Principles of Radiation Physics

Published on 09/04/2015 by admin

Filed under Hematology, Oncology and Palliative Medicine

Last modified 09/04/2015

Print this page

rate 1 star rate 2 star rate 3 star rate 4 star rate 5 star
Your rating: none, Average: 0 (0 votes)

This article have been viewed 4554 times

7 Principles of Radiation Physics

In this chapter, we present the basic radiation physics needed for the practice of radiation oncology. Many of the specific applications of these basic physics principles, such as brachytherapy and treatment planning, are presented elsewhere in this textbook. Since the primary audience for this textbook is assumed to be practicing radiation oncologists and resident clinicians, many topics are not treated in the detail that would be required by radiation physicists. Excellent physics textbooks that present the topics in more depth are available.14

Basic Concepts

Atomic and Nuclear Structure

Matter is composed of individual elements distinguishable from each other according to their physical and chemical properties. The fundamental building block for these elements is the individual atom, which consists of a small central core, called the nucleus, surrounded by diffuse clouds of electrons. The nucleus, containing almost all the mass in the atom, consists of uncharged neutrons (mass of 1.675 × 10−27 kg) and positively charged protons (mass of 1.673 ×10−27 kg), bound together by the strongly attractive nuclear force. The electrons (mass of 9.11 × 10−31 kg) are negatively charged particles with the same absolute value of charge as the protons. Since the atoms have the same number of electrons and protons, they are electrically neutral. The radius of the atom is about 10−10 m, whereas the radius of the nucleus is about 10−14 m.

The atom is completely specified by the atomic number Z, which is the number of protons in the nucleus (or electrons in the atom), and the mass number A, which is the number of neutrons plus protons in the nucleus. Obviously, the number of neutrons in the nucleus, N, is just A – Z. Element X is specified as image. Isotopes of an element have the same Z but different A and, therefore, have the same chemical properties but can have different physical properties. Isobars are atoms with the same A but different Z, and isotones are atoms with different A and Z but the same N.

In the planetary model of the atom, first described by Niels Bohr, the electron orbits make up discrete concentric shells, with a different binding energy associated with each shell. The shells are labeled K, L, M, and so on from the innermost to the outermost shell. The maximum number of electrons allowed in each shell is 2n2, where n = 1 for the K shell, n = 2 for the L shell, and so on. Therefore, there are up to 2 electrons in the K shell, 8 in the L shell, 18 in the M shell, and so on. A simple schematic drawing of the Bohr model is shown in Fig. 7-1.

The electrons are bound to the protons in the nucleus by the attractive electromagnetic force; therefore, the inner K orbit electrons are bound most tightly because they are closest. The electrons in the outer orbit are called valence electrons and they are typically bound very loosely. The process of excitation is the movement of an electron from an inner orbit to an outer orbit. This requires the addition of energy to the atom. If an electron is completely removed from an atom, the process is called ionization. The energy required to just barely remove an electron from an atom is called the binding energy. In a heavy atom such as tungsten, the binding energy of K shell electrons is about 70,000 electron volts (eV), whereas the valence electrons have binding energies of a few eV. Note that 1 eV is defined as the energy acquired by an electron accelerated through a potential difference of 1 volt, and 1 eV = 1.602 × 10−19 Joules.

Nuclear Transformations

Henri Becquerel discovered natural radioactivity in 1896 when he observed the blackening of wrapped photographic plates when they were placed in contact with certain elements. This event was preceded by the discovery of an invisible form of energy, dubbed x-rays, by Röntgen, who observed the glowing of fluorescent material placed near a gas discharge tube.

We now understand that Becquerel observed the natural decay of radioactive nuclei into one of three types of radiation: positively charged alpha particles (α), which we now know are helium nuclei; negatively charged beta particles (β), which are electrons; and uncharged gamma-rays (γ), which are a type of electromagnetic radiation emitted from nuclei. It is entirely possible that Becquerel also observed other types of elementary particles that can exist in the nucleus, or can be produced in nuclear decay processes.

The explanation of nuclear decay depends on understanding the interplay between the very strong, attractive nuclear force that binds the neutrons and protons together in the nucleus, and the moderately strong repulsive electromagnetic force between the protons. As the protons and neutrons move about within the nucleus, there is some probability that one of them will acquire enough kinetic energy to escape from the nuclear potential energy “well.” Large nuclei with many neutrons and protons tend to be less stable because of the increasing repulsion of the protons. For large nuclei to be stable, an excess of neutrons is required to provide the nuclear attraction to keep the protons from escaping. Fig. 7-2 shows the line of stable nuclei for the range of known elements.

image

FIGURE 7-2 • Plot of numbers of neutrons and protons in stable nuclei.

(From Khan FM: The physics of radiation therapy, ed 3, Philadelphia, 2003, Lippincott Williams & Wilkins, p 4.)

Decay Constant

The decay of a large collection of nuclei can be described as a statistical phenomenon. At any moment in time, the number of nuclei disintegrating per unit of time is proportional to the number of nuclei available. Mathematically,

image

where N is the number of radioactive nuclei and λ is the decay constant. The minus sign indicates that the number of nuclei is decreasing with time. Integration of this differential equation leads to the following equation:

image

where N0 is the number of radioactive nuclei initially present, N(t) represents the number remaining at time t, and e is the base of the natural logarithm (2.71828).

The rate of decay, previously referred to as dN/dt, is also called the activity of a radioactive substance, referred to by the symbol A. Then A = − λN and

image

where A0 is the initial activity and A(t) is the activity at time t. The unit of activity is the Curie (Ci) or the Becquerel (Bq), which is 1 disintegration per second:

image

The Ci was originally defined to be the activity associated with 1 g of radium, although the number of disintegrations per second from 1 g of radium is now known to be somewhat less than 3.7 × 1010.

Modes of Radioactive Decay

Just as electrons have energy levels related to the orbit or shell they occupy, nucleons (protons and neutrons) are thought to occupy shells that represent discrete energy levels within the nucleus. When nucleons are excited into higher levels, they can subsequently decay back to their initial energy level, emitting a photon in the process. This process is called γ decay. Because of the very strong nature of nuclear forces, these excited states are very short-lived, with half-lives on the order of 10−15 seconds. If, in the process of leaving the nucleus, the γ-decay photon transfers enough energy to an inner shell electron to eject it from the atom, the process is called internal conversion. The subsequent cascade of an outer shell electron into the vacated hole in the inner shell releases a characteristic x-ray photon with energy equal to the difference between the energy levels of the inner and outer shells. In some situations, the characteristic x-ray is absorbed and re-emitted by the atom in the form of an ejected orbital electron. Such electrons, emitted in lieu of x-rays, are called Auger electrons.

The decay of a nucleus accompanied by the ejection of an electron (e) or positron (e+) is called β decay. In β (−) decay, a nucleus that has an excess of neutrons can reduce its energy by converting a neutron into a proton, an electron, and an antineutrino (which has energy but no mass). In β (+) decay, a neutron-deficient nucleus can become more stable by converting a proton into a neutron, a positron, and a neutrino. The typical disintegration energy of 1 to 3 MeV is shared between the emitted particles, the neutrino (antineutrino), and positron (electron). For heavy nuclei, the average binding energy of the nucleons inside the nucleus may be less than 10 MeV. Half-lives for β decay can range from a few seconds to many years.

A second way in which neutron-deficient nuclei can gain stability is by the process of electron capture. In this process, an electron in one of the inner shells is captured by the nucleus, thus, transforming a proton into a neutron. That is, a proton plus the captured electron is converted to a neutron plus a neutrino. In general, the inner electron orbital vacancy is quickly filled with an electron from an outer shell, resulting in the emission of characteristic x-rays or Auger electrons. Very heavy nuclei with Z > 82 frequently decay with the emission of an α particle, which is identical to the helium nucleus and is composed of two protons and two neutrons tightly bound together. The binding energy of this configuration is particularly high, making this decay mode a very efficient way to decrease the energy and increase the stability of a heavy nucleus. From a specific nuclide, the α decay is monoenergetic, that is, the α particle is emitted with a single, well-defined energy. Typically, α particles are emitted with kinetic energies of between 5 and 10 MeV.

Radioactive Decay Series and Equilibrium

A total of 106 elements are known. Of these, the first 92 occur naturally; that is, they are either stable or radioactive with half-lives long enough that they can still be found naturally in trace amounts on the earth. Elements 93 to 106 are very unstable and can only be produced artificially in accelerators with heavy ion bombardment of heavy nuclei. Lead, with Z = 82 is the heaviest known element that has completely stable isotopes. All naturally occurring isotopes of elements with Z between 83 and 92 are members of one of three radioactive series referred to as the uranium (Z = 92) series, the actinium (Z = 89) series, and the thorium (Z = 90) series. The name of each series refers to the parent, a long-lived nuclide that decays into a daughter nuclide, that then decays into another daughter, followed by successive transformations that continue until a stable isotope of lead is reached.

In each series, the half-life of the parent is longer than any of the half-lives of any of the daughter nuclides. This phenomenon leads to a state of transient or secular equilibrium. Since each daughter is less stable than the original parent, the activity of the daughter will increase with time, until it approaches the activity of the long-lived parent. If the half-lives are very different, the activities will become essentially identical after a number of daughter half-lives, and this is called secular equilibrium. An example is shown in Fig. 7-3. If the half-lives of the parent and daughter are very similar, transient equilibrium is reached in a few half-lives, after which the daughter has an activity in excess of the parent activity but decays with the same rate as the limiting parent decay rate. An example of this phenomenon is shown in Fig. 7-4. For either case, after a time that is long compared to the half-life of the shorter-lived nuclide,

image

FIGURE 7-3 • Example of secular equilibrium in the decay of 226Ra to 222Rn.

(From Khan FM: The physics of radiation therapy, ed 3, Philadelphia, 2003, Lippincott Williams & Wilkins, p 19.)

image

where Ai and Ti refer to the activity and half-life of the daughter (d) or parent (p).

X-Ray Production

Most radiotherapy is delivered with beams of x-rays that are produced as a result of the interactions of accelerated electrons with matter. This can be either through excitation or ionization of the atom via interactions of the accelerated electrons with target electrons leading to the emission of characteristic x-rays, mentioned earlier, or through direct interaction between the electrons and the electromagnetic field of the target nuclei, leading to bremsstrahlung x-rays.

Characteristic X-Rays

When electrons are incident on target atoms, they can ionize those atoms by depositing sufficient energy to eject an inner shell electron. The inner shell vacancy is subsequently filled by an outer shell electron, causing the emission of a characteristic x-ray with energy equal to the difference between the binding energies of the inner and outer shells. This process is diagrammed in Fig. 7-5. Although this process can take place in both low-Z and high-Z atoms, only for high-Z atoms are the binding energies sufficient to produce radiation in the x-ray portion of the electromagnetic spectrum. For example, the binding energy of K-shell electrons in tungsten, a common target for x-ray tubes and accelerators, is about 70 keV, whereas for aluminum, it is only about 1.5 keV. As mentioned earlier, the characteristic x-ray energy is occasionally transferred directly to an orbital electron, leading to the production of an Auger electron.

image

FIGURE 7-5 • Illustration of the production of characteristic radiation.

(From Khan FM: The physics of radiation therapy, ed 3, Philadelphia, Lippincott Williams & Wilkins, p 35.)

Bremsstrahlung X-Rays

When high-energy electrons interact directly with the electromagnetic field of a target nucleus, they are deflected and lose a portion of their energy due to deceleration. The energy lost is then emitted in the form of radiation called bremsstrahlung (literally, braking radiation).

Since the incident electron can lose any portion of its energy in this process, the energy of these bremsstrahlung x-rays can vary continuously from nearly zero to the full energy of the incident electron. The angle of the x-rays, relative to the beam direction, depends on the energy of the electron. At low energies typical of those used for diagnostic radiology, the x-ray has equal probability of being emitted in any direction, whereas at high energies typically used for radiotherapy, the x-rays are emitted preferentially in the forward direction. These facts explain the use of thick targets and 90-degree angles between electrons and x-rays for diagnostic applications and thin transmission targets for radiotherapy.

The probability of bremsstrahlung production varies with Z2 of the target material, whereas the efficiency of x-ray production depends on the product of Z and E, the energy of the electrons. At 100 keV in a tungsten target, only about 1% of the incident energy of the beam is converted into x-rays, the remainder ending up in heat.

In general, both bremsstrahlung and characteristic x-rays are present when electrons strike a target. Fig. 7-6 shows x-ray spectra for a thick tungsten target for electron energies of from 65 keV to 200 keV. Note the appearance of the characteristic x-rays on top of the bremsstrahlung spectrum once the electron energy is in excess of the K-shell binding energy of 70 keV. Any material in the x-ray beam, including the target itself, will absorb some of the energy of the beam. The effect of such absorption is to preferentially filter out the low energy portion of the bremsstrahlung spectrum. The process of preferentially attenuating the low-energy component of the beam is referred to as “beam hardening.” When the beam is hardened, the average energy of the x-rays increases at the expense of reduced intensity.

Interaction of X-Rays and Particles With Matter

X-Ray Interactions

As described earlier, x-rays can be considered as packets of energy, called photons, for purposes of considering their interactions with matter. There are a number of different interaction processes that the photons can experience, and each of these has a probability described by an attenuation coefficient. Mathematically, this can be written as

image

where dN is the reduction in the number of photons due to interactions in a thickness dx of an absorber, N is the number of incident photons, and µi is the attenuation coefficient for the ith interaction process. After integration over the thickness, this equation becomes

image

where N(x) is the number of photons left in the beam (without interaction) as a function of the thickness x of material traversed, and µtot is the sum of the individual attenuation coefficients. (µi/ρ) is called the mass-attenuation coefficient for process i, where ρ is the density of the material. Dividing the conventional attenuation coefficients by density removes the dependence on the physical density of materials and the resulting mass attenuation coefficients are approximately equal from material to material. The corresponding thickness, ρx, is in units of grams per square centimeter (g/cm2). The total mass attenuation coefficients for water and lead are shown as functions of energy in Fig. 7-7. There are five different photon interaction processes, including coherent scattering, photoelectric effect, Compton scattering, pair production, and photodisintegration. These will be discussed in turn.

image

FIGURE 7-7 • Mass attenuation coefficients for lead and water. The sharp discontinuities in the lead curve are due to absorption edges.

(From Johns HE, Cunningham JR: The physics of radiology, ed 4, Springfield, IL, 1983, Charles C. Thomas.)

The thickness of material that reduces the number of transmitted photons to one half the initial value is referred to as the half-value layer (HVL) and depends on the nature of the material and the energy of the photons. This occurs when

image

implying that

image

Although not constant with depth, HVL is often used to describe the quality of the beam or its ability to penetrate material.

Compton Scattering

Compton scattering is a process in which the incident photon interacts with an orbital electron as if it were a free particle, since the binding energy is small compared to the photon energy. The dynamics of the interaction can be described as a typical particle–particle scattering interaction whereby the photon transfers some of its energy to the electron and is scattered at an angle φ relative to the incident direction. The electron is ejected at an angle θ relative to the forward direction. The energy of the outgoing Compton-scattered photon is equal to the difference between the incident photon energy and the energy transferred to the electron. A diagram of the Compton process is shown in Fig. 7-8.

image

FIGURE 7-8 • Illustration of the Compton effect.

(From Khan FM: The physics of radiation therapy, ed 3, Philadelphia, Lippincott Williams & Wilkins, p 67.)

As with any particle–particle scattering process, energy and momentum conservation can be applied to obtain relationships between energy and angle.

From conservation of energy:

image

From conservation of momentum:

image

and

image

where ve is the velocity of the electron, me is the rest mass of the electron, c is the speed of light, and

image

Solving these three simultaneous equations, we obtain the following relationships:

image

image

image

where α = hνo/ mec2 and f = (1 − cos φ).

From these equations, we can examine several limiting cases of the Compton effect. A grazing hit corresponds to θ = 90 degrees, in which case the photon is undeflected, and continues in the forward direction. A direct hit corresponds to θ = 0 degrees, meaning that φ = 180 degrees, or in other words, the photon scatters backwards. For intermediate situations, for example, when the photon is deflected by 90 degrees, the angle and energy of the electron depend on the energy of the photon. When α is much greater than 1 (high energy), θ is small and E is essentially equal to hνo. When α is much less than 1 (low energy), the photon energy remains nearly unchanged and the Compton process approaches that of coherent scattering.

The mass attenuation coefficient for Compton scattering is independent of Z, decreases slowly with photon energy, and is directly proportional to the number of electrons per gram, which varies by only 20% from the lightest to the heaviest elements (with the exception of hydrogen). Therefore, in the energy region where Compton processes dominate, the attenuation of the beam will vary according to the integrated density of the material traversed. This fact is responsible for the relatively poor contrast observed in portal verification films exposed with megavoltage x-rays exiting from the irradiated patient.

Relative Importance of Interaction Types

The total mass absorption coefficient is the sum of the coefficients for coherent, photoelectric, Compton, and pair production. For low-Z targets such as tissue (average Z = 7), Compton processes are overwhelmingly dominant throughout the range of energies used in therapy. Fig. 7-9 shows the relative importance of photoelectric, Compton, and pair production as a function of photon energy and the atomic number of the target material.4 For tissue, Compton dominates between approximately 30 keV and 30 MeV, whereas for lead, photoelectric dominates up to approximately 800 keV and pair production above about 5 MeV. Because of its calcium content, bone has a higher absorption by photoelectric and pair production. This leads to high bone doses for orthovoltage and kilovoltage and for higher energy accelerator beams.

image

FIGURE 7-9 • Relative importance of the three main modes of energy loss for photons as a function of energy and atomic number of the medium.

(From Hendee WR: Medical radiation physics: roentgenology, nuclear medicine and ultrasound, Chicago, 1979, Year Book Medical Publishers, p 115.)

Charged Particles

Electrons

Electrons lose energy to the medium by two processes, ionization and radiation. Ionization processes are interactions with the atomic electrons, whereas radiation losses are interactions with the field of the nucleus, leading to bremsstrahlung x-ray production. Whereas the microscopic interactions are well understood, the macroscopic description of the energy and range of the particles at any point in the medium is not simple. This is because the electrons are very much lighter than the atomic nuclei. Therefore, the electron can lose a very large fraction of its energy in a single process and can be deflected by very large angles. This means that even if the electron beam is monoenergetic when entering a medium, there will be a large amount of range-straggling, or a large variation from electron to electron as to where, in the phantom, the electron will stop. Fig. 7-10 shows a plot of absorbed dose as a function of depth for monoenergetic electrons incident on water. There is a depth beyond which the dose is almost zero, where all the incident electrons have been stopped and where the remaining dose is due to bremsstrahlung x-rays produced by the electrons in the medium. This is in sharp contrast to the exponential falloff of dose for x-rays. This ranging out of electrons in matter is responsible for the popularity of electrons for treatment of superficial disease.

image

FIGURE 7-10 • Depth-dose curve for monoenergetic electrons incident on water. R50 is the mean range and Rp is the extrapolated or practical range.

(From Johns HE, Cunningham JR: The physics of radiology, ed 4, Springfield, IL, 1983, Charles C. Thomas.)

Protons and Heavier Ions

The use of protons and heavy ions for radiotherapy was pioneered in the 1970s and 1980s at the Massachusetts General Hospital in collaboration with the now decommissioned Harvard Cyclotron Laboratory and at the Lawrence Berkeley National Laboratory in Berkeley, California.515 Since then, several proton and charged-particle facilities have been built around the world, a list of which can be found on the website for the Particle Therapy Co-Operative Group (PTCOG; http://ptcog.web.psi.ch/). Similar to electrons, protons lose energy primarily by electromagnetic interactions with the atomic electrons. A major difference between protons and electrons is that protons are much heavier than electrons and therefore lose only a very small fraction of their energy in an individual interaction, scattering minimally in the process. Protons lose energy at an increasing rate as they slow down, yielding an enhanced region of energy deposition called the Bragg Peak just before they stop. One can predict very accurately the depth at which the protons will come to rest if the initial energy of the protons and the electron density of the material traversed are known. The absence of exit dose beyond the intended target makes protons nearly ideally suited for optimal physical dose delivery. A typical depth dose distribution for a clinical, energy-modulated proton beam, compared to that of a 10-MV bremsstrahlung x-ray beam, is shown in Fig. 7-11. A small percentage of proton interactions is via nuclear interactions with the nucleus. These rare interactions are qualitatively different than electromagnetic interactions and are thought to be responsible for enhancing the cell kill per unit dose by about 10% to 20% over that of x-rays.16

Heavy ions such as helium, carbon, neon, and argon nuclei have also been used in radiotherapy.912,14,15,17,18 Similar to protons, they lose energy by interacting with the atomic electrons. Being even heavier than protons, they scatter even less and have even more rapid dose falloff outside the central beam and even faster falloff of dose beyond the end of range. In addition to these electromagnetic interactions, heavy ions have a probability of interacting with the nucleus via nuclear interactions that increases with the mass of the incident ion. In particular, as the heavy ions begin to slow down, they tend to suffer nuclear interactions in which they lose a very large amount of their energy in a single event. This high density of deposited energy kills cells in a far more efficient way per unit dose than x-rays. These particles are said to have a high radiobiologic efficiency (RBE). For more information on these particles, the reader is referred to the literature.13

Neutrons

Neutrons have also been used in clinical treatments.1924 These particles are neutral, so they cannot lose energy by any means other than the nuclear interaction. All of their interactions are “catastrophic,” because a significant portion of their energy is deposited in an individual event. Their primary interactions are with protons within the nucleus. These nuclear events result in recoil protons and charged nuclear fragments that have rather low energy and deposit large amounts of energy very close to the site of the original interaction. The falloff of neutron dose with depth is exponential and similar to that of low-energy x-rays, with the depth of the 50% dose being around 10 cm for typical treatment energies. Neutrons are very efficient in producing cell kill per unit dose (high RBE) relative to x-rays for both tumor tissue and normal tissues. Clinical trials have been conducted to investigate the areas where neutrons might have an advantage over x-rays. A summary of the physical characteristics and results of early clinical results is available.13

In addition, clinical trials are under way to investigate the use of slow neutrons, either thermal or epithermal in energy, to treat boronated tumor cells in patients.2532 The clinical efficacy of these neutrons for cancer therapy depends on the very high cross section for neutron capture by boron and on a high ratio of boron in the tumor cells to that in the surrounding normal cells. To date, clinical results of these trials have been disappointing.

Radiation Therapy Treatment Machines

Before about 1950, external beam radiation therapy was primarily carried out with x-ray beams produced in evacuated x-ray tubes by electrons accelerated with an electric field, impinging on a target and interacting with the nuclei of target atoms. The details of the interactions and the properties of the resulting x-ray beams are described elsewhere in this chapter. The maximum energy of these electrons and the photon beams that they produced was about 400 kV. Therefore, this period of radiation therapy history is called the kilovoltage era. Although most of these machines have been replaced in the intervening years with 60Co teletherapy machines and electron linear accelerators, low-energy x-ray generators still play a limited role in the treatment of superficial disease.

Kilovoltage Units

A schematic diagram of an x-ray tube suitable for radiation therapy is shown in Fig. 7-12. Since a high-voltage supply is used to generate the accelerating potential for the electron beams, a maximum energy of approximately 300 kV is achievable with this design. The resulting x-ray beams have a spectrum of photon energies with a maximum equal to the energy of the electron beam.

image

FIGURE 7-12 • Schematic diagram of an x-ray tube that could be used for radiation therapy.

(From Khan FM. The physics of radiation therapy, ed 3, Philadelphia, Lippincott Williams & Wilkins, p 29.)

Megavoltage Units

X-ray and γ-ray beams of energy greater than 1 MV are classified as megavoltage beams. In the 1930s a number of transformer-based and van de Graaff generator–based units working in the 800 KeV to 2 MeV range were installed around the world. However, it was not until the introduction of the 60Co teletherapy machine and the linear accelerator that the routine use of megavoltage x-ray and γ-ray beams occurred, significantly changing the practice of radiotherapy and leading to substantial improvements in clinical results.33

Teletherapy Machines

The megavoltage era began with the introduction of the 60Co teletherapy machine into radiotherapy clinics beginning in 1951.34,35 The development of nuclear reactors in the late 1940s made possible the production of small 60Co sources with specific activities (in Curies per gram) high enough to produce clinically acceptable dose rates of more than 1 Gray (Gy) per minute at a typical treatment distance of 80 cm from the source. These machines quickly became the standard of radiotherapy because of their simplicity of design and operation, low cost, and availability. The two photons produced by the decay of 60Co have energies of 1.17 and 1.33 MeV, yielding a depth-dose curve that falls to 50% of maximum in about 10 cm of water or tissue. In addition, the relatively high energy of the photons leads to “skin sparing” due to a reduced entrance dose that builds up for a distance of 5 mm (the maximum range of the secondary electrons from 60Co photons) to a maximum value. The half-life of 60Co is 5.27 years, making the source change a relatively infrequent requirement. With the highest specific activities currently achievable, dose rates of at least 1.5 Gy per minute for a field size of 40 cm × 40 cm can be achieved at 80 cm source-to-treatment distance.

Aside from the requirement to replace the sources every few years, additional disadvantages to 60Co isotope machines include the need to shield against continuous leakage radiation and the lack of field flatness for the largest field sizes. In addition, the physical size of the source capsule, typically 1.5 cm to 2.0 cm in diameter, is much larger than the effective source size available on any machine that uses an electron beam (typically 4 mm or less for linear accelerators). This large source size contributes to a geometric penumbra (the penumbra is the area at the edge of the beam where the dose falls from high dose to low dose) that can be significantly wider than with accelerators.

Before the development of 60Co isotope machines, both radium and cesium sources were used for teletherapy treatments. However, the limitations of low specific activity for radium and low energy and limited specific activity for cesium limited their usefulness for general radiotherapy.

Betatrons

The next chronological development in treatment machines is the betatron, first developed by D. W. Kerst at the University of Illinois36 primarily for physics experimentation. Electrons with energies of up to 45 MeV have been produced in betatrons for radiotherapy and used either directly or to produce bremsstrahlung x-ray beams. The massive size of these machines, their high cost, and relatively low dose rate combined to limit their usefulness in radiation therapy.

Linear Accelerators

The use of microwaves to accelerate electrons to high energies for radiotherapy was first demonstrated in Great Britain in 1953.37 This became possible primarily because of the development of high-power microwave generators for military radar use during World War II. The major components of a typical medical linear accelerator (linac) are shown in a block diagram in Fig. 7-13. The acceleration of the electron beam takes place in the accelerator wave guide, consisting of a stack of cylindrical cavities with a hole through the center, into which resonant electromagnetic waves of frequency in the microwave range (∼3000 MHz) have been coupled. The electron beam is created and preaccelerated to approximately 50 keV in a conventional electrostatic electron gun, injected into the resonating waveguide, spatially bunched and accelerated through interactions with the electromagnetic field in the individual cavities, and then emerges from the accelerating structure as a narrow pencil beam that can be magnetically bent or focused onto a scattering foil (for electron treatments) or a bremsstrahlung target (for x-ray treatments) in the treatment head of the linac. For low energies of 6 MeV or less, the beam can be magnetically focused in the forward (vertical) direction onto the axis of the treatment head, whereas at higher energies, the beam is usually bent through a 90-degree or 270-degree angle before entering the treatment head since the accelerating structure is much longer for high energies and is frequently oriented horizontally to save space. For more details about the acceleration process, the reader is referred to several excellent review articles in the literature.3840

image

FIGURE 7-13 • A block diagram of a typical medical linear accelerator.

(From Khan FM: The physics of radiation therapy, ed 3, Philadelphia, 2003, Lippincott Williams & Wilkins, p 43.)

Although the details of the acceleration process can be quite technical, the shaping of the beam in the treatment head is both important and conceptually easy to understand. Fig. 7-14

Buy Membership for Hematology, Oncology and Palliative Medicine Category to continue reading. Learn more here