5.1 A Zoo of Particles
In the beginning of the 1930s, just a few particles were
known: the proton and the electron, the charged constituents of the
atom; the neutron, which was ensuring the stability of the atomic
nuclei; the neutrino (predicted but by then not discovered yet),
whose existence was conjectured to guarantee the conservation of
energy and momentum in beta decays; and the photon, i.e., the
quantum of the electromagnetic field. Then, as we discussed in
Chap. 3, new and often unexpected particles
were discovered in cosmic rays: the positron, the antiparticle of
the electron; the muon, the heavy brother of the electron; the
charged pion, identified as the particle postulated by Yukawa as
the mediator of the strong interaction; the strange “V” particles
and
(called “V” from the topology of
their decays in bubble chambers).


Livingston plot, representing the maximum energies attained by accelerators as a function of the year: original (left) and updated (right). For colliders, energies are translated into the laboratory system.
Original figures from M.S. Livingston and J.P. Blewett, “Particle Accelerators,” MacGraw Hill 1962; A. Chao et al. “2001 Snowmass Accelerator R&D Report,” eConf C010630 (2001) MT100
Accelerators provide the possibility to
explore in a systematic way the energy scale up to a few TeV,
and thanks to this a huge number of new particles have been
discovered. Already in the 1950s, many discoveries of particles
with different masses, spins, charges, properties took place. Their
names almost exhausted the Greek alphabet: these particles were
called ,
,
,
,
,
,
,
,
,
Classifications had to be devised to put an order in such a complex zoology. Particles were first classified according to their masses in classes with names inspired, once again, to Greek words: heavy particles like the proton were called baryons ; light particles like the electron were called leptons ; and particles with intermediate masses were called mesons . The strict meaning of the class names was soon lost, and now we know a lepton, the tau, heavier than the proton. According to the present definition, leptons are those fermions (particles with half-integer spins) that do not interact strongly with the nuclei of atoms, while baryons are the fermions that do. Mesons are bosons (particles with integer spins) subject to strong interactions. Baryons and mesons interact thus mainly via the strong nuclear force and have a common designation of hadrons .
The detailed study of these particles
shows that there are several conserved quantities in their
interactions and decays. The total electric charge, for example, is
always conserved, but also the total number of baryons appears to
be conserved, and thus, the proton, being the lightest baryon,
cannot decay (the present experimental limit for the proton
lifetime is of about 10 years). Strange particles if they
decay by strong interactions give always birth to a lighter strange
particle, but the same is not true when they decay via weak
interactions. To each (totally or partially) conserved quantity, a
new quantum number was associated: baryons, for instance, have
“baryonic quantum number”
(antibaryons have baryonic quantum number
, and mesons have baryonic quantum
number 0).
As a consequence of baryon number
conservation, for example, the most economic way to produce an
antiproton in a proton–proton collision
is the reaction . A proton beam with energy above the
corresponding kinematic threshold is needed to make this process
possible. The Bevatron, a proton synchrotron at the Lawrence
Berkeley National Laboratory providing beams with energy of
6.2 GeV, was designed for this purpose and started operation
in 1954. In the following year, Chamberlain, Segrè, Wiegand, and
Ypsilantis announced the discovery of the antiproton; the new
particle was identified by measuring its momentum and mass using a
spectrometer with a known magnetic field, a Cherenkov detector, and
a time-of-flight system. This discovery confirmed that indeed, as
predicted by the Dirac equation, to each particle corresponds an
oppositely charged particle with the same mass and spin.
The existence of particles and antiparticles is an example of symmetry, and symmetries became more and more present in the characterization of particles and of their interactions. Particle physicists had to study or reinvent group theory in order to profit of the possible simplifications guaranteed by the existence of symmetries.
5.2 Symmetries and Conservation Laws: The Noether Theorem
Being a part of the Universe, it is difficult to imagine how humans can expect to understand it. But we can simplify the representation of the Universe if we find that its description is symmetrical with respect to some transformations. For example, if the representation of the physical world is invariant with respect to translation in space, we can say that the laws of physics are the same everywhere, and this fact greatly simplifies our description of Nature.








Symmetries of a Hamiltonian with respect to given operations entail conservation laws: this fact is demonstrated by the famous Noether theorem.1 In the opinion of the authors, this is one of the most elegant theorems in physics.

5.3 Symmetries and Groups



- 1.
It is closed: the product of any of two elements a, b is an element c of the group(5.8)
- 2.
There is one and only one identity element: the product of any element a by the identity element e is the proper element a(5.9)
- 3.
Each element has an inverse: the product of an element a by its inverse element b (designated also as a
) is the identity e
(5.10) - 4.
The associativity law holds: the product between three elements a, b, c can be carried out as the product of one element by the product of the other two or as the product of the product of two elements by the other element, keeping however the order of the elements:(5.11)

A symmetry is an invariance over a transformation or a group of transformations. In physics, there are well-known symmetries, some related to fundamental properties of space and time. For instance in mechanics, the description of isolated systems is invariant with respect to space and time translations as well as to space rotations. Noether’s theorem grants that for each symmetry of a system there is a corresponding conservation law and therefore a conserved quantity. The formulation of Noether’s theorem in quantum mechanics is particularly elegant.
5.3.1 A Quantum Mechanical View of the Noether’s Theorem









![$$\begin{aligned}{}[\hat{H}, U] = \hat{H}U-U\hat{H} = 0 \, . \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ18.png)
![$$\begin{aligned} \frac{d}{dt} \langle \psi |U|\psi \rangle = -\frac{i}{\hbar } \langle \psi |[U, H]|\psi \rangle = 0. \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ19.png)
5.3.1.1 Continuum Symmetries



![$$\begin{aligned}{}[H, I + i\epsilon G ] = 0 \Longrightarrow [H, G ] = 0 \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ23.png)
![$$[H, I] = 0$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_IEq29.png)
![$$\begin{aligned} \frac{d}{dt} \langle G \rangle = \frac{i}{\hbar } \langle [H, G] \rangle = 0 \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ24.png)

Continuum symmetries in quantum mechanics are thus associated to conservation laws related to the group generators. In the next section, we shall make examples; in particular, we shall see how translational invariance entails momentum conservation (the momentum operator being the generator of space translations).
Let us see now how Noether’s theorem can be extended to discrete symmetries.
5.3.1.2 Discrete Symmetries

![$$\begin{aligned}{}[ \hat{H}, \hat{P} ] = 0 \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ25.png)


Parity-invariant Hamiltonians represent interaction which conserve parity.
Let us examine now some examples of symmetries.
5.3.2 Some Fundamental Symmetries in Quantum Mechanics
5.3.2.1 Phase Shift Invariance






5.3.2.2 Space Translation Invariance







The operator commutes with the Hamiltonian of an
isolated system, and the associated conserved quantity is the
linear momentum
.


5.3.2.3 Rotational Invariance








![$$\begin{aligned} { R}_{x}\left( {\theta }_{x}\right) { R}_{{ y}}\left( {\theta }_{y}\right) { R}_{{ z}}\left( {\theta }_{z}\right) \ne { \exp }\left[ \frac{i}{\hbar }\left( {\theta }_{x}{\hat{L}}_{x}+{\theta }_{y}{\hat{L}}_{y}+{\theta }_{z}{\hat{L}}_{z}\right) \right] : \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ41.png)


![$$\begin{aligned} \exp \left( \hat{A} { + } \hat{B}\right) =\exp \left( \frac{1}{2}\left[ \hat{A},\hat{B}\right] \right) \exp { (}\hat{A}{ )\exp }(\hat{B}) \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ42.png)
![$$\left[ \hat{A},\hat{B}\right] $$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_IEq46.png)
![$$\begin{aligned} \left[ {\hat{L}}_{x},{\hat{L}}_{y}\right] =i\hbar {\hat{L}}_{z} \; ; \; \left[ {\hat{L}}_{y},{\hat{L}}_{z}\right] =i\hbar {\hat{L}}_{x} \; ; \; \left[ {\hat{L}}_{z},{\hat{L}}_{x}\right] =i\hbar {\hat{L}}_{y} \, . \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ43.png)
Once a basis is defined, operators can in most cases be associated to matrices; there is a isomorphism between vectors and states, matrices and operators.2 In the following, whenever there is no ambiguity, we shall identify operators and matrices, and we shall omit when there is no ambiguity the “hat” associated to operators.
5.3.3 Unitary Groups and Special Unitary Groups
Unitary groups U(n) and Special Unitary groups SU(n) of a generic rank n play a central role in particle physics both related to the classification of the elementary particles and to the theories of fundamental interactions.









The special unitary group SU(n) is the
group of unitary complex matrices of dimension and with determinant equal to 1. The
number of free parameters and generators of the group is thus
. Particularly important groups will
be the groups SU(2) and SU(3).
5.3.4 SU(2)












![$$\begin{aligned} \left[ {\hat{S}}_{i},{\hat{S}}_{j}\right] =i\hbar \ {\varepsilon }_{ijk}{\hat{S}}_{k} \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ48.png)


These commutation relations are
identical to those discussed above for the generators of space
rotations (the normal angular momentum operators) in three
dimensions, which form a SO(3) group (group of real orthogonal
matrices of dimension with determinant equal to 1). SU(2)
and SO(3) have
thus the same algebra, and there is a mapping between the elements
of SU(2) and elements of SO(3) which respect the respective group
operations. But, while in our example SO(3) operates in the real
space transforming particle wave functions, SU(2) operates in the
internal space of particle spinors.
































Graphical representation of SU(2) multiplets
5.3.5 SU(3)
SU(3) is the group behind the so-called eightfold way (the organization of baryons and mesons in octets and decuplets, which was the first successful attempt of classification of hadrons) and behind QCD (quantum chromodynamics, the modern theory of strong interactions). Indeed, SU(3) operates in an internal space of three-dimensional complex vectors and thus can accommodate at the same level rotations among three different elements (flavors u, d, s or colors Red, Green, Blue). The eightfold way will be discussed later in this chapter, while QCD will be discussed in Chap. 6; here, we present the basics of SU(3).






















-
leaves Y unchanged and changes
by ±1,
-
changes Y by
and changes
by ±1/2,
-
changes Y by
and changes
by ±1/2.
5.3.6 Discrete Symmetries: Parity, Charge Conjugation, and Time Reversal
Let us examine now in larger detail three fundamental discrete symmetries: parity, charge conjugation, and time reversal.
5.3.6.1 Parity








Electromagnetic and strong interactions appear to be invariant under parity transformations (and therefore the parity quantum number is conserved in these interactions) but, as it will be discussed in the next chapter, weak interactions, with the surprise of most of physicists in the 1950s, are not.







The photon has a negative parity: this can be seen by the fact that the basic atomic transition is characterized by the emission of a photon and a change of orbital angular momentum by one unit. All vector bosons have a negative parity, while the axial vector bosons have a positive parity.
Example: Experimental Determination of the Pion
Parity. By convention, we define that protons and neutrons
have positive intrinsic parity. The negative parity of pions can be
determined by assuming parity and angular momentum conservation in
the capture at rest of a by a deuterium nucleus producing two
neutrons in the final state (
. The parity of a system of two
particles is the product of the parities of the two particles
multiplied by a (
)
factor where l is the orbital angular momentum of the
system (
is the ground state). In the case of
the nn system discussed above,
and thus the final state parity is
. Pseudo-scalar mesons (like pions)
have negative parity, while scalar mesons have positive parity.

5.3.6.2 Charge Conjugation
Charge conjugation reverses the sign of
all “internal” quantum numbers (electric charge, baryon number,
strangeness, ...) keeping the values of mass, momentum, energy, and
spin. It transforms a particle in its own antiparticle. Applying
the charge conjugation operator twice brings the state back to its
original state, as in the case of parity. The eigenvalues of
are again
, but most of the elementary particles
are not eigenstates of
(particle and antiparticle are not
usually the same); this is trivial for electrically charged
particles.
Once again electromagnetic and strong
interactions appear to be invariant under charge conjugation, but
weak interactions are not (they are “almost” invariant with respect
to the product as it will be discussed later
on).
Electric charge changes sign under
charge conjugation and so do electric and magnetic fields. The
photon has thus a negative eigenvalue. The neutral pion
decays into two photons: its
eigenvalue is positive. Now you
should be able to answer the question: is the decay
possible?
5.3.6.3 Time Reversal and CPT





5.3.7 Isospin
In 1932, J. Chadwick discovered the
neutron after more than 10 years of intensive experimental
searches following the observation by Rutherford that to explain
the mass and charges of all atoms, and excluding hydrogen, the
nucleus should consist of protons and of neutral bound states of
electrons and protons. The particle discovered was not a bound
state of electron and proton—meanwhile, the uncertainty relations
demonstrated by Heisenberg in 1927 had indeed forbidden it. The
neutron wasindeed a particle like the
proton with almost the same mass ( MeV/
,
MeV/
), the same behavior with respect to
nuclear interaction, but with no electric charge. It was the
neutral “brother” of the proton.
Soon after neutron discovery, Heisenberg
proposed to regard proton and neutron as two states of a single
particle later on called the nucleon. The formalism was borrowed from
the Pauli spin theory and Wigner, in 1937, called “isospin”
symmetry this new internal symmetry with respect to rotations in
the space defined by the vectors (p, 0) and (0, n). Strong interactions should be
invariant with respect to an internal SU(2) symmetry group, the
nucleons would have isospin , and their states would be described
by isospin spinors. By convention, the proton is identified with
the isospin-up
projection and the neutron with the
isospin-down
projection.
As we have discussed in
Chap. 3, Yukawa postulated in 1935 that the
short-range nuclear interaction might be explained by the exchange
of a massive meson, the pion. The charge independence of nuclear
interactions suggested later on that three pions (,
,
) should exist. Nuclear interaction
could thus be seen as an interaction between the nucleon isospin
doublet
and a isovector
triplet of pions. Three spin 0 and
isospin 1 pions (
with
,
with
,
with
) with almost the same masses
(
139.6 MeV/
,
135.0 MeV/
) were indeed discovered in the late
1940s and in the beginning of the 1950s. The isospin theory of
nuclear interactions was established.
5.3.7.1 The Isospin of a System of Particles
The isospin of a system of particles can be expressed as a function of the isospin of the individual particles using the same addition rules valid for the sum of ordinary spins or angular momenta.





Clebsch–Gordan coefficients and spherical
harmonics. From K.A. Olive et al. (Particle Data Group), Chin.
Phys. C 38 (2014) 090001. Note: a square-root sign is to be
understood over every coefficient, e.g., for read







5.3.7.2 Isospin and Cross Section
Strong interactions are invariant under
SU(2) rotations in the internal isospin space and according to
Noether’s theorem, total isospin is conserved in such interactions.
The transition amplitudes between initial and final states are a
function of the isospin I and
can be labeled as .
- 1.
- 2.
- 3.
.
The deuteron d is a
pn bound state and must have
isospin ; otherwise, the bound states
pp and nn should exist (experimentally, they do
not exist). The isospin quantum numbers
of the final states are thus those of
the
, which means
,
,
, respectively. The isospins of the
initial states follow the rules of the addition of two isospin
1 / 2 states discussed above, and are, respectively,
,
and
. As the final state is a pure
state, only the transition amplitude
corresponding to
is possible. The cross section
(proportional to the square of the sum of the scattering
amplitudes) for the reaction
should then be half of each of the
cross sections of any of the other reactions.


- 1.
- 2.
- 3.
.







Total cross section for the collision of positive and negative pions with protons as a function of the pion kinetic energy.
Credit: E.M Henley and A. Garcia, Subatomic physics, World Scientific 2007




This resonance is now called the
; being a
state it has, as expected, four
projections, called
.
5.3.8 The Eightfold Way

Dharma wheel representing the Buddhist “eightfold path” to liberation from pain.
From Wikimedia Commons
As discussed in Chap. 3, strange particles had been
discovered in the late 1940s in cosmic rays, and later abundantly
produced in early accelerator experiments in the beginning of the
1950s. These particles were indeed strange considering the
knowledge at that time: they have large masses, and they are
produced in pairs with large cross sections, but they have large
lifetimes as compared with what expected for nuclear resonances.
Their production is ruled by strong interactions while they decay
weakly. A new quantum number, strangeness ,
was assigned in 1953 to these particles by Nakano and Nishijima
and, independently, by Gell-Mann. By convention, positive
K mesons (kaons) have
strangeness , while
baryons have strangeness
. “Ordinary” (nonstrange) particles
(proton, neutron, pions,
) have strangeness 0.

















Fundamental meson and baryon octets: on the
left spin 0, parity −1 (pseudo-scalar mesons); on the right the
spin 1/2 baryons. The axis is the abscissa, while the
Y axis is the ordinate

Spin 3/2, parity 1 baryon decuplet. The
axis is the abscissa, while the
Y axis is the ordinate. The
has
, and the
s have









Bubble chamber picture of the first
.
From V.E. Barnes et al., “Observation of a Hyperon with Strangeness Minus Three”, Physical Review Letters 12 (1964) 204
There was however an empty spot in the
decuplet: a baryon with ,
,
,
and a mass around 1670
MeV/
was missing. This particle, which we
call now the
, was indeed
discovered in the Brookhaven National Laboratory 2-meter hydrogen
bubble chamber in 1964 (Fig. 5.9). A
meson interacts with a proton in the
liquid hydrogen of bubble chamber producing a
, a
, and a
, which then decays according to the
following scheme:

Measuring the final state charged
particles and applying energy–momentum conservation, the mass of
the was reconstructed with a value of (1686
± 12) MeV/
, in agreement with the prediction of
Gell-Mann and Ne’eman.
This “exoteric” classification was thus widely accepted, but something more fundamental should be behind it!
5.4 The Quark Model
5.4.1
SU(3)



Fundamental representations and
of SU(3)
This classification of quarks into
u, d, and s states introduces a new symmetry called
flavor symmetry , and the
corresponding SU(3) group is labeled as SU(3) (shortly SU(3)
) whenever confusion is possible with
the group
(shortly
) of strong interactions that will be
discussed in the next chapter.
Q |
I |
|
S |
B |
Y |
|
---|---|---|---|---|---|---|
u |
|
1/2 |
|
0 |
1/3 |
1/3 |
d |
|
1/2 |
|
0 |
1/3 |
1/3 |
s |
|
0 |
0 |
|
1/3 |
|




“nonet” (octet singlet) of pseudo-scalars mesons
There are three states with and
both for pseudo-scalars (
,
,
) and vectors (
,
,
). The
and the
are the already well-known states
with
. The other states should then
correspond to the SU(3) symmetric singlet
, and to the octet isospin singlet,
orthogonal to the SU(3) singlet,
; however, the physically observed
states are mixtures of these two “mathematical” singlets. Due to
these mixings, there is in fact a combination of the SU(3) octet
and singlet which is commonly designated as the “nonet.”
The quark content of a meson can be accessed studying its decay modes.


5.4.2 Color
Color is at the basis of the present
theory of strong interactions, QCD (see Sect. 6.4), and its introduction solves
the so-called puzzle. The
is formed by three u quarks with orbital angular momentum
(it is a ground state) and in the
same spin projection state (the total spin is
). Therefore, its flavor, spin, and
orbital wave functions are symmetric, while the Pauli exclusion
principle imposes that the total wave functions of states of
identical fermions (as it is the case) should be antisymmetric.
In color space, quarks are represented
by complex three-vectors (the generalization of the two-dimensional
spinors). The number of colors in QCD is , as we shall see later in this
Chapter; the quark colors are usually designated as Red, Blue and Green, having the antiquark the
corresponding anticolors.
Quarks interact via the emission or absorption of color field bosons, the gluons. There are eight gluons corresponding to the eight generators of the SU(3) group (see Sect. 5.3.5). Gluons are in turn colored, and the emission of a gluon changes the color.
(Anti)baryons are singlet states obtained by the combination of three (anti)quarks; mesons are singlet states obtained by the combination of one quark and one antiquark. All stable hadrons are color singlets, i.e., they are neutral in color.




OZI favored (left) and suppressed (right)
decay diagrams

Baryon ground states in the quark model: the
spin 1/2 octet (left), and the spin 3/2 decuplet (right). The
vertical (S)-axis corresponds
to the Y-axis, shifted by 1
( corresponds to
).
By Trassiorf [own work, public domain], via Wikimedia Commons
In color space, the physical states are antisymmetric singlets (the total color charge of hadrons is zero, the strong interactions are short range, confined to the interior of hadrons and nuclei). The product of the spin wave function and the flavor wave function in ground states (angular orbital momentum = 0) must then be symmetric. The net result is that the ground-state baryons are organized in a symmetric spin 1/2 octet and in a symmetric spin 3/2 decuplet (Fig. 5.13).
5.4.3 Excited States (Nonzero Angular Momenta Between Quarks)
Hundreds of excited states have been discovered with nonzero orbital angular momentum; they can be organized in successive SU(3) multiplets.

invariant mass plot in
annihilations (left) and in
proton–beryllium interactions (right).
Credits: Nobel foundation
5.4.4 The Charm Quark





Left: 16-plets for the pseudo-scalar
(a) and vector
(b) mesons made of the
u, d, s, and c quarks as a function of isospin
, charm C, and hypercharge
. The nonets of light mesons occupy
the central planes to which the
states have been added. Right: SU(4)
multiplets of baryons made of u, d,
s, and c quarks: (a) The 20-plets including an SU(3)
octet; (b) The 20-plets
with an SU(3) decuplet.
From K.A. Olive et al. (Particle Data Group), Chin. Phys. C 38 (2014) 090001
In terms of the quark model, the
J/ can be interpreted as a
(where c stands for a new quark flavor, called
the charm, which has an
electric charge of 2/3) vector (being produced in
) meson. The possibility of the
existence of a fourth flavor was suggested in 1964 by, among
others, Bjorken and Glashow for symmetry reasons: at that time,
four leptons—the electron, the muon, and their respective
neutrinos—were known, and just three quarks. Later, in 1970,
Glashow, Iliopoulos, and Maiani demonstrated that a fourth quark
was indeed needed to explain the suppression of some neutral current weak processes—this is
the so-called GIM mechanism that will be discussed in
Sect. 6.3.6.
With the existence of a fourth flavor,
the flavor symmetry group changes from SU(3) to SU(4) giving rise
to more complex multiplets which were named “supermultiplets.”
Supermultiplets can be visualized (Fig. 5.15) as solids in a
three-dimensional space , Y, C,
where C is the new charm
quantum number.
A rich spectroscopy of charmed hadrons
was open. For instance, the pseudo-scalar SU(3) octet becomes a
15-particle SU(4) multiplet with seven new mesons with at least a
c quark (,
,
,
,
,
,
), and the spin 3/2 decuplet of
baryons becomes a 20-particle multiplet.
5.4.4.1 Quarkonia: The Charmonium







Hadronization: the string mechanism








Energy levels for (a) the positronium and (b) the charmonium states.
From S. Braibant, G. Giacomelli, and M. Spurio, “Particles and fundamental interactions,” Springer 2012
The right plot of
Fig. 5.17 shows the energy levels of charmonium;
they can be obtained inserting the potential (5.86) into the
Schrödinger equation. One obtains GeV/fm, and
The bottom quark, which will be introduced in the next section, has an even larger mass, and it gives rise to a similar spectroscopy of quarkonia.
5.4.5 Beauty and Top
A fifth quark was discovered a few
years later. In 1977, an experiment in Fermilab led by Leon
Lederman studied the mass spectrum of pairs produced in the interaction of
a 400 GeV proton beam on copper and platinum targets. A new heavy
and narrow resonance, named the upsilon
, was found, with a mass of around
9.46 GeV/
.
The was interpreted as a
vector meson where b stands for a new quark flavor, the
bottom or beauty
, which has,
like the d and the s, an electric charge of
. Several hadrons containing at least
a b quark were discovered. A
family of
mesons, called the bottomium and indicated by the letter
, was there, as well as mesons and
baryons resulting from the combination of b quarks with lighter quarks:
pseudo-scalar mesons like the
,
, the
, and the
; bottom baryons like
,
,
,
. Heavy mesons and baryons with a
single heavy quark are very interesting. The light quarks surround
the heavy quark in a strong force analogy of electrons around the
proton in the electromagnetically bound hydrogen atom.
The discovery of the b quark inaugurated a third generation or
family of quarks. Each family is formed
by two quarks, one with electric charge and the other with electric charge
(the first family is formed by quarks
u and d; the second by quarks c and s). In the lepton sector, as will be
discussed at the end of this chapter, there were at that time
already five known leptons (the electron, the muon, their
corresponding neutrinos, and a recently discovered heavy charged
lepton, the tau) . With the bottom quark,
the symmetry was restored between quarks and leptons but the sixth
partners both in the quark and in lepton (the tau neutrino) sector
were missing. The existence of a third family of quarks had indeed
been predicted already in 1973, before the discovery of the
, by Makoto Kobayashi and Toshihide
Maskawa, to accommodate in the quark model the CP violation observed in the
system (this will be discussed in
Chap. 6). The hypothetical sixth quark was
named before its discovery the “top” quark .
The top quark was missing and for many
years, and many people looked for it in many laboratories (in the
USA and in Germany, Japan, CERN), both at electron–positron and at
proton–(anti)proton colliders. A strong indication of a top with a
mass around 40 GeV/ was even announced in 1984 but soon
dismissed. Lower limits on the top mass were later established, and
indications on the value of the mass were derived from the standard
model (see Chap. 7); finally, in 1995, the discovery
of the top quark was published by the CDF experiment (and soon
after by the D0 experiment) at Fermilab, at a mass of (176 ± 18)
GeV/
. The present (2018) world average of
the direct measurements is (173.1 ± 0.6) GeV/
. The top is heavier than a gold
nucleus; with such a large mass, its decay phase space is huge and
its lifetime is very short, even if the decay is mediated by the
weak force. It is so short (the estimated value is around
s) that the top does not live
long enough to hadronize: there are no top hadrons.
5.4.6 Exotic Hadrons
It is evident that it is possible to
form other color singlet bound states than the ordinary mesons
() and baryons (qqq). Indeed many states are predicted
formed by: just gluons (glueballs); two
quarks and two antiquarks (tetraquarks);
four quarks and one antiquark (pentaquarks); or six quarks (hexaquarks). These hadrons had been searched for
long and many candidates did exist, but only recently (2014 and
2015) the LHCb collaboration at CERN confirmed the existence of a
tetraquark (the Z(4430), a
bound
state) and two pentaquarks (the
and the
, both bound
states). A rich spectroscopy can be
studied in the future.
5.4.7 Quark Families





5.5 Quarks and Partons
In the words of Murray Gell-Mann in 1967, quarks seemed to be just mathematical entities. This picture was deeply changed in a few years by the results of deep inelastic scattering experiments.
Indeed in the 1950s Robert
Hofstadter,6 in a series of Rutherford-like experiments
using a beam of electrons instead of particles (electrons have no strong
interactions), showed departures from the expected elastic point
cross section. Nucleons (protons and neutrons) are not point-like
particles. The proton must have a structure and the quarks could be
thought as its constituents.
5.5.1 Elastic Scattering

![$$\begin{aligned} \frac{d\sigma }{d\varOmega }=\frac{\alpha ^{2}\cos ^{2}\left( \frac{\theta }{2}\right) }{4 E^{2}\sin ^{4}\left( \frac{\theta }{2}\right) }\frac{{ E'}}{E}\left[ 1+\frac{Q^{2}}{2m_{p}^{2}} \tan ^{2}\left( \frac{\theta }{2}\right) \right] \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ87.png)






![$$\begin{aligned} \frac{d\sigma }{d\varOmega }=\frac{{\alpha }^{2}\cos ^{2}\left( \frac{\theta }{2}\right) }{4\ E^{2}\sin ^{4}\left( \frac{\theta }{2}\right) }\frac{{ E'}}{E}\left[ \frac{G_{E}^{2}\left( Q^{2}\right) +\frac{Q^{2}}{4m_{p}^{2}}G_{M}^{2}\left( Q^{2}\right) }{1+\frac{Q^{2}}{4m_{p}^{2}}}+2\frac{Q^{2}}{4m_{p}^{2}}\ G_{M}^{2}\left( Q^{2}\right) \tan ^{2}\left( \frac{\theta }{2}\right) \right] \nonumber \\ \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ89.png)



At low ,
and
can be interpreted as the Fourier
transforms of the electric charge and of the magnetization current
density inside the proton. In the limit
(
for the exchanged virtual photon),
the electron “sees” the entire proton and it could be expected that
. This is what the experiment tells
for
, but it is not the case for
. In fact, the measured value is
. The proton has an anomalous magnetic
moment
which reveals already that the proton
is not a Dirac point-like particle. The same is observed for the
neutron which has
.







5.5.2 Inelastic Scattering Kinematics










Deep inelastic scattering kinematics




-
the lost energy(5.96)
in the laboratory reference frame is the energy lost by the electron:
(5.97) -
the inelasticity y(5.98)
. In the laboratory frame is the fraction of the energy lost by the electron:
(5.99) -
the Bjorken variable, x(5.100)
. Using the definition of
, x can also be expressed as:
(5.101)(5.102)then
, the elastic scattering formula is recovered.
At a fixed center-of-mass energy,
, the inelastic scattering final state
can be characterized by the Lorentz invariant variables,
,
, x, y,
as well as by the scattered electron energy
and scattered angle
in the laboratory reference frame.
However, from all those variables, only two are independent. The
experimental choice is usually the directly measured variables
and
, while the theoretical interpretation
is usually done in terms of
and
or
and x.




5.5.3 Deep Inelastic Scattering




![$$\begin{aligned} \frac{d\sigma }{d\varOmega { d}E'}= \ \frac{{\alpha }^{2}{\cos }^{2}\left( \frac{\theta }{2}\right) }{4\ E^{2}\,{\sin }^{4}\left( \frac{\theta }{2}\right) }\frac{{ E'}}{E}\left[ W_{2}\left( Q^{2},\nu \right) +2W_{1}\left( Q^{2},\nu \right) \ {\tan }^{2}\left( \frac{\theta }{2}\right) \right] . \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ107.png)
























While Bjorken was suggesting the
scaling hypothesis, the groups lead by J. Friedman, H. Kendall, and
R. Taylor7 designed and built at SLAC electron
spectrometers able to measure energies up to 20 GeV for
different scattering angles. The strong dependence of the elastic form
factors was then confirmed up to
30 GeV
while, surprisingly, in the region
GeV, the inelastic cross section
showed a very mild
dependence (Fig. 5.19, left).






Left: Deep inelastic electron–proton
differential cross section normalized to the Mott cross section, as
measured by SLAC. From Ref. [F5.2] in the “Further readings.”
Adapted from Nobel Foundation, 1990. Right: scaling: measurements at different
scattered energies and angles. From W. Atwood, “Lepton Nucleon
Scattering,” Proc. 1979 SLAC Summer Institute (SLAC-R-224)

scaling for the proton: measurements
at fixed x and at different
.
From J.I. Friedman and H.W. Kendall, Annual Rev. Nucl. Science 22 (1972) 203

Representation of electron–parton scattering in Feynman’s Quark–Parton Model
The Feynman partons were soon after identified as the Gell-Mann and Ne’eman quarks. However, nowadays, the term “parton” is used often to denominate all the nucleon constituents, i.e., the quarks and antiquarks and even the gluons.
5.5.4 The Quark–Parton Model




























Validation of the Callan–Gross relation.
Adapted from S. Braibant, G. Giacomelli and M. Spurio, “Particles and fundamental interactions,” Springer 2012






Quarks in hadrons may have different
flavors and thus different charges and masses. The corresponding
PDFs are denominated according to the corresponding flavor:
u(x), d(x),
s(x), c(x),
b(x) t(x)
for quarks;
for antiquarks.

![$$\begin{aligned} {F_{2}}^{ep}\left( Q^{2}, x\right) \simeq x\ \left[ \frac{4}{9}\left( u(x)+\bar{u}(x)\right) +\frac{1}{9}\left( d(x)+\bar{d}(x)+s(x)+\bar{s}(x)\right) \right] . \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ121.png)












Parton distribution functions at
Q GeV
(left) and Q
GeV
(right). The gluon and sea
distributions are scaled down by a factor of 10. The experimental,
model, and parameterization uncertainties are shown.
From K.A. Olive et al. (Particle Data Group), Chin. Phys. C 38 (2014) 090001



![$$\begin{aligned} {F_{2}}^{ep}\left( Q^{2}, x\right) \simeq \ x\ \left[ \frac{4}{9} u_v^p(x) + \frac{1}{9} d_v^p(x) + \frac{10}{9} \bar{u}^p(x) + \frac{2}{9} \bar{s}^p(x) \right] \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ127.png)
![$$\begin{aligned} {F_{2}}^{en}\left( Q^{2}, x\right) \simeq \ x\ \left[ \frac{1}{9} u_v^p(x) + \frac{4}{9} d_v^p(x) + \frac{10}{9} \bar{u}^p(x) + \frac{2}{9} \bar{s}^p(x) \right] \end{aligned}$$](/epubstore/A/A-D-Angelis/Introduction-To-Particle-And-Astroparticle-Physics/OEBPS/images/304327_2_En_5_Chapter/304327_2_En_5_Chapter_TeX_Equ128.png)

















dependence of
.
From K.A. Olive et al. (Particle Data Group), Chin. Phys. C 38 (2014) 090001
5.5.5 The Number of Quark Colors






















Measurements of .
Adapted from [F5.2] in the “further readings”; the data are taken from K.A. Olive et al. (Particle Data Group), Chin. Phys. C 38 (2014) 090001
5.6 Leptons


The (tau) lepton, with its neutrino, was
the last discovered.
5.6.1 The
Discovery of the
Lepton



Left: An event observed at Mark I. The muon
moves upward and the electron downward. The numbers 13 and 113 give
the relative amount of the electromagnetic energy deposited.
Credit: Martin Perl et al., M. L. Perl et al., “Evidence for
Anomalous Lepton Production in
Annihilation,” Phys. Rev. Lett. 35
(1975) 1489. Right: Feynman diagram for the
decay in
,
,
.
By en:User: JabberWok and Time 3000 [GFDL http://www.gnu.org/copyleft/fdl.html], via Wikimedia Commons











5.6.2 Three Neutrinos





Feynman diagram for the decay in the case where just one neutrino
species exists
Around the same time, the possibility
to produce a neutrino beam from the decay of pions created in the
collision of GeV protons on a target was intensively discussed in
the cafeteria of the Columbia University. In 1962, a kind of a
neutrino beam was finally available at Brookhaven National
Laboratory (BNL): the idea was to put an internal target in a long
straight section of the proton accelerator and to drive with a
magnet the proton beam on it; the pions coming from the proton
interactions were then decaying into pions. An experiment led by
Leon Lederman, Melvin Schwartz, and Jack Steinberger was set to
observe the neutrino reaction within a 10-ton spark chamber.
Hundreds of millions of neutrinos were produced mostly accompanied
by a muon ( as it will be discussed in the next
chapter). Forty neutrino interactions in the detector were clearly
identified; in six of them, the final state was an electron, and in
thirty-four, the final state was a muon. The
and
are, thus, different particles,
otherwise the same number of events with one electron and with one
muon in the final state should have been observed.


Tau-neutrino event in DONUT. A tau neutrino produces several charged particles. Among them a tau particle, which decays to another charged particle with missing energy (at least one neutrino).
From K. Kodama et al., DONUT Collaboration, “Observation of tau-neutrino interactions,” Phys. Lett. B 504 (2001) 218
5.7 The Particle Data Group and the Particle Data Book
How can one manage all this information about so many particles? How can one remember all these names? The explosion of particle discoveries has been so great, that Fermi said, “If I could remember the names of all these particles, I’d be a botanist.”
Fortunately, a book, called the Review of Particle Physics (also known as the Particle Data Book) , can help us. It is edited by the Particle Data Group (in short PDG), an international collaboration of about 150 particle physicists, helped by some 500 consultants, that compiles and organizes published results related to the properties of particles and fundamental interactions, reviewing in addition theoretical advancements relevant for experimental physics. The PDG publishes the Review of Particle Physics and its pocket version, the Particle Physics Booklet, which are printed biennially in paper, and updated annually in the Web. The PDG also maintains the standard numbering scheme for particles in event generators (Monte Carlo simulations).
-
Particle physics summary tables—Brief tables with the properties of of particles.
-
Reviews, tables, and plots—Review of fundamental concepts from mathematics and statistics, tables related to the chemical and physical properties of materials, review of current status in the fields of standard model, cosmology, and experimental methods of particle physics, tables of fundamental physical and astronomical constants, summaries of relevant theoretical subjects.
-
Particle listings—Extended version of the Particle Physics Summary Tables, with reference to the experimental measurements.
The Particle Physics Booklet (about 300 pages) is a condensed version of the Review, including the summary tables, and a shortened section of reviews, tables, and plots.
The publication of the Review of Particle Physics in its present form started in 1970; formally, it is a journal publication, appearing in different journals depending on the year.
5.7.1 PDG: Estimates of Physical Quantities
The “particle listings” (from which the “summary tables” are extracted) contain all relevant data known to the PDG team that are published in journals. From these data, “world averages” are calculated.
-
it is superseded by or included in later results.
-
no error is given.
-
it involves questionable assumptions.
-
it has a poor signal-to-noise ratio, low statistical significance, or is otherwise of poorer quality than other data available.
-
it is clearly inconsistent with other results that appear to be more reliable.
-
OUR AVERAGE—From a weighted average of selected data.
-
OUR FIT—From a constrained or overdetermined multiparameter fit of data.
-
OUR EVALUATION—Not from a direct measurement, but evaluated from measurements of related quantities.
-
OUR ESTIMATE—Based on the observed range of the data, not from a formal statistical procedure.
5.7.2 Averaging Procedures by the PDG
The average is computed by a
minimization. There is an attempt to
use uncorrelated variables as much as possible, but correlations
are taken into account.
When the error for a measurement
x is asymmetric, the error used
is a continuous function of the errors and
. When the resultant average
x is less than
,
is used; when it is greater than
,
is used; in between, the error is a
linear function of x.



-
If
is less than or equal to 1, and there are no known problems with the data, the results are accepted.
-
If
is very large, the PDG may
-
not to use the average at all, or
-
quote the calculated average, making an educated (conservative) guess of the error.
-
-
If
is greater than 1, but not so much, the PDG still averages the data, but then also increases the error by
. This scaling procedure for errors does not affect central values.





An ideogram representing clearly inconsistent data—the measurements of the mass of the charged kaon. The arrow at the top shows the position of the weighted average, while the width of the shaded pattern shows the error in the average after scaling by the factor S.
From K.A. Olive et al. (Particle Data Group), Chin. Phys. C 38 (2014) 090001
A short summary of particle properties is also listed in the Appendix D of this book.
Further Reading
- [F5.1]
S. Haywood, “Symmetries and Conservation laws in Particle Physics: an introduction to group theory for experimental physicists,” Imperial College Press 2011. An excellent introduction to group theory and its application in particle physics.
- [F5.2]
A. Bettini, “Introduction to Elementary Particle Physics,” Cambridge University Press 2014. A very good introduction to Particle Physics at the undergraduate level putting together experimental and theoretical aspects and discussing basic and relevant experiments.
- [F5.3]
M. Thomson, “Modern Particle Physics,” Cambridge University Press, 2013. A recent, pedagogical and rigorous book covering the main aspects of Particle Physics at advanced undergraduate and early graduate level.
- [F5.4]
PDG 2017, C. Patrignani et al. (Particle Data Group), Chin. Phys. C, 40, 100001 (2016) and 2017 update. URL: http://pdg.lbl.gov/. Including the previous editions, this is the most quoted reference in this book.
- 1.
Kinematic thresholds and conservation laws. Compute the kinematic threshold of the reaction
in a fixed target experiment.
- 2.
Can neutron be a bound state of electron and proton? The hypothesis that the neutron is a bound state of electron and proton is inconsistent with Heisenberg’s uncertainty principle. Why?
- 3.
- 4.
Parity. Verify explicitly if the spherical harmonics
- 5.
Constructing baryons. How many different baryon combinations can you make with 1, 2, 3, 4, 5, or 6 different quark flavors? What’s the general formula for n flavors?
- 6.
Baryons. Using four quarks (u, d, s, and c), construct a table of all the possible baryon species. How many combinations carry a charm of
1? How many carry charm
2, and
3?
- 7.
Compositeness of quarks? M. Shupe [Phys. Lett. B 611, 87 (1979)] has proposed that all quarks and leptons are composed of two even more elementary constituents: c (with charge −1/3) and n (with charge zero) - and their respective antiparticles. You are allowed to combine them in groups of three particles or three antiparticles (ccn, for example, or nnn). Construct all of the eight quarks and leptons in the first generation in this manner. (The other generations are supposed to be excited states.) Notice that each of the quark states admits three possible permutations (ccn, cnc, ncc, for example)—these correspond to the three colors.
- 8.
Decays of the
baryon. Which decay do you think would be more likely:
- 9.
Decay of charmed mesons. Which decay do you think would be least likely:
- 10.
Cross sections and isospin. Determine the ratio of the following interactions cross sections at the
resonance:
- 11.
Decay branching ratios and isospin. Consider the decays of the
into
,
and
. Determine the ratios between the decay rates in these decay channels.
- 12.
Quantum numbers. Verify if the following reactions/decays are possible and if not say why:
- (a)
,
- (b)
,
- (c)
,
- (d)
,
- (e)
,
- (f)
,
- (g)
,
- (h)
- (a)
- 13.
Width and lifetime of the
. The width of the
meson is
keV. What is its lifetime? Could you imagine an experiment to measure it directly?
- 14.
mass. Verify the relations between the masses of all the particles lying in the fundamental baryon decuplet but the
and predict the mass of this one. Compare your prediction with the measured values.
- 15.
Decays and conservation laws. Is the decay
possible?
- 16.
Experimental resolution in deep inelastic scattering. Consider an
deep inelastic scattering experiment where the electron scattering angle is
6
. Make an estimation of the experimental resolution in the measurement of the energy of the scattered electron that is needed to distinguish the elastic channel (
) from the first inelastic channel (
).
- 17.
e
deep inelastic scattering kinematics. Consider the
deep inelastic scattering and deduce the following formulae:
- 18.
Gottfried sum rule. Deduce in the framework of the quark–parton model the Gottfried sum rule
and
deep inelastic scattering experiments is approximately 1/4.