Threshold displacement energy

Last updated

In materials science, the threshold displacement energy (Td) is the minimum kinetic energy that an atom in a solid needs to be permanently displaced from its site in the lattice to a defect position. It is also known as "displacement threshold energy" or just "displacement energy". In a crystal, a separate threshold displacement energy exists for each crystallographic direction. Then one should distinguish between the minimum (Td,min) and average (Td,ave) over all lattice directions' threshold displacement energies. In amorphous solids, it may be possible to define an effective displacement energy to describe some other average quantity of interest. Threshold displacement energies in typical solids are of the order of 10-50 eV. [1] [2] [3] [4] [5]

Contents

Theory and simulation

The threshold displacement energy is a materials property relevant during high-energy particle radiation of materials. The maximum energy that an irradiating particle can transfer in a binary collision to an atom in a material is given by (including relativistic effects)

where E is the kinetic energy and m the mass of the incoming irradiating particle and M the mass of the material atom. c is the velocity of light. If the kinetic energy E is much smaller than the mass of the irradiating particle, the equation reduces to

In order for a permanent defect to be produced from initially perfect crystal lattice, the kinetic energy that it receives must be larger than the formation energy of a Frenkel pair. However, while the Frenkel pair formation energies in crystals are typically around 5–10 eV, the average threshold displacement energies are much higher, 20–50 eV. [1] The reason for this apparent discrepancy is that the defect formation is a complex multi-body collision process (a small collision cascade) where the atom that receives a recoil energy can also bounce back, or kick another atom back to its lattice site. Hence, even the minimum threshold displacement energy is usually clearly higher than the Frenkel pair formation energy.

Each crystal direction has in principle its own threshold displacement energy, so for a full description one should know the full threshold displacement surface for all non-equivalent crystallographic directions [hkl]. Then and where the minimum and average is with respect to all angles in three dimensions.

An additional complication is that the threshold displacement energy for a given direction is not necessarily a step function, but there can be an intermediate energy region where a defect may or may not be formed depending on the random atom displacements. The one can define a lower threshold where a defect may be formed , and an upper one where it is certainly formed . [6] The difference between these two may be surprisingly large, and whether or not this effect is taken into account may have a large effect on the average threshold displacement energy. . [7]

It is not possible to write down a single analytical equation that would relate e.g. elastic material properties or defect formation energies to the threshold displacement energy. Hence theoretical study of the threshold displacement energy is conventionally carried out using either classical [6] [7] [8] [9] [10] [11] or quantum mechanical [12] [13] [14] [15] molecular dynamics computer simulations. Although an analytical description of the displacement is not possible, the "sudden approximation" gives fairly good approximations of the threshold displacement energies at least in covalent materials and low-index crystal directions [13]

An example molecular dynamics simulation of a threshold displacement event is available in 100_20eV.avi. The animation shows how a defect (Frenkel pair, i.e. an interstitial and vacancy) is formed in silicon when a lattice atom is given a recoil energy of 20 eV in the 100 direction. The data for the animation was obtained from density functional theory molecular dynamics computer simulations. [15]

Such simulations have given significant qualitative insights into the threshold displacement energy, but the quantitative results should be viewed with caution. The classical interatomic potentials are usually fit only to equilibrium properties, and hence their predictive capability may be limited. Even in the most studied materials such as Si and Fe, there are variations of more than a factor of two in the predicted threshold displacement energies. [7] [15] The quantum mechanical simulations based on density functional theory (DFT) are likely to be much more accurate, but very few comparative studies of different DFT methods on this issue have yet been carried out to assess their quantitative reliability.

Experimental studies

The threshold displacement energies have been studied extensively with electron irradiation experiments. Electrons with kinetic energies of the order of hundreds of keVs or a few MeVs can to a very good approximation be considered to collide with a single lattice atom at a time. Since the initial energy for electrons coming from a particle accelerator is accurately known, one can thus at least in principle determine the lower minimum threshold displacement energy by irradiating a crystal with electrons of increasing energy until defect formation is observed. Using the equations given above one can then translate the electron energy E into the threshold energy T. If the irradiation is carried out on a single crystal in a known crystallographic directions one can determine also direction-specific thresholds . [1] [3] [4] [16] [17]

There are several complications in interpreting the experimental results, however. To name a few, in thick samples the electron beam will spread, and hence the measurement on single crystals does not probe only a single well-defined crystal direction. Impurities may cause the threshold to appear lower than they would be in pure materials.

Temperature dependence

Particular care has to be taken when interpreting threshold displacement energies at temperatures where defects are mobile and can recombine. At such temperatures, one should consider two distinct processes: the creation of the defect by the high-energy ion (stage A), and subsequent thermal recombination effects (stage B).

The initial stage A. of defect creation, until all excess kinetic energy has dissipated in the lattice and it is back to its initial temperature T0, takes < 5 ps. This is the fundamental ("primary damage") threshold displacement energy, and also the one usually simulated by molecular dynamics computer simulations. After this (stage B), however, close Frenkel pairs may be recombined by thermal processes. Since low-energy recoils just above the threshold only produce close Frenkel pairs, recombination is quite likely.

Hence on experimental time scales and temperatures above the first (stage I) recombination temperature, what one sees is the combined effect of stage A and B. Hence the net effect often is that the threshold energy appears to increase with increasing temperature, since the Frenkel pairs produced by the lowest-energy recoils above threshold all recombine, and only defects produced by higher-energy recoils remain. Since thermal recombination is time-dependent, any stage B kind of recombination also implies that the results may have a dependence on the ion irradiation flux.

In a wide range of materials, defect recombination occurs already below room temperature. E.g. in metals the initial ("stage I") close Frenkel pair recombination and interstitial migration starts to happen already around 10-20 K. [18] Similarly, in Si major recombination of damage happens already around 100 K during ion irradiation and 4 K during electron irradiation [19]

Even the stage A threshold displacement energy can be expected to have a temperature dependence, due to effects such as thermal expansion, temperature dependence of the elastic constants and increased probability of recombination before the lattice has cooled down back to the ambient temperature T0. These effects, are, however, likely to be much weaker than the stage B thermal recombination effects.

Relation to higher-energy damage production

The threshold displacement energy is often used to estimate the total amount of defects produced by higher energy irradiation using the Kinchin-Pease or NRT equations [20] [21] which says that the number of Frenkel pairs produced for a nuclear deposited energy of is

for any nuclear deposited energy above .

However, this equation should be used with great caution for several reasons. For instance, it does not account for any thermally activated recombination of damage, nor the well known fact that in metals the damage production is for high energies only something like 20% of the Kinchin-Pease prediction. [4]

The threshold displacement energy is also often used in binary collision approximation computer codes such as SRIM [22] to estimate damage. However, the same caveats as for the Kinchin-Pease equation also apply for these codes (unless they are extended with a damage recombination model).

Moreover, neither the Kinchin-Pease equation nor SRIM take in any way account of ion channeling, which may in crystalline or polycrystalline materials reduce the nuclear deposited energy and thus the damage production dramatically for some ion-target combinations. For instance, keV ion implantation into the Si 110 crystal direction leads to massive channeling and thus reductions in stopping power. [23] Similarly, light ion like He irradiation of a BCC metal like Fe leads to massive channeling even in a randomly selected crystal direction. [24]

See also

Related Research Articles

<span class="mw-page-title-main">Crystallographic defect</span> Disruption of the periodicity of a crystal lattice

A crystallographic defect is an interruption of the regular patterns of arrangement of atoms or molecules in crystalline solids. The positions and orientations of particles, which are repeating at fixed distances determined by the unit cell parameters in crystals, exhibit a periodic crystal structure, but this is usually imperfect. Several types of defects are often characterized: point defects, line defects, planar defects, bulk defects. Topological homotopy establishes a mathematical method of characterization.

<span class="mw-page-title-main">Exciton</span> Quasiparticle which is a bound state of an electron and an electron hole

An electron and an electron hole that are attracted to each other by the Coulomb force can form a bound state called an exciton. It is an electrically neutral quasiparticle that exists mainly in condensed matter, including insulators, semiconductors, some metals, but also in certain atoms, molecules and liquids. The exciton is regarded as an elementary excitation that can transport energy without transporting net electric charge.

<span class="mw-page-title-main">Wannier function</span>

The Wannier functions are a complete set of orthogonal functions used in solid-state physics. They were introduced by Gregory Wannier in 1937. Wannier functions are the localized molecular orbitals of crystalline systems.

<span class="mw-page-title-main">PLATO (computational chemistry)</span>

PLATO is a suite of programs for electronic structure calculations. It receives its name from the choice of basis set used to expand the electronic wavefunctions.

<span class="mw-page-title-main">Crystallographic defects in diamond</span>

Imperfections in the crystal lattice of diamond are common. Such defects may be the result of lattice irregularities or extrinsic substitutional or interstitial impurities, introduced during or after the diamond growth. The defects affect the material properties of diamond and determine to which type a diamond is assigned; the most dramatic effects are on the diamond color and electrical conductivity, as explained by the electronic band structure.

<span class="mw-page-title-main">Mott insulator</span> Materials classically predicted to be conductors, that are actually insulators

Mott insulators are a class of materials that are expected to conduct electricity according to conventional band theories, but turn out to be insulators. These insulators fail to be correctly described by band theories of solids due to their strong electron–electron interactions, which are not considered in conventional band theory. A Mott transition is a transition from a metal to an insulator, driven by the strong interactions between electrons. One of the simplest models that can capture Mott transition is the Hubbard model.

<span class="mw-page-title-main">Spin density wave</span>

Spin-density wave (SDW) and charge-density wave (CDW) are names for two similar low-energy ordered states of solids. Both these states occur at low temperature in anisotropic, low-dimensional materials or in metals that have high densities of states at the Fermi level . Other low-temperature ground states that occur in such materials are superconductivity, ferromagnetism and antiferromagnetism. The transition to the ordered states is driven by the condensation energy which is approximately where is the magnitude of the energy gap opened by the transition.

<span class="mw-page-title-main">Stopping power (particle radiation)</span> Retarding force acting on charged particles due to interactions with matter

In nuclear and materials physics, stopping power is the retarding force acting on charged particles, typically alpha and beta particles, due to interaction with matter, resulting in loss of particle kinetic energy. Stopping power is also interpreted as the rate at which a material absorbs the kinetic energy of a charged particle. Its application is important in a wide range of thermodynamic areas such as radiation protection, ion implantation and nuclear medicine.

<span class="mw-page-title-main">Interstitial defect</span> Crystallographic defect; atoms located in the gaps between atoms in the lattice

In materials science, an interstitial defect is a type of point crystallographic defect where an atom of the same or of a different type, occupies an interstitial site in the crystal structure. When the atom is of the same type as those already present they are known as a self-interstitial defect. Alternatively, small atoms in some crystals may occupy interstitial sites, such as hydrogen in palladium. Interstitials can be produced by bombarding a crystal with elementary particles having energy above the displacement threshold for that crystal, but they may also exist in small concentrations in thermodynamic equilibrium. The presence of interstitial defects can modify the physical and chemical properties of a material.

Quantum dimer models were introduced to model the physics of resonating valence bond (RVB) states in lattice spin systems. The only degrees of freedom retained from the motivating spin systems are the valence bonds, represented as dimers which live on the lattice bonds. In typical dimer models, the dimers do not overlap.

The hexatic phase is a state of matter that is between the solid and the isotropic liquid phases in two dimensional systems of particles. It is characterized by two order parameters: a short-range positional and a quasi-long-range orientational (sixfold) order. More generally, a hexatic is any phase that contains sixfold orientational order, in analogy with the nematic phase.

<span class="mw-page-title-main">Collision cascade</span> Series of collisions between nearby atoms, initiated by a single energetic atom

In condensed-matter physics, a collision cascade is a set of nearby adjacent energetic collisions of atoms induced by an energetic particle in a solid or liquid.

Stopping and Range of Ions in Matter (SRIM) is a group of computer programs which calculate interactions between ions and matter; the core of SRIM is a program called Transport of Ions in Matter (TRIM). SRIM is popular in the ion implantation research and technology community, and also used widely in other branches of radiation material science.

Swift heavy ions are the components of a type of particle beam with high enough energy that electronic stopping dominates over nuclear stopping. They are accelerated in particle accelerators to very high energies, typically in the MeV or GeV range and have sufficient energy and mass to penetrate solids on a straight line. In many solids swift heavy ions release sufficient energy to induce permanently modified cylindrical zones, so-called ion tracks. If the irradiation is carried out in an initially crystalline material, ion tracks consist of an amorphous cylinder. Ion tracks can be produced in many amorphizing materials, but not in pure metals, where the high electronic heat conductivity dissipates away the electronic heating before the ion track has time to form.

<span class="mw-page-title-main">Binary collision approximation</span> Heuristic used in simulations of ions passing through solids

In condensed-matter physics, the binary collision approximation (BCA) is a heuristic used to more efficiently simulate the penetration depth and defect production by energetic ions in solids. In the method, the ion is approximated to travel through a material by experiencing a sequence of independent binary collisions with sample atoms (nuclei). Between the collisions, the ion is assumed to travel in a straight path, experiencing electronic stopping power, but losing no energy in collisions with nuclei.

The toric code is a topological quantum error correcting code, and an example of a stabilizer code, defined on a two-dimensional spin lattice. It is the simplest and most well studied of the quantum double models. It is also the simplest example of topological order—Z2 topological order (first studied in the context of Z2 spin liquid in 1991). The toric code can also be considered to be a Z2 lattice gauge theory in a particular limit. It was introduced by Alexei Kitaev.

Radiation materials science is a subfield of materials science which studies the interaction of radiation with matter: a broad subject covering many forms of irradiation and of matter.

Symmetry-protected topological (SPT) order is a kind of order in zero-temperature quantum-mechanical states of matter that have a symmetry and a finite energy gap.

<span class="mw-page-title-main">Interatomic potential</span> Functions for calculating potential energy

Interatomic potentials are mathematical functions to calculate the potential energy of a system of atoms with given positions in space. Interatomic potentials are widely used as the physical basis of molecular mechanics and molecular dynamics simulations in computational chemistry, computational physics and computational materials science to explain and predict materials properties. Examples of quantitative properties and qualitative phenomena that are explored with interatomic potentials include lattice parameters, surface energies, interfacial energies, adsorption, cohesion, thermal expansion, and elastic and plastic material behavior, as well as chemical reactions.

Multiscale Green's function (MSGF) is a generalized and extended version of the classical Green's function (GF) technique for solving mathematical equations. The main application of the MSGF technique is in modeling of nanomaterials. These materials are very small – of the size of few nanometers. Mathematical modeling of nanomaterials requires special techniques and is now recognized to be an independent branch of science. A mathematical model is needed to calculate the displacements of atoms in a crystal in response to an applied static or time dependent force in order to study the mechanical and physical properties of nanomaterials. One specific requirement of a model for nanomaterials is that the model needs to be multiscale and provide seamless linking of different length scales.

References

  1. 1 2 3 Andersen, H. H. (1979). "The depth resolution of sputter profiling". Applied Physics. Springer Science and Business Media LLC. 18 (2): 131–140. Bibcode:1979ApPhy..18..131A. doi:10.1007/bf00934407. ISSN   0340-3793. S2CID   54858884.
  2. M. Nastasi, J. Mayer, and J. Hirvonen, Ion-Solid Interactions - Fundamentals and Applications, Cambridge University Press, Cambridge, Great Britain, 1996
  3. 1 2 P. Lucasson, The production of Frenkel defects in metals, in Fundamental Aspects of Radiation Damage in Metals, edited by M. T. Robinson and F. N. Young Jr., pages 42--65, Springfield, 1975, ORNL
  4. 1 2 3 R. S. Averback and T. Diaz de la Rubia, Displacement damage in irradiated metals and semiconductors, in Solid State Physics, edited by H. Ehrenfest and F. Spaepen, volume 51, pages 281--402, Academic Press, New York, 1998.
  5. R. Smith (ed.), Atomic & ion collisions in solids and at surfaces: theory, simulation and applications, Cambridge University Press, Cambridge, UK, 1997
  6. 1 2 Malerba, L.; Perlado, J. M. (2 January 2002). "Basic mechanisms of atomic displacement production in cubic silicon carbide: A molecular dynamics study". Physical Review B. American Physical Society (APS). 65 (4): 045202. Bibcode:2002PhRvB..65d5202M. doi:10.1103/physrevb.65.045202. ISSN   0163-1829.
  7. 1 2 3 Nordlund, K.; Wallenius, J.; Malerba, L. (2006). "Molecular dynamics simulations of threshold displacement energies in Fe". Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms. Elsevier BV. 246 (2): 322–332. Bibcode:2006NIMPB.246..322N. doi:10.1016/j.nimb.2006.01.003. ISSN   0168-583X.
  8. Gibson, J. B.; Goland, A. N.; Milgram, M.; Vineyard, G. H. (15 November 1960). "Dynamics of Radiation Damage". Physical Review. American Physical Society (APS). 120 (4): 1229–1253. Bibcode:1960PhRv..120.1229G. doi:10.1103/physrev.120.1229. ISSN   0031-899X.
  9. Erginsoy, C.; Vineyard, G. H.; Englert, A. (20 January 1964). "Dynamics of Radiation Damage in a Body-Centered Cubic Lattice". Physical Review. American Physical Society (APS). 133 (2A): A595–A606. Bibcode:1964PhRv..133..595E. doi:10.1103/physrev.133.a595. ISSN   0031-899X.
  10. Caturla, M.-J.; De La Rubia, T. Diaz; Gilmer, G.H. (1993). "Point defect Production, Geometry and Stability in Silicon: a Molecular Dynamics Simulation Study". MRS Proceedings. Cambridge University Press (CUP). 316: 141. doi:10.1557/proc-316-141. ISSN   1946-4274.
  11. Park, Byeongwon; Weber, William J.; Corrales, L. René (16 October 2001). "Molecular-dynamics simulation study of threshold displacements and defect formation in zircon". Physical Review B. American Physical Society (APS). 64 (17): 174108. Bibcode:2001PhRvB..64q4108P. doi:10.1103/physrevb.64.174108. ISSN   0163-1829.
  12. Uhlmann, S.; Frauenheim, Th.; Boyd, K. J.; Marton, D.; Rabalais, J. W. (1997). "Elementary processes during low-energy self-bombardment of Si(100) 2 × 2 a molecular dynamics study". Radiation Effects and Defects in Solids. Informa UK Limited. 141 (1–4): 185–198. Bibcode:1997REDS..141..185U. doi:10.1080/10420159708211569. ISSN   1042-0150.
  13. 1 2 Windl, Wolfgang; Lenosky, Thomas J; Kress, Joel D; Voter, Arthur F (1998). "First-principles investigation of radiation induced defects in Si and SiC". Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms. Elsevier BV. 141 (1–4): 61–65. Bibcode:1998NIMPB.141...61W. doi:10.1016/s0168-583x(98)00082-2. ISSN   0168-583X.
  14. Mazzarolo, Massimiliano; Colombo, Luciano; Lulli, Giorgio; Albertazzi, Eros (26 April 2001). "Low-energy recoils in crystalline silicon: Quantum simulations". Physical Review B. American Physical Society (APS). 63 (19): 195207. Bibcode:2001PhRvB..63s5207M. doi:10.1103/physrevb.63.195207. ISSN   0163-1829.
  15. 1 2 3 Holmström, E.; Kuronen, A.; Nordlund, K. (9 July 2008). "Threshold defect production in silicon determined by density functional theory molecular dynamics simulations" (PDF). Physical Review B. American Physical Society (APS). 78 (4): 045202. Bibcode:2008PhRvB..78d5202H. doi:10.1103/physrevb.78.045202. ISSN   1098-0121.
  16. Loferski, J. J.; Rappaport, P. (15 July 1958). "Radiation Damage in Ge and Si Detected by Carrier Lifetime Changes: Damage Thresholds". Physical Review. American Physical Society (APS). 111 (2): 432–439. Bibcode:1958PhRv..111..432L. doi:10.1103/physrev.111.432. ISSN   0031-899X.
  17. Banhart, Florian (30 July 1999). "Irradiation effects in carbon nanostructures". Reports on Progress in Physics. IOP Publishing. 62 (8): 1181–1221. Bibcode:1999RPPh...62.1181B. doi:10.1088/0034-4885/62/8/201. ISSN   0034-4885. S2CID   250834423.
  18. P. Ehrhart, Properties and interactions of atomic defects in metals and alloys, volume 25 of Landolt-B"ornstein, New Series III, chapter 2, page 88, Springer, Berlin, 1991
  19. Partyka, P.; Zhong, Y.; Nordlund, K.; Averback, R. S.; Robinson, I. M.; Ehrhart, P. (27 November 2001). "Grazing incidence diffuse x-ray scattering investigation of the properties of irradiation-induced point defects in silicon". Physical Review B. American Physical Society (APS). 64 (23): 235207. Bibcode:2001PhRvB..64w5207P. doi:10.1103/physrevb.64.235207. ISSN   0163-1829. S2CID   16857480.
  20. Norgett, M.J.; Robinson, M.T.; Torrens, I.M. (1975). "A proposed method of calculating displacement dose rates". Nuclear Engineering and Design. Elsevier BV. 33 (1): 50–54. doi:10.1016/0029-5493(75)90035-7. ISSN   0029-5493.
  21. ASTM Standard E693-94, Standard practice for characterising neutron exposure in iron and low alloy steels in terms of displacements per atom (dpa), 1994
  22. "James Ziegler - SRIM & TRIM". www.srim.org.
  23. Sillanpää, J.; Nordlund, K.; Keinonen, J. (1 July 2000). "Electronic stopping of Si from a three-dimensional charge distribution". Physical Review B. American Physical Society (APS). 62 (5): 3109–3116. Bibcode:2000PhRvB..62.3109S. doi:10.1103/physrevb.62.3109. ISSN   0163-1829.
  24. K. Nordlund, MDRANGE range calculations of He in Fe (2009), public presentation at the EFDA MATREMEV meeting, Alicante 19.11.2009