Probabilistic proofs of non-probabilistic theorems: Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
en>Tsirel
 
Line 1: Line 1:
By investing in a premium Word - Press theme, you're investing in the future of your website. This means you can setup your mailing list and auto-responder on your wordpress site and then you can add your subscription form to any other blog, splash page, capture page or any other site you like. Your parishioners and certainly interested audience can come in to you for further information from the group and sometimes even approaching happenings and systems with the church. 2- Ask for the designs and graphics that will be provided along with the Word - Press theme. You can customize the appearance with PSD to Word - Press conversion ''. <br><br>Luckily, for Word - Press users, WP Touch plugin transforms your site into an IPhone style theme. After all, Word - Press is free, many of the enhancements for Word - Press like themes and plugins are also free, and there is plenty of free information online about how to use Word - Press. You are able to set them within your theme options and so they aid the search engine to get a suitable title and description for the pages that get indexed by Google.  Should you loved this post and you would love to receive more info with regards to [http://ad4.fr/wordpress_backup_5166915 wordpress backup plugin] i implore you to visit the internet site. This is identical to doing a research as in depth above, nevertheless you can see various statistical details like the number of downloads and when the template was not long ago updated. Once you've installed the program you can quickly begin by adding content and editing it with features such as bullet pointing, text alignment and effects without having to do all the coding yourself. <br><br>The least difficult and very best way to do this is by acquiring a Word - Press site. Word - Press has different exciting features including a plug-in architecture with a templating system. all the necessary planning and steps of conversion is carried out in this phase, such as splitting, slicing, CSS code, adding images, header footer etc. Every single Theme might be unique, providing several alternatives for webpage owners to reap the benefits of in an effort to instantaneously adjust their web page appear. Converting HTML to Word - Press theme for your website can allow you to enjoy the varied Word - Press features that aid in consistent growth your online business. <br><br>The disadvantage is it requires a considerable amount of time to set every thing up. Quttera - Quttera describes itself as a 'Saa - S [Software as a Service] web-malware monitoring and alerting solution for websites of any size and complexity. A higher percentage of women are marrying at older ages,many are delaying childbearing until their careers are established, the divorce rate is high and many couples remarry and desire their own children. If you choose a blog then people will be able to post articles on your site and people will be able to make comments on your posts (unless you turn comments off). Look for experience: When you are searching for a Word - Press developer you should always look at their experience level. <br><br>You will know which of your Word - Press blog posts are attracting more unique visitors which in turn will help you develop better products and services for your customers. Here's a list of some exciting Word - Press features that have created waves in the web development industry:. Must being, it's beneficial because I don't know about you, but loading an old website on a mobile, having to scroll down, up, and sideways' I find links being clicked and bounced around like I'm on a freaking trampoline. Web developers and newbies alike will have the ability to extend your web site and fit other incredible functions with out having to spend more. Press CTRL and the numbers one to six to choose your option.
The '''threshold displacement energy''' <math>T_d</math> is the minimum [[kinetic energy]]
that an atom in a [[solid]] needs to be permanently
displaced from its lattice site to a
[[crystallographic defect|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 [[Crystallography|crystallographic]]
direction. Then one should distinguish between the
minimum  <math>T_{d,min}</math> and average  <math>T_{d,ave}</math> 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 [[Electronvolt|eV]].
<ref name="And79">H. H. Andersen, The Depth Resolution of Sputter Profiling,
Appl. Phys. 18, 131 (1979)</ref>
<ref name="Nastasi">M. Nastasi, J. Mayer, and J. Hirvonen, Ion-Solid Interactions - Fundamentals and Applications, Cambridge University Press, Cambridge, Great Britain, 1996</ref>
<ref name="Luc75">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</ref>
<ref name="Ave98">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.</ref><ref name="Smith">R. Smith (ed.), Atomic & ion collisions in solids and at surfaces: theory, simulation and applications, Cambridge University Press, Cambridge, UK, 1997</ref>
 
== Theory and simulation ==
 
The threshold displacement energy is a materials property relevant during high-energy [[particle radiation]] of materials.
The maximum energy <math>T_{max}</math> that an irradiating particle can transfer in a
[[binary collision approximation|binary collision]]
to an atom in a material is given by (including [[Special relativity|relativistic]] effects)
 
<math>
T_{max} = {2 M E (E+2 m c^2) \over (m+M)^2 c^2+2 M E}
</math>
 
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 <math>m c^2</math> of the irradiating particle, the equation reduces to
 
<math>
T_{max} = E {4 M m  \over (m+M)^2 }
</math>
 
In order for a permanent defect to be produced from initially perfect [[crystal]] lattice, the kinetic energy that it receives <math>T_{max}</math> 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.<ref name="And79" /> 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
<math>
T_d(\theta,\phi) = T_d([hkl])
</math>
for all non-equivalent [[Miller index|crystallographic directions]] [hkl]. Then
<math>
T_{d,min} = \min(T_d(\theta,\phi))
</math>
and
<math>
T_{d,ave} = {\rm ave}(T_d(\theta,\phi))
</math>
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 <math>T^l_d</math>,  
and an upper one where it is certainly formed <math>T^u_d</math>
.<ref name="Mal02">L. Malerba and J. M. Perlado,
Basic mechanisms of atomic displacement production in cubic silicon carbide: A molecular dynamics study, Phys. Rev. B 65, 045202 (2002)</ref>
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.
.<ref name="Nor05c">K. Nordlund, J. Wallenius, and L. Malerba, Molecular dynamics simulations of threshold energies in Fe, Nucl. Instr. Meth. Phys. Res. B 246, 322 (2005)</ref>
 
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
<ref name="Mal02" />
<ref name="Nor05c" />
<ref name="Gib60">J. B. Gibson, A. N. Goland, M. Milgram, and G. H. Vineyard,
Dynamics of Radiation Damage, Phys. Rev 120, 1229 (1960)</ref>
<ref name="Erg64">C. Erginsoy, G. H. Vineyard, and A. Englert, Dynamics of Radiation Damage in a Body-Centered Cubic Lattice, Phys. Rev. 133, A595 (1964)</ref><ref name="Cat94">M.-J. Caturla, T. Diaz de la Rubia, and G. H. Gilmer, Point defect production, geometry and stability in Silicon: a molecular dynamics simulation study, Mat. Res. Soc. Symp. Proc. 316, 141 (1994)</ref>
<ref name="Par01">B. Park, W. J. Weber, and L. R. Corrales, Molecular-dynamics simulation study of threshold displacements and defect formation in zircon, Phys. Rev. B 64, 174108 (2001)</ref>
or quantum mechanical
<ref name="Uhl97">S. Uhlmann, T. Frauenheim, K. Boyd, D.Marton, and J. Rabalais, Radiat. Eff. Defects Solids 141, 185 (1997).</ref>
<ref name="Win98">W. Windl, T. J. Lenosky, J. D. Kress, and A. F. Voter, Nucl. Instr. and Meth. B 141, 61 (1998).</ref>
<ref name="Maz01">M. Mazzarolo, L. Colombo, G. Lulli, and E. Albertazzi, Phys.
Rev. B 63, 195207 (2001).</ref>
<ref name="Hol08">E. Holmström, A. Kuronen, and K. Nordlund, Threshold defect production in silicon determined by density functional theory molecular dynamics simulations, Phys. Rev. B (Rapid Comm.) 78, 045202 (2008).</ref>
[[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 <ref name="Win98" />
 
An example molecular dynamics simulation of a threshold displacement event is available in [http://www.youtube.com/watch?v=ZWu5Qf8y6iQ]. The animation shows how a defect ([[Frenkel defect|Frenkel pair]], i.e. an [[interstitial defect|interstitial]] and [[vacancy defect|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.<ref name="Hol08" />
 
Such simulations have given significant qualitative insights into the threshold displacement energy, but the quantitative results should be viewed with caution.
The classical [[molecular dynamics|interatomic potential]]s 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.<ref name="Nor05c" /><ref name="Hol08" /> 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 [[particle radiation|electron irradiation]]
experiments. Electrons with kinetic energies of the order of hundreds of [[keV]]s or a few [[MeV]]s 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
<math>T^l_{d,min}</math>
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 [[Miller index|crystallographic directions]] one can determine also direction-specific thresholds
<math>T_d^l(\theta,\phi)</math>.<ref name="And79" /><ref name="And79" /><ref name="Luc75" /><ref name="Ave98" /><ref name="Lof58">J. Loferski and P. Rappaport, Phys. Rev. 111, 432 (1958).</ref><ref name="Ban99">F.~Banhart, Rep. Prog. Phys. 62 (1999) 1181</ref>
 
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 [[point defect|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 T<sub>0</sub>, 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 pair]]s 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.<ref name="Ehr91">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</ref>
Similarly, in Si major recombination of damage happens already
around 100 K during ion irradiation and 4 K during electron irradiation
<ref name="Par00">P. Partyka, Y. Zhong, K. Nordlund, R. S. Averback, I. K. Robinson, and
      P. Ehrhart,
      Grazing incidence diffuse x-ray scattering investigation of the
      properties of irradiation-induced point defects in silicon,
      Phys. Rev. B 64, 235207 (2002)</ref>
 
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 T<sub>0</sub>.
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 [[Crystallographic defect|defects]] produced by higher energy irradiation using the Kinchin-Pease or NRT
equations<ref name="NRT">M. J. Norgett, M. T. Robinson, and I. M. Torrens,
      A proposed method of calculating displacement dose rates,
      Nucl. Engr. and Design 33, 50 (1975)</ref><ref name="ASTMFe">ASTM Standard E693-94,
Standard practice for characterising neutron exposure in iron and low
alloy steels in terms of displacements per atom (dpa), 1994</ref>
which says that the number of Frenkel pairs produced <math>N_{FP}</math>
for a [[stopping power (particle radiation)|nuclear deposited energy]] of <math>F_{Dn}</math> is
 
<math>
    N_{FP} = 0.8 {F_{Dn} \over 2 T_{d,ave}}
</math>
 
for any nuclear deposited energy above <math> 2 T_{d,ave}/0.8</math>.
 
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.<ref name="Ave98" />
 
The threshold displacement energy is also often used in
[[binary collision approximation]]
computer codes such as [[Stopping and Range of Ions in Matter|SRIM]]<ref>http://www.srim.org</ref> 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 implantation|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.<ref name="Sil00">
J. Sillanpää, K. Nordlund, and J. Keinonen, Electronic stopping of
Silicon from a 3D Charge Distribution, Phys. Rev. B 62, 3109 (2000)</ref>
Similarly, light ion like He irradiation of a BCC metal like Fe
leads to massive channeling even in a randomly selected
crystal direction.<ref name="Nor09">K. Nordlund, MDRANGE range
calculations of He in Fe (2009), public presentation
at the EFDA MATREMEV meeting, Alicante 19.11.2009</ref>
 
== See also ==
 
* [[Threshold energy]]
* [[Stopping power (particle radiation)]]
* [[Crystallographic defect]]
 
== References ==
 
<references/>
 
[[Category:Condensed matter physics]]
[[Category:Radiation effects]]

Revision as of 03:17, 21 December 2013

The threshold displacement energy is the minimum kinetic energy that an atom in a solid needs to be permanently displaced from its lattice site 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 and average 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]

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 [1]. 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][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

References

  1. 1.0 1.1 1.2 1.3 H. H. Andersen, The Depth Resolution of Sputter Profiling, Appl. Phys. 18, 131 (1979)
  2. M. Nastasi, J. Mayer, and J. Hirvonen, Ion-Solid Interactions - Fundamentals and Applications, Cambridge University Press, Cambridge, Great Britain, 1996
  3. 3.0 3.1 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. 4.0 4.1 4.2 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. 6.0 6.1 L. Malerba and J. M. Perlado, Basic mechanisms of atomic displacement production in cubic silicon carbide: A molecular dynamics study, Phys. Rev. B 65, 045202 (2002)
  7. 7.0 7.1 7.2 K. Nordlund, J. Wallenius, and L. Malerba, Molecular dynamics simulations of threshold energies in Fe, Nucl. Instr. Meth. Phys. Res. B 246, 322 (2005)
  8. J. B. Gibson, A. N. Goland, M. Milgram, and G. H. Vineyard, Dynamics of Radiation Damage, Phys. Rev 120, 1229 (1960)
  9. C. Erginsoy, G. H. Vineyard, and A. Englert, Dynamics of Radiation Damage in a Body-Centered Cubic Lattice, Phys. Rev. 133, A595 (1964)
  10. M.-J. Caturla, T. Diaz de la Rubia, and G. H. Gilmer, Point defect production, geometry and stability in Silicon: a molecular dynamics simulation study, Mat. Res. Soc. Symp. Proc. 316, 141 (1994)
  11. B. Park, W. J. Weber, and L. R. Corrales, Molecular-dynamics simulation study of threshold displacements and defect formation in zircon, Phys. Rev. B 64, 174108 (2001)
  12. S. Uhlmann, T. Frauenheim, K. Boyd, D.Marton, and J. Rabalais, Radiat. Eff. Defects Solids 141, 185 (1997).
  13. 13.0 13.1 W. Windl, T. J. Lenosky, J. D. Kress, and A. F. Voter, Nucl. Instr. and Meth. B 141, 61 (1998).
  14. M. Mazzarolo, L. Colombo, G. Lulli, and E. Albertazzi, Phys. Rev. B 63, 195207 (2001).
  15. 15.0 15.1 15.2 E. Holmström, A. Kuronen, and K. Nordlund, Threshold defect production in silicon determined by density functional theory molecular dynamics simulations, Phys. Rev. B (Rapid Comm.) 78, 045202 (2008).
  16. J. Loferski and P. Rappaport, Phys. Rev. 111, 432 (1958).
  17. F.~Banhart, Rep. Prog. Phys. 62 (1999) 1181
  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. P. Partyka, Y. Zhong, K. Nordlund, R. S. Averback, I. K. Robinson, and P. Ehrhart, Grazing incidence diffuse x-ray scattering investigation of the properties of irradiation-induced point defects in silicon, Phys. Rev. B 64, 235207 (2002)
  20. M. J. Norgett, M. T. Robinson, and I. M. Torrens, A proposed method of calculating displacement dose rates, Nucl. Engr. and Design 33, 50 (1975)
  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. http://www.srim.org
  23. J. Sillanpää, K. Nordlund, and J. Keinonen, Electronic stopping of Silicon from a 3D Charge Distribution, Phys. Rev. B 62, 3109 (2000)
  24. K. Nordlund, MDRANGE range calculations of He in Fe (2009), public presentation at the EFDA MATREMEV meeting, Alicante 19.11.2009