Chapter Four - One electron-capture in collisions of fast nuclei with biomolecules of relevance to ion therapy

https://doi.org/10.1016/bs.aiq.2021.03.001Get rights and content

Abstract

This work is on total cross sections for single electron capture by protons from atomic and molecular targets. Most of the included atomic targets are also the constituents of the examined molecular targets. The latter targets are selected to be biomolecules. Computations are carried out using the continuum distorted wave (CDW) method at intermediate and high impact energies. First computed are atomic cross sections. These are subsequently employed to obtain molecular cross sections in the independent atom model (the Bragg additivity). Atomic targets are the backbone of DNA/RNA, i.e. hydrogen H, carbon C, nitrogen N and oxygen O. Also considered are neon Ne, argon Ar and krypton Kr. The studied molecular targets are H2, water vapor H2O, carbon monoxide CO, carbon dioxide CO2, methane CH4, ethylene C2H4, ethane C2H6, butane C4H10 as well as the DNA/RNA nucleobases, uracil C4H4N2O2, adenine C5H5N5, guanine C5H5N5O, thymine C5H6N2O2 and cytosine C4H5N3O. Neon atoms are isoelectronic with water molecules. This permits gaining some useful insights into the role of molecular forces by analyzing the cross sections for neon and water targets. The cross sections for each atomic target shell are summed up and compared with the available experimental data for atomic and molecular targets. The usefulness of the Bragg sum rule is assessed for molecular charge-exchange cross sections. Besides atomic physics, interest in this problem area exists also in heavy ion transport physics of relevance to plasma physics, astrophysics, the fusion program and radiotherapy. For example, in ion therapy, when the primary beam particles reach their deepest penetration distance, they release almost all their energy and this is biologically and clinically of utmost importance. In this targeted region, ions slow down abruptly and move with energies 20 keV/amu ≤ E ≤ 150 keV/amu. Therein, capture and loss of electrons are dominant energy loss mechanisms yielding maximum dose depositions. The present databases for electron capture cross sections can be used in Monte Carlo simulation of ion energy losses in matter, including tissue.

Introduction

The origin of radiotherapy was the measurements by Bragg and Kleeman [1, 2] in 1904/1905. They showed that α-particles exhibit an increased ionization yield near the maximum penetration distance (the so-called range). The measured stopping power, as the distribution of ion energy loss per distance, was almost flat for most of the traversed path and exhibited a sharp maximum at the end of the particles’ track. Such a maximum (later called the Bragg peak) was at variance with the common conception that ion energy loss distribution should decay exponentially with distance.

Regarding electromagnetic interactions of ions with matter, energy losses of ions can occur in various ways. Different channels open up depending of the ion energy. Ionization dominates over e.g. electron capture at high ion energies E. However, when ion energies are considerably reduced, the dominant ion energy loss channels are electron capture and electron loss that occur concomitantly and interchangeably. It is because of this ‘open-channel-switch’ that most of the remaining ion energy is released at the Bragg peak, which corresponds to the lower edge of the intermediate energy region located at 20 keV/amu ≤ E ≤ 150 keV/amu.

In hadron therapy, the entrance energy of the primary beam is selected to reach the target location. The Bragg peak resides at the targeted tumor site. Around the Bragg peak, the radiation dose delivered by ions attains its maximum and this eradicates tumorous cells. Therefore, the immediate surrounding of the Bragg peak is biologically and clinically the most important area for treating the patient's tumor. Moreover, healthy tissue is largely spared in the flat domain prior to the Bragg peak. Such overall favorable dose-depth profile characteristics of ions optimize the net effect of hadron therapy. This is how the physics concept of ion transport in tissue gave impetus to the development of optimized radiotherapy in medicine.

Understandably, due to the unavailability of accelerators, this was not the original intention of Bragg and Kleeman [1, 2]. However, four decades after their experiments, Wilson [3, 4], the former head of the experimental part of the Manhattan Project, was the first to convey to physicians the far-reaching power of the Bragg peak for radiotherapy. To connect to physicians directly, he decided to publish his 1946 article [3] in a prestigious medical journal – Radiology. He explained convincingly that with the advantageous depth-dose profile of ions, physicians would be able to achieve a much better tumor control than with either electrons or photons. He emphasized that because of the optimal conformity of ion dose profiles to the target, tumor would receive maximum radiation dose, with a minimal damage to the surrounding healthy tissue. This Wilson's vision marks the beginning of optimized radiotherapy, the precision ion therapy.

Physicians wasted no time to implement this proposal in collaboration with physicists and already in the 1950s the first patients underwent ion therapy in Berkeley and Uppsala. Nowadays, some seven decades later, ion therapy is the workhorse of radiotherapy with more than 100 accelerators operating worldwide and more than 150000 patients treated by protons (85%) and carbon nuclei (13%) alone [5]. In the accompanying research, much attention has been paid to find the most appropriate ions for tumors at different sites in the human body and the main focus was on comparative analyses of protons and carbon ions [6, 7, 8, 9].

Ions through their direct hits produce about 30% of radiation damage to molecules of the targeted tumor tissue. The remaining damage is made by indirect hits via secondary particles (e.g. δ-electrons), free radicals and reactive oxygen species. In radiobiology, the main radiation targets are considered to be the deoxyribonucleic acid (DNA) molecules. Particularly within the Bragg peak, ions can massively disrupt the internal structure of DNA and inflict lethal, irreparable double strand breaks in these ‘molecules of life’.

Such radiation damage to DNA molecules is caused by energy deposited through electromagnetic and nuclear interactions of ions with tissue. Due to the exceeding complexity of these collisions, deterministic evaluations of energy losses of ions during their passage through tissue are of limited usefulness in practice. For this reasons, comprehensive computations of ion energy losses are usually performed by Monte Carlo (MC) simulations. The main input data to stochastic simulations are deterministically computed electronic and nuclear stopping powers (the former and the latter are concerned with electromagnetic and nuclear interactions, respectively) [9, 10, 11, 12, 13, 14]. On the other hand, the principal ingredient of stopping powers is the appropriate set of atomic, molecular and nuclear cross sections.

There is a number of MC codes in use for hadron therapy, e.g. PTRAN [15, 16, 17, 18], PETRA[19, 20], FLUKA [21], KURBUC [22, 23, 24], SRNA [25, 26, 27, 28, 29, 30], PATRAC [31, 32], SHIELD-HIT [33, 34, 35, 36, 37], SEICS [38], GEANT4-DNA [39], TILDA [40], etc. The first MC code for purposes of radiotherapy was PTRAN [15]. Three MC codes for ion therapy have been developed at the Karolinska Institute in Stockholm: PETRA [19], SHIELD-HIT [33] and KURBUC-carbon [24].

In particular, the MC code SHIELD has originally been designed and implemented by Sobolevsky [41, 42] with no inclination to radiotherapy whatsoever. Later, the first extension of SHIELD to heavy ion therapy (HIT), known as SHIELD-HIT [33], has been developed in a cooperative research project involving the Institute for Nuclear Research of the Russian Academy of Sciences (Moscow), The Nobel Medical Institute – Karolinska Institute, Stockholm (Division of Radiation Physics) and Stockholm University (Department of Medical Physics).

There are two parts that need to be substantially improved in the SHIELD-HIT code. One part is an appropriate and full upgrade of the Bethe-Bloch formula for electronic stopping powers that have, thus far, been empirically extrapolated below the quite high cut-off energy, Ecut-off ∼ 1000 keV/amu. No charge-exchange nor electron loss collisions are included, both of which dominate over ionization and excitation in the clinically most important region around the Bragg peak at 20 keV/amu ≤ E ≤ 150keV/amu. The other part is the essential inclusion of transport of electrons, ejected from the tissue traversed by ions. Especially δ-electrons are crucial as they are by far more effective in imparting radiation damage to the targeted tissue than any ion (primary, secondary or higher-generation ions due to nuclear transmutations of the tissue nuclei and/or the projectile nuclei heavier than protons).

For ion-atom and ion-molecule collisions, the main mechanisms for energy losses of the primary beam particles during their passage through matter are ionization, excitation, electron capture and electron loss. The present focus is on single electron capture by ions from atomic and molecular targets. Both atomic and molecular targets are selected to be of relevance for ion therapy. Thus, the main atomic targets of our interest are hydrogen H, carbon C, nitrogen N and oxygen O. Moreover, these atoms are also the constituents of the presently chosen biomolecules as molecular targets (water, carbon oxides, hydrocarbons and DNA/RNA nucleobases).

The total cross sections for ion-atom/molecule collisions are presently computed using the continuum distorted wave (CDW) method at intermediate and high energies. For charge-exchange and ionization, this most complete distorted wave theory has been introduced by Cheshire [43] and Belkić [44]. For charge-exchange, the CDW method has been shown to be in excellent agreement with measurements above about 50 keV/amu or so on the average [45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61].

In the first detailed applications of the CDW method for capture into the ground and excited final states, tedious parametric derivatives of the transition amplitude Tif(CDW) have been carried out. Therefore, this technique could be done for a handful of transitions. For example, in computations of cross sections Qif(CDW), Belkić et al. [49] used partial differentiations of the transition amplitude only for the orbital/quantum numbers 1 ≤ ni,f ≤ 3 (0 ≤ li,fni,f − 1, − li,fmi,f ≤ +li,f) and ni,fli,f = 4s. Later, based on these limited analytical results for Tif(CDW), a computer program for computations of Qif(CDW) has been published [62].

However, the breakthrough in a wider usage of the CDW method came with the availability of the analytical results of the general transition amplitudes Tif(CDW) in the case of arbitrary i and f. The necessary analytical results for the bound-free matrix elements with any quantum numbers for hydrogenlike and multi-electron atomic systems have been derived in Refs. [63, 64, 65, 66]. These closed expressions for Tif(CDW) have subsequently been programmed into the robust and automatic computer codes for any Qif(CDW) in the case of both hydrogenlike [67] and multi-electron [68] atomic targets. These programs supersede the earlier CDW code [62] based on parametric differentiations.

For multi-electron atomic targets, the available computer program in the CDW method [68] is flexible as it can employ the Roothaan-Hartree-Fock (RHF) wave functions in the analytical forms given by Clementi and Roetti [69] or Bunge et al. [70], both for atomic numbers 2–54, as well as by McLean and McLean [71] for atomic numbers 55–92. These computer codes from Ref. [68] are user-friendly since one can instantly switch from employing the RHF functions from Ref. [69] to Refs. [70, 71] merely by reading the input variational parameters from two different sets of tables. Importantly, the computer codes in the CDW method [67, 68] are remarkably expedient.

Even though the RHF wave functions of Clementi and Roetti [69] are widely used, they are not as accurate as those of Bunge et al. [70] for neutral atoms. The improvements made by Bunge et al. [70] relative to Clementi and Roetti [69] came from using a greater number of the Slater type orbitals (STOs) in the expansion basis set and from partially optimizing the principal quantum numbers n [72, 73].

The CDW computer programs [67, 68] have been built for atomic targets alone. However, they can also be used for molecular targets without any modification. To this end, all that is needed is to resort to the independent atom model (IAM) [74, 75, 76, 77, 78], or equivalently, the Bragg additivity [1, 2]. Atomic cross sections are computed first and then the sought molecular cross sections are deduced by the Bragg sum rule. This rule approximates a molecular cross section by the sum of the corresponding cross sections for each of the constituent atoms multiplied by the number of atoms in the given molecular target.

The underlying assumption of the Bragg additivity is that the molecular degrees of freedom (vibration, rotation, ...) can be ignored. In other words, by neglecting molecular forces, all the atoms in a molecule can be viewed as the independent centers of scattering. This is a reasonable hypothesis at higher impact energies at which the collision time is small which makes the projectiles unable to discern the finer internal structure of a molecular target. The Bragg sum rule has abundantly been used in the literature [79, 80, 81]. It met with success in e.g. charge-exchange involving protons and hydrocarbon molecular targets both at high energies (E ≤ 2500 keV) [82] and at intermediate energies (60–120 keV) [83].

In the case of atomic targets, besides the mentioned hydrogen H, carbon C, nitrogen N and oxygen O, we shall also consider neon Ne, argon Ar and krypton Kr. Molecular targets included in this study are molecular hydrogen H2, water H2O, two carbon oxides (carbon monoxide CO, carbon dioxide CO2), four hydrocarbons (methane CH4, ethylene C2H4, ethane C2H6, butane C4H10) and five DNA/RNA nucleobases (uracil C4H4N2O2, adenine C5H5N5, guanine C5H5N5O, thymine C5H6N2O2 and cytosine C4H5N3O). Uracil is from RNA (ribonucleic acid), whereas adenine, guanine, thymine and cytosine are from DNA. Neon atom is isoelectronic with water molecule. Therefore, it may prove instructive to compare the cross sections for these two targets to learn about the relative importance of molecular forces in H2O as a function of impact energy of projectiles [13, 14, 84, 85].

The atomic and molecular cross section databases to be supplied in this work are deemed to be of versatile use not only in atomic and molecular collision physics, but also in plasma physics and astrophysics [86, 87, 88, 89]. Likewise, they are very important in ion transport physics of relevance to several interdisciplinary areas ranging from the fusion research program for new energy to hadron therapy in medicine [9, 13, 14]. For example, in radiotherapy, dose-planning systems are always in need of improvement, which can come from better estimates of ion energy losses in tissue. For a long time now, much of the cross section input data to MC simulations has been computed using several empirical/phenomenological constructs based on either the asymptotic first Born approximation (as in the Bethe-Bloch formula) or on some mathematical ad hoc expressions (e.g. the Rudd formula) with adjustable parameters fitted to some experimental data from the given measurement. These drawbacks can advantageously become obsolete by employing e.g. the versatile CDW method, which is able to unequivocally provide the needed cross sections from the first principles of physics.

Atomic units will be used throughout unless stated otherwise.

Section snippets

Theory

Electron transfer from a hydrogenlike atomic target to an impinging multiple charged nucleus as a projectile of velocity υ is customarily symbolized by the process:ZP+(ZT,e)i(ZP,e)f+ZT,where the parentheses indicate the presence of the initial and final hydrogenlike bound states i and f of the electron (e). Here, ZP and ZT are the charges of the projectile and target nuclei, respectively. The masses of these three particles will be denoted by me, mP and mT, where mKme (K=P, T), with the

Results

The results from the CDW method on charge-exchange will now be presented in three separate sub-sections, 3.1-3.3. The first and the second sub-sections 3.1 and 3.2 deal with ion-atom and ion-molecule collisions, respectively. The third section 3.3 is concerned with both these collisions and it is devoted to the universal lineshapes of total cross sections. These scaled or ‘normalized’ cross sections are the cross sections from the first two sub-sections 3.1 and 3.2 divided by the number of

Conclusion

This study is on total cross sections computed by means of the continuum distorted wave method, CDW, for single electron transfer during proton-atom and proton-molecule collisions. The main focus is on intermediate and high impact energies, where the CDW method is known to be adequate. Regarding atomic targets, the present illustrations deal with hydrogen H, carbon C, nitrogen N, oxygen O, neon Ne, argon Ar and krypton Kr. As to molecular targets, several groups are investigated. Among these

Acknowledgments

This work was supported by the Research Funds from Radiumhemmet (via the Karolinska University Hospital) and from the Stockholm County Council (FoUU).

References (212)

  • RajD.

    A note on the use of the additivity rule for electron-molecule elastic scattering

    Phys. Lett. A

    ([object Object])
  • MożejkoP. et al.

    Elastic cross section calculations for electron collisions with XY4(X = Si, Ge; Y = H, F, Cl, Br, I) molecules

    Nucl. Instrum. Methods Phys. Res. B

    ([object Object])
  • YuY.C. et al.

    Energy loss straggling of energetic 3He and 6Li ions into polymer foils

    Nucl. Instrum. Methods Phys. Res. B

    ([object Object])
  • MančevI. et al.

    Theoretical state-selective and total cross sections for electron capture from helium atoms by fully stripped ions

    At. Data Nucl. Data Tables

    ([object Object])
  • MančevI. et al.

    State-selective and total cross sections for electron capture from the K-shell of multi-electron atoms by fully stripped projectiles

    At. Data Nucl. Data Tables

    ([object Object])
  • BraggW.H. et al.

    On the ionization curves of radium

    Phil. Mag.

    ([object Object])
  • BraggW.H. et al.

    On the α–particles of radium and their loss of range in their passing through various atoms and molecules

    Phil. Mag.

    ([object Object])
  • WilsonR.R

    Radiological use of fast protons

    Radiology

    ([object Object])
  • WilsonR.R.

    Range, straggling and multiple scattering of fast protons

    Phys. Rev.

    ([object Object])
  • PTCOG, Particle Therapy Co-Operative Group
  • Durante, M. (ed.), Focus on heavy ions in biophysics and medical physics (Topical Issue), New. J. Phys., 10 2008,...
  • SchardtD. et al.

    Heavy-ion tumor therapy: physical and radiobiological benefits

    Rev. Mod. Phys.

    ([object Object])
  • TsujiiH

    Overview of carbon-ion radiotherapy

    J. Phys. Conf. Ser.

    ([object Object])
  • Belkić

    Review of theories on ionization in fast ion-atom collisions with prospects for applications to hadron therapy

    J. Math. Chem.

    ([object Object])
  • Belkić

    The role of high-energy ion-atom/molecule collisions in radiotherapy

    J. Phys. Conf. Ser.

    ([object Object])
  • Belkić

    Theory of heavy ion collision physics

  • BergerM.J.

    Proton Monte Carlo transport program PTRAN

    ([object Object])
  • BergerM.J.

    Penetration of proton beams through water: I. Depth-dose distributions, spectra and LET distributions

    ([object Object])
  • BergerM.J.

    Penetration of proton beams through water: II. Three-dimensional absorbed dose distributions

    ([object Object])
  • CarlssonÅ.K. et al.

    A Monte Carlo and analytical calculation of proton pencil beams for computerized treatment plan optimization

    Phys. Med. Biol.

    ([object Object])
  • MedinJ. et al.

    PETRA, a Monte Carlo code for the simulation of proton and electron transport in water

    ([object Object])
  • AndreoP. et al.

    Absorbed dose determination in external beam radiotherapy: an International code of practice for dosimetry based on standards of absorbed dose to water

  • UeharaS. et al.

    Development of a Monte Carlo track structure code for low-energy protons in water

    Int. J. Radiat. Biol.

    ([object Object])
  • UeharaS. et al.

    Monte Carlo track structure code for low-energy alpha-particles in water

    J. Phys. Chem. B

    ([object Object])
  • LiamsuwanT. et al.

    A Monte Carlo track structure simulation code for the full-slowing-down carbon projectiles of energies 1 keVu 1-10 MeVu 1 in water

    Phys. Med. Biol.

    ([object Object])
  • IlićR.D.

    SRNA-2KG: Proton transport simulation by Monte Carlo techniques

    ([object Object])
  • IlićR.D. et al.

    SRNA-Monte Carlo codes for proton transport simulation in combined and voxelized geometries

    Nucl. Techn. Radiat. Protect.

    ([object Object])
  • IlićR.D. et al.

    The Monte Carlo SRNA code as the engine in the ISTAR proton dose planning software for the TESLA accelerator installation

    Nucl. Techn. Radiat. Protect.

    ([object Object])
  • IlićR.D. et al.

    The Monte Carlo SRNA-VOX code for 3D proton dose distribution in voxelized geometry using CT data

    Phys. Med. Biol.

    ([object Object])
  • Beebe-WangJ. et al.

    PET imaging of dose distributions in proton-beam cancer therapy

    Nucl. Techn. Radiat. Protect.

    ([object Object])
  • Beebe-WangJ. et al.

    Verifying radiation treatments in proton therapy via PET imaging of the induced positron emitters

  • FriedlandW. et al.

    Simulation of DNA damage after proton irradiation

    Radiat. Res.

    ([object Object])
  • AlloniD. et al.

    Track structure, radiation quality and initial radiobiological events: considerations based on the PATRAC code experience

    Int. J. Radiat. Biol.

    ([object Object])
  • GudowskaI. et al.

    Ion beam transport in tissue-like media using the Monte Carlo code SHIELD-HIT

    Phys. Med. Biol.

    ([object Object])
  • GeithnerO. et al.

    Calculation of stopping power ratios for carbon ion dosimetry

    Phys. Med. Biol.

    ([object Object])
  • HenknerK. et al.

    Monte Carlo simulations on the water-to-air stopping power ratio for carbon ion dosimetry

    Med. Phys.

    ([object Object])
  • HenknerK. et al.

    Test of the nuclear interaction model in SHIELD-HIT and comparison to energy distributions from GEANT4

    Phys. Med. Biol.

    ([object Object])
  • Garcia-MolinaR. et al.

    Proton beam interaction of liquid water: A combined molecular dynamics and Monte Carlo simulation study of the Bragg peak profile

  • Cited by (6)

    • The BCIS-4B method for state-selective and state-summed total cross sections: Proton–helium​ charge exchange at 10–4000 keV

      2023, Atomic Data and Nuclear Data Tables
      Citation Excerpt :

      These simulations include the critically important estimates of energy losses of ions during their passage through matter. Such applications are routinely encountered in e.g. astrophysics [1,2], thermonuclear fusion research [3–8], plasma physics [3,4,6,7] and medical physics [9–18]. On the theoretical side, the literature is abundant with a number of first- and second-order methods that have frequently been applied to heavy particle collisions involving single electron capture by bare nuclei from atomic and molecular targets.

    • Single-electron transfer from helium atoms to energetic multiply-charged nuclei

      2022, Atomic Data and Nuclear Data Tables
      Citation Excerpt :

      Therefore, accurate dosimetry for dose planning systems in hospitals ought to include the contributions from energy losses due to charge-exchange. The pertinent stopping powers rely upon electron-capture cross sections and this is precisely where the database from the present study can find its crucially important applications [15–20]. Note that in Refs. [30,31], the specific acronym CDW-4B was not mentioned.

    View full text