You are on page 1of 46

B.Sc. Physics(Hons.

)
Semester 5
Discipline Specific Elective (DSE)
HPHDS5021T
Module B: Particle Physics
BOOKS:
(1) Atomic Physics by J. Yarwood
(2) Nuclear Physics by S.N.Ghoshal
(3) Introductory Nuclear Physics by K.S. Krane
(4) Introduction to High Energy Physics by D.H. Perkins
(5) Introduction to Elementary Particle Physics by A. Bettini

Conduction of electricity in gases

X-rays .
.

Fig.1

current

0 p.d. across electrodes


Fig.2
This Photo by Unknown Author is licensed under CC BY-
SA

Theoretical explanation of the characteristic of an ionized gas:


If n2 positive ions exist among n1 negative ions, then R, the number of
recombinations per second, is given by
R = αrn1n2 (1)
If n1 = n2 = n, then
R = αrn2 (2)
The mobility k of a gaseous ion is defined as the velocity acquired by an ion in
unit electric field. The velocity acquired in an electric field E is known as the
drift or ionic velocity; it is in the direction of the applied field and is
superimposed on the much greater thermal agitation velocities of the mixture of
ions and molecules,
u = kE (3)
Suppose the ionizing agent produces q pairs of ions per cubic centimeter per
second, i.e. there are q positive ions and q electrons produced. If the electric
field E = 0, then the ions would recombine at a rate given by eqn.(2). At
equilibrium, the number of ions being produced per second equals the number
recombining per second, therefore,
q = R = αrn2
𝑞
So, n = √𝛼 (4)
𝑟

If the electric field applied across the gas is E, then the positive ions acquire a
velocity up = kpE towards the cathode, whilst the electrons move with a velocity
un = knE towards the anode.
Current density J = (kp + kn)Ene (5)
For small field strengths, it may be assumed that n is given by eqn. (4). Then
𝑞
J = (kp + kn)√𝛼 eE (6)
𝑟

So J = cE, where c is a constant (7)


If the electric field applied across the gas is increased, recombination becomes a
negligible effect, The saturation current Is will correspond to a case where all
the ions produced at the rate of q pairs per cubic centimeter per second, by the
ionizing agent, are collected by the electrodes. Therefore,
Is = qev , where v is the volume of the gas ionized (8)
Explanation of different regions of the curve:
The regions CD and DE of the curve was explained by Townsend on the basis
that gaseous ions produce further ions on collision with the neutral gas
molecules, if the field strength was sufficiently great.
Cathode C
x
dx d

Anode A
Fig.3
Theories of ionization by collision:
(a) Assuming that only electrons produce ionization by collision
Consider one electron released from the cathode and say, it produces α pair of
ions, i.e. α positive ions and α electrons by collision per centimeter length of
path of the electron in the gas. Suppose that at a distance x from the cathode, the
number of electrons existing (including the original electron) is n. During a
travel of a further infinitesimal distance dx, the number of positive ions and
hence additional number of electrons is
dn = nαdx
Integrating, ln n = αx + c
When x = 0, n = 1
When x = d, n = nd, ln nd = αd
So, nd = eαd
If N0 electrons are released from the cathode per second by the photoelectric
effect, the total number of electrons N reaching the anode per second is hence
given by
N = N0eαd (7)
And J = J0eαd (8)
Here J0 is the initial current density due to the photoelectric effect, J is the
current density on ionization by collision and J = Ne. Here α is called the first
Townsend coefficient. A possibility of positive ions creating more ion pairs by
collision was considered incorrect after much debate and experimentation.
Though the positive ions are incapable of producing ionization by collision,
they do release electrons from the material of the cathode when they impinge
upon it under the action of the electric field, i.e. they provide the work function
energy necessary to cause electron emission. The ratio of the number of
electrons released by the cathode to the number of positive ions impinging upon
it per second is called γ, known as the second Townsend coefficient. This γ is a
function of E/p (p being the pressure)
Methods of measuring mobilities of gaseous ions:
(a) Air-blast method, in which is determined the ratio of velocity of the ion
acquired in a stream of gas having a known velocity to the velocity
acquired by the ion in a known, transverse electric field.
(b) Methods which employ an alternating potential
(c) Methods employing an electrical gate or shutter, in which the time of
passage of a narrow strip of ions through a known distance in a known
electric field is found.
Gas filled tubes:
Ionisation chamber
Proportional Counter
Geiger Mȕller counter
A typical G.M. tube consists of a cylindrical copper cathode with a thin axial
tungsten wire anode supported inside a cylindrical glass envelope. The second
type is an end window counter. This pattern is useful for counting the less
penetrating radiations. These thin windows are demanded in counters for α
particles and the less energetic β particles which are unable to penetrate the
glass wall of the conventional counter. The anode is a tungsten wire and the
cathode is in the form of a graphite coating on the inside wall of the glass
envelope. Many sizes of counting tubes are made from 1 to 5 cm in diameter
and 2 to 100 cm in length.

Let a and b the radii of the anode wire and the cathode respectively. The
value of a varies between 0.01 and 0.03 cm, while b is frequently 1 cm. The
electric field Er at a distance r from the central axis is given by
𝑉
Er = 𝑏
𝑟𝑙𝑛 𝑎

The onset of ionization on collision depends on the ratio of E/p. When it


occurs, fresh electrons produced themselves create new ion pairs; the
cumulative effect resulting leads rapidly to the development of a Townsend
avalanche of ions. In an ionization chamber, the gas amplification factor m is 1.
In the proportional counter, m is between 103 and 104. In a G.M. counter, m is
about 108.
The electron avalanche is concentrated near the anode wire. The positive ions,
being comparatively heavier, travel comparatively slowly towards the cathode.
The time of flight of the positive ions is about 100 times the time required to
build an electron avalanche. This time of flight is about 100 μs in a counter with
a cathode radius of 1 cm.
The object of a polyatomic alcohol vapor or a diatomic bromine vapor is to
prohibit the production of spurious pulses following the main required pulse due
to a single particle entry, This vapor is called a quenching gas.
Ionisation potential Vi = 15.8 V for argon
Work function of cathode surface, φ is not more than 5 V
For molecular bromine, Vi = 12.8 V
An alcohol quenched counter has a lifetime of about 109 counts
If sufficient quenching gas is present, the inert gas ions, in their comparatively
long path to the cathode, will almost all collide with molecules of the quenching
gas and transfer their charges to them. The inert gas ions will then be
neutralized. Only molecular ions of the quenching gas will reach the cathode.
When a molecular ion is very near the cathode, it will capture an electron from
the cathode surface and become neutralized. The molecule will then be left with
𝑞 𝑞
an energy of (𝑉𝑖 – φ) eV, where 𝑉𝑖 is the ionization potential of the molecule.
However, unlike the inert gas atom with excess energy, the excess energy of the
neutralized molecule will cause it to dissociate.
A bromine quenched counter has unlimited life. The cathode material must be
resistant to chemically active bromine vapor. Stainless steel, nichrome and
graphite are suitable.
G.M. tubes can detect about 5000 particles per second.
For an alcohol quenched counter, the operating voltage is 1000 to 1200 V
For a bromine quenched counter, the operating voltage is 200 to 300 V

H.T. R

Pulse Amp. Scaler ckt. Register

Fig. 3
The dead time is defined as the time interval between production of the initial
pulse and the initiation of the second Geiger discharge. This arises due to the
slower movement of the heavier positive ions from the anode region to the
cathode. Since the presence of the positive ion sheath around the anode lowers
the electric field below the Geiger threshold, the counter can record another
ionizing particle only after the field has been restored to a value above the
Geiger threshold Vth. The actual resolving time of a counter is somewhat longer
than the dead time. Since a finite pulse must develop before it can be counted by
the counter circuit, though the two terms are interchangeable.
The recovery time of the counter is the time interval after which the
counter returns to its original state to produce the full sized pulses again.
Let resolving time of a counter = τ
Because of the finite resolving time τ, the counter does not respond to all the
events occurring in it. If N is the actual rate of arrival of the ionizing particles
within a counter, then the counting rate n that the counter is able to record will
be less than N because it is unable to respond to the particles for an interval nτ
per second. So the number of particles missed by the counter per second will be
Nnτ, which should thus be equal to N – n.
Thus, Nnτ = N – n
𝑛
And, N = 1−𝑛𝜏

If β particles enter a G.M. tube at a constant rate from a radioactive source and
the number of pulses recorded per second is plotted against the p.d. across the
tube, a curve like that shown in the figure below is obtained. The counter will
remain inoperative until the voltage is sufficient to cause the development of an
electron avalanche. At the Geiger threshold voltage, the pulse size is constant.
Beyond this threshold, further increase of voltage causes little increase of the
number of counts. This gives the plateau region which is over 200 V for a good
tube. The operating voltage is chosen to be on the plateau. The operating
voltage minus the threshold voltage is the over-voltage of the tube. The pulse
size increases in proportion to the over voltage; using bromine quenched tubes,
voltage pulses of 20 V can readily be achieved if required.
Neutron sources
Thermal neutrons have energy of 0.025 eV
Epithermal neutrons have energies of the order of 1 eV
Slow neutrons have energies of the order of 1 keV
Fast neutrons have energies of the order of 1 to 10 MeV
Detection of neutrons
Using the following reactions:
10
B (n,α) 7Li
10B (n,α) 7Li*
6Li (n,α) 3H
Boron trifluoride proportional counter:
The target nuclei may be present within the detector, either as a constituent
of the counter gas or as a coating on the inside walls. Boron trifluoride is a
compound of boron which is a gas at ordinary room temperature (B.P. -101 ºC).
It has a good proportional counter and ionization chamber characteristics if it is
free from impurities. The thermal neutrons absorbed by the 10B nucleus release
2.31 MeV energy which is shared by the two nuclei 7Li and 4He in the ratio 4:7.
The alpha particles have a range of about 0.8 cm in the BF3 gas filling the
counter at a pressure of 1 atm. The range of the lithium nucleus is much shorter.
Slow neutron detection through induced radioactivity measurement: Neutron
induced nuclear reaction in many nuclei result in the production of residual
nuclei which are radioactive. In particular, the (n,γ) reaction induced by slow
neutrons produce β-active end products in many cases. These have high thermal
neutron cross-section because of the low 1/v law. It may be noted that in
reactions involving low energy neutrons, like (n,α), (n,p), (n,γ), the cross section
σ is proportional to 1/v, where v is the speed of the neutron. Some examples are
given in this table:
Target Isotope Abundance(%) σ(n,γ) Activity τ
barns
Manganese 55Mn 100 13.3 β- 2.58 h
Copper 63Cu 69.2 4.4 β-,β+, E.C. 12.7 h
65Cu 30.8 2.2 β- 5.1 min

Detection of fast neutrons: The most common method of detecting and


measuring fast neutron fluxes is based on the observation of recoil protons in
neutron-proton scattering, The cross section is energy dependent. A simple
neutron detector based on this method consists of a proportional counter filled
with hydrogen or a hydrocarbon (e.g. methane or propane). Neutrons with
energies down to about 20 keV can be detected by such a counter.

Semiconductor detectors
Reverse biased pn junction. The energy required to produce an electron-hole
pair is about 3 eV (compared to about 30 eV to produce an electron-positive ion
pair in a gas filled tube). The width d of the depletion layer is given by
2𝑘𝜖0 (𝑉0 + 𝑉𝑏 ) 1/2
d = { }
𝑒(𝑛𝑎 − 𝑛𝑑 )

Here, Vb is the applied p.d. and V0 is the junction p.d. existing at the junction
due to the diffusion of charge carriers. Also, k is the ratio of permittivities of the
material and vacuum.
For silicon, V0 = 0.7 V, k = 11.8
For germanium, V0 = 0.2 V, k = 16
These detectors, known as diffused junction detectors, should always be kept
in complete darkness to prevent photo-electric current flow. Since the depletion
region in these detectors is only a few mm thick, they can be used for
spectrometry of particles of low and intermediate energies (about 1.5 MeV for
electrons, 20 MeV for protons, and 80 MeV for α-particles).
Diffused junction detectors are usually produced by diffusion of a high
concentration of an n-type impurity, e.g. phosphorus (pentavalent) on the
surface of a p-type silicon crystal. This forms a p-n junction at a depth of about
0.1 to 0.2 μm below the surface. The back contact is through a p+ boron
(trivalent) diffused layer. A coating of gold on either side provides the electrical
contact. As the n+ layer is heavily doped, the depletion region extends primarily
into the p-type silicon.
Surface barrier detectors are usually made of n-type silicon. These detectors
have a very thin dead layer at the window (about 0.1 μm). As aα result there is
very small energy loss for the incoming radiation at the window which is
convenient for relatively low energy α-particle studies. Light sensitivity is high
so that proper light shielding arrangement must be provided.
These are some important characteristics of surface barrier detectors:
(1) Linear energy response at different energies for different types of
particles
(2) High energy resolution which gives rise to narrow pulses. Typically, full
width at half maximum (FWHM) is about 10 keV for α-particles of
several MeV, which is somewhat higher than that for protons or
deuterons. Cooling improves the resolution.
(3) Time response is very quick being in the ns region because of the short
distance moved by the electron-hole pairs before collision
(4) High conversion efficiency due to much smaller energy required for
producing electron-hole pairs, which is less than one-tenth of that
required in case of gaseous ionization detectors.
d = (2μL Vbt)1/2
Where d = width of compensation zone (cm)
μL = mobility of lithium ions cm2/volt-sec
Vb = applied reverse bias (V)
t = drift time (s)
Si(Li) detectors: It is difficult to achieve depletion layers of more than 2 mm
thickness in silicon surface barrier detectors due to relatively higher impurity
concentration, as a result of which higher bias voltages cannot be applied.
Thicker detectors of silicon can however be made by lithium drifting technique.
Such detectors are known as lithium drifted silicon detectors or simply Si(Li)
detectors. Lithium drifting helps achieve very high degree of impurity
compensation which permits application of higher bias voltage. The thickness
of the compensation zone is proportional to the drifting time t which is the time
for the lithium ions to drift from the surface of the silicon wafer in the electric
field at the junction.
The Si(Li) detectors are usually employed for the spectroscopy of high energy
charged particles. For this purpose they are normally operated at room
temperature. For photons, they must be cooled to achieve higher energy
resolution.
Ge(Li) detectors: For X-ray and γ-ray spectrometry, lithium drifted germanium
detectors Ge(Li) are prepared since the atomic number of germanium is higher
(Z = 32) than that of silicon (Z= 14), the photopeak efficiency being
proportional to Z5. The procedure for preparing these detectors is is similar to
that for silicon, differing in details. Both planar and coaxial configurations are
available for Ge(Li) detectors.
Scintillation counters:
Inorganic scintillators: NaI(Tl)
Organic scintillator: anthracene
Inorganic scintillator: activator states in band gap

Conduction band

Activator states

Valence Band

To increase the probability for photon emission and to reduce self-absorption


of light, small amounts of impurities called activators are added to the crystal.
A commonly used activator is thallium, and so these detectors are indicated as,
for instance NaI(Tl). The activator provides states in the energy gap and the
light emission takes place between the activator states. In the case of NaI, the
wavelength of maximum emission is shifted from 303 nm in pure NaI to 410
nm in NaI(Tl). Absorption at this energy cannot occur in NaI(Tl), because the
activator ground states are not populated and the change in wavelength from the
ultraviolet to the visible gives a better overlap with the maximum sensitivity of
most photomultiplier tubes.
In aromatic hydrocarbons, such as those typified by the ring structure of
benzene, three of the four valence electrons are in the hybridized orbitals called
σ orbitals; these are strongly localized between each carbon, its two carbon
neighbors, and a single hydrogen. The fourth electron, which is in the so-called
𝜋 orbital, is not as well localized and does not participate in the bonding process
as strongly as the σ electrons. It is this 𝜋 electron that is most responsible for the
scintillation process.

This Photo by Unknown Author is licensed under CC BY-SA

Inorganic scintillators
Scintillator Wavelength Decay time Refractive Density Relative Remarks
of max. (ns) index at λ (g/c,c,) scintillation
emission efficiency
(nm)
NaI(Tl) 410 230 1.85 3.67 100 Hygroscopic,
good energy
resolution
CsI(Tl) 565 1000 1.80 4.51 45 Non-
hygroscopic,
hjgh γ
absorption,
useful for
charged
particle
spectroscopy
Organic scintillators
Scintillator Wavelength Decay time Refractive Density Relative Remarks
of max. (ns) index at λ (g/c.c.) scintillation
emiddion efficiency
(nm)
Anthracene 447 30 1.59 1.25 25 Useful for β
detection

Organic solutions
Scintillator Wavelength Decay time Refractive Density Relative Remarks
of (ns) index at λ (g/l) scintillation
maximum efficiency
emission
(nm)
p-terphenyl 350 1.5 1.50 5 10 Useful for
in toluene soft β
detection

This Photo by Unknown Author is licensed


under CC BY-SA
Anthracene
Photomultiplier tubes
A photomultiplier tube is a sealed tube made of glass or quartz at one end of
which is coated a semi-transparent layer of some photosensitive material, e.g.
CS3Sb which is an intermetallic compound of caesium and antimony. This is
known as the photocathode. When scintillation photons fall on the
photocathode, photoelectrons are emitted with a quantum efficiency of 20% to
30%. The latter is defined as the ratio of the number of electrons emitted to the
number of incident photons.
Facing the photocathode there is a series of secondary electron emitting
surfaces known as the dynodes. The presence of the dynodes gives rise to the
multiplication in the number of electrons finally collected at the anode for each
electron emitted from the photocathode. The first dynode is kept at a positive
potential of about 80 to 100 V with respect to the photocathode. The successive
dynodes are also kept at similar positive potentials with respect to the dynodes
immediately preceding them. The anode is at the same potential as the last
dynode.
When an energetic electron falls on the sensitive surface of a dynode,
secondary electrons are emitted from the latter by a process similar to the
process of photoelectron emission. The ratio of the number of electrons emitted
from the sensitive surface of a dynode for each electron striking it, is known as
the secondary electron emission coefficient. This is usually greater than unity,
which makes it possible to achieve a very large multiplication in the number of
electrons by the photomultiplier tube. If the secondary electron emission
coefficient is N and the number of stages of multiplication (i.e. the number of
dynodes) is n, the multiplication factor is Nn, provided there is no loss in the
electron number as they travel from one dynode to the next. With modern
photomultipliers using dynode materials e.g. BeO, MgO, Cs3Sb, the value of N
ranges from 5 to 10. Assuming N = 5 and n = 10, we get the multiplication
factor for the electron number
Nn = 510 = 107

This Photo by Unknown Author is licensed under CC BY-SA

Multiwire proportional counters (MWPC)


In MWPCs, the ionization produced by a charged particle in the gas of the
chamber (usually argon + 10% methane, or argon and isobutane), is amplified
by an ionization avalanche in the gas as the ions move towards the electrodes
under the influence of high electric fields. This device consists of many parallel
anode wires stretched in a plane between two cathode planes. The different
anode wires act as independent detectors. A typical structure has wires of 20 μm
diameter with 2 mm spacing between cathode planes 12 mm apart, operating at
a potential difference of 5 kV. In general, several electrons from the primary
ionization will drift an anode wire and create separate avalanches with negative
pulses with very fast rise times (~ 0.1 ns). The effective spatial resolution from
the anode pulses is of order 0.7 mm. (The geometry and electric field
distribution is shown in the figures below)

Cerenkov counter
When high energy charged particles traverse dielectric media, part of the light
emitted by excited atoms appears in the form of a coherent wavefront at fixed
angle with respect to the trajectory – a phenomenon known as the Cerenkov
effect, after its discoverer. Such a radiation is produced whenever the velocity v
= βc of the particle exceeds the phase velocity c/n of light in the medium (n
being the refractive index of the medium). From Huygens’ construction, we can
see from the figure below, that the wavefront forms the surface of a cone about
the trajectory as axis, such that
𝑐𝑡/𝑛 1 1
Cos θ = = ,β>𝑛
𝛽𝑐𝑡 𝛽𝑛

Wavefront
𝑐𝑡
𝑛

θ
Trajectory vt

Silicon microstrip detectors


These detectors are based on a silicon wafer, about 100 μm thick, with
surfaces of several cm2. A ladder of many n-p diodes is built on the surface of
the wafer in the shape of parallel strips with a pitch of few tens of μm. The
strips are the equivalent of the anode wires in an MWPC and are read out by
charge amplifiers. The device is reverse biased and fully depleted. A charged
particle produces electron-hole pairs that drift and are collected at the strips.
The spatial resolution is very good, of the order of 10 μm.

Particle accelerators
Cyclotron
In a cyclotron, the beam is bent in a circular path by a magnetic field, and the
particles orbit inside two semicircular metal chambers called ‘dees’ because of
their shape. The dees are connected to a source of alternating voltage. When the
particles are inside the dees, they feel no electric field and follow a circular path
under the influence of a magnetic field. In the gap between the dees, however,
the particles feel an accelerating voltage and gain a small energy each cycle.
The time it takes for a particle one semicircular path is independent of the
radius of the path – as the particles spiral to larger radii, they also gain energy
and move at a greater speed, and the gain in path length is compensated by the
increased speed. If the half-period of the ac voltage on the dees is set equal to
the semicircular orbital time, then the field alternates in exact synchronization
with the passage of particles through the gap, and the particle sees an
accelerating voltage each time it crosses the gap.

This Photo by Unknown Author is licensed


under CC BY
The Lorentz force in the circular orbit, qvB, provides the necessary centripetal
acceleration to maintain the circular motion at an instantaneous radius r, and
thus
𝑚𝑣 2
F = qvB = (1)
𝑟

The time necessary for a circular orbit is


𝜋𝑟 𝑚𝜋
t = = (2)
𝑣 𝑞𝐵

The frequency of the ac voltage is


1 𝑞𝐵
f = = (3)
2𝑡 2𝜋𝑚

The velocity increases gradually as the particle spirals outward, and the greatest
velocity occurs at the largest R
𝑞𝐵𝑅
vmax = (4)
𝑚

This leads to the maximum kinetic energy


1 2 𝑞 2 𝐵2 𝑅2
Tmax = m𝑣𝑚𝑎𝑥 = (5)
2 2𝑚

The first cyclotron built by Lawrence and used for nuclear physics research had
a magnet of diameter 37”(94 cm). Later, another cyclotron with a magnet of
diameter 60”(152 cm) was built under his supervision. This cyclotron could
produce deuterons of energy 20 MeV and α-particles of energy 40 MeV.
Limitations of cyclotron:
As the beam in a cyclotron travels outward toward the edge of the machine,
the magnetic field lines are diverted somewhat from the true vertical. There are
two effects of this fringing field, one beneficial and one harmful. The curvature
of the field lines gives a net force component toward the median plane, which
tends to provide focusing and to counteract the tendency of the beam to diverge.
At the same time, however, the field loses its uniformity and the resonance
condition can no longer be maintained of frequency is held constant.

There is a practical limit to the energy which can be attained in a fixed


frequency cyclotron due to the relativistic increase of ion mass. Since the mass
appears in the denominator in the expression for frequency, resonance condition
cannot be satisfied, as the velocity of the ions becomes high and the circulating
ions fall out of step with the applied rf voltage. Writing m = m0√(1 − 𝛽 2 ),
where β = v/c, eqn. (5) can be rewritten as
𝑞𝐵
f = √(1 − 𝛽 2 ) (6)
2𝜋𝑚0

To accelerate ions to relativistic energies by the cyclotron principle it is


necessary either to change the mode of acceleration or the distribution of
magnetic field in such a way that the resonance condition is obeyed in spite of
the relativistic increase of mass of the ion. Eqn.(6) shows that this is possible if
either
(1) The magnetic induction B is increased with time in such a way that the
factor √(1 − 𝛽 2 ) cancels out, the frequency of the rf field being held
constant
(2) If the frequency of the rf field is decreased with time in such a way that
the resonance condition (6) holds at all instants, the magnetic field B
remaining constant.
The first of these methods forms the basis of the working principle of
synchrotrons while the second method is applied in synchro-cyclotrons. Both
these machines utilize the principle of phase stability for their operation.
The principle of phase stability
This principle was discovered by E.M. McMillan and V.I. Veksler in 1945-
46. In a variable frequency cyclotron, a continuous beam is not possible, for the
time to travel the semicircular orbits will no longer be constant and equal to the
half-period (which is now variable). Thus the particles travel through the
cyclotron in bunches and the frequency is swept from its maximum value (when
the bunch is near the centre, the particles are only slightly accelerated and the
relativistic increase in mass is slight) to its minimum value (when the bunch is
ready to exit the cyclotron, the maximum energy is attained, and the mass has
its largest value). Particles in the bunches will arrive at the gap between the dees
at different times. Phase stability provides a sort of time-focusing effect, those
particles that arrive early are delayed somewhat, and on the next cycle, are
closer to the centre of the bunch, while those that arrive late get advanced and
are likewise pushed closer to the centre.
Consider a particle circulating at the centre of a bunch and arriving at the gap
at the instant the accelerating voltage passes through zero (see figure). Such a
particle would circulate forever in this stable orbit, called a synchronous orbit,
Now suppose another particle in the bunch arrives a bit earlier, at point b. This
particle sees an accelerating potential which will increase the energy and radius
of the orbit, but the mass will increase, thereby decreasing the orbital frequency.
Because its frequency is lower, it arrives at the next gap crossing later in phase,
that is, closer to the center of the bunch. Similarly, a particle arriving originally
later than the center of the bunch will be decelerated, and the decrease in mass
increases the angular frequency and pushes the particles closer to the center of
the bunch at the next gap crossing. Particles in a bunch therefore may perform
oscillations with respect to the synchronous orbit, moving first ahead, then
closer to, and then perhaps behind the particles in the synchronous orbit. In the
synchrocyclotron, as the frequency is slowly decreased, the radius of the
synchronous orbit will increase and with it the energy will increase.
For a typical synchrocyclotron, such as the 184-inch at Berkeley, the
cyclotron frequency range is from 36 to 18 MHz and the modulation rate is 64
Hz, the number of orbits is of order 105. The energy of the extracted protons is
740 MeV, and the field strength is about 2.3T. The Berkeley synchrocyclotron
was first operated in 1946, and provided a mean current of 0.1 μA.
Betatron
As the electron is very light, the relativistic increase of mass of the electron
becomes important at relatively much lower energies (> 10 keV). To
circumvent this difficulty, D.W. Kerst at the University of Illinois in USA,
devised a new instrument, known as the betatron (1940).
The velocity of electrons at 1 MeV or higher is close to the velocity of light, c.
So, with increasing energy, the electron velocity does not increase appreciably.
Only its mass increases. Thus the electrons at relativistic energies can be made
to revolve in orbits of constant radius by the action of an applied magnetic field
which varies with time. The acceleration of the electrons in the betatron takes
place due to changing magnetic flux through the area enclosed by the electron
trajectory. According to Faraday’s law, the changing magnetic flux B produces
an electric field E:
𝜕𝐵
∇×E = - (1)
𝜕𝑡

This induced electric field E accelerates the electrons which go on revolving


in an orbit of constant radius with a doughnut shaped accelerator tube. As the
electrons gain energy, their momentum p = mv (m = relativistic mass) increases
and hence the quantity BR must be increased in proportion by increasing the
magnetic induction B to make them rotate in an orbit of constant radius
determined by the relation
𝑝
BR = (2)
𝑒

Where p = mβc/√(1 − 𝛽 2 ) is the relativistic momentum.


Equation (1), on integration, gives the induced emf ε:
𝜕𝐵
ε = ∮ 𝐸. 𝑑𝑙 = ∫(∇ × 𝐸) . 𝑑𝑆 = -∫ 𝜕𝑡 .dS
𝜕𝜑
or, ε = - (3)
𝜕𝑡

where φ = ∫ 𝐵. 𝑑𝑆 is the magnetic flux, ε determines the energy gained per turn.
If R is the radius of the electron orbit, the induced emf is also given by
ε = 2𝜋RE
The force acting on the electron is
𝑒𝜀 𝑒 𝑑𝜑
F = eE = = (4)
2𝜋𝑅 2𝜋𝑅 𝑑𝑡

(neglecting -ve sign). Substituting from eqn.(2), we get,


𝑑𝑝 𝑑𝐵
F = = eR (5)
𝑑𝑡 𝑑𝑡

Comparing eqns. (4) and (5), we get,


𝑑𝜑 𝑑𝐵
= 2𝜋R2 (6)
𝑑𝑡 𝑑𝑡

Integration of eqn.(6) gives,


φ – φ0 = 2𝜋R2B (7)
For a uniform magnetic induction, on the other hand, the flux would be 𝜋R2B.
So for the betatron condition (6)the magnetic flux must be twice that for a
uniform magnetic induction. For this to be possible, the magnetic induction in
the central region of the area enclosed by the electron orbit must be higher than
that at the orbit itself. This is made possible by shaping the pole pieces suitably.
If <B> is the mean induction within the orbit, then the condition <B> = 2BR
must be satisfied during the acceleration. In addition, a few discs of high
permeability are placed between the pole pieces in the central region.

Electron synchrotron
The frequency of the rf field is unchanged in the synchrotron, the matching of
the frequency being achieved by changing the magnetic field with time. The
initial acceleration in the synchrotron is produced by betatron principle. The
changing magnetic field through the orbit of the electrons induces an electric
field along the orbit, which accelerates the electrons up to about 2 MeV. At this
point, the betatron action ceases and the electrons revolve in the orbit with
almost constant velocity v ≈ c. So the angular speed ω = c/R is also constant and
hence the orbit radius R remains practically constant.
The electrons are injected at fairly high speed from the electron gun so that
the orbit radius increases only marginally afterwards. After the initial betatron
acceleration, the electrons revolve in phase-stable orbits with the angular speed
ω = eB/m where m is the relativistic mass and gain energy due to acceleration
by the rf field within the accelerator tube. Since m increases with the gain of
Energy, B must be increased to maintain synchronism.
The main advantage of the synchrotron is that the magnet can be made much
lighter than in the betatron. The magnet used in the latter is of the iron-core type
since the total flux linked with the core is twice that for a uniform magnetic
field. In a synchrotron, no iron core is needed. The magnetic field is to act only
in the region where the doughnut shaped accelerator tube is placed, since the
primary function of the field is to keep the elctrons revolving in the phase-stable
orbit. The use of this type of magnet results in considerable saving in the cost of
the magnet since there is no iron in the core region, as in the case of the
betatron. The vertical section of the magnet has the shape of the letter ‘C’
shown in the figure. The pole-faces of the magnet are nearly parallel with slight
tapering outwards for focusing the beam. The betatron acceleration in the initial
stages is achieved with the help of a number of flux bars made of stel located in
the central region. At lower field strengths, the magnetization of the flux-bars
increases as the field increases, which induces the betatron accelerating field. At
higher field strength, their magnetization becomes saturated, so that the betatron
action stops.
The electrons revolve within a ring-shaped accelerator tube of toroidal section
(doughnut) made of glass or porcelain. A 70” section of this tube is silver-
plated, both inside and outside. There is a small gap (1/8” wide) in the plating
between which the rf field acts. As the electrons cross this gap, they are
accelerated by the rf field along the axis of the doughnut. The frequency of the
rf field is equal to the frequency of revolution of the electrons in the orbit. As
the electrons gain energy, their relativistic mass increases due to which they end
to move in orbits of larger radius. The increase of the magnetic field maintains
the orbit radius constant.
Proton synchrotron
Protons and other heavy ions can be accelerated by the synchrotron principle,
provided both the magnetic field and the rf frequency are varied to maintain the
resonance condition. This is necessary because, unlike in the case of electrons,
the velocity of the protons is not constant even upto several GeV. So the
frequency of revolution for the proton does not remain constant. There can be
resonance only if the frequency of the rf accelerating field is increased in unison
with the increase of revolution frequency. At the same time, the magnetic field
must be increased to keep the proton rotating in an orbit of fixed radius, in spite
of the increase of energy. Such an arrangement permits the use of an annular
magnet, thereby reducing the cost of the machine considerably.
Usually a ring-shaped magnet with a ‘C’ section is used. The C section can
face towards the outside or the inside of the machine. In some cases a ring
magnet with an ‘H’ section has been used. In case a, the field decelerates the
particle, while in b, it accelerates the particle. The core region is empty.

For magnets with C sections, the vacuum chamber (ring shaped) between the
poles is accessible around the entire outer periphery. In this case the magnetic
field at the orbit has a value of 1.4 T up to the point where the saturation effects
become significant. With the H section magnet, the maximum field of over 2.1
T can be used which results in a smaller radius for a given energy. However,
this severely restricts the accessibility to the vacuum chamber and hampers
research activities.
As the magnetic field is increased to keep the particles on track, there is some
gain of energy due to the betatron effect. This is however small compared to the
energy gain per turn by the particle in crossing the accelerating system.
The pole faces of the magnet are tapered so that the field decreases outwards.
The field index n lies between 0 and 1. This helps achieve orbit stability due to
betatron oscillations. In addition there are phase oscillations due to operation of
the principle of phase stability.
The choice of the magnet structure and the maximum field determines the
orbit radius for a given energy. If T be the kinetic energy, then for an induction
field B, the orbit radius R is given by
T(T + 2W0) = (BeRc)2 (1)
Where W0 = M0c2 is the particle rest energy. For T and W0 in GeV, and B in
Tesla, we get,
{𝑇(𝑇+2𝑊0 )}1/2
R = (2)
0.3𝐵
In the Cosmotron, the chosen energy T is 3 GeV. So for a maximum field of
1.4T, R = 9.1 m.
The magnet design, both at Brookhaven and Berkeley provides sectors in the
orbit free from the magnetic field. In both cases, the circular magnet is split into
four quadrants. The four sectors of the accelerating tube within these are joined
by four straight sections free from magnetic field. The straight sections have
been shown not to disturb orbit stability.
The four-field straight sections serve a number of purposes, such as housing
the rf accelerator, accommodation for the injector, the target, vacuum
manifolds, etc. The value of the field index n required in the case of orbits with
straight sections must lie between 0.5 to 0.8 for effective focusing of the ion
beam.

The particles injected into the proton synchrotron are pre-accelerated to a fairly
high energy. In the Cosmotron, this is done by a 4 MeV Van de Graff generator.
In the Bevatron, a 0.5 MeV Cockcroft-Walton generator feeds a 10 MeV proton
linac which was chosen as the injector.
Linear accelerator
A linear accelerator accelerates charged particles along a straight line in
multiple steps by an oscillating electric field. In all linear accelerators, the rf
field includes an e.m. wave whose phase velocity is equal to the velocity of the
accelerated particles. Since this latter velocity increases as the particle travels
through the accelerator, in the case of the atomic ions, the phase velocity of the
travelling wave must also increase with distance along the accelerating system.
Consider the principle of working of a linear accelerator for ions, such as
protons. The linear accelerator consists of a series of coaxial cylindrical drift
tubes along the axis of which the charged particles travel. One set of alternate
electrodes is connected to one terminal of the rf supply system while the other
set of alternate electrodes is connected to the other terminal. In a linear
accelerator for ions the successive electrode tubes are of gradually increasing
length.
Suppose the rf voltage between the first and second drift tube is near its
maximum negative value when a positive ion crosses the gap between them. As
a result the ions gain an amount of energy qV, where V is the amplitude of the
rf voltage and q is the charge of the ion. As the ions travel down the second drift
tube, they do not gain any energy because of the screening action of the tube. If
the frequency of the rf voltage is such that the phase changes by 𝜋 when the
ions emerge from the second drift tube into the gap between the second and
third tubes, they gain an additional amount of energy qV, so that their energy
becomes 2qV as they enter the third drift tube. Because of the increased energy,
the ions travel through the third tube with a higher constant velocity. The phase
of the accelerating voltage again changes by 𝜋 when the ions emerge into the
gap between the third and fourth drift tubes. So that they gain an energy qV
again as they cross the gap, and the total energy becomes 3qV. In this way, the
process of energy gain by qV continues at each successive gap crossing and the
ions finally emerge with an energy nqV after traversing the nth gap. The
velocity of the ions at this point is
2𝑛𝑞𝑉 1/2
vn = ( ) (1)
𝑀

Where M is the ion mass.


Since the ions travel with gradually increasing (constant) velocity through the
successive drift tubes, it is necessary to increase the length of the successive
tubes to maintain resonance condition. This condition is such that the time taken
by the ion to traverse through the drift tube at any stage must be equal to half
the period (T) of the accelerating voltage, i.e.
𝑇 𝐿𝑛 𝜆
= = (2)
2 𝑣𝑛 2𝑐

𝜆𝑣𝑛 𝜆 2𝑛𝑞𝑉 1/2


Ln = = ( ) (3)
2𝑐 2𝑐 𝑀

Upon crossing the gap, particles will experience a slight radial focusing (see
figure below). In the left half of the gap (region ab), the lines of force of the
electric field focus off-axis particles toward the axis, while in the region bc,
there is a defocusing effect. However, the acceleration of the particles means
they move more slowly, and thus spend more time, in the region ab so that the
focusing effect exceeds slightly the defocusing effect.
In their first linear accelerator, Lawrence and Sloan used rf voltage of
amplitude 42 kV at a frequency of 10 MHz. In 1947, L.W. Alvarez at the
University of California, Berkeley built a proton linear accelerator using 47 drift
tubes at an rf frequency of 202.5 MHz. The protons were initially accelerated to
4 MeV by a Van de Graaf generator and then injected into the first drift tube of
the accelerator. The final energy of the proton beam was 32 MeV.
Electrostatic accelerators
All electrostatic accelerators have two main components, viz., the HV
generator and the accelerating tube. Though different types of accelerators have
different HV generators, the accelerator tubes are the same in all of these. The
accelerator tube is made up of a series of cylinders of insulating materials (e.g.
glass) with a total length of a fraction of a meter to a few meters. A series of
metallic electrodes are placed within the cylinders with vacuum tight seals
coming out through the insulating cylinders. These are connected to a chain of
resistors with very high resistances to provide a uniform potential distribution
from one end of the tube to the other. The electrodes within the cylinders
protect the walls of the latter from the beam and have a focusing action on the
beam.
Two electrodes are located at the two ends of the accelerator tube kept in a
chamber which is highly evacuated. The target to be bombarded by the charged
particle beam forms one of the electrodes which is kept at a ground potential.
The other electrode is connected to the high voltage terminal of the HV
generator and the ion source is placed within it.
If V is the voltage generated, then a particle of charge q gains akinetic energy
qV on going through the tube. If q = +e, the electronic charge, then the energy is
V in MeV, the voltage being given in a million volts.
The limit of the voltage generated is set by the discharges between the two
electrodes outside the tube or between the HV terminals and ground or the walls
of the chamber accommodating the accelerator tube. To avoid such breakdown,
the accelerator tube is made as long as possible and kept as far away from the
chamber walls as possible. In practice, the upper limit of the voltage is only a
few million volts which can be increased to some extent by enclosing the
machine within a closed tank containing some dry insulating gas, such as a
mixture of CO2 and N2 at pressures up to ~ 15 atmospheres. Pure Sulphur
hexafluoride (SF6) gas is also used for this purpose. The use of a pressure tank
allows the machine to be housed in a much smaller space than would be
possible in air at atmospheric pressure.
Van de Graaf accelerator
There are two pulleys P1 and P2 over which an insulating belt is made to move
with the help of a motor connected to pulley P2 which is kept at ground
potential. Near this pulley, a row of metal points extends across the width of the
belt maintained at a potential of 20-30 kV.
Assume the points to be kept at a positive potential. A corona discharge
between the points and the belt ionizes the gas which helps to spray positive
charge on the moving belt.
The upper pulley P1 is inside a large hollow metal sphere S. Here another set
of sharp corona points connected to S removes the charge from the belt to the
latter. This charge gradually accommodates on the outer surface of S. Its
potential thus increases gradually. Because of the large size of the sphere S, it
can hold a large quantity of charge and its potential can rise to a very high
value.
The ion source and the top end of the accelerator tube are located inside the
HV terminal S.
The convective charging current is determined by the characteristics of the
belt and is given by I = σvw where σ is the charge density on the belt which has
a width w and moves with a velocity v. Charging currents up to several mA can
be produced which can be raised in various ways, e.g. by using a number of
parallel belts or by charging the belt during its downward motion with opposite
charge.
Van de Graaf generators can reach higher voltages than the Cockcroft-Walton
generators and above about 2 MV, these are the only practical electrostatic
generators.
Pelletron accelerators
In a pelletron, the charging belt is replaced by a charging chain, consisting of
steel cylinders joined by links of solid insulating material such as nylon. The
metal cylinders are charged as they leave a pulley at the ground potential and
the charge is removed as they pass over a pulley within a high voltage terminal.
Pelletrons range in the maximum operating voltage from 200 kV to 25 MV.
Beam current capabilities range from a few microamperes to ~ 0.8 mA.
Tandem accelerators
In these accelerators, singly charged positive ions are accelerated to twice the
energy corresponding to the energy of the HV terminal. The accelerating tube
crosses the HV terminal so that the ion source and the target at the opposite
ends of it are both at the ground potential.

The positive ions from the ion source passes through an electron-adding canal
where a flow of hydrogen adds two electrons successively to them, so that they
are transformed into negative ions. These are then accelerated to the positive
HV terminal. Here the beam passes through an electron stripper, in which
collisions with some neutral gas atoms reconverts the high energy negative ions
into a beam of positive ions, without change of energy. This positive ion beam,
in travelling down the remaining portion of the accelerator tube, is further
accelerated through an equal voltage when it reaches the target, so that the final
energy is twice the value corresponding to the voltage produced in the machine.
The accelerated beam intensity is rather low in these machines, being of the
order of a few microamperes. However, the energy available is quite high (~ 12
to 20 MeV) for singly charged ions.
Accelerators in India
(1) The biggest accelerator in India, which is in operation, is the 224 cm
sector-focused (AVF type) cyclotron at the Variable Energy Cyclotron
Center (VECC) at Bidhannagar, Kolkata. It is designed to deliver proton
beams of 6 to 60 MeV, deuteron beam of 12 to 65 MeV and α-particle
beam of 24 to 130 MeV. At present it gives α-particles of 80 MeV
energy. The beam current is a few microamperes. The magnet, which
weighs 262 tonnes, produces the highest field 2.1 T. Three spiral ridges
are used for azimuthal variation of the field. The rf frequency can be
changed from 5.5 to 16.5 MHz. The magnetic field is increased outwards
with the help of 17 auxiliary coils.
(2) A K-14 type pelletron has been set up at TIFR in Mumbai (1986).
According to definition, the energy available is E = kq2/A where q is the
charge and A is the mass number of the accelerated ion. This machine has
been purchased from High Voltage Corporation, USA. It gives 28 MeV
protons, 42 MeV α-particles, and correspondingly higher energies for
heavier ions. It is used for heavy ion reaction studies.
(3) An almost similar pelletron (K-15) giving the energy of 8 to 30 MeV per
nucleon, obtained from the same source, has been set up (1990) at the
Nuclear Science Center at Delhi under the aegis of the UGC. It is used for
heavy ion reaction studies.
(4) A number of electrostatic accelerators of lower energy are in operation in
different laboratories in India. These include the 5.5 MeV Van de Graaf
accelerator at TIFR, a 2 MeV Tandem Van de Graaf at BARC (Mumbai)
and similar machines at IGCAR, Kalpakkam, Tamil Nadu, IOP,
Bhubaneswar and IIT Kanpur.
(5) There is a synchrotron radiation facility at CAT, Indore. There is a 700
MeV electron synchrotron with a 20 MeV electron beam from a
microtron as injector.
Accelerators abroad
The first proton synchrotron was the ‘Cosmotron’ at Brookhaven National
Laboratory (USA), completed in 1952 and designed to produce protons at an
energy of 3 GeV. The mean radius of the orbit was about 10 m and the
maximum field strength was 1.4 T. Injection energy was 3.5 MeV and the ac
oscillator frequency varied from 0.37 to 4 MHz during the acceleration. A
competing proton synchrotron was built at the same time at the Lawrence
Radiation Laboratory at Berkeley, with a slightly larger radius (18 m) and field
strength (1.6 T). The Berkeley machine was called ‘Bevatron’ (BeV ≡ GeV); it
was completed in 1954 with a design energy of 6.4 GeV.
By 1960, two machines were in operation, the alternating gradient
synchrotron (AGS) at Brookhaven, and the CERN proton synchrotron (CERN is
the European Center for Nuclear research). The AGS reached an energy of 33
GeV, following injection at 50 MeV.
At Fermilab in Batavia, Illinois, is a 500 GeV separated function proton
synchrotron with an orbital radius of 1000 m. The accelerator began operations
in 1972 at an energy of 200 GeV. Injection at 8 GeV is provided by a sequence
of three accelerators, a 0.8 MeV Cockcroft-Walton, followed by a 200 MeV
drift tube linear accelerator and then an 8 GeV “booster” synchrotron. The peak
field in the bending magnets is 1.4 T and the resonant frequency in the ac
cavities is 53 MHz.
A note on sector focused AVF cyclotrons:
The method of maintaining the resonance condition at relativistic energies
involves the radial increase of the magnetic field. This however conflicts with
the requirement of focusing the ion beam for which the magnetic field should
decrease at larger radii. Thomas proposed the use of alternately high and low
regions of the magnetic field around the orbit which could be produced by
radial sectors of iron (ridges to the pole faces so that the pole gaps were
alternately wide and narrow. The average field around the orbit could then be
increased radially outwards to maintain resonance.
Such a cyclotron is also called an AVF (azimuthally varying field) cyclotron.
The stable orbits in an AVF cyclotron are not circles, the particles perform
radial oscillations about the circular orbit. At the boundaries between the high-
field and low-field sectors, there is an azimuthal component to the field and the
Lorentz force gives a vertical force that tends to keep the beam in focus in the
midplane.

The study of cosmic rays led to the birth of the subject of Particle Physics.
Many particles, e.g. mesons, hyperons, other baryons, were discovered in
cosmic rays. Many more were discovered subsequently in particle accelerator
experiments.
Hadrons

Baryons Mesons
Leptons (electron, muon, tau and their associated neutrinos)
Baryons have baryon number B = 1. Antibaryons (e.g. the antiproton) have
baryon number -1.
Lepton number Le (for electrons): Le = +1 for e-, νe, Le = -1 for e+, 𝜈̅𝑒 , Le = 0 for
others.
Lepton number Lτ (for τ lepton) can be similarly defined.
̅̅̅,
Lepton number Lμ (for muons): Lμ = +1 for μ-, νμ, Lμ = -1 for μ+,𝜈 𝜇 Lμ = 0 for
others.
An example of a decay:
μ+ → e+ + νe + ̅̅̅
𝜈𝜇
Lμ = -1 0 0 -1
Le = 0 -1 +1 0
Masses of leptons:
Mass (MeV) Lifetime
e- 0.511 stable
μ- 105.6 2.2 μs
τ- 1777 0.29 ps

Isospin
The charge independence of nuclear forces means that in most instances, we
do not need to distinguish in the formalism between neutrons and protons, and
this leads us to group them together as members of a common family, the
nucleon. The formalism for nuclear interactions may depend on the multiplicity
of nucleon states (two) but it is independent of whether the nucleons are protons
or neutrons. The exception of course, is the electromagnetic interaction, which
can distinguish between protons and neutrons. With respect to the strong
nuclear force alone, the symmetry between neutrons and protons remains valid.
This two-state degeneracy leads naturally to a formalism analogous to that of
1
the magnetic interaction of a spin- particle. The neutron and proton are treated
2
as two different states of a single particle, the nucleon. The nucleon is assigned
a fictitious spin vector, called the isospin. (This idea was introduced by
Heisenberg in 1932.) The two degenerate nuclear states of the nucleon in
absence of electromagnetic fields, like the two degenerate spin states of a
nucleon in absence of a magnetic field, are then “isospin-up”, which we
arbitrarily assign to the proton, and “isospin-down”, the neutron. That is, for a
1 1
nucleon with isospin quantum number I = 2 , a proton has I3 = +2 and a neutron
1
has I3 = -2 . These projections are measured with respect to an arbitrary axis
called the “3-axis” in a coordinate system whose axes are labelled 1,2,3. The
isospin obeys the usual rules for angular momentum vectors. Thus we use an
isospin vector I of length √𝐼(𝐼 + 1) ћ.
The concept of isospin is not only useful in classifying hadrons, but also in
constraining their dynamics in scattering and decay processes. If these proceed
through strong interactions, both the total isospin and its third component are
conserved; if they proceed through electromagnetic interactions, only the third
component is conserved; while if they proceed through weak interactions,
neither is conserved.
Cross section
Consider a reaction of the form
a+b→c+d (1)
with two particles in the initial state i and two in the final state f, If we regard b
as the target, and a as the projectile particle – usually in a well-collimated beam
– then the cross section for the above reaction is defined as the transition rate W
per unit incident flux per target particle. If na is the density of particles in the
incident beam, and vi the relative velocity of a and b, then the flux of particles
per unit time through unit area normal to the beam is
φ = navi (2)
If there are nb particles in the target per unit area, each of effective cross section
σ, the probability that any incident particle will hit a target is σnb, the number of
interactions per unit area per second will be nanbσvi. The transition rate per
target particle is therefore
W = σφ = σnavi (3)
The unit of cross section is barns. 1 barn = 10-28 m2.
The Breit-Wigner formula for elastic scattering cross section
𝛤2 /4
σei = 4𝜋λ2 (2ℓ + 1) (𝐸− 𝐸 2 + 𝛤 2 /4 (4)
𝑅)
The width Γ is defined so that the elastic cross section σel falls by a factor of 2
from the peak value when |E - ER| = + Γ/2.
Interactions and conservation laws
Interaction Gravity Electromagnetic Weak Strong

Field quantum Graviton Photon W+, W-, Z0 Gluon


Spin, Parity 2+ 1- 1-, 1+ 1-
Mass (GeV) 0 0 80 - 90 0
Typical cross 10-33 10-44 10-30
section,
m2(GeV)
Typical decay 10-20 10-8 10-23
lifetime, s

Conserved Interaction
Quantity Strong Electromagnetic Weak
Energy/momentum Yes Yes Yes
Charge Yes Yes Yes
Baryon number Yes Yes Yes
Lepton number Yes Yes Yes
Isospin Yes No No (ΔI = 1 or ½)
Strangeness Yes Yes No (ΔS = 1,0)
Parity Yes Yes No

Decay of pions: 𝜋- → μ- + ̅̅̅,


𝜈𝜇 𝜋+ → μ+ + νμ (lifetime = 26 ns)
𝜋0 → 2γ (lifetime = 0.8 × 10-16s)
The Λ0 hyperon is produced via the strong interaction: 𝜋- + p → Λ0 + K0
The Λ0 hyperon decays via the weak interaction: Λ0 → 𝜋- + p
The Λ0 hyperon has strangeness S = -1, the K0 meson has strangeness S = +1
The Σ0 hyperon, which has S = -1, decays via the electromagnetic interaction Σ0
→ Λ0 + γ. There are two other (charged) Σ hyperons, Σ+ and Σ-. Both have S = -
1. These three are members of an isospin triplet, corresponding to I = 1. The Σ
hyperon can be produced via the reaction: 𝜋- + p → Σ- + K+, 𝜋+ + p → Σ+ + K+.
The Σ+ decays weakly: Σ+ → n + 𝜋+.
The following relation was proposed by Gell-Mann and Nishijima:
𝑄 𝐵+𝑆
= + I3 (1)
𝑒 2

The cascade or xi, Ξ hyperon, was observed in cosmic ray studies in cloud
chambers and was assigned S = -2, for it was produced along with a pair of K
̅̅̅̅0 , and it decayed weakly:
mesons: K- + p → Ξ- + 𝜋+ + K0 + K0 + 𝐾
Ξ- → n + 𝜋 -
The existence of the Ω- baryon with S = -3, as well as its mass and decay modes
was predicted before it was observed in 1964, on the basis of unitary symmetry
and the quark model.

Masses of metastable strange hyperons:


S Mass (MeV) Lifetime (ps)
Λ -1 1116 263
Σ+ -1 1189 80
Σ0 -1 1193 7.4 × 10-8
Σ- -1 1197 148
Ξ0 -2 1315 87
Ξ- -2 1321 49

Masses of mesons:
S Mass (MeV) Lifetime
K+ +1 494 12 ps
K0 +1 (498) n.a.
K- -1 494 12 ps
̅̅̅̅
𝐾0 -1 (498) n.a.
𝜋+,𝜋- 0 139.6 26 ns
𝜋0 0 135 0.84 × 10-16 s
Parity
Pψ(r) = ψ(-r), P2ψ(r) = Pψ(-r) = ψ(r), so P2 = 1.
If ψ(-r) = ψ(r), the state has even parity
If ψ(-r) = - ψ(r), the state has odd parity
For spherical polar coordinates (r,θ,φ), the substitution r → -r implies r → r, θ
→ 𝜋 – θ, φ → 𝜋 + φ.
Consider spherical harmonics 𝑌𝑚𝑙 (θ,φ). Up to a constant, this is determined by
𝑃𝑚𝑙 (cos θ)eimφ. So under the transformation θ → 𝜋 – θ, φ → 𝜋 + φ, we have,
eimφ → eim(ℼ + φ) → (-1)meimφ
𝑃𝑙𝑚 (cos θ) → 𝑃𝑙𝑚 {cos(ℼ - θ)} = (-1)l+m 𝑃𝑙𝑚 (cos θ)
Thus, 𝑌𝑙𝑚 (θ,φ) → 𝑌𝑙𝑚 (ℼ - θ, ℼ + φ) = (−1)𝑙 𝑌𝑙𝑚 (θ,φ)
So the spherical harmonics have parity (−1)𝑙

Effect of P on some physical quantities:


Quantity Effect of P
r -r
p -p
σ (spin) σ
E(electric field) -E
B(magnetic field) B

Intrinsic parity
A single particle can be, but is not necessarily, in an eigenstate of P only if it
is at rest. The eigenvalue of P in this frame is called intrinsic parity which can
be positive or negative. Fermions have half-integer spin and angular momentum
conservation requires them to be produced in pairs. Therefore only relative
parities can be defined. Conventionally proton parity is assumed positive and
parities of other fermions are given relative to the proton.
Determination of parity of the pion
The parity of the 𝜋- is determined by observing its capture at rest by
deuterium nuclei.
𝜋- + d → n + n
In practice, one brings a 𝜋- beam of low energy into a liquid deuterium target.
The energy is so low that large fractions of pions come to rest in the liquid after
having suffered ionization energy loss. Once a 𝜋- is at rest, the following
processes take place. Since they are negative, the pions are captured, within a
time lag of a few picoseconds, in an atomic orbit, replacing an electron. The
system is called a ‘mesic atom’. The initial orbit has high values of both the
quantum numbers n and l, but again very quickly (1 ps), the pion reaches a
principal quantum number n of about 7. At these values of n the wave function
of those pions that are in S orbit largely overlaps with the nucleus. In other
words, the probability of the 𝜋- being inside the nucleus is large, and they are
absorbed.
The deuteron has spin 1 and the pion has spin 0. The orbital momentum ℓ = 0.
Therefore, initially, S = 0, L = 0, and thus, J = L + S = 1. Finally also, we must
have J = 1.
The deuterium nucleus contains two nucleons, of positive intrinsic parity, in an
S wave. Hence its parity is positive. In conclusion, its initial parity is equal to
that of the pion.
The final state consists of two identical fermions and must be antisymmetric
in their exchange. If the two neutrons are in a spin singlet state, which is
antisymmetric in the spin exchange, the orbital momentum must be even, and
vice versa, if the neutrons are in a triplet. The singlet state has S = 0 and the
triplet state has S = 1. The overall symmetry under interchange of the two
identical neutrons will be (-1)L+S+1, and this must be negative. So, L + S must be
even. The overall angular momentum J = 1 requires either L = 0, S = 1, or, L =
1, S = 0 or 1, or, L = 2, S = 1. Of these, only L = S = 1 has L + S even. So the
only possibility is the 3P1 state of two neutrons, with parity (-1)L = -1, which
requires negative parity also for the initial state. Thus the pion has odd parity.
The discovery of the neutrino: Reines and Cowans’ experiment:
The most intense source of neutrinos on earth are fission reactors. They
produce electron antineutrinos with a continuum energy spectrum up to several
MeV (on the average, 6 per fission). The flux is proportional to the reactor
power. The power of the Savannah river reactor in South Carolina (USA) was
0.7 GW. The 𝜈̅𝑒 flux was about Φ = 1017 m-2 s-1.
Electron antineutrinos can be detected by the inverse beta process but its cross
section is extremely small
𝜈̅𝑒 + p → n + e+ (1)
The cross section σ is for this process is given by
𝐸
𝜈
σ ≈ 10-47 (𝑀𝑒𝑉 )2 m2 (2)

an easily available material containing many protons is water. Let us calculate


in order of magnitude the quantity of water needed to have, for example, a rate,
10-3 Hz for the reaction (1) on free protons. Taking a typical energy Eν = 1
MeV, the rate per target proton is
W = Φσ = 1017 × 10-47 = 10-30 s-1.
Consequently, we need 1027 protons. Since a mole of H2O contains 2NA ≅ 1024
protons, we need 1000 moles, hence 18 kg. In practice, much more is needed,
taking all inefficiencies into account. Reines worked with 200 kg of water. The
reaction (1) was observed, using a target of cadmium chloride (CdCl 2) and
water. The positron produced in this reaction rapidly comes to rest by ionization
loss and forms positronium, which annihilates to γ-rays, in turn producing fast
electrons by the Compton effect. The electrons are recorded in a liquid
scintillation counter. The time scale for this process is of order 10-9 s, so the
positron gives a so-called prompt pulse. The function of the cadmium is to
capture the neutron after it has been moderated (i.e. reduced to thermal energy
by successive elastic collisions with protons) in water – a process which delays
by several microseconds the γ-rays coming from eventual radiative capture of
the neutron in cadmium. Thus the signature of an event consists of two pulses
microseconds apart.
The main difficulty of the experiment is not the rate but the discrimination of
the signal from the possibly more frequent background sources that can
simulate the signal. There are three principal causes: the neutrons that are to be
found everywhere near a reactor, cosmic rays and the natural radioactivity of
material surrounding the detector and in the water itself.
1
Baryon octet (JP = 2 + baryons)

n(udd), p (uud), Σ-(dds, 1197), Σ0(1192), Λ(uds), Σ+(uus, 1189), Ξ-(dss,


1321),Ξ0(uss, 1315)
3
Baryon decuplet (JP = 2 + baryons)

Δ-(ddd), Δ0(udd), Δ+(uud), Δ++(uuu), [Δ(1232)] Σ-(1387), Σ0(1384), Σ+(1383), Ξ-


(1535), Ξ0(1532),Ω-(1672)

Pseudoscalar mesons (JP = 0-)


K0(d𝑠̅, 498), K+(u𝑠̅ ,494), 𝜋-(d𝑢̅, 140), 𝜋0(135), η(548), η’(958), K-(s𝑢̅, 494),
̅̅̅̅
𝐾 0 (s𝑑̅ , 498)
Vector mesons (JP = 1-)
K*0, K*+(892), ρ(d𝑢̅, 776), φ(1020), ω(782), ρ0 ρ+(u𝑑, ), K*-, ̅̅̅̅̅
𝐾 ∗0 (892)

Gell-Mann Okubo mass formula for baryons:


𝑌2
M = M0 + M1Y + M2[I(I + 1) - ]
4

Or, M = a + bY, since Y = B + S = 2(I – 1)


For the baryon octet, the above equation predicts that
3MΛ + MΣ = 2MN - MΞ
The values are 4541 MeV and 4514 MeV.
Quark Model (Gell-Mann & Zweig, 1964)
All baryons are composed of three quarks. All mesons are composed of a quark
and an antiquark. Quarks carry fractional electric charge. The u quark has
2 1 1
charge + e, the d quark has charge - e, and the s quark has charge - e.
3 3 3
1
Quarks have baryon number 3 . Color is a new degree of freedom. All quarks
carry color charge. Color charge comes in three varieties: red, blue and green.
The theory of strong interactions is Quantum Chromodynamics (QCD). The
strong interactions are mediated by gluons. Gluons are massless and carry color
charge.
The fourth quark
In 1974, S. Ting and collaborators at the AGS proton accelerator at
Brookhaven built a spectrometer designed for the search for very heavy
particles with JPC = 1--. With the assumed quantum numbers, the particle, which
was to be called J, would decay into e+ and e-. The overall reaction to search for,
calling the undetected part of the final state X, is
p + N → J + X → e + + e- + X
The mass m of the e+e- system can be calculated. This m (e+e-) distribution
shows a spectacular resonance. Its mass is m J = 3100 MeV and has the
outstanding feature of being extremely narrow.
At the same time, the SPEAR e+e- collider, built by B. Richter and
collaborators, was operational at SLAC. When the energy was varied in small
steps about 3 GeV, a huge resonance appeared with all cross sections jumping
more than two orders of magnitude. The resonance was very high and extremely
narrow. Since the discovery at SLAC was independent, a different name was
given to the new particle, ψ. It is now called J/ψ.
After the discovery of the ψ, a systematic search started at SPEAR, scanning
the energy at very small steps. A second narrow resonance was found, which
was called ψ’. The quantum numbers of the ψ’ are JPC = 1--.
From the line shape of the resonance, one can extract accurate values of the
mass and width. The values obtained at SLAC were
m (ψ) = 3097 MeV, Γ (ψ) = 91 keV
m (ψ’) = 3686 MeV, Γ (ψ’) = 281 keV
The extremely small widths show that J/ψ and ψ’ are hadronic states of a
completely new type, they are called ‘hidden charm’ states. Both are made of a
c𝑐̅ pair in a 3S1 configuration. The former is the fundamental level (1 3S1) and
the latter is the first radial excited level (2 3S1 ). The charmed (c ) quark has
2
charge 3 e and mass about 1.25 GeV. It has C = +1. The meson D+ has the quark
content c𝑑̅ , with JP = 0- has mass 1869.3 MeV and lifetime 1.04 ps. The meson
D0 has quark content c𝑢̅, with JP = 0- has mass 1864.5 MeV and lifetime 0.4 ps.
The meson 𝐷𝑠+ has quark content c𝑠̅, with JP = 0- has mass 1968.2 MeV and
lifetime 0.5 ps.
The bottom (b) quark
In 1977, Lederman and collaborators built a two-arm spectrometer, at
Fermilab designed to study μ+μ- pairs produced by high energy hadronic
collisions. The reaction studied was
p + (Cu, Pt) → μ+ + μ- + X
The 400 GeV proton beam extracted from the Tevatron was aimed at a metal
target made of copper or platinum. The m(μ+,μ-) mass distribution, after
subtracting a continuum background, shows three barely resolved resonances,
which were generically called ϒ.
The precision study of the new resonances was made at the e+e- colliders at
DESY (Hamburg) and at Cornell in the USA. The measurement of the mass and
the widths of the ϒs was made. The situation is very similar to that of the ψs,
now with three vey narrow resonances, all with JPC = 1--. They are interpreted as
the states 3S1 of the b𝑏̅ ‘atom’, the bottomium, with increasing principle
quantum number. The bottom quark has a mass of about 4.2 GeV and charge of
1
- 3 e. It has B = -1

State Quarks M(MeV) Γ/τ JPC


ϒ (13S1) b𝑏̅ 9460 54 keV 1--
ϒ (23S1) b𝑏̅ 10023 32 keV 1--
B+ u𝑏̅ 5279 1.6 ps 0-
B0 d𝑏̅ 5279 1.5 ps 0-

The discovery of the top quark


The top quark was discovered at the Tevatron Collider in Fermilab in 1995.
At the Tevatron top production was a very rare event, it happened once in 10 10
collisions. Experimentally, one detects the top by observing its decay products.
The top decays most probably into final states containing a W boson and a b
quark or antiquark. Therefore, one searches for the processes
p + 𝑝̅ → t + 𝑡̅ + X
t → W+ + b
𝑡̅ → W- + 𝑏̅
The W boson, the mediator of the weak interactions, has a mass of 80 GeV and
a very short lifetime. It does not leave an observable track and must be detected
by observing its daughters. The W decays most frequently into a quark-
antiquark pair, but these decays are difficult to distinguish from the much more
common events with quarks directly produced by the proton-antiproton
annihilation. We must search for rare but clearer cases, such as those in which
both W bosons decay into leptons: W → e νe or, W → μ νμ
Another clean channel occurs when one decays into a lepton and the other
into a quark-antiquark pair, requiring the presence of a b and a 𝑏̅, from the t and
𝑡̅ decays. Namely, one searches for the following sequence of processes
p + 𝑝̅ → t + 𝑡̅ + X
t → W+ + b → W+ + jet (b)
𝑡̅ → W- + 𝑏̅ → W- + jet (𝑏̅ )
W → e νe or W → μ νμ
The requested ‘topology’ must have one electron or one muon, one neutrino,
four hadronic jets, two of which contain a bottom particle.
The mass of the top quark was measured to be 174.2 + 3.3 GeV
Natural units
We take the unit of time as the second.
We take the unit of length such that c = 1. Therefore, in dimensional equations,
we have [L] = [T].
We now define the unit of mass in such a way that ћ = 1. Mass energy and
momentum have the same dimensions. [M] = [P] = [E] = [L-1]
Luminosity of a collider
The reaction rate R is given by
R = σL (1)
Where σ is the cross section and L is the luminosity (in units of cm -2s-1).
For two oppositely directed beams of relativistic particles,
𝑓𝑛𝑁1 𝑁2
L = (2)
𝐴
Where N1 and N2 are the numbers of particles in each bunch, n is the number
of bunches in either beam around the ring, and A is the cross sectional area of
the beams, assuming them to overlap completely, f is the revolution frequency.
Examples:
Which of the following reactions are allowed by conservation laws and which
are forbidden? Give reasons in either case.
(1) 𝜋0 → e+ + e-
(2) K+ + n → Σ+ + 𝜋0
Solutions: (1) Allowed
(2)Initially, B = 1, finally B = 1. Charge is conserved. Initially, S = 1,
finally, S = -1. So there is a change of S by 2, ΔS = -2. Not allowed.
For each of these reactions, establish whether they are allowed or not, giving
reasons. Establish the type of reactions that allow it.
(3) 𝜋- + p → 𝜋0 + n
(4) 𝑝 + 𝑝̅ → Λ + Λ
Solutions: (3) Allowed
(4)Charge is conserved. Initially, B = 0, finally, B = 2. Initially, S =
0, finally, S = -2. Forbidden.
Cross section ( a short calculation)
Consider the process: a + b → c + d
Transition rate W = σnavi (1)
2𝜋
Formula for W: W = |Mif|2 ρf (2)
ћ

Mif can be interpreted as overlap integral over volume, Mif = ∫ 𝜓𝑓∗ 𝐻′𝜓𝑖 𝑑𝜏.
Here ρf is a density of states, or phase space factor. To calculate the phase space
factor, we restrict the particles to an arbitrary volume, which cancels in the
calculation and which we shall take as unity.
𝑑𝑛 𝑑𝛺 𝑑𝑝𝑓
Then ρf = 𝑑𝐸 = 𝑝𝑓2 gf (3)
0 ℎ3 𝑑𝐸0

Here E0 is the total energy in the CM frame, pf is the final-state momentum in


this frame ( |pc| = |pd| = pf) and dΩ is the solid angle containing the final-state
particles; gf is a spin multiplicity factor. With na = 1, we have, from eqn. (1)
𝑊 𝑝𝑓2 𝑑𝑝𝑓
σ = 𝑣 = |Mif |2 gf (4)
𝑖 𝑣𝑖 𝑑𝐸0

Numerical constants have been absorbed in |Miif|2.


Conservation of energy gives: Ec + Ed = E0

Thus, √𝑝𝑓2 + 𝑚𝑐2 + √𝑝𝑓2 + 𝑚𝑑2 = E0 (5)

𝑝𝑓 𝑝𝑓 𝑑𝑝𝑓
So, ( 𝐸 + 𝐸 ) 𝑑𝐸 = 1
𝑐 𝑑 0

𝐸𝑐 + 𝐸𝑑 𝑑𝑝𝑓
In other words, pf ( ) 𝑑𝐸 = 1
𝐸𝑐 𝐸𝑑 0

𝐸0 𝑑𝑝𝑓
Or, pf =1
𝐸𝑐 𝐸𝑑 𝑑𝐸0

𝑑𝑝𝑓 𝐸𝑐 𝐸𝑑
Thus, = (6)
𝑑𝐸0 𝐸0 𝑝𝑓

𝑑𝑝𝑓
Now, we can also write, from above, (vc + vd) 𝑑𝐸 = 1
0

Again, vc + vd = vf , the relative velocity of c and d.


𝑑𝑝𝑓 1
So, 𝑑𝐸 = 𝑣 (7)
0 𝑓

If sc and sd are the spins of the particles c and d, then the number of substates for
them are (2sc + 1)(2sd + 1), and thus gf = (2sc + 1)(2sd + 1)
(2𝑠𝑐 +1)(2𝑠𝑑 +1)
Then σ = |Mif|2 𝑝𝑓2 (8)
𝑣𝑖 𝑣𝑓

Implied in this expression, integrated over angles, is that |Mif|2 has been
averaged over all spin states of and b and summed over all orbital angular
momentum states involved.

Examples of interactions:
Σ0 → Λ + γ, τ = 10-19 s, electromagnetic decay
n → p + e- + 𝜈̅𝑒 weak decay
π- + p → π0 + n strong interaction
The neutral, massless carrier of the strong force is called the gluon. The theory
of strong interactions between quarks is called Quantum Chromodynamics
(QCD). The quarks carry color charge. There are three possible values of the
color charge – red, blue and green. Antiquarks carry anticolor. Gluons also
carry color charge. The quark-antiquark force is independent of the color
involved. The potential between two quarks is taken to be of the form
4 𝛼𝑠
V = -3 + kr
𝑟

Examples of decay of K mesons:


K+ → 𝜋0 + e+ + νe
K- → 𝜋0 + e- + 𝜈̅𝑒

You might also like