Hostname: page-component-848d4c4894-v5vhk Total loading time: 0 Render date: 2024-06-28T20:56:38.312Z Has data issue: false hasContentIssue false

Monte-Carlo techniques for radiotherapy applications II: equipment and source modelling, dose calculations and radiobiology

Published online by Cambridge University Press:  28 February 2023

Andrew L. Fielding*
Affiliation:
School of Chemistry and Physics, Queensland University of Technology (QUT), Brisbane, Australia
*
Author for correspondence: Andrew L. Fielding, School of Chemistry and Physics, Queensland University of Technology (QUT), Brisbane, Australia. E-mail: a.fielding@qut.eduau
Rights & Permissions [Opens in a new window]

Abstract

Introduction:

This is the second of two papers giving an overview of the use of Monte-Carlo techniques for radiotherapy applications.

Methods:

The first paper gave an introduction and introduced some of the codes that are available to the user wishing to model the different aspects of radiotherapy treatment. It also aims to serve as a useful companion to a curated collection of papers on Monte-Carlo that have been published in this journal.

Results and Conclusions:

This paper focuses on the application of Monte-Carlo to specific problems in radiotherapy. These include radiotherapy and imaging beam production, brachytherapy, phantom and patient dosimetry, detector modelling and track structure calculations for micro-dosimetry, nano-dosimetry and radiobiology.

Type
Literature Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press

Introduction

This is the second overview paper on the use of Monte-Carlo in radiotherapy. The first part gave an introduction to the Monte-Carlo technique and the different codes that are available for simulating radiotherapy treatments. In this second overview paper on the use of Monte-Carlo for radiotherapy, the main areas of application will be summarised. This includes modelling the production of beams of ionising radiation for radiotherapy and medical imaging, treatment verification, patient dosimetry and radiobiology.

Radiotherapy and imaging equipment modelling

A common application of the Monte-Carlo method is in the modelling of imaging and radiotherapy equipment such as X-ray tubes, linear accelerators and detectors. For comprehensive reviews of the use of Monte-Carlo techniques for radiotherapy, the reader is referred to the following literature. Reference Verhaegen and Seuntjens1Reference Verhaegen and Seco9 An overview of some of the published work in some of the different areas of radiotherapy now follows.

X-ray tube modelling

Monte-Carlo techniques have been used to model X-ray tubes in the context of kilo- and ortho-voltage radiotherapy as well as medical and industrial imaging. The modelling of the X-ray tube enables a deeper understanding of the Bremsstrahlung production process in the target as well as the influence of components such as windows, collimators and filters on the spectral and fluence properties of the X-ray beams that are generated. Reference Poludniowski and Evans10,Reference Poludniowski11 Figure 1 shows an example of a simple X-ray tube geometry implemented using the XTUBE component module of the BEAMnrc user code. Reference Rogers, Faddegon, Ding, Ma, We and Mackie12 Information from the manufacturer on the target material (s), thickness (es), angle, electron beam energy and radius and window details can be used to develop the model. Additional collimators, filters and applicators can also be included using other available component modules. X-ray tubes are used for a wide variety of applications, and Monte-Carlo simulation has been applied accordingly. These applications include X-ray tubes in clinical diagnostic CT systems. Reference Bazalova and Verhaegen13 As part of the Monte-Carlo collection of papers in this journal, Aghdam et al used the MCNP code to model the treatment head of an intraoperative electron radiotherapy accelerator in order to calculate the radiation contamination dose in the area around the device. Reference Aghdam, Siavashpour, Aghamiri, Mahdavi and Nafisi14 In another study, the same group investigated the effective SSD dependence on electron energy and applicator size. Reference Aghdam, Baghani, Mahdavi, Aghamiri and Akbari15 Al-Ghorabie used the BEAMnrc code to model the XStrahl 150 kV radiotherapy unit. Simulated X-ray spectra, first and second Half-Value Layers (HVLs) and dose profiles were compared to measured data. Measured and simulated spectra were found to agree to ∼1% and HVLs differed between 2·3 and 3·6%. Percentage depth doses agreed to within 2% and beam profiles in the range 3–5 % depending on applicator size. Reference Al-Ghorabie16 At this point, the reader is reminded that Monte-Carlo simulations are highly sensitive to the configuration parameters used for a particular simulation. For this reason, when seeking to reproduce simulations by others, it is important that the exact same settings are used. An important responsibility therefore falls on the authors of research involving Monte-Carlo simulation to include all relevant configuration parameters to enable reproducibility by others. Reference Sechopoulos, Rogers and Bazalova-Carter17

Figure 1. Simple X-ray tube model created using the XTUBE component module of the BEAMnrc user code. Grey indicates a vacuum. Blue indicates the tungsten target embedded on the red copper mounting. Axis dimensions are in centimetres.

Radiotherapy delivery equipment modelling

Monte-Carlo techniques have been widely used to model clinical electron and X-ray beams, patient dosimetry and imaging systems found in the radiotherapy treatment room. Padilla-Cabal et al compared the MCNPX and BEAMnrc Monte-Carlo models for an Elekta Precise SL-25 photon beam. The MC models were compared to measure dosimetry profile data in homogeneous water and homogeneous phantoms containing air, bone and lung equivalent materials. Reference Padilla-Cabal, Pérez-Liva, Lara, Alfonso and Lopez-Pino18 Toossi et al used the MCNPX code to simulate the 8, 12 and 14 MeV electron beams generated by a Siemens Primus linear accelerator. Simulated dosimetry profile data were compared to measured data for different field sizes including applicators as part of a commissioning process. Reference Bahreyni Toossi, Ghorbani, Akbari, Sabet and Mehrpouyan19 Monte-Carlo techniques can also be used to model more specialised radiotherapy equipment such as Tomotherapy Reference Zhao, Mackenzie, Kirkby and Fallone20,Reference Zhao, Mackenzie, Kirkby and Fallone21 and the Cobalt-60 based Gamma Knife. Reference Moskvin, DesRosiers, Papiez, Timmerman, Randall and DesRosiers22Reference Junios, Irhas and Novitrian24 Mahmoudi et al used the BEAMnrc code to model the beam profiles of the individual and 201 combined beams of the Gamma Knife stereotactic radiotherapy system. Simulated data were compared with measured film dosimetry in a plexiglass head phantom for four different collimator sizes, 4, 8, 14 and 18 mm diameter. Reference Mahmoudi, Shirazi, Geraily, Nia, Bakhshi and Maleki25 It is also worth noting that Monte-Carlo techniques have also been used widely for verifying treatment dosimetry through the use of Electronic Portal Imaging Detector (EPID)-based dosimetry. Reference Kairn, Cassidy, Sandford and Fielding26Reference Herwiningsih, Hanlon and Fielding29 The technique can offer an accurate calculation of the portal dose response of the detector for comparison to the measured portal dose.

Proton and Heavy Ion beam modelling

Monte-Carlo techniques have been extensively used to model clinical proton and heavy ion beams and their interactions and deposition of dose in patient geometries. Reference Rahman, Bruza, Lin, Gladstone, Pogue and Zhang30Reference Vedelago, Geser, Muñoz, Stabilini, Yukihara and Jäkel33 Zarifi et al used the GATE code to study the depth dose characteristics of mono-energetic proton pencil beams of energies 5–250 MeV in water and obtain the energy–range relationship. Further, the stopping powers of the proton pencil beams in a water phantom were compared to data from the NIST standard reference database. Reference Zarifi, Ahangari, Jia, Tajik-Mansoury, Najafzadeh and Firouzjaei34 Zafiri et al performed a study to compare the accuracy of the different physics lists (models) that are available in the GATE code when simulating monoenergetic therapeutic proton pencil beams with energies 5–250 MeV. Reference Zarifi, Ahangari, Jia and Tajik-Mansoury35 Chiang et al used TOPAS to model the treatment head of the Mevion HYPERSCAN pencil beam scanning system proton therapy system including the energy modulation system and Adaptive Aperture. Depth doses and in-air spot sizes were found to have good agreement with measured beam data (1 mm and 10 %, respectively). Full-width half maximums and lateral penumbra agreed to within 2 mm. Reference Chiang, Bunker, Jin, Ahmad and Chen36

Imaging system modelling

The GATE Monte-Carlo system, introduced in the companion part I review, has been extensively used to model CT and the emission imaging systems (nuclear medicine, PET and SPECT). The reader is directed to an excellent topical review by Sarrut et al that includes comprehensive lists of the commercially available emission tomography imaging systems that have been modelled and compared against experimental data. Reference Sarrut, Bała and Bardiès37 A number of authors have used Monte-Carlo techniques to model the EPID and cone beam CT imaging systems that are now found on contemporary linear accelerators. This work has included modelling of the X-ray tube and detector systems. A number of authors have modelled a linear accelerator clinical cone beam CT system for the purposes of evaluating patient dosimetry during dosimetry. Reference Downes, Jarvis, Radu, Kawrakow and Spezi38Reference Spezi, Downes, Radu and Jarvis40 Monte-Carlo has also been used to evaluate and correct for scatter contributions to both megavoltage and kV cone beam CT. Reference Spies, Ebert, Groh, Hesse and Bortfeld41Reference Swindell and Evans45 Monte-Carlo has been widely used for modelling and optimising MV portal imaging systems for both anatomical imaging and dosimetry. Flampouri et al used the EGSnrc codes to optimise a low atomic number (Z) X-ray Bremsstrahlung target and detector for radiotherapy MV imaging. Reference Flampouri, Evans, Verhaegen, Nahum, Spezi and Partridge46 Accurate modelling of the portal imaging detector response has been performed using Monte-Carlo techniques. The calculated portal dose image can then be compared to a measured image as a method of in-vivo treatment verification. Parent et al developed a Monte-Carlo model for the Elekta iView GT a-Silicon flat panel imager. Reference Parent, Fielding, Dance, Seco and Evans47,Reference Parent, Seco, Evans, Fielding and Dance48 A different technique based on calculation and measurement of radiological thickness was demonstrated by Kairn et al Reference Kairn, Cassidy, Sandford and Fielding26 A number of other authors performed similar studies for Varian liquid ionisation chamber and flat panel imaging detectors Reference Siebers, Kim, Ko, Keall and Mohan28,Reference Spezi and Lewis49

Brachytherapy

While the majority of this overview has focussed on external beam radiotherapy and associated imaging, it is worth noting that Monte-Carlo calculation techniques have been widely used to model the dosimetry of brachytherapy. Many of the Monte-Carlo codes have capabilities for defining the geometries and sources used in brachytherapy, including EGSnrc and its egs_brachy user code, Reference Chamberland, Taylor, Rogers and Thomson50 MCNP Reference Solberg, DeMarco and Chetty51,Reference Zaker, Zehtabian, Sina and Koontz52 and GEANT4· Reference Enger, Landry and D’Amours53Reference Landry, Reniers, Pignol, Beaulieu and Verhaegen55 For those without GEANT4 experience, brachytherapy simulation capabilities are also offered through the more user-friendly TOPAS and GATE toolkits. Reference Sarrut, Bardiès and Boussion56,Reference Berumen, Ma, Ramos-Méndez, Perl and Beaulieu57 As part of the Monte-Carlo collection in this journal, Dagli et al compared three different dose calculation algorithms for clinical HDR brachytherapy. The aim was to investigate the accuracy of the BrachyDose Monte-Carlo code in heterogeneous media through a comparison with the Eclipse TG-43 dose calculation tool and the Acuros BV model-based algorithm. Acuros BV and BrachyDose were found to have a good agreement, but significant dose differences were seen with the Eclipse TG-43 dose calculation. Reference Dagli, Yurt and Yegin58

Dosimetry and treatment verification

The first step in a Monte-Carlo simulation of a radiotherapy treatment involves modelling the production of the beam of ionising radiation as discussed in Sections 2·1, 2·2 and 2·3. The output from the simulation can then be used as an input to the simulation of the interaction of the beam(s) with a patient, phantom or detector. Computed tomography images of patients and phantoms in DICOM format can be used to accurately create a voxelised model of the geometry or specific anatomy, including the different segmented tissues and structures in a patient. Simulation of MV X-ray and electron beams used in radiotherapy requires the use of a CT-electron density calibration curve that provides the relationship between CT number and electron density for the segmented tissues in the voxelised patient model. As part of this process, it is common to simplify the assignment of electron densities so that the complex tissues existing in the human body are categorised into just a few basic types, e.g. soft tissue, lung, bone, and air.

As part of the Monte-Carlo collection of this journal, Zaman et al used the EGSnrc code to model enhanced dynamic wedge deliveries to heterogeneous slab phantoms containing lung and bone heterogeneities in water-equivalent material. The Monte-Carlo results were compared with calculations using the Anisotropic analytical algorithm (AAA) and the Acuros XB algorithms found in the Varian Eclipse treatment planning system. Reference Zaman, Kakakhel and Hussain59 Yabsantia et al. performed a comparison of 6 MV photon small field output factors measured following the TGS-483 code of practice with calculations performed using the EGSnrc Monte-Carlo code. Measured output factors for three different detectors, the IBA-CC01, Sun Nuclear EDGE and IBA-SFD, were corrected using the TG-483 formalism and compared with the Monte-Carlo simulations performed using the egs_chamber EGSnrc user code. Reference Yabsantia, Suriyapee, Phaisangittisakul, Oonsiri, Sanghangthum and Seuntjens60 Vassiliev et al used the BEAMnrc/DOSXYZnrc Monte-Carlo code to study the effect of motion on target dose coverage for lung stereotactic body radiation therapy patients treated with flattening filter-free beams. Fifteen patient plans were re-calculated using 4D-CT data with and without the flattening filter. The target coverage for treatment planning system calculations was compared with the Monte-Carlo for each of the patients. Reference Vassiliev, Peterson, Chang and Mohan61 Tanha et al investigated the accuracy of a collapsed cone convolution and ETAR algorithm through comparison with a BEAMnrc Monte-Carlo model of the Varian 2100 C/D 18V photon beam. Dosimetric comparisons were made for a pituitary gland treatment planned on a rando-phantom. Differences between the Monte-Carlo and collapsed cone and ETAR were found to be up to 6 and 10 %, respectively. Reference Tanha, Mahdavi and Geraily62 Nithiyanantham et al compared the XVMC Monte-Carlo dose calculation algorithm found in the commercial Monaco treatment planning system to measured doses for small field sizes (8 × 8 up to 40 × 40 mm) in soft tissue equivalent phantoms containing lung and bone heterogeneities. They observed deviations for the smallest field sizes in and around the soft-tissue heterogeneity interfaces. Reference Nithiyanantham, Mani and Raju63 Mamballikalam et al investigated the dosimetry of 6 MV flattening filter-free small fields measured with three different dosimeters, the IBA CC01 pinpoint chamber, the IBA stereotactic field diode and the PTW microDiamond. Measured doses were compared with calculated doses obtained using the PRIMO Monte-Carlo code. The work showed that for field sizes below 1 × 1 cm perturbation and volume averaging corrections should be applied. Reference Mamballikalam, Senthilkumar and Jayadevan64 Chow and Owrangi used the BEAMnrc code to investigate the mucosal dose in the oral and nasal cavity and its dependence on beam energy, beam angle, mucosal thickness and the backscatter from neighbouring bone for small 6 and 18 MV fields generated by a Varian 21 EX accelerator. Reference Chow and Owrangi65 Acar et al compared the accuracy of the Varian Eclipse electron Monte-Carlo dose calculation algorithm with the Gaussian pencil beam algorithm. The work focussed on the ability of both algorithms to calculate peripheral doses for 6 electron beam energies ranging from 6 to 22 MeV. Calculated peripheral doses were compared with measured (ion chamber and EBT film) for three field sizes, 6 × 6, 10 × 10 and 25 × 25 cm. Differences of up to 8·8 % between Gausssian pencil beam and measurements were reduced to less than 4·3 % for electron Monte-Carlo and measurement. Reference Acar, Caglar and Altinok66 Abdul Aziz et al used a BEAMnrc model of a 6 MV photon beam generated by a Siemens Primus linear accelerator to investigate the effect of artefacts due to a titanium hip prosthesis in CT data on the dose distribution in and around the prosthesis. Reference Abdul Aziz, Mohd Kamarulzaman, Mohd Termizi, Abdul Raof and Tajuddin67

Micro/Nano-dosimetry, radio-chemistry and radiobiology

The macroscopic modelling of radiation production and dose deposition Monte-Carlo can also be used to model radiation interactions on the micro and nano-length scale. For these applications, a track structure paradigm, rather than the condensed history technique, is used that models the individual low energy charged particle interactions at the much shorter length scales. GEANT4 has led the way with this through the development of the GEANT4-DNA toolkit, Reference Incerti, Baldacchino and Bernal68,Reference Incerti, Douglass, Penfold, Guatelli and Bezak69 an extension to the GEANT4 toolkit. GEANT4-DNA has been validated for the modelling of the low energy physics interactions in water as well as the production and subsequent interactions of the reactive chemical species following the radiolysis of water. Reference Bernal, Bordage and Brown70,Reference Peukert, Incerti and Kempson71 From a radiobiological perspective, the DNA damage produced by ionising radiation has also been modelled Reference Sakata, Belov and Bordage72 using GEANT4-DNA. Validation of these microscopic track structure simulation codes for different ionising radiations has been performed against laboratory experiments and in comparison with other independent Monte-Carlo codes. Reference Kyriakou, Sakata and Tran73 As has already been indicated, for those without C++ expertise, all these GEANT4-DNA capabilities for low energy physics, chemistry and radiobiology are accessible through the user-friendly TOPAS-nBio toolkit. Reference Schuemann, McNamara and Ramos-Méndez74Reference Zhu, McNamara and Ramos-Mendez77 Figure 2(a) shows a TOPAS simulation of the simulated tracks of 50 keV electrons traversing a spherical cell model containing mitochondria and gold nanoparticles. Similar models have been used to investigate the nano-scale and radiobiological interactions underpinning dose enhancement with nano-particles and more recently high dose-rate FLASH radiotherapy delivery, Reference Rudek, McNamara, Ramos-Méndez, Byrne, Kuncic and Schuemann78,Reference Ramos-Méndez, Domínguez-Kondo, Schuemann, McNamara, Moreno-Barbosa and Faddegon79 both using TOPAS-nBio. In the Monte-Carlo collection of this journal, Belamri et al used GEANT4 to investigate the potential dose enhancement effect of gold, silver or platinum nano-particles when irradiated with proton beams with energies typical of those used for brain treatments. Reference Belamri, Dib and Belbachir80

Figure 2. (a) Simple spherical cell model created using TOPAS-nBio. Red central sphere is the nucleus, green spheres are mitochondria and very small yellow spheres are gold nanoparticles. Red lines are the tracks of 50 keV electrons. (b) Free radical chemical species are generated along an electron track (shown in red) as it traverses water.

The TOPAS-nBio code also has the capability to model the chemistry following the physical processes. This includes the free radical chemical species production (e.g. H+ and OH− ions) plus the subsequent diffusion of these species through the cell geometry. Reference Ramos-Méndez, Perl, Schuemann, McNamara, Paganetti and Faddegon75 Figure 2(b) shows an example of the chemical species generated along an electron track as it traverses water. The motion of the chemical species can then be modelled as a function of time.

These radiobiological level simulations of the physics and chemistry of ionising radiation interactions are an extremely powerful tool that can play a crucial role in increasing our understanding of the radiobiology of ionising radiation at the cellular length scale.

Conclusions

Monte-Carlo is a powerful tool for simulating the transport of ionising radiation as it traverses matter. This feature has resulted in the statistical technique being widely and successfully applied to the modelling and simulation of radiotherapy treatments, with most modern radiotherapy treatment planning systems offering a Monte-Carlo algorithm for at least electron beams. A number of different codes are available for free download and require little prior programming experience. These codes enable the user to model the complex geometries in radiotherapy and imaging equipment and the beams they produce. These simulated beams can then be used to accurately model the dosimetry in phantoms, patients, detectors and the radiobiology at smaller micro- and nano-length scales. The Monte-Carlo method has a rich history in the development of new radiation production equipment and detector technology as well as improving our understanding of ionising radiation interactions in human tissues. The continued research into the development of Monte-Carlo codes and algorithms for radiotherapy applications would seem to indicate it will have an increasing role to play well into the foreseeable future.

References

Verhaegen, F, Seuntjens, J. Monte Carlo modelling of external radiotherapy photon beams. Phys Med Biol 2003; 48 (21): R107R164.CrossRefGoogle ScholarPubMed
Park, H, Paganetti, H, Schuemann, J, Jia, X, Min, CH. Monte Carlo methods for device simulations in radiation therapy. Phys Med Biol 2021; 66 (18): 18TR01. doi: 10.1088/1361-6560/ac1d1f CrossRefGoogle ScholarPubMed
Andreo, P. Monte Carlo simulations in radiotherapy dosimetry. Radiat Oncol 2018; 13 (1): 121.CrossRefGoogle ScholarPubMed
Chetty, IJ, Curran, B, Cygler, JE, et al. Report of the AAPM Task Group No. 105: issues associated with clinical implementation of Monte Carlo-based photon and electron external beam treatment planning. Med Phys 2007; 34 (12): 48184853.CrossRefGoogle ScholarPubMed
Brualla, L, Rodriguez, M, Lallena, AM. Monte Carlo systems used for treatment planning and dose verification. Strahlenther Onkol 2016; 193 (4): 243. doi: 10.1007/s00066-016-1075-8 CrossRefGoogle ScholarPubMed
Reynaert, N, van der Marck, SC, Schaart, DR, et al. Monte Carlo treatment planning for photon and electron beams. Radiat Phys Chem 2007; 76 (4): 643686.CrossRefGoogle Scholar
Rogers, DW. Fifty years of Monte Carlo simulations for medical physics. Phys Med Biol 2006; 51 (13): R287R301.CrossRefGoogle ScholarPubMed
Seco, J, Verhaegen, F. Monte Carlo Techniques in Radiation Therapy: Applications to Dosimetry, Imaging, and Preclinical Radiotherapy. Boca Raton: Taylor & Francis Group, 2021.CrossRefGoogle Scholar
Verhaegen, F, Seco, J. Monte Carlo Techniques in Radiation Therapy: Introduction, Source Modelling, and Patient Dose Calculations. Boca Raton: Taylor & Francis Group, 2021.CrossRefGoogle Scholar
Poludniowski, GG, Evans, PM. Calculation of x-ray spectra emerging from an x-ray tube. Part I. electron penetration characteristics in x-ray targets. Med Phys 2007; 34 (6): 21642174.CrossRefGoogle ScholarPubMed
Poludniowski, GG. Calculation of x-ray spectra emerging from an x-ray tube. Part II. X-ray production and filtration in x-ray targets. Med Phys 2007; 34 (6): 21752186.CrossRefGoogle ScholarPubMed
Rogers, DW, Faddegon, BA, Ding, GX, Ma, CM, We, J, Mackie, TR. BEAM: a Monte Carlo code to simulate radiotherapy treatment units. Med Phys 1995; 22 (5): 503524.CrossRefGoogle Scholar
Bazalova, M, Verhaegen, F. Monte Carlo simulation of a computed tomography x-ray tube. Phys Med Biol 2007; 52 (19): 59455955.CrossRefGoogle ScholarPubMed
Aghdam, SRH, Siavashpour, Z, Aghamiri, SMR, Mahdavi, SR, Nafisi, N. Evaluating the radiation contamination dose around a high dose per pulse intraoperative radiotherapy accelerator: a Monte Carlo study. J Radiother Pract 2020; 19 (3): 265276.CrossRefGoogle Scholar
Aghdam, MRH, Baghani, HR, Mahdavi, SR, Aghamiri, SMR, Akbari, ME. Monte Carlo study on effective source to surface distance for electron beams from a mobile dedicated IORT accelerator. J Radiother Pract 2017; 16 (1): 2937.CrossRefGoogle Scholar
Al-Ghorabie, FHH. Experimental measurements and Monte Carlo modelling of the XSTRAHL 150 superficial X-ray therapy unit. J Radiother Practice 2015; 14 (1): 4355.CrossRefGoogle Scholar
Sechopoulos, I, Rogers, DWO, Bazalova-Carter, M, et al. RECORDS: improved reporting of Monte Carlo radiation transport studies: report of the AAPM Research Committee Task Group 268. Med Phys 2018; 45 (1): e1e5.CrossRefGoogle ScholarPubMed
Padilla-Cabal, F, Pérez-Liva, M, Lara, E, Alfonso, R, Lopez-Pino, N. Monte Carlo calculations of an Elekta Precise SL-25 photon beam model. J Radiother Pract 2015; 14 (3): 112.CrossRefGoogle Scholar
Bahreyni Toossi, MT, Ghorbani, M, Akbari, F, Sabet, LS, Mehrpouyan, M. Monte Carlo simulation of electron modes of a Siemens Primus linac (8, 12 and 14 MeV). J Radiother Pract 2013; 12 (4): 352359.CrossRefGoogle Scholar
Zhao, YL, Mackenzie, M, Kirkby, C, Fallone, BG. Monte Carlo evaluation of a treatment planning system for helical tomotherapy in an anthropomorphic heterogeneous phantom and for clinical treatment plans. Med Phys 2008; 35 (12): 53665374.CrossRefGoogle Scholar
Zhao, YL, Mackenzie, M, Kirkby, C, Fallone, BG. Monte Carlo calculation of helical tomotherapy dose delivery. Med Phys 2008; 35 (8): 3491.CrossRefGoogle ScholarPubMed
Moskvin, V, DesRosiers, C, Papiez, L, Timmerman, R, Randall, M, DesRosiers, P. Monte Carlo simulation of the Leksell Gamma Knife: I. Source modelling and calculations in homogeneous media. Phys Med Biol 2002; 47 (12): 19952011.CrossRefGoogle ScholarPubMed
Moskvin, V, Timmerman, R, DesRosiers, C, et al. Monte carlo simulation of the Leksell Gamma Knife: II. Effects of heterogeneous versus homogeneous media for stereotactic radiosurgery. Phys Med Biol 2004; 49 (21): 48794895.CrossRefGoogle ScholarPubMed
Junios, J, Irhas, I, Novitrian, N, et al. Characterization of Gamma Knife PerfexionTM source based on Monte Carlo simulation. Radiol Phys Technol 2020; 13 (4): 398404.CrossRefGoogle Scholar
Mahmoudi, A, Shirazi, A, Geraily, G, Nia, TH, Bakhshi, M, Maleki, M. Penumbra width determination of single beam and 201 beams of Gamma Knife machine model 4C using Monte Carlo simulation. J Radiother Pract 2019; 18 (1): 8287.CrossRefGoogle Scholar
Kairn, T, Cassidy, D, Sandford, PM, Fielding, AL. Radiotherapy treatment verification using radiological thickness measured with an amorphous silicon electronic portal imaging device: monte Carlo simulation and experiment. Phys Med Biol 2008; 53 (14): 39033919.CrossRefGoogle ScholarPubMed
Chin, PW, Spezi, E, Lewis, DG. Monte Carlo simulation of portal dosimetry on a rectilinear voxel geometry: a variable gantry angle solution. Phys Med Biol 2003; 48 (16): N231N238.CrossRefGoogle ScholarPubMed
Siebers, JV, Kim, JO, Ko, L, Keall, PJ, Mohan, R. Monte Carlo computation of dosimetric amorphous silicon electronic portal images. Med Phys 2004; 31 (7): 21352146.CrossRefGoogle ScholarPubMed
Herwiningsih, S, Hanlon, P, Fielding, A. Sensitivity of an Elekta iView GT a-Si EPID model to delivery errors for pre-treatment verification of IMRT fields. Australas Phys Eng Sci Med 2014; 37 (4): 763770.CrossRefGoogle ScholarPubMed
Rahman, M, Bruza, P, Lin, Y, Gladstone, DJ, Pogue, BW, Zhang, R. Producing a beam model of the Varian ProBeam Proton Therapy System using TOPAS Monte Carlo Toolkit. Med Phys 2020; 47 (12): 65006508.CrossRefGoogle ScholarPubMed
Tourovsky, A, Lomax, AJ, Schneider, U, Pedroni, E. Monte Carlo dose calculations for spot scanned proton therapy. Phys Med Biol 2005; 50 (5): 971981.CrossRefGoogle ScholarPubMed
Paganetti, H, Jiang, H, Parodi, K, Slopsema, R, Engelsman, M. Clinical implementation of full Monte Carlo dose calculation in proton beam therapy. Phys Med Biol 2008; 53 (17): 48254853.CrossRefGoogle ScholarPubMed
Vedelago, J, Geser, FA, Muñoz, ID, Stabilini, A, Yukihara, EG, Jäkel, O. Assessment of secondary neutrons in particle therapy by Monte Carlo simulations. Phys Med Biol 2022; 67 (1): 015008. doi: 10.1088/1361-6560/ac431b CrossRefGoogle ScholarPubMed
Zarifi, S, Ahangari, HT, Jia, SB, Tajik-Mansoury, MA, Najafzadeh, M, Firouzjaei, MP. Bragg peak characteristics of proton beams within therapeutic energy range and the comparison of stopping power using the GATE Monte Carlo simulation and the NIST data. J Radiother Pract 2020; 19 (2): 173181.CrossRefGoogle Scholar
Zarifi, S, Ahangari, HT, Jia, SB, Tajik-Mansoury, MA. Validation of GATE Monte Carlo code for simulation of proton therapy using National Institute of Standards and Technology library data. J Radiother Pract 2019; 18 (1): 3845.CrossRefGoogle Scholar
Chiang, BH, Bunker, A, Jin, H, Ahmad, S, Chen, Y. Developing a Monte Carlo model for MEVION S250i with HYPERSCAN and Adaptive ApertureTM pencil beam scanning proton therapy system. J Radiother Pract 2021; 20 (3): 279286.CrossRefGoogle Scholar
Sarrut, D, Bała, M, Bardiès, M, et al. Advanced Monte Carlo simulations of emission tomography imaging systems with GATE. Phys Med Biol 2021; 66 (10): 10TR03. doi: 10.1088/1361-6560/abf276 CrossRefGoogle ScholarPubMed
Downes, P, Jarvis, R, Radu, E, Kawrakow, I, Spezi, E. Monte Carlo simulation and patient dosimetry for a kilovoltage cone-beam CT unit. Med Phys 2009; 36 (9): 41564167.CrossRefGoogle ScholarPubMed
Ding, GX, Duggan, DM, Coffey, CW. Accurate patient dosimetry of kilovoltage cone-beam CT in radiation therapy. Med Phys 2008; 35 (3): 11351144.CrossRefGoogle ScholarPubMed
Spezi, E, Downes, P, Radu, E, Jarvis, R. Monte Carlo simulation of an x-ray volume imaging cone beam CT unit. Med Phys 2009; 36 (1): 127136.CrossRefGoogle ScholarPubMed
Spies, L, Ebert, M, Groh, BA, Hesse, BM, Bortfeld, T. Correction of scatter in megavoltage cone-beam CT. Phys Med Biol 2001; 46 (3): 821833.CrossRefGoogle ScholarPubMed
Bootsma, GJ, Verhaegen, F, Jaffray, DA. The effects of compensator and imaging geometry on the distribution of x-ray scatter in CBCT. Med Phys 2011; 38 (2): 897914.CrossRefGoogle ScholarPubMed
Poludniowski, G, Evans, PM, Hansen, VN, Webb, S. An efficient Monte Carlo-based algorithm for scatter correction in keV cone-beam CT. Phys Med Biol 2009; 54 (12): 38473864.CrossRefGoogle ScholarPubMed
Mosleh-Shirazi, MA, Swindell, W, Evans, PM. Optimization of the scintillation detector in a combined 3D megavoltage CT scanner and portal imager. Med Phys 1998; 25 (10): 18801890.CrossRefGoogle Scholar
Swindell, W, Evans, PM. Scattered radiation in portal images: a Monte Carlo simulation and a simple physical model. Med Phys 1996; 23 (1): 6373.CrossRefGoogle Scholar
Flampouri, S, Evans, PM, Verhaegen, F, Nahum, AE, Spezi, E, Partridge, M. Optimization of accelerator target and detector for portal imaging using Monte Carlo simulation and experiment. Phys Med Biol 2002; 47 (18): 33313349.CrossRefGoogle ScholarPubMed
Parent, L, Fielding, AL, Dance, DR, Seco, J, Evans, PM. Amorphous silicon EPID calibration for dosimetric applications: comparison of a method based on Monte Carlo prediction of response with existing techniques. Phys Med Biol 2007; 52 (12): 33513368.CrossRefGoogle ScholarPubMed
Parent, L, Seco, J, Evans, PM, Fielding, A, Dance, DR. Monte Carlo modelling of a-Si EPID response: the effect of spectral variations with field size and position. Med Phys 2006; 33 (12): 45274540.CrossRefGoogle ScholarPubMed
Spezi, E, Lewis, DG. Full forward Monte Carlo calculation of portal dose from MLC collimated treatment beams. Phys Med Biol 2002; 47 (3): 377390.CrossRefGoogle ScholarPubMed
Chamberland, MJP, Taylor, REP, Rogers, DWO, Thomson, RM. Egs_brachy: a versatile and fast Monte Carlo code for brachytherapy. Phys Med Biol 2016; 61 (23): 82148231.CrossRefGoogle ScholarPubMed
Solberg, TD, DeMarco, JJ, Chetty, IJ, et al. A review of radiation dosimetry applications using the MCNP Monte Carlo code. Radiochim Acta 2001; 89 (4–5): 337355.CrossRefGoogle Scholar
Zaker, N, Zehtabian, M, Sina, S, Koontz, C, Meigooni1 AS. Comparison of TG-43 dosimetric parameters of brachytherapy sources obtained by three different versions of MCNP codes. J Appl Clin Med Phys 2016; 17 (2): 379390.CrossRefGoogle ScholarPubMed
Enger, SA, Landry, G, D’Amours, M, et al. Layered mass geometry: a novel technique to overlay seeds and applicators onto patient geometry in Geant4 brachytherapy simulations. Phys Med Biol 2012; 57 (19): 62696277.CrossRefGoogle ScholarPubMed
Poon, E, DeBlois, F, Devic, S, Vuong, T, Verhaegen, F. Monte Carlo modeling of a novel brachytherapy applicator for rectal cancer treatment using GEANT4. Radiother Oncol 2005; 76: S27S28.CrossRefGoogle Scholar
Landry, G, Reniers, B, Pignol, JP, Beaulieu, L, Verhaegen, F. The difference of scoring dose to water or tissues in Monte Carlo dose calculations for low energy brachytherapy photon sources. Med Phys 2011; 38 (3): 15261533.CrossRefGoogle ScholarPubMed
Sarrut, D, Bardiès, M, Boussion, N, et al. A review of the use and potential of the GATE Monte Carlo simulation code for radiation therapy and dosimetry applications. Med Phys 2014; 41 (6): 064301.CrossRefGoogle ScholarPubMed
Berumen, F, Ma, Y, Ramos-Méndez, J, Perl, J, Beaulieu, L. Validation of the TOPAS Monte Carlo toolkit for HDR brachytherapy simulations. Brachytherapy 2021; 20 (4): 911921.CrossRefGoogle ScholarPubMed
Dagli, A, Yurt, F, Yegin, G. Evaluation of BrachyDose Monte Carlo code for HDR brachytherapy: dose comparison against Acuros®BV and TG-43 algorithms. J Radiother Pract 2020; 19 (1): 7683.CrossRefGoogle Scholar
Zaman, A, Kakakhel, MB, Hussain, A. A comparison of Monte Carlo, anisotropic analytical algorithm (AAA) and Acuros XB algorithms in assessing dosimetric perturbations during enhanced dynamic wedged radiotherapy deliveries in heterogeneous media. J Radiother Pract 2019; 18 (1): 7581.CrossRefGoogle Scholar
Yabsantia, S, Suriyapee, S, Phaisangittisakul, N, Oonsiri, S, Sanghangthum, T, Seuntjens, J. Investigation of field output factors using IAEA-AAPM TRS-483 code of practice recommendations and Monte Carlo simulation for 6 MV photon beams. J Radiother Pract 2023; 22: 16.CrossRefGoogle Scholar
Vassiliev, ON, Peterson, CB, Chang, JY, Mohan, R. Monte Carlo evaluation of target dose coverage in lung stereotactic body radiation therapy with flattening filter-free beams. J Radiother Pract 2022; 21 (1): 8187.CrossRefGoogle ScholarPubMed
Tanha, K, Mahdavi, SR, Geraily, G. Comparison of CCC and ETAR dose calculation algorithms in pituitary adenoma radiation treatment planning; Monte Carlo evaluation. J Radiother Pract 2014; 13 (4): 447455.CrossRefGoogle Scholar
Nithiyanantham, K, Mani, GK, Raju, S, et al. Characterisation of small photon field outputs in a heterogeneous medium using X-ray voxel Monte Carlo dose calculation algorithm. J Radiother Pract 2018; 17 (1): 114123.CrossRefGoogle Scholar
Mamballikalam, G, Senthilkumar, S, Jayadevan, PM, et al. Evaluation of dosimetric parameters of small fields of 6 MV flattening filter free photon beam measured using various detectors against Monte Carlo simulation. J Radiother Pract 2021; 20 (3): 332339.CrossRefGoogle Scholar
Chow, JCL, Owrangi, AM. Monte Carlo study on mucosal dose in oral and naval cavity using photon beams with small field. J Radiother Pract 2011; 10 (4): 261271.CrossRefGoogle Scholar
Acar, H, Caglar, M, Altinok, AY. Experimental validatıon of peripheral dose distribution of electron beams for eclipse electron Monte Carlo algorithm. J Radiother Practice 2018; 17 (3): 279288.CrossRefGoogle Scholar
Abdul Aziz, MZ, Mohd Kamarulzaman, FN, Mohd Termizi, NA, Abdul Raof, N, Tajuddin, AA. Effects of density from various hip prosthesis materials on 6 MV photon beam: a Monte Carlo study. J Radiother Pract 2017; 16 (2): 155160.CrossRefGoogle Scholar
Incerti, S, Baldacchino, G, Bernal, M, et al. The Geant4-DNA project. Int J Model Simulat Sci Computing 2010; 1 (2): 157178.CrossRefGoogle Scholar
Incerti, S, Douglass, M, Penfold, S, Guatelli, S, Bezak, E. Review of Geant4-DNA applications for micro and nanoscale simulations. Phys Med 2016; 32 (10): 11871200.CrossRefGoogle ScholarPubMed
Bernal, MA, Bordage, MC, Brown, JMC, et al. Track structure modeling in liquid water: a review of the Geant4-DNA very low energy extension of the Geant4 Monte Carlo simulation toolkit. Phys Med 2015; 31 (8): 861874.CrossRefGoogle Scholar
Peukert, D, Incerti, S, Kempson, I, et al. Validation and investigation of reactive species yields of Geant4-DNA chemistry models. Med Phys 2019; 46 (2): 983998.CrossRefGoogle ScholarPubMed
Sakata, D, Belov, O, Bordage, MC, et al. Fully integrated Monte Carlo simulation for evaluating radiation induced DNA damage and subsequent repair using Geant4-DNA. Sci Rep 2020; 10 (1): 20788.CrossRefGoogle ScholarPubMed
Kyriakou, I, Sakata, D, Tran, HN, et al. Review of the Geant4-DNA Simulation Toolkit for Radiobiological Applications at the Cellular and DNA Level. Cancers 2021; 14 (1): 35. doi: 10.3390/cancers14010035 CrossRefGoogle ScholarPubMed
Schuemann, J, McNamara, AL, Ramos-Méndez, J, et al. TOPAS-nBio: an Extension to the TOPAS Simulation Toolkit for Cellular and Sub-cellular Radiobiology. Radiat Res 2019; 191 (2): 125138.CrossRefGoogle Scholar
Ramos-Méndez, J, Perl, J, Schuemann, J, McNamara, A, Paganetti, H, Faddegon, B. Monte Carlo simulation of chemistry following radiolysis with TOPAS-nBio. Phys Med Biol 2018; 63 (10): 105014.CrossRefGoogle ScholarPubMed
McNamara, A, Geng, C, Turner, R, et al. Validation of the radiobiology toolkit TOPAS-nBio in simple DNA geometries. Phys Med 2017; 33: 207215.CrossRefGoogle ScholarPubMed
Zhu, H, McNamara, AL, Ramos-Mendez, J, et al. A parameter sensitivity study for simulating DNA damage after proton irradiation using TOPAS-nBio. Phys Med Biol 2020; 65 (8): 085015.CrossRefGoogle ScholarPubMed
Rudek, B, McNamara, A, Ramos-Méndez, J, Byrne, H, Kuncic, Z, Schuemann, J. Radio-enhancement by gold nanoparticles and their impact on water radiolysis for x-ray, proton and carbon-ion beams. Phys Med Biol 2019; 64 (17): 175005.CrossRefGoogle ScholarPubMed
Ramos-Méndez, J, Domínguez-Kondo, N, Schuemann, J, McNamara, A, Moreno-Barbosa, E, Faddegon, B. LET-dependent intertrack yields in proton irradiation at ultra-high dose rates relevant for FLASH therapy. Radiat Res 2020; 194 (4): 351362.CrossRefGoogle ScholarPubMed
Belamri, C, Dib, ASA, Belbachir, AH. Monte Carlo simulation of proton therapy using bio-nanomaterials. J Radiother Practice; Cambridge 2016; 15 (3): 290295.CrossRefGoogle Scholar
Figure 0

Figure 1. Simple X-ray tube model created using the XTUBE component module of the BEAMnrc user code. Grey indicates a vacuum. Blue indicates the tungsten target embedded on the red copper mounting. Axis dimensions are in centimetres.

Figure 1

Figure 2. (a) Simple spherical cell model created using TOPAS-nBio. Red central sphere is the nucleus, green spheres are mitochondria and very small yellow spheres are gold nanoparticles. Red lines are the tracks of 50 keV electrons. (b) Free radical chemical species are generated along an electron track (shown in red) as it traverses water.