Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

DNA and Chromatin Damage Caused by Radiation
DNA and Chromatin Damage Caused by Radiation
DNA and Chromatin Damage Caused by Radiation
Ebook874 pages

DNA and Chromatin Damage Caused by Radiation

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Advances in Radiation Biology, Volume 17: DNA and Chromatin Damage Caused by Radiation outlines the different biological reactions to radiation. This book discusses the linear energy transfer and energy loss; DNA breaks and track structure; DNA radicals from water radicals; and radiation-induced strand breaks in isolated DNA. The radiation damage to DNA and its nearby environment; thiol radioprotectors and mechanism of action; radiolysis of water and track reactions; and computer simulation of higher order structure of DNA are also elaborated. This publication likewise covers the concept of chromatin structure; DNA supercoiling studied by sedimentation; measurement of radiation-induced DNA breakage; and analysis of damage in interphase cells. This volume is a useful reference to biologists and students concerned with DNA and chromatin damage caused by radiation.
LanguageEnglish
Release dateOct 22, 2013
ISBN9781483282275
DNA and Chromatin Damage Caused by Radiation

Related to DNA and Chromatin Damage Caused by Radiation

Medical For You

View More

Reviews for DNA and Chromatin Damage Caused by Radiation

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    DNA and Chromatin Damage Caused by Radiation - John T. Lett

    Advances in Radiation Biology

    DNA and Chromatin Damage Caused by Radiation

    John T. Lett

    Department of Radiological Health Sciences, Colorado State University, Fort Collins, Colorado

    Warren K. Sinclair

    National Council on Radiation Protection and Measurements, Bethesda, Maryland

    ISSN  0065-3292

    Volume 17 • Number Suppl. (C) • 1993

    Table of Contents

    Cover image

    Title page

    Advisory Board

    Copyright page

    Chapter 1: Linear Energy Transfer and Track Structure

    Publisher Summary

    I Introduction

    II Linear Energy Transfer and Energy Loss

    III δ-Electron Emission

    IV Condensed Phase Effects — Track Core

    V Electron Transport and the Track Halo

    VI DNA Breaks and Track Structure

    VII Summary

    Acknowledgments

    Chapter 2: Primary Free Radical Processes in DNA

    Publisher Summary

    I Introduction

    II DNA Radicals from Water Radicals—Recent Advances

    III One-Electron Oxidized Species of DNA in an Aqueous Environment

    IV Dynamics of Radiation-Induced Changes in Solid DNA

    V Radiation-Induced Strand Breaks in Isolated DNA

    VI Forward Look

    Acknowledgments

    Chapter 3: The Chemical Consequences of Radiation Damage to DNA

    Publisher Summary

    I Introduction

    II Radiation Damage to DNA and Its Nearby Environment: Direct and Quasi-direct Effects

    III The Indirect Effect

    IV The Confluence of Chemical Events for the Direct, Quasi-direct, and Indirect Effects

    V Thiol Radioprotectors and Mechanism of Action

    VI Conclusions and Directions for Future Efforts

    Acknowledgments

    Chapter 4: Computer Simulation of Initial Events in the Biochemical Mechanisms of DNA Damage

    Publisher Summary

    I Introduction

    II Energy Deposition Events and Creation of Tracks by Charged Particles

    III Radiolysis of Water and Track Reactions

    IV Computer Simulation of the Biochemical Stage: The Formation of Strand Breaks

    V Results of Yields on Strand Breaks

    VI Computer Simulation of Higher Order Structure of DNA

    VII Concluding Remarks and Future Directions

    Chapter 5: DNA Loop Structure and Radiation Response

    Publisher Summary

    I Introduction

    II The Concept of Chromatin Structure

    III DNA Supercoiling Studied by Sedimentation

    IV Alternative Methods

    V DNA Loop Structure and Growth State

    VI DNA Loop Structure, Anchoring, and Radiosensitivity

    VII Cell Cycle Effects

    VIII Conclusions

    Chapter 6: Radiation-Induced Damage in Chromosomal DNA Molecules: Deduction of Chromosomal DNA Organization from the Hydrodynamic Data Used to Measure DNA Double-Strand Breaks and from Stereo Electron Microscopic Observations

    Publisher Summary

    I Introduction

    II Methods for the Measurement of DNA Breakage

    III DNA Size, Shape, and Number Concentration Measurement

    IV Problems Unique to the Measurement of Large DNA Molecules

    V Measurement of Radiation-Induced DNA Breakage

    VI Size and Shape Determination of Mammalian Cell Chromosomal DNA Molecules

    VII Structure of the Mammalian Chromosome

    G Discussion of the Model

    VIII Summary

    Acknowledgments

    Chapter 7: Ionizing Radiation Damage and Its Early Development in Chromosomes

    Publisher Summary

    I Introduction

    II Chromosomal Aberrations at Mitosis

    III Analysis of Damage in Interphase Cells

    IV What Are PCC Breaks?

    V The Nature of Critical Cellular Structures

    VI Conclusions

    Acknowledgments

    Index

    Advisory Board

    G. E. ADAMS

    R. J. MICHAEL FRY

    U. HAGEN

    P. C. HANAWALT

    J. LINIECKI

    JOHN B. LITTLE

    JEAN R. MAISIN

    HIROMICHI MATSUDAIRA

    WILLIAM J. SCHULL

    C. STREFFER

    ARTHUR C. UPTON

    Copyright page

    Copyright © 1993 by ACADEMIC PRESS, INC.

    All Rights Reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopy, recording, or any information storage and retrieval system, without permission in writing from the publisher.

    Academic Press, Inc.

    1250 Sixth Avenue, San Diego, California 92101-4311

    United Kingdom Edition published by

    Academic Press Limited

    24–28 Oval Road, London NW1 7DX

    International Standard Serial Number: 0065-3292

    International Standard Book Number: 0-12-035417-9

    PRINTED IN THE UNITED STATES OF AMERICA

    93 94 95 96 97 98 EB 9 8 7 6 5 4 3 2 1

    1

    Linear Energy Transfer and Track Structure

    Gerhard Kraft and Michael KrÄMer,     Gesellschaft Für Schwerionenforschung mbH, D-6100, Darmstadt, Germany

    Publisher Summary

    This chapter discusses track structures and linear energy transfer. There are two types of energy dissipation of heavy ions—nuclear and electronic stopping. Nuclear stopping has high RBE values and predominates at extremely low specific energies of a few kiloelectronvolts per mass unit, corresponding to the last micrometers of the particle range. At higher energies, nuclear stopping contributes only a few percent to the total stopping; the predominant process is the electronic stopping. The electronic stopping is proportional to the square of the effective projectile charge and increases up to a maximum value. The predominant process of electronic stopping involves the interaction of the projectile with the target electrons. In the collision process, a spectrum of electrons is produced with different kinetic energies ranging from zero up to a maximum energy that depends on the velocity of the primary ion. From the energy loss of the primary ion, two-thirds are transformed into the kinetic energy of the electrons. The residual energy is used to overcome the binding energy of the electrons and for target excitations. In the center of the track, the ionized atoms are pushed out from their original positions by electrostatic repulsion. This process is called Coulomb explosion and is responsible for the formation of the latent track in nuclear track detectors. In DNA experiments, the influence of track structure on the induction of double- and single-strand breaks is observed for LET values greater than 5 keV/μm. In consequence, the ratio of DSB to SSB increases with decreasing track diameters, and the highest values are found at the very end of the tracks. However, if the radiosensitivity is increased by changing the DNA environment or the repair capacity, additional DSB are produced directly in single events.

    I Introduction

    Dose is currently regarded as the principal parameter in radiobiology. Usually, the different biological reactions to ionizing radiation are studied as a function of dose. From a comparison of the dose–effect curves, conclusions concerning the basic mechanism governing the radiobiological reactions are drawn. From its definition as energy deposited per mass unit, dose appears to be a continuous parameter and it should be possible to divide it into infinitesimal subunits of smaller mass elements receiving smaller amounts of energy. As it turns out, this is not true even for relatively large objects like subcellular structures where the size of the target exceeds the atomic dimensions by many orders of magnitude.

    For sparsely ionizing radiation like X- or γ-rays, the energy is transferred from the electromagnetic radiation via Compton or photo processes or pair production into more or less energetic electrons. It is not only these processes but also the secondary collision processes of the liberated electrons that result in ionization and consequently in biochemical and biological damage. With electrons as the primary radiation, the biochemical and biological effects are produced by the ionization processes of these electrons and their subsequent electron cascade.

    When masses or volumes containing only a few or one electron are considered, the grainy structure of dose becomes visible. Microdosímetry has demonstrated that for small volumes of tissue-equivalent materials this fine structure of the energy deposition plays an important role (Kellerer and Rossi, 1978).

    In general, the radiobiological effect of all types of sparsely ionizing radiation are caused by the action of liberated electrons and of the primary ionization events. With the exception of very low energy electrons produced, for example, by soft X-rays, the liberated electrons experience the same slowing down process in the target. In consequence, the biological efficiency differs only some 10% between electromagnetic radiation produced from different sources like Co-γ-rays or the Bremsstrahlung spectrum of an X-ray tube.

    For particle radiation like protons, a-particles, or heavier ions (and also for neutrons, which act via reaction and recoil ions), a much larger variation in the relative biological efficiency (RBE) is observed. Their biological effect is caused by primary ionizations occurring along the particle trajectories and the action of the liberated electrons and their secondaries. Differences in the biological efficiency of these particles therefore have to be attributed to the spatial and time correlations of the ionization events caused by the electrons and the primary particle.

    In radiobiology different concepts are used to characterize radiation quality: the LET concept, the track structure concept, and the concept of microdosimetry.

    In early studies of particle radiobiology, the specific ionization density, i.e., the number of ion–electron pairs created along the path of the particle, was used to characterize densely ionizing radiation. But assuming that for the creation of each ion–electron pair the same constant amount of energy is necessary, the number of ionization events per track follows exactly the electronic stopping power curve in its dependence on particle energy. For gases it is known that on the average an energy of approximately 31 eV is necessary to create an ionization event, but for material of tissue density, this quantity is not known. Therefore, Zirkle proposed in 1954 to characterize densely ionizing radiation by the linear energy transfer (LET) to the absorber material. The LET measured in kiloelectronvolts per micrometer is numerically identical with the stopping power of the particles in water as reference material as long as no restrictions to the δ-electron¹ energy are applied.

    Linear energy transfer has proven to be a useful parameter in the description of many radiobiological effects in different cellular and subcellular targets. In general, the RBE increases with increasing LET values up to a maximum at around 100 keV/μm. At higher LET values, RBE decreases to values much smaller than one (Barendsen et al., 1963). Exceptions from this general behavior are the induction of single-strand breaks and other biological effects related to single-strand breaks, but not to double-strand breaks (Kraft, 1988). In addition, radiation effects in very sensitive targets, for instance, in repair-deficient mutants (Lett et al., 1989), where the track structure seems to play only a minor role, do not show RBE maxima at LET values around 100 keV/μm. However, the same steep decrease of RBE for higher values of LET is also found for these effects.

    In the last decade it has been shown experimentally that LET is not a good parameter when different ions, having different atomic numbers, are used (Todd, 1965; Wulf et al., 1985; Kraft, 1987; Belli et al., 1989; Folkard et al., 1989). In this case, for the same LET, different biological efficiencies have been observed. In addition, it was also evident from atomic and nuclear track physics that LET cannot be a good parameter to describe the action of heavy charged particles independently from the particular particle species.

    From atomic physics it is known that the energy loss of the primary particle is not simply divided into very small packages just creating ion–electron pairs like pearls on a string. The μ-electrons are ejected with a spectrum of kinetic energies (Bethe, 1930) and can cause further ionization at reasonable distances from the primary track (Kobetich and Katz, 1968a). This was confirmed by the observation in nuclear track physics of extended tracks of several micrometer in diameter recorded in nuclear photo emulsions after their exposure to highly energetic particles from cosmic rays or from accelerators (Kobetich and Katz, 1968b). The diameter of these tracks can be much larger than the dimensions of critical structures inside cells or even larger than a complete cell. In consequence, the three-dimensional distribution of ionization events around the particle track and the effects of the elevated ionization density must be known for a complete description of radiation effects of particles. Up to now, this has not been achieved.

    There are two reasons for this failure. First, the collision process between the projectile and the target atom with its electron is not understood in detail; i.e., the energy and the angular dependence of the emitted δ-electrons has not been measured for heavy ion impacts on solid targets and only a few measurements for gas targets exist. Therefore, it is not known precisely how the primary ionizations events are formed along the particle track and what fraction of the energy is transported via the electrons outside the primary collision region. Second, the fate of the emitted electrons is also not known. These δ-electrons are scattered both elastically without significant energy loss but changing direction and inelastically creating new ionization events or excitations. Even very sophisticated Monte Carlo calculations (Paretzke, 1988) of track structure use mostly simplified cross sections for the creation of electrons and their angular and energy distribution, and have difficulties in estimating the correct elastic and inelastic cross sections for the electrons transport.

    In these calculations the radius of the track is determined by the maximum range of the high energy part of the continuous δ-electron distribution. Although these electrons are emitted preferably in the forward direction, the multiple scattering of the slowing down process homogenizes the angular distribution to a large extent. In a few extreme cases, electrons emitted initially in the forward direction are later found moving perpendicular to the projectile trajectory and even in backward direction after several collisions.

    The shape of the radial dose profile depends essentially on the energy spectrum of the primarily produced electrons. Highly energetic electrons having a long range produce the outer part of the radial dose profile, and low-energy electrons with a short range produce the inner part. Because much more low-energy electrons are produced, the dose deposited close to the center of the track is higher than the dose close to the maximum radius. In general, the dose decreases from the inner part of the track proportional to the square of the radial distance from the center over many orders of magnitude up to the maximum radius. Monte Carlo calculations starting from reasonable electron spectra and calculating the energy dissipation of a sufficient number of electrons are in good agreement with the few existing measurements of radial dose distributions (Krämer and Kraft, 1992). Both experiments and calculations produce the same radial dependence over the major part of the track, i.e., a 1/r² dose dependence. In the calculations no indication is found of an additionally increased energy deposition in the center of the track as assumed in some models. The same holds for the experiments where an indication of leveling off for very small radial distances is found.

    In some models (Chatterjee and Schaefer, 1976; Kellerer, 1977) the inner part of the track is called the track core and is treated separately from the outer part, which is called the penumbra. In these models an equipartition of the energy loss between the core and the penumbra is assumed with a homogenous dose distribution inside the core and an 1/r² dependence for the penumbra. More recent experiments in solid-state physics explored the structure of the inner part of the track in greater detail (Albrecht et al., 1985). Because of the high degree of ionization caused by the primary projectile, the track center can be described as a hot plasma at least for the heavier ions in which specific plasma reactions like the Coulomb explosion of the target atoms and the recombination of electrons and ions take place (Groeneveld et al., 1980; Klaumünzer et al., 1986). But these experiments did not indicate an elevated energy deposition in the track center, as postulated in the core – penumbra model. There is of course a higher efficiency of atomic displacement in the central region yielding a higher probability for the disruption of biochemical bonds. However, the volume affected by Coulomb explosion is very small compared to the region where radiobiological effects are caused by the electrons. Core effects could not be identified in radiobiological experiments.

    Up to now there has been no unique theory describing radiation effects on the basis of first principles, neither for the complex system of a cell nor for the much simpler system of pure DNA in a buffer solution. Consequently, this chapter does not provide such a theory, but the different steps of the energy deposition by electronic and nuclear collisions, the production of electrons in these collisions, and their angular and energy distribution will be described as examples and the basis of track structure calculations will be presented. Basic principles will be explained but it is not the purpose of this chapter to present all physical details. This chapter is for the fairly interested radiobiologist; the reader more interested in basic physics is referred to the original references.

    II Linear Energy Transfer and Energy Loss

    A Dose – LET Relation

    The dose deposited by particles in a target is determined by the number of traversing particles and the LET (Zirkle, 1954) of each particle. If the dimensions of a biological target are larger than the diameter of the individual tracks and if no restrictions on the energy transfer, whether coming from high- or low-energy electrons, are applied, then the LET is identical with the total energy loss dE/dx of the heavy charged particle. The dose can be calculated using the formula.

    (1)

    where LET = dE/dx is the linear energy transfer and equal to the energy loss dE/dx; F is the particle fluence; and p is the density of the stopping material, i.e., 1 g/cm³ for water.

    This expression can be used for track segment experiments, i.e., for experiments in which the LET does not change significantly during the passage of the particle through the biological target. For ions of low specific energies of a few megaelectronvolts per mass unit, the variation of LET is very large and therefore only very thin biological targets (Kraft-Weyrather et al., 1989) like cells in monolayer or thin DNA films can be used for track segment experiments. For high-energy ions of several 100 MeV/u, track segment conditions can be realized even for an extended target. In the case of thick samples—thick compared to the track length of constant LET—the values of LET have to be averaged and this mean value can be used to calculate the dose. Two different LET averages can be used: track average LET and dose average LET [International Commission on Radiation Units and Measurements (ICRU), 1970]. However, using mean values instead of the distribution functions introduces large errors, especially when the distributions are broad and asymmetric and the biological response functions are nonlinear. Therefore, in radiobiological experiments track segment conditions should always be considered in order to ensure unique interpretation of the data.

    B Energy – Velocity Relations

    Because energy, specific energy, and velocity are used very frequently for the calculation of LET, the physical relationship between these parameters will be given in the nonrelativistic and relativistic limit together with some examples of how to use these formulas.

    At low energies in the nonrelativistic limit, the particle energy E is related to the velocity v by the formula E = ½ mv² (m = mass of the ion). Frequently the velocity in terms of the velocity of light is used even for nonrelativistic energies. In this nonrelativistic limit, the velocity β = v/c can be approximated by

    (2)

    with the rest mass E0 = mc². For an ion of A nucleons, E0, = A × mc² = A × 935 MeV. Introducing the specific energy Espec = Eskin/A,

    (3)

    For example, for a 20 MeV/u ion,

    That is, for a 20 MeV/u particle, the velocity is 20% of the velocity of light.

    For relativistic energies this approximation is not valid and the relativistic istic expression for β has to be used:

    (4)

    with the total energy E = Ekin + E0 and the rest energy Mc². Again introducing the specific energy, Espec, and replacing the total mass by the proton mass times the number of nucleons M = mp × A,

    (5)

    For an ion with 20 MeV/u specific energy, the relativistic calculation yields

    (6)

    The difference between the relativistic and the nonrelativistic calculation is less than 2% for 20 MeV/u. In both cases the velocity of 20 MeV/u ions is approximately 20% of the velocity of light. The numerical identity between specific energy and velocity is just a coincidence, but it will help the reader remember this value. (But a specific energy of 50 MeV/u does not correspond to 50% of the velocity of light!) The correct relationship between specific energy and velocity β is graphed in Fig. 1. It should be noted that the velocity depends only on the specific energy and is independent of the mass or atomic number of the ion: Particles having the same specific energy have the same velocity. This is a major reason for the use of specific energy instead of the total energy in the context of energy loss calculations in radiobiology.

    Fig. 1 Relation between the velocity and the specific energy of relativistic particles.

    C Projectile Velocity and Charge State

    The particle velocity is the main parameter governing the stopping processes and determines, first, the charge state of the projectile and, second, the type of interaction mechanism between the projectile and the target.

    If the velocity of the projectile is greater than the orbital velocity of its own electrons, then the electrons of the projectiles will be stripped off by the interaction with the target material (Bohr, 1913).² Because of the shell structure of atoms, the weakly bound outer shell electrons are stripped at low projectile velocities, whereas at higher velocities electrons from the inner shells are increasingly involved in particle interaction. If the projectile velocity is much higher than the orbital velocity of the innermost shell (K-shell) all electrons are stripped off and the charge of the projectile, which is effective in the collision process, is the positive charge of the atomic nucleus alone. The effective charge as a function of energy is approximated by semiempirical formulas. Frequently used is an expression by Barkas et al. (1963):

    (7)

    In Fig. 2 the effective charge of some representative ions is plotted as a function of the specific energy. According to the higher orbital velocities of the electrons bound to the heavier ions, higher projectile energies are needed to strip off the electrons and to reach an effective charge equal to the atomic number. For uranium ions a velocity much larger than β = 0.45 is necessary to reach this state. This corresponds to specific energies of approximately 500 MeV/u or more.

    Fig. 2 Effective nuclear charge as function of the specific energy for different atomic numbers.

    In addition to the charge state of the projectile, the interaction mode between projectile and the target electrons and atoms is determined by the projectile velocity. When the projectile velocity is large compared to the velocity of the target electrons (mostly the outer electrons), the projectile will lose its kinetic energy mostly by ionization and excitation processes of these target electrons. In this case, the prominent stopping mechanism is the electronic energy loss. The electronic energy loss reaches its maximum between the Bethe–Bloch region (Bethe, 1930; Bloch, 1933a,b), where the stopping power is proportional to 1/E, and the Lindhard – Scharf – Schiott (Lindhard et al.

    If the velocity of the projectile is lower than the velocity of the target electrons, the elastic scattering on the screened Coulomb potential is the predominant process of energy dissipation, called nuclear stopping (Schiott, 1966).

    In Fig. 3 the dependence of the LET (energy loss) on projectile energy is shown qualitatively. The different maxima indicate two different mechanism of energy dissipation: nuclear stopping at low specific energies and electronic stopping for higher energies.

    Fig. 3 The two mechanisms of energy loss, nuclear and electronic energy loss, are shown schematically as function of specific energy. At 10 keV/u nuclear stopping predominates. The maximum of electronic stopping is located at around 1 to 10 MeV/u. In the Lindhard–Scharf–Schiott (LSS) region the stopping power is proportional to the velocity, in the Bethe–Bloch region 1/E.

    D Electronic Stopping Power

    The major part of the energy dissipation of a heavy particle is due to the interaction with the target electrons. In 1913 Bohr calculated the stopping power dE/dx using a classical approach for small energies:

    (8)

    For velocities comparable to the velocity of light, the stopping power in this classical calculation is given by

    (9a)

    where e is elementary charge, Z1/Z2 is projectile/target atomic number; m is electron mass; v is the average orbital frequency of the target electrons calculated using

    (9b)

    Bohr used two extreme situations for the derivation of this formula: For small impact parameters, i.e., for very direct hits (knock-on collisions), the energy transfer is large compared to the binding energy. Therefore, the target electrons can be regarded as quasi free. For large impact parameters (glancing collisions), the energy transfer is small compared to the binding energy and the electrons are regarded as bound harmonic oscillators. However, the approximation of the collision process by these two extreme situations is used only to find the correct formula. It does not mean that the reality of the collision process can be separated into these two categories of glancing and knock-on collisions only. In reality the impact parameters are statistically distributed and all radial distances occur according to their geometric probability.

    The necessity to average over the impact parameter in the classical calculation was one of the important motivations for Bethe in 1930 to calculate the stopping power using quantum mechanics. In the first Born approximation he yielded for the stopping power

    (10)

    with I = mean ionization potential and

    (11)

    where fn = oscillator strength of the nth state.

    Finally, Bloch (1933a,b) developed a universal expression for the stopping power, which covers both high and low energy regions,

    (12)

    where Ψ is the logarithmic derivative of the Γ function.

    For the ionization potential, Bloch used a Thomas Fermi approximation:

    (13)

    The main difference between the classical Bohr (1913, 1948) and the quantum mechanical Bethe-Bloch approximation is the logarithmic term, which contains no Z1-dependence in the Bethe formulation. However, in all these formulas, the ionization potentials enter in the denominator of the logarithmic term, which means that the outer electrons having larger quantum numbers, i.e., the weakly bound electrons, play the largest role in the stopping process.

    These formulas are calculated for a fixed projectile charge number. For their application to heavy ions, this number has to be replaced by the actual charge of the projectile inside the stopping material, i.e., the effective charge as defined in the last section: Z1→Zeff. In addition, density corrections, which describe the transition from gas to solid, as well as contributions from shell corrections have to be added to these formulas (Salamon, 1980).

    In practice, when radiobiological parameters are to be calculated or plotted as a function of the unrestricted linear energy transfer (= dE/dx), tables that are based on the Bethe – Bloch expression and adapted to the experimental data are normally used. Widely used tables are those of Northcliffe and Shilling (1970) for low energies (E ≤ 10 MeV/u), those of Hubert et al., (1980, 1990 for medium energies (E ≤ 500 MeV/u), and those of Benton and Henke (1967) and Heinrich et al. (1991) for high energies (up to 2 GeV/u). In Fig. 4 these tables are compared, including measurements of Geissel et al. (1982a, 1983).

    Fig. 4 Comparison of the energy loss in carbon (p = 2.26 g/cm³). (—), Heinrich et al. (1991); (—), Northcliffe and Shilling (1970); (x-x-x), Hubert et al. (1990); (•), experimental data of Geissel et al. (1983).

    In the maximum LET region around 2–10 MeV/u, 30% differences between the Northcliffe and Shilling extrapolation and the measured value of Geissel as well as the tables of Hubert et al. (1980, 1990) and Heinrich et al. (1991) exist for the heavier ions like uranium. For the relativistic energies the Heinrich data are also in good agreement to values calculated by Benton and Henke (1967). For the lighter ions like carbon or neon, the experimental data and all calculations fit very well to each other.

    Restricted and unrestricted LET values for water are given in Fig. 5. The restricted LET is defined as the part of the total linear energy loss of a charged particle that is due to energy transfers up to a specified energy cutoff value (ICRU, 1970). Frequently used energy cutoffs are 100 or 200 eV. Although this definition specifies an energy cutoff and not a range cutoff, the restricted LET values are frequently associated with the energy deposition close to the central region of the particle track (energy locally imparted).

    Fig. 5 Restricted and unrestricted energy loss in water according to Heinrich et al. (1991). For energies greater than 10 MeV/u LET,100 and LET500 parallels LET ∞ because the low-energy and high-energy part in the spectra of the δ-electrons are increasing nearly proportional to each other.

    E Braggs Rule for Complex Target Molecules

    The most frequently used absorber in radiobiology is water. However, some of the LET tables do not contain values for water or other molecules as an absorber material. The stopping powers in compound absorbers can be calculated according to the Bragg rule. There is good evidence that the Bragg rule of additivity relating the stopping power of a compound to that of its constituents is a fairly good approximation at least at energies above 5 MeV/μm (Northcliffe and Shilling, 1970). According to this rule, the stopping power of a compound of molecular weight M and containing Ni atoms of the atomic weight Ai is given by

    (14)

    This rule yields for water

    (15)

    The LET for DNA can be calculated in the same way. The Bragg rule, however, shows its major deviation from experiments for hydrocarbon compounds. Therefore, water, which is the major constituent of DNA itself and its surrounding environment, is always a good approximation in radiobiology. If only a first approximation with an accuracy of 10% of the LET is needed, the stopping power values of carbon can be used but scaled with the density of water. These values are consistently 10% lower than the water values for energies between 1 and 1000 MeV/μm.

    F Nuclear Stopping

    The nuclear stopping contributes only a small part to the total stopping process at medium or high energies but plays a significant role at low energies (E ≤ 10 keV/u) at the very end of the particle track. The nuclear stopping can be neglected in the analysis of most experiments and it is treated here more or less for the sake of completeness. The nuclear stopping process can be understood as the Rutherford scattering of the projectile at the screened potential of the target atom. On the basis of a Thomas Fermi-type screened potential, Lindhard, Scharf, and Schiott (LSS-model) found a universal dependence of the stopping cross section when energy and range are measured in reduced Thomas Fermi units (Schiott, 1966),

    (16)

    where

    (17)

    and

    (18)

    where M1, M2 are the masses of projectile and target, respectively; Z1, Z2 are their atomic numbers; N is the number of the target atoms; a is the Thomas Fermi screening parameter; a= 0.8853 a0(Z²/³1 + Z²/³2)¹/²; and R= range.

    In these reduced units the nuclear stopping is a universal function and is independent of the absorber material and the projectile as shown in Fig. 6. Numerical values can be found in the publication of Schiott (1966).

    Fig. 6 Nuclear stopping in reduced LSS units according to Schiott (1966).

    G Radiobiological Efficiency of Nuclear Stopping

    In a few experiments with protons and heavier ions, different biological endpoints like the inactivation of φX174 DNA and ribonuclease (Jung and Kürzinger, 1968, 1969), the induction of chromosome aberrations and the inactivation of mammalian cells (Kraft et al., 1983), and the inactivation of bacteria spores and yeast cells (Schneider et al., 1990) have been measured for very low particle energies. For all these different endpoints a decrease of the action cross section with decreasing energy that parallels the electronic stopping power was observed. However, for energies below 10 keV/u, a steep increase in the cross section toward lower energies is observed again. This second maximum of the cross section is attributed to the nuclear stopping power and it reaches absolute values comparable to the maximum values of the electronic stopping power. The high efficiency of nuclear stopping compared to electronic stopping is due to a different reaction mechanism. In the electronic stopping process, single electrons are ejected from biomolecules, but this does not necessarily lead to a destruction of the chemical binding and to the disintegration of the molecule. Missing electrons can be restituted before the chemical bonds are broken. In nuclear stopping processes the projectiles interact with the complete target atom and the target atom will be kicked out of the chemical surrounding if enough energy is transferred. Damage to the biomolecule is therefore very likely; RBE values greater than 4 have been reported in the range of nuclear stopping (Jung and Kiirzinger, 1968) and cross sections exceeding the size of cell nuclei (Schneider et al., 1990).

    III δ-Electron Emission

    A Classification of the Emission Processes

    If electrons were produced in the ionization processes with zero or nearly zero kinetic energy, then all the ionization events would occur in the trajectory of the primary particle like pearls on a string. The number of pearls per string length, i.e., ionization events per track length, may be calculated from the LET divided by the average energy needed to create an ionization event (ion – electron pair). As already pointed out by Bethe in 1930, the major fraction of the primary energy deposition is converted into the kinetic energy of the liberated electrons. These electrons leave their origin in the central part of the track and transport a large fraction of their energy close to the end of their range. The primary energy distribution of the electrons together with the energy loss of the electrons determines therefore the radial dose distribution of the particle track and the track diameter.

    A typical spectrum of electrons emitted in heavy ion collisions is shown in Fig. 7. The spectrum consists of a broad continuum overlayed by some line structures and has a very pronounced angular dependence, which is different for each of the features shown in the spectrum. The various structures originate from the following processes (Schmidt-Böcking et al., 1992):

    Fig. 7 Typical δ-electron spectrum of a heavy ion atom encounter according to Stolterfoht et al. (1974). For detailed explanation see text.

    1. Continuous δ-electron emission from the interaction of the projectile with quasi free electrons (two-body interaction).

    2. The continuous δ-electron emission from the interaction of the projectile with bound electrons (three-body interaction).

    3. Auger electrons from the multiply excited target atom.

    4. Auger electrons from the excited projectile atoms.

    5. Electron loss from the projectile.

    (1 + 2) originate from ionization processes caused in the target atom by the projectile and produce continuous δ-electron spectra; (3 + 4) are auto-ionization processes of target and projectile atoms and produce discrete electron maxima; (5) are the electrons traveling in a convoy with the projectile having the same velocity

    There are also other electron-emitting processes in heavy ion atom collision. For instance, for very heavy collision systems having Z1 + Z2 < 132, innershell vacancies can be filled by the production of an electron-positron pair from which one partner is emitted in the collision. Also, coincident electron-positron emission that might be attributed to the decay of collective electron excitation in the solid target has been found. However, these processes are not understood in detail (Kienle, 1988) and their cross sections are extremely small. Therefore they will not contribute to the biological effect.

    B Continuous Electron Emission

    The contributions to the different electron continua depend on the impact parameter, i.e., the distance of closest approach, between the primary projectile carrying an effective charge e’Zeff and target electrons. For small impact parameters, i.e., for small distances of closest approach, the energy transfer from the projectile to the target electrons is large compared to the binding energy of the target electrons. In this case the collision process can be treated in the binary encounter approximation (BEA) (Bonsen and Vriens, 1970; Folkmann et al., 1975).

    In a binary encounter, two unbound particles are scattered like classical billiard balls. The emission angle and particle energy are correlated as follows:

    (19)

    Frequently used as a rule of thumb for the maximum electron energy is Ee[KeV] = 2 × Espec [MeV/u]. In the scattering process of the bound target atoms by the projectile, the electrons of the target initially have a momentum distribution that depends on the quantum number of the initial state. Therefore, in the BEA calculation the cross section for electron emission at an angle θ with an energy transfer Δ = E + Ebind is given by

    (20)

    where f(ve) is the distribution function of the electrons in the different quantum states of the target (Bonsen and Vriens, 1970) and the scattering cross section is given by

    (21)

    where v’e is the velocity of the scattered electron and the indices e and p refer to electron and projectile, respectively. In Fig. 8 measured and calculated electron spectra are compared.

    Fig. 8 Comparison between measured spectrum (—) and BEA calculation [redrawn from Folkmann et al. (1975)].

    If the impact parameter between projectile and target electron is large (glancing collision), the energy transferred to the electron is small and the electron cannot be regarded as quasi-free. This collision process cannot be calculated with BEA and has to be treated as a three-body interaction between projectile, target electron, and target nucleus. In this three-body interaction the electron is scattered twice: first by the projectile and second by the target nucleus. The electrons are consequently emitted in all angles between 0° and 180° and the angular distribution is nearly isotropic in the lab system. For heavy ions, however, the three body electron emission can be anisotropic too because of the strong polarization introduced by the high projectile charge (Kelbch et aL, 1985; Stolterfoht et al., 1987).

    C Auger Electron Emission

    In contrast to the continuous electron emission caused by the direct impact of the projectiles, the Auger electrons have a discrete energy and are emitted in an auto-ionization process. Due to the collision process, vacancies in inner atomic shells can be produced in both atoms, target and projectile. When these vacancies are filled by outer electrons, the difference in the binding energies can be emitted either as photons or as the kinetic energy of electrons if the energy is larger than the ionization energy. The Auger process can occur in either the target or the projectile atom. Since the target atom is practically immobile after the collision, the target Auger electrons are emitted isotropically in the lab system. For the projectile Auger electrons, the isotropic emission holds only in the center of mass system. In the lab system, the center of mass is moving, which results in an anisotropic emission pattern, as can be seen in Fig. 7. A detailed review of Auger spectroscopy has been given by Stolterfoht (1987).

    D Loss Electrons

    Due to the characteristic lineshape of their energy distribution, the loss electrons are also termed cusp electrons (Varger and Gemmell, 1976; Groeneveld et al., 1980). These electrons either originate from the remaining bound electrons of the projectile that are lost into unbound states [electron loss to the continuum (ELC)] or are target electrons captured to low energy (unbound) states in the projectile continuum [electron capture to the continuum (ECC)]. In both cases, ECC and ELC, the electrons are emitted with nearly zero kinetic energy from the moving projectile. Therefore, they are emitted in the forward direction in the lab system with a peak energy of

    (22)

    with me = electron mass and vp = projectile energy.

    In reality, the cusp electrons have to be considered as being scattered by the impact of the target nuclei onto the electrons of the projectiles. Small deviations from a strong 0° emission due to this effect are observed. However, the distinct feature of the cusp electrons is the strong forward peaked emission line centered around the particle trajectory (see Fig. 7).

    E Electron Emission in Heavy Ion Atom Collisions

    For lighter ions the emission characteristics follow the rules outlined above nicely. For heavy ions, however, some deviations for these rules are important (Schmidt-Böcking et al., 1991).

    1. Electrons emitted in Auger or loss processes contribute only a small fraction to the total electron emission, with the majority of the electrons being found in the continua produced by the projectile interaction with the target electrons.

    2. Due to the polarization of the target atom by the projectile charge, the electrons are emitted preferably in forward direction and fewer electrons are found to be emitted perpendicular to the trajectory of the projectile. At intermediate angles (around 60°), a small maximum of the electron emission has been observed in very heavy ion atom collisions (U→Ar) at low energies, which has been interpreted as rainbow scattering.

    3. Due to multiple ionization processes, frequently two or more electrons can be emitted simultaneously.

    4. Finally, the mean energy of the emitted electrons shifts to higher values for heavy ion impact compared to the effects of light ions. For example, for 1 MeV/u the mean electron energy for proton impact is centered around 10 eV, whereas for uranium impact it is at around 100 eV.

    Heavy ion effects on the electron emission have not been measured systematically up to now. However, the few measurements that exist can be reproduced by an n-body classical Monte Carlo calculation (nCTMC), in which the trajectories of all electrons and nuclei are calculated in a nonquantum mechanical limit (Olson et al., 1989). The unusual behavior of the heavy ions, especially the fact that more electrons are emitted in the forward direction, leads to a higher ionization density in the forward cone than that for lighter ions. This might have some biological consequences.

    F Experiments and Model Calculations

    In many track structure calculations the electron emission is treated in a simplified way (Butts and Katz, 1967; Kiefer and Straaten, 1986; Paretzke, 1988). The collision process is regarded as a two-body interaction between the projectile effective charge and the unbound electron electrons of the target. A straightforward calculation of basic Coulomb scattering yields an expression for the energy distribution of the emitted electrons [for detailed derivation see, for instance, Spohr (1990)].

    (23)

    with

    (24)

    where dn is the number of electrons produced in the energy interval between ε and ε + ; Zeff is the effective charge of the projectile; β = v/c the velocity of the projectile; and N = the electron density in the target material.

    Together with the kinematics of the Coulomb scattering, this formula yields a strong maximum for electron emission perpendicular to the projectile trajectory. In some model calculations (Butts and Katz, 1967), it is therefore assumed that the electrons are always emitted only perpendicular to the projectile trajectory.

    In contrast to these simplifying assumptions, the experiments (Folkmann et al., 1975; Toburen, 1982; Kelbch et al., 1985) and the theories explaining these experiments yield a strong forward emission characteristic. In some cases (Schmidt et al., 1989; Schmidt-Böcking et al., 1992), more than 90% of all electrons are emitted into a cone of 60° in forward the direction (Fig. 9). The fairly good agreement of models of radial dose distributions with perpendicular emission characteristics with experiment is due to the fact that the electrons suffer many angular deflections during the slowing down process in the target material. This diffusion process homogenizes the angular distribution after a few collisions to a large extent and the propagation of the electrons does not depend too much on the starting conditions. In addition, the major fraction of the dose originates from low energy electrons, which are emitted more isotropically.

    Fig. 9 Angular plot of the double differential cross section for electron emission. In the extreme case of very heavy ion impact (U–Ar) the angular distribution shows a strong forward-directed emission characteristic. The maximum at 60° is due to multiple scattering in the target atom ( Schmidt et al., 1989).

    Therefore, dose profiles can be correct even if starting conditions with unrealistic angular distribution have been used.

    However, it is evident that the strong forward-peaked emission will yield higher ionization densities in the inner part of the track. This is important because the changes in the biological efficiency of high LET radiation are correlated with the variation of ionization densities in the particle tracks. In addition, differences between gas and condensed phase may also be relevant to the primary effects. But up to now nearly no experimental data exist for the electron emission from solids.

    IV Condensed Phase Effects — Track Core

    A Differences between Gas and Solid Targets

    Radiobiology deals with radiation damage in tissues or cells, i.e., in material that is not in the gas phase but in condensed phase. Therefore, differences in the track formation between the gas and the condensed phase are of basic interest. Tracks in gas can be visualized in a cloud chamber where the ionization events of the primary particles are center-points of condensation of small droplets of water. Isolated ionization points lead to isolated droplets, as have been photographed in many cases. High ionization density yields a continuous track, but on a microscopic scale there is no significant interaction between the different ionization events in the continuous track.

    This changes drastically when the density of the target is increased. The atomic configuration of the projectile, the interaction of the projectile with target atoms, and the target–target interaction are all different. The collision frequency increases with the target density and the average time between two consecutive collisions decreases. When this time is shorter than the time constant for deexcitation, the electrons remain in excited states and the electronic configuration of the projectile is different from a ground state configuration. This has been measured in X-ray experiments (Kraft et al., 1977) as well as in deexcitation by Auger emission when the projectile is leaving the solid target (Betz, 1972). Electrons of the projectile can be transferred to unbound states by step processes of two or more steps. This leads to ELC processes, as discussed previously. Therefore, the effective charge is different in a gas target than in a solid absorber and the corresponding differences in the stopping power have been measured (Geissel et al., 1983).

    Effective charges that are used for LET calculations and for the interpolation of LET values are mostly fitted to solid target data or calculated according to the Barkas formula, which is also valid for condensed targets only. For these data, no additional adjustments have to be made, for the application of the tabulated LET values to condensed matter.

    B Collective Excitations

    Highly charged particles induce ionization when they pass a neutral target atom. Even if no electrons are liberated by the heavy ion impact, the charge distribution of a neutral atom becomes polarized by the passage of a highly charged projectile. In the case of a gaseous target, the target atoms are insulated and the interaction of their polarization is neglectible. In a solid target, atoms of similar polarization states are close together and interact forming a common potential (Gemmell, 1979). Because the polarization is caused by the projectile ion, this potential travels like a wake behind the projectile. The existence of this plasmon wake has been confirmed experimentally using molecular ion beams that traverse solid carbon foils (Varger and Gemmell, 1976). Entering the foil, molecular bound electrons are stripped off and the two separated ions repulse each other via Coulomb interaction. However, after a short travel distance, a strong orientation of the system is observed. Due to the electronic polarization of the target atoms, one of the ions is focused into the minimum of the plasmon potential and both ions travel as a convoy. Although this collective excitation has been found in ion experiments and not in experiments where the emission of electrons has been observed, it is evident that the strong polarization behind the projectile also has a strong influence on the electron emission. One consequence is the enhanced emission of cusp electrons, which are called convoy electrons in the case of solid targets. It may also be possible that deviation from the recently observed angular emission characteristics observed in gases occurs in solids and that more electrons are emitted in the direction of the projectile. Such a forward-peaked emission characteristic could contribute to the Coulomb explosion mechanism, because these primary electrons together with their secondaries will produce a higher degree of ionization in the center of the track.

    C Coulomb Explosion

    When the valence electron of a gas molecule is stripped off, the remaining atoms separate via Coulomb repulsion. In a solid, in particular in a crystal, the atoms can only leave their regular positions if the energy ED necessary for a displacement is smaller than the Coulomb energy Ecoul (Fleischer et al., 1975; Groeneveld et al., 1980)

    (25)

    with q1 q2 being the product of the charges of the two ionized atoms and r0 their separation.

    Assuming for the displacement energy a reasonable value of 20 eV and for the interatomic distance r = 2 Å, a Coulomb explosion inside a solid can occur if the following condition is fulfilled

    (26)

    (27)

    This means that the Coulomb explosion occurs along a particle track if one or two electrons (in average 1.6) are ionized in subsequent ion atom collisions. A schematic illustration of the Coulomb explosion mechanism is given in Fig. 23 (see p. 47). For the lighter elements like carbon, oxygen, nitrogen, the valence electrons are located in the 2p-shell and have typical binding energies of 10 to 20 eV. For the liberation on average of 1.5 electrons, approximately 30 eV from the energy of the projectile is needed. Since a greater fraction of the

    Enjoying the preview?
    Page 1 of 1