harigel, Bubble chambers, technology and impact of high energy physics,
http://dbserv.ihep.su/pubs/aconf00/tconf00/ps/ c6-4.pdf.)

jectory in an applied magnetic field (an example of such


curved trajectories was seen in Figures 12 and 16). More
precisely, the momentum p of a particle of known charge
e is related to both the applied magnetic field B and the
radius of curvature of the tracks by:
FIG. 14: Schematic of a drift chamber.

p = eB.

(75)

Exercise 20: Derive equation (75).


4.

Particle identification

Means of determining the identity of a particle often


rely on measurements of the mass of the said particle.
If the momentum is known, and either the velocity or
the energy is also known, then the mass can be easily
inferred.
Exercise 21: Prove the claim made in the previous
sentence.
We defer a study of energy measurement until section
3.B.5, and here consider two dierent forms of velocity
measurement.
An obvious method for measuring velocity is to determine the time of flight between a pair of scintillators.
Unfortunately, this simple method has only limited usefulness, because of the fact that, in a given experiment,

the velocity of most of the particles is typically extremely


close to that of light.
The most common method for velocity measurement,
in the context of particle identification, makes use of

Cerenkov
radiation. This phenomenon is the visible-light
analogue of the sonic boom produced by aircraft which
travel faster than the speed of sound in air. Accordingly,

Cerenkov
radiation is produced when a charged particle
has speed v which obeys the inequality (c/n) < v < c,
when c/n is the speed of light in a medium of refractive
index n. With reference to Figure 17, we see the visiblelight wavefronts emitted by a charged particles at three
successive times t = 1, 2, 3.
In the above diagram, denote by t the time which has
passed since the particle was at position A. The cosine
of the angle in the diagram is given by:

cos =

distance from A to B
distance from A to C

19

FIG. 15: Image of the JADE jet chamber, while it was being
assembled (image taken from Martin and Shaw, page 61).

FIG. 17: Cerenkov


radiation due to a particle which travels
faster than the speed c/n of light in a medium of refractive
index n .

ther particles; the process continues until a cascade or


shower develops. This leads to almost all of the particles energy being deposited in the energy-measurement
device.
C.

The Brookhaven experiment

As a simple example of a hybrid detector which makes


use of some of the concepts we have discussed, we consider the 1974 Brookhaven experiment which led to the
discovery of the J/ particle (see equation (72)). This is
a short-lived particle (lifetime 1020 s) and so it must
be detected indirectly by looking at its decay products.
The Brookhaven experiment used a 28.5 GeV/c proton
beam produced by the synchrotron at Brook-haven National Laboratory, USA. This beam was directed onto a
fixed solid target made from beryllium, in order to study
the reaction:

FIG. 16: Computer reconstruction of a two jet event which


was recorded using the JADE jet chamber (image taken from
Martin and Shaw, page 62).

(c/n) t
c
1
=

,
vt
nv
n

(76)

where v is the velocity of the particle and v/c. Thus


measurement of can be used to infer the speed v of the
particle.

5.

Energy measurement

Energy measurement is typically done using the process of absorption, where the particle in question interacts with the material of an absorber, giving energy to
secondary particles which themselves give energy to fur-

p + p J/ + X,

(77)

where X denotes any reaction products which are allowed


by the relevant conservation laws. Because it is so shortlived, the J/ rapidly decays via the reaction:
J/ e+ + e ,

(78)

and it was the resulting e+ e pair which was detected in


the experiment.
Before going further, we need to briefly revise some
Special Relativity! The space-time coordinate vector
x , = 0, 1, 2, 3, is given by x (ct, x), where x denotes spatial position in a given (inertial) frame of reference, and all other symbols have their usual meaning. The four-momentum p is given by p (E/c, p),
where E is the energy in a given inertial frame of reference, and p is the ordinary three-momentum. The
square (E/c)2 p2 of the four momentum (please

20
note the negative sign!) is the same for all inertial
frames of reference, so therefore it may be evaluated in
the rest frame of the particle (where p = 0) , to give
E 2 /c2 = m2 c4 /c2 = m2 c2 , where m is the rest mass of
the particle. We close this bit of revision by noting that
the four-momentum is a conserved quantity in all interactions; for example, in the reaction A + B C + D, the
total four-momentum of the reactants A + B is the same
as the total four-momentum of the products C + D.

FIG. 19: Plot of the invariant mass W of the detected e+ e


pairs which arose from reaction (78) in the Brookhaven experiment (image taken from Martin and Shaw, page 72).

FIG. 18: Schematic of the Brookhaven experiment, with

dipole magnets M , Cerenkov


detectors C, shower counters
S, and multi-wire proportional counters D (image taken from
Martin and Shaw, page 72).

We now return to reaction (78). By the conservation


law of four-momentum as applied to reaction (78), we
have:

EJ/
, pJ/
c

Ee
Ee+
, pe +
, pe+
c
c

Ee + Ee+
, pe + pe+
c

2
ECM
= 2EA EB (1 + cos ),

(79)

where EJ/ denotes the energy of the J/, pJ/ denotes


its momentum, etc. Squaring both sides of equation (79)
- remember the negative sign! - and then evaluating the
frame-independent quantity on the left side in the rest
frame of the J/, we obtain:
m2J/ c2 = c2 (Ee + Ee+ )2 (pe + pe+ )2

(Ee + Ee+ )2
(p + p + )2
mJ/ =
e 2 e
4
c
c

Hence the rest mass mJ/ of the J/ particle may


be determined from the energy and momentum of each
member of the e+ /e pair produced when it decays via
(78).
Using the experimental setup shown in Figure 18, the
Brookhaven group obtained the plot shown in Figure 19.
This plot shows the number of events versus the righthand side of equation (80). There is a clear peak corresponding to the mass 3097 MeV/c2 , which was quoted in
equation (72).
Exercise 22: At a collider, two high-energy particles
A and B, with respective energies EA and EB which are
much greater than either of the rest masses, collide at
a crossing angle . Show that the total centre-of-mass
energy ECM is given by:

(80)

(81)

if the rest masses are neglected.


Exercise 23: At HERA, a 30 GeV electron beam collides with an 820 GeV proton beam at zero crossing angle.
Please evaluate the total centre-of-mass energy and show
that a fixed-target electron accelerator would require a
beam energy of approximately 50,000 GeV to achieve
the same total centre-of-mass energy. Briefly comment
on this result.
Exercise 24: Estimate the thickness of iron through
which a beam of neutrinos with energy 200 GeV must
travel if 1 in 109 of them are to interact. Assume that
the neutrino-nucleon cross section is 1038 E cm2 ,

21
where E is given in GeV. The density of iron is 7.9
g/cm3 .
Exercise 25: Two particles of masses m1 and m2 , and
common momentum p, travel between two scintillation
counters which are a distance L apart. (a) Show that
the dierence in their flight times decreases like p2 for
large momenta. (b) Hence calculate the minimum flight
path necessary to distinguish pions from kaons if they
have momentum 3 GeV/c, and the time of flight can be
measured with 200 ps accuracy.

IV.

SPACE-TIME SYMMETRIES

One of the more important concepts in physics is that


of the symmetry, or invariance, of the equations describing a system, under operations such as rotations or
translations in space. As we shall soon see, such symmetries are linked to associated conservation laws[10]. For
example, the invariance of a system under spatial translations implies the conservation in time of linear momentum for the said system, while the invariance of a system
under spatial rotations implies the conservation of angular momentum.
We conclude this brief introduction with a summary
of the remainder of this rather abstract chapter. Section
4.A shows that translational invariance leads to the conservation in time of linear momentum. In section 4.B,
we see that rotational invariance leads to the conservation of angular momentum. In 4.C, we show that invariance under spatial reflection implies the conservation in
time of a quantum number known as parity. Finally,
in 4.D we see that invariance under an operation known
as charge conjugation (which replaces all particle by
their antiparticles) implies the conservation of a quantum number known as C-parity.

A.

Translational invariance

If we consider a system of particles on which there


are no external forces acting, then the translation of the
whole system from one place to another leaves unaltered
all physical properties of the system. Stated more formally, this implies that the systems Hamiltonian is invariant under spatial translations.
Consider, then, a displacement of the system by some
constant vector a, so that the position coordinate xi of
each particle transforms to xi according to:
xi xi = xi + a.

(82)

In what follows, we need only consider infinitesimal


displacements a = x. Under such an infinitesimal transformation, the original Hamiltonian H(x1 , x2 , ) gets
changed into a new Hamiltonian:

H(x1 , x2 , ) H(x1 , x2 , )
= H(x1 + x, x2 + x, ).

(83)

Let us restrict ourselves to the case of a single free


particle of mass m, the Hamiltonian for which expresses
the fact that all energy is kinetic:
H = 2 /(2m), 2 = x2 + y2 + z2 .

(84)

Since spatial derivatives of position (rather than position itself) appear in (84), the displacement (82) leaves
the Hamiltonian unchanged, and so:
H(x ) H(x + x) = H(x).

(85)

Stated dierently, the Hamiltonian is invariant under


defined by:
the translation/displacement operator D

D(x)
(x + x),

(86)

where (x) is an arbitrary wavefunction. Next, expand


the right side of (86) in a first-order Taylor series, to give:

D(x)
= (x) + x (x)
= (1 + x )(x),

(87)

from which we infer that:


= 1 + x = 1 + ix p
,
D

(88)

i is the momentum operator in natural


where p
units.
To continue, see what happens when we apply the
Hamiltonian operator H(x), followed by the displace to a given wavefunction (x):
ment operator D,

DH(x)(x)
= H(x + x)(x + x)
= H(x)(x + x)(cf. (85))

= H(x)D(x).

(89)

This may be re-arranged to give:

(DH(x)
H(x)D)(x)
=0

H(x) (x) = 0,
D,

(90)

where [A, B] AB BA; (note that [A, B] is known as


the commutator of A and B ... also, if [A, B] = 0 then
we say that A commutes with B). Now, due to the
arbitrary nature of the wavefunction (x) appearing in
(90), we see that:

22

H(x) = 0.
D,

(91)

If we substitute (88) into (91), we find that:


, H(x)] = 0 = [
[1 + ix p
p, H] = 0.

(92)

We now invoke the following result from elementary


quantum mechanics: if an operator (which does not depend explicitly on time) commutes with a given Hamiltonian H, then it will be a constant of the motion.
Exercise 26: Prove the statement made in the previous sentence.
Hence we infer from (92) that, as a consequence of the
translational invariance of the Hamiltonian (84), linear
momentum is conserved for a quantum-mechanical system which is described by the said Hamiltonian.
This completes our first example of the link between
symmetries and associated conservation laws.
Exercise 27: Generalize the discussion which led to
(92), to the case of an N-particle system which is described by the wave-function (x1 , x2 , , xN ).
B.

Rotational invariance

We have just shown that translational invariance implies the conservation of linear momentum. We now show
that rotational invariance implies the conservation of angular momentum.
Consider an infinitesimal rotation by an angle , in
the anti-clockwise direction, about the z axis of an xyz
Cartesian coordinate system. Under this rotation, the
coordinates xi (xi , yi , zi ) of each particle transform to
xi (xi , yi , zi ) according to:
xi xi = xi cos yi sin ,
yi yi = xi sin + yi cos ,
zi zi = zi .

(93)

Since is infinitesimal, cos 1 and sin ,


and so (93) becomes:
xi xi = xi yi ,
yi yi = xi + yi ,
zi zi = zi .

(94)

Restricting ourselves to the case of a single (spinless)


particle in a central potential V (r), we drop the i subscript:
x x = x y,
y y = x + y,
z z = z.

introIn analogy with the displacement operator D


duced in the previous section, we now define the operator
z () as that which eects an infinitesimal rotation by
R
an angle about the z axis:
z ()(x) = (x y, x + y, z).
R

(96)

Expanding the right side in a first-order Taylor series,


we obtain:
z ()(x) = (x y, x + y, z)
R
(x, y, z) yx (x, y, z) + xy (x, y, z)
= (1 yx + xy )(x, y, z),
(97)

from which we see that:

z () = 1 + (xy yx )
R
= 1 + i{i(xy yx )}.

(98)

z () = 1 + iL
z = 1 + iz L,
R

(99)

H = 0,
L,

(100)

J = L + S.

(101)

The term in braces is equal to the quantum-mechanical


z of the orbital angular
expression for the z component L
and so:
momentum operator L,

where z is a unit vector which is parallel to the positive


z axis. The above equation is analogous to expression
(88), which was derived in the previous section. Using
similar logic to that which led from (88) to (92), one can
show that (99) leads to:

which expresses conservation of angular momentum[11]


for a spinless particle in a central potential.
Exercise 28: Obtain (100) from (99).
To sum up, we have shown that invariance under rotation of a given Hamiltonian (e.g. central potential)
implies the conservation of orbital angular momentum
(for a spinless particle).
Question: How does this finding generalize to the case
of a particle with spin, for which the total angular momentum J is the sum of the orbital angular momentum
L and the spin[12] angular momentum S:

We state, without proof, that for this case (i.e. a particle with spin in a central potential) the total angular
momentum J is conserved:
H] = 0,
[J,

(102)

whereas the orbital and spin angular momenta L, S are


not separately conserved:

(95)

H] = 0,
[L,

H] = 0.
[S,

(103)

23
C.

Parity

The previously-mentioned transformations, namely


spatial translations and rotations, were examples of continuous transformations. In this and the following section, we turn our attention to the discrete transformations of parity and charge conjugation, respectively.
The parity transformation is:
xi xi = xi ,

(105)

Before proceeding further, we note that parity is not


an exact symmetry of nature, because it is violated by the
weak interaction. However, parity is known to be a symmetry under the strong and electromagnetic interactions.
For most particle-physics scenarios, the strong and electromagnetic interactions completely dominate the weak
interaction, and so, despite the fact that parity is not
an exact symmetry which is respected by all laws of nature, it is a very useful approximate symmetry when one
considers processes which proceed via the strong or electromagnetic interactions. Accordingly, for the remainder
of this section, we examine the consequences of symmetry
under the parity transformation.
For a single particle a with wavefunction a (x, t), the
parity operator P is defined via:
P (x, t) Pa (x, t),

(106)

where Pa is an eigenvalue known as the intrinsic parity


of the particle a. Since acting with the parity operator
twice leaves the wave-function unchanged, we must have:
Pa2 = 1 = Pa = 1.

(107)

For a many-particle system described by the multiparticle wave-function (x1 , x2 , , t), (106) generalizes
to:
P (x1 , x2 , , t)
= P1 P2 (x1 , x2 , , t).

(108)

We now show that a particle with definite orbital angular momentum is an eigen-state of parity. The spatial
part nlm (r, , ) of the wave-function for such a particle
is:

(109)

Here, (r, , ) are spherical polar coordinates, Rnl (r) is


the radial wave-function, and the spherical harmonics
Ylm (, ) are:

(104)

where, as usual, xi is the position vector of the ith particle. A system, described by Hamiltonian H(x1 , x2 , ),
is invariant under the parity transformation if the said
Hamiltonian obeys:
H(x1 , x2 , ) H(x1 , x2 , )
= H(x1 , x2 , ).

nlm (r, , ) = Rnl (r)Ylm (, ).

Ylm (, )
(2l + 1)(l m)! m
Pl (cos )ei ,
4(l + m)

(110)

where:
(1)m sinm
Plm (cos ) =
2l l!

m+l
d

( sin2 )l , m l.
d(cos )

(111)

Exercise 29: Using the relations:


x = r sin cos ,
y = r sin sin ,
z = r cos ,

(112)

between Cartesian and spherical polar coordinates, show


that the parity transformation is:
r r = r,
= ,
= + .

(113)

Hence show that, under the parity transformation, the


spherical harmonics become:
Ylm (, ) Ylm ( , + )
= (1)l Ylm (, ).

(114)

Making use of the results of the previous exercise, we


see that the action of the parity operator on the wavefunction in (109) yields:
P nlm (r, , ) = P Rnl (r)Ylm (, )
= Pa Rnl (r)Ylm ( , + )
= Pa Rnl (r)(1)l Ylm (, )
= Pa (1)l nlm (r, , ).

(115)

This concludes our proof of the statement, made just


after equation (108), that a particle with definite orbital
angular momentum is an eigen-state of parity. More
precisely, equation (115) shows that the wave-function

24
nlm (r, , ) for a particle of definite orbital angular momentum, is an eigenstate of the parity operator P , with
eigenvalue Pa (1)l , where Pa is the intrinsic parity of the
particle.
As was the case with the previous sections, invariance
of the Hamiltonian under a certain transformation leads
to an associated conservation law[13]. For the case at
hand, then, invariance of a Hamiltonian H under the
parity transformation implies that parity P is conserved,
i.e.:

D.

P , H = 0.

(116)

Charge conjugation

Charge conjugation C is the operation which replaces


all particles by their antiparticles. While charge conjugation is an exact symmetry of both the electromagnetic
and strong forces, it is violated by the weak force (cf. the
discussion immediately following equation (105)). Thus,
if we ignore the weak interaction, the Hamiltonian H of
a given quantum-mechanical system is invariant under C
and hence obeys the associated conservation law:

H = 0.
C,

(117)

In the context of charge conjugation, we note that


there are two classes of particle, namely those which have
distinct antiparticles (such as the proton p and electron
e ) and those which do not (such as the photon and the
neutral pion 0 ). Denote the former class by a and the
latter class by . Going further with this notation, denote by |a, the wave-function of a particle which has
a distinct antiparticle, and denote by |, the wavefunction of a particle which does not have a distinct antiparticle. The operation of C on each of these two classes
of wave-function is:
= |a, ,
C|a,

= C |, ,
C|,

(118)
(119)

where a is the antiparticle of a, and C is an eigenvalue


of modulus unity. If we apply C to both sides of (119),
we see that C 2 = 1 and so:
C = 1.

(120)

Exercise 30: Why didnt we include a factor of Ca


on the right side of (118)?
The quantity C = 1 is known as the C-parity of
a given particle (the said particle must not possess a
distinct particle if its C-parity is to be defined).

These ideas can be extended to many-particle systems described by the multi-particle wave-function
|1 , 2 , ; a1 , a2 , ; :
1 , 2 , ; a1 , a2 , ;
C|
= C1 C2 |1 , 2 , ; a1 , a2 , ; .
V.

(121)

A CLOSER LOOK AT HADRONS

In this chapter, we lay the foundations for a more detailed study of the hadrons (strongly interacting particles). Recall that, in the standard model of elementary
particle physics, the hadrons are viewed as composite
particles constructed from quarks. The quarks themselves are structureless point particles.
After a quick piece of revision, we begin our detailed
study of the hadrons by looking at the so-called isospin
symmetry, an approximate symmetry which makes use of
the fact that the masses of the up and down quarks are
very nearly equal (c/f Table 2). We will then move on
to a back-of-the-envelope calculation which estimates the
small mass dierence between the up and down quarks.
Having done this, we then consider the so-called resonances, which are highly unstable hadrons that are
viewed as excited states of the underlying quark systems;
the short lifetimes of these resonances, which are typically of the order of 1023 seconds, can be determined
using a suitable extension of the techniques introduced
in section 3.C. We move onto a brief introduction to the
quark diagrams, which depict strong interactions at the
underlying quark level. The chapter closes with a description of the quark-antiquark potential.
A.

Revision

Before launching into a more detailed study of the


hadrons, we collect together some of the relevant results
from the previous lectures.
Hadrons are strongly-interacting particles which are
conveniently characterized by a number of parameters,
the most obvious of which is the (rest) mass. In addition
to the mass, there are the quantum numbers, obtained in
the context of space-time symmetries, which were introduced in the previous chapter - (a) spin, (b) parity, and,
for particles that do not possess a distinct anti-particle,
(c) the C-parity. The spin and parity of a given particle
are denoted by J P , where J is the spin and P = 1 is
+
the parity (example: the electron has J P = 12 , meaning
1
that it has spin 2 and parity +1. For particles with Cparity of C, we use the notation J P C (example: the 0
is a 0+ particle).
There are further quantum numbers, which are not
associated with space-time symmetries; such quantum
numbers are known as internal quantum numbers. These
and T (see section 2.B.2). Q and
include Q, B, S, C, B

25

B are conserved in all known interactions. While S, C, B


are conserved in strong and electromagnetic interactions,
they are not necessarily conserved in weak interactions.
The values, for quarks, of the quantum numbers
are listed in Table 5. The values of these
Q, B, S, C, B
quantum numbers for hadrons can be obtained by summing the relevant quantum numbers of their quark constituents. Note also that, to get the relevant quantum
numbers for the anti-quarks d, u, s, c, b, t, multiply the
said quantum numbers of the corresponding quark by
1.
T
Quark Q B S C B
d
-1/3 1/3 0 0 0 0
u
2/3 1/3 0 0 0 0
s
-1/3 1/3 -1 0 0 0
c
2/3 1/3 0 1 0 0
b
-1/3 1/3 0 0 -1 0
t
2/3 1/3 0 0 0 1
Table 5. Values, for the six quarks, of the quantum
T.
numbers Q, S, C, B,
We close this revision by recalling that all of the known
hadrons can be understood as belonging to three dierent
classes of quark bound state: (1) baryons qqq; (2) antibaryons qqq; (3) mesons qq.

B.

n(940) = udd,

(122)

while a less familiar example is:


K+ (494) = us,

K0 (498) = ds.

(124)

Evidently, the hypercharge Y is the same for all members of a given isospin multiplet (c/f the first sentence of
section 5.B). The second of our new quantum numbers
is:

I3 Q

Y
.
2

(125)

Since, as argued earlier, each member of a given isospin


multiplet diers in charge from its neighbour by 1, we
see that I3 takes on a dierent value for each member of
such a multiplet. Denote, by the isospin quantum number
I, the maximum value for I3 within a given multiplet:
I (I3 )max .

(126)

You will show, in the next exercise, that all isospin


multiplets contain (2I + 1) members, with:
I3 = I, I 1, , I.

Isospin

Regarding the hadrons, one notices that they occur


in families, each member of which has the same spin,
the members have very similar masses,
parity, B, S, C, B;
but dier in their electric charge Q.
The most familiar example of such a family is:
p(938) = uud,

+ T.
Y B+S+C +B

(127)

Exercise 31: (a) Deduce the entries listed in Table


6. (b) We now consider the isospin formalism in light of
the quark model of hadrons. In this context, note that
hadron isospins I3 are additive quantum numbers, so that
the I3 of a given hadron is simply the sum of the I3 s for
the constituent quarks:
(quark 1)
(quark 2)
I3 = I3
+ I3
+ .

(128)

Bearing this in mind, show that:


(123)

We see, immediately, that subsequent members of such


families are obtained by replacing up quarks with down
quarks. Since up and down quarks have approximately
equal masses, and dier in their charge by +1 (in units
of the electron charge ), we conclude that each member
of the family will have approximately equal masses and
dier only in their charge Q. These families are known
as isospin multiplets. Such multiplets reflect an approximate symmetry of hadrons known as isospin symmetry,
the theory for which was formulated by Heisenberg in the
early 1930s.
To study this theory in more detail, we need to introduce some new quantum numbers, both of which are
constructed from quantum numbers that are already familiar to us. The first of these quantum numbers, termed
the hypercharge Y , is:

I3 =

1
(Nu Nd )
2

(129)

for any hadron, where Nu is the number of up quarks in


the hadron, and Nd is the number of down quarks. (c)
Interpret the results of (b).
We leave the isospin formalism to one side for the moment, but will be making much use of this formalism in
later sections.
Quark B Y
Q I3
I
d
1/3 1/3 -1/3 -1/2 1/2
u
1/3 1/3 2/3 1/2 1/2
s
1/3 -2/3 -1/3 0
0
c
1/3 4/3 2/3 0
0
b
1/3 -2/3 -1/3 0
0
t
1/3 4/3 2/3 0
0
Table 6. Values, for the six quarks, of the quantum
numbers B, Y, Q, I3 , I.

26
C.

Exercise 32: Show, from (131), that:

Mass splittings of the baryons

As we have seen, the usefulness of the concept of


isospin symmetry relies on the fact that the up and down
quarks have approximately equal masses. However, there
is a small dierence between the mass of the up and the
down quarks. In this section, then, we give a very nice
back-of-the-envelope calculation, which estimates the difference in masses between the up and down quarks.
The three baryons form an isospin multiplet, with
the following values for mass (in MeV/c2 ) and quark
structure:
+ (1189) = uus,
0 (1193) = uds,
(1197) = dds.

the rest masses of the constituent quarks;

the energy due to the strong interactions between


the quarks in the baryon;
the energy due to the electromagnetic interaction
between the quarks in the baryon.
We assume that the energy M0 , due to the strong interactions between the quarks, is the same for all members
+ , 0 , of the baryon multiplet. Also, assume that
the electromagnetic energy between any pair of quarks is
proportional to the products of the charges of the said
quark pair, with a constant of proportionality . Hence
we write:

1
(m + m0 2m+ )
3
2
= 3.7MeV/c .

(132)

Bearing in mind the results of this exercise, we see that


we have obtained an estimate for the (small) mass difference between the up and down quarks, using a rough
calculation whose accuracy is vindicated in more sophisti2
cated treatments. Given that mu md 0.35 GeV/c ,
we see that the masses of the up and down quarks dier
by about 1 %.

(130)

Again, we see that consecutive members of an isospin


multiplet dier by the conversion of an up quark into
a down quark; since the up and the down quarks dier
in their electric charge by +1, consecutive members of
an isospin multiplet dier in their charges by the same
amount.
For the purposes of our back-of-the-envelope calculation, we consider the following three contributions to the
rest mass of a given baryon:

m M0 + ms + 2md
+(e2d + ed es + ed es )

= M0 + ms + 2md +
3
m0 M0 + ms + md + mu
+(eu ed + eu es + ed es )

= M0 + ms + mu + md
3
m+ M0 + ms + 2mu
+(e2u + eu es + eu es )
= M0 + ms + 2mu ,

md mu

D.

Resonances

Resonances are extremely short-lived particles (on human timescales), with characteristic lifetimes of the order
of 1023 seconds. As remarked earlier, this is approximately the time taken for light to travel a distance of
one proton width. Indeed, the resonances are the most
transient natural phenomena which have ever been studied.
Given their extreme transience, it is out of the question
to determine the masses and lifetimes of the resonances
by looking at their tracks in a position-sensitive detector.
Instead, one must rely on indirect techniques very similar
to those introduced in section 3.C, in the context of the
Brookhaven experiment. There, we saw a means of using
the products to which a short-lived particle decays, in
order to determine the rest mass of the said short-lived
particle.
As a concrete example of this chain of reasoning, consider the formation of an unknown resonance X 0 in the
reaction:
+ p X 0 + n,

(133)

where the resonance X 0 rapidly decays via:


X 0 + + .

(134)

The net reaction is:

(131)

where mu is the mass of the up quark, md and ms are


similarly defined, eu is the charge of the up quark, etc.

+ p + + + n,

(135)

Following the logic which led to equation (80), we can


calculate the invariant mass W as:
W =

c4 (E+ + E )2 c2 (p+ + p )2 ,

(136)

a quantity which, if the product + pair are due to the


decay of a short-lived particle, will peak strongly at the

27
rest mass of the said unstable particle (re-read section
3.C if this point is obscure).
Accordingly, Figure 20 plots a histogram, for the +
pairs produced in the net reaction (135), of the number
of times a given W occurs[14] (vertical axis), as a function of W (horizontal axis). We see three peaks, the
centroids of which yield the rest masses of three resonances, termed 0 (769),f20 (1275) and 0 (1700). Further,
the width W of each resonance peak is related to the
lifetime of the associated resonance, via the energy-time
uncertainty principle, which, in natural units, gives:
W. 1 1/W.

(137)

Exercise 33: Show that, if 1023 seconds, then


2
W 100MeV/c , which is similar to the width of the
peaks shown in Figure 20.
FIG. 21: Spectrum of the proton and its excited states. The
vertical axis gives the mass of the various particles, with the
horizontal axis listing their spin J and parity P, using the
notation introduced in section 5.A. The proton itself is represented by the line of lowest energy, in the bottom left corner
of the graph. Image taken from Martin and Shaw, page 123.

+ + p ++ + + p.

(138)

The quark diagram, corresponding to the above reaction, is given in Figure 22.

FIG. 20: Histogram of the invariant mass for pion pairs produced in reaction (135). The horizontal axis gives the invariant mass W (see equation (136)), while the vertical axis gives
the number of times that a given invariant mass W occurred.
Image taken from Martin and Shaw, page 117.

We close this section by noting that the resonances


can be viewed as excited states of the underlying quark
system, which rapidly decay to the ground state. For
example, the proton has a number of excited states (resonances!), as shown in Figure 21. Note that the lowest
energy state is the proton itself.
The processes of resonance formation and decay, together with other strong interactions, may be clearly visualized using the so-called quark diagrams. The quark
diagrams describe strong interactions at the quark level,
using quark lines to indicate the trajectories of the various quarks involved in the given interaction. As an example, consider the following process for the formation
and subsequent decay of the ++ resonance:

FIG. 22: Quark diagram representing, at the quark level, the


formation and subsequent decay of the ++ resonance, via
reaction (138).

E.

The quark-antiquark potential

We begin with a piece of terminology: the mesons cc


and bb are respectively known as charmonium and bot-

28
tomium. These systems have a number of measurable excited states (c/f the discussion on resonances as excited
states of quark systems). The energy states of the charmonium and bottomium systems are consistent with the
following expression for the quark-antiquark potential:
a
V (r) = + br,
r

(140)

In the following two sub-sections, we separately consider the previously-mentioned super-multiplets of (i) the
lightest mesons; and then (ii) the lightest hadrons.

(139)

where r is the distance between the quark and the antiquark, and a, b are constants whose value does not concern us. The above potential is somewhat bizarre in that
it is Coulomb-like at small r, but rises linearly with r
when r is large.
Exercise 34: Derive the allowed combinations of
charm C and electric charge Q for mesons and baryons,
in the context of the quark model of hadrons.
Exercise 35: In the simple model for the baryon
mass splittings given in section 5.C, the electromagnetic
interaction between a given pair of quarks was assumed to
be equal to ea eb , where ea and eb are the quark charges
in units of e. Extend that piece of reasoning to deduce the
value of , and hence make a rough estimate for the mean
distance r between the quarks in the baryons, assuming that their electromagnetic interaction is essentially
2
Coulomb-like. (Answers: 1.7 MeV/c , r 1 fm.)
F.

Y = B + S.

The lightest hadrons

Mendeleevs periodic table (1869) revealed a deep underlying order to the plethora of elements; this order, we
now know, arises from the fact that atoms are composite
particles made up of protons, electrons and neutrons.
Just as the elements can be arranged in the periodic
table, so too can sets of hadrons be arranged into socalled super-multiplets, such as those shown in Figures
23 and 24. The intriguing geometric structure of these
super-multiplets preceded the quark theory of hadrons;
indeed, the super-multipletss geometric structure was a
major piece of evidence which led to the formulation of
the quark hypothesis.
Rather than taking a historical point of view which
obtains the quark model from the super-multiplets, we
will do the opposite and use the quark model of hadrons
to deduce the super-multiplet structure.
Our discussions will rely heavily on the following two
assumptions:
Assumption #1: Since we are restricting ourselves to
a study of the lightest hadrons, we limit section 5.Fs
considerations to hadrons made from the three lightest
quarks only, namely: u,d,s (see Table 2).
Assumption #2: As a further restriction, we only
study hadrons whose constituent quark system has zero
orbital angular momentum L.
Evidently, for the light hadrons to which we are con = T = 0, and hence definition
fining our study, C = B
(124) for the hypercharge Y reduces to:

1.

The light mesons

Recall from section 2.B.2, equation (62), that the


Standard Model considers the mesons to be made
from a quark-antiquark pair. Since, in our study of
the lightest mesons, we are only considering u, d, s
as our building blocks, there are evidently only
nine quark-antiquark combinations to be considered:
uu, ud, us, du, dd, ds, su, sd, ss.
Now, we only care about states with definite orbital
angular momentum L = 0, which have l = 0 (where
l is the orbital angular momentum quantum number).
We saw in (115) that a single-particle wave-function
nlm of definite orbital angular momentum, has parity
P = Pa (1)l , where Pa is the intrinsic parity of the particle a. For a two-particle wave-function, this generalizes
to P = Pa Pb (1)l , where a and b are the labels for the
two particles. What does this have to do with mesons?
Well, the wave-function for a given meson ab is such an
example of a two-particle wave-function. Hence the parity for a meson with l = 0 (see Assumption #2) is given
by P = Pa Pb (1)l = Pa Pb . We take this further: it can
be shown, using the parity invariance of the Dirac equation (see section 1.B.1) that the intrinsic parity Pa of any
quark must be opposite in sign to that of any anti-quark;
by convention, we take Pu = Pd = Ps = Pc = Pt = Pb =
+1 and Pu = Pd = Ps = Pc = Pt = Pb = 1. Hence
P = Pa Pb =-1, and therefore the parity of all the light
mesons is -1.
Having looked at the parity of the light mesons, we now
enquire into their spins. Each of the quark constituents of
a meson have spin-1/2; these spins may be anti-parallel,
in which case the meson has spin S = 0, or the spins can
be parallel, in which case the meson has spin S = 1.
Using the notation introduced in section 5.A, therefore,
the spin-parity of the lightest mesons (of which there will
be 9 2 = 18) can be either J P = 0 or J P = 1 . These
18 particles are shown in Figure 23. Here, we see (a) the
nine J P = 0 mesons and (b) the nine J P = 1 mesons.
Each of these super-multiplets plots a given light meson
as a function of I3 and Y (see equations (124) and (125)).
In our super-multiplet diagram, dierent members of
a given isospin multiplet (see section 5.B) lie along horizontal lines.
Note also that the quark assignments of the 1 meson
nonet are the same as those for corresponding positions
on on the 0 nonet. This is because all members of the 1
nonet are resonances (see section 5.D), and may therefore
be viewed as excited states of the corresponding members
of the 0 nonet (also, c/f Figure 21).

29
two possibilities: (a) The spin of the baryon, which is
made up of three spin- 12 , quarks, will be equal to 32 if
all three quark spins are in the same direction, like so:
; (b) The spin of the baryon will be equal to 12 if
two quark spins are in one direction and the remaining
quark spin is in the other direction, like so: . These
two possibilities are consistent with the fact that there
are two super-multiplets of light baryons (See Figure 24),
with one super-multiplet having spin 12 and the remaining
super-multiplet having spin 32 .

FIG. 23: Super-multiplets for the light mesons: (a) the 0


meson nonet and (b) the 1 meson nonet. The quark structure of each particle is indicated in round brackets; the quark
assignments for particles in plot (b) are the same as for (a),
and are therefore not shown.

Lastly, you will have noticed the complicated quark


structure which was assigned to the three mesons with
I3 = Y = S = 0, which lie at the origin of each graph in
Figure 23. One might have thought that three simple
mesons, namely uu, dd and ss, would have belonged here,
rather than the complicated linear combinations of these
possibilities that we have presented. Unfortunately, we
do not have time to go suciently into the isospin formalism so as to properly derive these results. If interested in pursuing this point further, see pp. 155-157 of
D.H.Perkins, Introduction to High-Energy Physics, third
edition, Addison-Wesley (1987).
Exercise 36: Verify that each of the quark assignments in Figure 23(a) leads to the claimed values of the
coordinate (I3 , Y ), for each meson in the super-multiplet.

FIG. 24: Super-multiplets for the light baryons: (a) the 12


+
baryon octet and (b) the 32 baryon decuplet. The quark
structure of each particle is indicated.

How about the parity of the lightest baryons? Following similar logic to that used in the previous section,
we conclude that a baryon made of quarks abc will have
parity:
P = Pa Pb Pc (1)L12 +L3 ,

2.

The light baryons

We now look at the two super-multiplets for the light


baryons. We shall again makes use of assumptions #1
and #2, as listed in section 5.F.
Regarding the spin of the lightest baryons, there are

(141)

where L12 is the angular momentum quantum number for


a chosen pair of quarks in their centre-of-mass frame, and
L3 is the angular momentum of the third quark about
the centre of mass of the pair, in the overall centre-ofmomentum frame. Since we take L12 = L3 = 0, the
parity of our lightest baryons is, by (141), equal to:

30

P = Pa Pb Pc = 1,

(142)

where we have made of the fact that Pu = Pd = Ps = 1.


Summarizing the results of the previous two paragraphs, we expect the lighter baryons to fall into two
+
classes, namely (a) those with J P = 12 ; and (b) those
+
with J P = 32 . This is indeed the case, as shown in
Figure 24.
As was the case with the meson super-multiplets, the
positions of the particles in the baryon super-multiplets
are based on the coordinates (I3 , Y ). One can readily
verify that the quark assignments in Figure 24 are consistent with the coordinates (I3 , Y ) which have been assigned to each of the particles. Also, as was the case with
the meson super-multiplets, members of a given isospin
multiplet lie along the same horizontal lines in the baryon
super-multiplet.
Notwithstanding these similarities, there are two
striking dierences between the pair of baryon supermultiplets (Figure 24) and the pair of meson supermultiplets (Figure 23): (a) The pair of meson supermultiplets had the same shape, number of particles, and
quark assignments for each particle. This is not the case
for the pair of baryon super-multiplets. (b) Each meson
super-multiplet had a number of members equal to the
number of possible dierent combinations (i.e., nine) of
a quark and an anti-quark, chosen from the allowed set
u, d, s. Now, bearing in mind that baryons are composed
of three quarks, there are ten dierent allowed combinations of three quarks chosen from the set u, d, s, namely:
sss, ssd, ssu, sdd, sud, suu, ddd, ddu, duu, uuu (note
that it does not matter which order the quarks are listed
in, for the same reason that a bag filled with two red balls
and a green ball is the same as a bag filled with one green
ball and two red balls). Since there are ten allowed combinations, one might expect there to be ten particles in
each of the super-multiplets of Figure 24; instead, there
are only eight particles in Figure 24(a), while there are
the full ten in Figure 24(b).
These problems cannot be explained using logical deductions based on assumptions 1 and 2 (as listed in
section 5.F). Rather, we must introduce an additional
assumption. This additional requirement is ultimately
based on the Pauli exclusion principle, although this fact
will not become clear until the end of section 5.G.
In view of the somewhat twisted chain of logic which is
to follow, let me summarize the remainder of the chapter.
(a) We will begin with a brief explanation of the salient
points concerning the Pauli exclusion principle; (b) We
then explicitly violate this principle in a well-defined way;
(c) Next, we show how our explicit violation of the Pauli
exclusion principle leads to a correct prediction of the
structure of the baryon super-multiplets; (d) Lastly, we
introduce the concept of colour to show that our violation
of the Pauli principle is only an apparent one, and that
in fact our logic is consistent with the Pauli principle.

In the context of the quantum mechanics of multipleparticle systems with spin[15], note that the wavefunction of such a system will consist of the product of
two parts: a spatial wave-function (r1 , r2 , r3 , ) which
depends on the spatial coordinates r1 , r2 , of particles 1, 2, , and a spin wave-function which lists the
spin orientations of each particle. As a concrete example, suppose that we have a system of two spin-1/2
particles; a possible wave-function for such a system is
(r1 , r2 )| >, where |(r1 , r2 )|2 dr1 dr2 gives the probability for finding particle 1 in a volume dr1 of space
centered on the point r1 in space while simultaneously
finding particle 2 in a volume dr2 of space centered on
the point r2 , and | > means that both particles have
spin up. For our purposes, the important point in all of
this is that the required wave-function may be expressed
as a product of a spatial wave-function and a spin wavefunction.
We can now state the Pauli exclusion principle in a
form which is useful for us: multi-particle wave-functions
which include identical particles with half-integer spin,
must be anti-symmetric upon interchange of two identical spin-half particles. Stated dierently, when two identical spin-1/2 particles are interchanged, the new wavefunction must be the negative of the old wave-function.
Thus, if we have a wave-function (r1 , r2 )(1, 2) describing a system of two identical spin-1/2 particles, the wavefunction must be anti-symmetric on interchanging the
two particles[16][17].
Now that we have introduced the Pauli principle, we
are going to maximally violate it by assuming that only
overall symmetric space-spin wave-functions are allowed
for quark systems. This apparently bizarre piece of logic,
which will be resolved in the next section on colour, is
able to lead directly to the observed form of the baryon
decuplets.
Now, the lowest-lying states of a given quantum system
have spatial wave-functions (r1 , r2 , ) that are symmetric under the interchange of like particles. We shall
assume this to be the case for the low-energy quark systems (light baryons) that we are considering. Now, since
we want the overall space-spin wave-function to be symmetric under the interchange of like particles, this implies
that the spin wave-function must be symmetric under the
interchange of like particles.
Bearing all this in mind, we can finally attack the problem of the observed form of the baryon super-muliplets of
Figure 24. We will separately consider the three dierent
classes of baryons which are allowed by assumption #1:
Class 1 contains two like quarks and one unlike quark
(e.g. uud, dds etc.); Class 2 contains three like quarks
(e.g. uuu etc.); Class 3 contains three unlike quarks (i.e.,
uds).
Class 1 (Two like quarks) There are six allowed combinations of two like quarks and one unlike quark:
uud, uus, ddu, dds, ssu, ssd. We illustrate the case uud.
Since we are maximally violating the Pauli principle,
the two like quarks must have the same spin orientation:

31
u u . Adding the third, unlike quark, then gives u u d
(total spin = 12 + 12 + 12 = 32 ) or u u d (total spin =
1
1
1
1
2 + 2 2 = 2 ). Both spin-1/2 and spin-3/2 baryons
are allowed by this mechanism, which accounts for the
six particles which form a ring about the origins of
Figures 24(a) and 24(b).
Class 2 (Three like quarks) There are only three
allowed combinations of three like quarks, namely
uuu, ddd, sss. Since the spin wave-function must be symmetric with respect to the interchange of like spin-1/2
particles, this implies that the allowed spin states of
three like quarks are those for which all quarks have the
same spin orientation. Thus the only allowed values are
u u u , d d d and s s s , all of which have a total spin
of 12 + 12 + 12 = 32 . No spin-1/2 particles can be constructed
for Class 2. This accounts for the three particles on the
tips of the triangle in Figure 24(b), and also explains why
the same particles are missing from Figure 24(a).
Class 3 (No like quarks) There is one allowed combination of three unlike quarks, namely uds. The possible
spin states are u d s (total spin = 32 ), u d s (total spin
= 12 ) and u d s (total spin = 12 )[18]. Thus, bearing in
mind that the combination uds puts us at the origin of
a plot of Y versus I3 , we have two spin-1/2 particles at
the origin of Figure 24(a), and one spin-1/2 particle at
the origin of Figure 24(b).

G.

Colour

Our explanation of the properties of Figure 24


was contingent upon an apparent maximal violation of
the Pauli principle, since we assumed that the spacespin wave function was symmetric (rather than antisymmetric) under the interchange of like spin-half particles. We are now ready to justify this strange piece of
logic, by introducing the concept of colour.
The problem was resolved by Greenberg in 1964, who
suggested that, in addition to the space and spin components of a wave-function, there exists a so-called colour
wave-function. Thus the total wave-function is given
by the product:
= C ,

(143)

where is the spatial wave-function, is the spin wavefunction, and C is the colour wave-function. This allows
us to regain the Pauli principle: we took both and
to be symmetric under the interchange of like spin-1/2
quarks; if the colour wave-function C is anti-symmetric,
then the total wave-function C will be anti-symmetric
under the interchange of like spin-1/2 particles, in accordance with the Pauli principle.
Now, the spatial wave-function is a function of position because it describes how a given quantum object is
spread out in three-dimensional space. The spin wavefunction describes a dierent attribute of a quantum par-

ticle, namely the orientation of its spin. The colour wavefunction describes the state of a new attribute, namely
the colour. A given quark can exist in three colour states,
termed red,green and blue; a given anti-quark
can exist in three colour states, termed anti-red,antigreen and anti-blue. The attribute of colour exists in
addition to the other attributes (spatial extent and spin
orientation). Colour (also called colour charge) is to the
strong interaction what electric charge is to the electromagnetic interaction: just as like electric charges have an
electrostatic repulsion and unlike electric charges have an
electrostatic attraction, so too do like colour charges experience a strong-force repulsion (chromostatic repulsion), and unlike colour charges experience a strong-force
attraction. We shall have more to say about this in the
next chapter.
We close this chapter by noting that, for baryons, the
required totally-antisymmetric colour wave-function C
is:
1
C = (r1 g2 b3 g1 r2 b3 + b1 r2 g3
6
b1 g2 r3 + g1 b2 r3 r1 b2 g3 ),

(144)

where r, b, g stand for red/green/blue, r1 means that the


first quark is red, etc.
Exercise 37: Show that the colour wave-function
in equation (144) is indeed anti-symmetric upon interchange of any two quarks.
Exercise 38: We assumed that the space-spin wavefunction of a baryon was symmetric under the interchange of any pair of like quarks. Under assumptions
#1 and # 2 of section 5.F, what super-multiplet structure for the light baryons would have been obtained if
we took the space-spin wave-function of a baryon to be
anti-symmetric under the interchange of any pair of like
quarks?
Exercise 39: The lightest charmed baryons have the
quark composition cab with zero orbital angular momentum, where c is the charmed quark and a, b can be any
of the light quarks u, d, s. Show that the resulting states
+
can be classified into three super-multiplets: (a) J P = 12
baryons in which the light quark pair ab has spin-zero;
+
(b) J P = 12 baryons in which the light quark pair ab
+
has spin-one; (c) J P = 32 baryons in which the light
quark pair ab has spin-one. Please plot diagrams of each
super-multiplet.
VI.

QUANTUM CHROMODYNAMICS

Classical electrodynamics deals with the electromagnetic interactions of electric charges in motion; here, electromagnetic forces are considered to be due to the action
of electric and magnetic fields on charged matter. Quantum electrodynamics (QED) is a quantum version of this
theory, which works with a quantized electromagnetic

32
field and considers the electromagnetic force as due to
the exchange of virtual photons (see section 1.D).
We can build on this analogy to introduce quantum
chromodynamics (QCD). This is the quantum theory
of the strong interactions between colour charges, which
treats the strong force as due to the exchange of gluons (see section 1.A). Particles with colour charge (e.g.
any particle made from quarks) feel the strong force, just
as particles with electric charge feel the electromagnetic
force. Similarly, particles with no colour charge (e.g. leptons and photons) do not feel the strong force.
In this chapter, we give an introduction to some of the
fundamental ideas of quantum chromodynamics, which
describes strong interactions in the standard model of elementary particle physics. We begin with the concept of
confinement (section 6.A), which demands that all free
particles should have a neutral colour charge. We then
explore the role of gluons as the carriers of the strong
force (section 6.B). We move onto the concepts of antiscreening and asymptotic freedom (section 6.C), which
leads to the remarkable conclusion that the strong force
gets weaker at distances smaller than a hadron radius.
In the same section, we argue that the electromagnetic
interaction gets stronger at smaller distances. We close
with an introduction to the concept of hadron jets (section 6.D), which are narrow streams of hadrons that are
produced in certain strong interactions.

A.

Colour charges and confinement

We introduced the concept of colour in section 5.G, in


the context of a desire to correctly predict the structure
of the two super-multiplets of light baryons. This entry
of the colour concept (in 1964) was rather arbitrary, but
has since gained greater legitimacy on account of the fact
that the colour theory can correctly account for many
observed phenomena in elementary particle physics.
A given quark can exist in three colour states, termed
red,green and blue; a given anti-quark can exist in
three colour states, termed anti-red,anti-green and
anti-blue. As is the case for electrostatics, like colour
charges repel (via the strong force) while unlike colour
charges attract. Thus a red quark will be repelled from
a red quark, a red quark will be attracted to an anti-red
anti-quark, and a red quark will be attracted to green and
blue quarks. The strength of the force between colour
and anti-colour is stronger than the force between two
colours.
The colour theory demands that all particles have a
total colour charge of zero; this is known as colour confinement. For mesons, which are composed of a quarkantiquark pair, this colour neutrality is achieved by having the quark pair in colour states of red-antired, blueantiblue, or green-antigreen. For baryons, which are composed of three quarks, colour neutrality is achieved by
having one red quark, one blue quark and a green quark.
One consequence of the colour confinement hypothe-

FIG. 25: (a) A red quark is attracted to an anti-red quark,


forming a colour-neutral meson. (b) A red quark, green quark
and blue quark are all attracted to one another, forming a
colour-neutral baryon.

sis is that free quarks (which, being coloured, cannot be


colour neutral) cannot exist in nature. Despite many
searches for free quarks, based on looking for particles
with fractional electrical charge, no free quarks have been
observed. Note that, if a free light quark existed, it would
necessarily be stable: this is so because it has nothing to
decay to!
Another consequence of colour confinement is that the
strong force between nucleons is a pale shadow of the
strong force between quarks. Again, an analogy from
electrodynamics is useful. As shown in Figure 26(a), the
van der Waals force FV DW is a force between electrically neutral particles. This van der Waals force, which
varies as the inverse seventh power of the separation between neutral atoms or molecules, is much weaker than
the electrostatic force from which it is derived. Now, in
Figure 26(b) we see that quark-antiquark pairs in a given
meson are attracted to one another by the strong force.
The whole meson is colour neutral, by the hypothesis of
quark confinement. Two colour-neutral mesons may be
attracted to one another via the strong-force equivalent
of the van der Waals interaction. It is this pale shadow
of the strong force that binds neutrons and protons in the
nucleus of an atom.

FIG. 26: (a) The van der Waals force FV DW is a weak force
between electrically neutral particles. (b) The strong force
between colour-neutral hadrons is the strong-force equivalent
of the van der Waals interaction.

33
B.

Gluons as mediators of the strong force

Electric charge determines how particles are pushed


around by the carriers of the electromagnetic force,
namely photons; there are two types of electric charge,
namely positive and negative; like electric charges repel and unlike electric charges attract. Similarly, colour
charge determines how particles are pushed around by
the carriers of the strong force, namely gluons; there are
six types of colour charge, namely red, green, blue, antired, anti-green and anti-blue; like colour charges repel
and unlike colour charges attract. In this section we examine in more detail the notion that gluons are mediators/carriers of the strong force.
Whereas there is a single photon responsible for the
electromagnetic force, there are eight gluons responsible
for the strong force. There are two neutral gluons,
denoted G0 and G0 , and six colour-changing gluons,
GRB , GRG , GBR , GBG , GGR and GGB . Emission of a
colour-neutral gluon leaves unchanged the colour of the
quark that emitted the said gluon; however, emission of
a colour-changing gluon changes the colour of the quark
which emitted the gluon. Thus we have a picture in which
the quarks inside a given particle are continuously changing colour.
An example of quark-quark scattering is shown in Figure 27(a). Here, a red up quark emits a red-antiblue
gluon, thus being changed into a blue quark. Note that
colour charge is preserved at this vertex: red = blue +
red + anti-blue. The blue strange quark absorbs the
gluon, changing its colour to red. The resulting force is
an attractive one, because quarks of opposite colour are
attracted to one another.
Now, since the six colour-changing gluons are themselves coloured, gluons can scatter from other gluons,
as indicated in Figure 27(b). Again, note that colour
charge is conserved at each vertex of this figure. Such
gluon-gluon interactions have no analogue in quantum
electrodynamics, as the photon has zero electric charge.
Figure 28 shows a more complicated Feynman diagram
representing the interactions which bind two up and one
down quarks in the proton.
We close this section with a brief discussion of the socalled glueballs. QCD allows the formation of colourneutral glueball states composed entirely of gluons (see
Figure 29). Despite many experimental searches, such
states have never been observed.
C.

Screening, anti-screening, asymptotic freedom

The vacuum of classical physics is flat and featureless.


The quantum vacuum is pictured as being filled with virtual electron-positron pairs which exist for a suciently
small time for the uncertainty principle to be respected.
Four such virtual electron-positron pairs are shown in
Figure 30(a). As we shall now show, the existence of the
dynamic quantum vacuum leads to so-called vacuum po-

FIG. 27: (a) Feynman diagram for scattering of a red up quark


from a blue strange quark, via the exchange of a red-antiblue
colour-changing gluon GRB . (b) Since gluons themselves have
colour charge, they can scatter o one another, an example
of which is given here.

FIG. 28: Feynman diagram showing possible interactions of


the three quarks inside a proton.

larization eects which imply that the electric charge of


a given particle gets stronger the closer one approaches
the particle.
Notice that, as a consequence of electrostatic repulsion, Figure 30(a) has the virtual positrons closer to the
real electron at the centre of the figure, while the virtual
electrons are further away. This has the eect of immersing the bare electron in a cloud of virtual positrons
and electrons, with the sheath of virtual electrons being
further away from the real electron than the sheath of
virtual positrons (i.e. the vacuum is polarized). This
leads to a shielding eect (screening) when one is sufficiently far away from the electron, with its eective
electric charge being less that that of the bare particle. When one is very close to the electron, the sheath
of virtual particles becomes irrelevant, and one sees the

34

FIG. 29: Glueballs are colour-neutral states consisting of gluons. Such states have never been experimentally observed.

full or bare charge of the particle. This leads us to


the remarkable conclusion that the electric charge gets
larger at very small distances.
In Figure 30(a), we saw how electron-positron pairs
serve to screen the charge of an electron, making it appear weaker at larger distances. This explanation was
in the context of a single electron. In Figure 30(b), we
consider screening in the context of a pair of interacting
electrons. Figure 2(a) shows two electrons interacting via
the exchange of a virtual photon. Figure 30(b) shows an
electron-positron pair being created and then destroyed
by this virtual photon. The virtual electron-positron
pair shields the charges of the two scattering electrons
from one another; such vacuum-polarization Feynman diagrams serve to weaken the electromagnetic interaction.
Vacuum polarization eects also exist for quarks. The
quantum vacuum contains virtual gluon pairs and virtual
quark-antiquark pairs, as shown in Figure 30(c). (Virtual electron-positron pairs are also present, but since
they have no color charge their presence leads to no
strong interactions.) There are two eects at work here.
(i) The creation of the quark-antiquark pairs serves to
weaken the strong interaction at suciently-short distances (screening), using an analogous argument to that
given in the context of Figure 30(a). (ii) However,
the creation of the gluon pairs serves to strengthen the
strong interaction at suciently-short distances (antiscreening), for reasons that we do not go into. Detailed
calculations show that the anti-screening eect dominates over the screening eect, with the end result that
the strong interaction gets weaker at suciently-short
distances, and stronger at larger distances. The strong
force becomes weakest at the very smallest distances, a
phenomenon known as asymptotic freedom. The strongforce analogues of Figure 30(b) are shown in Figures
30(d) and 30(e) respectively.

FIG. 30: (a) Polarization of the vacuum by virtual electronpositron pairs in the vicinity of an electron. (b) An eect
of vacuum polarization on electron-electron scattering. (c)
Polarization of the vacuum by virtual quark-antiquark and
gluon-antigluon pairs in the vicinity of a quark. (d,e) Eects
of vacuum polarization on quark-quark scattering.

D.

Jets

Figure 31(a) shows the electric lines of force between


an electron and a positron. These lines of force fill all of
the space surrounding the electron-positron pair. Figure
31(b) shows the strong lines of force between the quark
and anti-quark inside a meson. These lines of force do not
fill space, because of the fact that these lines of force are
attracted to one another by gluon-gluon scattering (c/f
Figure 27(b)). Since these lines of force are attracted
to one another, an attempt to pull apart the quarkantiquark pair results in a stretching of lines of force
into a tight tube or string (Figure 31(c)). Eventually,
the tension in the gluon string is suciently strong for it
to be energetically favorable to create a quark-antiquark
pair at the centre of the string, with the net result that
we now have two mesons instead of one. This process is
known as fragmentation, and we shall now show that this
mechanism is closely related to the phenomenon of jets.
Jets, as the name implies, are directed streams of
hadrons that are produced in some high-energy collisions.
An example of a two-jet event was given in Figure 16.
The mechanism by which such jets are formed is shown
in Figure 32. Here, an electron and a positron annihilate
to form a high-energy photon. This photon leads to the
formation of a quark-antiquark pair which fly away from
one another. Since the string force is rather weak at distances smaller than 1 fermi or so, they quarks do not interact very strongly until their distance apart is around

35
VII.

WEAK INTERACTIONS
A.

Introduction

We now make a detailed study of the weak interaction,


this being the third force on our list of the four known
forces of nature (see section 1.A).
As mentioned in section 1.D.1, the existence of the
weak interaction is considered to be due to the exchange
of three dierent spin-1 bosons, namely W + , W and Z 0 .
We saw that these particles are very massive (with masses
almost 100 times that of the proton) and therefore have
a very short range (around 1/500 of a proton radius).
Terminology:
(a) Interactions involving the exchange of W bosons are known as charged current
interactions[19], while those involving exchange of Z 0
bosons are termed neutral current interactions. (b) In
addition to the neutral current versus charged current
distinction mentioned above, one may also classify weak
interactions as leptonic (i.e., involving only leptons),
semi-leptonic (i.e., involving both leptons and hadrons)
or hadronic (i.e., involving only hadrons). Examples:

FIG. 31: (a) Lines of electrostatic force between an electron


and a positron; (b) Lines of chromostatic force between a
quark and an anti-quark inside a meson; (c) Stretching of the
chomostatic lines of force as the quark and anti-quark are
separated; (d) Snapping of the lines of force into a quarkantiquark pair, yielding two mesons in the place of one (this
is the process of fragmentation).

1 fermi. By the process of fragmentation, as shown in


Figure 31(c) and (d), the initial quark-antiquark pair are
converted into two jets of hadrons.

FIG. 32: Mechanism for the formation of two jets in electronpositron annihilation.

e + e + (leptonic);
n p + e + e (semi-leptonic)
+ p (hadronic).
B.

(145)
(146)
(147)

On the importance of the weak interaction

The gravitational force is so weak that we ignore it


for the vast majority of discussions in elementary particle
physics, which typically deal with length scales and times
much smaller than those encountered in everyday life.
However, this does not imply that the gravitational force
is unimportant; indeed, this force plays an extremely important role over cosmological distances and times[20].
Similarly, the weakness of the weak interaction does
not imply that the consequences of this force are insignificant. Below, we give a very incomplete list demonstrating the importance of the weak force. Each member of
this list is extremely interesting and worth exploring in
much more depth than the superficial treatment below;
unfortunately, we dont have enough time to do so!
(a) All hadrons and leptons participate in the weak
interaction. When a given particle can also participate in strong and electro-magnetic interactions, these
stronger forces dominate. However, when both strong
and electromagnetic interactions are forbidden by the relevant conservation laws, the weak interaction becomes
dominant[21].
(b) Weak interactions control the thermonuclear reaction rate in main sequence[22] stars. For example, the
first (and therefore rate-limiting) step in hydrogen burning within the solar core is currently believed to be a
weak interaction, namely: p + p d + e+ + e . It is

36
precisely because the weak interaction is so weak that
a given proton in the extremely dense[23] solar core has
a mean wait of 1010 years before undergoing this ratelimiting weak interaction; if the weak interaction were
stronger, then the lifetime of the solar system would be
much shorter.
(c) Weak interactions are also important in stellar evolution beyond the main sequence. For example, once a
star has burned all of its hydrogen by fusing it into
helium, it can begin to burn helium and progressively
heavier elements. By the time it gets to burning carbon, the star is much hotter and denser than it was at
the hydrogen-burning stage. The density is too great for
photons to eciently transport heat from the core of the
star to the outer layers; neutrinos then play an important
role in heat transfer from the core to the outer layers of
the star. The weak interaction is therefore responsible
for shortening the lifetime of these older stars.
(d) There are believed to be about 109 neutrinos for
every proton. Thus, even though they have a very small
rest mass, neutrinos may still carry a significant amount
of the energy in the universe. In particular, their gravitational eects may be important in determining the largescale structure of the universe.
(e) The fact that neutrinos are so weakly interacting makes them excellent probes of the interior of the
sun, and of supernovae (both of which produce an abundance of neutrinos). Other forms of radiation, being more
strongly interacting, are multiply scattered in passing
from the core to the surface of such bodies.

C.

and momentum (c/f the discussion accompanying Figure


2).
Bearing the above points in mind, we now consider
the W -lepton interactions, for which there are 16 3 =
48 basic processes (16 for each of the three lepton
generations[26]: l e, , ). For a given lepton generation l, the 16 basic W lepton processes can all be
obtained by deforming the two basic vertices shown in
Figure 33. Eight basic processes can be obtained from
each of these basic vertices, yielding 16 for each lepton
generation l. For example, the eight possible deformations of Figure 33(a) yield the eight basic processes listed
in Figure 34. The eight possible deformations of Figure
33(b) can be obtained by replacing all particles in Figure 34 with their corresponding anti-particles (because
this is the manner in which Figures 33(a) and 33(b) are
related).

FIG. 33: The two basic vertices for the W -lepton interactions.

Charged-current weak interactions

We mentioned, in section 1.A, that the Standard Model


treats the physics of elementary particles in terms of
three basic sets of particles: leptons, quarks and forcecarrying bosons (i.e., gauge bosons). Now, since both
leptons and hadrons feel the weak force, we expect the
W + , W , Z 0 gauge bosons to interact with both leptons and quarks. Accordingly, the following two subsections discuss the weak charged-current interactions[24]
of leptons and quarks respectively. Consideration of the
neutral-current reactions[25] will be postponed until later
in the chapter.

1.

Charged currents #1: the W -lepton interactions

The basic processes


Before launching into a treatment of the chargedcurrent interactions of leptons, we briefly revise the basic
electro-magnetic processes sketched in Figure 1. In the
present context, there are two salient points: (a) Each of
the basic electro-magnetic processes in Figure 1 can be
arrived at by suitably deforming any one of the graphs
appearing in that figure; (b) As seen in exercise 2, none of
the basic electro-magnetic process conserve both energy

FIG. 34: The eight basic processes for W -lepton interactions, corresponding to the various deformations of Figure
33(a).

Exercise 40: (a) Show that both charge and lepton


numbers are conserved for the basic vertices appearing
in Figure 33. (b) Show that these are the only possible

37
W -lepton vertices which satisfy both charge and lepton
number conservation.
Now, the basic processes associated with the electromagnetic interaction (see Figure 1) were all seen to be
virtual, as none of them are able to satisfy both momentum and energy conservation. The situation is dierent
for the basic processes in Figure 34; only six out of these
eight processes are unable to satisfy both energy and momentum conservation (see exercise below). The six virtual processes may be combined in such a way that energy
conservation is conserved; an example of this is given in
Figure 35 (c/f Figure 2 and the accompanying text).

FIG. 35: Two dierent Feynman diagrams related by dierent


time ordering. By convention, only one of the possible time
orderings is drawn, with the other(s) left implied.

Exercise 41: (a) Show that the vertices of Figures


34(e) and 34(f) are consistent with both energy and momentum conservation[27]. Hint: Work in the rest frame
of the W boson. (b) Show that the remaining vertices of
Figure 34 are inconsistent with both energy and momentum conservation, and therefore represent virtual processes. (C/f exercise 2.)
Intrinsic strength of the weak interaction
We move on to a consideration of the intrinsic strength
of the weak interaction. We stated without proof, in the
last paragraph of 1.3.2, that the probability of occurrence for a basic electro-magnetic process is O(), where
1/137 is the fine structure constant. If, in addition, we recall the hypothesis of lepton universality[28],
we conclude that each of the vertices derived from Figure
33 is characterized by a single dimensionless parameter
W , which is to weak interactions what the fine-structure
constant is to electro-magnetic interactions. Remarkably, it can be shown (see the next exercise) that W ,
which suggests that the intrinsic strength of the weak interaction is about one-half that of the electro-magnetic
interaction (one might have expected that W ).
This subtle and unexpected result indicates that the apparently vast dierence in strength between the electromagnetic and weak interactions is to some extent illusory;
the fact that W indicates the the electro-magnetic
and weak forces have similar intrinsic strength; the latter
appears much weaker to us than the former because of
the massive nature of the gauge bosons associated with
the weak force.
Exercise 42: The purpose of this exercise is to estimate W , using similar dimensional arguments to
those used in obtaining equations (54)-(58). (a) Consider the following pair of weak W-boson decays: W +

e+ + e and W e + e . In natural units, the decay


rate associated with these reactions has the dimension
of energy, and is measured to be 0.2 GeV. Neglecting the final state lepton masses, use a simple dimensional
argument to obtain: W MW . (b) Hence show that
W 0.3. Note that a more detailed calculation gives
W 0.6.
Time ordering
We now move onto the notion of time ordering in the
context of weak interactions[29]. Consider Figure 35,
which shows two dierent time orderings for a given Feynman diagram. By convention, only one of the possible
time orderings is drawn, with the other(s) left implied.

FIG. 36: Origin of an approximately zero-range interaction


in inverse muon decay.

2.

Charged currents #2: the W -quark interactions

The previous subsection dealt with the charged-current


weak interactions of leptons. We now consider the
charged-current weak interactions of strongly interacting particles (hadrons). This will be done by considering
the emission and absorption of W bosons by the quarks
inside the hadrons, as shown in the examples of Figure
37. Our discussion of the W -quark interactions will be
based around two central ideas: (a) lepton-quark symmetry and (b) quark mixing.
(a) Lepton-quark symmetry
We confine our discussion of weak quark interactions
to the first two quark generations (see equation (60)):

u
d


c
,
,
s

(148)

from which the vast majority of known hadrons are constructed. For a reason that will soon become apparent,
we also write down the first two lepton generations (see
equation (40)):

e
e

,
.

(149)

We are now ready to consider the concept of leptonquark symmetry, which states that the two generations
(148) of quarks have identical weak interactions to the
two generations (149) of leptons. Therefore, one can

38

FIG. 38: The W -quark vertices obtained from the hypothesis of lepton-quark symmetry, ignoring the eects of quark
mixing.

FIG. 37: (a) Dominant quark diagram for reaction (146).


(b,c) Dominant quark diagrams for reaction (147).

convert the W lepton vertices of the previous section, to the desired W quark vertices of the present
section, via the following rules:
e u, e d, c, s.

(150)

Before continuing, let me point out my opinion that


this is a remarkable connection which surely has some
deep and fundamental origin, of which I am unaware.
Unfortunately, I can only blandly state that the remarkable assumption of lepton-quark symmetry is justified because it is consistent with the results of experiment[30].
Please let me know if you are aware of a deeper significance to this assumption!
Making use of the hypothesis of lepton-quark symmetry, the basic W -lepton vertices of Figure 33 yield the
basic W -quark vertices shown in Figure 38. Significantly, we see that the weak interaction does not conserve
flavour; this is consistent with the observation made

in the third paragraph of section 5.A: While S, C, B


are conserved in strong and electromagnetic interactions,
they are not necessarily conserved in weak interactions.
The results of the hypothesis of lepton-quark symmetry work well for interactions such as pion decay, as illustrated by the quark diagram of Figure 39(a). Note that
the first vertex of this Figure is obtained by suitably deforming the basic vertex of Figure 38(b).
Unfortunately, our scheme for lepton-quark symmetry
is unable to explain observed reactions such as the kaon
decay shown in Figure 39(b), as it does not allow the
coupling of the strange to the up quark. To explain such

FIG. 39: Feynman diagrams for two dierent semi-leptonic


decays. The existence of (a) can be explained without needing to invoke the hypothesis of quark mixing. However, the
existence of reaction (b) cannot be explained without the notion of quark mixing.

interactions, we need to invoke the idea of quark mixing, due to Cabibbo[31].


(b) Quark mixing
Cabibbos idea was to rotate or mixthe down and
strange quarks by an angle which is now known as the
Cabibbo angle C :
d = d cos C + s sin C ,
s = d sin C + s cos C .

(151)

One then replaces the quark doublets of (148) with the


new doublets:

u
d


c
,
,
s

(152)

and assumes that lepton-quark symmetry applies to (152)


and (149) (previously, we assumed lepton-quark symmetry applied to (148) and (149)).
As an example of the consequences of Cabibbos
scheme for quark mixing, the basic vertex of Figure 38(a)
has its d quark replaced with a d quark, as shown on the
left of Figure 40. Since the d quark is a superposition
of down and strange quarks, as given in the top row of
(151), the single vertex at the left of Figure 40 is equivalent to the two vertices shown on the right of Figure
40. The previously-forbidden usW coupling (see Figure

39
39(b)) is now allowed (Cabibbo allowed) and has intrinsic strength gus = gW sin C , while the previouslyallowed udW coupling (see Figure 29(a)) is still allowed
(but Cabibbo suppressed) and has intrinsic strength
gud = gW cos C .

Exercise 46: Please prove the statement made in the


previous sentence.
Restricting ourselves to the first two lepton generations, we denote the basic Z 0 -lepton vertices by:
e e Z 0 , Z 0 , e e Z 0 , Z 0 .

(153)

FIG. 40: Interpretation of the ud W vertex as a sum of udW


and usW vertices.

The hypothesis of quark mixing involves a single free


parameter, namely the Cabibbo angle C . This angle
has been determined, from experiment, to have the value
12.7 0.1 degrees.
Exercise 43: Which of the following six decays are
allowed in lowest-order[32] weak interactions?
(a) K + + + + + e + e
(b) K + + + e + e
(c) 0 + e+ + e
(d) 0 + e + e
(e) 0 p + + 0
(f) + + +
Exercise 44: Classify the following semi-leptonic
decays of the D+ (1869) = cd mesons as Cabibbosuppressed, Cabibbo-allowed, or forbidden in lowestorder weak interactions.
(a) D+ K + + + e+ + e
(b) D+ K + + + e+ + e
(c) D+ + + + + e + e
(d) D+ + + + e+ + e
Exercise 45: If the top quark were stable, the lowlying states of the meson tt (toponium) could be approximated by non-relativistic motion in a Coulomb potential V (r) = 4s /(3r), where s 0.1 and r denotes
radial distance. Use the simple Bohr model to calculate
the radius of the ground state and the time taken to complete a single Bohr orbit in the ground state. Compare
this with the expected lifetime of the top quark.

FIG. 41: Basic vertices for the Z 0 -lepton interactions.

To obtain the basic Z 0 -quark vertices from the basic


Z -lepton vertices, we once again make use of the hypotheses of lepton-quark symmetry and quark mixing,
which (as we have already seen) amounts to the following replacement scheme:
0

e u, e d , c, s ,
where d and s are as defined in equation (151).

FIG. 42: Basic vertices for the Z 0 -quark interactions.

Thus the basic Z 0 -quark vertices may be denoted by:


uuZ 0 , ccZ 0 , d d Z 0 , s s Z 0 .

Neutral-current weak interactions

The previous subsection (7.C) discussed the charged


current weak interactions, which involve the exchange
of W particles. We now move onto a consideration of
the neutral current weak interactions, which involve the
exchange of Z 0 particles.
The basic Z 0 -lepton vertices are shown in Figure 41,
where l denotes e, , . These basic vertices are the
only possible vertices that can be written down, which
conserve both electric charge and the lepton numbers
Le , L , L .

(155)

Now, we leave it as an exercise to show that:


d d Z 0 + s s Z 0 = ddZ 0 + ssZ 0 ,

D.

(154)

(156)

which implies that the basic Z 0 -quark vertices of equation (155) may be replaced by:
uuZ 0 , ccZ 0 , ddZ 0 , ssZ 0 .

(157)

These vertices evidently conserve quark numbers such


as strangeness and charm, unlike the charged current
weak interactions which do not conserve these quark
numbers (see paragraph of text immediately before Figure 38).
Exercise 47: Please derive equation (156).

40
E.

Electroweak unification: a very brief note

Prior to the work of Faraday and Maxwell, the electric and magnetic interactions were considered to be two
separate interactions. As is well known, these were conceptually unified into a single electro-magnetic set of interactions, as summarized by the Maxwell equations of
classical electrodynamics[33].
Now, to date we have considered the electro-magnetic
and the weak forces as separate interactions. Glashow,
Weinberg and Salam, in the early 1960s, unified the
electro-magnetic and weak forces into a single conceptual scheme which is now known as the electro-weak
interaction[34]. The W , Z 0 , are considered to be the
carriers of the electroweak force. Here, we give a very superficial sketch of the ideas underlying the electro-weak
theory; unfortunately, we do not have time for a more
detailed treatment!
The electroweak theory states that, whenever there exists an process involving exchange of a photon, then there
also exists a process involving the exchange of a Z 0 . An
example is given in Figure 43, for the case of an electron
scattering from a proton.

Now, the total cross section for this reaction, as a function of the centre-of-mass energy ECM , is sketched in
Figure 45. The two main features of this graph are: (i)
the fact that the cross section is proportional to the inverse square of the energy E, for suciently low E; (ii) the
sharp peak at 91.2 GeV/c2 . In the next exercise I have
asked you to show how these features arise from both
electro-magnetic and weak processes, without needing to
invoke detailed calculations using the electroweak theory.

FIG. 45: Cross section for the reaction e +e+ ++ , as


a function of the centre-of-mass energy ECM . (Figure taken
from Martin and Shaw, page 225.)

Exercise 48: (a) Using a rough dimensional argument


which neglects lepton masses, show that the contribution
of Figure 44(a) to the cross section of reaction (158),
is approximately given by:


FIG. 43: Contributions to electron-proton scattering from (a)
photon exchange; (b) Z 0 exchange. Rather than separately
considering these as electromagnetic and weak interactions
respectively, they are instead seen as two possible electroweak
processes. (Figure taken from Martin and Shaw, page 225.)

Another example, which we will look at in some detail,


is that of the reaction:
e + e+ + + .

(158)

The two main Feynman diagrams contributing to this


reaction are given in Figure 44.

2
.
E2

(159)

(b) Using a similar method of argument, show that


the contribution Z of Figure 44(b) to the cross section
of reaction (158), is approximately given by:
Z G2Z E 2 ,

(160)

where GZ is a constant of natural-units dimensions E 2


which characterises the strength of the weak interaction
at suciently-low energies (c/f equation (23)). (c) Show
that:

GZ

2
gZ
.
MZ2

(161)

Hint: re-read 1.D.3, and remember that we are working


in natural units. (d) Using your results from the previous
calculations, show that:

FIG. 44: The two main electroweak contributions to the reaction e + e+ + + . (Figure taken from Martin and
Shaw, page 219.)

Z
G2 E 4
E4
Z2 4 .

MZ

(162)

(e) Hence estimate the energy above which the contribution of Figure 44(b), to the cross section of reaction

41
(158), ceases to be negligible when compared to the contribution of Figure 44(a). (f) What is the origin of the
peak in Figure 45? In your answer, incorporate a precise

reformulation of the following correct but vague sentence:


The weak interaction can sometimes be stronger than
the electro-magnetic interaction.

[1] A hadron is defined to be a particle which feels the strong


interaction. Hadrons which are fermions (half-integral
spin particles) are called baryons, while hadrons which
are bosons (integral spin particles) are called mesons.
[2] Note that, for the remainder of this course, the term
mass will refer to rest mass.
[3] To get a feel for what these numbers mean, note that the
proton radius is around 0.85 0.02 fm (fermis)
[4] The approximation leading to (19) is known as the first
Born approximation, and corresponds to single-particle
exchange in the picture where forces are due to particle
exchange.
[5] More will be said about cross sections, soon!
[6] Very recently, in mid 2003, a new five-quark
baryon state was reported. This pentaquark is
a major new discovery in particle physics. See
http://www.phy.ohiou.edu/ hicks/thplus.htm for details.
[7] The existence of this particular neutrino can also be inferred from the conservation law of electron-lepton number, which was introduced in the previous section. However, these conservation laws were not known in 1930.
[8] The sole known exception, discovered in 2003, is known
as the penta-quark hadron - see an earlier footnote for
more details about this exotic state.
[9] We need to be clear about what is meant by a long lifetime, in the context of elementary particle physics. The
typical radius r of most hadrons is of the order of 1 fermi;
the associated characteristic timescale r/c is the time
taken for a hadron moving at the speed of light to travel
one hadron radius. Thus the characteristic timescale for
hadrons is r/c 1023 s. Most hadrons are highly unstable, and decay (via the strong interaction) over timescales
of the order of 1023 s. By a long lived particle, we
mean that the lifetime of the said particle is orders of
magnitude longer than 1023 s.
[10] This is the core concept of the present chapter.
[11] See exercise 26 if this point is obscure.
[12] The spin of a particle may be defined as the angular momentum of that particle in its own rest frame. It is therefore occasionally called by the more informative name of
intrinsic angular momentum.
[13] Evidently, this chapter studies particular cases of this
statement. For a general statement of the relation between symmetries and associated conservation laws, see
any suciently-advanced text on quantum mechanics or
classical mechanics which discusses Noethers theorem.
See, for example, Mandl and Shaws Quantum Field Theory or Goldsteins Classical Mechanics.
[14] This W is a function of the energy and momentum of
both the and the + , as indicated in equation (136).
[15] It may be useful to state here that a baryon is such a
system, for it consists of three particles (quarks) each of
which have a spin (of one-half).
[16] An example of such an anti-symmetric wave-function is
(r1 , r2 ) (| > | >), where (r1 , r2 ) has the property that (r1 , r2 ) = (r2 , r1 ). For, on interchang-

ing particles 1 and 2, our wave-function becomes multiplied by a negative sign, which indicates that it is
completely anti-symmetric: (r1 , r2 ) (| > | >)
(r2 , r1 ) (| > | >) = (r1 , r2 ) (| > | >).
This can be achieved in two ways: (i) (r1 , r2 ) can be
anti-symmetric (i.e., (r2 , r1 ) = (r1 , r2 )) and (1, 2)
can be symmetric (i.e., (1, 2) = (2, 1)); (ii) (r1 , r2 )
can be symmetric (i.e., (r2 , r1 ) = (r1 , r2 )) and (1, 2)
can be anti-symmetric (i.e., (1, 2) = (2, 1)).
The remaining possibilities (u d s , u d s , u d s ,
u d s and u d s ) are no dierent from the first three
cases (e.g. u d s and u d s describe the same particle
turned up-side down).
It was thought, for many years, that all weak interactions
were due to charged-current reactions only. This belief
disappeared with the observation of neutral-current weak
interactions at CERN in 1973.
This may be seen as a consequence of the cumulative
eects of the fact that gravitational charge (i.e., mass) is
always non-negative.
(a) For example, we saw in section 2.B.2 that both
the strong and electro-magnetic interactions conserve
B. Therefore, the lightthe quantum numbers Q, S, C, B,
est hadron with a given set of quantum numbers
B cannot decay via the strong or electroQ, S, C, B,
magnetic interactions; it can only decay by the weak interaction. A famous example is neutron decay (68/69).
(b) As another example, we mentioned in sections 4.C
and 4.D that parity and C-parity are conserved in the
strong and electro-magnetic interactions, but are not necessarily conserved by the weak interaction.
A Main Sequence star may be very crudely defined
as one that is burning hydrogen at the center of its
core. For more information, see e.g. B.W.Carroll and
D.A.Ostlie, An Introduction to Modern Astrophysics,
Addison-Wesley, Reading (1996), sections 8.2 (The
Hertzsprung-Russell Diagram) and 13.1 (Evolution on
the Main Sequence).
The density of the solar core is about 100 g/cm3 .
As mentioned in section 7.A, the charged-current weak
interactions involve exchange of W gauge bosons.
As mentioned in section 7.A, the neutral-current weak
interactions involve exchange of Z 0 gauge bosons.
See equation (40).
Indeed, Figure 34(f), together with its corresponding
anti-particle reaction, is the main mechanism via which
the W bosons were detected at CERN in 1983.
See section 2.A.3, which states the assumption of lepton
universality as the electro-magnetic reactions of tauons
and muons are identical to those of electrons, provided
that their mass dierences are taken into account. In the
present context, we widen this to: the electro-magnetic
and weak reactions of tauons and muons are identical to
those of electrons, provided that their mass dierences
are taken into account.
Before continuing, you may find it useful to re-read the

[17]

[18]

[19]

[20]

[21]

[22]

[23]
[24]
[25]
[26]
[27]

[28]

[29]

42
first paragraph of section 1.C.3.
[30] More precisely, the assumptions of lepton-quark symmetry and quark mixing (see part (b)) yield results that are
consistent with experiment.
[31] N. Cabibbo, Phys. Rev. Lett. 10 531 (1963).
[32] By lowest order we mean weak interactions which involve the exchange of a single W or Z particle.
[33] Other instances of conceptual unification in physics include: (a) The unification of celestial and terrestrial mechanics into the single scheme of classical mechanics (the
ancient Greeks considered celestial and terrestrial bodies
to obey dierent laws of mechanics); (b) Einsteins unifi-

cation, via General Relativity, of the theories of gravitation and Special Relativity. (c) The various attempts, as
yet unsuccessful, to achieve a Grand Unification of the
electro-magnetic, weak, strong and gravitational forces
into a single conceptual scheme.
[34] Earlier attempts, from the ashes of which rose the
Glashow/Weinberg/Salam theory, were made by the likes
of Klein and Schwinger. Compare the discussion on the
relative intrinsic strengths of the weak and electromagnetic interaction, given in 7.C.1.

Potrebbero piacerti anche