Sei sulla pagina 1di 223

The development of new anti-scatter grids

for improving x-ray image diagnostic quality


and reducing patient radiation exposure

Submitted by

Abel Zhou

A thesis submitted for the requirements of the degree of


Doctor of Philosophy
2018

Faculty of Health
The University of Canberra, Canberra, Australia
A sample grid with a cut

i
Contents

STATEMENT OF AUTHORSHIP ------------------------------------------------------------------ XIII

ACKNOWLEDGEMENTS ----------------------------------------------------------------------------- XV

ABSTRACT ---------------------------------------------------------------------------------------------- XVII

TABLE OF ABBREVIATIONS ---------------------------------------------------------------------- XXI

GLOSSARY OF TERMS ---------------------------------------------------------------------------- XXIII

CHAPTER 1: THE NATURE AND SCOPE OF THIS PROJECT ------------------------------ 1

1.1. INTRODUCTION ----------------------------------------------------------------------------------------- 1

1.2. BACKGROUND ------------------------------------------------------------------------------------------ 1

1.2.1. Overview of medical x-ray imaging ----------------------------------------------------------- 1


1.2.2. Ionising radiation -------------------------------------------------------------------------------- 3
1.2.2.1. Rayleigh scattering ------------------------------------------------------------------------------------------------- 3
1.2.2.2. Compton scattering ------------------------------------------------------------------------------------------------- 4
1.2.2.3. Photoelectric absorption ------------------------------------------------------------------------------------------- 5

1.2.3. Harmful effects of ionising radiation ---------------------------------------------------------- 7


1.2.3.1. Deterministic effects of ionising radiation---------------------------------------------------------------------- 8
1.2.3.2. Stochastic effects of radiation exposure ------------------------------------------------------------------------ 8
1.2.3.2.1. Evidence of stochastic effects -----------------------------------------------------------------------------9
1.2.3.2.2. Prediction of stochastic effects: linear-no-threshold ------------------------------------------------ 11

1.2.4. Radiation protection ---------------------------------------------------------------------------- 12


1.2.5. Effects of scatter radiation --------------------------------------------------------------------- 13
1.2.5.1. Subject contrast ---------------------------------------------------------------------------------------------------- 13
1.2.5.2. Image contrast and radiographic contrast ---------------------------------------------------------------------- 14
1.2.5.3. Expression of optical contrast ----------------------------------------------------------------------------------- 16
1.2.5.4. Effects of scatter radiation on image contrast ----------------------------------------------------------------- 18

1.2.6. Overview of scatter radiation reduction techniques --------------------------------------- 19

v
1.2.7. Anti-scatter grids -------------------------------------------------------------------------------- 21
1.2.7.1. Mechanism of scatter radiation reduction by grids ----------------------------------------------------------- 21
1.2.7.2. Reduction of primary and scatter radiation by grids --------------------------------------------------------- 23
1.2.7.3. Patient radiation dose increase attributed to grids ------------------------------------------------------------ 24

1.3. RESEARCH OPPORTUNITY FROM THE LITERATURE ------------------------------------------------ 25

1.4. MOTIVATIONS ----------------------------------------------------------------------------------------- 25

1.5. PURPOSE------------------------------------------------------------------------------------------------ 25

1.6. SIGNIFICANCE ----------------------------------------------------------------------------------------- 26

1.7. OVERVIEW OF METHODOLOGY ---------------------------------------------------------------------- 27

1.8. RESEARCH OBJECTIVES ------------------------------------------------------------------------------- 27

1.9. RESEARCH CONTRIBUTIONS-------------------------------------------------------------------------- 28

1.10. ORGANISATION OF THIS THESIS -------------------------------------------------------------------- 29

CHAPTER 2: GRID EVALUATION METHODOLOGY ---------------------------------------- 31

2.1. GRID EVALUATION CRITERIA ------------------------------------------------------------------------ 31

2.1.1. Brief history of grid evaluation criteria ------------------------------------------------------ 31


2.1.2. Basic grid evaluation criteria: Tp, Ts, and Tt----------------------------------------------- 32
2.1.3. Image contrast improvement factor: K------------------------------------------------------- 33
2.1.4. Bucky factor: B ---------------------------------------------------------------------------------- 33
2.1.5. Quantum signal-to-noise ratio improvement factor: KSNR -------------------------------- 34
2.1.6. Image quality improvement factor: Q -------------------------------------------------------- 34

2.2. MAXIMUM KSNR FOR DETERMINATION OF STRIP OPTIMAL THICKNESS ------------------------- 35

2.3. RADIATION MEASUREMENT FOR GRID EVALUATION ---------------------------------------------- 36

2.4. EXPERIMENTAL SETUP FOR GRID EVALUATION ---------------------------------------------------- 40

2.4.1. Narrow beam condition ------------------------------------------------------------------------ 40


2.4.2. Broad beam condition with a primary beam blocker -------------------------------------- 42

vi
2.4.3. Broad beam condition -------------------------------------------------------------------------- 43

2.5. SIMULATION SETUP FOR GRID EVALUATION ------------------------------------------------------- 44

SUMMARY --------------------------------------------------------------------------------------------------- 45

CHAPTER 3: INFORMATION FOR THE DEVELOPMENT OF THE MONTE CARLO

CODE SYSTEM -------------------------------------------------------------------------------------------- 47

3.1. INTRODUCTION ---------------------------------------------------------------------------------------- 47

3.2. PROCEDURE FOR SIMULATING THE PRODUCTION OF X-RAY IMAGES ---------------------------- 48

3.3. PHOTON TRANSPORTATION -------------------------------------------------------------------------- 53

3.3.1. Change in a photon’s location and direction ----------------------------------------------- 53


3.3.1.1. Determining the free path length -------------------------------------------------------------------------------- 56
3.3.1.2. Determining the polar angle, θ: differential cross-section -------------------------------------------------- 59

3.3.2. Change in a photon’s status and energy ----------------------------------------------------- 61


3.3.3. Random numbers: random number generators and sampling methods ----------------- 62
3.3.3.1. Random numbers -------------------------------------------------------------------------------------------------- 62
3.3.3.2. Sampling methods ------------------------------------------------------------------------------------------------- 63
3.3.3.2.1. Exponential sampling for free path length ------------------------------------------------------------ 64
3.3.3.2.2. Discrete indexed sampling ------------------------------------------------------------------------------ 65
3.3.3.2.2. Reject-accept sampling ---------------------------------------------------------------------------------- 66

3.4. IMPLEMENTATION OF INTERACTIONS BETWEEN PHOTONS AND MATTER ----------------------- 67

3.4.1. Rayleigh scattering------------------------------------------------------------------------------ 67


3.4.2. Compton scattering ----------------------------------------------------------------------------- 68
3.4.3. Photoelectric absorption ----------------------------------------------------------------------- 69

3.5. RADIATION DOSIMETRY IN THE MONTE CARLO SIMULATION ----------------------------------- 70

3.5.1. Photon fluence, photon flux, and energy fluence ------------------------------------------- 70


3.5.2. Mass energy transfer coefficient -------------------------------------------------------------- 71
3.5.3. Mass energy absorption coefficient----------------------------------------------------------- 71

vii
3.5.4. KERMA (kinetic energy released per unit mass) ------------------------------------------- 72
3.5.5. Absorbed dose ----------------------------------------------------------------------------------- 73
3.5.6. Radiation exposure------------------------------------------------------------------------------ 74

3.6. VIRTUAL PHANTOM FOR THE MONTE CARLO SIMULATION -------------------------------------- 75

SUMMARY --------------------------------------------------------------------------------------------------- 76

CHAPTER 4: NEW METHOD FOR DETERMINING RADIATION TRANSMISSION IN

GRIDS --------------------------------------------------------------------------------------------------------- 79

4.1. INTRODUCTION: RADIATION TRANSMISSION IN GRID DESIGNS ----------------------------------- 79

4.2. NEW METHOD FOR DETERMINING RADIATION TRANSMISSION IN GRID DESIGNS -------------- 82

4.2.1 Determination of radiation transmission ----------------------------------------------------- 83


4.2.2. Determination of variable ai in focused grids ----------------------------------------------- 86
4.2.3. Determination of variable ai in parallel grids ---------------------------------------------- 91

4.3. VALIDATION OF THE NEW METHOD ----------------------------------------------------------------- 92

4.3.1. Criteria for comparisons of grid performance and grid selection ----------------------- 93
4.3.2. Grid details, phantoms, radiation beam quality, and simulation setups ---------------- 93
4.3.3. Simulation of grid movement ------------------------------------------------------------------ 94
4.3.4. Simulation of photon transport ---------------------------------------------------------------- 96
4.3.5. Results -------------------------------------------------------------------------------------------- 96
4.3.5.1. Theoretical comparison: transmission in parallel grid ------------------------------------------------------- 96
4.3.5.2. Transmission of primary radiation in focused grid: Tp------------------------------------------------------ 99
4.3.5.3. Transmission of scatter radiation in focused grid: Ts ------------------------------------------------------ 100
4.3.5.4. Images behind a stationary grid and moving grid ---------------------------------------------------------- 103
4.3.5.5. Dependence of radiation transmission (Tp and Ts) -------------------------------------------------------- 104

4.3.6. Comparisons of Tp and Ts between simulation and the literature --------------------- 107
4.3.6.1. Tp comparison---------------------------------------------------------------------------------------------------- 107
4.3.6.1. Ts comparison ---------------------------------------------------------------------------------------------------- 109

viii
SUMMARY ------------------------------------------------------------------------------------------------- 111

CHAPTER 5: VALIDATION OF THE NEW MONTE CARLO CODE SYSTEM FOR

GRID EVALUATION ----------------------------------------------------------------------------------- 115

5.1. MATERIAL AND METHODS ------------------------------------------------------------------------- 115

5.1.1. Setups of experiment and simulation ------------------------------------------------------- 115


5.1.2. Radiation detector----------------------------------------------------------------------------- 116
5.1.3. X-ray beams ------------------------------------------------------------------------------------ 117
5.1.4. Grid details ------------------------------------------------------------------------------------- 118
5.1.5. Phantoms --------------------------------------------------------------------------------------- 118
5.1.6. Simulation of photon transportation ------------------------------------------------------- 118

5.2. RESULTS ---------------------------------------------------------------------------------------------- 119

5.2.1. Results obtained through simulation ------------------------------------------------------ 119


5.2.1. Results determined in experiment --------------------------------------------------------- 123

5.3. DISCUSSION ------------------------------------------------------------------------------------------ 126

SUMMARY ------------------------------------------------------------------------------------------------- 129

CHAPTER 6: CRITERIA FOR DESIGNING NEW GRIDS ----------------------------------- 131

6.1. OVERVIEW OF GRID DESIGN THEORY ------------------------------------------------------------- 131

6.2. GRID DESIGN FACTORS ----------------------------------------------------------------------------- 132

6.3. ISSUES IN GRID DESIGN ----------------------------------------------------------------------------- 134

6.4. CRITERIA FOR DESIGNING NEW GRIDS ------------------------------------------------------------ 135

SUMMARY ------------------------------------------------------------------------------------------------- 138

ix
CHAPTER 7: MONTE CARLO EVALUATION OF NEW GRID DESIGNS AND

PERFORMANCE RESULTS -------------------------------------------------------------------------- 141

7.1. MATERIAL AND METHODS ------------------------------------------------------------------------- 141

7.1.1. Grid details ------------------------------------------------------------------------------------- 141


7.1.1.1 Mammographic grid designs ----------------------------------------------------------------------------------- 141
7.1.1.2. General grid designs --------------------------------------------------------------------------------------------- 142

7.1.2. Virtual phantoms ------------------------------------------------------------------------------ 143


7.1.2.1. Virtual breast phantoms for mammographic grid evaluation --------------------------------------------- 143
7.1.2.2. Virtual water phantoms for general grid evaluation ------------------------------------------------------- 145

7.1.3. X-ray beam quality ---------------------------------------------------------------------------- 146


7.1.3.1. Mammographic x-ray beam quality -------------------------------------------------------------------------- 146
7.1.3.2. General x-ray beam quality ------------------------------------------------------------------------------------ 146

7.1.4. Virtual image receptor ----------------------------------------------------------------------- 147


7.1.5. Simulation setup and photon transportation ---------------------------------------------- 148
7.1.5. Data fitting ------------------------------------------------------------------------------------- 148

7.2. RESULTS ---------------------------------------------------------------------------------------------- 149

7.2.1. Mammographic grid results ----------------------------------------------------------------- 149


7.2.1.1. Mammographic grids: Tp -------------------------------------------------------------------------------------- 149
7.2.1.2. Mammographic grids: Ts --------------------------------------------------------------------------------------- 151
7.2.1.3. Mammographic grids: KSNR ------------------------------------------------------------------------------------ 153
7.2.1.4. Strip optimal thicknesses of mammographic grids --------------------------------------------------------- 155

7.2.2. General grid results --------------------------------------------------------------------------- 157


7.2.2.1. General grids: Tp ------------------------------------------------------------------------------------------------ 157
7.2.2.2. General grids: Ts ------------------------------------------------------------------------------------------------- 158
7.2.2.3. General grids: KSNR ---------------------------------------------------------------------------------------------- 160
7.2.2.4. Strip optimal thicknesses of general grids ------------------------------------------------------------------- 161

7.3. DISCUSSION ------------------------------------------------------------------------------------------ 163

7.3.1. Mammographic grids ------------------------------------------------------------------------- 163


7.3.1.1. New mammographic grid designs vs the literature: Tp and Ts ------------------------------------------ 163

x
7.3.1.2. New mammographic grid designs vs the literature: K SNR ------------------------------------------------- 167

7.3.2. General grids ---------------------------------------------------------------------------------- 169


7.3.2.1. New general grid designs vs the literature: Tp and Ts ----------------------------------------------------- 169
7.3.2.2. New general grid designs vs the literature: K SNR ----------------------------------------------------------- 172

SUMMARY ------------------------------------------------------------------------------------------------- 173

CHAPTER 8: CONCLUSIONS ----------------------------------------------------------------------- 177

8.1. RECOMMENDATIONS FOR FUTURE RESEARCH --------------------------------------------------- 186

8.2. SUMMARY -------------------------------------------------------------------------------------------- 187

REFERENCE ---------------------------------------------------------------------------------------------- 189

APPENDIX ------------------------------------------------------------------------------------------------- 203

xi
Acknowledgements
Exploring new things or discovering new knowledge is a relentless journey. Four years ago, when

I was doing my master’s at Charles Sturt University, Professor Rob Davidson inspired me to

undertake a research journey. I would like to thank Professor Davidson for his encouragement and

inspiration. Since then, I have received continuous support from many people.

Professor Rob Davidson has given me his endless support since the start of my research journey.

This doctoral project continued from an original master’s project which began from an interest

arising from my eight years’ experiences in medical x-ray imaging as a radiographer and later as

a clinical application specialist for the Dextroscope, a virtual reality system for neurosurgical

planning. Professor Davidson’s expertise in medical x-ray imaging, his advice on my research

progress, and his vision and direction for research inspired me to accomplish this doctoral project.

I would like to thank Professor Davidson for the support in allowing me to use the radiation science

labs both at Charles Sturt University and the University of Canberra, for organising experiments;

data collection at The Canberra Hospital, for a scholarship from the Faculty of Health at the

University of Canberra (without this scholarship, I could not have completed this project), for the

inspiration in developing the methods used in this project, for his friendship in supporting me

throughout my candidature, and for other support and help not yet mentioned.

Special thanks goes to Mr Yuming Yin, who helped me with the radiation transmission method

used in my Monte Carlo simulation. I would like to thank Dr Graeme L White for the comments

on and suggestions for my articles and thesis and for support in performing the experiments at The

Canberra Hospital. I would like to thank Mr Bill Shelley and his colleagues from the Faculty of

Arts and Design at the University of Canberra for helping make the radiation blockers. I would

xv
like to thank the friendly staff from the Department of Radiology at The Canberra Hospital for

their support in using the x-ray equipment.

I would like to thank my family for supporting me in undertaking a journey in research. Dora and

my children, Samantha and Albert, have understood my desire to explore new things and have

allowed me time that could have been spent with them. Mum and Dad are always encouraging me

exploring and learning. Sorry Mum, I should have spent more time with you in China during your

final moments of life in April 2016.

Finally, I would like to thank my supervisors once more for spending so much of their time to help

me. At Charles Sturt University, thank you, Dr Han Swan, for your help with statistics. Thank you,

Dr Kelly Spuur, for your expert help in mammography. Mr Warren Lusby, thank you for your

expert help in anatomy and radiographic positioning. Dr Allan Ernest, Dr Xiaocheng Zhu, Dr

Lihong Zheng, Dr De Li Liu, and Dr Liang Bin, thank you for your friendships. Dr Liang Bin,

thank you for helping me with MatLab. Following my transfer of candidature to the University of

Canberra, I would like to thank Associate Professor Stuart Semple, my supervisor, and Dr Richard

Keegan for supporting my candidature. Thank you, Dr Peter Copeman, for your expert help in

presentation and pitching.

xvi
Abstract
Medical imaging is commonly used to provide clinicians with a diagnosis for their patients, or to

follow up on a patient’s treatment. Medical imaging incorporates the use of x-rays, ionising

radiation, and non-ionising radiation. One of the major challenges in x-ray imaging is to

substantially remove scatter radiation reaching the image receptor without increasing a patient’s

exposure to ionising radiation, and without compromising image diagnostic quality.

X-ray image diagnostic quality is established by the differentiation attenuations of primary

radiation in imaging anatomy. Differentiation attenuations emanate from interactions between

primary radiation and the matter of the imaging anatomy. These interactions may scatter primary

radiation and produce scatter radiation, some of which moves towards the image receptor.

Attenuated primary radiation and scatter radiation reaching the image receptor form the image.

The intensities of the scatter radiation deform the differentiation attenuations in the attenuated

primary radiation. Consequently, the image diagnostic quality is reduced. A common solution for

removing scatter radiation reaching the image receptor is to use an anti-scatter grid technique. This

technique, however, increases patient exposure to ionising radiation.

This project investigated how a grid technique can be used in mammography and general

radiography without this increase in radiation, and without sacrificing image diagnostic quality.

The project addressed the transmission of primary radiation (Tp) and the transmission of scatter

radiation (Ts) in grid materials. In addition, it presented a quantitative relationship between these

transmissions and the stochastic effects of ionising radiation. This relationship suggests a strategy

for the reduction of these effects, an essential element for patient radiation protection outlined in

xvii
the ALARA (as low as reasonably achievable) principle. This project emanates from the scholarly

literature with regard to two areas:

1) Using grids fails to minimise radiation exposure delivered to patients because of the
compensation for the reduction of primary radiation in grid materials.

2) Using grids lacks adequate optimisation of image diagnostic quality because of the
remaining scatter radiation in the image.

In this project, a series of actions to investigate grids were performed in three distinct phases. The

first phase examined quantitative methodologies determining Tp and Ts. Moreover, this phase

developed a new Monte Carlo simulation code system. In addition, this phase proposed a novel

criterion for the determination of strip optimal thicknesses by using the first differential of the

quantum signal-to-noise ratio (SNR) improvement factor (KSNR). Furthermore, this phase also

established a new method to determine radiation transmissions in grid materials, overcoming

limitations in current radiation transmission methods. An article related to the development and

validation of this new method can be seen in appendix A.

The second phase of this project developed a new grid design method that can be used to design

grids to overcome the reduction of primary radiation in grid strips. This phase first analysed the

relationships between transmitting primary radiation and reducing scatter radiation. Factors

associated with these relationships were then revealed and issues for grid design were examined.

Finally, the criteria for designing new grids were evaluated by analysing the relationships between

these factors.

xviii
The third phase of this project determined a solution for reducing scatter radiation reaching the

image receptor. This solution neither increases patient exposure to radiation nor sacrifices image

diagnostic quality. In this phase, many new mammographic grids and general grids were designed

using the criteria identified in the second phase of the project. The strip optimal thickness of these

new grids was determined using the strip optimal thickness criterion proposed in the first phase.

The designs of these new grids were evaluated using the Monte Carlo simulation code system

developed and validated in the first phase. The results of these new grids were presented in terms

of Tp, Ts, and KSNR.

The results for these new grid designs showed that it is possible to have Tp approximately equal

to the Tp of a perfect grid (Tp = 1). The Ts of these new designs depends on their grid ratio: the

higher the grid ratio, the smaller the Ts. Compared to contemporary grids in the literature, new

general grid designs and new mammographic grid designs have 39% and 28% greater Tp,

respectively, and all of them have smaller Ts. Furthermore, the KSNR of these new designs showed

that new grids could have a benefit of improved image diagnostic quality, regardless of anatomical

thicknesses, either in mammography or general radiography. It was found that current

mammographic grids have improved image diagnostic quality only for thick breasts (greater than

approximately 5 cm). The new mammographic grid designs, however, will have improved image

diagnostic quality for all breasts, no matter whether they are thin or thick.

In conclusion, the new grid designs proposed in this project substantially remove scatter radiation

reaching the image receptor, do not increase radiation exposure to patients, and do not compromise

image diagnostic quality. Using new grids that use these new designs will not increase the

stochastic effects of ionising radiation in either mammography or general radiography. Using such

xix
new grids to replace current contemporary grids will not only result in less scatter radiation

reaching the x-ray image receptor, but the stochastic effects will also be reduced by more than

approximately 39% in general radiography and 28% in mammography. Furthermore, scatter

radiation reaching the x-ray image receptor can be further reduced by using such new grids with a

high grid ratio and without increasing stochastic effects.

Using grids built with these new designs, an effective ionising radiation reduction management

strategy in the implementation of the ALARA principle can be achieved. This means that the

lowest radiation exposure can be achieved using such new grids to produce x-ray images without

compromising image diagnostic quality.

xx
Table of Abbreviations
ALARA As low as reasonably achievable

BEIR Biological Effects of Ionisation Radiation

COV Coefficient of variation

CT Computed tomography

CTDI CT dose index

DDREF Dose rate effectiveness factor

DLP Dose-length product

DNA Deoxyribonucleic acid

EPSM Engineering & Physical Science in Medicine

ESD Entry skin dose

FOV Field of view

HVL Half-value layers

IAEA International Atomic Energy Agency

ICRP International Commission on Radiological Protection

KAP KERMA area product

KERMA Kinetic Energy Released per MAss

LNT Linear-no-threshold

MGD Mean/average glandular dose

MSAD Multiple scan average dose

MTF Modulation transfer function

RAM Random access memory

RMS Root mean square

SNR Signal-to-noise ratio

SPR Scatter-to-primary ratio

SSE Sum of squared error

xxi
Glossary of terms

absorbed dose the energy deposited through ionisation in unit mass of irradiated
material

air-gap technique a way of reducing scatter radiation reaching the image receptor by
keeping a gap in between the imaging object and the image receptor

ALARA (as low as a radiation protection principle, which requires x-ray examinations
reasonably to be performed with the lowest possible radiation dose delivered to
achievable) principle the patient and without compromising the image diagnostic quality

anti-scatter grid a device used in x-ray imaging systems to reduce scatter radiation
reaching the image receptor

area density also known as density thickness, it is calculated as the mass per unit
area. The SI derived unit is kilogram per square metre (kg.m-2)

attenuation (radiation) the gradual loss in the intensity of radiation

azimuth angle (vector an angle, in a 3-dimensional Cartesian system, with its z-axis
in 3-dimensional alinged in the vector’s direction and the tail of the vector fixed at
Cartesian system) the Cartesion system’s origin, between the positive x-axis and the
perpendicular projecton of the vector down onto the x-y plane

binding energy the energy that holds an electron to a nucleus


(electron)

bremsstrahlung a process of x-ray production through deceleration of a moving


electron. From German meaning breaking radiation

xxiii
cancerous risk a chance of developing cancer

collimator a device that is placed between the x-ray tube and the patient and
which uses high attenuation materials, typically lead, to control the
shape and size of the x-ray field

Compton scattering discovered by Arthur Holly Compton, the inelastic scattering of


photons by electrons, with a decrease in the photon energy

computed radiography a type of digital radiography that uses photostimulable phosphors


(PSPs) to capture a latent image on the image receptor

computed tomography an x-ray imaging system that uses x-rays to create cross-section
images of the body

contrast the difference in intensities between adjacent areas in an image

diagnosis (x-ray the identification of the nature of an illness or problem by


examination) examining the x-ray image of the anatomy

digital image the use of computer algorithms to perform image processing on


processing digital images

digital imaging system an x-ray imaging system that captures and displays images by
digital means. A digital radiographic system is an example of a
digital imaging system

direction cosines (of a the cosines of the angles between the vector and the three
vector) coordinate axes

xxiv
DNA a molecule that carries the genetic instructions used in the growth,
(deoxyribonucleic development, functioning, and reproduction of all known living
acid) organisms and many viruses. It has four bases – adenine (A),
cytosine (C), guanine (G), and thymine (T), which attach to a sugar-
phosphate to form a complete nucleotide

electromagnetic (EM) a type of radiation that consists of electric and magnetic fields at 90
radiation degrees to each other. x-rays, light, and radio waves are examples of
EM radiation

energy fluence the amount of energy passing through a unit cross-section area

exposure time the period during which x-rays are emitted from the x-ray tube

filtration the removal of x-ray photons from the beam by attenuation when
the beam is passed through a medium

focal spot the area on the anode of the x-ray tube where x-ray photons are
produced

free path length (x- a distance that the x-ray passes without interactions
ray)

Gray (Gy) a derived unit of ionising radiation dose

grid cut-off undesirable absorption of primary radiation by grid materials

half-value layer also known as half-value thickness, a measurement of the effective


energy of the x-ray beam

health risk a chance of developing diseases

heritable risk a chance of a child inheriting a transmissible disease from a parent

xxv
image clarity (x-ray) the quality of being easy to visualise anatomic structures in the x-
ray image

image contrast (digital the difference between the intensities of adjacent regions in the
image) image

image diagnostic the image quality that is formed by only the intensities of the
quality attenuated priamry radiation behind the imaging anatomy

image receptor an array of detectors or a film capable of converting x-ray


intensities to form an image

imaging object an anatomy or phantom undergoing the x-ray examination

intensity (energy) radiation energy per unit area

intensity (photons) number of photons per unit area

ionising radiation EM radiation of sufficient energy to remove an electron from its


binding atom

latent image the period between exposure to ionising radiation and the
appearance of diseases, such as cancer

mammographic an imaging system investigating breast internal tissues by using x-


system rays

mammography a projection imaging technology using x-rays to produce breast


images of breast, glandular, and adipose tissues

mantissa the part of a floating-point number that represents the significant


digits of that number

xxvi
mass energy the fraction of mass attenuation coefficient that gives rise to the
absorption coefficient total energy locally deposited at the site of interaction

mass energy transfer the fraction of mass attenuation coefficient that gives rise to the
coefficient kinetic energy of the charged particle, and, if any, the charged
particle binding energy

mono-energetic beam an x-ray beam comprising photons of all the same energy

Monte Carlo a computerised mathematical technique using a statistical and


simulation random sampling technique to solve problems using conventional
and deterministically mathematic approaches

non-cancerous risk a chance of developing diseases other than cancer

strip optimal thickness for a given grid ratio and strip height, the grid strip thickness with
which the grid has the highest value of the quantum-to-signal-noise
improvement factor

phantom a device that mimics radiographic characteristics of human


anatomy. Typically used to evaluate radiographic systems without
exposing humans to ionising radiation

photoelectric the emission of electrons when photons are absorbed in matter


absorption
(photoelectric effect)

photoelectric effect the emission of electrons when photons are absorbed in matter
(photoelectric
absorption)

photon a bundle of energy of EM radiation

xxvii
photon fluence the number of photons or particles passing through a unit cross-
section area

photon flux the rate of photon fluence

Poisson mottle the appearance of noise in an x-ray image because of low signal-to-
noise ratio in the x-ray intensities reaching the image receptor

polar angle (vector in the photon’s angle in a 3-dimensional Cartesian system that is
3-dimensional system) formed between the photon’s direction and the z-axis positive
direction

poly-energetic beam an x-ray beam that is composed of photons of many different


energies

primary radiation radiation that does not change direction during its passage

quantum signal-to- a grid performance quantity measuring the signal-to-noise-ratio


noise ratio (SNR) improvement attributed to the grid
improvement factor
(KSNR)

radiation dose an amount of ionising radiation that is received in an object or


anatomical region. Absorbed dose is a specific measure of radiation
dose

radiation exposure a measure of the ionisation of air ascribed to ionising radiation from
gamma rays or x-rays

radiograph a projection image of the anatomy, produced using x-rays and an


image receptor

xxviii
radiographer a qualified person who is responsible for undertaking a radiographic
examination

radiographic contrast the difference in optical densities between adjacent areas in an


(film) image

radiography an imaging technology using x-rays to produce projection images of


body anatomies to view the internal structures of human bodies

Rayleigh scattering the scattering of photons by atoms, without changing the photon
energy

scatter radiation radiation that changes direction during its passage

scatter-to-primary the ratio of the amount of scatter radiation to the amount of primary
ratio (SPR) radiation

signal-to-noise ratio a ratio of the amount of signal to the amount of noise in a system
(SNR)

slot technique a way of reducing scatter radiation reaching the image receptor by
constraining the x-ray beam in a narrow fan shape

somatic effect cell damage that is passed on to succeeding cell generations

spectrum (beam or x- a plot of the number of photons as a function of photon energy or


ray) photon frequency

stochastic effect health risk effect that occurs by chance after a latent period and
without a threshold dose, and whose probability is proportional to
the cumulative dose

xxix
subject contrast the difference in exit intensities between adjacent areas in the x-ray
beam that results from different properties within the anatomy

technique (x-ray) a way of performing an x-ray examination

technology a device or machinery developed from scientific knowledge

transmission of the ratio of the amount of primary radiation reaching the image
primary radiation (Tp) receptor with a grid to without a grid

transmission of scatter the ratio of the amount of scatter radiation reaching the image
radiation (Ts) receptor with a grid to without a grid

tube heat loading (x- energy deposition on the anode of the x-ray tube
ray)

tube voltage (kVp) the voltage or electrical potential difference that is applied between
the cathode and the anode of the x-ray tube to accelerate the
electrons between the cathode and the anode to generate x-ray
radiation, usually measured in peak kilovolts (kVp)

x-ray examination an investigative procedure using x-rays to produce images for


diagnostic purposes, such as a mammographic examination

x-ray tube a device that generates an x-ray beam and requires a high voltage, a
current, a means of first accelerating then decelerating the electrons,
and a vacuum to achieve the production of x-rays

x-rays a form of electromagnetic radiation having energy ranging from a


few hundred electron volts to several hundred kiloelectron volts

xxx
Chapter 1: The nature and scope of this project

1.1. Introduction

A significant challenge in x-ray imaging is to optimise radiation protection, taking into account

economic and social factors, using the guiding principle of ‘as low as reasonably achievable’

(ALARA) outlined by the International Commission on Radiological Protection (2007a). This

principle requires delivering the lowest radiation dose to patients when performing x-ray

examinations, without compromising the image diagnostic quality.

This project addresses the problem of an anti-scatter grid technique that is commonly used in many

x-ray imaging systems such as mammography and radiography. While this technique reduces

scatter radiation reaching the image receptor, it also reduces primary radiation reaching the image

receptor. Compensation for the reduction in primary radiation is needed to avoid increasing the

image quantum noise, but this compensation increases patients’ radiation exposure.

1.2. Background
1.2.1. Overview of medical x-ray imaging

X-rays were discovered by Röntgen (1896), a German physicist and mechanical engineer. He used

the mathematical designation ‘x’ for something unknown when he speculated on a new kind of ray

from electrical discharge passing through a vacuum tube. These new rays caused shimmering in

the barium platinocyanide screen. Röntgen used x-rays to produce an x-ray image of his wife’s

1
hand, and this image gave birth to a new tool used in the field of medical diagnostics. In 1901,

Röntgen received the Nobel Prize for his discovery of x-rays.

Since the last century, x-rays have been used for the diagnosis and treatment of many conditions.

In the area of medical diagnostics, x-rays1 are used in a wide variety of imaging technologies, for

example, mammography, to produce diagnostic images. Such technologies have been described in

detail in several textbooks on medical imaging physics (Bushberg et al., 2011, Flower, 2012,

Bushong, 2013). Some x-ray diagnostic imaging technologies are given in Table 1.1.

Table 1.1. Types of medical diagnostic x-ray imaging technologies (Bushberg et al.,
2011).
Imaging technology Characteristics of images

Mammography Single projection of breast tissue

Radiography Single projection of body or body part

Series of single projections of body or body part to display


Fluoroscopy
functional and anatomical motion

Cross-sectional views of body or body part reconstructed


Computed tomography (CT)
from multiple projections

In planar/projection x-ray imaging technologies (such as mammography or radiography), the three

basic components are the x-ray tube, imaging anatomy, and image receptor. First, primary

radiation is generated at the x-ray focal spot. Then the primary radiation is constrained using high

atomic material, such as lead diaphragms, in a cone beam, rectangle beam, or another beam shape.

The primary radiation is directed to the imaging anatomy.

1
In this thesis, the terms ‘X-rays’, ‘photons’, and ‘radiation’ are used interchangeably, unless otherwise stated.

2
When the primary radiation is travelling in the anatomy, some primary radiation in the beam

interacts with the anatomy’s matter. Differentiation attenuations emanate from the interactions

between the primary radiation and the anatomy’s matter. These interactions may scatter the

primary radiation and produce scatter radiation, some of which moves towards the image receptor.

Attenuated primary radiation and scatter radiation reaching the image receptor form the image.

1.2.2. Ionising radiation

In x-ray imaging, photon energy is often measured in kiloelectron volts (keV). Photons are

generated from the x-ray tube, and their energy is approximately 10–150 keV. These photons may

interact with tissue through Rayleigh scattering, Compton scattering, or photoelectric absorption

(Bushberg et al., 2011).

Atoms can be approximately described using the Rutherford-Bohr atomic model (Podgoršak,

2006). In this model, an electron is bound to the nucleus of an atom. The electron is orbiting the

nucleus in one of the atom’s shells. The electron’s orbiting force is the attraction force between

the positive charge of protons in the nucleus and the negative charge of the electron.

In Compton scattering or photoelectric absorption, a photon interacts with a bound electron in an

atom and causes the electron to be ejected from the atom. The photon’s ability to eject bound

electrons is known as the ionising effect. Photons that can cause the ionising effect are known as

ionising radiation. x-rays used in x-ray imaging are ionising radiation.

1.2.2.1. Rayleigh scattering

Rayleigh scattering is known as coherent interaction in that an incident photon is deflected and

this photon’s energy is preserved. As illustrated in Figure 1.1, in Rayleigh scattering, the electrons

3
of an atom act as a whole to interact with an incident photon and change the photon’s direction.

During this interaction, no electrons are ejected, and atom ionisation does not occur (Bushberg et

al., 2011, Carlton et al., 2012).

Figure 1.1. Rayleigh scattering. The electrons of an atom act as a whole to interact with an
incident photon and change this photon’s direction.

Rayleigh scattering dominates mainly in x-ray photons with very low energy, such as photon

energy of 10–35 keV used in mammography. In soft tissue, Rayleigh scattering accounts for less

than 5% of interactions for photon energy greater than 70 keV and for about 10% for photon energy

less than 30 keV (Bushberg et al., 2011, Carlton et al., 2012).

1.2.2.2. Compton scattering

Compton scattering is an incoherent interaction in that a photon interacts with an electron and,

after the interaction, this photon’s direction and energy are changed, as illustrated in Figure 1.2. In

Compton scattering, a photon is more likely to interact with a loosely bound or ‘free’ electron.

4
Figure 1.2. Compton scattering. An incident photon of energy E0 interacts with a bound
electron of binding energy Eb. After the interaction, the scattered photon with energy E c is
deflected away from this incident photon’s direction. The ejected electron is emitted with kinetic
energy Ee-. The direction of this electron forms an angle (β) with this incident photon’s direction.
β is less than 90 degrees.

During Compton scattering, part of the incident photon’s energy is transferred to the bound

electron (Bushberg et al., 2011, Carlton et al., 2012). This electron gains kinetic energy and escapes

from the atom. The kinetic energy plus this electron’s binding energy equals the energy lost by the

photon. The photon is scattered and becomes a scatter photon. This electron’s direction is confined

to an angle, which is not more than π/2 radians with respect to this photon’s direction before the

interaction (Attix, 2008). Compton scattering results in scatter radiation and ionisation.

1.2.2.3. Photoelectric absorption

Photoelectric absorption is known as the photoelectric effect. In photoelectric absorption, a photon

interacts with a bound electron. The electron absorbs all the photon’s energy and then escapes from

the atom. The electron’s kinetic energy equals the photon’s energy minus the electron’s binding

5
energy (Figure 1.3). Photoelectric absorption only occurs if the photon energy is greater than the

electron binding energy (Bushberg et al., 2011, Carlton et al., 2012).

In photoelectric absorption, a photon most likely interacts with the electrons whose binding

energies are the closest to, but less than, the photon’s energy. As illustrated in Figure 1.3, photons

with energy exceeding the K-shell binding energy are most likely to interact with the K-shell

electrons through photoelectric absorption.

Following photoelectric absorption with a K-shell electron, the atom is ionised with a K-shell

electron vacancy. This vacancy will be most likely filled with an electron from a nearby shell (in

this example, the L-shell) with a lower binding energy. The electron filling this vacancy, however,

can be from any shell (M-shell or N-shell) with a lower binding energy. As it escapes from its shell,

this electron creates another vacancy, which then is filled with an electron from an even lower

binding energy shell. Thus, an electron cascade from outer shells to inner shells occurs.

In an electron cascade, the difference in binding energy is released as photons known as

characteristic x-rays. A bound electron, typically from the same shell of the cascading electron,

may absorb the characteristic x-ray that is emitted by the cascading electron. If, after absorbing the

characteristic x-ray, this bound electron escapes from the shell, this electron then is known as an

Auger electron (Bushberg et al., 2011).

6
Figure 1.3. Photoelectric absorption. A 100-keV photon is undergoing an interaction with an
atom of 4-electron shell. In this example, a K-shell electron is ejected with a kinetic energy
equal to the incident photon’s energy minus the K-shell binding energy. The vacancy created
in the K-shell results in the electron’s transition from the L-shell to the K-shell. The difference
in the K-shell and L-shell binding energies results in the emission of a 28 keV characteristic x-
ray. The electron cascade continues and results in the production of other characteristic x-rays
of lower energies; for example, 4 keV and 1 keV. Auger electrons of various energies may be
emitted in lieu of the characteristic x-ray emissions (not shown in this diagram).

The emission of Auger electrons and the emission of characteristic x-rays are competing processes.

The probability of characteristic x-ray emission decreases as the atomic number of the interaction

material decreases. Soft tissue is mostly composed of materials of low atomic matter.

Characteristic x-ray emission does not frequently occur in soft tissue; Auger electron emission

predominates (Bushberg et al., 2011).

1.2.3. Harmful effects of ionising radiation

Photons may interact with anatomical matter when they travel, which may cause ionisation in the

anatomy’s tissues or organs. This ionisation has detrimental effects, for example, breaking

deoxyribonucleic acid (DNA) strands or killing cells (Graham et al., 2003, Bushberg et al., 2011),

7
which can be categorised as deterministic effects or stochastic effects (International Commission

on Radiation Protection, 2007b).

1.2.3.1. Deterministic effects of ionising radiation

Deterministic effects of ionising radiation are acute damage to organs and tissues. This damage

occurs as loss of tissue function or organ function, such as cell death and, in extreme cases, death

of the irradiated individual. Deterministic effects have a threshold dose, and the estimated

threshold doses for various tissues and organs can be found in the International Commission on

Radiological Protection (2012).

The threshold doses of the deterministic effects depend on several factors, including the type of

the ionising radiation, the type of irradiated organ/tissue, the volume of the tissue exposed, and the

clinical effect on the organ/tissue exposed. When the patient’s radiation dose exceeds the threshold

dose, the severity of the deterministic effect increases as the patient’s radiation dose increases

(Khong et al., 2013, International Commission on Radiation Protection, 2007b, Hall, 2006).

The estimates of threshold doses may be lower than what is recommended. Investigations have

shown that radiation-induced eye cataracts and circulatory disease occur at lower radiation doses

than their current threshold doses (International Commission on Radiological Protection, 2007b).

1.2.3.2. Stochastic effects of radiation exposure

Stochastic effects of ionising radiation are health risks, including cancerous risk, heritable risk,

and non-cancerous risk (Kadhim et al., 2013). These risks usually have a latent period, which may

8
be from several years to more than ten years (International Commission on Radiation Protection,

2007b).

The cancerous risk results from damage to genes by direct or indirect radiation energy deposited

in the nuclear DNA. Unrepaired or wrongly repaired damage in a cell’s DNA may result in the

development of cancer in the individual exposed (Kadhim et al., 2013), known as a somatic effect.

If the damage results in any disease in future generations of the individual exposed, this damage

is known as a heritable effect (International Commission on Radiation Protection, 2007b). The

heritable risk of radiation exposure is mostly evidenced by the offspring of Japanese atomic-bomb

survivors (Preston et al., 1994, Preston et al., 2003, Preston et al., 2007). The non-cancerous risk

induced by low-dose ionising radiation includes cataract, atherosclerotic disease, inflammatory

responses, and myocardial infarctions (Schultz-Hector and Trott, 2007, Little et al., 2008b, Little,

2010, Baker et al., 2011, Picano et al., 2012).

The probability of the occurrence of stochastic effects is proportionally related to the individual’s

cumulative radiation dose. The severity of the stochastic effects is not related to radiation dose.

For example, a cancer induced by 2 Gray (Gy) is no worse than a cancer induced by 0.01 Gy.

Stochastic effects have no threshold dose, which means a single instance of unrepaired damage in

a cell’s genetic material could cause cancer or a hereditary defect (Mossman, 2006, Tubiana et al.,

2006).

1.2.3.2.1. Evidence of stochastic effects

The stochastic effects of patient radiation exposure are evidenced by a wide range of investigations,

such as cancer incidence in the offspring of the Japanese atomic-bomb survivors (Little et al.,

9
2009b, Little, 2003), development of cancers from experimental animals (International

Commission on Radiological Protection, 2007b), and the significantly high rates of cancer among

irradiated populations. These irradiated populations include patients who have undergone long-

term fluoroscopy studies, been exposed to radiation at an early age, or been treated with radiation

therapy, as well as individuals working near accelerators and other sources of ionising radiation.

Among female tuberculosis patients who have undergone extensive diagnostic fluoroscopy, an

increase in breast cancer has been reported in a 30-year follow-up investigation of 4940 women

(Boice Jr et al., 1991). This investigation showed significantly higher rates of breast cancer (147

breast cancers occurred in contrast to the 113.6 expected), which was not apparent until about 10–

15 years after the initial fluoroscopy examinations. No higher rates of breast cancer (87 observed

versus 100.9 expected) were found among women treated without radiation exposure. The results

also showed that age at radiation exposure strongly influenced the risk of radiation-induced breast

cancer, with young women being at the highest risk and those over the age of 40 being at lowest

risk. This finding is supported by a United States scoliosis cohort study of 5573 females who were

exposed to radiation of a much lower cumulative dose before the age of 20 and had a statistically

significant increased risk of breast cancer (Doody et al., 2000). The estimated mean-cumulative

radiation dose to the breast was 108 mGy. When the scoliosis cohort study was limited to

individuals with breast doses between 10 mGy and 90 mGy, the significantly higher risk remained.

Higher cancer risks were reported for various types of cancers in patients who had undergone

radiation therapy. After treatment of haemangioma, a significantly higher risk for breast cancer

was found for a mean dose of 290 mGy to the breast with a range of 0–35.8 Gy at 95% confidence

interval (Eidemüller et al., 2009, Eidemüller et al., 2011). Similar results were found in females

10
who were treated for postpartum mastitis with doses typically from 1 to 6 Gy (Hall, 2006). An

increase in lung cancer had been reported in patients with Hodgkin’s lymphoma who had received

treatment by irradiation doses of 5 Gy or more (Travis et al., 2002, Dores et al., 2002).

Leukaemia is one of the malignant cancers most likely to be linked to radiation exposure. It is

commonly diagnosed in x-ray workers, physicists, and engineers working near accelerators and

other sources of ionising radiation (Little et al., 2009a, Hall and Giaccia, 2012).

The risk of cancer due to ionising radiation exposure has been widely reported and supported with

evidence. The stochastic effects of radiation exposure are unavoidable. One of the essential

elements in radiation protection is to quantitatively predict these effects.

1.2.3.2.2. Prediction of stochastic effects: linear-no-threshold

The linear-no-threshold (LNT) theory predicts that stochastic effects are proportional to the

cumulative radiation dose. A simple causative relationship between radiation dose and cancer risk

is found from most reliable animal data, excluding mouse data for thymic lymphoma and ovarian

cancers (Little et al., 2008a).

The International Commission on Radiation Protection (2007a) uses dose and dose rate

effectiveness factor (DDREF) values to model the risks of radiation-induced cancer. DDREF

values for doses at or below 2 Gy have a value of 2. In comparison, the Biological Effects of

Ionisation Radiation (BEIR) Committee uses the Bayesian statistics of the combination of the life

span study of atomic-bomb survivors and selected animal studies (National Research Council,

11
2006). The BEIR VII study (National Research Council, 2006) claimed that DDREF values in the

range of 1.1–2.3 are consistent with the Bayesian statistics.

The International Commission on Radiation Protection (2007a) determined DDREF values by

examining the subjectivity and probabilistic uncertainty of radiation-induced cancers. In contrast,

the BEIR VII (National Research Council, 2006) uses a DDREF value of 1.5 to predict cancer risk

by focusing on the subjectivity inherent in choices of DDREF values. The International

Commission on Radiation Protection (2007a) retained a DDREF summary value of 2 for

radiological protection. The predictions of sex-averaged nominal risks of the stochastic effects of

radiation exposure can be referred to in reports by the International Commission on Radiation

Protection (2007a) and the National Research Council (2006).

The stochastic effects of patient radiation exposures are unavoidable and are proportional to the

patient’s cumulative radiation dose. The International Commission on Radiation Protection (2007a)

has developed the radiation protection guidance for the practice of medical x-ray imaging.

1.2.4. Radiation protection

The prime goal of radiation protection is to prevent the occurrence of deterministic effects and to

minimise the stochastic effects, not only for patients but also for healthcare professionals. The

International Commission on Radiological Protection (ICRP, 2007a) developed principles for

medical x-ray examinations that relate to justification, optimisation, and dose minimisation. The

justification refers to the fact that every radiation exposure that a patient receives must be

associated with a positive net benefit. The optimisation and dose minimisation mean that all

radiation exposures must be kept ALARA without compromising image diagnostic quality.

12
A significant challenge with the ALARA principle is to produce the acceptable image diagnostic

quality with the lowest possible radiation dose delivered to patients.

1.2.5. Effects of scatter radiation


Clinicians rely on adequate image contrast to be able to visualise anatomical differences in the

image. Image contrast is the difference between adjacent intensities in an image. Adequate contrast

can increase clinicians’ confidence in making diagnoses. In the process of image production,

contrast can be described at each production stage. In x-ray imaging, two common concepts of

contrast are the subject contrast and image contrast.

Subject contrast is the fundamental contrast that emanates from the intensities of the x-ray beam

exiting the imaging object. Subject contrast is converted to a latent image stored in the image

receptor, and a final image is processed from the latent image. The difference observed in the final

image is the image contrast. In film-screen systems, the image contrast processed from the latent

image cannot be altered. In digital systems, however, the image contrast can be manipulated when

it is displayed on a viewing device, such as a monitor (Bushberg et al., 2011).

1.2.5.1. Subject contrast


Subject contrast originates from the interactions between photons and tissue and can be measured

as the difference in intensities between adjacent areas in the exit beam. Factors influencing the

subject contrast are photon energy and the tissues. A tissue’s physical properties, including the

element components, densities, thicknesses, and shapes, produce subject contrast.

A tissue’s physical properties generally cannot be altered. Compressions can be applied in limited

x-ray examinations to enhance the subject contrast by changing the tissue thickness; for example,

mammography with a compression technique (OLeary et al., 2011). Subject contrast can be

13
enhanced by altering the physical properties by using a contrast agent, such as potassium iodine-

sulphate or barium solutions. A natural way of enhancing the subject contrast is introducing air;

for example, chest x-ray taken in full inhalation.

Photon energy can be altered to enhance subject contrast because the probability of interactions

between photons and tissue depends on the photons’ energy. An x-ray beam generated with a peak

tube voltage has a range of photon energies. This beam is a poly-energetic beam and is

characterised by the anode materials, filtration materials, and peak kilovoltage (kVp) applied to

the x-ray generator. The beam’s properties can be measured in a spectrum. Most x-ray generators

are built with fixed anode materials and filtration materials, but some may be constructed with

selectable filtration materials, especially in mammographic systems.

1.2.5.2. Image contrast and radiographic contrast


Most x-ray images are presented in digital form as digital images. A digital image numerically

represents a two-dimensional image as a number of rows and columns of pixels. For digital x-ray

images, the image contrast is the difference between the pixel values of adjacent regions in the

image. Image contrast can be changed by changing pixel values.

A digital image’s pixel values can be altered by changing the window level and width of the

image’s display. A digital image window is the range of values for the pixels. A digital image

window’s level is the middle point of this range, and a digital image window’s width is the width

of this range.

Digital images’ pixel values can be changed by advanced image processing. There are many

advanced image processing algorithms, such as look-up tables, histograms, spatial enhancement,

14
unsharp masks, contrast-enhancement masks, dynamic range control, and multi-scale processing

(Davidson, 2006).

Radiographic contrast of x-ray films is a measure of the difference between the optical density of

adjacent areas in the film (Davidson, 2006). Radiographic contrast can be measured using

densitometers calibrated for the same luminous condition as that of an x-ray film lightbox.

Measuring two optical densities can be used to determine radiographic contrast.

The formation of film optical density is a chemical process. A film’s optical density is a function

of film radiation exposure and is often presented in graph form, which has the shape of a sigmoid

curve (s-curve), known as the characteristic curve. The characteristic curve is divided into a tow,

a latitude (in the middle), and a shoulder. In the latitude, the curve’s slope is approximately straight,

and a film’s radiation exposures within this latitude produces useful film optical contrast. Hence,

a limited range of film radiation exposure is able to produce a desirable optical density.

Radiographic contrast, by its definition, is unchangeable. However, the visual perception of the

radiographic contrast can be optimised to a limited extent by changing the visible luminance level,

such as viewing an (over-exposed) film with a bright spotlight.

The visual perception of x-ray films and x-ray digital images can be optimised by changing the

viewing conditions. The optimum perception of x-ray films or images can be explained by the

Rose (1948) model, which uses the statistics of photon Poisson noise in a background region to

determine an absolute scale for the performance of the human visual system. This model explored

the quantum nature of light to study the limits of human visual system performance. Sturm and

MORGAN (1949) successfully used this model to determine the fundamental role that photon

statistics would play in limiting the clarity of intensification images for visualisation.

15
The Rose (1948) model is valid for low-contrast signals, as pointed out by Burgess (1999).

However, it is only an adequate approximation for determining the optimum visual perception of

films’ optical density and digital images’ brightness and contrast.

1.2.5.3. Expression of optical contrast


Optical contrast is a visual perception of features. The expression of optical contrast can be made

in different ways, such as Weber contrast, Michelson contrast, Ratio contrast, or Root mean square

contrast. Each method may be used for different purposes.

Weber contrast is calculated as the ratio of the luminance of the feature to the luminance of the

feature’s background (Equation 1.1) and is also referred to as ‘Weber fraction’. Weber contrast is

suitable for small features that are present on a large uniform background. It can be used to

determine radiographic contrast of x-ray films (Webber et al., 1977, Koedooder and Venema,

1986).

|𝐼−𝐼𝑏 |
𝐶= (1.1)
𝐼𝑏

where C is the Weber contrast, I, is the luminance of the features, and, Ib is the
luminance of the background.

Michelson contrast is known as the visibility calculated as Equation 1.2. It is commonly used for

patterns where both bright and dark features are equivalent and take up similar fractions of the

area. Michelson contrast is used to determine x-ray image contrast (International Commission on

Radiation Units and Measurements, 1989, Strid, 1976, Sisniega et al., 2013) and is defined as:

16
𝐼𝑚𝑎𝑥 −𝐼𝑚𝑖𝑛
𝐶= (1.2)
𝐼𝑚𝑎𝑥 +𝐼𝑚𝑖𝑛

where C is the Michelson contrast, Imax is the highest luminance of an area, and Imin is
the lowest luminance of this area.

Ratio contrast (Equation 1.3) is the ratio of the transmitted luminance I1 of an area to the

transmitted luminance I2 of this area’s surrounding (Bonenkamp and Hondius Boldingh, 1959a,

Morgan, 1946b):

𝐼1
𝐶= (1.3)
𝐼2

where C is the ratio contrast, I1 is the transmitted luminance of an area, and I2 is the
transmitted luminance of this area’s surrounding.

Root mean square (RMS) contrast (Equation 1.4) is often used in image signal analysis. RMS

contrast is determined by the standard deviation of the pixel intensities (Peli, 1990):

1
𝐶= √ ∑𝑁−1 𝑀−1 ̅2
𝑖=0 ∑𝑗=0 (𝐼𝑖𝑗 − 𝐼 ) (1.4)
𝑀×𝑁

where intensities Iij are the ‘i-th’ and ‘j-th’ element of the two-dimensional image of
size M by N. Ī is the average intensity of all pixel values in the image. The image (I) is
assumed to have its pixel intensities normalised in the range [0, 1].

17
1.2.5.4. Effects of scatter radiation on image contrast

Scatter radiation, when it reaches the image receptor, together with primary radiation, creates a

latent image. During the process of converting the latent image to a digital image, scatter radiation

cannot be separated from primary radiation and therefore cannot be removed. The effects of scatter

radiation on image contrast can be quantified using Webber contrast. When imaging system noise,

which may also affect image contrast, is neglected, the image contrast without scatter radiation is

equal to the subject contrast, which can be calculated by Equation 1.5. When scatter radiation is

not removed, the image contrast can be calculated by Equation 1.6.

|𝐷𝑝𝑟𝑖𝑚𝑎𝑟𝑦 −𝐷𝑏𝑎𝑐𝑘𝑔𝑟𝑜𝑢𝑛𝑑 |
𝐶= (1.5)
𝐷𝑏𝑎𝑐𝑘𝑔𝑟𝑜𝑢𝑛𝑑

|𝐷𝑝𝑟𝑖𝑚𝑎𝑟𝑦+𝑠𝑐𝑎𝑡𝑡𝑒𝑟 −𝐷𝑏𝑎𝑐𝑘𝑔𝑟𝑜𝑢𝑛𝑑+𝑠𝑐𝑎𝑡𝑡𝑒𝑟 |
𝐶= (1.6)
𝐷𝑏𝑎𝑐𝑘𝑔𝑟𝑜𝑢𝑛𝑑+𝑠𝑐𝑎𝑡𝑡𝑒𝑟

where D is a measure of radiation in the region of interest, the subscript ‘primary’


refers to the primary radiation in this region, ‘background’ refers to the primary radiation
in this region’s surrounding, and ‘scatter’ refers to scatter radiation.

The reduction of image contrast attributed to scatter radiation is a function of scatter radiation to

primary radiation ratios (SPR). In radiography, SPR varies from less than 1 to more than 25 for

imaging anatomy thicknesses of 5–50 cm (Fetterly and Schueler, 2007), assuming the image

contrast without scatter radiation is 0.2. The reduction of this image contrast is illustrated in Figure

1.4. As SPR increases from 0 to 9, this image contrast reduces from 0.2 to approximately 0.02.

18
0.2
(Webber conrast)
Image contrast

With scatter radiation


Without scatter radiation
0.1

0
0 2 4 6 8
SPR

Figure 1.4. Reduction of the image contrast attributed to scatter radiation. The image contrast
reduces as the SPR increases.

Tromans et al. (2010) found the reduction of image contrast attributed to scatter radiation is

irreversible in the image. They attempted to remove the effect of scatter radiation on digital images

and found that the ‘noise’ ascribed to the scatter radiation cannot be separated from the primary

radiation or the signal.

The detrimental effects of scatter radiation on image contrast is permanent and cannot be undone.

Substantially reducing the scatter radiation reaching the image receptor is essential to create

adequate image contrast for making accurate diagnoses.

1.2.6. Overview of scatter radiation reduction techniques

In medical x-ray imaging, x-ray radiation is generated from a finite focal spot. This radiation is

called primary radiation. Scatter radiation is radiation that, during passage through the imaging

object, changes in direction.

The air-gap technique can also reduce scatter radiation reaching the image receptor. In the air-gap

technique, a large space is kept between the imaging object and the image receptor, by moving the

image receptor away from the imaging object. The reduction of scatter radiation increases as this

19
space increases (Sorenson and Floch, 1985). When the space increases, radiation exposures are

increased to compensate for the reduction of primary radiation intensities. This reduction is

ascribed to the distance increased between the focal spot and the image receptor (Neitzel, 1992,

Krol et al., 1996, Alzimami et al., 2009).

The slot technique reduces scatter radiation reaching the image receptor by keeping the exposing

region within a narrow slot region. This technique is like a single slice CT in which the primary

radiation beam is confined to form a narrow (for example, 1 cm) fan beam. Then the fan beam is

guided over the patient during radiation exposure (Jaffe and Webster, 1975, Barnes et al., 1976,

Barnes et al., 1993). The radiation exposure time for using the single slot technique is extremely

long because of the limited coverage of a single fan beam. X-ray tube heat loading is approximately

increased by the ratio of the imaging size to the slot width but can be reduced by using multiple

slots. The high tube heat loading, complex mechanical control system of multiple slots, long

radiation exposure time, and difficulty in controlling radiation exposure time makes the

implementation of the slot technique difficult (Barnes et al., 1993).

Current attempts to reduce scatter radiation focus on software solutions (Mentrup et al., 2014,

Takahiro Kawamura et al., 2015, Renger et al., 2016). These attempts study scatter radiation

distribution and subtract an estimate of scatter radiation to obtain a new image. Some applications

of these attempts have been found, such as in bedside x-ray examinations. They improve image

quality, but not as well as the anti-scatter grid technique. Furthermore, these attempts have a

fundamental challenge in that the scatter radiation distribution of an anatomical region may vary

between individuals. An inaccurate estimation of scatter radiation might introduce artefacts, which

can lead to misinterpretation of the image.

20
In 1913, Bucky (1913) introduced a grid in between an image receptor and an imaging object to

take an x-ray image of that object. He found that the improved clarity of this image was attributed

to the reduction of scatter radiation by the grid. The technique is known as an anti-scatter grid

technique and is the most common technique used in medical x-ray imaging systems to reduce

scatter radiation reaching the image receptor.

1.2.7. Anti-scatter grids

1.2.7.1. Mechanism of scatter radiation reduction by grids

Anti-scatter grids consist of a periodic array of radiopaque foil strips separated by radiolucent

material that fills the strip interspaces. The mechanism of scatter radiation reduction by grids was

first described by Bucky (1913). The description of this mechanism can be found in several

textbooks, such as Bushberg et al. (2011), Carlton et al. (2012), and Bushong (2013).

An illustration of an anti-scatter grid’s mechanism is given in Figure 1.5. After primary radiation

is generated at the focal spot, the primary radiation is confined to a shape, such as a rectangular

shape, and is directed to the irradiated object. Some primary radiation travels through the object

and reaches the grid. Some primary radiation is scattered by the object’s matter to produce scatter

radiation which may reach the grid. Some primary radiation is absorbed in the object’s matter.

Most scatter radiation reaching the grid is absorbed in the radiopaque strips. Most primary

radiation reaching the grid travels through the radiolucent interspaces that are aligned to the

primary radiation’s direction. Thus, most scatter radiation is stopped from reaching the image

receptor, and most primary radiation travels through the grid and reaches the image receptor.

21
Figure 1.5. Reduction of scatter radiation by an anti-scatter grid. Some primary radiation and
some scatter radiation travel through the grid and reach the image receptor. Some primary
radiation and some scatter radiation are absorbed in the grid’s strip materials.

There are different types of grids, such as focused grids, parallel grids, cross-hatch grids, and

tapered grids. Focused grids are linear grids with their strips aligned to a common line focus and

are the most common type of grids implemented in medical x-ray imaging systems.

Grid strips can be made from materials of high atomic number, such as tungsten or lead. High

atomic number materials absorb a great fraction of radiation travelling through them. The

interspace materials can be made from radiolucent materials, such as aluminium, polymethyl

methacrylate (PMMA), wood, or carbon fibre. These materials let a great fraction of radiation

travel through them.

22
1.2.7.2. Reduction of primary and scatter radiation by grids

The reduction of primary radiation and scatter radiation in grid materials can be quantified as the

transmissions of primary and scatter radiation. The transmission of primary radiation (Tp) is the

ratio of primary radiation reaching the image receptor with a grid in place to that without a grid in

place. The transmission of scatter radiation (Ts) is the ratio of scatter radiation reaching the image

receptor with a grid in place to that without a grid in place.

In the scholarly literature, the Tp of carbon-interspaced grids decreases from 0.764 to 0.640 as the

grid ratio increases from 8:1 to 21:1 (Fetterly and Schueler, 2007, Fetterly and Schueler, 2009). A

similar trend in Tp was shown in Kim et al. (2007), in which Tp decreases from 0.76 to 0.71 as the

grid ratio increases from 5:1 to 10:1. Generally, Tp varies from 0.64 to 0.77 with a mean of

approximately 0.72 (SD = 0.034) (Fetterly and Schueler, 2007, Kim et al., 2007, Fetterly and

Schueler, 2009) for general grids or from 0.74 to 0.80 with a mean approximately 0.78 (SD =

0.038) for mammographic grids (Carton et al., 2009, Salvagnini et al., 2012).

Ts decreases from 0.235 to 0.058 as the grid ratio increases from 8:1 to 21:1. A similar trend in Ts

was also shown in Kim et al. (2007), in which Ts decreases from 0.40 to 0.30 as the grid ratio

increases from 5:1 to 10:1. In general, Ts varies from 0.058 to 0.456 with a mean of approximately

0.212 (SD = 0.142) (Fetterly and Schueler, 2007, Kim et al., 2007, Fetterly and Schueler, 2009)

for general grids or from 0.125 to 0.196 with a mean of approximately 0.156 (SD = 0.046) for

mammographic grids (Carton et al., 2009, Salvagnini et al., 2012).

Grids reduce a large fraction of scatter radiation and a small fraction of primary radiation reaching

the image receptor. For example, a typical grid ratio 10:1 grid has Tp equal to 0.75 and Ts equal

to 0.17 (Fetterly and Schueler, 2007). When this grid is used in an x-ray examination of 20-cm

23
thick abdomen, which has an SPR of approximately 6:1 (Fetterly and Schueler, 2007), (Fetterly

and Schueler, 2007), the relative amount of scatter radiation to the amount of primary radiation

reaching the image receptor would be approximately 1.34:1. As illustrated in Section 1.2.5, this

relative amount scatter radiation is a major degradation factor reducing the image contrast.

1.2.7.3. Patient radiation dose increase attributed to grids

As photons carry specific amounts of energy and exhibit the properties of waves and particles

(Carlton et al., 2012), the photons’ statistics can be determined using Poisson distribution which

governs the statistics of random events that are independent of each other and time (International

Atomic Energy Agency, 2014). Poisson statistical variance of photons can be measured in the

coefficient of variation (COV) for COV = 1/sqrt(A), where ‘A’ is the number of photons and sqrt(A)

is the square root of ‘A’ (International Atomic Energy Agency, 2014, Bushberg et al., 2011). The

photons’ COV increases as the number of these photons decreases.

Grids reduce the number of primary photons and hence increase these photons’ COV. The effect

of this increase on the x-ray image can be perceived as Poisson mottle (Grangeat, 2009). An x-ray

image’s Poisson mottle is reduced by increasing radiation exposure, which determines the number

of photons generated from the x-ray focal spot.

An increase in radiation exposure, which is ascribed to the compensation for the reduction on the

primary radiation in the grid materials, increases patient radiation exposures. This increase is

approximately 39% for general grids (mean Tp = 0.72) and 28% for mammographic grids (mean

Tp = 0.78) and is equal to the increase in stochastic effects.

24
1.3. Research opportunity from the literature

The research opportunity for this project emanates from the scholarly literature with regard to two

areas:

1) Using grids fails to minimise radiation exposure delivered to patients because of the
compensation for the reduction of primary radiation in grid materials.

2) Using grids lacks adequate optimisation of image diagnostic quality because of the
remaining scatter radiation in the image.

1.4. Motivations

Grids have been used in x-ray imaging systems since the 1920s. Contemporary grids have average

Tp and Ts equal to 0.72 and 0.211 for general grids and 0.78 and 0.156 for mammographic grids,

respectively. Without using grids, especially x-ray images of anatomies thicker than 20 cm, images

will not be created with adequate contrast for making sensible diagnoses. When grids are used, the

stochastic effects are increased by 39% in general radiography and 28% in mammography.

This project was motivated by the fact that the use of grids will not change in the coming decades.

If primary radiation in grids is not reduced, increases in the stochastic effect attributed to using

grids will continue.

1.5. Purpose

The primary purpose of this project is to investigate anti-scatter grids and their Tp and Ts. In

addition, this project will also associate these Tp and Ts with the stochastic effects of ionising

radiation as a stochastic effect reduction management strategy the implementation of the ALARA

principle.

25
1.6. Significance

This project will make a significant and original contribution to the scholarly literature and

provides a potential solution in reducing stochastic effects. This solution will also be useful for

medical x-ray imaging systems, particularly in the areas of dose-image quality optimisation. First,

a significant contribution relates to the reduction of stochastic effects as a means of addressing the

problem of the radiation transmissions in grids, as a radiation dose management strategy, and as a

practical and important option in the implementation of the ALARA principle. Second, as an

original contribution, this study reveals new knowledge in grid design.

Finally, this project is worthy of research ascribed to the recent increasing concerns regarding the

stochastic effects of low-dose ionisation radiation exposure from medical imaging (Schultz-Hector

and Trott, 2007, Little et al., 2008b, Little, 2010, Baker et al., 2011, Picano et al., 2012). When

radiation doses fall, stochastic effects fall proportionally. However, as there is no threshold dose,

this means that any radiation dose, no matter how small, has the potential to cause harmful effects,

such as cancer or leukaemia. The chance of these harmful effects occurring increases

proportionally as the patient’s radiation dose increases.

The use of grids, which are of great benefit in patient disease management, increases the stochastic

effects of ionising radiation and is a serious and important problem, for which solutions have been

sought since 1913 (Wilsey, 1923, Morgan, 1949, Bonenkamp and Hondius Boldingh, 1959b, Strid,

1976, Kalender, 1982, Chan et al., 1985, Fetterly and Schueler, 2007, Fetterly and Schueler, 2009,

Cunha et al., 2010). A solution to this problem will be a significant step in the quest to reduce

ionising radiation delivered to patients without compromising image diagnostic quality. Thus, this

26
solution will minimise the probability of stochastic effects on patients undergoing x-ray

examinations.

1.7. Overview of methodology

This research has three distinct phases. In the first phase, this project will quantitatively evaluate

Tp and Ts. As the manufacture of many prototype grids is not only expensive but requires a lot of

time, experimental evaluation of many prototype grids is not feasible in this project. An alternative

grid evaluation method, the Monte Carlo simulation, will be developed and validated. In the second

phase, this project will explore the relationship between Tp and Ts. A new grid design method will

be established, which can assist in designing new grids with Tp increased and Ts reduced

concurrently. In the third and final phase, many new grids will be designed and evaluated in the

Monte Carlo simulation.

1.8. Research objectives

The objectives of this research are:

(1) to develop a Monte Carlo code system for grid evaluation

(2) to validate the Monte Carlo code system

(3) to develop a new grid design method

(4) to design grids using this new method

(5) to evaluate new grids using the Monte Carlo simulation code system.

27
1.9. Research contributions

The main goal of this research is to investigate grids, to address radiation transmission in grids,

and to associate these transmissions with the stochastic effects of radiation exposure as a stochastic

effect reduction management strategy in the ALARA principle. This goal was achieved through a

series of actions performed in three distinct phases and through original contributions to the

scholarly literature. These contributions are:

(1) A new grid design method: This new grid design method provides criteria to design grids
with negligible reduction of primary radiation and without compromising the reduction of
scatter radiation. This method solved the problems of the reduction of primary radiation in
grid materials. In other words, this method provided a solution for designing new grids with
the expectation that this will improve the transmission of primary radiation, for example, Tp >
0.99, without compromising the reduction of scatter radiation.

(2) A novel criterion to determine the strip optimal thickness: This novel criterion employs
the first differential of the quantum signal-to-noise ratio improvement factor (KSNR) to
determine a condition for the determination of strip optimal thicknesses. This criterion
assumed that if the maximum KSNR exists, it will exist at the zero value of its first differential.

(3) A new method for determining radiation transmissions in grid materials: This new
method for determining of radiation transmissions in grid materials was used in the Monte
Carlo simulation to determine radiation transmissions in grid materials. This method
overcame the limitations of those set out in the literature. The development and validation of
this method has been published (Zhou et al. (2016).

(4) A new Monte Carlo simulation code system: The new Monte Carlo simulation code system
can be used to simulate x-ray imaging systems. This system was developed in Matlab Ver.
R2015b (The MathWorks Inc., Natick, Massachusetts) and validated for the evaluation of
grids. This system has the advantage of simulating millions of photons’ transportation
simultaneously by using the array operation in the MatLab software environment. Validation

28
of this Monte Carlo code system was presented at the Engineering and Physical Science in
Medicine 2016 conference and is in preparation for publication.

(5) Many new grid designs: Many new grid designs were created and evaluated in this project
through Monte Carlo simulation. The application of new grids from these new designs will
reduce negligible amounts of primary radiation without sacrificing the reduction of scatter
radiation. These new grids can supersede current grids and be used as a stochastic effect
reduction management strategy.

1.10. Organisation of this thesis

This thesis is organised into eight chapters. Chapter 1 introduced the nature and scope of this study

by identifying, in the scholarly literature, problems with contemporary grids, the need for a

solution to these problems, and the research objectives.

Chapter 2 will review methodological approaches that can be used in this project to conduct

quantitative measurements of radiation transmission in grids. In addition, Chapter 2 will

investigate a novel criterion using the first differential of KSNR to determine the strip optimal

thickness for a given grid ratio and strip height.

Chapter 3 will elucidate information for the development of a new Monte Carlo simulation code

system for simulating x-ray imaging. These principles will be used to guide the development of a

new Monte Carlo simulation code system, which will be used in this project to evaluate many new

grid designs.

Chapter 4 will clarify the simulation of photon transportation in grid materials. It will first review

methods for determining radiation transmissions in grid materials and then explain the limitations

29
of these methods. Following that, this chapter will continue to establish and validate a new method

for determining radiation transmission in grid materials.

Chapter 5 will present the validation of the new Monte Carlo code system developed using the

principles reviewed in Chapter 3.

Chapter 6 will focus on a new grid design method to overcome the reduction of primary radiation

in grid strips. This chapter will first present an overview of current grid design, will then discuss

grid design factors and issues, and finally will examine the criteria for designing new grids.

Chapter 7 will apply the achievements made in Chapters 2 to 6 to determine a solution for the

reduction of primary radiation in grid strips. This chapter will present many new grid designs

which will be evaluated through Monte Carlo simulation.

Chapter 8 will present the interpretation of the results made in this project and draw conclusions

from these results. Furthermore, recommendations for future research and summary will be

outlined in this chapter.

30
Chapter 2: Grid evaluation methodology

This chapter marks the start of the first phase and highlights a methodology for quantitative

evaluation of the transmissions of primary (Tp) and scatter (Ts) radiation. It has five main sections:

the first reviews criteria for grid evaluation; the second proposes a novel criterion for determining

strip optimal thicknesses; the third focuses on radiation measurements in grid evaluation; the fourth

presents the setups for radiation measurements in the experiment; and the last discusses the

variation of these setups for simulation.

2.1. Grid evaluation criteria

Grid evaluation is used to quantify the reduction of scatter radiation, the reduction of primary

radiation, and their effects on image contrast. Scatter radiation reduces image contrast and leads

to difficulty in making accurate diagnoses whereas primary radiation carries information about

anatomical structure to form a useful image. The amount of reduction of scatter and primary

radiation are used to derive grid performance criteria. New grid criteria for digital imaging

technologies have been added in recent years.

2.1.1. Brief history of grid evaluation criteria

Approximately four decades after Bucky (1913) invented anti-scatter grids to improve x-ray image

clarity, Bonenkamp and Hondius Boldingh (1959a) carried out systematic investigations of grid

performance and presented grid performance criteria; these are widely used today to determine

grid performance.

31
The basic grid performance criteria are Tp, Ts, and the transmission of total radiation (Tt). All

three are fundamental quantities that indicate grid performance and are essential quantities for

derivations of grid performance criteria.

The derived grid performance criteria are the image contrast improvement factor (Morgan, 1946a,

Bonenkamp and Hondius Boldingh, 1959a), the Bucky factor, the quantum signal-to-noise ratio

improvement factor (Dick and Motz, 1978, Neitzel, 1992), and the image quality improvement

factor (International Electrotechnical Commission, 2013). The basic criteria and these derived

criteria are discussed in the following sections.

2.1.2. Basic grid evaluation criteria: Tp, Ts, and Tt

Tp, Ts, and Tt are characteristics of grids. Tp is the ratio of the amount of primary radiation with

a grid to without a grid and is calculated using Equation 2.1. Ts is the ratio of the amount of scatter

radiation with a grid to without a grid and is calculated using Equation 2.2. Tt is the ratio of the

amount of total radiation with a grid to without a grid and is calculated using Equation 2.3. Tp, Ts,

and Tt are used to derivate other grid performance factors, of which common ones are further

discussed below.

𝑃𝑟𝑖𝑚𝑎𝑟𝑦_𝑅𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛𝑤𝑖𝑡ℎ 𝑔𝑟𝑖𝑑
𝑇𝑝 = (2.1)
𝑃𝑟𝑖𝑚𝑎𝑟𝑦_𝑅𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛𝑤𝑖𝑡ℎ𝑜𝑢𝑡 𝑔𝑟𝑖𝑑

𝑆𝑐𝑎𝑡𝑡𝑒𝑟_𝑅𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛𝑤𝑖𝑡ℎ 𝑔𝑟𝑖𝑑
𝑇𝑠 = (2.2)
𝑆𝑐𝑎𝑡𝑡𝑒𝑟_𝑅𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛𝑤𝑖𝑡ℎ𝑜𝑢𝑡 𝑔𝑟𝑖𝑑

𝑇𝑜𝑡𝑎𝑙_𝑅𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛𝑤𝑖𝑡ℎ 𝑔𝑟𝑖𝑑
𝑇𝑡 = (2.3)
𝑇𝑜𝑡𝑎𝑙_𝑅𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛𝑤𝑖𝑡ℎ𝑜𝑢𝑡 𝑔𝑟𝑖𝑑

32
2.1.3. Image contrast improvement factor: K
The image contrast improvement factor (K) is also known as the image contrast improvement ratio.

K is a combination effect of Tp and Tt and is a measure of the transmission of a large fraction of

primary radiation and a small fraction of scatter radiation. In film-screen imaging systems, the

transmitted primary radiation carries anatomical information while the transmitted scatter radiation

helps ensure desirable film optical density. K is an acceptable measure of the compromise between

the film image diagnostic quality and patient radiation dose (Bonenkamp and Hondius Boldingh,

1959a). It is calculated using Equation 2.4.

𝑇𝑝
𝐾= (2.4)
𝑇𝑡

where Tp is the transmission of primary radiation, Tt is the transmission of total


radiation, and K is the image contrast improvement factor.

2.1.4. Bucky factor: B


When a grid is used, a large fraction of total radiation is stopped from reaching the image receptor.

In film-screen systems, an increase in radiation exposure to compensate for this fraction is needed

to ensure that the film’s radiation exposure is within the film’s latitude (discussed in Section

1.2.5.2). In digital imaging systems, an increase in radiation exposure to compensate for the

reduction of primary radiation is needed to avoid increasing the image Poisson mottle (discussed

in Section 1.2.7.2).

The relative increase in radiation exposure is described by the grid Bucky factor (B), which

accounts for the compensation for the reduction on both primary radiation and scatter radiation.

33
This factor is useful for screen-film systems but not for digital imaging systems (Fetterly and

Schueler, 2007). It is calculated using Equation 2.5.

1
𝐵= (2.5)
𝑇𝑡

where Tt is the transmission of total radiation and B is the Bucky factor.

2.1.5. Quantum signal-to-noise ratio improvement factor: KSNR


The quantum signal-to-noise ratio (SNR) improvement factor (KSNR) was introduced for digital

image detectors (Dick et al., 1978). Digital images’ contrast and brightness can be altered through

digital image processing algorithms (referred to in Section 1.2.5). A digital image’s diagnostic

quality mainly depends on the image’s SNR, and the SNR depends primarily on the ratio of

primary radiation to scatter radiation. KSNR is a measure of a digital image’s SNR improvement

(Dick and Motz, 1978, Neitzel, 1992, Fetterly and Schueler, 2007) and is calculated using Equation

2.6.

𝐾𝑆𝑁𝑅 = 𝐾 √𝑇𝑡 (2.6)

where Tt is the transmission of total radiation; K is the image contrast improvement


ratio; and KSNR is the quantum signal-to-noise ratio improvement factor.

2.1.6. Image quality improvement factor: Q


The image quality improvement factor (Q) is similar to KSNR and is another useful grid

performance indicator for digital image detectors (International Electrotechnical Commission,

2013). Q can be calculated using Equation 2.7.

34
𝑇𝑝2
𝑄= = 𝑇𝑝 × 𝐾 (2.7)
𝑇𝑡

where Tp and Tt are the transmissions of primary and total radiation, respectively; K
is the image contrast improvement ratio; and Q is the image quality improvement factor.

2.2. Maximum KSNR for determination of strip optimal thickness


KSNR can be calculated using Equation 2.8. As the strip thickness approaches zero, Tp, Ts, and Tt

approach one and KSNR approaches one. As the strip thickness approaches infinity, Tp, Ts, and Tt

approach zero and KSNR is likely approaching zero because the elements in the right side of

Equation 2.8 are approaching zero. However, the KSNR of grids as set out in the scholarly literature

is generally greater than one. The KSNR of a carbon-fibre-interspaced grid with a grid ratio of 8:1

is about 1.37 (Fetterly and Schueler, 2007). These KSNR values suggest that KSNR has a maximum

value for strip thicknesses from zero to infinity.

𝑃𝑔+𝑆𝑔 𝑇𝑝×𝑃+𝑇𝑠×𝑆
𝐾𝑆𝑁𝑅 = 𝐾 √𝑇𝑡 = 𝐾 √ = 𝐾√ (2.8)
𝑃+𝑆 𝑃+𝑆

Where Pg is primary radiation transmitted the grid; Sg is the scatter radiation


transmitted the grid; P is the primary radiation without the grid; S is the scatter radiation
without the grid with Tt = (Pg + Sg) / (P + S), Tp = Pg / P, and Ts = Sg / S; Tp, Ts, and Tt
are the transmissions of primary, scatter, and total radiation, respectively; K and KSNR are
the contrast improvement ratio and the quantum signal-to-noise ratio improvement factor,
respectively.

Strip optimal thickness is defined as the strip thickness with which, for a given grid ratio, strip

height, and cover material and thickness, the grid has the maximum KSNR. Directly finding the

35
strip optimal thickness can be problematic, but it can be indirectly determined by finding the

maximum KSNR.

KSNR can be expressed as a function of grid ratio, strip height, strip thickness, cover thickness, and

the linear attenuation coefficients of the strip, cover, and interspace materials (Equation 2.9). For

a given grid ratio, strip height, strip and interspace materials, cover material and thickness, the

maximum KSNR can be found at the condition in which the first differential of Equation 2.9, with

respect to the strip thickness, is equal to zero (Equation 2.10). After the maximum KSNR is found,

the strip optimal thickness is then determined at the maximum KSNR.

𝐾𝑆𝑁𝑅 = 𝑓(𝑟, ℎ, 𝑑, 𝑐𝑡ℎ 𝜇𝑠 , 𝜇𝑖 , 𝜇𝑐 ) (2.9)

𝛿𝐾𝑆𝑁𝑅 𝛿𝑓(𝑟,ℎ,𝑑,𝑐𝑡ℎ 𝜇𝑠 ,𝜇𝑖 ,𝜇𝑐 )


= =0 (2.10)
𝛿𝑑 𝛿𝑑

where, r is grid ratio; h is strip height; d is strip thickness; cth is cover thickness; μs, μi,
and μc are the strip, interspace, and cover materials’ linear attenuation coefficients,
𝛿𝑦
respectively; 𝛿𝑥 is the first differential of function y with respect to variable x; and KSNR is
the quantum signal-to-noise ratio improvement factor.

2.3. Radiation measurement for grid evaluation

The amount of radiation measured is proportional to the amount of signal that is produced when

the radiation interacts with the detector materials. This signal depends on the amount of energy

deposited in the detector materials. The energy deposited in a material depends on three factors:

the density of the material, the mass energy absorption coefficients, and the thickness of the

material (Hubbell and Seltzer, 1995, Hubbell, 2006).

36
As x-ray beams used in x-ray imaging are usually poly-energetic and have a spectrum of energy,

the energy deposition in the detector materials can be specified at each photon energy or expressed

as an ‘effective’ value over the beam spectrum reaching the detector. The x-ray beam spectrum

depends on the peak kilovoltage (kVp) applied to the generator and the characteristics of the

generator (such as anode material, anode angle, and filtrations). The x-ray beam spectrum is also

influenced by, if presented, the imaging anatomy which attenuates, or ‘hardens’, the beam and

gives the beam a higher average energy.

The amount of radiation energy deposited in a detector depends on the material thickness, mass

density, and mass energy absorption coefficients of the detector. The mass energy absorption

coefficient generally is higher at lower photon energy and decreases as photon energy increases.

However, if the material has an atomic absorption edge, which is always true for high atomic

materials (e.g. Gadolinium has absorption edges at 50.24 kilo-electron-volt (keV) and 8.37 keV),

the energy deposition increases dramatically above the absorption edge, which causes a local

minimum energy deposition immediately below the absorption edge.

The amount of signal produced by photon energy deposition in a detector depends on the

interactions between the photons and the detector materials. Though photons may interact with

matter through Rayleigh scattering, Compton scattering, or photoelectric absorption, the main

interaction with the material of a high atomic number detector is photoelectric absorption. In

photoelectric absorption, the interaction of a photon with the detector materials causes an electron

to be emitted from its atom. In the electron’s subsequent loss of kinetic energy, excitation and

ionisation occur and produce secondary signals, such as optical photons or electronic charges.

37
The amount of radiation can be measured in different quantities, such as total energy, absorbed

dose, photon fluence, and radiation energy fluence. For x-ray images, ideally all energy in the

beam should be converted to an image signal, but many radiation quantities do not measure the

total energy of the beam. The absorbed dose measures energy deposition characterised by the mass

energy absorption coefficient but the absorbed dose is not a measure of the total energy of the

beam. Photon fluence is a measure of the total number of photons per unit area but does not account

for the photon energy. Radiation energy fluence is a measure of the total energy per unit area and

is proportional to the total energy of the beam.

In grid evaluation, radiation measurements are made with a variety of detectors, for example x-ray

film (Wilsey, 1921), Ge detector (Ter-Pogossian et al., 1974), gas ionisation chamber (Strid, 1976),

NaI crystal (Barnes et al., 1976, Dick et al., 1978), fluorescent screen with photo-detector

(Bonenkamp and Hondius Boldingh, 1959a, Niklason et al., 1981), and photostimulable storage

phosphor plate (Floyd et al., 1991, Flynn and Samei, 1999, Fetterly and Hangiandreou, 2001,

Fetterly and Schueler, 2007, Fetterly and Schueler, 2009). The International Electrotechnical

Commission (1978, 2001) recommended a photo-detector, similar to the detector used in

Bonenkamp and Hondius Boldingh (1959a), with a calcium-tungstate-phosphor fluorescent screen.

As this screen is not commercially available anymore, it has been replaced with gadolinium

oxysulphide fluorescent screen (International Electrotechnical Commission, 2013). Using a photo-

detector, lights that are emitted from the fluorescent screen are measured. The results of radiation

transmission then depend on the detection efficiency of the photo-detector and the light-conversion

efficiency of the fluorescent screen, which is a function of the screen area density and radiation

beam quality.

38
In grid evaluation, differences between radiation detectors could not be excluded from affecting

grid evaluation results. Floyd et al. (1991) found that scatter fraction measured with

photostimulable storage phosphor plates was in acceptable agreement with those measured with

films and most other previously reported measurements. Floyd et al. (1991) did not account for

differences in experimental setups, such as phantom sizes, primary beam’s field of view (FOV),

and kVp. These differences have been found to affect the distribution of SPR (Boone and Seibert,

1988). The SPR in Floyd et al. (1991) for the 20 cm × 20 cm × 20 cm (W × L × H) phantom is

approximately the same as the SPR in Fetterly and Schueler (2007) for the 30 cm × 30 cm × 20

cm (W × L × H) phantom. Theoretically, these SPR should be different from each other. Difficulty

in experimental setups has been found to influence grid evaluation results (Fetterly and Schueler,

2007, Cunha et al., 2010).

In the Monte Carlo simulation, correction can be made for radiation measurements from different

detectors which depend on x-ray beam quality and photons’ entry angles. Shuping and Judy (1977)

found that for a constant radiation exposure (measured in air-absorbed dose), the energy deposited

in a fluorescent screen increased as the tube kVp increased. This result shows that fluorescent

screens yield different amounts of light for the same air-absorbed dose of different x-ray beam

quality. Photons’ entry angles to the detector entry surface affect the photons’ travelling distances

in the materials of the detector. These distances affect the probabilities of interactions between

photons and the materials in the detector. Hence, these angles affect the photons’ energy deposition

in the detector materials.

39
2.4. Experimental setup for grid evaluation

Grid evaluation requires the measurement of scatter, primary, and total radiation with and without

a grid. These measurements were obtained with a beam stop method.

The amount of scatter radiation can be quantified with a modulation transfer function (MTF)

method (Salvagnini et al., 2012). MTF measures how accurately an imaging system reproduces

the details of an object to its image (Boreman, 2001). In x-ray imaging, MTF is a measure of the

transfer of the subject contrast to the image contrast. Boone and Seibert (1988) demonstrated the

effect of scatter radiation on image contrast by comparing MTF with and without scatter radiation.

Saunders Jr and Samei (2006) further examined the effect of scatter radiation on MTF using Monte

Carlo simulation and found a similar result as Boone and Seibert (1988). Their result showed that

scatter radiation reduces low frequency in MTF. Salvagnini et al. (2012) extended this result to

quantify scatter radiation using the reduction of low frequency in MTF. The MTF method requires

the determination of the imaging system’s MTF.

In the beam stop method, measurements of the amount of primary, scatter and total radiation were

performed under three conditions: a narrow beam condition, a broad beam condition with a

primary beam blocker, and a broad beam condition (Bonenkamp and Hondius Boldingh, 1959a,

Floyd et al., 1991, Fetterly and Schueler, 2009).

2.4.1. Narrow beam condition

The amount of primary radiation was measured in the narrow beam condition with the radiation

beam tightly confined to the radiation detection area with diameter of approximately 6 mm (Figure

2.1). The phantom (water phantom for general grids, and Polymethyl methacrylate (PMMA)

phantom for mammographic grids) was placed as close as possible to the x-ray source/focal spot.

40
Lead diaphragms were placed above and below the phantom. The radiation beam entering the

phantom and reaching the detector only travelled through the area confined by these diaphragms.

Another set of lead diaphragms was placed in between the detector and the phantom to further

narrow the radiation beam to the radiation detection area.

The radiation beam was centred to the grid which was placed at its focal distance from the x-ray

source. The detector was placed underneath the grid without a space. Lead diaphragms were placed

underneath the detector to prevent back scatter radiation. The primary radiation was then measured

with and without the grid.

Figure 2.1. Arrangement of x-ray source, phantom, grid, and detector in the narrow beam
condition. (a) Front setup view of general grid evaluation. (b) Side setup view of
mammographic grid evaluation.

41
2.4.2. Broad beam condition with a primary beam blocker

The amount of scatter radiation was measured in the broad beam condition with a primary beam

blocker. In this condition, the centre of the primary beam was blocked by a lead blocker (Figure

2.2). The grid was placed at its focal distance from the x-ray source/focal spot. The detector was

placed underneath the grid without a space. Lead diaphragms were placed underneath the detector

to prevent back scatter radiation.

The phantom (water phantom for general grids, and Polymethyl methacrylate (PMMA) phantom

for mammographic grids) was placed above the grid and as close as possible to it. The projection

of the lead blocker was to just cover the entire radiation detection area with a diameter of

approximately 6 mm. The thickness of the lead blocker was adequate to substantially stop primary

radiation reaching the detection area. The primary beam was confined to an area of 30 cm × 30 cm

at the phantom exit-surface for general grid evaluation or 15 cm × 15 cm for mammographic grid

evaluation. The scatter radiation was then measured with and without the grid.

42
Figure 2.2. Arrangement of x-ray source, phantom, grid, and detector in the broad beam
condition with a blocker. (a) Front setup view of general grid evaluation. (b) Side setup view of
mammographic grid evaluation.

2.4.3. Broad beam condition

The setup for the broad beam condition (water phantom for general grids, and Polymethyl

methacrylate (PMMA) phantom for mammographic grids) was the same as for the broad beam

condition with a lead blocker, except that the lead blocker was removed (Figure 2.3). The amount

of total radiation was measured with and without the grid.

43
Figure 2.3. Arrangement of x-ray source, phantom, grid, and detector in the broad beam
condition. (a) Front setup view of general grid evaluation. (b) Side setup view of
mammographic grid evaluation.

2.5. Simulation setup for grid evaluation

The simulation setups for grid evaluation can be made the same as the experimental setups. These

setups do not account for the forward scatter radiation that arises in the narrow beam condition,

nor do they account for the scatter radiation that otherwise would arise in the phantom where

primary radiation is blocked in the broad beam condition with a lead blocker. These setups,

however, would require two times more simulation time because the measurements require three

different setups.

In this project, the simulation setup was made the same as the broad beam condition. With this

setup, primary radiation and scatter radiation were labelled by primary and scatter markers. The

energies of photons distinguished by markers were tallied for the calculations of Tp and Ts. This

44
setup accounted for the forward scatter radiation arising in the primary beam’s FOV in the narrow

beam condition and setup accounted for the scatter radiation which otherwise would not arise in

the blocked region underneath the lead blocker.

Summary

This chapter examined the criteria for grid evaluation with an emphasis on the basic criteria: Tp,

Ts, and Tt. These basic criteria were used to calculate the derivative grid evaluation criteria. The

derivative grid evaluation criterion, KSNR, was discussed for the determination of grid strip optimal

thickness. A novel criterion, the first derivative of KSNR, was proposed to determine the optimal

thickness for a given grid ratio and strip height.

Furthermore, this chapter reviewed the difference between the measurements of the amount of

radiation in grid evaluation and highlighted the beam-block method for the measurements of

primary, scatter, and total radiation. This beam-block method determined primary, scatter, and

total radiation using three measurement conditions: narrow beam condition for primary radiation,

broad beam condition with a primary beam blocker for scatter radiation, and broad beam condition

for total radiation. In addition, this chapter also discussed the differences in radiation measurement

between the experimental and simulation setups.

These grid evaluation criteria and the beam-block method for radiation measurement were used in

the experiment to determine grid performance or in the Monte Carlo simulation to determine grid

design performance. The next chapter will discuss the development of our Monte Carlo simulation

code system for grid evaluation.

45
Chapter 3: Information for the development of the
Monte Carlo code system

This chapter presents the second part of the first phase in this project. It discusses the simulation

of x-ray imaging and presents an overview of Monte Carlo simulation. It then delineates a

procedure for simulating the production of x-ray images. After that, the chapter details photon

transportation and interactions between photons and matter. Finally, the radiation dosimetry and

imaging phantom construction used in the Monte Carlo simulation are discussed. A dedicated

Monte Carlo simulation code system will be developed using the information discussed in this

chapter.

3.1. Introduction

Monte Carlo simulation was first used for exploring the behaviour of neutron chain reactions in

fission devices in the 1940s. At that time, physicists were unable to understand the behaviour of

neutron chain reactions using conventional and deterministically mathematical approaches.

Stanislaw Ulam and John von Neumann proposed a statistical and random sampling technique to

solve problems like the behaviour of neutron chain reactions. Simulation codes were developed

using this statistical and random sampling technique, and these codes were successfully executed

on an electronic computing machine at the Los Alamos Scientific Laboratory (Eckhardt, 1987).

Monte Carlo simulation has since been used to resolve a wide range of complex physical and

mathematic problems. In Monte Carlo simulation, a problem can be fit into a model in which an

event happens in a random manner that depends on a sequential sample of random numbers. For

47
different sequences of random numbers, the simulation does not yield identical results; however,

they agree with each other within some errors due to statistic variation (Landau and Binder, 2009).

Monte Carlo simulation can virtually replicate x-ray imaging technologies to simulate the

production of x-ray images. In addition, radiation energy deposited in the imaging phantom can

be determined to estimate the organ dose received by patients. If a phantom is constructed to

represent an anatomical structure, radiation energy deposited in this phantom can be used to

estimate radiation doses received by the organs or tissues in this anatomy.

Monte Carlo simulation is a preferable method used to study x-ray imaging systems. It has been

used to investigate scatter radiation (Kalender, 1981, Jing et al., 1998, and Boone and Cooper III,

2000), evaluate anti-scatter grid performance (Kyriakou and Kalender, 2007, Jang et al., 2008,

Lazos and Williamson, 2010, Tanaka et al., 2013), assess patient radiation dose (Kim et al., 2010,

Cunha et al., 2010), analyse image diagnostic quality (Tomal et al., 2013, Chen et al., 2015), and

validate imaging technologies’ implementation (Yu et al., 2009, Yan et al., 2010, Bornefalk et al.,

2010, Herrnsdorf and Petersson, 2016).

3.2. Procedure for simulating the production of x-ray images

The procedure for simulating the production of an x-ray image started with the generation of x-ray

photons at the x-ray focal spot and finished with the deposition of photon energy in the image

receptor. An x-ray imaging system with three basic imaging components (the x-ray generator, the

imaging object, and the image receptor) is illustrated in Figure 3.1.

48
Figure 3.1. Three basic components of an x-ray imaging system.

During the acquisition of an image, primary photons were produced at the focal spot. They were

constrained to travel towards the imaging object. When travelling inside the imaging object, some

photons might be absorbed in the imaging object’s matter, some might travel through the imaging

object without any interaction and reached the image receptor, and some might interact with the

imaging object’s matter and change their direction to become scatter photons. Some scatter

photons, in their subsequent interaction with the matter, might be absorbed in the matter, whereas

some might successfully travel through the imaging object and reach the image receptor.

The intensities of all photons that had reached the image receptor formed the image. Generally,

simulating the production of an image followed the tracks of photons from the x-ray focal spot to

the image receptor.

49
Following a photon’s track is just like transporting the photon from the focal spot to the image

receptor, or to where the photon energy is completely absorbed, or to where the photon meets

predefined conditions. The transportation of photons in the simulation of an x-ray imaging system

is illustrated using a flowchart in Figure 3.2. Photon transportation proceeded as follows:

1) Generate a photon at the focal spot

2) Determine the photon’s properties: energy, position, direction, and status (primary
or scatter)

3) Determine a free path length in the materials of the imaging object

4) Transport the photon by the distance of the free path length to a new location

5) Is the photon inside the imaging object?

(A) Yes.

What type of interaction?

a. If Rayleigh scattering

i. Change status to scatter

ii. Determine the photon’s new direction

iii. Go to step 3

b. If Compton scattering

i. Change status to scatter

50
ii. Determine the photon’s new energy and direction

iii. Tally energy deposition in the imaging object at the


site of interaction

iv. Go to step 3

c. If photoelectric absorption

i. Tally energy deposition in the imaging object at the


site of the interaction

ii. Go to step 9

(B) No.

i. Go to step 6

6) Transport the photon to the plane in which the detector’s detection area is

7) Is the photon inside the detection area?

(A) Yes. Go to step 8

(B) No. Go to step 9

8) Tally the photon’s energy in the detector

9) End the photon transportation

51
Figure 3.2. Flow chart of photon transportation in Monte Carlo simulation

52
3.3. Photon transportation

Photon transportation included changes in a photon’s location, direction, energy, and status

(primary or scatter). The change in location was determined using a free path length, which is the

distance that a photon travels without any interaction with matter. A free path length was a random

length determined by the linear attenuation coefficient and a random number. The change in

direction depended on the interaction type and two random angles. Each of these changes is

discussed in the following sections.

3.3.1. Change in a photon’s location and direction

Changes in a photon’s location and/or direction were performed in a three-dimensional Euclidean

space. A three-dimensional Cartesian coordinate system was defined by x-, y- and z-axes. A

photon’s location and direction were represented by a vector, which was determined using six

variables: coordinates (x, y, z) and direction cosines (α, β, γ). A constant relationship between

variables α, β, and γ is given in Equation 3.1. The positive z-axis direction was defined from the

x-ray focal spot to the image receptor.

The initial coordinates (x, y, z) of a photon were sampled from the x-ray focal spot area, where

photons are generated. The initial direction cosines (α, β, γ) of this photon were sampled in a

probability distribution function, such as an isotropic direction distribution. When this photon

changed its location by a free path length l, the new coordinates (x’, y’, z’) were calculated using

Equations 3.2, 3.3, and 3.4.

𝛼 2 + 𝛽2 + 𝛾 2 = 1 (3.1)

𝑥 ′ = 𝑥 + 𝑙𝛼 (3.2)

53
𝑦 ′ = 𝑦 + 𝑙𝛽 (3.3)

𝑧 ′ = 𝑧 + 𝑙𝛾 (3.4)

When the direction of a photon was changed, the new direction was determined using the direction

cosines and two random angles: a polar angle θ and an azimuth angle ω (Figure 3.3). The polar

angle was sampled from the interval (0, π]. The azimuth angle was uniformly sampled from the

interval (0, 2π]. All angles were measured with angle units in radians.

Figure 3.3 Angular deflection in an interaction.

If α and β of the direction cosines (α, β, γ) were equal to zero (α = 0 and β = 0), the new direction

cosines (α’, β’, γ’) were calculated using Equations 3.5, 3.6, and 3.7, similar to Salvat (2015p. 36).

54
If α and β of the direction cosines (α, β, γ) were not equal to zero, a polar angle (θ0) and an azimuth

angle (ω0) were calculated using Equations 3.8 and 3.9. The new direction cosines (α’, β’, γ’) were

calculated using Equations 3.10, 3.11, and 3.12. These equations are different from Salvat (2015p.

37) by an angle shift determined by the polar angle θ0. This polar angle θ0 may have a significant

effect on the probability distribution of scatter photons’ directions because the random polar angle

θ is not uniformly sampled. As the azimuth angle ω is uniformly sampled, it does not have any

effect on this probability. In the simulation, the calculation of ω0 was avoided.

𝛼 ′ = 𝑠𝑖𝑛(𝜃) 𝑐𝑜𝑠(𝜔) (3.5)

𝛽 ′ = 𝑠𝑖𝑛(𝜃) 𝑠𝑖𝑛𝑒(𝜔) (3.6)

𝛾 ′ = 𝑐𝑜𝑠(𝜃) (3.7)

𝜃0 = 𝑎𝑐𝑜𝑠(𝛾) (3.8)

𝛼
𝑎𝑐𝑜𝑠 ( ) 𝛽>0
√𝛼 2 +𝛽 2
𝛼
𝜔0 = − 𝑎𝑐𝑜𝑠 ( ) 𝛽<0 (3.9)
√𝛼 2 +𝛽 2
𝛼
𝑎𝑐𝑜𝑠 ( ) 𝛽 = 0, 𝛼 ≠ 0
{ √𝛼 2 +𝛽 2

𝑠𝑖𝑛(𝜃+𝜃0 )
𝛼 ′ = 𝛼 𝑐𝑜𝑠(𝜃 + 𝜃0 ) + (𝛼𝛾 𝑐𝑜𝑠(𝜔 + 𝜔0 ) − 𝛽 𝑠𝑖𝑛(𝜔 + 𝜔0 )) (3.10)
√1−𝛾2

𝑠𝑖𝑛(𝜃+𝜃0 )
𝛽 ′ = 𝛽 𝑐𝑜𝑠(𝜃 + 𝜃0 ) + (𝛽𝛾 𝑐𝑜𝑠(𝜔 + 𝜔0 ) + 𝛼 𝑠𝑖𝑛(𝜔 + 𝜔0 )) (3.11)
√1−𝛾2

55
𝛾 ′ = 𝛾 𝑐𝑜𝑠(𝜃 + 𝜃0 ) − 𝑠𝑖𝑛(𝜃 + 𝜃0 ) 𝑐𝑜𝑠(𝜔 + 𝜔0 ) √1 − 𝛾 2 (3.12)

When a photon changed its location and/or direction, the photon’s new location was calculated

from a free path length. The new direction of this photon was calculated from its direction cosines,

a polar angle and an azimuth angle that was uniformly sampled. The determination of the free path

length and the polar angle is discussed in the following sections.

3.3.1.1. Determining the free path length

A free path length (or free path) is the distance that a photon would travel in matter without any

interaction with the matter. The free path length was determined using the linear attenuation

coefficient and a random number. It could also be determined using the interaction cross-section

or the mass attenuation coefficient.

The interaction cross-section measures the same probability of interactions between x-ray radiation

and matter, as does the mass attenuation coefficient but in different types of units. The interaction

cross-section, the linear attenuation coefficient, and the mass attenuation coefficient all depend on

both the photon’s energy and its mass. The linear attenuation coefficient depends on mass density,

which can vary for a given element or compound. In contrast, the interaction cross-section and the

mass attenuation coefficient do not depend on mass density.

The interaction cross-section is a theoretical measure of interactions between a photon and an

object layer. This cross-section is a fictitious projection of a particle that can be better understood

by conceiving the particles like tiny spheres. No matter from which direction you approach the

spheres, you will always see the projections of the spheres as cross-sections through the centres of

the spheres. The collision between the photon and the electron can only be possible within the

56
electron’s fictitious cross-section. Each electron has a cross-section multiplied by the total number

of electrons of the atom to give the total cross-section of this atom.

The interaction cross-section data (also known as the atomic cross-section data) were calculated

and presented in units of cm2/atom or barns/atom (or b/atom, where 1 barn = 10–24 cm2) (Chantler,

2000). The connection between the interaction cross-section and the mass attenuation coefficient

is the conversion factor, A/NA, where A is the relative atomic mass of the target element and NA is

Avogadro’s number = 6.0221415 x 10–23 atoms mol-1 (Hubbell, 2006). As the connection between

the linear attenuation coefficient and the mass attenuation coefficient is a mass density factor, the

connections between the interaction cross-section and the linear attenuation coefficient are the

conversion and density factors.

The linear attenuation coefficient and mass attenuation coefficient were determined through an

experiment. In the experiment, measurements of beam intensity with and without an object layer

were used to determine the mass attenuation coefficient, which can be calculated using Equation

3.13 (Hubbell, 2006). The linear attenuation coefficient was then calculated by multiplying the

mass attenuation coefficient by the mass density using Equation 3.14. The fractional reduction of

beam intensity, -dI(l)/I, is proportional to the mass attenuation coefficient, µ/ρ, and the object

layer thickness, l (Equation 3.15). For a homogeneous object layer and mono-energetic beam, the

integration of Equation 3.15 reduces to the exponential attenuation law (Equation 3.16) for which

Equation 3.13 follows (Hubbell, 2006). Equation 3.16 calculates radiation attenuation.

𝐼0
𝜇/𝜌 = 𝑙 −1 𝑙𝑛 ( ) (3.13)
𝐼(𝑙)

57
𝐼0
𝜇 = 𝑙 −1 𝑙𝑛 ( )𝜌 (3.14)
𝐼(𝑙)

𝑑𝐼(𝑙) 𝜇
− = 𝑑𝑙 (3.15)
𝐼 𝜌

𝜇
𝐼(𝑡) = 𝐼0 𝑒𝑥𝑝(− 𝑙) (3.16)
𝜌

where µ/p is mass attenuation coefficient, usually in unit cm2 g-1, l is the matter
thickness of an object layer in unit of g cm-2, I0 is the intensity of the incident beam of
photons measured with the object layer removed from the beam, I(l) is the intensity of the
transmitted beam measured with the object layer interposed in the beam, µ is the linear
attenuation coefficient, and ρ is the density of the object layer.

Both the mass attenuation coefficient and the linear attenuation coefficient determined in the

experiment are limited by photon energy range. For a wide photon energy range, a theoretical

interaction cross-section has been compiled and maintained by various institutes, organisations, or

government agencies; for example, the National Bureau of Standards in the USA (name changed

in 1988 to the National Institute of Standards and Technology, NIST), the National Standard

Reference Data System in the USA, or the International Atomic Energy Agency (IAEA).

Theoretical interaction cross-section data for photon energy from 1 eV to 100 GeV were compiled

with hybrids of experimental data and theoretically calculated data. These data had been verified

with experimental data or validated by experiment (Cullen et al., 1997).

The datasets of the mass attenuation coefficients and linear attenuation coefficients for Rayleigh

scattering, Compton scattering, and photoelectric absorption of elements, compounds, or mixtures

58
were retrieved from the XCOM data bank (Berger and Hubbell, 1987) which is currently

maintained by National Institute of Standards and Technology (2016).

3.3.1.2. Determining the polar angle, θ: differential cross-section

The differential cross-section is a fictitious cross-section that represents the probability of a photon

being scattered in a new direction. This new direction forms an angle (polar angle, θ) with this

photon’s direction before being scattered. The differential cross-section can be understood as the

ratio of the number of the detected scatter photons to the total number of photons in the beam.

In an experiment in which a detector is placed to detect the scatter photon at directions that form

an angle θ with the photon’s direction before the interaction (Figure 3.3), if the incident photon is

replaced by a mono-energetic beam, the detector will detect all photons scattered into these

directions. The number of detected photons over the total number of photons in the beam

determines the differential cross-section in these directions. As differential cross-sections are

measures of the angular distributions of scatter photons’ directions, they are essential for Monte

Carlo simulation, to sample scatter photons’ directions.

Differential cross-sections of free electrons were theoretically determined using Rayleigh

scattering and Compton scattering models. These theoretical models assumed that electrons are

free and not moving; however, electrons in matter are bound in atoms and are moving with

momentum. The effect of bound electrons (also known as the binding effect) on scatter photons’

directions was accounted for using the atomic form factor for Rayleigh scattering or the incoherent

scattering function for Compton scattering (Hubbell et al., 1975).

59
The probability of the angular distribution of scatter photons’ directions was calculated using

Equation 3.17 for Rayleigh scattering (Hubbell et al., 1975) or Equation 3.18 for Compton

scattering (Heitler, 1954).

1+𝑐𝑜𝑠 2 𝜃
𝑃(𝜃) = [𝐹(𝑞, 𝑍)]2 (3.17)
2

𝐸 2 𝐸 𝐸
𝑃(𝜃) ≅ ( 𝐶 ) ( 𝐶 + − 𝑠𝑖𝑛2 𝜃) 𝑆(𝑞, 𝑍) (3.18)
𝐸 𝐸 𝐸 𝐶

𝐸
𝐸𝐶 = (3.19)
1+𝑘(1−𝑐𝑜𝑠𝜃)

where P(θ) is the probability of scattering at angle θ, E is the photon energy, F(q, Z)
and S(q, Z) are the atomic Form factor and incoherent scattering function, respectively, of
which values can be retrieved from NIST data bank, q is the momentum transfer variable
related to E and θ, and Z is the atomic (charge) number of the nucleus of the target atom,
𝐸
EC is the scatter photon energy, 𝑘 = , me is the electron mass, and c is a constant, the
𝑚𝑒 𝐶 2
speed of light in vacuum.

The atomic form factor only accounts for the effect of the bound electron which is undergoing the

interaction. It does not account for the effect of the other bound electrons of compounds on the

scatter photon (Morin, 1982, Poletti et al., 2002, Tartari et al., 2002, Poludniowski et al., 2009a,

Cunha et al., 2010). Investigations showed that this effect on the distribution of scatter radiation is

not negligible (Persliden and Carlsson, 1997, Cardoso et al., 2003, Poludniowski et al., 2009a,

Cardoso et al., 2009, Cunha et al., 2010) and it is referred to as the interference effect of compounds.

60
The interference effect can be accounted for using an interference function (Poludniowski et al.,

2009a). When the inference effect of compounds was accounted for, the probability of angular

deflection in Rayleigh scattering was calculated using Equation 3.20.

1+𝑐𝑜𝑠 2 𝜃
𝑃(𝜃) = [𝐹(𝑞, 𝑍)]2 𝐺(𝑞, 𝑍) (3.20)
2

where P(θ) is the probability of scattering at angle θ, F(q, Z) and G(q, Z) are the atomic
form factor and interference function, respectively, q is the momentum transfer variable
related to E and θ, Z is the atomic (charge) number of the nucleus of the target atom, the
values of F(q, Z) can be retrieved from NIST data bank, and the values of G(q, Z) of some
materials can be retrieved from Tartari et al. (2002), Royle and Speller (1995), or Evans et
al. (1991).

3.3.2. Change in a photon’s status and energy

A photon’s initial status was set to ‘primary’ when photons were generated at the focal spot.

Whenever a photon underwent an interaction, its status was changed to ‘scatter’.

A change in a photon’s energy was only made for Compton scattering. In Rayleigh scattering, a

photon was scattered without changing its energy. In photoelectric absorption, a photon was

absorbed in the matter. In Compton scattering, the energy change depended on the direction of the

scatter photon. The scatter photon energy was computed using the Compton (1923) equation

(Equation 3.19), which assumed that the electron was free and at rest (Bushberg et al., 2011).

Electrons in matter are orbiting their atoms which bind them. These electrons are moving with

momentum. The momentum of an orbiting electron, when this electron is undergoing an

interaction with a photon, can affect the scatter photon’s energy which is determined using the

Compton equation (Equation 3.19). The change in scatter photon energy ascribed to an electron’s

61
momentum is known as the Doppler broadening effect – the greater the Doppler broadening effect,

the lesser the photon energy lost. The Doppler broadening effect only happens in Compton

scattering.

In the above discussions, random numbers were essential and were extensively used in photon

transportation. The generation of random numbers and number sampling methods are discussed in

the following section.

3.3.3. Random numbers: random number generators and sampling methods

In Monte Carlo simulation, many decisions for events to occur depend on random numbers and

criteria (such as bigger than, equal to, or smaller than predefined values). The frequency of an

event’s occurrence is often dependent on the frequency of a range of number being sampled.

Sampling numbers for the occurrences of events is a vital component of Monte Carlo simulation.

3.3.3.1. Random numbers

Random numbers are numbers sampled/generated in a random manner. In computing, random

numbers are often generated using a deterministic algorithm, which means these random numbers

are generated in a predictable sequence. A deterministic algorithm that generates random numbers

can be written as a software function, which is commonly referred to as a random number generator.

When a random number generator is executed, numbers are generated in a sequence. For example,

when the nth time a random number generator is executed, the nth number in the sequence is

returned. When the last number in the sequence is returned, the next time the generator is executed,

the generator will return the 1st number in the sequence. Therefore, this sequence is periodic.

62
A random number generator constructed using Equation 3.21 produces a sequence with a period

of the order of 109 (Press and Teukolsky, 1992). This generator is like a random number generator

used in Matlab versions older than 5 (The MathWorks Inc., Natick, Massachusetts). In Matlab

version 7.4 and later, a new uniform random number generator uses an algorithm known as the

Mersenne Twister (M. Matsumoto and Nishimura, 1997). This generator has an incredible period

of length 219937 − 1. This generator was used in the Monte Carlo simulation in this project.

𝑥𝑛 𝑎𝑥𝑛−1 𝑚𝑜𝑑(231 −1)


𝑅𝑛 = , 𝑥𝑛 = (3.21)
231 −1 231 −1

where Rn is the nth random number in the sequence, a is a constant, and


a𝑥𝑛−1 mod(231 − 1) is the least nonnegative remainder from dividing a𝑥𝑛−1 by 231 − 1.

Random numbers generated using a deterministic algorithm are known as pseudo-random numbers

because they are not random but have excellent random properties. The advantage of using pseudo-

random numbers is that they are efficiently generated with tens of gigabit per second rates, which

is only limited by computing speed and resources (Press and Teukolsky, 1992).

True random numbers can be generated, such as using a hardware that generates random numbers

from a physical process. Hardware random number generators generally generate a limited number

of random numbers per second. Some fast hardware random number generators can generate

random numbers with multi-gigabit per second rates (Marandi et al., 2012).

3.3.3.2. Sampling methods

Sampling methods determine the frequency of a range of numbers being sampled. The most

common sampling method is uniform sampling, in which any number has an equal probability of

63
being sampled. In Monte Carlo simulation, many probabilities are not uniformly distributed.

Different distributions can be sampled more efficiently using different sampling methods.

3.3.3.2.1. Exponential sampling for free path length

In this project, the probability of a photon travelling a distance without any interaction with matter

was calculated using Equation 3.22, which was deduced from Equation 3.16. The probability

function (P(l)) (Equation 3.22) has a constant probability for any constant path length (l) in a

uniform media where µ/p is a constant. This probability function represents a uniform distribution.

As a free path length (l) is associated with the cumulative probability of interactions from path

length zero to l, the uniformly distributed probability function (P(l)) cannot be correctly sampled

by uniformly sampling the path length, which has a one-to-one relationship with the cumulative

probability of P(l). Equation 3.22 was rewritten as Equation 3.23, in which the random number ξ,

was uniformly distributed in the interval (0, 1). The free path length then can be sampled by

uniformly sampling ξ as given in Equation 3.23.

𝜇
𝜉 = 𝑃(𝑙) = 𝑒𝑥𝑝(− 𝑙) (3.22)
𝜌

1
𝑙= − 𝜇 𝑙𝑛(1 − 𝜉) (3.23)

𝜌

where ξ or P(l) is the probability, µ/p is mass attenuation coefficient, and l is the free
path length.

64
3.3.3.2.2. Discrete indexed sampling

A discrete indexed sampling method uniformly samples the indices of a list/table, and each time

the sampled index is used to retrieve the value at this index. Any probability distribution (either

discrete or continuous) of random events can be converted into a list or table, in which each of

these events is assigned with a range of indices. A sequence of these random events is then

represented by a random sequence of this list’s indices.

An arbitrary non-uniform probability distribution function, F(x), is illustrated in Figure 3.4(a),

where x is a random variable. F(x) is the frequency of x. Despite x and F(x) being continuous or

discrete, they can be divided into equal discrete intervals. The more the intervals, the better the

representation of this frequency; in other words, the more the intervals, the higher the accuracy.

Figure 3.4(b) illustrates the area under the curve divided into small rectangular areas that are

distinguished by grid lines. On the x-axis, a discrete interval of x determines a column. The first

column, x1, starts at the smallest x. The last column, xmax, ends at the largest x. On the y-axis, a

discrete interval of F(x) determines a row. The first row starts at zero; the last row ends at F(x).

The total number of the small rectangular areas under the curve is the sampling resolution; the

higher the resolution, the better the representation of F(x). A count of these areas can be established

starting at the first column from top to bottom, then at the second column from top to bottom, and

so on, until reaching the last row in the last column. A list of the probability F(x) is shown in Table

3.1.

65
Table 3.1. Indices and variable values

Index 1 2 3 4 5 6 7 8 9 10 … Countmax

Random variable x1 x2 x3 x3 x3 x4 x4 x4 x4 x4 … xmax

6 6 (b)
5 (a) 5

4 4

F(x)
F(x)

3 3
2 2
1 1
0 0
x
x

Figure 3.4. Indexed discrete sampling. (a) Plot of the non-uniform probability function F(x); (b)
Discrete areas in small rectangles under the curve.

3.3.3.2.2. Reject-accept sampling

A reject-accept sampling method uniformly samples a random number and tests the number under

a condition. If the condition is satisfied, the number is accepted; otherwise, the number is rejected.

If the number is rejected, this process continues until a random number is accepted.

Many non-uniform probability distributions can be sampled using reject-accept sampling. In many

situations, reject-accept sampling can be replaced by discrete indexed sampling and appropriate

results can still be achieved. The disadvantage of reject-accept sampling is that this sampling may

be very inefficient. The disadvantage of the discrete indexed sampling method can be that a very

long list may be needed for an appropriate representation of a probability distribution.

66
3.4. Implementation of interactions between photons and matter

The interactions between photons and matter were implemented using Rayleigh scattering,

Compton scattering, photoelectrical absorption, and pair production (in nuclear field and in

electron field) models. A pair production can only be possible if photon energy is at least twice the

energy of an electron mass; that is, approximately 512 × 2 kilo-electron-volt (keV). In x-ray

imaging, photons have energy approximately 10–150 keV. Therefore, the relevant interactions

between these photons and matter are Rayleigh scattering, Compton scattering, and photoelectrical

absorption (Hubbell, 2006).

Different types of interactions occur competitively. Their probabilities of occurrence can be

calculated as the ratios of the cross-section of each interaction type to the total interaction cross-

section. The simulation of these interactions is discussed in the following sections.

3.4.1. Rayleigh scattering

In Rayleigh scattering, a photon interacts with the electrons of an atom as a whole, and after the

interaction the photon changes its direction and does not lose its energy. Bushberg et al. (2011)

suggested that the photon interacts with and excites the whole atom through the energy from the

photon’s electric field, which causes all electrons of the atom to oscillate in phase. The oscillation

of these electrons immediately radiates a photon (scatter photon) of the same energy in a slightly

different direction. According to Hendee and Ritenour (2002), a photon is assumed to interact with

the atom’s electrons as a whole, and the photon is deflected or scattered with a negligible loss of

energy. Generally, Rayleigh scattering is implemented as such that after the interaction the photon

changes its direction without losing energy (Chan and Doi, 1982, Boone and Cooper III, 2000,

Nikolopoulos et al., 2007, Cunha et al., 2010, Khodajou-Chokami and Sohrabpour, 2015).

67
The direction of a scatter photon was determined using two angles: a polar angle θ and an azimuth

angle ω (Figure 3.3). The polar angle was determined per probability distribution function given

in Equation 3.17. Previous discussion (Section 3.3) had shown that this distribution does not

account for the interference effect of compounds. When the interference effect was accounted for,

the polar angle was determined per probability distribution function in Equation 3.20. The azimuth

angle was uniformly sampled in (0, 2π] with angle units in radians. The polar angle, the azimuth

angle, and the photon direction cosines before the interaction determined the direction of the scatter

photon. The scatter photon direction was then calculated using the equations given in Section 3.3.1.

3.4.2. Compton scattering

In Compton scattering, a photon interacts with a bound electron and causes the electron to be

emitted from its atom. The photon loses some energy and is scattered to become a scatter photon.

The scatter photon’s direction was determined using two random angels: a polar angle θ and an

azimuth angle ω (Figure 3.3). The polar angle was sampled from a probability distribution function

given in Equation 3.18. Like in Rayleigh scattering, the azimuth angle was uniformly sampled in

(0, 2π] with angle units in radian. The polar angle, the azimuth angle, and the photon direction

cosines before the interaction determined the direction of the scatter photon. This direction was

calculated using the equations given in Section 3.3.1.

The energy of the scatter photon was affected by not only the photon scattering direction but also

by the electron momentum before the interaction. The effect of the electron’s momentum on the

scatter photon energy is known as the Doppler broadening effect. Section 3.3.2 discussed that the

Doppler broadening effect may cause the energy of the scatter photon at a direction to change and

68
that for a fixed scattering angle, the energy of the scatter photon of a mono-energetic beam may

have a spectrum instead of a constant energy.

In x-ray imaging simulation, the Doppler broadening effect has been found to have negligible

effects on scatter radiation (Cunha et al., 2010) and can be accounted for by using an appropriate

distribution of energy broadening/changing. This project did not account for the Doppler

broadening effect. The scatter photon energy was computed using Equation 3.19.

Emitted electrons in tissues have low kinetic energy, less than a few keV, because of photons’

energy (<150 keV) as well as the low atomic number materials of tissues. An electron only travels

in soft tissues for a very short distance (<1 mm). Any secondary electrons or photons excited or

emitted by this electron also only survived a very short distance (<1 mm). The tracking of the

emitted electrons was neglected, and their kinetic energy was locally deposited at the interaction

site (Chan and Doi, 1982, Boone and Cooper III, 2000, Nikolopoulos et al., 2007, Cunha et al.,

2010, Khodajou-Chokami and Sohrabpour, 2015).

3.4.3. Photoelectric absorption

Photoelectric absorption is also known as the photoelectric effect. In photoelectric absorption, a

photon interacts with a bound electron. The photon loses all of its energy and causes the electron

to be emitted with kinetic energy equal to the photon energy minus the electron’s binding energy.

Photoelectric absorption only occurs if the photon energy is greater than the electron binding

energy. The tracking of emitted electrons was neglected and their kinetic energy was locally

deposited at the interaction site (Chan and Doi, 1982, Boone and Cooper III, 2000, Nikolopoulos

et al., 2007, Cunha et al., 2010, Khodajou-Chokami and Sohrabpour, 2015).

69
3.5. Radiation dosimetry in the Monte Carlo simulation

Radiation dosimetry is the measure of radiation. Measuring the amount of radiation is another

essential component of Monte Carlo simulation and is made in derived units from seven basic

units. The seven basic units are provided by the International System of Units (SI) for science and

technology and are metre for length, kilogram for mass, second for time, ampere for electric current,

kelvin for temperature, candela for luminous intensity and mole for substance. The detailed

derivations of radiation measurements are referred to in the International Commission on

Radiation Units and Measurements (2011). Several units that are commonly used in medical x-ray

imaging are discussed in the following sections.

3.5.1. Photon fluence, photon flux, and energy fluence

Photon fluence is the number of photons or particles passing through a unit cross-section area,

typically expressed in units of cm-2. The usual symbol Φ is given to photon fluence and calculated

using Equation 3.24 (International Commission on Radiation Units and Measurements, 2011).

Photon flux is the rate of photon fluence and is calculated using Equation 3.25. Photon flux is

useful for measuring photons over long periods (tens of seconds), such as in fluoroscopy. The unit

for photon flux is usually cm-2 s-1 (International Commission on Radiation Units and

Measurements, 2011).

Energy fluence is the amount of energy passing through a unit cross-section area. For a mono-

energetic beam of photons, energy fluence (Ψ) is the product of photon fluence (Φ) and the energy

per photon (E). The typical unit for energy fluence is J m-2. Energy fluence of a mono-energetic

beam is calculated using Equation 3.26 (International Commission on Radiation Units and

Measurements, 2011).

70
𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑝ℎ𝑜𝑡𝑜𝑛𝑠
𝛷= (3.24)
𝑎𝑟𝑒𝑎

𝛷
𝐹𝑙𝑢𝑥 = (3.25)
𝑡𝑖𝑚𝑒

𝛹 = 𝛷𝐸 (3.26)

where Φ is photon flux, Ψ is the energy fluence, and E is the photon energy.

3.5.2. Mass energy transfer coefficient

In an interaction between a photon and a charged particle, the photon may lose a fraction of its

energy to this particle as the particle’s kinetical energy plus, if any, its binding energy. The mass

energy transfer coefficient is the total mass attenuation coefficient multiplied by this fraction. The

mass energy transfer coefficient is given as μtr/ρ (International Commission on Radiation Units

and Measurements, 2011).

3.5.3. Mass energy absorption coefficient

The mass energy absorption coefficient accounts for the fraction of the mass energy transfer

coefficient that gives rise to the total energy locally deposited at the interaction site (International

Commission on Radiation Units and Measurements, 2011). After a charged particle gains kinetic

energy, its kinetic energy is deposited in matter through the subsequent excitation or emission of

other charged particles. Eventually, the kinetic energy of all charged particles is absorbed in the

matter. However, in interactions in which scatter photons are produced, the scatter photon energy

may not be deposited in the matter. Therefore, the mass energy absorption coefficient is always

less than the mass energy transfer coefficient.

71
With high atomic number materials, such as lead, several tens of keV of electrons emitted in

photoelectric absorption may produce bremsstrahlung radiation. Bremsstrahlung radiation is the

radiation emitted by a charged particle and is ascribed to this particle’s acceleration/deceleration

caused by an electric field of another charged particle (Bushberg et al., 2011). Some of the higher

energy (several tens keV) bremsstrahlung radiation may escape the interaction site. Therefore, the

mass energy absorption coefficient is generally smaller than the mass energy transfer coefficient

(International Commission on Radiation Units and Measurements, 2011).

Approximately 96% of the human body’s matter is composed of low atomic number elements

(oxygen 65.0%, carbon 18.5%, hydrogen 9.5%, and nitrogen 3.2%) (Tortora and Derrickson,

2008). In x-ray imaging, photons interact with body matter mainly through Compton scattering.

Electrons emitted by these photons in Compton scattering have low kinetic energy (less than 10

keV), and these electrons produce low energy bremsstrahlung radiation which excites or emits

other electrons with low kinetic energy. The kinetic energy of all these electrons is deposited within

less than one millimetre of the interaction site. For body tissues, the mass energy absorption

coefficient is approximately equal to the mass energy transfer coefficient (International

Commission on Radiation Units and Measurements, 2011).

3.5.4. KERMA (kinetic energy released per unit mass)

KERMA stands for kinetic energy released per unit mass and is the kinetic energy transferred to

charged particles by uncharged ionising radiation per unit mass. The SI unit for KERMA is J kg-1

with the special name of the Gray (Gy) or milliGray (mGy) (1 Gy = 1000 mGy = 1 J kg-1). KERMA

can be calculated using Equation 3.27 as the product of the mass energy transfer coefficient and

energy fluence (Kawrakow, 2000).

72
𝜇𝑡𝑟
𝐾 = 𝛹( ) (3.27)
𝜌

where K is KERMA, μtr/ρ is the mass energy transfer coefficient, and Ψ is energy
fluence.

When uncharged ionisation radiation, such as x-ray photons, passes through a matter, the

deposition of radiation energy in the matter happens in two steps. In the first step, energy of the

uncharged ionisation radiation is transferred to charged particles such as electrons. As in medical

x-ray imaging, photon energy is less than 150 keV, and the photon energy is transferred to electrons

through the relevant photoelectric absorption or Compton scattering. In the second step, the

charged particles deposit their energy in the matter by exciting and ionising other electrons.

Eventually, all the kinetic energy of electrons is deposited near the interaction site because of the

very short travelling distances (<1 mm) of these electrons and, if any, bremsstrahlung radiation

(less than a few keV) originating from acceleration/deceleration of these electrons (International

Commission on Radiation Units and Measurements, 2011).

3.5.5. Absorbed dose

The absorbed dose is the energy deposited through ionisation in unit mass of irradiated material.

The absorbed dose is calculated using Equation 3.28 and the SI unit is J kg-1 or Gy. The difference

between the absorbed dose and the KERMA is that KERMA is only defined for uncharged ionising

radiation, but the absorbed dose is defined for both charged and uncharged ionising radiation. A

historical unit, the rad, is often used for the absorbed dose and one rad is equal to 0.01 J kg-1 (1 rad

= 0.01 J kg-1 = 10 mGy) (International Commission on Radiation Units and Measurements, 2011).

73
𝐸
𝐷= (3.28)
𝑚

where D is the absorbed dose, E is the total energy departed, and m is the mass.

The absorbed dose is one of the basic dosimetric quantities. Different medical x-ray imaging

technologies may use specific dosimetric quantities. Computed tomography (CT) often uses the

CT dose index, the dose-length product, or multiple scan average dose. Fluoroscopy uses the entry

skin dose or KERMA area product (KAP). Mammography uses the mean/average glandular dose

(International Commission on Radiation Units and Measurements, 2011).

3.5.6. Radiation exposure

Radiation exposure is the amount of electrical charge that is produced by ionising radiation per

unit mass of air. Radiation exposure is calculated using Equation 3.29. The historical unit of

radiation exposure is the Roentgen, and the SI unit for radiation exposure is coulombs per kg (1

Roentgen = 2.58 x 10–4 C kg-1) (International Commission on Radiation Units and Measurements,

2011).

𝑄
𝑋= (3.29)
𝑚

where X is the radiation exposure, Q is the total amount of charges of one sign, and m
is the mass.

In medical diagnostic imaging, parameters given to an x-ray generator to generate photons are

often in milliampere second (mAs) and kilovoltage peak (kVp). The output intensity of an x-ray

generator can be measured and expressed as a radiation exposure per unit mAs under specific

74
operating conditions. An example of radiation exposure is 10 mR/mAs at 70 kVp at a source-

image distance of 100 cm with a 2 mm aluminium equivalent filtration (International Commission

on Radiation Units and Measurements, 2011).

Radiation exposure can be directly measured with air-filled radiation detectors (air-chamber

dosimeters). The effective atomic numbers of air and soft tissue are approximately the same.

Radiation exposure is nearly proportional to the absorbed dose in soft tissue over the range of

photon energies used in diagnostic radiology (International Commission on Radiation Units and

Measurements, 2011).

Radiation exposure can be calculated from an average energy deposition per ion pair in air, which

is approximately constant at 33.85 eV/ion pair or 33.85 J/C. The relationship between the absorbed

dose, radiation exposure, and air KERMA is given by Equation 3.30. For example, one Roentgen

of radiation exposure results in approximately 8.73 mGy absorbed dose in air (International

Commission on Radiation Units and Measurements, 2011).

𝐷 𝐸
𝑊= = (3.30)
𝑋 𝑄

where W is the average energy deposition per ion pair in air, D is the absorbed dose, X
is the radiation exposure, E is the total energy deposition in matter, and Q is the total amount
of charges of one sign.

3.6. Virtual phantom for the Monte Carlo simulation

In simulation, the details of phantoms are constructed as dummy anatomies that facilitate the study

of imaging systems, imaging procedures, or patient radiation dose. Phantoms can be constructed

75
using polynomial boundaries of materials or using Hounsfield units of volume data from CT.

Polynomial boundaries are best for uniform phantoms or phantoms with several types of uniform

materials. Hounsfield units of volume data from CT is best for heterogeneous phantoms mixed

with many materials. The heterogeneous phantom constructed in Badal and Badano (2009) is an

example of Hounsfield units of volume data from CT.

Summary

This chapter discussed details of the simulation of photon transportation in x-ray imaging. These

details were used to develop our Monte Carlo simulation code system. A simulation procedure for

photon transportation was developed, which followed a photon’s track from the x-ray focal spot

to the image receptor. Details of photon transportation were then examined, including: (1) changes

in a photon’s location, direction, and/or energy; (2) random number generation and sampling

methods; (3) relevant interactions between photons and matter; and (4) radiation dosimetry and

virtual phantom construction.

The methods for measuring the changes in a photon’s location, direction, and/or energy were

discussed and established. Photon location change was determined using a free path that was

determined using a random number and the linear attenuation coefficient. Photon direction change

was determined using the photon direction and two random angles, an azimuth angle uniformly

sampled in (0, 2π] and a polar angle sampled in (0, π] with all angle units in radian. The polar

angle depended on the atomic form factor for Rayleigh scattering or on the incoherent scattering

function for Compton scattering. In Rayleigh scattering, the interference effect of compounds on

the polar angle could be accounted for by using an interference function. Photon energy change

was calculated using the Compton equation. The Doppler effect on the scatter photons’ energy

76
broadening was ignored. This effect could be implemented using a probability distribution of the

energy broadening, however.

A random number generator and sampling methods were analysed for simulating photon

transportation. Uniform sampling was generally used in the simulation. An indexed sampling

method was introduced to sample non-uniformly distributed probability for improved sampling

efficiency. This indexed sampling method gave each probable event the proportional number of

indices to the event’s probability. When this method was executed, an index was uniformly

sampled from the indices. The sampled index was used to determine which event will happen. A

reject-accept sampling method was also discussed for sampling non-uniformly distributed

probability. In this method, a random number was uniformly sampled and tested under a condition.

If the condition was satisfied, the number was accepted; otherwise, the number was rejected. If the

number was rejected, this process continued until a random number was accepted

The relevant interactions (Rayleigh scattering, Compton scattering, and photoelectric absorption)

between photons and matter were implemented in our Monte Carlo simulation code system.

Rayleigh scattering was implemented with only change in photon direction. Compton scattering

was implemented with changes in photon direction and energy. Photoelectric absorption was

implemented as photon energy was completely deposited at the interaction site. In all the

interactions, the kinetic energy of emitted electrons was locally deposited in the matter. The

tracking of electrons was ignored.

In addition, radiation dosimetry was reviewed and its implementation was discussed. Two virtual

phantom construction methods were also discussed: one used the polynomial boundaries of

materials and the other used the Hounsfield units of volume data from CT.

77
The virtual phantom construction methods, however, were not suitable for constructing a grid for

the simulation of photon transportation in the grid materials. The next chapter will solve the

problem of photon transportation in grid materials.

78
Chapter 4: New method for determining radiation
transmission in grids

This chapter addresses photon transportation in grid materials. It reviews contemporary methods

for radiation transmission in grids and discusses these methods’ limitations. It then develops and

validates a new method, to overcome these limitations. The article related to the work of this

chapter is referred to Zhou et al. (2016).

4.1. Introduction: radiation transmission in grid designs

Basic grid performance parameters of transmission of primary, scatter, and total radiation are

discussed in Chapter 2. For physical grids, these transmissions are determined using the amounts

of radiation measured with or without grids (Bonenkamp and Hondius Boldingh, 1959a, Strid,

1976, Fetterly and Schueler, 2007, Fetterly and Schueler, 2009, Salvagnini et al., 2012).

A grid’s performance can be assessed through its design evaluation by ignoring manufacturing

defects, if any. Grid designs can be evaluated through Monte Carlo simulation, which are used to

determine grid designs’ radiation transmissions using analytical methods (Dance and Day, 1984,

Dance et al., 2000, Malusek et al., 2003, Ullman et al., 2006, Malusek et al., 2008, Star-Lack et

al., 2009, Lazos and Williamson, 2010, Sun and Star-Lack, 2010, Zbijewski et al., 2012, Sisniega

et al., 2013, Tomal et al., 2013, Chen et al., 2015).

Several analytic methods (Strid, 1976, Kalender, 1982, Bernstein et al., 1983, Day and Dance,

1983) have been developed to determine grid designs’ radiation transmissions. First, these methods

79
assumed that both primary radiation and scatter radiation were uniformly distributed at the grid

entry surface. Generally, primary radiation at the grid entry surface is approximately uniformly

distributed. However, it is well known that scatter radiation is non-uniformly distributed (Boone

et al., 2000, Johns and Yaffe, 1982). Second, these methods determined the radiation mean

transmission in each grid unit, which consists of a grid strip and an adjacent interspace region.

These methods, therefore, eliminate the details of grid lines resulted from stationary grids. Grid

lines in an x-ray image appear as white strips, which may interfere with the interpretation of the

image and make diagnosis more difficult.

Strid (1976) proposed an analytical method to calculate the transmission of scatter radiation, of

which angular distribution was determined using sample measurements in different directions.

These measurements were obtained using an ionisation chamber detector, which was equipped

with a long, slender, cylindrical collimator. The collimator was rotated to allow the scatter

radiation to be measured in different directions.

Strid (1976) assumed scatter radiation was distributed symmetrically about the centre of a parallel

grid and designated four exclusive regions for scatter radiation entering each grid unit. In the first

three regions, exact transmissions were determined. The transmission in the fourth region was

extrapolated by ignoring the radiation that would travel to at least another grid unit. For focused

grids, Strid (1976) determined an approximate transmission of scatter radiation by treating focused

strips as locally parallel.

Kalender (1982) determined radiation transmission by using a weighting factor and the total

attenuation of grid materials. The determination of the weighting factor depends on radiation path

lengths in the grid materials. Day and Dance (1983) suggested that this weighting factor could be

80
determined from the scatter-radiation-energy-angle spectrum, which can be either determined

through Monte Carlo simulation or in an experiment like that of Strid’s (1976).

Bernstein et al. (1983) extended the Strid (1976) method to obtain better approximation of

radiation transmissions in focused grids. Bernstein et al. (1983) designated three distinct regions

instead of the four distinct regions of Strid (1976). Furthermore, Bernstein et al. (1983) determined

the photon path lengths in focused grid material by increasing the photon’s polar angle by another

angle, which was equal to the angle of the strip relative to the grid normal (this grid normal is

parallel to the z-axis in Figure 4.1). However, Day and Dance (1983) found that Bernstein et al.

(1983) did not determine the correct radiation path lengths in focused grid materials.

Following these issues with the methods of Strid (1976), Kalender (1982), and Bernstein et al.

(1983), Day and Dance (1983) proposed a new method to determine the radiation transmissions in

oblique grids, which had their strips in parallel. Day and Dance (1983) also assumed that radiation

was uniformly distributed at the grid’s entry surface. The Day and Dance method precisely

determined the grid-unit mean transmissions of radiation in parallel grids and extrapolated

radiation mean transmissions in focused grids by replacing strip angles with new angles. These

new angles were calculated using two values: (1) the x-coordinate (Figure 4.1) of the radiation

entry points on the grid and (2) the grid focal distance. The Day and Dance method is used in the

Monte Carlo simulation to determine radiation transmissions in grid materials (Dance and Day,

1984, Dance et al., 2000, Malusek et al., 2003, Ullman et al., 2006, Malusek et al., 2008, Star-Lack

et al., 2009, Lazos and Williamson, 2010, Sun and Star-Lack, 2010, Zbijewski et al., 2012,

Sisniega et al., 2013, Tomal et al., 2013, Chen et al., 2015).

81
Figure 4.1. Strip alignments of focused grids showing the geometrical relationships of the strip
thickness (d), interspace (D), and strip height (h). A Cartesian system is defined by the x-, y-,
and z-axes perpendicular and parallel to the strip direction, and perpendicular to the plane of
the grid, respectively. The centre of the grid is located at C(0, 0, f0) where f0 is the grid focal
distance. The line segment AAexit illustrates a photon trajectory passing through the grid.

Three problems for all the above methods are: (1) inaccurate assumption of scatter radiation

distribution; (2) extrapolation of radiation transmissions by using grid-unit mean transmission,

which does not preserve the details of grid lines resulting from stationary grids; (3) extrapolation

of radiation transmissions for focused grids. In the following sections, a new method will be

developed to overcome these problems.

4.2. New method for determining radiation transmission in grid designs

A photon has a probability for travelling through grid materials without interacting with them. In

this case, the photon may have travelled in the grid’s strip material and/or its interspace material,

and the photon’s path lengths in each of these materials can be determined. This probability can

then be calculated using these lengths and these materials’ linear attenuation coefficients. In the

following sections, a new method was developed to determine this probability.

82
Contemporary methods assume that photons at the grid entry surface are uniformly distributed.

This new method is not constrained by the distribution of photons at the grid entry surface. In

addition, this and other methods ignore scatter radiation and fluorescent radiation arising from the

grid materials. because the amounts of those reaching the image receptor are negligible (Strid,

1976, Kalender, 1982, Bernstein et al., 1983, Day and Dance, 1983).

4.2.1 Determination of radiation transmission

Photon path lengths in grid materials were determined using a geometrical constraint. In a grid, a

strip surface is the interface surface between the strip and interspace materials (Figure 4.1). This

constraint was defined by all points on the same strip surface; that is, a constant x-coordinate for

parallel grids or a constant x-coordinate to z-coordinate ratio for focused grids.

For all types of grids, a three-dimensional Cartesian system was defined by X-, Y-, and Z-axes

perpendicular and parallel to the strip direction, and perpendicular to the plane of the grid,

respectively. The centre of grid entry plane was located at C(0, 0, f0). The f0 is the focal distance

for a focused grid or the distance from the focal spot to the grid entry plane of a parallel grid

(Figure 4.1). The x-ray focal spot was located at the origin O(0, 0, 0). When an arbitrary photon

entered the grid at point A(x, y, z) on the grid’s entry plane and moved in the direction given by

the direction cosines (Δx, Δy, Δz) without any interaction, this photon might intersect the grid’s

strip surfaces and exit the grid at point Aexit(xexit, yexit, zexit) on the grid’s exit plane (Figure 4.1).

Similar to the Day and Dance (1983) formula, the transmission of the photon along path AAexit was

given by Equation 4.1.

𝑇(𝑥, 𝑦, 𝑧, ∆𝑥, ∆𝑦, ∆𝑧) = 𝑒 −𝑙𝑠 𝑢𝑠 −𝑙𝑖 𝑢𝑖 (4.1)

83
where T is the transmission of a photon, ls and li are this photon’s path lengths in the
strip and interspace region, respectively, and us and ui are the linear attenuation coefficients
of the strip and interspace materials, respectively.

If this photon intersected any strip surfaces, the coordinate of the ith intersection point was

expressed as Mi(x+aiΔx, y+aiΔy, z+aiΔz), where ai was a variable. The value of the variable ai was

solved in the following sections. The coordinates of point A, point Aexit, and, if any, all Mi were

used to calculate ls and li. The linear attenuation coefficients (Section 3.5) can be retrieved from

the XCOM data bank (Berger and Hubbell, 1987) currently maintained by National Institute of

Standards and Technology (2016).

The spatial relationships between point A, point Aexit, grid strip, and grid interspace determined the

expression of the variable ai. These relationships were exhaustively classified by ten mutually

exclusive conditions (Table 4.1). An illustration of these relationships is shown in Figure 4.2, and

these conditions’ criteria are detailed in Table 4.2. The solution of variable ai is detailed in Section

4.2.2 (focused grids) and Section 4.2.3 (parallel grids).

84
Figure 4.2. Photon’s trajectories through the grid materials: C1, C2, C5, C6, and C7 illustrate
trajectories from an interspace region in the grid entry surface to a region in the grid exit surface;
C3, C4, C8, C9, and C10 illustrate trajectories from a strip region in the grid entry surface to a
region in the grid exit surface.

Table 4.1. Ten categories (Ci) of location relationships of points A(x, y, z) and A’(x’,
y’, z’). The conditions are restricted for x ≥ 0. For x ≤ 0, as it is mirror symmetric to x
≥ 0, results can be determined by applying a ‘minus’ sign to the x-coordinates.
C1 (x > x’) C2 (x > x’) C3 (x > x’) C4 (x > x’) C5 (x = x’)

A(x, y, z) i i s s i

A’(x’, y’, z’) s i s i i

C6 (x < x’) C7 (x < x’) C8 (x < x’) C9 (x < x’) C10 (x = x’)

A(x, y, z) i i s s s

A’(x’, y’, z’) s i s i s

where i is a location in an interspace; s is a location in a strip.

85
Table 4.2. The determination criteria of the ten categories (Ci) of the location
relationships of points A(x, y, z) and A’(x’, y’, z’). The conditions are restricted for x ≥
0. For x ≤ 0, as it is mirror symmetric to x ≥ 0, results can be determined by applying
a ‘minus’ sign to the x-coordinates.
Categories criteria of categories

C1 𝑟 < 𝐷, 𝑟′ ≥ 𝐷, and 𝑥 > 𝑥′


C2 𝑟 < 𝐷, 𝑟 ′ < 𝐷, and 𝑥 > 𝑥′
C3 𝑟 ≥ 𝐷, 𝑟′ ≥ 𝐷, and 𝑥 > 𝑥′
C4 𝑟 ≥ 𝐷, 𝑟 ′ < 𝐷, and 𝑥 > 𝑥′
C5 𝑟 < 𝐷 and 𝑥 = 𝑥′
C6 𝑟 < 𝐷, 𝑟′ ≥ 𝐷, and 𝑥 < 𝑥′
C7 𝑟 < 𝐷, 𝑟 ′ < 𝐷, and 𝑥 < 𝑥′
C8 𝑟 ≥ 𝐷, 𝑟′ ≥ 𝐷, and 𝑥 < 𝑥′
C9 𝑟 ≥ 𝐷, 𝑟 ′ < 𝐷, and 𝑥 < 𝑥′
C10 𝑟 ≥ 𝐷 and 𝑥 = 𝑥′

where r is the mantissa of (𝑥 + 𝐷/2)/(𝐷 + 𝑑); r’ is the mantissa of (𝑥′ + 𝐷/2)/(𝐷 + 𝑑).

4.2.2. Determination of variable ai in focused grids

In Figure 4.1, the photon entered the grid’s entry surface at A(x, y, z) and exited the grid at Aexit(xexit,

yexit, zexit). The straight line between the origin O(0, 0, 0) and Aexit(xexit, yexit, zexit) intersected the

grid’s entry plane at point A’(x’, y’, z’). Ten possible relationships between point A and point A’,

called categories (Ci, i = 1, 2,…, 10) (Table 4.2), were determined with two mantissas. A mantissa

is the positive fractional part of real number; for example, for x = 3.1415, the mantissa is 0.1415.
𝐷
𝑥+
One of the two mantissas was equal to the mantissa of 𝐷+𝑑, where x was the x-coordinate of point
2

𝐷
𝑥′+
A. The other mantissa was equal to the mantissa of , where 𝑥′ was the x-coordinate of point A’.
2
𝐷+𝑑

86
Points A and A’ might be located in different grid units. The difference between the different grid

units was determined in Equation 4.2.

𝐷 𝐷
𝑥+ 𝑥′+
𝑁 = |𝐼𝑁𝑇 (𝑑+𝐷2 ) − 𝐼𝑁𝑇 ( 𝑑+𝐷2 )| (4.2)

where N is the number of the different grid units, |a| is the absolute value of ‘a’, INT(a)
is the largest integer equal to or smaller than ‘a’, d is the strip thickness, D is the interspace
distance, x is the x-coordinate of point A(x, y, z), and 𝑥′ is the x-coordinate of point A’(x’,
y’, z’).

All points on the same strip surface had the ratios of their x-coordinates to their z-coordinates

equal; that is, x/z = constant = T, where T was the common ratio on a strip surface. Ti denoted the

common ratio of the strip surface when the photon on the ith intersected a strip surface. Only

photons with x-coordinates equal to or greater than zero (x ≥ 0) were discussed. Solutions for

photons when x < 0 were obtained by a mirror-symmetrical operation about the y-z plane. Photons

with a directional component equal to or smaller than zero (Δz ≤ 0) were excluded from the

discussion. These photons were not directly moving towards the image receptor. In this project,

these photons were neglected because the fraction of them reaching the image receptor is trivial.

In category C1, N was a whole number greater than zero. If N was equal to 1, only one intersection

point M1(x+a1*Δx, y+a1*Δy, z + a1*Δz) existed from point A to point Aexit. For category C1 these

results hold:

𝑥−𝑟
𝑇1 = (4.3)
𝑓0

𝐷
𝑥+
where r is the mantissa of 𝐷+𝑑. 2

87
𝑥+𝑎1 ×∆𝑥
𝑇1 = (4.4)
𝑧+𝑎1 ×∆𝑧

Equation (4.4) can be rearranged as

(𝑇1 × ∆𝑧 − ∆𝑥) × 𝑎1 = 𝑥 − 𝑇1 × 𝑧
(4.5)

Since A(x, y, z) is not in a strip, the following holds

𝑥
𝑇1 ≠ (4.6)
𝑧

Equations (4.5) and (4.6) suffice

𝑇1 × ∆𝑧 − ∆𝑥 ≠ 0 (4.7)

Equations (4.5) and (4.7) suffice

1 𝑥−𝑇 ×𝑧
𝑎1 = 𝑇 ×∆𝑧−∆𝑥 (4.8)
1

The straight lines AM1 and M1Aexit were the trajectories that the photon might travel in the

interspace material and strip material, respectively, where linear propagation of the photon within

the grid materials was assumed. From the definition of the cosine directions, the value of a1, always

greater than 0, was equal to the length of AM1. The value of a1 simplified the calculations of the

photon path lengths in the grid’s strips and interspaces.

If N was equal to 2, the first intersection M1(x+a1*Δx, y+a1*Δy, z + a1*Δz) was still valid. Two

more intersections points, M2(x+a2*Δx, y+a2*Δy, z + a2*Δz) and M3(x+a3*Δx, y+a3*Δy, z +

a3*Δz), existed on other strip surfaces. Let:

𝑥−𝑟−𝑑
𝑇2 = (4.9)
𝑓0

𝑥−𝑟−𝑑−𝐷
𝑇3 = (4.10)
𝑓0

88
𝐷
𝑥+
where r is the mantissa of 𝐷+𝑑. 2

Similar to a1, a2 and a3 were obtained as:

𝑥−𝑇 ×𝑧
2
𝑎2 = 𝑇 ×∆𝑧−∆𝑥 (4.11)
2

𝑥−𝑇 ×𝑧
3
𝑎3 = 𝑇 ×∆𝑧−∆𝑥 (4.12)
3

The values of a2 and a3 were equal to the lengths of AM2 and AM3, respectively.

For photon paths in category C1, N might be greater than 3 (indeed, without any interaction with

grid materials, N was limited to the number of grid strips). As N increased, the expressions of the

intersection points followed a general pattern. This pattern led to a general expression for photons

in category C1.

Similar to C1, general expressions for all the categories were obtained:


𝑥𝑒𝑥𝑖𝑡 = 𝑥 + ∆𝑥 × ∆𝑧 (4.13)


𝑦𝑒𝑥𝑖𝑡 = 𝑦 + ∆𝑦 × ∆𝑧 (4.14)


𝑧𝑒𝑥𝑖𝑡 = 𝑧 + ∆𝑧 × ∆𝑧 (4.15)

𝑓
𝑥 ′ = 𝑥𝑒𝑥𝑖𝑡 × ℎ+𝑓
0
(4.16)
0

𝑓
𝑦 ′ = 𝑦𝑒𝑥𝑖𝑡 × ℎ+𝑓
0
(4.17)
0

𝑓
𝑧 ′ = 𝑧𝑒𝑥𝑖𝑡 × ℎ+𝑓
0
(4.18)
0

𝑥𝑖 = 𝑥 + 𝑎𝑖 ∆𝑥 (4.19)

𝑦𝑖 = 𝑦 + 𝑎𝑖 ∆𝑦 (4.20)

𝑧𝑖 = 𝑧 + 𝑎𝑖 ∆𝑧 (4.21)
𝑥−𝑧×𝑇
𝑖
𝑎𝑖 = 𝑇 ×∆𝑧−∆𝑥 (4.22)
𝑖

89
𝐷
𝑥+
Given that r is the mantissa of 𝐷+𝑑: 2

For category C1: N > 0

𝑖 𝑖−1
𝑥−𝑟−𝑑×𝐼𝑁𝑇( )−𝐷×𝐼𝑁𝑇( )
𝑇𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 − 1 (4.23)
𝑓0

For category C2: if N = 0, the photons would only travel in interspace material. For N > 0,

𝑖 𝑖−1
𝑥−𝑟−𝑑×𝐼𝑁𝑇( )−𝐷×𝐼𝑁𝑇( )
𝑇𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 (4.24)
𝑓0

For category C3: if N = 0, the photons would only travel in strip material. For N > 0,

𝑖−1 𝑖
[ 𝑥−𝑟+𝐷 ]−𝑑×𝐼𝑁𝑇( )−𝐷×𝐼𝑁𝑇( )
𝑇𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 (4.25)
𝑓0

For category C4: N >= 0

𝑖 𝑖−1
[ 𝑥−r+𝐷]−𝐷×INT( )−𝑑×INT( )
𝑇𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 + 1 (4.26)
𝑓0

For category C5, the photon would only travel in interspace material.

For category C6: N >= 0

𝑖 𝑖−1
[ 𝑥−𝑟+𝐷]+𝑑×𝐼𝑁𝑇( )+𝐷×𝐼𝑁𝑇( )
𝑇𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 + 1 (4.27)
𝑓0

For category C7: if N = 0, the photon would only travel in interspace material. For N > 0,

𝑖 𝑖−1
[ 𝑥−𝑟+𝐷]+𝑑×𝐼𝑁𝑇( )+𝐷×𝐼𝑁𝑇( )
𝑇𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 (4.28)
𝑓0

For category C8: if N = 0, the photon would only travel in strip material. For N > 0,

𝑖 𝑖−1
[ 𝑥−𝑟+(𝑑+𝐷)]+𝐷×𝐼𝑁𝑇( )+𝑑×𝐼𝑁𝑇( )
𝑇𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 (4.29)
𝑓0

For category C9: N > 0

90
𝑖 𝑖−1
[ 𝑥−𝑟+(𝑑+𝐷)]+𝐷×𝐼𝑁𝑇( )+𝑑×𝐼𝑁𝑇( )
𝑇𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 − 1 (4.30)
𝑓0

For category C10, the photon would only travel in strip material.

4.2.3. Determination of variable ai in parallel grids

Parallel grids were made of parallel strips where points on the same strip surface had their x-

coordinates equal in the coordinate system in Figure 4.1. A constant x-coordinate was used to

determine intersections on each grid strip surface. The relationships given in Tables 4.1 and 4.2

were valid for parallel grids, except that point A’ was replaced by point Aexit (Figure 4.1). The

general expressions, which are different from that for focused grids, were given for Δx ≠ 0 (for Δx

= 0, see categories C5 and C10):

𝑥 ′ = 𝑥𝑒𝑥𝑖𝑡 (4.31)

𝑦 ′ = 𝑦𝑒𝑥𝑖𝑡 (4.32)

𝑧 ′ = 𝑧𝑒𝑥𝑖𝑡 (4.33)

For category C1: N >0

𝑖 𝑖−1
𝑟+𝑑×𝐼𝑁𝑇( )+𝐷×𝐼𝑁𝑇( )
𝑎𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 − 1 (4.34)
|∆𝑥|

For category C2: if N = 0, the photon would only travel in interspace material. For N > 0,

𝑖 𝑖−1
𝑟+𝑑×𝐼𝑁𝑇( )+𝐷×𝐼𝑁𝑇( )
𝑎𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 (4.35)
|∆𝑥|

For category C3: If N = 0, the photon would only travel in strip material. For N > 0,

𝑖−1 𝑖
( 𝑟−𝐷)+𝑑×𝐼𝑁𝑇( )+𝐷×𝐼𝑁𝑇( )
2 2
𝑎𝑖 = 𝑖 ∈ 1, 2, 3, … , 2𝑁
|∆𝑥|
(4.36)

For category C4: N >= 0

91
𝑖 𝑖−1
( 𝑟−𝐷)+𝐷×𝐼𝑁𝑇( )+𝑑×𝐼𝑁𝑇( )
2 2
𝑎𝑖 = 𝑖 ∈ 1, 2, 3, … , 2𝑁 + 1
|∆𝑥|
(4.37)

For category C5, the photon would only travel in interspace material.

For category C6: N>=0

𝑖 𝑖−1
( 𝐷−𝑟)+𝑑×𝐼𝑁𝑇( )+𝐷×𝐼𝑁𝑇( )
2 2
𝑎𝑖 = 𝑖 ∈ 1, 2, 3, … , 2𝑁 + 1
∆𝑥
(4.38)

For category C7: if N = 0, the photon would only travel in interspace material. For N > 0,

𝑖 𝑖−1
( 𝐷−𝑟)+𝑑×𝐼𝑁𝑇( )+𝐷×𝐼𝑁𝑇( )
2 2
𝑎𝑖 = 𝑖 ∈ 1, 2, 3, … , 2𝑁
∆𝑥
(4.39)

For category C8: if N = 0, the photon would only travel in strip material. For N > 0,

𝑖 𝑖−1
( 𝐷+𝑑−𝑟)+𝐷×𝐼𝑁𝑇( )+𝑑×𝐼𝑁𝑇( )
𝑎𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 (4.40)
∆𝑥

For category C9: N > 0

𝑖 𝑖−1
( 𝐷+𝑑−𝑟)+𝐷×𝐼𝑁𝑇( )+𝑑×𝐼𝑁𝑇( )
𝑎𝑖 = 2 2
𝑖 ∈ 1, 2, 3, … , 2𝑁 − 1 (4.41)
∆𝑥

For category C10, the photon would only travel in strip material.

4.3. Validation of the new method

The validation of this new method was performed by comparing grid performances determined

using this new method and current methods. Parallel and focused grids were selected from the

scholarly literature, and their performance was also retrieved for comparison. Theoretical

comparisons of radiation transmissions were made only for parallel grids and used mono-energetic

92
radiation. For the focused grids, Monte Carlo simulations were performed by keeping the

simulation setups the same as the experimental setups in the literature.

4.3.1. Criteria for comparisons of grid performance and grid selection

The transmissions of primary (Tp) and scatter (Ts) radiation were used as criteria for grid

performance comparisons, indicating the relative primary radiation and scatter radiation reaching

the image receptor with grid to without grid, respectively. Other grid performance factors, as

discussed in Chapter 2, were combination effects of Tp, Ts, and the scatter-to-primary ratio (SPR)

without grid. Using Tp and Ts for grid performance comparisons had the advantage of avoiding

the combination effects of Tp and Ts.

Focused grids were selected from the literature in which the grid geometric details, Tp, and Ts

were readily retrievable. A parallel grid was selected for theoretical comparisons of radiation

transmission between this new method and the current methods.

4.3.2. Grid details, phantoms, radiation beam quality, and simulation setups

In the simulation, the virtual mammographic grid was constructed using details of the

mammographic grid that was experimentally evaluated in Carton et al. (2009). This is a moving

grid manufactured by Smit Rontgen (Eindhoven, The Netherlands) and contained in a breast

support. The virtual general grid was constructed using the general grid that was experimentally

evaluated in Fetterly and Schueler (2009) and made with lead strips and carbon fibre interspace

materials. For theoretical comparisons, the design of a parallel grid was selected from Dick and

Motz (1978). This grid design was used by Bernstein et al. (1983) for theoretic comparison. The

details of all these grid designs are given in Table 4.3.

93
Table 4.3. Details of anti-scatter grids
Carton et al. Fetterly and Schueler Dick and Motz
(2009) grid (2009) grid (1978) grid
Strip thickness d (cm) 0.0020 0.0069 0.0076
Strip material Lead Lead Lead
Interspace distance D (cm) 0.0300 0.0330 0.0174
Interspace material Fibre Fibre Aluminium
Strip height h (cm) 0.1500 0.4950 0.2090
Grid total cover thickness
0.0400 0.0600 0
(cm)
Cover material Fibre Fibre Nil
Focal distance f0 (cm) 65 100
Strip frequency (cm-1) 31 25 40
Grid ratio 5:1 15:1 12:1

The simulations used the same radiation beam qualities as in the experiments in the literature. Rh-

target beams of 35 peak kilovoltage (kVp), 40 kVp, 45 kVp, and 49 kVp with a total 0.27 mm

copper filtration were generated in simulations for the mammographic grid. These beams were

calculated using the model of Boone et al. (1997). They had higher average photon energy than a

common mammographic x-ray beam because they were specially selected for contrast-enhanced

mammographic subtraction studies in Carton et al. (2009).

The RQR-8 x-ray beam (International Electrotechnical Commission, 2005) was used in

simulations for the general grid evaluation and was calculated using the model of Poludniowski et

al. (2009b).

4.3.3. Simulation of grid movement

When a grid moves during image acquisition, the grid reciprocates about its central line. The

simulation of a grid’s movement was performed in a way in that the grid’s x-coordinate (Figure
94
4.2) changed as a function of time while its y- and z-coordinates remained unchanged. The grid’s

maximum displacement from its central line was set at ±1 cm, the frequency of the grid’s

reciprocation was set at 300 per minute, and the radiation exposure duration was set at 50 ms. The

change of grid x-coordinate as a function of time was given by Equation 4.43.

∆𝑥𝑔𝑟𝑖𝑑 = (−1)𝑖𝑛𝑡(𝑅𝑃𝑀×𝑡) × 𝐿 × 𝑀𝑎𝑛𝑡𝑖𝑠𝑎(𝑅𝑃𝑀 × 𝑡) (4.43)

where Δxgrid is the change of grid x-coordinate, L is the maximum displacement from
grid central line, t is the time at which the photon reaches the grid entry plane, RPM is the
frequency of grid reciprocation, int(A) is the largest integer equal to or smaller than ‘A’,
and Mantissa(A) is the mantissa of ‘A’.

All focused grids were placed at their focal distances from the x-ray focal spot. The parallel grid

was placed at 100 cm from the x-ray focal spot. All grids were placed as close as possible to the

phantoms’ exit-planes.

Water phantoms with a 30 cm × 30 cm cross-sectional area and a range of thicknesses were used

for the general grid. Mammographic phantoms comprising 50% ICRU-44 adipose tissue and 50%

ICRU-44 breast tissue (ICRU, 1989) were used for the mammographic grid. These mammographic

phantoms had a 22 cm × 11.5 cm cross-sectional area and a range of thicknesses (2–8 cm in 2-cm

increments).

An array of detectors (40 cm × 40 cm for the general grid and 24 cm × 30 cm for the

mammographic grid) was placed behind the grid with a 1 cm space to allow the placement of a

moving grid. These detectors had a pixel pitch of 37.5 µm and had no dead space between adjacent

detectors. They were constructed as perfect detectors that convert all of a photons’ energy to signal.

95
4.3.4. Simulation of photon transport
Simulations of photon transport were performed using the dedicated Monte Carlo code system

developed using the details discussed in Chapter 3. The simulations were performed in Matlab Ver.

R2015b (The MathWorks Inc., Natick, Massachusetts) on a computer using an Intel® Core™ i7–

5600U CPU, 2 cores of 2.6 GHz processors and 16 Gigabyte (GB) random access memory.

Each simulation was repeated ten times to obtain the means and standard deviations of Tp, Ts, and

Tt. In each simulation, the number of primary photons generated at the x-ray focal spot increased

as the phantom thickness increased. In the general grid’s evaluation, approximately 100 million

primary photons were generated for the 10-cm thick phantom. The number of primary photons

was increased to approximately 100 billion for the 50-cm thick phantom.

4.3.5. Results

4.3.5.1. Theoretical comparison: transmission in parallel grid

The theoretical transmissions in the parallel grid are shown in Figure 4.3. They were determined

using a mono-energetic radiation beam uniformly distributed across the x-coordinate and are

functions of photon incident angles.

The radiation mean transmissions determined using this new method has excellent agreement with

the radiation transmissions determined using the methods of Bernstein et al. (1983) and Day and

Dance (1983). Bernstein et al. (1983) found a sign error in one of the equations of Strid (1976).

After correcting this error, we found that the radiation transmission of Strid (1976) is different

from the other methods at some regions. The difference starts at the angle of which the photon

travelled through one grid unit and has a period of one grid unit. Bernstein et al. (1983) suggested

that this difference was caused by the fourth region identified in the Strid (1976) method. However,

96
we found this difference is caused by ignoring the part of the radiation that was in the fourth region

and passed partially through the strip.

The maximum different transmission from the Strid (1976) method is approximately 14% for the

70-keV photons (Figure 4.3(b)) and more than 1000% for the 30-keV photons (Figure 4.3(a)). As

the absolute difference (less than 1.2%) in the radiation transmission was small, this difference

might have no effect on the outcome of the simulation.

Figure 4.3. Radiation transmission as a function of photon incident angle. (a) Transmission of
30-keV mono-energetic photons. (b) Transmission of 70-keV mono-energetic photons.

Figure 4.4 shows a magnified region of the radiation transmission of this new method and the Day

and Dance method. The radiation transmission of this new method is a square-wave form with a

minimum value of approximately 0.004, a maximum value of approximately 0.876, and a mean

value of approximately 0.323. The radiation transmission of the Day and Dance method is a

horizontal straight line, with a constant value of approximately 0.323.

97
Figure 4.4. Transmission of 70-keV mono-energetic beam in a parallel grid in a magnified
region. The entry angle of the beam is 3 degrees. The beam is uniformly distributed across
the grid strips.

The transmissions of non-uniform poly-energetic radiation in the parallel grid are about the same

in this new method and the Day and Dance method (Table 4.4). Tp is approximately 0.585 (SD <=

0.001) and Ts is approximately 0.082 (SD < 0.001).

A parallel grid can be considered symmetrical about any of its strips, assuming the grid has an

infinite size. Whether a parallel grid moves or not, there is always the same relationship between

the strips and the central ray of the primary radiation beam, and there is always grid cut-off in

parallel grids. Despite the parallel grid’s operation (moving or stationary), its radiation

transmissions were the same and therefore agreed with the theoretical transmissions.

Table 4.4. Radiation transmission in parallel grid

Tp Ts
Mean SD Mean SD
The new method (stationary grid) 0.585 0.001 0.082 < 0.001

The new method (moving grid) 0.585 0.001 0.082 < 0.001

Day and Dance method 0.585 < 0.001 0.082 < 0.001

98
4.3.5.2. Transmission of primary radiation in focused grid: Tp

Table 4.5 shows the results of the mammographic grid’s Tp for the 35-kVp beam. Regardless of

the grid’s operation (moving or stationary) and the methods, Tp increased marginally as the

phantom thickness increased. When this grid moved, Tp increased negligibly from 0.841 to 0.843

(SD < 0.001) in the new method. When the grid was stationary, Tp increased negligibly from 0.874

to 0.875 (SD ≤ 0.001) in the new method and from 0.874 to 0.875 (SD < 0.001) in the Day and

Dance method.

Table 4.6 shows the results of the general grid’s Tp. Regardless of the grid’s operation (moving or

stationary) and the methods, Tp increased marginally as the phantom thickness increased. When

this grid moved, Tp increased negligibly from 0.672 to 0.679 (SD ≤ 0.002) in the new method.

When the grid was stationary, Tp increased negligibly from 0.719 to 0.727 (SD ≤ 0.003) in the

new method and from 0.721 to 0.729 (SD ≤ 0.002) in the Day and Dance method.

The difference in Tp between the moving grid and the stationary grid was approximately 4% for

the mammographic grid (Table 4.5) or 7% for the general grid (Table 4.6). This reduction in Tp is

ascribed to grid’s defocusing that occurred during the grid’s movement. This reduction is

significant and is due to grid cut-off. Grid cut-off is undesirable absorption of primary radiation

by grid materials. The difference between the grid cut-offs in the general grid and in the

mammographic grid was primarily due to their different grid ratios.

99
Table 4.5. Transmission of primary radiation (Tp) in the mammographic grid

Tp

Phantom thickness The new method


Day and Dance Method
(cm) Stationary grid Moving grid
2 0.874 (SD < 0.001) 0.841 (SD < 0.001) 0.874 (SD < 0.001)

4 0.874 (SD < 0.001) 0.841 (SD < 0.001) 0.874 (SD < 0.001)

6 0.875 (SD < 0.001) 0.842 (SD < 0.001) 0.875 (SD < 0.001)

8 0.876 (SD 0.001) 0.843 (SD < 0.001) 0.875 (SD < 0.001)
Results of Tp were obtained with the 35-kVp beam.

Table 4.6. Transmission of primary radiation (Tp) in the general grid

Tp

Phantom thickness The new method


Day and Dance Method
(cm)
Stationary grid Moving grid
10 0.719 (SD 0.003) 0.672 (SD 0.001) 0.721 (SD ≤ 0.001)

20 0.721 (SD 0.002) 0.673 (SD 0.001) 0.723 (SD ≤ 0.001)

30 0.723 (SD 0.002) 0.676 (SD 0.001) 0.725 (SD 0.001)

40 0.727 (SD 0.002) 0.678 (SD 0.001) 0.728 (SD 0.001)

50 0.727 (SD 0.002) 0.679 (SD 0.002) 0.729 (SD 0.002)

4.3.5.3. Transmission of scatter radiation in focused grid: Ts

Table 4.7 shows the Ts of the mammographic grid with the 35-kVp beam. Regardless of whether

the grid was moving or stationary, the Ts in this new method is approximately the same as the Ts

100
in the Day and Dance method. As the phantom thickness increased from 2 cm to 8 cm, Ts increased

from 0.180 to 0.197 (SD ≤ 0.001).

Table 4.8 shows the Ts of the mammographic grid with the 49-kVp beam. The Ts in this new

method was approximately the same for the moving and stationary grids. As the phantom thickness

increased from 2 cm to 8 cm, the Ts in this new method increased from 0.224 to 0.248 (SD ≤ 0.001)

and Ts in the Day and Dance method increased from 0.219 to 0.233 (SD < 0.001). The difference

in Ts between the new method and the Day and Dance method is significant at two standard

deviations.

Table 4.7. Transmission of scatter radiation (Ts) with 35-kVp beam in the
mammographic grid
Ts
Phantom thickness The new method
Day and Dance Method
(cm) Stationary grid Moving grid
2 0.181 (SD 0.001) 0.180 (SD 0.001) 0.180 (SD 0.001)

4 0.186 (SD 0.001) 0.187 (SD < 0.001) 0.187 (SD < 0.001)

6 0.192 (SD 0.001) 0.192 (SD 0.001) 0.193 (SD 0.001)

8 0.197 (SD 0.001) 0.197 (SD < 0.001) 0.197 (SD < 0.001)

101
Table 4.8. Transmission of scatter radiation (Ts) with 49-kVp beam in the
mammographic grid.

Ts
Phantom thickness The new method
Day and Dance Method
(cm)
Stationary grid Moving grid
2 0.224 (SD 0.001) 0.224 (SD 0.001) 0.219 (SD < 0.001)

4 0.233 (SD 0.001) 0.233 (SD < 0.001) 0.220 (SD < 0.001)

6 0.241 (SD 0.001) 0.241 (SD 0.001) 0.227 (SD < 0.001)

8 0.248 (SD 0.001) 0.248 (SD < 0.001) 0.233 (SD < 0.001)

Table 4.9 shows the general grid’s Ts. Whether the grid was moving or stationary, Ts was

approximately the same for the new method and the Day and Dance method. As the phantom

thickness increased from 20 cm to 50 cm, Ts increased marginally from 0.063 to 0.066 (SD ≤

0.004). Ts, however, was approximately 0.065 (SD ≤ 0.002) for the 10-cm phantom in both

methods.

Table 4.9. Transmission of scatter radiation (Ts) in the general grid

Ts

Phantom thickness The new method


Day and Dance Method
(cm) Stationary grid Moving grid
10 0.065 (SD 0.002) 0.065 (SD 0.002) 0.065 (SD < 0.001)
20 0.063 (SD 0.004) 0.063 (SD 0.003) 0.063 (SD < 0.001)
30 0.064 (SD < 0.001) 0.064 (SD < 0.001) 0.064 (SD 0.001)
40 0.065 (SD 0.001) 0.065 (SD 0.001) 0.065 (SD 0.001)
50 0.066 (SD 0.001) 0.066 (SD 0.001) 0.066 (SD 0.001)

102
4.3.5.4. Images behind a stationary grid and moving grid

Figure 4.5 shows small portions of the 20-cm thick phantom’s images. The images in Figures 4.5(a)

and 4.5(b) were produced using this new method with the grid stationary and moving, respectively.

The image in Figure 4.5(c) was produced using the Day and Dance method with the grid stationary.

Figure 4.5(a) shows the image with grid lines. Figure 4.5(b) and 4.5(c) show the images without

grid lines.

When grids operated when stationary, this new method preserved the grid lines which were

apparent in the image (Figure 4.5). The Day and Dance method, however, eliminated these grid

lines because this method only calculated grid unit mean transmissions, which are not the radiation

transmissions of stationary grids. In previous grid investigations through Monte Carlo simulation

(Malusek et al., 2003, Ullman et al., 2006, Star-Lack et al., 2009, Lazos and Williamson, 2010,

Sun and Star-Lack, 2010, Cunha et al., 2010, Sisniega et al., 2013, Tomal et al., 2013, Chen et al.,

2015, Chen, 2016), the effect of grid lines and the effect of grid cut-off has been neglected.

103
Figure 4.5. Small portions of the 20-cm thick phantom’s images. (a) and (b) Created using the
new method with the grid stationary and moving, respectively. (c) Created using the Day and
Dance method with the grid stationary.

4.3.5.5. Dependence of radiation transmission (Tp and Ts)

The mammographic grid is a moving grid, so was evaluated while it was moving. Figures 4.6

shows the Tp of this mammographic grid as a function of tube voltage. Despite the phantom

104
thickness and tube voltage, the mammographic grid’s Tp determined using the new method was

less than that determined through the Day and Dance method. At the same tube voltage, the Tp in

both the new method and the Day and Dance method increased marginally (difference less than

0.4%) as the phantom thickness increased from 2 cm to 8 cm. At the same phantom thickness, the

Tp in both the new method and the Day and Dance method also increased marginally (difference

less than 1%) as the tube voltage increased from 35 kVp to 49 kVp. Regardless of the phantom

thickness, as the tube voltage increased from 35 kVp to 49 kVp, the Tp determined through this

new method and through the Day and Dance method increased from approximately 0.841 to 0.851

(SD < 0.001) and 0.874 to 0.883 (SD < 0.001), respectively. The difference in the Tp between the

new method and the Day and Dance method was due to that the grid cut-off effect that was

accounted for in the new method.

0.89
Tp
Tp - 2cm (this work)
Tp - 4cm (this work)
Tp - 6cm (this work)
Tp - 8cm (this work)
Tp - 2cm (Day & Dance method)
Tp - 4cm (Day & Dance method)

0.86
Tp - 6cm (Day & Dance method)
Tp - 8cm (Day & Dance method)

0.83
34 39 44 49
Tube voltage (kVp)

Figure 4.6. Mammographic grid’s Tp and Ts as a function of tube voltage.

105
For both methods, the mammographic grid’s Ts increased as the phantom thickness and/or tube

voltage increased (Figure 4.7). At the same tube voltage, as the phantom thickness increased from

2 cm to 8 cm, the Ts determined increased approximately 11% in the new method and

approximately 10% in the Day and Dance method. At the same phantom thickness, as the tube

voltage increased from 35 kVp to 49 kVp, the Ts increased approximately 26% in the new method

and approximately 20% in the Day and Dance method. Regardless of the phantom thickness, as

the tube voltage increased from 35 kVp to 49 kVp, the Ts determined through the new method and

the Day and Dance method increased from approximately 0.180 to 0.248 (SD ≤ 0.001) and 0.180

to 0.233 (SD ≤ 0.001), respectively.

0.26
Ts Ts - 2cm (this work)

Ts - 4cm (this work)

Ts - 6cm (this work)


0.23
Ts - 8cm (this work)

Ts - 2cm (Day & Dance method)

Ts - 4cm (Day & Dance method)


0.20
Ts - 6cm (Day & Dance method)

Ts - 8cm (Day & Dance method)

0.17
34 39 44 49
Tube voltage (kVp)

Figure 4.7. Mammographic grid’s Tp and Ts as a function of tube voltage.

The Day and Dance method underestimated the mammographic grid’s Ts for the higher tube kVp

(40, 45, and 49 kVp). At the 35 kVp tube voltage, the Ts in the new method was approximately

the same as the Ts in the Day and Dance method. From 40 kVp to 49 kVp, the Ts in the new

106
method was significantly different from the Ts in the Day and Dance method at two standard

deviations. This significant difference was attributed to two factors: (1) the Day and Dance method

extrapolated focused grids’ strip angles and (2) the Day and Dance method assumed that scatter

radiation was uniformly distributed.

4.3.6. Comparisons of Tp and Ts between simulation and the literature

4.3.6.1. Tp comparison

In general, the mammographic grid’s Tp increased as the tube voltage increased in both the

simulation and in Carton et al. (2009) (Figures 4.8 and 4.9). Regardless of the method, the

mammographic grid’s Tp obtained from the simulation increased marginally as the phantom

thickness and/or tube voltage increased (Figure 4.8). In Carton et al. (2009), the Tp increased

significantly as the tube voltage increased or the Tp varied marginally as the phantom thickness

increased (Figure 4.9).

The general grid’s Tp obtained through the simulation increased marginally (less than 1%) as the

phantom thickness increased from 10 cm to 30 cm (Figure 4.10). The general grid’s Tp was

approximately constant in Fetterly and Schueler (2009) (Figure 4.10). In the simulation, the Tp

increased as the phantom thickness increased because the beam-hardening effect increased as the

phantom thickness increased.

When these grids were operated when moving, the Tp determined through the new method was a

better approximation of Carton et al. (2009) and Fetterly and Schueler (2009) than the Tp

determined through the Day and Dance method (Figures 4.9 and 4.10). This was because the effect

of grid cut-off due to grid movement was accounted for in the new method. When the general grid

107
was operated when stationary, the Tp was approximately the same for the new method and the

Day and Dance method (Figure 4.10).

0.89
Tp
Tp - 2cm (this work)

Tp - 4cm (this work)

Tp - 6cm (this work)


0.86
Tp - 8cm (this work)

Tp - 2cm (Day & Dance method)

Tp - 4cm (Day & Dance method)

0.83 Tp - 6cm (Day & Dance method)


34 39 44 49
Tp - 8cm (Day & Dance method)
Tube voltage (kVp)

Figure 4.8. Mammographic grid’s Tp and Ts as a function of tube voltage.

0.86
Tp Tp - 2cm (this work)

Tp - 4cm (this work)

Tp - 6cm (this work)

0.83 Tp - 8cm (this work)

Tp - 2cm (Carton et al., 2009)

Tp - 4cm (Carton et al., 2009)

Tp -6cm (Carton et al., 2009)


0.80
34 39 44 49 Tp - 8cm (Carton et al., 2009)
Tube voltage (kVp)

Figure 4.9. Mammographic grid’s Tp and Ts as a function of tube voltage.

108
Tp Tp - Fetterly & Schueler, 2009

Tp - Stationary, This new method


0.70

Tp - Moving, This new method

Tp - Stationary, Day & Dance Method

0.60
10 20 30 40 50

Phantom thickness (cm)

Figure 4.10. General grid’s Tp as a function of phantom thickness.

The difference in the mammographic grid’s Tp between the new method and Carton et al. (2009)

decreased from approximately 4% to 1% as the tube voltage increased from 35 kVp to 49 kVp

(Figure 4.9). The difference in the mammographic grid’s Tp between the Day and Dance method

and Carton et al. (2009) decreased from approximately 8% to 4% as the tube voltage increased

from 35 kVp to 49 kVp. The difference in the general grid’s Tp between the new method with the

grid moving and Fetterly and Schueler (2009) was approximately 5.3% (SD = 0.004) (Figure 4.10).

The difference in Tp between the new method with the grid stationary and Fetterly and Schueler

(2009) was approximately 11.5% (SD = 0.004) (Figure 4.10).

4.3.6.1. Ts comparison

The mammographic grid’s Ts increased in both the simulation and Carton et al. (2009) as the tube

voltage and/or phantom thickness increased (Figures 4.11 and 4.12). The general grid’s Ts

remained approximately the same in the simulation and Fetterly and Schueler (2009) as the

phantom thickness increased (Figure 4.13). This trend of the mammographic grid’s Ts increasing

with increasing tube voltages agrees with Carton et al. (2009) and Salvagnini et al. (2012).

109
0.26
Ts Ts - 2cm (this work)
Ts - 4cm (this work)
0.23 Ts - 6cm (this work)
Ts - 8cm (this work)
Ts - 2cm (Day & Dance method)
0.20
Ts - 4cm (Day & Dance method)
Ts - 6cm (Day & Dance method)
0.17 Ts - 8cm (Day & Dance method)
34 39 44 49
Tube voltage (kVp)

Figure 4.11. Mammographic grid’s Tp and Ts as a function of tube voltage.

0.25 Ts Ts - 2cm (this work)


Ts - 4cm (this work)
0.22 Ts - 6cm (this work)
Ts - 8cm (this work)
0.19
Ts - 2cm (Carton et al., 2009)
0.16 Ts - 4cm (Carton et al., 2009)
Ts - 6cm (Carton et al., 2009)
0.13 Ts - 8cm (Carton et al., 2009)
34 39 44 49
Tube voltage (kVp)

Figure 4.12. Mammographic grid’s Ts as a function of tube voltage.

Ts Ts - Fetterly & Schueler, 2009

0.08 Ts - Stationary, This new method

Ts - Moving, This new method


0.04
Ts - Stationary, Day & Dance Method

0.00
10 20 30 40 50
Phantom thickness (cm)

Figure 4.13. General grid’s Ts as a function of phantom thickness.

110
The difference in the mammographic grid’s Ts between the simulation and the literature decreased

as the tube voltage and/or phantom thickness increased (Figure 4.12). The mean difference in the

mammographic grid’s Ts between this new method and Carton et al. (2009) decreased from

approximately 21% to 13% (SD ≤ 0.066) as the tube voltage and phantom thickness increased

from 35 kVp to 49 kVp and 2 cm to 8 cm, respectively. The mean difference in the mammographic

grid’s Ts between the Day and Dance method and Carton et al. (2009) decreased from

approximately 19% to 10% (SD ≤ 0.085) as the tube voltage and phantom thickness increased

from 35 kVp to 49 kVp and 2 cm to 8 cm, respectively. In Carton et al. (2009), the standard

deviation of the Ts at 40 kVp was greater than 0.03. At two standard deviations, the difference in

the mammographic grid’s Ts between the simulation and Carton et al. (2009) was not significant.

Regardless of the methods or the grid’s operation (moving or stationary), the difference in the

general grid’s Ts between the simulation and Fetterly and Schueler (2009) was approximately 16%

(Figure 4.13).

The difference in Tp or Ts between the simulation and the literature may be ascribed to several

causes, such as grids’ manufacture defects, and/or difficulty in the experimental setup. The exact

causes, however, cannot be determined through either the simulation or the experiment.

Summary

This chapter presented the development and validation of a new method to determine radiation

transmission in grids. This new method overcame the limitations of current methods as set out in

the scholarly literature. These limitations included: (1) assuming scatter radiation is uniformly

distributed in the grid entry surface; (2) extrapolating radiation transmission from grid-unit mean

111
transmission, which does not preserve the details of grid lines resulting from stationary grids; and

(3) extrapolating radiation transmission to focused grids.

The new method was developed using photon path lengths in the grid’s strip and interspace

materials. These lengths were determined using a geometrical constraint, which is a constant x-

coordinate for parallel grids, or a constant x-coordinate to z-coordinate ratio for focused grids. The

new method used these lengths and these materials’ linear attenuation coefficients to determine

the radiation transmission in the grid. This and the other methods all ignored the scatter radiation

and fluorescent radiation arising from the grid materials.

Validation of this new method was performed by comparing grid performance determined through

the simulation using the new method and contemporary methods, and that retrieved from the

literature. Most emphasis was made on the differences between the new method and the Day and

Dance method, which is the most appropriate contemporary method for determining radiation

transmission in focused grids.

The results of the parallel grid showed excellent agreement between the mean transmission in the

new method and the transmissions of other methods in the scholarly literature, except Strid’s (1976)

method. Strid’s method designated four mutually exclusive regions to determine the radiation

transmission in grids. Bernstein et al. (1983) suggested that Strid’s method wrongly designated the

fourth region. However, we found that Strid’s method did neglect transmission of radiation located

in the fourth region and only passed partially through the grid strip.

The comparisons of radiation transmissions in the focused grids were made only between the

literature, the new method, and the Day and Dance method. The results of the focused grids showed

112
that Tp was affected by grid operation. When the grids were operated when stationary, the Tp

determined using the new method was approximately the same as that obtained through the Day

and Dance method and was higher than that in the literature. When the grids were operated when

moving, the Tp determined using the new method was approximately 4% for the mammographic

grid and 7% for the general grid, which is lower than that obtained through the Day and Dance

method and higher than that in the literature. The difference in Tp between this new method and

the Day and Dance method was due to the grid-cut off effect of the grids’ movement. The Day and

Dance method did not account for the grid-cut off effect on Tp.

Ts was not affected by the grid’s operation. The general grid’s Ts determined using the new

method was approximately the same as that obtained using the Day and Dance method and was

higher than that in the literature. The mammographic grid’s Ts determined using the new method

was higher than that obtained using the Day and Dance method and was higher than that in the

literature. The Day and Dance method, however, underestimated the mammographic grid’s Ts.

The discrepancy of the mammographic grid’s Ts between the new method and the Day and Dance

method increased as the tube voltage increased.

When the grids were operated when stationary, the Day and Dance method did not preserve the

grid lines because their method only extrapolated radiation transmission from the grid-unit mean

transmission, which is not the radiation transmission of stationary grids. The new method

preserved these grid lines which were shown as white strips in the image.

The discrepancy in Tp and Ts between the simulation and the literature may be ascribed to several

causes, such as grids’ manufacturing defects, the difference between the simulation setup and the

113
experimental setup, and/or difficulty in the experimental setup. The exact causes, however, cannot

be determined in either the simulation or the experiment.

Chapter 2 to Chapter 4 presented the details for using our Monte Carlo simulation code system to

evaluate grid designs. Chapter 5 will show the validation of our Monte Carlo simulation code

system.

114
Chapter 5: Validation of the new Monte Carlo code
system for grid evaluation

This chapter presents the final part of the first phase of the project. It validates the Monte Carlo

simulation code system that was developed using the details presented in Chapter 3. The validation

is done through grid performance comparisons between simulation results and experiment results.

This validation was presented at the Engineering & Physical Science in Medicine 2016 conference.

5.1. Material and methods


5.1.1. Setups of experiment and simulation

The setups of radiation measurements were guided by the IEC60627 (International

Electrotechnical Commission, 2013) international standards, which are recommended for the

experimental evaluation of grids. The setups were made under three conditions: (1) a narrow beam

condition for primary radiation measurement, (2) a broad beam condition with a primary beam

blocker for scatter radiation measurement, and (3) a broad beam condition for total radiation

measurement. The details of these setups were discussed in Chapter 2. The measurements of the

amount of radiation obtained in these setups were used to calculate four quantities: The

transmissions of primary (Tp), scatter (Ts), total (Tt) radiation, and the scatter-to-primary ratio

(SPR) without grid. Tp, Ts and Tt were calculated using Equations 2.1 to 2.3 (see Chapter 2). SPR

was calculated using Equation 5.1.

𝑆𝑐𝑎𝑡𝑡𝑒𝑟_𝑅𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛𝑤𝑖𝑡ℎ𝑜𝑢𝑡 𝑔𝑟𝑖𝑑
𝑆𝑃𝑅 = (5.1)
𝑃𝑟𝑖𝑚𝑎𝑟𝑦_𝑅𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛𝑤𝑖𝑡ℎ𝑜𝑢𝑡 𝑔𝑟𝑖𝑑

115
In addition, a range of the primary beam’s field of views (FOVs) was simulated to investigate the

effect of the FOV on Tp. A range of primary beam blocker sizes was also simulated to investigate

the effect of blocker size on Ts. In the experiment, only one FOV of primary beam and one primary

beam blocker size were investigated.

In the simulation, logic labels were assigned to these primary photons which, without interaction,

would travel within predefined primary beam FOVs and/or within predefined primary beam

blocker areas. A logic label occupies one byte of random access memory (RAM). These labels

used only relatively negligible amounts of RAM during the simulation. By using photon labels,

the simulation setup was simplified to the broad beam condition, which was used for all

measurements.

The advantage of using labels for photons is that simulation time can be reduced by one-third

because one setup is sufficient to obtain measurements of total and scatter radiation, which

otherwise can only be obtained experimentally in two separate setups. Seven primary beam FOVs

and seven primary beam blocker sizes were investigated in the simulation.

5.1.2. Radiation detector

In the experiment, the amount of radiation was measured using a DRX Evolution Plus 3543 Image

Receptor – Gd2O2S:Tb Scintillators amorphous image receptor (Carestream Health, Inc.,

Rochester, NY, USA). The detailed information on the construction of this image receptor is

proprietary; as such, it could not be obtained. In the simulation, a Gd2O2S image receptor

(phosphor height 600 μm, phosphor area 50 µm × 50 µm, silicon structure thickness 40 µm) was

constructed for the approximation of the DRX Evolution Plus 3543 Image Receptor. In each

radiation measurement setup, the experiment and simulation were repeated three times to obtain

116
three sets of measurements. These repeated sets of measurements were used to calculate means

and standard deviations.

5.1.3. X-ray beams

To compare the experiment and the simulation, tube voltages were selected depending on phantom

thickness. These voltages were selected because x-ray imaging generally uses tube voltage

depending on the thickness of the imaging anatomy: the thicker the anatomy, the higher the tube

voltage. In the experiment and the simulation, voltages increased from 60 to 100 peak kilovoltage

(kVp) in 10 kVp increments as the phantom thickness increased from 10 cm to 30 cm in 5-cm

increments.

X-ray beam spectrums used in the simulation were calculated using the model of Poludniowski et

al. (2009b). The quality of radiation beams used in the experiment and the simulation was

measured (Table 5.1). The half-value layers of radiation beams used in the simulation were

approximately 0.1 mm lower than those used in the experiment, but this difference would have

had negligible effects on the simulation and experimental results. The beams’ mean energy used

in the simulation was measured, but not that used in the experiment.

Table 5.1. Details of x-ray beams


Half-value layer Aluminium
Mean energy (keV)
(mm)
kVp *Experiment Simulation Experiment Simulation
60 38.5 2.4 2.5
70 42.2 2.7 2.8
80 46.3 3.2 3.3
90 49.4 3.4 3.5
100 53.0 3.8 3.8
* Not available

117
5.1.4. Grid details

Grid details are listed in Table 5.2. The grid was received from Philips Medical Systems, Germany

and was made of lead strips, carbon fibre interspace, and aluminium covers. The paint materials

on the grid entry and exit surfaces have a thickness of approximately 1.44 mm. As recommended

by the grid’s manufacturer, in the simulation, these materials were weighted by increasing the strip

thickness by 3 µm and each cover by 450 µm.

Table 5.2. Details of anti-scatter grid


Strip thickness d (μm) 36
Strip material Lead
Interspace distance (μm) 238
Interspace material Fibre
Strip height h (mm) 3
Covers’ thickness (μm) 200
Cover material Aluminium
Focal distance f0 (cm) 110
Strip frequency (cm-1) 36
Grid ratio 12:1

5.1.5. Phantoms

Polymethyl methacrylate (PMMA) was used as the phantom material. Each piece of PMMA had

a 30 cm × 30 cm cross-section area and 1 cm thickness. The phantoms’ thicknesses were set from

10 cm to 30 cm in 5-cm increments. The phantoms’ designs were constructed in the simulation.

5.1.6. Simulation of photon transportation


The simulation of photon transportation was performed in Matlab Ver. R2015b (The MathWorks

Inc., Natick, Massachusetts) on a computer using an Intel® Core™ i7–5600U CPU, 2 cores of 2.6

118
GHz processors and 16 Gigabyte RAM. Radiation transmissions in the grid materials were

determined using the method developed and validated in Chapter 4.

The number of primary photons generated at the x-ray focal spot increased as the phantom

thickness increased. Approximately 100 million primary photons were generated for the 10-cm

thick phantom. The number of primary photons was increased to approximately 10 billion for the

30-cm thick phantom.

5.2. Results
5.2.1. Results obtained through simulation

As the radius of the primary beam’s FOV increased from 0.15 cm to 2 cm in the simulation, Tp

determined with the same kVp was approximately the same (Figure 5.1). As the kVp and phantom

thickness increases from 60 kVp and 10 cm to 100 kVp and 30 cm, respectively, Tp increased

from approximately 0.636 to 0.716. The statistic variation of Tp decreased as the radius of the

primary beam’s FOV increased because the larger the primary beam size resulted in more photons,

which the statistical variation is primarily dependent on.

119
Tp
60 kVp & 10 cm
0.74

70 kVp & 15 cm
0.72

0.70 80 kVp & 20 cm

0.68 90 kVp & 25 cm

0.66 100 kVp & 30 cm

0.64

0.62

0.60
0 0.5 1 1.5 2
Radius of primary beam's FOV at detector (cm)

Figure 5.1. Tp as a function of primary beam size. Tp remains approximately unchanged as


the radius of primary beam’s FOV increases. The error bar indicates two standard deviations.

In the simulation, Ts determined with the same kVp decreased as the radius of the primary beam’s

FOV increased from 0.15 cm to 2 cm (Figure 5.2). Ts decreased from approximately 0.048 to

0.033 for the 60 kVp and 10 cm phantom, 0.062 to 0.045 for the 70 kVp and 15 cm phantom, 0.071

to 0.060 for the 80 kVp and 20 cm phantom, 0.087 to 0.074 for the 90 kVp and 25 cm phantom,

and 0.097 to 0.090 for the 100 kVp and 30 cm phantom. The statistical variation of Ts decreased

as the radius of the primary beam’s FOV increased due to the same reason as that for the statistic

variation of Tp.

120
0.14
Ts 60 kVp & 10 cm

0.12 70 kVp & 15 cm

0.10 80 kVp & 20 cm

90 kVp & 25 cm
0.08
100 kVp & 30 cm
0.06

0.04

0.02

0.00
0 0.5 1 1.5 2
Radius of primary beam's FOV at detector (cm)

Figure 5.2. Ts as a function of primary beam size. Ts decreases as the radius of primary
beam’s FOV increases. The error bar indicates two standard deviations.

In the simulation, Tt determined with the same kVp was approximately the same as the radius of

the primary beam’s FOV increased from 0.15 cm to 2 cm (Figure 5.3). As the kVp and phantom

thickness increase from 60 kVp and 10 cm to 100 kVp and 30 cm, respectively, Tt decreased from

approximately 0.222 to 0.134. The statistic variation of Tt decreased as the radius of the primary

beam increased due to the same reason as that for the statistical variation of Tp.

121
Tt
60 kVp & 10 cm
0.24
70 kVp & 15 cm
0.22
80 kVp & 20 cm
0.20
90 kVp & 25 cm

0.18 100 kVp & 30 cm

0.16

0.14

0.12

0.10
0 0.5 1 1.5 2
Radius of primary beam's FOV at detector (cm)

Figure 5.3. Tt as a function of tube voltage and phantom thickness. Tt remains approximately
unchanged as the radius of primary beam’s FOV increases. The error bar indicates two
standard deviations.

In the simulation, as the radius of the primary beam increased from 0.15 cm to 2 cm (Figure 5.2),

SPR determined with the same kVp decreased from approximately 2.320 to 2.175 for the 60 kVp

and 10 cm phantom, 4.556 to 3.964 for the 70 kVp and 15 cm phantom, 7.129 to 6.297 for the 80

kVp and 20 cm phantom, 11.196 to 9.259 for the 90 kVp and 25 cm phantom, and 18.925 to 13.050

for the 100 kVp and 30 cm phantom. The statistical variation of SPR decreased as the radius of

the primary beam increased due to the same reason as for the statistical variation of Tp.

122
SPR
60 kVp & 10 cm
20

18 70 kVp & 15 cm

16
80 kVp & 20 cm
14

12 90 kVp & 25 cm

10
100 kVp & 30 cm
8

0
0 0.5 1 1.5 2
Radius of primary beam's FOV at detector (cm)

Figure 5.4. SPR as a function of tube voltage and phantom thickness. SPR decreases as the
radius of primary beam’s FOV increases. The error bar indicates two standard deviations.

5.2.1. Results determined in experiment

As the tube voltage and phantom thickness increased from 60 kVp and 10 cm to 100 kVp and 30

cm, respectively, Tp determined in the experiment increased from approximately 0.623 to 0.718

(Figure 5.5). Ts increased from approximately 0.047 to 0.088 (Figure 5.6), Tt decreased from

approximately 0.219 to 0.135 (Figure 5.7), and SPR increased from approximately 2.452 to 13.065

(Figure 5.8).

123
0.74 Tp

0.72

0.70

0.68

0.66

0.64

0.62

0.60
60 kVp & 10 cm 70 kVp & 15 cm 80 kVp & 20 cm 90 kVp & 25 cm 100 kVp & 30 cm

Tube voltage and phantom thickness

Figure 5.5. Tp determined in experiment as a function of tube voltage and phantom thickness.
Tp increases as the tube voltage and phantom thickness increase. The error bar indicates two
standard deviations.

0.10
Ts
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01
0.00
60 kVp & 10 cm 70 kVp & 15 cm 80 kVp & 20 cm 90 kVp & 25 cm 100 kVp & 30 cm

Tube voltage and phantom thickness

Figure 5.6. Ts determined in experiment as a function of tube voltage and phantom thickness.
Ts increases as the tube voltage and phantom thickness increase. The error bar indicates two
standard deviations.

124
Tt
0.24

0.22

0.20

0.18

0.16

0.14

0.12

0.10
60 kVp & 10 cm 70 kVp & 15 cm 80 kVp & 20 cm 90 kVp & 25 cm 100 kVp & 30 cm

Tube voltage and phantom thickness

Figure 5.6. Tt determined in experiment as a function of tube voltage and phantom thickness.
Tt decreases as the tube voltage and phantom thickness increase. The error bar indicates two
standard deviations.

14.00 SPR

12.00

10.00

8.00

6.00

4.00

2.00

0.00
60 kVp & 10 cm 70 kVp & 15 cm 80 kVp & 20 cm 90 kVp & 25 cm 100 kVp & 30 cm

Tube voltage and phantom thickness

Figure 5.8. SPR determined in experiment as a function of tube voltage and phantom thickness.
SPR increases as the tube voltage and phantom thickness increase. The error bar indicates
two standard deviations.

125
5.3. Discussion

As the tube voltage and phantom thickness increase, the trend of Tp increasing in the simulation

agrees with that in the experiment (Figure 5.9). This increasing trend is due to the increased

average photon energy because the higher the photon energy, the lower the radiation attenuation

in the grid materials. As the tube voltage and phantom increased from 60 kVp and 10 cm to 100

kVp and 30 cm, respectively, the difference in Tp between the simulation and the experiment

decreased from approximately 1.8% to 0.1%.

Tp experiment Simulation
0.74

0.72

0.70

0.68

0.66

0.64

0.62

0.60
60 kVp & 10 cm 70 kVp & 15 cm 80 kVp & 20 cm 90 kVp & 25 cm 100 kVp & 30 cm
Tube voltage and phantom thickness

Figure 5.9. Tp as a function of tube voltage and phantom thickness. Tp increases as the tube
voltage and phantom thickness increase. The error bar indicates two standard deviations.

As the tube voltage and phantom thickness increase, the trend of Ts increasing in the simulation

agrees with that in the experiment (Figure 5.10). This increasing trend is also due to the increased

126
average photon energy because the higher the photon energy, the lower the radiation attenuation

in the grid materials. As the tube voltage and phantom increased from 60 kVp and 10 cm to 102

kVp and 30 cm, respectively, the difference in Ts between the simulation and the experiment

decreased from approximately 20.0% to 1.1%. The difference in Ts between the simulation and

the experiment with the 60 kVp and 10 cm phantom and the 70 kVp and 15 cm phantom is

significant at two standard deviations. This difference may be caused by difficulty in the

experiment setup as shown in section 5.2.1 that the primary beam’s blocker size affects Ts (Figure

5.2).

0.10 experiment Simulation


Ts
0.09

0.08

0.07

0.06

0.05

0.04

0.03

0.02
60 kVp & 10 cm 70 kVp & 15 cm 80 kVp & 20 cm 90 kVp & 25 cm 100 kVp & 30 cm

Tube voltage and phantom thickness

Figure 5.10. Ts as a function of tube voltage and phantom thickness. Ts increases as the tube
voltage and phantom thickness increase. The error bar indicates two standard deviations.

Tt obtained through the simulation is approximately the same as that determined in the experiment

(Figure 5.11). As the tube voltage and phantom thickness increase, the trend of Tt decreasing in

127
the simulation agrees with that in the experiment. The difference in Tt between the simulation and

the experiment is, on average, approximately 1.2% (SD 0.012).

Tt experiment Simulation
0.24

0.22

0.20

0.18

0.16

0.14

0.12

0.10
60 kVp & 10 cm 70 kVp & 15 cm 80 kVp & 20 cm 90 kVp & 25 cm 100 kVp & 30 cm

Tube voltage and phantom thickness

Figure 5.11. Tt as a function of tube voltage and phantom thickness. Tt decreases as the tube
voltage and phantom thickness increase. The error bar indicates two standard deviations.

SPR is approximately the same in the simulation and the experiment (Figure 5.12). As the tube

voltage and phantom thickness increase from 60 kVp and 10 cm to 100 kVp and 30 cm, the SPR

increases from approximately 2.452 to 13.065 in the experiment and 2.278 to 12.592 in the

simulation. The difference in SPR between the simulation and the experiment is approximately,

on average, 3.6% (SD 0.025).

128
SPR experiment Simulation
14.00

12.00

10.00

8.00

6.00

4.00

2.00

0.00
60 kVp & 10 cm 70 kVp & 15 cm 80 kVp & 20 cm 90 kVp & 25 cm 100 kVp & 30 cm

Tube voltage and phantom thickness

Figure 5.12. SPR as a function of tube voltage and phantom thickness. SPR increases as the
tube voltage and phantom thickness increase. The error bar indicates two standard deviations

In general, the results obtained through the simulation are approximately the same as those

determined in the experiment. Some discrepancy in Ts between the simulation and the experiment

exists which, however, might be due to difficulty in the experiment setup.

In summary, our Monte Carlo simulation code system is appropriate for analysing and/or

evaluating the designs of grids which can be, or have been, manufactured with negligible defects.

Summary

This chapter presented the validation of our Monte Carlo simulation code system for grid

evaluation. This validation was made by comparing grid performance determined through the

simulation and the experiment. The grid performance was evaluated using the beam-block method,

which was reviewed in Chapter 2. The simulation was performed using approximately the same

conditions as that in the experiment. In addition, the simulation also investigated additional

129
conditions: a range of primary beam sizes for the measurement of primary radiation and a range

of primary beam blocker sizes for the measurement of scatter radiation.

Comparing grid performances between the simulation and the experiment showed that the Tp, Ts,

Tt, and SPR determined in our Monte Carlo simulation code system were approximately the same

as those determined in the experiment. The investigation of a range of primary beam sizes in the

simulation showed that the Tp was not affected by the primary beam’s FOV. The investigation of

a range of primary beam blocker size showed that the Ts decreased as the radius of the primary

beam blocker increased. This is because the more the primary beam was blocked, the less forward

scatter radiation that arose from the phantom’s matter underneath the blocker.

In this chapter, our Monte Carlo simulation code system was proven that it is appropriate for

analysing and/or evaluating the designs of grids that can be, or have been, manufactured with

negligible defects. In the next chapter, a new method to design new grids was developed to

overcome the reduction of primary radiation in grid materials.

130
Chapter 6: Criteria for designing new grids

This chapter describes the second phase of the project. It gives an overview of current grid designs

and then discusses factors and issues related to grid design. The final part of this chapter evaluates

the criteria for designing new grids.

6.1. Overview of grid design theory

Two grid design theories are reported in the scholarly literature, each of them focusing on different

radiation transmissions of grids. One theory (Strid, 1976) directs grid design towards the maximum

reduction of scatter radiation by optimising the grid strip thickness. In the previous chapter, the

transmission of scatter radiation (Ts) was found that it depends on tube voltage. This means an

‘optimal’ strip thickness at a certain tube voltage may no longer be ‘optimal’ when the tube voltage

changes.

The other theory (Bonenkamp and Hondius Boldingh, 1959b) presents a mechanism to maximise

the transmission of primary radiation (Tp) for a given grid lead content, which is measured as mass

per unit area (commonly in units g/cm2). In this theory, Tp is increased by reducing the strip

thickness and keeping the grid-unit length unchanged (a grid-unit length is the sum of strip

thickness and an adjacent interspace distance). As the grid lead content is constant, reducing the

strip thickness increases the strip height. Consequently, the grid ratio is increased. In this theory,

for a given lead content, the best grid would be the one constructed with the largest grid ratio.

131
These theories focus on optimising Ts or maximising Tp, but not both. In the following sections

of this chapter, a new grid design method was investigated to optimise Ts and maximise Tp

simultaneously.

6.2. Grid design factors

Grid design factors are these parameters that affect Tp and/or Ts. There are two types of factors.

One type relates to grid construction, including strip thickness, strip height, strip material,

interspace material, interspace distance, grid ratio, grid cover material, and grid cover thickness.

The other type relates to grid application, including x-ray beam quality, scatter radiation

distribution, and grid application limits. The relationship between these factors is illustrated in

Figure 6.1: the factor at each arrow’s end influences the factor at the arrow’s head.

Figure 6.1. Relationships between grid design factors for the transmission of scatter radiation
and primary radiation. The factor at the arrow’s head is influenced by the factor at this arrow’s
end.

132
Grid ratio is the determining factor for grid application limits, which are the distances at which Tp

is reduced by a predefined percentage (Hondius Boldingh, 1961). This predefined percentage is

40% for general application grids and 20% for mammographic grids (International

Electrotechnical Commission, 2013).

As the grid ratio determines a grid’s application limits, in the following analysis of the relationship

illustrated in Figure 6.1, it is assumed that the grid ratio is constant. Equivalently, the grid

application limits are constant.

The material and distance of the grid interspace are the essential factors for maximising Tp. These

factors’ effects on Ts should be negligible because allowing the primary radiation to travel through

the interspace region would also permit the scatter radiation to travel through this region.

The height, material, and thickness of the grid strip are crucial factors for optimising Ts. The

thickness affects Tp because primary radiation travelling inside strips is completely absorbed.

Since primary radiation does not travel through grid strips, the combination effect of strip material

and height on Tp can be ignored. Strip height, however, determines the interspace height which

may affect Tp.

In the previous chapter, Tp and Ts were found that they increase as the tube voltage increases.

Tube voltage is a determining factor for x-ray beam quality, which together with the imaging

anatomy determines the distribution of scatter radiation. The beam quality directly affects Tp. The

distribution of scatter radiation directly affects Ts.

133
6.3. Issues in grid design

Issues in grid design relate to the determinations of Tp and Ts, which are the basic grid

performance factors. These factors can be determined through experiments with prototype grids;

however, manufacturing prototype grids is not only expensive but also takes a long time. In many

situations, such as designing new grids or simulating x-ray imaging systems with grids,

determining Tp and Ts for grid designs is essential.

When the directions and locations of the photons on the grid entry surface are known, the

transmission of these photons in the grid’s materials can be determined. In a 3-dimension Cartesian

coordinate system, a photon (P) with energy ɛ enters a grid from its entry surface at an arbitrary

point (x, y, z) with a direction (Δx, Δy, Δz). This photon, without interaction with the grid, would

exit this grid and reach a point S(x0, y0, z0) in the image receptor’s area. The integration of the

energy of all photons without interacting with the grid materials and reaching point S can be

expressed as Equation 6.1. If the grid were not there, the integration of the energy of all these

photons reaching the image receptor area could be expressed as Equation 6.2.

Neglecting the effects of these photons’ directions and the radiation detector’s materials on energy

deposition in the image receptor, Equation 6.1 determines the total energy at the point S with the

grid and Equation 6.2 determines the total energy at the point S without the grid. These effects can

be accounted for in these equations by including the angular factors of these directions and the

materials’ energy absorption coefficients.

Primary photons’ coordinates and directions at the grid’s entry surface can be analytically

determined. However, scatter photons’ coordinates and directions at the grid’s entry surface cannot

be analytically determined. This is because scatter photons have randomly, non-uniformly arisen

134
from the imaging anatomy. Without knowing the scatter photons’ coordinates and directions at the

grid’s entry surface, Ts cannot be determined using Equations 6.1 and 6.2.

𝐸(𝑥0 , 𝑦0 , 𝑧0 ) =

∫ ∫ ∫ ∫ ⟨𝑃(𝜀, 𝑥, 𝑦, 𝑧, ∆𝑥, ∆𝑦, ∆𝑧)|𝑃(∆𝑥, ∆𝑦, ∆𝑧) ≡ 𝑆(𝑥0 , 𝑦0 , 𝑧0 )⟩ 𝑒 −𝑢𝑠 (𝜀)∗𝑙𝑠 −𝑢𝑖 (𝜀)∗𝑙𝑖 𝑑𝜀𝑑𝑧𝑑𝑦𝑑𝑥

(6.1)

𝐸(𝑥0 , 𝑦0 , 𝑧0 ) = ∫ ∫ ∫ ∫ ⟨𝑃(𝜀, 𝑥, 𝑦, 𝑧, ∆𝑥, ∆𝑦, ∆𝑧)|𝑃(∆𝑥, ∆𝑦, ∆𝑧) ≡ 𝑆(𝑥0 , 𝑦0 , 𝑧0 )⟩ 𝑑𝜀𝑑𝑧𝑑𝑦𝑑𝑥

(6.2)

where < A | B > is a function that returns ‘A’ if ‘B’ is true, otherwise returns zero,
P(ɛ,x, y, z, Δx, Δy, Δz) returns the photon P’s energy ɛ, P(Δx, Δy, Δz) ≡ S(x0, y0, z0) returns
true if and only if aΔx = x0, aΔy = y0, and aΔz = z0 with ‘a’ being a positive real number,
µs(E) and µi(E) are the linear attenuation coefficients of the strip material and interspace
material for photons of energy E, respectively, and ls and li are the distances that the photons
will travel in the strip material and interspace material without interaction, respectively.

The Monte Carlo simulation of x-ray imaging can determine an approximation of the distribution

of scatter photons at the grid entry surface. The accuracy of this approximation, however, can be

increased by increasing the number of total photons in the simulation (Landau and Binder, 2009).

In this project, the Tp and Ts of grid designs were determined through Monte Carlo simulations.

6.4. Criteria for designing new grids

Grids should be designed to substantially remove scatter radiation reaching the image receptor

(minimising Ts), and to also maximally retain primary radiation reaching the image receptor

(maximising Tp). As primary radiation is generated at a finite x-ray focal spot, the primary

135
radiation’s beam reaching the grid entry surface is a divergent beam. Since scatter radiation arises

randomly and non-uniformly from any point in the imaging anatomy, the scatter radiation reaching

the grid entry surface is non-uniformly distributed.

Focused grids permit maximal transmission of primary radiation as the strips of these grids are

aligned to the divergent primary beam. They may be linear or cross-hatched. A linear focused grid

has its strips parallel and aligned to focus on a common line. Contemporary grids are almost always

focused. In the following discussion, focused grids with a given grid ratio are assumed.

Tp depends on four factors: the strip thickness, the interspace material, the interspace distance, and

the strip height. Strid (1976) has found that Ts is optimised by an optimal strip thickness. In

Chapter 2, a criterion was proposed for determining strip optimal thickness. Strip thickness,

therefore, can be treated as a constant.

The interspace region is best as a vacuum. Alternatively, it can be air-filled for easy grid

construction.

Since the grid ratio is fixed, a change in the interspace distance must be paired with a proportional

change in the strip height. As the strip thickness is constant, when the interspace distance is

increased, the Tp will increase. Tp then can be increased by concurrently increasing the interspace

distance and the strip height without changing the grid ratio. Alternatively, Tp depends on the strip

height.

Ts depends on four factors: scatter radiation distribution, strip material, strip height, and strip

thickness. The scatter radiation distribution, which depends on both the imaging anatomy and the

x-ray beam quality, varies from one x-ray examination to another. The scatter radiation distribution

136
is set for a standard imaging condition (International Electrotechnical Commission, 2013, 2001,

1978). This distribution can be treated as a constant.

The strip material should be high atomic number material, such as lead or tungsten. However, lead

is highly poisonous; if it is inhaled or swallowed, it damages the nervous system and causes brain

disorders. Therefore, tungsten strips are more desirable for all types of grids. Molybdenum strips

are also desirable for mammographic grids.

As the strip height is proportionally increased to maximise Tp as the interspace distance is

increased, Ts should not depend on the strip height. Ts, then, primarily depends on the strip

thickness which, as discussed above, can be optimised. The strip optimal thickness can be

determined using the criterion proposed in Chapter 2.

After the interspace distance and strip height are proportionally increased, the volume of the

imaging anatomy (the green region in Figure 6.2) is increased. The scatter radiation arising in this

volume may travel directly towards the interspace underneath this volume and through the grid

without any interaction with the grid’s materials. If the grid ratio increases, this volume will

decrease. Therefore, this volume can be kept approximately unchanged or decreased by increasing

the grid ratio.

137
Figure 6.2. Relationships between strip thickness (d), interspace (D), strip height (h), and front
and back cover thicknesses (x1 and x2, respectively).

Based on the above, there are three essential criteria for designing new grids with a given grid ratio:

1) using air-interspace material

2) proportionally increasing interspace distance and strip height

3) optimising strip thickness.

Summary

The prime goal of the second phase of this research was to develop a new method to design new

grids to overcome the reduction of primary radiation in grid materials. In this chapter, an overview

of two current contemporary grid design theories was presented. One of these theories, the Strid’s

grid design theory, emphasises on optimising grid strip thickness to maximise grid performance.

The other theory, the Bonenkamp and Hondius Boldingh’s grid design theory, focuses on

maximising Tp for a given lead content.

138
In this chapter, the factors and issues in grid design were also analysed. Then new criteria were

determined for designing new grids: (1) using air-interspace materials; (2) increasing interspace

distance and the strip height, proportionally; (3) optimising strip thickness. These provide guidance

for designing new air-interspaced grids whose Tp and Ts will only depend on their strip height and

strip optimal thickness, respectively. Such new grids could overcome the problem of the reduction

of primary radiation in grid materials.

In the next chapter, such new grid designs were evaluated through Monte Carlo simulation to

determine a solution for the reduction of scatter radiation without increasing patients’ radiation

exposure and without compromising image diagnostic quality.

139
Chapter 7: Monte Carlo evaluation of new grid designs
and performance results

This chapter describes the last phase of the project, using all the achievements made in the previous

phases to evaluate a solution for avoiding the reduction of primary radiation reaching the image

receptor. A series of new grid designs was constructed and evaluated through Monte Carlo

simulations. This chapter has three sections: the first provides the details of material and methods

used for the evaluation; the second presents the evaluation results; and the third compares these

results with the literature.

7.1. Material and methods


7.1.1. Grid details

7.1.1.1 Mammographic grid designs

New mammographic grid designs were made with reference to scatter-to-primary ratio (SPR), tube

voltages, and grid ratios in current mammography, where SPR is approximately 0.2–1.3 (Barnes

and Brezovich, 1978, Boone and Cooper III, 2000, Carton et al., 2009, Zhou et al., 2016), tube

voltages may vary from approximately 24 peak kilovoltage (kVp) to 49 kVp (Carton et al., 2009,

Cunha et al., 2010), and mammographic grids have a lead strip around 20 μm thick and grid ratios

4:1 or 5:1 (Carton et al., 2009, Cunha et al., 2010).

A series of mammographic grid designs was made with a range of grid ratios and a range of strip

thicknesses. The combination of these ratios and thicknesses gave a total of 781 grids. The grids

were designed with tungsten strips and air interspace, and the strip height was fixed at 2 cm taller

141
than contemporary mammographic grids’ strip heights. This height would increase the penumbra

effect. Therefore, a justification was made to reduce the potential penumbra effect by increasing

the grid focal distance so that the image’s lack of sharpness ascribed to the penumbra effect was

reduced. The details of these grids were listed in Table 7.1.

Table 7.1. Details of mammographic grids

Grid ratio 5:1 to 15:1


Strip thickness d (µm) 2 to 50

Strip material Tungsten


Interspace material Air
Strip height h (cm) 2
Grid total cover thickness (µm) 100
Cover material Fibre
Focal distance f0 (cm) 70

7.1.1.2. General grid designs

New general grid designs were made with reference to SPR, tube voltages, and grid ratios in

current general x-ray examinations, where SPR might vary from less than 0.1 (such as for a finger

x-ray) to 26 (such as for a 50-cm thick abdominal x-ray) (Fetterly and Schueler, 2007), tube

voltages of general x-ray beams have a broad range (such as 50 kVp for wrist x-ray or 120 kVp

for a chest x-ray), general grid ratios may vary from 5:1 to 21:1 (Fetterly and Schueler, 2009), and

grid lead strip thickness may vary from 30 µm to more than 100 μm (Strid, 1976, Fetterly and

Schueler, 2009).

A series of general grid designs were made with a range of grid ratios and strip thicknesses. The

combination of these ratios and thicknesses gave a total of 1800 grids. These grids were designed

142
with tungsten strips and air interspace, and the grid height was fixed at 5 cm. This height was

justified for current contemporary Bucky systems, which have a space in between the exit-surface

of the imaging anatomy and the entry surface of the image receptor. This space is intentionally

designed using the air-gap technique to improve the efficiency of scatter radiation reduction (Bell

et al., 2003). It is approximately 10 cm and allows an adequate space for these new grids to be

successfully assembled in Bucky systems. The details of these grids are listed in Table 7.2.

Table 7.2. Details of general grids

Grid ratio 5:1 to 40:1


Strip thickness d (µm) 2 to 140
Strip material Tungsten
Interspace material Air
Strip height h (cm) 5
Grid total cover thickness (µm) 100
Cover material Fibre
Focal distance f0 (cm) 120

7.1.2. Virtual phantoms

7.1.2.1. Virtual breast phantoms for mammographic grid evaluation

Two virtual breast phantoms of semi-circular cross-section were used in the simulation. The cross-

section had an 8-cm radius. These phantoms were composed of, in sandwich structure, 50%

glandular tissue and 50% adipose tissue using ICRU (1989) standards (Table 7.3).

In the literature, a 5-cm thick breast phantom is often used for grid performance evaluation, a

thickness that is recommended by the International Electrotechnical Commission (2013). The

image diagnostic quality of thin breasts was found not to benefit from mammographic grids, but

143
the image diagnostic quality of thick (approximately 5 cm or more) breasts was found to benefit

(Bernhardt et al., 2006).

Table 7.3. Sandwich-structured breast-phantom compositions

Tissue type* (1 cm thickness for each layer)


1st layer Adipose Adipose
2nd layer Glandular Glandular
3rd layer Adipose
4th layer Glandular
5th layer Adipose
6th layer Glandular
7th layer Adipose
8th layer Glandular

Total thickness 2 cm 8 cm

*Tissues are of The International Commission on Radiation Units and Measurements


(1989) standards.

In this project, two phantoms were selected to represent the thinnest (2 cm) and thickest (8 cm)

breasts. If a grid could improve the image diagnostic quality of both the thinnest breast and the

thickest breast, this grid should benefit the image diagnostic quality of all breasts.

The justification for using 50% glandular tissue, 50% adipose tissue, and semi-circular cross-

section phantoms was that this phantom’s composition was the closest approximation of breasts

of various thicknesses from 2 cm to 8 cm (Cunha et al., 2010). SPRs from different phantom

material have been found to be different (Dance et al., 2000, Huda et al., 2003, Cunha et al., 2010,

Cunha et al., 2012, Dance et al., 2012, Tomal et al., 2013, Sisniega et al., 2013). Ts changes as

SPR changes (Carton et al., 2009, Salvagnini et al., 2012, Zhou et al., 2016). Using approximate

breast phantoms would minimise any potential effects on Ts due to the difference between the

breasts and the breast phantoms.

144
7.1.2.2. Virtual water phantoms for general grid evaluation

Two virtual water phantoms were used to evaluate new general grid designs. One phantom was 10

cm thick, which represents a thin anatomy such as an ankle. The other phantom was 30 cm thick,

which represents a thick anatomy such as an abdomen. These phantoms had a cross-section of 30

cm × 30 cm.

The justification for using water phantoms was that they approximated body soft tissue

composition. A 20-cm thick water phantom is recommended for general grid evaluation (Strid,

1976, Fetterly and Schueler, 2007, International Electrotechnical Commission, 2013, International

Electrotechnical Commission, 2001, International Electrotechnical Commission, 1978). The 10

cm and 30 cm thicknesses were selected in this study because if the images of both the thin and

thick anatomies could benefit from a grid, this grid should be of benefit to the image quality

improvement of all anatomies. In general x-ray examinations, an imaging anatomy might be

thicker than 50 cm for a very large person. However, this project did not use a phantom thicker

than 30 cm because the 30 cm is the thickness of most thick imaging anatomies: the thicknesses

of most lateral spines is approximately 30 cm and the thicknesses of extremities, head and body

trunk are less than 30 cm.

Another reason for not using a phantom thicker than 30 cm was to save simulation time. To achieve

a similar level of statistic error, the time needed for the simulation of a 40-cm thick anatomy can

be 20 times more than that needed for a 30-cm thick anatomy. In this project, one simulation of a

30-cm thick anatomy took approximately 10 hours.

145
7.1.3. X-ray beam quality

7.1.3.1. Mammographic x-ray beam quality

Ts increases as tube voltage increases (Carton et al., 2009, Zhou et al., 2016). In mammography,

tube voltages are selected depending on breast thickness. The highest possible tube voltages of x-

ray beams for breast thicknesses were selected for evaluating the new mammographic grid designs.

These x-ray beams were calculated using the model of Boone et al. (1997). The beams’ details

were given in Table 7.4.

Table 7.4. Details of x-ray beam


Ari KERMA half-value
kVp* Phantom thickness (cm) Mean energy (keV)
layer Al (mm)

28 2 17.7 0.311

36 8 19.5 0.404

*Rhodium anode and rhodium filtration (0.02 mm)

7.1.3.2. General x-ray beam quality

In general radiography, tube voltages are generally selected depending on the imaging anatomy

thickness: the greater the thickness, the higher the tube voltage. In the simulation, tube voltages

were increased as the phantom thickness increased. X-ray beams were calculated using TechnicVR

v2.0 (Shaderware, Darlington, England); details are listed in Table 7.5.

146
Table 7.5. Details of x-ray beam

kVp Phantom thickness (cm) Mean energy (keV) Half-value layer Al (mm)

60 10 cm 38.5 2.8

102 30 cm 53.0 4.8

7.1.4. Virtual image receptor

Computed radiography/mammography uses a photostimulate phosphor plate to record a latent

image, and the image is read using an imaging plate reader. The reader uses a specific laser to

stimulate the phosphors in the plate to emit light which then is detected by a photomultiplier tube.

Digital mammography/radiography uses various types of image receptors. These image receptors

form part of either a directly detectable imaging system or an indirectly detectable imaging system.

In indirectly detectable imaging systems, radiation interacts principally with a phosphor screen.

Immediately after an interaction with radiation, the phosphors in the screen emit visible light which

is then detected by a two-dimensional array of photodiodes (Yaffe and Rowlands, 1997). Image

receptors used in indirect detectable imaging systems are a Gd2O2S:Tb (Gadolinium oxysulphide)

or CsI (caesium iodide scintillator) phosphor screen coupled with photodiodes (Zhao et al., 2004).

In directly detectable imaging systems, a photoconductor, such as amorphous selenium (a-Se),

polycrystalline layers of HgI2, PbO, or CdZnTe, is used as the principal detecting element. The

photoconductor absorbs x-ray photons and converts their energy directly to electronic charges

(Kasap et al., 2011). The a-Se detectors are suitable for mammographic x-ray beams because they

have a high detection quantum efficiency for these beams (Kasap et al., 2011).

147
An amorphous selenium image receptor was constructed in the simulation for the mammographic

grid evaluation, with an 85 µm pitch and a 200 μm thick selenium layer. It had 2816 × 3584 pixels,

was 240 mm × 300 mm, and was like a flat-panel detector used in the Siemens MAMMOMAT

Inspiration (Siemens AG Healthcare, Erlangen, Germany).

A CsI/Tl (550 µm thickness) image receptor was constructed in the simulation for the general grid

evaluation. It was 43 cm × 43 cm, 3000 × 3000 pixels, and was like the Pixium 4600 CsI/Tl flat-

panel image receptor (Trixell, Moirans, France), whose performance was evaluated in Geijer et al.

(2001).

7.1.5. Simulation setup and photon transportation


The simulation setup was the same as that used in Chapter 5, except for the logic labels which

were not used. Primary or scatter photons were distinguished by their marker, primary or scatter,

respectively. The simulation of photon transportation was performed in Matlab Ver. R2015b (The

MathWorks Inc., Natick, Massachusetts) on a computer using an Intel® Core™ i7–5600U CPU,

2 cores of 2.6 GHz processors and 16 Gigabyte RAM. Radiation transmissions in the grid materials

were determined using the method developed and validated in Chapter 4.

The simulation of each grid evaluation for each phantom was repeated 10 times to obtain 10 sets

of measurements to calculate means and standard deviations.

7.1.5. Data fitting

All simulation results were fitted using an interpolant from Matlab Ver. R2015b (The MathWorks

Inc., Natick, Massachusetts) with a biharmonic (v4) method. This interpolant uses the “griddata”

function in the Matlab for the biharmonic method and fits an interpolating surface that passes

148
through every data point. The fitting statistics were summarised using two parameters: the sum of

squared error (SSE) and the coefficient of determination (R-squared). SSE is also known as the

sum of the squared difference between each measurement, and the measurements’ mean and is a

measure of variation within the measurements (Durbin and Watson, 1951). SSE can be determined

using Equation 7.1. R-squared is the square of the correlation coefficient, which measures the

degree of a linear relationship between variables. R-squared can be used to analyse the degree of

connection between variables (Sharma et al., 1981) and can be determined using Equation 7.2

SSE = ∑𝑛𝑖=1(𝑥𝑖 − 𝑥)2 (7.1)

Where ∑𝑛𝑖=1 𝑎𝑖 is the sum of 𝑎𝑖 for indices 𝑖 = 1 to 𝑖 = n, 𝑥𝑖 is the ith sample of the 𝑥
variable, and 𝑥 is the mean of the samples.

2
𝑛×∑𝑛𝑖=1(𝑥𝑖 ×𝑦𝑖 )−∑𝑛𝑖=1 𝑥𝑖 ×∑𝑛𝑖=1 𝑦𝑖
R_squared = [ 2 2
] (7.2)
√[∑𝑛𝑖=1 𝑥𝑖 2 −(∑𝑛𝑖=1 𝑥𝑖 ) ]×[∑𝑛𝑖=1 𝑦𝑖 2 −(∑𝑛𝑖=1 𝑦𝑖 ) ]

Where ∑𝑛𝑖=1 𝑎𝑖 is the sum of 𝑎𝑖 for indices 𝑖 = 1 to 𝑖 = n; 𝑥𝑖 is the ith sample of the 𝑥
variable; 𝑦𝑖 is the ith sample of the 𝑦 variable; the ith sample of the 𝑥 variable is paired with
the ith sample of the 𝑦 variable; n is the total number of paired samples.

7.2. Results
7.2.1. Mammographic grid results

7.2.1.1. Mammographic grids: Tp

The fitting statistics of the mammographic grid designs’ Tp are, for both phantoms, the SSE less

than 0.001 and the R-squared greater than 0.999.

149
Tp decreased as the grid ratio or/and the strip thickness increased (Figures 7.1 and 7.2). As the

grid ratio and the strip thickness increased from 5:1 to 15:1 and 2 μm to 50 μm, respectively, Tp

decreased in the 2-cm phantom from 0.993 to 0.958 (SD < 0.001), and in the 8-cm phantom from

0.995 to 0.959 (SD < 0.001).

Figure 7.1. Tp for the 2-cm phantom as a function of the strip thickness and grid ratio. Tp was
fitted with a surface plane.

150
Figure 7.2. Tp for the 8-cm phantom as a function of the strip thickness and grid ratio. Tp was
fitted with a surface plane.

7.2.1.2. Mammographic grids: Ts

The fitting statistics of the new mammographic grid designs’ Ts were calculated. The SSE was

0.004 for the 10-cm phantom and 0.001 for the 30-cm phantom. The R-squared was greater than

0.999 for both phantoms.

Ts decreased as the grid ratio or/and strip thickness increased (Figures 7.3 and 7.4). As the grid

ratio and strip thickness increased from 5:1 to 15:1 and 2 μm to 50 μm, respectively, Ts decreased

in the 2-cm phantom from 0.352 to 0.052 (SD < 0.001) and in the 8-cm phantom from 0.521 to

0.064 (SD < 0.001).

151
Figure 7.3. Ts for the 2-cm phantom as a function of the strip thickness and grid ratio. Ts was
fitted with a surface plane.

Figure 7.4. Ts for the 8-cm phantom as a function of the strip thickness and grid ratio. Ts was
fitted with a surface plane.

152
7.2.1.3. Mammographic grids: KSNR

KSNR is the quantum signal-to-noise ratio (SNR) improvement factor (KSNR). The fitting statistics

of the mammographic grid designs’ KSNR were calculated. The SSE was less than 0.001 for the

10-cm phantom and 0.002 for the 30-cm phantom. The R-squared was greater than 0.999 for both

phantoms.

For a given grid ratio, KSNR first increased then decreased as the strip thickness increased (Figures

7.5 and 7.6). For a thin strip (e.g. 2 µm), KSNR increased as the grid ratio increased from 5:1 to

15:1. For a thick strip (e.g. 40 µm), KSNR first increased then decreased as the grid ratio increased.

As the grid ratio increased from 5:1 to 15:1 at the strip optimal thicknesses, KSNR increased in the

2-cm phantom from 1.257 to 1.277 (SD < 0.001) and in the 8-cm phantom from 1.992 to 2.135

(SD < 0.001).

153
Figure 7.5. KSNR for the 2-cm phantom as a function of the strip thickness and grid ratio. KSNR
was fitted with a surface plane.

Figure 7.6. KSNR for the 8-cm phantom as a function of the strip thickness and grid ratio. KSNR
was fitted with a surface plane.

154
7.2.1.4. Strip optimal thicknesses of mammographic grids

In Section 2.2, a criterion was discussed for the determination of strip optimal thickness for a given

strip height and grid ratio. This criterion was used to determine the mammographic grid designs’

strip optimal thicknesses.

With the same phantom thickness, the strip optimal thickness decreased as the grid ratio increased

(Figures 7.7 and 7.8). With the same grid ratio, the strip optimal thickness increased as the phantom

thickness increased. As the phantom thickness and tube voltage increased together, the increase of

the optimal thickness was ascribed to the combination effect of the phantom thickness and the tube

voltage. It is anticipated that the strip optimal thickness is affected by the tube voltage, the phantom

thickness, and the grid ratio.

As the grid ratio increased from 5:1 to 15:1 at these strip optimal thicknesses, Tp changed less than

1%; Ts decreased in the 2-cm phantom from 0.168 to 0.058 and in the 8-cm phantom from 0.198

to 0.069; and KSNR increased in the 2-cm phantom from 1.257 to 1.277 and in the 8-cm phantom

from 1.992 to 2.135.

155
Figure 7.7. Tp and Ts for the 2-cm phantom as functions of the strip thickness and grid ratio.
Both Tp and Ts were fitted with surface planes.

Figure 7.8. Tp and Ts for the 8-cm phantom as functions of the strip thickness and grid ratio.
Both Tp and Ts were fitted with surface planes.

156
7.2.2. General grid results

7.2.2.1. General grids: Tp

The fitting statistics of these general grid designs’ Tp were calculated. The SSE was 0.007 for the

10-cm phantom and 0.005 for the 30-cm phantom. The R-squared was greater than 0.999 for both

phantoms.

Tp decreased as the grid ratio and/or the strip thickness increased (Figures 7.9 and 7.10). Tp

decreased from 0.994 to 0.894 (SD = 0.002) as the grid ratio and strip thickness increased from

5:1 to 40:1 and 2 μm to 140 μm, respectively. The difference in Tp between the two phantoms was

less than 0.1%.

Figure 7.9. Tp for the 10-cm phantom as a function of the strip thickness and grid ratio. Tp was
fitted with a surface plane.

157
Figure 7.10. Tp for the 30-cm phantom as a function of the strip thickness and grid ratio. Tp
was fitted with a surface plane.

7.2.2.2. General grids: Ts

The fitting statistics of these general grid designs’ Ts were calculated. The SSE was 0.006 for the

10-cm phantom and 0.001 for the 30-cm phantom. The R-squared was greater than 0.999 for both

phantoms.

Ts decreased as the grid ratio and/or the strip thickness increased. As the grid ratio and strip

thickness increased from 5:1 to 40:1 and 2 μm to 140 μm, respectively, Ts decreased in the 10-cm

phantom from 0.835 to 0.023 (SD < 0.001) and in the 30-cm phantom from 0.844 to 0.025 (SD <

0.001).

158
Figure 7.11. Ts for the 10-cm phantom as a function of the strip thickness and grid ratio. Ts
was fitted with a surface plane.

Figure 7.12. Ts for the 30-cm phantom as a function of the strip thickness and grid ratio. Ts
was fitted with a surface plane.

159
7.2.2.3. General grids: KSNR

The fitting statistics of these general grid designs’ KSNR were calculated. The SSE was 0.128 for

the 10-cm phantom and 1.277 for the 30-cm phantom. The R-squared was greater than 0.999 for

both phantoms.

For a given grid ratio, KSNR first increased then decreased as the strip thickness increased (Figures

7.13 and 7.14). For a thin strip (for example 2 µm), KSNR increased as the grid ratio increased from

5:1 to 40:1. For a thick strip (for example 130 µm), KSNR first increased then decreased as the grid

ratio increased. As the grid ratio increased from 5:1 to 40:1 at the strip optimal thicknesses, KSNR

increased in the 10-cm phantoms from 2.786 to 3.303 (SD = 0.01) and in the 30-cm phantom from

5.670 to 10.073 (SD = 0.02).

Figure 7.13. KSNR for the 10-cm phantom as a function of the strip thickness and grid ratio.
KSNR was fitted with a surface plane.

160
Figure 7.14. KSNR for the 30-cm phantom as a function of the strip thickness and grid ratio.
KSNR was fitted with a surface plane.

7.2.2.4. Strip optimal thicknesses of general grids

The strip optimal thicknesses of these general grid designs were determined using the criterion

used to determine the mammographic grids’ strip optimal thicknesses.

With the same phantom thickness, the general grid designs’ strip optimal thickness decreased as

the grid ratio increased (Figures 7.15 and 7.16). With the same grid ratio, the strip optimal

thickness increased as the phantom thickness increased. The trend of the strip optimal thickness

was the same for the mammographic grid designs and the general grid designs. As the tube voltages

for the general grid designs’ evaluation were also selected depending on phantom thickness, the

general grid designs’ strip optimal thickness should also be affected by the tube voltage, the

phantom thickness, and the grid ratio.

161
As the grid ratio increased from 5:1 to 40:1 at these strip optimal thicknesses (Figures 7.15 and

7.16), Tp decreased in the 10-cm phantom from 0.983 to 0.978 and in the 30-cm phantom from

0.974 to 0.971; Ts decreased in the 10-cm phantom from 0.206 to 0.029 and in the 30-cm phantom

from 0.311 to 0.029; KSNR increased in the 10-cm phantom from 2.786 to 3.303 and in the 30-cm

phantom from 5.670 to 10.073.

Figure 7.15. Tp and Ts for the 10-cm phantom as functions of the strip thickness and grid ratio.
Tp and Ts were fitted with surface planes.

162
Figure 7.16. Tp and Ts for the 30-cm phantom as functions of the strip thickness and grid ratio.
Tp and Ts were fitted with surface planes.

7.3. Discussion
7.3.1. Mammographic grids

7.3.1.1. New mammographic grid designs vs the literature: Tp and Ts

The Tp and Ts of the new mammographic grid designs with strip optimal thicknesses were

compared with the literature. Despite the grid ratios and/or phantom thicknesses, the Tp of the new

mammographic grid designs were approximately 23% higher than those reported by Carton et al.

(2009), or 33% higher than those reported by Salvagnini et al. (2012) (Figures 7.17 and 7.18).

163
Tp 1

Tp - 2cm phantom
0.9
Tp - 2cm phantom & 35kVp (Carton
et al. 2009)
Tp - 2cm phantom & Inspiration
0.8 system (Salvagnini et al., 2012)

0.7
4 5 6 7 8 9 10 11 12 13 14 15
Grid ratio

Figure 7.17. Tp as a function of the grid ratio.

Tp 1

Tp - 8cm phantom
0.9
Tp - 8cm phantom & 35kVp (Carton
et al. 2009)
Tp - 8cm phantom & Inspiration
0.8 system (Salvagnini et al., 2012)

0.7
4 5 6 7 8 9 10 11 12 13 14 15
Grid ratio

Figure 7.18. Tp as a function of the grid ratio.

164
The Ts of the new mammographic grid designs depended on the grid ratio and/or phantom

thickness. The dependence of Ts on phantom thickness agrees with the findings of Carton et al.

(2009). Contemporary mammographic grids are commonly built with a grid ratio of 5:1 (Carton

et al., 2009, Cunha et al., 2010, Salvagnini et al., 2012). The Ts of the grid ratio 5:1 for the new

mammographic grid designs was higher than those of Carton et al. (2009) and Salvagnini et al.

(2012) (Figures 7.19 and 7.20). For the 2-cm phantom, the new mammographic grid design’s Ts

was approximately 20% higher than Carton et al. (2009) and approximately 14% higher than

Salvagnini et al. (2012) (Figure 7.19). For the 8-cm phantom, the new mammographic grid

design’s Ts was approximately 18% higher than Carton et al. (2009) and 27% higher than

Salvagnini et al. (2012) (Figure 7.20).

Ts

Ts - 2cm phantom

0.15 Ts - 2cm phantom & 35kVp (Carton


et al. 2009)

Ts - 2cm phantom & Inspiration


system (Salvagnini et al., 2012)

0.05
4 5 6 7 8 9 10 11 12 13 14 15
Grid ratio

Figure 7.19. Ts as a function of the grid ratio.

165
Ts
Ts - 8cm phantom

Ts - 8cm phantom & 35kVp (Carton


0.15
et al. 2009)

Ts - 8cm phantom & Inspiration


system (Salvagnini et al., 2012)

0.05
4 5 6 7 8 9 10 11 12 13 14 15
Grid ratio

Figure 7.20. Ts as a function of the grid ratio.

The difference in Ts between the grid ratio of 5:1 for the new mammographic grid design and

those in Carton et al. (2009) and Salvagnini et al. (2012) may be due to several causes: the

difference in the x-ray beam quality, the difficulty in the experimental setups, the difference in

these grids’ strip height, or any combination of these. The effect of the strip height on Ts can be

counteracted by increasing the grid ratio: a grid ratio of 6:1 for the new mammographic grid design

has approximately the same Ts as those of Carton et al. (2009) and Salvagnini et al. (2012).

Despite the phantom thicknesses, the Ts of the grid ratio of 15:1 of the new mammographic grid

design was approximately 57% lower than Carton et al. (2009). For the 2-cm phantom, this grid

design’s Ts was approximately 60% lower than Salvagnini et al. (2012). For the 8-cm phantom,

this grid design’s Ts was approximately 46% lower than Salvagnini et al. (2012).

166
7.3.1.2. New mammographic grid designs vs the literature: KSNR

The KSNR of the new mammographic grid designs with strip optimal thicknesses were compared

with the literature. The KSNR of the new mammographic grid designs were greater than 1.26 for

the 2-cm phantom (Figure 7.21) or greater than 1.99 (1.99 to 2.14) for the 8-cm phantom (Figure

7.22). The KSNR of those grids in Carton et al. (2009) and Salvagnini et al. (2012) were 0.82 and

0.90, respectively, for the 2-cm phantom (Figure 7.21) and 1.005 and 1.06, respectively, for the 8-

cm phantom (Figure 7.22).

The KSNR of the new mammographic grid designs were 40 to 43% higher than (Carton et al., 2009)),

and 52 to 55% higher than Salvagnini et al. (2012) for the 2-cm phantom (Figure 7.21). The KSNR

of these new mammographic grid designs were 89 to 102% higher than (Carton et al., 2009)) and

98 to 113% higher than Salvagnini et al. (2012) for the 8-cm phantom (Figure 7.22).

In the literature, mammographic grids were found to improve only thick breasts’ image diagnostic

quality, when breast thickness was larger than approximately 5 cm (Bernhardt et al., 2006). The

KSNR of the new mammographic grid designs were greater than 1.26, regardless of the phantom

thickness and/or grid ratio. New mammographic grid designs, however, will improve image

diagnostic quality for all breasts, whether they are thin or thick.

167
1.4 KSNR

1.3

1.2
2cm phantom

1.1
2cm phantom & 35kVp
1 (Carton et al. 2009)

2cm phantom & Inspiration


0.9 system (Salvagnini et al.,
2012)
0.8

0.7
4 5 6 7 8 9 10 11 12 13 14 15
Grid ratio

Figure 7.21. KSNR as a function of the grid ratio.

KSNR
2.1

1.9
8cm phantom
1.7
8cm phantom & 35kVp
1.5 (Carton et al. 2009)

1.3 8cm phantom & Inspiration


system (Salvagnini et al.,
2012)
1.1

0.9
4 5 6 7 8 9 10 11 12 13 14 15
Grid ratio

Figure 7.22. KSNR as a function of the grid ratio.

168
7.3.2. General grids

7.3.2.1. New general grid designs vs the literature: Tp and Ts

The Tp and Ts of the new general grid designs with strip optimal thicknesses were compared with

the literature. Despite the phantom thickness and/or grid ratio, the Tp of the new general grid

designs were approximately the same (less than 1% difference) (Figures 7.23 and 7.24). The Tp of

Kim et al. (2007) and Fetterly and Schueler (2007, 2009) decreased linearly as the grid ratio

increased (Figures 7.23 and 7.24). The Tp of the new general grid designs were 0.974 to 0.983,

which was 31 to 38% higher than Kim et al. (2007) and 29 to 43% higher than Fetterly and Schueler

(2007, 2009).

Tp - 10cm phantom

Tp - 20cm phantom (Kim et al., 2007)

Tp - 20cm phantom (Fetterly et al., 2007, 2009)


Tp

0.95

0.85

0.75

0.65
5 15 25 35
Grid ratio

Figure 7.23. Tp as a function of the grid ratio.

169
Tp - 30cm phantom

Tp - 20cm phantom (Kim et al., 2007)

Tp - 20cm phantom (Fetterly et al., 2007, 2009)


Tp

0.95

0.85

0.75

0.65
5 15 25 35
Grid ratio

Figure 7.24. Tp is as a function of the grid ratio.

There was a trend for the Ts of the new general grid designs to decrease as the grid ratio increased.

As the grid ratio increased from 5:1 to 41:1, the Ts of these new grid designs decreased from

approximately 0.21 to 0.03 in the 10-cm phantom (Figure 7.25) and approximately 0.31 to 0.03 in

the 30-cm phantom (Figure 7.26). This trend agrees with Fetterly and Schueler (2007), Fetterly

and Schueler (2009) and Kim et al. (2007).

Regardless of the phantom thickness, for the same grid ratio, the new grid designs’ Ts were on

average 56% (23 to 64%) less than Kim et al. (2007), 32% (24 to 40%) less than Fetterly and

Schueler (2007) and 17% (7 to 27%) less than Fetterly and Schueler (2009) (Figures 7.25 and 7.26).

The best Ts of the new grid designs was on average 92% (91 to 94%) less than Kim et al. (2007),

80% (73 to 88%) less than Fetterly and Schueler (2007) and 61% (50 to 72%) less than Fetterly

and Schueler (2009).

170
Ts - 10cm phantom
Ts - 20cm phantom (Kim et al., 2007)
Ts - 20cm phantom (Fetterly et al., 2007, 2009)
Ts

0.5

0.4

0.3

0.2

0.1

0
5 15 25 35
Grid ratio

Figure 7.25. Ts as a function of the grid ratio.

Ts - 30cm phantom
Ts - 20cm phantom (Kim et al., 2007)
Ts - 20cm phantom (Fetterly et al., 2007, 2009)
Ts

0.5

0.4

0.3

0.2

0.1

0
5 15 25 35
Grid ratio

Figure 7.26. Ts as a function of the grid ratio.

171
7.3.2.2. New general grid designs vs the literature: KSNR

The KSNR of the new general grid designs with strip optimal thicknesses were compared with the

literature. The KSNR of the new general grid designs increased as the grid ratio increased (Figures

7.27 and 7.28). This trend of increasing KSNR agrees with Fetterly and Schueler (2007), Fetterly

and Schueler (2009) and Kim et al. (2007).

In the 10-cm phantom, KSNR of the new grids were 2.79 to 3.30 (Figure 7.27). In the 30-cm

phantom, KSNR of the new grids were 5.67 to 10.07) (Figure 7.28). Regardless of the phantom

thicknesses, the new general grid designs’ KSNR were 35 to 635% higher than Fetterly and Schueler

(2007, 2009). The new general grid designs’ KSNR were several times (2.1 to 11.4) more than Kim

et al. (2007).

10cm phantom
50cm phantom (Fetterly et al., 2007, 2009)
20cm phantom (Fetterly et al., 2007, 2009)
20cm phantom (Kim et al., 2007)

KSNR
3.5

2.5

1.5

0.5
5 10 15 20 25 30 35 40
Grid ratio

Figure 7.27. KSNR as a function of grid ratio.

172
30cm phantom
50cm phantom (Fetterly et al., 2007, 2009)
20cm phantom (Fetterly et al., 2007, 2009)
20cm phantom (Kim et al., 2007)

10 KSNR

8
6
4
2
0
5 10 15 20 25 30 35 40
Grid ratio

Figure 7.28. KSNR as a function of grid ratio.

Summary

The discoveries made in the first and second phases of this research were applied in this chapter

to determine a solution for the reduction of primary radiation in grid materials. In this chapter,

many new grid designs were created and evaluated through the Monte Carlo simulation.

The simulation results showed that it is possible to have the Tp of new grid designs approximately

equal to the Tp of a perfect grid (Tp = 1) without compromising the new grid designs’ Ts, which

depends on the grid ratio – the higher the grid ratio, the smaller the Ts. The application of the new

grid designs to scatter radiation reduction will reduce negligible amounts of primary radiation: Tp

is more than 0.99 for the mammographic grid designs and 0.974 to 0.983 for the general grid

designs. When compared to contemporary grids in the literature, new grids using these new grid

designs will reduce the stochastic effects of ionising radiation, due to the compensation for the

173
primary radiation reduction in the grid materials, by approximately 39% for general radiography

and 28% for mammography.

With the same grid ratios, new grids using these new grid designs will have a much smaller Ts

than those of contemporary grids in the literature, except for the grid ratio of 5:1 for the new

mammographic grid design. Furthermore, scatter radiation reaching the image receptor can be

further reduced by using such new grids with a high grid ratio, without increasing the stochastic

effects.

With all new grid designs, the KSNR increases as the imaging anatomy thickness increases and/or

the grid ratio increases. For the thinnest breast (2 cm thickness), the KSNR of the new

mammographic grid designs is 1.26 to 1.28. For the 10-cm thick anatomy, the KSNR of the new

general grid designs is 2.79 to 3.30, which is greater than the KSNR of those grids in the literature

for all phantom thicknesses (10 to 50 cm). For the same grid ratio and the same anatomical

thickness, the KSNR of the new general grid designs is more than three times greater than those in

the literature.

The KSNR of all the new grid designs implies that new grids using these new grid designs will have

the benefit of improved image diagnostic quality, regardless of anatomical thicknesses, either in

mammography or general radiography. In contrast, the contemporary mammographic grids were

found to only improve image diagnostic quality for breasts thicker than approximately 5 cm.

The KSNR of these new grid designs showed that for a given SPR, grid performance is a

combination effect of Tp and Ts. A grid with a high KSNR may not have the lowest Ts; however, a

grid with the lowest Ts among grids with the same Tp will always have the highest KSNR.

174
Building grids using these new designs can provide an effective ionising radiation reduction

management strategy by implementing the as low as reasonably achievable (ALARA) principle.

This means that the lowest radiation exposure can be achieved using such new grids to produce x-

ray images without compromising image diagnostic quality.

175
Chapter 8: Conclusions

In this research, new designs of anti-scatter grids were investigated. Grids are commonly used in

medical x-ray imaging systems to reduce scatter radiation reaching the image receptor. The strips

of grids reduce primary radiation transmitting through them and leading to increase patient

exposure to ionising radiation. New grid design criteria were revealed to maximise the

transmission of primary radiation (Tp) for a given grid ratio to avoid increasing patient exposure

to ionising radiation. A novel criterion for the determination of optimal thickness of grid strips was

investigated to maximise grid performance. A new method for radiation transmission in grid

materials was also developed for grid evaluation in Monte Carlo simulation. This method

determines the transmission probability of each photon instead of the grid-unit mean transmission

determined by the contemporary methods. These new grid designs were evaluated using a Monte

Carlo simulation code system which was developed and validated in this research. The results

showed that grids developed with these new designs would reduce scatter radiation reaching the

image receptor without increasing patient exposure to ionising radiation. Replacing current grids

by grids using these new designs, more scatter radiation would be stopped from reaching the image

receptor. In addition, the patient would be exposed to less ionising radiation, approximately

average 28% less in mammography and 39% less in general radiography.

The investigation of new grid designs was carried out in three distinct phases. In the first phase

(Chapters 2 to 5), grid evaluation methodology through Monte Carlo simulation was reviewed and

investigated. The development of a new grid design method was completed in the second phase

177
(Chapter 6). In the final phase (Chapter 7), the achievements made in Chapters 2 to 6 were applied

to determine the transmissions of primary (Tp), scatter (Ts), and total radiation (Tt) of those new

designs. Signal to noise ratio (SNR) improvement factor (KSNR) was derived from these

transmissions. The comparisons of Tp, Ts, Tt, and KSNR between these new grid designs and the

current grids investigated in the literature were also presented in the last phase.

In Chapter 1, an overview of the background and the opportunity for this research on the anti-

scatter grid technique were provided. This technique is used in most medical x-ray imaging

systems which incorporate the use of ionising radiation to provide clinicians with a diagnosis for

their patients. This research focused on grids used in mammography and general radiography.

Mammography and general radiography produce images of which diagnostic quality is formed by

the differentiation attenuation arising when primary radiation interacts with the matter of tissues

or organs.

An interaction between primary radiation and matter can scatter the primary radiation and produce

scatter radiation. Generally, through the interaction, primary radiation ionises the atoms in the

tissues and organs. Ionisation in living tissues and organs causes detrimental effects, including

deterministic effects and stochastic effects. The deterministic effects occur when a patient’s

radiation exposure exceeds the threshold dose. The stochastic effects have no threshold dose. The

probability of the occurrence of the stochastic effects increases as a patient’s cumulative radiation

dose increases. The linear-no-threshold (LNT) theory predicts that the probability of the

occurrence of the stochastic effects is proportional to the cumulative ionising radiation dose

received by a patient.

178
In mammography and general radiography, an image is formed by the intensities of both the

attenuated primary radiation beam and the scatter radiation reaching the image receptor. Images

are commonly stored and viewed in digital forms by many rows and columns of pixels. In a digital

x-ray image, the difference between the pixel values of adjacent regions is the image contrast,

which is reduced by scatter radiation reaching the image receptor. As clinicians rely on image

contrast to make confident and accurate diagnoses from images, scatter radiation in an image,

especially the images of thick imaging anatomies (such as an adult abdomen x-ray image), is a

major problem affecting clinicians’ confidence in making accurate diagnoses. This problem still

exists in mammography and general radiography, as well as other x-ray imaging technologies.

A common solution for removing scatter radiation reaching the image receptor is to use an anti-

scatter grid technique. This technique has been used for nearly a century in most x-ray imaging

systems. Other scatter radiation reduction techniques (the slot technique and air-gap technique) are

not commonly used because of their limitations. Some ‘gridless’ techniques (software solutions)

are currently under development where researchers attempt to study scatter radiation and subtract

it from an image to improve the image’s diagnostic quality. Some applications of these software

solutions have been used in bedside radiography. While the improvement in image diagnostic

quality by these software solutions is not as good as that from the anti-scatter grid technique, these

solutions are useful due to difficulties in using grids when performing bedside radiography.

Grids reduce a great portion of scatter radiation reaching the image receptor. Using grids, however,

still results a relatively great ratio of scatter radiation to primary radiation in images of thick

anatomies. When an x-ray image is undertaken for a 20-cm thick anatomy (such as an adult

abdomen radiography) without using a grid, the scatter-to-primary ratio (SPR) in the image could

179
be as much as 5.9. If a contemporary grid is used, the SPR could still be approximately 1.3. This

means that even if a grid is used, the intensities of the scatter radiation in this image are still about

1.3 times the intensities of the primary radiation.

Grids reduce primary radiation reaching the image receptor. The reduction of primary radiation

increases the image Poisson mottle, which makes it difficult to make a diagnosis from the image.

Radiation exposures are increased to avoid increasing the image Poisson mottle. The increased

radiation exposure, however, increases the stochastic effects of ionising radiation delivered to the

patients. When grids are used, the increase in the stochastic effects due to the increased radiation

exposure to compensate the reduction of primary radiation in the grid materials is approximately

28% in mammography and 39% in general radiography.

Patient radiation protection guides the practice of using ionising radiation in medical diagnostic or

treatment applications to prevent the occurrence of the deterministic effects and minimise

stochastic effects. Using grids that increase stochastic effects of ionising radiation raises concerns

for radiation protection in relation to the as low as reasonably achievable (ALARA) principle. This

principle requires x-ray examinations to be performed using the lowest possible radiation dose

delivered to the patients and without compromising the image diagnostic quality. A challenge in

scatter radiation reduction for mammography and general radiography is to substantially remove

scatter radiation reaching the image receptor, without increasing radiation exposure to the patients

and without compromising the image diagnostic quality.

This problem, that using grids to reduce scatter radiation reaching the image receptor increases the

stochastic effects of ionising radiation delivered to the patients, gave rise to the research in this

180
thesis. This project provided a solution to reduce patient exposure to ionising radiation without

compromising image diagnostic quality by developing and evaluating new grid designs.

In Chapter 2, grid evaluation criteria and methods were reviewed. These criteria and methods were

used to evaluate grid performance in the experiment performed in Chapter 5 and through the Monte

Carlo simulations performed in Chapters 4, 5, and 7. In Chapter 2,the basics grid criteria: Tp, Ts,

and Tt, and KSNR were discussed. Then the setups for the radiation measurements to determine

these criteria using a beam block method was examined. This method determines primary, scatter,

and total radiation using three measurement conditions: narrow beam condition for primary

radiation, broad beam condition with a primary beam blocker for scatter radiation, and broad beam

condition for total radiation. Furthermore, a novel criterion, the first differential of the KSNR, was

analysed to determine the optimal strip thickness for a given grid ratio and strip height. Analysis

showed that the KSNR increases as the strip thickness increases from zero and decreases as the strip

thickness increases further. This novel criterion assumed that there is a maximum KSNR for strip

thickness from zero to infinity. This criterion was used in Chapter 7 to determine the optimal strip

thicknesses of new grids’ designs.

In Chapter 3, details for developing our Monte Carlo simulation code system for grid evaluation

were reviewed and discussed. After the principles for photon transportation in matter were

reviewed, a simulation procedure for the simulation of x-ray imaging was developed. In this

procedure, photon transportation was tracked from the x-ray focal spot to the image receptor. The

details of the events in this procedure were analysed for: (1) changes in a photon’s location,

direction, and/or energy; (2) random number generation and sampling methods; and (3)

interactions between photons and matter; and (4) radiation dosimetry and virtual phantom

181
construction. Our Monte Carlo simulation code system was developed following this simulation

procedure.

The changes in a photon’s location, direction, and/or energy were determined using linear

attenuation coefficients and random numbers. Linear attenuation coefficients were retrieved from

XCOM data base maintained by the National Institute of Standards and Technology, USA.

Uniform random numbers were generated using the Mersenne Twister algorithm, which has a

period of length 219937 − 1. Non-uniform random numbers were sampled using either reject-accept

sampling method or discrete indexed sampling method.

Three types of interactions, Rayleigh scattering, Compton scattering, and photoelectric absorption,

were implemented in our Monte Carlo simulation code system. The scatter photon directions

depended on the atomic form factor for Rayleigh scattering and on the incoherent scattering

function for Compton scattering. If the interference effect of compounds on the scatter photon

direction was accounted for, the scatter photon’s direction was also dependent on the interference

function for Rayleigh scattering. The Doppler effect on the scatter photon’s energy broadening

was ignored as this effect was found negligible; however, this effect could be implemented using

a probability distribution of the energy broadening. In all the interactions, the kinetic energy of

emitted electrons was locally deposited in the matter. The tracking of electrons was ignored.

Virtual phantoms were constructed using either the polynomial boundaries of materials or the

Hounsfield units of volume data from computed tomography.

In Chapter 4, a new method was developed and validated to overcome the limitations set out in the

analytical methods in the literature for determining photon transportation in grid materials. These

analytical methods are used in Monte Carlo simulation of photon transportation to efficiently

182
determine radiation transmission in grid materials. These limitations included: (1) assuming scatter

radiation is uniformly distributed in the grid entry surface; (2) extrapolating radiation transmission

from grid-unit mean transmission, which does not preserve the details of grid lines resulting from

stationary grids; and (3) extrapolating radiation transmission for focused grids. This new method

used photons’ path lengths in strip and interspace materials and these materials’ linear attenuation

coefficients to determine the radiation transmission in the grid.

The validation of this new method was made by comparing grid performance determined through

simulation using this new method and contemporary methods, and that retrieved from the literature.

These comparisons showed that this new method and the Day and Dance method, which is the

most appropriate analytical method for determining radiation transmission in grid materials,

determined approximately the same Tp in both parallel grids and focused grids. However, they

also showed that the Day and Dance method underestimated the mammographic grid’s Ts. The

discrepancy in the mammographic grid’s Ts between this new method and the Day and Dance

method increased as the tube voltage increased. Grids can be operated while moving or stationary.

In the simulations, when the grids were operated when stationary, the Day and Dance method did

not preserve the grid lines because the method only extrapolated radiation transmission from the

grid-unit mean transmission which is not the radiation transmission of stationary grids. But our

new method preserved these grid lines which were shown as white strips in the image. The

comparisons also showed discrepancies in Tp and Ts between the simulation and the literature.

These differences may be ascribed to several causes, such as grids’ manufacturing defects, the

difference between the simulation setup and the experiment setup, and/or difficulty in the

experiment setup.

183
From Chapter 2 to Chapter 4, the details of using a Monte Carlo simulation code system to evaluate

grid designs were studied. These details were used to develop our Monte Carlo code system. This

system was validated through the comparisons of grid performance determined in a simulation and

an experiment. The validation of this system was presented in Chapter 5.

In Chapter 5, an experiment and a simulation were performed to determine SPR, Tp, Ts, and Tt of

one anti-scatter grid. Results showed that the Tp, Ts, Tt, and SPR determined in the simulation

were approximately the same as those determined in the experiment. In addition, through

investigating a range of primary beam’s field-of-view (FOV) and primary beam’s blocker size, it

showed that primary beam’s FOV only negligibly affects Tp. Ts, however, is affected by the

primary beam’s blocker size due to the scatter radiation that would have been arisen from the

phantom’s material underneath the blocker. These results demonstrated that our Monte Carlo

simulation code system is valid for evaluating the design of grids.

The primary goal set in the second phase of this research was accomplished in Chapter 6. A new

grid design method was developed to overcome the reduction of primary radiation in grid materials.

As found in Chapter 1, current grids have not been designed to maximise Tp and optimise Ts. In

Chapter 6, grid design factors and issues were analysed and then new criteria for designing new

grids were determined. These new criteria are: (1) using air-interspace materials; (2) increasing

interspace distance and the strip height, proportionally; and (3) optimising strip thickness. These

criteria provided guidance for designing new air-interspaced grids whose Tp and Ts will only

depend on their strip height and strip optimal thickness, respectively. Such new grids could

overcome the problem of the reduction of primary radiation in grid materials.

184
The evaluation results of the grid new designs were presented in Chapter 7. The achievements

made in the first and second phases were used to evaluate the designs of many new grids through

Monte Carlo simulation. The results from this evaluation were presented in terms of Tp, Ts, and

KSNR. These results affirmed a solution for reducing scatter radiation reaching the image receptor

without increasing patient exposure to ionising radiation.

When a grid is used during imaging, the grid’s Tp is an indication of a patient’s exposure to

ionising radiation: the less the Tp, the greater the patient’s exposure to ionising radiation. The Tp

of new grid designs were approximately equal to the Tp of a perfect grid (Tp = 1) without

compromising these new grid designs’ Ts. Generally, Ts depends on the grid ratio, the higher the

grid ratio, the smaller the Ts. The application of these new grid designs to scatter radiation

reduction would reduce negligible amounts of primary radiation as Tp is more than 0.99 for the

mammographic grid designs and more than 0.97 (0.974 to 0.983) for the general grid designs.

Using new grids made from these new designs to replace contemporary grids would reduce the

stochastic effects of ionising radiation due to the compensation for the primary radiation reduction

in the grid materials. The reduction of these stochastic effects would be approximately 39% for

general radiography and 28% for mammography.

KSNR is an indication of the improvement in image quality. A KSNR equal to 1 means that the grid

would not improve the image quality. A KSNR less than 1 means that the grid would reduce the

image quality; the smaller the KSNR, the greater the reduction. A KSNR greater than 1 means that

the grid would improve the image quality; the higher the KSNR, the greater the image quality

improvement.

185
The KSNR of new mammographic grid designs increased from 1.257 to 1.277 in the 2-cm phantom

and 1.992 to 2.135 in the 8-cm phantom, as grid ratio increased from 5:1 to 15:1. The KSNR of new

mammographic designs was more than 40 to 43% higher than that in the literature for the 2-cm

phantom and more than 89 to 102% higher than that in the literature for the 8-cm phantom. With

the same grid ratios, new grids made from these new mammographic designs would improve the

image quality for all breast thicknesses. In contrast, current mammographic grids were found they

only improve the image quality for breast thicknesses approximately greater than 5 cm.

The KSNR of new general grid designs increased from 2.786 to 3.303 in the 10-cm phantom and

5.670 to 10.073 in the 30-cm phantom, as the grid ratio increased from 5:1 to 40:1. Regardless the

phantom thicknesses, with the same grid ratio, the KSNR of new grids from these new general grid

designs was more than 35% to ten times higher than that in the literature.

The Tp (approximately equal to 1) and KSNR of all new grid designs implied that new grids with

these new grid designs would have the benefit of improved image diagnostic quality, regardless of

anatomy thicknesses, either in mammography or general radiography. Using grids built with these

new designs, an effective ionising radiation reduction management strategy in the implementation

of the ALARA principle can be achieved. This means that the lowest radiation exposure can be

achieved using such new grids to produce x-ray images without compromising image diagnostic

quality.

8.1. Recommendations for future research

New grids from these new designs require further investigation of possible manufacturing defects

and implementation issues. This investigation can be performed through experimental evaluation

186
of a prototype of these new grid designs. Experimental evaluation of a prototype grid will help not

only verify any problems but will also refine these designs.

As these grid designs were evaluated under mammographic and general radiographic conditions,

these designs’ application to cone-beam computed tomography (CT) or fluoroscopy need further

investigation. Current grids were not found to improve image diagnostic quality of cone-beam or

flat-panel CT. As these new grids reduce a negligible amount of primary radiation, they should

improve image diagnostic quality and reduce patient radiation exposures in cone-beam or flat-

panel CT. The investigation of these new grid designs in cone-beam or flat-panel CT can be

undertaken using similar methods to that have been used in this project.

Clinical evaluation of these new grids from the new designs is not within the scope of this project.

The work undertaken in this project has not evaluated these new grids for improvements in

sensitivity and specificity of x-ray examinations. The new grids from the new designs have a

smaller Ts than that from contemporary grids. The new grids would improve sensitivity and

specificity of x-ray examinations much more than these contemporary grids. In addition, further

reduction of scatter radiation can be achieved without increasing patient radiation exposures by

using the new grid with a high grid ratio. Clinical evaluation of these new grids can be undertaken

on an individual x-ray examination basis, tailored to the anatomical region under investigation.

8.2. Summary

On commercial and clinical acceptance of these new grids using these new designs, image

diagnostic quality can be improved with reduced patient radiation exposure. As the operation of

these new grids will not be different from contemporary grids, staff training for using these grids

will not be needed.

187
These new grids from the new designs are expected to produce the highest image diagnostic quality.

The application of these new grids should ultimately not only improve image diagnosis but also

provide an effective strategy for reducing patient radiation exposure.

188
Reference
ALZIMAMI, K., SASSI, S., ALKHORAYEF, M., BRITTEN, A. J. & SPYROU, M. M. 2009. Optimisation
of computed radiography systems for chest imaging. Nuclear Instruments and Methods in Physics
Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 600, 513-
518.

ATTIX, F. H. 2008. Introduction to radiological physics and radiation dosimetry, John Wiley & Sons.

BADAL, A. & BADANO, A. 2009. Accelerating Monte Carlo simulations of photon transport in a
voxelized geometry using a massively parallel graphics processing unit. Medical Physics, 36, 4878-
80.

BAKER, J. E., MOULDER, J. E. & HOPEWELL, J. W. 2011. Radiation as a risk factor for cardiovascular
disease. Antioxidants & Redox Signaling, 15, 1945-1956.

BARNES, G. T. & BREZOVICH, I. A. 1978. The intensity of scattered radiation in mammography.


Radiology, 126, 243-7.

BARNES, G. T., CLEARE, H. M. & BREZOVICH, I. A. 1976. Reduction of scatter in diagnostic radiology
by means of a scanning multiple slit assembly. Radiology, 120, 691-4.

BARNES, G. T., WU, X. & WAGNER, A. J. 1993. Scanning slit mammography. Medical Progress
through Technology, 19, 7-12.

BELL, N., ERSKINE, M. & WARREN-FORWARD, H. 2003. Lateral cervical spine examinations: an
evaluation of dose for grid and non-grid techniques. Radiography, 9, 43-52.

BERGER, M. & HUBBELL, J. 1987. XCOM: Photon Cross Sections, NBSIR.

BERNHARDT, P., MERTELMEIER, T. & HOHEISEL, M. 2006. X-ray spectrum optimization of full-
field digital mammography: simulation and phantom study. Medical Physics, 33, 4337-49.

BERNSTEIN, H., MUNTZ, E. P., SCHRECKENDGUST, J., KLEIN, D. J. & LEE, K. 1983. A detailed
experimental and theoretical comparison of the angular and energy dependencies of grid
transmission. Medical Physics, 10, 218-23.

BOICE JR, J. D., PRESTON, D., DAVIS, F. G. & MONSON, R. R. 1991. Frequent chest X-ray fluoroscopy
and breast cancer incidence among tuberculosis patients in Massachusetts. Radiation research, 125,
214-222.

BONENKAMP, J. G. & HONDIUS BOLDINGH, W. 1959a. Quality and choice of Potter Bucky grids. I.
A new method for the unambiguous determination of the quality of a grid. Acta Radiologica, 51,
479-89.

189
BONENKAMP, J. G. & HONDIUS BOLDINGH, W. 1959b. Quality and choice of Potter Bucky grids. III.
The choice of Bucky grid. Acta Radiologica, 52, 241-53.

BOONE, J. M. & COOPER III, V. N. 2000. Scatter/primary in mammography: Monte Carlo validation.
Medical Physics, 27, 1818-31.

BOONE, J. M., FEWELL, T. R. & JENNINGS, R. J. 1997. Molybdenum, rhodium, and tungsten anode
spectral models using interpolating polynomials with application to mammography. Medical
Physics, 24, 1863-1874.

BOONE, J. M., LINDFORS, K. K., COOPER III, V. N. & SEIBERT, J. A. 2000. Scatter/primary in
mammography: comprehensive results. Medical Physics, 27, 2408-16.

BOONE, J. M. & SEIBERT, J. A. 1988. Monte Carlo simulation of the scattered radiation distribution in
diagnostic radiology. Medical physics, 15, 713-720.

BOREMAN, G. D. 2001. Modulation transfer function in optical and electro-optical systems, SPIE Press
Bellingham, WA.

BORNEFALK, H., XU, C., SVENSSON, C. & DANIELSSON, M. 2010. Design considerations to
overcome cross talk in a photon counting silicon strip detector for computed tomography. Nuclear
Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors
and Associated Equipment, 621, 371-378.

BUCKY, G. 1913. A Grating-Diaphragm to Cut off Secondary Rays from the Object. Archives of the
Roentgen Ray, 18, 6-9.

BURGESS, A. E. 1999. The Rose model, revisited. Journal of the Optical Society of America A, 16, 633-
646.

BUSHBERG, J. T., SEIBERT, J. A., EDWIN M. LEIDHOLDT, J. & BOONE, J. M. 2011. The essential
physics of medical imaging, Lippincott Williams & Wilkins.

BUSHONG, S. C. 2013. Radiologic science for technologists: physics, biology, and protection, Elsevier
Health Sciences.

CARDOSO, S., GONÇALVES, O. & SCHECHTER, H. 2009. Evaluation of scatter-to-primary ratio in


radiological conditions. Applied Radiation and Isotopes, 67, 544-548.

CARDOSO, S. C., GONÇALVES, O. D., SCHECHTER, H. & EICHLER, J. 2003. Modelling the elastic
scattering in diagnostic radiology: the importance of structure form factors. Physics in Medicine
and Biology, 48, 1907.

CARLTON, R. R., ADLER, A. M. & FRANK, E. D. 2012. Principles of radiographic imaging : an art
and a science, Clifton Park, N.Y., Thomson Delmar Learning.

190
CARTON, A. K., ACCIAVATTI, R., KUO, J. & MAIDMENT, A. D. 2009. The effect of scatter and glare
on image quality in contrast-enhanced breast imaging using an a-Si/CsI(TI) full-field flat panel
detector. Medical Physics, 36, 920-8.

CHAN, H. P. & DOI, K. 1982. Investigation of the performance of antiscatter grids: Monte Carlo simulation
studies. Physics in medicine and biology, 27, 785-803.

CHAN, H. P., FRANK, P. H., DOI, K., IIDA, N. & HIGASHIDA, Y. 1985. Ultra-high-strip-density
radiographic grids: a new antiscatter technique for mammography. Radiology, 154, 807-15.

CHANTLER, C. 2000. Detailed tabulation of atomic form factors, photoelectric absorption and scattering
cross section, and mass attenuation coefficients in the vicinity of absorption edges in the soft X-ray
(Z= 30–36, Z= 60–89, E= 0.1 keV–10 keV), addressing convergence issues of earlier work. Journal
of Physical and Chemical Reference Data, 29, 597-1056.

CHEN, H. 2016. PhD Disertation: Characterization and Optimization of Silicon-strip Detectors for
Mammography and Computed Tomography. KTH VETENSKAP OCH KONST.

CHEN, H., DANIELSSON, M., XU, C. & CEDERSTRÖM, B. 2015. On image quality metrics and the
usefulness of grids in digital mammography. Journal of Medical Imaging, 2, 013501-013507.

COMPTON, A. H. 1923. A quantum theory of the scattering of X-rays by light elements. Physical review,
21, 483.

CULLEN, D. E., HUBBELL, J. H. & KISSEL, L. 1997. EPDL97: the evaluated photon data library,’97
version. UCRL-50400, 6, 1-28.

CUNHA, D. M., TOMAL, A. & POLETTI, M. E. 2010. Evaluation of scatter-to-primary ratio, grid
performance and normalized average glandular dose in mammography by Monte Carlo simulation
including interference and energy broadening effects. Physics in medicine and biology, 55, 4335-
59.

CUNHA, D. M., TOMAL, A. & POLETTI, M. E. 2012. Optimization of x-ray spectra in digital
mammography through Monte Carlo simulations. Physics in medicine and biology, 57, 1919-35.

DANCE, D. R. & DAY, G. J. 1984. The computation of scatter in mammography by Monte Carlo methods.
Physics in Medicine and Biology, 29, 237-247.

DANCE, D. R., STRUDLEY, C. J., YOUNG, K. C., ODUKO, J. M., WHELEHAN, P. J. &
MUNGUTROY, E. L. 2012. Comparison of breast doses for digital tomosynthesis estimated from
patient exposures and using PMMA breast phantoms. Breast Imaging. Springer.

DANCE, D. R., THILANDER, A. K., SANDBORG, M., SKINNER, C. L., CASTELLANO, I. A. &
CARLSSON, G. A. 2000. Influence of anode/filter material and tube potential on contrast, signal-
to-noise ratio and average absorbed dose in mammography: a Monte Carlo study. British Journal
of Radiology, 73, 1056-67.

191
DAVIDSON, R. A. 2006. Radiographic contrast-enhancement masks in digital radiography. PhD doctoral
thesis, The University of Sydney.

DAY, G. J. & DANCE, D. R. 1983. X-ray transmission formula for antiscatter grids. Physics in medicine
and biology, 28, 1429-33.

DICK, C. E. & MOTZ, J. W. 1978. New method for the experimental evaluation of x-ray grids. Medical
Physics, 5, 133-40.

DICK, C. E., SOARES, C. G. & MOTZ, J. W. 1978. X-ray scatter data for diagnostic radiology. Physics
in medicine and biology, 23, 1076.

DOODY, M. M., LONSTEIN, J. E., STOVALL, M., HACKER, D. G., LUCKYANOV, N. & LAND, C.
E. 2000. Breast cancer mortality after diagnostic radiography: findings from the US Scoliosis
Cohort Study. Spine, 25, 2052-2063.

DORES, G. M., METAYER, C., CURTIS, R. E., LYNCH, C. F., CLARKE, E. A., GLIMELIUS, B.,
STORM, H., PUKKALA, E., VAN LEEUWEN, F. E., HOLOWATY, E. J., ANDERSSON, M.,
WIKLUND, T., JOENSUU, T., VAN’T VEER, M. B., STOVALL, M., GOSPODAROWICZ, M.
& TRAVIS, L. B. 2002. Second Malignant Neoplasms Among Long-Term Survivors of Hodgkin’s
Disease: A Population-Based Evaluation Over 25 Years. Journal of Clinical Oncology, 20, 3484-
3494.

DURBIN, J. & WATSON, G. S. 1951. Testing for serial correlation in least squares regression. II.
Biometrika, 38, 159-178.

ECKHARDT, R. 1987. Stan ulam, john von neumann, and the monte carlo method. Los Alamos Science,
15, 30.

EIDEMÜLLER, M., HOLMBERG, E., JACOB, P., LUNDELL, M. & KARLSSON, P. 2009. Breast cancer
risk among Swedish hemangioma patients and possible consequences of radiation-induced
genomic instability. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis,
669, 48-55.

EIDEMÜLLER, M., HOLMBERG, E., JACOB, P., LUNDELL, M. & KARLSSON, P. 2011. Breast cancer
risk after radiation treatment at infancy: potential consequences of radiation-induced genomic
instability. Radiation protection dosimetry, 143, 375-379.

EVANS, S. H., BRADLEY, D. A., DANCE, D. R., BATEMAN, J. & JONES, C. H. 1991. Measurement
of small-angle photon scattering for some breast tissues and tissue substitute materials. Physics in
medicine and biology, 36, 7.

FETTERLY, K. A. & HANGIANDREOU, N. J. 2001. Effects of x-ray spectra on the DQE of a computed
radiography system. Medical Physics, 28, 241-249.

192
FETTERLY, K. A. & SCHUELER, B. A. 2007. Experimental evaluation of fiber-interspaced antiscatter
grids for large patient imaging with digital x-ray systems. Physics in medicine and biology, 52,
4863-80.

FETTERLY, K. A. & SCHUELER, B. A. 2009. Physical evaluation of prototype high-performance anti-


scatter grids: potential for improved digital radiographic image quality. Physics in medicine and
biology, 54, N37-42.

FLOWER, M. A. 2012. Webb's physics of medical imaging, CRC Press.

FLOYD, C. E., LO, J. Y., CHOTAS, H. G. & RAVIN, C. E. 1991. Quantitative scatter measurement in
digital radiography using a photostimulable phosphor imaging system. Medical Physics, 18, 408-
413.

FLYNN, M. J. & SAMEI, E. 1999. Experimental comparison of noise and resolution for 2k and 4k storage
phosphor radiography systems. Medical Physics, 26, 1612-23.

GEIJER, H., BECKMAN, K.-W., ANDERSSON, T. & PERSLIDEN, J. 2001. Image quality vs radiation
dose for a flat-panel amorphous silicon detector: a phantom study. European Radiology, 11, 1704-
1709.

GRAHAM, D. T., CLOKE, P. J. & VOSPER, M. 2003. Principles of radiological physics, Sydney,
Churchill Livingstone.

GRANGEAT, P. 2009. Tomography, John Wiley & Sons.

HALL, E. J. 2006. Radiobiology for the Radiologist, Lippincott Williams & Wilkins.

HALL, E. J. & GIACCIA, A. J. 2012. Radiobiology for the Radiologist, Lippincott Williams & Wilkins.

HEITLER, W. 1954. The quantum theory of radiation, Courier Dover Publications.

HENDEE, W. R. & RITENOUR, E. R. 2002. Medical imaging physics, New York, wiley-Liss, Inc.

HERRNSDORF, L. & PETERSSON, H. 2016. BEAM QUALITY CORRECTION FACTORS FOR KAP
METERS FOR LIGHTLY AND HEAVILY FILTERED X-RAY BEAMS. Radiation Protection
Dosimetry.

HONDIUS BOLDINGH, W. 1961. Quality and Choice of Potter Bucky Grids Parts IV and V. Acta
Radiologica, 55, 225-235.

HUBBELL, J. H. 2006. Review and history of photon cross section calculations. Physics in Medicine and
Biology, 51, R245-R262.

193
HUBBELL, J. H. & SELTZER, S. M. 1995. Tables of X-ray mass attenuation coefficients and mass energy-
absorption coefficients 1 keV to 20 MeV for elements Z= 1 to 92 and 48 additional substances of
dosimetric interest. National Inst. of Standards and Technology-PL, Gaithersburg, MD (United
States). Ionizing Radiation Div.

HUBBELL, J. H., VEIGELE, W. J., BRIGGS, E. A., BROWN, R. T., CROMER, D. T. & HOWERTON,
R. J. 1975. Atomic form factors, incoherent scattering functions, and photon scattering cross
sections. Journal of physical and chemical reference data, 4, 471-538.

HUDA, W., SAJEWICZ, A. M., OGDEN, K. M. & DANCE, D. R. 2003. Experimental investigation of
the dose and image quality characteristics of a digital mammography imaging system. Medical
Physics, 30, 442-8.

ICRU 1989. Tissue substitutes in radiation dosimetry and measurement. Report 44 of the International
Commission on Radiation Units and Measurements (Bethesda, MD).

INTERNATIONAL ATOMIC ENERGY AGENCY 2014. Quantitative Nuclear Medicine Imaging:


Concepts, Requirements And Methods. IAEA human health reports No. 9. Vienna: International
Atomic Energy Agency.

INTERNATIONAL COMMISSION ON RADIATION PROTECTION 2007a. Annex A. Biological and


Epidemiological Information on Health Risks Attributable to Ionising Radiation: A Summary of
Judgements for the Purpose of Radiological Protection of Humans. Annals of the ICRP, 37, 137-
246.

INTERNATIONAL COMMISSION ON RADIATION PROTECTION 2007b. Annex B and All references.


Annals of the ICRP, 37, 247-332.

INTERNATIONAL COMMISSION ON RADIATION UNITS AND MEASUREMENTS 1989. Tissue


Substitutes in Radiation Dosimetry and Measurement. ICRU Report 44.

INTERNATIONAL COMMISSION ON RADIATION UNITS AND MEASUREMENTS, ICRU 2011.


FUNDAMENTAL QUANTITIES AND UNITS FOR IONIZING RADIATION (Revised).
Journal of the American College of Radiology. Oxford University Press.

INTERNATIONAL COMMISSION ON RADIOLOGICAL PROTECTION 2007a. The 2007


Recommendations of the International Commission on Radiological Protection, ICRP Publication
103. Annals of the ICRP, 37.

INTERNATIONAL COMMISSION ON RADIOLOGICAL PROTECTION 2007b. Annex A. Biological


and Epidemiological Information on Health Risks Attributable to Ionising Radiation: A Summary
of Judgements for the Purpose of Radiological Protection of Humans. Annals of the ICRP.

INTERNATIONAL ELECTROTECHNICAL COMMISSION 1978. Characteristics of antiscatter grids


used in X-ray equipment: Characteristics of general purpose and mammographic antiscatter grids.
Geneva: IEC.

194
INTERNATIONAL ELECTROTECHNICAL COMMISSION 2001. Medical Diagnostic X-ray equipment:
Characteristics of general purpose and mammographic antis-catter grids. Geneva: IEC.

INTERNATIONAL ELECTROTECHNICAL COMMISSION 2005. Radiation conditions for use in the


determination of characteristics. Gen eve, IEC 61267.

INTERNATIONAL ELECTROTECHNICAL COMMISSION 2013. Diagnostic X-ray imaging equipment:


characteristics of general purpose and mammographic anti-scatter grids. 3.0. ed. Geneva: IEC.

JAFFE, C. & WEBSTER, E. W. 1975. Radiographic contrast improvement by means of slit radiography.
Radiology, 116, 631-5.

JANG, D. Y., KIM, D. I., PARK, C. J., CHOE, B. Y., SUH, T. S., LEE, H. K., JUNG, N. K., KIM, J. S.,
SEO, S. W. & SEONG, W. 2008. Performance evaluation of carbon-interspaced anti-scatter grids
in mammography: Empirical formula and Monte Carlo simulation studies. Journal of the Korean
Physical Society, 53, 863-867.

JING, Z., HUDA, W. & WALKER, J. K. 1998. Scattered radiation in scanning slot mammography. Medical
Physics, 25, 1111-7.

JOHNS, P. C. & YAFFE, M. 1982. Scattered radiation in fan beam imaging systems. Medical Physics, 9,
231-9.

KADHIM, M., SALOMAA, S., WRIGHT, E., HILDEBRANDT, G., BELYAKOV, O. V., PRISE, K. M.
& LITTLE, M. P. 2013. Non-targeted effects of ionising radiation—Implications for low dose risk.
Mutation Research, 752, 84-98.

KALENDER, W. 1981. Monte Carlo calculations of x-ray scatter data for diagnostic radiology. Physics in
medicine and biology, 26, 835-849.

KALENDER, W. A. 1982. Calculation of x-ray grid characteristics by Monte Carlo methods. Physics in
medicine and biology, 27, 353-61.

KASAP, S., FREY, J. B., BELEV, G., TOUSIGNANT, O., MANI, H., GREENSPAN, J., LAPERRIERE,
L., BUBON, O., REZNIK, A. & DECRESCENZO, G. 2011. Amorphous and polycrystalline
photoconductors for direct conversion flat panel x-ray image sensors. Sensors, 11, 5112-5157.

KAWRAKOW, I. 2000. Accurate condensed history Monte Carlo simulation of electron transport. II.
Application to ion chamber response simulations. Medical physics, 27, 499-513.

KHODAJOU-CHOKAMI, H. & SOHRABPOUR, M. 2015. Design of linear anti-scatter grid geometry


with optimum performance for screen-film and digital mammography systems. Physics in Medicine
and Biology, 60, 5753.

195
KHONG, P. L., RINGERTZ, H., DONOGHUE, V., FRUSH, D., REHANI, M., APPELGATE, K. &
SANCHEZ, R. 2013. ICRP PUBLICATION 121: Radiological Protection in Paediatric Diagnostic
and Interventional Radiology. Annals of the ICRP.

KIM, K. Y., OH, J. E., LEE, S. Y., CHOI, S. I., CHO, H. S., KIM, J. S., CHUNG, N. G., KIM, J. W. &
KIM, J. G. 2007. Performance evaluate of carbon-interspaced and aluminum-interspaced
antiscatter grids based upon the IEC standard fixture. Nuclear Science Symposium Conference
Record, 2007. NSS '07. IEEE.

KIM, S., YOSHIZUMI, T. T., TONCHEVA, G., FRUSH, D. P. & YIN, F.-F. 2010. Estimation of absorbed
doses from paediatric cone-beam CT scans: MOSFET measurements and Monte Carlo simulations.
Radiation Protection Dosimetry, 138, 257-263.

KOEDOODER, K. & VENEMA, H. W. 1986. Filter materials for dose reduction in screen-film
radiography. Physics in medicine and biology, 31, 585-600.

KROL, A., BASSANO, D. A., CHAMBERLAIN, C. C. & PRASAD, S. C. 1996. Scatter reduction in
mammography with air gap. Medical Physics, 23, 1263-70.

KYRIAKOU, Y. & KALENDER, W. 2007. Efficiency of antiscatter grids for flat-detector CT. Physics in
medicine and biology, 52, 6275-93.

LANDAU, D. P. & BINDER, K. 2009. A guide to Monte Carlo simulations in statistical physics,
Cambridge university press.

LAZOS, D. & WILLIAMSON, J. F. 2010. Monte Carlo evaluation of scatter mitigation strategies in cone-
beam CT. Medical physics, 37, 5456-5470.

LITTLE, M. P. 2003. Risks associated with ionizing radiation. British Medical Bulletin, 68, 259-75.

LITTLE, M. P. 2010. Do non-targeted effects increase or decrease low dose risk in relation to the linear-
non-threshold (LNT) model? Mutation Research, 687, 17-27.

LITTLE, M. P., HOEL, D. G., MOLITOR, J., BOICE, J. D., WAKEFORD, R. & MUIRHEAD, C. R.
2008a. New Models for Evaluation of Radiation-Induced Lifetime Cancer Risk and its Uncertainty
Employed in the UNSCEAR 2006 Report. Radiation Research, 169, 660-676.

LITTLE, M. P., TAWN, E. J., TZOULAKI, I., WAKEFORD, R., HILDEBRANDT, G., PARIS, F., TAPIO,
S. & ELLIOTT, P. 2008b. A systematic review of epidemiological associations between low and
moderate doses of ionizing radiation and late cardiovascular effects, and their possible mechanisms.
Radiation Research, 169, 99-109.

LITTLE, M. P., WAKEFORD, R. & KENDALL, G. M. 2009a. Updated estimates of the proportion of
childhood leukaemia incidence in Great Britain that may be caused by natural background ionising
radiation. Journal of Radiological Protection, 29, 467.

196
LITTLE, M. P., WAKEFORD, R., TAWN, E. J., BOUFFLER, S. D. & DE GONZALEZ, A. B. 2009b.
Risks associated with low doses and low dose rates of ionizing radiation: Why linearity may be
(almost) the best we can do. Radiology, 251, 6-12.

M. MATSUMOTO & NISHIMURA, T. 1997. Mersenne Twister Home Page [Online].


http://www.math.sci.hiroshima-u.ac.jp/~m-mat/MT/emt.html. [Accessed 10 Sep 2016].

MALUSEK, A., SANDBORG, M. & CARLSSON, G. A. Simulation of scatter in cone beam CT: Effects
on projection image quality. In: YAFFE, M. J. & ANTONUK, L. E., eds. Procedings of SPIE Vol
5030, 2003. Physics of Medical Imaging, 740-751.

MALUSEK, A., SANDBORG, M. & CARLSSON, G. A. 2008. CTmod—A toolkit for Monte Carlo
simulation of projections including scatter in computed tomography. Computer methods and
programs in biomedicine, 90, 167-178.

MARANDI, A., LEINDECKER, N. C., VODOPYANOV, K. L. & BYER, R. L. 2012. All-optical quantum
random bit generation from intrinsically binary phase of parametric oscillators. Optics express, 20,
19322-19330.

MENTRUP, D., JOCKEL, S., MENSER, B., NEITZEL, U., D. E. HAMBURG & EINDHOVEN, N. L.
2014. Grid-like contrast restoration for non-grid chest radiographs by software-based scatter
correction.

MORGAN, R. H. 1946a. An analysis of the physical factors controlling the diagnostic quality of roentgen
images: Part IV. Contrast and the film contrast factor. The American journal of roentgenology &
radium therapy, 55, 627-33.

MORGAN, R. H. 1946b. An analysis of the physical factors controlling the diagnostic quality of Roentgen
images: Part III. Contrast and the intensity distribution function of a roentgen image. The American
journal of roentgenology & radium therapy, 55, 67-89.

MORGAN, R. H. 1949. An analysis of the physical factors controlling the diagnostic quality of roentgen
images; unsharpness. The American journal of roentgenology & radium therapy, 62, 870-80.

MORIN, L. R. M. 1982. Molecular form factors and photon coherent scattering cross sections of water.
Journal of Physical and Chemical Reference Data, 11, 1091-1098.

MOSSMAN, K. L. 2006. Radiation risks in perspective, CRC Press.

NATIONAL INSTITUTE OF STANDARDS AND TECHNOLOGY 2016. XCOM: Photon Cross Sections
Database. http://www.nist.gov/pml/data/xcom/index.cfm.

NATIONAL RESEARCH COUNCIL 2006. Health Risks from Exposure to Low Levels of Ionizing
Radiation BEIR VII Phase 2, Washington, DC, National Research Council of the National
Academies.

197
NEITZEL, U. 1992. Grids or air gaps for scatter reduction in digital radiography: a model calculation.
Medical Physics, 19, 475-481.

NIKLASON, L. T., SORENSON, J. A. & NELSON, J. A. 1981. Scattered Radiation in Chest Radiography.
Medical Physics, 8, 677-681.

NIKOLOPOULOS, D., LINARDATOS, D., VALAIS, I., MICHAIL, C., DAVID, S., GONIAS, P.,
BERTSEKAS, N., CAVOURAS, D., LOUIZI, A. & KANDARAKIS, I. 2007. Monte Carlo
validation in the diagnostic radiology range. Nuclear Instruments and Methods in Physics Research
Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 571, 267-269.

OLEARY, D. D., GRANT, M. T. & RAINFORD, D. L. 2011. Image quality and compression force the
forgotten link in optimisaion of digital mammography. UCD School of Medicine and Medical
Science, Dublin, Ireland: health research board.

PELI, E. 1990. Contrast in complex images. Optical Society of America, 7, 2032-2040.

PERSLIDEN, J. & CARLSSON, G. A. 1997. Scatter rejection by air gaps in diagnostic radiology.
Calculations using a Monte Carlo collision density method and consideration of molecular
interference in coherent scattering. Physics in medicine and biology, 42, 155.

PICANO, E., VANO, E., DOMENICI, L., BOTTAI, M. & THIERRY-CHEF, I. 2012. Cancer and non-
cancer brain and eye effects of chronic low-dose ionizing radiation exposure. BMC Cancer, 12,
157.

PODGORŠAK, E. B. 2006. Rutherford—Bohr Atomic Model. Radiation Physics for Medical Physicists,
42-85.

POLETTI, M., GONÇALVES, O. & MAZZARO, I. 2002. X-ray scattering from human breast tissues and
breast-equivalent materials. Physics in medicine and biology, 47, 47.

POLUDNIOWSKI, G., EVANS, P. & WEBB, S. 2009a. Rayleigh scatter in kilovoltage x-ray imaging: is
the independent atom approximation good enough? Physics in medicine and biology, 54, 6931.

POLUDNIOWSKI, G., LANDRY, G., DEBLOIS, F., EVANS, P. M. & VERHAEGEN, F. 2009b.
SpekCalc: a program to calculate photon spectra from tungsten anode x-ray tubes. Physics in
medicine and biology, 54, N433-8.

PRESS, W. H. & TEUKOLSKY, S. A. 1992. Portable random number generators. Computers in physics,
6, 522-524.

PRESTON, D., RON, E., TOKUOKA, S., FUNAMOTO, S., NISHI, N., SODA, M., MABUCHI, K. &
KODAMA, K. 2007. Solid cancer incidence in atomic bomb survivors: 1958-1998. Radiation
research, 168, 1-64.

198
PRESTON, D. L., KUSUMI, S., TOMONAGA, M., IZUMI, S., RON, E., KURAMOTO, A., KAMADA,
N., DOHY, H., MATSUI, T. & NONAKA, H. 1994. Cancer incidence in atomic bomb survivors.
Part III: Leukemia, lymphoma and multiple myeloma, 1950-1987. Radiation research, 137, S68-
S97.

PRESTON, D. L., SHIMIZU, Y., PIERCE, D. A., SUYAMA, A. & MABUCHI, K. 2003. Studies of
mortality of atomic bomb survivors. Report 13: Solid cancer and noncancer disease mortality:
1950-1997. Radiation research, 160, 381-407.

RENGER, B., BRIESKORN, C., TOTH, V., MENTRUP, D., JOCKEL, S., LOHÖFER, F., SCHWARZ,
M., RUMMENY, E. J. & NOËL, P. B. 2016. EVALUATION OF DOSE REDUCTION
POTENTIALS OF A NOVEL SCATTER CORRECTION SOFTWARE FOR BEDSIDE CHEST
X-RAY IMAGING. Radiation protection dosimetry, 169, 60-67.

RÖNTGEN, W. C. 1896. On a new kind of rays. Science, 227-231.

ROSE, A. 1948. The sensitivity performance of the human eye on an absolute scale. JOSA, 38, 196-208.

ROYLE, G. J. & SPELLER, R. D. 1995. Quantitative x-ray diffraction analysis of bone and marrow
volumes in excised femoral head samples. Physics in medicine and biology, 40, 1487.

SALVAGNINI, E., BOSMANS, H., STRUELENS, L. & MARSHALL, N. W. 2012. Quantification of


scattered radiation in projection mammography: Four practical methods compared. Medical
Physics, 39, 3167-3180.

SALVAT, F. PENELOPE-2015, A code system for Monte Carlo simulation of electron and phonon
transport. Workshop proceedings, 2015 Barcelona, Spain. OECD Nuclear Energy Agency.

SAUNDERS JR, R. S. & SAMEI, E. A Monte Carlo investigation on the impact of scattered radiation on
mammographic resolution and noise. Proc. SPIE, Medical Imaging: Physics of Medical Imaging,
2006. 61423A-7A.

SCHULTZ-HECTOR, S. & TROTT, K. R. 2007. Radiation-induced cardiovascular diseases: is the


epidemiologic evidence compatible with the radiobiologic data? International Journal Of
Radiation Biology, 67, 10-8.

SHARMA, S., DURAND, R. M. & GUR-ARIE, O. 1981. Identification and analysis of moderator variables.
Journal of marketing research, 291-300.

SHUPING, R. E. & JUDY, P. F. 1977. Energy absorbed in calcium tungstate x‐ray screens. Medical
Physics, 4, 239-243.

SISNIEGA, A., ZBIJEWSKI, W., BADAL, A., KYPRIANOU, I. S., STAYMAN, J. W., VAQUERO, J. J.
& SIEWERDSEN, J. H. 2013. Monte Carlo study of the effects of system geometry and antiscatter
grids on cone-beam CT scatter distributions. Medical Physics, 40, 051915.

199
SORENSON, J. A. & FLOCH, J. 1985. Scatter rejection by air gaps: an empirical model. Medical Physics,
12, 308-16.

STAR-LACK, J., SUN, M., KAESTNER, A., HASSANEIN, R., VIRSHUP, G., BERKUS, T. &
OELHAFEN, M. Efficient scatter correction using asymmetric kernels. SPIE Medical Imaging,
2009. International Society for Optics and Photonics, 72581Z-72581Z-12.

STRID, K. G. 1976. Analysis of secondary screening with special reference to grids for abdominal
radiography. Acta Radiologica, 351, 1-113.

STURM, R. E. & MORGAN, R. H. 1949. Screen intensification systems and their limitations. American
Journal of Roentgenology and Radium Therapy, 62, 617-634.

SUN, M. & STAR-LACK, J. M. 2010. Improved scatter correction using adaptive scatter kernel
superposition. Physics in medicine and biology, 55, 6695-6720.

TAKAHIRO KAWAMURA, SATOSHI NAITO, KAYO OKANO & YAMADA, M. 2015. Improvement
in Image Quality and Workflow of X-Ray Examinations Using a New Image Processing Method,
"Virtual Grid Technology". FUJIFILM RESEARCH & DEVELOPMENT (No. 60-2015).

TANAKA, N., NAKA, K., SAITO, A., MORISHITA, J., TOYOFUKU, F., OHKI, M. & HIGASHIDA, Y.
2013. Investigation of optimum anti-scatter grid selection for digital radiography: physical imaging
properties and detectability of low-contrast signals. Radiological physics and technology, 6, 54-60.

TARTARI, A., TAIBI, A., BONIFAZZI, C. & BARALDI, C. 2002. Updating of form factor tabulations
for coherent scattering of photons in tissues. Physics in Medicine and Biology, 47, 163.

TER-POGOSSIAN, M. M., PHELPS, M. E., HOFFMAN, E. J. & EICHUNG, J. O. 1974. The Extraction
of the Yet Unused Wealth of Information in Diagnostic Radiology. Radiology, 113, 515-520.

THE INTERNATIONAL COMMISSION ON RADIATION UNITS AND MEASUREMENTS 1989.


Tissue Substitutes in Radiation Dosimetry and Measurement. In: BETTHESDA, M. D., , (ed.)
Report 44 of the International Commission on Radiation Units and Measurements

THE INTERNATIONAL COMMISSION ON RADIOLOGICAL PROTECTION 2012. ICRP Statement


on Tissue Reactions / Early and Late Effects of Radiation in Normal Tissues and Organs –
Threshold Doses for Tissue Reactions in a Radiation Protection Context. ICRP publication 118.
Ann. ICRP, 41(1/2).

TOMAL, A., CUNHA, D. M. & POLETTI, M. E. 2013. Optimal X-Ray Spectra Selection in Digital
Mammography: A Semi-Analytical Study. Nuclear Science, IEEE Transactions on, 60, 728-734.

TORTORA, G. J. & DERRICKSON, B. H. 2008. Principles of anatomy and physiology, John Wiley &
Sons.

200
TRAVIS, L. B., GOSPODAROWICZ, M., CURTIS, R. E., CLARKE, E. A., ANDERSSON, M.,
GLIMELIUS, B., JOENSUU, T., LYNCH, C. F., VAN LEEUWEN, F. E. & HOLOWATY, E.
2002. Lung cancer following chemotherapy and radiotherapy for Hodgkin's disease. Journal of the
National Cancer Institute, 94, 182-192.

TROMANS, C., DIFFEY, J. & BRADY, S. 2010. Investigating the Replacement of the Physical Anti-
scatter Grid with Digital Image Processing. In: MARTÍ, J., OLIVER, A., FREIXENET, J. &
MARTÍ, R. (eds.) Digital Mammography. Springer Berlin Heidelberg.

TUBIANA, M., AURENGO, A., AVERBECK, D. & MASSE, R. 2006. The debate on the use of linear no
threshold for assessing the effects of low doses. Journal of Radiological Protection, 26, 317-24.

ULLMAN, G., SANDBORG, M., DANCE, D. R., HUNT, R. A. & CARLSSON, G. A. 2006. Towards
optimization in digital chest radiography using Monte Carlo modelling. Physics in medicine and
biology, 51, 2729-2743.

WEBBER, R. L., YOUMANS, H. D. & NAGEL, R. N. 1977. A nonlinear model for predicting radiographic
contrast. Oral Surgery, Oral Medicine, Oral Pathology, 43, 798-811.

WILSEY, R. B. 1921. The intensity of scattered x-rays in radiography. Am. J. Roentgenol, 8, 328.

WILSEY, R. B. 1923. The Design of Potter-Bucky Diaphragm Grids. Radiology, 1, 47-49.

YAFFE, M. & ROWLANDS, J. 1997. X-ray detectors for digital radiography. Physics in Medicine and
Biology, 42, 1.

YAN, H., MOU, X., TANG, S., XU, Q. & ZANKL, M. 2010. Projection correlation based view
interpolation for cone beam CT: primary fluence restoration in scatter measurement with a moving
beam stop array. Physics in medicine and biology, 55, 6353-75.

YU, C., LIU, B., O'CONNOR, J. M., DIDIER, C. S. & GLICK, S. J. 2009. Characterization of scatter in
cone-beam CT breast imaging: Comparison of experimental measurements and Monte Carlo
simulation. Medical Physics, 36, 857-869.

ZBIJEWSKI, W., SISNIEGA, A., VAQUERO, J. J., MUHIT, A., PACKARD, N., SENN, R., YANG, D.,
YORKSTON, J., CARRINO, J. A. & SIEWERDSEN, J. H. Dose and scatter characteristics of a
novel cone beam CT system for musculoskeletal extremities. SPIE Medical Imaging, vol 8313,
2012.

ZHAO, W., RISTIC, G. & ROWLANDS, J. 2004. X-ray imaging performance of structured cesium iodide
scintillators. Medical physics, 31, 2594-2605.

ZHOU, A., YIN, Y., WHITE, G. L. & DAVIDSON, R. 2016. A new method for radiation transmission of
anti-scatter grids. Biomedical Physics & Engineering Express, 2, 055011.

201
Appendix
Publications arising from this work are listed below. They can be found from their DOI links.

Zhou, Abel, Yuming Yin, Graeme L. White, and Rob Davidson. 2016. "A new method for radiation
transmission of anti-scatter grids." Biomedical Physics & Engineering Express 2 (5):055011. doi:
https://doi.org/10.1088/2057-1976/2/5/055011.

Zhou, Abel, Graeme L. White, and Rob Davidson. 2018. "Validation of a Monte Carlo code
system for grid evaluation with interference effect on Rayleigh scattering." Physics in Medicine &
Biology 63 (3). doi: https://doi.org/10.1088/1361-6560/aaa44b.

203

Potrebbero piacerti anche