Sunteți pe pagina 1din 31

Exposition on Particle Calorimetry

Michael Kossin
Fall 2010

Contents
I

Introduction

1 Particles and their classes

II

Showers

2 Electromagnetic showers
2.1 Electron and positron collision events and energy loss mechanisms
2.1.1 High energy . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.2 Low energy . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.3 Critical energy . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Photon collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1 Pair production . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2 Mean free path . . . . . . . . . . . . . . . . . . . . . . . . .
2.3 Electromagnetic shower propagation . . . . . . . . . . . . . . . . .
2.3.1 Shower maximum . . . . . . . . . . . . . . . . . . . . . . .
2.3.2 Energy deposit as a function of depth . . . . . . . . . . . .
2.3.3 Moli`ere radius . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.4 Shower profiles . . . . . . . . . . . . . . . . . . . . . . . . .
3 Hadronic showers
3.1 Electron collisions . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Spallation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Hadronic shower components . . . . . . . . . . . . . . . . . .
3.4 Spallation cross sections . . . . . . . . . . . . . . . . . . . . .
3.5 Consider the neutron . . . . . . . . . . . . . . . . . . . . . . .
3.6 Invisible energy . . . . . . . . . . . . . . . . . . . . . . . . . .
3.7 Hadronic shower profiles and their relation to electromagnetic

III

Calorimetry in general

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

5
5
5
5
6
7
7
10
10
10
11
11
12

. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
shower profiles

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

12
12
14
14
15
15
15
16

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

16

4 Measuring end results


16
4.1 Recording scintillation events . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5 Homogeneous and sampling calorimeters and energy response
18
5.1 Energy response of homogeneous calorimeters . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.2 Energy response of sampling calorimeters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1

6 Energy resolution
19
6.1 Sampling fluctuations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6.2 Fluctuations due to invisible energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
7 Compensation
20
7.1 Decreasing electromagnetic shower energy response . . . . . . . . . . . . . . . . . . . . . . . . 20
7.2 Increasing hadronic shower energy response . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

IV

Calorimeters at the Large Hadron Collider

21

8 The Large Hadron Collider

21

9 Geometry

23

10 The Compact Muon Solenoid


10.1 Electromagnetic calorimeter .
10.1.1 Main barrel . . . . . .
10.1.2 Endcaps . . . . . . . .
10.1.3 Preshower detector . .
10.2 Hadron calorimeter . . . . . .
10.3 Zero degree calorimeter . . .
10.4 Calibration of the CMS . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

24
25
25
26
26
26
27
27

11 ATLAS
28
11.1 Electromagnetic calorimeters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
11.2 Hadron calorimeters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

Part I

Introduction
Perhaps the most common use of the term calorimeter is in reference to a device used to measure heat
transferred to a medium during various processes, most simply through the measurement of temperature
change of the medium. In the field of particle physics, a calorimeter is a device which, ideally, measures the
total energy of a particle. Analogous to the fact that traditional calorimeters can only measure a change
in thermal energy if that energy is eventually converted into heat which is transferred to the calorimeter,
particle calorimeters can only measure the energy associated with a particle if that energy is transferred
from the particle to the calorimeters sampling medium; since energy is defined as the ability to do work,
the energy of a particle can only be measured as that energy is used in a demonstration of the work the
particle is capable of. The purpose of the calorimeter is to cause particles to perform such demonstrations
and measure their final products [12].

Particles and their classes

In order to examine how different particles interact, it is important to examine what properties are different
between them. We use the term interact to talk about the mechanisms through which particles affect each
other. These fundamental mechanisms are the strong force, the weak force, electromagnetism, and gravity.
Mankinds scientific knowledge has recently grown to the point where we are able to explain how some these
forces affect particles on a subatomic level at low energy levels: particles belonging to a class called bosons
carry forces between other particles. Photons are the carriers of the electromagnetic force, gluons are the
2

Figure 1: The different discovered and theorized particles, the forces through which they interact, and related
theoretical systems. By Gaetan Landry.

carriers of the strong force, and the W and Z bosons carry the weak force. A hypothetical particle called
the graviton may carry the gravitational force. The forces vary in their strengths and the way their effects
carry across a distance. For example, gravity is much weaker than electromagnetism, but the effects of both
gravity and electromagnetism may permeate through an infinite distance, while the effects of the nuclear
force, an artifact of the strong force, are limited to subatomic scales. Not all forces may affect all particles.
In fact, there is a theoretical form of matter designated as dark matter which interacts with other matter
through gravity, and is theorized to interact through the weak force, but is not affected by any other force.
More detailed discussion of the mechanisms of force propagation involve studies of topics such as quantum
field theory, and this paper shall proceed no further in this regard. Particles that interact through the
electromagnetic force are said to be charged, and a particle may be either positively charged or negatively
charged. Particles that interact through the strong force are said to be color charged. Particles also carry
different values of a property called spin.
Electrons, muons, neutrinos and a few other particles make up the class of particles known as leptons.
All leptons carry mass and a spin value of 1 /2 . Electrons carry negative charge. The muon has the same
electric charge as the electron but is much more massive. The neutrino carries much less mass than the
electron and no electric charge.
Hadrons are particles made up of quarks. Quarks come in several flavors and may carry positive or
negative charge with a magnitude of 1 /3 or 2 /3 of the charge of an electron. The different quarks are
differentiated by their electric charges, their masses, and their weak isospins (a property responsible for how
they interact with the weak force). The different quarks are known as up, down, strange, charm, top, and
bottom. Hadrons can be made of two or three quarks; those made of three quarks are called baryons, while
those made of one quark and one antiquark (a form of antimatter, which shall be discussed shortly) are called
mesons. Hadrons include protons, neutrons, kaons, and pions. The proton carries a positive electric charge
whose magnitude equals that of the electrons and a spin of 1 /2 . The neutron carries no electric charge, a
spin of 1 /2 , and a mass slightly greater than that of a proton. Pions are mesons and come in several types
that carry different electric charges and masses. A pion may carry either a positive or negative charge equal
in magnitude to that carried by an electron, or no electric charge, and carries a mass (differently-charged
pions carry slightly different masses) greater than that carried by an electron and less than that carried by
a proton. A kaon may also carry a positive or negative charge equal in magnitude to that of an electron, or
no electric charge, and carries a mass greater than that of a pion but less than that of a proton. Kaons are
mesons, but, unlike pions, they carry a strange quark in addition to an up quark or a down quark.
Antiparticles are those which exhibit the same properties as certain particles except for the fact that
the electric charge that they hold is opposite in sign. The positron, which is the antiparticle compliment of
the electron, for example, has the same mass and spin as an electron, but holds a positive charge equal in
magnitude to that of the electron. There exist antiparticle counterparts to particles which carry no electric
charge as well, and, in the case of neutral antihadrons, contain antiquarks which have electric charges that
are opposite to their quark counterparts. Mesons are composed of both quarks and antiquarks. Kaons are
composed of either an up quark and a strange antiquark, a down quark and a strange antiquark, a strange
quark and a down antiquark, or a strange quark and an up quark. A pion can be composed of an up quark
and a down antiquark, which is the antiparticle of the version that is composed of a down quark and an up
antiquark. Pions may also be composed of an up quark and an up antiquark or a down quark and a down
antiquark.

Part II

Showers
We shall now discuss the events that occur when high energy particles enter a medium.
As particles move through a medium, they are affected by and affect the medium through collisions. A
collision is an event in which particles or groups of particles come close enough together to affect each other.

Collisions may result in one or more events depending on the types of particles involved, their energies, and
other factors. Showers occur when these events result in subsequent collisions. There are two categories
of showers that may develop: electromagnetic showers result from the penetration of high energy electrons,
positrons, or photons, while hadronic showers result from the penetration of hadrons.

Electromagnetic showers

Electromagnetic showers are those that involve electrons, positrons, and photons traveling through a medium.
We shall first examine the individual collisions events, then look at how these events are perpetuated through
showers.

2.1

Electron and positron collision events and energy loss mechanisms

Electrons traveling through a medium lose energy as they collide with atomic electrons through interactions
involving the electromagnetic force. The energy lost during each collision can be calculated from the impulse
effect of the electromagnetic force, but this calculation is complicated by the deflection this impulse imparts
upon electrons and positrons. Later, we shall examine in greater detail the energy lost by much heavier
particles through calculations simplified by the fact that these particles are hardly deflected during collisions
with electrons.
2.1.1

High energy

Bremsstrahlung Bremsstrahlung, which is German for braking radiation, arises from the fact that
accelerating electric charge sources produce electromagnetic radiation. Since particles much heavier than
electrons do not undergo significant acceleration when colliding with electrons except when traveling at very
high speeds due to relativistic effects, such collisions do not produce significant bremsstrahlung. However,
a large portion of the energy lost by high energy electrons and positrons during collisions is due to this
phenomenon. The wavelength of the radiation produced depends on the magnitude of the acceleration of
the charge source and is continuous: the wavelength can take on an uncountable infinite number of values.
Quantum mechanical theory explains that electromagnetic radiation is carried by photons. Photons each
carry an amount of energy that depends only on the frequency of electromagnetic radiation.
Radiation length relates the energy lost by a high energy electron or positron through bremsstrahlung to
the distance it travels. It is the distance such a particle travels as it loses 1 1e of its energy through this
mechanism. The radiation length varies with the material the electron or positron travel through: less dense
materials have a longer radiation length than denser materials. When discussing calorimetry, we sometimes
use radiation length as a unit of thickness. When thickness is described in this way, the energy lost by a
high energy electron or positron due to bremsstrahlung through a given thickness does not depend on the
density of the penetrated material.
2.1.2

Low energy

At lower energy levels, energy loss is dominated by mechanisms other than bremsstrahlung.
Ionization If an electron imparts enough energy upon an atomic electron to completely overcome the
coulomb force binding that electron to the atom, ionization occurs. The atomic electron leaves the atom,
becoming a free electron, and the atom becomes an ion.
Excitation When a collision between electrons does not involve energy levels high enough for ionization
to occur, atomic electrons involved in a collision may simply enter an exited state. According to quantum
mechanical theory, electrons orbiting an atomic nucleus occupy one of a number of states depending on their
energy. These states have discreet energy requirements: an electron cannot jump from one state to another

Figure 2: This graph illustrates two definitions of critical energy. The points where the two bremsstrahlung
lines cross the ionization line on this graph are the critical energies given by the second definition suggested
by experimental physicist Bruno Rossi. The dotted line shows the approximate bremsstrahlung values given
by the equation which makes the definitions match. From Particle data group.

200

l
ta
o
T

Rossi:
Ionization per X0
= electron energy

50
40
30

em
Ex
s
ac
tb E
re
m
ss
tr
ah

70

Br

dE/dx X0 (MeV)

100

lu
ng

Copper
X0 = 12.86 g cm2
Ec = 19.63 MeV

Ionization

20

Brems = ionization
10

10
20
50
Electron energy (MeV)

100

200

unless it has absorbed a certain amount of energy. Conversely, an electron cannot absorb any energy unless
that energy is enough to cause the electron to rise to the next highest state. The lowest state an electron
can occupy given an atoms configuration is called its ground state. An electron exited to a state beyond its
ground state quickly fall back to its ground state, emitting radiation (photons) in the process. The time it
takes an electron to fall from an exited state to its rest state is called the response time. The frequency of the
radiation emitted and the energy of the photons carrying it depend on the difference between the energies
associated with the atomic electrons ground state and the exited state.
2.1.3

Critical energy

Critical energy is the energy level at which energy lost by a particle through bremsstrahlung equals that
lost through ionization and excitation during collisions [4]. The reason the relationship between energy
lost through these two mechanisms depends on the energy of the particle is that energy loss rate per radiation length through ionization rises logarithmically while energy loss rate per radiation length due to
bremsstrahlung rises linearly. In one definition, the critical energy is the energy level at which the energy
lost by an electron through bremsstrahlung equals that lost through ionization. Another definition gives the
critical energy as the energy at which the all the energy held by an electron would be lost through ionization
in just one radiation length. These definitions are actually equivalent if we approximate the energy lost
through bremsstrahlung as the total particle energy divided by the radiation length in the first definition, an
approximation which breaks down at low energy levels. A comparison of these two definitions is illustrated
by figure 2.

Figure 3: Pair production occurring, with energy being absorbed by a nucleus. We see the electron and
positron diverge from the path of the photon by opposite angles with magnitude . By David Horman.

Nucleus

2.2

Photon collisions

Photons carry energy equal to the frequency of the radiation associated with it multiplied by Plancks
constant. At low energy, photons deposit energy through the photoelectric effect, but at higher energies a
phenomenon known as pair production may occur when a photon enters the vicinity of another particle.
2.2.1

Pair production

Photons, under certain conditions, can be replaced by pairs of electrons and positrons. The fact that this
event is possible may be counterintuitive: a massless particle disappears, and two particles, both carrying
mass and electric charge, suddenly begin to exist. This feeling can be combatted by answering a simple
question: are any physical laws actually broken when such an event takes place? More specifically, are laws
that hold that mass, energy, charge, momentum, or angular momentum (which is the sum of the quantum
mechanical concepts of spin and orbital angular momentum) cannot be created or destroyed being violated?
Obviously, since an electron has an electric charge that is opposite that of a positron, the amount of electric
charge of the system is conserved during such a transformation: the combined charges of the electron and the
positron equal the charge of the original photon, or zero. Likewise, the spins of the electron and positron,
each equalling 1 /2 , add up to equal the spin of the original photon, or one. It is true that neither mass
nor energy are conserved if we adhere to the classical understandings of both. However, Einsteins famous
equation E = mc2 tells us that an amount of energy has an equivalence to an amount of mass. A massive
particle, even one at rest, is said to carry rest energy equal to the particles mass multiplied by the square of
the speed of light. Thus, if a photon becomes an electron and a positron, energy conservation is not violated
since the energy of the photon can become mass held by the electron and positron, but only if the photon
had an amount of energy that was equal to or greater than the combined rest energies of these two particles.
Therefore, if a photon carries less energy than this, pair production cannot occur. If a photon has more
energy than the combined rest energies of the electron and positron, the electron and positron, upon their
creations, will each have kinetic energies equal to half the difference between the photons energy and the
combined rest energies of the particles. Finally, momentum may be conserved during pair production under

certain conditions as well. Photons carry a positive amount of momentum related linearly to their energies.
This momentum can be conserved during production of electron and pairs that are at rest (due to their
geneses from a photon whose energy equalled the combined rest masses of the positron and electron) only
if such pair production occurs in the vicinity of a mass-carrying particle or group of particles. Momentum
conservation in the event of pair production from a photon possessing more energy than the combined
energies of the created electron and positron is also impossible in the absence of additional material which
may absorb momentum, but this concept is a bit less intuitive. We shall examine this concept by looking
at the special case in which a high energy photon produces an electron and a positron that each have the
same velocity relative to the velocity the photon (and the nucleus involved in the collision), and in which
the photon has more energy than the sum of the rest masses of the electron and positron that are created.
Based on relativistic considerations, a photons energy Ep is related to its momentum pp by the equation
Ep = pp c,

(1)

where c is the speed of light (and the speed at which photons must travel). Also, when considering consequences of relativity, the momentum pe of particles with rest mass m traveling at velocity v is given by the
equation
mv
pe = p
1 (v/c)2
and the energy Ee of such particles is given by
Ee = p

mc2
1 (v/c)2

In this special case, the two massive particles move in paths with opposite angles to the original photons
path. Their momenta in the directions normal to the original photons path are therefore canceled if they
move with the same speed, which preserves the direction of the momentum vector if is between 90 and
zero degrees, and their combined momentum becomes
mv
2 cos p
.
1 (v/c)2
The cosine of must obviously be less than one due to the above mentioned constraints of . For energy
to be conserved, we must have the combined energies of the electron and positron equal the energy of the
photon:
2mc2
Ep = p
,
1 (v/c)2
but if we try to determine the momenta of the photon and the particles created during pair production using
equation 1 we find that
Ep
2mc
mv
=p
> 2 cos p
.
2
c
1 (v/c)
1 (v/c)2
We arrive to the result that, in order for energy to be conserved, excess momentum is produced. Pair
production must involve matter besides the photon and the created pair to carry off excess momentum.
This is why pair production occurs in the vicinity of atomic nuclei: it is an event that results from collisions
between high-energy photons and atomic nuclei. It is also why, during pair production, the system containing
just the photon and the pair of particles produced loses energy.
When certain particles and antiparticles collide, annihilation occurs: both particles disappear and are
replaced by photons that carry the same amount of energy that the annihilating particles did, including the
particles summed rest energies. Therefore energy, like charge, is conserved during annihilation of particles
and antiparticles. Momentum must also be conserved during annihilation events. Similar to the situation
above concerning pair production, and due to the facts that photons carry no mass and may travel at no
speed other than c, momentum conservation places a restriction on the result of annihilation: it must result
8

Figure 4: These graphs show the cross sections of several collision events between photons and atoms of
two different absorbing media with respect to photon energy. We are interested in nuc , the cross section of
nuclear pair production. Other cross sections shown here are those for collisions involving the photoelectric
effect (p.e. ), Rayleigh scattering (Rayleigh ), Compton scattering (Compton ), pair production that transfers
momentum to atomic electrons (e ), and photonuclear interactions (g.d.r. ). From Particle data group.
(a) Carbon (Z = 6)
- experimental tot

Cross section (barns/ atom)

1 Mb

p.e.
1 kb

Rayleigh

1b

nuc
Compton

10 mb

(b) Lead (Z = 82)

Cross section (barns/ atom)

1 Mb

- experimental tot

p.e.

Rayleigh
1 kb

nuc
g.d.r.
1b

Compton
10 mb
10 eV

1 keV

1 MeV

Photon Energy

1 GeV

100 GeV

in the creation of more than one photon. As an example, if an electron and positron collide at a very low
velocity, two photons will be produced that travel in opposite directions. The photons will have a radiation
frequency such that their energies equal that of the rest energies of the electron and positron, but the total
momentum between them will be zero.
There are other processes that effect photons traveling through matter, but pair production is dominant
at the high photon energy levels involved in shower production. The cross-section of a pair-production event
rises with the atomic weight A and the atomic number Z of the atomic nucleus involved. As seen in figure
4, the cross-section of pair-production events approaches a limit at high photon energy levels. This limit
can be related to the radiation length X0 defined in subsubsection 2.1.1 by the equation
=

7 A
9 X0 NA

(2)

where NA is Avogadros number, being used here as a conversion constant [5].


2.2.2

Mean free path

A photon carrying radiation of a certain frequency can only exist carrying a certain amount of energy. A
photon cannot deposit only part of its energy. During any collision event, the photon deposits all of its energy
to another entity and vanishes. Thus, a definition similar to to that of radiation length, that is, the distance
through which a photon loses 1 1e of its energy, cannot be meaningful. Instead, we discuss a photons mean
free path, which, for photons traveling with a certain energy traveling through a certain material, is defined
as the mean distance those photons travel before being involved in collisions and vanishing. From equation
2, we find that the mean free path is 79 of the radiation length associated with a high-energy electron moving
through the same material [12].

2.3

Electromagnetic shower propagation

We have now examined the individual processes that perpetuate electromagnetic showers. The fact that
these chain-reactions occur is intuitively true based on the results of bremsstrahlung, pair production, and
annihilation. The initiation of electromagnetic showers only requires a high-energy photon, electron, or
positron to enter a medium. If it is a high-energy electron that enters the medium, the photons produced
through bremsstrahlung will produce positrons and more electrons through pair production when they collide
with atomic nuclei. Both the newly-created positrons and electrons travel through the medium producing
more photons through bremsstrahlung. These photons may produce even more electrons and positrons
through pair production, but most photons produced through bremsstrahlung do not carry enough energy
to effect pair production, and are destroyed in events such as the photoelectric effect. The positrons continue
shower perpetuation when they annihilate, usually with an atomic electron, resulting in the production of
more high-energy photons. These reactions continue to produce each for some time. It should be obvious
how the penetration of a positron or high-energy photon will result in such a chain reaction as well.
2.3.1

Shower maximum

The particles that perpetuate electromagnetic showers lose energy as they do so. Electrons and positrons
lose energy through bremsstrahlung whenever they interact with an electric field. Pair production involves
a loss of energy to the atomic nucleus that carries off the excess momentum of the event, and so the total
energy of the electron and positron that the event produces is slightly less than the energy carried by the
initial photon. The annihilation of electrons and positrons often produces more than two photons, and in
this case the photons each carry less energy than either of the annihilated particles. Eventually, particles
are produced that in no way can increase the number of particles traveling in the shower. We come to see
photons whose energies are less than the combined rest energies of the electron and positron pairs which
could otherwise be produced, and electrons that move too slowly to produce high-energy photons through
bremsstrahlung. The depth within the medium where particle multiplication in a certain shower ceases is
10

Figure 5: This graph shows the total energy deposited by a shower propagating through iron as a fraction of
the energy of the particle that initiated the shower as a function of the depth in radiation lengths. We also
see the numbers of particles carrying energy above a certain cutoff value. This data was produced through
a computer simulation. From Particle data group.

0.125

30 GeV elect r on
incident on ir on

(1/ E 0) dE/ dt

0.100

80

0.075

60
Ener gy

0.050

40
Phot ons
1/ 6.8

0.025
0.000

20

Elect r ons
0

5
10
15
t = dept h in r adiat ion lengt hs

20

Number cr ossing plane

100

called the shower maximum. After this point, we see that the number of particles within the shower stops
increasing. Intuitively, the shower maximum depends on the energy of the particle that initiates the shower;
a particle carrying more energy will, in general, produce a shower that propagates further before the energies
of its particles fall below critical levels. The shower maximum, when measured in units such as meters,
depends on the material through which the shower propagates. Particles traveling through materials made
of atoms with high atomic numbers will likely experience energy-stripping collisions more often. However,
shower maximum is less dependent on the medium when it is given in units of radiation length. The shower
maximum occurs where the average energy of the showers particles is less than the critical energy.
2.3.2

Energy deposit as a function of depth

As showers develop and involve particles that hold less energy, the rates through which they lose energy
change as well as the proportions of the different energy loss mechanisms. Typically, as showers first begin
to develop, the amount of particles involved increases at such a rate that the rate at which energy is deposited
to the medium increases with depth. Beyond a certain depth, however, the tendency of the shower energy
loss rate to increase due to the increase in the particles involved is overtaken by the tendency of the particles
involved to carry less energy and deposit less during collision events. As the shower progresses, it loses less
energy due to bremsstrahlung and pair-production and more energy to other events.
2.3.3

Moli`
ere radius

The Moli`ere radius M is a proportion of the radiation length X0 of a high energy electron moving through
a certain medium to that electrons critical energy c in the same medium scaled by a factor called the
scale energy Es which is related to the electrons mass and a constant which holds quantum mechanical
information concerning the electric force. It is given by the equation
M = Es
11

X0
.
c

The Moli`ere radius of an absorption medium relates to the maximum width of an electromagnetic shower
propagating through it.
2.3.4

Shower profiles

Since both radiation length and critical energy are a related to the atomic number Z of the atoms that
compose the medium, the Moli`ere radius is calculated from an equation which roughly cancels out Z, and
so the Moli`ere possesses a low dependence on Z compared to that possessed by the radiation length. As a
result, the shape of a shower depends heavily on the material through which it propagates: long and narrow
showers develop in low-Z materials while shorter yet wider showers develop in high-Z materials.
If we measure depth in units of radiation length and the distance between a particle and the shower
center in units of Moli`ere radii, shapes of showers traveling through different media look more similar than
they do when we measure such values in meters. However, differences still remain in spite of the adjustments
made by these units. We see that higher Z materials host electromagnetic showers that have higher shower
maximum depths, and that decay more slowly beyond their shower maxima. More radiation lengths of lower
Z materials are therefore needed to contain a given proportion of a showers energy. We also find that the
length of a showers profile (again, measured in radiation lengths) scales logarithmically with the energy of
the particle that initiates the shower.
The spread of a shower is caused the paths of electrons being altered, an effect which is to be expected
due to the electrons low masses, and by photons produced by bremsstrahlung, which may travel at high
angles relative to the axis of the shower. Transverse properties are those related to the relationship between
shower components and the distance from the axis of shower development. Transverse shower properties can
be discussed in two ways: lateral shower properties are discussed in relation to a function of a unit volumes
distance from the shower axis, while radial properties are those discussed in terms of a radial (ring-shaped)
slice and a function of that slices radius. According to experimental particle physicist Richard Wigmans,
however, these two terms are often used interchangeably in papers, presenting a situation of ambiguity which
leads to significant confusion. Radial energy density tends to decrease exponentially as a function of the
distance from the showers center. We find that for high Z materials, the radial energy density decreases
more rapidly as a function of distance from the shower axis than for low Z materials.

Hadronic showers

Showers initiated by the penetration of hadrons through a medium develop differently than those initiated by
electrons, positrons, and photons. While we shall first discuss inelastic collisions between penetrating protons
and atomic electrons, we are most interested in phenomena that occur at high energy levels, such as spallation.
While the showers discussed previously are perpetuated solely through electromagnetic interactions, hadronic
showers involve phenomena related to the strong nuclear force.

3.1

Electron collisions

If a particle is much heavier than an electron and carries an electric charge, the majority of the energy
it loses as it travels through matter is through collisions between the particle and the electrons of the
mediums atoms. While observing these low-energy collisions is not the goal of modern high energy particle
calorimeters, these collisions are one of the results that is directly measured, and shower processes can be
extrapolated from these measurements. The amount of energy lost by the traveling particle per unit of path
length, dE/dx, is called the stopping power. It can be roughly calculated using classical assumptions, but
precise calculation of stopping power requires quantum mechanical and relativistic considerations. We shall
proceed through the classical calculation of energy loss, as the process illuminates several aspects of these
types of collisions. The following calculations are based on the book by Leo [9].
We shall first calculate the impulse on an atomic electron as a charged particle moving at velocity v
collides with it, then calculate the energy change from that impulse. The minimum distance the electron

12

comes to that particle will be given as b. We start with the definition of impulse:
Z
I = Fdt
We may find the force F due to the electric field E generated by the traveling particle with charge given as
a multiple of that of an electron as ze using Coulombs law. We only need to consider E , the component of
the electric force perpendicular to the particles path at all points, since any force produced by the particle
parallel to its path as it approaches the atomic electron is canceled out by an equal force in the opposite
direction as the particle recedes from the atomic ion:
Z
Z
Z
dx
dt
dx = e E .
I = e E dt = e
dx
v
The above equations are only valid if the atomic electron is initially at rest and is only moved slightly during
the collision, and if the penetrating particle is not significantly deviated from its path, a condition which
is satisfied since the penetrating particle we are discussing is very massive compared to the electron. To
calculate the integral, we use the integral form of Gausss Law, which, when expressed using Gaussian units1 ,
is
Z
EdA = 4Qenc ,
s

which means that the total electric field experienced by the entirety of a surface (the electric flux through
that surface) surrounding the source of that field is 4Q, with Q being the magnitude of the enclosed charge.
We imagine that the atomic electron moves (in the reference frame of the penetrating particle) along the
outside of an infinitely long cylinder with radius b and find the electric field experienced by that surface to
be
Z
E 2bdx = 4ze.
So to find the total electric field experienced by the electron, we find the electric flux through a line along
that cylinder by removing 2b from both sides of the above equation:
Z
2ze
.
E dx =
b
We arrive to the result

2ze2
bv
Since
Z
Z
Z
dv
I = F dt = m dt = mdv = mdv = mv
dt
2
when mass is not a function of velocity , and since the change in a electrons energy is given by
Z
Z
Z
Z
dv
dv
1
U = F dx = m dx = m vdt = mvdt = m(v)2 ,
dt
dt
2
I=

which again assumes that mass is not a function of velocity, we arrive at a relation between mass and
impulse, and therefore the amount of energy gained by the electron and lost by the traveling particle during
the collision:
I
2z 2 e4
U =
=
.
2me
me v 2 b2
This calculation requires modification due to quantum mechanical and relativistic effects which will not
be discussed [9].
1 In Gaussian units, distance is expressed in centimeters and mass is expressed in grams. Unlike SI units, Gaussian units
allow Gausss law and many other theories to be expressed without scaling factors, explaining their popularity among theoretical
physicists.
2 An assumption that does not hold and therefore requires that these equations be refined for particles traveling at relativistic
speeds.

13

3.2

Spallation

Spallation is the most common result of collisions between high energy hadrons and nuclei of a mediums
atoms. Generally, the word spallation refers to a process in which an object enters a medium and produces
ejecta from the medium, referred to as spall.
Hadrons and the quarks that compose them interact with each other through the strong force. The strong
force has three aspects, called colors. Full understanding of the concepts of aspects and colors requires a
study of quantum chromodynamics, which is beyond the scope of this paper, but we shall undergo a brief
discussion using an analogy of the electromagnetic force. The electromagnetic force holds one such aspect: a
charge that can either be positive or negative. The strong force, however, has three such degrees of freedom,
referred to as red, green, and blue. The strong force is such that its strength increases between particles
as a function of the difference between them. As a result of this property, we never find free quarks. The
strong forces interaction with quarks produces a residual effect on the hadrons that they compose, called the
nuclear force. However, the nuclear force between two particles which may be affected by it does not increase
with distance, and in fact decreases more rapidly than any other force does as a function of distance. At
very short distances, however, the nuclear force is stronger than any other force which may act on hadrons
by far. As mentioned in section 1, gluons are the mediators of the strong force.
Spallation is facilitated by the strong force. When a hadron collides with an atomic nucleus, it may cause
the components of the nucleus to travel through the nucleus itself, colliding with other nuclear hadrons,
which may move to produce even more collisions. This phenomenon is known as a fast intranuclear cascade.
In a process similar to pair production, certain particles may come into existence if particles move through
a nucleus with enough kinetic energy to exceed the rest energies of new particles, and if other requirements
related to conservation of mass, momentum, charge, and angular momentum are met. These particles are
usually mesons, those which are made of one particle and one antiparticle and that therefore quickly decay
into high energy photons (those that have wavelengths in the gamma radiation spectrum) or other particles.
The penetrating hadron itself may transform into new hadrons. Also, the struck nucleus may eject a portion
of its neutrons and protons, entering an excited state, and emits hadrons and photons in the process of
returning to a lower energy state. This process, which follows the fast intranuclear cascade, is known as the
evaporation stage.

3.3

Hadronic shower components

Spall particles that do not have short decay times, such as protons, neutrons, and charged pions, may
proceed to undergo spallation events with other nuclei of a medium, and a hadronic shower exists. Along
with propagation behaviors of these types of showers, we must also consider the proportions of the many
different particles that they may involve (remember, when we discussed electromagnetic showers we only
had to worry about the propagation of protons, electrons, and positrons). An important consideration, for
example, is the proportion of mesons in a shower at different levels, since mesons quickly decay into photons
(or, sometimes, electron and positron pairs) which produce electromagnetic showers. This consideration is
important since, as we shall discuss later, particle calorimeters almost exclusively measure the effects of
electromagnetic showers directly. About a third of the mesons ejected from nuclei during a showers initial
spallation event are neutral pions, and the average number of all mesons produced from spallation increases
logarithmically with the energy of the collision. To find the portion of shower that is in the form of neutral
pions, and therefore the portion of a hadronic shower that propagates as an electromagnetic shower, at a
shower generation n, we add the portion of the shower propagating electromagnetically before the generation
n to the portion of the shower not propagating electromagnetically multiplied by the fraction f0 of spallation
particles that are neutral pions at the energy level of collisions typically occurring during the generation in
question. For example, if, for simplicity, one third of particles produced through spallation are neutral
pions at every generation, then one third of the shower would be an electromagnetic shower after the first
generation, 13 + 23 31 of the shower would propagate electromagnetically after the second generation, and
so on. The portion of a hadronic shower that has become an electromagnetic shower by a generation n is

14

therefore
n

fe m = 1 (1 f0 ) .
The average fraction of neutral pions produced through spallation f0 depends on the atomic number of
atoms in the absorbing medium. The fraction throughout the entire shower also depends on whether the
initial spallation event was started by a pion or a proton: electromagnetic fractions of showers started when
a pion impacts a nucleus are about 15% lower than the fractions of showers started when a proton hits a
nucleus. This is because the type of particles that are leading particles, the particles that leave the nucleus
first during spallation, depend on the type of particle that started the spallation reaction. A reaction started
by a pion will produce leading particles in the form of pions of types that often interact with absorber
medium atoms before decaying, while a reaction started by a proton will produce leading particles that are
protons, and these effects propagate through the showers. We can determine the percentages of energies
carried by electromagnetic showers from our examinations of the portions of hadron shower particles that
are those which decay electromagnetically.

3.4

Spallation cross sections

Because of the low range of the nuclear force, the cross section of a spallation reaction is extremely small
when compared to the cross sections of collisions events discussed previously. The relation for the cross
section of such a collision between a particle with energy E and a nucleus that initially has an atomic mass
At that produce a spallation event that results in the atomic nucleus losing particles such that its atomic
number becomes Zf and its atomic mass becomes Af is
Zf ,Af eP (AT Af ) eR|Zf SAf +T Af |
2

3
2

where P , R, S, and T are constants. The total spallation cross section is a sum of the cross sections of all
spallation reactions that may occur.

3.5

Consider the neutron

Because they do not carry charge, low-energy neutrons are able to come very close to atomic nuclei and
affect elastic collisions- those in which the sum of the kinetic energies of the neutron and the nucleus remains
constant and kinetic energy is transferred between the neutron and the nucleus. Depending on the location
of the collision on the nucleus, the energy transferred from the neutron to the nucleus is some portion of the
maximum possible transfer amount, which is a function of the nucleuss atomic number A and is given by
Emax =

4A
.
(A + 1)2

Since the energy transferred per collision decreases with atomic number, it is obvious that a medium that
has an abundance of hydrogen will absorb kinetic energy from low-energy neutrons most efficiently.
Free neutrons may spontaneously decay (they have a half-life of about 15 minutes), but they are much
more likely to be absorbed when traveling through a medium at very low speeds. The neutron is absorbed
into a nucleus of the medium, which puts that nucleus into an excited state. The nucleus then proceeds
to emit energy as gamma rays or, in the case of lithium or boron nuclei, an alpha particle, which is a
mass-carrying boson.
At high energies neutrons may inelastically collide with an atomic nucleus, putting the nucleus into an
excited state without being absorbed.

3.6

Invisible energy

Spallation requires that nuclear particles overcome the nuclear force, and the energy required and expelled
in this process is called nuclear binding energy. When spallation occurs, the sum of the energies of the spall
15

particles is less than the that of the particle that produced the spallation event. This difference is the energy
that was expended in overcoming the nuclear binding energy. It is not detectable by any means, although it
may be replaced if some of the spall particles join other nuclei in events that produce gamma radiation.

3.7

Hadronic shower profiles and their relation to electromagnetic shower profiles

The inertias of high-energy hadrons are such that they propagate through an absorbtion medium virtually
unaffected until they are involved in a nuclear collision. Just as the mean free path tells the average distance
a high energy photon will travel in some medium before being destroyed in a collision event, the average
distance that high energy hadrons travel through a medium is known as the mediums nuclear interaction
length, and is given by the symbol int , and the probability that a particle traverses some distance z in this
medium without colliding with a nucleus is
P = ez/int .
The nuclear interaction length is inversely proportional to the sum of the cross sections of all possible
spallation events and proportional to the weights of atoms with which a penetrating hadron may collide:
tot

A
.
int

Nuclear interaction lengths are generally far greater than the mean free path of photons and the radiation
length of electrons in a given medium. Hadronic showers, therefore, propagate much further into an absorbing
medium than electromagnetic showers in general. Similarly to an electromagnetic shower, the particles
involved in a hadronic shower increase in number as the shower progresses through the medium until their
number reaches a maximum. The energy carried by the hadronic shower decays less per (SI) unit of distance
than that carried by electromagnetic showers after their respective shower maxima. Hadronic showers are
also larger laterally. We find that as a function of the atomic numbers of the atoms that make up the
absorbtion medium, radiation length drops faster than nuclear interaction length: the cross section of a
bremsstrahlung event is proportional to Z 2 of the medium, while the cross section of hadronic spallation
is proportional to Z 2/3 . High Z materials start bremsstrahlung reactions faster, and are used to separate
hadronic showers from electromagnetic showers in media [12].

Part III

Calorimetry in general
4

Measuring end results

Whether the focus is electromagnetic showers or hadronic showers, a common mechanism that modern
particle calorimeters use to measure energy deposit is scintillation. Scintillation is the light emitted when
an atomic electron moves from an excited state to a less energetic state. Given materials may produce
scintillation radiation having one of several possible discrete wavelength values depending on the material
composing the medium, and photodetectors sensitive to these frequencies may therefore be used to detect
scintillation events. Many of the processes that we have discussed in previous sections result, sometimes
eventually and indirectly, in scintillation.
More often, though, calorimeters measure the effects of penetrating particles on the charge of their media.
A calorimeter designed to measure charge may utilize liquified nobel gasses as its medium, and measures the
amount of ionization of the atoms of the element due to charged particle and photon collisions. Electrons
freed from the mediums atoms are attracted to an anode on one side of a chamber containing the sampling
medium, while the positively charged ions are attracted to a cathode in another location, and current flow

16

Figure 6: A photomultiplier tube. By Colin Eberhardt.

through a conduit linking the anode and cathode due to resultant voltage changes between them is measured.
Charge-measuring calorimeters may also utilize semiconductors as the sampling medium. When atoms of the
semiconductor are excited, they become conductors, and electric fields may be used to collect free positive and
negative charges created by the penetration of ionizing particles. Calorimeters that utilize this mechanism
are called solid-state calorimeters. The sampling medium of solid state calorimeters may be constructed out
of silicon chips.

4.1

Recording scintillation events

All the energy of particles involved in an electromagnetic shower propagating through a homogeneous
calorimeter may absorbed in the excitation of medium in a way that produces scintillation. Most of the
photons produced through bremsstrahlung accomplish this when the collide with electrons of the medium;
low energy electrons may produce excitation of an atomic electron through the coulomb force, or may ionize
the atom resulting in a free electron that eventually causes excitation in other atomic electrons. High energy
photons affect pair production, producing particles that cause scintillation through the effects described
above.
The sampling medium of calorimeters that utilize scintillation for energy measurement consists of a material that allows light to pass through. There exist boundaries between sections of the sampling medium
through which scintillation light may not pass. Light detection systems, tuned to activate upon the observation of radiation with any of the frequencies that may be produced through scintillation of the sampling
medium, are attached to each of the sections separated by opaque boundaries. This setup enables a determination concerning the location of a scintillation event. The detector may record the division in which a
scintillation event takes place, but there is no way to determine exactly where the event took place within
a single division. Higher granularity results from smaller divisions of the sampling medium and enables the
location of a scintillation event to be determined with greater precision. A photomultiplier tube (figure 6) is
a device which may be attached to a sampling medium division in order to detect light from scintillation. At
the front of a photomultiplier tube is a plate. When photons interact with this plate, electrons are ejected
from the side of the plate facing away from the scintillation medium through the photoelectric effect. These
electrons, which are too few in number to be detected directly, are attracted to a positively charged plate
called a dynode. When they collide with this plate, they knock off more electrons, which are then attracted
to a second plate, producing a similar result. The dynodes are angled in such a way so as to allow knocked-off
and original electrons to continue onto dynodes further back in the tube. The process continues, multiplying
the number of electrons traveling through the tube, until they produce a measurable current pulse upon
17

an anode at the back end of the tube. Another such device is an avalanche photodiode, which allows a free
electron ionized through the photoelectric effect to knock other electrons from atoms by accelerating it under
a high voltage, producing a current.
Some problems may occur with photodetectors that produce inaccurate energy readings. That is, even
if all the energy of a shower is absorbed by a sampling medium to produce scintillation light, not all of that
energy will produce recordable results. Photomultiplier tubes may suffer from saturation: electric current
created within the tubes upon detection events lowers the potential difference between the dynode plates,
meaning that electrons knocked off from the plates during subsequent detection events will not impact plates
further back in the tube with as much energy as would be expected.

Homogeneous and sampling calorimeters and energy response

Homogeneous calorimeters are those in which the material intended to stop incoming particles is also the
material that produces scintillation light for energy measurement. Sampling calorimeters, on the other hand,
are those that utilize one material, the passive medium, to absorb large amounts of energy from penetrating
particles, and another material, the active medium, to measure the energies of particles at different points
along their paths so as to ascertain the effect the passive medium has upon them. Sampling calorimeters
are designed so that a single particle passes through several boundaries between active and passive media,
making use of many possible patterns for staggering the two media classes. Sampling calorimeters are used
to measure the energy content of hadronic showers because the stopping power of scintillation media is too
low to contain hadronic showers in any reasonable volume of the material.

5.1

Energy response of homogeneous calorimeters

Energy response is the portion of the energy carried by a shower that is actually recorded. Phenomena
that affect the energy responses of homogeneous calorimeters are also applicable to sampling calorimeters.
In addition to instrumental effects on energy response discussed in subsection 6, the energy response to
electromagnetic showers is decreased by particles that escape the detector. Energy response to hadronic
showers is mitigated by additional problems. For one, the invisible energy, discussed in subsection 3.6, has
an obvious impact. Also, the energy of hadronic showers is mostly detected by scintillation caused by the
showers electromagnetic components. This component only makes up a portion of the hadronic shower, and
the size of this portion is dependent on the showers energy, as discussed in subsection 3.3.

5.2

Energy response of sampling calorimeters

The sampling fraction, fsamp , of a sampling calorimeter is defined by Wigmans as the fraction of energy
deposited by minimum ionizing particles in the sampling medium of the total energy deposited by such
particles. Minimum ionizing particles are hypothetical particles that have just enough energy to affect
ionization of the absorbing medium. As mentioned before, scintillation calorimeters, at least in theory, have
the ability to translate all the energy carried by an electromagnetic shower into scintillation light. This is
obviously not the case with a sampling calorimeter in which some energy is absorbed in the passive medium.
In fact, the portion of the energy contained in an electromagnetic shower that is recorded is less for showers
that carry higher energies in sampling calorimeters in which the sampling medium has a lower Z value than
that of the active medium, and larger differences in Z produce larger differences in portions of energies
recorded between higher energy showers and lower energy showers. This is called the transition effect, and
is due to the different proportions of mechanisms through which photons and electrons deposit energy in a
medium at different energy levels, and the different Z dependencies of these different mechanisms. These
dependencies are generally examined in terms of the ratio of the energy deposited by an electromagnetic
shower component to the energy that would be deposited by a minimum ionizing particle: e/mip for electrons
or /mip for gamma photons.

18

6
6.1

Energy resolution
Sampling fluctuations

Sampling fluctuations are inconsistencies in energy measured from showers propagating through sampling
calorimeters, and lower the energy resolution of these devices. Energy resolution is the precision with
which energy deposited in a calorimeter may be measured. Resolution of sampling calorimeters is lowered
by fluctuations in the number of the low energy electrons that would deposit energy into the sampling
calorimeter as scintillation that exist in the sampling and passive portions. Sampling fluctuations are the
primary concern of resolution of sampling calorimeters. There are several equations that offer approximations
of sampling fluctuations effects on energy resolution, based on both empirical and theoretical considerations.
One such equation was derived by Amaldi and relates the contribution of sampling fluctuations to energy
resolution to the average energy lost by a minimum ionizing particle , and F (Z)hcos i which is a factor
that accounts for empirical observations related to Z dependence:
s

,
asamp = 3.2%
F (Z)hcos i
and we may find that the energy resolution of a shower with energy E due to sampling fluctuations is
r

(/E)samp =
E
[1]
This equation produces reasonably accurate results concerning energy resolutions of calorimeters with
an iron-based passive medium and a plastic active medium, and also for resolutions of those with an iron
passive medium and a liquid argon active medium, but the results are not consistent with those observed
in calorimeters utilizing other materials or those which do not contain very thick scintillator plates. Experimentally, we see that energy resolution depends on the relative thickness of the active and passive media.
The thickness of the scintillating portions of the calorimeter determines the portion of photons which deposit
their energies in a way that is observed. Apparently, an equation that accounts for the sampling
p fraction of
a calorimeter is necessary, and empirically, energy resolution is found to scale proportional to 1/fsamp . We
also find that energy resolution depends on the geometry of the boundaries between passive and sampling
media. For example, for a given sampling fraction, a calorimeter that is made of alternating plates of passive
and sampling media has less energy resolution than one made of a passive medium interspersed with scintillating fibers with thicknesses on the order of a millimeter, and thinner fibers allow for better resolutions.
This has to do with the differences in the sum of the areas of the boundaries between sampling and active
media; higher surface areas allow a higher fraction of low energy electrons to produce scintillation light. We
now find that, due to sampling fluctuations, the energy resolution of sampling calorimeters scales as
r
d/fsamp
(/E)samp = a
,
E
where d is the thickness of the sampling components.
Sampling fluctuations also affect readings from hadronic showers. The energy resolution of hadronic
showers is decreased by the fact that not all of the energy of hadronic showers is detected through scintillation,
the fact that charged hadrons affect scintillation in active media, and the fact that individual particles in
a hadronic shower generally produce higher energy signals than individual components of electromagnetic
showers. Since hadrons penetrate further into media than electromagnetic shower components, hadronic
shower components are likely to cross more boundaries between active media and passive media, andthe
statistical nature of particle collisions means that sampling fluctuations are increased by a factor of N ,
where N is the number of boundaries crossed. We also find that the increase in sampling fluctuations between
hadronic shower and electromagnetic shower readings is proportional to the thicknesses of the sampling media
measured in mean interaction lengths (int ). Experimentally, the energy resolution of hadronic showers was
found to be 0.28 times that of electromagnetic shower based solely on sampling fluctuations.
19

6.2

Fluctuations due to invisible energy

As mentioned in subsection 3.6, spallation involves the generation of invisible energy, that which is used in
the overcoming of nuclear forces when particles are ejected from nuclei and which cannot be detected by
calorimeters. The amount of energy lost as this invisible energy varies between reactions, depending on the
type of particle striking the atom and the nature of the process in which energy propagates through the
nucleus as a result of the collision; it is inherently unpredictable.

Compensation

Electromagnetic showers suffer from no effects comparable to the invisible energy loss that guarantees that
the energy response of a detector to hadronic showers will be less than the minimum ionizing energy response. In order to compensate for the inherent inefficiencies of measuring the energies of hadronic showers,
adjustments should be made so that the energy response to hadronic showers equals the energy response
to electromagnetic showers. This task is obviously impossible in homogeneous calorimeters: the energy
response to electromagnetic showers cannot be decreased below the energy response to minimum ionizing
particles using any mechanism that does not equivalently decrease the energy response of hadronic showers.
Likewise, the energy response to hadronic showers cannot possibly be increased above the energy response
to electromagnetic showers. Sampling calorimeters, however, which do not have perfect energy response to
electromagnetic showers and which present opportunities to adjust relative energy responses to hadronic
and electromagnetic showers, offer mechanisms through which the energy responses to electromagnetic and
hadronic showers may be equalized.
The ratio e/h gives the ratio of the energy response of a calorimeter to electromagnetic showers to the
energy response to hadronic showers. Without exception, e/h of homogeneous calorimeters is greater than
1 due to invisible energy phenomena. In the case where e/h is greater than one, a calorimeter is said to be
non-compensating or undercompensating. If e/h is equal to one, a calorimeter is compensating. Finally, if
e/h is less than one, the calorimeter is overcompensating. Compensation can either be performed by reducing
the energy response to electromagnetic showers, or by increasing the energy response to hadronic showers.
Methods that adjust signal strength of detectors to electromagnetic and hadronic shower components also
exist, and are called off-line compensation methods. It is important to note that no method of compensation
affects replacement of unobserved energy. As we shall soon find, there are mechanisms which may add extra
energy to the response to hadronic shower components, but there is no way to replace the energy lost.

7.1

Decreasing electromagnetic shower energy response

As mentioned in subsection 5.2, passive media made of atoms with high atomic numbers reduce the energy
response to electromagnetic showers more profoundly than they reduce the energy response to hadronic
showers. This is because the high Z materials absorb low-energy photons which may produce free electrons
through the photoelectric effect. The effectiveness of using high Z material to lower the electromagnetic
shower energy response is increased if the passive material is coated with a certain thickness of metal foil,
which prevents free electrons liberated through the photoelectric effect close to the boundaries between
sampling and passive media from escaping and depositing energy into the sampling medium. Changing the
thickness of this coating is a way to fine-tune calorimeter compensation.

7.2

Increasing hadronic shower energy response

A sampling medium containing hydrogen atoms can increase the energy response to neutrons freed from
atomic nuclei during spallation events. This is because neutrons lose their kinetic energy through elastic
collisions with atomic nuclei, while it is extremely rare for components of electromagnetic showers to lose
energy through such collisions. If the neutrons collide with a hydrogen nucleus in the sampling medium, the
nucleus will begin to move through the medium, perhaps producing signals through coulomb effects.

20

Figure 7: The Large Hadron Collider and the pre-accelerators that feed it. The linear accelerators that
impart initial accelerations upon protons and lead ions are labeled p and pb. By Arpad Horvath.

Part IV

Calorimeters at the Large Hadron Collider


8

The Large Hadron Collider

This paper shall examine calorimeters observing collisions at the Large Hadron Collider. The Large Hadron
Collider is a particle accelerator at CERN commissioned in 2008. It is the largest particle accelerator ever
constructed, and its capabilities towards energy transfer to particles exceed those of any other such device
by a factor of seven. As its name suggests, the Large Hadron Collider accelerates and collides protons, and
also lead ions. The particles are accelerated using electric forces, and guided using magnetic forces, through
a series of accelerators before being fed into the ring-shaped Large Hadron Collider as shown in figure 7. The
hadrons are produced and travel in bunches- the necessity of this will be explained below. Protons begin
their journey in a duoplasmatron, a device which creates a plasma from a gas then accelerates particles
from that plasma [10]. The protons are then accelerated through Linac-2, a linear particle accelerator,
leaving with 50 MeV of energy. They are then guided to the Proton Synchrotron Booster a ring accelerator
which accelerates them until they possess 1.4 GeV of energy and are then guided to the Proton Synchrotron
and further accelerated so that each particle possesses 26 GeV. The protons then enter the Super Proton
Synchrotron, where they receive further acceleration so as to carry 450 GeV of energy before finally being
fed into the Large Hadron Collider, which itself accelerates the particles until they possess energies of up to
7 TeV. When lead nuclei are used instead of individual protons, an ECR machine strips electrons from lead
atoms in a sample of vapor by free electrons accelerated using electromagnetic radiation. These lead ions
are accelerated through Linac 3, fed to the Low Energy Ion Ring with energies of 4.2 MeV, then continue
through the Proton Synchrotron, the Super Proton Synchrotron, and on to the Large Hadron Collider in a
21

Figure 8: A magnetic quadrupole. Electrically charged particles moving through the space in the middle of
the four magnets will experience a force that is in a direction perpendicular to that of the magnetic field
illustrated by the black lines. The result is that the particles are focused toward a point at the center of the
space surrounded by the field sources. By Wikimedia Commons user Geek3.

manner similar to that of single protons [3].


The particle bunches receive their linear acceleration from R-F cavities installed along the particle paths.
R-F cavities produce electric fields that are parallel to the hadron path and oscillate in magnitude at a
frequency such that the fields are negative as a positively charged particle bunch approaches the R-F cavity
and are positive as the bunch leaves in the opposite direction. The R-F cavities begin producing a negative
electric field again as a new bunch of particles approaches. To optimally deal with particles with higher
speeds, the R-F cavities are spaced farther apart. This is one reason why hadrons pass through several
accelerators before they are fed into the LHC: each of the accelerators is optimized for to accelerate particles
through a particular energy range. The nature of this acceleration mechanism also leads to the necessity
of the particles to be produced and to travel in bunches: the R-F cavities can only impart force onto the
particles in the same direction in which they are already traveling as they move through the cavities if the
particles move through the center of the cavity the moment its electric field changes direction. The majority
of particles in a continuous beam of particles would receive forces in alternating directions. The particles are
guided through and between the particle accelerators using dipole magnets. The areas of particle bunches
normal to their direction of travel are minimized using quadrupole magnets, as illustrated in figure 8.
The particles travel around the Large Hadron Collider in two parallel beams in opposite directions guided
by 1232 dipole magnets and focused by 392 quadrupole magnets. They are kept at a constant speed (after the
initial acceleration) by R-F cavities canceling out the energy lost by the particles through electromagnetic
radiation associated with their centripetal accelerations through the curve of the ring. The two beams
traveling through the LHC intersect at four points in the vicinities of four detectors designed to collect
data concerning the results of particle collisions. Even though the particle bunches contain extremely high
numbers of particles focused into beams as narrow as one millimeter, the sizes3 of the particles are such that
3 The definition of size as it is used here involves a discussion of quantum mechanics and is therefore beyond the scope of
this paper.

22

the probability that a given particle will be involved in a collision as it passes through a beam intersection
point is extremely low. That probability is further reduced by the fact that the particles tend to deflect
away from each other due to the electric charges they carry. Therefore, collision events are distributed along
a time range of several hours. However, due to the large number of hadrons inserted into the LHC, about
20 particle collisions occur every time bunches intersect.
The term luminosity is important in the study of collision frequency. It is defined as the number of
particles passing through a unit of area per some unit of time multiplied by the number of particles in that
same area with which collisions may be experienced per the same unit of time. For the situation described
above, wherein n bunches of N particles travel in opposite directions around a ring with a revolutionary
frequency of f and having areas normal to the directions of travel of A, the luminosity L is given by the
equation
N2
L = fn
A
In the Large Hadron Collider, normal operation involves 2808 bunches per beam initially containing
1.1 1011 protons each traveling around the ring 11245 times per second. At the collision points, the beams
are focused such that their areas normal to their directions of travel are 7 micrometers. According to the
above equation, the luminosity of the Large Hadron Collider is therefore
The collision frequency can be calculated from the luminosity. To do this, we must know the how small
the area two particles must simultaneously occupy in order for a collision to occur. This area is known as
cross section and depends on the energy possessed by the particles. The cross section of a collision is a
hypothetical area around a particle through which another particle must pass through in order for a collision
event to take place, and collisions involving larger cross-sections are more likely to occur given a certain
luminosity. When protons travel through the Large Hadron Collider at the maximum speed the device is
able to accelerate them to, the cross section of all types of collisions between two protons traveling in opposite
directions is 110 millibarns4 , but the cross section of inelastic collisions, the only collisions the instruments
at the Large Hadron Collider are designed to detect, is only 60 millibarns. The number of inelastic collisions
that occur per second can be found by multiplying the luminosity by the cross section. The Large Hadron
Collider produces 6 108 collisions per second, or one collision every 1.66 nanoseconds on average, or about
20 collisions per bunch crossing [2].
When the occurrences of collisions within the collider fall below an acceptable frequency, the magnetic
dipoles controlling the path of the beam are adjusted so as to cause any hadrons that may remain in the
beam to be jettisoned from the ring and to be safely absorbed by a mass of concrete in the vicinity of the
ejection point. The LHC may then be prepared for a new batch of particles.
ATLAS (A Toroidal LHC ApparatuS) and CMS (Compact Muon Solenoid) are detectors that collect
data at two of the LHCs beam intersection points and the remainder of this paper includes discussions
concerning them. The other beam intersection points exist to produce collisions to be observed by detectors
known as ALICE (A Large Ion Collider Experiment) and LHCb (Large Hadron Collider beauty), neither of
which are discussed in this paper. The unprecedented energy at which the Large Hadron Collider collides
particles and frequency of those collisions combine with the goals of the experiments involving the detectors
to produce a unique set of requirements on the detectors.

Geometry

Pseudorapidity is a measurement of the angle from the beams traveling through a particle accelerator. It is
given by the symbol and is defined

 

= ln tan
.
2
4 A barn is a unit of area equal to 1028 square meters, and derives its name from comments made by Manhattan Project
researchers that the uranium nucleus was as big as the side of a barn [11].

23

Figure 9: The components of the Compact Muon Solenoid. From JINST.

Angles between rays that share an endpoint located on the beam and are perpendicular to the beam are
given by the symbol . We shall use r for the distance from the beam, and for the cartesian distance from
the collision point. A vector r is one that points radially outward from the beam (in a direction perpendicular
to the beam).

10

The Compact Muon Solenoid

The Compact Muon Solenoid features several particle detection mechanisms surrounding a point of intersection between the two beams traveling through the Large Hadron Collider. The individual detectors are
mostly cylinders whose axes of extrusion are parallel to the beams, and are concentric so that a single line
drawn at any angle perpendicular to the beam paths passes through most of the detectors, as seen in figure
9. The detector closest to the collision point is the inner tracking system. This systems purpose is to record
the paths of particles while affecting those particles as little as possible. Surrounding the inner tracking
system is the electromagnetic calorimeter, a homogeneous calorimeter designed to measure the energy of
electromagnetic showers only. Outside the electromagnetic calorimeter is the hadron calorimeter, a sampling calorimeter that measures the energy of hadronic showers. The component surrounding the hadron
calorimeter is a superconducting magnet which allows the detectors to obtain momentum information from
charge-carrying particles based on the amount of curvature the magnetic field effects upon them. The coil of
the superconducting magnet is surrounded by the muon system, components designed to detect properties

24

Figure 10: The components of CMSs electromagnetic calorimeter. From JINST.

of muons that pass through the inner components. The muon system is interspersed with sections of the
superconducting magnets return yoke, the iron sections that connect to the poles of the magnet. There
are also detectors that do not surround the collision point. Like the other detectors, these are cylinders
surrounding the beam paths, but they are situated so that they are intersected only by particles that leave
the collision point on paths that are close to being parallel to the original beam paths. These detectors are
CASTOR (Centauro And Strange Object Research), a quartz calorimeter, and the zero degree calorimeter, a
set of two hadron and electromagnetic calorimeters situated between the beam pipes that are able to examine
particles that leave the collision point in paths that are truly parallel to the beams. We shall, of course,
focus on the calorimetric components of the CMS. Components of the CMS must be designed to resist the
effects of radiation and magnetic fields.

10.1

Electromagnetic calorimeter

The electromagnetic calorimeter includes a barrel that is a cylinder that surrounds the collision point and
ranges in pseudorapidity from -1.479 to 1.479, two flat endcaps normal to the beam ranging in pseudorapidity
from 1.479 to 3.0 and -1.470 to -3.0, and two preshower detectors, with one ranging ranging in pseudorapidity
from 1.653 to 2.6 and the other ranging from -1.653 to -2.6.
10.1.1

Main barrel

The main barrel of the electromagnetic calorimeter is a homogeneous calorimeter that utilizes lead tungstate
(PbWO4 ) crystals as a stopping and scintillation medium. The crystals were chosen for many reasons.
Their short radiation length of 0.89 centimeters and their small Moli`ere radius of 2.2 centimeters allow
large portions of electromagnetic showers to be contained in relatively small volumes. The time it takes for
scintillation light to decay to very low levels is on the same order of magnitude as the LHCs bunch crossing
period, which facilitates resolution between events separated in time. This consideration mitigates pileup
noise, effects from the extremely frequent penetration of particles through parts of the CMS. The crystals

25

are transparent to visible light, which allows scintillation light to pass through the crystals to the detection
mechanisms.
When viewed from the side facing the beam, the crystals are roughly square in shape. The crystals
are tightly packed in a cylindrical array, which necessitates that their broadnesses increase as r does. The
degree of tapering varies slightly with their position in the coordinate. In order to reduce the probability
of particles falling through the cracks between crystals, they are twisted three degrees on axes parallel to r
so that the cracks between them do not create planes parallel to the beam, and are also angled such that
lines normal to their faces that most nearly face toward the beam are three degrees from r. The tapering of
the crystals has been shown experimentally to reduce energy resolution because it causes a dependence of
energy response to showers on the position of the showers in the crystals. This problem, which is related to
focusing of scintillation light, is mitigated by depolishing one face of each crystal. Another contributor to
such a dependence is the energy lost by light as it travels from the scintillation points to the photodetectors
at the back ends of the crystals: there would be less energy response to showers that develop farther from the
photodetectors [6]. There are 360 crystals arrayed through and 170 crystals arrayed through , meaning
that the barrel houses 61,200 crystals. The positional granularity of this portion of the calorimeter is the
volume of one crystal (detectors cannot discern where an event takes place within a crystal). The crystals
are 23.0 centimeters long, and so they are 23.0 centimeters 0.89 centimeters, or 25.8 radiation lengths
long.
The number of photons emitted by these crystals during a scintillation event caused by a given energy
deposit depends on the temperature of the crystals: less scintillation light is produced at higher temperatures.
In order for this phenomenon to not affect energy resolution, the crystals must be kept at a constant
temperature, even as heat is introduced into the crystals by electronics attached to them. This requirement
is met through insulating foam separating the crystals from readout electronics, and a thermal screen through
which water kept at a constant temperature is pumped separating the crystals from the silicon tracker.
Avalanche photodiodes detect the scintillation light from the lead tungstate crystals. Two such devices
are attached to each crystal. Readings received from the photodiodes are also affected by temperature:
higher temperatures result in less amplification. Therefore, the photodiodes must also be kept at a constant
temperature, and one in every ten photodiode has a temperature sensor embedded within it. The photodiodes
are designed to avoid the production of false readings as particles that pass through them on their way to
the detectors surrounding the electromagnetic calorimeter. The photodiodes amplifications of their received
signals is dependent on the voltage used to propel electron avalanches, and so the photodiodes require a
custom high voltage power supply.
10.1.2

Endcaps

The endcaps utilize scintillators made of the same material as those in the barrel. The endcaps each house
7324 crystals arranged in a flat rectangular grid, the boundary of which is roughly a circle. The crystals in
the endcaps are not tapered. The photodetectors attached to these crystals are vacuum photodiodes, which
are similar to photomultiplier tubes with a single dynode.
10.1.3

Preshower detector

The preshower detectors are located between the electromagnetic calorimeter endcaps and the collision point.
Their purpose is to facilitate the identification of neutral pions in the endcaps.

10.2

Hadron calorimeter

The hadron calorimeter is set of calorimeters used to measure the energy of showers initiated by particles
that penetrated the electromagnetic calorimeter completely.
The main part of the hadron calorimeter is the barrel, which is located in the cylindrical volume between
the electromagnetic calorimeter and the superconducting magnet. This barrel is a sampling calorimeter,
which uses layers brass plates oriented so that their normal vectors are perpendicular to the particle beam

26

as passive media, and plastic scintillators as active media. There are 16 layers of scintillators separated by
brass plates. The flat scintillation and brass layers create concentric polygons when viewed from the ends
of the barrel cylinder, and layers cover larger areas as r increases. The outermost layer is thicker to prevent
leakage of late-developing showers. The innermost layer is a scintillator, and its purpose is to detect showers
that developed in the material between the electromagnetic calorimeter and the hadron calorimeter. In total,
there are roughly 70,000 separate scintillation tiles in the barrel of the hadron calorimeter, each connected
to a separate photodetector.
Outside the superconducting magnet, at low pseudorapidity values, is the outer hadronic calorimeter,
in place to capture shower components not absorbed by any of the calorimeters closer to the beam and to
ensure that the calorimeters are hermitic. The outer hadronic calorimeter uses plastic scintillator plates
as its sampling medium, and the iron of the superconducting magnet as its passive medium. It is a sampling calorimeter with two sampling layers: the first sampling layer is located outside the solenoid of the
superconducting magnet, and the second layer is located outside the magnets first return yoke. The outer
calorimeter extends through a low pseudorapidity region because particles leaving the collision point at low
pseudorapidities are those which traverse through less material as they travel through a cylindrical volume
extending between two constant r values. Mathematically, a particle leaving a collision point on a path that
makes an angle of with the particle beam passing through a cylinder with an inner radius of r1 and an
outer radius of r2 is (r2 r1 ) sin .
The endcaps of the hadron calorimeter are, again, sampling calorimeters that utilize plastic plates as the
scintillation medium and brass as the passive medium.
Light from the plastic scintillator plates sandwiched between the brass passive media in the hadron
calorimeter is carried to photodetectors though optical fibers run through grooves in the scintillator plates.
These fibers are made of light carrying media coated with reflective cladding. The fibers carry the light to
photodiodes for detection.

10.3

Zero degree calorimeter

To ensure that the CMS is a hermetic detector, there exist electromagnetic and hadron calorimeters that
span pseudorapidity between 8.3 and infinity and between -8.3 and negative infinity: they are able to detect
showers that develop parallel to the beam path. This is possible because the zero degree calorimeters are
located between the two beam pipes. The importance of the hermetic nature of the CMS in all directions
lies in one of the missions of the system. In order to detect particles which are practically invisible to any
of the detectors, their existences may be inferred by detecting unequal energy and momentum distributions
of detectable particles. As a crude example, if we observe that more energy was carried by detectable
particles into a certain half of a sphere with a center at the beam intersection point, then we would know
that invisible particles carried energy into that spheres opposite hemisphere. This type of analysis is only
possible if detectable particles do not escape from the CMS without being detected [8].

10.4

Calibration of the CMS

The Compact Muon Solenoid was calibrated using a relatively low-energy test beam of particles pointed
at production plates which produced particles directed into the detectors designed to mimic trajectories
of particles moving away from the beam intersection point. The test beams were produced in the SPS,
and entered the detectors in two modes: a high energy mode, ranging from 15 to 350 GeV/c, and a very
low energy mode, carrying particles with energy less than 9 GeV/c. The particles were first passed through
several counters to identify their types with certainty. Time of flight counters measured the speeds of particles
by simply measuring the time it took for the particles to traverse a certain distance.
The barrel sections of the electromagnetic and hadron calorimeters were calibrated using beams of 50
GeV/c electrons, and the outer hadron calorimeter was calibrated using a 150 GeV/c electron beam.
The relative responses to different particles throughout the collection of CMSs detectors was measured.
It was found that the response to protons is 47 percent of the response to electrons, and that the response to
charged pions is 62 percent of the response to electrons. It was found that the response to positively charged
27

pions was larger than the response to negatively charged pions; this difference is due to the decay products
of the particles. Furthermore, the difference in the responses to these two particles increases as their energies
increase. While the electromagnetic calorimeter was designed to measure showers started by electrons,
positrons, and photons, it was expected that a portion of protons and hadrons would initiate showers in this
detector. It was found that 59 percent of pions that penetrated the electromagnetic calorimeter barrel and
65 percent of the protons that penetrated started a shower reaction. However, these results were produced
using the low energy test beam, and it was found that the portion of energy deposited by hadrons in the
electromagnetic calorimeter barrel decreased as the energy of the particles was increased.
The differences in the electromagnetic/hadronic shower energy responses between the electromagnetic
calorimeter barrel, the hadron calorimeter barrel, and the outer hadron calorimeter were measured so that
corrections may be applied to determine the total energy of particle systems passing through all of these
detectors. Corrections were made for non-linear energy responses: those that changed as a function of energy,
once these functions were determined through calibration.
An important step was to parameterize the ratio of energy response to pions to the energy response to
electrons in the hadron calorimeter barrel as a function of the average energy of the particles in the barrel
[13].

11

ATLAS

Unlike the calorimeters of the Compact Muon Solenoid, the calorimeters that are part of the ATLAS system are mainly designed to measure ionization. This is done by detecting the effect that hadronic and
electromagnetic showers have on the electric charge of samples of liquid argon.

11.1

Electromagnetic calorimeters

The innermost calorimeter of ATLAS is its electromagnetic calorimeter. It features a liquid argon sampling
medium housed by accordion-shaped lead plates. The waves of the accordion shape are oriented so that
particles traveling in in the r direction pass through several waves, and so that the crest of a wave makes a
circle around the particle beam in the barrel portion of the electromagnetic calorimeter. In the electromagnetic calorimeters endcaps, the waves of the accordion geometry run so that the crests of waves extend in
lines perpendicular to the particle beam. In the endcaps, the gaps between the accordion-shaped plates vary
with so that the energy response of this part of the calorimeter is not dependent on . Each accordion
layer consists of a lead stopping medium, and a liquid argon sampling medium. Between the lead and the
liquid argon is a stainless steel coating glued to the lead. Housed in the middle of the liquid argon sampling
chambers are plates that hold the readout electronics. The outer layer of these plates is copper which is
connected to a high-voltage power supply, and which attracts electrons freed through ionization by particles
passing though the liquid argon. This layout is seen in figure 12.

11.2

Hadron calorimeters

The ATLAS hadron calorimetry system features both scintillator-based hadronic shower detectors and liquid
argon based ionization detectors.
A scintillator tile-based sampling hadron calorimeter surrounds the electromagnetic accordion calorimeter
barrel. Steel is used as the stopping medium. Light from the scintillator tiles is carried to photomultiplier
tubes by optical fibers running along the edges of the tiles. There are eleven layers of scintillating tiles, and
the area of the tiles increases as r does. This barrel features a total of 460,000 separate scintillating tiles.
The endcaps of the hadron calorimeter system are liquid argon ionization detectors, but unlike those of the
electromagnetic calorimeter, they feature flat absorber plates. There are also forward calorimeters attached
to ATLAS that utilize liquid argon ionization detectors, and detect particles traveling at pseudorapidities
ranging to infinity. Like the CMS, ATLAS is intended to be a hermetic detector for detectable particles. In

28

Figure 11: The accordion sampling geometry of ATLASs calorimeters. From JINST.

=0

Square cells in
Layer 2

Strip cells in Layer 1

29

Figure 12: A closeup of a liquid argon chamber. From ATLAS Technical Design Report CERN/LHCC 96-40.

30

the forward calorimeter, the liquid argon medium surrounds electrodes in the shape of rods that run through
the medium parallel to the beam direction [7].

References
[1] Ugo Amaldi. Fluctuations in calorimetry measurements. Physica Scripta, 23(4a), 1981.
[2] Unknown authorship.
LHC Collisions.
lhc-machine-outreach/collisions.htm.

http://lhc-machine-outreach.web.cern.ch/

[3] Michel Chanel. LEIR: THE LOW ENERGY ION RING AT CERN.
[4] K. Nakamura et al. Particle data group. Review of Particle Physics, 2010.
[5] K. Nakamura et al. Passage of particles through matter. Review of Particle Physics, 2010.
[6] D.J. Graham and C. Seez. Simulation of Longitudinal Light Collection Uniformity in PbWO4 Crystals.
[7] Institute of Physics Publishing and SISSA. The ATLAS Experiment at the CERN Large Hadron Collider,
August 2008.
[8] Institute of Physics Publishing and SISSA. The CMS experiment at the CERN LHC, August 2008.
[9] William Leo. Techniques for Nuclear and Particle Physics Experiments. Springer-Verlag, New York
Berlin Heidelberg, 1994.
[10] Richard Scrivens. Cern hadron linacs. http://linac2.home.cern.ch/linac2/default.htm, January
2008.
[11] Doreen Wackeroth. Cross section. High Energy Physics made painless, 2002.
[12] Richard Wigmans. Calorimetry: Energy Measurement in Particle Physics. Clarendon Press, Oxford,
2000.
[13] Efe Yazgan. The CMS barrel calorimeter response to particle beams from 2 to 350 gev/c. In Journal
of Physics: Conference Series 160, 2009.

31

S-ar putea să vă placă și