119
Queensland University of Technology School of Physical and Chemical Sciences The Development of Normoxic Polymer Gel Dosimetry using High Resolution MRI Christopher Hurley M.App.Sci.(Med Phys), M.Ed.Admin., Grad.Dip.Ed., B.Eng.(Elec) A thesis submitted at the Queensland University of Technology, in the School of Physical and Chemical Sciences, in fulfillment of the requirements of the Doctor of Philosophy. 2006

The Development of Normoxic Polymer Gel Dosimetry using ...eprints.qut.edu.au/16442/1/Christopher_Hurley_Thesis.pdf · The Development of Normoxic Polymer Gel Dosimetry using High

  • Upload
    lamnhan

  • View
    215

  • Download
    1

Embed Size (px)

Citation preview

Queensland University of Technology

School of Physical and Chemical Sciences

The Development of Normoxic Polymer Gel

Dosimetry using High Resolution MRI

Christopher Hurley M.App.Sci.(Med Phys), M.Ed.Admin., Grad.Dip.Ed., B.Eng.(Elec)

A thesis submitted at the Queensland University of Technology, in the School of

Physical and Chemical Sciences, in fulfillment of the requirements of the Doctor

of Philosophy.

2006

ii

Keywords

Polymer gel dosimetry, radiotherapy, brachytherapy, radiation dosimetry, PAG,

MAGIC, MAGAT, PAGAT, normoxic polymer gel dosimeters, high-resolution

MRI.

iii

Abstract

Dosimetry is a vital component of treatment planning in radiation therapy.

Methods of radiation dosimetry currently include the use of: ionization chambers,

thermoluminescent dosimeters (TLDs), solid-state detectors and radiographic

film. However, these methods are inherently either 1D or 2D and their use

involves the perturbation of the radiation beam. Although the dose distribution

within tissues following radiation therapy treatments can be modeled using

computerized treatment planning systems, a need exists for a dosimeter that can

accurately measure dose distributions directly and produce 3D dose maps. Some

radiation therapy and brachytherapy treatments require mapping the dose

distributions in high-resolution (typically < 1 mm). A dosimetry technique that is

capable of producing high resolution 3D dose maps of the absorbed dose

distribution within tissues is required.

Gel dosimetry is inherently a 3D integrating dosimeter that offers high spatial

resolution, precision and accuracy. Polymer gel dosimetry is founded on the basis

that monomers dissolved in the gel matrix polymerize due to the presence of free

radicals produced by the radiolysis of water molecules. The amount of

polymerization that occurs within a polymer gel dosimeter can be correlated to the

absorbed dose. The gel matrix maintains the spatial integrity of the polymers and

hence a dose distribution can be determined by imaging the irradiated polymer gel

dosimeter using an imaging modality such as MRI, x-ray computed tomography

(CT), ultrasound, optical CT or vibrational spectroscopy. Polymer gel dosimeters,

however, suffer from oxygen contamination. Oxygen inhibits the polymerization

reaction and hence polymer gel dosimeters must be manufactured, irradiated and

scanned in hypoxic environments.

iv

Normoxic polymer gel dosimeters incorporate an anti-oxidant into the formulation

that binds the oxygen present in the gel and allows the dosimeter to be made under

normal atmospheric conditions. The first part of this study was to provide a

comprehensive investigation into various formulations of polymer and normoxic

polymer gel dosimeters. Several parameters were used to characterize and assess

the performance of each formulation of polymer gel dosimeter including: spatial

resolution and stability, temporal stability of the R2-dose response, optimal R2-

dose response for changes in concentration of constituents and the effects of

oxygen infiltration. This work enabled optimal formulations to be determined that

would provide greater dose sensitivity. Further work was done to investigate the

chemical kinetics that take place within normoxic polymer gel dosimeters from

manufacture to post-irradiation. This study explored the functions that each of

the constituent chemicals plays in a polymer gel dosimeter. Although normoxic

polymer gel dosimeters exhibit very similar characteristics to polyacrylamide

polymer gel dosimeters, one important difference between them was found to be a

decrease in R2-dose sensitivity over time in the normoxic polymer gel dosimeter

compared to an increase in the polyacrylamide polymer gel dosimeters.

From an investigation into the function of anti-oxidants in normoxic polymer gel

dosimeters, alternatives were proposed. Several alternative anti-oxidants were

explored in this study that found that whilst some were reasonably effective,

tetrakis (hydroxymethyl) phosphonium chloride (THPC) had the highest reaction

rate. THPC was found not only to be an aggressive scavenger of oxygen, but also

to increase the dose sensitivity of the gel. Hence, a formulation of normoxic

polymer gel dosimeter was proposed, called MAGAT, that comprised:

methacrylic acid, gelatin, hydroquinone and THPC. This formulation was

examined in a similar fashion to the studies of the other formulations of polymer

and normoxic polymer gel dosiemeters. The gel was found to exhibit spatial and

temporal stability and an optimal formulation was proposed based on the R2-dose

response.

Applications such as IVBT require high-resolution dosimetry. Combined with

high-resolution MRI, polymer gel dosimetry has potential as a high-resolution 3D

integrated dosimeter. Thus, the second component of this study was to

v

commission a micro-imaging MR spectrometer for use with normoxic polymer

gel dosimeters and investigate artifacts related to imaging in high-resolutions.

Using high-resolution MRI requires high gradient strengths that, combined with

the Brownian motion of water molecules, was found to produce an attenuation of

the MR signal and hence lead to a variation in the measured R2. The variation in

measured R2 was found to be dependent on both the timing and amplitude of

pulses in the pulse sequence used during scanning. Software was designed and

coded that could accurately determine the amount of variation in measured R2

based on the pulse sequence applied to a phantom. Using this software, it is

possible to correct for differences between scans using different imaging

parameters or pulse sequences.

A normoxic polymer gel dosimeter was irradiated using typical brachytherapy

delivery and the resulting dose distributions compared with dose points predicted

by the computerized treatment planning system.The R2-dose response was

determined and used to convert the R2 maps of the phantoms to dose maps. The

phantoms and calibration vials were imaged with an in-plane resolution of 0.1055

mm/pixel and a slice thickness of 2 mm. With such a relatively large slice

thickness compared to the in-plane resolution, partial volume effects were

significant, especially in the region immediately adjacent the source where high

dose gradients typically exist. Estimates of the partial volume effects at various

distances within the phantom were determined using a mathematical model based

on dose points from the treatment planning system. The normalized and adjusted

dose profiles showed very good agreement with the dose points predicted by the

treatment planning system.

vi

Table of Contents

ABSTRACT................................................................................................................................................. III TABLE OF CONTENTS................................................................................................................................. VI LIST OF PUBLICATIONS:............................................................................................................................. IX LIST OF ABBREVIATIONS: ........................................................................................................................... X STATEMENT OF ORIGINAL AUTHORSHIP:................................................................................................... XI ACKNOWLEDGEMENTS ............................................................................................................................. XII CHAPTER 1 INTRODUCTION .........................................................................................................................1

1.1 Description of Research Problem Investigated..................................................................1 1.2 Overall Objective of the Study ...........................................................................................3 1.3 The Specific Aims of the Study ...........................................................................................3 1.4 Account of Scientific Progress Linking the Scientific Papers ............................................4 References................................................................................................................................6

CHAPTER 2 LITERATURE REVIEW................................................................................................................9

2.1 Radiotherapy and Brachytherapy ......................................................................................9 2.2 Radiation Dosimetry ........................................................................................................12

2.2.1 Clinical Dosimetry Requirements ...........................................................................................12 2.2.2 Dosimetry Detectors................................................................................................................13

2.2.2.1 Ionization Chambers .......................................................................................................13 2.2.2.2 Solid-State Detectors.......................................................................................................14 2.2.2.3 Thermoluminescent Dosimeters ......................................................................................14 2.2.2.4 Radiographic Film ..........................................................................................................15 2.2.2.5 Chemical Dosimeters ......................................................................................................15

2.3 Gel Dosimeters ................................................................................................................16 2.3.1 Ferrous Sulfate (Fricke) Gels ..................................................................................................17 2.3.2 Polymer Gel Dosimeters .........................................................................................................19 2.3.3 Normoxic Polymer Gel Dosimeters.........................................................................................23

2.4 Characteristics of Polymer Gel Dosimeters.....................................................................24 2.4.1 Effects of Oxygen....................................................................................................................24 2.4.2 Effect of Light .........................................................................................................................24 2.4.3 Temperature ............................................................................................................................25 2.4.4 Concentration of monomers ....................................................................................................26 2.4.5 Ageing of the gel .....................................................................................................................26

2.5 Evaluation of Polymer Gel Dosimeters............................................................................27 2.5.1 Magnetic Resonance Imaging .................................................................................................27 2.5.2 X-Ray Computed Tomography ...............................................................................................29 2.5.3 Optical Computed Tomography ..............................................................................................31 2.5.4 Ultrasound ...............................................................................................................................32 2.5.5 Vibrational Spectroscopy ........................................................................................................33

2.6 High-Resolution MRI in Polymer Gel Dosimetry ............................................................34 2.7 Brachytherapy Applications of Gel Dosimetry ................................................................37 2.8 Sources of Uncertainty in Polymer Gel Dosimeters ........................................................40 2.9 Conclusion .......................................................................................................................42 References..............................................................................................................................44

CHAPTER 3 DOSE-RESPONSE STABILITY AND INTEGRITY OF THE DOSE DISTRIBUTION OF VARIOUS POLYMER GEL DOSIMETERS ......................................................................................................63

Abstract..................................................................................................................................64 3.1 Introduction .....................................................................................................................64 3.2 Materials and methods.....................................................................................................65

vii

3.2.1 Gel fabrication......................................................................................................... 65 3.2.2 Irradiation................................................................................................................ 66 3.2.3 Scanning................................................................................................................... 66

3.3 Results.............................................................................................................................. 67 3.3.1 R2-Dose stability...................................................................................................... 67 3.3.2 Integrity of dose distribution.................................................................................... 68

3.4 Discussion........................................................................................................................ 71 3.4.1 R2-Dose stability...................................................................................................... 71 3.4.2 Integrity of the dose distribution.............................................................................. 73

3.5 Conclusions ..................................................................................................................... 74 References.............................................................................................................................. 75

CHAPTER 4 A BASIC STUDY OF SOME NORMOXIC POLYMER GEL DOSIMETERS .......................................77

Abstract.................................................................................................................................. 78 4.1 Introduction ..................................................................................................................... 79 4.2 Materials and Methods .................................................................................................... 79

4.2.1 MAGIC gel components ........................................................................................... 80 4.2.2 Dose distribution of half-blocked field..................................................................... 81 4.2.3 Anti-oxidants ............................................................................................................ 82 4.2.4 Potentiometric oxygen measurements...................................................................... 82

4.3 Results.............................................................................................................................. 82 4.3.1 MAGIC gel components ........................................................................................... 82 4.3.2 Dose distribution of half-blocked field..................................................................... 86 4.3.3 Anti-oxidants ............................................................................................................ 87 4.3.4 Potentiometric oxygen measurements...................................................................... 89

4.4 Discussion........................................................................................................................ 90 4.4.1 Acrylic polymer gels ................................................................................................ 92 4.4.2 Normoxic polymer gels ............................................................................................ 93

4.4 Conclusions ..................................................................................................................... 98 References.............................................................................................................................. 99

CHAPTER 5 THE EFFECTS OF MOLECULE SELF-DIFFUSION OF WATER ON QUANTITATIVE MRI MEASUREMENTS IN HIGH-RESOLUTION POLYMER GEL DOSIMETRY ......................................................101

Abstract................................................................................................................................ 102 5.1 Introduction ................................................................................................................... 103 5.2 Theory............................................................................................................................ 104 5.3 Methods ......................................................................................................................... 106

5.3.1 Sample preparation................................................................................................ 106 5.3.2 R2 measurements ................................................................................................... 107 5.3.3 Self-diffusion coefficient measurements................................................................. 107 5.3.4 Computer simulations of the diffusion effect on R2 ............................................... 107 5.3.5 Computer simulations of the diffusion effect on phase encoding ........................... 108

5.4 Results............................................................................................................................ 108 5.4.1 R2 measurements ................................................................................................... 108 5.4.2 Computer simulations of the diffusion effect on R2 ............................................... 111 5.4.3 Computer simulations of the diffusion effect on phase encoding ........................... 112

5.5 Discussion...................................................................................................................... 115 5.6 Conclusions ................................................................................................................... 116 References............................................................................................................................ 116

CHAPTER 6 A STUDY OF A NORMOXIC POLYMER GEL DOSIMETER COMPRISING METHACRYLIC ACID, GELATIN, AND TETRAKIS (HYDROXYMETHYL) PHOSPHONIUM CHLORIDE (MAGAT)...................119

Abstract................................................................................................................................ 120 6.1 Introduction ................................................................................................................... 120 6.2 Materials and Methods .................................................................................................. 122

6.2.1 Formulation ........................................................................................................... 122 6.2.1.1 Investigation of Concentration of THPC and HQ ........................................... 122 6.2.1.2 Investigation of Concentration of Gelatin and MAA....................................... 122

6.2.2 R2-Dose Response ................................................................................................. 123 6.2.3 Spatial Stability...................................................................................................... 123

viii

6.2.4 Scanning and Processing .......................................................................................123 6.3 Results and Discussion ..................................................................................................123

6.3.1 Formulation ...........................................................................................................123 6.3.2 Concentrations of THPC and HQ ..........................................................................123 6.3.3 Concentrations of Gelatin and MAA......................................................................124 6.3.4 R2-Dose Response..................................................................................................125 6.3.5 Spatial Stability ......................................................................................................129

6.4 Conclusions....................................................................................................................132 References............................................................................................................................132

CHAPTER 7 HIGH-RESOLUTION GEL DOSIMETRY OF A HDR BRACHYTHERAPY SOURCE USING NORMOXIC POLYMER GELS ....................................................................................................................135

Abstract................................................................................................................................136 7.1 Introduction ...................................................................................................................136 7.2 Materials and Methods ..................................................................................................138 7.3 Results and Discussion ..................................................................................................140

7.3.1 Calibration .............................................................................................................140 7.3.2 Dose maps ..............................................................................................................142 7.3.3 Dose profiles ..........................................................................................................143 7.3.4 Agreement between dose maps and treatment plans ..............................................143

7.4 Conclusion .....................................................................................................................144 References............................................................................................................................145

CHAPTER 8 GENERAL DISCUSSION ..........................................................................................................147

8.1 The Principal Significance of the Findings....................................................................149 8.1.1 Analyzing and Optimizing Polymer Gel Dosimeter Formulations........................................149 8.1.2 Chemical Properties of Normoxic Polymer Gel Dosimeters .................................................150 8.1.3 A Normoxic Polymer Gel Dosimeter Using THPC...............................................................152 8.1.4 Evaluating Polymer Gel Dosimeters using High-Resolution MRI ........................................153 8.1.5 Application of Normoxic Polymer Gel Dosimeters using High-Resolution MRI to Brachytherapy Treatment Plans .....................................................................................................154

8.2 Conclusions and Future Work .......................................................................................155 References............................................................................................................................159

APPENDIX A ............................................................................................................................................161 LISTING OF THE CODE FOR DETERMINING VARIATIONS IN R2 DUE TO THE APPLICATION OF PULSE SEQUENCES DURING HIGH-RESOLUTION MRI

ix

List of Publications:

1. De Deene, Y., Venning, A., Hurley, C., Healy, B. J., Baldock, C., Dose-

response stability and integrity of the dose distribution of various polymer

gel dosimeters, Phys Med Biol, 2002. 47(14) 2459-2470.

2. De Deene, Y., Hurley, C., Venning, A., Vergote, K., Mather, M., Healy,

B.J., Baldock, C., A basic study of some normoxic polymer gel

dosimeters. Phys Med Biol, 2002. 47, 3441-63.

3. Hurley, C., De Deene, Y., Meder, R., Pope, J.M., Baldock, C., The effects

of molecular self-diffusion of water on quantitative MRI measurements in

high-resolution polymer gel dosimetry, Phys Med Biol, 2003. 48: 3043–

3058.

4. Hurley, C., Venning, A. and Baldock, C., A Study of a Normoxic Polymer

Gel Dosimeter comprising Methacrylic Acid, Gelatin and Tetrakis

(Hydroxymethyl) Phosphonium Chloride (MAGAT), App Radiat and Iso,

2005. 63: 443 - 456.

5. Hurley, C., McLucas, C., Pedrazzini, G., and Baldock, C., High-

Resolution Gel Dosimetry of a HDR Brachytherapy Source Using

Normoxic Polymer Gels: Preliminary Study, Nucl Instr Meth Phys Res A,

2006. 565: 793 – 803 (In Press).

x

List of Abbreviations: AAPM American Association of Physicists in Medicine ABAGIC Acrylamide, methylene-Bis-acrylamide, Ascorbic acid, Gelatin,

Hydroquinone, and copper(II) sulphate gel CT Computed Tomography HDR High Dose Rate HEA 2-Hydroxyethyl Acrylate gel ICRU International Commission on Radiation Units IMRT Intensity Modulated Radiotherapy IVBT Intravascular Brachytherapy LDR Low Dose Rate MAGAS Methacrylic Acid, Gelatin, AScorbic acid MAGAT Methacrylic Acid, Gelatin, Ascorbic Acid and THPC MAGIC Methacrylic Acid, Gelatin, Initiated by Copper Sulphate (the first

normoxic polymer gel proposed by Fong et al (2001). MRI Magnetic Resonance Imaging PAG PolyAcrylamide Gelatin gel PAGAS PolyAcrylamide, Gelatin, AScorbic acid PDR Pulsed Dose Rate R1 MRI Longitudinal Relaxation Rate (measured in s-1) R2 MRI Transverse Relaxation Rate (measured in s-1) T1 MRI Longitudinal Relaxation Time (measured in s) T2 MRI Transverse Relaxation Time (measured in s) TE MRI Echo Time ΔTE MRI Inter-Echo Time THPC Tetrakis (Hydroxymethyl) Phosphonium Chloride TLD Thermoluminescent Dosimeter TR MRI Relaxation Time

xi

Statement of Original Authorship:

The work contained in this thesis has not been previously submitted to meet

requirements for an award at this or any other higher education institution. To the

best of my knowledge and belief, the thesis contains no material previously

published or written by another person except where due reference is made.

Christopher A. Hurley

17th August, 2006

xii

Acknowledgements I would like to sincerely thank the following persons for their invaluable

contributions to the project:

Clive Baldock, my supervisor, for inspiring me to excellence, providing

invaluable advice and for the countless hours of support in helping me to see the

bigger picture in research. I am deeply and sincerely grateful for your endless

enthusiasm and perseverance as you challenged me to achieve ever greater heights

in this project!

Yves De Deene for the incredible knowledge and experiences that you shared

with me, helping me to understand and appreciate the complex world of MRI.

Your ability to open my mind to understanding the world of MRI and Physics is

extraordinary.

Jim Pope for your exceptional ability to understand exactly what was going on

when I was exploring the field of high-resolution MRI. Your advice was always

exactly what I needed.

Thanks to Cameron McLucas and Greg Pedrazzini for your patient work in the

irradiation of the phantoms and your expert technical advice in brachytherapy and

dosimetry. Thanks to Southern X-Ray Clinic of the Wesley Hospital for the use

of MRI scanner, brachytherapy afterloader and linear accelerators and the Wesley

Research Institute. Your commitment to research in medicine is highly

commendable.

Brian Thomas and Elizabeth Stein for your assistance and guidance throughout

the course. Your willingness to sit and talk at any time, encouraged me to see the

course through to the end.

1

Chapter 1

Introduction

1.1 Description of Research Problem Investigated

The use of radiation to supply a lethal dose to tissue affected by disease has been a

practice used for many years. A common part of most major hospitals in the

world, the radiation therapy (or radiation oncology) department has formed an

integral component of medicine’s fight against cancer. The aim of radiation

therapy has been to deliver a dose of ionizing radiation to a tumour or lesion,

whilst minimizing the dose that may be delivered to the surrounding healthy

tissue. Radiation therapy has advanced significantly over the past decades and

now offers a wide range of techniques that can conform a radiation beam to

maximize dose to a targeted tissue whilst minimizing dose to surrounding tissues.

Conformal radiotherapy and intensity-modulated radiotherapy (IMRT) have made

considerable advances in shaping the doses in three dimensions (3D) delivered to

tissues. In a similar fashion, brachytherapy, which involves a radioactive source

being placed directly inside target tissues within a patient’s body, has enabled the

localization of ionizing radiation to a small volume of tissue, again effectively

minimizing the dose being delivered to healthy tissues that may surround the

tumour or lesion. With these advances in dose delivery, the target volumes can

now incorporate complex geometries with high dose gradients.

Although the amount of radiation delivered to the patient’s body can been easily

determined through well-known and derived equations, it is more difficult to

accurate determine the distribution of absorbed dose within the patient’s body.

Absorbed dose distributions are typically determined using computerized

2

treatment planning software that is based on radiation models and simulations of

the absorbed dose. Direct measurements of the absorbed dose have traditionally

been determined using ionization chambers, thermoluminescent dosimeters

(TLDs), solid-state detectors and radiographic film. However, ionization

chambers, TLDs and solid-state detectors usually exhibit poor spatial resolution

(due to the size of the measuring device), radiographic film is inherently 2D and

all these detectors perturb the radiation beam. Gel dosimetry provides a method

by which the spatial, 3D distribution of the absorbed dose can be determined

[1,2].

Currently there are several different formulations for gel dosimeters under

investigation. Each comprises an aqueous gel matrix to provide spatial stability,

cross-linker monomers that polymerize when irradiated and other constituents that

function to maintain the chemical stability or improve the performance of the gel

as a dosimeter [3-6]. Polyacrylamide polymer gel dosimeters suffer from the

effects of oxygen infiltration which prevents the polymerization process that can

be correlated to absorbed dose. Normoxic polymer gel dosimeters, that use an

anti-oxidant to bind free oxygen, have recently been proposed but require further

investigation and development before use in clinical practice.

Following irradiation, phantoms of polymer gel dosimeter are evaluated using

imaging modalities such as MRI [2], x-ray computed tomography (CT) [7],

optical CT [8] and ultrasound [9]. To date, MRI has been the most frequently

used scanning technique for gel dosimetry. Using high-resolution MRI to

evaluate polymer gel dosimeters has the potential to provide dose distributions

with high spatial resolutions in the order of sub-millimeters (~ 100 microns).

However, the evaluation of polymer gel dosimeters using high-resolution MRI is

only in its infancy and there is still much work to be done. To date, high-

resolution MRI has not been used to evaluate normoxic polymer gel dosimeters.

3

1.2 Overall Objective of the Study

The objective of this study was to further develop normoxic polymer gel

dosimetry using high-resolution magnetic resonance imaging and assess its

feasibility in the verification of high-resolution treatment plans, such as those used

in intravascular brachytherapy.

1.3 The Specific Aims of the Study

The specific aims of this study included:

• Exploration of different formulations of polymer gel dosimeters and

normoxic polymer gel dosimeters in order to obtain normoxic polymer gel

dosimeters with optimal characteristics for use in gel dosimetry.

• Measurement of the effects of physical and chemical properties of various

formulations of normoxic polymer gel dosimeters on dose maps obtained

using gel dosimetry with high-resolution MRI to assess their performance

for use in radiotherapy dosimetry.

• Investigation of high-resolution magnetic resonance imaging, to achieve

in-plane spatial resolutions ~ 100 microns, for its potential use in

evaluating polymer gel dosimeters accurately and efficiently.

• Examination of the effects of molecular diffusion on images produced

using high-resolution MRI in order to eliminate the effects causing errors

in the calculated dose maps of polymer gel dosimeters.

• Application and assessment of normoxic polymer gel dosimeters

irradiated using typical brachytherapy deliveries and evaluated using high-

resolution MRI.

4

1.4 Account of Scientific Progress Linking the Scientific

Papers

Normoxic polymer gel dosimeters present a significant advance in gel dosimetry

and show good potential to the verification of radiation therapy and brachytherapy

dose distributions in 3D. However, their development is still in its infancy and

normoxic polymer gel dosimeters have yet to be incorporated into clinical

practice. Using high-resolution MRI provides the potential for extending the

applications of gel dosimetry to include radiation therapy and brachytherapy

treatments that require verification with resolutions at the sub-millimeter level

(typically ~ 100 microns). This study has been broken into two aspects: the

analysis and development of normoxic polymer gel dosimeters and secondly, the

investigation of the use of high-resolution MRI to evaluate normoxic polymer gel

dosimeters.

Significant changes of the polymer structure are known to occur in polymer gel

dosimeters following irradiation [10]. These changes ultimately affect the

chemical and physical properties of a gel and hence its suitability for use in gel

dosimetry. Chapter 3 investigates some different formulations of polymer gel

dosimeter, including polyacrylamide gel (PAG), polymer gel dosimeters made

with 2-hydroxyethyl acrylate (HEA), and normoxic polymer gel dosimeters

including the MAGIC gel (methacrylic and ascorbic acid in gelatin initiated by

copper) [11]. The effects of varying the concentrations of the constituent

components of the gel on the R2-dose response was explored to produce an

optimal formulation. The temporal stability of the R2-dose response was also

examined by relating changes in the R2-intercept and the R2-dose sensitivity

(slope) to reactions within the gel. The spatial stability was investigated using

dose profiles through a phantom exposed to a half-blocked field.

Chapter 4 further investigates normoxic polymer gel dosimeters through a

chemical analysis of the MAGIC gel. The role of the different chemicals and

reactions kinetics are explored as they affect the R2-dose response and spatial and

5

temporal stability of the MAGIC gel. A comprehensive investigation of the

chemical reactions that take place within a normoxic polymer gel dosimeter is

presented in order to explain the role that each constituent component plays in the

overall gel. An understanding of these reactions assists in the development of

more optimal formulations. In addition, alternative anti-oxidants to ascorbic acid

are proposed and investigated as to their effectiveness for oxygen scavenging in

normoxic polymer gel dosimeters.

High-resolution MRI requires high gradient strengths that can significantly vary

the R2 values obtained due to molecular diffusion of water molecules within the

gel itself. The degree of variation is affected by the imaging parameters chosen

by an operator. Chapter 5 investigates the extent to which the imaging parameters

alter the R2 values and techniques that can be incorporated to provide an accurate

dose map using high-resolution MRI in polymer gel dosimetry. Software is

developed that can predict the variation in R2 due to the application of MRI pulse

sequences that are used.

Chapter 6 investigates a normoxic polymer gel dosimeter comprising tetrakis

(hydroxymethyl) phosphonium chloride (THPC) as an alternative anti-oxidant to

the ascorbic acid that is used in MAGIC polymer gel dosimeters. This gel,

composed of methacrylic acid, gelatin, and THPC, was called MAGAT. This

chapter evaluates the R2-dose response, the temporal stability of the R2-dose

response, the spatial stability and provides an optimal formulation for the

MAGAT polymer gel dosimeter using high-resolution MRI.

Finally, chapter 7 examines the application of a normoxic polymer gel dosimeter

using high-resolution MRI to typical brachytherapy deliveries. Using a line and a

point irradiation pattern as a plan, dose distribution maps were produced and

compared with dose points predicted by the treatment planning system. Adjusting

for partial volume effects, the dose profiles show good agreement to dose points

predicted by the computerized treatment planning system.

6

References

[1] Maryanski, M.J., Gore, J.C., Kennan, R.P. and Schulz, R.J., NMR relaxation

enhancement in gels polymerized and cross-linked by ionizing radiation: a

new approach to 3D dosimetry by MRI. Magn Reson Imaging, 1993. 11(2)

253-8.

[2] Maryanski, M.J., Gore, J.C. and Schulz, R.J., US Patent: Three-dimensional

detection, dosimetry and imaging of an energy field by formation of a

polymer in a gel. 1994, Patent Number 5,321,357. United States.

[3] Baldock, C., Burford, R.P., Billingham, N., Cohen, D. and Keevil, S.F.,

Polymer gel composition in MRI dosimetry. Med Phys, 1996. 23 1070.

[4] De Deene, Y., Hanselaer, P., De Wagter, C., Achten, E. and De Neve, W.,

An investigation of the chemical stability of a monomer/polymer gel

dosimeter. Phys Med Biol, 2000. 45(4) 859-78.

[5] Lepage, M., Whittaker, A.K., Rintoul, L., Back, S.A. and Baldock, C.,

Modelling of post-irradiation events in polymer gel dosimeters. Phys Med

Biol, 2001. 46(11) 2827-39.

[6] Lepage, M., Whittaker, A.K., Rintoul, L., Back, S.A. and Baldock, C., The

relationship between radiation-induced chemical processes and transverse

relaxation times in polymer gel dosimeters. Phys Med Biol, 2001. 46(4)

1061-74.

[7] Trapp, J.V., Back, S.A., Lepage, M., Michael, G. and Baldock, C., An

experimental study of the dose response of polymer gel dosimeters imaged

with x-ray computed tomography. Phys Med Biol, 2001. 46(11) 2939-51.

[8] Gore, J.C., Ranade, M., Maryanski, M.J. and Schulz, R.J., Radiation dose

distributions in three dimensions from tomographic optical density scanning

of polymer gels: I. Development of an optical scanner. Phys Med Biol,

1996. 41(12) 2695-704.

[9] Mather, M.L., Whittaker, A.K. and Baldock, C., Ultrasound evaluation of

polymer gel dosimeters. Phys Med Biol, 2002. 47(9) 1449-58.

7

[10] Lepage, M., Whittaker, A.K., Rintoul, L. and Baldock, C., 13C-NMR, 1H-

NMR, and FT-Raman study of radiation-induced modifications in radiation

dosimetry polymer gels. J App Poly Sci, 2001. 79 1572-1581.

[11] Fong, P.M., Keil, D.C., Does, M.D., and Gore, J.C., Polymer gels for

magnetic resonance imaging of radiation dose distributions at normal room

atmosphere. Phys Med Biol, 2001. 46(12) 3105-13.

8

9

Chapter 2

Literature Review

2.1 Radiotherapy and Brachytherapy Since the development of radiation therapy and related fields, patients have been

exposed to radiation for the treatment of malignant disease. The approach taken

to achieve this has involved the exposure of the affected tissue area to a radiation

beam of sufficient energy to cause significant damage to the cells in the affected

area without affecting the surrounding normal tissue to any great extent [1].

Exposing tissues affected by cancer to sufficient levels of radiation causes

irreversible cell damage therefore preventing the cancerous cells from further

growth and metastasis [2].

There are currently several methods by which radiation can be delivered to a

targeted area. The most widely used method is external beam irradiation. In this

approach, a linear accelerator (linac) generates megavoltage energy photon or

electron beams that are focused and aimed onto a patient’s body. Using

collimation and rotation, linacs are able to confine the radiation exposure to a

particular region of the patient’s body. Using x-ray computed tomography (CT) it

is possible to obtain scans of a patient enabling greater anatomical delineation.

This information can then be used to ensure that high doses of radiation are

delivered to target volume whilst sparing the surrounding healthy tissues [3].

Magnetic resonance imaging (MRI) and positron emission tomography (PET) can

enhance the process of tumour identification giving a more precise geometric

definition of the tumour. A more precise tumour definition may lead to improved

irradiation of the true extent of the tumour. This process is known as conformal

radiation therapy (CRT). Multi-leaf collimators (MLCs) can now be used to

10

control the shape of the radiation beam used to treat a patient [4]. Current MLCs

typically have between 40 to 120 leaves of varying widths (0.5 to 1 cm) across the

leaf range. These leaves, made of tungsten, can be moved in front of the beam at

specific lengths to define the radiation beam. In this way, treatment exposures can

be made to conform more precisely to target volumes using sophisticated

computer software that manipulates radiotherapy equipment in real time. One key

objective of the clinical implementation of conformal radiotherapy is assuring that

the complex manipulations of the radiotherapy equipment required for the therapy

are actually performed, and that the dose distributions calculated by treatment

planning systems and delivered by treatments are correct [5]. Correctly

measuring the delivered dose is central to the process of assuring the accuracy of

treatment plans and treatment deliveries.

Intensity modulated radiation therapy (IMRT) provides the ability to vary the

radiation fluence within each radiation beam during treatment. IMRT can further

enhance the ability of linac to control the radiation distribution within a targeted

volume. The dose distribution in this case is non-uniform across several radiation

beams, which summate to produce an optimal dose distribution [6]. Combined

with conformal radiation therapy using multi-leaf collimators, treatment plans can

now incorporate complex geometries with non-uniform dose distributions to be

delivered to a patient.

With the increase in treatment complexity, the need for verification of computer

generated treatment plans is most significant. The dose distribution required by a

treatment plan is calculated by complex computer algorithms that model the

radiation system. Monte Carlo models have the potential to predict the dose in

complicated geometries using a variety of different types of radiation (electrons,

photons, scatter from collimators, scatter within the patient body, etc.) as applied

to the internal geometry of the region of the patient body [7]. Despite the

advances in Monte Carlo modeling of radiotherapy deliveries, there is still a need

to verify the absorbed dose at points within the distribution using direct

measurements.

11

Brachytherapy involves the use of radioactive sources that are either placed

directly on the patient’s skin or, more commonly, inserted into the patient and

positioned close to the affected tissues [1]. In this way, high radiation doses can

be delivered directly to the tumour site thereby concentrating dose in regions

requiring treatment and effectively minimizing the irradiation of surrounding

healthy tissues. Brachytherapy treatments can be classified as either high dose

rate (HDR) or low dose rate (LDR) depending on the radioactive source used.

HDR sources must be inserted into the target tissue for a short period of time,

whereas LDR sources must be inserted for substantially longer periods of time to

deliver sufficient dose to the targeted tissues. New brachytherapy source designs

are often commissioned using a Monte Carlo-based analysis of the dose

distribution surrounding the source [8-15].

More recently, a combination of the HDR and LDR brachytherapy has been used

called pulsed dose rate brachytherapy (PDR) [14,16,17]. PDR involves the use of

stronger radiation sources than those used in LDR brachytherapy given in a series

of short exposures of 10 to 30 minutes every hour to approximate the same overall

dose as with LDR brachytherapy. Typically sources include gamma and beta

emitters such as 192Ir, 32P, 90Sr/Y and 125I. Each source has its own advantages

and is generally chosen to match a specific treatment. These sources may be used

either as wire stents, seeds inserted through catheters or liquids that can be

injected into balloons during angioplasties.

Intravascular Brachytherapy (IVBT) emerged from the need to resolve the

problem of restenosis, a re-closing of arteries following angioplasty [18-20]. In

1995, the American Association of Physicists in Medicine (AAPM) established a

task group committee to investigate and report on the dosimetry of interstitial

brachytherapy sources. Their findings were presented in a report, called TG-43,

that examined the role of photon-emitting sources used for interstitial

brachytherapy [21]. This work was followed-up by a second report in 1999 by

TG-60 that investigated the physics in intravascular brachytherapy [22]. This

report examined the relative advantages of different sources available for use in

brachytherapy. It found that beta emitters, such as 32P and 90Sr/Y have advantages

in terms of high activity and dose rate, good radiation safety and a long half-life.

12

Gamma emitters, such as 192Ir, have advantages in terms of radial dose uniformity,

high dose rate, and reasonably long half-life. One of the key concerns to be raised

in this report was dose inhomogeneity due to non-centering of the source. This

concern is reflected in the need to assure the positional accuracy of the radiation

delivery system that is particularly significant in brachytherapy [23]. Although

Monte Carlo calculations have been effective in the determination of dose

distributions surrounding brachytherapy sources, these must be verified for further

optimization of procedures to be possible.

2.2 Radiation Dosimetry

Dosimetry has been at the core of the development of radiation therapy and there

currently exists many different methods of measuring the absorbed dose delivered

to tissue (and other mediums) [24,25]. Today, complex computer algorithms are

used to determine the dose distributions required in a particular treatment plan.

These algorithms aim to assure that the calculated dose distributions accurately

reflect the dosimetry requirements of a treatment plan. In addition, with computer

controlled radiation delivery, software must assure that the complex

manipulations of radiotherapy equipment are actually performed correctly [5,26].

The determination of absorbed dose in 3D is fundamental to clinical

environments, but few methods exist by which 3D measurements can be made

easily and accurately. Computer algorithms model the effective dose absorbed

into specific volumes of tissue and other structures in the patient’s body [1,22].

However useful the computer calculations might be in the prediction of absorbed

dose, the problem of easily and accurately measuring the actual absorbed dose

distribution in 3D still remains, even if only to provide a verification of the

computed calculations [27].

2.2.1 Clinical Dosimetry Requirements

Advances in conformal radiation therapy treatments have enabled target volume

to be defined with complex geometries. Similarly, radiation delivery can consist

13

of varied and non-uniform irradiation fields. These complexities make radiation

dosimetry a difficult process that must meet stringent criteria in order to be

effective in producing accurate and detailed maps of the dose distribution. The

following parameters are typically significant in the design of dosimetry systems:

measurement sensitivity, accuracy, precision, spatial resolution, energy and dose-

rate insensitivity, tissue equivalence, non-directionality, ease of use and able to

produce three-dimensional, integrated dose maps [7].

The International Commission on Radiation Units (ICRU) recommends that the

overall accuracy in delivered dose be within 5 % of the true dose [28,29]. To be

effective in a variety of radiation therapy applications, dosimeters need to measure

dose accurately and to have high spatial resolution. This is particularly true for

brachytherapy, where high resolution is essential since steep dose gradients exist

close to the source. These high dose gradients can occur within very small

regions of interest, typically < 2 mm [30]. Detectors should also exhibit tissue

equivalence in order to not perturb the radiation field and hence have a significant

effect on the measured dose. Non-tissue equivalent dosimeters require the

application of correction factors in order to determine absorbed dose. These may

introduce uncertainties. Dosimeters should also be able to integrate dose for a

number of sequential fields to accommodate the time varying doses delivered to a

patient.

2.2.2 Dosimetry Detectors

There are currently many different dosimetry detectors that are used to determine

radiation dose distribution delivered in radiotherapy treatments each with its own

advantages and disadvantages based on criteria listed above.

2.2.2.1 Ionization Chambers

Currently the most commonly used dosimeter for external beam measurements is

the ionization chamber [31,32]. Ionization chambers provide one dimensional

point dose measurements in radiation therapy applications. Ionization chambers

14

consist of a cavity containing gas, usually air, in which charge is liberated by the

ionization of a gas within the chamber by radiation. This charge is collected by

electrodes, typically composed of aluminum or carbon based material. The charge

can then be correlated to the delivered dose through calibration factors traceable

to a standards laboratory. Although these detectors have high accuracy and

practicality, they do introduce a perturbation of the photon beam which must be

corrected for. The active volumes of ionization chambers are typically between

0.1 and 0.6 cm3 resulting in poor spatial resolution making them unsuitable for

dosimetry in the near-zone of brachytherapy sources. Measuring the dose

distribution in 3D using ionization chambers is a laborious process. Dose

distributions resulting highly conformal external beam delivery typically require

spatial resolution beyond the capabilities of ionization detectors.

2.2.2.2 Solid-State Detectors

Solid-state detectors, such as semiconductor diodes, can be manufactured with an

active volume around 0.1 mm3, and hence are more suited to applications such as

brachytherapy. Solid-state detectors function by measuring the amount of charge

liberated by the passage of ionizing radiation in solid semiconductors [33-36].

Arrays of solid-state detectors can be spatially arranged or translated around static

fields to achieve 2D and 3D maps of dose distributions. These detectors have the

ability to measure dose distributions with higher resolutions than ionization

chambers, in real-time and are easy to use, however, they are not tissue equivalent

and require independent calibration. Solid-state detectors also suffer from an

energy dependence and their response drifts over time due to radiation damage.

2.2.2.3 Thermoluminescent Dosimeters

Thermoluminescence refers to the emission of light from an irradiated crystalline

material following heating. The amount of light emitted from a crystal can be

correlated to absorbed dose. Thermoluminescent dosimeters (TLDs) consist of

small crystals available in a range of sizes, some as small as 1 mm2, that act as

point detectors. When exposed to radiation, these crystals store a small amount of

the energy in the crystal lattice. Upon heating, these crystals release this stored

energy as light, which can be detected using a photomultiplier tube. They can be

used to provide higher resolutions using a spatially arranged array of closely

15

packed TLDs. The main advantages of TLDs include: their wide useful dose

range, small physical size, reuseability and economy for most radiation types [37-

40].

2.2.2.4 Radiographic Film

Radiographic films are used to capture a 2D image of a dose distribution.

Conventional films are based on silver-halide (typically silver bromide) and have

a strong energy dependence at photon energies in the 10 to 200 keV range [41], an

effect derived from the high atomic number of the silver in the film and

absorption due to the photoelectric effect that is significant in this energy range.

In addition, silver-halide films are non-tissue equivalent. They function by

converting silver ions to silver upon irradiation. The bromine is removed during

developing leaving opaque clusters of silver on the film in irradiated regions.

Radiochromic films are relatively tissue equivalent and do not require chemical

developing, however, they exhibit significant temperature dependence and their

sensitivity varies photon energy. They develop a specific colouring in irradiated

regions as a result of a dye-forming or polymerization process in which energy is

transferred from an energetic photon or particle to the receptive part of a leuko-

dye or colourless photomonomer molecule [41]. Gaf-chromic films, a type of

radiochromic film, were developed with a more uniform response with photon

energy. Gaf-chromic films are popular in clinical radiation dosimetry as they also

exhibit high sensitivity and high spatial resolution [42,43].

Film dosimeters are inherently 2D and can be used to obtain dose information by

relating the absorbed dose to the optical density of the film. Although sheets of

film can be stacked between tissue equivalent material to provide a 3D dosimetry

system, the process is cumbersome and time consuming. The use of radiographic

films in dosimetry is complicated by their non-tissue equivalence, uncertainties in

film processing and their inherent 2D nature [44].

2.2.2.5 Chemical Dosimeters

Chemical dosimeters function on the phenomenon that chemical changes occur in

the dosimeters when exposed to ionizing radiation. The absorbed dose delivered

16

to the dosimeter can be correlated to the extent of radiation-induced chemical

change within the dosimeter itself. These chemical changes can be measured as a

change in the spin-lattice or spin-spin relalaxation rates, the change in

concentration of ions present in solution or the optical turbidity within the sample.

One of the more widely known chemical dosimeters is the Fricke dosimeter,

proposed by Fricke and Morse in 1927 [45]. The Fricke dosimeter functions by

the conversion of ferrous ions (Fe2+) to ferric ions (Fe3+) in solution when

irradiated.

A second type of chemical dosimeter uses the radiolysis of water within a gel

matrix to initiate polymerization reactions. The degree of polymerization can be

correlated to the absorbed dose [46]. Chemical dosimeters are capable of high

precision dose measurements and can provide spatial information when dissolved

into an aqueous gel matrix.

2.3 Gel Dosimeters

As outlined above, there is a need for a dosimeter that does not perturb the

radiation beam, is inherently three-dimensional and has the ability to integrate

radiation doses over time. The use of radiation sensitive gels to fulfill these

requirements has great potential. Radiation sensitive gels were first considered

for use in radiation dosimetry in the 1950s by Day et al. who were investigating

radiation induced colour changes in dyes [47,48]. With the addition of gelling

agents, a chemical dosimeter could be made to be spatially stable and hence to

provide spatial dose information. With the development of methods to image the

chemical changes within gel dosimeters, this became the basis of modern gel

dosimetry.

Gel dosimetry has advanced significantly over the past two decades. Now gel

dosimeters have the ability to measure three-dimensional dose distributions with

high resolution (less than 1 mm in-plane resolution) [49] and the dose sensitivity

of the gel is independent of the energy of the irradiating beam and the dose rate

17

used to irradiate the gel [50]. Most importantly, the gel is the dosimeter and thus

does not perturb the radiation beam like conventional dosimetry techniques [51].

In addition, gel dosimeters are also tissue equivalent. The gel dosimeter can be

used to simulate the tissue of the human body undergoing radiation therapy, by

pouring into anthropomorphic phantoms [52].

There are two main varieties of gel dosimeter: ferrous sulfate gel dosimeters and

polymer gel dosimeters. More recently, normoxic polymer gel dosimeters have

become popular due to their ability to be manufactured under normal atmospheric

conditions.

2.3.1 Ferrous Sulfate (Fricke) Gels

In 1984, Gore et al. proposed the use of nuclear magnetic resonance imaging of

ferrous sulfate gel dosimeters, also called a Fricke gel, which exhibit a change in

their paramagnetic species as a result of exposure to ionizing radiation [53]. Gore

et al. also proposed the potential of this chemical response as a dosimeter capable

of producing 3D dose distributions of the nature required for use in radiation

therapy. The use of a Fricke gel has been developed into a method founded on the

principle that ferrous ions (Fe2+) are oxidized to ferric ions (Fe3+) when subjected

to free radicals produced by exposing water to ionizing radiation. To achieve this,

a gel consisting of an aerated dilute solution of ammonium ferrous sulfate is

suspended in an aqueous gel, such as agarose or gelatin. The gel matrix provides

the support structure by which the dosimeter can maintain a spatial arrangement,

and hence provide spatial information about the dose distribution within the

irradiated gel dosimeter. This was significant for gel dosimetry as Gore et al.

were able to show that the radiation-induced change could be detected using

nuclear magnetic resonance (NMR) [53].

The change from ferrous to ferric ions that occurs in regions exposed to the

radiation, provides changes in the gel dosimeter’s nuclear magnetic resonance

(NMR) spin-lattice relaxation rate, R1, and spin-spin relaxation rate, R2, as both

18

ferrous and ferric ions are paramagnetic species capable of reducing the proton

relaxation times of water. In particular, the ferric ions exhibit a stronger

paramagnetic enhancement of the water-proton NMR relaxation rates. Gore et al.

were also able to show that by spatially distributing the ferrous ions, the spatial

distribution of dose could be imaged using MRI [53]. This was a significant

advance in dosimetry as this was the first evidence for a truly 3D dosimeter for

clinical radiation therapy. Since then, many studies have investigated combining

Fricke chemical dosimeters with gelling agents and imaging using MRI [54-56].

Optical tomography, an alternative to MRI, has also been investigated by many

authors as an imaging modality for Fricke gel dosimetry [57-60].

A great deal of research has been performed to investigate various aspects of

Fricke gel dosimetry, including the effects of: ferrous ion concentration, radiation

dose rate, beam energy, oxygenation, agarous concentration and acid content [61-

64]. Additionally, the use of alternative gelling agents, such as gelatin [65-67]

and polyvinyl alcohol (PVA) [68], have been explored.

There are two major drawbacks with the use of Fricke gels for 3D dosimetry. The

first limitation of ferrous gels is the continual diffusion of the ferric ions through

the gel post-irradiation (see references cited in Baldock et al. 2001 [69]). This

diffusion leads to a blurring of dose distributions over time, and hence a

degradation of spatial integrity of the dosimetry. However, with the use of

chelating agents, this diffusion can be reduced to better maintain the spatial

information over extended time periods [69]. The addition of saccarides to

ferrous-agarose-xylenol orange gels have also improved dose sensitivity [70].

The problems experienced with Fricke gel dosimeters, have prompted an

alternative gel formulation that could provide a more stable dosimeter that would

provide both a better spatial resolution as well as being relatively insusceptible to

problems caused by ageing [71]. A more detailed discussion of Fricke gel

dosimeters can be found in Back et al. 1999 [72] and Schreiner 2004 [73].

19

2.3.2 Polymer Gel Dosimeters

In 1993, Maryanski et al. introduced the use of radiation initiated polymerization

into gel dosimetry [74]. It was well-known that polymerization and cross-linking

could be initiated by irradiation and that the degree of polymerization could be

correlated to the amount of radiation delivered [75-78]. Maryanski et al. used this

knowledge to construct a gel dosimeter based on the polymerization [74]. This

polymer gel dosimeter used an agarose gel infused with acrylamide and N,N’-

methylene-bis-acrylamide (bis) co-monomers. It functioned on the premise that

ionizing radiation would initiate the polymerization of the co-monomers and

induce cross-linking by way of the bis forming a connection between two

acrylamide chains. Nitrous oxide was used to saturate the solution in order to

remove any oxygen. Likewise, the manufacture of these dosimeters had to occur

in a hypoxic environment as oxygen was shown to inhibit the polymerization

process [74]. The use of polymer gel dosimeters provided solutions to some of

the problems that had been encountered with Fricke gels and hence provided the

ability to conduct improved dosimetry in 3D. Polymer gel dosimeters do not

suffer the diffusion problems observed in Fricke gels and the range of doses that

the polymer gel dosimeters responded to can be engineered by varying the

chemical constituents that make up the gel [5]. Another desirable property of

polymer gel dosimeters is their optical characteristics, in particular the observable

transparency of regions of the gel that are unexposed to ionizing radiation and the

opaque regions where ionizing radiation has affected the gel provided immediate

visual clues as the distribution of absorbed dose. The opacity of regions within

the irradiated regions of gel are due to the formation of polymer aggregates

initiated by free radicals formed by the radiolysis of water molecules.

Polymer gel dosimeters were found to exhibit significant changes in both the

NMR spin-lattice relaxation rate (R1) and the spin-spin relaxation rate (R2).

However, unlike the Fricke gel, in polymer gel dosimeters, a change in R2 is

much more pronounced than the change in R1 [74,79-81]. Therefore, the spin-

spin relaxation rates (R2 = 1/T2) determined from a suitable magnetic resonance

20

image and are correlated to the absorbed dose. One of the main aspects to emerge

from the investigation of polymer gel dosimetry using MRI was a quantitative

method of measuring the performance of the polymer gel as a dosimeter. The

slope of the linear region of the R2-dose response over the range 0 to 10 Gy,

provides a measure called the R2-dose sensitivity of the particular gel dosimeter

(see figure 1). The R2-dose sensitivity is a useful quantity to compare different

gel formulations and MRI imaging techniques. The original formulation of the

polymer gel exhibited a dose sensitivity of 0.28 s-1 Gy-1 [74].

0 2 4 6 8 10

1

2

3

4

R2 = 0.96146 + 0.28574*Dose(r2 = 0.9953, p < 0.0001)

R2 (

s-1)

Absorbed Dose (Gy)

Figure 1. A typical R2-dose response from a polyacrylamide gel dosimeter

showing a linear fit with corresponding formula from which the dose sensitivity

can be determined.

Maryanski et al. then went on to further develop the formulation of polymer gel

dosimeters. A new polymer gel dosimeter, based on Bis, Acrylamide, Nitrogen

and Gelatin (called BANG) was produced [82-84]. The reason for changing the

gel matrix from agarose to gelatin lay in attempting to reduce the component of

the R2 that was due to the gel matrix itself. The gelatin produced an R2 an order

of magnitude lower than the agarose gel. This reduction in the magnitude of R2

significantly reduced the zero dose baseline in MRI measurement [82]. The R2

magnitude then became dominantly controlled by the polymerization that had

occurred in the gel due the absorbed dose itself. The R2-dose sensitivity of the

polymer gel dosimeter based on gelatin was found to be only 0.25 s-1 Gy-1 over

21

the range of 0 to 8 Gy for a magnetic field strength of 1.5 T, but it did demonstrate

good reproducibility.

Since this initial work various concentrations of polymer gel dosimeters have been

explored. The replacement of acrylamide with acrylic acid led to the development

of a polymer gel dosimeter partially suited to the verification of stereotactic

radiosugery and high dose rate (HDR) brachytherapy dose delivery [50]. The

study demonstrated that the dose response was independent of irradiation

conditions. The R2-dose sensitivity of this polymer gel dosimeter was found to be

0.335 s-1 Gy-1 for doses up to 12 Gy. The response of the gel was independent of

the energy, up to 15 MeV, and of the dose rate, over the range 0.003 to 0.067 Gy

s-1.

The optical characteristics of gelatin based polymer gel dosimeters based on

gelatin were further explored using optical tomographic densitometry by

Maryanski and Gore [85,86]. The method is founded on the scattering of light

that occurs due to the presence of micro-particles that are produced due to the

polymerization of co-monomers initiated by the irradiation. The measurement

procedure consisted of the use of a specially designed optical scanner that

comprised a He-Ne laser that imaged the gel in much the same fashion as first

generation x-ray computed tomography using filtered back-projection to

reconstruct images of the absorbed dose within the gel. This method of imaging

gels was further developed into a method of dosimetry for complex stereotactic

radiosurgery [87]. This paper detailed the benefits that polymer gel dosimetry had

over conventional dosimetry techniques that had been used previously in

radiosurgery, including: thermoluminescent dosimeters (TLDs) and small volume

ionization chambers. These devices suffer from the inherent problems of poor

spatial resolution due to the size of the device and the perturbation of the ionizing

radiation due to the physical presence of the device when measuring. Knisley et

al. showed the polymer gel dosimeter to be effective 3D dosimeters exhibiting

high resolution, precision and accuracy [87].

Pappas et al. investigated the use of N-vinylpyrrolidone combined with bis in the

manufacture of an alternative polymer gel dosimeter called VIPAR [88,89].

22

Nitrogen was replaced with argon in this study as argon is heavier than air and

thus decreased the likelihood of air diffusion through the seals on the vessels

containing the dosimeter. Formulations including either gelatin or agarose were

explored, however, the agarose produced an increase in the turbidity of the gel

and, thus, gelatin was prefered. The R2-dose sensitivity of this polymer gel

dosimeter was found to be ~ 0.1 s-1 Gy-1 which, although less than half that of

acrylamide gels, remained constant with time between irradiation and imaging and

showed good reproducibility. The R2-dose sensitivity of VIPAR exhibited a

quasi-linearity over the range 0 to 12 Gy and thus validated the use of polymer gel

dosimeters based on N-vinylpyrrolidone as a satisfactory formulation for

applications in gel dosimetry.

The replacement of acrylamide with sodium methacrylate was investigated by

Murphy et al. [90]. The sodium methacrylate formulation exhibited several

advantages over the acrylamide formulations including: reduced toxicity, a higher

R2-dose response and the inclusion of a distinct NMR signal due to the presence

of a methyl group in the monomer. The methyl group is later consumed in the

polymerization process. Proton spectroscopy had been previously used to study

polyacrylamide gel dosimeters and had shown that the overall loss of monomer

could be determined spectroscopically as a function of dose [91]. Murphy et al.

were able to demonstrate the same or enhanced possibility in polymer gel

dosimeters composed of sodium methacrylate [90]. It was also found that the pH

of the gel had a major effect on the overall dose sensitivity of the gel. If the pH

was left unchanged during the manufacture of the gel, the R2-dose sensitivity was

found to be only half that of polyacrylamide gel dosimeter, however, by adding

sodium hydroxide and raised the pH to 7.7, the R2-dose sensitivity was made

equivalent to that of the polyacrylamide gels [90].

In summary, polymer gel dosimeters possess many of the desired characteristics

that were required for dosimetry in radiation therapy. Polymer gel dosimeters

have the ability to integrate dose without perturbing the radiation beam, they are

tissue-equivalent, independent of radiation energy over a wide range of photon

energies and inherently 3D. However, clinical acceptance has been limited in part

23

because they were not easy to manufacture, or use, due to the requirement of a

strict hypoxic environment during their manufacture [92].

2.3.3 Normoxic Polymer Gel Dosimeters

The next significant step in development of gel dosimetry came with the advent of

polymer gel dosimeters that could be manufactured under normal atmospheric

(normoxic) conditions. The first normoxic polymer gel dosimeter was proposed

by Fong et al. in 2001 [93]. Called MAGIC, it comprised: methacrylic, ascorbic

acid, hydroquinone, gelatin and copper(II) sulphate. The main feature introduced

in normoxic polymer gel dosimeters was the addition of an anti-oxidant, in this

case ascorbic acid, into the gel formulation. As noted earlier, the polymerization

process within polymer gel dosimeters is inhibited by the presence of oxygen

which scavenges the free radicals produced by the radiolysis of water. It is

usually these free radicals that initiate the polymerization reaction. With the

inclusion of an anti-oxidant in the formulation of a polymer gel dosimeter, oxygen

present in the gel dosimeter can be bound into metallo-organic complexes. Once

the oxygen is bound, it is prevented from binding the free radicals and hence

inhibiting polymerization reaction essential for polymer gel dosimetry.

More recently, various studies have investigated the addition of other anti-

oxidants to a range of different polymer gel dosimeter formulations. Tetrakis

(hydroxymethyl) phosphonium chloride (THPC) has been added as an anti-

oxidant to various formulations including methacrylic acid gelatin gel dosimeters

with copper(II) sulphate and hydroquinone (MAGAT), polyacryamide gelatin gel

dosimeters (PAGAT) [94-96], and just methacrylic acid, gelatin gel dosimeters

(MAGAS) [97]. Venning et al. have performed an extensive analysis of

radiological properties the MAGIC, MAGAS and MAGAT gels using Monte

Carlo modelling [94]. They were able to ascertain that the gel exhibited many of

the characteristics necessary for use in radiotherapy gel dosimetry, including

tissue equivalence. There are still many variations of formulation for normoxic

24

polymer gel dosimeters yet to be fully investigated. The further characterization

of normoxic polymer gel dosimeters is a major component of this thesis.

2.4 Characteristics of Polymer Gel Dosimeters 2.4.1 Effects of Oxygen

The process of polymerization is initiated by free radicals formed from the

radiolysis of water in the gel composition. These free radicals combine with the

monomers making them reactive. Molecular oxygen, however, acts as a

scavenger of these free radicals and hence prevents them from initiating the

polymerization process [51,74,82,92]. Even trace amounts of oxygen in the gel

mixture can lead to the failure of the gel as an effective dosimeter. An important

component of the manufacture of polymer gel dosimeters is the removal of

oxygen from either a reaction flask or a glove box by the bubbling of an inert gas,

for example nitrogen or argon, through the water that is to be used in the

formulation before mixing the other ingredients [98,99]. It is, therefore, important

to ensure the type and quality of the seals used on the vessels do not allow the

diffusion of oxygen into the vessel. Maintaining a strict hypoxic environment has

been a significant drawback of polymer gel dosimeters in the past and made the

process of polymer gel dosimetry awkward to implement into clinical practice.

However, with the advent of normoxic polymer gel dosimeters, as described

above, the strict hypoxic environment is no longer required. Normoxic polymer

gel dosimeters can be manufactured under normal atmospheric conditions on the

bench top. However, the development of normoxic polymer gel dosimeters is in

their infancy and much work is still to be done to be able to fully understand and

integrate normoxic polymer gel dosimeters into clinical practice.

2.4.2 Effect of Light

The initiation of the polymerization process should be caused by the radiolysis of

water that leads to the production of free radicals, as discussed above. However, a

25

number of alternative initiators exist. Bright light, especially sunlight, can initiate

photopolymerisation of the gel before it is irradiated and consequently degrade the

sensitivity of the gel [51,100]. Polymer gel dosimeters should, therefore, be

manufactured, irradiated and stored away from strong light sources.

2.4.3 Temperature

There are several places where temperature plays a significant role in the

manufacture of the gel. The first step in the manufacturing procedure requires

high temperature to facilitate mixing of the gelatin and water. The gelatin must be

added whilst the water is at room temperature to avoid the gelatin forming lumps.

Once the gelatin has soaked into the water, the mixture is then heated to ~ 50 °C

to ensure that the gelatin has completely dissolved into the water [98]. The

temperature of the mixture must be kept below 55 °C when mixing the monomers

to avoid pre-polymerisation that may be caused due to the temperature of the

solution. Following the manufacture, the temperature of the gel should be kept

low to ensure the gel sets in the vessel it has been placed in. Salomons et al.

(2002) have showed that a temperature increase occurs within a polyacrylamide

gel dosimeter during and immediately after irradiation due to the exothermic

polymerization reactions [101]. This temperature change can affect

polymerization reactions within the gel dosimeter and hence may lead to

inaccurate calibration of gel dosimeter images. In order to minimize the effect of

this artifact on the dose maps produced by gel dosimeters, the size, shape and

temperature of the gel dosimeters must be controlled.

During magnetic resonance imaging, the temperature of the polymer gel has a

very significant effect of the overall R2-dose sensitivity of the polymer gel

dosimeter. Several authors have found an increase in the R2-dose sensitivity of

the gel as temperature decreases [102,103]. This effect is thought due to a change

in the proton exchange rates in the gel as the temperature is varied. As the

temperature is decreased, the motion of the water protons becomes slower. This

increases the exchange rate of energy between protons [51]. It has also been

found that a change of even 1 °C within a phantom can give rise to dose

26

uncertainties of approximately 50 cGy in dose maps derived from gel dosimeters

imaged using MRI [102]. The temperature of a gel undergoing the imaging

process must be kept constant to avoid changes in the relaxation rates over the

time of imaging. The gel should be kept in temperature controlled conditions,

such as those of the MRI room, for at least 12 hours before imaging to allow time

for the dosimeter to equilibrate to the scanning temperature.

2.4.4 Concentration of monomers

The R2-dose sensitivity of a gel may be increased by increasing the total

monomer content of the gel, typically symbolized using %T [102,104]. The

typical concentrations of monomers range from 3 % to 9 % of the total weight.

An increase in the monomer concentration is limited, though, by the low solubility

of the bis and crystallization of the gel that can occur during storage [74]. Also,

as some monomers used in polymer gel dosimetry may be strong acids, high

concentrations can adversely affect the gelatin within the gel dosimeter over time.

In order to produce the highest R2-dose sensitivity, the relative proportion of each

of the individual co-monomers, typically symbolized using %C, was found to be

50 % of the total co-monomer content [102,105]. This result was supported by

investigating the effect of chemical exchange on magnetization transfer in

polyacrylamide gels [106].

2.4.5 Ageing of the gel

Unlike the problems encountered by Fricke gels where there is a diffusion of the

ferric ions over time thus degrading the spatial information contained in the gel

dosimeter, polymer gel dosimeters are not as limited by time constraints [107].

One exception is a time evolution of the dose response that occurs following the

irradiation of the gel as polymerisation processes are occurring at the greatest rate

and may lead to errors in the use of separate calibration vials and phantom [51].

De Deene et al., in an investigation into the stability of the polymer gel dosimeter

structure, found that the initial 12 hours post-irradiation yield significant errors

due to the chemical instability of the polymer gel [108]. After around 12 hours,

27

most polymer gel dosimeters maintain a reasonable temporal stability over a

period of several days.

2.5 Evaluation of Polymer Gel Dosimeters

Following irradiation, polymer gel dosimeters must be imaged to produce maps of

the absorbed dose distribution. To evaluate polymer gel dosimeters, several

imaging modalities have been explored and include: MRI, x-ray CT, optical CT

and ultrasound. In order to be effective in evaluating polymer gel dosimeters, an

imaging modality must be capable of sufficiently detecting the radiation-induced

changes that occur within the polymer gel dosimeter. To date, MRI has been the

most commonly investigated imaging modality used to evaluate polymer gel

dosimeters.

2.5.1 Magnetic Resonance Imaging

To date, MRI has been the most common imaging modality used to evaluate gel

dosimeters. The polymer chains within a polymer gel dosimeter, initiated by

irradiation, affect the mobility of the water molecules. Water molecules attached

to polymer chains undergo very slow and restricted motions compared to bulk

water and water molecules in hydrated monomer. MRI detects this difference by

measuring the larger relaxation rates exhibited by water molecules attached to the

polymer chains. The extent of the polymer network within a polymer gel

dosimeter is related to the absorbed dose. As the polymer network is spatially

arranged within an aqueous gel matrix, the dose distribution within the gel matrix

can be imaged using MRI. In order to relate the measured relaxation rates to

absorbed dose, a R2-dose calibration curve is used. This method provides an

absolute dosimetry. In relative dosimetry, a calibration curve is not required

provided the delivered dose is within the linear R2-dose response range of the

dosimeter.

28

The radiation-induced changes that occur within a polymer gel dosimeter have an

effect on the spin-lattice relaxation rate, R1, and the spin-spin relaxation rate, R2.

These changes can be correlated to the absorbed dose within a polymer gel

dosimeter and hence a calibration curve can be calculated. The change in R2 due

to irradiation of polymer gel dosimeters has been found to be more pronounced

than that of R1 and hence evaluations of polymer gel dosimeters are usually based

on R2 [109]. MRI has been widely demonstrated to be capable of imaging simple

and complex dose distributions in polymer gel dosimeters [49,51,74,82,103,110-

112].

Relaxation rates are obtained in MRI by applying a radio frequency (RF) pulse

exciting the magnetization of the spin system within water protons. As the

magnetization of the water protons returns to its equilibrium state, it can be

sampled. A pulse sequence is a set of RF pulses that are applied to a sample to

produce a nuclear magnetic resonance signal, called an echo. A multiple spin-

echo pulse sequence acquires a train of equally spaced spin-echoes. In this case,

the magnetization signal is refocused using 180° pulses. Each echo that results

occur at a specific echo time that can then be used to determine the spin-lattice or

spin-spin relaxation times, T1 and T2 respectively, within the sample. In 2D, a set

of base images can be attained through a specified cross-section of the sample

being imaged. T2 maps can then be calculated using an exponential fit of the time

course decay of each pixel value within consecutive base images.

Kaurin et al. (1999) demonstrated the effectiveness of a polymer gel dosimeter for

verifying conformal radiotherapy treatments in a four-field oblique and lateral

wedged field technique [113]. Likewise, polymer gel dosimeters have been used

to verify treatments involving IMRT and treatments using dynamic multi-leaf

collimators. De Deene et al. (2000) demonstrated the use of a polymer gel

dosimeter for the verification of dose distributions produced by conformal

radiation of a mediastinal tumour located near the oesophagus [111]. Oldham et

al. (1998) showed good results were obtained using polymer gel dosimeters for

the verification of a nine field tomotherapy irradiation [49]. The use of MRI to

evaluate dose distributions within polymer gel dosimeters has show good spatial

29

agreement with doses specified by treatment planning calculations. MRI is

known for its high contrast and ability to generate images in 3D.

MRI is limited, however, by practical and technical complications; in particular,

MRI equipment is expensive and often access is limited. Technical limitations

relate to the accuracy of dose distributions determined via MRI being susceptible

to temperature variations during scanning, as discussion earlier. Further

complications exist due to artifacts inherent to MRI. De Deene et al. have

demonstrated the spatial accuracy of dose distributions can be affected by

geometrical distortions caused by eddy currents, magnetic field inhomogeneities

and magnetic field gradient inhomogeneities [114,115]. The dose response was

also found to be dependent on the scanning orientation and the size of the field of

view [116]. Typically, MRI pulse sequences used in conventional diagnostic

imaging are not optimal for imaging polymer gel dosimeters. Sub-optimal pulse

sequences have been reported to be another source of uncertainty in dose

distributions produced by MRI equipment [117,118]. These studies have

demonstrated that it is possible, however, to optimize pulse sequences for

producing dose maps from imaging gel dosimeters [118,119]. These

complications make the routine use of MRI and polymer gel dosimetry in clinics

difficult.

2.5.2 X-Ray Computed Tomography

The potential of x-ray computed tomography (CT) to evaluate polymer gel

dosimeters has also been investigated [119-122]. The radiation-induced changes

within a polymer gel dosimeter lead to density changes which, in turn, affect the

x-ray linear attenuation coefficients [120,121]. Investigations have explored the

correlation between Hounsfield Units (HU) from x-ray equipment and absorbed

dose. These investigations have also explored the relationship between the x-ray

linear attenuation coefficient and absorbed dose given changes in the composition

of polymer gel dosimeters.

30

Audet et al. (2002) have more recently investigated the use of x-ray CT to

evaluate complex dose distributions in a polymer gel dosimeter. In their study, a

dose delivery based on a stereotactic treatment plan was used to expose a polymer

gel dosimeter. The dosimeter was later imaged using a clinical x-ray CT scanner

[122]. The results showed that a polymer gel dosimeter could be used to delineate

irradiated regions within a volume of unirradiated gel. There was good agreement

found between the measured dose distribution and the treatment plan, with spatial

agreement for the 50 % and 80 % isodose curves to within 3 mm. One key

advantage of using x-ray CT to evaluate polymer gel dosimeters is its relative

temperature insensitivity. Unlike the highly-temperature sensitivity of MRI, a

temperature change of 4 °C to 23 °C within a polymer gel dosimeter evaluated

using x-ray CT results in a change in dose sensitivity of less than 10 % [119].

Other advantages are economy and accessibility, especially compared to MRI and

its high spatial accuracy capable of sub-millimeter spatial resolutions.

More recently, there have been studies evaluating normoxic polymer gel

dosimeters using x-ray CT [96,123-125]. Brindha et al. (2004) have investigated

the linear attenuation coefficients for normoxic polymer gel dosimeters. In

particular, they have investigated normoxic polymer gel dosimeters made from

polyacrylamide or methacrylic acid and tetrakis (hydroxymethyl) phosphonium

chloride as an anti-oxidant, respectively PAGAT and MAGAT gels [96]. They

found that the CT-dose response was linear up to 15 Gy for the PAGAT gel and

up to 10 Gy for the MAGAT gel and concluded the two normoxic polymer gel

dosimeters were suitable for use in radiotherapy dosimetry. Hill et al. have

investigated the dose response of a MAGIC gel dosimeter using x-ray CT using

doses up to 150 Gy [123,124]. They then applied the methodology to the noval

measurement of computer tomography dose indices on diagnostic x-ray CT

scanners [125]. They concluded that normoxic polymer gel dosimetry was a valid

technique for use in x-ray CT quality assurance and hence showed promise to be

incorporated into clinical practice.

There are, however, a couple of limitations to the effective use of x-ray CT for

evaluating polymer gel dosimeters. X-ray CT suffers from a lower dose

sensitivity and a lower contrast-to-noise ratio (CNR) compared to the signal-to-

31

noise ratio (SNR) of MRI [119]. The CNR of an image can be improved by

averaging multiple scans, however, this leads to a dramatic increase in the

scanning time and tube loading. This in turn, increases the cost associated with

imaging gel dosimeters. Also, the linear effective dose range of a polymer gel

dosimeter evaluated using x-ray CT is relatively small compared to MRI, being

linear only up to around 10 Gy.

2.5.3 Optical Computed Tomography

As the optical properties of a polymer gel dosimeter change when irradiated, it has

been possible to detect absorbed dose within the dosimeter using optical methods

and correlating the optical density to absorbed dose. In addition by using a

computed tomography (CT) system, it is possible to reconstruct an image of the

dose distribution within a polymer gel dosimeter [85,86,126,127]. An optical CT

system functions by passing light through the polymer gel dosimeter via differing

angles encircling the dosimeter. The change in opacity within a polymer gel

dosimeter is a result of the radiation induced polymerization. As the polymer

aggregates form in the dosimeter, they become scatter centres causing the optical

attenuation to increase with dose [86]. The dose distribution within a polymer gel

dosimeter can be determined by reconstructing the tomographic images of optical

density and correlating to absorbed dose.

The advantages of an optical CT system to evaluate polymer gel dosimeters

include their low noise and hence a superior SNR compared to MRI images,

simple low cost equipment and capability to produce 3D dose maps with

sufficient spatial resolution, accuracy and precision. Oldham et al. (2001)

demonstrated the use of optical CT to evaluate a polymer gel dosimeter for the

verification of complex radiotherapy treatment plans such as those used in

radiosurgery and IMRT [126]. They found that 3D dose maps could be produced

in under an hour with sub-millimeter spatial resolution, 3 % accuracy and less

than 1 % precision. Xu et al. (2002) have also shown the capability of optical CT

32

to image steep dose gradients, such as those encountered in brachytherapy, within

a polymer gel dosimeter [128].

There are, however, significant technical limitations to the use of optical CT for

polymer gel dosimetry. Loss of signal can occur at container boundaries due to

reflection or refraction of laser light used for scanning. Likewise, with increasing

depth of the phantom, the attenuation of light can dramatically reduce the

available information. This places a significant restriction on the size of phantoms

that can be evaluated using optical CT. Additionally, with increasing dose, the

optical density of the gel becomes so significant that scanning is ineffective,

essentially narrowing the potential dynamic range of absorbed dose that can be

imaged to between 0 and 10 Gy. Also, the total optical density across a phantom

cannot exceed a certain maximum as determined by the dynamic range of the light

detection system and its corresponding SNR [128]. When imaging of polymer gel

dosimeters is restricted to only low doses (< 10 Gy ), the contrast-to-noise ratio is

also reduced, which can diminish the dose accuracy.

2.5.4 Ultrasound

It is well-known that ultrasound can be used to determine the characteristics of

materials [129]. The ultrasonic parameters of most interest include: ultrasonic

attenuation, ultrasonic speed of propagation and ultrasonic impedance [129].

Ultrasound has been well studied in its capability to characterize polymer

structures [129-132] and the changes in polymer structures due to irradiation [133-

135]. With changes in polymer structure due to irradiation being quantifiable via

a number of different ultrasound parameters, it has been possible to correlate

parameters to dose and hence build an effective dosimeter.

Maryanski et al. (1999) were the first to propose the use of ultrasound to evaluate

polymer gel dosimeters [136]. Extensive research into the use of ultrasound to

evaluate polymer gel dosimeters was conducted by Mather et al. [137] who found

that several ultrasound characteristics including: speed of propagation, attenuation

33

and transmitted signal intensity all display a strong variation with absorbed dose

that continues beyond doses of 15 Gy. Ultrasound was applied successfully to

both polyacrylamide polymer gel dosimeters (PAG) and normoxic polymer gel

dosimeters. However, correlated to ultrasonic speed, the dose response in PAG

and normoxic polymer gel dosimeters was found to be fundamentally different

due to differences in the dependence of the gel’s elastic properties on dose at

ultrasonic frequencies. It is from this investigation that Mather et al. were able to

construct a prototype ultrasound system capable of imaging absorbed dose

distributions in polymer gel dosimeters.

Currently, the use of ultrasound to evaluate polymer gel dosimeters is in its

infancy and further work is required to validate reproducibility and to determine

the sources of inconsistencies found in the absolute values of ultrasonic speed and

attenuation in polymer gel dosimeters [137].

2.5.5 Vibrational Spectroscopy

Fourier Transform (FT) Raman spectroscopy can be a useful tool for investigating

radiation induced changes in polymer gel dosimeters [138-142]. Typical

monomers, such as acrylamide and bis-acrylamide, have been identified through

vibrational bands. Changes to these vibrational bands can be correlated to

absorbed dose. Rintoul et al. used Raman microscopy to investigate dose

distributions with spatial resolutions approaching 1 μm [142]. Gustavsson et al.

have developed and optimized a polymer gel dosimeter based on 2-hydroxyethyl

acrylate (HEA) using FT-Raman spectroscopy [143]. From the FT-Raman

spectroscopy results, Gustavsson et al. were able to examine the chemical

structure and properties of the HEA gel and study how these effects contribute to

the relaxation process that is imaged using MRI.

The use of vibration spectroscopy to evaluate 3D dose in polymer gel dosimeters

is also in its early stages. There is still significant work that must be done for

vibrational spectroscopy to become an effective evaluation tool for dosimetry.

34

Although it has the potential for providing dose distributions with very high

spatial resolutions, it may be limited by the achievable penetration depth of light

into the dosimeter.

2.6 High-Resolution MRI in Polymer Gel Dosimetry MRI has been the gold standard of the many techniques that have been used to

evaluate polymer gel dosimeters post-irradiation. Some applications, such as

brachytherapy, require high resolution dose maps that are beyond the capabilities

of many clinical MRI scanners. To achieve high resolution in MRI, gradient

strengths must be increased, which, in turn, have a significant effect on the

relaxation rates of samples being imaging leading to unwanted artifacts. The

gradient strengths applied to a sample are dictated by the specific pulse sequence

used.

In MRI there are many different approaches in scanning a gel dosimeter, including

measuring the spin-lattice relaxation time, T1, the spin-spin relaxation time, T2,

and the apparent diffusivity, Dapp. T2 is predominantly measured in polymer gel

dosimetry. In measuring T2, however, the pulse sequence, and in particular the

echo spacing can be varied to optimize the imaging for the particular gel. Over

the lifetime of the application of polymer and Fricke gels to radiation therapy

dosimetry, many different pulse sequences have been investigated. Gore et al.,

the first to propose the use of MRI to image gel dosimeters, used an inversion

recovery pulse sequence to measure T1 and a Carr Purcell Meiboom Gill (CPMG)

pulse sequence to measure T2 [53]. Figure 2 shows a typical CMPG sequence

that images samples in three dimensions. As changes in T2 dominate those in T1

for polymer gel dosimeters, CPMG was selected by Maryanski et al. to image a

polymer gel dosimeter [74]. With a long recovery time (6000 ms) to allow the

longitudinal signal to completely recover, Maryanski et al. varied the echo time in

order to assess the most optimal echo spacing for the particular gel being scanned.

35

Figure 2. A typical CPMG pulse sequence where GRO, GSS and GPE are the

pulse amplitudesin each of the three dimensios, NT is the pulse duration, TE the

inter-echo time and n the number of echoes acquired.

Using a polyacrylamide polymer gel, Maryanski et al. used a series of Hahn spin

echo pulse sequences [82]. The Hahn spin echo pulse sequence is a single echo

sequence that is usually executed many times using different τ to obtain data to

determine the T2 of a sample. A multi-echo Carr-Purcell pulse sequence, where a

[τ-180°-τ-acquire] sequence is performed repeatedly following a 90° pulse to

obtain several images, was discussed as a seemingly more appropriate option,

however, it had previously been established that this pulse sequence gave

significant errors in T2 caused by the imperfect refocusing of 180° pulses. A

Hahn spin echo pulse sequence avoids the RF inhomogeneities caused by standing

wave effects and attenuation that leads to errors in T2 measurements [50,144].

However, the diffusion of spins can occur between the application of the pulses

and the echo formation thus leading to a loss in the signal, especially for longer

echo times [145]. The optimization of MRI for the purpose of gel dosimetry has

become one of the focuses of research into the improvement and implementation

of MRI gel dosimetry in the clinical setting [103,117,145].

A circularly polarized head coil is often used as both transmitter and receiver to

provide a more adequate signal-to-noise ratio (SNR) [146]. Ideally the main body

36

coil would be used for excitation (as the magnetic field is more uniform) and the

head coil used for detection. However, many clinical scanners will not allow this

approach [146]. The use of a circularly polarized head coil and a standard

multiple spin echo sequence, based on the CPMG pulse sequence, seems to be

favoured by many polymer dosimetry groups [49,88,90,147,148]. A multiple

spin-echo with phase alternating phase shift (PHAPS) pulse sequence has been

used by De Deene et al. [103,108,109,114,115]. This pulse sequence involves the

sampling of k-space (the raw data matrix) twice, once with a Carr-Purcell (CP)

encoding and once with a Carr-Purcell-Meiboom-Gill (CPMG) encoding [111].

The PHAPS scheme compensates for ghosting and mirror artifacts due to

stimulated echoes [103]. The application of a gradient train preceding each pulse

cycle of the sequence will reduce deviations in the measured T2 and consequently

the calculated absorbed dose [114]. The effect of this gradient train is to bring the

eddy current field offset into a steady state. Variations in the eddy current offset

are the cause of many errors in the measurement of T2 and artifacts in the

corresponding images.

It is well-known that the Brownian motion of water molecules in the presence of a

strong magnetic field gradient will have an attenuating effect on an MR signal

[149,150]. When applying a multiple spin-echo sequence to a sample, the echoes

are attenuated differently and hence a change in the measured R2 of the sample

results. The diffusion weighting factor, commonly referred to as the b-factor, is

used to characterize the effects of self-diffusion of water molecules on the MR

signal [151-153]. Polymer gel dosimetry usually involves very small echo times

of the order 10 to 30 ms [117,118] and, to achieve high-resolutions, very high

gradients strengths. These conditions can introduce significant artifacts into the

use of high-resolution MRI. For high-resolution MRI to be effective in polymer

gel dosimetry, an investigation is required into the effects of molecular self-

diffusion within polymer gel dosimeters due to the application of high gradient

strengths and the use of small echo times in the pulse sequence. Once validated,

high-resolution evaluation of polymer gel dosimeters will provide a valuable

dosimetry tool capable of assessing applications such as intravascular

brachytherapy that typically require resolutions less than 100 μm.

37

Ertl et al. (2000) used high-resolution MRI to evaluate a polyacrylamide polymer

gel dosimeter used in stereotactic radiation therapy [154]. In this study, Ertl et al.

irradiated a head phantom filled with a polyacrylamide polymer gel dosimeter

using a Leksell Gamma Knife. The phantom was then scanned using a 3 T whole-

body MR-tomograph to achieve an in-plane resolution of 234 μm and slice

thickness of 1 mm. Their results showed good agreement between film dosimetry

and calculated data. Berg et al. (2001) attempted high-resolution imaging on a 3

T whole-body scanner with a methacrylic polymer gel dosimeter [155]. In this

study, Berg et al. used T1, T2 and diffusivity to measure the dose response and

found that T2 produced the highest sensitivity. Berg et al. achieved an in-plane

resolution of less than 117 μm with a slice thickness of 0.8 to 1 mm [155]. Their

results from gel measurements were in good agreement with those predicted by

the treatment planning system at high dose levels.

2.7 Brachytherapy Applications of Gel Dosimetry The use of gel dosimeters has been applied to many areas of radiation therapy

[112,154,156-158]. However, its application to brachytherapy has been of

particular interest due to the difficulties faced by other dosimetric methods when

attempting to produce maps of the absorbed dose distribution surrounding a

radioactive source. There are two predominant issues that make dosimetry

difficult in brachytherapy: the high spatial resolution required due to the typical

dimensions of target volumes and the high dose gradients that exist in close

proximity to the source. These issues are more significant in intravascular

brachytherapy where radiation doses are delivered to artery walls only millimeters

away. In conventional brachytherapy tumours of the order of centimeters are

treated, while in intravascular brachytherapy arteries with diameters on the order

of millimeters are treated [159].

The use of gel dosimetry in brachytherapy has been investigated by a number of

authors [147,160-171]. In 1994, Schreiner et al. investigated the use of Fricke-

gelatin dosimeters using NMR to evaluate HDR brachytherapy dose distributions

[160]. They used a microSelectron HDR remote afterloader with a 192Ir source to

expose Fricke gel dosimeters to a variety of irradiation patterns. Results from the

38

Fricke-gelatin dosimeters showed good agreement with dose points predicted by

the treatment planning system.

The use of ferrous sulphate gel dosimeters in intracavitary brachytherapy was

investigated by Knutsen, et. al. in 1997 [163]. This investigation involved the use

of an 192Ir source in a remote afterloading device, however, the investigation was

limited by the diffusion of the ferric ions in the gel following irradiation. It was

shown, however, that the isodose lines produced by both computerized treatment

planning software and the measured experimental results were within 2 mm of

each other. The results of the two methods, computerized treatment planning and

Fricke gel dosimetry, were superimposed on each other to provide a means of

comparison.

Hafeli et al. (2000) measured the dose distribution of a 188W / 188Re beta line

source in an endovascular brachytherapy application using a number of different

methods, one of which was a polymer gel dosimeter [162]. Although there was

approximate agreement between the methods used (TLD, Gafchromic film,

polymer gel dosimeter and computer simulation), the polymer gel dosimeter did

not show the same trends as the other methods. However, it was noted that the

polymer gel dosimeter provided 3D data with good resolution.

Brachytherapy may be classed as either low dose rate (LDR) using sources that

deliver around 30 to 90 cGy.h-1 or high dose rate (HDR) where the sources deliver

around 60 to 300 cGy.h-1 [24,159]. The use of a LDR 137Cs source was explored

by Farajollahi et al. (1999) using a polymer gel dosimeter [161]. It was concluded

that polymer gel dosimeters were a suitable dosimeter for verifying LDR

brachytherapy and that good spatial resolution could be achieved. In this study,

the experiment simulated treatment of the intrauterine tube and ovoids, but the

potential of the gel to study more complex geometries was mentioned. The

susceptibility of the gel to oxygen contamination was seen as a main drawback for

the use of polymer gel dosimetry in clinical practice.

High dose rate brachytherapy was explored by McJury et al. (1999) using an 192Ir

source [147]. Again, the results indicated that gel dosimetry is appropriate for

39

measuring the dose distribution surrounding a brachytherapy source. Cross-

section planes were used to examine the distributions around the radioactive

source and good agreement existed between the gel dosimetric measurements and

computer treatment plans. One of the main concerns with regards to the

experiment performed was the possible heating of the gel over time during

imaging. The RF fields used inside MRI equipment deposit energy in the object

being imaged. Changes in the temperature of an object during the process of

imaging lead to changes in the measured T2 and hence lead to inaccuracies in the

measured dose. Papagiannis et al. (2001) investigated the dose distributions close

to an 192Ir source using a N-vinylpyrrolidone-based polymer gel dosimeter and

found good agreement with Monte Carlo dose calculations down to within 3 mm

of the source [166]. In their study, Papagiannis et al. found that the volume

averaging effects close to the source were significant and suggested that a

reduction in slice thickness should improve on the accuracy of the measured dose

distributions.

De Deene et al. (2001) conducted a comprehensive study of the accuracy of

polymer gel dosimetry on a phantom irradiated using brachytherapy [168]. In this

study, De Deene et al. examined the problems associated with oxygen permeation

into polymer gel dosimeters and the subsequent effects on the measured dose. It

was found that oxygen causes a dose threshold in the dose-R2 curve that was

linearly correlated with oxygen concentration. Oxygen was found to permeate

through the catheter and hence De Deene et al. concluded that the inhibitory

effects of oxygen could only be prevented by keeping the phantom in a nitrogen

environment during manufacture and irradiation [168]. A second effect was

found to be the diffusion of monomers post-irradiation in regions where high dose

gradients exist. This diffusion leads to dose overshoots in regions of high dose or

high dose gradients [169]. Susceptibility effects were found to deform the dose

map at locations very close to the catheter. The extent of the deformation was

related to catheter size, the position of the catheter with respects to the direction of

the main field of the MRI scanner and the receiver bandwidth of the imaging

sequence [168]. In using polymer gel dosimetry, it was also found that

discrepancies between where the centre of the source was assumed to be and

where it actually was could lead to significant errors in the measured dose

40

distribution. In addition, partial volume effects were found to be significant in

regions immediately adjacent the central catheter. The combination of these two

effects led to extreme variations in the measured dose distribution.

Amin et al. (2003) compared the results of polyacrylamide polymer gel

dosimeters with radiochromic film to investigate an intravascular brachytherapy 90Sr/90Y source [170]. They found good agreement between the gel dosimeter and

film with a resolution of 0.4 mm/pixel. However, when the resolution was

increased to 0.2 mm/pixel, Amin et al. found significant variations in the

measured dose of the gel dosimeter. It is possible that partial volume effects over

steep dose gradients may have led to these variations. They also found that

oxygen permeation into the gel caused considerable problems and hence made the

technique of gel dosimetry using polyacrylamide polymer gel dosimeters

unsuitable for implementation into clinical practice. It was concluded that

although polymer gel dosimetry did produce good dose distributions,

radiochromic film was preferred in clinical practice due to ease of use.

2.8 Sources of Uncertainty in Polymer Gel Dosimeters

There are potentially many sources of error in each of the steps involved in gel

dosimetry: the manufacture of the gel, the irradiation, the storage of the gel, the

scanning, the post-processing of the images and the calibration of the images.

Each stage in this process introduces stochastic and deterministic uncertainties

that can affect the results. The application of gel dosimetry to a clinical setting

requires the uncertainty in the overall procedure be well known and accounted for

[172,173].

Oldham et al. (1998) discussed a method of improving the calibration accuracy in

gel dosimetry by using depth dose data to obtain a number of average T2 values

for each dose level [174]. They showed that when a large number of calibration

points are obtained, the uncertainty in the slope of the linear R2 response, α, can

be lowered by a factor of about 4 and the uncertainty in the y-intercept, R0, can be

41

lowered by a factor of 10. Baldock et al. (1999) discussed the implications of this

measure of the uncertainty [172]. They pointed out that the r2 fitting parameter is

not a good method of indicating the relative quality of the calibrations. It was

shown that the most significant reductions in the overall uncertainty would be

achieved by reducing the noise in the R2 map.

An investigation into the noise in MRI polymer gel dosimetry was also conducted

by Low et al. (2000) [175]. They concluded that there was little difference in dose

maps produced from small and large vessels of gel imaged using MRI, however, it

was noted that this was inconsistent with some previous research [176]. Volume

averaging reduced dose errors at higher doses, but the systematic errors due to the

MRI scanning artifacts limited the reduction of the overall error.

The concept of dose resolution, pDΔ , was introduced by Baldock et al. (2001)

[117]. It is defined as the minimal separation between two absorbed doses so that

they may be distinguished with a given level of confidence, p. The selection of an

appropriate echo spacing for a given range of doses was determined by the echo

spacing that maintained the lowest dose resolution. This method of evaluating the

quality of gel dosimeters represented an improvement of the conventional dose

sensitivity, as it incorporates the uncertainty of the imaging and the calibration

curve into the measure. Figure 3 shows the 95 % dose resolution for a

polyacrylamide gelatin gel dosimeter.

0 1 2 3 4 5 6 7 8 9 10

0.2

0.4

0.6

0.8 20 ms 25 ms 30 ms

DΔ95

% (G

y)

Absorbed Dose (Gy)

42

Figure 3. The 95% Dose Resolution of a polyacrylamide gelatin gel measured

using a CPMG pulse sequence with inter-echo times of 20, 25 and 30

milliseconds.

Lepage et al. (2001) have investigated the variations in T2 obtained using MRI

due to non-uniformity in the main magnetic field [146]. These effects have also

been studied in ferrous sulphate gel dosimeters and have been found to give

significant uncertainties in T2 [177]. The sources of non-uniformity of the

measured T2 can occur due to imperfect main, or B0, magnetic fields, gradient

eddy currents, the inhomogeneities in the RF signal and the receiver filters.

Lepage et al. proposed that a correction matrix be constructed that would measure

the amount of non-uniformity in a particular MRI scanner for a given echo

spacing by making multiple scans of a number of uniform phantoms with varying

T2 values [146]. Using a linear interpolation, a large matrix could be constructed

of the measured differences between different regions within the field of view and

future images corrected using the matrix. It was found that within a region of

around 78 cm2 in the center of the coronal plane, there was good consistency in

the measured T2 indicating good uniformity in this the central region of this MRI

scanner.

Watanabe et al. (2005) investigated the use of linear and non-linear equations in

the process of calibrating dose distributions from polymer gel dosimeters and

radiographic film [178]. They found that linear equations could be used without

knowing the actual calibration equation derived from the R2-dose response curve.

It was found that using a transformation method, relative dosimetry contained less

uncertainty than the dose distribution determined by gel dosimetry or film,

provided the dosimetry method was at least quasi-linear or could undergo a

variable transformation using logarithmic functions to exhibit a quasi-linearity.

2.9 Conclusion

Normoxic polymer gel dosimetry shows great promise as an effective dosimeter

specifically in its potential to verify irradiation treatment plans. This study

43

provides further exploration of the physical and chemical properties of various

formulations of normoxic polymer gel dosimeters and seeks to apply the

technology to applications requiring high resolution, such as intravascular

brachytherapy.

44

References

[1] Khan, F.M., The physics of radiation therapy. 2nd ed. 1994, Baltimore:

Williams & Wilkins, pp 542-560.

[2] Johns, H.E. and Cunningham, J.R., The Physics of Radiology. 1983,

Springfield: Charles C Thomas.

[3] Webb, S., The Physics of Conformal Radiotherapy. 1997, Bristol: Institute

of Physics Publishing.

[4] Powlis, W.D., Smith, A.R., and Cheng, E., Initiation of multileaf

collimator conformal radiation therapy. Int J Radiat Oncol Biol Phys,

1991. 25: p. 171-179.

[5] Schreiner, L.J., Gel Dosimetry: motivation and historical foundations, in

DosGel'99 Proceedings of the 1st International Workshop on Radiation

Therapy Gel Dosimetry. 1999, Canadian Organization of Medical

Physicists: Lexington, Kentucky.

[6] Nutting, C., Dearnaley, D.F., and Webb, S., Intensity modulated radiation

therapy: a clinical review. Br J Radiol, 2000. 73: p. 459-469.

[7] Kron, T., Radiation therapy requirements: what do we expect from gel

dosimetry, in Proceedings of DosGel 2001: The second international

conference on radiotherapy gel dosimetry, Baldock, C. and De Deene, Y.,

Editors. 2001, Queensland University of Technology: Brisbane. p. 2-9.

[8] Karaiskos, P., Angelopoulos, A., Pantelis, E., Papagiannis, P., Sakelliou,

L., Kouwenhoven, E., and Baltas, D., Monte Carlo dosimetry of a new

192Ir pulsed dose rate brachytherapy source. Med Phys, 2003. 30: p. 9-16.

[9] Papagiannis, P., Angelopoulos, A., Pantelis, E., Sakelliou, L., Karaiskos,

P., and Shimizu, Y., Monte Carlo dosimetry of 60Co HDR brachytherapy

sources. Med Phys, 2003. 30: p. 712-721.

[10] Perez-Calatayud, J., Ballester, F., Serrano-Andres, M.A., Lluch, J.L.,

Puchades, V., Limami, Y., and Casal, E., Dosimetric characteristics of the

45

CDC-type miniature cylindrical 137Cs brachytherapy sources. Med Phys,

2002. 29: p. 538-543.

[11] Ballester, F., Lluch, J.L., Serrano-Andres, M.A., Casal, E., Puchades, V.,

and Limami, Y., Monte Carlo calculations of dose rate distributions

around the Walstam CDC.K-type 137Cs sources. Phys Med Biol, 2001.

46: p. 2029-2040.

[12] Williamson, J.F., Monte Carlo modeling of the transverse-axis dose of the

Model 200 103Pd interstitial brachytherapy source. Med Phys, 2000. 27:

p. 643-654.

[13] Williamson, J.F., Monte Carlo-based dose-rate tables for the Amersham

CDCS.J and 3M model 6500 137Cs tubes. Int J Radiat Oncol Biol Phys,

1998. 41: p. 959-970.

[14] Williamson, J.F. and Li, Z., Monte Carlo aided dosimetry of the

microselectron pulsed and high dose-rate 192Ir sources. Med Phys, 1995.

22(6): p. 809-19.

[15] Williamson, J.F., Comparison of measured and calculated dose rates in

water near I-125 and Ir-192 seeds. Med Phys, 1991. 18: p. 776-786.

[16] Valicenti, R.K., Kirov, A.S., Meigooni, A.S., Mishra, V., Das, R.K., and

Williamson, J.F., Experimental validation of Monte Carlo dose

calculations about a high-intensity Ir-192 source for pulsed dose-rate

brachytherapy. Med Phys, 1995. 22(6): p. 821-9.

[17] Skowronek, J. and Piotrowski, T., Pulsed dose rate brachytherapy: a

method description and review of clinical application. Przegl Lek, 2002.

59(1): p. 31-36.

[18] Balter, S., Chan, R.C., and Shope, T.B., Introduction and Overview:

Intravascular Brachytherapy - Fluoroscopically Guided Interventions, in

Intravascular Brachytherapy, Fluoroscopically Guided Interventions,

Balter, S., Chan, R.C., and Shope, T.B., Editors. 2002, American

Association of Physicists in Medicine: Montreal. p. 1-16.

[19] Teirstein, P.S., Massullo, V., Jani, S., Popma, J.J., Mintz, G.S., Russo,

R.J., Schatz, R.A., Guarneri, E.M., Steuterman, S., Morris, N.B., Leon,

M.B., and Tripuraneni, P., Catheter-based radiotherapy to inhibit

restenosis after coronary stenting. N Engl J Med, 1997. 336: p. 1697-1703.

46

[20] Waksman, R. and Schwartz, R.S., Vascular Brachytherapy for Prevention

of Restenosis: A Brief History and Overview. J Invasive Cardiol, 1999.

11(1): p. 33-35.

[21] Nath, R., Anderson, L.L., Luxton, G., Weaver, K.A., Williamson, J.F., and

Meigooni, A.S., Dosimetry of interstitial brachytherapy sources:

recommendations of the AAPM Radiation Therapy Committee Task

Group No. 43. American Association of Physicists in Medicine. Med

Phys, 1995. 22(2): p. 209-34.

[22] Nath, R., Amols, H., Coffey, C., Duggan, D., Jani, S., Li, Z., Schell, M.,

Soares, C., Whiting, J., Cole, P.E., Crocker, I., and Schwartz, R.,

Intravascular brachytherapy physics: report of the AAPM Radiation

Therapy Committee Task Group no. 60. American Association of

Physicists in Medicine. Med Phys, 1999. 26(2): p. 119-52.

[23] Chakri, A. and Thomadsen, B., A quality management program in

intravascular brachytherapy. Med Phys, 2002. 29(12): p. 2850-2860.

[24] Thwaites, D.I. and Williams, J.R., Radiotherapy physics in practice. 2nd

ed. 2000, Oxford: Oxford University Press,. xxii, 332 , 2 of plates.

[25] Attix, F.H., Roesch, W.C., and Tochilin, E., Radiation Dosimetry. 2nd ed.

1966, New York: Academic Press.

[26] Leavitt, D.D. and Starkschall, G. Proceedings of the XIIth International

Conference on the Use of Computers in Radiation Therapy. 1998.

Madison WI: Medical Physics Publ.

[27] Back, S.A., Magnusson, P., Olsson, L.E., Montelius, A., Fransson, A., and

Mattsson, S., Verification of single beam treatment planning using a

ferrous dosimeter gel and MRI (FeMRI). Acta Oncol, 1998. 37(6): p. 561-

6.

[28] (ICRU), I.C.o.R.U.a.M., Determination of Absorbed Dose in a Patient

Irradiated by Beams of X or Gamma Rays in Radiotherapy Procedures.

Technical Report 24. 1979: ICRU Publishing.

[29] Brahme, A., Dosimetric precision requirements in radiation therapy. Acta

Radiol Oncol, 1984. 23: p. 379-391.

[30] Van Dyk, J., The Modern Technology of Radiation Oncology. 1999,

Wisconsin: Medical Physics Publishing.

47

[31] (IAEA), I.A.E.A., Absorbed dose determination in photon and electron

beams, in An international code of practice. 1987, IAEA Technical

Reports Series No. 277: Vienna, Austria.

[32] Svensson, H. and Brahme, A., Recent advances in electron and photon

dosimetry, in Radiation Dosimetry: Physical and Biological Aspects.

1986, Plenum Press: New York.

[33] Rikner, G. and Grusell, E., General specifications for silicon

semiconductors for use in radiation dosimetry. Phys Med Biol, 1987. 32:

p. 870-873.

[34] Rikner, G. and Grusell, E., Patient dose measurements in photon fields by

means of silicon semiconductor detectors. Med Phys, 1987. 17: p. 870-

873.

[35] Nilsson, B., Ruden, B., and Sorcini, B., Characteristics of silicon diodes as

patient dosimeters in external radiation therapy. Radiother Oncol, 1988.

11: p. 279-288.

[36] Li, C., Lamel, L.S., and Tom, D., A patient dose verification program

using diode detectors. Med Dosim, 1995. 20: p. 209-214

[37] Horowitz, Y.S., ed. Thermoluminescence and Thermoluminescent

Dosimetry. Vol I - III. 1984, CRC Press: Florida.

[38] Kron, T., Thermoluminescence dosimetry and its applications in medicine

- Part 1: Physics, materials and equipment. Australas Phys Eng Sci Med,

1994. 17: p. 175-199.

[39] Kron, T., Thermoluminescence dosimetry and its applications in medicine

- Part 2: History and applications. Australas Phys Eng Sci Med, 1995. 18:

p. 1-25.

[40] Kron, T., Clinical thermoluminescence dosimetry: how do expectations

and results compare? Radiother Oncol, 1993. 26: p. 151-161.

[41] Niroomand-Rad, A., Blackwell C.R., Coursey, B.M., Gall, K.P., Galvin,

J.M, McLaughlin, W.L., Meigooni, A.S., Nath, R., Rodgers, J.E., Soares,

C.G.,Radiochromic Film Dosimetry: Recommendations of the AAPM

Radiation Therapy Committee Task Group No. 55. American Association

of Physicists in Medicine. Med Phys, 1998. 25(11): pp. 2093-2115.

[42] McLaughlin, W., Soares, C., Sayeg, J., McCullough, E., Kline, R., Wu, A.,

and Maitz, A., The use of a radiochromic detector for the determination of

48

stereotactic radiosurgery dose characteristics. Med Phys, 1994. 21(3): p.

379-388.

[43] Dini, S.A., Koona, R.A., Ashburn, J.R., and Meigooni, A.S., Dosimetric

evaluation of GAFCHROMIC XR tyhpe T and XR type R films. J. Appl

Clin Med Phys, 2005. 6(1): p. 114-134.

[44] Back, S.A.J., Implementation of MRI Gel Dosimetry in Radiation

Therapy, PhD Thesis, Department of Radiation Physics. 1998, Lund

University: Malmo, Sweden.

[45] Fricke, H. and Morse, S., The chemical action of roentgen rays on dilute

sulfate solutions as a measure of radiation dose. Am J Roentgenol Radium

Therapy Nucl Med, 1927. 18: p. 430-432.

[46] Chapiro, A., Radiation chemistry of polymeric system. High polymers ; v.

15. 1962, New York: Interscience,. xvii, 712.

[47] Day, M.J. and Stein, G., Chemical effects of ionising radiation in some

gels. Nature, 1950. 166: p. 146-147.

[48] Day, M.J., Andrews, H.L., Murphy, R.E., and Lebrun, E.J., Gel dosimeter

for depth dose measurements. Review of Scientific Instruments, 1957. 28:

p. 329-332.

[49] Oldham, M., I. Baustert, C. Lord, T.A. Smith, M. McJury, A.P.

Warrington, M.O. Leach, and S. Webb, An investigation into the

dosimetry of a nine-field tomotherapy irradiation using BANG-gel

dosimetry. Phys Med Biol, 1998. 43(5): p. 1113-32.

[50] Maryanski, M.J., G.S. Ibbott, P. Eastman, R.J. Schulz, and J.C. Gore,

Radiation therapy dosimetry using magnetic resonance imaging of

polymer gels. Med Phys, 1996. 23(5): p. 699-705.

[51] McJury, M., M. Oldham, V.P. Cosgrove, P.S. Murphy, S. Doran, M.O.

Leach, and S. Webb, Radiation dosimetry using polymer gels: methods

and applications. Br J Radiol, 2000. 73(873): p. 919-29.

[52] Keall, P. and C. Baldock, A theoretical study of the radiological properties

and water equivalence of Fricke and polymer gels used for radiation

dosimetry. Australas Phys Eng Sci Med, 1999. 22(3): p. 85-91.

[53] Gore, J.C., Kang, Y.S., and Schulz, R.J., Measurement of radiation dose

distributions by nuclear magnetic resonance (NMR) imaging. Phys Med

Biol, 1984. 29(10): p. 1189-97.

49

[54] Back, S.A., J. Medin, P. Magnusson, P. Olsson, E. Grusell, and L.E.

Olsson, Ferrous sulphate gel dosimetry and MRI for proton beam dose

measurements. Phys Med Biol, 1999a. 44(8): p. 1983-96.

[55] Reichl, B., D. Matthaei, J. Richter, and A. Haase, An automated fast MR-

imaging method for localized measurements of dose distributions using

NMR-Fricke gel dosimetry. Evaluation of influences on the measurement

accuracy. Strahlenther Onkol, 1996. 172(6): p. 312-9.

[56] Chu, W.C., W.Y. Guo, M.C. Wu, W.Y. Chung, and D.H. Pan, The

radiation induced magnetic resonance image intensity change provides a

more efficient three-dimensional dose measurement in MRI- Fricke-

agarose gel dosimetry. Med Phys, 1998. 25(12): p. 2326-32.

[57] Healy, B., Brindha, S., Zahmatkesh, M., Baldock, C., Characterisation of

the ferrous-xylenol orange-gelatin (FXG) gel dosimeter, in Third

International Conference on Radiotherapy Gel Dosimetry, De Deene, Y.

and Baldock, C., Editors. 2004, J Phys Conf Ser 3: Ghent. p. 142-145.

[58] Tarte, B.J., Jardine, P.A., and van Doorn, T., Laser scanned agarose gel

sections for radiation field mapping. Int J Radiat Oncol Biol Phys, 1996.

36: p. 175-179.

[59] Appleby, A. and Leghrouz, A., Imaging of radiation dose by visible colour

development in ferrous agarose xylenol orange gels. Med Phys, 1991. 18:

p. 309-312.

[60] Kelly, R.U., Jordan, K.J., and Battista, I., Optical CT reconstruction of 3D

dose distributions using the ferrous benzoic-xylenol (FBX) gel dosimeter.

Med Phys, 1998. 25: p. 1741-1750.

[61] Olsson, L.E., Petersson, S., Ahlgren, L., and Mattsson, S., Ferrous

sulphate gels for determination of absorbed dose distributions using MRI

techniques: Basic studies. Phys Med Biol, 1989. 34: p. 43-52.

[62] Olsson, L.E., Fransson, A., Ericsson, A., and Mattsson, S., MR imaging of

absorbed dose distributions from radiotherapy using ferrous sulphate gels.

Phys Med Biol, 1990. 35: p. 1623-1631.

[63] Olsson, L.E., Radiation Dosimetry using Magnetic Resonance Imaging, in

Department of Radiation Physics. 1990, Lund University: Malmo,

Sweden.

50

[64] Schulz, R.J., deGuzman, A.F., Nguyen, D.B., and Gore, J.C., Dose-

response curves for Fricke-infused agarose gels as obtained by nuclear

magnetic resonance. Phys Med Biol, 1990. 35(12): p. 1611-22.

[65] Hazle, J.D., Hefner, L., Nyerick, C.E., et al., Dose-response characteristics

of a ferrous-sulphate-doped gelatin system for determining radiation

absorbed dose distributions by magnetic resonance imaging (FeMRI).

Phys Med Biol, 1991. 36: p. 1117-1125.

[66] Schreiner, L.J., Parker, W., Evans, M.D.C., et al., Brachytherapy dose

verification using MRI and Fricke gelatin: a quality assurance technique.

Proc. Of the Conf. Brachytherapy: Today and the Future, 1994, Scotsdale

AZ, 127-133.

[67] Duzenli, C., Sloboda, R., Robinson, D., Material properties of gelatin

affecting spin-spin relaxation rate in the ferrous sulphate gel NMR

dosimeter. Phys Med Biol, 1994. 39: p. 1577-1592.

[68] Chu, K.C., Jordan, K.J., Battista, J.J., Van Dyk, J., and Rutt, B.K.,

Polyvinyl alcohol-Fricke hydrogel and cryogel: two new gel dosimetry

systems with low Fe3+ diffusion. Phys Med Biol, 2000. 45(4): p. 955-69.

[69] Baldock, C., Harris, P.J., Piercy, A.R., and Healy, B., Experimental

determination of the diffusion coefficient in two- dimensions in ferrous

sulphate gels using the finite element method. Australas Phys Eng Sci

Med, 2001. 24(1): p. 19-30.

[70] Healy, B.J., Zahmatkesh, M.H., Nitschke, K.N., and Baldock, C., Effect of

saccharide additives on response of ferrous-agarose-xylenol orange

radiotherapy gel dosimeters. Med Phys, 2003. 30(9): p. 2282-2291.

[71] Schreiner, L.J., Gel Dosimetry: motivation and historical foundations, in

DosGel'99 Proceedings of the 1st International Workshop on Radiation

Therapy Gel Dosimetry, Schreiner, L.J., Editor. 1999, Canadian

Organisation of Medical Physicists: Lexington, Kentucky.

[72] Back, S.A.J. and L.E. Olsson. Ferrous gels (FeGel) and MRI for absorbed

dose measurements: 15 years of development. in DosGel'99 Proceeings of

the 1st International Workshop on Radiation Therapy Gel Dosimetry.

1999. Lexington, Kentucky: Canadian Organisation of Medical Physicists.

51

[73] Schreiner, L.J., Review of Fricke gel dosimeters, in Third International

Conference on Radiotherapy Gel Dosimetry, De Deene, Y. and Baldock,

C., Editors. 2004, J Phys Conf Ser 3: Ghent. p. 9-21.

[74] Maryanski, M.J., J.C. Gore, R.P. Kennan, and R.J. Schulz, NMR

relaxation enhancement in gels polymerized and cross-linked by ionizing

radiation: a new approach to 3D dosimetry by MRI. Magn Reson Imaging,

1993. 11(2): p. 253-8.

[75] Dole, M., The Radiation Chemistry of Macromolecules Vol 1 and 2. 1973,

New York: Academic Press.

[76] Platzer, N., Irradiation of Polymers. 1967, Washington, DC: American

Chemical Society.

[77] Wilson, J.E., Radiation Chemistry of Monomers, Polymers and Plastics.

1974, New York: Marcel Dekker Inc.

[78] Charlesby, A., Radiation Effects in Materials. 1960, Oxford: Pergamon

Press.

[79] Audet, C. The NMR relaxometry of radiotherapy gel dosimeters. in

DosGel'99 Proceedings of the 1st International Workshop on Radiation

Therapy Gel Dosimetry. 1999. Lexington, Kentucky: Canadian

Organization of Medical Physicists.

[80] Lepage, M., A.K. Whittaker, L. Rintoul, S.A. Back, and C. Baldock, The

relationship between radiation-induced chemical processes and transverse

relaxation times in polymer gel dosimeters. Phys Med Biol, 2001. 46(4): p.

1061-74.

[81] Gochberg, D.F., P.M. Fong, and J.C. Gore, Studies of magnetization

transfer and relaxation in irradiated polymer gels--interpretation of MRI-

based dosimetry. Phys Med Biol, 2001. 46(3): p. 799-811.

[82] Maryanski, M.J., R.J. Schulz, G.S. Ibbott, J.X. Gatenby, D. Horton, and

J.C. Gore, Magnetic resonance imaging of radiation dose distributions

using a polymer-gel dosimeter. Phys Med Biol, 1994. 39: p. 1437-1455.

[83] Maryanski, M.J., J.C. Gore, and R.J. Schulz, US Patent: Three-

Dimensional Detection, Dosimetry and Imaging of an Energy Field by

Formation of a Polymer in a Gel, in US Patented Number 5,321,357. 1994:

United States.

52

[84] Maryanski, M.J., Ibbott, G.S., Schulz, R.J., Xie, J., and Gore, J.C.,

Magnetic resonance imaging of radiation dose distributions in tissue-

equivalent polymer-gel dosimeters, in Proc Soc Magn Reson Med. 1994.

p. 204.

[85] Maryanski, M.J., Y.Z. Zastavker, and J.C. Gore, Radiation dose

distributions in three dimensions from tomographic optical density

scanning of polymer gels: II. Optical properties of the BANG polymer gel.

Phys Med Biol, 1996. 41(12): p. 2705-17.

[86] Gore, J.C., M. Ranade, M.J. Maryanski, and R.J. Schulz, Radiation dose

distributions in three dimensions from tomographic optical density

scanning of polymer gels: I. Development of an optical scanner. Phys Med

Biol, 1996. 41(12): p. 2695-704.

[87] Knisely, J.P.S., L. Liu, M.J. Maryanski, M. Ranade, R.J. Schulz, and J.C.

Gore, Three-dimensional dosimetry for complex stereotactic radiosurgery

using a tomographic optical density scanner and BANG polymer gels.

Radiosurgery, 1997. 2: p. 251-260.

[88] Pappas, E., T. Maris, A. Angelopoulos, M. Paparigopoulou, L. Sakelliou,

P. Sandilos, S. Voyiatzi, and L. Vlachos, A new polymer gel for magnetic

resonance imaging (MRI) radiation dosimetry. Phys Med Biol, 1999.

44(10): p. 2677-84.

[89] Pappas, E., I. Seimenis, A. Angelopoulos, P. Georgolopoulou, M.

Kamariotaki-Paparigopoulou, T. Maris, L. Sakelliou, P. Sandilos, and L.

Vlachos, Narrow stereotactic beam profile measurements using N-

vinylpyrrolidone based polymer gels and magnetic resonance imaging.

Phys Med Biol, 2001. 46(3): p. 783-97.

[90] Murphy, P.S., V.P. Cosgrove, M.O. Leach, and S. Webb, A modified

polymer gel for radiotherapy dosimetry: assessment by MRI and MRS.

Phys Med Biol, 2000a. 45(11): p. 3213-23.

[91] Murphy, P.S., V.P. Cosgrove, A.J. Schwarz, S. Webb, and M.O. Leach,

Proton spectroscopic imaging of polyacrylamide gel dosimeters for

absolute radiation dosimetry. Phys Med Biol, 2000. 45(4): p. 835-45.

[92] Hepworth, S.J., M.O. Leach, and S.J. Doran, Dynamics of polymerization

in polyacrylamide gel (PAG) dosimeters: (II) modeling oxygen diffusion.

Phys Med Biol, 1999. 44(8): p. 1875-84.

53

[93] Fong, P.M., Keil, D.C., Does, M.D., and Gore, J.C., Polymer gels for

magnetic resonance imaging of radiation dose distributions at normal

room atmosphere. Phys Med Biol, 2001. 46(12): p. 3105-13.

[94] Venning, A.J., Nitschke, K., Keall, P.J., and Baldock, C., Radiological

properties of normoxic polymer gel dosimeters. Med Phys, 2005. 32(4): p.

1047-1053.

[95] Venning, A.J., Hill, B., Brindha, S., Healy, B.J., and Baldock, C.,

Investigation of the PAGAT polymer gel dosimeter using magnetic

resonance imaging. Phys Med Biol, 2005. 50: p. 3875-3888.

[96] Brindha, S., Venning, A.J., Hill, B., and Baldock, C., Experimental study

of attenuation properties of normoxic polymer gel dosimeters. Phys Med

Biol, 2004. 49: p. N353-361.

[97] Venning, A.J., Mather, M.L., and Baldock, C., Investigation of vacuum

pumping on the dose response of the MAGAS normoxic polymer gel

dosimeter. Australas Phys Eng Sci Med, 2005. 28(2): p. 105-110.

[98] Baldock, C., R.P. Burford, N. Billingham, G.S. Wagner, S. Patval, R.D.

Badawi, and S.F. Keevil, Experimental procedure for the manufacture and

calibration of polyacrylamide gel (PAG) for magnetic resonance imaging

(MRI) radiation dosimetry. Phys Med Biol, 1998. 43(3): p. 695-702.

[99] Haraldsson, P., S.A. Back, P. Magnusson, and L.E. Olsson, Dose response

characteristics and basic dose distribution data for a polymerization-based

dosemeter gel evaluated using MR. Br J Radiol, 2000. 73(865): p. 58-65.

[100] Seymour, R.B., C.E. Carraher, and R.B. Seymour, Seymour/Carraher's

polymer chemistry : an introduction. 4th , rev. and expanded ed.

Undergraduate chemistry ; 13. 1996, New York: M. Dekker,. xxvii, 688.

[101] Salomons, G.J., Park, Y.S., McAuley, K.B., Schreiner, L.J., Temperature

increases associated with polymerization of irradiated PAG dosimeters.

Phys Med Biol, 1992. 47: p. 1435-1448.

[102] Maryanski, M.J., C. Audet, and J.C. Gore, Effects of crosslinking and

temperature on the dose response of a BANG polymer gel dosimeter. Phys

Med Biol, 1997. 42(2): p. 303-11.

[103] De Deene, Y., C. De Wagter, B. Van Duyse, S. Derycke, W. De Neve, and

E. Achten, Three-dimensional dosimetry using polymer gel and magnetic

54

resonance imaging applied to the verification of conformal radiation

therapy in head-and-neck cancer. Radiother Oncol, 1998. 48(3): p. 283-91.

[104] Baldock, C., R.P. Burford, N. Billingham, D. Cohen, and S.F. Keevil,

Polymer gel composition in MRI dosimetry. Med Phys, 1996. 23: p. 1070.

[105] Fuxman, A.M., McAuley, K.B., Schreiner, L.J., Modeling of free-radical

crosslinking copolymerization of acrylamide and N,N’-

methylenebis(acrylamide) for radiation dosimetry. Macromol Theory

Simul, 2003. 12: p. 647-662.

[106] Kennan, R.P., K.A. Richardson, J. Zhong, M.J. Maryanski, and J.C. Gore,

The effects of cross-link density and chemical exchange on magnetisation

transfer in polyacrylamide gels. J Magn Reson, 1996. 110(B): p. pp. 267 -

277.

[107] McJury, M., M. Oldham, M.O. Leach, and S. Webb, Dynamics of

polymerization in polyacrylamide gel (PAG) dosimeters: (I) ageing and

long-term stability. Phys Med Biol, 1999b. 44(8): p. 1863-73.

[108] De Deene, Y., P. Hanselaer, C. De Wagter, E. Achten, and W. De Neve,

An investigation of the chemical stability of a monomer/polymer gel

dosimeter. Phys Med Biol, 2000. 45(4): p. 859-78.

[109] De Deene, Y., Fundamentals of MRI measurements, in Proceedings of 2nd

International Conference on Radiotherapy Gel Dosimetry, Baldock, C. and

De Deene, Y., Editors. 2001, Queensland University of Technology:

Brisbane, Australia.

[110] Hasson, B.F., Chemical dosimetry in the near-zone of brachytherapy

sources. Med Phys, 1996. 23: p. 803.

[111] De Deene, Y., De Wagter, C., Van Duyse, B., Derycke, S., Mersseman,

B., De Gersem, W., Voet, T., Achten, E., and De Neve, W., Validation of

MR-based polymer gel dosimetry as a preclinical three dimensional

verification tool in conformal radiotherapy. Magn Reson Med, 2000.

43(1): p. 116-25.

[112] Ibbott, G.S., Maryanski, M.J., Eastman, P., Holcomb, S.D., Zhang, Y.,

Avison, R.G., Sanders, M., and Gore, J.C., Three-dimensional

visualization and measurement of conformal dose distributions using

magnetic resonance imaging of BANG polymer gel dosimeters. Int J

Radiat Oncol Biol Phys, 1997. 38(5): p. 1097-103.

55

[113] Kaurin, D.G., Maryanski, M.J., Duggan, D.M., Morton, K.C., and Coffey,

C.W., Use of polymer gel dosimetry, pelvic phantom and virtual

simulation to verify setup and calculated three-dimensional dose

distributions for a prostate treatment, in Proceedings of 1st International

Workshop on Radiation Therapy Gel Dosimetry, Schreiner, L.J., Editor.

1999, Canadian Organization of Medical Physicists: Lexington, Kentucky.

p. 184-189.

[114] De Deene, Y., De Wagter, C., De Neve, W., and Achten, E., Artefacts in

multi-echo T2 imaging for high-precision gel dosimetry: I. Analysis and

compensation of eddy currents. Phys Med Biol, 2000. 45(7): p. 1807-23.

[115] De Deene, Y., De Wagter, C., De Neve, W., and Achten, E., Artefacts in

multi-echo T2 imaging for high-precision gel dosimetry: II. Analysis of

B1-field inhomogeneity. Phys Med Biol, 2000. 45(7): p. 1825-39.

[116] De Deene, Y. and De Wagter, C., Artefacts in multi-echo T2 imaging for

high-precision gel dosimetry: Effects of temperature drift during scanning.

Phys Med Biol, 2001. 46: p. 2697-2711.

[117] Baldock, C., Lepage, M., Back, S.A.J., Murry, P.J., Jayasekera, P.M.,

Porter, D., and Kron, T., Dose resolution in radiotherapy polymer gel

dosimetry: effect of echo spacing in MRI pulse sequence. Physics in

Medicine and Biology, 2001. 46: p. 449-460.

[118] De Deene, Y. and Baldock, C., Optimization of multiple spin-echo

sequences for 3D polymer gel dosimetry. Phys Med Biol, 2002. 47: p.

3117-3141.

[119] Hilts, M., Audet, C., Duzenli, C., and Jirasek, A., Polymer gel dosimetry

using x-ray computed tomography: a feasibility study. Phys Med Biol,

2000. 45(9): p. 2559-71.

[120] Trapp, J.V., Back, S.A.J., Lepage, M., Michael, G., and Baldock, C.,

Investigation of the x-ray computed tomography absorbed dose response

characteristics of various polymer gel dosimeter compositions. Phys Med

Biol, 2001. 46: p. 2939-2951.

[121] Trapp, J.V., Michael, G., De Deene, Y., and Baldock, C., Attenuation of

diagnostic energy photons by polymer gel dosimeters. Phys Med Biol,

2002. 47: p. 4247-4258.

56

[122] Audet, C., Hilts, M., Jirasek, A., and Duzenli, C., CT gel dosimetry

technique: Comparison of a planned and measured 3D stereotactic dose

volume. J Appl Clin Med Phys, 2002. 3: p. 110-118.

[123] Hill, B., Venning, A.J., and Baldock, C., Use of normoxic polymer gel

dosimeters for measuring diagnostic doses on CT scanners, in Third

International Conference on Radiotherapy Gel Dosimetry, De Deene, Y.

and Baldock, C., Editors. 2004, J Phys Conf Ser: Ghent. p. 224-227.

[124] Hill, B., Venning, A., and Baldock, C., The dose response of normoxic

polymer gel dosimeters measured using X-ray CT. Br J Radiol, 2005. 78:

p. 623-630.

[125] Hill, B., Venning, A.J., and Baldock, C., A preliminary study of the novel

application of normoxic polymer gel dosimeters for the measurement of

CTDI on diagnostic x-ray CT scanners. Med Phys, 2005. 32(6): p. 1589-

1597.

[126] Oldham, M., Siewerdsen, J.H., Shetty, A., and Jaffray, D.A., High

resolution gel-dosimetry by optical-CT and MR scanning. Med Phys,

2001. 23: p. 699-705.

[127] Maryanski, M.J. and Ranade, M.K., Laser micro-beam scanning of

dosimetry gels. Proc SPIE Int Soc Optic Eng, 2001. 4320: p. 764-767.

[128] Xu, Y., Wuu, C., and Maryanski, M.J., An algorithm for optimized CT

scanning of gel dosimeters. Med Phys, 2002. 29: p. 1205.

[129] Krstelj, V., Ultrasonic testing of polymer materials. Insight, 1996. 38: p.

14-20.

[130] Luprano, V.A.M., Ramires, P.A., Montagna, G., and Milella, E., Non-

destructive characterization of hydrogels. J Mater Sci Mater Med, 1997. 8:

p. 175-178.

[131] Taylor, K.J.W. and Wells, P.N.T., Tissue characterization. Ultrasound in

Med and Biol, 1989. 15: p. 421-428.

[132] Maffezzoli, A., Luprano, V.A., Montagna, G., Esposito, F., and Nicolais,

L., Ultrasonic wave attenuation during water sorption in poly(2-

hydroxyethyl methacrylate) hydrogels. Poly Eng Sci, 1996. 36: p. 1832-

1838.

57

[133] Goldshteins, A.Y., Dolatsis, A., and Zavitska, R.F., Ultrasonic wave

propagation in a polymerizing medium. Mekhanika Polimerov, 1974. 2: p.

204-208.

[134] Zahran, R.R., Kandeil, A.Y., Higazy, A.A., and Kassem, M.E., Ultrasonic

and thermal properties of gamma irradiated low-density polyethylene. J

Appl Poly Sci, 1993. 49: p. 1291-1297.

[135] Zahran, R.R., Effects of gamma irradiation on the ultrasonic and structural

properties of polyoxymethylene. Materials Letters, 1998. 37: p. 83-89.

[136] Maryanski, M.J., Radiation-sensitive polymer gels: properties and

manufacturing, in Proceedings of the First International Workshop on

Radiation Therapy Gel Dosimetry. 1999, Canadian Organsiation of

Medical Physicists: Lexington, Kentucky.

[137] Mather, M.L., Whittaker, A.K., and Baldock, C., Ultrasound evaluation of

polymer gel dosimeters. Phys Med Biol, 2002. 47(9): p. 1449-58.

[138] Baldock, C., Rintoul, L., Keevil, S.F., Pope, J.M., and George, G.A.,

Fourier transform Raman spectroscopy of polyacrylamide gels (PAGs) for

radiation dosimetry. Phys Med Biol, 1998. 43(12): p. 3617-27.

[139] Lepage, M., Whittaker, A.K., Rintoul, L., Back, S.A., and Baldock, C.,

Modelling of post-irradiation events in polymer gel dosimeters. Phys Med

Biol, 2001. 46(11): p. 2827-39.

[140] Jirasek, A. and Duzenli, C., Relative effectiveness of polyacrylamide gel

dosimeters applied to proton beams: Fourier transform Raman

observations and track structure calculations. Med Phys, 2002. 29(4): p.

569-77.

[141] Jirasek, A.I., Duzenli, C., Audet, C., and Eldridge, J., Characterization of

monomer/crosslinker consumption and polymer formation observed in FT-

Raman spectra of irradiated polyacrylamide gels. Phys Med Biol, 2001.

46(1): p. 151-65.

[142] Rintoul, L., Lepage, M., and Baldock, C., Radiation dose distribution in

polymer gels by Raman spectroscopy. Appl Spectrosc, 2003. 57(1): p. 51-

7.

[143] Gustavsson, H., Back, S.A.J., Lepage, M., Rintoul, L., and Baldock, C.,

Development and optimization of a 2-hydroxyethylacrylate MRI polymer

gel dosimeter. Phys Med Biol, 2004. 49: p. 227-241.

58

[144] Duzenli, C. and Robinson, D., Correcting for rf inhomogeneities in

multiecho pulse sequence MRI dosimetry. Med Phys, 1995. 22(10): p.

1645-50.

[145] Baustert, I.C., Oldham, M., Smith, T.A., Hayes, C., Webb, S., and Leach,

M.O., Optimized MR imaging for polyacrylamide gel dosimetry. Phys

Med Biol, 2000. 45(4): p. 847-58.

[146] Lepage, M., Tofts, P.S., Back, S.A., Jayasekera, P.M., and Baldock, C.,

Simple methods for the correction of T2 maps of phantoms. Magn Reson

Med, 2001. 46(6): p. 1123-9.

[147] McJury, M., Tapper, P.D., Cosgrove, V.P., Murphy, P.S., Griffin, S.,

Leach, M.O., Webb, S., and Oldham, M., Experimental 3D dosimetry

around a high-dose-rate clinical 192Ir source using a polyacrylamide gel

(PAG) dosimeter. Phys Med Biol, 1999. 44(10): p. 2431-44.

[148] Lepage, M., Whittaker, A.K., Rintoul, L., and Baldock, C., 13C-NMR,

1H-NMR, and FT-Raman study of radiation-induced modifications in

radiation dosimetry polymer gels. Journal of Applied Polymer Science,

2001. 79: p. 1572-1581.

[149] Brandl, M. and Haase, A., Molecular diffusion in NMR microscopy. J

Magn Reson, 1994. 103(Series B): p. 162-167.

[150] Carr, H.Y. and Purcell, E.M., Effects of diffusion on free precession in

nuclear magnetic resonance experiments. Physical Review, 1954. 94: p.

630-638.

[151] Mattiello, J., Basser, P.J., LeBihan, D., Analytical expressions for the b

matrix in NMR diffusion imaging and spectroscopy. Journal of Magnetic

Resonance, 1994. 108(Series A): p. 131-141.

[152] Karger, J., Pfeifer, H., Heink, W., Principles and applications of self-

diffusion measurements by nuclear magnetic resonance. Advances in

Magnetic Resonance, 1988. 12: p. 1 - 89.

[153] Stejskal, E.O. and Tanner, J.E., Spin diffusion measurements: spin echoes

in the presence of a time-dependent field gradient. J Chem Phys, 1965. 42:

p. 288 - 292.

[154] Ertl, A., Berg, A., Zehetmayer, M., and Frigo, P., High-resolution dose

profile studies based on MR imaging with polymer BANG(TM) gels in

59

stereotactic radiation techniques. Magn Reson Imaging, 2000. 18(3): p.

343-9.

[155] Berg, A., Ertl, A., and Moser, E., High-resolution polymer gel dosimetry

by parameter selective MR- microimaging on a whole body scanner at 3T.

Med Phys, 2001. 28(5): p. 833-43.

[156] Hepworth, S.J., McJury, M., Oldham, M., Morton, E.J., and Doran, S.,

Dose mapping of inhomogeneities positioned in radiosensitive polymer

gels. Nuclear Instruments and Methods in Physics Research, 1999.

422(A): p. 756-760.

[157] Farajollahi, A.R., Bonnett, D.E., Tattam, D., and Green, S., The potential

use of polymer gel dosimetry in boron neutron capture therapy. Phys Med

Biol, 2000. 45(4): p. N9-14.

[158] Ramm, U., Weber, U., Bock, M., Kramer, M., Bankamp, A., Damrau, M.,

Thilmann, C., Bottcher, H.D., Schad, L.R., and Kraft, G., Three-

dimensional BANG gel dosimetry in conformal carbon ion radiotherapy.

Phys Med Biol, 2000. 45(9): p. N95-102.

[159] Soares, C., Dosimetric Issues in Vascular Brachytherapy (TG-43/60), in

Intrascular Brachytherapy - Fluoroscopically Guided Interventions, Balter,

S., Chan, R.C., and Shope, T.B., Editors. 2002, American Association of

Physicists in Medicine: Montreal. p. 321-372.

[160] Schreiner, L.J., Crooks, I., Evans, M.D.C., Keller, B.M., Parker, W.A.,

Imaging of HDR brachytherapy dose distributions using NMR Fricke-

gelatin dosimetry. Magn Reson Imaging, 1994. 12: p. 901-907.

[161] Farajollahi, A.R., Bonnett, D.E., Ratcliffe, A.J., Aukett, R.J., and Mills,

J.A., An investigation into the use of polymer gel dosimetry in low dose

rate brachytherapy. Br J Radiol, 1999. 72(863): p. 1085-92.

[162] Hafeli, U.O., Roberts, W.K., Meier, D.S., Ciezki, J.P., Pauer, G.J., Lee,

E.J., and Weinhous, M.S., Dosimetry of a W-188/Re-188 beta line source

for endovascular brachytherapy. Med Phys, 2000. 27(4): p. 668-75.

[163] Knutsen, B.H., Skretting, A., Hellebust, T.P., and Olsen, D.R.,

Determination of 3D dose distribution from intracavitary brachytherapy of

cervical cancer by MRI of irradiated ferrous sulphate gel. Radiother

Oncol, 1997. 43(2): p. 219-27.

60

[164] Ibbott, G.S. and Nath, R., Dose-rate constant for Imagyn 125I

brachytherapy source. Med Phys, 2001. 28(4): p. 705.

[165] Baras, P., Seimenis, I., Kipouros, P., Papagiannis, P., Angelopoulos, A.,

Sakelliou, L., Pappas, E., Baltas, D., Karaiskos, P., Sandilos, P., and

Vlachos, L., Polymer gel dosimetry using a three-dimensional MRI

acquisition technique. Med Phys, 2002. 29(11): p. 2506-16.

[166] Papagiannis, P., Pappas, E., Kipouros, P., Angelopoulos, A., Sakelliou, L.,

Baras, P., Karaiskos, P., Seimenis, I., Sandilos, P., and Baltas, D.,

Dosimetry close to an 192Ir HDR source using N-vinylpyrrolidone based

polymer gels and magnetic resonance imaging. Med Phys, 2001. 28(7): p.

1416-26.

[167] Wuu, C.S., Schiff, P., Maryanski, M.J., Liu, T., Borzillary, S., and

Weinberger, J., Dosimetry study of Re-188 liquid balloon for intravascular

brachytherapy using polymer gel dosimeters and laser-beam optical CT

scanner. Med Phys, 2003. 30(2): p. 132-7.

[168] De Deene, Y., Reynaert, N., and De Wagter, C., On the accuracy of

monomer/polymer gel dosimetry in the proximity of a high-dose-rate

192Ir source. Phys Med Biol, 2001. 46(11): p. 2801-25

[169] Fuxman, A.M., McAuley, K.B., Schreiner, L.J., Modelling of

polyacrylamide gel dosimeters with spatially non-uniform radiation dose

distribution. Chem Eng Sci, 2005. 60: p. 1277-1293.

[170] Amin, M.N., Horsfield, M.A., Bonnett, D.E., Dunn, M.J., Poulton, M., and

Harding, P.F., A comparison of polyacrylamide gels and radiochromic

film for source measurements in intravascular brachytherapy. Br J Radiol,

2003. 76(911): p. 824-31.

[171] Gore, E., Gillin, M.T., Albano, K., and Erickson, B., Comparison of high

dose-rate and low dose-rate dose distributions for vaginal cylinders. Int J

Radiat Oncol Biol Phys, 1995. 31(1): p. 165-70.

[172] Baldock, C., Murry, P., and Kron, T., Uncertainty analysis in polymer gel

dosimetry. Phys Med Biol, 1999. 44(11): p. N243-6.

[173] (IAEA), I.A.E.A., Absorbed dose determination in photon and electron

beams, in An international code of practice for dosimetry. 2000, Technical

Report Series 400 IAEA: Vienna.

61

[174] Oldham, M., McJury, M., Baustert, I.B., Webb, S., and Leach, M.O.,

Improving calibration accuracy in gel dosimetry. Phys Med Biol, 1998.

43(10): p. 2709-20.

[175] Low, D.A., Markman, J., Dempsey, J.F., Mutic, S., Oldham, M.,

Venkatesan, R., Haacke, E.M., and Purdy, J.A., Noise in polymer gel

measurements using MRI. Med Phys, 2000. 27(8): p. 1814-7.

[176] Low, D.A., Dempsey, J.F., Venkatesan, R., Mutic, S., Markman, J., Mark

Haacke, E., and Purdy, J.A., Evaluation of polymer gels and MRI as a 3-D

dosimeter for intensity- modulated radiation therapy. Med Phys, 1999.

26(8): p. 1542-51.

[177] Magnusson, P., Back, S.A., and Olsson, L.E., MRI image plane

nonuniformity in evaluation of ferrous sulphate dosimeter gel (FeGel) by

means of T1-relaxation time. Magn Reson Imaging, 1999. 17(9): p. 1357-

70.

[178] Watanabe, Y., Variable transformation of calibration equations for

radiation dosimetry. Phys Med Biol, 2005. 50(6): p. 1221-34.

62

63

Chapter 3

Dose-response Stability and Integrity of

the Dose Distribution of Various

Polymer Gel Dosimeters

Full-text available at:

http://dx.doi.org/10.1088/0031-9155/47/14/307

halla
This article is not available online. Please consult the hardcopy thesis available from the QUT Library

77

Chapter 4

A Basic Study of Some Normoxic

Polymer Gel Dosimeters

Full-text available at: http://dx.doi.org/10.1088/0031-9155/47/19/301

halla
This article is not available online. Please consult the hardcopy thesis available from the QUT Library

101

Chapter 5

The Effects of Molecule Self-Diffusion

of Water on Quantitative MRI

Measurements in High-Resolution

Polymer Gel Dosimetry

Full-text available at:

http://dx.doi.org/10.1088/0031-9155/48/18/306

halla
This article is not available online. Please consult the hardcopy thesis available from the QUT Library

119

Chapter 6

A Study of a Normoxic Polymer Gel Dosimeter

comprising methacrylic acid, gelatin, and

Tetrakis (Hydroxymethyl) Phosphonium

Chloride (MAGAT)

Full-text available at: http://dx.doi.org/10.1016/j.apradiso.2005.03.014

halla
This article is not available online. Please consult the hardcopy thesis available from the QUT Library

135

Chapter 7

High-Resolution Gel Dosimetry of a

HDR Brachytherapy Source Using

Normoxic Polymer Gels

Full-text available at: http://dx.doi.org/10.1016/j.nima.2006.05.167

halla
This article is not available online. Please consult the hardcopy thesis available from the QUT Library

147

Chapter 8

General Discussion

Radiation therapy has seen significant advances in the effective irradiation of

tumours using methods that can conform the irradiation pattern to the affected

target tissue more precisely than previously possible [1]. The approach is to

deliver the maximum dose to the target volume whilst sparing the healthy

surrounding tissues. Conformal radiotherapy, IMRT, dynamic multi-leaf

collimation and on-line portal imaging have all aided this development and hence

treatment delivery can now comprise very complex geometries. Likewise,

brachytherapy treatments can now be achieved with sub-millimeter scales using

high dose gradients. Currently dose distributions are calculated using treatment

planning systems that incorporate complex computer algorithms. With the

increase in complexity there needs to be an effective means by which to verify

that the planned treatment is delivered through direct measurements. Polymer gel

dosimetry has been shown to be an effective technique for accurate 3D dosimetry

with high spatial resolution. It is capable of integrating dose throughout a

dynamic delivery without perturbing the radiation beam [2,3]. One key clinical

limitation to polymer dosimeters is lack of ease of manufacture and use due to

strict hypoxic requirements that are necessary to prevent oxygen from infiltrating

the gel, hence inhibiting the polymerization process that is used to correlate to

absorbed dose to produce dose distributions [4].

The work of this thesis has involved the investigation of normoxic polymer gel

dosimeters with high-resolution MRI. Normoxic polymer gel dosimeters include

an anti-oxidant in the formulation to bind dissolved oxygen preventing it from

inhibiting the polymerization process [5]. The first aspect of this study involved a

comprehensive investigation of different formulations of polymer and normoxic

148

polymer gel dosimeters, in order to determine the optimal composition for use in

polymer gel dosimetry. The second aspect of this research was to investigate

high-resolution MRI for normoxic polymer gel dosimetry with high spatial

resolution. The key application of this work has been intravascular brachytherapy

(IVBT) that aims to deliver doses to the walls of arteries to prevent restenosis

following percutaneous transluminal coronary angioplasty (PTCA). Annually,

there are 450 000 coronary angioplasties performed on de novo and restonotic

lesions in the United States alone [6,7]. Approximately 70 to 80 % of cases

require stent implants to reduce restenosis and 15 to 35 % of these patients still

develop restenosis [6,8]. IVBT has been found to be effective in the treatment of

restenosis.

The following aspects were studied to enable a full assessment of normoxic

polymer gel dosimeters using high-resolution MRI as an effective medical

dosimeter:

• An exploration of different formulations of polymer gel dosimeters and

normoxic polymer gel dosimeters to obtain optimal characteristics for use

in gel dosimetry.

• The determination of the physical and chemical properties of various

formulations of normoxic polymer gel dosimeters on dose maps obtained

using gel dosimetry with MRI.

• An examination of the chemical mechanisms that occur when normoxic

polymer gel dosimeters are manufactured, irradiated and subsequently

evaluated.

• The investigation of high-resolution magnetic resonance imaging and its

potential for use in evaluating polymer gel dosimeters. This involved the

commissioning a micro-imaging magnetic resonance spectrometer for

high-resolution imaging of polymer gel dosimeters.

149

• An examination of the effects that high gradient strengths combined with

molecular diffusion, have on R2 and hence dose maps produced using

high-resolution MRI. It was found that with higher resolution MRI, there

were significant changes in the measured R2. Software was developed

that could use differences in the MRI parameters applied during imaging

to calculate quantitatively the effects of molecular self-diffucion of water

on the measured R2 within a polymer gel dosimeter. A listing of the

developed software is included in appendix A.

• The application and evaluation of normoxic polymer gel dosimeters

irradiated using high-resolution MRI on typical brachytherapy deliveries.

In particular a point source and line irradiation patterns were explored and

compared with predictions from computer treatment planning software.

8.1 The Principal Significance of the Findings

8.1.1 Analyzing and Optimizing Polymer Gel Dosimeter Formulations

Polymer gel dosimeters have the potential to provide accurate, integrated 3D maps

of dose distributions with high spatial resolution. However, several different

formulations of polymer gel dosimeter have been investigated and proposed over

the years. In order to be able to compare different polymer gel dosimeters,

specific criteria are used that relate to requirements for clinical dosimetry of

radiation therapy treatment planning. Before polymer and normoxic polymer gel

dosimetry can be implemented in clinical or routine practice, comprehensive

studies must be conducted to explore, in detail, the chemical and physical aspects

of gel dosimetry. This thesis is an investigation of the common formulations of

polymer gel dosimeters and the recently proposed normoxic polymer gel

dosimeters.

The investigation of each polymer gel dosimeter involves a temporal study of the

changes in both the slope and R2-intercept of the R2-dose response curve. The

150

changes in the R2-intercept at zero dose are related to the ageing of the gelatin

network. It was found that the ageing process of the gelatin network was related

to the total concentration of gelatin molecules. An increase in gelatin was also

found to result in a decrease in the characteristic time for the R2-dose sensitivity

(slope) to saturate. However, the total degree of post-irradiation polymerization

or restructuring was found to be unaffected in the long term by the concentration

of gelatin. The ratio of cross-linker monomer to linear monomer was found to

have a significant effect on the stability of the R2-dose sensitivity but was found

not to affect the R2 intercept. Additionally, whilst the R2-dose sensitivity

increased over time for PAG gel dosimeter, it was found to decrease for normoxic

polymer gel dosimeters. The temporal stability of polyacrylamide gel (PAG)

dosimeters was also found to differ from that of normoxic polymer gel

dosimeters. These differences indicate that normoxic polymer gel dosimeters

have a different polymer structure to PAG gel dosimeters.

The spatial stability of the gel was also explored by examining the measured dose

at the dose edge in a gel exposed to a half-beam blocked field. This enabled

monomer diffusion to be measured and hence used as an indicator of stability.

Both the PAG gel dosimeters and the normoxic polymer gel dosimeters were

found to be spatially stable for doses up to 12 Gy, whilst the HEA gel dosimeter

demonstrated an instability in the form of a dose overshoot that varied over time.

The spatial stability of the PAG and normoxic polymer gel dosimeters was found

to be stable with an apparent penumbra between 3.1 mm and 4.4 mm for gels of

different formulation. The normoxic polymer gel dosimeters exhibited a smaller

apparent penumbra (3.2 to 4.3 mm) than the PAG gel dosimeters (4.05 to 4.4

mm), however, the HEA gel exhibited the lowest apparent penumbra at around 3.1

mm. It was concluded that waiting periods of 10 h and 30 h be respected between

irradiating and scanning PAG and normoxic polymer gel dosimeters respectively.

8.1.2 Chemical Properties of Normoxic Polymer Gel Dosimeters

151

The first normoxic polymer gel dosimeter was proposed by Fong et al. in 2001

[5]. It was comprised of methacrylic acid, ascorbic acid, hydroquinone, copper

(II) sulphate and gelatin, and it was called MAGIC. Although specific

concentrations were provided by Fong et al., no explanation or justification was

given for the choice of concentrations. This study involved varying the

concentration of each constituent in order to map a dose response between 0 and 5

Gy. A region of optimal dose response was determined by examining a difference

R2 map between unirridiated and irradiated gels given varying concentration of

each constituent.

Advances in the understanding of the chemical behaviour and properties of

normoxic polymer gel dosimeters have been made through the identification of

the reactions that occur from manufacture to post-irradiation. It was found that

the reaction between two constituents, copper (II) sulphate and ascorbic acid,

leads to the formation of a ascorbate-copper complex. Some initial

polymerization takes place due to the creation of radicals in the ascorbate-copper-

oxygen complex. Ascorbic acid (an anti-oxidant) was found to scavenge and bind

the free oxygen within the gel dosimeter, whilst copper acts as a catalyst in the

oxygen scavenging. With high concentrations of copper (II) sulphate the

polymerization reaction is terminated by a redox reaction in which copper is

reduced. Hydroquinone, when used in low quantities, facilitates the radiation-

induced polymerization reaction, however, at high concentrations it acts as an

inhibitor.

In this thesis, five anti-oxidants were explored for potential use in normoxic

polymer gel dosimetry; only three were found to be effective. The oxygen

scavenging ability of an anti-oxidant could be measured independently and

provided basis for determining the characteristic time for a normoxic polymer gel

dosimeter to become radiation sensitive. Of the three viable anti-oxidants, tetrakis

(hydroxymethyl) phosphonium chloride (THPC) was found to have the highest

reaction rate. THPC was also found to increase the R2-dose sensitivity of the

normoxic polymer gel dosimeter.

152

8.1.3 A Normoxic Polymer Gel Dosimeter Using THPC

Tetrakis (hydroxymethyl) phosphonium chloride (THPC) was found to be a very

aggressive scavenger of dissolved oxygen. Hence, a normoxic polymer gel

dosimeter made using methacrylic acid, gelatin, hydroquinone and THPC (called

MAGAT) was proposed and investigated. In a similar study to section 8.1.2, the

constituents of the MAGAT normoxic polymer gel dosimeter were optimized by

varying their concentrations and exploring the effect on the dose response of the

gel. Firstly, the concentrations of THPC and hydroquinone were varied and

assessed for optimal R2-dose response. Secondly, concentrations of methacrylic

acid and gelatin were also varied to determine the optimal concentration that

produced the highest dose response. The R2-dose response was found to varying

slightly over a 7 day period. The R2-intercept steadily increased over this time,

whilst the R2-dose sensitivity (slope) steadily decreased. These results concurred

with those of the previous study that examined the R2-dose response of normoxic

polymer gel dosimeter made with ascorbic acid as the anti-oxidant.

In additional, the spatial stability of the MAGAT polymer gel dosimeter was

investigated by exposing a phantom to a half-beam blocked field and examining

the dose profile through the boundary between the irradiated and unirradiated

sections of the phantom. The spatial stability of the MAGAT polymer gel

dosimeter was found to exhibit an apparent penumbra of around 3.5 mm.

Previous studies of the normoxic polymer gel dosimeters, found that the

penumbra was between 3.2 mm and 3.7 mm for normoxic polymer gel dosimeter

with different formulations (see section 8.1.1). It was found that the addition of

small quantities of hydroquinone aided in maintaining the spatial stability.

It was concluded that the MAGAT normoxic polymer gel dosimeter exhibited a

R2-dose response that was stable over a 7 day period and spatially stable when

imaged 24 h after irradiation. This concurs with the previous study which

concluded that normoxic polymer gel dosimeters should be scanned about 30 h

after irradiation.

153

8.1.4 Evaluating Polymer Gel Dosimeters using High-Resolution MRI

In order to evaluate polymer gel dosimeters at high-resolution, a 4.7 T magnetic

resonance spectrometer was commissioned for use in scanning calibration vials

and phantoms designed for simulating typical brachytherapy treatments. This

commissioning involved investigating various pulse sequences and the pulse

parameters of various optimized pulse sequences to obtain suitable images. Using

MRI to evaluate polymer gel dosimeters at high-resolution introduces artifacts

into the images that are often negligible in MRI at typical clinical resolutions.

The predominant effect was found to be the self-diffusion through steep magnetic

field gradients of water molecules caused by Brownian motion. This self-

diffusion had a significant effect on measured R2 values at high-resolution. It was

found that changes in the imaging parameters, including echo time, which is often

used to optimize pulse sequences in polymer gel dosimetry [9,10], led to

significant changes in measured R2 values due to the self-diffusion.

Software was written to evaluate the extent of the variation in measured R2 values

for specific parameters of a MRI pulse sequence (see appendix A). The software

is capable of incorporating changes in timing parameters and gradient strengths

and subsequently predicting the variation in measured R2 as a result of the water

self-diffusion given the particular pulse sequence. Gel dosimetry typically

involves imaging a set of calibration vials and a phantom. This is usually done

separately, and thus, any differences in the imaging parameters between the

calibration vials and phantoms may introduce significant differences in the

measured R2 values when scanned at high-resolution. These variations in the

measured R2 values will give rise to significant errors in the measured absorbed

dose. Correction factors can be determined using the software and applied to R2

maps of calibration vials and phantoms thereby adjusting for errors that arise due

to self-diffusion at high-resolution. On a simpler level, keeping the imaging

parameters the same for both the calibration vials and the phantoms prevents any

differences in R2 occuring due to self-diffusion of water molecules in high-

resolution MR imaging.

154

8.1.5 Application of Normoxic Polymer Gel Dosimeters using High-

Resolution MRI to Brachytherapy Treatment Plans

An optimal formulation of a normoxic polymer gel dosimeter was manufactured

and irradiated using a brachytherapy HDR remote afterloader system. The dose

deliveries were set by typical brachytherapy plans that included a single dwell

position simulating a point source irradiation and a line irradiation pattern. A set

of calibration vials was also filled with normoxic polymer gel dosimeter from the

same batch. Post-irradiation, both the calibration vials and the phantoms were

evaluated using high-resolution MRI to produce R2 maps. These R2 maps were

then converted to relative dose maps and compared to the dose distribution

predicted by the computer treatment planning system.

The investigation found differences between the dose distributions measured in

the normoxic polymer gel dosimeter and the predicted dose distributions from the

computer treatment planning system. However, an exploration of various artifacts

revealed that partial volume averaging had a significant effect on the measured

dose given the relatively higher slice thickness compared to in-plane resolution

and the high dose gradients at radial distances close to the source (< 10 mm). An

approximation of the uncertainty due to partial volume effects could be

determined from dose points predicted by the computerized treatment planning

system. These approximations of the amount of variation due to partial volume

effects were used to calculate an adjustment that was applied to measured dose

profiles. Very good agreement was found between the adjusted, normalized dose

distributions measured using the gel dosimeter and dose points predicted by the

computerized treatment planning system. The results indicate that normoxic

polymer gel dosimeters evaluated with high-resolution MRI can be effective in

applications requiring high-resolution such as intravascular brachytherapy.

Likewise they could be the tool of choice for verifying treatment plans involving

complex geometries or new source designs.

155

8.2 Conclusions and Future Work

Polymer gel dosimeters have been developed to map dose distributions resulting

from irradiations such as radiation therapy and brachytherapy treatments.

Radiation therapy and brachytherapy applications that incorporate high dose

gradients, extremely asymmetric fields or complex geometries are difficult to

measure directly using conventional dosimetry techniques. However, polymer gel

dosimetry offers a technique capable of producing 3D integrated dose maps with

high spatial resolution. Whilst polymer gel dosimeters based on polyacrylamide

suffer from the effects of oxygen infiltration, normoxic polymer gel dosimeters

can easily be fabricated on the bench top under normal atmospheric conditions

and show great potential as effective dosimeters.

An integral part of the polymer gel dosimetry process is the evaluation using an

imaging modality such as MRI or optical CT. Applications such as intravascular

brachytherapy, however, require high-resolution dose information due to the small

dimensions of the target volume. Hence, the application of high-resolution MRI

techniques for evaluating normoxic polymer gel dosimeters has been investigated.

At high-resolution, MRI suffers from a number of artifacts, the most significant of

which is the self-diffusion of molecules during scanning.

The aims of this study have been achieved and the main conclusions are as

follows:

• Various formulations of polymer gel dosimeters and normoxic polymer

gel dosimeters have been optimized to achieve specific characteristics

desired for use in radiation therapy and brachytherapy dosimetry. These

included: R2-dose response, temporal stability of the R2 intercept and the

R2-dose sensitivity (slope) of the R2-dose response, spatial stability and

reduced oxygen effects. Optimal formulations included the MAGIC

normoxic polymer gel dosimeter (9 % methacrylic acid, 1 mM ascorbic

acid, 8 % gelatin, 0.01 mM copper(II) sulphate, 0.01 mM hydroquinone).

156

• The chemical mechanisms that take place in normoxic polymer gel

dosimeter formulation have been extensively explored and used to explain

the role that constituents of a specific formulation have. It has been

shown using chemical models that only low concentrations of copper are

effective in catalyzing the oxygen scavenging ability of ascorbic acid.

Hydroquinone in low concentrations was also found to facilitate the

radiation-induced polymerization reaction. This information assists in the

process of developing optimal formulations and explaining phenomena

associated with normoxic polymer gel dosimetry.

• The desired chemical and physical properties listed above were used to

propose a new formulation of normoxic polymer gel dosimeter. This

normoxic polymer gel dosimeter, called MAGAT, uses tetrakis

(hydroxymethyl) phosphonium chloride (THPC) as an anti-oxidant. The

optimal composition of the MAGAT gel was found to be 6 % methacrylic

acid, 6 % gelatin, 10 mM THPC and 0.05 mM hydroquinone based on R2

response. This dosimeter was found to exhibit promising temporal

stability of its R2-dose response and spatial stability.

• High-resolution magnetic resonance imaging has been implemented on a

micro-imaging spectrometer using a pulse sequence optimized for use

with polymer gel dosimetry. A multi-slice, multi-echo CPMG sequence

was found to be suitable for use with gel dosimetry. Small inter-echo

times are essential for imaging normoxic polymer gel dosimeters due to

relatively small characteristic T2 times. Likewise, to achieve high-

resolution imaging, a spectrometer must be capable of large magnetic

gradients.

• Artifacts were found to occur due to scanning at high-resolutions, the most

significant of which was found to be the effect on the measured R2 due to

self-diffusion of water molecules during scanning. Using knowledge of

the gradient strengths applied to a sample, the diffusion weighting could

157

be calculated and the R2 corrected for this effect. Software was

developed to analyze the timing and magnitude of pulses within a pulse

sequence in order to calculate correction factors that could be applied (a

listing is included in appendix A).

• A normoxic polymer gel dosimeter was fabricated and used to verify

typical brachytherapy treatments plans. In particular a line irradiation

pattern and irradiation due to a single dwell position effectively simulating

a point source. Dose distributions were produce using high-resolution

MRI (achieving an in-plane resolution of 0.1055 mm/pixel) and compared

to predictions of dose points from the computerized treatment planning

system.

• Partial volume effects were found to be significant at high resolution or

when large voxel sizes are used that have been exposed to large dose

gradients. By approximating the amount of variation within each voxel

due to partial volume effects, adjustments could be applied to dose

profiles that produced very good agreement between measured and

predicted doses from the computerized treatment planning system.

This thesis is the first study to comprehensively investigate normoxic polymer gel

dosimeters and to apply them to high-resolution brachytherapy applications. The

investigation of the theory behind the chemical kinetics and polymerization

processes that take place in polymer and normoxic polymer gel dosimetry support

the experimental findings. The implementation of normoxic polymer gel

dosimetry in the verification of high-resolution applications has been

demonstrated with good agreement between polymer gel dosimeter dose

measurements and dose calculations using a computerized treatment planning

system. It is therefore concluded that normoxic polymer gel dosimetry using

high-resolution MRI can be effectively applied to the verification of applications

requiring high-resolution, such as intravascular brachytherapy. Likewise,

normoxic polymer gel dosimetry could be used in applications where complex

geometries or severely asymmetric fields exist. Of particular significance is the

potential for normoxic polymer gel dosimetry using high-resolution MRI to be

158

applied to new source designs and more complex treatment plans given the

advances that have occurred in conformal radiation therapy over the past decade.

Future work could include an investigation into the reproducibility of the results.

Of particular interest would be an investigation of increasing resolution in the

MRI evaluation of normoxic polymer gel dosimeters. In this study, a resolution

of 0.1055 × 0.1055 × 2.0 mm3 was used. With smaller slice thickness (< 2 mm),

it would be possible to further reduce the partial volume effect and hence decrease

the uncertainty of the measured doses. Differences between MRI scans of the

calibration tubes and the phantoms may also have an effect on the measure dose

values and hence require further investigation.

Although the anti-oxidants used in this study to reduce the oxygen sensitivity of

polymer gel dosimeters exhibited good results, there are many anti-oxidants that

are yet to be explored for their effectiveness in normoxic polymer gel dosimetry.

The nature of the chemical processes that occur in normoxic polymer gel

dosimetry using these other anti-oxidants would need to be explored in order to

propose alternative formulations. It was noted in this study that the polymer

structure of normoxic polymer gel dosimeters differs from that of polyacrylamide

polymer gel dosimeters. The different structure still requires further exploration

in order to understand the differences in the temporal stability of normoxic

polymer gel dosimeters.

It has been concluded that the use of normoxic polymer gel dosimeters with high-

resolution MRI for the verification of radiation therapy and brachytherapy

treatment plans, is valid and shows great potential as an effective 3D integrating

dosimeter. The ease of manufacture of normoxic polymer gel dosimeters on the

bench top makes them a viable contender for implementation in routine clinical

treatment planning.

159

References

[1] Kron, T., Radiation therapy requirements: what do we expect from gel

dosimetry, in Proceedings of DosGel 2001: The second international

conference on radiotherapy gel dosimetry, Baldock, C. and De Deene, Y.,

Editors. 2001, Queensland University of Technology: Brisbane. p. 2-9.

[2] Maryanski, M.J., Ibbott, G.S., Schulz, R.J., Xie, J., and Gore, J.C.,

Magnetic resonance imaging of radiation dose distributions in tissue-

equivalent polymer-gel dosimeters, in Proc Soc Magn Reson Med. 1994.

p. 204.

[3] Schreiner, L.J., Gel Dosimetry: motivation and historical foundations, in

DosGel'99 Proceedings of the 1st International Workshop on Radiation

Therapy Gel Dosimetry. 1999, Canadian Organisation of Medical

Physicists: Lexington, Kentucky.

[4] Maryanski, M.J., Radiation-sensitive polymer gels: properties and

manufacturing, in Proceedings of the First International Workshop on

Radiation Therapy Gel Dosimetry. 1999, Canadian Organsiation of

Medical Physicists: Lexington, Kentucky.

[5] Fong, P.M., Keil, D.C., Does, M.D., and Gore, J.C., Polymer gels for

magnetic resonance imaging of radiation dose distributions at normal

room atmosphere. Phys Med Biol, 2001. 46(12): p. 3105-13.

[6] Baim, D.S., Cutlip, D.E., Midei, M., Linnemeier, T.J., Schreiber, T., Cox,

D., Kereiakes, D., Popma, J.J., Robertson, L., Prince, R., Lansky, A.J., Ho,

K.K., and Kuntz, R.E., Final results of a randomized trial comparing the

MULTI-LINK stent with the Palmaz-Schatz stent for narrowing in native

coronary arteries. Am J Cardiol, 2001. 87: p. 157-162.

[7] Waksman, R., Vascular Brachytherapy For Restenosis: Unresolved Issues,

in Intravascular Brachtherapy, Fluoroscopically Guided Interventions,

160

Balter, S., Chan, R.C., and Shope, T.B., Editors. 2002, American

Association of Physicists in Medicine: Montreal, Quebec.

[8] Baim, D.S., Cutlip, D.E., O'Shaughnessy, C.D., Hermiller, J.B., Kereiakes,

D., Giambartolomei, A., Katz, S., Lansky, A.J., Fitzpatrick, M., Popma,

J.J., Ho, K.K., Leon, M.B., and Kuntz, R.E., Final results of a randomized

trial comparing the NIR stent to the Palmaz-Schatz stent for narrowing in

native coronary arteries. Am J Cardiol, 2001. 87(152-156).

[9] Baldock, C., Lepage, M., Back, S.A.J., Murry, P.J., Jayasekera, P.M.,

Porter, D., and Kron, T., Dose resolution in radiotherapy polymer gel

dosimetry: effect of echo spacing in MRI pulse sequence. Phys Med Biol,

2001. 46: p. 449-460.

[10] De Deene, Y. and Baldock, C., Optimization of multiple spin-echo

sequences for 3D polymer gel dosimetry. Phys Med Biol, 2002. 47: p.

3117-3141.

161

Appendix A

Listing of the Code for Determining

Variations in R2 due to the Application

of Pulse Sequences during High-

Resolution MRI

The following Java code was written in order to quantitatively determine the

magnitudes of variation in R2 values when imaging a polymer gel dosimeter using

high-resolution MRI. The code uses information about the pulse sequence, in

particular the pulse timing and strengths in order to calculate the self-diffusion

weightings that would be applied to sample during high-resolution MRI imaging.

162

package diffcalc3; import javax.swing.UIManager; import java.awt.*; public class diffcalc3 { boolean packFrame = false; public diffcalc3() { Frame1a frame = new Frame1a("Diffusion Weighting Calculator"); if (packFrame) { frame.pack(); } else { frame.validate(); } Dimension screenSize =

Toolkit.getDefaultToolkit().getScreenSize(); Dimension frameSize = frame.getSize(); if (frameSize.height > screenSize.height) { frameSize.height = screenSize.height; } if (frameSize.width > screenSize.width) { frameSize.width = screenSize.width; } frame.setLocation((screenSize.width - frameSize.width) / 2,

(screenSize.height - frameSize.height) / 2); frame.setVisible(true); } public static void main(String[] args) { try { UIManager.setLookAndFeel(UIManager.getSystemLookAndFeelClassName()); } catch(Exception e) { e.printStackTrace(); } new diffcalc3(); } }

Class: diffcalc3 Author: Christopher Hurley Date: June, 2002

163

public class Echo { public int te = 0; public int echotime = 0; public double height = 0; public boolean enabled = false; public Echo() { } public Echo(int te, int echotime, double height) { this.te = te; this.echotime = echotime; this.height = height; enabled = true; } public void show() { enabled = true; } public void hide() { enabled = false; } } import java.io.*; import java.util.*; import java.awt.*; import java.awt.event.*; import javax.swing.*; public class Frame1a extends JFrame implements ActionListener{ // Constants.... public int nEchoes = 128; public Vector graphData = new Vector(); public int TE = 5050; // in useconds public double Gmax; // in T/m public double gamma; // in rad/s/T public double D; // in m^2/s public double TimeInc = 0.000001; // in useconds public int P0; public int P1; public int aqq;

Class: Frame1a Author: Christopher Hurley Date: June, 2002

Class: Echo Author: Christopher Hurley Date: June, 2002

164

public int d2; public int d3; public int d4; public int d5; public int d12; public int d14; public int d18; public double GROa; public double GROb; public double GSSa; public double GSSb; public double GSSc; public double GSSd; public double GSSe; public double GSSf; public double GPHa; public double GPHb; int c0, c1, c2, c3, c4, c4_5, c5, c6, c7, c8, c8_5, c9, c10, c11,

c12; int cTE; // Cycle TE calculated for repeat section int p0, p1, p2; // Pulse centres for P0 and P1 and Echo (time) double h0 = 3.0; // Arbitrary height for P0 double h1 = 6.0; // Arbitrary height for P1 int inclGrad = 5; // Default to include SS & RO gradients MyInt Loc = new MyInt(5); surface1 canvas = new surface1(); JPanel contentPane; BorderLayout borderLayout1 = new BorderLayout(); JPanel jPanel1 = new JPanel(new BorderLayout()); JTextArea output = new JTextArea(); JMenuBar jMenuBar1 = new JMenuBar(); JMenu jMenu1 = new JMenu(); JMenuItem jMenuItem1 = new JMenuItem(); JMenuItem jMenuItem2 = new JMenuItem(); JMenuItem jMenuItem3 = new JMenuItem(); JMenuItem jMenuItem4 = new JMenuItem(); JMenu jMenu2 = new JMenu(); JMenuItem jMenuItem5 = new JMenuItem(); JMenuItem jMenuItem6 = new JMenuItem(); JMenuItem jMenuItem7 = new JMenuItem(); JMenu jMenu3 = new JMenu(); JMenuItem jMenuItem8 = new JMenuItem(); JMenuItem jMenuItem9 = new JMenuItem(); JMenuItem jMenuItem10 = new JMenuItem(); JMenu jMenu4 = new JMenu(); JMenuItem jMenuItem11 = new JMenuItem(); JMenuItem jMenuItem12 = new JMenuItem(); public Frame1a(String title) { super(title); enableEvents(AWTEvent.WINDOW_EVENT_MASK); try {

165

jbInit(); } catch(Exception e) { e.printStackTrace(); } } private void jbInit() throws Exception { contentPane = (JPanel) this.getContentPane(); contentPane.setLayout(borderLayout1); this.setSize(new Dimension(700, 500)); jMenu1.setText("File"); jMenuItem1.setActionCommand("OpenFile"); jMenuItem1.setText("Open"); jMenuItem1.setAccelerator(javax.swing.KeyStroke.getKeyStroke(79,

0, false)); jMenuItem2.setActionCommand("SaveFile"); jMenuItem2.setText("Save"); jMenuItem2.setAccelerator(javax.swing.KeyStroke.getKeyStroke(83,

0, false)); jMenuItem3.setActionCommand("CloseFile"); jMenuItem3.setText("Close"); jMenuItem3.setAccelerator(javax.swing.KeyStroke.getKeyStroke(67,

0, false)); jMenuItem4.setText("Constants"); jMenu2.setText("Parameters"); jMenuItem5.setActionCommand("EditParam"); jMenuItem5.setText("Edit"); jMenuItem6.setActionCommand("LoadParam"); jMenuItem6.setText("Load"); jMenuItem7.setActionCommand("SaveParam"); jMenuItem7.setText("Save"); jMenu3.setText("Gradients"); jMenuItem8.setActionCommand("EditGrad"); jMenuItem8.setText("Edit"); jMenuItem9.setActionCommand("LoadGrad"); jMenuItem9.setText("Load"); jMenuItem10.setActionCommand("SaveGrad"); jMenuItem10.setText("Save"); jMenu4.setText("Calculation"); jMenuItem11.setActionCommand("GradCalculate"); jMenuItem11.setText("Calculate"); jMenuItem12.setActionCommand("GradInclude"); jMenuItem12.setText("Inclusions"); jMenuBar1.add(jMenu1); jMenuBar1.add(jMenu2); jMenuBar1.add(jMenu3); jMenuBar1.add(jMenu4); jMenuBar1.add(jMenuItem4); jMenu1.add(jMenuItem1); jMenu1.add(jMenuItem2); jMenu1.add(jMenuItem3);

166

jMenuItem1.addActionListener(this); jMenuItem2.addActionListener(this); jMenuItem3.addActionListener(this); jMenuItem4.addActionListener(this); jMenu2.add(jMenuItem5); jMenu2.add(jMenuItem6); jMenu2.add(jMenuItem7); jMenuItem5.addActionListener(this); jMenuItem6.addActionListener(this); jMenuItem7.addActionListener(this); jMenu3.add(jMenuItem8); jMenu3.add(jMenuItem9); jMenu3.add(jMenuItem10); jMenuItem8.addActionListener(this); jMenuItem9.addActionListener(this); jMenuItem10.addActionListener(this); jMenu4.add(jMenuItem11); jMenu4.add(jMenuItem12); jMenuItem11.addActionListener(this); jMenuItem12.addActionListener(this); jPanel1.add(output,BorderLayout.NORTH); jPanel1.add(canvas,BorderLayout.CENTER); output.setRows(5); contentPane.add(jPanel1, BorderLayout.CENTER); contentPane.add(jMenuBar1, BorderLayout.NORTH); readConstants(); displayGrad(); } public void displayGrad() { canvas.repaint();

Pulse pulP0 = new Pulse(TE,p0-20,p0+20,h0); canvas.pulseList.addElement(pulP0);

Pulse pulP1 = new Pulse(TE,p1-20,p1+20,h1); canvas.pulseList.addElement(pulP1); Pulse groA = new Pulse(TE,c2,c3,GROb); canvas.gradRO.addElement(groA); Pulse groB = new Pulse(TE,c8,c9,GROa); canvas.gradRO.addElement(groB); Pulse gssA = new Pulse(TE,c0,c1,GSSa); canvas.gradSS.addElement(gssA); Pulse gssB = new Pulse(TE,c2,c3,GSSb); canvas.gradSS.addElement(gssB); Pulse gssC = new Pulse(TE,c4,c5,GSSc); canvas.gradSS.addElement(gssC); Pulse gssD = new Pulse(TE,c5,c6,GSSd); canvas.gradSS.addElement(gssD); Pulse gssE = new Pulse(TE,c7,c8,GSSe); canvas.gradSS.addElement(gssE);

167

Pulse gssF = new Pulse(TE,c9,c10,GSSf); canvas.gradSS.addElement(gssF); canvas.firstEcho = new Echo(TE, p2, 6); canvas.setBounds(200,200,100,100); canvas.setForeground(Color.white); canvas.setBackground(Color.blue); } // END DisplayGrad public void CalculateDiffusion() { // Setup gradient lists... // i measured in microseconds. double integ1; double integ1prev = 0; double integ2; double integ2prev = 0; double sum = 0; int pn = 1; long i; // Front section... (with 90 pulse) for(i = c0; i < c4; i++) { // for(2) if((i>=c0) && (i<c1)) sum = getGrad(1); else if((i>=c2) && (i<c3)) sum = getGrad(3); else sum = getGrad(4); sum = sum*pn*Gmax/100; integ1 = integ1prev + (sum*TimeInc); integ1prev = integ1; integ2 = integ2prev + (integ1*integ1*TimeInc); integ2prev = integ2; } // END for(2) // Recirculating section... (no 90 pulse) // Start at c4 which is set to zero. for(i = 0; i < (cTE*nEchoes); i++) { // for(3) if((i%cTE) == (c5-c4)) pn *= -1; if(((i%cTE)>=(c4_5-c4)) && (((i%cTE) < (c5-c4)))) sum = getGrad(5); else if (((i%cTE)>=(c5-c4)) && (((i%cTE) < (c6-c4)))) sum = getGrad(6); else if (((i%cTE)>=(c7-c4)) && (((i%cTE) < (c8-c4)))) sum = getGrad(8); else if (((i%cTE)>=(c8-c4)) && (((i%cTE) < (c9-c4)))) sum = getGrad(9); else if (((i%cTE)>=(c9-c4)) && (((i%cTE) < (c10-c4)))) sum = getGrad(10); else sum = getGrad(11); sum = sum*pn*Gmax/100; integ1 = integ1prev + (sum*TimeInc); integ1prev = integ1; integ2 = integ2prev + (integ1*integ1*TimeInc); integ2prev = integ2;

168

if ((i%cTE) == (c8_5-c4)) { graphData.addElement(new Double(integ2*gamma*gamma*D)); } // END if } // END for(3) // Print results of bD to screen... output.append(" \n"); for(i=0; i<nEchoes; i++) { output.append(String.valueOf(((Double) graphData.elementAt((int)i)).doubleValue()) + " "); } // END for try { FileOutputStream stream = new FileOutputStream("calcs.dat"); PrintWriter pW = new PrintWriter(stream,true); for(i=0; i<nEchoes; i++) { pW.println(String.valueOf(((Double) graphData.elementAt((int)i)).doubleValue())); } stream.close(); } catch(IOException ie) { JOptionPane.showMessageDialog(this,"Error opening 'calcs.dat': "

+ ie.toString()); } graphData.removeAllElements(); } // END method: calculateDiffusion. public double getGrad(int loc) { switch(loc) { case 1: switch(inclGrad) { case 1: case 2: case 4: return 0; case 3: case 5: case 6: case 7: return GSSa; } case 2: return 0; case 3: switch(inclGrad) { case 1: case 4: return GROa; case 2: return 0; case 3: case 6: return GSSb; case 5: case 7: return GROa + GSSb; }

169

case 4: return 0; case 5: case 6: switch(inclGrad) { case 1: case 2: case 4: return 0; case 3: case 5: case 6: case 7: return GSSc; } case 7: return 0; case 8: switch(inclGrad) { case 1: return 0; case 2: case 4: return GPHa; case 3: case 5: return GSSe; case 6: case 7: return GSSe + GPHa; } case 9: switch(inclGrad) { case 1: case 4: case 5: case 7: return GROb; case 2: case 3: case 6: return 0; } case 10: switch(inclGrad) { case 1: return 0; case 2: case 4: return GPHb; case 3: case 5: return GSSf; case 6: case 7: return GPHb + GSSf; } case 11: return 0; } // end switch(loc) return 0;

170

} // END getGrad protected void processWindowEvent(WindowEvent e) { super.processWindowEvent(e); if (e.getID() == WindowEvent.WINDOW_CLOSING) { System.exit(0); } } public void actionPerformed(ActionEvent e) { String source = e.getActionCommand(); if(source == "OpenFile") { JFrame fileframe = new frame2(); fileframe.setSize(400,300); fileframe.setLocation(200,100); fileframe.setVisible(true); } else if (source == "SaveFile") { } else if (source == "CloseFile") { } else if (source == "GradCalculate") { readConstants(); CalculateDiffusion(); } else if (source == "GradInclude") { JDialog incGrad = new frame6(Loc); inclGrad = Loc.getValue(); incGrad.setSize(200,300); incGrad.setLocation(200,100); incGrad.setVisible(true); output.append(" Inc: " + Loc.getValue()); } else if (source == "EditParam") { JDialog paramFrame = new frame4(); paramFrame.setSize(200,300); paramFrame.setLocation(200,100); paramFrame.setVisible(true); readConstants(); displayGrad(); } else if (source == "EditGrad") { JDialog gradFrame = new frame5(); gradFrame.setSize(200,300); gradFrame.setLocation(200,100); gradFrame.setVisible(true); readConstants(); displayGrad(); } else if (source == "Constants") { JDialog constFrame = new frame3(); constFrame.setSize(200,150); constFrame.setLocation(200,100); constFrame.setVisible(true); readConstants(); displayGrad(); } } public void readConstants() { // Read in constants...

171

output.setText(""); try { File constsFile = new File("consts.dat"); BufferedReader reader = new BufferedReader(new

FileReader(constsFile)); String line = reader.readLine(); D = Double.parseDouble(line); line = reader.readLine(); Gmax = Double.parseDouble(line); line = reader.readLine(); gamma = Double.parseDouble(line); reader.close(); output.append(String.valueOf(D) + "m^2/s " +

String.valueOf(Gmax) + "T/cm " + String.valueOf(gamma));

} catch (IOException ie) { output.append(ie.toString()); } // Read in parameters... try { File paramsFile = new File("params.dat"); BufferedReader reader = new BufferedReader(new

FileReader(paramsFile)); String line = reader.readLine(); TE = Integer.parseInt(line); line = reader.readLine(); P0 = Integer.parseInt(line); line = reader.readLine(); P1 = Integer.parseInt(line); line = reader.readLine(); aqq = Integer.parseInt(line); line = reader.readLine(); d2 = Integer.parseInt(line); line = reader.readLine(); d3 = Integer.parseInt(line); line = reader.readLine(); d4 = Integer.parseInt(line); line = reader.readLine(); d5 = Integer.parseInt(line); line = reader.readLine(); d12 = Integer.parseInt(line); line = reader.readLine(); d14 = Integer.parseInt(line); line = reader.readLine(); d18 = Integer.parseInt(line); reader.close(); output.append("\n" + String.valueOf(TE) + " " + String.valueOf(P0) + " " + String.valueOf(P1) + " " + String.valueOf(aqq) + " " +

172

String.valueOf(d2) + " " + String.valueOf(d3) + " " + String.valueOf(d4) + " " + String.valueOf(d5) + " " + String.valueOf(d12) + " " + String.valueOf(d14) + " " + String.valueOf(d18)); } catch (IOException ie) { output.append(ie.toString()); } // Read in gradients... try { File gradisFile = new File("gradis.dat"); BufferedReader reader = new BufferedReader(new

FileReader(gradisFile)); String line = reader.readLine(); GROa = Double.parseDouble(line); line = reader.readLine(); GROb = Double.parseDouble(line); line = reader.readLine(); GSSa = Double.parseDouble(line); line = reader.readLine(); GSSb = Double.parseDouble(line); line = reader.readLine(); GSSc = Double.parseDouble(line); line = reader.readLine(); GSSd = Double.parseDouble(line); line = reader.readLine(); GSSe = Double.parseDouble(line); line = reader.readLine(); GSSf = Double.parseDouble(line); reader.close(); output.append("\n" + String.valueOf(GROa) + " " +

String.valueOf(GROb) + " " + String.valueOf(GSSa) + " " + String.valueOf(GSSb) + " " + String.valueOf(GSSc) + " " + String.valueOf(GSSd) + " " + String.valueOf(GSSe) + " " + String.valueOf(GSSf)); } catch (IOException ie) { output.append(ie.toString()); } // Recalculated delays... // Start at half way through 90 pulse... c0 = 0; c1 = P0/2; c2 = c1 + d4; c3 = c2 + d3; c4 = c3 + d4 + d5; // Repeat section for Number of Echoes c4_5 = c4 + 1; c5 = c4_5 + d4 + d18 + P1/2; // First half of 180 pulse c6 = c4_5 + d4 + d18 + P1 + d18; // Second half of 180 pulse c7 = c6 + d4 + 1 + d5; c8 = c7 + d2;

173

c8_5 = c8 + d14 + d14 + aqq/2; c9 = c8 + d14 + d14 + aqq; c10 = c9 + d12 + d12; c11 = c10 + d4 + d5; // END of repeat section. cTE = c11 - c4; output.append(" Calculated TE: " + String.valueOf(cTE) + " "); p0 = (c0 + c1)/2; p1 = (c4 + c6)/2; p2 = (p0 + p1 + p1); } // END ReadConstants } import javax.swing.*; import java.awt.*; import java.awt.event.*; public class frame2 extends JFrame { JFileChooser jFileChooser1 = new JFileChooser(); public frame2() { try { jbInit(); } catch(Exception e) { e.printStackTrace(); } } private void jbInit() throws Exception { this.setTitle("Open File"); jFileChooser1.addActionListener(new

java.awt.event.ActionListener() { public void actionPerformed(ActionEvent e) { jFileChooser1_actionPerformed(e); } }); this.getContentPane().add(jFileChooser1, BorderLayout.NORTH); } void jFileChooser1_actionPerformed(ActionEvent e) { String source = e.getActionCommand(); this.setTitle(source); if (source == "CancelSelection") { this.dispose(); } else if (source == "ApproveSelection") { }

Class: frame2 Author: Christopher Hurley Date: June, 2002

174

} } import javax.swing.*; import java.awt.*; import java.awt.event.*; import java.io.*; public class frame3 extends JDialog { JPanel contentPane; JPanel jPanel1 = new JPanel(); JPanel jPanel2 = new JPanel(); JButton jButton1 = new JButton(); JButton jButton2 = new JButton(); JLabel labelD = new JLabel("D"); JTextField fieldD = new JTextField(" "); JLabel unitsD = new JLabel("m^2/s"); JLabel labelGM = new JLabel("Gmax"); JTextField fieldGM = new JTextField(" "); JLabel unitsGM = new JLabel("T/m"); JLabel labelGA = new JLabel("Gamma"); JTextField fieldGA = new JTextField(" "); JLabel unitsGA = new JLabel("rad/s/T"); BorderLayout borderLayout1 = new BorderLayout(); FlowLayout flowLayout1 = new FlowLayout(); GridBagLayout gridBagLayout1 = new GridBagLayout(); public frame3() { try { jbInit(); } catch(Exception e) { e.printStackTrace(); } } private void jbInit() throws Exception { this.setTitle("Set Constants"); contentPane = (JPanel) this.getContentPane(); contentPane.setLayout(borderLayout1); jButton1.setText("Accept"); jButton1.addActionListener(new java.awt.event.ActionListener() {

Class: frame3 Author: Christopher Hurley Date: June, 2002

175

public void actionPerformed(ActionEvent e) { jButton1_actionPerformed(e); } }); jButton2.setText("Cancel"); jButton2.addActionListener(new java.awt.event.ActionListener() { public void actionPerformed(ActionEvent e) { jButton2_actionPerformed(e); } }); jPanel1.setLayout(flowLayout1); jPanel1.add(jButton1, null); jPanel1.add(jButton2, null); contentPane.add(jPanel2, BorderLayout.CENTER); contentPane.add(jPanel1, BorderLayout.SOUTH); jPanel2.setLayout(gridBagLayout1); GridBagConstraints c[] = new GridBagConstraints[9]; for(int i=0; i<9;i++) { c[i] = new GridBagConstraints(); c[i].gridwidth = 1; c[i].gridheight = 1; c[i].ipadx = 10; c[i].ipady = 1; c[i].gridx = i%3; c[i].gridy = i/3; } jPanel2.add(labelD,c[0]); jPanel2.add(fieldD,c[1]); jPanel2.add(unitsD,c[2]); jPanel2.add(labelGM,c[3]); jPanel2.add(fieldGM,c[4]); jPanel2.add(unitsGM,c[5]); jPanel2.add(labelGA,c[6]); jPanel2.add(fieldGA,c[7]); jPanel2.add(unitsGA,c[8]); } void jButton1_actionPerformed(ActionEvent e) { // Accept // save values... try { FileOutputStream stream = new FileOutputStream("consts.dat"); PrintWriter pW = new PrintWriter(stream,true); pW.println(fieldD.getText().trim()); pW.println(fieldGM.getText().trim()); pW.println(fieldGA.getText().trim()); stream.close(); } catch(IOException ie) { JOptionPane.showMessageDialog(this,"Error opening 'consts.dat':

" + ie.toString()); } // exit

176

this.dispose(); } void jButton2_actionPerformed(ActionEvent e) { // Cancel // exit without saving this.dispose(); } void jTextField1_actionPerformed(ActionEvent e) { } } import javax.swing.*; import java.awt.*; import java.awt.event.*; import java.io.*; public class frame4 extends JDialog { JPanel contentPane; JPanel mainPanel = new JPanel(new BorderLayout()); JPanel buttonPanel = new JPanel(new FlowLayout()); JPanel paramPanel = new JPanel(new GridLayout(11,2)); JPanel descPanel = new JPanel(new FlowLayout()); JButton jButton1 = new JButton(); JButton jButton2 = new JButton(); JLabel labelTE; JLabel labelP0; JLabel labelP1; JLabel labelaqq; JLabel labeld2; JLabel labeld3; JLabel labeld4; JLabel labeld5; JLabel labeld12; JLabel labeld14; JLabel labeld18; JTextField fieldTE; JTextField fieldP0; JTextField fieldP1; JTextField fieldaqq; JTextField fieldd2; JTextField fieldd3; JTextField fieldd4; JTextField fieldd5; JTextField fieldd12; JTextField fieldd14;

Class: frame4 Author: Christopher Hurley Date: June, 2002

177

JTextField fieldd18; public frame4() { try { jbInit(); } catch(Exception e) { e.printStackTrace(); } } private void jbInit() throws Exception { this.setTitle("Set Parameters"); contentPane = (JPanel) this.getContentPane(); mainPanel = (JPanel) this.getContentPane(); jButton1.setText("Accept"); jButton1.addActionListener(new java.awt.event.ActionListener() { public void actionPerformed(ActionEvent e) { jButton1_actionPerformed(e); } }); jButton2.setText("Cancel"); jButton2.addActionListener(new java.awt.event.ActionListener() { public void actionPerformed(ActionEvent e) { jButton2_actionPerformed(e); } }); labelTE = new JLabel("TE"); labelP0 = new JLabel("P0"); labelP1 = new JLabel("P1"); labelaqq = new JLabel("aqq"); labeld2 = new JLabel("d2"); labeld3 = new JLabel("d3"); labeld4 = new JLabel("d4"); labeld5 = new JLabel("d5"); labeld12 = new JLabel("d12"); labeld14 = new JLabel("d14"); labeld18 = new JLabel("d18"); fieldTE = new JTextField(); fieldP0 = new JTextField(); fieldP1 = new JTextField(); fieldaqq = new JTextField(); fieldd2 = new JTextField(); fieldd3 = new JTextField(); fieldd4 = new JTextField(); fieldd5 = new JTextField(); fieldd12 = new JTextField(); fieldd14 = new JTextField(); fieldd18 = new JTextField(); paramPanel.add(labelTE); paramPanel.add(fieldTE); paramPanel.add(labelP0); paramPanel.add(fieldP0); paramPanel.add(labelP1); paramPanel.add(fieldP1); paramPanel.add(labelaqq); paramPanel.add(fieldaqq);

178

paramPanel.add(labeld2); paramPanel.add(fieldd2); paramPanel.add(labeld3); paramPanel.add(fieldd3); paramPanel.add(labeld4); paramPanel.add(fieldd4); paramPanel.add(labeld5); paramPanel.add(fieldd5); paramPanel.add(labeld12); paramPanel.add(fieldd12); paramPanel.add(labeld14); paramPanel.add(fieldd14); paramPanel.add(labeld18); paramPanel.add(fieldd18); buttonPanel.add(jButton1); buttonPanel.add(jButton2); JLabel desc = new JLabel("Enter in the Delays in useconds ..."); descPanel.add(desc); mainPanel.add(buttonPanel,BorderLayout.SOUTH); mainPanel.add(paramPanel,BorderLayout.CENTER); mainPanel.add(descPanel,BorderLayout.NORTH); } void jButton1_actionPerformed(ActionEvent e) { // Accept // save values... try { FileOutputStream stream = new FileOutputStream("params.dat"); PrintWriter pW = new PrintWriter(stream,true); pW.println(fieldTE.getText().trim()); pW.println(fieldP0.getText().trim()); pW.println(fieldP1.getText().trim()); pW.println(fieldaqq.getText().trim()); pW.println(fieldd2.getText().trim()); pW.println(fieldd3.getText().trim()); pW.println(fieldd4.getText().trim()); pW.println(fieldd5.getText().trim()); pW.println(fieldd12.getText().trim()); pW.println(fieldd14.getText().trim()); pW.println(fieldd18.getText().trim()); stream.close(); } catch(IOException ie) { JOptionPane.showMessageDialog(this,"Error opening 'params.dat':

" + ie.toString()); } // exit this.dispose(); // exit this.dispose(); } void jButton2_actionPerformed(ActionEvent e) { // Cancel // exit without saving this.dispose(); } void jTextField1_actionPerformed(ActionEvent e) { }

179

} import javax.swing.*; import java.awt.*; import java.awt.event.*; import java.io.*; public class frame5 extends JDialog { JPanel contentPane; JPanel mainPanel = new JPanel(new BorderLayout()); JPanel buttonPanel = new JPanel(new FlowLayout()); JPanel paramPanel = new JPanel(new GridLayout(8,2)); JPanel descPanel = new JPanel(new FlowLayout()); JButton jButton1 = new JButton(); JButton jButton2 = new JButton(); JTextField fieldGROa; JTextField fieldGROb; JTextField fieldGSSa; JTextField fieldGSSb; JTextField fieldGSSc; JTextField fieldGSSd; JTextField fieldGSSe; JTextField fieldGSSf; public frame5() { try { jbInit(); } catch(Exception e) { e.printStackTrace(); } } private void jbInit() throws Exception { this.setTitle("Set Parameters"); contentPane = (JPanel) this.getContentPane(); mainPanel = (JPanel) this.getContentPane(); jButton1.setText("Accept"); jButton1.addActionListener(new java.awt.event.ActionListener() { public void actionPerformed(ActionEvent e) { jButton1_actionPerformed(e); } }); jButton2.setText("Cancel"); jButton2.addActionListener(new java.awt.event.ActionListener() { public void actionPerformed(ActionEvent e) {

Class: frame5 Author: Christopher Hurley Date: June, 2002

180

jButton2_actionPerformed(e); } }); JLabel labelGROa = new JLabel("GROa"); JLabel labelGROb = new JLabel("GROb"); JLabel labelGSSa = new JLabel("GSSa"); JLabel labelGSSb = new JLabel("GSSb"); JLabel labelGSSc = new JLabel("GSSc"); JLabel labelGSSd = new JLabel("GSSd"); JLabel labelGSSe = new JLabel("GSSe"); JLabel labelGSSf = new JLabel("GSSf"); fieldGROa = new JTextField(); fieldGROb = new JTextField(); fieldGSSa = new JTextField(); fieldGSSb = new JTextField(); fieldGSSc = new JTextField(); fieldGSSd = new JTextField(); fieldGSSe = new JTextField(); fieldGSSf = new JTextField(); paramPanel.add(labelGROa); paramPanel.add(fieldGROa); paramPanel.add(labelGROb); paramPanel.add(fieldGROb); paramPanel.add(labelGSSa); paramPanel.add(fieldGSSa); paramPanel.add(labelGSSb); paramPanel.add(fieldGSSb); paramPanel.add(labelGSSc); paramPanel.add(fieldGSSc); paramPanel.add(labelGSSd); paramPanel.add(fieldGSSd); paramPanel.add(labelGSSe); paramPanel.add(fieldGSSe); paramPanel.add(labelGSSf); paramPanel.add(fieldGSSf); buttonPanel.add(jButton1); buttonPanel.add(jButton2); JLabel desc = new JLabel("Enter in the Gradient Strength % ..."); descPanel.add(desc); mainPanel.add(buttonPanel,BorderLayout.SOUTH); mainPanel.add(paramPanel,BorderLayout.CENTER); mainPanel.add(descPanel,BorderLayout.NORTH); } void jButton1_actionPerformed(ActionEvent e) { // Accept // save values... try { FileOutputStream stream = new FileOutputStream("gradis.dat"); PrintWriter pW = new PrintWriter(stream,true); pW.println(fieldGROa.getText().trim()); pW.println(fieldGROb.getText().trim()); pW.println(fieldGSSa.getText().trim()); pW.println(fieldGSSb.getText().trim()); pW.println(fieldGSSc.getText().trim()); pW.println(fieldGSSd.getText().trim()); pW.println(fieldGSSe.getText().trim()); pW.println(fieldGSSf.getText().trim()); stream.close(); }

181

catch(IOException ie) { JOptionPane.showMessageDialog(this,"Error opening 'gradis.dat':

" + ie.toString()); } // exit this.dispose(); } void jButton2_actionPerformed(ActionEvent e) { // Cancel // exit without saving this.dispose(); } void jTextField1_actionPerformed(ActionEvent e) { } } import javax.swing.*; import java.awt.*; import java.awt.event.*; import java.io.*; public class frame6 extends JDialog { JPanel contentPane; JPanel mainPanel = new JPanel(new BorderLayout()); JPanel buttonPanel = new JPanel(new FlowLayout()); JPanel paramPanel = new JPanel(new GridLayout(3,1)); JPanel descPanel = new JPanel(new FlowLayout()); JButton jButton1 = new JButton(); JButton jButton2 = new JButton(); JCheckBox RO; JCheckBox PH; JCheckBox SS; JLabel labelRO; JLabel labelPH; JLabel labelSS; MyInt Loc; public frame6(MyInt Loc) { this.Loc = Loc; try { jbInit(); } catch(Exception e) { e.printStackTrace();

Class: frame6 Author: Christopher Hurley Date: June, 2002

182

} } // frame6 private void jbInit() throws Exception { this.setTitle("Inclusions"); contentPane = (JPanel) this.getContentPane(); mainPanel = (JPanel) this.getContentPane(); jButton1.setText("Accept"); jButton1.addActionListener(new java.awt.event.ActionListener() { public void actionPerformed(ActionEvent e) { jButton1_actionPerformed(e); } }); jButton2.setText("Cancel"); jButton2.addActionListener(new java.awt.event.ActionListener() { public void actionPerformed(ActionEvent e) { jButton2_actionPerformed(e); } }); JLabel labelGSSf = new JLabel("GSSf"); RO = new JCheckBox("Read Out Gradient"); PH = new JCheckBox("Phase Encode Gradient"); SS = new JCheckBox("Slice Select Gradient"); buttonPanel.add(jButton1); buttonPanel.add(jButton2); paramPanel.add(RO); paramPanel.add(PH); paramPanel.add(SS); JLabel desc = new JLabel("Tick the gradients to include..."); descPanel.add(desc); mainPanel.add(buttonPanel,BorderLayout.SOUTH); mainPanel.add(paramPanel,BorderLayout.CENTER); mainPanel.add(descPanel,BorderLayout.NORTH); } // END jbInit void jButton1_actionPerformed(ActionEvent e) { // Accept if (!RO.isSelected() && !PH.isSelected() && !SS.isSelected()) Loc.setValue(0); else if (RO.isSelected() && !PH.isSelected() && !SS.isSelected()) Loc.setValue(1); else if (!RO.isSelected() && PH.isSelected() && !SS.isSelected()) Loc.setValue(2); else if (!RO.isSelected() && !PH.isSelected() && SS.isSelected()) Loc.setValue(3); else if (RO.isSelected() && PH.isSelected() && !SS.isSelected())

183

Loc.setValue(4); else if (RO.isSelected() && !PH.isSelected() && SS.isSelected()) Loc.setValue(5); else if (!RO.isSelected() && PH.isSelected() && SS.isSelected()) Loc.setValue(6); else if (RO.isSelected() && PH.isSelected() && SS.isSelected()) Loc.setValue(7); else Loc.setValue(0); this.dispose(); } void jButton2_actionPerformed(ActionEvent e) { // Cancel // exit without saving this.dispose(); } void jTextField1_actionPerformed(ActionEvent e) { } } public class GradLine { int time; int data; int integ1; int square; public GradLine() { data = 0; time = 0; } public GradLine(int time, int data) { this.time = time; this.data = data; } } public class MyInt {

Class: GradLine Author: Christopher Hurley Date: June, 2002

Class: MyInt Author: Christopher Hurley Date: June, 2002

184

int value; public MyInt() { value = 0; } public MyInt(int i) { value = i; } public void setValue(int i) { value = i; } public int getValue() { return value; } } public class Pulse { public int start = 0; // in microseconds public int finish = 0; // in microseconds public double height = 0; // in percentage public int te = 0; // in microseconds public boolean enabled = false; public Pulse() { } public Pulse(int te, int start, int finish, double height) { this.te = te; this.start = start; this.finish = finish; this.height = height; enabled = true; } public void show() { enabled = true; } public void hide() { enabled = false; } } public class surface1 extends Canvas { Vector pulseList = new Vector();

Class: Pulse Author: Christopher Hurley Date: June, 2002

Class: surface1 Author: Christopher Hurley Date: June, 2002

185

Vector gradRO = new Vector(); Vector gradSS = new Vector(); Echo firstEcho; public int timeOffset = 50; public surface1() { } public void paint(Graphics g) { Dimension d = getSize(); int start = 0; int finish = 0; g.setColor(Color.yellow); int w = d.width; int h = d.height; g.drawString("PULSE",5,h/4); g.drawString("GRO",5,h/2); g.drawString("GSS",5,3*h/4); if(pulseList.isEmpty()) { g.drawLine(0, h/4, w, h/4); } else { for(int i=0; i<pulseList.size();i++) {

finish = drawPulse(g,(Pulse)pulseList.elementAt(i),h,w,start,h/4);

start = finish; } // Draw Echo... int eSF = drawEcho(g,w,h/4); g.drawLine(start,h/4,eSF-30,h/4); g.drawLine(eSF+30,h/4,w,h/4); } start = 0; if(gradRO.isEmpty()) { g.drawLine(0, h/2, w, h/2); } else { for(int i=0; i<gradRO.size();i++) { finish =

drawPulse(g,(Pulse)gradRO.elementAt(i),h,w,start,h/2); start = finish; } g.drawLine(start,h/2,w,h/2); } start = 0; if(gradSS.isEmpty()) { g.drawLine(0, 3*h/4, w, 3*h/4); } else { for(int i=0; i<gradSS.size();i++) { finish =

drawPulse(g,(Pulse)gradSS.elementAt(i),h,w,start,3*h/4);

start = finish; } g.drawLine(start,3*h/4,w,3*h/4);

186

} } public int drawPulse(Graphics g, Pulse pulse, int h, int w, int

lastF, int line) { int dh = (int)((h*pulse.height)/100); int te = pulse.te; int dstart = pulse.start; int dfinish = pulse.finish; int ds = w*dstart/((int) (1.5*te)) + timeOffset; int df = w*dfinish/((int)(1.5*te)) + timeOffset; g.drawLine(lastF,line,ds,line); g.drawLine(ds,line,ds,line - dh); g.drawLine(ds,line-dh,df,line-dh); g.drawLine(df,line,df,line-dh); return df; } public int drawEcho(Graphics g, int w, int line) { int x = (firstEcho.echotime*w/((int) (1.5 * firstEcho.te))) +

timeOffset - 60; int y = line - 60; g.drawArc(x-30,y+60,60,60,0,90); g.drawArc(x-30,y,60,60,0,-90); g.drawArc(x+30,y,60,60,-90,-90); g.drawArc(x+30,y+60,60,60,90,90); return firstEcho.echotime*w/((int)(1.5*firstEcho.te))+timeOffset-

30; } }