+ All Categories
Home > Documents > Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of...

Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of...

Date post: 20-Aug-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
21
This is an electronic reprint of the original article. This reprint may differ from the original in pagination and typographic detail. Powered by TCPDF (www.tcpdf.org) This material is protected by copyright and other intellectual property rights, and duplication or sale of all or part of any of the repository collections is not permitted, except that material may be duplicated by you for your research use or educational purposes in electronic or print form. You must obtain permission for any other use. Electronic or print copies may not be offered, whether for sale or otherwise to anyone who is not an authorised user. Laakkonen, J.; Nieminen, Risto Computer simulations of radiation damage in amorphous solids Published in: Physical Review B DOI: 10.1103/PhysRevB.41.3978 Published: 01/01/1990 Document Version Publisher's PDF, also known as Version of record Please cite the original version: Laakkonen, J., & Nieminen, R. (1990). Computer simulations of radiation damage in amorphous solids. Physical Review B, 41(7), 3978-3997. https://doi.org/10.1103/PhysRevB.41.3978
Transcript
Page 1: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

This is an electronic reprint of the original article.This reprint may differ from the original in pagination and typographic detail.

Powered by TCPDF (www.tcpdf.org)

This material is protected by copyright and other intellectual property rights, and duplication or sale of all or part of any of the repository collections is not permitted, except that material may be duplicated by you for your research use or educational purposes in electronic or print form. You must obtain permission for any other use. Electronic or print copies may not be offered, whether for sale or otherwise to anyone who is not an authorised user.

Laakkonen, J.; Nieminen, RistoComputer simulations of radiation damage in amorphous solids

Published in:Physical Review B

DOI:10.1103/PhysRevB.41.3978

Published: 01/01/1990

Document VersionPublisher's PDF, also known as Version of record

Please cite the original version:Laakkonen, J., & Nieminen, R. (1990). Computer simulations of radiation damage in amorphous solids. PhysicalReview B, 41(7), 3978-3997. https://doi.org/10.1103/PhysRevB.41.3978

Page 2: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

PHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990

Computer simulations of radiation damage in amorphous solids

J. Laakkonen' and R. M. NieminenLaboratory of Physics, Helsinki Uniuersity of Technology, Otakaari l, SF 02l-50Espoo, Finland

(Received 24 March 1989)

Microscopic radiation damage in a Lennard-Jones amorphous solid is investigated by computersimulations of collision cascades. Molecular-dynamics simulations with various primary knock-onatom (PKA) energies and directions are carried out. Energy outflow from the computational box isaccounted for but electronic losses are neglected. The simulations show the PKA energy to spreadrapidly among the nearby atoms, and the atomic trajectories disclose such features as replacementcollision sequences and focused chains. Vacancies are created in the central region of the cascadeand are surrounded by a cloud of interstitials. The defects mainly disappear independently of eachother, the vacancies faster than the interstitials. Recombination plays a minor role. At the end of asimulation all the defects created have vanished and little change in the sample volume or structureis observed. The threshold energy for a permanent displacement is found for various PKA direc-tions. The validity of the modified Kinchin-Pease model for an amorphous solid is discussed.

I. INTRODUCTION

When a fast moving particle traverses a solid, it strikesthe host atoms, transferring large amounts of kinetic en-

ergy. The energy received by the primary knock-onatoms (PKA's) usually far exceeds the bond energy, andthus a PKA initiates a sequence of successive collisions, acollision cascade. The outcome of a collision cascade is alarge amount of point defects, i.e., vacancies and intersti-tials. A vast majority of the defects vanish during orsoon after the cascade is over, but some may survive, in-creasing the defect content of the solid over thethermodynamic-equilibrium concentration. Under suit-able circumstances point defects can subsequently ag-glomerate, creating larger defect formations, and thus acomplex defect structure may follow, giving rise tochanges in many physical properties.

Formation of a collision cascade is a complex many-particle event and no comprehensive, analytic frameworkincluding, for instance, the effects of the structure of thesolid or the forces between the atoms does exist. The fewattempts made treat the cascade, e.g. , as a suddenly heat-ed region (a thermal spike) or a volume, where a minia-ture "explosion" has taken place at the position of thePKA (a displacement spike). The advance of the cascadeis then followed by assuming the solid to be a homogene-ous medium and by applying the classical laws of heatconduction or shock waves. Obviously, these kind ofmodels cannot disclose the real structure or the many de-tails of the cascade. If not the form of the cascade itselfbut its outcome (the number of defects) instead is con-sidered, then the statistical models first introduced byKinchin and Pease' have been applied to some extent.One computational method capable of yielding a wealthof information about the collision cascade is direct com-puter simulation. The applicability of this method is nowrapidly improving along with the progress in computercapacity and the advances made in the fields of simula-tion algorithms and potential functions for interactions

between particles.For collision cascades there are two possible simula-

tion methods: in the case of low and moderate PKAenergies the full molecular-dynamics (MD) simulationis preferable, whereas for high PKA energies[Et,K~ ) (100 1000—)Ed, where Ed is the displacement en-

ergy] the binary-collision approximation (BCA) is ap-propriate. In the MD simulations one considers all theparticles, while in the BCA method a lower limit is set forthe kinetic energy of the atoms that are followed. Bothmethods have been widely applied. The precedingmethods (especially MD) compute the actual trajectoriesof the particles, and they are used to simulate the col-lisional and the cooling phases of the cascade duringwhich most of the defects produced have already disap-peared. This corresponds to real time of approximately10 " s. During later stages the defects left will diffusecausing recombination and clustering, or they may be-come trapped into sinks such as dislocations and grainboundaries. These effects take much longer times,rendering the MD simulation impractical since it wouldrequire vast amounts of computer time. The solution isto apply for the diffusion of defects of a MonteCarlo —type simulation.

The foundations of MD simulations of collision cas-cades were laid by the classic work of Gibson et al. , whostudied radiation damage in crystalline copper. Sincethen basically the same methods have been repeatedly ap-plied. Along with experimental data, these studies havecontributed much to our understanding of the formationradiation damage in crystalline materials. This makes itsurprising that radiation damage in the other type ofsolids, amorphous materials, has hardly been studied atall. These materials have many properties distinctlydifferent from those typical for crystalline structures.Often these exceptional characteristics are also of techno-logical importance. The few studies of radiation damagein amorphous metals show that metal-metalloid glasseshave a good resistance against radiation effects. ' Then,

41 3978 1990 The American Physical Society

Page 3: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

. . . RADIATION DAMAGE IN AMORPHOUS SOLIDS 3979

II. METHOD

The sample for computer simulations here is the sameas studied before in two earlier papers (Refs. 9 and 10, re-ferred to hereafter as I and II, respectively). This is aone-component amorphous noble-gas solid of 1372atoms. The potential experienced by an atom is obtainedas a sum of pairwise interactions of the Lennard-Jonesform:

P(r ) =4e12 ' 6

o'

This same functional form applies to all noble gases, andthus the results describe all noble gases for appropriatevalues of c. and o.. To exploit this scaling, here we use re-duced units (abbreviated in text as r.u. ): the units oflength and energy are o. and c, respectively, the unit fortemperature is e/k~ and the unit for time (Mcr /e)'~, Mbeing the atomic mass. As a specific example here weconsider neon, for which" c./k& =35.3 K, a=2. 8509 A,and M=3.351X10 kg. With these values the timeunit is 2.364 X 10 ' s.

In MD simulation one numerically integrates the clas-sical Newtonian equations of motion. Here the algorithmby Beeman' is used. To compute the forces the potential(1) is truncated at r =2.61 r.u. (=7.45 A for Ne), which

combined with their high mechanical strength, such ma-terials would be most useful as construction materials innuclear technology.

In this work the MD simulation is used to study radia-tion damage in a prototype amorphous solid. The sampleis a one-component system, where the interactions be-tween the atoms are described by the Lennard-Jones po-tential. This potential best approximates the noble-gasatoms, whereas real metallic glasses are alloys of two ormore constituents and the potentials are not necessarilyof the two-particle type. However, we expect to exposesome of the characteristics generic to radiation damage inamorphous solids. Although the absolute energy scalesof noble gas and metallic systems are very different, col-lision cascades corresponding to impact energies scaledby the displacement thresholds enable qualitative com-parison between different materials. In many respects theone-component Lennard-Jones system is also a muchstudied reference case. It is thus of interest to know itsirradiation behavior. In this report we present the resultsof 30 collision-cascade simulations done for a sample of1372 atoms. The PKA energies studied range from about1Ed to 80Ed. Ten different PKA's and directions areused. In Sec. II we describe the simulation method andconsider the definition and identification of point defectsin amorphous materials. Before starting with the radia-tion damage simulations we first determine in Sec. III thevalue of Ed for the amorphous Lennard-Jones system.Section IV deals with the energy outQow through thecomputational box surfaces during a cascade simulation.The collision-cascade simulations are presented in Sec. Vwith analyses of the cascade structure, point defects, andother related data. The results are discussed in Sec. VI,and Sec. VII contains a summary of the main findings.

for a crystalline fcc lattice is the distance up to betweenthe fifth- and sixth-nearest neighbors. The integrationtime step 4t varies depending on the stage of the collisioncascade: it is changed continuously during the cascadeaccording to the criterion that during one time step thefastest atom moves at the most 0.035 r.u. (0.1 A for Ne).At the end of a cascade simulation, when the system isnear equilibrium at a low temperature and the atoms canbe considered as vibrating around their equilibrium posi-tions, the time step is set by the requirement that oneatomic vibration should last (20—30) b, t Ac. cordingly,the maximum value of b t used in the simulation is 0.0296r.u. (=7X10 ' s for Ne). The use of periodic boundaryconditions is a standard procedure in MD simulation toextend the computational cell into a pseudoinfinite ar-rangement, removing thereby the effect of the boundaries.This is used here also except for the initial stage of a cas-cade, when a large amount of energy is injected in thecomputational cell and the energy fiow out of the cellmust be accomplished in some way. We shall return tothis question later. The temperature and the pressure ata given time are determined in the usual way from theaverage kinetic energy per particle and the virialtheorem, respectively.

The amorphous sample was prepared by a computer-simulated rapid quench from a liquid state (for details seeI). The pair distribution function of the amorphousstructure obtained is in very good agreement with otherstudies, as discussed in I and II. The sample temperaturehas been lowered in this study to 0.00148 r.u. (0.0522 Kfor Ne) from the previous 0.122 r.u. used in I and II. Theexternal pressure is zero. The identification of vacanciesand interstitials is an integral part of the analyses of anyradiation damage study. For amorphous solids, however,this is not straightforward and therefore a detailed studywas made in II to create working methods to be used inextracting point defect information from MD data. Forconsistency we briefly repeat here the procedure given inII for characterization of point defects in an amorphousLennard-Jones solid.

To find vacancies a cavity analysis is used: At first asphere of radius r is attached with each atom:

r =(4&2n ) (2)

where n is the mean atomic density. Then the space be-tween the atomic spheres is filled with nonoverlapping,empty spheres (cavities) as large as possible. Nearby cav-ity spheres will form a cluster, if (i) all the cavity spheresin the cluster have a volume larger than 0.076 r.u. (ii) atleast one of the spheres of the cluster has a volume largerthan 0.129 r.u. , and (iii) the surface of each cavity sphereis closer than 0.274 r.u. to some other cavity sphere be-longing to the same cluster. Given this, a monovacancyis identified as a single cavity with volume larger than0.26 r.u. or as a cavity cluster with volume in the range0.32—0.49 r.u. The volume V, of a cavity cluster for a di-

vacancy varies between 0.49» V, »0.74, for a 3 vacan-

cy between 0.74» V, » 0.98 and for a 4 vacancyV, ~ 0.98 r.u.

The interstitials are found by computing the local hy-drostatic press p; at each atom from

Page 4: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3980 J. LAAKKONEN AND R. M. NIEMINEN 41

(3)

a, =2

—2f]J

(5)

The summation in (5) extends just beyond the nearestneighbors. The interstitials have large negative pressureand in II a limiting value for the interstitial pressure powas proposed. This depends on the external pressure p as

po = 1.736p —12.230 . (6)

To exclude the borderline cases, here we increase thevalue given earlier by 10% and identify interstitials asthose atoms having a pressure p; &1.1po. The tempera-ture dependence of po was found to be weak up to halfthe glass transition temperature. Beyond this ~po~ startsto increase roughly linearly with temperature. In a col-lision cascade the temperature first increases sharply, butthe increase, however, is transient and applies only to thecascade atoms. Under these conditions the use of an une-quivocal temperature is doubtful, and ag a result the tem-perature dependence ofpo is not considered here.

where r, - is the distance between atoms i and j. Theatomic volume 0; is computed as

0 = a4m

1 3 l

where

2 -t gl 0

0

1.03, 1.08, and 1.09 r.u. , while the first maximum of thepair distribution function of the sample occurs at 1.07r.u. We take the atom A as origin and divide the line BCinto four equal parts. Then the vectors from A to B, C,and the three dividing points define five directions. Otherdirections are found by creating a plane, which is perpen-dicular to the plane formed by atoms A, B, and C andwhich includes the atom A and the center point of theline BC. The positions of the atoms at this plane andnearby are shown in Fig. 1. We scan this plane for thefull circle in steps of 15': the direction 0' is the atom A

moving towards the center of line BC, and positive andnegative angles are from this in the counterclockwise andclockwise directions, respectively.

The value of Ed for a given direction is found by givingthe atom A (called the test atom) impulses of varying en-

ergy and looking for the threshold value, at which atomicdisplacements of the order of one nearest-neighbor dis-tance emerge. Each run simulates 510 time steps (about3.6 r.u. of real time) and the difference of the positions atthe first and last time steps are computed for each atom.The distribution of the position differences is found to be

III. DISPLACEMENT THRESHOLD ENERGY

The threshold energy for producing permanently dis-placed atoms (displacement threshold in short) Ed isdefined as the kinetic energy an atom must have to leaveits equilibrium site and not to return. For a crystal thismeans formation of a vacancy-interstitial pair (a Frenkelpair) and the value of Ed is the minimum energy for sucha defect to be stable against recombination. In the case ofan amorphous structure the situation is different: whenan atom is displaced, the empty space at the originalatomic position can readily be filled by several smallmovements of the neighbor atoms or by one large leap ofa single atom. Likewise the displaced atom can easily"dissolve" in the material at its new position and is notnecessarily counted as an interstitial. For this reason it isobvious that as a result of an atomic displacement in anamorphous structure anything analogous to a Frenkelpair of a crystal does not necessarily form. Therefore wemonitor Ed from the changes in the atomic positions in-stead of the appearance of a vacancy-interstitial pair.

The displacement threshold depends on the directionof the atomic movement. In amorphous systems there isno crystalline order to define directions and local con-cepts must be used instead. We do this by seeking threeatoms in the middle of the computational box, all ofwhich are nearest neighbors to each other. The distancesbetween the atoms found (labeled as A, 8, and C) are

-2 -I

-2

0

FIG. 1. Positions of atoms at the plane, where the displace-ment threshold energy is studied. The test atom ( A) is shown asa solid circle and is situated in the middle of the volume plotted.Direction 0' is indicated by an arrow. Atoms in this and the fol-lowing figures are plotted as follows: the volume plotted is arectangular box that has two edges of equal lengths and that liein the plane of the paper (x and y directions). The length ofthese varies from plot to plot and can be seen from the figure.The third edge (z direction) is always 2.806 r.u. long. The box isdivided in the z direction into three equal smaller boxes. Anatom is now shown as a triangle, circle, or a square dependingwhether it is in the uppermost, middle, or lowest box, respec-tively. The length 2.806/3 r.u. in the z direction makes each ofthe boxes to be slightly less than the nearest-neighbor distanceand thus each of the boxes show only one "atomic layer. " Re-duced units (r.u. ) are used in this and all the other figures.

Page 5: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

41 . . . RADIATION DAMAGE IN AMORPHOUS SOLIDS 3981

continuous and reaching up to -0.63 r.u. However, ifthe energy has been large enough distinct displacementsof 0.8 r.u. and larger can be seen. Initially the sample hasthe temperature of 0.00148 r.u. (0.0522 K for Ne) andthe pressure is zero. Since only small energies are con-sidered, we apply here the regular constant volume simu-lation with periodic boundary conditions.

Figure 2 shows the displacement of the test atom as afunction of the kinetic energy given to it. The directionof the impulse is 16.7 from the 0' direction towards atom8 (i.e., in the ABC plane perpendicular to the plane ofFig. 1). It is seen that initially the displacement is verysmall, but at 292 r u .(0..8875 eV for Ne) it suddenlyjumps up to 0.83 r.u. and changes only slightly when theenergy is further increased. The test atom has clearly ex-perienced a displacernent from one equilibrium positionto another. The atomic movements are shown in Fig. 3,where the test atom is seen to initiate an actual sequenceof replacement collisions. A supplement run simulatingan additional 3000 time steps confirms that the replace-ments of the test atom and the next three atoms along thechain are permanent. Figure 4 shows that the kinetic en-

ergy is transferred from atom to atom very efficientlyalong the chain and, for example, the test atom itself isseen to give off nearly all of its kinetic energy. Note thatthe maxima of kinetic energy for the other permanentlydisplaced atoms are considerably less than what is neededfor the test atom to get displaced. Figures 3 and 4 are ex-amples of the property of a collision cascade to formchains that resemble focused collisions in crystallinestructures and, correspondingly, in the following we usethe concept of a focused collision also for amorphous sys-tems. Figure 3 is seen to exhibit also a replacement col-lision sequence.

Figure 5 shows the energy dependence of the displace-ment for 8=45' in the plane of Fig. l. As shown earlierin Fig. 2, the change in the displacement is very clear.Now this happens at a much lower energy of 133 r.u. , yetthe displacement (1.46 r.u. ) is larger than before. It is in-teresting to see that the test atom displacement stays ap-proximately constant up to rather high energies of 1200r.u. , when another jump to 2.5 r.u. occurs. The stepwisestructure is obviously a result of the form of the

)( ~

0r4~ r2,

03

-2—..J2$ g

-2

FIG. 3. Atom movements for the threshold energy (292 r.u. )

of Fig. 2. In the plot 3.98 r.u. of real time is simulated. Thelarge symbols show the initial positions of the atoms (cf. Fig. 1).Trajectory of an atom is shown by a line, the type of which indi-cates the box (z coordinate), where the atom is moving: dashedfor the uppermost, solid for the middle, and a dotted line for thelowest box. A small symbol shows that an atom has left (or en-tered) the plotting volume; the symbol type has the same mean-

ing as for initial positions.

potential-energy surface: an atom must be given enoughenergy to leap from one local minimum to another; if theenergy is not enough the atom stays at its position andgives the excess energy to its neighbors. A study of theatomic movements reveals that in this case the test atomdoes not force any atom to leave its site, but the test atomfinds itself a new site between the neighbor atoms. Theformer site of the test atom is filled by one atom, whoseempty position is then taken over by another atom.

In the two preceding cases discussed, the test atom ex-periences a sharp increase in the displacement when the

300

1.0— 200

0e 0.5-MO

100

2II

II

4

00 100

I

200ENERGY

I

300I

400 0.04TlME

0.08 0.12

FIG. 2. Displacement of the test atom as a function of kinet-ic energy given to it. For the direction see the text. The linesconnecting the symbols are drawn only to guide the eye.

FIG. 4. Kinetic energy of the labeled atoms of Fig. 3 as afunction of time. Other atoms of Fig. 3 receive much less ener-

gy than the labeled ones.

Page 6: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3982 J. LAAKKONEN AND R. M. NIEMINEN

1200

800-

2—ZLU

UJC3

CLCO

O 1

400—

-180 -90

8 (deg)

90 180

o'0 1000 2000

ENERGY

3000

FIG. 5. Displacement of the test atom as a function of kinet-ic energy given to it. Direction of movement is 45' in the planeof Fig. 1 ~

kinetic energy exceeds a certain value. This, however,need not always be the case, and a few directions havebeen found where the displacement has no stepwise form.In these cases the energy of the test atom is distributedrather evenly by many atoms, and as a result it is dimcultfor any atom to make a distinct displacement. By in-creasing the energy large displacements finally emerge,but because of the complex energy distribution the dis-placement may fall down with increasing energy whennew scattering channels become possible.

The preceding discussion shows that when the testatom is given an energy impulse large enough, then theatom, one or more of its neighbors, or both will get dis-placed by about one nearest-neighbor distance. Wedefine the displacement threshold Ed in a given directionto be the lowest test atom energy, when any atom makessuch a displacement. Figure 6 shows the results of thedisplacement threshold simulations for directions in theplane of Fig. 1. Also the five directions in the plane ABCare shown. The inaccuracy of the points is 5% or less.One immediately observes the large variation of Ed from118 to 1040 r.u. The minima occur at 30' and —90', theformer of which corresponds to an "easy or open direc-tion" (see Fig. l) and the only displacement is due to thetest atom itself. In the latter case the displacement ismade by one of the neighbors, which moves into one ofits easy directions after receiving energy from the testatom. The local maxima of Ed at —30', 120', and 150are all cases, where many atoms receive a sizable portionof the test atom energy. Note that a direct hit towards anearest-neighbor atom or just between a pair of themdoes not result in a maximum nor a minimum of Ed. Theaverage displacement threshold energy computed fromthe figure is 440 r.u.

Using the definitions given in Sec. II for point defects,it is possible to study their existence during an atomicdisplacement in amorphous solids. For a vacancy we findthat when an atom is being displaced, the former atomic

FIG. 6. Displacement threshold energy as a function ofdirection in the plane of Fig. 1. The triangles show threshold

energy for directions at a plane perpendicular to the plane ofFig. 1; the rightmost and leftmost triangles are head-on direc-tions towards the two adjacent nearest neighbors (atoms 8 and

C) of the test atom.

site is initially identified as a vacancy but disappears soon(in about 0.8 r.u. ) through the movements of nearbyatoms. The displaced atoms themselves behave as inter-stitials, but only for a very short time (about 0.08 r.u. ).After this time their new neighborhood has readjusted it-self and the displaced atoms are no longer identified as in-terstitials. The stability of the atomic arrangements aftera displacement has been tested by extending some of theregular 550 time step runs another 3000 time steps more.These show no novel features, and the displaced atomsare seen to stay at their new positions.

IV. ENERGY DISSIPATION AT BOUNDARIES

In a collision event the energy received by the PKA isfirst transferred into kinetic energy of the cascade atomsand the cascade volume heats up considerably. Afterthis, heat is quickly transported away to other parts ofthe solid and the temperature of the cascade volume de-creases. From the computer simulation point of view thesolid away from the cascade body represents a heat sinkand, unless the PKA energy is small or the sample large,this must be accounted for somehow. A natural way is touse boundary conditions that make possible energy toflow out of the computational box through the surfaces.Obviously the traditional periodic boundary conditionsdo not fulfill this, since they do not allow a particle aswell as energy to leave the computational box.

One way to accomplish energy outflow at surfaces is toadd a dissipative force component to the equations ofmotion for atoms near the surface. The dissipative forceis proportional to the particle velocity and provides ameans of energy consumption. The proportionality con-stant is chosen to minimize boundary reflection. 'Beeler even gives a formula based on an analogy of adamped, one-dimensional oscillator. In addition to thedissipation, other considerations are needed for the sur-face atoms. Gibson et al. take the atoms outside thecomputational box into account by applying two otheradditional forces on the surface atoms: a constant force

Page 7: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

41 . . . RADIATION DAMAGE IN AMORPHOUS SOLIDS 3983

and a spring force. The former accounts for the cohesiondue to conduction electrons, and the latter is used to de-scribe the material resistance for slow deformations ofthe solid. King and Benedek just keep the surface layerof atoms 6xed. ' Beeler suggests representing the atomsbeyond the boundary region as an elastic continuum. Aslight disadvantage is that each of the methods describedcalls for modifying the integration schemes for positionand velocity of the surface atoms. Also the choice of thevarious force constants is not unequivocal. For thesereasons we use another method, which requires minimalchanges in the simulation algorithm and yet provides forthe energy outflow in a reasonable way.

We start with the periodic boundary conditions but, inaddition, define a boundary region at the surface so thatthe atoms inside the boundary region are treated in a spe-cial way: the kinetic energy of these atoms is computedperiodically, and if this exceeds a given limit then, retain-ing the direction, the atomic velocity is scaled down tosome low value. In this way the integration algorithmsneed not be changed, and no new forces are needed, sincethe boundary atoms feel the interaction of the atoms atopposite surfaces because of the repetition of the compu-tational box through the periodic boundary conditions.We call this procedure an application of a heat bath. Pa-rameters in this model are the thickness of the boundaryregion, a scaled down limit for kinetic energy, kinetic en-

ergy after scaling, and time between successive scalings.These are chosen so that (i} the energy outflow approxi-mates an infinite sample, (ii) reflection at cell boundariesis at minimum, and (iii} the volume of the boundary re-

4 & ~ ~ohp8 p pa

h CLI od $ g ae 6 8p b o

h- e P ',' ~- ad d

sQ. p'

d-h h 0 8

p

o " e g

(8)p p 'p:. o

I CL d s8

ph

ph

h0

-4 o. I-

gion is as small as possible. The first two criteria abovefavor a thick boundary with few scalings, which is inconflict with the third requirement, and thus a comprom-ise is needed.

The Srst parameter set studied has a thin boundary re-

gion, the thickness of which is only 0.35 r.u. , this beingless than the nearest-neighbor distance. %'ithin theboundary region the velocity of atoms having kinetic en-

ergy above 0.0078 r.u. is scaled down to a Maxwellian ve-locity distribution with average kinetic energy of 0.0014r.u. (this corresponds to temperature 0.05 K for Ne). The

30

10

0

20 -':.

l~

10—

ated particles

Other particles

average T

NO HEAT BATH

WITH HEAT BATH

4 0

tr —$—0

h

' ph

4 0

6I 0 h P

h0S ~~ 0

pgl

I

o+ "'

(k-

h

o-.,Q,

Qi

ho 6

0po

h6 o'Qne 0h

o

Q0 D

Q0

0 (b)p

t 0

0 I 1

o o h t)0

o oh

o

00 0.5 1 1.5 2

TIME

FIG. 7. Cooling of the 300 heated atoms in the cases of noheat bath and with heat bath. The average temperature andtemperatures of the heated and not heated atoms are shown.

FIG. 8. Atomic trajectories showing the eftect of the heatbath. The initial position of the PKA is indicated by an arrowand the energy given is 10000 r.u. The plotted volume is orient-ed parallel with the computational cell. Solid lines mark thecomputational box surfaces, and the dashed line in (b) indicatesthe boundary of the heat bath region. In (a) no heat bath isused, and trajectories span time 0—0.191 r.u. (b) is obtainedwith a heat bath; the time interval is 0—0.208 r.u.

Page 8: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3984 J. LAAKKONEN AND R. M. NIEMINEN 41

scaling is done every fifth time step. The energy outflowis tested by setting the kinetic energy of 300 atoms in thecenter of the computational box to 28.3 r.u. (1000 K forNe), while the rest of the 1372 atoms are initially at tem-perature 0.0015 r.u. , and by monitoring the cooling of theheated atoms. The 300 atoms are situated within a radiusof 4.10 r.u. from the center of the computational box.The upper part of Fig. 7 shows the various temperaturesas a function of time with regular periodic boundary con-ditions and no heat bath applied. An interesting featureis the bumps in the temperature of the heated atoms ap-pearing first from 0.5 to 1.2 r.u. with the second onestarting around 1.5 r.u. The bumps result from theperiodic boundary conditions, which return the heatpulse back to the center. This, of course, is a spuriouseffect. A real effect, on the contrary, is that a part of theheat pulse can be reflected back from the cold atomsaffecting the cooling of the hot atoms, but this wouldhappen sooner and not until a time lag of 0.5 r.u. ; in fact,some delay is seen in the temperature decrease of the hot

atoms during a period from 0.07 to 0.2 time units. Thedip in the average temperature at 0.035 r.u. marks thetime when the initial excessive kinetic energy is parti-tioned in correct proportions into kinetic and potentialenergies. The effect of the heat bath is evident in thelower part of Fig. 7, which shows the results of a simula-tion like the preceding one but using a heat bath with theparameters given. The bumps in temperature have nowvanished and the temperature decreases steadily. Aslight reminiscence of a bump around 0.8 r.u. is probablydue to some energy reflecting from the boundary region.

In addition to making energy outflow possible, the heatbath should not distort the cascade structure inside thecomputational box. To study this we have made two cas-cade simulations, giving a kinetic energy of 10000 r.u. toan atom in the center of the box and recording the atomictrajectories with and without the heat bath. The atomictrajectories are shown in Fig. 8(a) for no heat bath and inFig. 8(b) with heat bath used. The surface of the compu-tational box is also plotted, and the dashed line in Fig.

t = 0.0214

FIG. 9. Spreading of kinetic energy during the simulation of Fig. 8(a) (i.e., no heat bath is used). The contour lines show the con-stant kinetic energy values of the atoms inside the plotting volume, which is the same as in Fig. 8. The time instant (t) and the con-tour spacing (b T) are shown in each plot. Note that 6T refers to the average kinetic energy per atom, which is now not the same asthe temperature because of a component of net outward motion in the cascade.

Page 9: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

41 . . . RADIATION DAMAGE IN AMORPHOUS SOLIDS 3985

8(b) shows the boundary of the heat bath region. FromFig. 8(a) one clearly sees how the cascade proceedsbeyond the computational box, but because of the period-ic boundary conditions it in fact enters the same box fromthe opposite side. When the heat bath is used [Fig. 8(b)],the cascade terminates very effectively at the box surface;only two small branches succeed in escaping through theheat bath region. The atomic trajectories in the heat bathregion and in its immediate vicinity differ, of course, butfor trajectories of the other atoms the effect of the heatbath is remarkably small. Figures 9 and 10 describe thespreading of kinetic energy for the cascades of Figs. 8(a)and 8(b), respectively. Each contour plot shows the con-stant kinetic energy distributions of the atoms inside theplotting volume at the time given. In Fig. 9(a) a few col-lisions have occurred, and the energetic atoms are shownas a number of concentric contour lines. Here the cas-cade begins to form. In Fig. 9(b) the front of the cascadeis just passing the computational box surface. Figure 9(c)shows the energy distribution long after the cascade fronthas left the plotted volume but yet before the cascade hasstarted to interact with itself. Inclusion of the heat bath

does not change the start of the cascade and the energydistributions are identical to those of Figs. 9(a) and 9(b).When reaching the surface region, the heat bath starts toaffect the moving cascade front [Fig. 10(a)] and absorbsthe cascade energy blocking its propagation. Figure10(b), which is for about the same tiine as the end of thetime period plotted in Fig. 8(b), reveals the possible ener-

gy reflection at the boundary to be minimal indeed. Thelast figure [10(c)] shows the energy distribution at thesame time as Fig. 9(c), and, as can be seen, the distribu-tions inside the computational box are of the same mag-nitude but the detailed forms are not quite alike. Thelargest kinetic energy encountered in these figures isabout 17 r.u. This is much less than the lowest displace-ment threshold 118 r.u. found in Sec. III, and thus theatoms cannot cause any displacements in the cold partsof the solid. The path of the main cascade is seen to heatup to and beyond the melting point and cooling of this issomewhat different for the two cases, at least near the boxboundaries.

The chosen width of the boundary region results in areasonable energy outflow at the surfaces with little

t = 0.0681

S S-~I

go~~o

gh(C

FIG. 9. (Continued).

Page 10: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3986 J. LAAKKONEN AND R. M. NIEMINEN 41

reflection, while the advance of the cascade is effectivelycutoff at the boundary. Possibly the most essentialdifference when compared with an infinite sample is inthe initial stages of the radiation damage self-annealing,as especially near the boundary region the defect struc-ture can be affected. The effect of the boundaries dimin-ishes when going inside the computational box and alsoduring the course of time, when the average temperaturerapidly decreases. Doubling the width of the boundaryregion with less frequent scaling results in a slightlybetter energy dissipation and smaller distortion of thetrajectories near the boundary. The small improvementis, however, outweighed by the larger number of the spe-cially treated boundary region atoms.

V. RADIATION DAMAGE SIMULATIONS

We have studied the development of radiation damagein amorphous solids by performing collision-cascadesimulations for ten different PKA energies ranging from330 to 26300 r.u. For each energy three different PKA

positions and directions are used and thus a total of 30collision cascades have been simulated. The PKA posi-tions are within 2 r.u. from the computational box sur-face, and the energy impulse directions are chosen so thatthe computational box volume is utilized in the best way.In a real collision event the atoms will lose energy also inelectronic interactions. A standard way to incorporatethis effect into computer simulations is to subject theatoms to some energy loss per path length, which at lowenergies is proportional to the atomic velocity. ' Elec-tronic loss, however, is prominent for energies larger thanthose considered here, and the effect is not included inthe present simulations.

Each simulation run consists of two parts: the firstpart is 1500 time steps long and simulates the initial stageof a cascade. During this the heat bath explained in Sec.IV is used. For the second part no heat bath is appliedbut instead at first (during 2000 time steps) the atomic ve-locities and the computational box volume are adjustedso that the sample temperature is about 0.028 r.u. withzero external pressure; finally a 1000-time-step-long equi-

I

1htI

l(' )

t = 0.316 I

0)%J /~

\

'L

L„

FIG. 9. {Continued).

Page 11: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

41 . . . RADIATION DAMAGE IN AMORPHOUS SOLIDS 3987

librium period follows when no scalings are done. Thefirst part simulates 10—16 r.u. of real time and thesecond one about 90 r.u. Initially the sample temperatureis 0.0015 r.u. During the cascade it rises up to 12 r.u. forthe highest PKA energies but because of the heat baththe temperature then decreases to 0.03 r.u. or less by theend of the first part.

Figure 11 shows a typical example of the initial stagesof one collision cascade. It is clearly seen how the PKAenergy is at first distributed to atoms situated in a cone,which has its tip at the PKA position and opens in thePKA direction. The cone formation is observed for allcascades; the cone angle is near 45' and appears to in-crease slightly with increasing PKA energy. In the figureone can also find a focusing sequence. Though focusedsequences do not occur for every cascade, they are stillvery common, and one or more are seen for about half ofthe cascades. For a fixed PKA direction the focusing se-quence usually stays the same for a large range of thePKA energies studied, but in one case we find the focus-

ing sequence changes when the PKA energy is varied.The importance of the focused collisions lies in their abil-ity to increase the extent of the radiation damage. In Fig.11 one sees that a great number of atoms (that are notcontained in the focusing sequence) experience large dis-placements, and thus they have considerable amounts ofkinetic energy. The same applies to all focused collisionsfound. %e correspondingly expect a focusing collision inan amorphous solid to receive less of the available energythan in a crystal. Also, because of the lack of periodicstructure, the range of a focusing sequence is obviouslyshorter in amorphous materials. These reasons renderfocusing collisions to be of lesser importance in amor-phous solids.

Figure 12 shows the number of vacancies and intersti-tials as a function of time for two PKA energies, 2000and 11500 r.u. The immediate observation is that thegeneration of point defects as a result of a collision cas-cade is a transient event and the defects vanish soon afterthey are created. The vacancies are seen to disappear in

t = 0.0681

&o 4 eO/

a

~l

FIG. 10. The same as Fig. 9, but for the case of Fig. 8{b),where a heat bath is used.

Page 12: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3988 J. LAAKKONEN AND R. M. NIEMINEN 41

1 and 3 time units for PKA energies of 2000 and 11 500r.u. , respectively. For interstitials these values are 7 and—15 time units. Figure 12 is typical for all the cascadesstudied, with appropriate scaling by PKA energy.Specifically, at the end of a simulation run we find no va-cancy left for any cascade and only one or two intersti-tials for nine of the 30 cascades. The existence of intersti-tials is not connected with the PKA energy and, in fact,refiects the typical equilibrium state of the sample (seeII). Both defect types start to form at the same time butthe interstitial maximum occurs slightly sooner than thevacancy maximum. The interstitials survive longer thanthe vacancies. The number of interstitials reaches max-imum at about a time 0.19 r.u. , and the vacancy numberis at maximum around 0.29 r.u. Unexpectedly, thesetime instants turn out not to depend appreciably on thePKA energy; for vacancies, a very slight increase with in-creasing PKA energy is observed.

It has been explained in II that an interstitial can showup as increased local pressure for one to three atoms.

Therefore in Fig. 12 we distinguish "all" and "distinct"interstitials: interstitials being nearest neighbors arecounted as one in 'distinct" values. It is seen that initial-ly a large number of interstitial neighbors are created, butthe local pressures are soon smoothed, and most of thesurviving interstitials are distinct. The value of the PKAenergy has an effect on the maximum interstitial numberand the level at which the interstitial number sets afterthe first peak. The average maximum numbers of (dis-tinct) interstitials and vacancies are plotted in Fig. 13 as afunction of PKA energy. One sees that the number of in-terstitials grows very fast initially, but when the PKA en-

ergy increases the growth rate decreases, reaching satura-tion at around 15000 energy units. The number of va-cancies created is for low PKA energies smaller than thatof interstitials. In this range the vacancy number in-creases linearly with PKA energy up to 11500 r.u. , afterwhich the growth is somewhat slower but still faster thanfor interstitials. Beyond 16500 r.u. more vacancies thaninterstitials are created. However, it is conceivable that

t = 0.201

0

FIG. 10. (Continued).

Page 13: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

41 . . ~ RADIATION DAMAGE IN AMORPHOUS SOLIDS 3989

the sample is too small for the largest PKA energies() 15000 r.u. ).

Figure 14 shows the energy dependence of the (distinct)interstitial number right after the first peak is over and aslow decrease has set in. This happens at time t ~0.42r.u. and can be determined rather accurately as seen, e.g.,from Fig. 12(b). Now the interstitial value increases near-ly linearly with the PKA energy up to 11 500 r.u. and, in-terestingly enough, this increase closely resembles thatfor the vacancies. After 11 500 r.u. the growth rate clear-ly changes. The slow decrease of the interstitial numberafter the first peak is approximately linear with time, ascan be seen from Fig. 12(b).

Lindhard et al. ' have given the following integralequation for the average number v(E) of displacementsproduced by a PKA of energy E:

I K(E, T)[v(T)+v(E —T) v(E)]—dT=O, (7)0

where E(E,T), the scattering law, is the probability thata particle with initial kinetic energy E will transfer kinet-ic energy T to another particle in a single collision.When analytically solving this equation it is usually as-

sumed that (i) the cascade consists of binary elastic col-lisions, (ii) the displacement probability P(E) is a stepfunction P(E ) =8(E E—z ), where Ez is the displacementthreshold, and (iii) the energy Ez, consumed when mak-ing a displacement, can be neglected. Robinson' ' andSigmund' have studied this equation in detail for variousforms of the scattering law and the equation has beengeneralized to allow the particle also to move withoutmaking any collisions (describing channeling and replace-ment chains). In practice, however, the simple Kinchin-Pease model' has been widely applied instead. In addi-tion to the preceding assumptions (i) —(iii) the Kinchin-Pease model considers only hard-sphere scattering anddoes not account for the crystal structure. With these as-sumptions one obtains, for the number of displacements,

0 for E &E&

1 for E~&E~2E&vE =.for ZE„&E .

d

L1'

t = 0.316

-i-q-~~(

~ /

pb it+f h/ f y

(C) (

FIG. 10. (Continued).

Page 14: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3990 J. LAAKKONEN AND R. M. NIEMINEN 41

The original Kinchin-Pease model has been later im-

proved (the modified Kinchin-Pease model' ' ) so thatthe number of displacements reads

cause displacements, and thus the end of the intervalmarks the time when the cascade has ceased to expandand the number of defects created is about at maximum.For the minimum value of an acceptable displacement we

0 for 0&E &E„

l for Ed&1 & E—d

2v(E)= a

(9) 12—

K A.for —Ed &E .

2Ed

Above, P is the energy actually left for damage produc-tion when the part spent for electronic excitations hasbeen removed. The displacement eSciency K is used toallow for deviation of the atomic interaction from ahard-core potential.

Excluding defect clustering, a displaced atom in a crys-tal along with the vacancy left behind forms a Frenkelpair and often one uses the number of Frenkel pairs as asynonym for the number of displacements. For amor-phous solids, however, there is no more such correspon-dence. We thus consider explicitly the number of dis-placements when making a comparison with theKinchin-Pease model. Since the displacement is de6nedhere as a difference in the atomic position at two timesteps instead of referring to a misalignment with respectto the periodic structure, it is important to properlychoose the time instants, when the displacements arecomputed. The initial positions are, of course, those atthe beginning of a simulation run. The end of the inter-val we set at a time when the kinetic energy of the fastestatom has been less than 118 r.u. at least for 30 time steps.This energy value is the lowest energy found in Sec. III to

8

Z 4

090 —'

60—

30

0

60-

vacancies

interstitials

(a)

2 3

4 ih h1

0 JC38

40

Z20

0—Obr+

3

0

o0

00 g-4 — 0

pQ

Qr Qg ~ CP

0

8

e&

0 ~

0180 -'

00

00

cr 120- p

60

TIME

interstitials

b)

0

FIG. 11. Collision cascade caused by a PKA with the energyof 8200 r.u. The initial PKA position is shown by an arrow.The plot is made for the first 0.25 r.u. of time from the cascadestart.

FIG. 12. The number of vacancies and interstitials as a func-tion of time during a collision cascade. The PKA energy is 2000r.u. in (a) and 11 500 r.u. in (b). For interstitials the open sym-bols show the number of all interstitial atoms and the solid sym-bols give the number of distinct interstitials (interstitials that arenot nearest neighbors) ~

Page 15: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

41 . . . RADIATION DAMAGE IN AMORPHOUS SOLIDS 3991

90 60—

60

Z',

30

00 10000

ENERGY

20000

MI-Z'.

40-O

CO

020-

CQ

Z

FIG. 13. The average maximum number of vacancies (cir-cles) and distinct interstitials (triangles) as a function of thePKA energy. The values correspond to the maxima of curveslike those in Fig. 12. The average of the three runs for each en-

ergy is given. The lines are drawn only to guide the eye.

00 10000

ENERGY

20000

use 1.07 r.u. , i.e., the position of the first maximum of thepair distribution function. Figure 15 shows the numberof displacements as a function of the PKA energy. Thedependence is seen to be linear in energy up to 15000r.u. ; thereafter the number of displacements appears toincrease faster. A least-squares fit to a straight line forE &15000 r.u. gives the energy prefactor 1.30X10r. u. . This is close to the Kinchin-Pease factor—,'Ed =1.14X10 r. u. ', where for Ed the average value440 r.u. from Sec. III is used. Using these results we ob-tain, for the displacement eSciency a of the modifiedKinchin-Pease model, an energy-independent value at1.14. As a comparison, for crystals the constant valueK=O. 8 has been generally accepted. ' ' At large energiesmore defects are produced than the linear increase wouldindicate. But, in accordance with the preceding discus-sion, this is obviously an artifact due to the small sample:

60—

FIG. 15. Number of displacements as a function of energy ofthe PKA atom. The circles show the average and the bars indi-

cate the standard deviation of the three separate collision-cascade simulations done for each energy. The line is a least-squares fit for energies below 15 000 r.u.

the boundary conditions apparently do not allow for thelargest PKA energies to spread out fast enough but con-strain the energy in a relatively small volume for too longa time and thereby increase the displacement number.Note that the criterion used here for the end of the dis-placement time interval corresponds to the moment,when all the possible displacements caused by the PKAhave been made. This choice is consistent with theKinchin-Pease model, where an atom after being dis-placed is not assumed to recombine with a vacancy.King and Benedek find, too, that the number of Frenkelpairs in their crystal simulation corresponds to theKinchin-Pease value when the Frenkel pairs are countedin the end of the collisional or defect generation phase(when the number is at maximum). '

The Kinchin-Pease model sets a sharp cutofF for v(E)

& 30-z'

00 10000

ENERGY

I

20000

FIG. 14. The number of distinct interstitials at the time whenthe maximum defect number has been passed and the rapid de-crease is changing to a more slowly vanishing rate of intersti-tials.

500ENERGY

1000

FIG. 16. Number of displacements at low PKA energies asdetermined by Eq. (10).

Page 16: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3992 J. LAAKKONEN AND R. M. NIEMINEN 41

at the displacement energy Ed. However, as is evidentfrom Fig. 6, the displacement threshold actually dependson the direction: Ed =Ed(Q). Correspondingly, Kinget al.~2 define v(E) for energies near Ed by

v(E}=I e[E—E (0)]dQ4m

(10)

In Fig. 6 values of Ed(Q) for one atom have been present-ed. Using this data Fig. 16 shows v(E }for v 1 as deter-mined according to Eq. (10). For a crystal, the qualita-tive features of v(E ) are a sharp increase at the minimumthreshold energy and an extended plateau at a valuev(E) ( 1.' ' Both of these are assumed to be a result of

the periodic structure. Figure 16 indicates that in thecase of an amorphous solid v(E) does not rise abruptlyfrom zero, nor is there a distinct plateau before v(E) at-tains the value 1.

The actual positions of vacancies and interstitials for atypical cascade are depicted in Fig. 17. This is the samecollision event as considered in Fig. 12(b}. The PKA en-

ergy is 11 500 r.u. , and Fig. 17(a) shows the atomic inove-ments during the first 0.44 time units. At the very begin-ning of the cascade vacancies appear near the initial posi-tion of the PKA, while interstitials are created furtheraway by the advancing cascade front. Right after this[Fig. 17(b}, time is 0.095 r.u.] vacancies are also formed

5 =''0

e

iP g G

S0'g

L$yP ag

pa-.p ~g gt

h 8G

8 @ Q

ib ae

o 'P8

-5 -~ ~

0

g }, po:'g ",I~gf~ +.. ~Q

~~b g~ l3

~gP, '~' *"pO(i

~ Clh-.- g

(8) gar

iP

O er

p

h0

0h

ho h

h0

-5 -h 0

0h ~0 h

op 0

0

o oh h

0 ~ h 0p h ~0

Q~.W~ Q

8

I ~ o 'g I~C0 0 0 h D

h 0 h 00h Do

h 0 ho D~ 0 0

h ~ eo ~ o h (P

00 h

a ho0 ~ D 00 h

0h

O 0g 0

oh

0k4o

I

~ ~h h

D hp~ G

h ~ a op& o o

0 pk

0 (b) 8

OD

10

5 I 0 a D t~a O

D h a (~ a o D 0 a 0

a 0 h D Op 4 h 0 h0 ~ 0 0 a ~a h oh ~ h 0

0 8 a CQ h 0 ~ D(} oa~+0 D D 0

CI 0

0 h ()Q 4 + 0 ~ 00 p 0 h o D o0 s- ~ ', ', g+U' '

0 ~ ~a0 o hO ~ ~ a ho D ao

0 h o 0o 0 0p 0 ~ 4 D Dh oD 0

h ho o Q 0 Ro 0 o ~ o&

0 h h0 ~ D

Dp0 o a 0 o D R Oo o0 0 o D (c)o D 0

0 0 Q 0D 8V A a I CI l h~n n Se

FIG 17. A collision cascade lnltlated by a PKA atom with an energy 11500 r.u. (a) shows atomic tra3ectorles wlthln 0.44 timeunits from the beginning of the cascade. The initial PKA position is indicated by an arrow. (b) and (c) show the locations of defectsand atoms at times Q.Q95 and 0.67 r.u. , respectively. Noninterstitial atoms are plotted as described earlier. The interstitial atoms areindicated as solid symbols. Cavities of various sizes are represented by circles of different diameters d depending on the cavityvolume V: if V &0.49 r.u. , then d =3dp (dp is the diameter of the sphere representing an atom), for 0.49 V&0.98 r.u. , d =5.4dp(this is the volume range for clusters of two or three vacancies) and for V) 0.98 r.u. , d =9dp. A + or —sign inside a cavity sphereindicates that the cavity is in the upper or lower third of the plotted volume. All the plotted volumes in (a)—(c) are oriented in thesame way.

Page 17: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

41 . . . RADIATION DAMAGE IN AMORPHOUS SOLIDS 3993

along the cascade, making some of the initially createdinterstitials vanish. The majority of the interstitials arefound at the outer parts of the cascade. At the time whenthe numbers of vacancies and interstitials have passedtheir maxima [Fig. 17(c), time is 0.67 r.u. ; see also Fig.12(b)] the vacancies are seen to be located close to eachother near the beginning of the cascade, and the intersti-tials are further away surrounding the vacancies. In spiteof the vacancies being close to each other, we typically donot Snd them to form large clusters. The vacancy distri-bution also keeps changing somewhat, while the defectnumbers decrease continuously. Later, at time 2 r.u. ,only a few defects exist; the vacancies are still approxi-mately in the central region of the cascade volume andthe interstitials in the outer parts. No distinct defectcomplexes are found.

The correlations between the defects can be studiedquantitatively by computing the pair distribution func-tions for cavity-cavity [denoted as g„(r)), interstitial-interstitial [g;;(r)], and cavity-interstitial [g„(r)] pairs.We have done this for the cascade considered earlier inFigs. 12(b) and 17. Figure 18(a) shows the pair distribu-tion functions in the beginning of the simulation at time0.066 r.u. , when the cascade is rapidly growing and the

defect numbers have not reached their maxima [cf. Fig.12(b)]. At the time of Fig. 18(b) (time 1.26 r.u. ) the so-called short-term annealing has started and m.ost of thedefects created have already vanished. The cavity-cavitycorrelation does not change considerably during thecourse of time: a prominent peak is seen at 0.75 r.u. witha rapidly decreasing tail. The peak is due to several prox-imate cavities forming a vacancy and the short tail indi-cates that the cavities (vacancies) created are situatedwithin a limited region. The interstitial-interstitial distri-bution, too, has a clear peak at short distances but thetail is clearly longer than for cavities and becomes moreuniform with time. The long tail manifests quantitativelythe fact evident in Fig. 17, i.e., that the interstitials arerather evenly distributed in a large volume. Previously inII it has been stated that the interstitials appear often asnearest-neighbor pairs. This can also be seen from theinterstitial-interstitial distribution function in Fig. 18 asthe sharp peak at 0.9 r.u. The cavity-interstitial correla-tion function follows the spreading of the interstitialsalong with the cascade front. An interesting feature isthe prominent peak at 0.94 r.u. indicating of the ex-istence of some close cavity-interstitial pairs. Moreover,unlike g„and g;;, the cavity-interstitial distribution func-

40

cavity-cavity

80

cavity-cavity

La 40U)

10—interstitial-interstitial interstitial-interstitial

0

cawty-interstitial cavity-interstitial

00 4

DISTANCE

4DISTANCE

FIG. 1g. The different defect pair distribution functions of the collision cascade in Fig. 17 at (a) t =0.066 r.u. and (b) t = 1.26 r.u.

Page 18: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3994 J. LAAKKONEN AND R. M. NIEMINEN 41

tion is seen to vary to some extent during the annealing,and one case has been found where the peak at 0.94 r.u. isdisappeared.

From the pair distribution functions one can easilycompute the average distances between various defectpairs. This is done for the cascade considered earlier andthe average distances are plotted in Fig. 19 as a functionof time. After the rapid initial increase (following thecascade expansion) at time t =0.4 r.u. the averagecavity-cavity distance is 3.8 r.u. , whereas the interstitial-interstitial and cavity-interstitial distances ( -5.2 r.u. forboth) are considerably larger. Apart from fluctuations,the average distances are seen to decrease slightly withtime, and no abrupt changes take place.

A point defect in an amorphous structure can vanish,in principle, by two different mechanisms: a vacancy andan interstitial may annihilate each other through recom-bination or the immediate neighborhood of a defect canslightly rearrange itself and thereby dissolve the extravolume of a vacancy or ease the pressure of an interstitialatom (we call this mechanism the collective mode). Forcrystals the recombination (excluding the diffusion to thesurface and trapping at grain boundaries, etc.) is the onlypossibility. Considering the relative importance of thetwo mechanisms in short-term annealing of collision cas-cades some conclusions can be drawn from the pair dis-tribution functions. If the recombination were the princi-pal annealing mechanism then one might expect the firstpeak of the cavity-interstitial distribution function to di-minish as the defects annihilate, and the average cavity-interstitial distance should increase. In order to quantita-tively study the number of close cavity-interstitial pairswe consider the integral

4m f g„(r )dr1C 7

C

N-d

where N, is the number of cavities and R =1.24 r.u. sothat the integration region just includes the first peak ofg„. N;, is thus proportional to the number of interstitialsper one cavity that are closer than 1.24 r.u. , and we use itto estimate the prerequisites for recombination. At thefirst tiine instant studied (t =0.032 r.u. ) N;, = 1, but when

the cascade expands ¹,drops to 0. 1, . . . ,0.2 within 0.2time units and subsequently does not change consider-ably. Obviously a recombination event would show up asa simultaneous decrease in N;, and the number of cavitiesand interstitials. A few such events are found, but theycannot account for the most of the defects vanished. Asa result we conclude the collective mode to be the mainannealing mechanism. This is also supported by the ob-servation that g;;(r) changes only very slightly in time,indicating a stable interstitial distribution. Yet asignificant recombination rate would require a fast inter-stitial diffusion, we have found that the cavities do notmove appreciably before dissolving.

Computer simulation makes it possible to study wheth-er changes in volume and structure because of only onePKA occur. Since small changes are considered, it is im-portant that the equilibrating procedure described previ-ously is adequate. To study this we have simply per-formed another 3000-time-step-long equilibrium run afterthe first one for some of the cascades. All of the resultsshow no changes in the volume of the sample or in thestructure as disclosed by the pair distribution function,and thus the cascades can be considered as havingreached quasiequilibrium in the end of a simulation run.The change 6V in the sample volume is seen to lie in therange 0.46 —0.12-% of the initial volume, and its averageis negative. However, the sign and magnitude of hVvaries randomly with increasing PKA energy. Thus theobserved changes in 5 V are to be considered only as sta-tistical variation, and no definite conclusions of thevolume change can be drawn. Changes in the structureare studied by comparing the pair distribution functionbefore and after a collision cascade. The changes turnout to be very small; mainly the position and the heightof the first peak may vary in minute amounts, and, as be-fore, the changes are of a statistical rather than a sys-tematic nature.

~p:~ 0 ,.R' ''~a

~ ' '0..~0 0

0.0 0.5 1.0TIME

1.5 2.0 2.5

FIG. 19. Average distances between the defect pairs as afunction of time. Squares show cavity-cavity distance, trianglesare for interstitial-interstitial pairs, and circles denote cavity-interstitial distance. The data are from the collision simulationof Fig. 17.

VI. DISCUSSION

A common question concerning the quantitative relia-bility of MD simulation results is what the role of the po-tential used is. Besides the detailed form of the potential,the applicability of the pair potential formulation in gen-eral has also been discussed. Also the density dependenceof the potential has been scrutinized. Metals can be di-vided into two broad classes: simple and nonsimple met-als. The former are nearly-free-electron metals withtightly bound electron cores, which do not overlap appre-ciably with the cores of neighboring atoms. For these thepair potential concept finds justification via pseudopoten-tial perturbation theory. The situation is different for thenonsimple metals, a typical example of which are thetransition metals. Because of the d-band effects it is ques-tionable whether the energy of a transition metal can be

Page 19: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

41 . . . RADIATION DAMAGE IN AMORPHOUS SOLIDS 3995

expressed as a sum of pair potentials and recipes formany-atom interactions have been proposed. MDpractitioners, however, have used serniempirical pair po-tentials for transition metals with considerable successfor selected problems. The Lennard-Jones potential usedhere is physically suitable for solid noble gases, but theresults obtained can help elucidate the generic propertiesof disordered systems.

Formation of radiation damage is accompanied withchanges in the local density of the solid, and correspond-ingly one is faced with the possible density dependence ofthe potential and its effect on the defect structure. Ac-cording to Taylor the density dependence of a potentialaffects the energy changes associated with a defect morethan its structure. This applies even for extended defectssuch as grain boundaries (but not free surfaces) and Tay-lor concludes that a density-independent potential can beused to give a reasonable defect structure.

The question of density dependence also touches uponthe constraints of the simulation. de Leeuw et al. havedirectly compared the results obtained using constant-pressure and constant-volume simulations. They find theensembles yield the same static and short-term dynamicalproperties but observe slight differences at large times(e.g., different diffusive behavior). In the present workthese findings are accounted for by adjusting the samplevolume (i.e, using in effect a constant-pressure simulation}for the second part of the cascade when the short-termannealing is taking place. For the initial collisional andfinal equilibrium states the constant-volume simulation isapplied.

In this work a new method of accomplishing the ener-

gy dissipation at boundaries has been introduced. Themethod uses a boundary region and energy scaling, whichare easy to include in the simulation algorithm, and theappropriate parameter values can be readily determinedapplying the criteria given. We consider the parameterset used here a fair cornprornise between the various re-quirements, and the physically important requirementsfor the boundary conditions, reasonable energy outAowfrom the sample and little affect on the cascade structure,were seen to be attained. Obviously, there is always roomfor discussion of the optimal parameter choice or the bestway to include the energy outflow in the simulations, but,on the other hand, there are indications that the simula-tion results, after all, are not very sensitive to the bound-ary effects. Another question is the size of the computa-tional box (the number of atoms}. The results of thesimulations indicate that the present sample size may betoo small for large PKA energies (E) 15000 r.u. ) andcorrespondingly one should keep this possibility in mindwhen examining the outcome of high-energy cascades.

Figure 6 shows the displacement threshold energy todepend sharply on the displacement direction. This isslightly surprising, since a disordered structure mightlead one to expect more smooth variations. On the otherhand, King and Benedek find a rather similar rapidlyvarying threshold energy curve for crystalline Cu in theirsimulations, ' and Audouard et al. experimentally reachthe conclusion that amorphous and crystalline Fe75Bq5have the same displacernent threshold energy. King and

Benedek also discover similar easy and hard directionsand a stepwise type of displacement as found in thiswork. Likewise the atomic trajectories for both struc-tures exhibit replacement collision sequences witheffective energy transfer from atom to atom. It seems, infact, that the displacernent threshold energy has manyfeatures common for both the crystalline and amorphousstructures. This indicates that the displacement thresh-old is determined by the immediate vicinity of an atomand does not probe the long-range order of the material.The ratio of Ed minimum (118 r.u. , 0.36 eV for Ne) to thecohesive energy of a solid noble gas [0.02 eV for crystal-line Ne (Ref. 31)] is about two times as much as that forcrystalline Cu [Ed;„-—25 eV," cohesive energy is ap-proximately 3.5 eV (Ref. 32)]. Allowing for the definitionof a displacement and the exact value of cohesive energyfor the amorphous structure, these numbers suggest thatthe displacement threshold is slightly larger for the disor-dered than for the crystalline structure.

The collision cascades were seen to exhibit featuressuch as replacement and focused collision sequences. Inthis case a focused chain means collisions propagating viaa string of densely packed neighboring atoms rather thanatoms scattering into definite angles. In addition, nowthe focused chain has a curved trajectory and gives off en-

ergy more rapidly than in a crystal. In general, the cas-cade was observed to expand initially into a cone-shapedform, which results in a fast spreading of cascade energythereby reducing the extent of the affected region. Re-placernent collisions and rapid branching of the cascadehave also been seen in simulations of amorphous iron byYarnamoto et al.

A study of the temporal properties of the point defectnumbers revealed that the interstitial number is at itsmaximum slightly before that for vacancies, and thesetime instants change only little with the PKA energy.The minor energy dependence of the time correspondingthe maximum number can also be seen in cascade simula-tions for a crystalline structure. ' The time instant at amaximum number appears to be a fraction of a picosecond, the exact value depending on the potentialstrength and the atomic mass.

The PKA energy directly affects the maximum num-bers of defects. The vacancy number increases nearlylinearly with the PKA energy, for high energies slowerthan for small energies. The prompt interstitial numberinitially rises very fast but levels off at large energies. Thevanishing of defects is slower for interstitials than for va-cancies. Finally, after equilibrating, no vacancies andonly 0—2 interstitials were observed. The interstitialswere further connected with the normal, semiequilibriumstate of the sample rather than resulting from the col-lision cascade. These findings are clearly in contrast withsimulations for crystalline samples, where stable Frenkelpairs or displacernents after cascades are routinely seen.

Comparing the present results with those obtained for asimilar sample by Chaki and Li, the time of reachingthe maximum interstitial number is the same for both,but otherwise the time development of the defect num-bers are different; for example, Chaki and Li find that thevacancies disappear more slowly than the interstitials.

Page 20: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

3996 J. LAAKKONEN AND R. M. NIEMINEN 41

The reason for this probably lies in the different descrip-tion of point defects and the simulation algorithm itself.(Chaki and Li use Voronoi volumes for defectidentification, and they do not allow for energy outflow atthe boundaries. ) At the end of a collision simulationChaki and Li do not find any new defects, either.

A study of the defect positions revealed that the vacan-cies are created near the PKA position at the center ofthe cascade and the interstitials spread around them.This is an interesting observation, as the same is knownalso to happen with crystals. ' ' In spite of being closeto each other no vacancy clustering was seen; on the oth-er hand, the vacancies exist only a short time before van-ishing. The defect pair correlation functions showed theinterstitials to be distributed rather evenly around the va-cancies. The computed average vacancy-vacancy dis-tance appears to be considerably less than the vacancy-interstitial or interstitial-interstitial distances supportingthe suggested picture. The pair distribution functionsalso helped to differentiate between the two main possi-bilities of short-term annealing: a mutual recombinationof a vacancy and an interstitial or individual vanishing ofa defect. We suggest the latter; recombination wouldpresume a significant mobility of the interstitials, which isimprobable because of the stable interstitial distribution.Low mobility of interstitials is also supposed by Chakiand Li. On the other hand, individual vanishing of de-fects is clearly needed to make the vacancies and intersti-tial disappear at different rates.

In order to compare the results with the predictions ofthe modified Kinchin-Pease model, we determined thenumber of displacements at the moment, when the cas-cade has ceased to expand and the defect number isaround its maximum. Because of the missing periodicity,a displacement is defined here directly as a difference ofan atom's position at two time instants. Following theKinchin-Pease model the displacement number turnedout to increase linearly with PKA energy, except forlarge energies where the increase is more rapid. With thedisplacement threshold Ed =440 r.u. we obtained, for thedisplacement efficiency, a=1.14, which should be com-pared with the value 0.8 often used for metals. We findthis correspondence to further support the value pro-posed here for Ed. Obviously, an exact value of Ed can-not be set unequivocally, and, e.g., for Cu the valueEd =25 eV is used often even if Gibson et al. find the dis-placement threshold to vary between 25 and 85 eV.Note that a smaller value of Ed would correspondinglydecrease ~, too. The curve for the number of displace-ments near the threshold (v ~ 1) turned out not to clearlyexhibit the features typical for crystals: a sharp increaseand a distinct plateau. This we consider to follow fromthe lack of periodicity in an amorphous structure so thatthe existing short-range order is not enough to producethese details. Further studies would be in order to collectmore statistics for other directions.

It is well known that heavy irradiation causes swellingespecially in crystalline but to some extent also in amor-phous metals. Computer simulation with periodicboundary conditions is capable of disclosing changes inthe sample volume as observed, e.g. , in publication I

when preparing the amorphous structure from a crystal.We checked this for swelling due to a collision event butfound no systematic changes in the sample volume or thepair distribution function. This is not surprising, sinceonly one PKA in the sample corresponds to a veryminute dose, and many more cascades (one following theother) would be needed to simulate the doses observed tocause swelling. However, Yamamoto et al. report dis-tinct changes in the pair distribution function after onlyone cascade.

VII. CONCLUSIONS

A computer simulation study of radiation damage inan amorphous solid involves a proper simulation tech-nique and a practical way to identify point defects in thestructure. The latter subject has been studied in our ear-lier publication (II). In this work the usual MD methodhas been modified to allow for energy outflow throughthe computational box surfaces and constant-pressuresimulation has been applied during defect annealing. The30 collision-cascade simulations made show the energy ofa PKA to be rapidly distributed to nearby atoms, and asa result the affected cascade volume is limited. Collisionsequences such as replacement collisions and focusedchains are observed. The defect structure is found to becomposed of mostly vacancies in the center with evenlydistributed interstitials around. All the defects turn outto be unstable: the vacancies disappear in 1 —3 time units;interstitials have a longer lifetime. We conclude that thedefects vanish independently of each other. The easinessof point defect annealing and the limited cascade volumeare obvious reasons for the abnormally good radiationresistance of amorphous solids. The number of defectscreated increases with the PKA energy, but the time in-stant when the defect number is at maximum is not ap-preciably affected by it. The displacement threshold en-

ergy as a function of ejection angle is determined and isfound to show considerable variation. The number ofdisplacements computed from the cascades is seen to in-crease linearly with the PKA energy and with an averagedisplacement threshold energy reasonable compatibilitywith the modified Kinchin-Pease model is obtained. Thesample volume and pair distribution function show nosystematic changes before and after a collision cascade.

In this work a one-component system with the simpleLennard-Jones pair potential has been studied and noelectronic losses are included. These are obviously notthe conditions prevailing in a real amorphous metal.However, in addition to increasing the knowledge of theproperties of amorphous Lennard-Jones systems we ex-pect these studies to reveal many of the general proper-ties of amorphous solids.

ACKNOWLEDGMENTS

We are grateful to Dr. M. Manninen for useful discus-sions and Mr. T. Siili for his extensive participation increating the computer programs for graphical presenta-tion of collision-cascade data.

Page 21: Aalto UniversityPHYSICAL REVIE%' B VOLUME 41, NUMBER 7 1 MARCH 1990 Computer simulations of radiation damage in amorphous solids J.Laakkonen' and R. M. Nieminen Laboratory ofPhysics,

. . . RADIATION DAMAGE IN AMORPHOUS SOLIDS

'Present address: Okmetic Ltd. , P.O. Box 44, SF-02631 Espoo,Finland.

G. H. Kinchin and R. S. Pease, Rep. Prog. Phys. 18, 1 (1955).~J. R. Beeler, Radiation Effects Computer Experiments (North-

Holland, Amsterdam, 1983).H. L. Heinisch, J. Nucl. Mater. 117,46 (1983).

4T. Muroga and S. Ishino, J. Qucl. Mater. 117, 36 (1983).~J. B. Gibson, A. N. Goland, M. Milgram, and G. H. Vineyard,

Phys. Rev. 120, 1229 (1960); see also C. Erginsoy, G. H. Vine-yard, and A. Englert, ibid. 133, A595 (1964).

J. Yli-Kauppila, Ph.D. thesis, Helsinki University of Technolo-

gy, 1983.7A. Audouard, J. Balogh, J. Dural, and J. C. Jousset, J. Non-

Cryst. Solids 50, 71 (1982).S. Cui, R. E. Johnson, and P. Cummings, Surface Sci. 207, 186

(1988).J. Laakkonen and R. M. Nieminen, J. Non-Cryst. Solids 75,

237 (1985).'oJ. Laakkonen and R. M. Nieminen, J. Phys. C 21, 3663 (1988).

D. O. Welch, G. J. Dienes, and A. Paskin, J. Phys. Chem.Solids 39, 589 (1978).D. Beeman, J. Comput. Phys. 20, 130 (1976).W. E. King and R. Benedek, J. Nucl. Mater. 117, 26 (1983).J. Lindhard, M. Scharff, and H. E. Schi/tt, Mat. Fys. Medd.Dan. Vid. Selsk. 33, No. 14 (1964).

'5J. Lindhard, V. Nielsen, M. Scharff, and P. V. Thomsen, Mat.Fys. Medd. Dan. Vid. Selsk. 33, No. 10 (1963).

' M. T. Robinson, Philos. Mag. 12, 741 (1965).' M. T. Robinson, Philos. Mag. 17, 639 (1968).

M. T. Robinson and O. S. Oen, J. Nucl. Mater. 110, 147(1982).

P. Sigmund, Radiat. Eff. 1, 15 (1969).M. J. Norgett, M. T. Robinson, and I. M. Torrens, Nucl. Eng.Des. 33, 50 (1975).M. W. Guinan and J. H. Kinney, J. Nucl. Mater. 1034104,1319(1981).

W. E. King, K. L. Merkle, and M. Meshii, Phys. Rev. B 23,6319 (1981).

23K. L. Merkle, W. E. King, A. C. Baily, K. Haga, and M.Meshii, J. Nucl. Mater. 117,4 (1983).F. Ercolessi, E. Tosatti, and M. Parrinello, Phys. Rev. Lett.57, 719 (1986).S. M. Foiles, M. I. Baskes, and M. S. Daw, Phys. Rev. B 33,7983 (1986).K. W. Jacobsen, J. K. Ngfrskov, and M. J. Puska, Phys. Rev.B 35, 7423 (1987).R. N. Barnett, C. L. Cleveland, and U. Landman, Phys. Rev.Lett. 55, 2035 (1985).R. Taylor, J. Phys. (Paris) Colloq. 46, C4-309 (1985).S. W. de Leeuw, M. Dixon, and R. J. Elliott, Philos. Mag. A46, 677 (1982).A. Audouard, J. Balogh, J. Dural, and J. C. Jousset, Radiat.Eff. 62, 161 (1982).N. W. Ashcroft and N. D. Mermin, Solid State Physics (Holt,Rinehart and Winston, New York, 1976).

32W. A. Harrison, Electronic Structure and the Properties ofSolids (Freeman, San Francisco, 1980).R. Yamamoto, H. Shibuta, and M. Doyama, J. Nucl. Mater.85486, 603 (1979).

34T. K. Chaki and J. C. M. Li, Philos. Mag. B 51, 557 (1985).B.T.-A. Chang and J. C. M. Li, Scripta Metall. 11,933 (1977).


Recommended