Sie sind auf Seite 1von 147

Part G Treatment Planning

1 Part G Treatment Planning


2 Introduction

3 Treatment planning is the process of determining the most appropriate way to irradiate
4 the patient. It is a combination of five essential steps:

5 1. Choosing an appropriate patient positioning and immobilisation method so that


6 treatments will be reproducible;

7 2. Identification of the target tissue;

8 3. Selecting a suitable beam arrangement;

9 4. Evaluating the resulting dose distribution;

10 5. Calculating the treatment machine settings to deliver the required absolute dose.

11 This results in the flow of data illustrated in Part H Figure H3.1.

12 Prior to Treatment Planning it will be necessary to determine the nature and extent of
13 the tumour using various diagnostic techniques. Occasionally it will be possible to conduct the
14 diagnostic investigation in such a way as to provide appropriate data for treatment planning, but
15 normally it will be necessary to carry out a separate imaging study. This is because the
16 requirements of diagnostic imaging and treatment planning are largely incompatible. For
17 diagnosis the priority is to obtain artefact free images with appropriate enhancement of the
18 tumour. Treatment planning requires a geometrically accurate 3-dimensional model of the
19 patient in the precise treatment position, together with a means of transferring the 3-dimensional
20 coordinate system to the treatment machine. For accurate tissue inhomogeneity corrections the
21 electron density distribution of the patient is also required.

22 For common treatment situations a standard beam arrangement will be established as a


23 “class solution” and planning will consist of establishing the correct beam geometry followed
24 by a dose computation to determine the monitor unit settings required. However, it is important
25 to realise that the arrangement of the beams relative to each other can have a significant effect
26 on the dose gradient at the edge of the target volume. Thus where the edge of the target volume
27 is determined by the penumbra of only one beam the dose gradient will be shallower than when

PartG.doc Page 1 of 147 13 January, 2004


Part G Treatment Planning

1 more than one beam edge is involved – as in the superior/inferior direction for coplanar
2 transverse beams. For this reason purely geometric planning is not sufficient. In determining
3 the field margins to be applied geometrically in such class solutions and for non-standard
4 treatments the relationship between the edge of the target volume and the edge of the dose
5 distribution will need to be considered in detail.

6 The size of the target volume and the margins around what is visible on an image should
7 be determined as suggested in ICRU 50 (ICRU 1993) and ICRU 62 (ICRU 1999). This is
8 covered in Chapter G1. Chapter G2 and Chapter G3 describe how to determine the planning
9 target volume (PTV). Chapter G4 considers the geometric aspects of planning and Chapter G5
10 and Chapter G6 look at ways of combining beams to achieve an appropriate dose distribution.
11 The algorithms used to calculate these dose distributions have already been described in Part F.
12 It is possible to calculate the effect of simple beam combinations using manual graphical
13 methods to estimate the summed contributions. While such methods can be instructive and can
14 help the student to understand the effect of combining beams, modern treatment planning is
15 almost exclusively carried out using computer algorithms and so manual methods will not be
16 discussed. However, it is essential that the radiotherapy physicist has a good grasp of the
17 concepts discussed in Part F that enable the absolute dose to be calculated. Finally, in Chapter
18 G7 and Chapter G8 methods of evaluating the dose distributions to establish the physical
19 relationship with the target volume and to obtain an estimate of the biological effect of the
20 treatment are considered.

21 Treatment planning requires a combination of a large amount of data and computer


22 systems to produce the final result. It is essential therefore that there is a detailed quality
23 assurance programme to minimise the risk of error. This is covered in Part H Chapter 3.

PartG.doc Page 2 of 147 13 January, 2004


Part G Treatment Planning

1 Chapter G1. Target Definition


2 Anthony J. Neal

3 G1.1. Introduction
4 The therapeutic use of radiation (radiotherapy) is an established method of treating
5 malignant tumours. Radiotherapy is a local treatment, complementary to systemic treatments
6 such as chemotherapy and hormone therapies. It is important to appreciate that the desired anti-
7 cancer activity and unwanted toxicity of radiotherapy is only seen in the tissues directly
8 irradiated by the primary beam(s).

9 Definition of the tumour, adjacent organs at risk and other anatomical structures is
10 therefore an essential part of the planning process, ensuring that the beam size, number,
11 trajectory and weighting are optimised in relation to these structures. This will in turn ensure
12 that the prescribed radiation dose is delivered to the appropriate volume of tumour bearing
13 tissue to achieve the endpoint of tumour cure or palliation (symptom control) whilst incurring
14 an acceptable rate and severity of both early- and late radiation morbidity.

15 In recent years radiotherapy has become increasingly complex. The current trend is
16 towards the implementation of conformal radiotherapy and intensity modulated radiotherapy
17 (IMRT). These aim to reduce normal tissue toxicity by limiting irradiation of normal tissues
18 whilst still irradiating the tumour in its entirety. This should in turn produce reduced morbidity
19 for a given absorbed dose compared with standard treatment techniques, or allow dose
20 escalation for an equivalent level of morbidity. These technologically demanding treatments can
21 only fulfil their promise if the therapeutic absorbed dose is delivered to the tumour reliably and
22 reproducibly from day to day during a protracted course of treatment.

23 The aim of this chapter is to outline the basic principles of target definition, exploring
24 the associated problems and uncertainties, and how they can be addressed in everyday clinical
25 practice. The detailed procedures for acquiring the anatomical data needed for the definition of
26 target volumes and other anatomical structures of interest will be described in Chapter G2 with
27 special reference to Computerised Tomography (CT) imaging. The specific role of Magnetic
28 Resonance (MR) imaging for treatment planning will be emphasized in Chapter G3.

PartG.doc Page 3 of 147 13 January, 2004


Part G Treatment Planning

1 G1.2. The ICRU volumes:


2 The ICRU 50 report (ICRU 1993) stipulates standard protocols for recording and
3 reporting radiotherapy treatments of all degrees of complexity. These include the adoption of
4 standard terminology to describe the volumes relevant to radiotherapy treatment planning. The
5 adoption of these terms serves several purposes:

6 • Improves clarity of thought and a logical approach to planning.

7 • Promotes consistency in physics planning and clinical practice.

8 • Allows standardisation of clinical trial protocols, particularly for complex, multi-phase


9 treatments.

10 • Facilitates communication between different centres and within clinical trials.

11 The volumes defined in the ICRU 50 report are described below.

12 G1.2.1. Gross tumour volume (GTV)

13 This is the gross palpable, visible and demonstrable extent and location of the malignant
14 growth. This is determined by physical examination by the oncologist and the results of
15 radiological investigations relevant to the site of the tumour. As the term suggests, tumours have
16 a length, breadth and depth and the GTV must therefore be identified using orthogonal 2 D
17 images or 3 D imaging. It may be clearly delineated for some tumours (e.g. cerebral metastases
18 or lung cancers) and very ill-defined for others (e.g. high grade gliomas of the brain or tumours
19 diffusely involving an organ). Even the imaging modalities with the best tissue contrast (CT and
20 MR – see above) may not resolve the tumour edge with precision, and indeed may not
21 distinguish between tumour and inflammatory tissue/host reaction.

22 G1.2.2. Clinical target volume (CTV)

23 This is a tissue volume encompassing the GTV but with an additional margin
24 accounting for any microscopic extension of the primary tumour or regional lymph node spread.
25 It is impossible to ascertain the degree of microscopic tumour invasion around a tumour unless
26 it is removed completely. Therefore this volume may not be defined separately but considered
27 when defining the planning target volume (see below). When specified, is more usual for an
28 empirical margin to be added to the GTV concentrically. The results of clinico-pathological

PartG.doc Page 4 of 147 13 January, 2004


Part G Treatment Planning

1 studies of primary tumours where microscopic extent can be measured, and perhaps correlated
2 with tumour type and other pathological markers of invasiveness, are not available for most
3 tumours. However, it is imperative that the CTV receives the prescribed radiation dose,
4 otherwise, tumour cure is unlikely.

5 G1.2.3. Planning target volume (PTV)

6 This is usually the final volume defined by the clinician during the planning process. It
7 encompasses fully the GTV and CTV. It is in effect a safety margin added to the GTV/CTV to
8 take account of the random and systematic errors inherent in radiotherapy set-up and treatment
9 delivery. Once again, it is often represented as a concentric margin around the GTV/CTV,
10 although data may be available for certain tumours (e.g. prostate) to demonstrate the vectors of
11 the most likely errors and their magnitudes. The PTV size, shape and position are used to
12 determine beam size, shape, number and orientation.

13 G1.2.4. Treated volume (TV)

14 This is the volume of tissue enclosed by an isodose surface selected and specified by
15 the clinician as being appropriate to achieve the aim of treatment i.e. cure or palliation. This
16 may, for example, be the 95% isodose volume for a curative treatment plan. Quite often, the TV
17 is significantly larger than the PTV – this is undesirable. The use of 3-D treatment planning and
18 shaping the radiation fields to the shape of the PTV using conformal radiation delivery
19 techniques ensures that the TV encloses the PTV with as narrow a margin as possible. This
20 ensures minimal irradiation of surrounding organs at risk whilst coverage of the PTV is assured.

21 G1.2.5. Irradiated volume (IV)

22 This is the tissue volume receiving a radiation absorbed dose that is considered
23 significant in relation to normal tissue tolerance. This concept is in practice not often considered
24 but may be useful when comparing one or more competing treatment plans. Clearly, it would be
25 preferable to accept the plan with the smallest IV, all else being equal.

26 G1.2.6. Organs at risk (OAR)

27 These are organs adjacent to the PTV which are non-target as they do not contain
28 malignant cells. The aim should therefore be to minimise irradiation of OARs as they are often

PartG.doc Page 5 of 147 13 January, 2004


Part G Treatment Planning

1 relatively sensitive to the effects of ionising radiation and, if damaged, may lead to substantial
2 morbidity. The OARs to be considered will vary greatly according to the anatomical region
3 being treated, the size of the PTV and the location of the PTV in these regions. The following
4 are examples of the most common OARs that must be considered:

5 • Brain – lens of eye, optic chiasm, brain stem.

6 • Head & neck – lens of eye, parotid glands.

7 • Thorax – spinal cord, lungs

8 • Abdomen – spinal cord, large bowel, small bowel, kidneys.

9 • Pelvis – bladder, rectum, femoral heads, large bowel, small bowel.

10 Many OARs have a defined tolerance dose at which a given level of late radiation
11 morbidity can be expected. These are summarised well in the article by {Emami et al. (1991)}.
12 Some OARs such as lung, brain and kidney demonstrate a volume effect with increasing loss of
13 function with increasing proportion of the organ irradiated.

14 Controversy exists concerning the outlining of hollow OARs e.g. colon or rectum.
15 Delineation of the outer surface of the wall of the structure is adequate for determining the
16 anatomical position. However, when a Dose Volume Histogram is to be constructed, this may
17 be inadequate as the volume of the wall will be overestimated. Several approaches to this
18 problem have been applied, with some centres outlining the outer and inner walls and others
19 calculating “dose-surface histograms” (see G7.3).

20 The ICRU 50 volumes and their relationship to each other are summarised in Figure G
21 1.1.

PartG.doc Page 6 of 147 13 January, 2004


Part G Treatment Planning

2 Figure G 1.1 Schematic representation of the volumes defined by ICRU demonstrating the
3 relationship of different volumes to each other. Redrawn from ICRU (1993)

4 G1.3. Clinical issues in target definition:


5 The acquisition of high quality images, and the definition of the ICRU 50 volumes are
6 integral parts of treatment planning. Outlining of these volumes (segmentation) is a time
7 consuming and tedious process, and therefore prone to random errors due to lapses in
8 concentration and uncertainties. Other steps in the planning process may introduce other
9 random and systematic errors. Failure to consider these issues will lead to errors being
10 implemented during the patient’s treatment. It is therefore vital to appreciate and if possible
11 eliminate sources of error and uncertainty. Some recognised problems with volume definition
12 are discussed below.

13 G1.3.1. Poor organ/tumour definition

14 For some malignancies, it is the usual practice to treat the whole of the organ of origin
15 of a tumour (e.g. the prostate or the bladder. In these instances, the primary tumour may not be
16 clearly seen, but this is irrelevant due to the large margin for error.

17 Some tumours are well circumscribed with an easily defined margin of demarcation
18 from the surrounding normal tissue. For example, benign tumours (e.g. meningiomas of the
19 brain and acoustic neuromas) have little propensity to invade the surrounding tissues and are

PartG.doc Page 7 of 147 13 January, 2004


Part G Treatment Planning

1 often encapsulated, making the tumour stand out clearly from the tissue of origin. The same can
2 be said for some malignant tumours such as brain metastases, low grade gliomas of the brain
3 and some soft tissue sarcomas. In these examples, there will be little dispute regarding
4 delineation of the GTV and in turn the PTV.

5 Inevitably, there are some tumours that are far less well defined. This will be a reflection
6 of diffuse infiltration at the tumour periphery and/or a similar radiographic density of the
7 tumour compared with the surrounding normal tissue. Thus it is possible that the tumour size
8 and shape will not be fully appreciated by the clinician. This is a particular problem with high
9 grade gliomas arising within the brain (Figure G 1.2), where clinico-pathological studies have
10 shown tumour extending for some centimetres beyond the edge of the main tumour mass. This
11 is usually accounted for by allowing a large (e.g. 3 cm.) margin between the GTV and PTV.
12 Imaging with MR is deemed superior to CT due to the superior soft tissue contrast for neural
13 tissue which is rich in lipids (and therefore protons) but it does not always allow clear tumour
14 delineation.

15

16

17

18

19

20

21

22

23 Figure G 1.2 High grade glioma of the brain. Note the poor demarcation between
24 tumour and the surrounding brain and in turn the difficulty in delineating the GTV in
25 each of these examples (a & b).

26 Comparison of CT and MR volumes suggests that in some cases there is a substantial


27 difference in the tumour volumes, with MR consistently leading to a larger volume {Ten
28 Haken et al. 1992}.

PartG.doc Page 8 of 147 13 January, 2004


Part G Treatment Planning

1 The appreciation of the position and extent of the GTV and of OAR benefits clearly
2 from the combination of several complementary imaging modalities. This requires that the
3 various images data sets are mapped onto a common coordinate system. This process, called
4 image “registration”, is mostly used to combine CT and MR images and has been developed
5 initially for the brain (see G3.6.2 and G3.7). The CT images are used as the reference (see
6 G2.3.1) and the MR images are recalculated and registered with respect to the CT coordinate
7 system, using one of the methods described succinctly in G3.6.2.

8 The practical use of registered multimodality data sets requires specific tools which
9 must be made available as part of a specific radiotherapy imaging console or be included in the
10 treatment planning system used for calculation of dose distributions. One solution can be to
11 display the reference and the registered images side by side and have a cursor linked pixel for
12 pixel, so that the clinician can move from one imaging modality to the other, modifying the
13 volume according to the image which demonstrates it best. The images can also be overlaid on
14 the top of each other with possibility to switch instantaneously from one modality to another
15 one. Finally the images can be blended in a single resulting image with possibility of
16 interactively change the relative weight of each of them. This process is commonly called image
17 fusion.

18 (a)

19 (b) (c)

20

21 Figure G 1.3a Small perspex fiducial marker. The chamber can be filled with an
22 appropriate radio-opaque contrast e.g. gadolinium for MR, iodinated contrast for CT or
23 positron emitting isotope for PET. It can then be placed on the surface of the patient at a
24 defined anatomical point b) CT image with markers c) PET image with markers

PartG.doc Page 9 of 147 13 January, 2004


Part G Treatment Planning

1 G1.3.2. Inter-observer variation

2 Some tumours are well circumscribed, and in these instances, there can be little debate
3 regarding the extent of the GTV. However, for more ill-defined tumours, there can be some
4 considerable inter-observer variation. It is clear that if the same set of images are presented to a
5 panel of specialists, one would expect some degree of concordance. In the study by {Leunens
6 et al. (1993}, CT images from several glioma patients were presented to a multi-disciplinary
7 group of clinicians including radiation oncologists, diagnostic radiologists and neurosurgeons,
8 varying in seniority. Comparison of the GTVs and PTVs suggested considerable variation from
9 person to person in the GTV contour, in some cases with a substantial difference in tumour
10 volumes. The only way to eliminate such subjectivity and improve consistency may be to use
11 computer image analysis although there are no data to suggest that this would be superior to the
12 average oncologist.

13 G1.3.3. Internal organ movements

14 For many years, it was assumed that internal organs did not move greatly from day to
15 day, or hour to hour throughout the day. Accordingly, little attempt was made to account for
16 these variations. This assumption is satisfactory for some sites where the tissue containing the
17 tumour has very little scope for movement e.g. the brain. However, with the emergence of
18 precision radiation delivery techniques such as conformal radiotherapy, these issues have had to
19 be considered. Conformal radiotherapy ensures conformation of the shape of the TV to the PTV
20 with as small a margin as possible. Therefore there is little room for error, and slight movement
21 of the prostate may result in under-treatment resulting in failure to control the tumour, or over-
22 treatment of an OAR leading to excessive radiation morbidity. As conformal radiotherapy has
23 been most widely used for treatment of the prostate, it is for this indication that most data are
24 available. The prostate is moderately fixed within the bony pelvis by a series of ligaments. The
25 bladder lies anteriorly, a hollow, muscular sac that undergoes regular distension and contraction
26 according to urine volume contained within it. The rectum lies posteriorly, a similarly hollow,
27 muscular sac that also undergoes regular distension and contraction according to the amount of
28 gas and faeces contained within it. Laterally lie the relatively immobile, inflexible bony wall of
29 the pelvis and femoral heads. It is clear that unless care is taken to ensure consistency of bladder
30 and rectal filling at the time of imaging, and subsequently at the time of treatment delivery,
31 diurnal variations could lead, not only to changes in the position of the prostate CTV, but also to
32 intrusion of the bladder and rectal walls (each OARs) into the TV. This hypothesis has been

PartG.doc Page 10 of 147 13 January, 2004


Part G Treatment Planning

1 elegantly demonstrated by {Ten Haken et al. (1991} using beam’s eye view images of the
2 prostate, bladder and rectum and 3D dose distributions. Other investigators have since
3 demonstrated the vectors and magnitudes of these movements using radio-opaque marker seeds
4 {Balter et al. 1995 a}. Such information can be used to optimise the margin allowed between
5 the CTV and the PTV.

6 Whilst it has been conventional practice to add a uniform margin around the GTV to
7 yield a PTV for planning, some authors have advocated the use of non-uniform margins to take
8 account of the particular treatment technique and anatomical relations between the PTV and
9 OARs. For example, Pickett et al. {1995} have proposed the use of margins varying from less
10 than 1 cm up to more than 2 cm for the prostate gland treated with a co-planar 6 field conformal
11 radiotherapy technique. This approach may provide the best compromise between CTV
12 coverage and avoidance of OARs. Similar data are available for other internal organs
13 susceptible to motion. For example, in the abdomen the kidneys are often in close proximity to
14 the radiation beam trajectory and occasionally are themselves susceptible to malignant disease.
15 They are also very susceptible to the late effects of ionising radiation and essential for the day to
16 day metabolic function of vertebrates. Being close to the thoracic cavity, the kidneys are
17 susceptible to respiratory motion, moving inferiorly with inspiration and superiorly with
18 expiration. The study by Schwartz et al. {1994} suggests that the inferior poles of the kidneys
19 may move by more than 4 cm in the superior-inferior direction between maximum inspiration
20 and expiration (mean 1.7 cm), while other studies have indicated that the displacement may be
21 even greater {Moerland et al. 1994}.

22 It has been noted that respiration leads to measurable variations in the CT volume
23 relationships of both the lungs and the liver, which can be appreciated by changes in the dose
24 volume histograms for these organs {Balter et al. 1996}. For those organs which move with
25 respiration, it is of course impractical to plan and deliver treatment during a breath hold without
26 some form of gating system. The relatively slow image acquisition of CT images in transverse
27 mode does at least give a time averaged picture that is more representative of the situation
28 during treatment. Empirical corrections in the superior-inferior direction can be used when
29 planning intrapulmonary tumours to take account of respiratory excursion, but it is recognised
30 that this may lead to increased normal tissue complication probabilities for the OARs {Ten
31 Haken et al. 1997}. Other options include the use of a stereotactic body frame to restrict lung
32 movement or to employ a “respiratory-gated” treatment in which the CT scanner and

PartG.doc Page 11 of 147 13 January, 2004


Part G Treatment Planning

1 subsequently the therapeutic radiation beam is switched on only at a predetermined phase of the
2 respiratory cycle {Ohara et al. 1989, Ramsey et al. 1999}. The latter is made more efficient by
3 being combined with Assisted Breathing Coordinator (ABC) (Remouchamps et al 2003a,
4 2003b, Wilson et al 2003)in which the patient is taught to regulate their breathing cycle.

5 The ultimate solution to the problem of internal organ movement subsequent to


6 treatment planning is to image the organ at the time of treatment. The technical issues are
7 considerable but have been overcome for some organs. For example, the placement of metallic
8 radio-opaque seeds at the periphery of the prostate allows the prostate to be located at the time
9 of treatment and the patient position corrected accordingly {Balter et al. 1995b}.

10 To address some of these issues ICRU report 62 has been introduced as a supplement to
11 ICRU report 50 {ICRU 1999}. This addresses some of the contemporary issues regarding target
12 definition and reporting for 3D treatment planning. The definitions of GTV and CTV remain
13 unchanged as these are oncological concepts independent of any technical developments. The
14 main changes of relevance include:

15 • The margin between the CTV and PTV is segregated into an “Internal Margin” (IM) to take
16 account of variations in the size, shape and position of the CTV relative to anatomical
17 reference points, and a “set-up” margin (SM) to take account of uncertainties related to set
18 up errors, reproducibility of patient position and mechanical stability of the treatment unit.
19 The report reflects the obvious uncertainty as to how these two margins should be added
20 together to produce the composite margin and therefore allow definition of the PTV. At
21 least it may serve to clarify thought and lead to a custom-defined PTV that takes into
22 account the relative vectors of these uncertainties, accepting that the whole process is a
23 compromise that relies on the experience and judgement of the radiotherapy team.

24 • Organs at risk (OARs) are defined according to their pattern of response to radiation
25 damage. “Serial” OARs have an architecture whereby damage to any one functional subunit
26 in a linear chain leads to sudden organ failure and in turn manifestation of morbidity (e.g.
27 spinal cord injury leading to radiation myelitis). “Parallel” OARs require damage to a
28 critical number of functional subunits to manifest radiation damage, and increasing radiation
29 damage leads to progressive loss of organ function (e.g. lung, kidney). “Serial-parallel”
30 OARs demonstrate attributes of both types (e.g. heart).

31 • Introduction of the concept of “Planning organ at risk volumes” (PORVs) which take
32 account of changes in the size, shape, volume and position of OARs. The PORV’s

PartG.doc Page 12 of 147 13 January, 2004


Part G Treatment Planning

1 derivation from the OAR is therefore analogous to the PTV relative to the CTV. The PTV
2 and PORV(s) may of course overlap, in which case judgement and experience must prevail.

3 • Introduction of the concept of a “conformity index”, defined as the quotient of the TV and
4 PTV, where the TV encompasses the PTV (see Chapter G7).

5 G1.4. Conclusions and future trends


6 The increasing sophistication of treatment planning and radiation delivery has
7 outstripped our capability to delineate the ICRU 50 volumes with precision and consistency.
8 Such uncertainties are added to systematic and day to day random errors in patient external
9 position and internal organ orientation {ref to IOP(?) publication on uncertainties}. Failure to
10 consider these issues will undermine the treatment, ultimately leading to an undesirable shift in
11 the balance between tumour control and normal tissue morbidity.

12 Future improvements in medical imaging such as functional imaging (PET and SPECT)
13 and their incorporation into the planning process will contribute to delineation of the GTV for
14 ill-defined tumours. More clinico-pathological data is necessary to relate tumour type, size and
15 grade to the degree of microscopic tumour extension around the GTV. These data may be used
16 to determine the optimum margins to be allowed around tumours to give the CTV; particularly
17 for those tumours that have reduced in size due to pre-treatment with chemotherapy and/or
18 hormone therapy. The acquisition of more data regarding internal organ motion may also permit
19 a better estimation of the margin that is necessary between the CTV and PTV.

20 ICRU report 62 has the potential to ensure that the description and reporting of target
21 definition keeps pace with the rapidly advancing field of 3-D treatment planning and the
22 complex issues that must be addressed by physicists and clinicians alike.

PartG.doc Page 13 of 147 13 January, 2004


Part G Treatment Planning

1 Chapter G2. Patient Data Acquisition


2 Anthony J. Neal

3 G2.1. Introduction
4 Patient data acquisition is the initial step of the treatment planning process. It is required
5 for three different purposes :

6 • to assess the position and extent of the target volume in relationship with the other
7 anatomical structures and particularly to organs at risk (see Chapter G1).

8 • to acquire the data required for accurate computation of the dose distribution (e.g. shape and
9 composition of the body)

10 • to acquire the information necessary for accurate set up of the patient (e.g. landmarks or
11 reference structures)

12 In all cases, it is essential that the patient position is defined at the very beginning and
13 kept throughout the whole treatment planning and delivery process. This will be addressed in
14 G2.2. The patient data acquisition by itself is nowadays mostly based on imaging devices which
15 obtain in a single phase the information required to achieve the goals listed above. The most
16 common methods are described in G2.3. A full chapter is dedicated to the use of MR imaging
17 for treatment planning which is increasingly implemented (see Chapter G3). In some instances
18 it may be necessary to combine several imaging modalities. This possibility will be discussed in
19 ?? and more specifically in G3.6.2 for MR.

20 G2.2. Patient positioning and immobilisation


21 External patient movements during imaging may introduce substantial random errors
22 into the planning process. Differences between positioning during imaging and subsequent
23 simulation and/or treatment may introduce further errors. Even if the deviation is in the order of
24 millimetres, it may be sufficient to lead to under-dosage of the CTV and in turn lead to
25 treatment failure. This is especially the case for conformal radiotherapy. Some of this
26 movement can be prevented by positioning patients comfortably and sympathetically, together
27 with an explanation of the importance of keeping still. Use of the immobilisation device that
28 will be utilised throughout treatment is essential. Despite this, imaging protocols can be lengthy,

PartG.doc Page 14 of 147 13 January, 2004


Part G Treatment Planning

1 particularly for 3 D treatment planning where such a large set of imaging data has to be
2 acquired.

3 Immobilisation during treatment has therefore become routine during imaging for
4 planning treatments where precision is important. Immobilisation devices are usually
5 individually fashioned for a particular patient and are used by them alone.

6 Examples of immobilisation devices include:

7 • Simple sand/bolus bags for general support.

8 • External arm supports and body “boards” e.g. breast and thoracic treatments.

9 • Thermoplastic casts e.g. for head and neck treatments, pelvic treatments.

10 • Vacuum bags e.g. for breast and pelvic treatments.

11 • Thermochemical polystyrene devices (headrests and knee supports) e.g. head and neck and
12 pelvic treatments.

13 • Stereotactic frames e.g. brain treatments.

14 Positioning studies have shown conclusively that such devices reduce the mean
15 positioning error for the anatomical sites for which they are used and improve consistency of
16 set-up.

17 It is a fundamental requirement of any imaging modality that patient positioning is


18 identical to that adopted for the proposed radiotherapy technique. This is usually best assured by
19 the therapy radiographers personally supervising the imaging following instructions from the
20 radiation oncologist.

21 G2.3. Imaging for treatment planning:


22 The clinician must be able to define the size, shape and location of the gross tumour
23 volume and organs at risk in three dimensional (3D) space. Small, superficial tumours (e.g. skin
24 cancers and tumours of the oral cavity) are easily visualised and palpated, and therefore may be
25 fully appreciated by the clinician without resorting to complex medical imaging methods.
26 However, for less well defined tumours, particularly those arising within the deeper parts of the
27 body, accurate information can only be obtained by one or more methods of medical imaging.

PartG.doc Page 15 of 147 13 January, 2004


Part G Treatment Planning

1 Quality assurance at each stage is essential to ensure integrity between planning and the
2 implementation of the treatment plan.

3 The ideal imaging modality for radiotherapy treatment planning would have the
4 following attributes:

5 • Well tolerated and atraumatic.

6 • Quick to acquire images (less movement artefact)

7 • Excellent soft tissue and bone contrast with high resolution.

8 • No geographical image distortion.

9 • Able to record a 3 D imaging data-set.

10 • Data can be used for radiotherapy dose calculations.

11 • Widely available and cost effective.

12 The most widely used imaging modalities will now be compared and contrasted in the
13 context of radiation treatment planning:

14 G2.3.1. Computed tomography (CT)

15 Computed tomography should be considered the ‘gold standard’. CT is widely available


16 at all cancer centres and many small ‘district’ hospitals. An example of a CT image of ?? is
17 shown in Figure G 2.1. The images are acquired reasonably quickly and comfortably, and
18 therefore movement artefact is not a problem. The images of the tumour can be augmented by
19 the injection of radio-opaque iodinated contrast media, particularly useful for tumours of the
20 central nervous system. Overall, soft tissue contrast is very good with high resolution due to the
21 small pixel size attainable. Bone contrast is excellent, a useful attribute when bone invasion
22 must be visualised e.g. for tumours arising within the base of skull, tumours invading adjacent
23 bone. The modest radiation exposure is not a problem for the cancer patient but it does mean
24 that children and anxious adults must remain unaccompanied during imaging.

25

26

27

PartG.doc Page 16 of 147 13 January, 2004


Part G Treatment Planning

10

11

12 Figure G 2.1 Example of a CT image of ??

13 CT produces a full 3 D set of imaging data which is easily imported to - and handled by
14 all commercial treatment planning systems. The high resolution is important when small surface
15 landmarks (e.g. a small radio-opaque point marker as a reference point) must be resolved. There
16 is no inherent image distortion, and therefore the 3 D positional data is reliable provided that
17 adequate quality assurance is performed. One particular advantage of CT is that the images are
18 an indirect representation of the various tissue electron densities within the images. As
19 megavoltage radiation interacts with tissue predominantly according to electron density
20 (Compton scattering), the CT data can be used directly for radiation absorbed dose calculations
21 providing a correction factor is applied to take account of the kilovoltage photon energy used
22 for CT data acquisition. This is particularly relevant when the radiation beam(s) traverse tissue
23 inhomogeneities such as lung and bone which attenuate the beam very differently compared
24 with a homogeneous volume of soft tissue. This is far preferable to manual corrections for
25 different tissue densities.

PartG.doc Page 17 of 147 13 January, 2004


Part G Treatment Planning

1 ADD REFERENCE TO CORRESPONDANCE BETWEEN HOUNSFIELD UNITS AND


2 DENSITY + CT CALIBRATION = TO BE COMPLETED ACCORDING TO CONTENT
3 OF QUALITY ASSURANCE PART.

4 CT is particularly useful for acquiring the 3 D external contour of the patient which
5 would be very tedious to acquire manually (although there are a number of laser devices
6 commercially available for this purpose). The external contour is essential for planning,
7 representing the interface between air and tissue and therefore allowing the planning computer
8 to take account of variations in beam attenuation and obliquity of incidence to calculate dose
9 distributions. Fortunately, simple segmentation algorithms can be used on the 3 D data to draw
10 an appropriate outline automatically on each transverse image.

11 In the context of acquisition of treatment planning images, CT is sufficiently versatile to


12 permit imaging in the radiotherapy treatment position. This will entail substitution of the convex
13 couch insert with a flat one so that the set-up emulates a radiotherapy simulator or treatment
14 couch. Unfortunately, CT is unsuitable for patients who are unable to lie reasonably flat. As
15 with conventional, simulator-based planning, it is important that the patient is positioned by a
16 therapy radiographer in the position for treatment. This should be both comfortable and
17 reproducible, preferably following a standard departmental protocol employing the placement
18 of radio-opaque fiducial reference points and the use of immobilisation devices as appropriate.

19 It has been customary to acquire data as a sequence of transverse images. However,


20 many modern scanners have the capability to use a helical data acquisition mode. This is
21 usually a far quicker protocol but is yet to gain favour for radiotherapy treatment planning due
22 to problems with reconstruction of organs and image software issues with treatment planning
23 computers. The area of imaging will be determined by the anatomical region occupied by the
24 tumour, the extent of the likely PTV and the likely trajectories of the beams that will be used to
25 cover it. Any organs at risk must be imaged in their entirety if dose volume histograms (DVHs)
26 or biological endpoints such as normal tissue complication probability (NTCP) will be
27 calculated. The imaging protocol will require that contiguous transverse images are obtained to
28 ensure that a full 3 D appreciation of the anatomy is obtained. The image slice separation will
29 not normally be greater than 10 mm, and frequently 5 mm or less, particularly when the tumour
30 region itself is imaged and even more so when the GTV is very small e.g. in the brain. The use
31 of narrower image acquisitions through the region of the tumour allows accurate delineation of
32 the superior and inferior extent of the GTV, and in turn increases the accuracy of the final CTV

PartG.doc Page 18 of 147 13 January, 2004


Part G Treatment Planning

1 and PTV. Acquisition of thin contiguous slices (e.g. with thickness smaller than 5 mm) over a
2 long supero-inferior distance carries with it the disadvantages of computer storage space and the
3 tedium of outlining the therapy relevant outlines on a large number of individual images. On the
4 other hand, since large series of thin images are useful for the generation of good quality
5 digitally reconstructed radiographs (see G4.4.5),a reasonable compromised must be reached.

6 The restricted aperture and field of view of a conventional diagnostic CT scanner can be
7 a problem for treatment planning for breast cancer patients, who would normally be treated with
8 the arm elevated.. This problem does not exist if the CT option of a simulator is used (see Part C
9 Section C4.3.3) but in such a case, the acquisition is restricted to a small number of slices and it
10 is not possible to perform a full 3D reconstruction and to apply conformal radiotherapy
11 technique. One other disadvantage of CT is seen when there is a metallic implant such as joint
12 prosthesis (see Figure G 2.2), surgical clips or large masses of dental amalgum. Such objects
13 can lead to considerable streaking of the CT images which may make it impossible to undertake
14 planning.
15

16

17

18

19

20

21

22

23

24

25

26

27

28

29 Figure G 2.2 CT image of the pelvis with artefacts from a metallic hip prosthesis.

PartG.doc Page 19 of 147 13 January, 2004


Part G Treatment Planning

1 G2.3.2. Magnetic resonance imaging (MR)

2 MR imaging provides an important added value to CT imaging in particular for target


3 delineation (see G1.3.1). Although often presented in axial sections as for CT images, MR
4 images have a very different appearance which depends widely on the acquisition sequence and
5 on the display settings. Because of the MR principle and acquisition method, the bony
6 structures do not exhibit any signal (black areas) whereas small differences in soft tissue
7 composition can be easily seen . A typical MR image of the pelvis which exhibits some
8 distortion is shown on Figure G 2.3. A comparison of registered CT and MR images of the skull
9 is shown on Figure G 2.4.

10

11 Figure G 2.3 MR image of the pelvis demonstrating non-linear distortion. Note the radio-
12 opaque point markers at the edge of the image which were linearly arranged rather than a
13 concave arrangement as recorded on the image.

PartG.doc Page 20 of 147 13 January, 2004


Part G Treatment Planning

2 Figure G 2.4 Side by side display of a CT image (left) and of a MR image (right) of the skull.
3 The bony structures cannot be seen on the MR image where the dotted line corresponds to the
4 registered bony outline of the CT image. The tumour delineation (blue) is much easier on the
5 MR image (from Khoo et al 2000)

6 The increasing importance of MR imaging for treatment planning is such that a full
7 chapter is devoted to this topic. The reader is therefore encouraged to refer to Chapter G3 for
8 complementary information.

9 G2.3.3. Ultrasound (US)

10 Ultrasound is cheap, simple to perform, non-invasive and well tolerated. It provides


11 reasonable images of tumours, but only when they are in close proximity to the probe. It is
12 therefore unsuitable for tumours situated deep within the thorax or abdomen. Bone significantly
13 limits the depth of penetration of ultrasound and is itself not resolved with much detail at all.
14 The images are generally difficult for the non-expert to interpret, and open to subjectivity.
15 Image quality has been improved in recent years by the introduction of transducers that can be
16 introduced into body cavities to visualise adjacent structures e.g. rectal US probe to image
17 prostate. Because it relies on assumptions about the speed of sound in tissue it may be subject to
18 geometrical distortion.

19 The main limitation of ultrasound for radiotherapy is that it does not produce a full 3 D
20 dataset for processing by a treatment planning computer. It is more usual to merely have a
21 succession of 2 D images obtained at the discretion of the radiologist.

22 The role of ultrasound is therefore currently confined to providing staging information


23 complementary to other imaging modalities that can indirectly guide the clinician with the task

PartG.doc Page 21 of 147 13 January, 2004


Part G Treatment Planning

1 of volume definition. Simple depth measurements of the chest wall are useful for planning
2 electron boosts in breast cancer patients as they help select the optimum electron energy.

3 G2.3.4. Positron emission tomography (PET)

4 PET utilises positron emitting, biologically active molecules so that PET images reflect
5 the biochemical function of tissues and organs rather than their structure. It therefore produces
6 information complementary to that obtained from CT and MR. It yields 3D data sets which can
7 be imported to the treatment planning computer. It may have a useful role in helping to define
8 the edge of ill-defined, metabolically highly active tumours such as gliomas of the brain. Both
9 CT and MR imaging may not yield much information regarding involvement of regional lymph
10 nodes, and often involvement can only be inferred when the node is significantly enlarged. PET
11 may again help to determine whether a node of borderline size is likely to be involved, and can
12 even detect tumour in nodes that were previously normal on CT/MR. One specificity of PET
13 imaging is therefore the fact that it can be thought as a mean to evaluate the directly the CTV
14 since it is likely to provide useful information on the spread of the microscopic disease.

15 As with MR, the images must be registered using CT as a ‘gold standard’. The images
16 are unfortunately rather low in resolution compared to CT and MR. One of the main
17 disadvantages is the fact that PET imaging facilities are few and far between and the technology
18 carries a heavy cost penalty with it. Even fewer centres have a dedicated cyclotron for
19 generating the precious, short half-life positron emitting isotopes. Reference to (Figure G 2.5).

20

21

22

23

24

25

26

27

PartG.doc Page 22 of 147 13 January, 2004


Part G Treatment Planning

5 Figure G 2.5 PET image (to be obtained)

6 G2.3.5. Planar radiographs

7 Planar radiographs have the advantage of being easy to obtain and interpret. They may
8 be used alone to define primary tumours of the lung or secondary tumours of bone where there
9 is sufficient contrast between the tumour and surrounding tissues. Reliance on such 2 D images
10 may be sufficient for the planning of non-curative, palliative treatments where a geographical
11 tumour miss may not affect symptom relief and where normal tissue toxicity is not so great an
12 issue due to the lower radiation doses used. A single film may be sufficient when an anterior-
13 posterior parallel opposed pair of beams is to be used. Orthogonal films are used when the
14 tumour has to be more precisely defined. An anterior film will allow delineation of the tumour
15 in the superior-inferior and left-right directions, while a lateral film will allow delineation in the
16 superior-inferior and anterior-posterior directions.

17 Contrast enhanced radiographs are useful in non-3D planning where it is important


18 to identify the position of a hollow viscus:

19 • Barium may be swallowed to outline the oesophagus, stomach and small bowel.

20 • Barium may be given as an enema to outline the rectum and lower colon.

21 • Radio-opaque tampons may be inserted into the vagina to outline this structure.

22 • Intravenous iodinated contrast may delineate the kidneys and allow them to be avoided
23 when the abdomen is treated.

24 • Iodinated contrast may be introduced into the bladder to outline it and suggest the position
25 of the top of the prostate.

26 • Iodinated contrast can be introduced into the urethra to suggest the position of the lowest
27 extent (apex) of the prostate.

PartG.doc Page 23 of 147 13 January, 2004


Part G Treatment Planning

1 • Metallic seeds can be introduced per rectum into the pelvis under guidance of trans-rectal
2 ultrasound to define the prostatic apex and base of seminal vesicles Figure G 2.6.
3

4
5 Figure G 2.6 a) Individual gold seeds cut from a length of wire.
6 b) CT topogram showing 3 seeds within the pelvis of a humanoid phantom.

7 G2.4. Conclusions and future trends


8 Acquisition of high quality 3D imaging data is becoming increasingly necessary to
9 exploit recent advances in the precision delivery of radiation. CT and MR are the cornerstones
10 of any imaging protocol. The development of PET imaging, preferably combined with CT
11 within the same piece of equipment, is likely to become standard practice in the future. Care
12 must be taken to ensure consistency of patient set-up for each imaging modality and for each
13 subsequent step of the planning process. This requires the implementation of a quality assurance
14 program as discussed in Part G Chapter 3.

PartG.doc Page 24 of 147 13 January, 2004


Part G Treatment Planning

1 Chapter G3. Magnetic Resonance (MR) Imaging in


2 treatment planning
3 Vincent S. Khoo

4 G3.1. Introduction
5 Computed tomography (CT) has made a dramatic impact on the accuracy of
6 radiotherapy treatment planning in the past two decades by improving the localisation of target
7 volumes which will allow higher local control rates with less geographical misses and permit
8 better awareness of the surrounding critical organs thereby potentially minimising normal tissue
9 complications (Goitein et al. 1979, Goitein 1979). The development of magnetic resonance
10 (MR) has introduced several added imaging benefits that may confer an advantage over the use
11 of CT in treatment planning.

12 G3.2. Principles of MR imaging


13 The basis of magnetic resonance (MR) techniques is the measurement of radiofrequency
14 radiation resulting from transitions induced between nuclear spin states of tissue hydrogen
15 atoms (protons) in the presence of a strong external magnetic field. Whereas CT depends on X-
16 ray attenuation by tissues which is a function of atomic number and electron density, relative
17 pixel intensities in MR images are a function of proton densities and tissue relaxation times.
18 MR imaging does not use ionising radiation.

19 The major source of contrast in clinical MRI is the difference in relaxation times
20 between different tissue types. The two parameters that are often used to characterise the
21 behaviour of an MR signal and which can be used as a basis for generating MR contrast are the
22 Spin-Lattice (T1) and Spin-Spin (T2) relaxation times. MR images are calculated from signals
23 acquired by applying a particular '
imaging or pulse sequence'comprised of one or more
24 radiofrequency pulses and magnetic field gradients and the rate at which the transitions occur or
25 relaxation times. The nature of the acquired MR signal and hence the image contrast depends
26 on the relative timings of the applied pulse(s), gradients and signal acquisition in relation to the
27 proton density structure of the tissue(s) being imaged.

28 A feature of MRI is that there is no standard universally applied imaging sequence;


29 there being an infinite set of combinations of possible timings and arrangements of the various

PartG.doc Page 25 of 147 13 January, 2004


Part G Treatment Planning

1 imaging sequence components. In practice, sequences are often designed to give images that
2 are weighted according to a particular relaxation time (i.e. T1- or T2-weighted). Even when
3 utilising a particular type of weighting, there is flexibility and variability in the imaging
4 parameters used. Thus T1 weighted images acquired on different MR scanners are often
5 obtained using slightly different values of the echo- (TE) and repetition- (TR) times resulting in
6 images with subtly different contrast. Pixel intensities also vary from measurement to
7 measurement, depending on the instrument design and calibration. MRI therefore has a greater
8 flexibility in varying contrast compared to CT where essentially fixed imaging parameters are
9 used. Using CT, each tissue type, therefore, gives image intensities that are relatively constant
10 and which can be quantified in terms of '
Hounsfield Units'
.

11 The flexibility inherent in MRI can be used to some extent to tailor the image contrast to
12 meet clinical requirements. Choice of contrast is then influenced by both clinical imperatives
13 and scanner constraints. Some of the clinical factors include the need for varying relaxation
14 times locally by administering contrast agent; the choice of the optimal imaging plane (i.e.
15 axial, sagittal, coronal etc), the scanning volume and patient tolerance. Other factors that
16 influence image quality include image Signal to Noise and Contrast to Noise ratios, the image
17 resolution and the scanning time. The acquired MR image is often a compromise between
18 interdependent and mutually exclusive imaging requirements.

19 G3.3. Rationale for the use of MR in treatment planning


20 The main advantage of MR compared to CT, is its ability to better demonstrate and
21 characterise tumours and soft tissues.

22 Contrast between healthy and malignant tissue can be obtained from differences in the
23 T1 and T2 relaxation times exploited by using the appropriate imaging sequences. Imaging
24 sequences can be conveniently characterised in terms of Spin-Echo (SE) or Gradient-Echo (GE)
25 sequences. A common patient scanning protocol involves both T1-weighted spin-echo imaging
26 (short TE and TR) for good anatomical delineation and T2-weighted spin-echo imaging (long
27 TE and TR) to differentiate pathological from normal tissues. An example of this can be seen
28 for the prostate gland (Figure G 3.1).
29
30
31
32
33

PartG.doc Page 26 of 147 13 January, 2004


Part G Treatment Planning

1
2
3
4
5
6
7
8
9
10
11
12
13
14

15 Figure G 3.1 (a) T1 weighted MR image of the prostate; (b) T2 weighted MR image of the prostate.
16 T1 weighted images are useful for highlighting the periprostatic tissues which lie between the prostatic
17 capsule and the puborectalis muscle (small arrows). T2 weighted images show the internal prostatic
18 structure better with the central zone appearing dark (large arrows) and the peripheral zone appearing
19 brighter (medium arrow). The prostatic capsule is also easily visualised (small black arrow heads)

20 Anatomical regions surrounded by thick bone will absorb x-rays from CT, reducing soft
21 tissue image quality, producing artefacts and decreasing the visualisation of nearby tumours.
22 This effect is absent with MR because cortical bone does not provide a MR signal and appears
23 as a dark area. Thus tumours within the posterior fossa or brainstem, and tumours centred at
24 bony prominences or enclosures such as the spinal cord are better defined.

25 The use of MR contrast agents such as Gadopentetate dimeglumine (GDPA) may


26 further enhance visualisation of tumour or disease process under investigation. MR has been
27 shown to be superior to CT in the staging of soft tissue (Herrlin et al. 1990) and pelvic tumours
28 (Hricak 1991), and sensitive in detecting and defining the extent of bone marrow disease
29 (Kattapuram et al. 1990). It is the imaging modality of choice for brain, spinal cord, and some
30 head and neck tumours (De Vries and Bydder 1988, Sze 1988).

31 Detailed MR images can be acquired in any orthogonal or non-orthogonal plane without


32 loss of spatial resolution. This multiplanar capability provides greater functionality for 3D
33 treatment planning compared to CT. Furthermore, 3D volumetric data sets can be obtained
34 thereby avoiding the need for separate axial, sagittal, and coronal plane acquisitions. Recent
35 developments have produced faster imaging and real time applications. MR imaging in cine
36 mode display can provide '
beam-eye-views’ capability (Ogino et al. 1993) and may be used to
37 assess organ motion (Padhani et al. 1999). These developments may allow the development of
38 '
virtual or real-time'MR treatment simulation (Mizokawi et al. 1996).

PartG.doc Page 27 of 147 13 January, 2004


Part G Treatment Planning

1 By combining the high resolution anatomical information from MR with functional


2 information available from PET and SPECT, a clinically useful structure-function map can be
3 produced. This procedure has been used in glioma radiotherapy (Pardo et al. 1994) and applied
4 with radiolabeled monoclonal antibodies or radioimmunotherapy to evaluate the 3D dosimetry
5 of the target-to-surrounding normal tissue absorbed-dose ratios (Sgouros et al. 1993). Another
6 advantage is that MR can record the spatial T1-induced changes from irradiation within doped
7 gel phantoms allowing complex radiotherapy dose distributions such as those from 3D intensity
8 modulated treatments and brachytherapy (Schreiner et al. 1994) to be mapped.

9 MR can also provide physiological and biochemical tumour information. MR


10 angiography can assess cerebral vasculature, providing a non-invasive method to plan
11 stereotactic radiotherapy for cerebral vascular malformations and angiomas. MR Spectroscopy
12 can non-invasively measure the behaviour of a range of clinically significant biochemical
13 compounds and may offer new methods to study morphological alterations, physiological and
14 metabolic changes non-invasively within tumours during and post-treatment (Leach 1994). This
15 may allow individual modification of radiation dose or prompt alternative treatments for non-
16 responders. Thus, MR imaging provides a much wider range of applications not only for
17 treatment planning but for cancer management. CT is limited in these aspects showing only
18 anatomical images with contrast based solely on differences in intrinsic or contrast agent
19 generated X-ray attenuation. Differences between CT and MR imaging are listed in Table F 3.1.

20 G3.4. Problems with the use of MR imaging in treatment planning


21 Although the imaging quality and versatility of MR is ideal for acquiring the necessary
22 information for treatment planning, MR has not yet seriously challenged CT for treatment
23 planning in most sites because of the following reasons:

24 G3.4.1. Electron density information

25 Whereas CT depends on X-ray attenuation by tissues which is a function of atomic


26 number and electron density, relative pixel intensities in MR images are a function of proton
27 densities and tissue relaxation times. For CT data, electron density can be automatically
28 calibrated from Hounsfield units but MR signal intensity has no such correlation. For dose
29 calculations, MR images will need either to have absorption coefficients assigned to the relevant

PartG.doc Page 28 of 147 13 January, 2004


Part G Treatment Planning

1 anatomical structures/regions or to register the MR and CT images so that the superior tumour
2 definition of MR can be transferred to the CT data for planning.
3

4 Table F 3.1 General comparisons between MRI and conventional CT imaging


5

Subject Parameters MRI CT

Patient Magnetic safety concerns Present Nil


Radiofrequency heat deposition Present Nil
Ionizing radiation dose Nil Present
Claustophobia in scan tube More marked Minimal
Scanning noise Moderate Minimal
Contrast materials allergy
-iodinated contrast Not applicable Present
- Gd DTPA (gadolinium) Minimal Not applicable

Characteristics Soft tissue contrast Excellent Moderate


Cortical bone contrast Poor Excellent
Detection of calcification Poor Excellent
Metallic artefacts
- non-ferromagnetic material Some Marked
- ferromagnetic material Marked Marked

Machine Size of '


patient tube' Smaller Larger
Image resolution Good Better
Scanning time *Moderate *Short
Electron density information Nil Present
Functionality and technical Large Limited
sequences
Geometrical image accuracy Distortion present - Excellant
object, sequence and
machine dependent
Multiplannar imaging Any plane Limited
Multiplanner reconstructions Available Available
Cost Higher Lower
Availability Restricted Widely available
6 *scanning times are dependant on the imaging volume, number of slices and for MRI , the pulse
7 sequences used.
8

9 G3.4.2. Imaging of bone

10 Cortical bone in MR imaging is shown as regions of very low signal intensity. The
11 presence, type (compact bone vs. spongy bone), shape and extent of bone are important in
12 assessing bone/tissue or bone/air inhomogeneities for dose calculations. Bony boundaries and
13 landmarks are not clearly visible on MR images which can limit image registration.

PartG.doc Page 29 of 147 13 January, 2004


Part G Treatment Planning

1 G3.4.3. MR Image Distortion

2 A highly uniform magnetic field with perfectly linear orthogonal gradients is desired for
3 MR imaging. However, inhomogeneities of the main magnetic field, and non-linearities in the
4 gradients together with the presence of objects within the magnet field will cause both
5 geometric and signal intensity distortion in clinical practice. The presence of any imaging
6 distortion presents a major hindrance in the utilisation of MR images for treatment planning as
7 it will provide inaccurate spatial information and preclude accurate image correlation. In
8 contrast, CT data are spatially accurate and can be used directly in treatment planning systems
9 without the need for correction. Sources of MR image distortion can be conveniently considered
10 as system-related distortions and object-induced (i.e. patient-dependant) effects Figure G 3.2.
11
12

13
14 Figure G 3.2 Various forms of MR-related distortion can be seen in this image of a coplanar
15 array of water-filled tubes supported within a circular solid plastic block. System-related
16 distortion effects are seen in the apparent curvature of the tubes at A and their disappearance
17 at B, which was due to warping distortion of the imaging plane. Magnetic susceptibility
18 differences due to the presence of the plastic support block at C give rise to object-induced
19 distortions in the form of discontinuities at the point where each tube enters the support block.
20 The resultant displacements have occurred along the readout (horizontal) direction.

21 G3.4.3.1. System-related distortions

22 System-related image distortions arise due to inhomogeneities in the main magnetic


23 field, non-linearity of magnetic field gradients and the presence of eddy current effects. All
24 magnetic fields possesses inhomogeneities which result from imperfections in the magnet
25 windings or stray external magnetic fields and cause image slice distortion and in-plane
26 distortions in the read-out gradient direction. For a given magnet inhomogeneity, the magnitude
27 of the resulting image distortion is inversely proportional to the gradient strength. The
28 homogeneity of the static magnetic field (Bo) decreases with increasing distance from the centre
29 of the magnet bore and therefore the degree of distortion increases for larger fields-of-view.

PartG.doc Page 30 of 147 13 January, 2004


Part G Treatment Planning

1 MR images can also be distorted by non-linearities in the applied gradients and by the
2 presence of eddy currents which cause both slice and in-plane distortion in both the read-out
3 and phase-encoding directions. Eddy currents arise when the applied gradients are turned on or
4 off resulting in the presence of transient magnetic fields which contribute to image distortion.
5 Newer MR scanners utilise shielded and compensated gradients to reduce eddy currents. In
6 practice, these distortions remain constant if the imaging sequence and its parameters (TE, TR,
7 etc) are unchanged. Hence eddy current induced distortions may be quantified and corrected
8 provided the same imaging sequence is used for mapping distortions and imaging patients.

9 For brain imaging with head coils or similar fields-of-view (i.e. 10-15 by 10-15 cm),
10 investigators have reported distortion errors of 3-4 mm at the periphery and less than 1 mm at
11 the centre of the image and these images were considered satisfactory for planning purposes
12 (Schad et al. 1987, Pötter et al 1992, Hill et al. 1994). With larger fields-of-view between 20
13 and 40 cm, phantom studies have revealed point displacements of up to 10 mm and 16 mm
14 respectively at the edge of the image (Kokoves 1991, Finnigan et al. 1997). Uncorrected
15 images with distortions of this magnitude cannot be used for the generation of accurate patient
16 outlines in treatment planning. The system-related distortion are strongly dependent on the
17 design of the MR system. They are also influenced by the imaging sequence, the size of the
18 gradients employed and the position within the field-of-view.

19 G3.4.3.2. Object-induced distortions

20 Object-induced distortions arise principally from magnetic susceptibility and chemical


21 shift effects. The magnetic susceptibility effect results because any object placed within a
22 magnetic field changes the value of that field. This distortion is particularly marked at
23 boundaries of structures having different intrinsic magnetic susceptibilities and tends to be most
24 pronounced at air/tissue boundaries, for example, near the outline of the head and at internal
25 cavities such as the sinuses. These effects can be complex as they depend on the orientation of
26 the boundary to the main magnetic field and can influence the value of the field at substantial
27 distances from the susceptibility interface.

28 Another major source of object-induced distortion is the chemical shift effect caused by
29 the different resonance frequencies of protons in fat and water which results in an apparent
30 positional shift of fatty tissues relative to water based tissue along the read-out gradient
31 direction. Chemical shift effects are more noticeable in the abdomen and pelvis because of the
32 presence of extensive fatty tissues. They have also been documented in head and neck regions

PartG.doc Page 31 of 147 13 January, 2004


Part G Treatment Planning

1 (Tien et al. 1991, Sakurai et al. 1992) but may not be as obvious for MR images of the central
2 nervous system due to the relatively limited presence of fat.

3 The functional dependence of susceptibility and chemical shift effects is similar to that
4 for main field inhomogeneities in that the resultant distortions are inversely proportional to
5 gradient strength. Phantom studies of susceptibility effects at Royal Marsden Hospital have
6 demonstrated shifts of the fiducial markers in the read-out gradient by 2-3 mm with respect to
7 the phantom (Finnigan et al. 1997). When the same markers were imaged on the pelvic skin
8 surface of patients, the apparent shift had increased to 5 mm. This additional shift was attributed
9 to the contribution of chemical shift from subcutaneous fatty tissues and correlated well with the
10 expected chemical shift of 3.2 ppm for fat.

11 G3.4.3.3. MR correction of image distortion

12 The effect of MR distortions needs to be considered prior to utilising MR images for


13 treatment planning. At the Royal Marsden Hospital, a protocol has been developed to correct
14 for MR-related distortions (Finnigan et al. 1997) Tanner et al 2000. Furthermore, this protocol
15 incorporates measures to maintain the quality, accuracy, and safety of the imaging and
16 correction procedures. System-related distortions may be corrected using dedicated MR
17 phantoms of known geometry (Tanner et al 2000). Royal Marsden Hospital studies using these
18 phantoms have provided comprehensive maps of system distortions in three orthogonal planes
19 and by applying these measurements, observed distortions of up to 9 mm may be reduced to 1
20 mm or less (Finnigan et al. 1997). However, object- or patient-induced distortions cannot be
21 corrected simply by phantom measurements. At Royal Marsden Hospital, a procedure based on
22 gradient reversals has been shown to be effective in correcting for object induced distortions
23 which include effects due to chemical shift (Finnigan et al. 1997). This correction procedure
24 requires acquisition of one additional image for each slice acquired and whilst this increases the
25 overall patient scanning time, such correction is needed for accurate definition of target
26 volumes, image correlation and data registration for treatment planning.

PartG.doc Page 32 of 147 13 January, 2004


Part G Treatment Planning

1 G3.5. Other considerations affecting utilisation of MR imaging for


2 treatment planning

3 G3.5.1. General considerations

4 It must be noted that MR imaging using ‘closed’ MR systems that resemble narrow
5 tunnels can be a formidable procedure for patients due to the noise and enclosed space of the
6 magnet bore. Current generation closed MR systems have now been designed with ‘flanged’
7 openings at either end of the magnet to create a better sense of ‘space’ for the patient if they can
8 be positioned towards the end of the magnet bore. Newer ‘open’ C-shaped MR imaging systems
9 may overcome the claustrophobia of patients to some extent but are not as widely available as
10 closed MR imaging systems. Alternatively patient sedation may be provided or it is possible to
11 accompany and reassure the patient during MR imaging as ionising radiation is not used.

12 A small proportion of patients will be unsuitable for MRI by virtue of:

13 • Claustrophobia

14 • Presence of a cardiac pacemaker.

15 • Previous history of metallic particle eye injury.

16 The latter two scenarios are excluded by careful screening with a questionnaire prior to patients
17 entering the room housing the scanner. If doubt does exist regarding metallic foreign bodies in
18 the eyes, plain radiographs of the orbits can help exclude the possibility of a very small ferro-
19 magnetic object within the eye.

20 The reason for not allowing patients with ferro-magnetic objects close to critical
21 structures is that these objects may be moved by strong magnetic fields. External objects such
22 as hair clips, scissors and the like must also be excluded because of the potential danger from
23 impact injuries as well as the introduction of distortions secondary to these objects within the
24 system.

25 The MR scanning table must reproduce the flat bed design of the radiotherapy simulator
26 and treatment couch. Any immobilisation device used in planning needs to be assessed for its
27 magnetic safety as well as being able to fit into the patient tunnel and not interfere with the MR
28 examination coils. Where this is not practical, e.g. for patients in stereotactic head frames, MR
29 scans cane be registered to planning CT scans.

PartG.doc Page 33 of 147 13 January, 2004


Part G Treatment Planning

1 G3.5.2. Motion

2 Any patient or organ motion will give rise to image blurring and artefacts thereby
3 potentially masking pathology. Prolonged scan times will increase the potential for both patient
4 and internal organ movement. Physiological movements such as cardiac and respiratory
5 movements have a greater effect on the imaging of the thorax, abdomen, and pelvis compared
6 to the limbs or the brain. Blood vessel motion can be significant and can result in severe
7 artefacts that propagate through the images. Blood flow and motion of abdominal fat due to
8 respiration can result in '
phase mismapping'during data acquisition. This mismapping will give
9 rise to image ghosting on image reconstruction (White et al. 1992). Peristaltic bowel
10 movements, rectal and bladder filling are more subtle but can still alter the position of the target
11 volumes within the pelvis during imaging.

12 Diagnostic methods exist to compensate for the problems of motion. However, it


13 remains to be seen if these compensatory methods such as ECG gating, blood flow
14 compensation techniques, presaturation methods, and respiratory gating methods will be
15 applicable for treatment planning. Some methods such as respiratory gating have been less
16 successful because the imaging times have substantially increased (Wood 1988). Techniques
17 such as scan averaging and breath-holding imaging, using fast “turbo” sequences, for example,
18 considerably reduce artefacts due to respiration (Mirowitz et al. 1990, Flentje et al. 1993). Anti-
19 muscarinic agents such as hyoscine butylbromide can be used to decrease gastrointestinal
20 motility. Ultra-fast MR sequences such as Magnetisation Prepared Rapid Acquisition Gradient
21 Echo (MP-RAGE), fast Spin-Echo and Echo Planar Imaging (EPI) have much shorter scanning
22 times, and may produce acceptable image quality. Echo planar images have been obtained in
23 under 100 ms, effectively '
freezing'some physiological motion (Edelman et al. 1994). The
24 possible distortion effects associated with these newer sequences need to be assessed before use.

25 G3.6. Methods to allow the use of MR imaging in treatment planning


26 The lack of electron density information and the presence of MR distortion has meant
27 that image processing methods must be employed if MR imaging is to be utilised for treatment
28 planning. Once MR distortion and radiofrequency non-uniformity effects have been quantified
29 and corrected, image segmentation and image registration or correlation techniques can be used
30 to overcome the lack of electron density data from MR images. These procedures are not

PartG.doc Page 34 of 147 13 January, 2004


Part G Treatment Planning

1 mutually exclusive as segmentation methods can be used to preprocess data for multimodality
2 image registration. These concepts will only be briefly outlined.

3 G3.6.1. Image Segmentation

4 One method of performing dose calculations on MR images is to segment the relevant


5 tissues in order to assign the appropriate attenuation factors to these tissues. Manual
6 segmentation methods are labour intensive and time consuming. Fully automatic segmentation
7 methods can be more efficient but the segmentation process needs to be accurate. Often these
8 algorithms do not have adequate intelligence to distinguish reliably between subtle differences
9 in soft tissue structures. Semi-supervised automatic methods are often needed to ensure quality
10 control for these procedures. A review of the many segmentation methods is not within the
11 scope of this chapter and further information may be obtained for instance from the review by
12 Clarke et al (1995).

13 In general, parameters such as MR pixel intensities can be used to threshold regions of


14 interest, and define other features such as the edges and texture of anatomical organs to augment
15 the segmentation process. However, there can be difficulty in determining the maximum
16 contrast difference for the different soft tissues in order to select the optimal threshold for
17 segmentation. Furthermore, non-uniformities of the radiofrequency field can vary the pixel
18 intensities of a given tissue type within different regions of the same MR image, reducing the
19 effectiveness of automatic pixel-oriented techniques. These issues may often require substantial
20 operator input to provide reliable segmentation and may also result in both inter- and intra-
21 observer variations. Attention needs to be paid to the increased noise associated with the use of
22 faster modern MR imaging sequences which may result in reduced image quality and increased
23 variable image grey scale shading that may result from radiofrequency non-uniformities and
24 patient-induced distortion effects. MR segmentation methods appear to be successful for normal
25 brain images (Clarke et al. 1993). The greater soft tissue contrast obtainable with MR imaging
26 for regions such as the abdomen and pelvis may permit more reliable segmentation compared to
27 CT (Khoo et al. 1999), as segmentation in these anatomical regions remains time-consuming
28 (Neal et al. 1994). Segmentation algorithms for MR imaging need to minimise the level of
29 experienced operator supervision and provide reliable accurate results before they can be
30 adopted as standard procedure.

PartG.doc Page 35 of 147 13 January, 2004


Part G Treatment Planning

1 G3.6.2. Image Registration

2 The registration of MR images with CT images allows the use of CT-calculated isodose
3 contours with the superior tumour definition of MR for improved treatment planning. Manual
4 registration methods have been inefficient and error-prone. Automated approaches may allow
5 more accurate registration.

6 There are a number of algorithms for image registration, but generally, they fall into one
7 of several categories:

8 • Surface matching {e.g. Levin et al. 1988}. One or more 3 D contours are outlined (e.g. the
9 external contour and specific anatomical structures that are well visualised on the 2 image
10 sets) and then brought together in 3 D space until the best mathematical match is made. The
11 acquisition of contiguous and narrow CT images ensures good spatial definition and
12 therefore optimises outcome.

13 • Point matching {e.g. Hill et al. 1991}. A number of (usually at least 5) 3 D anatomical
14 reference points are nominated on each of the 2 image sets. These points are preferably
15 easily identified, small and immobile e.g. bony landmarks. The more points nominated, the
16 more accurate the final mathematical match (and in turn anatomical match) of the images is
17 likely to be. An alternative is to use external fiducial markers applied to the patient for each
18 of the imaging sessions. Whilst it is easier to use a large number of points with this method
19 compared with internal landmarks (which can be surprisingly difficult to nominate),
20 reproducibility of marker position between different imaging sessions on different days
21 presents a challenge. It is therefore vital that the marker positions are documented with
22 photographs, diagrams and measurements of their position relative to reproducible
23 anatomical landmarks. The problems associated with external fiducial markers are
24 eliminated for brain imaging by using a stereotactic frame which is firmly and reproducibly
25 located on the patient’s head. This is usually fixed by surgical pins which firmly locate the
26 frame against the skull, but relocatable frames have been devised for treatment planning
27 {Kooy et al. 1994}. See Part H Chapter 2.

28 For a more detailed description of the various methodologies and issues, the reader is
29 referred to {Kessler et al. 1991}.

30 Image registration does have limitations:

PartG.doc Page 36 of 147 13 January, 2004


Part G Treatment Planning

1 • It is time consuming.

2 • Patient positioning can be difficult to reproduce during acquisition of different types of


3 image at different scanning sessions. For example, the degree of neck flexion is very
4 different for CT versus MR and may present difficulties in registering images of the
5 head/neck region when the body surfaces are used for matching the images to each other.

6 • There may be a problem with reproducibility of the position of different internal organs to
7 each other between imaging sessions. This is a particular problem with the prostate which
8 may shift position depending on differential filling of the bladder and rectum (see below).
9 This may compromise the position of internal reference points.

10 • External fiducial markers (if used) must be large enough to be visualised but small enough
11 to give a precise 3-D reference point (? Cross reference).

12 • An adequate number of common points must be identified to provide the registration


13 algorithm with enough spatial information to perform an accurate geometric match between
14 image sets. This can be difficult when internal anatomical reference points are used, as
15 insufficient points may be identifiable as being common to the 2 sets of imaging data.
16 Differences in slice orientation between CT and MR data-sets can make it difficult to
17 nominate consistent internal reference points.

18 Simple registration procedures are usually 2D methods and include the use of a
19 projection system with an appropriate magnification which allows MR images to be
20 superimposed over simulation films or isodose plans. Bony landmarks such as vertebrae and
21 pelvic bones are used as correlation points. Registration methods in 3D are more complex and a
22 multitude of parameters need to be considered in the matching procedure such as the
23 reproducibility of patient positioning, the angle of the imaging plane, image contrast and
24 resolution, and the volume of the image set including the width and gap distance of each slice.
25 Registration techniques may be applied either prospectively or retrospectively and include point
26 or line matching; iterative matching, principal axes and surface matching. There are many
27 different methods of image registration. A review of current techniques has been published by
28 van den Elsen et al. (1993).

29 In general, prospective methods involve the use of immobilisation devices and external
30 fiducial markers to acquired image data in a predetermined controlled setting for later
31 correlation based on these markers. This coordinate system of fiducial markers can also be used

PartG.doc Page 37 of 147 13 January, 2004


Part G Treatment Planning

1 for the assessment of distortion, verification of patient set-up and image data QA. Retrospective
2 registration techniques have the advantage of not requiring image data to be acquired in a
3 predetermined controlled manner but rely on the ability to identify corresponding features in the
4 images to be registered and then calculating the transformation which matches these features.
5 Registration accuracy is an important issue. Assuming MR distortions have been corrected,
6 brain registration methods have reported a mean correlative value of 1-3 mm for CT-MR
7 matching (Schad et al. 1987, Hill et al. 1994, Pelizzari et al. 1989, Van Herk and Kooy 1994).
8 More difficulty can be expected for imaging of the thorax, abdomen and pelvis where
9 movement and distortion effects are greater.

10 G3.7. Clinical Sites of MR imaging applications in treatment planning


11 MR imaging has been used extensively for the brain as it is the modality of choice for
12 this site. Intracranial movement is minimal, the skull can be rigidly immobilised and the head is
13 easily imaged in the centre of the magnetic field, thus MR distortions are generally smaller
14 compared to larger body regions. These factors, coupled with effective head registration
15 methods have allowed easier integration of MR/CT for treatment planning. MR imaging has
16 been shown to quantitatively improve target volume definition in 62-82% of CT-based brain
17 treatment plans (Pötter et al. 1992, Heester et al. 1993, Gademann et al. 1993) as well as provide
18 complementary information not visualised by CT imaging {Khoo et al 2000}. MR imaging has
19 been used alone for treatment planning by assuming a homogenous attenuation value inside the
20 brain with an error of less than 2% compared to CT-based dosimetry (Schad et al. 1994).
21 Registered MR/CT images have also been used for brain brachytherapy planning to simulate 3D
22 catheter/radioisotope seed arrays to confirm and/or adjust tumour dosimetry (Hardy et al 1992).
23 High resolution 3-D vascular and neuro-anatomical information for stereotactic radiotherapy of
24 cerebral arterio-venous malformations can be obtained using flow compensated gradient echo
25 and fast 3D MR sequences. This avoids the invasiveness of angiography and allows frequent
26 follow-up scans to assess the occlusion of the treated feeding blood vessels (Ehricke and Schad
27 1992).

28 MR has provided improved delineation of target volumes for radiotherapy of


29 nasopharyngeal and head/neck cancers (Curren et al. 1986, Kovacs et al.1992). Fast breath-held
30 MR sequences and coronal MR angiography techniques have provided similar improvements in
31 supradiaphragmatic lymphoma treatment planning with better localisation and avoidance of
32 partial volume imaging effects (Müller-Schimpfle et al. 1992). MR imaging has also been

PartG.doc Page 38 of 147 13 January, 2004


Part G Treatment Planning

1 shown to provide improved definition of target volumes in prostate radiotherapy, especially in


2 the localisation of the prostatic apex, compared to CT (Khoo et al. 1999, Kagawa et al. 1997).
3 For further radiotherapy applications of MR, the reader is referred to a review by Khoo et al
4 (1997).

5 G3.8. Conclusions
6 MR imaging has considerable potential for treatment planning. The superior soft tissue
7 contrast provided by MR and the ability to infinitely vary contrast by manipulation of the
8 imaging parameters facilitate optimal tumour evaluation. Together with its 3D multiplanar
9 imaging capability, MR can provide advantages over reconstructed CT images. Used in
10 conjunction with CT, MR imaging can provide both complementary and supplementary
11 information for the localisation and characterisation of tumours and surrounding normal tissues.

12 MR distortions can be geometrically corrected to provide accurate spatial information.


13 For distortion correction, it important that both patient images and distortion data are acquired
14 using the same set of imaging conditions since MR distortion is strongly influenced by the
15 parameters of the imaging sequence.

16 The appropriate MR imaging parameters for optimum tumour visualisation and


17 treatment planning need to be identified and faster MR sequences will reduce the effect of
18 motion artefacts. Non-operator dependant segmentation and image registration methods which
19 are accurate and reliable are required so that MR can be either used alone or integrated with CT
20 for dose calculations. Whilst the full potential of MR in treatment planning remains to be
21 realised, these methods will greatly enhance the utilisation and integration of MR applications
22 in treatment planning.

PartG.doc Page 39 of 147 13 January, 2004


Part G Treatment Planning

1 Chapter G4. Beam definition - Virtual simulation


2 Vibeke Nordmark Hansen

3 G4.1. Introduction
4 The Planning Target Volume (PTV) defined in G1.2 must receive a dose large enough
5 to reach the therapeutic goal. It must therefore be covered by the beams that are used for
6 treating that volume. For skin tumours the beam definition is straightforward, one beam only is
7 used and the beam shape directly reflects the tumour lesion plus a margin. Treating tumours not
8 visible from the exterior of the patient requires the use of x-rays to define the beams. The
9 simulator (see part C chapter 4) provides the geometric set-up of the accelerators together with a
10 diagnostic x-ray tube. Hence the bony anatomy is easy to identify, and beam definition can be
11 made relative to that. However, more sophisticated approaches are required to match the shape
12 of the PTV precisely. In these cases the beam definition is given by projection of the PTV and a
13 margin. Two main simulator techniques are used : anatomical set-up and computer based set-
14 up. The computer based set-up relies either on a crude definition of anatomy or on a full pre-
15 simulation CT scan of the patient in the treatment position. In the latter, the simulation is
16 performed by a computer, using the CT data as a 3D model of the patient. It is currently referred
17 to as "virtual simulation". In the following discussion these three methods of simulation will be
18 covered.

19 G4.2. Beam definition using anatomical set-up

20 G4.2.1. Direct simulation of the beam arrangement

21 In cases where the precise shape of the PTV is not considered, the beam arrangement
22 will be decided according to the diagnostic information that exists. This may be in the form of
23 X-rays (possibly with contrast), CT scans, Radioisotope imaging, Ultrasound, or MR scans, and
24 possibly biological tests to determine the stage of disease. The geometry of the beams is then
25 defined during simulation on the basis of the anatomy that is identified using the image
26 intensifier.

27 A number of points need to be considered during simulation, some of which may have
28 been decided prior to simulation, either on the basis of the specific patient, or on the basis of
29 protocols for the given tumour site :

PartG.doc Page 40 of 147 13 January, 2004


Part G Treatment Planning

1 • Beam modality

2 • Beam energy

3 • Beam direction

4 • Beam divergence, or the arrangements to avoid/ compensate for beam divergence at one
5 beam edge

6 • Beam shape / lead shielding

7 In most radiotherapy departments the available beam modalities are photons and
8 electrons. Electrons will be discussed in Chapter G6. Photons are, in most departments,
9 available with different energies ranging from 4 MV up to about 20 MV. The choice of energy
10 is dependent on patient separation and the depth of tumour. If the skin or superficial tissue less
11 than 0.5 cm thick is part of the target, bolus may be used to increase the dose to this region. The
12 preferred energy range for targets close to and involving the skin is 4-6 MV. If, however, the
13 target is deeply situated higher energies offer better dose distributions. In some cases a
14 combination of e.g. 20 MV and 6 MV will yield the best plan. For more detail on depth dose
15 curves and distributions see Chapter G5. Beam direction, shape, and blocking are site specific,
16 and will be briefly discussed for a limited number of sites.

17 Finally, at simulation all physical parameters required to calculate the dose need to be
18 acquired. These parameters are, in addition to field size, and possible lead shielding, (when the
19 shape of the beam is not rectangular), the dose prescription point, and possible additional dose
20 calculation points (the depth and patient separation at those points, and the source to skin
21 distances (SSD) for all the beams). Some of these parameters may be acquired indirectly by
22 taking a patient outline on which the isocentre of the beams is indicated. From the outline, the
23 SSD and depth to the point of interest can be measured for each beam.

24 Outlines can be obtained in a number of ways e.g. an outline may be acquired manually
25 using the SSD indicator and polar graph paper, or by using a pantograph, or with one of a
26 number of commercial optical outlining devices. At least one outline should be taken in the
27 plane of treatment through the centre of the treated area. Further outlines are recommended if
28 the patient shape changes significantly within the treatment area.

29 Optical devices generally make use of laser projections on the patient skin, the images
30 being acquired by two (or more) fixed cameras or by a single camera mounted on the simulator

PartG.doc Page 41 of 147 13 January, 2004


Part G Treatment Planning

1 gantry and operated at different gantry angles (Wilks 1993). These images are converted into
2 outlines, which may be scaled and printed or exported to a treatment planning computer.
3 Because of patient curvature, several images from different angles need to be acquired and
4 merged to obtain a surface outline. Further parallel outlines may be obtained using either
5 discrete or continuous couch movement. The resulting outlines cover the visible anterior surface
6 of the patient, hence the posterior of the patient is inaccurate and often approximated by a
7 straight line. Thus, care must be taken if beams are to enter the patient from posterior angles.
8 Measurements of the patient separation can be taken, using calipers to check points on the
9 contour. Systems for measuring surface shapes lack information on internal structures. One
10 exception is the system described by Wilks (1993) which uses an algorithm to draw lungs on
11 breast contours by using the depth of lung seen on the simulated field radiograph.

12 If one has access to a simulator equipped with a CT option (see Part C Chapter 4), the
13 acquisition of external outlines is no longer necessary. These outlines, as well as the outlines of
14 internal structures, are extracted directly from the CT images which are transferred to the
15 treatment planning console.

16 G4.2.2. Specific examples of anatomical set-up.

17 As explained above, anatomical set-ups should be restricted to cases where it is


18 considered acceptable to define the PTV on the basis of general diagnostic information, without
19 strict delineation of this volume on a series of axial slices. It must also be realised that such an
20 approach is possible only if the beam arrangement is simple enough to be able to anticipate
21 what the dose distribution will look like. Some typical examples are listed below. For a more
22 comprehensive overview of radiotherapy treatment set-ups see (Dobbs et al 1999).

23 G4.2.2.1. Parallel opposed pelvic fields. (Mainly used for palliative treatments).

24 In the case of simple open pelvic fields, the field area will be defined at simulation
25 according to bony landmarks. In some cases contrast will be used to indicate the location of the
26 bladder or rectum, or a tampon to locate the vagina.

27 G4.2.2.2. Mantle fields

28 It has been the practice to treat Hodgkins and non-Hodgkins lymphoma with large
29 anterior and posterior fields covering the whole chest and neck. Since the aim is to treat tumour
30 and lymphatic drainage while avoiding critical structures, such as lung and heart, these fields

PartG.doc Page 42 of 147 13 January, 2004


Part G Treatment Planning

1 need carefully designed shaping. Knowledge of the pre chemotherapy extent of the solid
2 tumour (generally in the form of a diagnostic CT scan of the patient) and an anatomical atlas
3 data for the position of lymph nodes will enable the shielding to be determined. The field size
4 required is usually larger than the maximum available on the linear accelerator so these
5 treatments are planned at an extended SSD (Source Surface Distance), typically between 120
6 and 140 cm. Extending the SSD has the added advantage of reducing the size and weight of the
7 lung blocks.

8 G4.2.2.3. Tangential breast beams

9 To simulate tangential breast treatments the gantry angle that covers the breast volume,
10 whilst minimising the amount of underlying lung tissue in the fields, can be found by screening
11 the patient. The opposing beam should be such that the beam edges make a non-divergent edge
12 skimming the lung. The relative gantry angles can be worked out by trigonometry, using the
13 field width.

14 For tangential breast beams it is also advantageous to keep a non divergent or vertical
15 beam edge at the superior border in order to allow for the addition of an axillary field (even if
16 that is not prescribed as primary treatment). This can be achieved either by using trigonometry
17 to compensate by adjusting the couch angle and the collimator angle of the tangential fields [4],
18 or one can use "half beam blocks" (asymmetric diaphragms) provided the field length is not
19 restricted.

20 In most situations, an anatomical set-up does not dispense with the need to confirm the
21 plan by computation of the dose distribution, using the data acquired during the simulation
22 process. Small adjustments of the plan (i.e. beam weights, beam modifiers, shielded area) might
23 be performed without resuming the simulation. However large changes (i.e. beam directions)
24 could necessitate a repeat simulation.

25 G4.3. Computer based beam definition - the "standard" approach.


26 This approach is recommended when the choice of beam arrangement is too difficult to
27 make an anatomical set-up practical. Beam characteristics will be defined during the dose
28 planning process. In these cases, the "simulation" process is split up in two parts :

29 The acquisition of anatomical data (before dose planning)

PartG.doc Page 43 of 147 13 January, 2004


Part G Treatment Planning

1 The verification simulation of the selected plan (after dose planning)

2 Most of the methods used for data acquisition have been explained previously. They
3 combine the acquisition of a limited set of external outlines, reference skin marks, and
4 information on internal structures obtained from orthogonal x-rays (with the simulator), from
5 other diagnostic images (CT, MR, etc.) or from the simulator CT option.

6 After transferring these data to the treatment planning system and indicating the limits
7 of the PTV, the beams are graphically defined on transverse sections in order to cover the PTV
8 and to avoid the organs at risks. Their characteristics (number, energy, incidence, position, field
9 size, etc.) are interactively edited until a satisfactory dose distribution is obtained.

10 The beam parameters are finally transferred back to the simulator in order to check the
11 plan feasibility and consistency on the patient, before progressing with treatment. At this stage,
12 the beam entry points can be marked on the patient’s skin (e.g. tattooed markers) and
13 radiological images taken to serve as a reference.

14 This "standard" approach is appropriate for a plan which makes use of a limited number
15 of cross-sections (i.e. less than 10). When a large series of thin adjacent slices is acquired, a full
16 3D reconstruction can take place and the beam definition can be made with the "virtual
17 simulation" approach.

18 G4.4. Computer based beam definition - "virtual simulation"

19 G4.4.1. Principle of virtual simulation

20 When a full CT data set is available for a patient in the radiotherapy treatment position,
21 the CT data serves as a 3D anatomical model of the patient. The CT data provides the basic
22 electron density for the dose calculation, hence allowing corrections for tissue heterogeneities.
23 In 3D planning systems (Goitein et al 1983, Mohan et al 1988, Galvin et al 1995) the CT data
24 can be viewed in transaxial (i.e. the original CT slices), sagittal and coronal views. Generally
25 the sagittal and coronal views show the coarser resolution in the slice spacing direction,
26 however, the basic anatomy is well defined.

27 The Target and Organs at Risk (OAR) should, as a minimum, be outlined on all slices
28 (see Chapter G1). The projections of these outlines are shown on the transverse planes (where
29 they were defined), and on the sagittal and coronal plans, either as contours or as colour wash.

PartG.doc Page 44 of 147 13 January, 2004


Part G Treatment Planning

1 In many 3D planning systems the user can toggle any structures on or off, which can be very
2 helpful, as overlaying structures obscure each other, and the underlying greyscale CT data. In
3 addition to the 3 orthogonal 2D views it is helpful to have a 3D surface rendered view of the
4 structures outlined. In particular it is useful to view the 3D volumes from different angles to
5 decide on the beam directions.

6 The planning process for computer based set-up is taken away from the simulator and
7 performed at the planning computer or at the virtual simulator. At present, however, it is still
8 common, as for the "standard" approach, to send the patient to the simulator for a “plan check”,
9 i.e. to get reference images of the planned fields and possibly to mark the patient up for
10 treatment, and hence limit the set-up time for the first treatment. In some departments, the
11 simulator is interfaced to the record and verify system so that the set-up parameters (except the
12 Monitor Units) are transferred automatically to the accelerator, thereby saving time and
13 lessening the risk of errors incurred by manual data entry. Where there is couch angulation, the
14 plan check also checks for collision risks, as that is not usually incorporated into planning
15 systems. Some groups have developed graphical simulation whereby collision risks between
16 the gantry and couch are predicted (Kessler et al 1995, Muthuswamy 1999). However, the
17 graphical simulator does not fully allow for the risk of collision between patient and gantry.
18 Assessing the collision risk is not relevant for coplanar beams.

19 G4.4.2. Beam axis direction

20 Except where protocols prescribe the beam angles (e.g. the treatment of prostate
21 tumours with 4 or 6 coplanar fields), choosing the incident direction for the beams is generally
22 done manually by the treatment planner. The planner needs to find the beam direction that treats
23 the PTV while best avoiding irradiation of the OAR. This may be done by rotating 3D views or
24 as described below:

25 One treatment planning system manufacturer has developed a useful tool to enable the
26 planner to predict combinations of gantry and couch angles that will best avoid the OAR. This
27 is illustrated in Figure G 4.1(a). This “spherical view” shows a flattened globe, with the
28 geometric centre of the PTV, also the treatment isocentre, at its centre and the OAR projected
29 onto its surface. The colours of the OAR are as in Figure G 4.1(b) and the grey shade indicates
30 forbidden couch and gantry combinations. Beams with couch and gantry angles that are in the
31 clear regions will not be passing through an OAR in order to reach the PTV. The gantry angle

PartG.doc Page 45 of 147 13 January, 2004


Part G Treatment Planning

1 is the meridian, with a range of values from 0° at the "north pole" to 180° at the "south pole" in
2 the central portion of the sphere, and from 180° at the "south pole" to 360° at the "north pole" in
3 the outer portions. The couch angle is the latitude and ranges from 0° both at the centre and the
4 periphery to ± 90° at the dividing lines. The projected organs in the spherical view are
5 continuous but the couch and gantry combinations are discontinuous at the dividing lines.

6 Figure G 4.1(b) is called a Beam’s Eye View (BEV). It is generated by taking a plane
7 perpendicular to the beam direction and showing the projected PTV and OAR relative to the
8 beam edges (the yellow lines). The main restriction of the spherical view is that it only shows
9 the projection of the central axis and this shortcoming is illustrated in Figure G 4.1(b) where
10 part of the beam passes through the bladder even though this is not apparent from the spherical
11 view.

12
13 Figure G 4.1 (a) A spherical view produced from Voxelplan. The gantry angle is depicted as
14 meridians and the couch angle as latitudes. (b) The 3D BEV of the same patient showing
15 PTV (yellow), bladder (green), right (blue) and left (red) femoral head. This view is from 0°
16 gantry and 0° couch. The yellow cross in (a) indicates the couch and gantry angle for this
17 BEV.

18 Software exists which semi-automatically optimises the beam direction. However, as


19 has been discussed in the literature such optimisation will always be limited (Rowbottom et al
20 1999a, 1999b).

21 G4.4.3. Field shaping

22 Having determined the incident beam axes, a Beam’s Eye View (BEV) can be generated
23 showing the projected extent of the PTV and the OAR in a plane perpendicular to the beam
24 axis. The field shape is then generated by adding a margin to allow for the penumbra of the
25 beam. The ideal width of this margin depends on the field size and depth and whether the fields
26 are coplanar or not. In a coplanar situation the margin will need to be 2-3 mm greater in the
27 directions in which all the field edges coincide. The shielding can either be a custom-made

PartG.doc Page 46 of 147 13 January, 2004


Part G Treatment Planning

1 low-melting-point alloy block or can be achieved with a multi-leaf collimator (see next
2 Section). A custom made block conforms to the field shape and is made either by sending the
3 shape from the treatment planning system to a block cutter, or having the blocks made manually
4 using a printed template.

5 G4.4.4. Fitting MLC leaves in Beam’s Eye Views

6 Multileaf collimators do not produce smoothly rounded isodose distributions as


7 discussed in Part E Section E3.5.4 where strategies for choosing leaf positions are discussed.
8 Because the geometric projection of the MLC leaves defines approximately the 50% isodose a
9 margin on the PTV must first be applied and it is this desired “field” shape that is fitted by the
10 MLC.

11 To fit to the field shape several fitting algorithms exist. Generally treatment planning
12 systems support fitting leaves to outside, inside, or to middle of the MLC leaf (see (a)
13 (b)

14 Figure G 4.2a and Part E Section E3.5.4). In practice the corners of the MLC leaves
15 may be considered to be cut off as shown in (a) (b)

16 Figure G 4.2b (Heisig et al. 1994) and this concept provides a good fit to the 50%
17 isodose line.

18 In practice the choice of which model to use for fitting the MLC’s will be restricted by
19 what is available on the planning computer, and, for the sake of consistency, there should be a
20 departmental policy on the margin size and fitting mode for different sites of treatment. How
21 well the MLC conforms to the target shape will also depend on the collimator angle. Most
22 planning systems provide the planner with a graph of either the total volume or the normal
23 tissue volume irradiated with collimator angle. This enables the planner to choose a collimator
24 angle that results in the best fit of the MLC’s to the PTV. If, however, a wedge is used, the
25 collimator angle and hence the MLC orientation is determined by the wedge direction. In such
26 cases the MLC fit may be improved by shifting the isocentre for that beam (Fernandez et al
27 1995).

28

PartG.doc Page 47 of 147 13 January, 2004


Part G Treatment Planning

1 (a) (b)

2 Figure G 4.2 MLC fitting to the field shape. Here a schematic corner of an elliptical PTV
3 are shown. Solid line: MLC fitting outside the field shape; Grey shaded: MLC fitting to
4 middle of leaf; Dashed lines: fitting to shield all outside of field shape. In (b) the fitting
5 method of Heisig et al (1994) is shown.

6 The standard MLC leaf width is 10 mm at the isocentre. However, recently standard or
7 micro MLCs have been designed with smaller leaf widths (see Part C Section C2.4.3). The
8 micro MLCs were originally designed for treatment of brain tumours, but are also shown to
9 give a closer fit for prostate treatments [Kubo et al, 1999] significantly sparing more bladder
10 and rectum.

11 G4.4.5. Digitally Reconstructed Radiographs (DRR)

12 Just as anatomical simulation uses flouroscopic or conventional radiographs, so Virtual


13 Simulation uses Digitally Reconstructed Radiographs (DRR’s) to enable treatment fields to be
14 visualised in relation to the patient’s anatomy. The advantage of a DRR produced from a
15 therapy CT data set with the patient in the treatment position is that the data used for target
16 localisation is derived from the same data set as that used for verification of the treatment beam
17 direction. This means that the DRR can be used as the reference image, and hence eliminate the
18 simulation process. In addition the PTV outlines can be directly projected onto the DRR.
19 However, in order to be useful tools in the clinic DRR’s must be of comparable quality to
20 conventional radiographs. An understanding of how DRR’s are produced will enable the
21 physicist to ensure that best use is made of the software used to create and enhance DRR’s.

PartG.doc Page 48 of 147 13 January, 2004


Part G Treatment Planning

1 DRR’s are generated by summing the radiological thicknesses along each ray line from
2 the virtual source of the beam through the 3-D patient model to a point in the plane of the DRR .
3 The geometry used to make the DRRs and the mathematical transformations required are
4 described in Siddon, 1981 and 1985. The attenuation along the ray lines can be derived from
5 the definition of the Hounsfield Unit, HU, as follows (Killoran et al, 2001)

µ − µw
6 HU = 1000
µw

HU
7 µ = µw ( − µw )
1000

8 where µ and µw are the linear attenuation coefficients of the tissue and water respectively.

9 The linear attenuation along a ray line will be the sum of the attenuation in the voxels in
10 the path of the ray line. It is important to note that CT scans are acquired at about twice the
11 energy used for conventional radiographs so that the ratio of the photoelectric effect to the
12 Compton effect is much reduced. Consequently bone contrast is inherently much lower in
13 DRR’s.

14 Using the relationship of linear attenuation coefficient with energy it is possible to


15 generate DRR’s that resemble simulator images (photoelectric), megavoltage images
16 (Compton) or are simply radiological thickness maps. (For more detail see Sherouse et al 1990,
17 Cheng et al 1987, and Cullip et al 1993.)

18 For reference images the Photoelectric effect is most often used, as that gives the
19 highest contrast and best visualisation of bony landmarks. It is debatable whether Compton
20 images are of any use as treatment verification becomes a comparison of two low contrast
21 images (Killoran et al 2001). Radiological thickness maps contain the information required to
22 make dose compensators (see G5.2.3 ) in situations where fields are opposed (eg tangential
23 breast treatments)

24 DRR quality is influenced by a number of parameters, both to do with the data


25 acquisition, and with the reconstruction algorithms:

26 • The CT scan width determines the resolution in the longitudinal plane and provides an
27 objective limit to DRR quality. Slice widths of 10mm will result in poor DRR’s. Images
28 that are suitable for verification in the pelvic region can be produced with 5mm slices while

PartG.doc Page 49 of 147 13 January, 2004


Part G Treatment Planning

1 2-3mm is recommended in the chest and head and neck regions (Dong et al, 1995, Gilhuijs
2 et al 1996). It is also important that the CT data set extends to the whole treatment area,
3 which can be some distance beyond the tumour if non co-planar beam arrangements are
4 used.

5 • The ray tracing mode, which refers to the way the summation is performed, either uses the
6 nearest CT pixel only, for a fast reconstruction, or interpolates between the CT pixels of
7 adjacent slices, or uses a full interpolation of all 8 nearest pixels. Figure G 4.3a and Figure
8 G 4.3b compares nearest pixel with interpolation.

9 • The step size, i.e., the size of the increment when summing up the DRR, that is used.

10 • The number of pixels that there are in the output DRR.

11 • The magnification of the DRR: DRR’s will need to be rescaled to be the same size as the
12 simulator verification image or the portal image. However this increases the pixel size and
13 therefore reduces the spatial resolution.

14 • Manipulating the relative contrast of the structures in the DRR. It is possible to enhance
15 DRR’s by selectively altering the visibility of different structures according to their relative
16 densities. Several planning computers incorporate this feature which in principle enables
17 airways and bone to be simultaneously enhanced, something that is impossible with
18 conventional radiographs.

19 If the PTV and other structures have been outlined on the CT data set they can be
20 overlaid on the DRR. In this way features, such as the cribiform plate, or optic chiasm are
21 identified. The beam shape and conformal blocks or MLC’s may also be added as overlays, as
22 shown in Figure G 4.3c and Figure G 4.3d.

23 To summarise, using DRR’s as the reference images for treatment verification


24 eliminates the need to use the simulator to verify a computer plan and therefore reduces one
25 source of error. It is important that the patient has the same position at treatment as at CT
26 scanning.

27 DRR’s cannot match the spatial resolution of conventional radiographs but it is possible
28 to enhance and manipulate the digital images to ensure that they are useful to the clinician. It is
29 therefore important that DRR’s are produced by skilled and experienced staff who have had the
30 time to test the enhancing tools available and liaise with the clincians who will use the images.

PartG.doc Page 50 of 147 13 January, 2004


Part G Treatment Planning

2 (a) (b)
3

4 (c) (d)
5

7 Figure G 4.3 Examples of a DRR for a pelvic field (a and b) and a brain treatment (c and d).
8 The DRR in (a) is a thickness map generated from 10mm thick CT slices using “nearest
9 pixel” interpolation mode which yields stepped bony structures. (b) is the same as (a) but
10 calculated as a photoelectric image with interpolation between slices. Different overlays can
11 be seen eg the beam axis (cross in a), projections of the PTV contours (black dotted lines in
12 a,b,c) or the shape of the field as defined from main collimator jaws, customized blocks (in b
13 and d) or MLC (in a).

PartG.doc Page 51 of 147 13 January, 2004


Part G Treatment Planning

2 Chapter G5. Photon-Beam Treatment Planning


3 Techniques
4 Peter Childs and Christine Lord

5 G5.1. Introduction
6 The goal of radiotherapy treatment planning is to design a beam configuration which
7 will deliver a homogeneous dose to the specified planning target volume (PTV) ensuring that
8 normal tissue receives a minimum dose and critical organs receive less than their tolerance
9 doses. ICRU 50 recommends an acceptable homogeneous dose of +7%/-5% of the prescribed
10 dose, the prescribed dose being to the 100% isodose which may be the centre of the volume, the
11 isocentre, mid-plane or some other suitable point (ICRU 1993). This is achieved by choice of
12 treatment modality, typically photons or electrons, beam energy, beam arrangement, use of
13 wedges or compensators and blocks if necessary, and methods such as conformal planning
14 where appropriate. These objectives are not always met and sometimes compromise may be
15 necessary depending on the aims of radiotherapy. For example, if the treatment is palliative,
16 long term radiation effects may not be a consideration and a simple beam arrangement will
17 suffice. On the other hand, if the treatment is radical, i.e. curative, long term effects need to be
18 taken into account and this may result in a more complex radiotherapy regime. Centres may
19 also be limited in the choice of photon beam energy or the availability or otherwise of electron
20 therapy.

21 Single beams are easily visualised with the use of isodose curves, sets of which may be
22 produced for a range of field sizes for each available treatment energy. This is particularly the
23 case for electron and kilovoltage treatments. In the past, when beams were combined, isodose
24 curves were amalgamated manually i.e. hand planning. This was a slow process and placed
25 limits on the optimisation of treatment plans. Modern treatment planning computers provide a
26 fast way of viewing dose distributions. The effect of adding beams, their positioning, changing
27 beam weightings and adding wedges can be seen almost instantaneously, thus facilitating the
28 planning process and allowing the planner to investigate various options before deciding on the
29 final plan. The ability to view dose distributions in three dimensions also adds to the armoury of
30 the radiotherapy treatment planner as does the use of dose volume histograms.

PartG.doc Page 52 of 147 13 January, 2004


Part G Treatment Planning

1 G5.2. Single Beam Treatments

2 G5.2.1. Electron Beams

3 Electron beams provide a well defined region of high dose with some skin sparing and a
4 steep dose fall off. Most electron treatments are with single beams and they will be dealt with in
5 the next chapter.

6 G5.2.2. Kilovoltage Beams

7 Treatment with kilovoltage beams are also normally single fields, used for superficial
8 disease such as skin lesions. With these energies there is a large dose fall off with depth and the
9 maximum dose is at or very near the surface. Typical depth doses at 1 cm depth are 80% for
10 120 kV, 90 % for 150 kV and 95% for 300 kV. There is differential absorption of dose in bone
11 due to the photoelectric effect and lesions over bone or calcified cartilage should therefore be
12 treated with electrons rather than with kilovoltage beams. (The exception is in the region of the
13 eye because of the need for a corneal shield (Amdur et al 1992)).

14 G5.2.3. Megavoltage Beams

15 Single, high energy photon fields provide skin sparing due to the build-up effect and
16 increased dose with depth compared to kilovoltage beams. Both the build-up effect and the
17 depth dose increase with energy. Energies available range from 60Co sources to 50 MV linear
18 accelerator beams, although more typical megavoltage energies would be 4 to 25 MV. See
19 Table G 5.1 for typical beam parameters. Combinations of high and low energy photon beams
20 can produce an equivalent intermediate energy if required and not available.

21 The dose gradient across the PTV for a single photon beam is large, the fall off for a
22 6MV beam being approximately 4 % per cm. Dose homogeneity will probably fall short of the
23 recommendations in ICRU 50.

24 Normalisation to the required treatment depth will give the prescribed dose at this point,
25 but will result in a high dose at the point of dose maximum, dmax, and any normal tissue in the
26 beam overlaying the PTV. The use of a higher energy will improve the distribution, see Figure
27 G 5.1 (a) and (b).

28

PartG.doc Page 53 of 147 13 January, 2004


Part G Treatment Planning

1 Table G 5.1 Typical Beam parameters for a 10 x 10 field at 100cm SSD. From BJR
2 Supplement 25 (BJR 1996)
3
Nominal dmax % Depth dose % Depth dose
Energy (MV) (cm) at 5cm at 10cm
Cobalt 0.5 80.4 58.7
4 1.0 83.9 63.0
6 1.5 86.9 67.5
8 2.0 89.6 71.0
10 2.3 91.4 73.0
15 2.9 94.5 77.0
25 3.8 98.5 83.0
4

5 Single fields are therefore not suitable for treating deep seated volumes.

6 For volumes near the surface, bolus can be added to ensure that none of the PTV lies
7 within the build up region.

8 The field size normally defines the 50% isodose relative to the central axis, therefore a
9 margin must be added to the PTV when designing a beam to fit a target volume so that it does
10 not lie within the beam’s penumbral region. This margin will increase with energy, but a typical
11 value is 6 mm at the isocentre. In practice, the actual amount added to the field will depend on
12 beam divergence relative to the shape of the PTV.

13 Single fields provide a simple mode of treatment which may suit the clinical needs of
14 the patient.

15 Single fields are used to treat breast nodes (in combination with tangential fields to the
16 breast, see below) and the lower neck and supraclavicular nodes (in combination with an
17 opposed pair to the upper cervical nodes).

18

PartG.doc Page 54 of 147 13 January, 2004


Part G Treatment Planning

2 Figure G 5.1 Dose distributions from single photon beams (a) 6MV photon beam, (b) 25MV
3 beam (c) 45 ° wedged beam. 100% is at dmax, 10cm depth is indicated

4 One example of the use of single beams is that of a long narrow posterior field in the
5 treatment of the spine, with the anterior of the spinal cord at approximately 5 cm deep. The
6 variation in dose to the cord, due to the change in source surface distance, (SSD), and depth,
7 along the treatment field, can be greater than 10%. It is possible to obtain a uniform dose with
8 the use of purpose built compensators, shaped blocks of aluminium or other suitable material,
9 placed in the beam accessory tray, below the treatment head. Alternatively, this compensation
10 of the single field may be achieved with smaller “top up” fields overlaid on the single field,
11 using asymmetric collimator settings.

PartG.doc Page 55 of 147 13 January, 2004


Part G Treatment Planning

1 G5.3. Two Beam Arrangements

2 G5.3.1. Parallel Pair

3 For deeper volumes, two opposed beams may be used, a “parallel opposed pair”. This
4 will produce a more uniform dose in the PTV but will also give a high dose to overlying normal
5 tissue in the path of the beam, the region of highest dose being near the surface. The shape of
6 the isodoses in a parallel pair set up is that of an hour glass, see Figure G 5.2a. This can be seen
7 in both the transverse or axial and sagittal planes. Parallel pair treatments are used widely in all
8 areas of the body, and may be for the entire treatment course, especially when the intent is
9 mainly palliative, or as the first phase of a two phase radical treatment. As the separation of the
10 patient increases, the uniformity of the dose distribution decreases with an increase in dose to
11 normal tissue, or an unacceptable dose inhomogeneity within the PTV . This can be offset if a
12 higher energy is chosen for treatment. For a prescription of 100% at mid plane depth, the
13 maximum dose should ideally not exceed 107%.

14

15 Figure G 5.2 6MV parallel pair distributions normalised at mid-plane showing the effect of
16 beam weighting, separation is 20cm (a) equal weighting, (b) 2:1 weighting

PartG.doc Page 56 of 147 13 January, 2004


Part G Treatment Planning

1 G5.3.2. Beam Weighting

2 If the PTV is not at the centre of the patients’ cross-section, i.e. in the mid-plane of the
3 patient, then the opposed beams can be given different weightings (see Figure G 5.2b). Beam
4 weighting depends on the planning system or calculation technique used. It can be proportional
5 to the “beam on” time (e.g. set Monitor Units for linear accelerators or time for Cobalt units), or
6 proportional to the contribution of that beam to a specific point, usually the isocentre. For a
7 parallel pair with the isocentre at mid-plane the two weight point definitions will be the same (in
8 the absence of inhomogeneities), so a 2:1 weighting will give the ratio of beam on time and of
9 contributions to the isocentre. However, a different weight point will show up differences
10 between the two weighting systems and an equal “beam on” time will not mean an equal
11 contribution to the isocentre. The isocentre dose from each beam will depend on SSD, isocentre
12 depth below surface and heterogeneities in the beam. Increasing the weighting of the beam
13 entering the side of the patient nearest the PTV will increase the PTV dose relative to normal
14 tissue on the opposing side. Weighting can also be used to reduce the dose to sensitive
15 structures located either in the entrance path of the beam, or the exit path beyond the PTV.

16 G5.3.3. Wedge Filters

17 Where opposed beams are applied to sloping surfaces, for example in larynx or breast
18 treatments, it is necessary to use wedge filters to maintain a uniform dose distribution
19 perpendicular to the beam direction (see Figure G 5.4a). The effect of the wedge is to tilt the
20 isodoses and in these situations it compensates for “missing tissue”. Details of wedge filter
21 designs can be found in Part C Section C2.4.4.1 and of practical issues relating to them in Part E
22 Sections E3.4.1 and E3.6.1.

23 Wedges are normally utilised in the axial plane, in the anterior-posterior or right-left
24 direction, but superior-inferior wedges can also be employed to compensate for the slope of the
25 patient contour in the sagittal plane. Wedges aligned in the superior-inferior direction may also
26 improve the dose distribution if the PTV is inclined relative to the patient outline in the sagittal
27 plane.

28 G5.3.4. Tangential Beams

29 Radiotherapy of the breast and chest wall (post mastectomy) is usually carried out using
30 two beams. Tangential fields are used which may be set up so that the posterior border of each

PartG.doc Page 57 of 147 13 January, 2004


Part G Treatment Planning

1 field matches. This ensures that the beams do not diverge into the lung which would result in an
2 increase in lung dose. Wedges are added, again to achieve a more uniform dose distribution,
3 compensating for the differential tissue thickness across the beam width, see Figure G 5.3. The
4 amount of lung in the beam, a major heterogeneity, needs also to be considered. Additionally,
5 not all treatment planning systems account for the lack of scatter in air, a further consideration
6 when beams enter the patient obliquely. The choice of suitable beam weightings to account for
7 asymmetry of the breast or positioning of the isocentre may be needed to optimise the dose
8 distribution.

10 Figure G 5.3 6MV tangential breast fields, normalised to 100% at the isocentre

11 Breast radiotherapy may involve node irradiation using an anterior field with a possible
12 posterior boost field. Suitable collimator angles and couch twists can be set up on the tangential
13 beams so that a vertical match plane is produced for the node fields. Alternatively, asummetric
14 fields can be matched, again to achieve a vertical match plane with no divergence into adjacent
15 fields.

16 G5.3.5. Wedged Pair beams

17 Wedged pair beam arrangements are common in head and neck radiotherapy e.g.
18 parotid, oral cavity and maxillary antrum, the angle between the beams, often referred to as the
19 “hinge angle”, being less than 180°. In this case the wedge is used to compensate not for
20 “missing tissue” but also because the beam entry points are close together on one side of the

PartG.doc Page 58 of 147 13 January, 2004


Part G Treatment Planning

1 patient. The smaller the hinge angle, the larger the required wedge angles needed to obtain a
2 uniform dose across the PTV. An example is shown in Figure G 5.4b. This is a good beam
3 arrangement for volumes near the surface as it minimises dose to normal tissue.

5 Figure G 5.4 6MV dose distributions from two wedged fields

6 G5.3.6. Photon and Electron beams

7 Photon and electron beams can be overlaid to treat the same site. The result of this
8 arrangement is an increase in skin sparing over an electron only treatment, and a lower dose at
9 depth than that obtained from a photon only treatment. Examples include parotid treatments,
10 reducing the dose to the contralateral parotid and spinal cord and the intra mammary chain in
11 the mediastinum.

12 G5.4. Multiple Coplanar Beams


13 The use of three or more beams reduces the dose level to normal tissue within the
14 irradiated volume compared with a parallel pair arrangement, as the entrance and exit doses do
15 not necessarily coincide. This is important when the beam traverses a sensitive structure such as
16 the bladder or rectum or a dose limiting structure such as the spinal cord or lung. Beams
17 converge on the PTV and the use of suitable weighting and wedges produces high uniform dose
18 within the PTV. It is possible to choose beam weightings and wedges so that the dose to normal
19 tissue in the path of the beam is spread equally. This may not necessarily be desirable when the

PartG.doc Page 59 of 147 13 January, 2004


Part G Treatment Planning

1 beams are traversing sensitive structures and beam parameters can be manipulated so that the
2 dose to these structures is reduced. This will of course increase the dose to normal tissue in the
3 path of the other beams. Therefore selection of the best field arrangement for a specific site can
4 depend upon many factors with varying degrees of clinical importance (Bedford et al 1999).
5 Beam angles can also be chosen to reduce the dose to critical structures such as the spinal cord.

6 G5.4.1. Three Beam Arrangements

7 Three field treatments are used in the pelvis (prostate, bladder), abdomen and thorax as
8 well as the brain.

9 The beam arrangement depends on treatment site but a typical configuration in the
10 pelvic area will be an anterior or posterior beam plus two lateral or lateral oblique fields. See
11 Figure G 5.5. In the pelvis (treatment of the prostate or bladder), the three field arrangement
12 ensures that no beam enters through the rectum. Wedges on the lateral or lateral oblique fields
13 ensure a uniform dose to the PTV, compensating for the uneven spacing of the beams around
14 the patient. Figure G 5.6 illustrates a three field bronchus treatment, the beams being arranged
15 to avoid the spinal cord and the contralateral lung.

16 Figure G 5.5 6MV 3 field dose distribution to treat the bladder

PartG.doc Page 60 of 147 13 January, 2004


Part G Treatment Planning

2 Figure G 5.6 6MV 3 field dose distribution to the bronchus, beams are arranged to avoid the
3 contra-lateral lung and keep dose to the spinal cord (SC) below tolerance, RL = right lung,
4 LL = left lung

5 Oesophagus treatments may be delivered in two phase treatments with an anterior and
6 posterior parallel pair as the first phase followed by a three field arrangement, an anterior and
7 two posterior oblique fields, possibly with wedges. The first phase delivers minimum dose to
8 the lungs, the second phase minimum dose to the spinal cord. The number of fractions for each
9 phase is chosen so that the dose to the lungs and spinal cord is kept below tolerance.

10 It is possible to use different beam energies in multi-field plans and a lower beam
11 energy may be chosen for one beam to reduce the exit dose e.g. the anterior beam in prostate
12 treatment which exits through the rectum. It should be noted, though that a lower beam energy
13 will increase the surface and entrance doses.

PartG.doc Page 61 of 147 13 January, 2004


Part G Treatment Planning

1 G5.4.2. Four Beam Arrangements

2 Four field arrangements can be used in the pelvic area, consisting of an anterior,
3 posterior and two lateral beams, see Figure G 5.7. This is a straightforward beam arrangement
4 and wedges may again be used to compensate for “missing tissue”. Different weightings may
5 also be used.

7 Figure G 5.7 6MV four field brick arrangement in the pelvic area, treating the same patient
8 and PTV as Figure G 5.5

9 G5.4.3. Arrangements with more than four beams

10 Arrangements using more than four fields are not common. As the number of fields
11 increases, the mean dose to normal tissue within the irradiated volume decreases but the volume
12 of normal tissue that is irradiated increases. The volume of the 50% isodose takes on a star
13 shaped appearance and the dose to any critical structures adjacent to the PTV may increase.
14 Whilst there may be some advantages in using a large number of fields for specific cases,
15 overall treatment time will be greater due to increased setting up and delivery time.

16 Rotational “arc” therapy is the ultimate end point for large numbers of fields but this is
17 seldom practised in conventional radiotherapy today. The main reason is that in conventional

PartG.doc Page 62 of 147 13 January, 2004


Part G Treatment Planning

1 rotational therapy (i.e. with no dynamic conformal shielding), it is not possible to produce
2 conformal dose distributions so the technique is limited to cylindrical or spherical PTVs.

3 One example of a six field treatment is that of prostate boost treatments, one of the arms
4 of the RT01 trial in the UK (MRC), see figure 8. This is the second phase of a two phase
5 prostate treatment. The first phase is a conventional 3 field technique to the PTV with a
6 prescribed dose of 64 Gy. The second phase consists of two lateral beams with right and left
7 anterior and posterior oblique fields to treat the GTV only to 10 Gy. Both phases include
8 conformal blocking.

10 Figure G 5.8 Six field phase 2 prostate boost distribution. R = rectum, B = bone. Note the
11 lower dose to normal tissue relative to that in Figure G 5.5 and Figure G 5.7, but the
12 consequent increase in the amount of normal tissue irradiated

13 G5.5. Non coplanar Beams


14 Even in relatively simple treatments such as treatment of the larynx by a wedged pair of
15 anterior oblique beams, non-coplanar fields may be required. The couch may need to be
16 rotated typically 5° or 10° in order to avoid the beams passing through the shoulders, and thus

PartG.doc Page 63 of 147 13 January, 2004


Part G Treatment Planning

1 making them non-coplanar to one another, i.e. the central axes of the beams no longer lie on a
2 single plane.

3 If a PTV has been defined in a plane angled from the true transverse, and oblique beams
4 are required (i.e. beams not at a cardinal angle), then these will not be coplanar in the treatment
5 plane. A plan can be produced in the inclined plane, assuming it to be a true transverse plane
6 and the beam parameters corrected to be coplanar in the angled plane by adjusting the gantry
7 angle, floor rotation, and collimator rotation (Casebow 1976).

8 Although the use of multiple fields can significantly reduce dose to normal tissue,
9 invariably the exit dose of one field contributes to the entrance dose of another, opposing field.
10 More importantly, the ability to direct several treatment beams at the PTV whilst avoiding
11 critical structures may be impossible if all the beam axes are in one plane (normally the
12 transverse plane). To overcome these problems, non coplanar field arrangements can be
13 employed. Below the neck, the body can be considered as a cylinder, and therefore, any beams
14 entering out of the transverse plane will invariably result in increased path lengths to reach the
15 PTV resulting in an increased irradiated volume. Also, the physical limitations of conventional
16 treatment units prevent couch rotation/ gantry angle combinations that are required for
17 obliquely incident fields without collision with the patient or couch. i.e. the beams will fall into
18 the “collision zones” of the set up.

19 However, the head can be seen as approximately spherical and thus the increased path
20 length issue does not apply. It is also possible to direct beams from a wider range of angles
21 around the top of the head without entering the collision zones, although constraints such as the
22 patient support system, be it a shell or stereotactic frame, can restrict the choice of beam
23 arrangements. Because the critical organs can now potentially be irradiated from any direction,
24 it is essential to be able to visualise these structures in three dimensions during the planning
25 process rather than viewing single dose planes. Analysis of the dose distribution must also be
26 performed in three dimensions as simply considering a transverse, sagittal and coronal
27 doseplane through the isocentre may miss important information outside these axes. Therefore,
28 the use of tools such as dose volume histograms and 3-D graphics are important in the planning
29 process. The ideal four field arrangement, minimising beam overlap would be a tetrahedron.
30 Constraints, for example, not exiting into radiosensitive structures such as the thyroid and
31 lenses, can limit the freedom of beam orientations, while for conformally shaped beams, the
32 beam configuration may be influenced as much by the optimum shielding of adjacent critical

PartG.doc Page 64 of 147 13 January, 2004


Part G Treatment Planning

1 structures as by achieving the ideal dose distribution. It has been found that increasing the
2 number of static beams beyond four does not improve normal brain sparing (Perks et al. 1999)
3 for lesions such as the pituitary. Although arcing non-coplanar fields may have a role in
4 treating spherical brain lesions (using circular fields), where the PTV is of irregular shape, fixed
5 conformal fields offer the best solution for normal tissue sparing (Laing et al. 1993).

6 G5.6. Fixed Source Surface Distance (SSD) beams vs. Isocentric


7 treatments
8 An essential component of any treatment plan are the set up instructions that relate the
9 dose distribution and field arrangement to the external reference marks of the patient. For
10 multiple field arrangements, it is advantageous to be able to set up the treatment at the outset,
11 and therefore only have to move the gantry and collimator between fields, rather than adjusting
12 the patient position and couch movements. This is illustrated in Figure G 5.9 for a four field
13 treatment in which the isocentric arrangement has been set up to the skin reference mark using
14 the isocentre lasers and the optical distance indicator, (ODI). The treatment unit lasers are
15 centred on the isocentre and therefore intersect with the patient enabling lateral marks to be
16 used to ensure that the patient is not rotated prior to setting up the remaining fields. Although
17 the ODI would not be necessary for setting up the remaining fields, it may be useful as a
18 secondary check as the SSD for each beam can be obtained from the treatment plan. For the
19 fixed SSD set-up, the ODI and lasers must be set up to each beam entry point by moving the
20 couch laterally and vertically, and separate measurements are required to ensure that the patient
21 is not rotated. Isocentric plans are therefore easier to set up, as only the isocentre has to be set
22 once, and this patient position then applies to all fields.

PartG.doc Page 65 of 147 13 January, 2004


Part G Treatment Planning

1 (a) Isocentric set up for a “four field brick” treatment

3 (b) Fixed SSD set-up of (a).

4 Figure G 5.9 A comparison between a 4 field isocentric set-up and a 3 field fixed SSD set-up

PartG.doc Page 66 of 147 13 January, 2004


Part G Treatment Planning

1 However, maintaining a constant source to surface distance, SSD, in a multiple beam


2 treatment does simplify the dose calculation as the inverse square dependence of the beam
3 output is constant and may be looked up directly from tables. This may be preferable in “mark
4 on set” set-ups where the fields are planned at the time of treatment and the dose calculated and
5 delivered while the patient remains in the treatment room.

6 There are potential disadvantages in the use of isocentric arrangements but for the vast
7 majority of treatments this is the preferred method. For modern megavoltage machines, the
8 standard source axis distance, SAD, is 100 cm and therefore the change in distance (which is
9 unlikely to exceed 20 cm), will not have a great effect on the depth dose characteristics
10 (although ideally the treatment planning computer should be able to correct for this). The
11 proximity of lead trays, and even the treatment head, to the patient can be a significant problem
12 in physically setting up isocentric treatments deep in the pelvis of large patients, and even if
13 possible to set–up, the surface dose may be increased due to electron contamination from the
14 close head or lead tray.

15 G5.7. Extended SSD


16 It may not be possible to cover larger PTVs with a field at the standard SSD and an
17 extended SSD is therefore required to increase the field size available on a particular treatment
18 machine. Examples of long volumes include the spine in treatments of the whole Central
19 Nervous System, and sarcomas, some of which extend from the pelvis to the knee and may be
20 in excess of 50 cm. Increasing the SSD increases the field size geometrically, with a dose
21 output that falls by inverse square dependence. For example, a 40 cm long field will become 52
22 cm long at 130 cm SSD. However, the depth dose characteristics may be significantly different
23 at greatly extended SSDs (i.e. greater than 130 cm SSD) and the beam data acquired at the
24 standard SSD should be corrected, or data measured at, or close to, the proposed treatment SSD.
25 The option to treat at extended SSD depends upon the flexibility of the treatment unit itself, in
26 that a large range of couch movements are required. Raising and lowering the couch will
27 facilitate increased field sizes in the anterior-posterior direction, but the range of lateral couch
28 movements available will usually be less. As patients are generally treated lying either supine
29 or prone (to obtain a reproducible set-up) extended SSD fields will therefore be greater for
30 fields principally in the AP plane. Extended SSD fields can be set at oblique gantry angles,
31 which by combining both couch lateral and vertical movements, provides the greatest SSD
32 available for a given treatment unit couch. When oblique fields at extended SSD are employed,

PartG.doc Page 67 of 147 13 January, 2004


Part G Treatment Planning

1 the machine isocentre is located considerably distant from the patient surface and set-up
2 instructions are required to move the couch both vertically and laterally from a skin reference
3 point. Figure G 5.10 shows a beam configuration to treat a sarcoma in the thigh. The extent of
4 the lesion in the transverse plane shown means that only a small channel of normal tissue is
5 spared irradiation. The preservation of this channel is essential in preventing amputation of the
6 lower limb, so the beams are angled to present a non diverging edge to the dose distribution and
7 therefore a sharp dose gradient from the PTV to the normal tissue. However, this means that
8 the accuracy of the set-up is critical and demonstrates the importance of accurate and
9 reproducible lateral and vertical couch movements and display scales over a range perhaps
10 exceeding 80 cm.

11
12 Figure G 5.10 An extended SSD treatment of the thigh, showing the couch lateral and
13 vertical movements required to set up each field individually from a skin reference point.

14 G5.8. Matching fields


15 The need to match adjacent fields can arise when treating a volume greater in length
16 than can be covered by a single field even at extended SSD. However, the most likely, need for

PartG.doc Page 68 of 147 13 January, 2004


Part G Treatment Planning

1 beam matching is when treating an area abutting a previously treated volume and usually this
2 will require a compromise in producing an ideal dose distribution. It is important to establish
3 whether overdosing or under dosing normal tissue is the deciding clinical factor when
4 determining gaps between fields.

5 When it is expected that abutting fields will be required, the treatment can be planned
6 using asymmetric collimators to eliminate beam divergence at the junction and enable the two
7 fields to provide a uniform dose at all depths (Figure G 5.11b). In principal, it would be possible
8 to match diverging field edges with asymmetric fields as shown in Figure G 5.11c, but this
9 could be restricted by the limitations of the collimator travel and the maximum field size
10 available on the treatment unit. The risk with matching fields is that a small error in the set-up
11 can result in a large overlap volume where dose is contributed from both fields. Another
12 drawback in matching fields asymmetrically with no divergence at the junction is that the field
13 edge not being abutted will diverge more and consequently irradiate more normal tissue. This
14 effect is illustrated inFigure G 5.10 where the non-diverging field edge in this case is producing
15 a sharp dose fall off into the normal tissue region. However in this treatment, the excess
16 divergence arises in air so there is no penalty.

17 Figure G 5.11 Matching fields, (a) with a specified skin gap or C/A separation, (b) using
18 asymmetric collimators to eliminate divergence at the junction, or (c) using asymmetric
19 collimators to match the divergence at the junction.

PartG.doc Page 69 of 147 13 January, 2004


Part G Treatment Planning

1 Where symmetric fields are to be abutted it is necessary to calculate the separation


2 required between the field edges at the skin surface to prevent overdose/underdose of the
3 underlying tissue. Due to the divergence of the beam, exact matching can only occur at one
4 depth (see Figure G 5.11a). Therefore when calculating gaps, a depth must be specified at
5 which the fields will be matched. Frequently the fields are matched at the midplane depth.
6 Alternatively, matching may be to a critical structure e.g. the cord and the gap calculated to
7 ensure that no part of the cord receives more than 50% of dose from each field.

8 The distance between the fields may be defined as a gap at the skin surface between the
9 field edges, or by the distance between the central axes of the fields (Figure G 5.11a). The
10 depth at which the field is defined must be known i.e. how deep the isocentre is for each beam.

11 If an isocentric field is to be matched, the depth of the isocentre must be known so that
12 the field size at the skin surface can be determined. For example, a field 20 cm long with an
13 isocentre depth of 7 cm on a machine with standard SSD of 100 cm will have a length at the
14 surface of 18.6 cm.

15 In Figure G 5.12 it can be seen that the 50% isodose of an 80 cm SSD 60Cobalt beam
16 does not follow the geometric divergence of the defined field size. However, it is more relevant
17 to consider the 50% decrement line, which for the most part does follow the geometric beam
18 divergence. For example, if the depth dose at 10 cm (60%) was prescribed to, then the
19 dosimetric 50% field edge at this depth would be 30% i.e. on the geometric field edge.

PartG.doc Page 70 of 147 13 January, 2004


Part G Treatment Planning

60
2 Figure G 5.12 Co beam (80 cm SSD) showing 50% isodose and 50% decrement lines

3 NB. Although the use of asymmetric fields and collimator rotation/floor rotation to
4 improve field matching is theoretically possible, for many treatments requiring matching to
5 previously treated fields (often palliative and with modest prescribed doses), techniques of this
6 complexity are rarely justified.

7 G5.8.1. Matching fields to isocentric planned distributions

8 Unlike the example of matching two isolated fields, it is not possible to provide a perfect
9 50% match between a planned isocentric dose distribution and an abutting field due to the
10 divergence of the planned fields. The clinician must decide where the tissue may be overdosed
11 or must be underdosed before a gap can be determined. Figure 13 shows a field abutting the
12 50% level at the tumour but overdosing at the posterior field entry point.

PartG.doc Page 71 of 147 13 January, 2004


Part G Treatment Planning

1
2 Figure G 5.13 Matching a single field to a previously treated isocentric plan

3 G5.8.2. Matching electron to photon fields

4 The shape of the electron field isodoses makes exact matching impossible, although the
5 50% isodose (i.e. the field size defined) will not be as bowed as the lower value isodose levels.
6 A small gap of typically 2 mm offers the best compromise for most clinical electron energies,
7 but depending on whether overdosing or underdosing is the main clinical concern, no gap or up
8 to 5 mm gap at the skin surface may be employed. The use of surface bolus and an increased
9 electron energy may offer a better solution as the high dose region can then be outside the
10 patient. As with photon beam matching, concerns arising over the beam junctions may be
11 alleviated by considering that modest doses may be involved, for example when a 20 Gy neck
12 node electron field contributes to a total dose of 60 Gy.
13

PartG.doc Page 72 of 147 13 January, 2004


Part G Treatment Planning

1 Chapter G6. Electron-Beam Treatment Planning


2 Techniques
3 Alan Mc Kenzie and David Thwaites

4 G6.1. Introduction
5 The feature which makes electron beams very useful for a range of clinical applications
6 is the shape of the depth dose curve. This allows relatively uniform dose delivery to relatively
7 well-defined regions from the surface to the therapeutic range, with sparing of underlying tissue
8 due to the rapid reduction of dose towards the practical range. At about 20 MeV and above, the
9 gradient of the depth dose curve begins to lose the characteristic lower-energy steepness, and
10 the separation of the high-dose and spared regions becomes increasingly blurred. Energies
11 higher than this can be used for deep-seated tumours, but the techniques begin to approach
12 those used for megavoltage photons. However the majority of linear accelerators which provide
13 clinical electron beams produce energies within the range 4 to 20 MeV and this section
14 concentrates primarily on their use. Electron beams of these energies deliver high doses
15 (greater than about 90% depth dose) to depths from around 1 to 6 cm. Common applications
16 include skin and lip cancers, chest wall and peripheral lymphatic areas in breast cancer,
17 additional boost doses to limited volumes such as scar areas and nodes, various head and neck
18 cancers and other sites lying within these depths of the surface. Typically, the most commonly
19 used beams are those of medium energy.

20 The basic physics of electron beams used in clinical practice has been described earlier,
21 in Part E Chapter E5. This section discusses how these principles may be applied to techniques
22 and problem solving in electron treatment planning. In the following discussion, the term
23 "treatment planning" is taken to include treatment using single fields; indeed, such treatments
24 constitute the bulk of electron therapy in clinical practice.

25 When presented with an electron treatment planning problem, it is helpful to use a


26 check-list as an aid, and the following list contains items which should be considered.

27 G6.2. Check-list for electron treatment planning


28 Special techniques - does the problem warrant special techniques such as electron arcing?
29

PartG.doc Page 73 of 147 13 January, 2004


Part G Treatment Planning

1 Treatment SSD - does any part of the patient'


s anatomy, or do the available applicator
2 sizes, prevent the treatment being given at the standard treatment distance (nominal
3 SSD) on the central axis? This nominal SSD is generally 100 cm, which implies a gap
4 of 50 mm between the applicator and the skin surface for many models of linear
5 accelerator. (Note: this SSD is a nominal value dictated by the geometry of the machine,
6 the effective SSD will be variable with energy and field size as discussed in section 5.2)

7 Obliquity - If electrons are to enter any part of the patient with significant obliquity, then
8 additional considerations apply as described in Example D.

9 Surface irregularities - If these are present in the area to be treated, side-scatter bolus may be
10 required or bolus to smooth the entrance surface to be flat and perpendicular to the
11 beam axis,

12 Choice of energy - This is determined by the amount of surface bolus and the required depth of
13 the treatment isodose beneath the surface.

14 Field size - account needs to be taken of the decrease of dose near the field edges.

15 Calculation of monitor units - the isodose to which the treatment is prescribed must be carefully
16 specified (it should be 100%) and the monitor unit calculation must take into account
17 any field shaping or modification to the applicator.

18 Vulnerable organs - These may need to be shielded during treatment.

19 G6.3. Field size and coverage of the target volume


20 Electron applicator sizes, or insert (cut-out) sizes, if the applicator carries an insert with
21 an aperture cut in it, are generally quoted in terms of the geometrical size found by measuring
22 the extent of the light beam projected through the applicator (or insert) onto the skin at the
23 normal treatment distance. This will also be, to a close approximation, the size of the field
24 defined by the 50% isodose at the skin surface.

25 The International Commission on Radiation Units and Measurements (ICRU 1993)


26 recommends that planning target volumes should be covered by the 95% isodose or better, but
27 to insist on this same criterion in electron therapy will often mean unacceptably large margins
28 around the planning target volume. For electron therapy, the International Atomic Energy

PartG.doc Page 74 of 147 13 January, 2004


Part G Treatment Planning

1 Agency (IAEA 1987) recommends an 85% coverage, and the American Association of
2 Physicists in Medicine (AAPM 1991) recommends 90% coverage. Pragmatically, in electron
3 planning, it is common practice to aim to cover the planning target volume with the 90%
4 isodose. In some instances lower-value isodoses may be chosen, as for example, in chest wall
5 irradiation where the 80% isodose may be chosen to lie at the lung surface in order to reduce
6 penetration into lung and hence overall lung dose. If any critical structure, such as cord, lies
7 beneath the target volume, then the relative doses, both at the inner limit of the target and at this
8 structure, must be carefully considered to make an appropriate clinical choice of the optimum
9 beam energy.
10

11
12 Figure G 6.1 The flat floor of the 90% isodose curve is smaller than the field width by a
13 margin of 4E mm where E is the incident energy in MeV. In practice, a margin of 10 mm is
14 commonly used, which will treat a target volume with rounded edges (shown shaded).
15 Approximate depths of key percentage doses are also illustrated. Isodose values are shown in
16 intervals of 10%.

17 It will be seen from Figure G 6.1 that the width of the "flat floor" of the 90% isodose
18 surface is considerably less than the 50% field width at the tissue surface. The 90% isodose
19 surface generally begins to flatten out at a distance of about 2E mm inside the 50% isodose (see
20 Figure G 6.1) where E is in MeV. Hence, to guarantee that a target volume is treated uniformly
21 to the same 90% depth as the central axis, one would need to use a field size about 4E mm
22 larger than the width of the target volume. In clinical practice, this is too large a margin, and,
23 instead, the addition of 10 mm around the target volume is commonly used, which allows for

PartG.doc Page 75 of 147 13 January, 2004


Part G Treatment Planning

1 some rounding of the bottom edges of the target volume. However, the significance of the 4E
2 mm dimension is that, for applicators smaller than 4E mm, there can be no flat floor to the 90%
3 isodose surface. This means that the depth of the 90% isodose for applicators less than 4E mm
4 wide is less than that in the large-applicator limit, and great care should be exercised when
5 planning with these relatively small applicators,or where any dimensions of a shaped field
6 approach this critical value. Figure G 6.1 also illustrates the 2 , 3, 4, 5 rule as discussed in Part E
7 Section E5.2 to estimate the depths of various isodose parameters from the beam energy.

8 Rules should not be applied blindly, and it is always incumbent upon the treatment
9 planner to take account of the clinical requirements for target coverage. Attention must be
10 given to the dose outside the geometric beam edge, particularly where the beam is to be placed
11 close to a critical structure (organ at risk). In this situation, some compromise on field size and
12 positioning may be necessary.
13

14
15 Figure G 6.2 In this example of treating the parotid, a 90 mm wide field is sufficient to treat
16 a target volume measuring 80 mm across. Isodose values are shown in intervals of 10%.

PartG.doc Page 76 of 147 13 January, 2004


Part G Treatment Planning

1 G6.4. Examples of Electron Planning

2 G6.4.1. Example A - single electron field with no complications

3 As an example of the application of the check-list to a relatively simple plan, consider


4 the treatment of a parotid tumour (Figure G 6.2) which is to be treated with 20 fractions of 2.75
5 Gy in four weeks. The target volume measures 80 mm in the ant-post direction, 80 mm in the
6 superior-inferior direction and extends 49 mm beneath the skin surface. Considering the check-
7 list, no issues are raised by the need for special techniques, treatment FSD or obliquity. Other
8 items on the check-list are:

9 Surface irregularities - Since it is not the primary intention to treat the ear, it is unnecessary to
10 provide side-scatter bolus, and the ear is merely taped as flat as possible. However, a
11 wax ear-plug is required to prevent electrons tunnelling down the external auditory
12 meatus.

13 Choice of energy - The depth of the planning target volume is 49 mm, which will be reached by
14 an 18 MeV beam, for which the 90% isodose is at 55 mm deep with a 200 mm x 200
15 mm applicator (see Table G 6.1) and 53 mm for a 10 cm x 10 cm applicator. The
16 surface dose of an 18 MeV beam is 88 %, from Table G 6.1. However, it is not the
17 intention to treat to the skin surface, and the target volume does not come closer to the
18 skin surface than 4 mm, which, from Table G 6.1, is more than the 1 mm required to
19 raise the dose to 90%.

20 Field size - Using the criterion of a 10 mm margin around the planning target volume, the field
21 size should be 100 mm x 100 mm. However, because the target volume has been drawn
22 with rounded corners at depth, it may be treated with a 90 mm x 90 mm field in this
23 case.

24 Calculation of monitor units - The dose prescription point is on the central axis of the treatment
25 beam, at the depth of dmax, viz 2.2 cm in this linear accelerator. This means that a dose
26 of 2.75 Gy will be given to 100% on the central axis, 2.2 cm deep.

27 If a 90 mm x 90 mm insert is not available, one may be produced by attaching two lead


28 strips onto the edges of a 100 mm x 100 mm insert to reduce the field size appropriately. The
29 output (in Gy per monitor unit) of the newly fashioned insert should be measured, although

PartG.doc Page 77 of 147 13 January, 2004


Part G Treatment Planning

1 experience shows that where the field edges are brought in only minimally, the measured output
2 may not show significant change from that for the unmodified applicator.

3 Vulnerable organs - The ear plug mentioned above is used to protect the external auditory
4 meatus. The dose to the spinal cord is clinically acceptable.
5

6
7 Figure G 6.3 Illustration of the irradiation situation for the use of an internal shield behind
8 the cheek as described in example B

9 G6.4.2. Example B - use of internal shielding

10 A basal cell carcinoma on the cheek is to be treated with electrons using a 4 cm diameter
11 circular applicator. The full thickness of the cheek is to be treated, which measures 20mm from
12 the surface of the lesion to the inner surface of the cheek (Figure G 6.3).

13 Surface irregularities - The lesion itself is raised from the normal skin surrounding it. Wax
14 bolus will be applied, which will eliminate the effects of the irregular surface (Figure G
15 6.3)

16 Choice of energy – From Table G 6.1, it will be seen that, at normal incidence, the 90% depth
17 dose of a 9 MeV beam is 27 mm. The surface of the lesion must be covered with 5 mm
18 bolus (using Table G 6.1) in order to raise the dose there to 90%. With this
19 arrangement, 9 MeV is adequate for the purpose.

PartG.doc Page 78 of 147 13 January, 2004


Part G Treatment Planning

1 Table G 6.1 For each energy shown, these data relate the skin surface dose and depth of the
2 90% isodose (at 45o incidence and 105 cm FSD), to the maximum dose (ie at dmax)
3 measured at normal incidence and 100 cm FSD using a 200 mm x 200 mm applicator (see
4 Figure 6). Thicknesses of bolus required to bring the surface doses up to 90% have been
5 included. While the parameters for individual machines and different angles of obliquity may
6 vary from those listed above, the Table serves as a first approach to designing a treatment
7 plan.

Incident beam energy 6 MeV 9 MeV 12 MeV 15 MeV 18 MeV 22 MeV

Surface dose, Do,s, at normal


incidence relative to dose at 75% 86% 83% 87% 88% 88%
dmax at normal incidence, 100
FSD
Surface dose, D45,s, at 45o
incidence, 105 FSD, relative to 83% 85% 87% 89% 90% 90%
dose at dmax at normal
incidence, 100 FSD
Surface build-up at normal
incidence, 100 FSD, required to 6 mm 5 mm 4 mm 1 mm 1 mm 1 mm
raise skin surface to 90% of dose
at dmax, at normal incidence
Surface build-up at 45o
incidence, 105 FSD, required to 2 mm 2 mm 1 mm 1 mm 0 mm 0mm
raise skin surface to 90% of dose
at dmax at normal incidence, 100
FSD
Depth, do,90%, of dose at
normal incidence, 100 FSD, 17 mm 27 mm 38 mm 45 mm 55 mm 63 mm
which is 90% of dose at dmax at
normal incidence
Depth, d45,90%, of dose at 45o
incidence, 105 FSD, which is 7 mm 11 mm 17 mm 23 mm 26 mm 30 mm
90% of dose at dmax at normal
incidence, 100 FSD
Depth of dose beneath skin
surface, covered with build-up as 5 mm 9 mm 16 mm 22 mm 26 mm 30 mm
in row 4, at 45o incidence, 105
FSD, which is 90% of dose at
dmax at normal incidence, 100
FSD
8

9 Field size - Using the 4E rule, the minimum field size without affecting the central axis depth
10 doses is 4 x 9 = 36 mm. Hence, the 4 cm diameter applicator should be appropriate,
11 provided that the radiotherapy oncologist accepts the rounded edges of the treatment
12 volume at depth.

13 Vulnerable organs - It is decided to protect the gums with a lead shield. However, because of
14 backscatter from the lead, which would raise the dose on the inner mucosal surfaces of
15 the cheek, a layer of wax is needed between the lead and the cheek. Saunders and Peters

PartG.doc Page 79 of 147 13 January, 2004


Part G Treatment Planning

1 (1974) have suggested that a lead shield covered with 5 mm wax to “dampen” the
2 backscattering is sufficient for electrons with residual energy of 3.5 MeV or less at the
3 lead interface. In practice, this arrangement is suitable for most applications, bearing in
4 mind that lead shields will not generally be used for beams of incident energy much
5 above 9 MeV.

6 A calculation for the lead-5mm-wax combination is instructive. With the 5 mm wax


7 protection, plus the 20 mm thickness of cheek and the 5 mm bolus on top, the lead is at a depth
8 of 30 mm. For this rough calculation, it is sufficiently accurate to suppose a linear decrease of
9 energy from the surface downwards, so that the energy of the electron beam when it hits the
10 lead shield is 9 (1 - 30/45) where the range of the electron beam is estimated to be 45 mm from
11 the relation: practical range (mm) = 5 x incident energy (MeV) (see Figure G 6.1). Hence, the
12 electron energy at the lead shield is about 3 MeV.

13 Using the relation given by Klevenhagen et al (1982), the electron backscatter factor is:

14 EBS(at shield) = 1 + 0.735e −0.052 Es = 1 + 0.735e −0.052×3 = 1.63 G6.1

15 where Es is the energy in MeV of the electrons at the shield. In other words, the backscatter
16 fluence is 63% of the incident fluence at the lead surface. For energies at the shield of between 1
17 MeV and 9 MeV, Lambert and Klevenhagen (1982) give an exponential expression for the
18 electron backscatter “upstream” from the shield:

19 EBS(at t mm “upstream”) = e (- k t)

20 where k is given by the relation

21 k = 0.61Es−0.62

22 Hence, in this case, k = 0.31 and the backscatter 5 mm upstream from the shield is exp (-
23 0.31 x 5) = 0.21, that is, 21% of the backscatter found at the shield itself. Hence, the
24 backscatter 5mm upstream is, finally, 21% of 63%, or 13% of the incident fluence at the shield.
25 Since the incident fluence at the shield (at a depth of 30 mm) is 80% (from depth-dose curves),
26 the mucosal surface of the cheek will experience 13% x 80% = 10% backscatter. However, the
27 dose from the direct incident beam (at the depth of 25 mm) is 95% (from depth-dose tables),
28 and so the total dose to the mucosa, with the lead shield plus 5 mm wax in place, is 105%,
29 which is quite acceptable.

PartG.doc Page 80 of 147 13 January, 2004


Part G Treatment Planning

1 In general, incident beams will be of this energy or lower when shields need to be used,
2 so that the 5 mm wax protection over the lead shield will usually be adequate. A 2 mm
3 thickness of lead, by the same token, will also normally provide adequate shielding (in this case,
4 it is perfectly able to absorb the residual 3 MeV energy).
5

6
7 Figure G 6.4 With the pinna in this position, the necessary side scatter is provided by a well
8 fitting wax plug. Lead sheet protects the surrounding skin, but a plug in the external auditory
9 meatus is still necessary to prevent possible tunnelling of electrons. Isodose values shown
10 (both above and beneath the lead shielding) are 10%, 20%, 50%, 80% and 90%.

11 G6.4.3. Example C - single electron field with surface inhomogeneities

12 Figure G 6.4 illustrates the treatment of the pinna of the ear. Referring to the check-list,
13 there is no need for a special technique and there are no problems with FSD or obliquity. Other
14 items on the list are:

15 Surface irregularities - In order to avoid the problems of inhomogeneous dose distribution


16 which would arise from irradiating the protruding pinna, a well-fitting wax block is
17 fashioned to enclose the pinna on both the anterior and posterior surfaces. In practice,
18 this may be arranged by producing the wax block in two halves so that the pinna is
19 sandwiched between them. Vaseline helps to eliminate any remaining air gaps, but
20 some centres prefer not to use such an aid because it is messy.

21 Choice of energy - The pinna extends 20 mm from the tip to the root. FromTable G 6.1, the
22 minimum appropriate energy is 9 MeV, which has a 90% depth of 27 mm, and so the
23 wax block must extend 7 mm beyond the tip of the pinna. This is also greater than the 5
24 mm required, fromTable G 6.1, to build up the dose to 90% on the tip of the pinna.

PartG.doc Page 81 of 147 13 January, 2004


Part G Treatment Planning

1 Field size - The antero-posterior dimension of the pinna to be treated is of the order of only 10
2 mm. However, from the 4E rule, a flat 90% isodose is only achieved in this case with
3 surface field sizes of 4 x 9 mm + 10 mm = 46 mm. In practice, a 60 mm x 60 mm
4 applicator is used, and the pre- and post-auricular areas are shielded by lead. Since the
5 lead is placed at the depth of the 90% isodose, the small aperture in the lead does not
6 affect the 90% isodose itself, although it will constrain the lower-value isodoses of
7 electrons transmitted through the aperture to take an almost semi-circular shape as
8 shown in Figure G 6.4.

9 Vulnerable organs - Although lead has been used to shield tissue from the incident electron
10 beam, it will be noticed in Figure G 6.4 that electrons can leak through the aperture
11 under the lead. It is therefore sensible to place a plug in the external auditory meatus to
12 prevent possible tunnelling. The plug may be made from wax, or it could be produced
13 by soaking cotton wool in olive oil. A thickness of 2 mm of lead will afford sufficient
14 protection in this instance, because the overlying thickness of the wax block will have
15 reduced the energy of electrons to half that of the incident 9 MeV beam. In general, the
16 thickness (in millimetres) of lead shielding required is given by half of the energy (in
17 MeV) of the electrons incident upon the shielding.
18
19
20
21
22
23
24
25
26
27
28
29

PartG.doc Page 82 of 147 13 January, 2004


Part G Treatment Planning

1
2 Figure G 6.5 In order to treat a curved chest wall with a single electron field, it is generally
3 necessary to provide bolus which is thick near the central axis of the field and tapers to the
4 minimum required for surface build-up at the field edges. Isodose values shown are 50%,
5 80%, 90% and 100%. The maximum dose within the two "islands" defined by 100% isodoses
6 at either edge of the field is 105%.

7 G6.4.4. Example D - single electron field with obliquity

8 Consider, now, the treatment of chest-wall irradiation after mastectomy. In Figure G


9 6.5, the posterior surface of the planning target volume has not been extended as far as the
10 surface of the lung in order to minimise the lung dose as discussed below. The lung extends
11 relatively close to the skin surface, and the target volume extends considerably to the posterior
12 of the patient, so that the technique of treating the chest wall with an opposed pair of tangential
13 photon fields would mean that an unacceptably large portion of lung would be treated to a high
14 dose, with the risk of radiation pneumonitis. As before, we consider the points in the check-list
15 in turn, and use them to illustrate general principles where appropriate. There are no surface
16 irregularities.

17 Special techniques - Ideally, this would be treated using electron arc therapy. However, for
18 reasons discussed later, this technique is not available in many centres, and so the
19 simplest approach - a static beam - is considered in this case.

20 Treatment FSD - The applicator may be used at the standard distance of 50 mm on the central
21 axis of the field. However, curvature of the area means that skin at the periphery of the
22 field may lie considerably further from the applicator than the standard distance (see

PartG.doc Page 83 of 147 13 January, 2004


Part G Treatment Planning

1 next point). In practice, it is the bolus applied over the skin surface which is placed at
2 the standard distance of 50 mm.

3 Obliquity - The curvature of the area means that electrons at the edges of the field will enter
4 with considerable obliquity, that is, at an angle which is significantly removed from the
5 local perpendicular. Since the problems of obliquity become greater as the angle of
6 obliquity increases (McKenzie 1979), it is nearly always best to arrange the angle of
7 incidence of the central axis so that the angles of obliquity at the two extreme edges of
8 the field, as shown in the transverse plane, are about equal.provided that this does not

9 conflict with other clinical considerations. In this case, a gantry angle of 30o is
10 appropriate, and the beam may be said to be “in apposition” to the skin.
11

12
13 Figure G 6.6 This illustrates the geometry used to construct Table G 6.1. Measurements
14 were made with a Varian Clinac 2500 linear accelerator using a 200 mm x 200 mm
15 applicator, and the data in Table G 6.1 must be regarded only as a guide when applied to other
16 linear accelerators.

17 Figure G 6.6 shows the geometry for the data in Table G 6.1, which are representative in
18 the energies, FSD and angle of obliquity useful in a manual, "first pass" approach to treatment
19 planning with oblique fields.
20

21 Choice of energy - In oblique incidence, it is generally the depth of the planning target volume
22 beneath the skin at the edges of the field (16 mm in this case at the lateral edge of the

23 field) which determines the required incident energy. The penetration of a 45o oblique
24 beam at 12 MeV, (bottom row of Table G 6.1) is 16 mm from skin to 90% depth dose,

PartG.doc Page 84 of 147 13 January, 2004


Part G Treatment Planning

1 and so this energy is used, together with 1 mm build up on the skin surface to raise the
2 skin dose from 87% to 90% (again, using Table G 6.1). At the central axis, the
3 thickness of bolus needed is determined from Table G 6.1 by subtracting the depth of
4 the target volume on the central axis (16 mm) from the 90% penetration depth at normal
5 incidence (38 mm from Table G 6.1).

6 With these constraints on the bolus, with a treatment planning computer, it becomes a
7 relatively easy matter to design the exact shape of the bolus required to shape the isodoses to
8 conform more closely to the target volume, using perhaps one or two trial shapes. The shape of
9 bolus in Figure G 6.5 gives rise to the computed isodose distribution shown, which leaves an
10 acceptably low incursion of high-value isodoses into the lung. Notice that the maximum dose at
11 the field edges is 105%. If the increased distance from the source had not mitigated the effect
12 of obliquity, this value would have been nearer 120%.

13 Field size - At 107 cm FSD at the field edges, the geometric (50%) coverage afforded by a 200
14 mm x 200 mm applicator is 214 mm. This is sufficient to provide the 200 mm coverage
15 required between the 90% isodoses at the medial and lateral field edges.

16 Calculation of monitor units - It would normally be unnecessary to calculate the monitor units
17 required to deliver the prescribed dose for this treatment since this should be computed
18 automatically by the treatment planning system when producing the isodose
19 distribution. Nevertheless, at 100 cm FSD to the surface of the bolus on the central axis,
20 the monitor unit setting will be given simply by that determined locally for the 200 mm
21 x 200 mm applicator at the standard distance of 100 cm FSD. This setting may not
22 agree exactly with that produced by the computer (because of the influence of the
23 curved surface) but it will be acceptably close.

24 Vulnerable organs - The lung inhomogeneities do not perceptibly influence the distribution of
25 the isodoses which lie above them, but the low density of lung nevertheless plays a part
26 in the design of the treatment plan. In this plan, a margin has been left between the
27 target volume and the lung surface so that it is the 70-80% isodose which is coincident
28 with the lung surface rather than the 90%. If the 90% isodose were to lie only a few
29 millimetres closer to the surface of the lung, this would push the 80% isodose
30 considerably further into the lung volume, leaving little room for error and an increased
31 risk of radiation pneumonitis.

PartG.doc Page 85 of 147 13 January, 2004


Part G Treatment Planning

1 G6.5. Field matching


2 Occasionally, situations arise where adjacent fields must be matched in order to avoid
3 an over- or under-dose. Examples of where such beam matching may be required are

4 (i) providing varying penetration by using different energy electron beams for
5 adjoining areas.

6 (ii) treating a larger area than standard applicators allow.

7 (iii) treating an area where the adjacent area has had previous treatment.

8 (iv) reducing the extreme effects of obliquity when treating a curved surface.

9 In beam matching, the join should be positioned away from critical areas. Staggered
10 joins (ie. shifting joins at different treatment factions) may be considered if this is practical.
11 One method of improving the matching is to use extended SSD to utilise the broadened
12 penumbra regions produced. This necessitates on-surface shaping of the other field edges to
13 restore sharp penumbra, and may well lead to a loss of beam flatness. Several solutions to the
14 problem have emerged over the years, for example, Kalend et al 1985, Kurup et al 1992, Kurup
15 et al 1993, Feygelman 1994.

16 Ulnin and Palisca (1996) and Lachance et al (1997) proposed techniques which broaden
17 the penumbra on the edge to be matched. However, the designs of the modifications to the
18 applicators depend upon the applicator cones, and the overlap or gap which must be left
19 between matching fields is again dependent upon the detail of the design. Furthermore,
20 measurements are required for a range of applicator sizes and energies if computer treatment
21 planning is to be used.

22 An alternative, and simple solution is to add a slab of material such as


23 polymethylmethacrylate (PMMA) to cover completely the end of the electron applicator
24 (McKenzie 1998). Such a slab is called a "spoiler" because it spoils the originally narrow
25 penumbra by broadening it. With a spoiler:

26 (i) there is no dependence of the properties of the spoiler design upon the applicator
27 cone, provided that the spoiler is large enough to cover the aperture completely;

PartG.doc Page 86 of 147 13 January, 2004


Part G Treatment Planning

1 (ii) no design-dependent overlap or gap is required between adjacent fields, as


2 matching is always to the light beam edge regardless of applicator size or FSD;

3 (iii) computer treatment planning programs can be used without modification or extra
4 beam measurement to calculate the effect of the spoiler.

6
7 Figure G 6.7 This illustrates the characteristics of the isodose distribution under an electron
8 beam spoiler. The light-beam edge marks the position of the 50% isodose at the surface, and
9 depths of the flat portions of isodose surfaces are given by subtracting the spoiler thickness
10 from the depths of corresponding isodoses in a beam with no spoiler. The dimension L is
11 defined in Figure G 6.8. Isodoses are shown at 10% intervals.
12

PartG.doc Page 87 of 147 13 January, 2004


Part G Treatment Planning

1
2 Figure G 6.8 This illustrates the geometry used for beam-spoiler treatment in field matching
3 and the dimensions used in equation G6.2

4 Figure G 6.7 illustrates two important characteristics of the dose distribution in tissue
5 irradiated with an electron beam through a spoiler. These characteristics are general in that they
6 are independent of FSD, applicator size, electron energy incident upon the spoiler and spoiler
7 thickness (provided that the thickness is not so great that the electron beam is becoming totally
8 diffuse, which is avoided in practice). In summary, these characteristics are:

9 (i) The light-beam edge closely coincides with the 50% isodose at the tissue
10 surface.

11 (ii) The depths of the isodoses in the flat portion of the distribution in tissue
12 irradiated through a spoiler may be found by subtracting the thickness of the spoiler
13 from the depths of the corresponding isodoses in tissue irradiated without a spoiler.

14 These simple features are particularly useful in planning electron therapy with spoilers.
15 One direct consequence of the first characteristic is that electron beams should be matched by
16 ensuring that the light-beam edges coincide at the tissue surface.

17 Another feature of using spoilers is that, in general, they increase the surface dose so
18 that additional skin bolus may be unnecessary. A further advantage is that the number of
19 monitor units required to deliver a given dose to the tissue is the same as that calculated for a
20 treatment without a spoiler, except for a small inverse-square decrease to allow for the fact that

PartG.doc Page 88 of 147 13 January, 2004


Part G Treatment Planning

1 the point of maximum dose within the tissue has moved closer to the surface by an amount
2 equal to the thickness of the spoiler.

3 McKenzie (1998) has shown that the 20% - 80% penumbra width, p, is given by

1
4 p = 1.2 Lz 2 E −1 G6.2

5 where p is in mm, L is the distance in mm from the inner face of the spoiler (the one nearer the
6 electron source) to the point where the penumbra is measured (see Figure G 6.8), z is the
7 thickness of the spoiler in mm and E is the energy (in MeV) of the electrons incident upon the
8 front face of the spoiler. This dependence on L, z and E is in accordance with theory predicting
9 the broadening of pencil beams through thick absorbers. This equation can be used to indicate
10 the tolerance of abutting fields to an inadvertent overlap or gap.
11

12
13 Figure G 6.9 Even when two beams with a zero hinge angle are made to overlap (in this case
14 by 5 mm), the overdosing is slight, because of the penumbra broadening introduced by the
15 spoiler. Isodoses are shown at 10% intervals.

16 Figure G 6.9 shows the isodose distribution at the junction of two 9 MeV beams

17 transmitted through a 10 mm spoiler with a hinge angle of 0o. Although the two beams have
18 been made to overlap by 5 mm, the measured hot spot is only 110%.

19 An advantage of using spoilers for matching is that the facilities required for treatment
20 planning with spoiler beams, at least in two dimensions, are already available on planning
21 computers which use Fermi-Eyges theory (Hogstrom et al 1981) or more recent variants. If the

PartG.doc Page 89 of 147 13 January, 2004


Part G Treatment Planning

1 computer treatment planning program is the kind which allows beam modifiers, then planning is
2 straightforward with spoiler beams. If the beam modifier facility is not available, then it is
3 nearly as simple to digitise an air gap under a thickness of tissue corresponding to that of the
4 spoiler. This has been done in Figure G 6.10(a) to predict the effect of two beams incident

5 through 10 mm spoilers at the extreme hinge angle of 90o, treating a cylindrical phantom. The
6 phantom diameter was 160 mm, taken to be typical of that of a scalp.

7 For comparison, Figure G 6.10 (b) shows the measured dose distribution for the
8 arrangement in Figure G 6.10 (a). Even at this extreme angle, it is noteworthy that (i) the
9 maximum dose measured in the phantom is only 116% (at a point inside the 110% isodose
10 curve) and (ii) the computer treatment plan is a relatively faithful representation of the measured
11 dose.
12

13
14 (a) (b)
15 Figure G 6.10 If the treatment planning computer does not have the facility to introduce
16 beam modifiers, it is still possible to model the spoiler by the simple expedient of creating an
17 air gap under a thickness of tissue equal to that of the beam spoiler. (a) This shows the
18 distribution computed using the air-gap technique with a 160 mm diameter cylindrical
19 phantom and the extremely large hinge angle of 90o. The maximum dose is 110%, just
20 beneath the intersection of the two beam edges. (b) This shows the distribution obtained by
21 measuring the arrangement in (a). The maximum dose (inside the 110% isodose curve) is
22 116%. Isodoses in both cases are shown at 10% intervals.
23

24 It should be noted that, if field matching is used in electron irradiation of chest wall, the
25 overdose at the matching point will generally be lower, partly because the hinge angles involved
26 will be less, and partly because the larger radius of curvature of chest walls means that matching
27 points are further from the electron source, which reduces the overdose compared to the
28 maximum dose on the central axes of the beams. However, in cases of field matching on
29 curved surfaces, it is always important either to test the predicted doses in a phantom before

PartG.doc Page 90 of 147 13 January, 2004


Part G Treatment Planning

1 treatment, or on the patient during treatment, because of the possibility of electrons from one
2 field being scattered well into the middle of the abutting field, contributing to the total dose.

3 When using spoilers, only the minimum thickness of spoiler necessary to produce an
4 adequately wide penumbra should be used, and if that leads to too great a depth for the 90%
5 dose, then skin bolus should be applied in addition. A 10 mm spoiler works well with 9 MeV
6 for chest wall irradiation. Since a spoiler broadens penumbrae at all edges of the field,
7 consideration should be given to shielding tissue beyond field borders which are not involved in
8 the matching.

9 G6.6. Electron Arc Therapy


10 The problems of dose inhomogeneity associated with obliquity and beam matching in
11 the electron treatment of curved surfaces are effectively eliminated with the use of narrow-beam
12 electron arcing techniques. However, these techniques are not yet standard in every
13 radiotherapy centre and, unlike other techniques which have been discussed earlier in this
14 section, commissioning an electron arc therapy service requires considerably more preparation,
15 starting from the technology, through data gathering to implementing a system for dose
16 computation. In view of these considerations, it is not practicable to provide enough
17 information here to cover the topic completely, but only to highlight the key points to be
18 considered in establishing electron arc therapy in a radiotherapy centre. Reviews of electron arc
19 therapy can be found in Leavitt et al (1985) and Leavitt (1996).

20 In electron arc therapy, it is important to try to minimise the variation in focus-skin


21 distance over the treatment arc, and this is achieved by placing the linear accelerator isocentre
22 as closely as possible to the centre of curvature of the arc of the patient'
s surface. The depth of
23 the centre of curvature beneath the patient’s surface should be at least Rp, the electron range,
24 and preferable around twice this distance, in order to avoid possible overdosing from
25 superposition of doses in the region of beam “crossfire”. In the treatment of surfaces with
26 relatively large radii of curvature, such as chest wall, it is clear that the treatment cannot be
27 delivered through the electron applicator normally used to provide secondary collimation in
28 stationary-beam therapy, because there would not be adequate clearance for the patient.

29 Secondary collimation of the arc therapy beam therefore generally has to be positioned
30 at a significant distance from the patient'
s skin, so that the beam profile in the plane of the arc
31 extends well beyond the limits indicated by the light field (Leavitt et al 1985 and Lam et al

PartG.doc Page 91 of 147 13 January, 2004


Part G Treatment Planning

1 1987). (The term "primary beam" collimation is reserved here for that imposed upon the
2 electron beam by the x-ray collimators.) Because of this beam spread, the aim, at least for the
3 middle sections of the arc, is for the complete beam profile to pass over any given point, so that
4 the dose received by the point is integrated over the whole beam profile. This is more
5 practicable with narrower beams, of around 5 cm width when projected to the isocentre, which
6 will also minimise problems of obliquity at the edges of the beam profile.
7

8
9 Figure G 6.11 This attempts to explain the effect of radius of curvature of the patient on dose
10 delivered during arc therapy. In a complete rotation, the dose delivered at depth d is
11 proportional to (r - d)-1 (f - r + d)-1 (see text).

12 In order to see the effect of radius of surface curvature upon dose delivered by arc
13 therapy at a given depth, d, consider the irradiation of a cylindrical phantom of radius r (Figure
14 G 6.11). In particular, consider the disk-shaped volume (containing the central treatment plane)
15 defined by a sector of the beam which extends an angle in the direction perpendicular to the
16 plane of rotation (see Figure G 6.11). At depth d, the electron energy delivered to an
17 incremental depth d through the sector of angle during one complete rotation will be spread
18 evenly within the volume 2 (r - d) (f - r + d) d . Since the dose will depend upon the reciprocal
19 of the mass, and, hence, the volume throughout which the energy is distributed, the variation of
20 dose D(r,d) at depth d beneath a surface of radius r will be given by:

PartG.doc Page 92 of 147 13 January, 2004


Part G Treatment Planning

1 D(r,d) α (r - d)-1 (f - r + d)-1 G6.3

2 It will be seen from this dependence of dose upon radius that, in the competing effects
3 of decrease in radius of curvature and consequent increase in distance from the source, the
4 former effect wins. In summary, a decrease in radius of curvature leads to an increase in dose.

5 It may well be necessary to arrange for a treatment to be based on more than one centre
6 of curvature in order to accommodate the different skin curvatures in the central plane. In
7 planes parallel to the central plane, lying superior or inferior to it, the skin curvature may again
8 be different. The solution is, rather than to use a rectangular treatment beam, instead to use a
9 beam which varies in width at different distances superior or inferior to the central plane,
10 depending upon how the skin curvature varies away from the central plane. The resultant beam
11 shape may be fixed, and generally trapezoidal in shape, or it may vary during treatment using
12 multileaf collimators (Klein et al 1996, Leavitt et al 1989). Such field shaping with secondary
13 collimation, whether static or dynamic, may also be used to neutralise the relatively poor beam
14 profile in the superior-inferior direction (Leavitt et al 1985).

15 The dose distribution at the ends of the arc is naturally compromised because of
16 incomplete integration across the dose profile. The dose distribution may be sharpened by
17 placing tertiary collimating lead or low-melting-point alloy shaped to the treatment field on the

18 s skin and extending the treatment arc some 10o - 15o over the tertiary collimator.
patient'
19 Alternatively, dynamic collimation may be used (Swalec et al 1994) where the jaws in the plane
20 of the arc are opened and closed progressively at the beginning and end of the arc respectively,
21 which mimics the effect of tertiary shielding.

22 Dosimetry in electron arc therapy treatment planning should be done using an


23 appropriate pencil-beam computer algorithm. It is characteristic of electron arc therapy that the
24 depth-dose curves tend to display a lower surface dose, a greater depth of maximum dose,
25 dmax, and higher percentage depth-doses beyond dmax than they do for stationary electron
26 beams of the same incident energy. This can largely be explained using the same reasoning
27 underlying proportionality (Khan et al 1977): at greater treatment depths in arc therapy, the
28 electrons are distributed over diminishing lengths of arc and so the fractional depth dose is
29 greater than that for stationary beams.

PartG.doc Page 93 of 147 13 January, 2004


Part G Treatment Planning

1 The low surface doses in arc therapy can be an obstacle to achieving dose homogeneity,
2 and Leavitt et al (1990) have suggested combining electron energies, including using build up
3 on the lowest energy, to achieve a more uniform depth dose distribution.

4 It is always important in arc therapy to remember that the bremsstrahlung contribution is


5 additive around the isocentre (Kase and Bjärngard 1979). Clearly, as the angle of the treatment
6 arc increases, so will the relative bremsstrahlung contribution, which may become a significant
7 fraction of the treatment dose, particularly at higher incident electron energies.

8 G6.7. Inhomogeneities and electron planning algorithms


9 Inhomogeneities in composition and in shape modify dose distributions in two ways,
10 both of which need to be considered in the clinical situation. The first is the effect on
11 absorption and the consequent shift in isodose lines and is most pronounced through and
12 beyond large inhomogeneities. The second is due to scatter differences between different
13 materials (interface or edge effects) and is most pronounced for small inhomogeneities or near
14 the edges of larger heterogeneities. The physics of the latter and some basic effects were
15 discussed in Part E Section E5.2. The resultant distribution depends on the size shape and
16 composition (atomic number and electron density) of the inhomogeneity, as well as on the
17 energy and field size of the beam. In current practice they are best considered by using pencil
18 beam algorithms as available on most modern treatment planning systems. However,
19 increasingly, Monte Carlo techniques are being applied to these problems. Simple manual
20 methods can be used to provide initial rough estimates of the effects or approximate checks of
21 more sophisticated approaches (e.g. see Klevenhagen 1985, Thwaites 2000).

22 Pencil beam algorithms are discussed in Part F Section F3.2sections 7.8.2 and Monte
23 Carlo methods in Part F Chapter 4. Comprehensive reviews of the development of electron
24 beam algorithms can be found in Jette (1996) and of Monte Carlo modelling of electron beams
25 in Ma and Jiang (1999). A concise summary of electron beam algorithms is given in Thwaites
26 (2000), along with a brief review of their accuracy. Currently implemented pencil-beam
27 algorithms have generally been shown to be accurate to within around ±5% or ±5 mm over a
28 range of applications, with the more recent algorithms showing improvements in performance.
29 In the simpler situations, accuracies of ± 2 % or ± 2 mm are achievable in accordance with the
30 requirements laid down for photon algorithms (ICRU 1987). However larger discrepancies of
31 the order of 10% and with the potential to be significantly larger have been observed at larger

PartG.doc Page 94 of 147 13 January, 2004


Part G Treatment Planning

1 depths for narrow inhomogeneities with long edges parallel to the beam. Pencil-beam
2 algorithms are currently the most widely implemented approach for electron treatment planning
3 on commercial plannning systems. However there are a number of different algorithms and
4 implementations in use and the user must ensure that the system is well-tested locally and that
5 any limitations are clearly understood. Commissioning recommendations and test methods are
6 provided in Shiu et al (1992), van Dyk et al (1993), IPEMB (1994), IPEM (1999) and AAPM
7 (1998).

8 G6.8. Treatment prescription


9 Prescription depends on clinical requirements but is generally to the position of dose
10 maximum, although other levels may be used. It should be noted that, in some circumstances,
11 low energy scatter components may give a dose maximum quite close to the surface and this
12 may influence the position of dose prescription. Electron prescription recording and reporting is
13 dealt with in ICRU report 29 (1978); the ICRU is currently considering recommendations to
14 replace this, in follow up to the newer recommendations for photon beams included in ICRU
15 50(1993) and ICRU 62 (1999).

16 G6.9. Current and future developments


17 Electron therapy is set to benefit from the same developments which are currently
18 adding to the power of megavoltage photon therapy. Inverse treatment planning will
19 increasingly be used to design solutions to difficult problems and these solutions will be
20 implemented in clinical practice using electron-beam shaping with multileaf collimators both in
21 static and in dynamic mode, eg. Klien 1998, Korevaar et al 1998. This same technology also
22 offers the possibility of mixing the modalities of photon and electron beams, or different energy
23 electron beams, to design individually tailored beam characteristics of depth dose, flatness,
24 penumbra width and skin dose, eg. Karlsson and Zackrisson 1997, Korevaar et al 1998.

25 Whilst pencil-beam based treatment planning algorithms will continue to develop,


26 Monte-Carlo programs are now available commercially which will handle complex dose
27 distributions in acceptable times, and these will be especially useful for dealing with the effects
28 of inhomogeneities. Figure G 6.4 was drawn freehand with an eye to the physics of
29 backscattering from lead and diffusion through a narrow aperture. Monte-Carlo programs will
30 underpin the art of prediction in cases like this and will add significantly to the confidence
31 placed on calculated distributions. This is turn will lead to an increase in the use of intensity-

PartG.doc Page 95 of 147 13 January, 2004


Part G Treatment Planning

1 modulated electron beams and their combination with photon beams to tailor dose distributions
2 in appropriate clinical situations.

3 Chapter G7. Dosimetric Evaluation of Treatment Plans


4 Margaret Bidmead

5 G7.1. Introduction
6 The ideal radiotherapy treatment plan produces a uniform coverage of the target volume
7 without giving significant dose to surrounding normal tissue. The use of CT planning, MR and
8 fast computer hardware (for both planning and treatment delivery) has meant that 3D planning
9 is a realistic option. More degrees of freedom in the use of noncoplanar beam arrangements are
10 achievable, and the introduction of intensity modulation, both make plans more difficult to
11 compare and evaluate. Good, fast plan evaluation tools are essential to simplify this process.

12 Several such tools have been designed, and will be discussed in this chapter.

13
14
G7.2. Isodose display

15 Isodose information can be demonstrated on a treatment plan as lines of equal absolute


16 dose, where the dose prescription has been incorporated into the treatment plan, or relative dose
17 (%) which can then be applied to any prescription dose. The modality chosen for display is
18 selected locally and may depend on whether more than one phase of radiotherapy treatment is to
19 be delivered.

20 To calculate a dose distribution requires a lattice of points, spread over the volume of
21 interest, where dosimetrtic informations are considered as important. These points are contained
22 within the external surface of the patient and their number and spatial resolution are the result of
23 a trade-off between the speed of calculation which is dependent upon the dose calculation
24 algorithm used, and the accuracy aimed at. The X and Y co-ordinates of these points are within
25 the transverse planes and the Z co-ordinate is limited by the spacing between slices. The isodose
26 contours in a given calculation plane are usually produced by linear interpolation between this
27 matrix of points. The separation of the points should therefore be small enough to allow
28 accurate linear interpolation. The resolution of a dose distribution for visual display and
29 hardcopy could be different if the dose calculation algorithm is too slow for quasi real time
30 display, with a higher resolution chosen for the hardcopy information. This gives a final and

PartG.doc Page 96 of 147 13 January, 2004


Part G Treatment Planning

1 more accurate representation of the isodose curve chosen to cover the PTV, isodoses through
2 critical organs and the location of hot or cold spots in the dose distribution.

3 Although it is possible to display the dose distribution as total absolute doses over the
4 whole treatment course, some form of normalisation is usually employed. The beam weights
5 can be adjusted to give 100% to a pre-defined point, such as the ICRU reference point (ICRU
6 1993). Alternatively, the weights can be arbitrary chosen, the dose being renormalized
7 afterwards at this point. This allows for instance to assess the dose uniformity within the PTV
8 on the basis of a tolerable % deviation from prescribed dose (i.e. 95-105%).

9 The dose distribution can be displayed as multi-planar views as shown in Figure G 7.1,
10 particularly in transverse (Fig. 1), sagittal (Fig. 2) and coronal (Fig. 3) sections. It can also be
11 displayed as a 3D representation of anatomy and isodose surfaces (fig. ??)

12 Different display methods for dose distribution and anatomical information should be a
13 feature of any modern treatment planning system. Some useful features are listed as follows:
14

15 • Point dose contributions from all fields, especially for critical organ dose evaluation Figure
16 G 7.1

17 • Opaque colour wash displays can give a good overall impression of the dose distribution,
18 but can obscure target volumes, so should be interchangeable with conventional isodose
19 lines.

20 • Transluscent Ccolour wash displays superimposed on CT gray-scale information which can


21 be windowed

22 • Zoom and pan facilities, with variable calculation resolution over the zoomed areas.

23 3-D surface rendering of isodose surfaces displayed for instance as wire frames,
24 superimposed on organ volume surface displays, on a graphics device capable of real-time
25 image manipulation.

26 • “Volume of regret” is a dose reduction technique applied to the dose distribution, using
27 surface displays. A window of acceptability of dose level is chosen and portions of organs
28 outside this window are highlighted. For critical organs this is a one-sided test, where
29 regions of dose above the acceptable level are displayed. For target volumes regions of

PartG.doc Page 97 of 147 13 January, 2004


Part G Treatment Planning

1 either too high or too low dose can be shown. This is particularly useful when checking the
2 coverage of PTV when using arcing beams e.g. for stereotactic radiotherapy. [ref ??]

5 (a) Transverse display

Figure G 7.1 (b) Sagittal section (c) Coronal section

7 Ideally most of the plan evaluation should be done on real time displays of dose
8 distributions. It is useful to instantaneously display the transverse, sagittal and coronal planes on
9 an interactive VDU to help overcome the limitations of a 3D display on a 2D medium. This
10 requires special graphics hardware in which all the images are loaded in graphics memory and
11 rapidly accessed. The real time capability gives the illusion of the user moving through the 3D
12 display (Emami et al 1991)

13 G7.3. Dose Volume Histograms


14 Volume calculations performed by 3D treatment planning systems provide a large
15 amount of dose information which can be difficult to interpret and evaluate when it is displayed
16 as isodose curves on several transverse, sagittal and coronal planes. It is much simpler to

PartG.doc Page 98 of 147 13 January, 2004


Part G Treatment Planning

1 condense the 3D dose distribution data to a graph, which displays the radiation distribution
2 within a specifically defined volume of interest, so that summarising and analysing the data is
3 possible. Such a graphical representation is called a Dose Volume Histogram (DVH).

4 The DVH can be expressed as:

5 The summed volume of elements receiving dose in a specified dose


6 interval, against a set of equally spaced dose intervals.

7 This is a differential dose-volume histogram and shows absolute or relative volume in each dose
8 interval (bin) directly.

9 More useful, however are cumulative dose volume frequency distributions, which are
10 plots of the volume receiving a dose greater than, or equal to, a given dose, against dose. The
11 volume accumulates starting at the highest dose bin continuing towards zero dose, eventually
12 reaching 100% of the total dose. In most cases the volume is specified as the percentage of the
13 total volume of a particular structure receiving dose within each interval, however it may be
14 expressed as absolute volume in some cases.

15 DVH’s can be used during the planning process to check whether the dose is adequate
16 and uniform throughout the target volume, and the extent and value of any hot spots in adjacent
17 normal tissue. However because they do not display positional information they should not be
18 the only method of plan evaluation used.

19 Their main use, however, is as a plan evaluation tool. They can be used as a method of
20 comparing different treatment plans on a single graph, for specifically identified structures and
21 target volumes. Use of dose volume histograms in specific clinical sites has been extensively
22 reported in the literature (Chen et al 1984, Austin-Seymour et al 1986, Brown et al 1991, Coia
23 et al 1991, Kutcher et al 1991, Munzenrider et al 1991, Shank et al 1991, Shipley et al 1991,
24 Simpson et al 1991, Solin et al 1991, Drzymala et al 1994). DVH’s are also used as the input for
25 calculation of tumour control probabilities (TCP) and normal tissue complication probabilities
26 (NTCP) which allows quantitative evaluation of treatment plans. DVH’s need to be accurately
27 calculated because of the sensitivity of TCP and NTCP calculations to small changes in the
28 shape of the DVH.

29 In order to calculate DVH’s various parameters have to be considered (Dryzmala 1991):

PartG.doc Page 99 of 147 13 January, 2004


Part G Treatment Planning

1 Volume

2 The boundaries of the anatomical structures of the patient are defined and the anatomy
3 is subdivided into a volume grid of chosen, appropriate resolution. The 3D image is made up
4 of a sequence of CT scans. This can be represented as a 3D grid with cubic volume elements,
5 equally spaced in the X and Y directions, but not necessarily in the Z direction. The volume of
6 the voxel can then be defined by the product of the X and Y grid spacing and the slice thickness.
7 . In order to calculate the total volume of a particular structure, the volume elements (voxels)
8 lying within the structure are summed. A reasonable resolution for sampling the volume of
9 typical anatomical structures is 2-3mm.

10 Dose

11 The dose for each voxel can be determined at the same time for the outlined volume of
12 interest, by linearly interpolating the dose to the center of the voxel from the dose matrix. Dose
13 resolution then depends solely upon dose gradients in the region. All contributing voxels are
14 collected into the appropriate dose bins of the histogram, for each structure. The user can
15 specify the number and dose interval of the bins to customise the resolution of the DVH plots.
16 These bin values can then be expressed in graphical form.

17 Grid spacing and Size

18 The 3D dose matrix need not be coincident with the 3D volume calculation matrix, but
19 the accuracy of volume calculation should not be limited by the dose matrix resolution. (Often a
20 finer grid is used for volume estimation than for dose calculation). It is important that the dose
21 calculation window is made large enough to encompass all the voxels in any particular
22 structure, otherwise the dose to that voxel will be incorrect.

23 Volume Resolution

24 Another inaccuracy occurs when the CT scan information only partly covers the volume
25 of interest. Assuming that the CT data covers the region receiving any significant dose , then a
26 possible solution is to use a set of standard volumes of particular organs against which to
27 compare the computed volume. If the computed volume is significantly smaller than the
28 standard size the additional volume can be added to the zero dose bin.

PartG.doc Page 100 of 147 13 January, 2004


Part G Treatment Planning

1 Dose Matrix Resolution

2 Grid spacing for dose calculation is a compromise. The finer the grid, the greater the
3 calculation accuracy in regions of steep dose gradients for example close to the beam edges.
4 The disadvantages of a fine grid are the increased computation time and the creation of large
5 data files. In general two criteria are specified, the desired dose accuracy and the maximum
6 acceptable distance between the estimated and actual isodose contours. 2% dose accuracy or
7 2mm isodose positional accuracy will be achieved with a grid spacing of 5mm

8 Binning

9 The range of expected dose values is divided into equal intervals when constructing a
10 histogram. For each interval the volumes of the voxels receiving dose within that interval are
11 accumulated in the appropriated element of an array, or bin. Cumulative DVH’s are obtained by
12 adding volumes accumulated in each bin with the volumes in all bins corresponding to higher
13 dose intervals. Differential DVH’s treat the volume in each bin as a separate quantity, and a
14 bin interval of between 2 and 5 Gy is reasonable. For cumulative DVH’s the appropriate dose
15 interval for the bins depends on the dose response curve for the structure of interest. 0.5 Gy has
16 been shown to be reasonable, whereas 2Gy is too wide an interval.

17 G7.4. Addition and subtraction of DVH’s


18 It is sometimes useful to have DVH information of structures that are completely or
19 partially contained within another structure. e.g. rectal wall, as opposed to volume encompassed
20 by solid definition of rectum, or for a rectal tumour where normal rectum needs to be separated
21 from tumour in the DVH calculation, or a combined DVH of both lungs etc. In order to do this,
22 a system of logical operators is useful, such as ‘and’, ‘or’, ‘not’, ‘inside’ and ‘outside’. The
23 DVH calculation then looks at each point of the volume grid and determines within which
24 logically specified region the point lies.

25 Another method, useful when one structure is contained within another is to assign
26 hierarchy numbers to each structure , the highest number representing the most important
27 structure. Each element in the dose matrix is given a “tag” value, with the more important points
28 given a higher value, so that any points occurring in both structures will have the highest
29 priority for the structure of interest under calculation. For example, a structure within another
30 can be subtracted out by assigning the inner structure a higher priority than the outer one.

PartG.doc Page 101 of 147 13 January, 2004


Part G Treatment Planning

1 G7.5. Examples of Cumulative DVH’s


2 When evaluating several plans it is much easier to display the DVH curves of the
3 structure for comparison on the same graph. DVH PTV comparisons should show a uniformly
4 high dose throughout the volume , the shape approximates to a step function and a steep drop in
5 the “y” values show that a large percentage of the volume has a similar dose. An ideal curve
6 would show this step around the prescription dose, (Figure G 7.2) , where the DVH’s for the
7 PTV in 3 different treatment beam configurations are compared.

Figure G 7.2 DVH for PTV Figure G 7.3 DVH for Rectum

9 DVH’s for structures have a concave appearance. Figure G 7.3 shows comparative
10 DVH’s for a rectal volume where 50% of the volume is receiving 63, 66 or 68 Gy depending on
11 the different treatment plan. It is also useful to display both absolute dose and absolute volume,
12 together with normalised values, on the same graph. For observing small differences in DVH’s
13 it can be of benefit to use a non-linear volume axis.

14 G7.6. Errors in DVH calculation


15 Errors in DVH calculation are primarily dependent on :

16 The accuracy of estimating structure boundaries

17 These are sometimes difficult to identify and may be unreproducible. Contours are often
18 entered with mouse, tracker-ball, digitiser etc, which are difficult to control and introduce
19 additional error. (Improvements provided by computer automation are necessary). Structures
20 are often irregular shapes and “clever algorithms” are required to identify certain features.

PartG.doc Page 102 of 147 13 January, 2004


Part G Treatment Planning

1 The accuracy of dose calculations and interpolations of the dose matrix.

2 A good way to test the accuracy of dose calculations is to use test phantoms of known
3 dimensions and dose. Useful phantom designs include a set of concentric cubes and circles of
4 known volume.

5 G7.7. Limitations of DVH’s


6 It is not strictly correct to compare plans with different dose computational parameters
7 because different dose calculation algorithms may not handle heterogeneities, or penumbra etc.
8 in the same way. The dose matrix resolution can affect the results of the DVH, especially in
9 regions of steep dose gradient.

10 (i) DVH’s summarise the dose distribution within a structure graphically, they do
11 not give spatial information. They show that hot and cold spots exist, but do not indicate
12 where they occur, nor whether there are several of them. Dose Area Frequency
13 distributions can be used for skin surfaces and some use can be made of Dose-length
14 histograms for long thin structures (Dryzmala 1991).

15 (ii) Interpretation of the DVH plot is fairly subjective and implications of small
16 differences between DVH’s are not well understood. This indicates the usefulness of an
17 objective numeric score, such as TCP or NTCP to provide a “ranking” when comparing
18 plans.

19 (iii) DVH’s do not indicate the complexity of the field arrangements and will not
20 show any use of “illegal” couch, collimator angles etc. DVH’s should be used in
21 conjunction with other methods for treatment plan evaluation.

22 In summary the DVH is a very useful tool for 3D plan comparisons and as an input
23 parameter for TCP and NTCP calculations. TCP and NTCP are quantitative methods of
24 predicting the likelihood of local control. DVH’s and the predicted behaviour of the cell
25 population in question are used as input data. There is however large uncertainty in the clinical
26 data, so absolute probabilities are not attainable, but relative probabilities can be used to assess
27 different treatment plans. DVH’s should however be used in conjunction with other plan
28 evaluation tools, especially those which will give spatial information.

PartG.doc Page 103 of 147 13 January, 2004


Part G Treatment Planning

1 G7.8. Dose Statistics


2 Dose statistics as outlined in Table G 7.1 give a simplified view of the dose in a
3 specified structure for evaluation purposes to compare 3D plans.
4

5 Table G 7.1 Definitions of the dose statistics


6
Statistic Definition
Total volume The sum of all DVH voxels found within a set of
boundary contours.
ICRU 50 dose The specification dose as defined in the ICRU50
Report (ICRU 1993).
Mean dose The sum of the doses assigned to each voxel divided by
the total number of voxels
Min dose The dose minimum in the volume.
Max dose The maximum dose in the volume.
Volume greater than or The sum of all voxels within DVH bins corresponding to
equal to the prescribed a dose greater than or equal to the prescription dose.
dose

Volume greater than or The total volume within DVH bins corresponding to a
equal to the reference dose greater than or equal to the user supplied reference
dose dose.
7

8 G7.9. Figure of Merit (FOM)


9 FOM’s are described in detail in Jain and Khan (1992) and Jain et al (1993). The FOM
10 is a ranking device , calculated from inputs of TCP, NTCP and oncologist preferences, and can
11 rank up to three competing treatment plans in order of “merit” in order to achieve a single
12 numerical score for each treatment plan.

13 G7.10. Neural networks


14 A publication by Willoughby et al (1996) describes a system of evaluating and scoring
15 radiotherapy treatment plans using an artificial neural network. Treatment plans were assigned a
16 figure of merit by a radiation oncologist using a five-point rating scale. DVH data extracted
17 from a large training set was correlated to the physician-generated figure of merit using an
18 artificial neural network, and the net was tested on another set of plans. The accuracy of the

PartG.doc Page 104 of 147 13 January, 2004


Part G Treatment Planning

1 neural net in scoring plans compared well with the reproducibility of the clinical scoring and the
2 system is promising for the reliable generation of a clinically relevant figure of merit.

3 G7.11. Dose-volume distribution


4 Data is binned in volumes, rather than dose which can be especially useful when a
5 small number of dose calculation points are used (Niemerko and Goitein 1994).

6 G7.12. Other tools


7 Other tools that are available include:

8 • Real time use of BEV’s with dose contour displayed (McShan et al IJROBP Vol 18)

9 • Video loop stepping through two plans displayed side by side, particularly with colour wash
10 displaying isodoses.

11 • Dose at a point interactively integrated and displayed simultaneously on both plans.

12 • Dose difference displays

13 • Surface dose display, e.g. on the surface of the spinal cord (McShan et al 1979)

14 The other aspect of plan evaluation which has not been addressed is the accuracy of the
15 actual dose delivery, which is dependent on the complexity of the plan, size of margins used,
16 reproducibility of set-up etc. Some sort of uncertainty analysis should be developed to include
17 these other aspects of treatment delivery into treatment plan evaluation.

PartG.doc Page 105 of 147 13 January, 2004


Part G Treatment Planning

1 Chapter G8. Biological Evaluation of Treatment Plans


2 Alan Nahum and Gerald J. Kutcher

3 G8.1. Tumour Control Probability (TCP)

4 G8.1.1. Introduction

5 This section deals with the modelling of Tumour Control Probability; the emphasis is on
6 the effect of the distribution of the absorbed dose within the target volume rather than on the
7 difference between different fractionation schemes. Some of the reasons why a model for TCP
8 is desirable are listed below:

9 • As a way of estimating the effect of non-uniformities in the tumour dose distribution


10 (Brahme 1984; Sanchez-Nieto and Nahum 1999)

11 • Enables one to make estimates of the effect of dose and patient position uncertainties on
12 therapy outcome (Brahme 1984; Boyer and Schultheiss 1988; Mackay et al 1999)

13 • Optimisation of treatment plans is beginning to be done in terms of TCP and NTCP


14 (Källman 1992; Mohan et al 1992; DeGersem et al 1999)

15 • As a way of using the results of biological assays for α (tumour) etc. (Peters et al 1989;
16 Deacon et al 1984; West et al 1991; Bentzen 1997)

17 However, perhaps of greatest importance is that such modelling can serve as an aid to
18 clarity of thought in external-beam radiotherapy.

19 G8.1.2. A simple biologically-based TCP model

20 G8.1.2.1. General

21 It is well-known that the so-called Dose-Response curve has a sigmoid shape e.g.
22 Brahme (1984). Several authors have fitted mathematical functions to this curve. However, it is
23 not easy to see how changes in basic parameters such as tumour cell radiosensitivity,
24 inhomogeneities in the dose distribution, variation in tumour volume, in clonogenic cell density
25 etc. can be accommodated by empirical curve-fitting approaches. In the case of Tumour
26 Control, in contrast to that for Complications in Normal Tissues, it is actually possible to
27 develop a model starting from the response of cells to radiation. Niemierko and Goitein (1993)

PartG.doc Page 106 of 147 13 January, 2004


Part G Treatment Planning

1 described such a model, which is very similar to the one given here (Nahum and Tait 1992;
2 Webb and Nahum 1993).

3 G8.1.2.2. A simple Poissonian model

4 Numerous radiobiological experiments have demonstrated beyond doubt that the killing
5 of cells by radiation can be described by an expression of the form

(
S = exp − αD − βD 2 ) G8.1

6 where S is the surviving fraction after a (uniform) dose D of radiation to a population of cells.
7 The parameters α and β characterise the initial slope and degree of curvature, respectively, of
8 the survival curve. This is known as the Linear-Quadratic or LQ model of cell killing (e.g.
9 Fowler 1989; Steel 1997).

10

11 Figure G 8.1 Modification of the cell-survival curve when the radiation is given in 2-Gy
12 fractions

13 When the irradiation is fractionated as in external-beam radiotherapy (Figure G 8.1: full


14 curve), for small fractions such as 2 Gy (per day) fractions, the effective slope of the survival
15 curve is very nearly given by the value of α alone. Thus one can write:

PartG.doc Page 107 of 147 13 January, 2004


Part G Treatment Planning

N S ≈ N o exp(− αD ) G8.2

1 where No is initial number and Ns the surviving number of clonogenic cells, assumed here to be
2 irradiated uniformly and to have uniform radiosensitivity α(Gy-1); the latter is sometimes
3 denoted by αεφφ. Note it is straightforward to reinstate the β term; Niemierko and Goitein (1993)
4 include it in their model, for example, as do Sanchez-Nieto and Nahum (1999).

5 There is considerable radiobiological evidence for the statement that a tumour is only
6 "dead" when every single clonogenic cell has been eliminated (Ref?). Thus the probability of
7 Tumour Control (TCP) is the Probability that No Single Clonogenic Cell Survives. Exploiting
8 the Poisson Statistics Relation e.g. Porter (1980), Deasy (1996):

TCP = exp(− N S ) G8.3

9 If we now substitute Equation G8.2 for Ns into Equation G8.3 we arrive at

TCP = exp(− N o exp(−αD ) ) G8.4

10 Using a realistic value for the number of initial clonogenic cells No of the order of 109
11 (Steel 1997) and realistic values of from 0.1 to 1 Gy-1 (e.g. Deacon et al 1984) one obtains
12 the family of curves shown in Figure G 8.2.

13
14 Figure G 8.2 TCP curves from Equation G8.4 for clinically realistic values of No = 109 and α
15 ranging from 0.1 to 1.0 Gy-1

PartG.doc Page 108 of 147 13 January, 2004


Part G Treatment Planning

1 G8.1.2.3. Consistency with Clinical Data

2 A number of Local Control vs Tumour Dose studies have been published (e.g. Morrison
3 1975; Batterman et al 1981; Moore et al 1983; Hanks et al 1988). Despite the limitations
4 associated with such data i.e. uncertainties in the dosimetry, inadequate patient numbers,
5 imprecise clinical definition of Local Control etc. there are almost no Tumour Control vs Dose
6 curves with slopes anything like as steep as the ones in Figure G 8.2. This has led some
7 investigators to favour an empirical model to fit these clinical Dose-Response curves e.g.
8 Brahme (1984).

9 The TCP model described here, which is based on meaningful radiobiological


10 parameters, explicitly incorporates inter-patient variation by assuming that αis distributed
11 normally amongst the patient population, with standard deviation σα(Nahum and Tait 1992;
12 Webb and Nahum 1993). As one increases the value of σαthe slope of the dose-response curve
13 decreases. One way of thinking about this is to regard the resulting curve as the sum of the
14 curves for different ? values in Figure G 8.3. There is thus a group of patients with low α values
15 who will never be cured (TCP = zero), another group with high αvalues who will always be
16 cured, and a group with intermediate αvalues for whom the term stochastic fraction has been
17 coined as the outcome for these patients is literally a matter of chance (Zagars et al 1987).

18

19
20 Figure G 8.3 Tumour Control Probability (TCP) as a function of target dose, derived from
21 Equations G8.5 and G8.6, with a = 0.35, ρcl = 107 for a 320 cc volume, for σα = 0.0 and for the
22 clinically more realistic σα = 0.08 (adapted from Nahum and Tait, 1992)

PartG.doc Page 109 of 147 13 January, 2004


Part G Treatment Planning

1 The present model was first applied to the case of bladder tumours. A mean α value of
2 0.35 Gy-1 determined by Deacon et al (1984) for human bladder tumour cells grown in vitro
3 was adopted. Local control vs dose curves were then computed from

TCP( D) = g TCP(α ,D, N o) G8.5


ι i

4 where TCP (α, D, No) is given by Equation G8.4 and a fraction gi of the patients have α = αι
5 such that

gi ∝ exp - (α i - α ) 2 / 2 σ 2
α G8.5

6 and Σgi = 1. The initial number of clonogenic cells, No, has been estimated from the product ρcl x
7 Vtgt with the clonogenic cell density taken to be 107 [19] and the mean value of the target volume
8 Vtgt = 320 cm3; this latter value was derived from an analysis of the actual target volumes, as
9 outlined on CT, of patients entered into a clinical trial of conformal vs. conventional pelvic
10 radiotherapy (Tait et al 1997).

11 The two curves in Figure G 8.3 have been calculated using the above data. In both cases
12 the TCP at the actual clinical dose used, 64 Gy (32 x 2-Gy fractions), comes out at just below
13 0.5. This is consistent with clinical findings and lends some confidence to the model. The value
14 of σα= 0.08 is entirely empirical; it was arrived at by adjusting σαuntil the curve "fitted" the
15 dose-response data in Batterman et al (1981). It has been used in subsequent analyses e.g. Webb
16 and Nahum (1993). It can be noted that the more usual way of denoting the slope is by the
17 parameter γ = (∆TCP/∆D) D at either the 50% or 37% point on the TCP curve (Brahme 1984).
18 Values of 1.7 to 2.0 are typical for human tumours (Fischer and Moulder 1975; Stewart and
19 Jackson 1975), compared to around 7 for the conditions depicted in Figure G 8.2.

20 G8.1.2.4. Inhomogeneous Dose Distributions

21 The model developed thus far has assumed that all cells receive exactly the same dose.
22 Thus some way is needed to incorporate dose distributions into the TCP model. The data that is
23 required is the number of clonogenic cells No,i that receive a dose Di. Conveniently this is
24 exactly the information contained in a Dose-Volume Histogram (DVH) generated by the
25 planning computer. Strictly what is required is the differential dose-volume distribution, dV/dD

PartG.doc Page 110 of 147 13 January, 2004


Part G Treatment Planning

1 from which the more familiar cumulative DVH is calculated. Thus one generalises Equation
2 G8.2 to:

n
Ns = N o ,i exp[−α Di ]
i =1

3 where the summation is carried out over the n bins in the DVH. This expression should be also
4 used in Equations G8.4 and G8.5 in order to take account of the effect of both dose
5 inhomogeneities and inter-patient α variability (Webb and Nahum 1993).

6 The full expression for TCP can then be written as follows:

[ ]

1
7 TCP =
σ α 2π
∏ exp[− ρ V exp(−αD )]
i
cl i i exp − (α − α ) 2 / 2σ α2 dα
o

8 where the summation (of the indices of the exponential) over the i dose bins in the DVH has
9 been replaced by a multiplication of exponentials, ρcl is the clonogenic cell density, Vi is the
10 volume of tissue in the ith dose bin, and is the mean value of the radiosensitivity in the
11 interpatient distribution, characterised by standard devation σα .

12 G8.1.3. The predicted effect of dose non uniformity

13 Brahme (1984) applied his TCP model to the question of the effect on TCP of both
14 inhomogeneities in the target dose distribution and also uncertainties in the absolute absorbed
15 dose determination. A similar exercise has been carried out here for the particular case of the
16 bladder tumour parameters given above (Figure G 8.4).

17 The dose inhomogeneity in the target volume, consisting of 109 clonogenic cells of
18 uniform radiosensitivity, was assumed to follow a normal distribution i.e. No,i was varied
19 normally as a function of Di (Equation G8.7); with its width expressed in terms of σD/D (x-axis
20 in Figure G 8.4). The results of this exercise are shown in Figure G 8.4. The different curves
21 correspond to different values of σα The mean dose was set to 60 Gy.

22

PartG.doc Page 111 of 147 13 January, 2004


Part G Treatment Planning

1
2 Figure G 8.4
3 σ
4 σ ! " #$% &
5 "' ( ) & '(( *

6 One immediate observation is that the TCP always decreases as σD/D increases, even
7 though the mean dose does not change. Thus the maximum value of the TCP is for perfectly
8 uniform dose, and “hot” and “cold” spots do not compensate each other, as has been pointed out
9 by many workers (e.g. Sanchez-Nieto and Nahum 1999). The figure shows us that for a group
10 of patients with tumours of exactly the same radiosensitivity i.e. σα= 0.0 even small
11 inhomogeneities in dose have a disastrous effect on the TCP; this corresponds to the very steep
12 dose-response curves in Figure G 8.2. More realistic values of σα e.g. 0.10, result in a much less
13 dramatic reduction in TCP as the dose inhomogeneity is increased. The message is that the
14 appreciable inter-patient variability in radiosensitivity indicated clinically for many types of
15 tumours considerably reduces the consequences of deviations from target dose uniformity. The
16 corollary of this is the conclusion reached by Brahme (1984) that for certain classes of tumours
17 with steep dose-response curves, notably in the larynx (the normalised dose gradient γ > 4), only
18 very small uncertainties in the absolute dose determination can be tolerated.

19 Sanchez-Nieto and Nahum (1999) have suggested a method whereby a TCP model
20 could be used directly in a treatment planning system. This is done by assigning each dose bin
21 in the (differential) DVH the quantity ∆TCP (or deltaTCP) which is the amount by which the

PartG.doc Page 112 of 147 13 January, 2004


Part G Treatment Planning

1 TCP is changed as a result of the volume of tumour/target in this dose bin not being equal to the
2 prescribed dose. In this way the effect of “cold” bins can be seen directly.

3 G8.1.4. Variation in clonogenic cell density

4 If the model as described thus far is applied to the DVH of the target volume (the PTV in
5 ICRU50 terminology) then implicitly the assumption is made that the clonogenic cell density is constant
6 over the whole of the PTV i.e. one calculates the number of clonogenic cells at dose Di, No,i, in Equation
7 G8.7 from the product of Vo,i and ρcl (see Equation G8.8). However, the PTV actually involves a margin
8 for microscopic spread plus a second margin for geometrically inaccuracies. Thus clearly the
9 assumption of constant ρcl is quite unrealistic. Figure G 8.5 illustrates this point.

10
11 Figure G 8.5 Schematic drawing illustrating the problem of the variation of clonogenic cell
12 density at the edge of the PTV; a hypothetical dose and cell density profile through the centre of
13 the PTV is shown

14 Whilst there is presently no clinical data on exactly how the cell density does vary
15 throughout a radiotherapy target volume one can use the model to assess the effect that such
16 variations might have on the predicted TCP. One way of looking at this is to calculate the
17 change in dose D that corresponds to a change in cell density ρcl when one requires that the
18 voxel control probability VCP remains unchanged for a given volume element of cells. Figure
19 G 8.6, taken from Webb and Nahum (1993), shows the iso-VCP dose corresponding to a
20 decrease in clonogenic cell density given by the (entirely hypothetical) dashed curve. The main
21 message is that for a considerable decrease in ρ the allowable dose decrease is very modest.
22 Similar conclusions were drawn in Brahme and Ågren (1987).

PartG.doc Page 113 of 147 13 January, 2004


Part G Treatment Planning

3 Figure G 8.6 The variation in the dose corresponding to the iso-VCP condition for a tumour
4 with a variation in clonogenic cell density given by the full curve (from Webb and Nahum
5 1993 with permission)
6

7 G8.1.5. The TCP model applied to a body of patient data

8 The TCP model described here has been applied to the DVHs for the Gross Target
9 Volumes (GTV; ICRU 1993) for prostate patients in a clinical trial on conformal prostate
10 therapy. Parameters appropriate to prostate tumours have been obtained fitting the TCP model
11 to the Hanks et al (1988) clinical data set; patients free from recurrence at 7-year follow-up
12 were used (Sanchez-Nieto and Nahum 1999). Two curves were calculated, corresponding to
13 (uniformly irradiated) tumour volumes of 10 cm and 100 cm which correspond very roughly to
14 stage B and C respectively. The parameters used to compute the curves (shown in Figure G
15 8.7a) are given in Figure G 8.7b (Sanchez-Nieto and Nahum 1999). Note that here the β term in
16 the LQ expression (Equation G8.1) has been included, where it has been assumed that α/β = 10
17 Gy (e.g. Steel 1997; Part B Chapter 1).

PartG.doc Page 114 of 147 13 January, 2004


Part G Treatment Planning

Figure G 8.7a TCP curves resulting Figure G 8.7b The distribution of TCP
from parameters chosen to fit the values for 30 patients treated in the
Hanks (1988) prostate survival data RMT prostate trial; using Parameters
(from Sanchez-Nieto and Nahum derived from the fit in Figure G 8.7a.
(1999) with permission).
3

4 Figure G 8.7b is a frequency distribution of the (population-averaged) TCP values for


5 30 patients treated in the non-escalated arm, at 64 Gy (in 2-Gy fractions). The overall mean
6 TCP came out at 64% with a range from around 55% to 70%. This spread is due to a
7 combination of the different sizes of the GTVs and the non-uniformity in the dose distributions,
8 though the former dominates in this particular case. It is emphasised that each TCP value really
9 represents a patient-averaged value, for σα = 0.07 Gy-1 and thus is not actually appropriate for
10 that individual patient, whose α−value is unknown. In principle one could simulate the «true»
11 situation by choosing an α value for each patient randomly from the normal distribution with
12 α = 0.29 and σα = 0.07; in this case the range of TCP values obtained would be greater.

13 It is sometimes believed that the probability of controlling a tumour depends on the


14 minimum dose in the target volume. The model can be used to examine this hypothesis. An
15 analysis was carried out on 43 typical treatment plans for carcinoma of the bladder and the TCP
16 calculated for the GTV of each plan using the same parameter values as for the “realistic” curve
17 in Figure G 8.3. For each plan the mean dose, the minimum dose and the effective dose Deff
18 were recorded. This latter dose was that value, which if uniform in the tumour, would give the
19 same TCP as the actual dose distribution; it follows from Figure G 8.4 that it must lie between
20 Dmin and Dmean . It was found that, on the average, Deff was much closer to Dmean than to Dmin .

PartG.doc Page 115 of 147 13 January, 2004


Part G Treatment Planning

1 From this it can be concluded that, at least for treatment plans where the tumour dose is
2 reasonably uniform (as was the case here), the mean dose is much better than the minimum
3 dose as a single indicator of clinical effectiveness.

4 A very interesting analysis has recently been published by Terahara et al (1999) on the
5 effect of tumour dose inhomogeneity on local control; 132 patients with skull base chordoma
6 were treated with combined photons and protons, with the prescribed doses ranging from 66.6
7 to 79.2 Cobalt-Gray-Equivalent. It was found that there was a poor correlation between local
8 control and measures such as mean dose and prescribed dose but a significant one (low p-value)
9 with minimum dose and with the quantity Equivalent Uniform Dose (EUD). The EUD,
10 proposed by Niemierko (1997), is close in concept to TCP; it is the uniform dose in the target
11 that gives the same level of cell kill and involves summing up the contributions of the dose bins
12 in the DVH in the same way as in Equation G8.7. Thus this Terahara et al study effectively
13 provides clinical support for the effect of inhomogeneous dose distributions on the TCP given
14 here.

15 G8.1.6. Concluding remarks

16 The TCP model described here is mathematically very simple and yet it is a reasonably
17 complete description of the process of tumour eradication by irradiation, apart from
18 proliferation (Deasy 1996) and re-oxygenation effects. The main limitations in its use are the
19 lack of clinical data on radiosensitivity and clonogenic cell density. Thus the absolute values of
20 TCP predicted must be treated with caution. There have been some recent attempts to fit clinical
21 data on patient survival and doses (often corrected for differences in fractionation) with TCP
22 models and hence to derive clinically reasonable parameter values. There is enormous variation
23 in the parameters obtained depending on, for instance, whether or not interpatient heterogeneity
24 is incorporated (Brenner 1993, Webb 1994, Fenwick 1998, Webb and Nahum 1998, Buffa et al
25 1999).

26 Despite the uncertainties in the parameters, this and similar models can be used to gain
27 insight into the effect on TCP of certain features of dose distributions. Such features include:

28 • dosimetrical uncertainties (Boyer and Schultheiss 1988),

29 • the width of field margins including the setting of MLC leaves (Brahme 1984, Webb 1993),

PartG.doc Page 116 of 147 13 January, 2004


Part G Treatment Planning

1 • dose inhomogeneities in the target (Goitein 1986, Goitein and Niemierko 1996, Niemierko
2 1997, Sanchez-Nieto and Nahum 1999, Terahara et al 1999) and in particular the effect of a
3 boost to part of the target volume (Goitein et al 1995),

4 • patient movement (Mackay et al 1999)

5 • differences in patient radiosensitivities.

6 It is to be hoped that the biological assays for the latter currently under development (Peters et al
7 1989; West et al 1991; Bentzen 1997) will provide data to enable treatments to be individualised
8 both biologically as well as physically (e.g. Nahum and Tait 1992; Sanchez-Nieto et al 1999).

9 G8.2. Normal Tissue Complication Probability (NTCP)

10 G8.2.1. Introduction

11 Normal tissue complication probability models have evolved in recent years mainly in
12 order to evaluate treatment plans. The notion of using some type of biological index to
13 characterize one or more features of a treatment plan was expounded early on by Dritchillo who
14 applied what he called a complication probability factor to rank rival treatment plans (Dritchillo
15 1978). In spite of early attempts such as this, the increasing interest in NTCPs has in large part
16 been a consequence of the development of 3D treatment planning systems and their use in
17 conformal therapy. The possibility of designing treatments with new and complex dose
18 distributions to reduce normal tissue toxicity, and possibly escalate the tumour dose, has
19 encouraged attempts at characterizing the merits of a treatment plan. This has become
20 especially important because the limited clinical experience with innovative dose distributions
21 makes it difficult to choose between one or another candidate treatment plan. NTCP is a
22 particularly important index for ranking treatment plans. Like other quantitative indices (e.g.
23 TCP), it reduces the large amount of data characterizing a three-dimensional dose distribution to
24 a limited number of numerical indices, and makes it more efficacious to compare a number of
25 rival plans. In addition, the NTCP nominally represents an endpoint which often determines the
26 acceptability of a treatment plan, that is whether the tolerance of normal tissues have been
27 respected or exceeded.

28 In spite of its importance, NTCPs should be used with great caution when evaluating
29 treatment plans, since the models suffer from a number of problems. To begin with, the models
30 are very crude and hardly even try to represent the multiple and interrelated toxicities observed

PartG.doc Page 117 of 147 13 January, 2004


Part G Treatment Planning

1 in clinical practice. In this respect, modeling of NTCP is much more complex than TCP
2 because of the large and complex taxonomy of treatment toxicities. And even when clinical
3 responses are reduced to the barest essentials, perhaps to one or a few critical endpoints, the
4 paucity of clinical data and their large uncertainties makes the it very difficult to rely on the
5 calculated complication probabilities. These uncertainties not only can lead to large variations
6 in the absolute values of the calculated NTCPs (Lebesque 1995) but they can also affect the
7 relative ranking of candidate treatment plans. However, this situation is changing; the increased
8 use of 3D planning by many institutions is beginning to provide large data bases of three-
9 dimensional dose distributions potentially correlated with clinical endpoints. Although the
10 endpoints characterising clinical complications, as is well known, are difficult to define and the
11 data are painstaking to collect, nevertheless, there is a continuing accumulation of such data.
12 There are also indications that the pooling of clinical data and the ability to share dose
13 distributions between institutions electronically will provide more robust data bases where
14 models can be refined. Some consensus may then be reached on the models and their
15 parameters, at least for some important organs and endpoints.

16 The NTCP models currently in use stem from two main lines of development, both of
17 which aim to predict the probability of a complication as a function of the dose (or biologically
18 equivalent dose) and volume. These models can be distinguished by their descriptions of the
19 volume effect, that is how the probability of a complication changes with volume irradiated at a
20 fixed dose. One of the lines of development has led to a phenomenological model which was
21 first formulated by Lyman (Lyman 1985; Lyman and Wolbarst 1987a; Lyman and Wolbarst
22 1987b) and later augmented by Kutcher and Burman (1989, Kutcher et al. 1991). This LKB
23 model seeks to describe complication probabilities using tolerance doses for different irradiated
24 volumes as input data. The tolerance doses for different volumes are related through a power
25 law, and thus the clinical data fixes the volume effect which can vary widely. In spite of this
26 flexibility, the model does not exhibit a threshold effect with volume and so the model would
27 not strictly apply to endpoints like radiation pneumonitis. It is possible, however, to augment
28 the model by introducing a 5th parameter, the so-called critical volume, below which there is no
29 complication.

30 A second line of approach begins by making certain assumptions about the functional
31 and anatomic organization of tissues and organs such that different volume effects are specific
32 to different functional organizations. For example, serial (critical element) models assume that

PartG.doc Page 118 of 147 13 January, 2004


Part G Treatment Planning

1 certain organs are arranged like links on a chain, and when one link (a functional subunit) is
2 damaged the entire chain is broken and a complication ensues (Schultheiss et al. 1983; Wolbarst
3 1984; Withers 1988; Niermerko and Goitein, 1991). Organs with this architecture reveal a
4 small volume effect. Parallel models (critical volume models) assume that a complication does
5 not occur until a significant portion of the independent functional sub-units (the functional
6 reserve) of an organ have been incapacitated (Wolbarst et al 1982, Jackson et al 1993, Yorke et
7 al 1993, Niermerko and Goitein 1993). The volume effect in these tissues is large, since a
8 complication does not occur if less than the functional reserve is irradiated.

9 G8.2.2. Lyman Kutcher Burman (LKB) empirical model

10 G8.2.2.1. Homogeneous Irradiation

11 Lyman (1985) argued that normal tissue complication probabilities depend upon volume
12 as well as dose, and that they could be conveniently represented by an error function in dose and
13 volume. For organs that are irradiated by a uniform dose to part of the volume and no dose to
14 the rest, the NTCP can be represented by:
15

exp (−t 2 2 )dt


t
16 NTCP = 1 G8.9
2π −∞
17

18 t = (D − TD50 (V / Vref )) / (m • TD50 (V / Vref )) G8.10


19

20 TD50 (1) = TD50 (V / V ref ) • (V / Vref )n G8.11


21

22 The 4 parameters in this model are:

23 • TD50 (1), the dose to the whole organ which would lead to complication in 50% of the
24 population;

25 • m, a parameter representing the steepness of the dose-response curve;

26 • Vref, a reference volume, which in many cases will be the organ volume;

27 • n, the exponent of volume in the power law that relates the tolerance doses for uniform
28 whole and uniform partial organ irradiation.

29 This latter parameter represents the volume effect: when n is near unity, the volume effect is
30 large and when it is near zero, the volume effect is small. As we shall find, a large volume effect
31 implies that the NTCP correlates with the mean dose, while a small volume effect implies that it

PartG.doc Page 119 of 147 13 January, 2004


Part G Treatment Planning

1 correlates with the peak organ dose. Moreover, for the architectural models that are described
2 below, the analogue of a parallel architecture organ would be one with a large value of n, and for
3 a serial organ, one with small value of n.

4 The NTCP for partial organ irradiation in the Lyman model is based upon clinical
5 estimates of organ tolerance for partial uniform irradiation at different doses and volumes. Such
6 data is difficult to come by. Nevertheless, Emami et al. (1991) estimated tolerance doses from
7 the literature and unpublished experience. They chose to transform the data into one-third, two-
8 thirds and whole organ tolerances, rather than using the fractions reported in the literature, and
9 this forced them to interpolate the published data. The Emami partial organ tolerances were fit
10 to Equation G8.3 by Burman et al. (1991). The values of n, the volume effect, varied from near
11 zero for some organs, e.g. the spinal cord, up to about 0.7 for large organs like the liver and
12 lung. Since then, there have been a number of publications which have suggested changes in
13 the model parameters based upon new dose distributions and toxicity data, some of which will
14 be described in a later section.

15 G8.2.2.2. Inhomogeneous Irradiation

16 The above approach will apply only to a limited number of treatments, namely, those in
17 which the organ is treated with parallel opposed fields. For example, in head and neck cancer a
18 large portion of the parotid glands are often irradiated to near uniform doses with opposed
19 lateral fields, while the remainder receives a low (although non-zero) scatter dose. Multi-field
20 treatment techniques usually produce dose distributions which are inhomogeneous throughout
21 the organ so that the above modeling would not be applicable. Indeed, one of the difficulties
22 with the data culled from the literature in the Emami compilation is the condition of uniformity
23 was not necessarily met.

24 One way of extending the Lyman model to the more general inhomogeneous case is to
25 convert the organ'
s DVH into an "equivalent" uniform one using either interpolation (Lyman
26 and Wolbarst 1987b) or the so-called effective volume method (Kutcher and Burman 1989). If
27 this step is valid, then the dose and volume for the transformed uniform histogram can be used
28 to “read-off” the complication probability from the Lyman representation of the uniform partial
29 organ data.

30 In most instances both procedures lead to the same NTCP (Kutcher et al. 1991) and so
31 we describe the latter method. In the effective volume method, the normal organ’s DVH is

PartG.doc Page 120 of 147 13 January, 2004


Part G Treatment Planning

1 transformed into one in which the volume, Veff (which is equal to or less than the whole organ
2 volume) receives a dose equal to the peak organ dose (Dpeak). Veff is calculated from the
3 formula:
4
V i • ( Di / D peak )
1/ n
Veff =
5 i G8.12
6

7 Moreover, there is a family of equivalent uniform DVHs with effective volume and dose
8 related through the defining power law relationship. Thus there is an effective (Deff) or
9 equivalent uniform dose (EUD) which if applied to the whole organ (VT) would yield the same
10 NTCP.
11
1/ n Vi n
Deff = [Di • VT ]
12 i G8.13
13

14 This effective volume transformation is self-consistent with the power law model for
15 uniform irradiation in that it can be derived from just two hypotheses: the organ is
16 homogeneous in its response to radiation; and each element of the organ obeys the same power
17 law relationship as the whole organ (Kutcher and Burman 1989). The former condition is
18 clearly problematic. Most normal organs are not homogeneous in response. For example, the
19 nephrons of the kidney are not uniformly distributed, and the apex is less sensitive than the rest
20 of the lung.

21 G8.2.3. Organ Architecture Models

22 There have been a number of papers, quoted earlier, that have suggested that organ
23 architecture is related to complication probability. However, it was perhaps Withers who most
24 forcefully argued that it was worth considering that the response of organs to radiation might be
25 related to the organisation of its more basic units (Withers et al. 1988). The simplest
26 representation of an organ was that it was composed of independent functioning subunits
27 (FSUs) which can be defined either structurally (e.g. the nephrons of the kidney) or
28 operationally (e.g., the FSUs of the skin). Moreover, it was suggested that the FSUs had either
29 serial or parallel organizations. Since the FSUs are considered (by definition) to be independent
30 of one another, at least in their response to radiation damage, the probability of a complication
31 is amenable to statistical methods. In his paper, Withers applied these ideas to serial organs
32 with uniform partial irradiation and derived the complication probability (see Section G8.2.3.1)

PartG.doc Page 121 of 147 13 January, 2004


Part G Treatment Planning

1 The NTCP for inhomogeneous dose distributions for both serial or parallel organs has
2 been developed by Jackson et al. (1993) and Niermerko and Goitein (1991, 1993). We briefly
3 follow the reasoning of Jackson’s solution. Starting with a general inhomogenous dose
4 distribution, Jackson derived a risk histogram (and extension of a dose volume histogram)
5 which represented the family of fractional volumes of the organ in each of which the probability
6 of eradicating its FSUs was equal to or greater than a certain probability. Using the risk
7 histogram, he was able to calculate, by statistical methods, the probability of eradicating M and
8 only M FSUs, P(M). P(M) is a complicated function of the organ architecture and the radiation
9 conditions and the interested reader should refer to Jackson for further details. An important
10 point, however, is that once this function is known, it can be used to derive the complication
11 probability for serial or parallel normal organs as well as the control probability for tumours.
12 For example, NTCP is calculated by summing P(M) from a lower limit L (the minimum
13 number of eradicated FSUs required to realize a complication) to N (the total number of FSUs
14 in the organ).
15
N
16 NTCP = M=L
P(M) G8.14
17

18 If L=1, then we obtain a serial complication model where at least one FSU must be
19 eradicated to realize a complication. If L is greater than unity, we have a parallel complication
20 model in which at least L FSUs must be eradicated for a complication. Since the number of
21 FSUs is usually quite large, typically 104 to 109 (Thames and Hendry 1987), the derived dose
22 response curves are very steep. If we interpret this result as the dose response of an individual,
23 then it is possible to obtain more realistic dose response curves by averaging over a population
24 of patients. For example, we can average the NTCPs over a distribution of functional reserves
25 of an organ or we can average over intra or inter-organ distributions of radiosensitivities. Such
26 an approach leads to models with at least 4 parameters, two for the local response function, and
27 two for each population distribution, although less are possible for a serial model (Niermerko
28 and Goitein 1991).

29 G8.2.3.1. Serial Organs

30 The description above is somewhat formal and general. We can obtain some insight by
31 following Wither’s argument for a serial chain model. In the case of uniform irradiation of N of
32 the FSUs, and no radiation to the remainder, the probability of eradicating at least one is given
33 by one minus the probability of not eradicating any:

PartG.doc Page 122 of 147 13 January, 2004


Part G Treatment Planning

1
NTCP = 1− (1 − p)
N
2 G8.15
3
4 where p is the probability of eradicating a single FSU. This relationship yields a sigmoid curve
5 of NTCP versus dose. As the number of irradiated FSUs increases (that is, as the volume of
6 irradiated organ increases) the sigmoid response shifts to the left and the complication
7 probability increases. In addition, the partial volume dose response curves become steeper as the
8 volume increases. Niermerko and Goitein (1991) have also shown that the serial model can be
9 extended in a rather straightforward fashion for non-homogeneous irradiation, and that the serial
10 and the LKB models agree at low complication probability, that is in the clinically significant
11 domain. Finally, for the serial case, the whole organ dose response curve can be shown to
12 uniquely determine the volume effect of the organ.

13 G8.2.3.2. Parallel Organs

14 In parallel element tissues, which have also been modeled using binomial statistics for
15 homogenous irradiation (Yorke et al. 1993) and for inhomogeneous irradiation (Jackson et al.
16 1993, Niermerko and Goitein 1993), a complication occurs if the fraction of eradicated FSUs
17 exceeds a threshold fraction, the functional reserve of the organ. The kidney, liver, and lung are
18 conjectured to behave as parallel organs. Because the number of FSUs is always large in these
19 organs (Thames and Hendry 1987), the functional reserve of an organ can be defined by the
20 fraction rather than the number of eradicated FSUs. Moreover, as mentioned above it is likely
21 that this functional reserve varies over a population of patients. For example, liver function will
22 differ depending, for example, on the age and alcohol consumption of the patients. It is also
23 possible that radiosensitivities may vary over a population of patients or that there are intra-
24 organ variations, although Jackson (Jackson et al. 1993) has shown that intra-organ distributions
25 have negligible effects. For simplicity, if we assume that there is only a population distribution
26 of functional reserves, then the observed NTCP would be proportional to the fraction of patients
27 in which the local damage exceeds the functional reserve, that is
28
29 NTCP = H(f) G8.16
30

31 H is the cumulative distribution of functional reserves for the population and f is the fractional
32 damage of FSUs, that is
33
f = p(di) • (Vi / VT )
34 i G8.17
35

PartG.doc Page 123 of 147 13 January, 2004


Part G Treatment Planning

1 where Vi is the volume of the organ receiving dose di with probability of irradiation p(di).

2 Equations G8.16 and G8.17 can be interpreted as follows. To obtain NTCP we first
3 calculate the fractional damage from equation 16 using the dose volume histogram for the organ
4 and the local response function, p. The NTCP is obtained by calculating the fraction of the
5 population in which the fractional damage exceeds the functional reserve. This is nothing more
6 than the area under the differential distribution of functional reserves from zero to the fractional
7 damage, that is H(f) (see Figure G 8.8). To take one example, f might represent the fraction of
8 acini of the liver that have been irradiated, while H would represent the fraction of the patients
9 whose functional reserve for radiation hepatitis was fully depleted. In this representation, we
10 have a 4 parameter model where two of the parameters represent the distribution H (mean and
11 width) and two p(di).

12
13 Figure G 8.8 The calculation of NTCP in the parallel model in the limit where the
14 distribution of functional reserves is used to obtain NTCP for a population.

15 G8.2.3.3. Relationships Between LKB Serial and Parallel Models

16 Although the LKB and architecturally based models are derived in different ways they
17 yield similar results for NTCP in certain limits. For example, as we already mentioned, for
18 small values of n, the NTCP for the LKB model approaches that of the serial model. Thus, we
19 can think of organs with a small n as serial like in their behaviour. Moreover, for small n the
20 NTCP is determined primarily by the peak dose and is thus weakly correlated with the
21 irradiated volume. As n increases so does the influence of volume on the probability of a
22 complication. In the extreme, n=1, the NTCP correlates with the mean dose delivered to the

PartG.doc Page 124 of 147 13 January, 2004


Part G Treatment Planning

1 organ. For the two extremes, n=0 and n=1, to first approximation, the NTCP is correlated with
2 dosimetric variables. When evaluating treatment plans in this limit, we do not have to resort to
3 making any further assumptions about the biological models and the values of the parameters.
4 It therefore possible to use these two extremes as a starting point or rough approximation when
5 evaluating a treatment plan. We might, for example, use the mean dose when the lung, liver or
6 parotid glands are irradiated, and the peak dose for the spinal cord or small bowel.

7 It is important to note that organs with serial or parallel architecture have very different
8 volumes effects, that is how the NTCP changes with volume at fixed dose, as shown in Figure
9 G 8.9. For serial organs, the NTCP increases linearly with the irradiated volume, while for
10 parallel organs, the NTCP is zero up to a critical volume and then increases supralinearly.
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32

33 Figure G 8.9 Comparison of the volume effect, the change in NTCP with volume at fixed
34 dose, for serial (critical element) and parallel (critical volume) tissues with and without
35 accounting for variations in the model parameters. Th e NTCP for serial tissues show a
36 linear response with volume irradiated, while parallel tissues show a threshold behaviour.
37 (from Niermerko and Goitein 1993b)

38 Finally, Kwa has demonstrated a relationship between the LKB and parallel models
39 (Kwa et al. 1998a). For the parallel model if we assume that for moderate doses the probability
40 of killing an FSU can be represented by a power law in dose
41
1/ r
42 p(d) = const • d G8.18
43

PartG.doc Page 125 of 147 13 January, 2004


Part G Treatment Planning

1 then it is possible to show that Deff, or the effective uniform dose (EUD), can be represented by:
2
r Vi 1 / rn
Defff = [Di • VT ]
parallel

3 i G8.19
4

5 This expression is the same as that for the LKB model (Equation G8.16), if n= 1/r. The
6 assumption of a power law for p(d) is quite reasonable in the limit where the dose delivered is
7 smaller than that required for killing half the FSUs. We would expect that the above
8 correspondence between parallel and LKB models would break down at dose levels where there
9 is very high cell killing.

10 G8.2.4. Relationship of Models to Clinical Complications

11 Since the models are in an early stage of development, and the analysis of clinical
12 outcomes with different NTCP models has been rapidly changing, it would not be realistic to
13 present a detailed review of the applicability of the models. It is also not the purpose of this
14 section to describe the types and quality of data nor their uncertainties. Nevertheless it is
15 worthwhile to indicate the general direction work in this area has taken and to give some
16 indication of the use of clinical data in NTCP modeling. We will only review three sites rectum,
17 liver and lung since these had been importance for dose escalation studies. For an excellent
18 recent broader review for a number of sites see Lebesque (1999).

19 G8.2.4.1. Radiation Proctitits

20 Quantitative evidence for a volume effect in the rectum has changed rather significantly
21 due to interest in conformal therapy and dose escalation in prostate cancer. Earlier, Emami
22 estimated that the whole volume tolerance dose for 5% complications in the rectum was 60 Gy
23 (Emami et al. 1991). In addition, it was suggested that the same value be used no matter what
24 volume was irradiated, that is that there was a minimal volume effect at best. In contrast, Benk
25 et al. (1993) published a clinical study of 41 patients treated for prostate cancer with photons
26 and protons, and found a statistically significant increase in the probability of Grade II rectal
27 bleeding when more than 40% of the anterior portion of the rectal wall received at least 75
28 Cobalt Gray Equivalent. The prostate and pelvis was irradiated with photons, while the prostate
29 was boosted with a single proton field (Shipley et al. 1979) Since for the latter phase, the
30 rectum was immobilised with a stent, it was possible to know the position of the rectum and
31 therefore more accurately determine dose volume histograms for the rectal wall.

PartG.doc Page 126 of 147 13 January, 2004


Part G Treatment Planning

1 The presence of a volume effect has also been confirmed in other studies. For example,
2 the results of the dose escalation study by the group at Memorial Sloan-Kettering in New York
3 makes a compelling case for a rectal volume effect. They have been able to raise the dose to the
4 prostate to 81 Gy and more while keeping rectal bleeding below levels commonly found at
5 much lower doses ( Leibel et al. 1994; Zelefsky et al. 1995). This has been possible since the
6 dose escalation has been constrained by limiting rectal irradiation such that no more than 30%
7 of the rectal wall exceeds 75 Gy (Kutcher et al. 1996). Moreover, Hanks has also found that
8 rectal bleeding is reduced when some of the rectum is blocked in the boost phase of the
9 treatment (Schultheiss et al. ?). There still remain a number of questions. How does rectal
10 motion influence the correlation of dose and volume with rectal bleeding? What is the relative
11 influence on rectal bleeding between increasing the length of the circumference of irradiated
12 rectal wall? If we consider the rectum in two extremes, one in which the FSUs of the rectum
13 are arranged as in Figure ?a, then full circumferential radiation would be required for a
14 complication, while if the FSUs are arranged as in Figure ?b, then even partial circumferential
15 irradiation would be sufficient. See Dale et al. (1999) for further discussion on this point..

16 G8.2.4.2. Radiation Hepatitis

17 One of the most informative examples of how NTCPs models and data can be
18 successfully used is by Jackson and colleague for radiation hepatitis. They used clinical
19 outcomes in 93 patients treated for tumours of the liver, dose volumes histograms, the parallel
20 model and the method of maximum likelihood to determine the model parameters (Lawrence et
21 al .1990; Lawrence et al. 1992; Jackson et al. 1995). They assumed that the organ has a
22 functional reserve described by two parameters (whose values were obtained from the
23 maximum likelihood fit) and that NTCP is given by the integral of the functional reserve up to
24 the mean fraction of eradicated FSUs, that is up to the fraction damaged. For each patient the
25 fraction damaged was calculated by summing the product of the fractional volume of each
26 voxel of the organ times the probability of damage. The latter was calculated using each
27 patient'
s DVH and an assumed local response function with two parameters (whose values were
28 also obtained from the maximum likelihood fit).

29 The method of maximum likelihood was used by first assigning a best guess for the
30 model parameters. The predicted probability of a complication for each patient was then
31 compared against the observed grade of complication in that patient. The overall likelihood L of
32 the observations was taken as:

PartG.doc Page 127 of 147 13 January, 2004


Part G Treatment Planning

1
2 L( γ 1 , γ 2 ,..) = ∏ Om (t m , γ 1 , γ 2 ,..) ∏ (1- On (t n ,γ 1 , γ 2 ,..)) (9)
m n
complication no complication
3

4 where
5
6 O m (t m , γ 1 , γ 2 ,.. )= o(t m )NTCPm ( γ 1 , γ 2 ,..) , (10)
7

8 where L is the likelihood, o(tm) is the probability that a complication will manifest itself after the
9 follow-up time tm for the mth patient, calculated from complication and follow up time data with
10 the Kaplan-Meier method (Kaplan and Meier 1958), and γ 1 , γ 2 ,.. indicates the model
11 parameters. The likelihood is a product over all patients of the chance that there is a
12 complication multiplied by the chance there is no complication, which is maximised with respect
13 to the model parameters.

14 The results of applying such an analysis to the hepatitis data is shown in figure 11 for
15 observed complication rate as a function of calculated fraction damaged of the liver. The
16 observed complications show a threshold effect with a steep response which is described within
17 the parallel model. The best fit for the functional reserve distribution (and confidence intervals)
18 predict a threshold volume of about one-third must be exceeded before complications are
19 observed. Moreover, complications were seen only in the cohort of patients who received
20 whole liver irradiation as part of the treatment course. The lack of complications among
21 patients given only partial volume irradiation suggests that an attempt to increase local control
22 through dose escalation is feasible for these patients. Moreover the observation of
23 complications in this subset of patients would likely reduce the correlated uncertainties found
24 for the model parameters.

25 G8.2.4.3. Radiation Pneumonitis

26 Convincing evidence of a large volume effect for pneumonitis in mouse lung has been
27 published by Liao, Travis and Tucker (Liao et al. 1995)). They demonstrated that mice with
28 70% of the lung irradiated to 20 Gy in one fraction were all dead after 28 weeks. In contrast,
29 those with 40% of the lung irradiated to 20 Gy all survived. Additionally, lethality occurred at
30 lower doses for mice given whole lung irradiation than for mice with 70% of the lung irradiated.
31 This data strongly suggests that a parallel model may be appropriate for pneumonitis, since a
32 tolerance dose for death after 28 weeks was not found for mice given 40% lung irradiation.

PartG.doc Page 128 of 147 13 January, 2004


Part G Treatment Planning

1 They have also argued that the base of mouse lung is more radiosensitive than the apex.
2 If such a large difference were seen in clinical data, then for equal numbers of tumours in the
3 base and apex, when the overall complication rate is 20%, we would estimate that the
4 complication rate for tumours in the base of the lung would be on the order of 30% higher that
5 for tumours in the apex. Clinically, evidence to support this observation comes from Martel et
6 al. (1994), who reported that all nine lung cancer patients with pneumonitis ( ≥ Grade I) had
7 tumours in the middle or lower lobes, and Graham et al. (1994), who found that incidence of
8 ≥ Grade III pneumonitis was 6% in 49 patients with upper lobe tumours and 29% in 21 patients
9 with middle or lower lobe tumours.

10 The clinical evidence for lung tolerance prior to conformal radiation therapy has been
11 reviewed by Travis (1992) and Marks (1994). They discuss, for example, that whole lung
12 volume tolerance doses which were established for patients treated with hemi-body irradiation
13 (Van Dyk et al. 1981), after corrections for fractionation, are generally consistent with tolerance
14 doses deduced from patients treated with various fractionation schedules to volumes in excess
15 of 75% (Wara et al. 1973). Few quantitative data for partial volume tolerances were available
16 before the advent of conformal treatment techniques. However, there were indications that the
17 probability of pneumonitis is related to the volume of lung irradiated (Perez et al. 1980; Seydel
18 et al. 1985, Rothwell et al. 1985, Brady et al. 1965a, Brady et al. 1965b). Emami et al. (1991)
19 in their review of the literature concluded that the volume effect in the lung was very large, and
20 the Burman et al. (1991) fit yielded an n of 0.87.

21 Kwa et al. (1998b) has looked at the incidence of pneumonitis in 540 patients from 12
22 institutions. Since the patients with lung cancer showed higher complication rates at lower
23 doses, the patients were divided between lung cancer and others, mainly breast and lymphoma.
24 They found that the rate of grade 2 and higher complications for each group could be fit to the
25 Lyman model with n=1 (Figure G 8.5), suggesting that there was a large volume effect in lung,
26 and that the complications were correlated with mean dose. However, Niermerko has argued
27 that mean lung dose is a weak predictor of radiation pneumonitis (Niermerko, 1999). Overall,
28 the situation remains fluid; different groups for different applications suggest different
29 endpoints for predicting damage. For example, the EORTC suggests mean dose, the RTOG,
30 the percent volume receiving greater than 20 Gy, the University of Michigan, the effective
31 volume, for choosing dose escalation bins, and the group at Memorial has used NTCP (Fowler
32 1991, Graham and Purdy, 1998).

PartG.doc Page 129 of 147 13 January, 2004


Part G Treatment Planning

1 G8.2.5. Conclusions

2 We have described two major methods of describing complication probabilities, namely,


3 an empirical method and secondly an architectural description. The volume effect is central to
4 how normal organs behave in response to inhomogeneous distributions of radiation which has
5 important implications for the design of treatment strategies. In tissues with a small volume
6 effect, it may be advantageous to use many fields in order to distribute the dose over a larger
7 volume and thereby reduce the peak dose. On the other, in hand in parallel organs, it would
8 accordingly be better to use a small number of fields keep fractional damage below the organ'
s
9 functional reserve. Intensity modulation may prove useful in adjusting the volumetric
10 distribution to normal organs while maintaining the intended dose to the large volume. In the
11 future we may find that organ architecture will play a larger role in designing, in addition to
12 evaluating treatment plans (Kutcher et al. 1994)

13 Since the NTCP and TCP models described here are still rather crude, some of the
14 discussion in this chapter should be seen as preliminary and suggestive for further investigation.
15 These models retain a certain plausibility and ability to connect the radiation conditions at the
16 local level to responses of whole organs that can be measured in the clinic. In this sense, it is
17 worthwhile to continue to develop and refine the models using the outcomes from a number of
18 trials of conformal therapy and dose escalation. The models also suggest the value of altered
19 dose distributions in both tumour and normal tissues whose value can be determined by further
20 clinical studies.

21 Note the following figure captions exist and have not been referred to:

22 Figure 10. Two possible arrangements of FSUs for the rectum: (a) serially arranged along the
23 length and (b) serially arranged around the circumference of the rectal wall. (from Dale et al. 1999)
24
25 Figure 11. The observed complication rate for radiation hepatitits as a function of calculated
26 fraction of damaged liver (from Jackson et al. 1995)
27
28 Figure 12. The incidence of radiation pnemonitisin as a function of mean dose corrected
29 for fractionation (NTDmean) for a population of 504 patients from 5 institutions. (a) all
30 lymphoma and breast patients and lung patients from one of the institutions (N=264). (b) all
31 other lung patients (N=276) (from Kwa et al. 1998)

PartG.doc Page 130 of 147 13 January, 2004


Part G Treatment Planning

1 References
2 Sonoda M, Akano M, Miyahara J and Kato H 1983 Computed radiography utilising scanning laser stimulated
3 luminescence Radiology 148: 833-8
4 Alpert, N. M., Bradshaw, J. F., Kennedy, D. and Correia, J. A. 1990 The principal axes transformation - A
5 method for image registration J Nucl Med 31 1717-1722
6 American Association of Physicists in Medicine, AAPM Task Group 25 1991, Clinical electron beam
7 dosimetry Med Phys 18 73-109
8 American Association of Physicists in Medicine, AAPM Task Group 53 1998, Quality assurance for clinical
9 radiotherapy treatment planning Med Phys 25 1723-1829
10 Ashtari, M., Zito, J. L., Gold, B. I., Lieberman, J. A., Borenstein, M. T., et al. 1990 Computerised volume
11 measurement of brain structure Invest Radiol 25 798-805
12 Austin-Seymour MM, Chen GTY, Castro JR, Saunders WM, Pitluck S, Woodruff KH, Kessler M (1986) Dose
13 volume histogram analysis of liver radiation tolerance. Int.J.Radiat.Oncol.Biol.Phys. 12:31-35
14 Bakker, C. J. G., Moerland, M. A., Bhagwandien, R. and Beersma, R. 1992 Analysis of machine-dependant and
15 object-induced geometric distortion in 2DFT MR imaging Magn Reson Imaging 10 597-608
16 Balter J. M., Kwok L. L., Sandler H. M., Littles J. F., Bree R. L., Ten Haken R. K. Automated localisation of the
17 prostate at the time of treatment using implanted radio-opaque markers: technical feasibility. Int. J. Radiat.
18 Oncol. Biol. Phys. 1995a 33: 1281 – 1286.
19 Balter J. M., Ten Haken R. K., Lawrence T. S., Lam K. L., Robertson J. M. Uncertainties in CT-based radiation
20 therapy treatment planning associated with patient breathing. Int. J. Radiat. Oncol. Biol. Phys. 1996 36: 167
21 – 174.
22 Balter J.M., Sandler H.M., Lam K., Bree R.L., Lichter A.S., Ten Haken R.K. Measurement of prostate
23 movement over the course of routine radiotherapy using implanted markers. Int. J. Radiat. Oncol. Biol.
24 Phys. 1995b 31: 113 – 118.
25 Barentsz, J. O., Jager, G., Mugler, J. P. I., Oosterhof, G., Peters, H., et al. 1995 Staging urinary bladder cancer:
26 Value of T1-weighted three-dimensional magnetisation prepared-rapid gradient-echo and two-dimensional
27 spin-echo sequences Am J Roentgenol 164 1 109-115
28 Batterman J J, Hart A A M and Breur K (1981) Dose-effect relations for tumour control and complication rate after
29 fast neutron therapy for pelvic tumours. Br. J. Radiol. 54 899-904
30 Battista, J. 1980 Computed tomography for radiotherapy planning Int J Radiat Oncol Biol Phys 6 99-104
31 Bedford J L, Khoo V S, Oldham M, Dearnaley D P, Webb S. A comparison of coplanar four-field techniques for
32 conformal radiotherapy of the prostate. Radiother. Oncol. 51 225 – 235. 1999.
33 Benk V A, Adams J A, Shipley W U, Urie M M, McManus P L, Efird J T, Willett J C, Goitein M, (1993) Late
34 rectal bleeding following combined x-rays and proton high dose irradiation for patients with stages T3-T4
35 prostate carcinoma Int. J. Radiat. Oncol. Biol. Phys. 26 551-557
36 Bentzen S M (1997) Potential clinical impact of normal-tissue intrinsic radiosensitivity testing Radioth. Oncol. 43
37 121-131.
38 Bentzen S M, Thames H D and Overgaard J (1990) Does variation in the in vitro cellular radiosensitivity explain the
39 shallow clinical dose-control curve for malignant melanoma? Int. J. Radiat. Biol. 57 117-126
40 Bhanalaph, T., Varkarakis, M. J. and Murphy, G. P. 1974 Current status of bilateral adrenalectomy on advanced
41 prostatic carcinoma Annals of Surgery 179 17-23
42 BJR Supplement 25 Central Axis Depth Dose Data for use in Radiotherapy British Institute of Radiology 1996.
43 Bomans, M., Hohne, K. H., Tiede, U. and Riemer, M. 1990 3-D segmentation of the head for 3-D display. IEEE
44 Transactions on Medical Imaging 9 177-183

PartG.doc Page 131 of 147 13 January, 2004


Part G Treatment Planning

1 Boyer A L and Schultheiss T (1988) Effects of dosimetric and clinical uncertainty on complication-free local tumor
2 control. Radioth. Oncol. 11 65-71
3 Brady L W, Cancer L, Evans G C, Forest D S (1965) Carcinoma of the lung: results of supervoltage radiation
4 Archives of Surgery 90 90-94
5 Brady L W, German P A and Cancer L (1965) The effects of radiation therapy on pulmonary function in
6 carcinoma of the lung Radiology 85 130-134
7 Bragg, D. G. and Osborn, A. G. 1991 CNS imaging of neoplasms International Journal of Radiation Oncology
8 Biology Physics 21 841-845
9 Brahme A (1984) Dosimetric precision requirements in radiation therapy. Acta Radiol. Oncol. 23 379-3910
10 Brahme A and Ågren A-K (1987) Optimal Dose Distribution for Eradication of Heterogeneous Tumors Acta Oncol.
11 26 377-385
12 Brant-Zawadzki, M. N., Gillian, G. D., Atkinson, D. J., Edalatpur, A. and Jensen, M. 1993 Three-dimensional
13 MR imaging and display of intracranial disease: improvements with the MP-RAGE sequence and
14 gadolinium J Magn Reson Imaging 3 4 656-662
15 Brenner D J (1993) Dose, volume and tumour-control predictions in radiotherapy Int. J. Radiat. Oncol. Biol. Phys.
16 26 171-179.
17 Brown AP, Urie MM, barest G, Cheng E, Coia L, Emami BN, Galvin J, Kutcher J, Manolis J, Wong JW,
18 Yahalom J (1991) Three-dimensional treatment planning for Hodgkin’s disease.
19 Int.J.Radiat.Oncol.Biol.Phys. 21:205-215
20 Brown, A. P., Urie, M. M., Barest, G., Cheng, E., Coia, L., et al. 1991 Three-dimensional photon treatment
21 planning for Hodgkin' s disease Int J Radiat Oncol Biol Phys 21 1 205-215
22 Buchali, A., Dinges, S., Koswig, S., Rosenthal, P., Salk, S., et al. 1998 [Virtual simulation. First clinical results
23 in patients with prostate cancer] Strahlenther.Onkol. 174 2 88-91
24 Buffa FM, Fenwick JD and Nahum AE (1999) Letter to the Editor «A realistic closed-form radiobiological model
25 of clinical tumor-control data incorporating inter-tumor heterogeneity» by Roberts and Hendry Int. J. Rad.
26 Onc. Biol. Phys 41 689-699 1998.
27 Burman C, Kutcher G J, Emami B, Goitein M (1991) Fitting normal tissue tolerance data to an analytic function
28 Int. J. Radiat. Oncol. Biol. Phys. 21 123-135
29 Casebow M P. The angulation of radiotherapy machines in the treatment of inclined lesions. Br. J. Radiol. 49
30 278 – 280 1976.
31 Casebow MP, 1984 “Matching of adjacent radiation beams for isocentric radiotherapy” Brit.J.Rad. 57: 515-518.
32 Chaney EL, Thorn JS, Tracton G, Cullip TJ, Rosenman JG, and Tepper JE, 1995 “A portable software tool for
33 computing digitally reconstructed radiographs” Int. J. Oncol. Biol. Phys. 32 : 491-497
34 Chang, H. and Fitzpatrick, J. M. 1992 A technique for accurate magnetic resonance imaging in the presence of
35 field inhomogeneities IEEE Transactions on Medical Imaging 11 319-329
36 Chen GTY, Austin-Seymour MM, Castro JR, Collier JM, Lyman JT, Pitluck S, Saunders WM, Zink SR Dose
37 volume histograms in treatment planning evaluation of carcinoma of the pancreas In: Proceedings, Eight
38 International Conference on Uses of Computers in Radiation Therapy, IEEE, ISBN 0-8186-0559-6
39 1984:264-268
40 Cheng CW, Chin LM and Kijewski PKA, 1987 “Coordinate Transfer of Anatomical Information from CT to
41 Treatment Simulation” Int. J. Oncol. Biol. Phys. 13:1559-1569
42 Clarke, L. P., Velthuizen, R. P., Camacho, M. A., Heine, J. J., Vaidyanathan, M., et al. 1995 MRI segmentation:
43 Methods and application Magn Reson Imaging 13 343-368
44 Clarke, L. P., Velthuizen, R. P., Phuphanich, S., Schellenberg, J. D., Arrington, J. A., et al. 1993 MRI: Stability
45 of three supervised segmentation techniques Magn Reson Imaging 11 95-106

PartG.doc Page 132 of 147 13 January, 2004


Part G Treatment Planning

1 Coia LR, Galvin J, Sontag M, Blitzer P, Brenner H, Cheng E, Doppke K, Harms W, Hunt M, Mohan R,
2 Munzenrider J, Simpson J (1991) Three-dimensional photon treatment planning in carcinoma of the larynx.
3 Int.J.Radiat.Oncol.Biol.Phys. 21:183-192
4 Consensus Conference 1988 Magnetic resonance imaging JAMA 259 2132-2138
5 Conway J and Robinson MH, 1997 “CT virtual simulation” Brit. Journ. Rad., Special Issue, Nov.
6 Cullip TJ, Symon JR, Rosenman JG, and Chaney EL, 1995 “Digitally reconstructed fluoroscopy and other
7 interactive volume visualisations in 3D treatment planning” Int. J. Oncol. Biol. Phys. 27: 145-151
8 Curren, W. J., Hackney, D. B., Blitzer, P. H. and Bilaniuk, L. 1986 The value of magnetic resonance imaging in
9 treatment planning of nasopharyngeal carcinoma International Journal of Radiation Oncology Biology
10 Physics 12 2189-2196
11 Dale E, Olsen D R, Fossa S D (1999) Normal tissue complication probabilities correlated with late effects in the
12 rectum after prostate conformal radiotherapy Int. J. Radiat. Oncol. Biol. Phys 43 385-391
13 Dawant, B. M., Zijdenbos, A. P. and Margolin, R. A. 1993 Correction of intensity nonuniformity in MR images
14 for computer-aided tissue classification IEEE Trans Med Imaging 12 770-781
15 De Gersem W.R.T., Derycke S., Colle C.O., De Wagter C. and De Neve W.J. (1999) Inhomogeneous Target-Dose
16 Distributions: a Dimension more for Optimization? Int. J. Radiat. Oncol. 44 461-468.
17 De Salles, A. W. T., Abe, M. and Kjellberg, R. N. 1987 Transposition of target information from the magnetic
18 resonance and computed tomography scan images to conventional X-ray stereotactic space Appl
19 Neurophysiol 50 1-6 23-32
20 De Vries, L. S. and Bydder, G. M. 1988. Tumors of the central nervous system. Magnetic Resonance Imaging
21 (MRI). C. C. Partain, R. R. Price and J. A. Patton. Philadelphia, W. B. Saunders: 144-168
22 Deacon J, Peckham M J and Steel G G (1984) The radioresponsiveness of human tumours and the initial slope of the
23 cell survival curve. Radiother. Oncol. 2 317-323.
24 Deasy J (1996) Poisson formulas for tumor control probability with clonogen proliferation Rad. Res. 145 382-384.
25 Delannes, M., Malavaud, B., Douchez, J., Bonnet, J. and Daly, N. J. 1992 Iridium-192 interstitial therapy for
26 squamous cell carcinoma of the penis Int J Radiat Oncol Biol Phys 24 3 479-483
27 Dobbs J, Barratt A and Ash D Practical Radiotherapy Planning (Edward Arnold) 3rd Edition, 1999
28 Dobbs J, Barrett A, and Ash D, 1999 “Practical Radiotherapy Planning” third edition, Edward Arnold.
29 Dong L and Boyer A 1995 ”An image correlation procedure for DRRs and electronic portal images” Int. J.
30 Oncol. Biol. Phys. 33 : 1052-1060
31 Dritschilo A, Chaffer J T, Bloumon W D, Marck A (1978) The complication probability factor: A method of
32 selection of treatment plans. Brit. J. Radiol. 51 370-374
33 Drzymala R, Holman MD, Yan D, Harms WB, Vain NL, Kahn MG, Emami B and Purdy J (1994) Integrated
34 software tools for the evaluation of radiotherapy treatment plans. Int.J.R.O.B.P. 30 (4): 909-919
35 Drzymala RE, Mohan R, Brewster MS, Chu J, Goitein M, Harms W, Urie M (1991) Dose volume histograms.
36 Int.J.R.O.B.P. 21:71-78
37 Dunscombe, P., Leszczynski, K., Cosby, S. and Krispel, F. 1994 Technical note: an aid to radiation therapy
38 simulation Br.J.Radiol. 67 799 649-650
39 Dutreix J, Tubiana M and Dutreix A (1988) An approach to the interpretation of clinical data on the tumour control
40 probability-dose relationship Radiother. Oncol. 11 239-248
41 Ed. Williams J R and Thwaites D I Radiotherapy Physics in Practice (Oxford Medical Publications)
42 Edelman, R. R., Wielopolski, P. and Schmitt, F. 1994 Echo-planar MR imaging Radiology 192 600-612
43 Ehricke, H. H. and Schad, L. R. 1992 MRA-guided stereotactic radiation treatment planning for cerebral
44 angiomas Computer Medical Imaging and Graphics 16 65-71
45 Emami B, Lyman J T, Brown A, Coia L, Goitein M, Munzenrider J E, Shank B, Solin L J, Wesson M (1992)
46 Tolerances of normal tissues to theraputic radiation Int. J. Radiat. Oncol. Biol. Phys. 21 109-122

PartG.doc Page 133 of 147 13 January, 2004


Part G Treatment Planning

1 Emami B, Purdy JA, Manolis J, Barest G, Cheng E, Coia L, Doppke K, Galvin J, LoSasso T, Matthews J,
2 Munzenrider J, Shank B (1991) Three-dimensional treatment planning for lung cancer.
3 Int.J.Radiat.Oncol.Biol.Phys. 21:217-227
4 Emami B., Lyman J., Brown A., Coia L., Goitein M., Munzenrider J. E. Int. J. Tolerance of normal tissue to
5 therapeutic radiation. Radiat. Oncol. Biol. Phys. 1991 21: 109 – 122.
6 Ende, G., Treuer, H. and Boesecke, R. 1992 Optimisation and evaluation of landmark-based image correlation
7 Phys Med Biol 37 261-271
8 Evans, P. M., Gildersleve, J. Q., Morton, E. J., Swindell, W., Coles, R., et al. 1992 Image comparison techniques
9 for use with megavoltage imaging system British Journal of Radiology 65 776 701-709
10 Fenwick J D (1998) Predicting the radiation control probability of heterogeneous tumour ensembles: data analysis
11 and parameter estimation using a closed-form expression Phys. Med. Biol. 43 2159-2178
12 Fernandes EM, Shentall GS, Mayles WPM, and Dearnaley DP, 1995 “The acceptability of multileaf collimator
13 as a replacement for conventional blocks” Radioth. and Oncol. 36: 65-74
14 Feygelman V, Mandelzweig Y and Baral E 1994 Matching electron beams without secondary collimation for
15 treatment of extensive recurrent chest-wall carcinome Med. Dosimetry 19 23-27
16 Finnigan, D. I., Tanner, S. F., Dearnaley, D. P., Edser, E., Horwich, A., et al. 1997 Distortion-corrected magnetic
17 resonance images for pelvic radiotherapy treatment planning Quantitative Imaging in Oncology 3 72-76
18 Fischer J J and Moulder J E (1975) The Steepness of the Dose-Response Curve in Radiation Injury Radiology 117
19 179-184.
20 Flentje, M., Zierhut, D., Schraube, P. and Wannenmacher, M. 1993 Integration of coronal magnetic resonance
21 imaging (MRI) into radiation treatment planning of mediastinal tumors Strahlentherapie und Onkolgie 169
22 6 351-357
23 Forman, J. D., Mesina, C. F., He, T., Devi, S. B., Ben-Josef, E., et al. 1993 Evaluation of changes in the location
24 and shape of the prostate and rectum during a seven week course of conformal radiotherapy Int J Radiat
25 Oncol Biol Phys 27 222
26 Fortier, G. A. and Pellitteri, M. R. 1990 MRI treatment planning for CNS neoplasm: graphic display of target
27 volume using boxed cursor Int J Radiat Oncol Biol Phys 19 179-182
28 Fowler J (1999) Personal communication
29 Fowler, J.F. (1989) The linear quadratic formula and progress in fractionated radiotherapy. The British Journal of
30 Radiology, 62: 679-694.
31 Fraass, B. A., McShan, D. L., Diaz, R. F., Ten Haken, P. K., Aisen, A., et al. 1987 Integration of magnetic
32 resonance imaging into radiotherapy treatment planning. I. Technical considerations Int J Radiat Oncol
33 Biol Phys 13 1897-1908
34 Gademann, G., Schad, L. R. and Schlegel, W. 1993. The definition of target volume in tumours of the brain,
35 based on skull and facial area by means of MRI: its impact on precision radiotherapy. Three Dimensional
36 Treatment Planning. P. Minet. Liege, European Association of Radiology: 47-55
37 Gademann, G., Schlegel, W., Debus, J., Schad, L., Bortfeld, T., et al. 1993 Fractionated stereotactically guided
38 radiotherapy of head and neck tumours: A report on clinical use of a new system in 195 cases Radiotherapy
39 and Oncology 29 2 205-213
40 Galvin, J.M. et al. “The use of digitally reconstructed radiographs for three dimensional treatment planning and
41 CT simulation” Int. J. Oncol. Biol. Phys. 31 : 935-942; 1995
42 Gilhuijs KGA, Drukker K, Touw A, Van de Ven PJH, and Van Herk M, 1996 “Interactive three dimensional
43 inspection of patient setup in radiation therapy using digital portal images and computed tomography data”
44 Int. J. Oncol. Biol. Phys. 34 :873-885
45 Gilhuijs KGA, Van de Ven PJH, and Van Herk M, 1996 “Automatic three dimensional inspection of patient
46 setup in radiation therapy using portal images simulator images, and computed tomography data”
47 Med.Phys. 23 :389-399

PartG.doc Page 134 of 147 13 January, 2004


Part G Treatment Planning

1 Glatstein, E., Lichter, A. S., Fraass, B. A. and van de Geijn, J. 1985 The imaging revolution and radiation
2 oncology: Use of CT, ultrasound and NMR for the localisation, treatment planning and treatment delivery
3 Int J Radiat Oncol Biol Phys 11 1299-1311
4 Goitein M and Niemierko A (1996) Intensity modulated therapy and inhomogeneous dose to the tumour: a note of
5 caution. Int. J. Radiat. Oncol. Biol. Phys. 36 519-522.
6 Goitein M, Abrams M, Rowell D. Pollari H, and Wiles J,1983, “Multidimensional treatment planning: II Beam’s
7 eye-view, back projection, and projection through CT sections” Int. J. Oncol. Biol. Phys. 9:789-797
8 Goitein, M. (1986) Causes and consequences of inhomogeneous dose distributions in radiation therapy. Int. J.
9 Radiat. Oncol. Biol. Phys. 12 701-704.
10 Goitein, M. 1979 The utility of computed tomography in radiation therapy; an estimate of outcome International
11 Journal of Radiation Oncology Biology Physics 5 1799-1807
12 Goitein, M., Niemierko, A. and Okunieff, P. (1995) The probability of controlling an inhomogeneously irradiated
13 tumour: a stratagem for improving tumour control through partial tumour boosting. In: Quantitative Imaging in
14 Oncology. Proceedings of the 19th L.H.Gray conference. Ed. Kaulner, K., Carey, B., Crellin, A. and Harrison,
15 R.M. p:25-32.
16 Goitein, M., Wittenberg, J., Mendiondo, M., Doucette, J., Friedberg, C., et al. 1979 The value of CT scanning in
17 radiation therapy treatment planning: a prospective study International Journal of Radiation Oncology
18 Biology Physics 5 1787-1798
19 Graham M and Purdy J A (1998) Int. J. Radiat. Oncol. Biol. Phys. 41973
20 Graham M V, Drzymala R E, Jain N L, Purdy J A (1994) Confirmation of dose-volume histograms and normal
21 tissue complication probability calculations to predict pulmonary complications after radiotherapy for lung
22 cancer 33rd Astro Meeting
23 Hanks G, Martz K L and Diamond J J (1988) The Effect of Dose on Local Control of Prostate Cancer Int. J. Radiat.
24 Oncol. Biol. Phys. 15 1299-1305
25 Hardy, T. L., Brynildson, L. R., Gray, J. G. and Spurlock, D. 1992 Three-dimensional imaging for brachytherapy
26 planning Stereotactic and Functional Neurosurgery 59 1-4 179-181
27 Hazle, J. D., Hefner, L., Nyerick, C. E., Wilson, L. and Boyer, A. L. 1991 Dose-response characteristics of a
28 ferrous-sulphate-doped gelatin system for determining radiation absorbed dose distributions by magnetic
29 resonance imaging (Fe MRI) Phys Med Biol 36 8 1117-1125
30 Heester, M. A., Wijrdeman, H. K., Strukmans, H., Witkamp, T. and Moerland, M. A. 1993 Brain tumor
31 delineation based on CT and MR imaging Strahlentherapie und Onkolgie 169 12 729-733
32 Heisig,S. Shental, G.S, Mirza K, Mayles WPM, 1994 “Applcation of the GE target planning computer to multi-
33 leaf collimator treatments”, ICCR ISBN 0 9523146 0 6. Ed. by Hounsell, Wilkinson and Williams
34 Hendrick, R. E. and Raff, U. 1992. Image contrast and noise. Magnetic Resonance Imaging. D. D. Stark and W.
35 G. Bradley. St Louis, Mosby Year Book: 109-145
36 Herrlin, K., Bi-Ling, L., Pettersson, H., Willen, H. and Rydholm, A. 1990 Gadolinium-DPTA enhancement of
37 soft tissue tumors in magnetic resonance imaging Acta Radiology 31 3 233-236
38 Hill D.L.G., Hawkes D.J., Crossman J.E., Gleeson M.J., Cox T.C.S., Bracey E.E., Strong A.J., Graves P.
39 Registration of MR and CT images for skull base surgery using point-like anatomical features. Br. J.
40 Radiol. 1991 64: 1030 – 1035.
41 Hill, D. L. G., Hawkes, D. J., Gleeson, M. J., Cox, T. C., Strong, A. J., et al. 1994 Accurate frameless
42 registration of MR and CT images of the head: applications in planning surgery and radiation therapy
43 Radiology 191 2 447-454
44 Hogstrom K R, Mills M D and Almond P R 1981 Electron beam dose calculations Phys. Med. Biol. 26 445-
45 449
46 Hricak, H. 1991. The role of imaging in the evaluation of pelvic cancer. Important Advances in Oncology. V. T.
47 DeVita, S. Hellmann and S. A. Rosenberg. Philadelphia, J B Lippincott: 103-133

PartG.doc Page 135 of 147 13 January, 2004


Part G Treatment Planning

1 IAEA 1987 Technical Report Series 277. Absorbed dose determination in photon and electron beams.
2 International Atomic Energy Agency, Vienna.
3 ICRU (1992) Prescribing, Recording, and Reporting Photon Beam Therapy ICRU Report 50 (Bethesda, MD:
4 International Commission on Radiation Units and Measurements)
5 ICRU 1978 Dose specification for reporting external beam therapy with photons and electrons. Report No. 29
6 International Commission on Radiation Units and Measurements, Bethesda, Maryland
7 ICRU 1984 Radiation dosimetry: electron beams with energies between 1 and 50 MeV Report No 35
8 International Commission on Radiation Units and Measurements, Bethesda, Maryland
9 ICRU 1993 Prescribing , recording and reporting photon beam therapy Report No 50 International
10 Commission on Radiation Units and Measurements, Bethesda, Maryland
11 ICRU 1999 Prescribing , recording and reporting photon beam therapy (supplement to ICRU Report 50). Report
12 No. 62. International Commission on Radiation Units and Measurements, Bethesda, Maryland
13 ICRU 50 1993: Prescribing, recording, and reporting photon beam therapy. Bethesda, Maryland.
14 ICRU 62 1999: Prescribing, recording, and reporting photon beam therapy (Supplement to ICRU report 50).
15 Bethesda, Maryland.
16 ICRU Report 50 Prescribing, Recording and Reporting Photon Beam Therapy.
17 IPEM 1999 Physics aspects of quality control in radiotherapy. IPEM, York
18 IPEM(B) (Institute of Physics and Engineering in Medicine (and Biology) 1994 A guide to commissioning and
19 quality control of treatment planning systems. IPEM, York
20 Jackson A, Kutcher G J, Yorke E D (1993) Probability of radiation induced complications for normal tissues
21 with parallel architecture subject to non-uniform irradiation Med. Phys. 20 613-625
22 Jackson A, Ten Haken R K, Robertson J M, Kessler M L, Kutcher G J, Lawrence T S (1995) Analysis of
23 Clinical Complication Data for Radiation Hepatitis using a Parallel Architecture Model Int. J. Radiat.
24 Oncol. Biol. Phys. 31 883-891
25 Jain NL, Kahn MG, Drzymala R E, Emami B and Purdy J A. (1993) Objective evaluation of 3D radiation
26 treatment plans: A decision-analytic tool incorporating treatment preferences of radiation oncologists.
27 Int.J.Radiat.Oncol.Biol.Phys 26: 321-333
28 Jain NL, Kahn MG. (1992) Ranking radiotherapy treatment plans using decision-analytic and heuristic
29 techniques Comp. Biomed. Res 25:374-383
30 Jette D 1996 Electron beam dose calculations. in Radiation therapy physics (ed. Smith A R) Springer-Verlag,
31 Berlin.
32 Judnick, J. W., Kessler, M. L., Fleming, T., Petti, P. and Castro, J. R. 1992 Radiotherapy technique integrates
33 MRI into CT Radiol Technol 64 2 82-89
34 Just, M., Rosler, H. P., Higer, H. P., Kutzner, J. and Thelen, M. 1991 MRI-assisted radiation therapy planning of
35 brain tumors - clinical experiences in 17 patients Magn Reson Imaging 9 2 173-177
36 Kagawa, K., Lee, W. R., Schultheiss, T. E., Hunt, M. A., Shaer, A. H., et al. 1997 Initial clinical assessment of
37 CT-MRI image fusion software in localization of the prostate for 3D conformal radiation therapy
38 International Journal of Radiation Oncology Biology Physics 38 2 319-325
39 Kalend A M, Zwicker R D, Wu A and Sternick E S 1985 A beam edge modifier for abutting electron fields Med
40 Phys 12 793-798
41 Källman P (1992) Optimization of Radiation Therapy Planning using Physical and Biological Objective Functions
42 PhD Thesis (Department of Radiation Physics, Stockholm University).
43 Kaplan E L and Meier P (1958) Nonparametric estimation from incomplete observations J. Am. Stat. Assoc. 53
44 457-816
45 Karlsson M and Zackrisson B 1997 Exploration of new treatment modalities offered by high energy (up to 50
46 MeV) electrons and photons Radioth Oncol 43 303-309

PartG.doc Page 136 of 147 13 January, 2004


Part G Treatment Planning

1 Karlsson, M. and Zackrisson, B. 1993 Matching of electron and photon beams with a multi-leaf collimator
2 Radiother.Oncol. 29 3 317-326
3 Kase K R and Bjärngard B E 1979 Bremsstrahlung dose to patients in rotational electron therapy Radiology
4 133 531-532
5 Kattapuram, S. V., Khurana, J. C., Scott, J. A. and El Khoury, G. Y. 1990 Negative scintigraphy with positive
6 magnetic resonance imaging in bone metastasis Skeletal Radiology 19 113-116
7 Kessler M L, McShan D L, and Fraass B A, 1995: “A computer-controlled conformal radiotherapy system III:
8 Graphical simulation and monitoring of treatment delivery” Int.J.Radiation Oncology Biol. Phys. 33 1173-
9 1180.
10 Kessler M.L, Pitluck S., Petti P., Castro J.R. Integration of multimodality imaging data for radiotherapy
11 treatment planning. Int. J. Radiat. Oncol. Biol. Phys. 1991 21: 1653 – 1667.
12 Kessler, M. L., Ten Haken, R., Fraass, B. and McShan, D. 1991 Expanding the use and effectiveness of dose-
13 volume-histograms for 3D treatment planning Med Phys 18 611
14 Khan F M, Fullerton G D, Lee J M F, Moore V C and Levitt S H 1977 Physical aspects of electron-beam arc
15 therapy Radiology 124 497-500
16 Khan, F. M. 1992. The Physics of Radiation Therapy. Baltimore, Williams and Wilkins
17 Khoo, V. S., Adams, E. J., Saran, F., Bedford, J. L., Perks, J. R., et al. 2000 A comparison of planning volumes
18 determined by CT versus MRI for meningiomas of the base of the skull International Journal of Radiation
19 Oncology Biology Physics 46 1309-1317
20 Khoo, V. S., Dearnaley, D. P., Finnigan, D. J., Padhani, A., Tanner, S. F., et al. 1997 Magnetic resonance
21 imaging (MRI): considerations and applications in radiotherapy treatment planning Radiotherapy and
22 Oncology 42 1 1-15
23 Khoo, V. S., Padhani, A. R., Tanner, S. F., Finnigan, D. J., Leach, M. O., et al. 1999 Comparison of MRI
24 sequences with CT for the radiotherapy planning of prostate cancer: a feasibility study British Journal of
25 Radiology 72 1-8
26 Klein E E, Zuofeng L and Low D A 1996 Feasibility study of multileaf collimated electrons with a scattering
27 foil based accelerator Radiotherapy and Oncology 41 189-196
28 Klevenhagen S C 1985 Physics of electron beam therapy Adam Hilger, Bristol
29 Klevenhagen S C, Lambert G D and Arbabi A 1982 Backscattering in electron beam therapy for energies
30 between 3 and 35 MeV Phys. Med. Biol. 27 363-373
31 Klien E E 1998 Modulated electron beams using multi-segmented multileaf collimation Radioth Oncol 48 307-
32 311
33 Kokoves, L. 1991 Assessment of geometrical distortion in MR images - investigation for the future applications
34 in radiotherapy treatment planning University of Surrey
35 Kooy H.M., Dunbar S.F., Tarbell N.J., Mannarino E., Ferarro N., Shusterman S., Bellerive M., Finn L.,
36 McDomough C.V., Loeffler J.S. Adaptation and verification of the relocatable Gill-Thomas-Cosman frame
37 in stereotactic radiotherapy. Int. J. Radiat. Oncol. Biol. Phys. 1994 30: 685 – 691.
38 Kooy, H. M., van Herk, M., Barnes, P. D., Alexander, E. I., Dunbar, S. F., et al. 1994 Image fusion for
39 stereotactic radiotherapy and radiosurgery treatment planning Int J Radiat Oncol Biol Phys 28 1229-1234
40 Korevaar E W, van Vliet R J, Woudstra E, Heijmen B and Huizenga H 1998 Sharpening the penumbra of high
41 energy electron beams with low weight narrow photon beams Radioth Oncol 48 213-220
42 Kovacs, G., Pötter, R., Prott, F. J., Lenzen, B. and Knocke, T. H. 1992. The Münster experience with magnetic
43 resonance imaging assisted treatment planning used for high dose rate after loading therapy of
44 gynaecological and nasopharyngeal cancer. Advanced Radiation Therapy Tumour Response Monitoring
45 and Treatment Planning. A. Breit. Berlin, Heidelberg, New York, London, Paris, Tokyo, Hong Kong,
46 Barcelona, Budapest, Springer-Verlag,: 661-665
47 Kubo HD, Wilder RB, Conrad TE, 1999 “Impact of collimator leaf width on stereotactic radiosurgery and 3D
48 conformal radiotherapy treatment plans” Int.J.Radiation Oncology Biol. Phys. 44 937-945.

PartG.doc Page 137 of 147 13 January, 2004


Part G Treatment Planning

1 Kurup R G, Glasgow G and Leybovich L B 1993 Design of electron beam wedges for increasing the penumbra
2 of abutting fields Phys Med Biol 38 667-673
3 Kurup R G, Wang S and Glasgow G 1992 Feasibility study of multileaf collimated electrons with a scattering
4 foil based accelerator Phys Med Biol 37 145-153
5 Kutcher G J, and Burman C (1989) Calculation of complication probability factors for non uniform normal
6 tissue irradiation: The effective volume method Int. J. Radiat. Oncol. Biol. Phys. 16 1623-1630
7 Kutcher G J, Burman C, Brewster L, Goitein M, Mohan R (1991) Histogram reduction method for calculating
8 complication probabilities for three dimensional treatment planning evaluations Int. J. Radiat. Oncol. Biol.
9 Phys. 21137-146
10 Kutcher GJ, Fuks Z, Brenner H, Browwn AP, Burman C, Cheng E, Coia L, Krippner K, Manolis JM, Mohan R,
11 Simpson JR, Urie M, Vikram B, Wallace R (1991) Three-dimensional photon treatment planning for
12 carcinoma of the nasopharynx. Int.J.Radiat.Oncol.Biol.Phys. 21:164-182
13 Kutcher GJ, Leibel S A, Ling C C, Zelefsky M, Fuks Z (1996) New wine in an old bottle? Dose escalation under
14 dose-volume constraints: a model of conformal therpay of the prostate Int. J. Radiat. Oncol. Biol. Phys. 35
15 415-416
16 Kutcher GJ, Niehaus A, Yorke ED (1994) The Effect of Normal Organ Architecture on 3D Conformal
17 Strategies, in Hounsell AR, Wilkinson JM, Williams PC (ed) XIth Internation Conference on The Use of
18 Computers in Radiotherapy. Manchester pp 10-11
19 Kwa S L S, Lebesque J V, Theuws J C M, Marks L B, Munley M T, Bentel G, Oetzel D, Spahn U, Graham M V,
20 Dryzmala R E, Purdy J A, Lichter A A, Martel M K, Ten Haken R K (1998b) Radiation pnemonitis as a
21 function of mean lung dose: an analysis of pooled data of 540 patients Int. J. Radiat. Oncol. Biol. Phys. 42
22 1-9
23 Kwa S L S, Theuws J C M, Wagenaar A, Damen E M F, Boersma L J Baas P, Muller S H Lebesque J V (1998a)
24 Evaluation of two dose-volume histogram reduction models for the prediction of radiation pnemonitis
25 Radiother. Oncol. 48 33-44
26 Lachance B, tramblay D and Pouliot J 1997 A new penumbra generator for electron field matching Med Phys 24
27 485-495
28 Laing R W, Bentley R E, Nahum A E, Warrington A P, Brada M. Stereotactic radiotherapy of irregular targets:
29 a comparison between static conformal beams and non-coplanar arcs. Radiother. Oncol. 28 241 – 246.
30 1993.
31 Lam K S, Lam W C, O' Neill M J, Lee D J and Zinreich E 1987 Electron arc therapy: beam data requirements
32 and treatment planning Clinical Radiology 38 379-383
33 Lambert G D and Klevenhagen S C 1982 Penetration of backscattered electrons in polystyrene for energies
34 between 1 and 25 MeV Phys. Med. Biol. 27 721-725
35 Lawrence T S, Ten Haken R K, Kessler M L, Robertson J M, Lyman J T, Lavigne M L, Brown M B, DuRoss D
36 J, Andrews J C, Ensminger W D, Lichter A S (1992) The use of 3-D dose volume analysis to predict
37 radiation hepatitis Int. J. Radiat. Oncol. Biol. Phys. 23 781-788
38 Lawrence T S, Tesser R J Ten Haken R J (1990) An aplication of dose volume histograms to the treatment of
39 intrahepatic malignancies with radiation therapy Int. J. Radiat. Oncol. Biol. Phys. 19 1041-1047
40 Leach, M. O. 1992 Magnetic resonance imaging and spectroscopy: An introduction to theory, hardware, current
41 applications and safety J Radiol Prot 12 137-158
42 Leach, M. O. 1994 Magnetic resonance spectroscopy applied to clinical oncology Technology and Health Care 2
43 235-246
44 Leavitt D D 1996 Physics of electron arc therapy. in Radiation therapy physics (ed Smith A R) Springer-Verlag,
45 Berlin
46 Leavitt D D, Loverd M P, Gibbs F A and Stewart J R 1985 Electron arc therapy: physical measurement and
47 treatment planning techniques Int J Radiation Biol. Phys. 11 987-999

PartG.doc Page 138 of 147 13 January, 2004


Part G Treatment Planning

1 Leavitt D D, Stewart J R and Earley E 1990 Improved dose homogeneity in electron arc therapy achieved by a
2 multiple-energy technique Int J Radiation Oncology Biol. Phys. 19 159-165
3 Leavitt D D, Stewart J R, Moeller J H and Earley L 1989 Optimization of electron arc therapy doses by multi-
4 vane collimator control Int J Radiation Oncology Biol. Phys. 16 489-496
5 Lebesque J V (1999) Normal tissue complication probability models ESTRO Teaching Course on Conformal
6 Therapy in Clinical Practice Amsterdam, June 20-24 285-297
7 Lebesque, J V, Bruce A, Kroes G, Shouman T, van Herk M (1995) Variation in volumes, dose-volume
8 histograms and estimated normal tissue complication prbabilities of rectum and bladder during conorformal
9 radiotherapy of T3 prostate cancer Int. J. Radiat. Oncol. Biol. Phys. 33 251-257
10 Leibel S A. Zelefsky M J, Kutcher G J, Burman M, Kelson S M, Fuks Z (1994) Three dimsensional conformal
11 radiation therapy in localised carcinoma of the prostate: interim report of phase I dose-escalation study J.
12 Urol. 152 1792-1798
13 Leunens G, Menten J, Weltens C, Verstraete J and van der SE 1993 Quality assessment of medical decision
14 making in radiation oncology: variability in target volume delineation for brain tumours Radiother.Oncol.
15 29 169-175.
16 Levin D.N., Pelizzari C.A., Chen G.T.Y., Chen C.T., Cooper M.D. Retrospective geometric correlation of MR,
17 CT and PET images. Radiology 1988 169: 817 – 823.
18 Li, S., A. Boyer, Y. Lu and G. T. Chen (1997). "Analysis of the dose-surface histogram and dose-wall histogram
19 for the rectum and bladder." Med.Phys. 24(7): 1107-1116.
20 Liao Z X, Travis E L, and Tucker S L (1995) Damage and morbidity from pneumonitis after irradiation of partial
21 volumes of mouse lung Int. J. Radiat. Oncol. Biol. Phys. 321359-1370
22 Ling, C. C., Rogers, C. C. and Morton, R. J. 1983. Computed Tomography in Radiation Therapy. New York,
23 Raven Press
24 Lohr F, Schramm O, Schraube P, Sroka-Perez G, Seeber S, Schlepple G, Schlegel W, and Wannenmacher M,
25 1997 “Simulation of 3D-treatment plans in head and neck tumors aided by matching of digitally
26 reconstructed radiographs (DRR) and on-line distortion corrected simulator images” Radioth. and Oncol
27 45 : 199-207
28 Low, N. N., Vijayakumar, S., Rosenberg, I., Rubin, S., Virudachalam, R., et al. 1990 Beam's eye view based
29 prostate treatment planning: is it useful? Int.J.Radiat.Oncol.Biol.Phys. 19 3 759-768
30 Lunsford, L. D., Martinez, A. J. and Latchaw, R. E. 1986 Stereotaxic surgery with a magnetic resonance and
31 computerised tomography - compatible system J Neurosurg 64 6 872-878
32 Lyman J T (1985) Complication probabilities as assessed from dose-volume histograms Radiat. Res. 104 S13-
33 S19
34 Lyman J T and Wolbarst A B (1987a) Optimization of radiation therapy. III. A method for assessing
35 complication probabilities from dose-volume histograms Int. J. Radiat. Oncol. Biol. Phys. 13 103-109
36 Lyman J T and Wolbarst A B (1987b) Optimization of radiation therapy. IV. A dose volume reduction algorithm
37 Int. J. Radiat. Oncol. Biol. Phys. 13
38 Ma C-M and Jiang S B 1999 Monte Carlo modelling of electron beams from medical accelerators Phys Med
39 Biol 44(12) R157-189
40 Mackay R I, Graham P A, Moore C J, Logue J P and Sharrock P J (1999) Animation and radiobiological analysis of
41 3D motion in conformal radiotherapy. Radiother. Oncol. 52 43-49.
42 Mackenzie, R. and Dixon, A. K. 1995 Measuring the effects of imaging: An evaluative framework Clinical
43 Radiology 50 513-518
44 Mageras, G. S., Podmaniczky, K. C. and Mohan, R. 1992 A model for computer-controlled delivery of 3-D
45 conformal treatments Med.Phys. 19 4 945-953
46 Marchal, G., Bosmans, H. and Van Fraeyenhoven, L. 1990 Intracranial vascular lesions: optimisation and
47 clinical evaluation of three-dimensional time-of-flight MR angiography Radiology 175 443-448

PartG.doc Page 139 of 147 13 January, 2004


Part G Treatment Planning

1 Marks L B, (1994) The pulmonary effects of thoracic irradiation Oncology 8 89-100


2 Martel M K, Ten Haken R K, Hazuka M B, Turrisi A T, Frass B A, Lichter A S (1994) Dose-volume histogram
3 and 3-D treatment planning evaluation of patients with pneumonitis Int. J. Radiat. Oncol. Biol. Phys. 28
4 575-581
5 Maryanski, M. J. 1994 Magnetic resonance imaging of radiation dose distributions using a polymer-gel
6 dosimeter Phys Med Biol 39 1437-1455
7 McKenzie A L 1979 Air-gap correction in electron treatment planning Phys. Med. Biol. 24 628-635
8 McKenzie A L 1998 A simple method for matching electron beams in radiotherapy Phys. Med. Biol. 43 3456-
9 3478
10 McShan D L, Fraass B A, Lichter M D; Full integration of the beam’s eye view concept into computerized
11 treatment planning. Int. J. Radiat. Oncol. Biol. Phys. 18: 1485-1499
12 McShan D L, Silverman A, Lanza D M, Reinstein L E (1979); A computerized three-dimensional treatment
13 planning system utilizing interactive colour graphics. Br. J. Radiol. 52
14 Milosevic, M., Voruganti, S., Blend, R., Alasti, H., Warde, P., et al. 1998 Magnetic resonance imaging (MRI)
15 for localization of the prostatic apex: comparison to computed tomography (CT) and urethrograpy
16 Radiotherapy and Oncology 47 277-284
17 Mirowitz, S. A., Lee, J. K. T., Brown, J. J., Eilenberg, S. S., Heiken, J. P., et al. 1990 Rapid acquisation spin-
18 echo (RASE) MR imaging: a new technique for reduction of artefacts and acquisation time Radiology 175
19 1 131-135
20 Mizokawi, T., Nagata, Y., Okajima, K. and al, e. 1996 Development of an MR simulator: experimental
21 verification of geometric distortion and clinical application Radiology 199 855-860
22 Moerland M. A., van den Bergh A. C. M., Bhagwandien R., Janssen W. M., Bakker C. J. G., Langendijk J. J.
23 W., Battermann J. J. The influence of respiration induced motion of the kidneys on the accuracy of
24 radiotherapy treatment planning, a magnetic resonance imaging study. Radiother. Oncol. 1994 30: 150 –
25 154.
26 Mohan R, Bareest G, Brewster L, Chui C, Kutcher G, Laughlin J and Fuks Z, 1988 “A Comprehensive three-
27 dimensional Radiation Treatment Planning System” Int. J. Oncol. Biol. Phys. 15:481-495;
28 Mohan R, Mageras G S, Baldwin B, Brewster L J and Kutcher G J (1992) Clinically relevant optimization of 3-D
29 conformal treatments Med. Phys. 19 933-944
30 Moore J V, Hendry J H and Hunter R D (1983) Dose incidence curves for tumor control and normal tissue injury in
31 relation to the response of clonogenic cells Radiother. Oncol. 1 143-157
32 Morrison R (1975) The results of treatment of cancer of the bladder - a clinical contribution to radiobiology Clin.
33 Radiol. 26 67-75
34 Müller-Schimpfle, M., Layer, G., Köster, A., Brix, G., Kimig, B., et al. 1992 MRI and MRA in treatment
35 planning of subdiaphragmatic radiation therapy Journal of Computer Assisted Tomography 16 1 110-119
36 Munzenrider JE, Doppke KP, Brown AP, Burman C, Cheng E, Chu J, Chui C, Drzymala RE, Goitein M,
37 Manolis JM, Nori D, Simpson JR, Solin L, Urie MM (1991) Three-dimensional treatment planning for
38 para-aortic node irradiation in patients with cervical cancer. Int.J.Radiat.Oncol.Biol.Phys. 21:229-242
39 Muthuswamy MS, 1999 “A method of beam-couch intersection detection”Ohys.Med. 26: 229-235
40 Nahum A E (1996) «Converting Dose Distributions into Tumour Control Probability» in IAEA-TECDOC-896
41 «Radiation dose in radiotherapy from prescription to delivery» Proceedings of a seminar held in Rio de Janeiro,
42 Brazil, 27-30 August 1994 (IAEA, Vienna) pp 27-40.
43 Nahum A E and Tait D M (1992) Maximising Local Control by Customised Dose Prescription for Pelvic Tumours
44 Advanced Radiation Therapy: Tumour Response Monitoring and Treatment Planning ed. Breit A. (Heidelberg:
45 Springer) pp 425-431
46 Neal, A. J., Sivewright, G. and Bentley, R. 1994 Evaluation of a region growing algorithm for segmenting pelvic
47 computed tomography images during radiotherapy planning Br J Radiol 67 392-395

PartG.doc Page 140 of 147 13 January, 2004


Part G Treatment Planning

1 Niemierko A and Goitein M (1993) Modeling of normal tissue response to radiation: The critical volume model
2 Int. J. Radiat. Oncol. Biol. Phys. 25135-145
3 Niemierko A and Goitein M. Dose-volume distributions: A new approach to dose-volume histograms in three-
4 dimensional treatment planning. Med Phys 21 (1) 3-11
5 Niemierko, A. (1997) Reporting and analysing dose distributions: A concept of equivalent uniform dose. Med. Phys.
6 24 103-110.
7 Niemierko, A. and Goitein, M. (1993) Implementation of a model for estimating tumor control probability for an
8 inhomogeneously irradiated tumour. Radiother. Oncol. 29 140-147.
9 Niermerko A (1999) Prediction of overall pulmonary function loss in relation to 3D dose distribution for patients
10 with breast cancer and malignant lymphoma Radiother. Oncol. 53 195-197
11 Niermerko A and Goitein M (1991) Calculation of normal tissue complication probability and dose-volume
12 histogram reduction schemes for tissues with critical element architecture Int. J. Radiat. Oncol. Biol. Phys.
13 25 135-145
14 O'
Donnell, M. and Edelstein, W. A. 1985 NMR imaging in the presence of magnetic field inhomogeneities and
15 gradient field nonlinearities Med Phys 12 20-26
16 Ogino, T., Nawano, S., Shimizu, W. and Moriyama, N. 1993 Cine MRI in radiotherapy treatment planning of
17 brain tumors Radiation Medicine 11 5 201-205
18 Ohara K., Okumura T., Akisada M., Inada T., Mori T., Yokota H., Calaguas M.J. Irradiation synchronised with
19 respiration gate. Int. J. Radiat. Oncol. Biol. Phys. 1989 17: 853 – 857.
20 Olsen, D. R. and Hellesnes, J. 1994 Absorbed dose distribution measurements in brachytherapy using ferrous
21 sulphate gel and magnetic resonance imaging Br J Radiol 67 1121-1126
22 Padhani, A. R., Khoo, V. S., Suckling, J., Husband, J. E., Leach, M. O., et al. 1999 Evaluating the effect of rectal
23 distension and movement on prostate gland position using cine MRI International Journal of Radiation
24 Oncology Biology Physics 44 525-533
25 Pardo, F. S., Aronen, H. J., Kennedy, D., Moulton, G., Paiva, K., et al. 1994 Functional cerebral imaging in the
26 evaluation and radiotherapeutic treatment planning of patients with malignant gliomas International
27 Journal of Radiation Oncology Biology Physics 30 3 663-669
28 Pelizzari, C. A., Chen, G. T. Y., Spelbring, D. R., Weichselbaum, R. R. and Chen, C. T. 1989 Accurate three-
29 dimensional registration of CT, PET and MR images of the brain Journal of Computer Assisted
30 Tomography 13 20-27
31 Perez C A, Stanley K, Rubin P et al (1980) A prospective randomized study of various irradiation doses and
32 fractionation schedules in the treatment of inoperable non-oat-cell carcinoma of the lung Cancer 45 2744-
33 2753
34 Perks J R, et al. Optimisation of stereotactically guided conformal treatment planning of sellar and parasellar
35 tumours; the practical solution. Int. J. Radiat. Oncol. Biol. Phys. 45 (2) 507 – 513. 1999.
36 Peters L J, Brock W A, Chapman J D, Wilson G and Fowler J F (1989) Response predictors in radiotherapy: a
37 review of research into radiobiologically based assays Br. J. Radiol. Suppl 22 69-108
38 Petti P L, Siddon L S, Effective Wedge Angles with a Universal Wedge, Phys. Med. Biol. 30 985-991. 1985.
39 Phillips, M. H., Kessler, M. L., Chuang, F. Y. S., Frankel, K. A., Lyman, J. T., et al. 1991 Image correlation of
40 MRI and CT in treatment planning for radiosurgery of intracranial vascular malformations International
41 Journal of Radiation Oncology Biology Physics 20 4 881-889
42 Pickett B., Roach III M., Verhey L., Horine P., Malfatti C., Akazawa C. et al. The value of non-uniform margins
43 for six field conformal irradiation of localised prostate cancer. Int. J. Radiat. Oncol. Biol. Phys. 1995 32: 211 –
44 218.
45 Planskoy B, Bedford A M, Davis F M, Tapper P D, Loverock L T. Physical aspects of total-body irradiation at
46 the Middlesex Hospital (UCL group of hospitals), London 1988-1993 I. Phantom measurements and
47 planning methods. Phys. Med. Biol. 41 2307 – 2326. 1996.

PartG.doc Page 141 of 147 13 January, 2004


Part G Treatment Planning

1 Planskoy B, Tapper P D, Bedford A M, Davis F M. Physical aspects of total-body irradiation at the Middlesex
2 Hospital (UCL group of hospitals), London 1988-1993 II. In Vivo planning and dosimetry. Phys. Med.
3 Biol. 41 2327 – 2343. 1996.
4 Porter E H (1980) The statistics of dose/cure relationships for irradiated tumours Br. J. Radiol. 53 336-345
5 Pötter, R., Heil, B., Schneider, L., Lenzen, H., Al-Dandashi, C., et al. 1992 Sagittal and coronal planes from MRI
6 for treatment planning in tumors of brain, head and neck: MRI assisted simulation Radiotherapy and
7 Oncology 23 2 127-130
8 Pötter, R., Kelker, M., Prott, F. J. and Lenzen, B. 1992. Impact of magnetic resonance imaging assisted
9 simulation on target, treatment and irradiation volume in treatment planning of prostate cancer. Advanced
10 Radiation Therapy Tumour Response Monitoring and Treatment Planning. A. Breit. Berlin, Heidelberg,
11 New York, London, Paris, Tokyo, Hong Kong, Barcelona, Budapest, Springer-Verlag: 667-673
12 Pötter, R., Prott, F. J., Jaiser, C., Stöber, U. and Westrick, D. 1992. Technique of magnetic resonance imaging
13 assisted simulation based on direct coronal imaging for treatment planning of supradiaphragmatic
14 malignant lymphoma. Advanced Radiation Therapy Tumour Response Monitoring and Treatment
15 Planning. A. Breit. Berlin, Heidelberg, New York, London, Paris, Tokyo, Hong Kong, Barcelona,
16 Budapest, Springer-Verlag: 631-636
17 Prasad, P. V., Nalcioglu, O. and Rabbani, B. 1991 Measurement of three-dimensional radiation dose distribution
18 using MRI Radiat Res 128 1-13
19 Prince, M. R., Yuccel, E. K., Kaufman, J. A., Harrison, D. C. and Gellar, S. C. 1993 Dynamic gadolinium
20 enhanced three-dimensional abdominal MR arteriography J Magn Reson Imaging 3 877-881
21 Prott, F. J., Pötter, R., Kovacs, G., Lenzen, B. and Schaefer, U. 1992. Magnetic resonance imaging assisted
22 localisation of tumour and applicator for brachytherapy of cancer of the oesophagus. Advanced Radiation
23 Therapy Tumour Response Monitoring and Treatment Planning. A. Breit. Berlin, Heidelberg, New York,
24 London, Paris, Tokyo, Hong Kong, Barcelona, Budapest, Springer-Verlag: 607-611
25 Ramsey C.R., Scaperoth D., Arwood D., Oliver A.L. Clinical efficacy of respiratory gated conformal radiation
26 therapy. Med. Dosim. 1999 24: 115 – 119.
27 Roach, M., Faillace-Akazawa, P., Malfatti, C., Holland, J. and Hricak, H. 1996 Prostate volumes defined by
28 magnetic resonance imaging and computerized tomographic scans for 3-dimensional conformal
29 radiotherapy International Journal of Radiation Oncology Biology Physics 35 5 1011-1018
30 Robb, R. A. 1990. A software system for interactive and quantitative analysis of biomedical images. 3-D
31 Imaging in Medicine. K H Hohn, H Fuchs and S M Pitzer, NATO ASI Series. F 60: 333-361
32 Rothwell R I, Kelly S A and Joslin C A F (1985) Radiation pneumonitis in patients treated for breast cancer
33 Radiother. Oncol. 4 9-14
34 Rowbottom CG, Oldham M, and Webb S, 1999 “Constrained customization of non-coplanar beam orientations
35 in radiotherapy of brain tumours”, Phys.Med.Biol. 44: 383-399
36 Rowbottom CG, Webb S, and Oldham M, 1999 “Is it possible to optimize s radiotherapy treatment plan?” Int. J.
37 Oncol. Biol. Phys. 43: 698-699
38 Rudoltz, M. S., Ayyangar, K. and Mohiuddin, M. 1993 Application of magnetic resonance imaging and three-
39 dimensional treatment planning in the treatment of orbital lymphoma Med Dosim 18 129-133
40 Saini, S., Modic, M. T., Hamm, B. and Hahn, P. F. 1991 Advances in contrast-enhanced MR imaging Am J
41 Radiol 156 235-236
42 Sakurai, K., Fujita, N., Harada, K., Kim, S. W., Nakanishi, K., et al. 1992 Magnetic susceptibility artefact in
43 spin-echo MR imaging of the pituitary gland American Journal of Neuroradiology 13 1301-1308
44 Sanchez-Nieto B and Nahum A E (1999) The Delta-TCP concept: a clinically useful measure of tumour control
45 probability. Int. J. Rad. Onc. Biol. Phys. 44 369-380
46 Sanchez-Nieto B S, Nahum A E and Dearnaley D P (1999) The customisation of dose prescription based on
47 individual dose-volume and radiosensitivity data. Submitted to Radiotherapy and Oncology September
48 1999

PartG.doc Page 142 of 147 13 January, 2004


Part G Treatment Planning

1 Saunders J E and Peters V G 1974 Back-scattering from metals in superficial therapy with high energy
2 electrons Brit. J. Radiol. 47 467-470
3 Schad, L. R., Bluml, S., Hawighorst, H., Wenz, F. and Lorenz, W. J. 1994 Radiosurgical treatment planning of
4 brain metastases based on a fast three dimensional MR imaging technique Magnetic Resonance Imaging 12
5 5 811-819
6 Schad, L. R., Gademann, G., Knopp, M., Zabel, H. J., Schlegel, W., et al. 1992 Radiotherapy treatment planning
7 of basal meningiomas: improved tumor localization by correlation of CT and MR imaging data
8 Radiotherapy and Oncology 25 1 56-62
9 Schad, L., Boesecke, R., Schlegel, W., Hartmann, G. H., Sturm, V., et al. 1987 Three-dimensional image
10 correlation of CT, MR and PET studies in radiotherapy treatment planning of the brain tumours Journal of
11 Computer Assisted Tomography 11 948-1054
12 Schad, L., Lott, S., Schmitt, F., Sturm, V. and Lorenz, W. J. 1987 Correction of spatial distortion in MR
13 imaging: A prerequisite for accurate stereotaxy Journal of Computer Assisted Tomography 11 499-505
14 Schiebe, M. and Hoffmann, W. 2000 CT-based virtual simulation using the AdvantageSim 4.1. system.
15 Description of reliability and accuracy Strahlenther.Onkol. 176 8 377-380
16 Schmitt, F. 1985. Correction of geometrical distortions in MR-imaging. Computer Assisted Radiology. H. U.
17 Lemke, M. L. Rhodes, C. C. Jaffee and R. Flex. Berlin, New York, Springer-Verlag: 15-24
18 Schreiner, L. J., Crooks, I., Evans, M. D. C., Keller, B. M. and Parker, W. A. 1994 Imaging of HDR
19 brachytherapy dose distributions using NMR Fricke-gelatin dosimetry Magnetic Resonance Imaging 12
20 901-907
21 Schultheiss T E, Hanks G E, Hunt M A, Lee W R (199 ) Incidence of and factors related to late complications in
22 conformal and conventional radiation treatment of cancer of the prostate Int. J. Radiat. Oncol. Biol. Phys.
23 Schultheiss T E, Orton C G and Peck R A (1983) Models in radiation therapy: volume effects Med. Phys. 10
24 410-415
25 Schwartz L.H., Richaud J., Buffat L., Touboul E., Schlienger M. Kidney mobility during respiration. Radiother.
26 Oncol. 1994 32: 84 – 86.
27 Seydel H G, Diener-West M, Urtasun R et al. (1985) Radiation Therapy Oncology Group (RTOG),
28 "Hyperfractionation in the radiation therapy of unresectable non-oat cell carcinoma of the lung: preliminary
29 report of an RTOG pilot study Int. J. Radiat. Oncol. Biol. Phys. 111841-1847
30 Sgouros, G., Chiu, S., Pentlow, K. S., Brewster, L. J., Kalaigian, H., et al. 1993 Three-dimensional dosimetry for
31 radioimmunotherapy treatment planning Journal of Nuclear Medicine 34 9 1595-1601
32 Shank B, LoSasso T, Brewster L, Burman C, Cheng E, Chu JCH, Drzymala RE, Manolis J, Pilepich MV, Solin
33 LJ, Tepper JE, Urie MM (1991) Three-dimensional treatment planning for post-operative treatment of
34 rectal carcinoma. Int.J.Radiat.Oncol.Biol.Phys. 21:253-265
35 Shellock, F. G., Morisoli, S. and Kanal, E. 1993 MR procedures and biochemical implants, materials and devices
36 Radiology 189 587-599
37 Sherouse GS, Novins K and Chaney EL, 1990 “Computation of Digitally Reconstructed Radiographs for use in
38 Radiotherapy Treatment Design”, Int. J. Oncol. Biol. Phys. 18:651-658
39 Shipley W U., Tepper J E, Prout G R, Verhey L J, Mendiondo O A, Goitein M, Koehler A M, Suit H D (1979)
40 Proton radiation as boost therapy for localized prostatic carcinoma. JAMA 1912-1915.
41 Shipley WU,. Tepper JE, Prout GR, Verhey LJ, Mendiondo OA, Goitein M, Koehler Am, Suite HD (1979)
42 Proton radiation as boost therapy for localized prostatic carcinoma. JAMA 241:1912-1915
43 Shiu A S, Tung S, Hogstrom K R et al 1992 Verification data for electron beam dose algorithms Med Phys 19
44 623-636
45 Shuman, W. P., Griffin, B. R., Haynor, D. R., Johnson, J. S., Jones, D. C., et al. 1985 MR imaging in radiation
46 therapy planning Radiology 156 143-147
47 Shuman, W. P., Griffin, B. R., Haynor, D. R., Jones, D. C., Johnson, J. S., et al. 1987 The utility of MR in
48 planning the radiation therapy of oligodendroglioma Am J Radiol 148 3 595-600

PartG.doc Page 143 of 147 13 January, 2004


Part G Treatment Planning

1 Siddon RL 1981 “Solution to treatment planning problems using coordinate transforms”, Med.Phys. 8, pp. 766-
2 774,
3 Siddon RL, 1985 “Fast Calculation of the exact radiological path for a three-dimensional CT array”, Med.Phys.
4 12, pp. 252-255.
5 Simpson JR, Purdy JA, Manolis JM, Pilepich MV, Burman C, Forman J, Fuks Z, Cheng E, Chu J, Matthews J,
6 Mohan R, Solin L, Tepper J, Urie M (1991) Three-dimensional treatment planning considerations for
7 prostate cancer Int.J.Radiat.Oncol.Biol.Phys. 21:243-252
8 Solin LJ, Chu JCH, Sontag MR, Brewster L, Cheng E, Doppke K, Drzymala RE, Hunt RE, Kuske R, Manolis
9 JM, McCormick B, Munzenrider J E (1991) Three-dimensional treatment planning of the intact breast.
10 Int.J.Radiat.Oncol.Biol.Phys. 21:193-203
11 Steel G G (Ed.) (1997) Basic Clinical Radiobiology, 2nd Edition (Edward Arnold, Sevenoaks, UK)
12 Stehling, M. K., Turner, R. and Mansfield, P. 1991 Echo-planar imaging: Magnetic resonance imaging in a
13 fraction of a second Science 254 43-50
14 Stephenson, J. A. and Wiley, A. L., Jr. 1995 Current techniques in three-dimensional CT simulation and
15 radiation treatment planning Oncology (Huntingt) 9 11 1225-32, 1235
16 Stewart J G and Jackson A W (1975) The steepness of the dose response curve both for tumor cure and normal tissue
17 injury The Laryngoscope 85 1107-1111
18 Suciu, S., Sylvester, R., Iversen, P., Christensen, I. and Denis, L. 1990 Comparability of EORTC and
19 DAPROCA studies in advanced prostatic cancer Cancer 66 1029-1034
20 Suit H, Skates S, Taghian A, Okunieff P and Efird J T (1992) Clinical implications of heterogeneity of tumor
21 response to radiation therapy Radioth. Oncol. 25 251-260
22 Sumanaweera, T. S., Glover, G. H., Binford, T. O. and Adler, J. R. 1993 MR susceptibility misregistration
23 correction IEEE Trans Med Imaging 12 251-259
24 Suzuki, H. and Toriwaki, J. I. 1991 Automatic segmentation of head MRI images by knowledge guided
25 thresholding Comput Med Imaging Graph 15 233-240
26 Swalec J J, Leavitt D D and Moeller J H 1994 Improved field edge definition in electron arc therapy with
27 dynamic collimation techniques Int J Radiation Oncology Biol. Phys. 30 205-210
28 Sze, G. 1988 Gadolinium-DPTA in spinal disease Radiologic Clinics of North America 26 1009-1024
29 Tait D M, Nahum A E, Meyer L, Law M, Dearnaley D P, Horwich A, Mayles W P, and Yarnold J R (1997)
30 Acute Toxicity in pelvic radiotherapy; a randomized trial of conformal versus conventional treatment.
31 Radiotherapy and Oncology 42 121-136
32 Tanner, S. F., Finnigan, D. J., Khoo, V. S., Dearnaley, D. P. and Leach, M. O. 2000 Radiotherapy planning of
33 the pelvis using distortion corrected MR images: the removal of system distortions Physics in Medicine and
34 Biology 45 2117-2132
35 Ten Haken R. K., Forman J. D., Heimburger D. K., Gerhardsson A., McShan D. L., Perez-Tomayo C. Treatment
36 planning issues related to prostate movement in response to differential filling of the rectum and bladder.
37 Int. J. Radiat. Oncol. Biol. Phys. 1991 20: 1317 – 1324.
38 Ten Haken R. K., Thornton A. F., Sandler H. M., LaVigne M. L., Quint D. J., Fraass B. A. et al. A quantitative
39 assessment of the addition of MRI to CT-based 3D treatment planning of brain tumours. Int. J. Radiat.
40 Oncol. Biol. Phys. 1992 25: 121 – 133.
41 Ten Haken R.K., Balter J.M., Marsh L.H., Robertson J.M., Lawrence T.S. Potential benefits of eliminating
42 planning target volume expansions for patient breathing in the treatment of liver tumours. Int. J. Radiat.
43 Oncol. Biol. Phys. 1997 38: 613 – 617.
44 Ten Haken, R. K., Thornton, A. F., Sandler, H. M., La Vigne, M. L., Quint, D. J., et al. 1992 A quantitative
45 assessment of the addition of MRI to CT-based, 3-D treatment planning of brain tumors Radiother Oncol
46 25 121-133
47 Tepper, J. E. and Padikal, T. N. 1983. The role of computed tomography in treatment planning. Radiation
48 Therapy Planning. N. M. Bleehan, E. Glastein and J. L. Haybittle. New York, Marcel Dekker Inc: 139-158

PartG.doc Page 144 of 147 13 January, 2004


Part G Treatment Planning

1 Terahara A, Niemierko A, Goitein M et al (1999) Analysis of the relationship between tumor dose inhomogeneity
2 and local control in patients with skull base chordoma Int. J. Radiation Oncology, Biol., Phys. 45 351-358.
3 Thames H D and Hendry J H (1987) Fractionation in Radiotherapy (Taylor and Francis)
4 Thames, H. D., Schultheiss, T. E., Hendry, J. H., Tucker, S. L., Dubray, B. M., et al. 1992 Can modest
5 escalations of dose be detected as increased tumor control? Int J Radiat Oncol Biol Phys 22 241-246
6 Thornton, A. F., Sandler, H. M., Ten Haken, R. K., McSchan, D. L., Fraass, B. A., et al. 1992 The clinical utility
7 of magnetic resonance imaging in the 3-dimensional treatment planning of brain tumors International
8 Journal of Radiation Oncology Biology Physics 24 767-775
9 Thwaites, D I 2000 Electron beam treatment-planning techniques. Chapter 10 of Radiotherapy physics in
10 practice. (eds. Williams J R and Thwaites D I ) second edition Oxford University Press, Oxford.
11 Tien, R. D., Buxton, R. B., Schwaighofer, B. W. and Chu, P. K. 1991 Quantitative of structural distortion of the
12 neural foramina in geadient-echo MR imaging Journal of Magnetic Resonance Imaging 1 683-687
13 Toonkel, L. M., Soila, K., Gilbert, D. and Sheldon, J. 1988 MRI assisted treatment planning for radiation therapy
14 of the head and neck Magn Reson Imaging 6 3 315-319
15 Travis E L (1991) Lung morbidity of radiotherapy in Complications of Cancer Management edited by P.N.
16 Plowman, TJ. McElwin, and A.T. Meadows (Butterworth and Heinemann, Stonehan MA) 232-249
17 Ulnin K and Palisca M 1996 The scattering foil compensators in electron beam therapy Int J Radiat Oncol Biol
18 Phys 35 785-792
19 Unterweger, M., Debatin, J. F., Leung, D. A., Wildermuth, S., McKinnon, G. C., et al. 1994 Cardiac volumetry:
20 Comparison of echo-planar and conventional cine-magnetic resonance data-acquisition strategies Invest
21 Radiol 29 11 994-1000
22 Vaidyanathan, M., Clarke, L. P., Velthuizen, R. P., Phuphanich, S., Bensaid, A. M., et al. 1995 Comparisons of
23 supervised MRI segmentation methods of tumour volume determination during therapy Magn Reson
24 Imaging 13 719-728
25 Valicenti, R. K., Waterman, F. M., Croce, R. J., Corn, B., Suntharalingam, N., et al. 1997 Efficient CT
26 simulation of the four-field technique for conformal radiotherapy of prostate carcinoma
27 Int.J.Radiat.Oncol.Biol.Phys. 37 4 953-957
28 van Dyk J, Barnett R B, Cygler J E and Shragge P C 1993 Commissioning and quality assurance of treatment
29 planning computers Int J Radiat Oncol Biol Phys 26 261-273
30 Van Dyk J, Keane T J, Khan S et al. (1981) Radiation pneumonitis following large single dose irradiation: a re-
31 evaluation based on absolute dose to lung Int. J. Radiat. Oncol. Biol. Phys. 7 461-467
32 Van Herk, M. and Kooy, H. M. 1994 Automatic three-dimensional correlation of CT-CT, CT-MRI and CT-
33 SPECT using chamfer matching Medical Physics 21 1163-1178
34 van-den-Elsen, P. A., Pol, E. J. and Viergever, M. A. 1993 Medical image matching - A review with
35 classification IEEE Transactions on Biomedical Engineering 12 26-39
36 Wahl, R. L., Quint, L. E., Cieslak, R. D., Aisen, A. M., Koeppe, R. A., et al. 1993 Anatometabolic tumour
37 imaging: Fusion of FDG PET with CT or MRI to localise foci of increase activity J Nucl Med 34 1190-
38 1197
39 Wambersie A, Landberg T, Chavaudra J, Dobbs J, Hanks G, Johansson K-A, Moller T, Akanuma A, Gerard J-P,
40 Horiot J-C, Suntharalingam N. (1991) ICRU Report 50: Prescribing, recording, and reporting photon
41 beam therapy
42 Wara W M, Phillips T L, Margolis L W, Smith V (1973) Radiation pneumonitis: a new approach to the
43 deterioration of time dose factors Cancer 32 547-552
44 Webb S (1993) The effect on tumour control probability of varying the setting of a multileaf collimator with respect
45 to the planning target volume, Phys. Med. Biol. 38 1923-1936.
46 Webb S (1994) Optimum parameters in a model for tumour control probability including interpatient heterogeneity
47 Phys. Med. Biol. 39 1895-1914.

PartG.doc Page 145 of 147 13 January, 2004


Part G Treatment Planning

1 Webb S and Nahum A (1998) Regarding Wu et al. Tumor-Control Probability of Nasopharyngeal Carcinoma: a
2 comparison of different mathematical models, IJROBP 37(4): 913-920; 1997 Int. J. Rad. Onc. Biol. Phys. 40
3 1009-1010
4 Webb S and Nahum A E (1993) A model for calculating tumour control probability in radiotherapy including the
5 effects of inhomogeneous distributions of dose and clonogenic cell density Phys. Med. Biol. 38 653-666.
6 West C M L, Hendry J H, Scott D, Davidson S E and Hunter R D (1991) 25th Paterson Symposium - is there a
7 future for radiosensitivity testing? Br. J. Cancer 64 197-199
8 White, R. D., Ehman, R. L. and Weinreb, J. C. 1992 Cardiovascular MR imaging: Current level of clinical
9 activity Journal of Magnetic Resonance Imaging 2 365-370
10 Wilks R J, 1993 : “An optical system for measuring surface shapes for radiotherapy planning”, B.J.R., 6 351-
11 359.
12 Willoughby TR, Starkschall G, Janjan NA, Rosen II (1996) Evaluation and scoring of radiotherapy treatment
13 plans using an artificial neural network. Int.J.Radiat.Oncol.Biol.Phys 34(4) 923-30
14 Withers H R, Taylor J M G and Maciejewski B (1988) Treatment volume and tissue tolerance Int. J. Radiat.
15 Oncol. Biol. Phys. 14 751-759
16 Wolbarst A B (1984) Optimization of radiation therapy. II. The critical voxel model Int. J. Radiat. Oncol. Biol.
17 Phys. 10 741-745
18 Wolbarst A B, Chin L M and G.K. Svensson (1982) Optimization of radiation therapy: integral-response of a
19 model biological system Int. J. Radiat. Oncol. Biol. Phys. 8 1761-1769
20 Wood, M. L. 1988. Thoracic and abdominal motion artefacts. Magnetic Resonance Imaging. D. D. Stark and W.
21 G. Bradley. St Louis, Mosby: 792-803
22 Yorke E D, Kutcher G J, Jackson A, Ling C C (1993) Probability of radiation induced complications in normal
23 tissues with parallel architecture under conditions of uniform whole or partial organ irradiation Radiother.
24 Oncol. 26 226-237
25 Zagars G K, Schultheiss T E and Peters L J (1987) Inter-tumour heterogeneity and radiation dose-control curves
26 Radiother. Oncol. 8 353-362.
27 Zelefsky M J, Leibel S A, Kutcher G J, Kelson S, Ling C C Fuks Z (1995) The feasibility of dose escalation with
28 three dimensional conformal radiotherapy in patients with prostatic carcinoma The Cancer Journal 1 142-
29 150

PartG.doc Page 146 of 147 13 January, 2004


Part G Treatment Planning

2 Figure G 1.1 Different steps of the treatment planning process (from CART) ____ Error! Bookmark not defined.
3 Figure G 2.1 Schematic representation of the volumes defined by ICRU demonstrating the relationship of
4 different volumes to each other. Redrawn from {ref} ______________________________________________ 7
5 Figure G 2.2 High grade glioma of the brain. Note the poor demarcation between tumour and the surrounding
6 brain and in turn the difficulty in delineating the GTV in each of these examples (a & b). _________________ 8
7 Figure G 2.3 a) Small perspex fiducial marker. The chamber can be filled with an appropriate radio-opaque
8 contrast e.g. gadolinium for MR, iodinated contrast for CT or positron emitting isotope for PET. It can then be
9 placed on the surface of the patient at a defined anatomical point b) CT image with markers c) PET image with
10 markers _________________________________________________________________________________ 9
11 Figure G 3.1 Example of a CT image of ?? ____________________________________________________ 17
12 Figure G 3.2 CT image of the pelvis with artefacts from a metallic hip prosthesis. ______________________ 19
13 Figure G 3.3 MR image of the pelvis demonstrating non-linear distortion. Note the radio-opaque point markers
14 at the edge of the image which were linearly arranged rather than a concave arrangement as recorded on the
15 image. _________________________________________________________________________________ 20
16 Figure G 3.4 Side by side display of a CT image (left) and of a MR image (right) of the skull. The bony
17 structures cannot be seen on the MR image where the dotted line corresponds to the registered bony outline of
18 the CT image. The tumour delineation (blue) is much easier on the MR image (from Khoo et al 2000) ______ 21
19 Figure G 3.5 PET image (to be obtained)______________________________________________________ 23
20 Figure G 3.6 a) Individual gold seeds cut from a length of wire. b) CT topogram showing 3 seeds within the
21 pelvis of a humanoid phantom. ______________________________________________________________ 24
22 Figure G 4.1 (a) T1 weighted MR image of the prostate; (b) T2 weighted MR image of the prostate. T1
23 weighted images are useful for highlighting the periprostatic tissues which lie between the prostatic
24 capsule and the puborectalis muscle (small arrows). T2 weighted images show the internal prostatic
25 structure better with the central zone appearing dark (large arrows) and the peripheral zone
26 appearing brighter (medium arrow). The prostatic capsule is also easily visualised (small black
27 arrow heads) ___________________________________________________________________________ 27
28 Figure G 4.2 Various forms of MR-related distortion can be seen in this image of a coplanar array of
29 water-filled tubes supported within a circular solid plastic block. System-related distortion effects are
30 seen in the apparent curvature of the tubes at A and their disappearance at B, which was due to
31 warping distortion of the imaging plane. Magnetic susceptibility differences due to the presence of the
32 plastic support block at C give rise to object-induced distortions in the form of discontinuities at the
33 point where each tube enters the support block. The resultant displacements have occurred along the
34 readout (horizontal) direction. ____________________________________________________________ 30
35

PartG.doc Page 147 of 147 13 January, 2004

Das könnte Ihnen auch gefallen