You are on page 1of 26

Radiation Dosimetry

Subhalaxmi Mishra and T. Palani Selvam

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
Interaction Coefficients and Related Quantities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Cross Section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Linear Attenuation Coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Mass Attenuation Coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Mass Energy-Transfer Coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Mass Energy Absorption Coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Mass Stopping Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Linear Energy Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Radiation Chemical Yield . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Ionization Yield in a Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Mean Energy Expended in a Gas Per Ion Pair Formed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Dosimetric Quantities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Exposure (X) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Kerma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Energy Deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Energy Imparted . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Mean Energy Imparted . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Absorbed Dose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Charged Particle Equilibrium (CPE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
General Principles of Absorbed Dose Determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Cavity Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Detectors Used in Radiation Dosimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Ionization Chambers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Calorimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Chemical Dosimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Film Dosimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Solid-Sate Detectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Properties of Dosimeters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Accuracy and Precision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

S. Mishra (*) · T. P. Selvam


Radiological Physics and Advisory Division, Bhabha Atomic Research Centre, Mumbai, India
e-mail: subha@barc.gov.in; pselvam@barc.gov.in

© Springer Nature Singapore Pte Ltd. 2023 1


D. K. Aswal et al. (eds.), Handbook of Metrology and Applications,
https://doi.org/10.1007/978-981-19-1550-5_116-1
2 S. Mishra and T. P. Selvam

Linearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Dose Rate Dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Energy Dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

Abstract
Radiation dosimetry addresses determination of energy deposited in a given
medium quantitatively by either calculations or measurements. The objective of
radiation dosimetry is to determine absorbed dose in a medium using a detection
device, i.e., a radiation detector. Toward this goal, this chapter presents general
principles of radiation dosimetry. Interaction coefficients and quantities related to
radiation dosimetry of ionizing radiation are included. The chapter also presents
various cavity theories in detail that relate absorbed dose measured in one
medium to other. Principles of radiation dose measurements involving various
dosimeters are also discussed in this chapter.

Introduction

Ionizing radiation is classified as directly ionizing and indirectly ionizing. Directly


ionizing radiations are charged particles such as electrons, positrons, protons, neg-
ative π mesons (π), alpha particles, and heavy ions such as 12C, 14N, 16O, 56Fe, etc.
Indirectly ionizing radiations are neutral particles such as photons (X-rays, gamma
rays) and neutrons. Directly ionizing radiation interacts with the orbital electrons of
atoms in the medium through direct Coulomb interactions and deposits energy in the
medium. Energy deposition of indirectly ionizing radiation in matter involves two
steps. Step 1: A charged particle is produced in the medium. For example, photon
interactions with the medium result in production of electrons or positrons, and
neutron interactions produce protons, alpha particles, or heavier ions. Step 2: The
secondary charged particles produced in Step 1 deposit energy through collisions
with the orbital electrons of the atoms in the medium. Radiation dosimetry addresses
determination of energy deposited in a given medium quantitatively by either
calculations or measurements. The objective of radiation dosimetry is to determine
absorbed dose in a medium using a detection device. Dosimetric quantities provide a
physical measure of the region of interest which is correlated to the actual or
potential effects of ionizing radiation. Dosimetric measurements or calculations
require various specifications of the radiation field at the point of interest.
Definitions of the fundamental radiation quantities and their units are the core
mission of the International Commission on Radiation Units and Measurement
(ICRU). Several reports on the quantities and units in radiation physics, dosim-
etry, and radiological protection such as Report 10a, Radiation Quantities and
Units (1962), Report 51, Quantities and Units in Radiation Protection Dosimetry
(1993), Report 60, Fundamental Quantities and Units for Ionizing Radiation
(1998), and Report 85, Fundamental Quantities and Units for Ionizing Radiation
(Revised) (2011) are made available by the ICRU. ICRU Report 85 (2011)
supersedes the previously published reports related to radiation quantities and
Radiation Dosimetry 3

units. The definitions of some of the radiometric, dosimetric, and protection


quantities of interest are described in the following sections.

Interaction Coefficients and Related Quantities

Cross Section

The cross section, σ, of a target entity, for a particular interaction produced by


incident directly or indirectly ionizing radiation (charged or uncharged particles)
of a given type and energy, is defined as (ICRU 2011):

N
σ¼ ð1Þ
Φ
where N is the mean number of interactions per target entity and Φ is fluence of
particles. σ is expressed in m2.

Linear Attenuation Coefficient

Linear attenuation coefficient (ICRU 2011), μ, is defined as the probability of


interaction of S-ray and gamma photons per unit path length dl that they travel in
an absorber and can be expressed as follows:

1 dN
μ¼ ð2Þ
dl N
where N is the number of photons traversing a distance dl in the absorber and dN is
number of photons that undergo interactions in the absorber. μ depends upon the
incident energy of photons, atomic number, and mass density of the absorber. μ is
expressed in unit of cm1. Inverse of μ is mean free path τ.
μ is a useful parameter to determine the transmitted intensity of X-ray or gamma
photons when they traverse an absorber thickness of x. The differential equation for
attenuation can be expressed as below:

I ¼ I 0 eμx ð3Þ

where I is the transmitted intensity for an absorber of thickness x and Io is the initial
intensity of the radiation incident on the absorber. It can be observed that the
transmitted intensity can never be equal zero for any thickness of absorber.

Mass Attenuation Coefficient

As μ is directly proportional to the mass density, ρ, of the absorber material, to


eliminate this density dependence another coefficient, mass attenuation coefficient,
4 S. Mishra and T. P. Selvam

μ/ρ, is used. μ/ρ of a material, for uncharged particles of a given type and energy, is
defined as (ICRU 2011):

μ 1 1 dN
¼ ð4Þ
ρ ρ dl N
where dN/N is the mean fraction of the particles that experience interactions in
traversing a distance dl in the material of density ρ. The unit of μ/ρ is cm2/g.

Mass Energy-Transfer Coefficient

Let R be the incident radiant energy of the uncharged ionizing particles (X-rays,
gamma rays, and neutrons) incident on a material of density ρ and dRtr the mean
energy that is transferred to kinetic energy of charged particles by interactions of
the uncharged particles in traversing a distance dl in the material, and then the
mass energy-transfer coefficient, μtr/ρ, is defined as (ICRU 2011):

μtr 1 1 dRtr
¼ ð5Þ
ρ ρ dl R
μtr/ρ is expressed in cm2/g.
Let hν be the energy of the photon beams traversing in a material having density ρ
and Etr is the mean fraction of photon energy transferred per unit thickness of the
absorber to the charged particles that appears as kinetic energy. Energy-transfer
coefficient, μtr, and mass energy-transfer coefficient, μtr/ρ, can be written as follows:

μtr μ Etr
¼ ð6Þ
ρ ρ hυ
Etr
ðμtr Þ ¼ μ

In the case of neutrons,

μtr μ ΔEk
¼ ð7Þ
ρ ρ Ek
where ΔE
Ek is the fraction of energy transferred to the charged particles by the incident
k

neutrons of kinetic energy Ek.

Mass Energy Absorption Coefficient

The secondary charged particles generated by the photons may lose their energy
through radiative processes such as bremsstrahlung, in-flight annihilation, and
Radiation Dosimetry 5

fluorescence radiations while slowing down in the medium (ICRU 2011). Such loss
is not part of energy deposition. Hence, a separate quantity mass energy absorption
coefficient μen/ρ is defined as follows:

μen μ
¼ ð1 gÞ tr ð8Þ
ρ g

where g is the fractional energy loss through abovementioned radiative processes.


g depends on the type of material and energy of photons. g is negligible for low-energy
photons but significant for high-energy photons and high atomic number materials.

Mass Stopping Power

The quantity stopping power is used in the context of charged particles. The mass
stopping power, S/ρ, of a material, for charged particles of a given type and energy, is
defined as (ICRU 2011):

S 1 dE
¼ ð9Þ
ρ ρ dl
where dE is the mean energy lost by the charged particles in traversing a distance dl
in the material of density ρ (ICRU 2011):

S/ρ is expressed in MeV-cm2/g or equivalently J-m2/kg.


The quantity S ¼ dE/dl is linear stopping power (MeV/cm or J/m).

The charged particles lose energy in the absorber through (a) ionization (includes
both soft and hard collisions) and excitation with orbital electrons; (b) emission of
bremsstrahlung photons in the fields of nucleus or orbital electrons, in-flight anni-
hilation, and fluorescence radiations; and (c) elastic Coulomb interactions in which
recoil energy is imparted to atoms. Hence, S/ρ has three components as below:
     
S 1 dE 1 dE 1 dE 1 dE
¼ ¼ þ þ ð10Þ
ρ ρ dl ρ dl el ρ dl rad ρ dl nuc

where
 
1 dE
ρ  dl el is mass electron or mass collision stopping power
1 dE
ρ  dl rad is mass radiative collision stopping power
1 dE
ρ dl nuc is mass nuclear stopping power

Linear Energy Transfer

When charged particles propagate in a medium, energy is transferred to orbital


electrons through soft and hard collisions. In soft collisions, the energy transfer is
6 S. Mishra and T. P. Selvam

minimum, whereas in hard collisions energy transfer is maximum. The linear energy
transfer (LET) or restricted linear electronic stopping power, LΔ, of a material, for
charged particles of a given type and energy, is defined as (ICRU 2011):

dE Δ
LΔ ¼ ð11Þ
dl
where dEΔ is the mean energy lost by the charged particles due to electronic
interactions in traversing a distance dl. Here, the mean energy lost excludes the
mean sum of the kinetic energies in excess of Δ. If there is no threshold or energy
cutoff is imposed, it is called unrestricted stopping power. This is denoted as L1.
Note that L1 is always greater than LΔ.

Radiation Chemical Yield

The radiation chemical yield, G(x), of an entity, x, is defined as (ICRU 2011):

nð x Þ
G ðxÞ ¼ ð12Þ
e
where n(x) is the mean amount of substance of that entity produced, destroyed, or
changed in a system by the mean energy imparted, e, to the matter of that system.
G(x) is expressed in units of mol/J (ICRU 2011). The mole is the amount of
substance of a system that contains as many elementary entities as there are
atoms in 0.012 kg of 12C (ICRU 2011). Generally, the value of e is 100 eV, and
therefore the G value is expressed in (100 eV)1. A G value of 1 (100 eV)1
corresponds to a radiation chemical yield of approximately 0.1036 μmol/J
(ICRU 2011).

Ionization Yield in a Gas

The ionization yield in a gas, Y, is defined as (ICRU 2011):

N
Y¼ ð13Þ
E
where E is the initial kinetic energy of a charged particle of a given type which is
completely dissipated in the gas and N ¼ q =q. Here, q is the mean total liberated
charge of either sign and q is the elementary charge. Y is expressed in J1 and is
sometimes expressed in eV1 (ICRU 2011). From the definition of Y, it may be noted
that (a) N includes charge produced by bremsstrahlung or other secondary radiations
emitted by the initial and secondary charged particles and (b) charge of the initial
charged particle is not included in N, as this charge is not liberated in the energy
dissipation process.
Radiation Dosimetry 7

Mean Energy Expended in a Gas Per Ion Pair Formed

The mean energy expended in a gas per ion pair formed, W, is reciprocal of Y and is
defined as (ICRU 2011):

E
W¼ ð14Þ
N
where E and N have the usual meaning as mentioned above (section “Ionization
Yield in a Gas”). W is expressed in J and is sometimes expressed in eV (ICRU 2011).

Dosimetric Quantities

When radiation interacts with matter through different processes, energy is imparted
to the medium. Some of the radiation dosimetric quantities are described below
(ICRU 2011).

Exposure (X)

Exposure (X) is related to the ability of photons (X-rays and gamma rays) to ionize
air medium; thus, it is defined only for photons and air as the medium. When photon
is incident on a mass dm of dry air, it produces an ion pair (electron and positive ion)
through ionization process. X is defined as (ICRU 2011):

dQ
X¼ ð15Þ
dm
where dQ is the absolute value of the mean total charge of the ions of one sign
produced when all the electrons and positrons liberated or created by photons
incident on a mass dm of dry air are completely stopped in dry air. The unit of X is
Coulomb per kilogram (C/kg). Roentgen (R) is the commonly used unit of X. The
relation between SI unit of X and R is 1 R ¼ 2.58  104 C/kg.

Kerma

KERMA (kinetic energy released per unit mass), K, is defined for uncharged
ionizing radiations such as photons and neutrons for any medium. It is defined as
(ICRU 2011):

dE tr
K¼ ð16Þ
dm
8 S. Mishra and T. P. Selvam

where dEtr is the mean sum of the initial kinetic energies of all the charged particles
liberated in a mass dm of a material by the uncharged particle incident on dm (ICRU
2011). KERMA refers to the kinetic energy of the charged particles such as protons,
electrons, and positrons generated from indirectly ionizing radiation such as photons
or neutrons. The quantity dEtr includes the kinetic energy of the charged particles
emitted in the decay of excited atoms/molecules (e.g., Auger, Coster-Kronig, shake-
off electrons) or in nuclear de-excitation or disintegration (ICRU 2011). KERMA is
a non-stochastic quantity and is applicable only for indirectly ionizing radiations
such as photons and neutrons. The special unit of K is Gray (Gy).
Let E be the energy of uncharged particle and Φ be the fluence; then K is related to
Φ as follows:
   
μtr μ
K ¼ ΦE ¼ Ψ tr ð17Þ
ρ ρ

where μtr/ρ has the usual meaning as defined in section “Mass Energy-Transfer
Coefficient” and Ψ ¼ ΦE is the energy fluence of these uncharged particles. In
radiation dosimetry, K is generally expressed in terms of fluence of uncharged
particles Φ(E) as below:
ð
K ¼ ΦðEÞ E ðμtr ðEÞ=ρÞdE ð18Þ

The total kerma, K, is the summation of two components, i.e., collision kerma,
Kcol, and the radiative kerma, Kcol, and can be expressed as:

K ¼ K col þ K rad ð19Þ

Kcol is due to energy loss through ionization and Krad is due to energy loss through
bremsstrahlung photons.

Energy Deposit

The energy deposit, εi (a stochastic quantity, expressed in J or eV), is the energy


deposited in a single interaction, i, which is defined as (ICRU 2011):

ei ¼ ein  eout þ Q ð20Þ

where εin is the energy of the incident ionizing particle, εout is the sum of the energies
of all charged and uncharged ionizing particles leaving the interaction, and Q is the
change in the rest energies of the nucleus and of all elementary particles involved in
the interaction (Q > 0, decrease of rest energy; Q < 0, increase of rest energy). Note
that (a) εin and εout do not include the rest energies and (b) interactions with atomic
electrons resulting in atomic excitation do not result in a change in rest energies of
the nucleus or of elementary particles and thus Q ¼ 0, which is due to the fact that
there is subsequent de-excitation process.
Radiation Dosimetry 9

Energy Imparted

The energy imparted, ε (a stochastic quantity, expressed in J or eV), to the matter in a


given volume is sum of all energy deposits in the volume (ICRU 2011):
X
e¼ ei ð20Þ
i

where summation is performed over all the energy deposits, εi, in that volume. Note
that the term energy deposition event denotes the imparting of energy to matter by
correlated particles (e.g., a proton and its secondary electrons, an electron-positron
pair, or the primary and secondary particles in nuclear reactions).

Mean Energy Imparted

The mean energy imparted, e (a non-stochastic quantity, expressed in eV or J), to the


matter in a given volume is defined as (ICRU 2020):
X
e ¼ Rin  Rout þ Q ð21Þ

where Rin is the mean radiant energy of all charged and uncharged ionizing particles
that enter the volume, Rout is the mean radiant energy of all charged and uncharged
ionizing particles that leave the volume, and Q is the mean sum of all changes of
the rest energy of nuclei and elementary particles that occur in the volume. Note that
Q > 0, decrease of rest energy, and Q < 0, increase of rest energy.

Absorbed Dose

Absorbed dose, D (non-stochastic quantity), applicable to both indirectly and


directly ionizing radiations is defined as (ICRU 2011):

de
D¼ ð22Þ
dm
where de is the mean energy imparted by ionizing radiation to matter of mass dm.
D is expressed in J/kg1. D is also expressed in Gray (Gy).

Charged Particle Equilibrium (CPE)

When indirectly ionizing radiations such as photons and neutrons propagate in a


medium, they transfer their energy to the charged particles (electrons in the case of
photons and protons in the case of neutrons). Due to finite range of these charged
10 S. Mishra and T. P. Selvam

particles, energy deposition does not take place at the same location where these
charged particles are generated. In the case of photons, the secondary electrons lose
energy through ionization and production of bremsstrahlung photons (known as
radiative loss). Collision kerma, Kc, does not include radiative loss as the brems-
strahlung photons escape from the point where they are produced. For photons, Kc at
a point in the medium depends on the photon fluence at that point. When a parallel
beam of photons is incident on the medium, the photon fluence decreases with depth
due to exponential attenuation. Hence, Kc also decreases with depth. In the case of
absorbed dose, D, secondary electrons generated at earlier depths will deposit their
energy at later depths. Hence, D increases with depth initially and decreases after
some depth. This decrease is due to the fact that photon fluence decreases due to
exponential attenuation in the medium. Hence, there exists a depth, which is equal to
the maximum range of secondary electrons, denoted as xm, at which D is maximum.
At this point, charged particle equilibrium exists where both Kc and D are equal, i.e.,
D/Kc ¼ 1. Beyond xm, at any depth x, D is slightly larger than Kc because D depends
on photon fluence at earlier depth, i.e., x  x0 , whereas Kc depends on photon fluence
at depth x; here, x0 represents maximum range of the secondary electron. For depths
beyond xm, the ratio D/K > 1. The degree of difference between D and Kc depends
on photon energy and the type of medium. The depth dose curve of photons
therefore shows D/K < 1 in the buildup region (0 to xm), D/K ¼ 1 at xm, and D/
K > 1 at depths beyond xm. At the 60Co photon energies, xm ¼ 5 mm in water. At
very low-energy photons such as 25 keV, there will not be a buildup region, and
therefore both D and Kc will be same at all depths. This is because range of
secondary electrons is too small and they deposit energy at the spot where they are
generated. At such low-energy photons and since the medium is of low atomic
number, there is no radiative loss, and hence K equals Kc.

General Principles of Absorbed Dose Determination

Cavity Theories

In dosimetry of ionizing radiation, one needs a device that has a sensitive medium to
measure response in terms of absorbed dose to the incident ionizing radiation. This
device is known as radiation dosimeter. Cavity theory is then applied that relates
absorbed dose to the cavity to absorbed dose to the medium of interest. Here, cavity
implies that a small mass of material which is sensitive to ionizing radiation
embedded in the medium whose elemental composition and mass density are
generally different from that of the medium. The cavity may be in the form of
solid (diode, TLD, OSL etc.) or gaseous form (ionization chamber filled with air; air
here is known as cavity).

Bragg Gray Cavity Theory


It was W. H. Bragg who first discussed the relation between absorbed doses of two
different media in 1910 (Bragg 1910). Later, L. H. Gray made quantitative
Radiation Dosimetry 11

statements on cavity ionization theory (Gray 1929, 1936). Gray went on to prove that
when a gas-filled cavity was small enough with respect to the range of the electrons,
the presence of cavity did not perturb the electron fluence spectrum.
Let us consider a cavity (medium c) is embedded in the uniform medium m.
Assume the medium is irradiated by an ionizing radiation. According to the Bragg-
Gray cavity theory, the cavity should necessarily satisfy the following two condi-
tions (Palani Selvam et al. 2021):

(a) The presence of the cavity should not perturb the fluence of charged particles in
the medium. For this purpose, the cavity needs to be small as compared with the
range of charged. This implies that the charged particle fluence spectra in both
m and c are the same (Attix 1986; Palani Selvam et al. 2021).
(b) The energy deposited in the cavity is assumed to be solely by the charged
particles produced outside the cavity and they cross the cavity (Attix 1986).
This further implies that no secondary electrons are produced inside the cavity
and the electrons do not stop within the cavity (Attix 1986, Palani Selvam et al.
2021).

According to the Bragg-Gray cavity theory:


 m
Dm S
¼ ð23Þ
Dc ρ c

where Dm is dose to mat a point


m of measurement without the cavity being present, Dc
is dose to cavity, and S=ρ c is mean mass collision stopping power ratio of m to c.


max

Φm ðEÞðS=ρÞm dE
 m
S 0
¼ E ð24Þ
ρ c ðmax
Φm ðEÞðS=ρÞc dE
0

where Φm(E)dE is the unperturbed primary fluence of charged particles in medium


m with energies between E and E þ dE and (S/ρ)m and (S/ρ)c are the unrestricted
mass collision stopping powers of media m and c (in MeVcm2/g), respectively. The
upper limit of the integrals in Eq. (24) is given by the maximum energy, Emax, of the
electrons in the fluence spectrum, and the lower limit corresponds to the lowest
energy in the spectrum, here indicated by a zero.
When a photon beam is incident on m, it generates charged particles such as
electrons and positrons. These charged particles are known as primary charged
particles and are part of Φm(E)dE. These electrons and positron will further generate
secondary electrons through collisions with atomic electrons of the medium. How-
ever, the δ-rays produced by these charged particles as a result of hard collisions in
12 S. Mishra and T. P. Selvam

the slowing down of the primary electrons are not part of Φm(E)dE. In the case of a
charged particle beam incident on m, the δ-rays produced by this beam are again not
part of Φm(E)dE. Note that whatever may be the incident beam (electron, positron or
photon), bremsstrahlung photons can always be generated by the charged particles.
Here, bremsstrahlung photons can be those generated by both primary charged
particles and δ-rays. Note that Φm(E)dE includes charged particles generated by
bremsstrahlung photons that are generated by primary charged particles. In general,
irrespective of the incident beam, -rays are not part of whichever the ways δ-rays are
produced. Hence, Eq. (23) utilizes unrestricted mass collision stopping powers. This
means that the δ-rays are assumed to be stopped on the spot where they are
generated. However, the energetic δ-rays may leave the cavity resulting in a partial
energy deposition in the cavity. Therefore, the Bragg-Gray cavity theory is
insensitive to the cavity size. This is in contradiction to the abovementioned two
Bragg-Gray conditions. Although large number of knock-on electrons is created, the
Bragg-Gray cavity theory still works reasonably well.

Spencer-Attix Cavity Theory


The Spencer-Attix cavity theory also works on two conditions as defined for the
Bragg-Gray theory (see section “Bragg Gray Cavity Theory”). However, these
conditions are extended to secondary particle fluence in addition to primary particle
fluence. Therefore, it can be inferred that the Spencer-Attix cavity theory (Attix et al.
1958; Attix 1986; Spencer and Attix 1955; ICRU 1984) finite range of δ-rays is
accounted as opposed to the Bragg-Gray cavity theory. In the Spencer-Attix cavity
theory, cavity size is therefore characterized by a parameter Δ that corresponds to
approximately the energy of an electron that is just able to cross the cavity (Palani
Selvam et al. 2021). Note that Δ is a parameter which defines the limit between
transported and absorbed electrons. Thus, the Spencer-Attix cavity theory is based
on formulation of a two-group schematization: (a) fast group, δ-rays with initial
energy greater than a cutoff Δ are assumed to dissipate their energy outside the cavity
and therefore are part of fluence spectrum, and (b) slow group, δ-rays of energy less
than Δ, which are assumed to have zero range, dissipate their energy “locally,” i.e., in
the cavity. Under these conditions (Attix 1986; Spencer and Attix 1955; ICRU
1984):
 m
Dm LΔ
¼ ð25Þ
Dc ρ c

where


max

Φm ðEÞðLΔ ðEÞ=ρÞm dE þ Φm ðΔÞðSðΔÞ=ρÞm Δ


 m

¼ EΔ ð26Þ
ρ c ðmax
Φm ðEÞðLΔ ðEÞ=ρÞc dE þ Φm ðΔÞðSðΔÞ=ρÞc Δ
Δ
Radiation Dosimetry 13

Here,

Φm(E)dE is the unperturbed primary fluence of charged particles in medium m with


energies between E and E þ dE, including δ-rays.
(LΔ/ρ)m is the restricted mass collision stopping powers (in MeVcm2/g) at electron
energy E for the medium m.
(LΔ/ρ)c is the restricted mass collision stopping powers (in MeVcm2/g) at electron
energy E for the cavity c.
(S(Δ)/ρ)m is the unrestricted stopping powers for the cavity medium c at E ¼ Δ.
(S(Δ)/ρ)c is the unrestricted stopping powers for the cavity medium c at E ¼ Δ.

Note that (LΔ/ρ)m and (LΔ/ρ)c are based on restricted collisional energy losses
below an energy threshold Δ.
The terms in Eq. (26), i.e., Φm(Δ)(S(Δ)/ρ)mΔ and Φm(Δ)(S(Δ)/ρ)cΔ, are the track-
end terms which account for energy dissipation of electrons with E  Δ through an
evaluation of the restricted stopping powers for Δ  E  2Δ (Nahum 1978).
In ionization chamber-based dosimetry, is traditionally given by the kinetic
energy of the electron whose residual CSDA range in air is equal to the mean
chord length of electrons crossing the cavity, L. For a convex cavity and an isotropic
field, L ¼ 4 V ¼ S where V is the cavity volume and S the cavity surface area (Attix
1986). Note that L ¼ 4 V/S applies to a convex cavity that is front-back symmetrical,
even if the electron field arrives isotropically only from the front and is absent from
the back (Attix 1986).

Re-Examination of Spencer-Attix Cavity Theory


As described by Bouchard (2012), for clinical radiotherapy beams (electron and
photon beams), Spencer-Attix cavity theory (Attix et al. 1958; Attix 1986; Spencer
and Attix 1955; ICRU 1984) further adapted by Nahum (1978) is considered to be
the accepted method for converting absorbed dose in a detector (wall-less) to
absorbed dose in the medium of interest. This approach is in use for several years
and is widely used. Bouchard (2012) re-examined the Spencer-Attix cavity theory
from first principles, using a new general cavity theory derived from radiation
transport equations. Bouchard (2012) highlighted the drawbacks of the Spencer-
Attix theory as below:
In practice, the ideal conditions needed for the Spencer-Attix theory are not
generally met. Also, there is a neglect of physical effects such as electron binding
energy, positron transport, etc., hence the introduction of fluence perturbation factors
(AAPM 1983; IAEA 1987) to correct for departure from the idealistic conditions.
For a standard dosimetry, the departure from the Spencer-Attix theory is corrected by
applying fluence perturbation correction, Kfl. However, for a nonstandard beam
dosimetry such as IMRT (intensity modulated radiation therapy) practice, such
corrections are easily predictable (Bouchard et al. 2009). Kfl is usually calculated
using the Monte Carlo techniques. For example, the quantity of interest in radio-
therapy is absorbed dose to water Dw, whereas the ion chamber measures dose to air
14 S. Mishra and T. P. Selvam

Da. Kf l is calculated as below to account for deviation from the Spencer-Attix cavity
theory:

½Dw =Da 
Kf l ¼ ð27Þ
½LðΔÞ=ρw =½LðΔÞ=ρa

Limitations of the Spencer-Attix theory as put forth by Bouchard (2012) are given
below:

(a) The response of a detector is evaluated using the condensed history-based Monte
Carlo methods (Kawrakow 2000). Therefore, one can argue that the theoretical
ground on which stopping power ratio calculation relies is arbitrary.
(b) The choice of Δ itself is arbitrary that may result in theoretical and practical
inconsistencies. For example, the secondary electrons set in motion close to the
wall with energy slightly smaller than Δ such that they can exit the volume are
ignored considering that Δ is the kinetic energy of the electrons which is just
sufficient to cross the cavity.
(c) Δ defined for the cavity volume is assumed to be the same for the cavity filled
with the medium of interest which is always a dense material to which the
stopping power ratio is calculated.
(d) The electron impact ionization differential cross sections generally increase with
decreasing energy loss (Kim and Rudd 1994), and therefore the statement that Δ
should be equal to the energy necessary for an electron to cross the cavity is
unfounded. This is because the electrons with kinetic energy less than Δ have on
average a kinetic energy much lower than Δ.

Despite the abovementioned shortcomings associated with the Spencer-Attix


cavity theory, the validity of this theory (Ma and Nahum 1991; Borg et al. 2000)
and the choice of Δ (Borg et al. 2000; La Russa and Rogers 2009; Rogers and
Kawrakow 2003; Sempau et al. 2004) have been revised over the years. For
example, a study by Wang and Rogers (2007) on Monte Carlo-simulated response
of a silicon detector shows unrealistic values of Δ for the purpose of matching the
calculated response. La Russa and Rogers (2009) discussed that there is still no
proper method to determine Δ for ionization chambers to avoid significant correc-
tions in the air kerma formalism for 60Co beams.
The basic approach adapted by Bouchard (2012) is the same as Spencer-Attix
cavity theory in terms of categorization of knock-on electrons, i.e., slow and fast
electrons. The Nahum formulation (Nahum 1978) of Spencer-Attix theory (applica-
ble for electrons, positrons, and photons in the energy range of radiation therapy) is
shown to be a special case of this model. The model can also be adapted to the class
II condensed history Monte Carlo technique. The model demonstrates that Δ is only
associated with the error on the solution of the transport equation, rather than the
electron kinetic energy necessary to cross the cavity. Numerically, it is shown that
there is no prescription for the value of Δ yielding either independence of the beam
Radiation Dosimetry 15

energy or a fulfillment of Bragg-Gray conditions. The inconsistencies with Spencer-


Attix theory recently reported in the literature are explained, and the overall
approach is the use of the Monte Carlo approach to calculate detector dose response.
Thus, the Monte Carlo-based approach is no longer reflected by Spencer-Attix
theory and that an appropriate formalism could be adopted as a standard theory for
radiation dosimetry.

Burlin Cavity Theory


Burlin (1966, 1968) proposed a theory that includes all cavity sizes. The Burlin
cavity theory for photons that relates dose to medium (Dm) and dose to cavity (Dc)
can be written as follows:

Dm
¼ d ðS=ρÞm m
c þ ð1 d Þðμen =ρÞc ð28Þ
Dc
where d is a parameter related to cavity size (d ¼ 1 for small cavities, d ¼ 0 for large
cavities), ðS=ρÞm
c is the mean ratio of the restricted mass stopping powers of medium
and cavity, and ðμen =ρÞmc is the mean ratio of the mass-energy absorption coefficients
for the medium and cavity. Burlin and Snelling (1969) derived the expression for
electron beams from the general cavity theory of ionization for photons as below:

Dm  
¼ d 1=ðS=ρÞm
c ð29Þ
Dc
where ðS=ρÞm c is calculated by Spencer-Attix theory and d is a dimensionless
weighting factor which accounts for the decrease of the electron spectrum charac-
teristic of the medium inside the cavity. Almond and McCray (1970) extended
Burlin’s cavity theory by introducing an additional term as below. This is to include
the production of delta rays within the cavity:

Dm  
¼ d 1=ðS=ρÞm
c þ ð1 dÞðZ=AÞm
c ð30Þ
Dc
where ðZ=AÞm c is ratio of the electron densities of the cavity to the wall medium.
It is interesting to note that for small cavities, Burlin’s cavity theory approaches
the Spencer-Attix stopping power ratio and (μen/ρ) ratio for large size cavities. The
main drawback of Burlin cavity theory is that it ignores the scattering effects of
secondary electrons that result in large discrepancies in experimental measurements
in high atomic number media (Horowitz et al. 1983; Kearsley 1984). Janssens (1983)
argues that the modifications to Burlin theory proposed by Horowitz and Dubi
(1982) were factually incorrect.

Kearsley Cavity Theory


A new general cavity theory was developed by Kearsley (1984) for photon. The
model includes the following components:
16 S. Mishra and T. P. Selvam

(a) The model takes into account scattering of secondary electron at the cavity
boundaries.
(b) The model treats the wall contribution to the energy deposited in the cavity as a
perturbation of the matched wall cavity case.
(c) The model takes into account the secondary electron scattering at the cavity
boundaries using the analysis and experimental data of Dutreix and
Bernard (1965).
(d) The model reduces to the Burlin theory (i.e., a linear combination of the mass
stopping power ratios and the mass energy absorption coefficient ratio).

The advantages of the model are (i) it is easy to use, (ii) it provides a better fit to
available experimental data, and (iii) it can provide dose distribution information
inside the cavity. The principal disadvantage of the model is the knowledge on the
secondary electron backscatter factors produced by photons. Such factors for high-
energy X-rays (11, 15, and 20 MV) and 6oCo gamma rays for a variety of materials
have been made available by Dutreix and Bernard (1965).
Haider et al. (1997) mentioned that an outstanding feature of Kearsley model
(1984) is the ability to calculate dose distributions inside plane parallel cavities
surrounded by medium. In the absence of secondary electron backscattering, the
Kearsley theory converges to Burlin’s theory. However, the Kearsley theory has poor
correlation with experimental results in high-Z media. For example, Kearsley theory
has numerous parameters and dose to cavity depends on choice of parameters.

Haider Cavity Theory


Haider et al. (1997) addressed shortcomings of Kearsley theory through a new general
cavity theory. In their model, the authors chose X-ray photons (500 kV–20 MV) in
which the Compton scattering is the dominant interaction. The electron density is
proportional to the energy loss cross section of electron stopping power which is
mainly determined by the excitation energy of the molecules and the Compton scatter
cross section (Haider et al. 1997). According to Janssens (1983), when the Compton is
the predominant interaction and the difference of mean excitation energies per electron
is small, then the stopping power ratio is equal to the ratio of electron densities,
independent of the cavity size. In the approach, the total electron fluence in the cavity
of thickness is divided into three groups (Haider et al. 1997). According to this
approach, the electron fluence is as follows:

(a) That originated in the cavity including the backscattered electrons generated by
it in the cavity travelling in the opposite direction
(b) That originated in the front wall medium including the backscattered electrons
generated by it in the cavity travelling in the opposite direction
(c) In the cavity resulting from the difference in the backscattering coefficients of the
cavity medium and the back wall medium

In the theory, medium g (known as cavity medium) of thickness t is sandwiched


between medium w and medium m (Haider et al. 1997). The final expression, i.e.,
Radiation Dosimetry 17

ratio of dose to cavity (g) and the medium (w) surrounding the cavity, is given as
(Haider et al. 1997):
 
f ¼ d 1 ðμen =ρÞgw þ d2 Fw 1 þ bg ðS=ρÞgw þ d 3 d4 Fg ðμen =ρÞgw
 
þ d5 Fw ðS=ρÞgw bm bg ð31Þ

Here, d1, d2, d3, d4, and d5 are weighting factors, Fw is the fractional forward
electron fluence originating in the medium w, Fg is the fractional forward electron
fluence originating in the medium g, bg is the probability of backscattering of the
medium g, bm is the backscattering coefficient of the medium m, ðμen =ρÞgw is ratio of
mean mass energy absorption coefficient of medium g and medium m, and ðS=ρÞgw is
ratio of mean mass collision stopping power of medium g and medium m.
When the cavity size is very large compared with the range of the electrons,
Eq. (31) reduces to:

f ¼ ðμen =ρÞgw ð32Þ

When the cavity size is very small compared to the range of the electrons and
when the font wall and back wall are:

(a) Identical medium (i.e., w), hence, f reduces to the Bragg-Gray theory, i.e., f ¼
ðS=ρÞgw .
(b) Not identical, hence, f does not reduce to the Bragg-Gray theory. This is because
the contributions from the front wall medium, the cavity medium, and the
backscattering from the back wall medium are calculated separately and all
three media could be of different atomic composition (Haider et al. 1997).

Detectors Used in Radiation Dosimetry

The general principles of absorbed dose determinations can be applied when differ-
ent types of dosimeters are to be used in the determination of absorbed dose. In this
section, dosimeters such as calorimeters, chemical dosimeters, and ionization cham-
bers that can be used for the absolute absorbed dose measurements and dosimeters
such as liquid ionization chambers, solid-state dosimeters, and films that are gener-
ally used for relative absorbed dose measurements are discussed.

Ionization Chambers

Air-filled ionization chambers are used for primary air kerma strength standardiza-
tion of 60Co beam from the charge collected by the chamber free in air. The charge
collected in the air is related to absorbed dose to the medium or air kerma (La Russa
and Rogers 2009; Rogers 1992):
18 S. Mishra and T. P. Selvam

Qair    wall μ air


W L
K air ¼ en
K K K K K ð33Þ
mair ð1  gair Þ e air ρ air ρ wall h wall an comp

where Qair/mair is the charge measured in humid air per unit mass of dry air that
would occupy the cavity at standard temperature and pressure, (W/e)air is the mean
energy deposited in dry air per unit charge of one sign released, gair is the fraction of
charged particle kinetic energy lost to radiative processes, ðμen =ρÞair
wall is the spectrum-
averaged ratio of mass energy absorption coefficients, and ðL=ρÞwall air is the Spencer-
Attix ratio of restricted mass collision stopping powers. The correction factors due to
humidity is Kh, attenuation and scatter by the wall is Kwall, axial nonuniformity of the
beam is Kan, and material inhomogeneity in the chamber is Kcomp.
Under certain ideal conditions, Kh, Kan, and Kcomp corrections are unity by
definition. Furthermore, the correction for additional nonideal conditions, K, reduces
to a fluence correction, denoted here as Kf l, to correct for a potential perturbation of
the wall’s charged particle fluence in the cavity gas.
Under the assumption that photon interactions in the cavity is negligible (which is
not true for low-energy photons) and that the departure from Spencer-Attix condi-
tions is solely due to the perturbation of the charged particle fluence by the presence
of the cavity:
 wall  air
Dair L μen
K air ¼ K K ð34Þ
ð1  gair Þ ρ air ρ wall wall fl

As ðK c Þair ¼ K air ð1 gair Þ, the above equation is rewritten as:


 wall  air
L μen
ðK c Þair ¼ Dair K K ð35Þ
ρ air ρ wall wall fl

The values of ðL=ρÞwall


air , Kwall, and Kfl are generally calculated using Monte Carlo
methods. Note that these values are chamber-specific.

Calorimetry

Calorimetry method can be applied for absorbed dose measurements in a small


volume of an irradiated absorber. In this method, the small volume having mass dm
medium is thermally isolated. The mean absorbed dose D is given by (ICRU 1954):

Δe ΔEh ΔEs
D¼ ¼ þ ð36Þ
Δm Δm Δm
where Δe is the mean energy imparted to the absorber by the incident ionizing
radiation, ΔEh is the energy appearing as heat, and ΔEs is the chemical defect, which
Radiation Dosimetry 19

may be positive or negative. An example of ΔEs is the energy produced or absorbed


in induced radiochemical reactions. When there is no change of state,

ΔEh
¼ cp ΔT ð37Þ
Δm
where cp is specific heat capacity at constant pressured and ΔT is the change in
temperature.

Chemical Dosimetry

Chemical dosimetry involves determination of dose by the measurement of chemical


changes induced in materials by ionizing radiation. The mean absorbed dose to
dosimeter solution, D, using a chemical dosimeter can be determined as follows:

ΔA
D¼ ð38Þ
ρ l em G
where ΔA is increase in absorbance due to irradiation at a temperature “t” during the
spectrometric measurement, ρ is the density of dosimeter solution (g/cm3), l is
optical length in the photometer cell (m), εm is the molar linear absorption coefficient
of the product (m2/mol) at the wavelength used for spectrophotometric measure-
ments, and G is radiation chemical yield (in mol/J).
G is the fundamental quantitative characteristic of any chemical transformation
induced by ionizing radiation. Earlier G values were reported as the number of
molecules of product formed (X), as G(X), or of starting material Y changed G(Y ),
per 100 eV energy absorbed. Radiation chemical yield has the same symbol G(X) or
G(Y ) as the case may be. G(X) is the molar concentration n(X), produced/changed
for a known energy E (in J) is imparted to it:

GðXÞ ¼ nðXÞ=E ð39Þ

G-values in terms of number of species per 100 eV can be converted to SI units by


multiplying by 1.036  107 or in other words 1 mol/J ¼ 9.649  106 mol/100 eV.
The value of the radiation chemical yield may depend on the type of radiation and
the chemical reactions involved. When the radiation-induced reaction is a chain
reaction, then the G value is high. For nonchain reactions, it is not large. The value of
G can be calculated if the absorbed dose and the concentration of the product being
formed (or decomposed) are both known.
     
G mol J1 ¼ n mol kg1 =Dose J kg1

In actual practice, the G value is found from kinetic curves, relating product
concentration and absorbed dose. If the relationship is linear, the slope of the curve is
20 S. Mishra and T. P. Selvam

the G value. If the experimental points show deviation from linearity, G is calculated
by drawing a tangent to the curve from the point of origin.

Film Dosimetry

Radiographic Film
Radiographic film finds applications in diagnostic radiology, radiotherapy, and
radiation protection. The film consists of a base of thin plastic with a radiation-
sensitive emulsion, usually silver bromide (AgBr) grains suspended in gelatine,
coated uniformly on one or both sides of the base.
When ionizing radiation is incident on the film, it forms a latent image in the film.
This latent image becomes visible due to film blackening and is permanent after
processing. Hence, light transmission is a function of the film opacity which can be
measured in terms of optical density (OD) using densitometers. The OD is defined as:

OD ¼ log 10 ðI 0 =I Þ ð40Þ

where I0 is the initial light intensity and I is the transmitted intensity. Although films
provide qualitative information of absorbed dose, using a calibration curve dose vs
OD, dose can be determined. Ideally, the relationship between the dose and OD
should be linear, but some emulsions exhibit linear response over a limited dose
range, and some are nonlinear. The dose vs OD curve should therefore be established
for each film before it is used for dosimetry work.

Radiochromic Film
Radiochromic film (RCF) is a planar dosimeter for applications in radiotherapy and
food irradiation. RCF film consists of a radiosensitive (7–23 μm thick) colorless
leuco dye bonded to a 100- m-thick Mylar base. These films are colorless before
irradiation. Upon irradiation, the sensitive material, dye, is polymerized and turns a
deep blue color without physical, chemical, or thermal processing. Thus, RCF is
self-developing and requires neither developer nor fixer. As RCF is grainless, it has a
high spatial resolution of 1200 lines/mm (McLaughlin et al. 1991). The polymer
absorbs light, and the transmission of light through the film can be measured with a
suitable densitometer. Response of RCF is stable and is reproducible if protected
from UV light, unstable temperatures, and humidity. Since the dye is nearly tissue-
equivalent, its deviation from tissue-equivalent energy response is much smaller.
There are various types of radiochromic films such as EBT, EBT2, EBT3,
HD-V2, B3, etc. It is important that each film is subjected to different doses and
the absorbance is determined using suitable system that has range of a light source
(different wavelengths). Suitable light source should be identified based on maxi-
mum absorbance through the absorbance vs wavelength. The color intensity of the
film is proportional to the radiation absorbed dose. For dosimetry applications
involving B3 film, optical density of the exposed films can be measured at 552 nm
Radiation Dosimetry 21

using Genesys 20 portable spectrophotometer (Thermo Electron Corp, USA). The


color formation in the irradiated B3 film is not complete at the end of the irradiation.
An important aspect in dosimetry involving RCF is the generation of calibration
curves for the film and the reading system. For example, if HD-V2 film is intended
for dose measurements during irradiation of onion, potato, mango, etc. and the
reading system to be used is a spectrophotometer at 550-nm wavelength, then a
calibration curve for the film and the reading system should be generated. Similarly,
if B3 film is used along with Genesys 20 portable spectrophotometer (552 nm) for
measuring doses in the range of kGy, it is required to generate a calibration curve for
this dosimetry system.
Dosimetry using RCFs has a few advantages over radiographic films, such as
elimination of the need for darkroom facilities, film cassettes or film processing, dose
rate independence, better energy response characteristics, and insensitivity to ambi-
ent conditions. RCFs are generally less sensitive than radiographic films and are
generally useful at higher doses, and nonlinearity should be corrected for in the very
high-dose region.

Solid-Sate Detectors

Usage of gas-filled detectors (e.g., ionization chamber) is somewhat limited due to a


number of factors. For example, the small number of electron-ion pairs is generated
in a gas, and the number of target atoms per unit volume in the gas that the incident
radiation encounters is smaller. Solid-state detectors such as semiconductors, scin-
tillation, and thermoluminescent dosimeters have some advantages over gas-filled
detectors.

Semiconductor Detector
Semiconductors are most commonly used when best energy resolution is desired.
They are called semiconductors because their electrical conduction properties lie
between those of insulators and conductors. In semiconductors, the fundamental
information carriers are electron-hole pairs, which are collected. Germanium
(Ge) and silicon (Si) are two of the most commonly used semiconductor materials.
The choice of semiconductor depends on many factors:

(a) Resistivity
(b) Mobility of charges
(c) Drift velocity
(d) Purity
(e) Operating temperature
(f) Cost

Scintillation Detectors
As opposed to gas-filled detectors, the scintillation detector works on a different
principle. When ionizing radiation is incident on scintillation material, energy is
22 S. Mishra and T. P. Selvam

transferred to the atoms of the material, which takes the atoms into short-lived
excited states. When these excited atoms return to their ground states, they result
in emission light photons. These photons are known as scintillation photons. The
basic steps involved in detection of these scintillation photons are:

• Transfer of energy to the bound states of the material


• Relaxation of the excited state to the ground state, resulting in the emission of
light photons
• Collection of photons by a suitable photodetector
• Detection of the photodetector signal by associated electronics system

The amount of light produced in the scintillator is small. It must be amplified


before it can be recorded as a pulse or in any other way. The amplification or
multiplication of the scintillator’s light is achieved with a device known as the
photomultiplier tube (PMT). Its name denotes its function: it accepts a small amount
of light, amplifies it many times, and delivers a strong pulse at its output. Amplifi-
cations of the order of l06 are common for many commercial PMTs.
Alpha, beta, gamma, and neutron radiations can be detected using scintillation
detectors. Scintillator detectors, which can be solid, liquid, and gas, are generally
made of plastic, organics, or inorganic materials. They can be made in any shape and
size. NaI(Tl) detector is commonly used for gamma detection and gamma spectrom-
etry. The detector is generally shielded to reduce background radiation level before
use as it gives high background radiation levels due its high sensitivity.

Thermoluminescent Dosimeters
When ionizing radiation is incident on certain solids doped with appropriate impu-
rities, it creates an electron-hole pair in the crystalline structure. The liberated
electron migrates to the conduction band where it further migrates to an electron
trap, while the hole left migrates to a hole trap. The charge carriers are maintained for
a time that depends on the potential energy of the trap (the trap depth). If the traps are
deep, upon heating, the electrons and holes in the shallowest traps are released first.
The released charge carriers recombine with their counterparts at a luminescence
center or at a trap that is closely coupled to such a luminescence center. As a result of
this recombination, there is the release of visible light, which is characteristic of the
impurity recombination center. The light emitted is quantified and is related to
absorbed dose in the material. The material can be reused after annealing. Annealing
is application of appropriate heating to the material. As the methodology involves
application of thermal energy to stimulate emission of luminescence, it is known as
“thermoluminescence,” and the associate dosimetry is called thermoluminescence
dosimetry (TLD). Note that pure TL crystal does not exhibit considerable lumines-
cence and therefore is not quantifiable. Addition of trace amounts of impurities
(about ppm level) in the crystal lattice enhances its TL properties. Several TL
materials with dopants such as LiF:Mg,Ti,, LiF:Mg,Cu,P, CaF2:Mn, and CaSO4:
Dy are being used in radiation dosimetry.
Radiation Dosimetry 23

Properties of Dosimeters

The dosimeters must be characterized before they are used for the intended purposes.
The properties of a dosimeter should be characterized by accuracy and precision,
linearity, dose or dose rate dependence, absorbed dose and intrinsic energy response,
directional dependence, spatial resolution, etc.

Accuracy and Precision

The dosimetric quantity measured through a dosimeter will have an uncertainty. Uncer-
tainty is a parameter that describes the dispersion of the measured values of a quantity.
Uncertainty is expressed in terms of accuracy and precision. Precision of dosimetry
measurements specifies the reproducibility of the measurements under similar condi-
tions. Therefore, precision can be estimated through repeated measurements. High
precision means the standard deviation associated with the measured value is small.
The accuracy of dosimetry measurements is the proximity of their expectation value to
the “true value” of the measured quantity. Note that measured values are not absolutely
accurate and the inaccuracy of a measurement result is characterized as uncertainty.

Linearity

Linearity is a parameter that describes the relationship between the dosimeter


response and the dose delivered as a function of dose. Generally, the dosimeter
reading is reliable in certain range of doses as beyond a certain dose range a
nonlinearity may set in. The dosimeters may also exhibit saturation after certain
dose. The linearity range, the nonlinearity behavior, and the saturation behavior
depend on the type of dosimeter and its physical characteristics. Hence, it is
important to know whether the response of the detector is directly proportional to
the dose being measured. In general, one can write:

Ddet ðDÞ ¼ α kl ½Mdet ðDÞ Mdet ðDÞ ð41Þ

where kl [Mdet(D)] is the intrinsic linearity whose value is unity for some reference dose
D0 and α is a quantity that relates Mdet(D0) to Ddet(D0) (Rogers 2009). If kl is indepen-
dent of D and Mdet(D), then the detector’s response is said to be linear (Rogers 2009).
This means that if the dose is doubled, the response of the detector doubles. However,
some detectors may not exhibit this behavior which should be taken into account.

Dose Rate Dependence

The integrating systems that measure the dosimeric quantity should be independent
of the dose rate which means that the response of a dosimetry system at two different
24 S. Mishra and T. P. Selvam

dose rates should be constant. However, in practical applications, the dose rate may
influence the dosimeter readings. Hence, appropriate corrections must be
incorporated.
 
Let Mraw _ at dose rate D_ and then
det D be the raw response of the detector
 measured

0 _
dose rate corrected response of the detector Mdet D is:
     raw  
M0det D_ ¼ kdr Mraw _
det D Mdet D
_ ð42Þ
  
where kdr Mraw _ corrects for the dose rate dependence.
det D

Energy Dependence

Energy dependence of radiation detectors is an important property for determination


of absorbed dose as the detectors are exposed to a beam quality Q which is often
different from the calibration beam. Hence, it is important that such a dependency is
corrected for. The overall energy dependence of a detector is composed of two parts,
namely, intrinsic energy dependence and absorbed dose energy dependence
(DeWerd et al. 2009; Rogers 2009), which are explained below.

Intrinsic Energy Dependence


The intrinsic energy dependence, kbq(Q), relates the reading of the detector to the
dose to the sensitive volume of the detector (Rogers 2009). It is defined as:

kbq ðQÞ ¼ Ddet ðQÞ=Mset ðQÞ ð43Þ

where Ddet(Q) is the absorbed dose to the sensitive volume of the detector at a given
beam quality, Q, and Mdet(Q) is the meter reading of the detector at the same Q.
In principle, the intrinsic energy dependence could also be dependent on other
factors such as dose, dose rate, temperature and pressure, direction, etc. General-
purpose Monte Carlo codes that model radiation transport can be used to calculate
the absorbed dose energy dependence. However, such codes cannot be used to
calculate the intrinsic energy dependence as the thermoluminescence mechanism is
not modeled. Hence, intrinsic energy dependence can be determined only through
measurements.

Absorbed Dose Energy Dependence


The absorbed dose energy dependence relates the dose to the detector material to the
dose to the medium at the point of measurement of the detector (Rogers 2009). The
absorbed dose energy dependence at the beam quality Q, f(Q), is also known as
extrinsic energy dependence. It is defined as:

Dmed ðQÞ ¼ f ðQÞDdet ðQÞ ð44Þ


Radiation Dosimetry 25

where Dmed(Q) is the absorbed dose to the medium at the point of measurement of
the detector in the absence of the detector at the beam quality Q and Dmed(Q) is the
absorbed dose to the sensitive material of the detector at the same Q. The factor f(Q)
encompasses the component-specific and replacement perturbation effects as well as
the absorbed dose energy dependence of the detection material. It is a function of
beam quality Q but can also be a function of dose, dose rate, geometry of source and
detector location, etc.

References
AAPM Task Group 21 Radiation Therapy Committee (1983) A protocol for the determination of
absorbed dose from high-energy photon and electron beams. Med Phys 10:741–771
Almond PR, McCray K (1970) The energy response of LiF, CaF2 and Li2B4O7: Mn to high energy
radiations. Phys Med Biol 15:335–342
Attix FH (1986) Introduction to radiological physics and radiation dosimetry new. Wiley, New York
Attix FH, DeLa LV, Ritz VH (1958) Cavity ionization as a function of wall material. J Res Natl Bur
Stand 60:235–243
Borg J, Kawrakow I, Rogers DWO, Seuntjens JP (2000) Monte Carlo study of correction factors for
Spencer–Attix cavity theory at photon energies at or above 100 keV. Med Phys 27:1804–1813
Bouchard H (2012) A theoretical re-examination of Spencer-Attix cavity theory. Phys Med Biol 57:
3333–3358
Bouchard H, Seuntjens JP, Carrier JJ, Kawrakow I (2009) Ionization chamber gradient effects in
nonstandard beams. Med Phys 36:4654–4663
Bragg WH (1910) Studies in radioactivity. Macmillan, New York
Burlin TE (1966) A general theory of cavity ionisation. Br J Radiol 39:727–734
Burlin TE (1968) Cavity – chamber theory. In: Attix FH, Roesch WC (eds) Radiation dosimetry.
Academic Press, New York/London
Burlin TE, Snelling RJ (1969) The application of general cavity theory to the dosimetry of electron
fields. In: Proceedings of the 2nd symposium on micro dosimetry, pp 455–473
DeWerd LA, Bartol LJ, Davis SD (2009) Thermoluminescence dosimetry in clinical dosimetry
measurements in radiotherapy. Medical Physics Publishing, Madison. Available from http://
www.aapm.org/meetings/09SS/documents/2.pdf
Dutreix JJ, Bernard M (1965) Etude du flux des électrons secondaires et de leur rétrodiffusion.
Biophysik 2:179–192
Glossary of Terms and Definitions of Basic Quantities (2020) J ICRU 20:9–12
Gray LH (1929) The absorption of penetrating radiation. Proc R Soc Lond A122:647–668
Gray LH (1936) An ionization method for the absolute measurement of g-ray energy. Proc R Soc
Lond A156:578–596
Haider JA, Skarsgardz LD, Lamz GKY (1997) A general cavity theory. Phys Med Biol 42:491–500
Horowitz YS, Dubi A (1982) Proposed modifications of Burlin’s cavity theory for photons. Phys
Med Biol 27:867–872
Horowitz YS, Moscovitch M, Dubi A (1983) Modified general cavity theory applied to the
calculation of gamma dose in Co-60 thermoluminescence dosimetry. Phys Med Biol 28:
829–840
IAEA (1987) Absorbed dose determination in photon and electron beams: an international code of
practice technical report TRS-277. IAEA, Vienna
International Commission on Radiation Units and Measurement (ICRU) (1984) Radiation Dosim-
etry: Electron Beams with Energies between 1 and 50 MeV, ICRU Report 35. ICRU, Bethesda
International Commission on Radiation Units and Measurements (ICRU) (1962) Radiation quan-
tities and units, ICRU Report 10a. ICRU, Bethesda
26 S. Mishra and T. P. Selvam

International Commission on Radiation Units and Measurements (ICRU) (1993) Quantities and
units in radiation protection dosimetry, ICRU Report 51. ICRU, Bethesda
International Commission on Radiation Units and Measurements (ICRU) (1998) Quantities and
units for ionising radiation, ICRU Report 60. ICRU, Bethesda
International Commission on Radiation Units and Measurements (ICRU) (2011) Fundamental
quantities and units for ionising radiation (revised): ICRU Report 85. ICRU, Bethesda
Janssens A (1983) A proposed modification of Burlin’s general cavity theory for photons. Phys Med
Biol 29:455–456
Kawrakow I (2000) Accurate condensed history Monte Carlo simulation of electron transport:
II. Application to ion chamber response simulations. Med Phys 27:499–513
Kearsley E (1984) A new general cavity theory. Phys Med Biol 29:1179–1187
Kim YK, Rudd ME (1994) Binary-encounter-dipole model for electron-impact ionization. Phys
Rev A 50:3954–3967
La Russa DJ, Rogers DWO (2009) Accuracy of Spencer–Attix cavity theory and calculations of
fluence correction factors for the air kerma. Med Phys 36:4173–4183
Ma CM, Nahum AE (1991) Bragg–Gray theory and ion chamber dosimetry for photon beams. Phys
Med Biol 36:413–428
McLaughlin WL, Chen YD, Soares CG, Miller A, Van DG, Lewis DF (1991) Sensitometry of the
response of a new radiochromic film dosimeter to gamma radiation and electron beams. Nucl
Instrum Methods Phys Res 35:165–172
Nahum E (1978) Water/air stopping-power ratios for megavoltage photon and electron beams. Phys
Med Biol 23:24–38
Palani Selvam T, Shrivastava V, Bakshi AK (2021) Monte Carlo calculation of Spencer-Attix and
Bragg-Gray stopping-power ratios of tissue-to-air for ISO reference beta sources – an EGSnrc
study. JINST 16:P03006
Rogers DWO (1992) In: Purdy J (ed) Advances in radiation oncology physics, Medical physics
monograph no. 19. AAPM, New York, pp 181–223
Rogers DWO (2009) General characteristics of radiation dosimeters and a terminology to describe
them in clinical dosimetry measurements in radiotherapy. Medical Physics Publishing, Madison
Rogers DWO, Kawrakow I (2003) Monte Carlo calculated correction factors for primary standards
of air kerma. Med Phys 30:521–532
Sempau J, Andreo P, Aldana J, Mazurier J, Salvat F (2004) Electron beam quality correction factors
for plane-parallel ionization chambers: Monte Carlo calculations using the PENELOPE system.
Phys Med Biol 49:4427–4444
Spencer LV, Attix FH (1955) A theory of cavity ionization. Radiat Res 3:239–254
Wang LL, Rogers DWO (2007) Monte Carlo study of Si diode response in electron beams. Med
Phys 34:1734–1742

You might also like