SlideShare a Scribd company logo
1 of 90
Download to read offline
Development of a multi­knife­edge slit collimator for prompt gamma ray imaging  
during proton beam cancer therapy 
 
 
By 
 
John Francis Ready III 
 
 
A dissertation submitted in partial satisfaction of the 
 
requirements for the degree of 
 
Doctor of Philosophy 
 
in 
 
Engineering ­ Nuclear Engineering 
 
in the 
 
Graduate Division 
 
of the 
 
University of California, Berkeley 
 
 
 
 
 
Committee in charge: 
 
Professor Kai Vetter, Chair 
Professor Karl van Bibber 
Professor Steven Boggs 
 
 
Summer 2016 
Development of a multi­knife­edge slit collimator for prompt gamma ray imaging  
during proton beam cancer therapy 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
© Copyright 2016 by John Ready 
 
All Rights Reserved 
Abstract 
Development of a multi­knife­edge slit collimator for prompt gamma ray imaging during 
proton beam cancer therapy 
by 
John Francis Ready III 
Doctor of Philosophy in Nuclear Engineering 
University of California, Berkeley 
Professor Kai Vetter, Chair 
 
Proton beam usage to treat cancer has recently experienced rapid growth, as it                         
offers the ability to target dose delivery in a patient more precisely than traditional x­ray                             
treatment methods. Protons stop within the patient, delivering the maximum dose at the                         
end of their track—a phenomenon described as the Bragg peak. However, because a large                           
dose is delivered to a small volume, proton therapy is very sensitive to errors in patient                               
setup and treatment planning calculations. Additionally, because all primary beam particles                     
stop in the patient, there is no direct information available to verify dose delivery. These                             
factors contribute to the range uncertainty in proton therapy, which ultimately hinders its                         
clinical usefulness. A reliable method of proton range verification would allow the clinician                         
to fully utilize the precise dose delivery of the Bragg peak. 
 
Several methods to verify proton range detect secondary emissions, especially                   
prompt gamma ray (PG) emissions. However, detection of PGs is challenging due to their                           
high energy (2–10 MeV) and low attenuation coefficients, which limit PG interactions in                         
materials. Therefore, detection and collimation methods must be specifically designed for                     
prompt gamma ray imaging (PGI) applications. In addition, production of PGs relies on                         
delivering a dose of radiation to the patient. Ideally, verification of the Bragg peak location                             
exposes patients to a minimal dose, thus limiting the PG counts available to the imaging                             
system. 
 
An additional challenge for PGI is the lack of accurate simulation models, which limit                           
the study of PG production characteristics and the relationship between PG distribution                       
and dose delivery. Specific limitations include incorrect modeling of the reaction cross                       
sections, gamma emission yields, and angular distribution of emission for specific photon                       
energies. While simulations can still be valuable assets in designing a system to detect and                             
image PGs, until new models are developed and incorporated into Monte Carlo simulation                         
packages, simulations cannot be used to study the production and location of PG emissions                           
during proton therapy. 
1 
 
This work presents a novel system to image PGs emitted during proton therapy to                           
verify proton beam range. The imaging system consists of a multi­slit collimator paired                         
with a position­sensitive LSO scintillation detector. This innovative design is the first                       
collimated imaging system to implement two­dimensional (2­D) imaging for PG proton                     
beam range verification, while also providing a larger field of view than compared to                           
single­slit collimator systems. Other, uncollimated imaging systems have been explored for                     
PGI applications, such as Compton cameras. However, Compton camera designs are                     
severely limited by counting rate capabilities. A recent Compton camera study reported                       
count rate capability of about 5 kHz. However, at a typical clinical beam current of 1.0 nA,                                 
the estimated PG emission rate would be 6 x 10​8
per second. After accounting for distance                               
to the detector and interaction efficiencies, the detection system will still be overwhelmed                         
with counts in the MHz range, causing false coincidences and hindering the operation of the                             
imaging system. 
 
Initial measurements using 50 MeV protons demonstrated the ability of our system                       
to reconstruct 2­D PG distributions at clinical beam currents. A Bragg peak localization                         
precision of 1 mm (2σ) was achieved with delivery of (1.7 ± 0.8) x 10​8
protons into a                                   
poly(methyl methacrylate) (PMMA) target, suggesting the ability of the system to detect                       
relative shifts in proton range while delivering fewer protons than used in a typical                           
treatment fraction. This is key, as the ideal system allows the clinician to verify proton                             
range when delivering only a small portion of the prescribed dose, preventing the                         
mistreatment of the patient. Additionally, the absolute position of the Bragg peak was                         
identified to within 1.6 mm (2σ) with 5.6 x 10​10​
 protons delivered.  
 
These promising results warrant further investigation and system optimization for                   
clinical implementation. While further measurements at clinical beam energy levels will be                       
required to verify system performance, these preliminary results provide evidence that 2­D                       
image reconstruction, with 1–2 mm accuracy, is possible with this design. Implementing                       
such a system in the clinical setting would greatly improve proton therapy cancer                         
treatment outcomes. 
2 
Dedication  
 
To my wife Miriam—for your unending support, patience, and love. I wouldn’t have                         
gotten through this doctorate if not for you. I strive each day to make you as proud of me as                                       
I am of you. 
 
To my son Logan and daughter Lexi—you both came into the world during my                           
graduate school career and have completely changed my perspective on life. Your smiles                         
and love have provided me with more motivation than you will ever know. 
i 
 
Table of Contents 
    List of Figures  iv 
    List of Tables  vii 
    Acknowledgements   viii 
   
Chapter 1—Introduction  1 
Protons for cancer treatment  1 
Physics of proton therapy  2 
Range uncertainty  4 
Motivation & challenges  5 
   
Chapter 2—Techniques and Challenges of Range Verification  7 
Positron emission tomography  7 
PG emission  8 
Detection and imaging of PGs  10 
Detector materials  10 
Detection techniques  11 
Compton cameras  11 
Electron­track Compton cameras  11 
Gamma­electron vortex imaging  12 
PG timing  12 
Collimated cameras  12 
Conclusion  13 
   
Chapter 3—A Monte Carlo Study of Prompt Gamma Production  14 
Background  14 
PG production cross sections  15 
Methods  15 
Results  16 
Angular distribution of PGs  18 
Methods  20 
Results  21 
Conclusion  23 
   
Chapter 4—A Novel Multi­Slit Collimated Imaging System  24 
Introduction  24 
Collimator design  25 
Detector configuration  27 
Conclusion  30 
   
ii 
Chapter 5—Data acquisition, data processing, and image reconstruction  31 
Acquisition hardware  31 
Detectors  32 
High­voltage power supply  32 
Analog­to­Digital Converter  33 
Acquisition software and data processing  35 
Acquisition  35 
Data processing  37 
Image reconstruction  40 
   
Chapter 6—Simulation Study of Collimator Performance  43 
Introduction  43 
Methods and materials  43 
Results and discussion  46 
Conclusion  48 
   
Chapter 7—Experimental Characterization of System Performance  50 
Introduction  50 
Methods and materials  50 
Results and discussion  52 
Image reconstruction  52 
Range measurement  54 
Conclusion  59 
   
Chapter 8—Conclusion  61 
Summary  61 
Key challenges remaining  62 
Directly applicable to this PGI system  62 
Detector design  62 
Optimization of the image reconstruction algorithm  63 
Collimator design  63 
PGI field at large  64 
Monte Carlo simulations of PG production  64 
Effect of neutrons on imaging system  64 
Development of correlation between PG distribution and  
      Bragg peak location 
64 
Inhomogeneous target materials  64 
Clinical implementation of proton range verification  65 
Final thoughts  65 
   
References  67 
 
iii 
 
List of Figures 
Figure 1  Depth­dose distributions for a Spread­Out Bragg Peak (SOBP), its                 
constituent pristine Bragg peaks, and a 10 MV photon beam.  3 
Figure 2  Comparison of treatment planning dose distributions for proton therapy                 
treatment of prostate cancer with one single anterior field and two                     
parallel­opposed lateral fields.  5 
Figure 3  Simulated profiles for dose, PG, and PET integrated over the entire beam                       
for an abdomen irradiated with pencil beams.  8 
Figure 4  Production cross sections for PGs resulting from inelastic collisions of                   
protons with ​16​
O and ​12​
C.  10 
Figure 5  Comparison of simulation results with experimental data for 1.63 MeV                   
gamma production during proton irradiation of thin nitrogen target.  16 
Figure 6  Comparison of simulation results with experimental data for 2.31 MeV                   
gamma production during proton irradiation of thin nitrogen target.  17 
Figure 7  Comparison of simulation results with experimental data for 4.44 MeV                   
gamma production during proton irradiation of thin carbon target.  17 
Figure 8  Examples of angular distributions of gamma rays.  19 
Figure 9  Angular distributions for five separate gamma ray transitions resulting                 
from bombardment of Mylar, Mg, Si, and Fe targets with 33 MeV protons.  20 
Figure 10  Angular distribution of 4.4 MeV PGs (Geant4).  22 
Figure 11  Comparison of theoretical isotropic PG distribution and expected gamma                 
distribution as detected based on angular distribution.  22 
Figure 12  CAD model of imaging system showing 4 x 4 grid of LSO detector modules,                           
tungsten collimator with individually cut pieces held together with 3­D                   
printed plastic, and a proton beam incident on a plastic target.  24 
Figure 13  Fan beam collimator design modeled in TOPAS.  25 
Figure 14  Arrangement and dimensions of tungsten collimator pieces.  26 
Figure 15  CAD model to demonstrate the position and layout of LSO detector                     
modules.  28 
Figure 16  Photographs of bottom two rows of detector modules.  28 
Figure 17  Photograph of a detector module and connections.  29 
Figure 18  Gain adjustment on PMTs is performed by making slight adjustments in the                       
respective gain adjustment potentiometer.  29 
Figure 19  Photo of acquisition system hardware.  31 
Figure 20  Electronics in VME crate.  32 
Figure 21  Screenshot of HV power supply control interface.  33 
Figure 22  Image of the four SIS33316 Struck ADCs.  34 
iv 
 
Figure 23  The 4 x 4 detector array forms the ​L​S​O ​GA​mma detection i​N​strument                       
(LOGAN) that is powered by the HV power supply to transfer the light                         
signal produced in the scintillation crystal to an electrical signal generated                     
in the PMTs which is then sent to the ADCs.  34 
Figure 24  Back view of the ​L​EMO­to­​E​thernet for ​X​­ray ​I​maging (LEXI) apparatus.  35 
Figure 25  Firmware schematic of 4 Channel Sum Trigger sequence.  36 
Figure 26  Screenshot of raw signal data from a detector module.   37 
Figure 27  Image representing the PMT configuration in relation to the segmented                   
scintillator crystal.  38 
Figure 28  Example map of detector interactions as calculated using Equations 5 and                     
6.  38 
Figure 29  Histogram of gain correction factors used to ensure a more uniform                     
response across the detector array pixels.  39 
Figure 30  Map of gain correction factors used to account for variations in counting                       
rates of individual detector pixels.  39 
Figure 31  Example of ray tracing result for a point source projection through the                       
collimator onto the detector plane.  41 
Figure 32  Sensitivity map generated by the ray tracing algorithm.  42 
Figure 33  Simulation of collimator and flat detector panel in TOPAS.  44 
Figure 34  Simulated point sources shifted in 1 cm steps along the x­direction.  45 
Figure 35  1­D reconstructed image of point sources at varying locations which                   
demonstrates, via FWHM,the spatial resolution across one dimension.  46 
Figure 36  2­D reconstruction of a simulated point source at (x, y) position (15, 5).  47 
Figure 37  2­D reconstruction of a simulated point source at (x ,y) position (5, 10).  47 
Figure 38  Input and reconstructed response of a 4.4 MeV line source.   48 
Figure 39  Proton range in tissue­equivalent plastic (PMMA) for proton energy 0–170                   
MeV and 0­50 MeV.  51 
Figure 40  Photo of experimental setup at 88­Inch Cyclotron.  51 
Figure 41  Photo of PMMA target in place for Bragg peak measurements with 50 MeV                         
proton beam.  52 
Figure 42  2­D image reconstruction of 50 MeV proton beam in PMMA target.  53 
Figure 43  1­D image of PG distribution plotted with the simulated Bragg curve with                       
dose normalized to the maximum integral depth dose.  53 
Figure 44  Range retrieval precision (2 ) versus the number of delivered protons.  54 
Figure 45  Estimated target position and deviation from trend line.  55 
Figure 46  Estimated Bragg peak location and residuals.  56 
Figure 47  2­D image reconstruction of 50 MeV proton beam in the thick PMMA                       
target, separated by photon energy.  57 
v 
 
Figure 48  1­D images of PG distribution, sorted by photon energy, plotted with the                       
simulated Bragg curve with dose normalized to the maximum integral                   
depth dose.  57 
Figure 49  2­D image reconstruction of 50 MeV proton beam in the thick PMMA                       
target, separated by photon energy with target location shifted 3 mm to the                         
right relative to Figure 47.  58 
Figure 50  1­D images of PG distribution, sorted by photon energy, plotted with the                       
simulated Bragg curve with dose normalized to the maximum integral                   
depth dose with target location shifted 3 mm to the right relative to Figure                           
48.  58 
Figure 51  Positional uniformity of detection system.  59 
Figure 52  Energy­integrated and discrete PG emissions along the path of proton                   
pencil­beams in water.   60 
Figure 53  Age­adjusted invasive cancer incidence rates in 2012 for the ten primary                     
sites with the highest rates in men.  66 
 
vi 
List of Tables 
Table 1  Reaction channels that produce the most prominent gamma               
energies.  9 
Table 2  Basic properties of LSO scintillation detectors.  27 
 
vii 
 
Acknowledgements 
 
"​... to know even one life has breathed easier because you have lived—this is to                             
have succeeded.​"—Bessie Anderson Stanley, as abridged by Albert Edward                 
Wiggam 
 
My time at Berkeley has been filled with opportunities for personal and professional                         
growth. I have been humbled and honored by the chance to serve my communities: in                             
particular, veterans and graduate students through involvement in the Graduate Assembly.                     
The time and effort I have committed to fostering a vibrant community at Cal has been                               
returned to me many times over in emotional and social support and encouragement, and                           
it introduced me to mentors such as Ron Williams, Dean Joseph Greenwell, and Vice                           
Chancellor Harry Le Grande. While a PhD is largely viewed as an accomplishment of the                             
individual, as with any major personal endeavor, my PhD and dissertation would not be                           
possible without the support of a loving community. I therefore attribute the successful                         
closure of this chapter of my life to the vast array of humans that I have had the great                                     
pleasure of knowing and working with over the past six years. 
 
I would like to express my deepest gratitude to my supervisor Professor Kai Vetter                           
for his unwavering support, collegiality, and mentorship throughout this project. 
 
I wish to thank my committee members who were more than generous with their                           
expertise and precious time. 
 
I would like to extend my thanks to those who offered invaluable research guidance                           
and support over the years: Victor Negut; Rebecca Pak; Ryan Pavlovsky; Sam Huh; Tim                           
Aucott; Andy Haefner; Ross Barnowski; Don Gunter; Lucian Mihailescu; Justin Ellin; Joseph                       
Perl; Bill Moses; fellow graduate students, postdocs, and staff scientists of the Applied                         
Nuclear Physics Program; and countless others who have had a positive impact on my                           
work. 
 
Thanks to my dad, John; mom, Micki; mother­in­law, Lucy; brothers, Joey and 
Jimmy; and all my family members, friends, fellow student veterans, Graduate Assembly 
colleagues, and anyone else that has knowingly or unknowingly helped me along my path. 
 
Finally, I offer my sincere appreciation to the staff and facilities at Lawrence                         
Berkeley National Laboratory; the faculty, staff, and students who supported my                     
educational experience at UC Berkeley; and my funding source—the Nuclear Science and                       
Security Consortium—for providing me with the opportunity and freedom to contribute to                       
the scientific knowledge of the important field of nuclear science. 
 
viii 
This material is based upon work supported by the Department of Energy National                         
Nuclear Security Administration under Award Number: DE­NA0000979 through the                 
Nuclear Science and Security Consortium. 
 
This report was prepared as an account of work sponsored by an agency of the                             
United States Government. Neither the United States Government nor any agency thereof,                       
nor any of their employees, makes any warranty, express or limited, or assumes any legal                             
liability or responsibility for the accuracy, completeness, or usefulness of any information,                       
apparatus, product, or process disclosed, or represents that its use would not infringe                         
privately owned rights. Reference herein to any specific commercial product, process, or                       
service by trade name, trademark, manufacturer, or otherwise does not necessarily                     
constitute or imply its endorsement, recommendation, or favoring by the United States                       
Government or any agency thereof. The views and opinions of authors expressed herein do                           
not necessarily state or reflect those of the United States Government or any agency                           
thereof. 
 
 
ix 
  
Chapter 1—Introduction 
Protons for cancer treatment 
The use of radiation for the treatment of cancer began very shortly after the                           
discovery of X­rays in 1895 ​[1]​. With advances in technology and an increasing                         
understanding of radiation, radiotherapy has now become a standard treatment option for                       
a wide range of malignancies ​[2]​. Today, approximately 50% of all patients with localized                           
malignant tumors are treated with radiation ​[3]​. In theory, any tumor can be killed with a                               
large enough dose. However, the tolerance of healthy tissue surrounding the tumor volume                         
limits the radiation dose permitted in practice. Technical advances in radiation therapy                       
have been aimed mainly at reducing dose to healthy tissue while maintaining or increasing                           
the dose to the tumor volume. In addition to technical innovations in the “traditional”                           
photon and electron radiotherapy methods, physicists have searched for alternative                   
particles that offer advantages in their dose deposition and biological characteristics ​[4]​. 
 
First described in 1946 by Robert Wilson ​[5]​, the clinical potential of proton beams                           
lies in the basic physics of heavy charged particle interactions in matter. Wilson recognized                           
that the depth­dose profile of protons in a patient holds dosimetric advantages over                         
traditional photon (X­ray) therapy, which sparked his interest in developing protons for                       
use in tumor treatment. 
 
The first clinical use of protons occurred at Lawrence Berkeley Laboratory in 1957                         
[6]​. The work at Berkeley confirmed the predictions of Wilson and led to the use of heavy                                 
charged particles in treating human diseases associated with the malfunctioning of the                       
pituitary gland ​[7,8]​. The following decades saw gradual growth in both the types of                           
cancers treated with protons and the number of medical centers offering proton therapy                         
[9]​. More recently, there has been a rapid proliferation of proton therapy centers, growing                           
from approximately 30 to 60 centers worldwide from 2010 to 2015 (with another 30                           
under construction in 2016) ​[10]​. This rapid expansion of proton therapy has been met                           
with questions and controversy, as clinicians try to balance efficacy and rising costs of                           
treatment ​[11,12]​. 
 
The properties of proton therapy introduce unique challenges to the field of medical                         
physics ​[13]​. As the peak dose is delivered over a relatively small volume, proton therapy is                               
very sensitive to errors in radiation delivery. Mistakes in patient setup or uncertainty in the                             
treatment planning process can have drastic consequences in the final dose location.                       
Additionally, because all primary beam particles stop in the patient, there is no current                           
method available to verify dose is delivered as planned. A solution to address these main                             
challenges limiting the full optimization of proton therapy is the subject of the following                           
sections and chapters of this work. 
1 
 
 
  
Physics of proton therapy 
Whereas a photon beam has an exponentially decreasing dose distribution after a                       
short build­up, the distribution of the proton beam approaches a maximum near the end of                             
its range. A heavy charged particle (e.g., proton) traveling through matter loses energy                         
primarily through the ionization and excitation of atoms. Except at low velocities, protons                         
lose a negligible amount of energy in nuclear collisions ​[14]​, and a proton can transfer only                               
a small fraction of its energy in a single electronic collision. Thus, a proton travels an almost                                 
straight path through matter, losing small amounts of energy almost continuously through                       
collisions with atomic electrons, leaving ionized and excited atoms in its wake. 
 
The average linear rate of energy loss for a proton in a medium is called the                               
stopping power—also termed the linear energy transfer (LET). Quantum mechanically, the                     
stopping power is the mean, or expectation, value of the linear rate of energy loss ​[14]​. This                                 
energy distribution can be described by the Bethe­Bloch equation ​[15,16]​. Mathematically,                     
the rate of energy loss per unit length for a proton with velocity, ​v​, can be formulated as 
, − dx
dE
= 4π
m ce 2 ∙ n
β2 ∙ ( e2
4πε0
)
2
∙ ln[ (2m c βe
2 2
I∙ 1−β( 2 
))− β2
]  (1) 
where E ​= proton energy 
x​ = proton depth in material 
e​ = electron charge 
m​e​ = electron rest mass 
c​ = speed of light in vacuum 
n​ = electron density 
  β = c
υ
 
I​ = mean excitation potential 
ε​0​ = vacuum permittivity. 
 
Driven by the term in the denominator, the stopping power increases in      β2
                 
proportion to 1/E as the proton slows down and approaches zero velocity ​[17]​. In other                             
words, as a proton slows down in tissue, it interacts with more electrons. These                           
interactions reach a maximum at the end of beam range, where the proton slows down,                             
collects electrons, and delivers the final dose over a very small area. The profile of the                               
energy loss as a function of distance is termed the Bragg peak. 
 
As shown in Figure 1, the Bragg peak offers two advantages over traditional photon                           
therapy: 1) the entry (proximal) dose delivered to the patient remains relatively low up                           
until the Bragg peak region, and 2) the peak dose delivery is found in a small, precise area                                   
near the end of the proton range. The absence of an exit dose offers the opportunity for                                 
highly conformal dose distributions, while simultaneously limiting the irradiation of                   
normal tissue. 
 
2 
 
 
  
 
 
 
Figure 1​. Depth­dose distributions for a Spread­Out Bragg Peak (SOBP, dashed blue line), its                           
constituent pristine Bragg peaks (thin blue lines), and a 10 MV photon beam (red). In a typical                                 
treatment plan for proton therapy, the SOBP is the therapeutic radiation distribution. The SOBP is the                               
sum of several individual Bragg peaks at staggered depths. The pink area represents additional doses                             
of photon therapy—which can damage normal tissues and cause secondary cancers, especially of the                           
skin ​[18]​. 
 
 
The depth of the Bragg peak in a patient is directly related to the initial energy,                               
E​initial​, of the charged particle. We can formulate the range, ​R​, of a proton in a homogenous                                 
material by assuming it enters the material with ​E​initial and summing the energy loss in very                               
thin slabs until the energy reaches some very low final value, ​E​final​, as 
.R (E )initial = ∫
Efinal
Einitial
(dx
dE
)
−1
  (2) 
This energy dependence allows the clinician to place the Bragg peak, and thus the dose,                             
anywhere in the patient. For irradiation of a tumor, the proton beam energy and intensity                             
are varied in order to achieve the desired dose over the tumor volume. A single clinical                               
proton field, in contrast to a single photon field, can achieve dose conformation to the                             
target volume ​[19]​. In general, proton therapy reduces irradiation to normal tissue, while                         
permitting dose escalation to levels not achievable with standard techniques, improving                     
clinical outcomes ​[20]​. 
 
By superimposing several pristine proton beams with different proton energies                   
(and hence different proton beam ranges), a Spread­Out Bragg Peak (SOBP) can be shaped                           
to precisely match tumor tissue. The total energy deposited in a patient (integral dose) for                             
3 
 
 
  
a given target dose is always lower for proton treatments when compared to photon                           
treatment techniques ​[21]​. The high level of control afforded by proton beams is preferred                           
for treating tumors located close to critical organs, such as the spinal cord, eye, and brain,                               
where even a small dose can cause very serious consequences. 
Range uncertainty 
One major factor limiting the clinical effectiveness of proton therapy is range                       
uncertainty. With current technology, there is no method to verify that the proton dose was                             
delivered as planned. Thus, clinicians are completely dependent on treatment planning                     
calculations for dose delivery. In principle, one should be able to use a proton beam pointed                               
directly at a critical structure, with the beam energy tuned to stop just short of the critical                                 
organ while delivering maximum dose to the tumor volume. However, range uncertainties                       
require the addition of substantial safety margins to over­irradiate the tumor. 
 
Paganetti ​[21] gives a thorough review of the types and causes of proton range                           
uncertainties. In summary, uncertainties result from four main categories: 1) patient                     
motion; 2) variations in setup and anatomy; 3) approximations used in dose calculations;                         
and 4) biological effects. With no way to verify dose delivery, clinicians add safety margins                             
to ensure total irradiation of the target volume. For example, at the Massachusetts General                           
Hospital (MGH), treatment plans add 3.5% of the range plus an additional 1 mm ​[21]​.                             
Although this guarantees the coverage of the distal aspect of the target volume, it also risks                               
overdosing the normal tissue behind the target volume. The potential for overdosing                       
changes the irradiation strategies used for certain cancers, potentially limiting the                     
usefulness of the treatment. 
 
The most prevalent example of the limitations of current methods is the treatment                         
of prostate cancer. Naturally, the most effective approach irradiates the tumor from the                         
anterior (Figure 2a) so that the sharp distal falloff of the proton beam can be used to treat                                   
the target volume while sparing the rectum. This would require, however, a precise control                           
of the beam range in the patient with millimeter accuracy, which is not currently possible.                             
As a result, anterior fields have never been used, despite the fact that such fields can utilize                                 
sharp distal penumbra (~4 mm for 50–95% falloff) ​[22]​. Instead, only lateral fields are                           
used (Figure 2b), relying solely on the much broader lateral beam penumbra (>10 mm for                             
50–95% falloff) and delivering a larger integral dose to the patient ​[22]​. 
 
As a result of these additional margins, tissues distal to the target volume receive a                             
substantial dose, and the dosimetric benefits of proton therapy are lost. Subsequently,                       
clinical outcomes of some common treatment sites, such as the prostate, are essentially                         
equivalent between proton and X­ray modalities ​[23]​. With no clear clinical benefit in cases                           
such as the prostate, questions arise as to whether proton therapy is worth the cost                             
(approximately double that of treatment with photons). 
 
 
 
4 
 
 
  
 
Figure 2​. Comparison of treatment planning dose distributions for proton therapy treatment of                         
prostate cancer with (a) one single anterior field and (b) two parallel­opposed lateral fields. Structures                             
of bladder, bladder wall, prostate, anterior rectal wall, rectum, and femoral heads are shown by cyan                               
lines. Adapted from ​[24]​. 
 
 
As a note of additional concern, the sharp distal dose falloff of the Bragg curve also                               
makes the dose distribution extremely sensitive to uncertainties in treatment planning and                       
patient setup. If planning calculations are off by 1 cm, the location of the distal falloff will                                 
change by 1 cm, causing either an undershoot, which is to under­irradiate the distal portion                             
of the target volume, or an overshoot, where the full dose is delivered to normal tissue                               
behind the target volume. 
Motivation and challenges 
Debates amongst medical physicists regarding the cost and efficacy of proton                     
therapy start with consideration for the limitations of currently available technologies                     
[25,26]​. As described above, the underlying physics of proton therapy offer clear potential                         
benefits in the radiotherapy treatment of cancer. Because of range uncertainty and the lack                           
of method to verify range at delivery, the benefits of proton therapy suggested by the                             
physical principles of charged particle interactions cannot be realized. When a reliable                       
method of range verification is incorporated into the cancer treatment process, the full                         
potential of proton therapy will be available for utilization by clinicians. 
 
The work described herein presents a system for imaging prompt gamma rays                       
emitted during proton therapy as a method of range verification. This innovative design is                           
the first collimated imaging system to implement two­dimensional (2­D) imaging for                     
prompt gamma proton beam range verification. This work is timely because as proton                         
treatment facilities continue to proliferate, accurate and effective range verification is                     
essential to ensuring the efficacy and optimization of the treatment method. 
5 
 
 
  
 
Due to treatment modalities currently available clinically, this work is focused                     
mainly on proton therapy. However, there is similarly much interest in other heavy ion                           
beams for cancer treatment, such as carbon ​[27–32]​. With physical and biological                       
differences that offer potential benefits over protons, carbon ions will likely find their way                           
into the clinical mainstream in the near future ​[33,34]​. However, the same challenges                         
regarding range verification exist with heavier ions. The method of range verification                       
proposed in this work could be directly applied to the case of carbon ions or any other                                 
heavy ion treatment. As with proton therapy, a reliable range verification technique would                         
have a positive impact on the clinical efficacy of any heavy ion therapy. The importance and                               
potential impact of this research cannot be overstated. 
 
In this work, Chapter 2 discusses the technical challenges associated with proton                       
range verification, such as high photon energy, high background environment, and a lack of                           
imaging devices suitable for proton therapy application. Other methods of range                     
verification under investigation are also discussed. Chapter 3 describes the benefits and                       
challenges in using simulation tools to aid with research efforts related to prompt gamma                           
imaging. Chapters 4–7 characterize the design and performance of an imaging system                       
developed to address these challenges, while Chapter 8 summarizes the implications of this                         
imaging system. The ultimate goal of this work is to determine proton beam range to an                               
accuracy of a few millimeters, which has the potential to make a significant impact on                             
proton therapy and the entire radiation oncology field. 
 
   
6 
 
 
  
Chapter 2—Techniques and Challenges of Range 
Verification 
 
A reliable method of range verification would improve the clinical capability of                       
proton therapy by allowing for more precise targeting during treatments and increase our                         
understanding of range uncertainties. Because all primary particles are stopped in­patient,                     
research has focused on using secondary emissions to verify proton range. Of the several                           
techniques currently being studied for in vivo range monitoring during proton therapy, the                         
two most developed methods involve the imaging of secondary coincident positron                     
annihilation photons and characteristic prompt gamma (PG) rays emitted as a result of                         1
proton beam irradiation of the patient ​[35]​. 
Positron emission tomography 
Positron emission tomography (PET) imaging has been widely studied for use in                       
proton range verification. This clinically tested technique involves moving the patient to a                         
PET scanner immediately after proton irradiation ​[36]​. The method has shown some ability                         
to accurately measure the in vivo proton beam range. However, its implementation has                         
been limited due to scanner resolution, positron energy, and biological washout ​[37]​.                       
Additionally, because proton beams lose energy mainly via electromagnetic interactions,                   
the activation image from positron emitters generated by nuclear reactions is not directly                         
correlated to the dose distribution. The established method uses a Monte Carlo calculated                         
distribution of the positron emitters and compares this predicted image with a measured                         
image. The accuracy of the Monte Carlo calculation depends on the underlying cross section                           
data ​[21]​. 
 
Efforts are underway to develop in­beam PET which would add an imaging system                         
to the proton gantry ​[38–40]​. This technique could eliminate problems associated with                       
biological washout; however, there are additional problems created with the geometry of a                         
conventional PET scanner, which, due to its large size limits its installation in the proton                             
treatment room, as well as a poor signal­to­noise ratio experienced with attempts of PET                           
verification during beam delivery ​[41]​. The positron emitting nuclides most produced                     
during proton therapy (​15​
O, ​11​
C, ​30​
P, and ​38​
K) have radioactive decay half­lives between                         
2–20 mintues, providing time for biological washout effects over the few minutes of                         
required image acquisition time. Alternatively, the positron endpoint energy of some                     
short­lived positron emitters of interest for PET imaging during the proton treatment,                       
produced in high concentrations and of clinical interest, is too high to offer any benefit in                               
1
Recent literature uses the terms “prompt gamma ray” and “prompt gamma” interchangeably. For                           
consistency and clarity, the acronym ​PG is used throughout the text in place of both terms. In either case, PG                                       
refers to the near­instantaneous photon emitted following the de­excitation of a nucleus. ​PGI refers to the                               
imaging of prompt gamma rays. 
7 
 
 
  
proton range verification applications (such as the 16.3 MeV positron endpoint for ​12​
N,                         
with a range of ~2 cm) ​[42]​. 
PG emission 
The drawbacks of PET have fueled interest in the development of prompt gamma                         
imaging (PGI) during proton therapy. PGs are emitted instantaneously (on the order of a                           
few tens of femtoseconds) following an inelastic collision of a proton with a target nucleus.                             
Like PET, PGs are produced by nuclear reactions. However, in PGI the measured photons                           
are emitted directly from the nucleus as opposed to PET photons emitted some distance                           
from their origin following an electron­positron annihilation. Thus, the effects of biological                       
washout and decay time seen with the PET method are nonexistent in PGI. In a direct                               
comparison with PET, PG was found to have an approximately 10­fold larger production                         
rate and to have a distribution physically much closer to the Bragg peak ​[43]​. Thus, the                               
spatial distribution (Figure 3) of the induced activity correlates better with absorbed dose                         
for PG as compared to PET ​[44]​. The PG production is highest at low proton energies, as                                 
shown in Figure 4. This results in more PG emissions in the vicinity of the Bragg peak. 
 
 
 
Figure 3. Simulated profiles for dose (black), PG (blue), and PET (red) integrated over the entire beam                                 
for an abdomen irradiated with pencil beams. All profiles are normalized to unity for easier                             
comparison. The x­axis represents the position along the beam direction relative to the isocenter.                           
Adapted from ​[43]​. 
 
 
Unlike PET, which uses a well­established imaging modality for photon energy of                       
511 keV, PGs pose an imaging challenge due to their high energy (2–10 MeV). The                             
advantage of focusing detection efforts on high­energy gamma rays is the ability to filter                           
out lower energy photons that are also produced by the proton beam, such as the 511 keV                                 
8 
 
 
  
photons, which are not produced as close to the Bragg peak as PGs. Classical gamma                             
cameras used in nuclear medicine are not adapted for detection of high­energy gammas in                           
the presence of an important neutron background, so dedicated cameras are needed.                       
Imaging high­energy photons requires specialized materials and methods. 
 
 
Table 1.​ Reaction channels that produce the most prominent gamma energies ​[45]​. 
 
Target  Reaction  E​γ​
 ​
(keV)  Half­life 
C  12​
C(p,p’)​12​
C  4443  45 fs 
  13​
C(p,d)​12​
C  4443  45 fs 
  12​
C(p,n)​12​
N  4443  11 ms (β​+​
 ➔ ​12​
C​*​
) 
  13​
C(p,2p)​12​
N  4443  11 ms (β​+​
 ➔ ​12​
C​*​
) 
N  14​
N(p,p’)​14​
N  1635  4.8 fs 
  14​
N(p,n)​14​
O  1635  71 s (β​+​
 ➔​14​
N​*​
) 
  14​
N(p,p’)​14​
N  2313  68 fs 
  14​
N(p,n)​14​
O  2313  71 s (β​+​
 ➔ ​14​
N​*​
) 
O  16​
O(p,pα)​12​
C  4443  45 fs 
  16​
O(p,nα)​12​
N  4443  11 ms (β​+​
 ➔ ​12​
C​*​
) 
  16​
O(p,d)​15​
O  5241  122 s (β​+​
 ➔ ​15​
N​*​
) 
  16​
O(p,2p)​15​
N  5270  17 fs 
  16​
O(p,p’)​16​
O  6129  18 ps 
 
 
A sampling of characteristic gamma rays produced during proton therapy is listed in                         
Table 1. It is important to note the many different proton­induced reaction channels                         
available to produce the same discrete­energy photons. This complexity adds to the                       
challenge of cross section measurements and accurate simulations (discussed in Chapter                     
3). In addition to PGs emitted immediately following a nuclear interaction, there are                         
contributions from beta­delayed gamma rays. Depending on the half­life of the beta decay,                         
the gamma rays emitted as a result of these reaction pathways may not be distinguishable                             
from the PGs. 
 
 
 
 
9 
 
 
  
(a) (b) 
 
Figure 4. Production cross sections for PGs resulting from inelastic collisions of protons with ​16​
O (a)                               
and ​12​
C (b). Cross sections are largest at low proton energy. Data from: ​[45–47]​. 
Detection and imaging of PGs 
Detector materials 
Particular care must be taken in choosing an appropriate material for detecting PGs.                         
Each method of detection has characteristic energy, spatial, and time resolution                     
requirements. For example, collimated systems require excellent spatial resolution and                   
counting rate capability. A Compton camera design requires good spatial, energy, and time                         
resolution to perform the necessary reconstruction calculations. If a detection design relies                       
on timing, such as for background subtraction, excellent timing resolution will be required. 
 
Scintillation detectors are optimal materials for PGI due to their relatively high                       
stopping power and good time and energy resolutions. Roemer ​et al. reported a detailed set                             
of measurements and comparisons with several scintillator materials ​[48]​. Their analysis of                       
energy and time resolution found sufficient performance amongst several scintillation                   
materials, with CeBr​3 performing best. At clinically relevant gamma energies and count                       
rates, they found CeBr​3 to have an energy resolution of 2.2% and a time resolution of 190                                 
ps full­width half­maximum (FWHM). 
 
In a similar effort, Hueso­González ​et al​. performed comparison measurements with                     
block detectors of LSO (Lu​2​SiO​5​:Ce) and BGO (Bi​4​Ge​3​O​12​), both commonly found on PET                         
detection systems ​[49]​. They focused on optimization for a Compton camera design, which                         
requires excellent spatial, energy, and time resolution. LSO had the better performance in                         
their tests, but BGO may be a suitable alternative based on its lower cost. 
 
10 
 
 
  
Detection techniques 
A sampling of the variety of PGI techniques currently under development is detailed                         
below. Each method makes a unique attempt at solving the same problem. Namely, to                           
achieve millimeter­range accuracy in determining the location of the Bragg peak using PGs. 
Compton cameras 
Multiple efforts are underway to utilize Compton imaging for PG detection ​[50–54]​.                       
This method uses the energy and angle dependence of a Compton scatter interaction to                           
produce an image, in a technique first described by Everett ​et al. ​[55]​. This detection                             
method requires a complex setup of two or more “stages.” If a photon undergoes a                             
Compton scatter in the first detector and subsequently deposits the remainder of its energy                           
in a second detector, a cone of possible incident angle can be back­projected into the image                               
space. The collection of such cones over time allows a reconstruction algorithm to                         
determine the source origin. 
 
A recent study of such a system was reported by Polf ​et al. ​[56]​. They produced                               
some promising results, demonstrating that their Compton camera system can detect                     
relative shifts of 3 mm and 5 mm in the Bragg peak location. However, the beam current,                                 
and thus the PG production rate, used in their measurements was much lower than that                             
used at most proton therapy treatment facilities. As a conservative estimate, let us assume                           
a beam current of 1.0 nA (typical beam currents are 1–3 nA ​[57]​) delivered to the patient                                 
(or 6.2 x 10​9
protons/s). If a representative treatment field of a daily treatment is 125 cGy                                 
(2.25 x 10​8
protons), the total beam­on time for the dose is 36 ms. Assuming a PG                                 
production rate of 0.1 per proton ​[58]​, there will be 6.2 x 10​8
PGs emitted per second (or                                   
620 MHz). A 100 cm​2
detector placed 10 cm away would thus experience a count rate of                                 
~50 MHz. The highest achievable double­scatter count rate for a Compton camera reported                         
by McCleskey ​et al​. ​[59] was 5 kHz, well below that needed for clinical delivery rates.                               
Because the Compton camera is severely limited by count rate (too many counts results in                             
false coincidences), the authors concede their system is not yet clinically viable. Further                         
work will continue on the Compton cameras, but until issues of timing and detection                           
efficiency are solved, the clinical application of such systems will be hindered. 
Electron­track Compton cameras 
A variant of the traditional Compton camera, electron­track Compton cameras                   
utilize a gas chamber detector to measure the energy and direction of the scattered                           
electron instead of measuring the energy and angle of the Compton­scattered photon. This                         
imaging method has also been demonstrated with solid­state detectors ​[60]​. Combining the                       
energy and direction of the electron with the position and energy deposited by the incident                             
gamma ray provides sufficient information to determine a cone of incident angles from the                           
source. Thus, the electron­track Compton camera only requires one interaction in the                       
detector to determine the origin direction of a photon.  
 
11 
 
 
  
Such a system has been described and analyzed via simulation by Kurosawa ​et al.                           
[61]​. Kurosawa ​et al​. limited proton beam current to 2.5 pA due to detector dead time                               
considerations, as opposed to the normal 1 nA beam current used in clinical treatments.                           
While the concept is interesting, there remains much work to gain feasibility, particularly                         
due to the limited angular resolution and limited count rate capabilities of current                         
technology. 
Gamma­electron vortex imaging 
A method proposed by Kim ​et al​. involves tracing Compton electrons, similar to                         
electron tracking Compton imaging ​[62]​. An “electron converter stage” converts the                     
high­energy gamma rays to electrons via Compton scatter. The system then determines the                         
direction of the electrons using a pair of hodoscopes. Based on the assumption that                           
electrons scatter in the forward direction, the lines of travel of the electrons are                           
back­projected to the image space. Like other non­collimated systems, the potential benefit                       
of this method is higher detection efficiency. However, the authors do not provide specifics                           
on spatial resolution, although they state that a prototype system is under development for                           
further study.  
PG timing 
Protons travel very quickly through tissues, but they still have a finite transit time:                           
approximately 1–2 ns for a proton with a 5–20 cm range. Because the transit time increases                               
with range, a time­resolved PG measurement could indicate the range of proton travel. This                           
PG timing approach was proposed and explored by Golnik ​et al. ​[58]​. Based on simulations                             
and initial measurements, the authors suggest that proton range could be determined                       
within 2 mm using a single scintillation crystal. More work remains to demonstrate the                           
clinical feasibility of the PG timing method, but the initial report is very promising ​[57]​. 
Collimated cameras 
The most straightforward PGI method is a collimated imaging system. Efforts on                       
collimated systems for PGI have focused on variations of a single­slit collimator design.                         
This method provides a one­dimensional (1­D) profile of PG emission, which would be                         
most suitable for identifying the PG falloff location by placing the imaging system                         
perpendicular to the beam axis. The single­slit is a very simple and straightforward concept                           
to implement, and often uses a knife­edge design to increase the field of view and allow for                                 
image magnification. A prototype knife­edge slit camera tested by Smeets ​et al. ​[63]                         
demonstrated 1–2 mm accuracy in determining the location of a Bragg peak at near                           
clinically relevant beam currents. Similarly, Bom ​et al. ​[64] reported better than 1 mm                           
accuracy in determining the PG distribution using a single­slit collimator. Realizing the                       
clinical potential and relative simplicity of the single­slit collimator design, a prototype PGI                         
system has been developed for clinical application ​[65]​. 
 
Options for alternative collimator designs, such as multi­slit, are limited due to the                         
properties of PGs. Collimators must be much thicker than those used in diagnostic imaging                           
12 
 
 
  
due to the high photon energy ​[66]​. Therefore, multi­slit collimators, offering a larger field                           
of view or potentially greater detection efficiency, are a challenge to design ​[67]​. 
 
A high level of neutron background radiation can also limit the effectiveness of a                           
collimated camera (as well as Compton cameras). Some attempts have successfully                     
demonstrated adjustments in the technique to overcome the high background limitation.                     
Using the timing characteristics of the cyclotron­produced proton beam, the PG signature                       
can be separated from background radiation with sufficient time resolution, thus                     
improving PGI performance ​[44]​. Time­of­flight (TOF) measurements have also been used                     
to suppress neutron background ​[68–70]​. 
Conclusion 
Due to production rates and proximity to the actual Bragg peak, PGI offers more                           
potential than PET for successful implementation as a proton range verification method.                       
However, there has yet to be a study published with results of PGI during patient                             
irradiations, and thus the clinical effectiveness of PGI remains to be shown. Non­collimated                         
systems, such as the Compton camera, are plagued by the high count rates experienced                           
during clinical treatment conditions. Additionally, the detection efficiency of systems                   
requiring multiple detector interactions for a single event is too low for clinical                         
consideration. Until solved, the detection efficiency and count rate problems make these                       
methods infeasible.  
 
The closest system to clinical use is the relatively simple knife­edge slit collimated                         
system. Multiple experiments have shown the potential to determine PG distributions with                       
1–2 mm accuracy. As prototype systems are developed and optimized for clinical use, we                           
may see promising results and approach clinical implementation. However, single­slit                   
systems are limited to providing 1­D information about the PG distribution. 
 
This work proposes a new multi­slit collimated imaging design to improve upon the                         
single­slit concept. As the PGI field moves forward in determining the dose distribution in a                             
patient, a 2­D image of PGs will be needed. A Compton camera system would be capable of                                 
providing a 2­D distribution of PGs, however the count rate limitation is not likely to be                               
overcome. The system introduced in this study actualizes the benefits offered by PG                         
detection, while addressing and overcoming several key challenges associated with PGI. 
 
Finally, unless a straightforward relationship between PG emission yield and dose                     
deposition can be established, PGI (similar to PET) will have to rely on the comparison of                               
the measured PG signal with a previously calculated or modeled expectation. Thus, clinical                         
applicability will require extensive experimental validation of PG yields independent of the                       
proton beam properties and the irradiated tissue type ​[22]​. Given the range uncertainties                         
associated with proton therapy and the high costs of building the many new proton therapy                             
centers across the globe, it is likely that research and development of range verification                           
methods will continue until an effective solution is found.   
13 
 
 
  
Chapter 3—A Monte Carlo Study of Prompt 
Gamma Production 
Background 
As computing power has increased per unit cost in recent decades, Monte Carlo                         
simulation methods have become more prevalent in scientific work. The field of medical                         
physics, in particular, has extensively used Monte Carlo simulations to model radiation                       
transport and to study radiation treatment modalities ​[71]​. Monte Carlo simulations offer                       
an opportunity to study complex or experimentally difficult processes, with a level of detail                           
and reproducibility not available to the experimentalist. For the PGI application, Monte                       
Carlo methods would be an invaluable tool for the purposes of 
 
1) optimizing an imaging system for PG detection; 
2) identifying and mapping the prompt gamma ray emission profile,                 
particularly in the inhomogeneous environment of a patient; and 
3) providing the means to convert a detected PG profile to a dose distribution. 
 
While proton therapy and proton range in tissue has been studied and validated in                           
Monte Carlo systems, the production of PGs has only recently become a topic of interest                             
and simulation due to range uncertainty problems in proton therapy. For dose calculations                         
of proton therapy, the electromagnetic interactions of protons are well known and can be                           
reliably simulated ​[72]​. The simulation of PG emission, however, relies on complex reaction                         
models that were initially developed for high­energy physics applications. While Monte                     
Carlo methods have been adapted to medical applications successfully, there are still                       
limitations on the modeling of nuclear excitation and de­excitation. Initial validation                     
studies of PG emission yields have not shown a consensus between simulated and                         
experimental results ​[73]​. Furthermore, different Monte Carlo codes produce different                   
results, as demonstrated in the case of production of positron emitters during proton                         
therapy ​[74]​. And unfortunately, the production of PGs depends on a far greater number of                             
reaction channels than for positron emitters due to the complex nuclear physics associated                         
with the excitation and de­excitation of the nuclei. 
 
As described by Verburg ​et al​. ​[75]​, the Monte Carlo nuclear reactions producing                         
PGs are modeled in three stages: 
 
1) Direct reactions: protons interacting directly with one or two                 
nucleons of the target. 
2) Pre­equilibrium: protons interact with parts of the nucleus before the                   
target has reached equilibrium. 
3) Compound reactions: energy of the proton is shared statistically                 
among target nucleons.  
14 
 
 
  
  
In the proton energy range of concern for therapeutic use (<200 MeV), all three                           
reaction stages are relevant and nuclear excitation with subsequent PG emission can occur                         
in any stage ​[75]​. The complex nature of these interactions, along with the multiple                           
pathways available (see Table 1), makes it difficult for a theoretical model to accurately                           
describe the results. This also complicates simulation of these processes. Considering that                       
the models used in Monte Carlo software were developed without consideration of the PG                           
productions at lower proton energies, it is understandable that Monte Carlo methods might                         
not generate accurate simulations. 
 
At the clinical proton energy range, Geant4 uses the Axen­Wellisch model ​[76] to                         
calculate total reaction cross sections. This model uses a general formula for the range of                             
proton energies from 6.8 MeV to 10 GeV. The model compares relatively well with                           
experimental data across this large energy range; however, the data set they use ​[77] is                             
very sparse at the lower energies. For example, nitrogen cross section data only goes down                             
to 23 MeV. A more comprehensive set of experimental data for lower proton energies                           
would allow us to validate this model or make adjustments to it. 
 
The work below adds to evidence that current Monte Carlo systems do not                         
adequately simulate the PG production during proton therapy, particularly to the level of                         
accuracy demanded when designing and validating systems for clinical use on patients.                       
First is a study and comparison of the simulated yield of PGs. Next, the angular distribution                               
of PGs is simulated and compared with experimental expectations. 
PG production cross sections 
Methods  
To analyze the production rates of PGs, simulations were run using TOPAS, ​TO​ol for                           
PA​rticle ​S​imulation ​[78]​, a user­friendly platform that interfaces with and runs the Geant4                         
Monte Carlo particle transport package ​[79]​. Beams of varying proton energies were                       
directed incident on thin targets of carbon and nitrogen. At each proton energy level, 10​11
                             
protons were simulated, with resulting photons scored by energy. The resulting PG                       
emissions were scored and used to calculate a gamma production cross section for each                           
gamma energy level according to Equation 3:  
,Σ = R
I∙N   (3) 
where: 
Σ = cross section (cm​2​
) 
R = total number of gammas produced (specific gamma energy for reaction of interest) 
I = total number of protons incident on the target 
N = number of target nuclei presented to the beam per unit area (cm​­2​
). 
15 
 
 
  
Results 
Figures 5–7 show gamma production cross sections for carbon and nitrogen targets                       
for both simulations and experimental data . For the 1.63 MeV gamma produced in                         2
nitrogen, the simulated cross section peaks around 20 mb as opposed to the experimental                           
value which reaches well over 80 mb. The 2.31 MeV gamma from nitrogen shows a similar                               
peak in simulated and experimental results; however, the simulated cross section results                       
are shifted to much higher proton energies. The results of the 4.4 MeV gamma in carbon                               
also show much lower production cross sections in the simulated results than the                         
experimental measurements. 
 
 
Figure 5. Comparison of simulation results with experimental data for 1.63 MeV gamma production                           
during proton irradiation of thin nitrogen target ​[45,47,80]​. 
 
 
These results demonstrate that the Geant4 Monte Carlo simulation package, in this                       
specific TOPAS configuration, does not adequately simulate PG production. Furthermore,                   
the lack of discernable pattern across each respective photon energy precludes the                       
identification of a simple correction or scaling factor that could make the package suitable                           
for PG simulation. Simulations generate unreliable total gamma counts because of                     
differences in total reaction production cross sections; this makes them unreliable to                       
produce and study PG emissions. Additionally, as seen with Figure 6, various proton                         
energies produce PGs in different locations of the simulated target than would be seen                           
experimentally. Thus, there are two main components of error: 1) the absolute cross                         
section and magnitude of PG production, and 2) the relative shape of the yield as a function                                 
2
Because of timing resolution limitations in experimental measurements, the number of photons                         
counted includes components from PG reactions as well as β­delayed reaction channels, as listed in Table 1.                                 
These reaction channels were also included in the simulation results. 
16 
 
 
  
of energy (i.e., depth). Until a Monte Carlo system can more accurately model the                           
production of PGs in tissue­related materials, simulation data will not be a reliable tool in                             
assessing and characterizing PG production. 
 
 
Figure 6. Comparison of simulation results with experimental data for 2.31 MeV gamma production                           
during proton irradiation of thin nitrogen target ​[45,47,80]​. 
 
 
 
Figure 7. Comparison of simulation results with experimental data for 4.44 MeV gamma production                           
during proton irradiation of thin carbon target ​[45,47]​. 
 
17 
 
 
  
Angular distribution of PGs 
Depending on the quantum properties of the excited nuclear level, PG emission is                         
not generally isotropic ​[75]​. The angular distribution of gamma rays with respect to the                           
beam direction is given by the Legendre polynomial expansion: 
,(θ) P (cosθ), (l even)W = ∑
l=lmax
l=0
al l     (4) 
where ​l​max​ is the smaller of the following two quantities: 
 
1. twice the spin of the decaying state and 
2. twice the multipolarity of the gamma ray ​[81,82]​. 
 
Measured gamma ray angular distributions produced by protons of up to 13 MeV                         
are shown in Figure 8 ​[47]​. With the exception of the 6.13 MeV gamma rays from ​16​
O, the                                   
gamma rays considered in this example have a multipolarity of 2 or less, so that ​l​max is 4 or                                     
less, and there are at most three terms (​l = 0, 2, 4) in the above expansion. In a similar study                                         
using 33 MeV protons, Lesko ​et al​. ​[45] found the angular distribution of PGs to be more                                 
isotropic than with the 8–13 MeV example. This evidence, provided in Figure 9 for                           
comparison, suggests an angular distribution dependence on incident proton energy. At                     
higher proton energies, more nuclear reaction channels are available and there are                       
contributions from a larger variety of excitation pathways. This explains the different                       
angular profile seen at higher proton energies, and re­emphasizes the complexity of the PG                           
modeling challenge. 
 
The angular distribution of PGs is important to consider when benchmarking Monte                       
Carlo simulation results to experimental measurements. As described above, and                   
corroborated by Verburg ​et al. ​[75] with Geant4 and MCNP ​[83] codes, Monte Carlo                           
simulations do not produce accurate PG yields within the acceptable range of experimental                         
data. However, these analyses included only angle­integrated cross sections, which is a                       
noted limitation of the study because PG emission is not usually isotropic. Robert ​et al​. ​[84]                               
compared two Monte Carlo code results for the angular distribution of secondary particles,                         
showing that both Geant4 and FLUKA ​[85] codes provide essentially isotropic gamma                       
emissions. The purpose of their study was to compare the two Monte Carlo packages, but                             
they do not include any discussion as to the absolute veracity of their simulation results. 
 
The angular distribution of PG emission is equally important as the total PG                         
production for consideration in actual measurements. Polf ​et al. ​[86] described some initial                         
results using PG detection to quantify the oxygen content in proton­irradiated tissue. Their                         
proposed application is quite promising, but without a correction for the angular                       
distribution of gammas (which in some cases may vary by a factor of ~2), the precision of                                 
tissue composition determination is severely limited. If a quantitative measure of PGs is to                           
be performed, an accounting of the angular distribution will be required. 
 
18 
 
 
  
 
 
Figure 8​. Examples of angular distributions of gamma rays. The solid curves are fit with an expansion                                 
in Legendre polynomials of even order through zero for the ​14​
N case, through 6 for the ​16​
O case, and                                     
through 4 for the other cases. The gamma ray energies (top to bottom) are 4.44, 2.31, 6.13, 1.63, 1.37,                                     
1.78, and 0.847 MeV. Adapted from ​[47]​. 
 
19 
 
 
  
 
Figure 9. Angular distributions for five separate gamma ray transitions resulting from bombardment                         
of Mylar, Mg, Si, and Fe targets with 33 MeV protons. The solid lines are the results of least squares fit                                         
of Legendre polynomials to the data. Adapted from ​[45]​. 
 
Methods 
To assess the ability of the Geant4 Monte Carlo package to simulate the PG angular                             
distribution, a simple simulation was performed with TOPAS using the default physics                       
models. A spherical scorer (i.e., perfect detector) was placed around a 1 cm­thick graphite                           
target. A beam of protons (10​9
total) at 13 MeV was directed incident on the target, and the                                   
resulting 4.4 MeV gammas were scored by the detecting sphere and binned in 3°                           
increments with respect to the beam direction. The proton energy and target material were                           
chosen to provide a comparison to the experimental data shown in Figure 6. 
 
Given that many researchers falsely assume isotropic PG emission, with some                     
simulation results reinforcing that misconception ​[84]​, an additional analysis was                   
performed to compare PG distributions as would be detected for isotropic emission with                         
the experimentally demonstrated angular distributions of emissions. 
 
20 
 
 
  
Using the TOPAS simulation, a poly(methyl methacrylate) (PMMA; (C​5​O​2​H​8​)​n​; and ⍴                     
= 1.18 g/cm​3​
) target was irradiated with 160 MeV protons to simulate a typical clinical                             
proton beam energy setting. The tissue­equivalent PMMA target was binned into 0.2                       
mm­thick slices, and the average proton energy in each slice was determined by simulation.                           
The slice thickness was chosen to avoid significant variation of average proton energy                         
across any individual target slice. Using data presented in Figure 2 ​[45–47]​, cross sections                           
for production of the 6.129 MeV gamma line from ​16​
O were calculated via interpolation.                           
Then, an analytical calculation was performed assuming a detector located at the Bragg                         
peak location (15.2 cm) and 90° relative to the beam direction. First, the PG emissions were                               
assumed isotropic and the proton energy in a slice was used in combination with the cross                               
section at that energy to determine the number of 6.129 MeV gammas produced. Next,                           
using an interpolation of the data presented in Figures 6 and 7, the angular distribution of                               
gamma emission was considered along with the angular position of the detector with                         
respect to the given target slice. The results approximate the difference to be expected                           
between an actual measurement and a measurement with an assumption of isotropic                       
gamma emission. 
Results 
Figure 10 shows the angular distribution of 4.4 MeV gammas resulting from the                         
simulation of 13 MeV protons incident on a graphite target. The gamma distribution                         
appears symmetric about a peak at 90°. This is quite different than the experimental results                             
shown in Figure 8 for the same proton energy and target nucleus. The experimentally                           
measured gamma distribution shows 90° at the lowest point in the PG emission. In this                             
case, the simulation model results are less accurate than if the model assumed an isotropic                             
distribution. 
 
In Figure 11, the blue line represents the 6.129 MeV gamma production in a PMMA                             
target, assuming that the yield is isotropic. This count total would be measured at the                             
detector if the gamma production was actually isotropic. The orange line is a more realistic                             
estimate of the gamma count measured at the detector given the angle of detection. As can                               
be inferred from Figure 8, at lower proton energies there is a higher yield of 6.13 MeV                                 
gamma as the detection angle moves from 90°. With a detector setup 90° from the Bragg                               
peak location, the resulting PG distribution measured would be broader than for the                         
isotropic case—a lower peak near the end of proton range, with some area of higher                             
magnitude proximal to the peak. 
 
21 
 
 
  
 
Figure 10. Angular distribution of 4.4 MeV PGs (Geant4). This is the result of 10​9
total protons incident                                   
on a graphite target. 13 MeV p + ​12​
C. 
 
 
 
Figure 11​. Comparison of theoretical isotropic PG distribution (blue) and expected gamma                       
distribution as detected based on angular distribution (orange). Each gamma profile represents the                         
6.129 MeV gamma production in a PMMA target as would be measured.   
22 
 
 
  
Conclusion 
These results indicate that current Monte Carlo methods do not accurately simulate                       
the excitation and de­excitation of the nuclei of interest in PGI for proton therapy (C, N, and                                 
O). The reaction cross sections, gamma emission yields, and the angular distribution of                         
emission for the specific photon energies are not correctly modeled. This important finding                         
identifies a limitation in the Monte Carlo simulation method—an important tool in studying                         
the problem at hand. While simulations can still be a valuable asset in designing a system to                                 
detect and image PGs, it is important to understand the limitations of the modeling                           
packages. Until new models are developed and incorporated into a Monte Carlo simulation                         
package, the simulation toolkit cannot be used to study the production and location of PG                             
emissions during proton therapy. This hinders the necessary development of 1) PG distal                         
falloff verification, and 2) PG­to­dose profile conversion. 
 
While building a new model of PG emissions is outside the scope of this project, it is                                 
noteworthy to identify this as an issue of importance to the topic. Future work to improve                               
the simulation modeling of the phenomenon would have a substantial positive impact in                         
the research community. 
 
 
 
 
   
23 
 
 
  
Chapter 4—A Novel Multi­Slit Collimated 
Imaging System 
Introduction 
As discussed in Chapter 2, the most promising method of range verification is PGI                           
using a collimated imaging system. While there are multiple challenges and limitations to                         
the use of collimated cameras due to the nature of PGs, the alternatives to collimated                             
systems have not proven clinically viable ​[56]​. 
 
Attempts to improve on the performance of single­slit collimators resulted in the                       
development of a novel multi­slit collimator concept ​[87]​. The complete imaging system                       
(Figure 12) consists of an array of cerium­doped lutetium oxyorthosilicate (LSO)                     
position­sensitive scintillation detectors paired with a multi­knife­edge slit collimator,                 
which has been designed, constructed, and characterized at Lawrence Berkeley National                     
Laboratory (LBNL). Each component of the imaging system is described in the following                         
sections. 
 
 
 
Figure 12. CAD model of imaging system showing 4 x 4 grid of LSO detector modules, tungsten                                 
collimator with individually cut pieces held together with 3­D printed plastic, and a proton beam                             
incident on a plastic target. 
 
 
 
 
24 
 
 
  
Collimator design 
A multi­knife­edge slit collimator has been designed for PGI applications. The 20 x                         
20 x 7.5 cm tungsten (Hevimet—90% W, 6% Ni, 4% Cu) collimator acts as a coded aperture                                 
system ​[88–90] for high­energy gamma rays that casts a unique projection for each point in                             
the imaging plane. An adaptation to the single­slit camera design similar to those used in                             
previous studies ​[63,64]​, the multi­slit concept was developed to improve detection                     
efficiency, provide a larger field of view, and offer higher resolution spatial information. To                           
provide adequate collimation, the material and thickness of the collimator were set such                         
that 95% of PG emissions would be attenuated if traveling through the thickest part of the                               
collimator. Knife­edge slits provide a larger field of view than parallel slits and also allow                             
the opportunity for magnification of the image. Multiple other collimator designs were                       
considered (for example, a parallel slit design and a fan beam design shown in Figure 13);                               
however, the minimum septa thickness required to provide proper collimation is limited by                         
the penetration depth of gammas in the MeV range. Thus, collimation schemes used in                           
diagnostic imaging require thicker septa, which precludes any significant advantage in                     
efficiency or spatial resolution over single­slit designs. 
 
 
 
Figure 13. ​Fan beam collimator design modeled in TOPAS. Tungsten slits of 1 mm thickness are                               
arranged with 2 mm wide openings that get wider as the distance from the object increases. Individual                                 
BGO detectors are placed at the top of the collimator. The design, originally proposed by Andy Haefner                                 
[91]​, did not offer significant detection efficiency advantages over a single slit collimator. 
 
 
25 
 
 
  
By placing slits at varying angles relative to the central vertical slit, we achieve an                             
imaging system similar to the 2­D coded aperture systems. A linear discriminant analysis                         
performed on simulated proton paths determined that such geometry has the ability to                         
discriminate with 10­fold higher sensitivity between proton paths with distal ends                     
separated by 1 mm, as compared with a single slit ​[87]​. After several iterations optimizing                             
for image contrast, uniformity in imaging sensitivity, and instrument cross­correlation                   
sidelobes, the design presented here was determined to be the best candidate for further                           
investigation (see ​[87]​).  
 
The collimator is composed of 22 individually cut blocks, arranged to provide a                         
predetermined pattern of knife­edge slits with 2 mm aperture at the center (Figure 14).                           
The tungsten blocks are held in place by a plastic frame. The collimator pieces are arranged                               
to imaging sensitivity across the image space that is closer to uniform than an unoptimized                             
system. Increasing the number of slits provides a greater geometric efficiency, while also                         
allowing the aperture size to decrease; in our case, 2 mm versus the 3 mm or greater                                 
proposed in other studies. Additionally, the angled slits offer the capability of the                         
collimator to produce 2­D images. The image reconstruction technique and initial                     
characterizations of the collimator are provided in Chapters 5–7. 
 
Figure 14.​ Arrangement and dimensions of tungsten collimator pieces. 
26 
 
 
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016
Ready-Dissertation-29June2016

More Related Content

What's hot

nuclear medicine and its application in dentistry
nuclear medicine and its application in dentistrynuclear medicine and its application in dentistry
nuclear medicine and its application in dentistrygaurav katheriya
 
20 medical physics techniques
20 medical physics techniques20 medical physics techniques
20 medical physics techniquesmrrayner
 
Performance Characteristics of the MIT Epithermal Neutron Irradiation Facility
Performance Characteristics of the MIT Epithermal Neutron Irradiation FacilityPerformance Characteristics of the MIT Epithermal Neutron Irradiation Facility
Performance Characteristics of the MIT Epithermal Neutron Irradiation Facilitykent.riley
 
Principles and application of PET CT & PET MR
Principles and application of PET CT & PET MRPrinciples and application of PET CT & PET MR
Principles and application of PET CT & PET MRcharusmita chaudhary
 
Radiotherapy With Protons
Radiotherapy  With  ProtonsRadiotherapy  With  Protons
Radiotherapy With Protonsfondas vakalis
 
pranaya ppt Nuclear medicine in surgery
pranaya ppt Nuclear medicine in surgerypranaya ppt Nuclear medicine in surgery
pranaya ppt Nuclear medicine in surgeryPRANAYA PANIGRAHI
 
Nuclear medicine 2
Nuclear medicine 2Nuclear medicine 2
Nuclear medicine 2Rad Tech
 
Sapphire - Interphex Presentation
Sapphire - Interphex PresentationSapphire - Interphex Presentation
Sapphire - Interphex PresentationJT Benavides
 
Berbeco et al, Low Z target switching to increase tumor endothelial cell dose...
Berbeco et al, Low Z target switching to increase tumor endothelial cell dose...Berbeco et al, Low Z target switching to increase tumor endothelial cell dose...
Berbeco et al, Low Z target switching to increase tumor endothelial cell dose...David Parsons
 
Bio-optical Sensing Dissertation
Bio-optical Sensing DissertationBio-optical Sensing Dissertation
Bio-optical Sensing DissertationRicky Hennessy
 
PET and SPECT Scanning: Functional Brain Imaging
PET and SPECT Scanning: Functional Brain ImagingPET and SPECT Scanning: Functional Brain Imaging
PET and SPECT Scanning: Functional Brain ImagingBrendan Quinn
 
Particle beam – proton,neutron & heavy ion therapy
Particle beam – proton,neutron & heavy ion therapyParticle beam – proton,neutron & heavy ion therapy
Particle beam – proton,neutron & heavy ion therapyAswathi c p
 
Sapphire - Interphex Presentation
Sapphire - Interphex PresentationSapphire - Interphex Presentation
Sapphire - Interphex PresentationSapphire USA Inc.
 
Photon Vs proton beam therapy
Photon Vs proton beam therapyPhoton Vs proton beam therapy
Photon Vs proton beam therapyDr Vijay Raturi
 
16 09-2013 technetium nuclear medicine
16 09-2013 technetium nuclear medicine16 09-2013 technetium nuclear medicine
16 09-2013 technetium nuclear medicineKonstantin German
 
FINAL PLASMON RESONANCE
FINAL PLASMON RESONANCEFINAL PLASMON RESONANCE
FINAL PLASMON RESONANCEGuillermo Lash
 

What's hot (20)

nuclear medicine and its application in dentistry
nuclear medicine and its application in dentistrynuclear medicine and its application in dentistry
nuclear medicine and its application in dentistry
 
20 medical physics techniques
20 medical physics techniques20 medical physics techniques
20 medical physics techniques
 
ISI
ISIISI
ISI
 
Performance Characteristics of the MIT Epithermal Neutron Irradiation Facility
Performance Characteristics of the MIT Epithermal Neutron Irradiation FacilityPerformance Characteristics of the MIT Epithermal Neutron Irradiation Facility
Performance Characteristics of the MIT Epithermal Neutron Irradiation Facility
 
Principles and application of PET CT & PET MR
Principles and application of PET CT & PET MRPrinciples and application of PET CT & PET MR
Principles and application of PET CT & PET MR
 
Radiotherapy With Protons
Radiotherapy  With  ProtonsRadiotherapy  With  Protons
Radiotherapy With Protons
 
pranaya ppt Nuclear medicine in surgery
pranaya ppt Nuclear medicine in surgerypranaya ppt Nuclear medicine in surgery
pranaya ppt Nuclear medicine in surgery
 
Nuclear medicine 2
Nuclear medicine 2Nuclear medicine 2
Nuclear medicine 2
 
Sapphire - Interphex Presentation
Sapphire - Interphex PresentationSapphire - Interphex Presentation
Sapphire - Interphex Presentation
 
Radionuclide imaging
Radionuclide imagingRadionuclide imaging
Radionuclide imaging
 
Nuclear Medicine - Rivin
Nuclear Medicine - RivinNuclear Medicine - Rivin
Nuclear Medicine - Rivin
 
Berbeco et al, Low Z target switching to increase tumor endothelial cell dose...
Berbeco et al, Low Z target switching to increase tumor endothelial cell dose...Berbeco et al, Low Z target switching to increase tumor endothelial cell dose...
Berbeco et al, Low Z target switching to increase tumor endothelial cell dose...
 
Bio-optical Sensing Dissertation
Bio-optical Sensing DissertationBio-optical Sensing Dissertation
Bio-optical Sensing Dissertation
 
PET and SPECT Scanning: Functional Brain Imaging
PET and SPECT Scanning: Functional Brain ImagingPET and SPECT Scanning: Functional Brain Imaging
PET and SPECT Scanning: Functional Brain Imaging
 
Particle beam – proton,neutron & heavy ion therapy
Particle beam – proton,neutron & heavy ion therapyParticle beam – proton,neutron & heavy ion therapy
Particle beam – proton,neutron & heavy ion therapy
 
Sapphire - Interphex Presentation
Sapphire - Interphex PresentationSapphire - Interphex Presentation
Sapphire - Interphex Presentation
 
Photon Vs proton beam therapy
Photon Vs proton beam therapyPhoton Vs proton beam therapy
Photon Vs proton beam therapy
 
Dual energy case study
Dual energy case studyDual energy case study
Dual energy case study
 
16 09-2013 technetium nuclear medicine
16 09-2013 technetium nuclear medicine16 09-2013 technetium nuclear medicine
16 09-2013 technetium nuclear medicine
 
FINAL PLASMON RESONANCE
FINAL PLASMON RESONANCEFINAL PLASMON RESONANCE
FINAL PLASMON RESONANCE
 

Similar to Ready-Dissertation-29June2016

Radiotherapy With Protons
Radiotherapy  With  ProtonsRadiotherapy  With  Protons
Radiotherapy With Protonsfondas vakalis
 
Tesi Master Judilka Bermudez
Tesi Master Judilka Bermudez Tesi Master Judilka Bermudez
Tesi Master Judilka Bermudez thinfilmsworkshop
 
Proton vda brochure
Proton vda brochureProton vda brochure
Proton vda brochureMiha Ulčar
 
Parsons and Robar, An investigation of kV CBCT image quality and dose reducti...
Parsons and Robar, An investigation of kV CBCT image quality and dose reducti...Parsons and Robar, An investigation of kV CBCT image quality and dose reducti...
Parsons and Robar, An investigation of kV CBCT image quality and dose reducti...David Parsons
 
7-RADIATION PROTECTION IN DENTISTRY===7=2023.pdf
7-RADIATION PROTECTION IN DENTISTRY===7=2023.pdf7-RADIATION PROTECTION IN DENTISTRY===7=2023.pdf
7-RADIATION PROTECTION IN DENTISTRY===7=2023.pdfNASERALHAQ
 
20.pet scan in oncology
20.pet scan in oncology20.pet scan in oncology
20.pet scan in oncologyArnab Bose
 
BMES_Handheld_SPECT_poster_2015
BMES_Handheld_SPECT_poster_2015BMES_Handheld_SPECT_poster_2015
BMES_Handheld_SPECT_poster_2015Surabhi Nair
 
molecular imaging with PET & SPECT
molecular imaging with PET & SPECTmolecular imaging with PET & SPECT
molecular imaging with PET & SPECTShatha M
 
PH40003-4_A1_Poster Adam McMurchie
PH40003-4_A1_Poster Adam McMurchiePH40003-4_A1_Poster Adam McMurchie
PH40003-4_A1_Poster Adam McMurchieAdam Mcmurchie
 
Projection Radiography and its Applications
Projection Radiography and its ApplicationsProjection Radiography and its Applications
Projection Radiography and its ApplicationsKofiDeh1
 
Getting Ahead of the Expanding Landscape: Radiopharmaceutical Dosimetry
Getting Ahead of the Expanding Landscape: Radiopharmaceutical DosimetryGetting Ahead of the Expanding Landscape: Radiopharmaceutical Dosimetry
Getting Ahead of the Expanding Landscape: Radiopharmaceutical DosimetryMedpace
 
Vcu Chemistry Reasearch Facilities
Vcu Chemistry Reasearch FacilitiesVcu Chemistry Reasearch Facilities
Vcu Chemistry Reasearch FacilitiesJoseph Turner 'Jody'
 
Proteomics Practical (NMR and Protein 3D software
Proteomics Practical (NMR and Protein 3D softwareProteomics Practical (NMR and Protein 3D software
Proteomics Practical (NMR and Protein 3D softwareiqraakbar8
 
A biologically inspired cmos image sensor
A biologically inspired cmos image sensorA biologically inspired cmos image sensor
A biologically inspired cmos image sensorSpringer
 
A biologically inspired cmos image sensor
A biologically inspired cmos image sensorA biologically inspired cmos image sensor
A biologically inspired cmos image sensorSpringer
 

Similar to Ready-Dissertation-29June2016 (20)

Radiotherapy With Protons
Radiotherapy  With  ProtonsRadiotherapy  With  Protons
Radiotherapy With Protons
 
Tesi Master Judilka Bermudez
Tesi Master Judilka Bermudez Tesi Master Judilka Bermudez
Tesi Master Judilka Bermudez
 
Proton vda brochure
Proton vda brochureProton vda brochure
Proton vda brochure
 
Parsons and Robar, An investigation of kV CBCT image quality and dose reducti...
Parsons and Robar, An investigation of kV CBCT image quality and dose reducti...Parsons and Robar, An investigation of kV CBCT image quality and dose reducti...
Parsons and Robar, An investigation of kV CBCT image quality and dose reducti...
 
7-RADIATION PROTECTION IN DENTISTRY===7=2023.pdf
7-RADIATION PROTECTION IN DENTISTRY===7=2023.pdf7-RADIATION PROTECTION IN DENTISTRY===7=2023.pdf
7-RADIATION PROTECTION IN DENTISTRY===7=2023.pdf
 
20.pet scan in oncology
20.pet scan in oncology20.pet scan in oncology
20.pet scan in oncology
 
BMES_Handheld_SPECT_poster_2015
BMES_Handheld_SPECT_poster_2015BMES_Handheld_SPECT_poster_2015
BMES_Handheld_SPECT_poster_2015
 
Positron emission tomography
Positron emission tomographyPositron emission tomography
Positron emission tomography
 
molecular imaging with PET & SPECT
molecular imaging with PET & SPECTmolecular imaging with PET & SPECT
molecular imaging with PET & SPECT
 
PH40003-4_A1_Poster Adam McMurchie
PH40003-4_A1_Poster Adam McMurchiePH40003-4_A1_Poster Adam McMurchie
PH40003-4_A1_Poster Adam McMurchie
 
Projection Radiography and its Applications
Projection Radiography and its ApplicationsProjection Radiography and its Applications
Projection Radiography and its Applications
 
Getting Ahead of the Expanding Landscape: Radiopharmaceutical Dosimetry
Getting Ahead of the Expanding Landscape: Radiopharmaceutical DosimetryGetting Ahead of the Expanding Landscape: Radiopharmaceutical Dosimetry
Getting Ahead of the Expanding Landscape: Radiopharmaceutical Dosimetry
 
Vcu Chemistry Reasearch Facilities
Vcu Chemistry Reasearch FacilitiesVcu Chemistry Reasearch Facilities
Vcu Chemistry Reasearch Facilities
 
Proteomics Practical (NMR and Protein 3D software
Proteomics Practical (NMR and Protein 3D softwareProteomics Practical (NMR and Protein 3D software
Proteomics Practical (NMR and Protein 3D software
 
IGRT APP.pdf
IGRT APP.pdfIGRT APP.pdf
IGRT APP.pdf
 
A biologically inspired cmos image sensor
A biologically inspired cmos image sensorA biologically inspired cmos image sensor
A biologically inspired cmos image sensor
 
A biologically inspired cmos image sensor
A biologically inspired cmos image sensorA biologically inspired cmos image sensor
A biologically inspired cmos image sensor
 
2D NMR.pptx
2D NMR.pptx2D NMR.pptx
2D NMR.pptx
 
Siena16_Bennati
Siena16_BennatiSiena16_Bennati
Siena16_Bennati
 
Digital cephalometry
Digital cephalometryDigital cephalometry
Digital cephalometry
 

Ready-Dissertation-29June2016

  • 3. Abstract  Development of a multi­knife­edge slit collimator for prompt gamma ray imaging during  proton beam cancer therapy  by  John Francis Ready III  Doctor of Philosophy in Nuclear Engineering  University of California, Berkeley  Professor Kai Vetter, Chair    Proton beam usage to treat cancer has recently experienced rapid growth, as it                          offers the ability to target dose delivery in a patient more precisely than traditional x­ray                              treatment methods. Protons stop within the patient, delivering the maximum dose at the                          end of their track—a phenomenon described as the Bragg peak. However, because a large                            dose is delivered to a small volume, proton therapy is very sensitive to errors in patient                                setup and treatment planning calculations. Additionally, because all primary beam particles                      stop in the patient, there is no direct information available to verify dose delivery. These                              factors contribute to the range uncertainty in proton therapy, which ultimately hinders its                          clinical usefulness. A reliable method of proton range verification would allow the clinician                          to fully utilize the precise dose delivery of the Bragg peak.    Several methods to verify proton range detect secondary emissions, especially                    prompt gamma ray (PG) emissions. However, detection of PGs is challenging due to their                            high energy (2–10 MeV) and low attenuation coefficients, which limit PG interactions in                          materials. Therefore, detection and collimation methods must be specifically designed for                      prompt gamma ray imaging (PGI) applications. In addition, production of PGs relies on                          delivering a dose of radiation to the patient. Ideally, verification of the Bragg peak location                              exposes patients to a minimal dose, thus limiting the PG counts available to the imaging                              system.    An additional challenge for PGI is the lack of accurate simulation models, which limit                            the study of PG production characteristics and the relationship between PG distribution                        and dose delivery. Specific limitations include incorrect modeling of the reaction cross                        sections, gamma emission yields, and angular distribution of emission for specific photon                        energies. While simulations can still be valuable assets in designing a system to detect and                              image PGs, until new models are developed and incorporated into Monte Carlo simulation                          packages, simulations cannot be used to study the production and location of PG emissions                            during proton therapy.  1 
  • 4.   This work presents a novel system to image PGs emitted during proton therapy to                            verify proton beam range. The imaging system consists of a multi­slit collimator paired                          with a position­sensitive LSO scintillation detector. This innovative design is the first                        collimated imaging system to implement two­dimensional (2­D) imaging for PG proton                      beam range verification, while also providing a larger field of view than compared to                            single­slit collimator systems. Other, uncollimated imaging systems have been explored for                      PGI applications, such as Compton cameras. However, Compton camera designs are                      severely limited by counting rate capabilities. A recent Compton camera study reported                        count rate capability of about 5 kHz. However, at a typical clinical beam current of 1.0 nA,                                  the estimated PG emission rate would be 6 x 10​8 per second. After accounting for distance                                to the detector and interaction efficiencies, the detection system will still be overwhelmed                          with counts in the MHz range, causing false coincidences and hindering the operation of the                              imaging system.    Initial measurements using 50 MeV protons demonstrated the ability of our system                        to reconstruct 2­D PG distributions at clinical beam currents. A Bragg peak localization                          precision of 1 mm (2σ) was achieved with delivery of (1.7 ± 0.8) x 10​8 protons into a                                    poly(methyl methacrylate) (PMMA) target, suggesting the ability of the system to detect                        relative shifts in proton range while delivering fewer protons than used in a typical                            treatment fraction. This is key, as the ideal system allows the clinician to verify proton                              range when delivering only a small portion of the prescribed dose, preventing the                          mistreatment of the patient. Additionally, the absolute position of the Bragg peak was                          identified to within 1.6 mm (2σ) with 5.6 x 10​10​  protons delivered.     These promising results warrant further investigation and system optimization for                    clinical implementation. While further measurements at clinical beam energy levels will be                        required to verify system performance, these preliminary results provide evidence that 2­D                        image reconstruction, with 1–2 mm accuracy, is possible with this design. Implementing                        such a system in the clinical setting would greatly improve proton therapy cancer                          treatment outcomes.  2 
  • 5. Dedication     To my wife Miriam—for your unending support, patience, and love. I wouldn’t have                          gotten through this doctorate if not for you. I strive each day to make you as proud of me as                                        I am of you.    To my son Logan and daughter Lexi—you both came into the world during my                            graduate school career and have completely changed my perspective on life. Your smiles                          and love have provided me with more motivation than you will ever know.  i 
  • 6.   Table of Contents      List of Figures  iv      List of Tables  vii      Acknowledgements   viii      Chapter 1—Introduction  1  Protons for cancer treatment  1  Physics of proton therapy  2  Range uncertainty  4  Motivation & challenges  5      Chapter 2—Techniques and Challenges of Range Verification  7  Positron emission tomography  7  PG emission  8  Detection and imaging of PGs  10  Detector materials  10  Detection techniques  11  Compton cameras  11  Electron­track Compton cameras  11  Gamma­electron vortex imaging  12  PG timing  12  Collimated cameras  12  Conclusion  13      Chapter 3—A Monte Carlo Study of Prompt Gamma Production  14  Background  14  PG production cross sections  15  Methods  15  Results  16  Angular distribution of PGs  18  Methods  20  Results  21  Conclusion  23      Chapter 4—A Novel Multi­Slit Collimated Imaging System  24  Introduction  24  Collimator design  25  Detector configuration  27  Conclusion  30      ii 
  • 7. Chapter 5—Data acquisition, data processing, and image reconstruction  31  Acquisition hardware  31  Detectors  32  High­voltage power supply  32  Analog­to­Digital Converter  33  Acquisition software and data processing  35  Acquisition  35  Data processing  37  Image reconstruction  40      Chapter 6—Simulation Study of Collimator Performance  43  Introduction  43  Methods and materials  43  Results and discussion  46  Conclusion  48      Chapter 7—Experimental Characterization of System Performance  50  Introduction  50  Methods and materials  50  Results and discussion  52  Image reconstruction  52  Range measurement  54  Conclusion  59      Chapter 8—Conclusion  61  Summary  61  Key challenges remaining  62  Directly applicable to this PGI system  62  Detector design  62  Optimization of the image reconstruction algorithm  63  Collimator design  63  PGI field at large  64  Monte Carlo simulations of PG production  64  Effect of neutrons on imaging system  64  Development of correlation between PG distribution and         Bragg peak location  64  Inhomogeneous target materials  64  Clinical implementation of proton range verification  65  Final thoughts  65      References  67    iii 
  • 8.   List of Figures  Figure 1  Depth­dose distributions for a Spread­Out Bragg Peak (SOBP), its                  constituent pristine Bragg peaks, and a 10 MV photon beam.  3  Figure 2  Comparison of treatment planning dose distributions for proton therapy                  treatment of prostate cancer with one single anterior field and two                      parallel­opposed lateral fields.  5  Figure 3  Simulated profiles for dose, PG, and PET integrated over the entire beam                        for an abdomen irradiated with pencil beams.  8  Figure 4  Production cross sections for PGs resulting from inelastic collisions of                    protons with ​16​ O and ​12​ C.  10  Figure 5  Comparison of simulation results with experimental data for 1.63 MeV                    gamma production during proton irradiation of thin nitrogen target.  16  Figure 6  Comparison of simulation results with experimental data for 2.31 MeV                    gamma production during proton irradiation of thin nitrogen target.  17  Figure 7  Comparison of simulation results with experimental data for 4.44 MeV                    gamma production during proton irradiation of thin carbon target.  17  Figure 8  Examples of angular distributions of gamma rays.  19  Figure 9  Angular distributions for five separate gamma ray transitions resulting                  from bombardment of Mylar, Mg, Si, and Fe targets with 33 MeV protons.  20  Figure 10  Angular distribution of 4.4 MeV PGs (Geant4).  22  Figure 11  Comparison of theoretical isotropic PG distribution and expected gamma                  distribution as detected based on angular distribution.  22  Figure 12  CAD model of imaging system showing 4 x 4 grid of LSO detector modules,                            tungsten collimator with individually cut pieces held together with 3­D                    printed plastic, and a proton beam incident on a plastic target.  24  Figure 13  Fan beam collimator design modeled in TOPAS.  25  Figure 14  Arrangement and dimensions of tungsten collimator pieces.  26  Figure 15  CAD model to demonstrate the position and layout of LSO detector                      modules.  28  Figure 16  Photographs of bottom two rows of detector modules.  28  Figure 17  Photograph of a detector module and connections.  29  Figure 18  Gain adjustment on PMTs is performed by making slight adjustments in the                        respective gain adjustment potentiometer.  29  Figure 19  Photo of acquisition system hardware.  31  Figure 20  Electronics in VME crate.  32  Figure 21  Screenshot of HV power supply control interface.  33  Figure 22  Image of the four SIS33316 Struck ADCs.  34  iv 
  • 9.   Figure 23  The 4 x 4 detector array forms the ​L​S​O ​GA​mma detection i​N​strument                        (LOGAN) that is powered by the HV power supply to transfer the light                          signal produced in the scintillation crystal to an electrical signal generated                      in the PMTs which is then sent to the ADCs.  34  Figure 24  Back view of the ​L​EMO­to­​E​thernet for ​X​­ray ​I​maging (LEXI) apparatus.  35  Figure 25  Firmware schematic of 4 Channel Sum Trigger sequence.  36  Figure 26  Screenshot of raw signal data from a detector module.   37  Figure 27  Image representing the PMT configuration in relation to the segmented                    scintillator crystal.  38  Figure 28  Example map of detector interactions as calculated using Equations 5 and                      6.  38  Figure 29  Histogram of gain correction factors used to ensure a more uniform                      response across the detector array pixels.  39  Figure 30  Map of gain correction factors used to account for variations in counting                        rates of individual detector pixels.  39  Figure 31  Example of ray tracing result for a point source projection through the                        collimator onto the detector plane.  41  Figure 32  Sensitivity map generated by the ray tracing algorithm.  42  Figure 33  Simulation of collimator and flat detector panel in TOPAS.  44  Figure 34  Simulated point sources shifted in 1 cm steps along the x­direction.  45  Figure 35  1­D reconstructed image of point sources at varying locations which                    demonstrates, via FWHM,the spatial resolution across one dimension.  46  Figure 36  2­D reconstruction of a simulated point source at (x, y) position (15, 5).  47  Figure 37  2­D reconstruction of a simulated point source at (x ,y) position (5, 10).  47  Figure 38  Input and reconstructed response of a 4.4 MeV line source.   48  Figure 39  Proton range in tissue­equivalent plastic (PMMA) for proton energy 0–170                    MeV and 0­50 MeV.  51  Figure 40  Photo of experimental setup at 88­Inch Cyclotron.  51  Figure 41  Photo of PMMA target in place for Bragg peak measurements with 50 MeV                          proton beam.  52  Figure 42  2­D image reconstruction of 50 MeV proton beam in PMMA target.  53  Figure 43  1­D image of PG distribution plotted with the simulated Bragg curve with                        dose normalized to the maximum integral depth dose.  53  Figure 44  Range retrieval precision (2 ) versus the number of delivered protons.  54  Figure 45  Estimated target position and deviation from trend line.  55  Figure 46  Estimated Bragg peak location and residuals.  56  Figure 47  2­D image reconstruction of 50 MeV proton beam in the thick PMMA                        target, separated by photon energy.  57  v 
  • 10.   Figure 48  1­D images of PG distribution, sorted by photon energy, plotted with the                        simulated Bragg curve with dose normalized to the maximum integral                    depth dose.  57  Figure 49  2­D image reconstruction of 50 MeV proton beam in the thick PMMA                        target, separated by photon energy with target location shifted 3 mm to the                          right relative to Figure 47.  58  Figure 50  1­D images of PG distribution, sorted by photon energy, plotted with the                        simulated Bragg curve with dose normalized to the maximum integral                    depth dose with target location shifted 3 mm to the right relative to Figure                            48.  58  Figure 51  Positional uniformity of detection system.  59  Figure 52  Energy­integrated and discrete PG emissions along the path of proton                    pencil­beams in water.   60  Figure 53  Age­adjusted invasive cancer incidence rates in 2012 for the ten primary                      sites with the highest rates in men.  66    vi 
  • 11. List of Tables  Table 1  Reaction channels that produce the most prominent gamma                energies.  9  Table 2  Basic properties of LSO scintillation detectors.  27    vii 
  • 12.   Acknowledgements    "​... to know even one life has breathed easier because you have lived—this is to                              have succeeded.​"—Bessie Anderson Stanley, as abridged by Albert Edward                  Wiggam    My time at Berkeley has been filled with opportunities for personal and professional                          growth. I have been humbled and honored by the chance to serve my communities: in                              particular, veterans and graduate students through involvement in the Graduate Assembly.                      The time and effort I have committed to fostering a vibrant community at Cal has been                                returned to me many times over in emotional and social support and encouragement, and                            it introduced me to mentors such as Ron Williams, Dean Joseph Greenwell, and Vice                            Chancellor Harry Le Grande. While a PhD is largely viewed as an accomplishment of the                              individual, as with any major personal endeavor, my PhD and dissertation would not be                            possible without the support of a loving community. I therefore attribute the successful                          closure of this chapter of my life to the vast array of humans that I have had the great                                      pleasure of knowing and working with over the past six years.    I would like to express my deepest gratitude to my supervisor Professor Kai Vetter                            for his unwavering support, collegiality, and mentorship throughout this project.    I wish to thank my committee members who were more than generous with their                            expertise and precious time.    I would like to extend my thanks to those who offered invaluable research guidance                            and support over the years: Victor Negut; Rebecca Pak; Ryan Pavlovsky; Sam Huh; Tim                            Aucott; Andy Haefner; Ross Barnowski; Don Gunter; Lucian Mihailescu; Justin Ellin; Joseph                        Perl; Bill Moses; fellow graduate students, postdocs, and staff scientists of the Applied                          Nuclear Physics Program; and countless others who have had a positive impact on my                            work.    Thanks to my dad, John; mom, Micki; mother­in­law, Lucy; brothers, Joey and  Jimmy; and all my family members, friends, fellow student veterans, Graduate Assembly  colleagues, and anyone else that has knowingly or unknowingly helped me along my path.    Finally, I offer my sincere appreciation to the staff and facilities at Lawrence                          Berkeley National Laboratory; the faculty, staff, and students who supported my                      educational experience at UC Berkeley; and my funding source—the Nuclear Science and                        Security Consortium—for providing me with the opportunity and freedom to contribute to                        the scientific knowledge of the important field of nuclear science.    viii 
  • 13. This material is based upon work supported by the Department of Energy National                          Nuclear Security Administration under Award Number: DE­NA0000979 through the                  Nuclear Science and Security Consortium.    This report was prepared as an account of work sponsored by an agency of the                              United States Government. Neither the United States Government nor any agency thereof,                        nor any of their employees, makes any warranty, express or limited, or assumes any legal                              liability or responsibility for the accuracy, completeness, or usefulness of any information,                        apparatus, product, or process disclosed, or represents that its use would not infringe                          privately owned rights. Reference herein to any specific commercial product, process, or                        service by trade name, trademark, manufacturer, or otherwise does not necessarily                      constitute or imply its endorsement, recommendation, or favoring by the United States                        Government or any agency thereof. The views and opinions of authors expressed herein do                            not necessarily state or reflect those of the United States Government or any agency                            thereof.      ix 
  • 14.    Chapter 1—Introduction  Protons for cancer treatment  The use of radiation for the treatment of cancer began very shortly after the                            discovery of X­rays in 1895 ​[1]​. With advances in technology and an increasing                          understanding of radiation, radiotherapy has now become a standard treatment option for                        a wide range of malignancies ​[2]​. Today, approximately 50% of all patients with localized                            malignant tumors are treated with radiation ​[3]​. In theory, any tumor can be killed with a                                large enough dose. However, the tolerance of healthy tissue surrounding the tumor volume                          limits the radiation dose permitted in practice. Technical advances in radiation therapy                        have been aimed mainly at reducing dose to healthy tissue while maintaining or increasing                            the dose to the tumor volume. In addition to technical innovations in the “traditional”                            photon and electron radiotherapy methods, physicists have searched for alternative                    particles that offer advantages in their dose deposition and biological characteristics ​[4]​.    First described in 1946 by Robert Wilson ​[5]​, the clinical potential of proton beams                            lies in the basic physics of heavy charged particle interactions in matter. Wilson recognized                            that the depth­dose profile of protons in a patient holds dosimetric advantages over                          traditional photon (X­ray) therapy, which sparked his interest in developing protons for                        use in tumor treatment.    The first clinical use of protons occurred at Lawrence Berkeley Laboratory in 1957                          [6]​. The work at Berkeley confirmed the predictions of Wilson and led to the use of heavy                                  charged particles in treating human diseases associated with the malfunctioning of the                        pituitary gland ​[7,8]​. The following decades saw gradual growth in both the types of                            cancers treated with protons and the number of medical centers offering proton therapy                          [9]​. More recently, there has been a rapid proliferation of proton therapy centers, growing                            from approximately 30 to 60 centers worldwide from 2010 to 2015 (with another 30                            under construction in 2016) ​[10]​. This rapid expansion of proton therapy has been met                            with questions and controversy, as clinicians try to balance efficacy and rising costs of                            treatment ​[11,12]​.    The properties of proton therapy introduce unique challenges to the field of medical                          physics ​[13]​. As the peak dose is delivered over a relatively small volume, proton therapy is                                very sensitive to errors in radiation delivery. Mistakes in patient setup or uncertainty in the                              treatment planning process can have drastic consequences in the final dose location.                        Additionally, because all primary beam particles stop in the patient, there is no current                            method available to verify dose is delivered as planned. A solution to address these main                              challenges limiting the full optimization of proton therapy is the subject of the following                            sections and chapters of this work.  1     
  • 15.    Physics of proton therapy  Whereas a photon beam has an exponentially decreasing dose distribution after a                        short build­up, the distribution of the proton beam approaches a maximum near the end of                              its range. A heavy charged particle (e.g., proton) traveling through matter loses energy                          primarily through the ionization and excitation of atoms. Except at low velocities, protons                          lose a negligible amount of energy in nuclear collisions ​[14]​, and a proton can transfer only                                a small fraction of its energy in a single electronic collision. Thus, a proton travels an almost                                  straight path through matter, losing small amounts of energy almost continuously through                        collisions with atomic electrons, leaving ionized and excited atoms in its wake.    The average linear rate of energy loss for a proton in a medium is called the                                stopping power—also termed the linear energy transfer (LET). Quantum mechanically, the                      stopping power is the mean, or expectation, value of the linear rate of energy loss ​[14]​. This                                  energy distribution can be described by the Bethe­Bloch equation ​[15,16]​. Mathematically,                      the rate of energy loss per unit length for a proton with velocity, ​v​, can be formulated as  , − dx dE = 4π m ce 2 ∙ n β2 ∙ ( e2 4πε0 ) 2 ∙ ln[ (2m c βe 2 2 I∙ 1−β( 2  ))− β2 ]  (1)  where E ​= proton energy  x​ = proton depth in material  e​ = electron charge  m​e​ = electron rest mass  c​ = speed of light in vacuum  n​ = electron density    β = c υ   I​ = mean excitation potential  ε​0​ = vacuum permittivity.    Driven by the term in the denominator, the stopping power increases in      β2                   proportion to 1/E as the proton slows down and approaches zero velocity ​[17]​. In other                              words, as a proton slows down in tissue, it interacts with more electrons. These                            interactions reach a maximum at the end of beam range, where the proton slows down,                              collects electrons, and delivers the final dose over a very small area. The profile of the                                energy loss as a function of distance is termed the Bragg peak.    As shown in Figure 1, the Bragg peak offers two advantages over traditional photon                            therapy: 1) the entry (proximal) dose delivered to the patient remains relatively low up                            until the Bragg peak region, and 2) the peak dose delivery is found in a small, precise area                                    near the end of the proton range. The absence of an exit dose offers the opportunity for                                  highly conformal dose distributions, while simultaneously limiting the irradiation of                    normal tissue.    2     
  • 16.          Figure 1​. Depth­dose distributions for a Spread­Out Bragg Peak (SOBP, dashed blue line), its                            constituent pristine Bragg peaks (thin blue lines), and a 10 MV photon beam (red). In a typical                                  treatment plan for proton therapy, the SOBP is the therapeutic radiation distribution. The SOBP is the                                sum of several individual Bragg peaks at staggered depths. The pink area represents additional doses                              of photon therapy—which can damage normal tissues and cause secondary cancers, especially of the                            skin ​[18]​.      The depth of the Bragg peak in a patient is directly related to the initial energy,                                E​initial​, of the charged particle. We can formulate the range, ​R​, of a proton in a homogenous                                  material by assuming it enters the material with ​E​initial and summing the energy loss in very                                thin slabs until the energy reaches some very low final value, ​E​final​, as  .R (E )initial = ∫ Efinal Einitial (dx dE ) −1   (2)  This energy dependence allows the clinician to place the Bragg peak, and thus the dose,                              anywhere in the patient. For irradiation of a tumor, the proton beam energy and intensity                              are varied in order to achieve the desired dose over the tumor volume. A single clinical                                proton field, in contrast to a single photon field, can achieve dose conformation to the                              target volume ​[19]​. In general, proton therapy reduces irradiation to normal tissue, while                          permitting dose escalation to levels not achievable with standard techniques, improving                      clinical outcomes ​[20]​.    By superimposing several pristine proton beams with different proton energies                    (and hence different proton beam ranges), a Spread­Out Bragg Peak (SOBP) can be shaped                            to precisely match tumor tissue. The total energy deposited in a patient (integral dose) for                              3     
  • 17.    a given target dose is always lower for proton treatments when compared to photon                            treatment techniques ​[21]​. The high level of control afforded by proton beams is preferred                            for treating tumors located close to critical organs, such as the spinal cord, eye, and brain,                                where even a small dose can cause very serious consequences.  Range uncertainty  One major factor limiting the clinical effectiveness of proton therapy is range                        uncertainty. With current technology, there is no method to verify that the proton dose was                              delivered as planned. Thus, clinicians are completely dependent on treatment planning                      calculations for dose delivery. In principle, one should be able to use a proton beam pointed                                directly at a critical structure, with the beam energy tuned to stop just short of the critical                                  organ while delivering maximum dose to the tumor volume. However, range uncertainties                        require the addition of substantial safety margins to over­irradiate the tumor.    Paganetti ​[21] gives a thorough review of the types and causes of proton range                            uncertainties. In summary, uncertainties result from four main categories: 1) patient                      motion; 2) variations in setup and anatomy; 3) approximations used in dose calculations;                          and 4) biological effects. With no way to verify dose delivery, clinicians add safety margins                              to ensure total irradiation of the target volume. For example, at the Massachusetts General                            Hospital (MGH), treatment plans add 3.5% of the range plus an additional 1 mm ​[21]​.                              Although this guarantees the coverage of the distal aspect of the target volume, it also risks                                overdosing the normal tissue behind the target volume. The potential for overdosing                        changes the irradiation strategies used for certain cancers, potentially limiting the                      usefulness of the treatment.    The most prevalent example of the limitations of current methods is the treatment                          of prostate cancer. Naturally, the most effective approach irradiates the tumor from the                          anterior (Figure 2a) so that the sharp distal falloff of the proton beam can be used to treat                                    the target volume while sparing the rectum. This would require, however, a precise control                            of the beam range in the patient with millimeter accuracy, which is not currently possible.                              As a result, anterior fields have never been used, despite the fact that such fields can utilize                                  sharp distal penumbra (~4 mm for 50–95% falloff) ​[22]​. Instead, only lateral fields are                            used (Figure 2b), relying solely on the much broader lateral beam penumbra (>10 mm for                              50–95% falloff) and delivering a larger integral dose to the patient ​[22]​.    As a result of these additional margins, tissues distal to the target volume receive a                              substantial dose, and the dosimetric benefits of proton therapy are lost. Subsequently,                        clinical outcomes of some common treatment sites, such as the prostate, are essentially                          equivalent between proton and X­ray modalities ​[23]​. With no clear clinical benefit in cases                            such as the prostate, questions arise as to whether proton therapy is worth the cost                              (approximately double that of treatment with photons).        4     
  • 18.      Figure 2​. Comparison of treatment planning dose distributions for proton therapy treatment of                          prostate cancer with (a) one single anterior field and (b) two parallel­opposed lateral fields. Structures                              of bladder, bladder wall, prostate, anterior rectal wall, rectum, and femoral heads are shown by cyan                                lines. Adapted from ​[24]​.      As a note of additional concern, the sharp distal dose falloff of the Bragg curve also                                makes the dose distribution extremely sensitive to uncertainties in treatment planning and                        patient setup. If planning calculations are off by 1 cm, the location of the distal falloff will                                  change by 1 cm, causing either an undershoot, which is to under­irradiate the distal portion                              of the target volume, or an overshoot, where the full dose is delivered to normal tissue                                behind the target volume.  Motivation and challenges  Debates amongst medical physicists regarding the cost and efficacy of proton                      therapy start with consideration for the limitations of currently available technologies                      [25,26]​. As described above, the underlying physics of proton therapy offer clear potential                          benefits in the radiotherapy treatment of cancer. Because of range uncertainty and the lack                            of method to verify range at delivery, the benefits of proton therapy suggested by the                              physical principles of charged particle interactions cannot be realized. When a reliable                        method of range verification is incorporated into the cancer treatment process, the full                          potential of proton therapy will be available for utilization by clinicians.    The work described herein presents a system for imaging prompt gamma rays                        emitted during proton therapy as a method of range verification. This innovative design is                            the first collimated imaging system to implement two­dimensional (2­D) imaging for                      prompt gamma proton beam range verification. This work is timely because as proton                          treatment facilities continue to proliferate, accurate and effective range verification is                      essential to ensuring the efficacy and optimization of the treatment method.  5     
  • 19.      Due to treatment modalities currently available clinically, this work is focused                      mainly on proton therapy. However, there is similarly much interest in other heavy ion                            beams for cancer treatment, such as carbon ​[27–32]​. With physical and biological                        differences that offer potential benefits over protons, carbon ions will likely find their way                            into the clinical mainstream in the near future ​[33,34]​. However, the same challenges                          regarding range verification exist with heavier ions. The method of range verification                        proposed in this work could be directly applied to the case of carbon ions or any other                                  heavy ion treatment. As with proton therapy, a reliable range verification technique would                          have a positive impact on the clinical efficacy of any heavy ion therapy. The importance and                                potential impact of this research cannot be overstated.    In this work, Chapter 2 discusses the technical challenges associated with proton                        range verification, such as high photon energy, high background environment, and a lack of                            imaging devices suitable for proton therapy application. Other methods of range                      verification under investigation are also discussed. Chapter 3 describes the benefits and                        challenges in using simulation tools to aid with research efforts related to prompt gamma                            imaging. Chapters 4–7 characterize the design and performance of an imaging system                        developed to address these challenges, while Chapter 8 summarizes the implications of this                          imaging system. The ultimate goal of this work is to determine proton beam range to an                                accuracy of a few millimeters, which has the potential to make a significant impact on                              proton therapy and the entire radiation oncology field.        6     
  • 20.    Chapter 2—Techniques and Challenges of Range  Verification    A reliable method of range verification would improve the clinical capability of                        proton therapy by allowing for more precise targeting during treatments and increase our                          understanding of range uncertainties. Because all primary particles are stopped in­patient,                      research has focused on using secondary emissions to verify proton range. Of the several                            techniques currently being studied for in vivo range monitoring during proton therapy, the                          two most developed methods involve the imaging of secondary coincident positron                      annihilation photons and characteristic prompt gamma (PG) rays emitted as a result of                         1 proton beam irradiation of the patient ​[35]​.  Positron emission tomography  Positron emission tomography (PET) imaging has been widely studied for use in                        proton range verification. This clinically tested technique involves moving the patient to a                          PET scanner immediately after proton irradiation ​[36]​. The method has shown some ability                          to accurately measure the in vivo proton beam range. However, its implementation has                          been limited due to scanner resolution, positron energy, and biological washout ​[37]​.                        Additionally, because proton beams lose energy mainly via electromagnetic interactions,                    the activation image from positron emitters generated by nuclear reactions is not directly                          correlated to the dose distribution. The established method uses a Monte Carlo calculated                          distribution of the positron emitters and compares this predicted image with a measured                          image. The accuracy of the Monte Carlo calculation depends on the underlying cross section                            data ​[21]​.    Efforts are underway to develop in­beam PET which would add an imaging system                          to the proton gantry ​[38–40]​. This technique could eliminate problems associated with                        biological washout; however, there are additional problems created with the geometry of a                          conventional PET scanner, which, due to its large size limits its installation in the proton                              treatment room, as well as a poor signal­to­noise ratio experienced with attempts of PET                            verification during beam delivery ​[41]​. The positron emitting nuclides most produced                      during proton therapy (​15​ O, ​11​ C, ​30​ P, and ​38​ K) have radioactive decay half­lives between                          2–20 mintues, providing time for biological washout effects over the few minutes of                          required image acquisition time. Alternatively, the positron endpoint energy of some                      short­lived positron emitters of interest for PET imaging during the proton treatment,                        produced in high concentrations and of clinical interest, is too high to offer any benefit in                                1 Recent literature uses the terms “prompt gamma ray” and “prompt gamma” interchangeably. For                            consistency and clarity, the acronym ​PG is used throughout the text in place of both terms. In either case, PG                                        refers to the near­instantaneous photon emitted following the de­excitation of a nucleus. ​PGI refers to the                                imaging of prompt gamma rays.  7     
  • 21.    proton range verification applications (such as the 16.3 MeV positron endpoint for ​12​ N,                          with a range of ~2 cm) ​[42]​.  PG emission  The drawbacks of PET have fueled interest in the development of prompt gamma                          imaging (PGI) during proton therapy. PGs are emitted instantaneously (on the order of a                            few tens of femtoseconds) following an inelastic collision of a proton with a target nucleus.                              Like PET, PGs are produced by nuclear reactions. However, in PGI the measured photons                            are emitted directly from the nucleus as opposed to PET photons emitted some distance                            from their origin following an electron­positron annihilation. Thus, the effects of biological                        washout and decay time seen with the PET method are nonexistent in PGI. In a direct                                comparison with PET, PG was found to have an approximately 10­fold larger production                          rate and to have a distribution physically much closer to the Bragg peak ​[43]​. Thus, the                                spatial distribution (Figure 3) of the induced activity correlates better with absorbed dose                          for PG as compared to PET ​[44]​. The PG production is highest at low proton energies, as                                  shown in Figure 4. This results in more PG emissions in the vicinity of the Bragg peak.        Figure 3. Simulated profiles for dose (black), PG (blue), and PET (red) integrated over the entire beam                                  for an abdomen irradiated with pencil beams. All profiles are normalized to unity for easier                              comparison. The x­axis represents the position along the beam direction relative to the isocenter.                            Adapted from ​[43]​.      Unlike PET, which uses a well­established imaging modality for photon energy of                        511 keV, PGs pose an imaging challenge due to their high energy (2–10 MeV). The                              advantage of focusing detection efforts on high­energy gamma rays is the ability to filter                            out lower energy photons that are also produced by the proton beam, such as the 511 keV                                  8     
  • 22.    photons, which are not produced as close to the Bragg peak as PGs. Classical gamma                              cameras used in nuclear medicine are not adapted for detection of high­energy gammas in                            the presence of an important neutron background, so dedicated cameras are needed.                        Imaging high­energy photons requires specialized materials and methods.      Table 1.​ Reaction channels that produce the most prominent gamma energies ​[45]​.    Target  Reaction  E​γ​  ​ (keV)  Half­life  C  12​ C(p,p’)​12​ C  4443  45 fs    13​ C(p,d)​12​ C  4443  45 fs    12​ C(p,n)​12​ N  4443  11 ms (β​+​  ➔ ​12​ C​*​ )    13​ C(p,2p)​12​ N  4443  11 ms (β​+​  ➔ ​12​ C​*​ )  N  14​ N(p,p’)​14​ N  1635  4.8 fs    14​ N(p,n)​14​ O  1635  71 s (β​+​  ➔​14​ N​*​ )    14​ N(p,p’)​14​ N  2313  68 fs    14​ N(p,n)​14​ O  2313  71 s (β​+​  ➔ ​14​ N​*​ )  O  16​ O(p,pα)​12​ C  4443  45 fs    16​ O(p,nα)​12​ N  4443  11 ms (β​+​  ➔ ​12​ C​*​ )    16​ O(p,d)​15​ O  5241  122 s (β​+​  ➔ ​15​ N​*​ )    16​ O(p,2p)​15​ N  5270  17 fs    16​ O(p,p’)​16​ O  6129  18 ps      A sampling of characteristic gamma rays produced during proton therapy is listed in                          Table 1. It is important to note the many different proton­induced reaction channels                          available to produce the same discrete­energy photons. This complexity adds to the                        challenge of cross section measurements and accurate simulations (discussed in Chapter                      3). In addition to PGs emitted immediately following a nuclear interaction, there are                          contributions from beta­delayed gamma rays. Depending on the half­life of the beta decay,                          the gamma rays emitted as a result of these reaction pathways may not be distinguishable                              from the PGs.          9     
  • 23.    (a) (b)    Figure 4. Production cross sections for PGs resulting from inelastic collisions of protons with ​16​ O (a)                                and ​12​ C (b). Cross sections are largest at low proton energy. Data from: ​[45–47]​.  Detection and imaging of PGs  Detector materials  Particular care must be taken in choosing an appropriate material for detecting PGs.                          Each method of detection has characteristic energy, spatial, and time resolution                      requirements. For example, collimated systems require excellent spatial resolution and                    counting rate capability. A Compton camera design requires good spatial, energy, and time                          resolution to perform the necessary reconstruction calculations. If a detection design relies                        on timing, such as for background subtraction, excellent timing resolution will be required.    Scintillation detectors are optimal materials for PGI due to their relatively high                        stopping power and good time and energy resolutions. Roemer ​et al. reported a detailed set                              of measurements and comparisons with several scintillator materials ​[48]​. Their analysis of                        energy and time resolution found sufficient performance amongst several scintillation                    materials, with CeBr​3 performing best. At clinically relevant gamma energies and count                        rates, they found CeBr​3 to have an energy resolution of 2.2% and a time resolution of 190                                  ps full­width half­maximum (FWHM).    In a similar effort, Hueso­González ​et al​. performed comparison measurements with                      block detectors of LSO (Lu​2​SiO​5​:Ce) and BGO (Bi​4​Ge​3​O​12​), both commonly found on PET                          detection systems ​[49]​. They focused on optimization for a Compton camera design, which                          requires excellent spatial, energy, and time resolution. LSO had the better performance in                          their tests, but BGO may be a suitable alternative based on its lower cost.    10     
  • 24.    Detection techniques  A sampling of the variety of PGI techniques currently under development is detailed                          below. Each method makes a unique attempt at solving the same problem. Namely, to                            achieve millimeter­range accuracy in determining the location of the Bragg peak using PGs.  Compton cameras  Multiple efforts are underway to utilize Compton imaging for PG detection ​[50–54]​.                        This method uses the energy and angle dependence of a Compton scatter interaction to                            produce an image, in a technique first described by Everett ​et al. ​[55]​. This detection                              method requires a complex setup of two or more “stages.” If a photon undergoes a                              Compton scatter in the first detector and subsequently deposits the remainder of its energy                            in a second detector, a cone of possible incident angle can be back­projected into the image                                space. The collection of such cones over time allows a reconstruction algorithm to                          determine the source origin.    A recent study of such a system was reported by Polf ​et al. ​[56]​. They produced                                some promising results, demonstrating that their Compton camera system can detect                      relative shifts of 3 mm and 5 mm in the Bragg peak location. However, the beam current,                                  and thus the PG production rate, used in their measurements was much lower than that                              used at most proton therapy treatment facilities. As a conservative estimate, let us assume                            a beam current of 1.0 nA (typical beam currents are 1–3 nA ​[57]​) delivered to the patient                                  (or 6.2 x 10​9 protons/s). If a representative treatment field of a daily treatment is 125 cGy                                  (2.25 x 10​8 protons), the total beam­on time for the dose is 36 ms. Assuming a PG                                  production rate of 0.1 per proton ​[58]​, there will be 6.2 x 10​8 PGs emitted per second (or                                    620 MHz). A 100 cm​2 detector placed 10 cm away would thus experience a count rate of                                  ~50 MHz. The highest achievable double­scatter count rate for a Compton camera reported                          by McCleskey ​et al​. ​[59] was 5 kHz, well below that needed for clinical delivery rates.                                Because the Compton camera is severely limited by count rate (too many counts results in                              false coincidences), the authors concede their system is not yet clinically viable. Further                          work will continue on the Compton cameras, but until issues of timing and detection                            efficiency are solved, the clinical application of such systems will be hindered.  Electron­track Compton cameras  A variant of the traditional Compton camera, electron­track Compton cameras                    utilize a gas chamber detector to measure the energy and direction of the scattered                            electron instead of measuring the energy and angle of the Compton­scattered photon. This                          imaging method has also been demonstrated with solid­state detectors ​[60]​. Combining the                        energy and direction of the electron with the position and energy deposited by the incident                              gamma ray provides sufficient information to determine a cone of incident angles from the                            source. Thus, the electron­track Compton camera only requires one interaction in the                        detector to determine the origin direction of a photon.     11     
  • 25.    Such a system has been described and analyzed via simulation by Kurosawa ​et al.                            [61]​. Kurosawa ​et al​. limited proton beam current to 2.5 pA due to detector dead time                                considerations, as opposed to the normal 1 nA beam current used in clinical treatments.                            While the concept is interesting, there remains much work to gain feasibility, particularly                          due to the limited angular resolution and limited count rate capabilities of current                          technology.  Gamma­electron vortex imaging  A method proposed by Kim ​et al​. involves tracing Compton electrons, similar to                          electron tracking Compton imaging ​[62]​. An “electron converter stage” converts the                      high­energy gamma rays to electrons via Compton scatter. The system then determines the                          direction of the electrons using a pair of hodoscopes. Based on the assumption that                            electrons scatter in the forward direction, the lines of travel of the electrons are                            back­projected to the image space. Like other non­collimated systems, the potential benefit                        of this method is higher detection efficiency. However, the authors do not provide specifics                            on spatial resolution, although they state that a prototype system is under development for                            further study.   PG timing  Protons travel very quickly through tissues, but they still have a finite transit time:                            approximately 1–2 ns for a proton with a 5–20 cm range. Because the transit time increases                                with range, a time­resolved PG measurement could indicate the range of proton travel. This                            PG timing approach was proposed and explored by Golnik ​et al. ​[58]​. Based on simulations                              and initial measurements, the authors suggest that proton range could be determined                        within 2 mm using a single scintillation crystal. More work remains to demonstrate the                            clinical feasibility of the PG timing method, but the initial report is very promising ​[57]​.  Collimated cameras  The most straightforward PGI method is a collimated imaging system. Efforts on                        collimated systems for PGI have focused on variations of a single­slit collimator design.                          This method provides a one­dimensional (1­D) profile of PG emission, which would be                          most suitable for identifying the PG falloff location by placing the imaging system                          perpendicular to the beam axis. The single­slit is a very simple and straightforward concept                            to implement, and often uses a knife­edge design to increase the field of view and allow for                                  image magnification. A prototype knife­edge slit camera tested by Smeets ​et al. ​[63]                          demonstrated 1–2 mm accuracy in determining the location of a Bragg peak at near                            clinically relevant beam currents. Similarly, Bom ​et al. ​[64] reported better than 1 mm                            accuracy in determining the PG distribution using a single­slit collimator. Realizing the                        clinical potential and relative simplicity of the single­slit collimator design, a prototype PGI                          system has been developed for clinical application ​[65]​.    Options for alternative collimator designs, such as multi­slit, are limited due to the                          properties of PGs. Collimators must be much thicker than those used in diagnostic imaging                            12     
  • 26.    due to the high photon energy ​[66]​. Therefore, multi­slit collimators, offering a larger field                            of view or potentially greater detection efficiency, are a challenge to design ​[67]​.    A high level of neutron background radiation can also limit the effectiveness of a                            collimated camera (as well as Compton cameras). Some attempts have successfully                      demonstrated adjustments in the technique to overcome the high background limitation.                      Using the timing characteristics of the cyclotron­produced proton beam, the PG signature                        can be separated from background radiation with sufficient time resolution, thus                      improving PGI performance ​[44]​. Time­of­flight (TOF) measurements have also been used                      to suppress neutron background ​[68–70]​.  Conclusion  Due to production rates and proximity to the actual Bragg peak, PGI offers more                            potential than PET for successful implementation as a proton range verification method.                        However, there has yet to be a study published with results of PGI during patient                              irradiations, and thus the clinical effectiveness of PGI remains to be shown. Non­collimated                          systems, such as the Compton camera, are plagued by the high count rates experienced                            during clinical treatment conditions. Additionally, the detection efficiency of systems                    requiring multiple detector interactions for a single event is too low for clinical                          consideration. Until solved, the detection efficiency and count rate problems make these                        methods infeasible.     The closest system to clinical use is the relatively simple knife­edge slit collimated                          system. Multiple experiments have shown the potential to determine PG distributions with                        1–2 mm accuracy. As prototype systems are developed and optimized for clinical use, we                            may see promising results and approach clinical implementation. However, single­slit                    systems are limited to providing 1­D information about the PG distribution.    This work proposes a new multi­slit collimated imaging design to improve upon the                          single­slit concept. As the PGI field moves forward in determining the dose distribution in a                              patient, a 2­D image of PGs will be needed. A Compton camera system would be capable of                                  providing a 2­D distribution of PGs, however the count rate limitation is not likely to be                                overcome. The system introduced in this study actualizes the benefits offered by PG                          detection, while addressing and overcoming several key challenges associated with PGI.    Finally, unless a straightforward relationship between PG emission yield and dose                      deposition can be established, PGI (similar to PET) will have to rely on the comparison of                                the measured PG signal with a previously calculated or modeled expectation. Thus, clinical                          applicability will require extensive experimental validation of PG yields independent of the                        proton beam properties and the irradiated tissue type ​[22]​. Given the range uncertainties                          associated with proton therapy and the high costs of building the many new proton therapy                              centers across the globe, it is likely that research and development of range verification                            methods will continue until an effective solution is found.    13     
  • 27.    Chapter 3—A Monte Carlo Study of Prompt  Gamma Production  Background  As computing power has increased per unit cost in recent decades, Monte Carlo                          simulation methods have become more prevalent in scientific work. The field of medical                          physics, in particular, has extensively used Monte Carlo simulations to model radiation                        transport and to study radiation treatment modalities ​[71]​. Monte Carlo simulations offer                        an opportunity to study complex or experimentally difficult processes, with a level of detail                            and reproducibility not available to the experimentalist. For the PGI application, Monte                        Carlo methods would be an invaluable tool for the purposes of    1) optimizing an imaging system for PG detection;  2) identifying and mapping the prompt gamma ray emission profile,                  particularly in the inhomogeneous environment of a patient; and  3) providing the means to convert a detected PG profile to a dose distribution.    While proton therapy and proton range in tissue has been studied and validated in                            Monte Carlo systems, the production of PGs has only recently become a topic of interest                              and simulation due to range uncertainty problems in proton therapy. For dose calculations                          of proton therapy, the electromagnetic interactions of protons are well known and can be                            reliably simulated ​[72]​. The simulation of PG emission, however, relies on complex reaction                          models that were initially developed for high­energy physics applications. While Monte                      Carlo methods have been adapted to medical applications successfully, there are still                        limitations on the modeling of nuclear excitation and de­excitation. Initial validation                      studies of PG emission yields have not shown a consensus between simulated and                          experimental results ​[73]​. Furthermore, different Monte Carlo codes produce different                    results, as demonstrated in the case of production of positron emitters during proton                          therapy ​[74]​. And unfortunately, the production of PGs depends on a far greater number of                              reaction channels than for positron emitters due to the complex nuclear physics associated                          with the excitation and de­excitation of the nuclei.    As described by Verburg ​et al​. ​[75]​, the Monte Carlo nuclear reactions producing                          PGs are modeled in three stages:    1) Direct reactions: protons interacting directly with one or two                  nucleons of the target.  2) Pre­equilibrium: protons interact with parts of the nucleus before the                    target has reached equilibrium.  3) Compound reactions: energy of the proton is shared statistically                  among target nucleons.   14     
  • 28.       In the proton energy range of concern for therapeutic use (<200 MeV), all three                            reaction stages are relevant and nuclear excitation with subsequent PG emission can occur                          in any stage ​[75]​. The complex nature of these interactions, along with the multiple                            pathways available (see Table 1), makes it difficult for a theoretical model to accurately                            describe the results. This also complicates simulation of these processes. Considering that                        the models used in Monte Carlo software were developed without consideration of the PG                            productions at lower proton energies, it is understandable that Monte Carlo methods might                          not generate accurate simulations.    At the clinical proton energy range, Geant4 uses the Axen­Wellisch model ​[76] to                          calculate total reaction cross sections. This model uses a general formula for the range of                              proton energies from 6.8 MeV to 10 GeV. The model compares relatively well with                            experimental data across this large energy range; however, the data set they use ​[77] is                              very sparse at the lower energies. For example, nitrogen cross section data only goes down                              to 23 MeV. A more comprehensive set of experimental data for lower proton energies                            would allow us to validate this model or make adjustments to it.    The work below adds to evidence that current Monte Carlo systems do not                          adequately simulate the PG production during proton therapy, particularly to the level of                          accuracy demanded when designing and validating systems for clinical use on patients.                        First is a study and comparison of the simulated yield of PGs. Next, the angular distribution                                of PGs is simulated and compared with experimental expectations.  PG production cross sections  Methods   To analyze the production rates of PGs, simulations were run using TOPAS, ​TO​ol for                            PA​rticle ​S​imulation ​[78]​, a user­friendly platform that interfaces with and runs the Geant4                          Monte Carlo particle transport package ​[79]​. Beams of varying proton energies were                        directed incident on thin targets of carbon and nitrogen. At each proton energy level, 10​11                               protons were simulated, with resulting photons scored by energy. The resulting PG                        emissions were scored and used to calculate a gamma production cross section for each                            gamma energy level according to Equation 3:   ,Σ = R I∙N   (3)  where:  Σ = cross section (cm​2​ )  R = total number of gammas produced (specific gamma energy for reaction of interest)  I = total number of protons incident on the target  N = number of target nuclei presented to the beam per unit area (cm​­2​ ).  15     
  • 29.    Results  Figures 5–7 show gamma production cross sections for carbon and nitrogen targets                        for both simulations and experimental data . For the 1.63 MeV gamma produced in                         2 nitrogen, the simulated cross section peaks around 20 mb as opposed to the experimental                            value which reaches well over 80 mb. The 2.31 MeV gamma from nitrogen shows a similar                                peak in simulated and experimental results; however, the simulated cross section results                        are shifted to much higher proton energies. The results of the 4.4 MeV gamma in carbon                                also show much lower production cross sections in the simulated results than the                          experimental measurements.      Figure 5. Comparison of simulation results with experimental data for 1.63 MeV gamma production                            during proton irradiation of thin nitrogen target ​[45,47,80]​.      These results demonstrate that the Geant4 Monte Carlo simulation package, in this                        specific TOPAS configuration, does not adequately simulate PG production. Furthermore,                    the lack of discernable pattern across each respective photon energy precludes the                        identification of a simple correction or scaling factor that could make the package suitable                            for PG simulation. Simulations generate unreliable total gamma counts because of                      differences in total reaction production cross sections; this makes them unreliable to                        produce and study PG emissions. Additionally, as seen with Figure 6, various proton                          energies produce PGs in different locations of the simulated target than would be seen                            experimentally. Thus, there are two main components of error: 1) the absolute cross                          section and magnitude of PG production, and 2) the relative shape of the yield as a function                                  2 Because of timing resolution limitations in experimental measurements, the number of photons                          counted includes components from PG reactions as well as β­delayed reaction channels, as listed in Table 1.                                  These reaction channels were also included in the simulation results.  16     
  • 30.    of energy (i.e., depth). Until a Monte Carlo system can more accurately model the                            production of PGs in tissue­related materials, simulation data will not be a reliable tool in                              assessing and characterizing PG production.      Figure 6. Comparison of simulation results with experimental data for 2.31 MeV gamma production                            during proton irradiation of thin nitrogen target ​[45,47,80]​.        Figure 7. Comparison of simulation results with experimental data for 4.44 MeV gamma production                            during proton irradiation of thin carbon target ​[45,47]​.    17     
  • 31.    Angular distribution of PGs  Depending on the quantum properties of the excited nuclear level, PG emission is                          not generally isotropic ​[75]​. The angular distribution of gamma rays with respect to the                            beam direction is given by the Legendre polynomial expansion:  ,(θ) P (cosθ), (l even)W = ∑ l=lmax l=0 al l     (4)  where ​l​max​ is the smaller of the following two quantities:    1. twice the spin of the decaying state and  2. twice the multipolarity of the gamma ray ​[81,82]​.    Measured gamma ray angular distributions produced by protons of up to 13 MeV                          are shown in Figure 8 ​[47]​. With the exception of the 6.13 MeV gamma rays from ​16​ O, the                                    gamma rays considered in this example have a multipolarity of 2 or less, so that ​l​max is 4 or                                      less, and there are at most three terms (​l = 0, 2, 4) in the above expansion. In a similar study                                          using 33 MeV protons, Lesko ​et al​. ​[45] found the angular distribution of PGs to be more                                  isotropic than with the 8–13 MeV example. This evidence, provided in Figure 9 for                            comparison, suggests an angular distribution dependence on incident proton energy. At                      higher proton energies, more nuclear reaction channels are available and there are                        contributions from a larger variety of excitation pathways. This explains the different                        angular profile seen at higher proton energies, and re­emphasizes the complexity of the PG                            modeling challenge.    The angular distribution of PGs is important to consider when benchmarking Monte                        Carlo simulation results to experimental measurements. As described above, and                    corroborated by Verburg ​et al. ​[75] with Geant4 and MCNP ​[83] codes, Monte Carlo                            simulations do not produce accurate PG yields within the acceptable range of experimental                          data. However, these analyses included only angle­integrated cross sections, which is a                        noted limitation of the study because PG emission is not usually isotropic. Robert ​et al​. ​[84]                                compared two Monte Carlo code results for the angular distribution of secondary particles,                          showing that both Geant4 and FLUKA ​[85] codes provide essentially isotropic gamma                        emissions. The purpose of their study was to compare the two Monte Carlo packages, but                              they do not include any discussion as to the absolute veracity of their simulation results.    The angular distribution of PG emission is equally important as the total PG                          production for consideration in actual measurements. Polf ​et al. ​[86] described some initial                          results using PG detection to quantify the oxygen content in proton­irradiated tissue. Their                          proposed application is quite promising, but without a correction for the angular                        distribution of gammas (which in some cases may vary by a factor of ~2), the precision of                                  tissue composition determination is severely limited. If a quantitative measure of PGs is to                            be performed, an accounting of the angular distribution will be required.    18     
  • 32.        Figure 8​. Examples of angular distributions of gamma rays. The solid curves are fit with an expansion                                  in Legendre polynomials of even order through zero for the ​14​ N case, through 6 for the ​16​ O case, and                                      through 4 for the other cases. The gamma ray energies (top to bottom) are 4.44, 2.31, 6.13, 1.63, 1.37,                                      1.78, and 0.847 MeV. Adapted from ​[47]​.    19     
  • 33.      Figure 9. Angular distributions for five separate gamma ray transitions resulting from bombardment                          of Mylar, Mg, Si, and Fe targets with 33 MeV protons. The solid lines are the results of least squares fit                                          of Legendre polynomials to the data. Adapted from ​[45]​.    Methods  To assess the ability of the Geant4 Monte Carlo package to simulate the PG angular                              distribution, a simple simulation was performed with TOPAS using the default physics                        models. A spherical scorer (i.e., perfect detector) was placed around a 1 cm­thick graphite                            target. A beam of protons (10​9 total) at 13 MeV was directed incident on the target, and the                                    resulting 4.4 MeV gammas were scored by the detecting sphere and binned in 3°                            increments with respect to the beam direction. The proton energy and target material were                            chosen to provide a comparison to the experimental data shown in Figure 6.    Given that many researchers falsely assume isotropic PG emission, with some                      simulation results reinforcing that misconception ​[84]​, an additional analysis was                    performed to compare PG distributions as would be detected for isotropic emission with                          the experimentally demonstrated angular distributions of emissions.    20     
  • 34.    Using the TOPAS simulation, a poly(methyl methacrylate) (PMMA; (C​5​O​2​H​8​)​n​; and ⍴                      = 1.18 g/cm​3​ ) target was irradiated with 160 MeV protons to simulate a typical clinical                              proton beam energy setting. The tissue­equivalent PMMA target was binned into 0.2                        mm­thick slices, and the average proton energy in each slice was determined by simulation.                            The slice thickness was chosen to avoid significant variation of average proton energy                          across any individual target slice. Using data presented in Figure 2 ​[45–47]​, cross sections                            for production of the 6.129 MeV gamma line from ​16​ O were calculated via interpolation.                            Then, an analytical calculation was performed assuming a detector located at the Bragg                          peak location (15.2 cm) and 90° relative to the beam direction. First, the PG emissions were                                assumed isotropic and the proton energy in a slice was used in combination with the cross                                section at that energy to determine the number of 6.129 MeV gammas produced. Next,                            using an interpolation of the data presented in Figures 6 and 7, the angular distribution of                                gamma emission was considered along with the angular position of the detector with                          respect to the given target slice. The results approximate the difference to be expected                            between an actual measurement and a measurement with an assumption of isotropic                        gamma emission.  Results  Figure 10 shows the angular distribution of 4.4 MeV gammas resulting from the                          simulation of 13 MeV protons incident on a graphite target. The gamma distribution                          appears symmetric about a peak at 90°. This is quite different than the experimental results                              shown in Figure 8 for the same proton energy and target nucleus. The experimentally                            measured gamma distribution shows 90° at the lowest point in the PG emission. In this                              case, the simulation model results are less accurate than if the model assumed an isotropic                              distribution.    In Figure 11, the blue line represents the 6.129 MeV gamma production in a PMMA                              target, assuming that the yield is isotropic. This count total would be measured at the                              detector if the gamma production was actually isotropic. The orange line is a more realistic                              estimate of the gamma count measured at the detector given the angle of detection. As can                                be inferred from Figure 8, at lower proton energies there is a higher yield of 6.13 MeV                                  gamma as the detection angle moves from 90°. With a detector setup 90° from the Bragg                                peak location, the resulting PG distribution measured would be broader than for the                          isotropic case—a lower peak near the end of proton range, with some area of higher                              magnitude proximal to the peak.    21     
  • 35.      Figure 10. Angular distribution of 4.4 MeV PGs (Geant4). This is the result of 10​9 total protons incident                                    on a graphite target. 13 MeV p + ​12​ C.        Figure 11​. Comparison of theoretical isotropic PG distribution (blue) and expected gamma                        distribution as detected based on angular distribution (orange). Each gamma profile represents the                          6.129 MeV gamma production in a PMMA target as would be measured.    22     
  • 36.    Conclusion  These results indicate that current Monte Carlo methods do not accurately simulate                        the excitation and de­excitation of the nuclei of interest in PGI for proton therapy (C, N, and                                  O). The reaction cross sections, gamma emission yields, and the angular distribution of                          emission for the specific photon energies are not correctly modeled. This important finding                          identifies a limitation in the Monte Carlo simulation method—an important tool in studying                          the problem at hand. While simulations can still be a valuable asset in designing a system to                                  detect and image PGs, it is important to understand the limitations of the modeling                            packages. Until new models are developed and incorporated into a Monte Carlo simulation                          package, the simulation toolkit cannot be used to study the production and location of PG                              emissions during proton therapy. This hinders the necessary development of 1) PG distal                          falloff verification, and 2) PG­to­dose profile conversion.    While building a new model of PG emissions is outside the scope of this project, it is                                  noteworthy to identify this as an issue of importance to the topic. Future work to improve                                the simulation modeling of the phenomenon would have a substantial positive impact in                          the research community.              23     
  • 37.    Chapter 4—A Novel Multi­Slit Collimated  Imaging System  Introduction  As discussed in Chapter 2, the most promising method of range verification is PGI                            using a collimated imaging system. While there are multiple challenges and limitations to                          the use of collimated cameras due to the nature of PGs, the alternatives to collimated                              systems have not proven clinically viable ​[56]​.    Attempts to improve on the performance of single­slit collimators resulted in the                        development of a novel multi­slit collimator concept ​[87]​. The complete imaging system                        (Figure 12) consists of an array of cerium­doped lutetium oxyorthosilicate (LSO)                      position­sensitive scintillation detectors paired with a multi­knife­edge slit collimator,                  which has been designed, constructed, and characterized at Lawrence Berkeley National                      Laboratory (LBNL). Each component of the imaging system is described in the following                          sections.        Figure 12. CAD model of imaging system showing 4 x 4 grid of LSO detector modules, tungsten                                  collimator with individually cut pieces held together with 3­D printed plastic, and a proton beam                              incident on a plastic target.          24     
  • 38.    Collimator design  A multi­knife­edge slit collimator has been designed for PGI applications. The 20 x                          20 x 7.5 cm tungsten (Hevimet—90% W, 6% Ni, 4% Cu) collimator acts as a coded aperture                                  system ​[88–90] for high­energy gamma rays that casts a unique projection for each point in                              the imaging plane. An adaptation to the single­slit camera design similar to those used in                              previous studies ​[63,64]​, the multi­slit concept was developed to improve detection                      efficiency, provide a larger field of view, and offer higher resolution spatial information. To                            provide adequate collimation, the material and thickness of the collimator were set such                          that 95% of PG emissions would be attenuated if traveling through the thickest part of the                                collimator. Knife­edge slits provide a larger field of view than parallel slits and also allow                              the opportunity for magnification of the image. Multiple other collimator designs were                        considered (for example, a parallel slit design and a fan beam design shown in Figure 13);                                however, the minimum septa thickness required to provide proper collimation is limited by                          the penetration depth of gammas in the MeV range. Thus, collimation schemes used in                            diagnostic imaging require thicker septa, which precludes any significant advantage in                      efficiency or spatial resolution over single­slit designs.        Figure 13. ​Fan beam collimator design modeled in TOPAS. Tungsten slits of 1 mm thickness are                                arranged with 2 mm wide openings that get wider as the distance from the object increases. Individual                                  BGO detectors are placed at the top of the collimator. The design, originally proposed by Andy Haefner                                  [91]​, did not offer significant detection efficiency advantages over a single slit collimator.      25     
  • 39.    By placing slits at varying angles relative to the central vertical slit, we achieve an                              imaging system similar to the 2­D coded aperture systems. A linear discriminant analysis                          performed on simulated proton paths determined that such geometry has the ability to                          discriminate with 10­fold higher sensitivity between proton paths with distal ends                      separated by 1 mm, as compared with a single slit ​[87]​. After several iterations optimizing                              for image contrast, uniformity in imaging sensitivity, and instrument cross­correlation                    sidelobes, the design presented here was determined to be the best candidate for further                            investigation (see ​[87]​).     The collimator is composed of 22 individually cut blocks, arranged to provide a                          predetermined pattern of knife­edge slits with 2 mm aperture at the center (Figure 14).                            The tungsten blocks are held in place by a plastic frame. The collimator pieces are arranged                                to imaging sensitivity across the image space that is closer to uniform than an unoptimized                              system. Increasing the number of slits provides a greater geometric efficiency, while also                          allowing the aperture size to decrease; in our case, 2 mm versus the 3 mm or greater                                  proposed in other studies. Additionally, the angled slits offer the capability of the                          collimator to produce 2­D images. The image reconstruction technique and initial                      characterizations of the collimator are provided in Chapters 5–7.    Figure 14.​ Arrangement and dimensions of tungsten collimator pieces.  26