23
HAL Id: hal-01703022 https://hal.archives-ouvertes.fr/hal-01703022 Submitted on 9 Feb 2018 HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers. L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés. Elastic modeling of point-defects and their interaction Emmanuel Clouet, Céline Varvenne, Thomas Jourdan To cite this version: Emmanuel Clouet, Céline Varvenne, Thomas Jourdan. Elastic modeling of point-defects and their interaction. Computational Materials Science, Elsevier, 2018, 147, pp.49 - 63. 10.1016/j.commatsci.2018.01.053. hal-01703022

Elastic modeling of point-defects and their interaction

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Elastic modeling of point-defects and their interaction

HAL Id: hal-01703022https://hal.archives-ouvertes.fr/hal-01703022

Submitted on 9 Feb 2018

HAL is a multi-disciplinary open accessarchive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come fromteaching and research institutions in France orabroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, estdestinée au dépôt et à la diffusion de documentsscientifiques de niveau recherche, publiés ou non,émanant des établissements d’enseignement et derecherche français ou étrangers, des laboratoirespublics ou privés.

Elastic modeling of point-defects and their interactionEmmanuel Clouet, Céline Varvenne, Thomas Jourdan

To cite this version:Emmanuel Clouet, Céline Varvenne, Thomas Jourdan. Elastic modeling of point-defectsand their interaction. Computational Materials Science, Elsevier, 2018, 147, pp.49 - 63.10.1016/j.commatsci.2018.01.053. hal-01703022

Page 2: Elastic modeling of point-defects and their interaction

Elastic modeling of point-defects and their interaction

Emmanuel Cloueta,∗, Celine Varvenneb, Thomas Jourdana

aDEN-Service de Recherches de Metallurgie Physique, CEA,Paris-Saclay Univ., F-91191 Gif-sur-Yvette, FrancebCentre Interdisciplinaire des Nanosciences de Marseille, UMR 7325 CNRS - Aix Marseille Univ., F-13008 Marseille,

France

Abstract

Different descriptions used to model a point-defect in an elastic continuum are reviewed. The emphasisis put on the elastic dipole approximation, which is shown to be equivalent to the infinitesimal Eshelbyinclusion and to the infinitesimal dislocation loop. Knowing this elastic dipole, a second rank tensor fullycharacterizing the point-defect, one can directly obtain the long-range elastic field induced by the point-defectand its interaction with other elastic fields. The polarizability of the point-defect, resulting from the elasticdipole dependence with the applied strain, is also introduced. Parameterization of such an elastic model,either from experiments or from atomic simulations, is discussed. Different examples, like elastodiffusionand bias calculations, are finally considered to illustrate the usefulness of such an elastic model to describethe evolution of a point-defect in a external elastic field.

Keywords: Point-defects, Elasticity, Elastic dipole, Polarizability

1. Introduction

Point-defects in crystalline solids, being eitherintrinsic like vacancies, self-interstitial atoms, andtheir small clusters, or extrinsic like impurities anddopants, play a major role in materials propertiesand their kinetic evolution. Some properties ofthese point-defects, like their formation and migra-tion energies, are mainly determined by the regionin the immediate vicinity of the defect where thecrystal structure is strongly perturbed. An atomicdescription appears thus natural to model theseproperties, and atomic simulations relying eitheron ab initio calculations [1] or empirical potentialshave now become a routine tool to study point-defects structures and energies. But point-defectsalso induce a long-range perturbation of the hostlattice, leading to an elastic interaction with otherstructural defects, impurities or an applied elasticfield. An atomic description thus appears unnec-essary to capture the interaction arising from thislong-range part, and sometimes is also impossiblebecause of the reduced size of the simulation cell in

∗Corresponding authorEmail address: [email protected] (Emmanuel

Clouet)

atomic approaches. Elasticity theory becomes thenthe natural framework. It allows a quantitative de-scription of the point-defect interaction with otherdefects.

Following the seminal work of Eshelby [2], thesimplest elastic model of a point-defect correspondsto a spherical inclusion forced into a sphericalhole of slightly different size in an infinite elasticmedium. This description accounts for the point-defect relaxation volume and its interaction with apressure field (size interaction). It can be enrichedby considering an ellipsoidal inclusion, thus leadingto a interaction with also the deviatoric componentof the stress field (shape interaction), and by as-signing different elastic constants to the inclusion(inhomogeneity) to describe the variations of thepoint-defect “size” and “shape” with the strain fieldwhere it is immersed. Other elastic descriptions ofthe point-defect are possible. In particular, it canbe modeled by an equivalent distribution of point-forces. The long-range elastic field of the point-defect and its interaction with other stress sourcesare then fully characterized by the first moment ofthis force distribution, a second-rank tensor calledthe elastic dipole. This description is rather natu-ral when modeling point-defects and it can be used

Preprint submitted to Computational Materials Science February 9, 2018

Page 3: Elastic modeling of point-defects and their interaction

to extract elastic dipoles from atomic simulations.These different descriptions are equivalent in thelong-range limit, and allow for a quantitative mod-eling of the elastic field induced by the point-defect,as long as the elastic anisotropy of the matrix isconsidered.

This article reviews these different elastic modelswhich can be used to describe a point-defect and il-lustrates their usefulness with chosen examples. Af-ter a short reminder of elasticity theory (Sec. 2), weintroduce the different descriptions of a point-defectwithin elasticity theory (Sec. 3), favoring the elas-tic dipole description and showing its equivalencewith the infinitesimal Eshelby inclusion as well aswith an infinitesimal dislocation loop. The nextsection (Sec. 4) describes how the characteristics ofthe point-defect needed to model it within elasticitytheory can be obtained either from atomistic simu-lations or from experiments. We finally give someapplications in Sec. 5, where results of such an elas-tic model are compared to direct atomic simulationsto assess its validity. The usefulness of this elasticdescription is illustrated in this section for elastod-iffusion and for the calculation of bias factors, aswell as for the modeling of isolated point-defects inatomistic simulations.

2. Elasticity theory

Before describing the modeling of a point-defectwithin elasticity theory, it is worth recalling themain aspects of the theory [3], in particular the un-derlying assumptions, some definitions and usefulresults.

2.1. Displacement, distortion and strain

Elasticity theory is based on a continuous de-scription of solid bodies. It relates the forces, ei-ther internal or external, exerting on the solid toits deformation. To do so, one first defines theelastic displacement field. If ~R and ~r are the po-sition of a point respectively in the unstrained andthe strained body, the displacement at this point isgiven by

~u(~R) = ~r − ~R.

One can then define the distortion tensor ∂ui / ∂Rjwhich expresses how an infinitesimal vector

# »

dR inthe unstrained solid is transformed in

# »

dr in thestrained body through the relation

dri =

(δij +

∂ui∂Rj

)dRj ,

where summation over repeated indices is implicit(Einstein convention) and δij is the Kronecker sym-bol.

Of central importance to the elasticity theory isthe dimensionless strain tensor, defined by

εij(~R) =1

2

[(δin +

∂un∂Ri

)(δnj +

∂un∂Rj

)− δij

]=

1

2

(∂ui∂Rj

+∂uj∂Ri

+∂un∂Ri

∂un∂Rj

).

This symmetric tensor expresses the change of sizeand shape of a body as a result of a force acting onit. The length dL of the infinitesimal vector

# »

dR inthe unstrained body is thus transformed into dl inthe strained body, through the relation

dl2 = dL2 + 2εijdRidRj .

Assuming small deformation, a common assump-tion of linear elasticity, only the leading terms ofthe distortion are kept. The strain tensor then cor-responds to the symmetric part of the distortiontensor, as

εij(~R) =1

2

(∂ui∂Rj

+∂uj∂Ri

). (1)

The antisymmetric part of the distortion tensor cor-responds to the infinitesimal rigid body rotation. Itdoes not lead to any energetic contribution withinlinear elasticity in the absence of internal torque.

With this small deformation assumption, thereis no distinction between Lagrangian coordinates ~Rand Eulerian coordinates ~r when describing elasticfields. One can equally write, for instance, ~u(~r) or

~u(~R) for the displacement field, which are equiva-lent to the leading order of the distortion.

2.2. Stress

The force# »

δF acting on a volume element δV ofa strained body is composed of two contributions,the sum of external body forces ~f and the internalforces arising from atomic interactions. Because ofthe mutual cancellation of forces between particlesinside the volume δV , only forces corresponding tothe interaction with outside particles appear in thislast contribution, which is thus proportional to thesurface elements

# »

dS defining the volume elementδV . One obtains

δFi =

∫δV

fidV +

∮δS

σijdSj ,

2

Page 4: Elastic modeling of point-defects and their interaction

where σ is the stress tensor defining internal forces.Considering the mechanical equilibrium of the

volume element δV , the absence of resultant forceleads to the equation

∂σij(~r)

∂rj+ fi(~r) = 0, (2)

whereas the absence of torque ensures the symme-try of the stress tensor.

At the boundary of the strained body, internalforces are balanced by applied forces. If ~T adS is theforce applied on the infinitesimal surface elementdS, this leads to the boundary condition

σijnj = T ai , (3)

where ~n is the outward-pointing normal to the sur-face element dS.

The work δw, defined per volume unit, of theseinternal forces is given by

δw = −σijδεij ,

where δεij is the strain change during the deforma-tion increase, and the sign convention is δw > 0when the energy flux goes outwards the elasticbody. This leads to the following thermodynamicdefinition of the stress tensor

σij =

(∂e

∂εij

)s

=

(∂f

∂εij

)T

,

where e, s, and f = e− Ts are the internal energy,entropy, and free energy of the elastic body definedper volume unit.

2.3. Hooke’s law

To go further, one needs a constitutive equationfor the energy or the free energy. Taking as a ref-erence the undeformed state corresponding to theelastic body at equilibrium without any externalforce, either body or applied stress, the energy is ata minimum for ε = 0 and then

σij(ε = 0) =∂e

∂εij

∣∣∣∣ε=0

= 0.

The leading order terms of the series expansion ofthe energy are then

e(T, ε) = e0(T ) +1

2Cijklεijεkl,

where e0(T ) = e(T, ε = 0) is the energy of the un-strained body at temperature T . The elastic con-stants Cijkl entering this expression are thus de-fined by

Cijkl =∂2e

∂εij∂εkl.

This is a fourth-rank tensor which obeys minorsymmetry Cijkl = Cjikl = Cijlk because of thestrain tensor symmetry and also major symmetryCijkl = Cklij because of allowed permutation ofpartial derivatives. This leads to at most 21 inde-pendent coefficients, which can be further reducedby considering the symmetries of the solid body [4].

This series expansion of the energy leads to alinear relation, the Hooke’s law, between the stressand the strain

σij = Cijklεkl, (4)

which was summarized in 1678 by Robert Hooke asUt tensio, sic vis.1

2.4. Elastic equilibrium, superposition principle

Combining Hooke’s law (4) with the small defor-mation definition (1) of the strain tensor and theequilibrium condition (2), one obtains the equationobeyed by the displacement at equilibrium

Cijkl∂2uk(~r)

∂rj∂rl+ fi(~r) = 0. (5)

The elastic equilibrium is given by the solutionwhich verifies the boundary conditions, σijnj = T a

i

for imposed applied forces and ui = uai for imposed

applied displacements.

As elastic equilibrium is defined by the solutionof a linear partial differential equation (Eq. 5),the superposition principle holds. If two elasticfields, characterized by their displacement ~u1(~r)and ~u2(~r), correspond to equilibrium for the re-

spective body forces ~f1 and ~f2 and the respectiveboundary conditions (~ua1, ~T a1) and (~ua2, ~T a2), then

the elastic equilibrium for the body forces ~f1 + ~f2

and the boundary conditions (~ua1 +~ua2, ~T a1 + ~T a2)is given by the sum of these two elastic fields. Thetotal elastic energy is composed of the contribu-tions of each elastic field taken separately and an

1As the extension, so the force.

3

Page 5: Elastic modeling of point-defects and their interaction

interaction energy given by

Eint =

∫V

σ1ij(~r) ε

2ij(~r) dV

=

∫V

σ2ij(~r) ε

1ij(~r) dV .

(6)

This equation can be used to define interaction en-ergy between two defects.

The superposition principle allows making useof Green’s function. The elastic Green’s functionGkn(~r) is the solution of the equilibrium equationfor a unit point-force

Cijkl∂2Gkn(~r)

∂rj∂rl+ δin δ(~r) = 0, (7)

where δ(~r) is the Dirac delta function, i.e. δ(~r) = 0if ~r 6= ~0 and δ(~0) = ∞. Gkn(~r) therefore corre-sponds to the displacement along the rk axis fora unit point-force applied along the rn axis at theorigin. The solution of elastic equilibrium for theforce distribution ~f(~r) is then given by

uk(~r) =

∫V

Gkn(~r − ~r ′)fn(~r ′)dV ′,

σij(~r) = Cijkl

∫V

Gkn,l(~r − ~r ′)fn(~r ′)dV ′,

where we have introduced the notation Gkn,l =∂Gkn / ∂rl for partial derivatives.

An analytical expression of the Green’s functionexists for isotropic elasticity. Considering the elas-tic constants Cijkl = λδijδkl + µ(δikδjl + δilδjk),where λ and µ are the Lame coefficients, the Green’sfunction is given by

Gkn(~r) =1

8πµ

[λ+ 3µ

λ+ 2µδkn +

λ+ µ

λ+ 2µηkηn

]1

r,

with r = ‖~r‖ and ~η = ~r/r. No analytical expressionexists in the more general case of elastic anisotropy,but the Green’s function, and its successive deriva-tives, can be calculated efficiently from the elasticconstants using the numerical scheme of Barnett[5, 6]. Whatever the anisotropy, the Green’s func-tion and its derivatives will show the same variationwith the distance r,2 leading to the general expres-sions

Gkn(~r) = gkn(~η)1

r, Gkn,l(~r) = hknl(~η)

1

r2, . . .

2The scaling with the distance r is a consequence of Eq.(7), given that the δ(~r) function is homogeneous of degree−3.

where the anisotropy enters only in the angular de-pendence gkn(~η), hknl(~η), . . .

3. Elastic model of a point-defect

Different models can be used to describe a point-defect within elasticity theory. One such model isthe elastic dipole. We first describe this model andthen demonstrate the analogy with a description ofthe point-defect as an infinitesimal Eshelby inclu-sion or an infinitesimal dislocation loop. We finallyintroduce the polarizability of the point-defect.

3.1. Elastic dipole

A point-defect can be described in a continu-ous solid body as an equilibrated distribution ofpoint-forces [6–9]. Considering a point-defect lo-cated at the origin modeled by such a force distri-bution ~f(~r) =

∑Nq=1

~F q δ (~r − ~aq), i.e. consisting of

N forces ~F q each acting at position ~aq, the elasticdisplacement field of the point-defect is, accordingto linear elasticity theory, given by

ui(~r) =

N∑q=1

Gij(~r − ~aq)F qj ,

where we have used the elastic Green’s function.Far from the point-defect, we have ‖~r‖ ‖~aq‖ andwe can make a series expansion of the Green’s func-tion:

ui(~r) = Gij(~r)

N∑q=1

F qj − Gij,k(~r)

N∑q=1

F qj aqk

+ O(‖~aq‖2

).

As the force distribution is equilibrated, its resul-tant

∑q~F q is null. The displacement is thus given,

to the leading order, by

ui(~r) = −Gij,k(~r)Pjk, (8)

and the corresponding stress field by

σij(~r) = −CijklGkm,nl(~r)Pmn, (9)

where the elastic dipole is defined as the first mo-ment of the point-force distribution,

Pjk =

N∑q=1

F qj aqk. (10)

4

Page 6: Elastic modeling of point-defects and their interaction

This dipole is a second rank tensor which fullycharacterizes the point-defect within elasticity the-ory [6–9]. It is symmetric because the torque∑q~F q × ~aq must be null for the force distribution

to be equilibrated.Equations (8) and (9) show that the elastic dis-

placement and the stress created by a point-defectare long-ranged, respectively decaying as 1/r2 and1/r3 with the distance r to the point-defect.

The elastic dipole is directly linked to the point-defect relaxation volume. Considering a finite vol-ume V of external surface S enclosing the point-defect, this relaxation volume is defined as

∆V =

∮S

ui(~r) dSi,

where ~u(~r) is the superposition of the displacementcreated by the point-defect (Eq. 8) and the elasticdisplacement due to image forces ensuring null trac-tions on the external surface S. Use of the Gausstheorem, of the equilibrium condition (5) and of theelastic dipole definition (10) leads to the result [8]

∆V = SiiklPkl, (11)

where the elastic compliances Sijkl are the in-verse of the elastic constants, i.e. SijklCklmn =12 (δimδjn + δinδjm). For a crystal with cubic sym-metry, this equation can be further simplified [8]to show that the relaxation volume is equal to thetrace of the elastic dipole divided by three times thebulk modulus. More generally, as it will becomeclear with the comparison to the Eshelby’s inclu-sion, this elastic dipole is the source term definingthe relaxation volume of the point-defect. Its tracegives rise to the size interaction, whereas its de-viator, i.e. the presence of off-diagonal terms anddifferences in the diagonal components, leads to theshape interaction.

Of particular importance is the interaction en-ergy of the point-defect with an external elastic field~uext(~r). Considering the point-forces distributionrepresentative of the point-defect, this interactionenergy can be simply written as [6]

Eint = −N∑q=1

F qi uexti (~aq).

If we now assume that the external field is slowlyvarying close to the point-defect, one can make aseries expansion of the corresponding displacement

~uext(~r). The interaction energy is then, to first or-der,

Eint = −uexti (~0)

N∑q=1

F qi − uexti,j (~0)

N∑q=1

F qi aqj .

Finally, using the equilibrium properties of thepoint-forces distribution, one obtains

Eint = −Pij εextij (~0), (12)

thus showing that the interaction energy is simplythe product of the elastic dipole with the value atthe point-defect location of the external strain field.Higher order contributions to the interaction energyinvolve successive gradients of the external strainfield coupled with higher moments of the multipoleexpansion of the force distribution, and can be gen-erally safely ignored. This simple expression of theinteraction energy is the workhorse of the modelingof point-defects within linear elasticity in a multi-scale approach.

Instead of working with the elastic dipole tensor,one sometimes rather uses the so-called λ-tensor[10] which expresses the strain variation of a matrixvolume with the point-defect volume concentrationc,

λij =1

Ωat

∂εij∂c

, (13)

where ε is the homogeneous strain induced by thepoint-defects in a stress-free state and Ωat is theatomic volume of the reference solid. As it willbecome clear when discussing parameterization ofthe elastic dipole from experiments (§4.2), these twoquantities are simply linked by the relation

Pij = Ωat Cijkl λkl. (14)

Using this λ-tensor to characterize the point-defect,Eq. (12) describing its elastic interaction with anexternal elastic field becomes

Eint = −Ωat λij σextij (~0),

where σextij (~0) is the value of the external stress field

at the point-defect position.

3.2. Analogy with Eshelby’s inclusion

The Eshelby’s inclusion [11, 12] is anotherwidespread model which can be used to describea point-defect in an elastic continuum. As it willbe shown below, it is equivalent to the dipole de-scription in the limit of an infinitesimal inclusion.

5

Page 7: Elastic modeling of point-defects and their interaction

In this model, the point-defect is described as aninclusion of volume ΩI and of surface SI, having thesame elastic constants as the matrix. This inclu-sion undergoes a change of shape described by theeigenstrain ε∗ij(~r), corresponding to the strain thatwould adopt the inclusion if it was free to relax andwas not constrained by the surrounding matrix. Es-helby proposed a general approach [11] to solve thecorresponding equilibrium problem and determinethe elastic fields in the inclusion and the surround-ing matrix. This solution is obtained by consideringthe three following steps:

1. Take the inclusion out of the matrix and let itadopt its eigenstrain ε∗ij(~r). At this stage, thestress is null everywhere.

2. Strain back the inclusion so it will fit the holein the matrix. The elastic strain exactly com-pensates for the eigenstrain, so the stress inthe inclusion is −Cijklε∗kl(~r). This operation isperformed by applying to the external surfaceof the inclusion the traction forces correspond-ing to this stress

dTi(~r) = −Cijkl ε∗kl(~r) dSj ,

where# »

dS is an element of the inclusion exter-nal surface at the point ~r.

3. After the inclusion has been welded back intoits hole, the traction forces are relaxed. UsingGreen’s function, the corresponding displace-ment in the matrix is then

un(~r) =

∮SI

Gni(~r − ~r ′) dTi(~r′),

= −∮SI

Gni(~r − ~r ′)Cijkl ε∗kl(~r ′) dS ′j .

Applying Gauss theorem and the equilibrium condi-tion satisfied by the eigenstrain ε∗ij(~r), one obtainsthe following expression for the elastic displacementin the matrix

un(~r) = −∫

ΩI

Gni,j(~r − ~r ′)Cijkl ε∗kl(~r ′) dV ′,

(15)and for the corresponding stress field

σpq(~r) = −∫

ΩI

CpqmnGni,jm(~r − ~r ′)

Cijkl ε∗kl(~r

′) dV ′. (16)

Inside the inclusion, one needs to add the stress−Cijklε∗kl(~r) corresponding to the strain applied instep 2.

Far from the inclusion, we have ‖~r‖ ‖~r ′‖. Wecan therefore neglect the variations of the Green’sfunction derivatives inside Eqs. 15 and 16. Thiscorresponds to the infinitesimal inclusion assump-tion. For such an infinitesimal inclusion located atthe origin, one therefore obtains the following elas-tic fields

un(~r) = −Gni,j(~r)Cijkl ΩI ε∗kl, (17)

σpq(~r) = −CpqmnGni,jm(~r)Cijkl ΩI ε∗kl, (18)

where we have defined the volume average of theinclusion eigenstrain, ε∗ij = 1

ΩI

∫ΩIεij(~r) dV . Com-

paring these expressions with the ones describingthe elastic field of an elastic dipole (Eqs. 8 and9), we see that they are the same for any ~r valueprovided the dipole tensor and the inclusion eigen-strain check the relation

Pij = ΩI Cijkl ε∗kl. (19)

The descriptions of a point-defect as an elasticdipole, i.e. as a distribution of point-forces keep-ing only the first moment of the distribution, or asan infinitesimal Eshelby inclusion, i.e. in the limitof an inclusion volume ΩI → 0 keeping the prod-uct ΩI ε

∗ij constant, are therefore equivalent. The

point-defect can be thus characterized either by itselastic dipole tensor Pij or by its eigenstrain tensorQij = ΩI ε

∗ij [13].

Of course, the same equivalence is obtained whenconsidering the interaction energy with an externalstress field. For a general inclusion, Eshelby showedthat this interaction energy is simply given by

Eint = −∫

ΩI

ε∗ij(~r)σextij (~r) dV , (20)

where the integral only runs on the inclusion vol-ume. In the limiting case of an infinitesimal inclu-sion, one can neglect the variations of the externalstress field inside the inclusion. One thus obtainsthe following interaction energy,

Eint = −ΩI ε∗ij σ

extij (~0), (21)

which is equivalent to the expression (12) for anelastic dipole when the equivalence relation (19) isverified.

3.3. Analogy with dislocation loops

A point-defect can also be considered as an in-finitesimal dislocation loop. This appears natu-ral as dislocation loops are known to be elasticallyequivalent to platelet Eshelby’s inclusions [14, 15].

6

Page 8: Elastic modeling of point-defects and their interaction

The elastic displacement and stress fields of a dis-location loop of Burgers vector ~b are respectivelygiven by the Burger’s and the Mura’s formulae [16]

ui(~r) = Cjklm bm∫A

Gij,k(~r − ~r ′)nl(~r ′) dA ′,(22)

σij(~r) = Cijkl εlnhCpqmnbm∮L

Gkp,q(~r − ~r ′) ζh(~r ′) dl ′.(23)

The displacement is defined by a surface integralon the surface A enclosed by the dislocation loop,with ~n(~r ′) the local normal to the surface elementdA ′ in ~r ′, and the stress by a line integral alongthe loop of total line length L. ~ζ is the unit vectoralong the loop, and εlnh is the permutation tensor.

Like for the Eshelby’s inclusion, far from the loop(‖~r‖ ‖~r ′‖), we can use a series expansion of theGreen’s function derivatives and keep only the lead-ing term. Considering a loop located at the origin,we thus obtain

ui(~r) =Cjklm bmAlGij,k(~r), (24)

σpq(~r) =Cpqin Cjklm bmAlGij,kn(~r), (25)

where ~A is the surface vector defining the area ofthe loop. These expressions are equal to the onesobtained for an elastic dipole (8) and (9), withthe equivalent dipole tensor of the dislocation loopgiven by

Pjk = −Cjklm bmAl. (26)

Looking at the interaction with an external stressfield, the interaction energy with the dislocationloop is given by

Eint =

∫A

σextij (~r) bi nj dA. (27)

For an infinitesimal loop, it simply becomes

Eint = σextij (~0) biAj , (28)

which is equivalent to the expression (12) obtainedfor an elastic dipole when the equivalent dipole ten-sor of the dislocation loop is given by Eq. (26).

3.4. Polarizability

The equivalent point-forces distribution of apoint-defect can be altered by an applied elasticfield [17]. This applied elastic field thus leads to aninduced elastic dipole and the total elastic dipole of

the point-defect now depends on the applied strainεext:

Pij(εext) = P 0

ij + αijklεextkl , (29)

where P 0ij is the permanent elastic dipole in absence

of applied strain and αijkl is the point-defect diae-lastic polarizability [18–20]. Considering the anal-ogy with the Eshelby’s inclusion, this polarizabilitycorresponds to an infinitesimal inhomogeneous in-clusion, i.e. an inclusion with different elastic con-stants than the surrounding matrix. It describesthe fact that the matrix close to the point-defecthas a different elastic response to an applied strainbecause of the perturbations of the atomic bondingcaused by the point-defect. For the analogy withan infinitesimal dislocation loop, the polarizabilitycorresponds to the fact that the loop can change itsshape by glide on its prismatic cylinder (or in itshabit plane for a pure glide loop) under the actionof the applied elastic field.

Following Schober [18], the interaction of a point-defect located at the origin with an applied strainis now given by

Eint = −P 0ij ε

extij (~0)− 1

2αijkl ε

extij (~0) εext

kl (~0). (30)

This expression of the interaction energy, whichincludes the defect polarizability, has importantconsequences for the modeling of point-defects asit shows that some coupling is possible betweentwo different applied elastic fields. Considering thepoint-defect interaction with the two strain fieldsε(1) and ε(2) originating from two different sources,the interaction energy is now given by

Eint = − P 0ij

(1)ij + ε

(2)ij

)− 1

2αijkl

(1)ij + ε

(2)ij

)(ε

(1)kl + ε

(2)kl

),

= − P 0ijε

(1)ij −

1

2αijkl ε

(1)ij ε

(1)kl

− P 0ijε

(2)ij −

1

2αijkl ε

(2)ij ε

(2)kl

− αijkl ε(1)ij ε

(2)kl .

The last line therefore shows that, without thepolarizability, the interaction energy of the point-defect with the two strain fields will be simply thesuperposition of the two interaction energies witheach strain fields considered separately. A couplingis introduced only through the polarizability. Sucha coupling is for instance at the origin of one of the

7

Page 9: Elastic modeling of point-defects and their interaction

mechanisms proposed to explain creep under irra-diation. Indeed, because of the polarizability, theinteraction of point-defects, either vacancies or self-interstitial atoms, with dislocations under an ap-plied stress depends on the dislocation orientationwith respect to the applied stress. This stronger in-teraction with some dislocation families leads to alarger drift term in the diffusion equation of thepoint-defect and thus to a greater absorption ofthe point-defect by these dislocations, a mechanismknown as Stress Induced Preferential Absorption(or SIPA) [21–24]. This polarizability is also thecause, in alloy solid solutions, of the variation ofthe matrix elastic constants with their solute con-tent.

This diaelastic polarizability caused by the per-turbation of the elastic response of the surroundingmatrix manifests itself at the lowest temperature,even 0 K, and whatever the characteristic time ofthe applied strain. At finite temperature there maybe another source of polarizability. If the point-defect can adopt different configurations, for in-stance different variants corresponding to differentorientations of the point-defect like for a carbon in-terstitial atom in a body-centered cubic Fe matrix,then the occupancy distribution of these configu-rations will be modified under an applied stress orstrain. This possible redistribution of the point-defect gives rise to anelasticity [10], the most fa-mous case being the Snoek relaxation in iron alloyscontaining interstitial solute atoms like C and N[25]. When thermally activated transitions betweenthe different configurations of the point-defect arefast enough compared to the characteristic time ofthe applied stress, the distribution of the differentconfigurations corresponds to thermal equilibrium.Assuming that all configurations have the same en-ergy in a stress-free state and denoting by Pµij theelastic dipole of the configuration µ, the averagedipole of the point-defect is then given by

〈Pij〉 =

∑µ exp (Pµklε

extkl / kT )Pµij∑

µ exp (Pµklεextkl / kT )

.

As a consequence, the average elastic dipole of thepoint-defect distribution is now depending on theapplied stress and on the temperature, an effectknown as paraelasticity [17]. At temperatures highenough to allow for transition between the differentconfigurations, the interaction energy of the con-figurations with the applied strain is usually smallcompared to kT . One can make a series expansion

of the exponentials to obtain

〈Pij〉 =1

nv

nv∑µ=1

Pµij

−(

1

nv2

nv∑µ,ν=1

PµijPνkl −

1

nv

nv∑µ=1

PµijPµkl

)εextkl

kT,

where nv is the number of configurations. Thisleads to the same linear variation of the elasticdipole with the applied strain as for the diaelasticpolarizability (Eq. 29), except that the paraelasticpolarizability is depending on the temperature.

4. Parameterization of elastic dipoles

To properly model a point-defect with continuumelasticity theory, one only needs to know its elasticdipole. It is then possible to describe the elasticdisplacement (Eq. 8) or the stress field (Eq. 9)induced by the point-defect, and also to calculateits interaction with an external elastic field (Eq.12). This elastic dipole can be determined eitherusing atomistic simulations or from experiments.

4.1. From atomistic simulations

Different strategies can be considered for theidentification of elastic dipoles in atomistic simu-lations. This elastic dipole can be directly deducedfrom the stress existing in the simulation box, orfrom a fit of the atomic displacements, or finallyfrom a summation of the Kanzaki forces. We ex-amine here these three techniques and discuss theirmerits and drawbacks.

Definition from the stress

Let us consider a simulation box of volume V ,the equilibrium volume of the pristine bulk mate-rial. We introduce one point-defect in the simula-tion box and assume periodic boundary conditionsto preclude any difficulty associated with surfaces.Elasticity theory can be used to predict the vari-ation of the energy of the simulation box submit-ted to a homogeneous strain ε. Using the interac-tion energy of a point-defect with an external straingiven in Eq. (12), one obtains

E(ε) = E0 + EPD +V

2Cijklεijεkl − Pijεij , (31)

with E0 the bulk reference energy and EPD thepoint-defect energy, which can contain a contribu-tion from the interactions of the point-defect with

8

Page 10: Elastic modeling of point-defects and their interaction

its periodic images (see section 5.4). The averageresidual stress on the simulation box is obtained bysimple derivation as3

〈σij(ε)〉 =1

V

∂E

∂εij,

= Cijklεkl −1

VPij .

(32)

In the particular case where the periodicity vectorsare kept fixed between the defective and pristinesupercells (ε = 0), the elastic dipole is proportionalto the residual stress weighted by the supercell vol-ume:

Pij = −V 〈σij〉. (33)

This residual stress corresponds to the stress in-crease, after atomic relaxation, due to the intro-duction of the point-defect into the simulation box.When this equation is used to determine the elas-tic dipole in ab initio calculations, one should payattention to the spurious stress which may exist inthe equilibrium perfect supercell because of finiteconvergence criteria of such calculations. This spu-rious stress has to be subtracted from the stress ofthe defective supercell, so the residual stress enter-ing Eq. 33 is only the stress increment associatedwith the introduction of the point-defect.

One can also consider the opposite situationwhere a homogeneous strain ε has been applied tocancel the residual stress. The elastic dipole is thenproportional to this homogeneous strain:

Pij = V Cijklεkl. (34)

One would nevertheless generally prefer workingwith fixed periodicity vectors (ε = 0) as σ = 0 cal-culations necessitate an increased number of forcecalculations, as well as an increased precision for abinitio calculations. In the more general case where ahomogeneous strain is applied and a residual stressis observed, the elastic dipole can still be derivedfrom these two quantities using Eq. (32).

This definition of the elastic dipole from theresidual stress (Eq. 33), or more generally fromboth the applied strain and the residual stress (Eq.32), is to be related to the dipole tensor measure-ment first proposed by Gillan [28, 29], where theelastic dipole is equal to the strain derivative of theformation energy, evaluated at zero strain. Insteadof doing this derivative numerically, one can simply

3See also Refs. [26] and [27] for other proofs.

use the analytical derivative, i.e. the stress on thesimulation box, which is a standard output of anyatomistic simulations code, including ab initio cal-culations. This technique to extract elastic dipolesfrom atomistic simulations has been validated [30–32], through successful comparisons of interactionenergies between point-defects with external strainfields, as given by direct atomistic simulations andas given by the elasticity theory predictions usingthe elastic dipole identified through Eq. (33). Theresidual stress therefore leads to quantitative esti-mates of the elastic dipoles.

Definition from the displacement field

The elastic dipole can also be obtained from thedisplacement field, as proposed by Chen et al. [33].

Using the displacement field ~uat(~R) obtained afterrelaxation in atomistic simulations, a least-squarefit of the displacement field ~uel(~R) predicted byelasticity theory can be realized, using the dipolecomponents of the dipole as fit variables. A reason-able cost function for the least-square fit is

f(Pij) =∑~R

‖~R‖>rexcl

∥∥∥R2[~uel(~R)− ~uat(~R)

]∥∥∥2

, (35)

with rexcl the radius of a small zone around thepoint-defect, so as to exclude from the fit the atomicpositions where elasticity does not hold. The R2

factor accounts for the scaling of the displacementfield with the distance to the point-defect, thus giv-ing a similar weight to all atomic positions includedinto the fit. For atomistic simulations with periodicboundary conditions, one needs to superimpose theelastic displacements of the point-defect with its pe-riodic images, which can be done by simple summa-tion, taking care of the conditional convergence ofthe corresponding sum [32]. With large simulationboxes (≥ 1500 atoms), the obtained elastic dipolecomponents agree with the values deduced from theresidual stress, and the choice of rexcl is not critical.The number of atomic positions included in the fit,and for which elasticity is valid, is sufficiently highto avoid issues arising from the defect core zone [32].In contrast, for small simulation boxes of a few hun-dred atoms, i.e. typical of ab initio simulations, theobtained Pij values are highly sensitive to rexcl, andtheir convergence with rexcl cannot be guaranteed.This fit of the displacement field appears thereforeimpractical to obtain precise values of the elasticdipole in ab initio calculations.

9

Page 11: Elastic modeling of point-defects and their interaction

(a) unrelaxed vacancy (b) relaxed vacancy

(c) 0th order approx. (d) 1st order approx.

Figure 1: Procedure for the computation of the Kanzakiforces in the case of a vacancy. The white spheres corre-spond to atoms at their perfect bulk positions, i.e. beforerelaxation, the white square to the vacancy, and the blackspheres to the atoms at their relaxed position around thedefect.

Definition from the Kanzaki forces

The definition given in Eq. (10) of the elasticdipole as the first moment of the point-force dis-tribution offers a third way to extract this elasticdipole from atomic simulations. This correspondsto the Kanzaki force method [8, 34–41]. Kanzakiforces are defined as the forces which have to beapplied to the atoms in the neighborhood of thepoint-defect to produce the same displacement fieldin the pristine crystal as in the defective supercell.Computation of these Kanzaki forces can be per-formed following the procedure given in Ref. [39],which is illustrated for a vacancy in Fig. 1. Start-ing from the relaxed structure of the point-defect(Fig. 1b), the defect is restored in the simulationcell, e.g. the suppressed atom is added back for thevacancy case (Fig. 1c). A static force calculationis performed then and provides the opposite of thesearched forces on all atoms in the obtained simula-tion cell. These atomic forces are used to computethe elastic dipole Pij =

∑q F

qj a

qi , with ~F q the op-

posite of the force acting on atom at ~aq, assumingthe point-defect is located at the origin. The sum-mation is usually restricted to atoms located insidea sphere of radius r∞.

As Kanzaki’s technique is valid only in the har-monic approximation, one checks that the atomicforces entering the elastic dipole definition are inthe harmonic regime by restoring larger and larger

8

12

16

0 2 4 6

approximation 0

Pij (e

V)

(a)

r∞

/ a

0 2 4 6

approximation 16(b)

r∞

/ a

P11P33

Figure 2: Elastic dipole components of the SIA octahedralconfiguration in hcp Zr, as a function of the cutoff radius r∞of the force summation normalized by the lattice parametera. Values are obtained by the Kanzaki’s forces approach ona simulation box containing 12800 atoms, restoring (a) onlythe point-defect, and (b) up to 16 defect neighbor shells.The horizontal lines are the values deduced from the residualstress. Calculations have been performed with the EAM #3potential of Ref. [42] (see Ref. [32] for more details).

defect neighboring shells to their perfect bulk posi-tions [39] (Fig. 1c-d), computing the forces on theobtained restored structures, and then the elasticdipole. The case where n defect neighbor shellsare restored is referred to as the nth order approx-imation. As the restored zone becomes larger, theatoms remaining at their relaxed positions are morelikely to sit in an harmonic region. The convergenceof the resulting elastic dipole components with re-spect to n thus enables to evaluate the harmonicityaspect.

Fig. 2 provides the elastic dipole values as a func-tion of the cutoff radius r∞, for the octahedral con-figuration of the self-interstitial atom (SIA) in hcpZr. Only the point-defect has been restored in Fig.2a (approximation 0), whereas the restoration zoneextends to the 16th nearest-neighbors in Fig. 2b.Constant Pij values are reached for a cutoff radiusr∞ ∼ 2.5 a and ∼ 4 a, respectively, showing thatthe defect-induced forces are long-ranged [32, 41].As a result, the supercell needs to be large enoughto avoid convolution of the force field by periodicboundary conditions and a high precision on theatomic forces is required. Comparing with the elas-tic dipole deduced from the residual stress, one can-not only restore the point-defect (approximation 0in Fig. 2a) to obtain a quantitative estimate withthe Kanzaki method. A restoration zone extendingat least to the 16th nearest neighbors is necessaryfor this point-defect to obtain the correct elasticdipole. As the anharmonic region depends on thedefect and on the material, one cannot choose a pri-

10

Page 12: Elastic modeling of point-defects and their interaction

ori a radius for the restoration zone, but one needsto check the convergence of the elastic dipole withthe size of this restoration zone.

Discussion

These three approaches lead to the same valuesof the elastic dipole when large enough supercellsare used, thus confirming the consistency of thiselastic description of the point-defect. This hasbeen checked in Ref. [32] for the vacancy and var-ious configurations of the SIA in hcp Zr. But forsmall simulation cells typical of ab initio calcula-tions, both the fit of the displacement field and thecalculation from the Kanzaki forces are usually notprecise enough because of the too large defect coreregion, i.e. the region which has to be excluded fromthe displacement fit or the restoration zone for theKanzaki forces. This is penalizing for ab initio cal-culations, even for point-defects as simple as the Hsolute or the vacancy in hcp Zr [32, 43]. Besides,the Kanzaki’s technique requires additional calcu-lations to obtain the defect-induced forces and tocheck that the forces entering the dipole definitionare in the harmonic regime. As this restorationzone is extended, the defect-induced forces becomesmaller and the precision has to be increased. Thedefinition from the residual stress appears indeedas the only method leading to reliable Pij valueswithin ab initio simulations. It is also easy to apply,as it does not require any post treatment nor ad-ditional calculations: it only uses the homogeneousstress on the simulation box and the knowledge ofthe defect position is not needed.

All these methods can be of course also used todetermine the diaelastic polarizability. One onlyneeds to get the elastic dipole for various appliedstrains. The linear equation (29) then leads thestress-free elastic dipole P 0

ij and the polarizabilityαijkl. The most convenient method remains a def-inition from the residual stress. Considering thepolarizability, Eq. (32) now writes

〈σij(ε)〉 =

(Cijkl −

1

Vαijkl

)εkl −

1

VPij , (36)

thus showing that the polarizability is associatedwith a variation of the elastic constants propor-tional to the point-defect volume fraction. Thislinear variation of the elastic constants arising fromthe point-defect polarizability has been character-ized for vacancies and SIAs in face-centered cubic(fcc) copper [44], or various solute atoms in body-centered cubic (bcc) iron [45, 46].

8

12

16

20

0 2×10−4 4×10−4 6×10−4

432 250 128

Pij (

eV)

1/V (Å−3)

Number of Fe atoms

PxxPzz

Figure 3: Elastic dipole of a C atom lying in a [001] octa-hedral interstitial site in bcc Fe as a function of the inversevolume V of the supercell. The elastic dipole has been de-duced from the residual stress in ab initio calculations (seeRef. [47] for more details).

One consequence of the diaelastic polarizabilityis that the elastic dipole may depend on the sizeof the supercell with periodic boundary conditions.The strain at the point-defect position is indeed thesuperposition of the homogeneous strain εij and thestrains created by the periodic images of the point-defect εp

ij . In the ε = 0 case for instance, the ob-tained elastic dipole is then

Pij = P 0ij + αijklε

pkl. (37)

As the strain created by a point-defect varies as theinverse of the cube of the separation distance (Eq.9), the last term in Eq. (37) scales with the inverseof the supercell volume. Therefore, when homoth-etic supercells are used, one generally observes thefollowing volume variation

Pij = P 0ij +

δPijV

,

which can be used to extrapolate the elastic dipoleto an infinite volume, i.e. to the dilute limit [26, 32,47]. An example of this linear variation with theinverse volume is shown in Fig. 3 for an interstitialC atom in a bcc Fe matrix.

4.2. From experiments

From an experimental perspective, when tryingto extract elastic dipole of point-defects, both thesymmetry and the magnitude of the componentsof the elastic dipole tensor are a priori unknown,and possibly also the number of defect-types intothe material. We first restrict ourselves to the casewhere only one single type of point-defect with aknown symmetry is present.

If the point-defect has a lower symmetry thanthe host crystal, then it can adopt several variants

11

Page 13: Elastic modeling of point-defects and their interaction

which are equivalent by symmetry but possess dif-ferent orientations. The energy of such a volume Vcontaining different variants of the point-defect andsubmitted to a homogeneous strain is

E(ε) = E0 + EPD +V

2Cijklεijεkl

− Vnv∑µ=1

cµPµijεij , (38)

with nv the total number of different variants andcµ the volume concentration of variant µ. This rela-tion assumes that the different point-defects are notinteracting, which is valid in the dilute limit. Forzero stress conditions, as usually the case in experi-ments, the average strain induced by this assemblyof point-defects is

εij = Sijkl

nv∑µ=1

cµPµkl, (39)

with Sijkl the inverse of the elastic constants Cijkl.This linear relation between the strain and thepoint-defect concentrations corresponds to a Veg-ard’s law and allows for many connections with ex-periments. It generalizes Eq. 34 to the case of avolume containing a population of the same point-defect with different variants. As mentioned in §3.1,point-defects in experiments are sometimes rathercharacterized by their λ-tensor [10]. Combining thedefinition of this λ-tensor (Eq. 13) with Eq. (39),one shows the equivalence of both definitions:

λµij =1

ΩatSijkl P

µkl,

or equivalently Eq. (14).When the point-defect has only one variant or

when only one variant is selected by breaking thesymmetry – through either a phase transformation(e.g. martensitic [48, 49]) or the interaction with anapplied strain field for instance – the variations ofthe material lattice constants with the defect con-centration follow the defect symmetry. If the point-defect concentration is known, the elastic dipolecomponents are therefore fully accessible by mea-suring lattice parameter variations, e.g. by dilatom-etry or X-ray diffraction using the Bragg reflections.

On the other hand, for a completely disorderedsolid solution of point-defects with various variants(nv > 1), the average distortion induced by thepoint-defect population does not modify the parent

crystal symmetry [10]. Each variant is equiproba-ble, i.e. cµ = c0/nv with c0 the nominal point-defectconcentration. The stress-free strain induced by thepoint-defect (Eq. 39) thus becomes

εij = c0 Sijkl 〈Pkl〉 with 〈Pkl〉 =1

nv

nv∑µ=1

Pµkl.

Measurements of the lattice parameter variationswith the total defect concentration give thus accessonly to some sets of combinations of the Pij com-ponents. For instance, if we consider a point-defectin a cubic crystal, like a C solute in an octahedralsite of a bcc Fe crystal, one obtains the followingvariation of the lattice parameter with the soluteconcentration

a(c0) = a0

(1 +

Tr (P )

3 (C11 + 2C12)c0

), (40)

with C11 and C12 the elastic constants in Voigt no-tation. This variation can again be characterizedusing dilatometry or X-ray diffraction. But know-ing Tr (P ) is not sufficient for a point-defect witha lower symmetry than the cubic symmetry of thecrystal, as the elastic dipole has several indepen-dent components (two for the C solute atom in bccFe). Additional information is therefore needed tofully characterize the point-defect.

For those defects having a lower symmetry thantheir parent crystal, the anelastic relaxation ex-periments may provide such supplementary data[10, 50]. By applying an appropriate stress, a split-ting of the point-defect energy levels occurs, and aredistribution of the defect populations is operated.The relaxation of the compliance moduli then givesaccess to other combinations of the elastic dipolecomponents. Not all of the relaxations are allowedby symmetry, as illustrated for the C solute in bccFe, where only the quantity |P11 − P33| is accessi-ble [51]. The number of parameters accessible fromanelastic measurements is lower than the indepen-dent components of the defect elastic dipole. Thistechnique must then be used in combination withother measurements, like the variations of the lat-tice parameter.

Alternatively, a useful technique working witha random defect distribution is the diffuse Huangscattering. The diffuse scattering of X-rays nearBragg reflections [52–54] reflects the distortion scat-tering caused by the long-range part of the defect-induced displacement field. It thus provides infor-mation about the strength of the point-defect elas-tic dipole. The scattered intensity is proportional –

12

Page 14: Elastic modeling of point-defects and their interaction

in the dilute limit – to the defect concentration andto a linear combination of quadratic expressions ofthe elastic dipole components. The coefficients ofthis combination are functions of the crystal elasticconstants and of the scattering vector in the vicin-ity of a given reciprocal lattice vector. Therefore,by an appropriate choice of the relative scatteringdirection, the quadratic expressions can be deter-mined separately.

Except for simple point-defects like a substitu-tional solute atom or a single vacancy, the defectsymmetry may be unknown. Both anelastic re-laxation and Huang scattering experiments provideimportant information for the determination of thedefect symmetry. The presence of relaxation peaksin anelasticity is a direct consequence of the defectsymmetry [10, 50]. Within Huang scattering ex-periments, information about the defect symmetryis obtained either by the analysis of the morphol-ogy of iso-intensity curves or through an appropri-ate choice of scattering directions to measure theHuang intensity.

To conclude, when extracting elastic dipoles fromexperiments, one must usually rely on a combina-tion of several experimental techniques to obtain allcomponents.

5. Some applications

5.1. Solute interaction with a dislocation

This elastic modeling can be used for instance todescribe the interaction of a point-defect with otherstructural defects. To illustrate, and also validate,this approach, we consider a C interstitial atom in-teracting with a dislocation in a bcc iron matrix.This interstitial atom occupies the octahedral sitesof the bcc lattice. As these sites have a tetragonalsymmetry, the elastic dipole Pij of the C atoms hastwo independent components and gives thus rise toboth a size and a shape interaction. The interactionenergy of the C atom with a dislocation is givenby Eq. (12) where the external strain εext

ij is thestrain created by the dislocation at the position ofthe C atom. This has been compared in Ref. [55]to direct results of atomistic simulations, using forthe C elastic dipole and for the elastic constantsthe values given by the empirical potential used forthe atomistic simulations. Results show that elas-tic theory leads to a quantitative prediction whenall ingredients are included in the elastic model, i.e.when elastic anisotropy is taken into account to cal-culate the strain field created by the dislocation and

(a) Screw dislocation (h = 4d110 ' 8.1 A)

(b) Edge dislocation (h = −9d110 ' −18.2 A)

Figure 4: Binding energy Ebind = −Eint between a screwor an edge dislocation and a C atom in bcc iron for differentpositions x of the dislocation in its glide plane. The C atomis lying in a [100] octahedral interstitial site at a fixed dis-tance h of the dislocation glide plane. Symbols correspond toatomistic simulations and lines to elasticity theory, consid-ering all components of the stress created by the dislocationor only the pressure, and using isotropic or anisotropic elas-ticity.

when both the dilatation and the tetragonal distor-tion induced by the C atom are considered (Fig.4). The agreement between both techniques is per-fect except when the C atom is in the dislocationcore. With isotropic elasticity, the agreement withatomistic simulations is only qualitative, and whenthe shape interaction is not considered, i.e. whenthe C atom is modeled as a simple dilatation center(Pij = P δij), elastic theory fails to predict this in-teraction (Fig. 4). The same comparison betweenatomistic simulations and elasticity theory has beenperformed for a vacancy and a SIA interacting witha screw dislocation still in bcc iron [41]. The agree-ment was not as good as for the C atom. But in thiswork, the elastic dipoles of the point-defects were

13

Page 15: Elastic modeling of point-defects and their interaction

obtained from the Kanzaki forces, using the 0th or-der approximation, which is usually not as preciseas the definition from the stress (cf. § 4.1) and mayexplain some of the discrepancies.

One can also use elasticity theory to predict howthe migration barriers of the point-defect are mod-ified by a strain field. The migration energy is theenergy difference between the saddle point and thestable position. Its dependence with an appliedstrain field ε(~r) is thus described by

Em[ε] = Em0 + P ini

ij εij(~rini)− P sadij εij(~rsad), (41)

where P iniij and P sad

ij are the elastic dipoles of thepoint-defect respectively at its initial stable position~rini and at the saddle point ~rsad, and Em

0 is the mi-gration energy without elastic interaction. Still fora C atom interacting with a dislocation in a bcc Fematrix, comparison of this expression with resultsof direct atomistic simulations show a good agree-ment [56], as soon as the C atom is far enough fromthe dislocation core. Similar conclusions, on the va-lidity of equation (41) to describe the variation ofthe solute migration energy with an applied strain,have been reached for a SIA diffusing in bcc Fe [33],a vacancy in hcp zirconium [30] or a Si impurity infcc nickel [31].

5.2. Elastodiffusion

This simple model predicting the variation of themigration energy with an applied strain field (Eq.41) can be used to study elastodiffusion. Elastodif-fusion refers to the diffusion variations induced byan elastic field [57], either externally applied or in-ternal through the presence of structural defects.Important implications exist for materials, such astransport and segregation of point-defects to dislo-cations leading to the formation of Cottrell atmo-spheres [58], irradiation creep [59], or anisotropicdiffusion of dopants in semiconductor thin films[60, 61].

At the atomic scale, solid state diffusion oc-curs through the succession of thermally activatedatomic jumps from stable to other stable positions,with atoms jumping either on vacancy sites or on in-terstitial sites of the host lattice. Within transitionstate theory [62], the frequency of such a transitionis given by

Γα = ν0α exp (−Em

α / kT ), (42)

where ν0α is the attempt frequency for the transition

α and Emα is the migration energy.

Considering the effect of a small strain field onthis bulk system, the diffusion network and thesite topology will not be modified. On the otherhand, the presence of this small strain field modi-fies the migration energies and the attempt frequen-cies. As shown in the previous section, the elasticdipole description of the point-defect can predictthe modification of the stable and saddle point en-ergies, and thus of the migration energy (Eq. 41).Ignoring the strain effect on attempt frequencies,the incorporation of the modified energy barriersinto stochastic simulations like atomistic or objectkinetic Monte Carlo (OKMC) methods enables tocharacterize the point-defect elastodiffusion effect.This approach has been used, for instance, to studythe directional diffusion of point-defects in the het-erogeneous strain field of a dislocation, correspond-ing to a biased random walk [30, 56, 63].

Diffusion in a continuous solid body is character-ized by the diffusion tensor Dij which expresses theproportionality between the diffusion flux and theconcentration gradient (Fick’s law). The effect ofan applied strain is then described by the elastodif-fusion fourth-rank tensor dijkl [57], which gives thelinear dependence of the diffusion tensor with thestrain:

Dij = D0ij + dijkl εkl. (43)

This elastodiffusion tensor obeys the minor symme-tries dijkl = djikl = dijlk, because of the symmetryof the diffusion and deformation tensors, and alsothe crystal symmetries. Starting from the atom-istic events as defined by their transition frequen-cies (Eq. 42), the diffusion coefficient, and its varia-tion under an applied strain, can be evaluated fromthe long time evolution of the point-defect trajec-tories in stochastic simulations [64]. Alternatively,analytical approaches can be developed to provideexpressions [65, 66]. The elastodiffusion can thusbe computed by a perturbative approach, startingfrom the analytical expression of the diffusion ten-sor [57, 67]. This results in two different contri-butions: a geometrical contribution caused by theoverall change of the jump vectors and a contri-bution due to the change in energy barriers as de-scribed by Eq. (41). This last contribution is thusa function of the elastic dipoles at the saddle pointand stable positions. It is found to have an impor-tant magnitude in various systems [57, 67], being forinstance notably predominant for interstitial impu-rities in hcp Mg [68]. It is temperature-dependent,sometimes leading to complex variations with non-

14

Page 16: Elastic modeling of point-defects and their interaction

monotonic variations and also sign changes for someof its components [68]. As noted by Dederichs andSchroeder [57], the elastic dipole at the saddle pointcompletely determines the stress-induced diffusionanisotropy in cubic crystals. Experimental mea-surement of the elastodiffusion tensor componentscan therefore provide useful information about thesaddle point configurations.

Both approaches, relying either on stochasticsimulations or analytical models, are now usuallyinformed with ab initio computed formation andmigration energies, and attempt frequencies. Theelastic modeling of a point-defect through its elasticdipole offers thus a convenient way to transfer theinformation about the effects of an applied strain,as obtained from atomistic simulations, to the dif-fusion framework.

5.3. Bias calculations

Point-defect diffusion and absorption by elementsof the microstructure such as dislocations, cavities,grain boundaries and precipitates play an impor-tant role in the macroscopic evolution of materi-als. It is especially true under irradiation, since inthis case not only vacancies but also self-interstitialatoms (SIAs) migrate to these sinks. Owing totheir large dipole tensor components, SIAs gener-ally interact more than vacancies with the stressfields generated by sinks. This leads to a differ-ence in point-defect fluxes to a given sink knownas the “absorption bias”. For example, in the “dis-location bias model” [69], which is one of the mostpopular models to explain irradiation void swelling,dislocations are known as biased sinks: they absorbmore interstitials than vacancies. Voids, which pro-duce shorter range stress fields, are considered asneutral sinks, meaning that their absorption bias iszero. Since SIAs and vacancies are produced in thesame quantity, the preferential absorption of SIAsby dislocations leads to a net flux of vacancies tovoids and thus to void growth. Similar explanationsbased on absorption biases have been given to ratio-nalize irradiation creep [21] and irradiation growthin hexagonal materials [70]. In order to predict thekinetics of such phenomena, a precise evaluation ofabsorption biases is necessary.

Following the rate theory formalism [69], the ab-sorption bias of a given sink can be written as therelative difference of sink strengths for interstitials(k2i ) and vacancies (k2

v) [71]. The strength of a sinkfor a point-defect θ (θ = i, v) is related to the loss

rate φθ through

φθ = k2θDθcθ, (44)

where Dθ is the diffusion coefficient free of elasticinteractions and cθ is the volume concentration ofθ.

The sink strength can be calculated with differ-ent methods, for example by solving the diffusionequation around the sink [57, 69] or an associatedphase field model [72], or by performing object ki-netic Monte Carlo simulations (OKMC) [73, 74]. Itshould be noted that analytical solution of the diffu-sion equation is limited to a few cases and often re-quires the defect properties or the stress field to besimplified [75–77], so in general numerical simula-tions are necessary [78–81]. In the following we con-sider the OKMC approach, due to its simplicity andits flexibility to introduce complex diffusion mech-anisms and the effect of stress fields [30, 82, 83].

In OKMC simulations of sink strengths, a sinkis introduced in a simulation box where periodicboundary conditions are used and point-defects aregenerated at a given rate K. They diffuse in thebox by successive atomic jumps until they are ab-sorbed by the sink. For each defect in the simula-tion box, the jump frequencies of all jumps from thecurrent stable state to the possible final states arecalculated and the next event is chosen accordingto the standard residence time algorithm [84, 85].The jump frequency of event α is given by Eq. (42),considering the strain dependence of the migrationenergy through Eq. (41).

The sink strength is deduced from the averagenumber of defects in the box Nθ at steady state bythe following equation [83]:

k2θ =

K

DθNθ

, (45)

from which the bias is deduced:

B =k2i − k2

v

k2i

. (46)

Another method is often used for the calculationof sink strengths with OKMC [73, 74]. For eachdefect, the number of jumps it performs before itis absorbed by the sink is registered. The sinkstrength is then deduced from the average num-ber of jumps. Although this method is equivalentto the method based on the average concentrationin the non-interacting case, it is no more valid ifelastic interactions are included. In this case the

15

Page 17: Elastic modeling of point-defects and their interaction

4 6 8 10 12 14 16 18 20 22d (nm)

0.0

0.5

1.0

1.5

2.0

2.5

3.0

k2 v(n

m−

2 )

(a) Vacancies

Anisotropic saddleIsotropic saddleAnalytical solution

4 6 8 10 12 14 16 18 20 22d (nm)

0.0

0.5

1.0

1.5

2.0

2.5

3.0

k2 i(n

m−

2 )

(b) SIAs

4 6 8 10 12 14 16 18 20 22d (nm)

−4

−3

−2

−1

0

1

B

(c) Bias

Figure 5: Sink strengths of a twist grain boundary (θ = 7.5°)for (a) vacancies and (b) SIAs, and (c) absorption bias, asa function of the layer thickness d. (see Ref. [83] for moredetails).

average time before absorption should be measuredinstead of the average number of jumps, since jumpfrequencies now depend on the location of the de-fect and are usually higher. Therefore, applyingthis method in the interacting case often leads toan underestimation of sink strengths.

As an illustration, we consider the study pub-lished in Ref. [83], where sink strengths ofsemi-coherent interfaces have been calculated withOKMC, taking into account the effect of the strainfield generated by the interfaces. The strain is thesum of the coherency strain and of the strain dueto interface dislocations. It has been calculated bya semi-analytical method within the framework ofanisotropic elasticity [83, 86, 87]. We consider thecase of a twist grain boundary in Ag, which pro-duces a purely deviatoric strain field. Two grainboundaries distant from each other by d are intro-duced in the box and periodic boundary conditionsare applied.

Dipole tensors of vacancies and SIAs in Ag havebeen computed by DFT for both stable and saddlepositions [83], using the residual stress definition(Eq. 33). At the ground state, the elastic dipoleof the vacancy is isotropic and the one of the SIAalmost isotropic. On the other hand, the elasticdipole tensors have a significant deviatoric compo-nent for both point-defects at their saddle point.

Sink strengths of the twist grain boundary are

shown in Fig. 5-(a,b) as a function of the layer thick-ness d and compared to the analytical result withno elastic interactions k2 = 12/d2. Sink strengthsfor both vacancies and SIAs are significantly in-creased when elastic interactions are included andwhen anisotropy at saddle point is taken into ac-count, especially for thinner layers. However, ifthe saddle point is considered isotropic, the non-interacting case is recovered. This is due to thedeviatoric character of the strain field: since thedipole tensor of the vacancy in its ground state ispurely hydrostatic, the interaction energy of a va-cancy with the strain field is zero and there is nothermodynamic driving force for the absorption ofthe vacancy. A similar result is obtained for SIAs,because of their almost purely hydrostatic dipole fortheir ground state. Fig. 5c shows the evolution ofthe bias. For this interface, saddle point anisotropyleads to a negative bias, meaning that vacanciestend to be more absorbed than interstitials.

This approach has also been recently used for thecalculation of the sink strength of straight disloca-tions and cavities in aluminum [88]. In both cases,saddle point anisotropy appears to have a signifi-cant influence on the sink strengths. This confirmsanalytical results obtained with various levels of ap-proximation [77, 89, 90].

5.4. Isolated defect in atomistic simulations

The elastic modeling of point-defects is also use-ful in the context of atomistic simulations. Suchsimulations, in particular ab initio calculations, arenow unavoidable to obtain the point-defects ener-getics, like their formation and migration energies[1]. However, an ongoing issue is their difficulty toobtain the properties of isolated defects. One canuse atomistic simulations with controlled surface tomodel an isolated point-defect [91–95], but then,the excess energy associated with the point-defectcould be exactly set apart from the one of the exter-nal surfaces or interfaces only for interatomic poten-tials with a cutoff interaction radius, correspond-ing to short-range empirical potentials like EAM.For more complex potentials or for ab initio cal-culations, the absence of any interaction cutoff pre-vents an unambiguous definition of the point-defectenergy. A supercell approach relying on periodicboundary conditions is therefore usually preferred.The combined effect of periodic boundary condi-tions and of the limited size of such calculations, fornumerical cost reasons, makes the computed prop-erties difficult to converge for defects inducing long-

16

Page 18: Elastic modeling of point-defects and their interaction

range effects. This problem is well-known in thecontext of charged point-defects, where long-rangeCoulombian interactions exist between the defectand its periodic images and for which correctiveschemes have been developed [96–98]. For neutraldefects, interactions between periodic images alsoexist. These interactions are of elastic origin anddecay like the inverse cube of the separation dis-tance. Consequently, the computed excess energiesare those of a periodic array of interacting point-defects, and converge with the inverse of the su-percell volume to the energy of the isolated defect.This can be penalizing for defects inducing largedistortions, like SIAs or clusters, or for atomic cal-culations where only small supercells are reachable.The elastic description of a point-defect allows cal-culating this spurious elastic interaction associatedwith periodic boundary conditions to obtain the en-ergy properties of the isolated point-defect [99].

After atomic relaxation, the excess energy of asupercell containing one point-defect is given by:

EPDPBC(ε = 0) = EPD

∞ +1

2Eint

PBC, (47)

where EPD∞ is the excess energy of the isolated de-

fect and EintPBC is the interaction energy of the de-

fect with its periodic images. The factor 1/2 arisesbecause half of the interaction is devoted to the de-fect itself and the other goes to its periodic images.Continuous linear elasticity theory can be used toevaluate this elastic interaction. If the point-defectis characterized by the elastic dipole Pij , followingEq. 12, this interaction energy is given by

EintPBC = −Pij εPBC

ij , (48)

with εPBCij the strain created by the defect periodic

images. It can be obtained by direct summation

εPBCij = −

∑n,m,p

′Gik,jl(n~a1 +m~a2 + p~a3)Pkl. (49)

with ~a1, ~a2 and ~a3 the periodicity vectors of the su-percell. The prime sign indicates that the divergingterm (n = m = p = 0) has been excluded from thesum. As the second derivative of the Green’s func-tion Gik,jl(~r) is decaying like 1/r3, this sum is onlyconditionally convergent. It can be regularized fol-lowing the numerical scheme proposed by Cai [100].

After computing the point-defect energy with anatomistic simulation code, this energy can be cor-rected by subtracting the interaction energy with

the periodic images (Eq. 47) to obtain the proper-ties of the isolated defect. This interaction energy iscomputed from the elastic constants of the perfectcrystal, which are needed to evaluate the Green’sfunction and its derivative (cf. § 2.3), and from theresidual stress of the defective supercell to deter-mine the point-defect elastic dipole (cf. § 4.1). Thisis therefore a simple post-treatment, which does notinvolve any fitting procedure and which can be per-formed using the Aneto program provided as sup-plemental material of Ref. [99].

We have assumed in Eq. (47) that the supercellcontaining the point-defect has the same periodic-ity vector than the perfect supercell, i.e. the appliedhomogenous strain ε is null. This corresponds tothe easiest boundary conditions in atomistic simu-lations of point-defects. But sometimes, one prefersto relax also the periodicity vectors to nullify thestress in the supercell. Both these ε = 0 and σ = 0conditions converge to the same energy EPD

∞ in thethermodynamic limit but different energies are ob-tained for too small supercells. The elastic modelcan be further developed to rationalize this differ-ence [26, 99]. For σ = 0 conditions, a strain ε is ap-plied to the defective supercell to nullify its stress.Eq. (47) therefore needs to be complemented withthe energy contribution of this deformation

∆E(ε) =V

2Cijklεij εkl − Pij εij .

This applied strain ε in zero stress calculations islinked to the elastic dipole by Eq. (34). The excessenergy of the supercell containing one point-defectis thus now given by

EPDPBC(σ = 0) = EPD

∞ +1

2Eint

PBC −1

2VSijklPijPkl

= EPDPBC(ε = 0)− 1

2VSijklPijPkl,

(50)

where the elastic compliances of the bulk materialSijkl are the inverse tensor of the elastic constantsCijkl. This equation shows that ε = 0 and σ = 0conditions lead to point-defect excess energies dif-fering by a factor proportional to the inverse of thesupercell volume and to the square of the elasticdipole. This difference will be therefore importantfor small supercells and/or point-defects inducingan important perturbation of the host lattice. Butonce corrected through Eqs. (47) or (50), both ap-proaches should lead to the same value. σ = 0calculations appear therefore unnecessary.

17

Page 19: Elastic modeling of point-defects and their interaction

16

18

20

22

Ef

(eV

)

EAM

ε = 0ε = 0 corr.σ = 0σ = 0 corr.

GGA

(a )

16

20

24

28

0 1500 3000

Ef

(eV

)

(b )

Number of atoms

Figure 6: Formation energy of a SIA cluster containing eightinterstitials in bcc iron calculated for fixed periodicity vec-tors (ε = 0) or at zero stress (σ = 0) for different sizes of thesimulation cell: (a) C15 aggregate and (b) parallel-dumbellconfiguration with a 〈111〉 orientation. Atomistic simula-tions are performed either with the M07 empirical potential[101] (EAM) or with ab initio calculations (GGA). Filledsymbols refer to uncorrected results and open symbols tothe results corrected by the elastic model (see Ref. [99] formore details).

We illustrate the usefulness of this elastic post-treatment on an atomistic study of SIA clusters inbcc iron. These clusters appear under irradiationand can adopt different morphologies [101]. In par-ticular, some clusters can have a 3D structure withan underlying crystal symmetry corresponding tothe C15 Laves’ phase, and others have a planarstructure corresponding to dislocation loop clusterswith 1/2 〈111〉 Burgers vectors.

The formation energies of two different configura-tions of a cluster containing 8 SIAs, a C15 aggregateand a planar aggregate of parallel-dumbells with a〈111〉 orientation, are shown in Fig. 6 for differ-ent supercell sizes. They have been first calculatedwith an empirical EAM potential [101]: with fixedperiodicity vectors (ε = 0), one needs at least 2000atoms for the C15 aggregate and 4000 atoms forthe 〈111〉 planar configuration to get a formationenergy converged to a precision better than 0.1 eV.The convergence is slightly faster for zero stress cal-

culations (σ = 0) in the case of the C15 aggregate(Fig. 6a), but the opposite is true in the case ofthe 〈111〉 planar configuration (Fig. 6b). Whenwe add the elastic correction, the convergence isimproved for both cluster configurations. The cor-rected ε = 0 and σ = 0 calculations lead then tothe same formation energies, except for the small-est simulation cell (128 lattice sites) in the case ofthe 〈111〉 cluster. These formation energies havebeen also obtained with ab initio calculations for asimulation cell containing 250 lattice sites (Fig. 6).Uncorrected ε = 0 calculations lead to an energydifference ∆E = −5.6 eV between the C15 and the〈111〉 planar configuration, whereas this energy dif-ference is only ∆E = −0.6 eV in σ = 0 calculations.This variation of the energy difference is rational-ized once the elastic correction is added, and a goodprecision is obtained with this approach coupling abinitio calculations and elasticity theory, with an en-ergy difference of ∆E = 3.5 ± 0.2 eV. This elasticcorrection has been shown to accelerate the conver-gence of the point-defect formation and/or migra-tion energies obtained from atomistic simulations,in particular from ab initio calculations, in numer-ous other cases like SIA in hcp Zr [99, 102], vacancyin diamond silicon [99], or solute interstitials in bcciron [103].

6. Conclusions

Elasticity theory provides thus an efficient frame-work to model point-defects. Describing the point-defect as an equilibrated distribution of point-forces, the long range elastic field of the defect andits interaction with other elastic fields are fully char-acterized by the first moment of this force distri-bution, a second rank symmetric tensor called theelastic dipole. This description is equivalent to aninfinitesimal Eshelby inclusion or an infinitesimaldislocation loop. Knowing only the elastic con-stants of the matrix and the elastic dipole, a quan-titative modeling of the point-defect and its inter-actions is thus obtained. The value of this elas-tic dipole can be either deduced from experimen-tal data, like Vegard’s law parameters, or extractedfrom atomistic simulations. In this latter case, caremust be taken to avoid finite-size effects, in particu-lar for ab initio calculations. The definition throughthe residual stress appears as the most precise oneto obtain the dipole tensors.

The elastic description offers a convenient frame-work to bridge the scales between an atomic and

18

Page 20: Elastic modeling of point-defects and their interaction

a continuum description so as to consider the in-teraction of the point-defects with various complexelastic fields. This upscaling approach has alreadyproven its efficiency in the modeling of elastodif-fusion or in the calculation of absorption bias un-der irradiation. As the numerical evaluation of theelastic Green’s function and its derivatives does notpresent nowadays any technical difficulty, such anelastic model offers also a nice route to simulate theevolution of a whole population of point-defects ina complex microstructure, considering their mutualinteraction and their interaction with other struc-tural defects, in the same spirit as dislocation dy-namics simulations are now routinely used to modelthe evolution of a dislocation microstructure.

Acknowledgements - This work was performed us-ing HPC resources from GENCI-CINES and -TGCC(Grants 2017-096847). The research was partly fundedby the European Atomic Energy Community’s (Eu-ratom) Seventh Framework Program FP7 under grantagreement No. 604862 (MatISSE project) and in theframework of the EERA (European Energy ResearchAlliance) Joint Program on Nuclear Materials.

References

[1] C. Freysoldt, B. Grabowski, T. Hickel, J. Neuge-bauer, G. Kresse, A. Janotti, C. G. Van de Walle,First-principles calculations for point defects in solids,Rev. Mod. Phys. 86 (2014) 253—-305. doi:10.1103/

revmodphys.86.253.[2] J. Eshelby, The continuum theory of lattice defects,

in: F. Seitz, D. Turnbull (Eds.), Solid State Physics,Vol. 3, Academic Press, 1956, pp. 79–144. doi:10.

1016/S0081-1947(08)60132-0.[3] L. D. Landau, E. M. Lifshitz, Theory of Elasticity,

2nd Edition, Vol. 7 of Course of Theoretical Physics,Pergamon Press, 1970.

[4] J. F. Nye, Physical Properties of Crystals - Their rep-resentation by tensors and matrices, Oxford UniversityPress, 1957.

[5] D. M. Barnett, The precise evaluation of derivativesof the anisotropic elastic Green’s functions, Phys. Sta-tus Solidi B 49 (1972) 741–748. doi:10.1002/pssb.

2220490238.[6] D. J. Bacon, D. M. Barnett, R. O. Scattergood,

Anisotropic continuum theory of lattice defects,Prog. Mater. Sci. 23 (1980) 51–262. doi:10.1016/

0079-6425(80)90007-9.[7] R. Siems, Mechanical interactions of point defects,

Phys. Status Solidi 30 (1968) 645–658. doi:10.1002/

pssb.19680300226.[8] G. Leibfried, N. Breuer, Point Defects in Met-

als I, Vol. 81 of Springer Tracts in ModernPhysics, Springer-Verlag, Berlin, 1978. doi:10.1007/

BFb0045966.[9] C. Teodosiu, Elastic Models of Crystal Defects,

Springer-Verlag, Berlin, 1982.

[10] A. S. Nowick, B. S. Berry, Anelastic Relaxation inCrystalline Solids, Academic Press, New York, 1972.

[11] J. D. Eshelby, The determination of the elastic fieldof an ellipsoidal inclusion, and related problems, Proc.Roy. Soc. Lond. A 241 (1957) 376–396. doi:10.1098/

rspa.1957.0133.[12] J. D. Eshelby, The elastic field outside an ellipsoidal

inclusion, Proc. Roy. Soc. Lond. A 252 (1959) 561–569.doi:10.1098/rspa.1959.0173.

[13] M. Lazar, Micromechanics and theory of point de-fects in anisotropic elasticity: Eshelby factor meetsEshelby tensor, Journal of Micromechanics and Molec-ular Physics 02 (2017) 1750005. doi:10.1142/

S2424913017500059.[14] F. R. N. Nabarro, Theory of Crystal Dislocations, Ox-

ford Univ. Press, London, 1967.[15] T. Mura, Micromechanics of Defects in Solids, Kluwer

Academic, Dordrecht, 1987.[16] J. P. Hirth, J. Lothe, Theory of Dislocations, 2nd Edi-

tion, Wiley, New York, 1982.[17] E. Kroner, Dia- and para-elasticity, in: F. Kroupa

(Ed.), Theory of crystal defects, Academia, 1964, pp.215–230.

[18] H. R. Schober, Polarizabilities of point defects in met-als, J. Nucl. Mater. 126 (1984) 220–225. doi:10.1016/0022-3115(84)90032-1.

[19] M. P. Puls, C. H. Woo, Diaelastic polarizabilitiesdue to vacancies and interstitials in metals, J. Nucl.Mater. 139 (1986) 48–59. doi:10.1016/0022-3115(86)90163-7.

[20] A. Granato, The temperature-independent componentof the diaelastic effect, J. Alloys Compd. 211-212(1994) 503–508. doi:10.1016/0925-8388(94)90553-3.

[21] P. T. Heald, M. V. Speight, Steady-state irradia-tion creep, Philos. Mag. 29 (1974) 1075–1080. doi:

10.1080/14786437408226592.[22] P. T. Heald, M. V. Speight, Point defect behaviour

in irradiated materials, Acta Metall. 23 (1975) 1389–1399. doi:10.1016/0001-6160(75)90148-0.

[23] T. Bullough, J. R. Willis, The stress-induced pointdefect-dislocation interaction and its relevance to irra-diation creep, Philos. Mag. 31 (1975) 855–861. doi:

10.1080/14786437508229635.[24] R. Bullough, M. Hayns, Irradiation-creep due to point

defect absorption, J. Nucl. Mater. 57 (3) (1975) 348 –352. doi:10.1016/0022-3115(75)90220-2.

[25] J. L. Snoek, Effect of small quantities of carbonand nitrogen on the elastic and plastic propertiesof iron, Physica 8 (1941) 711–733. doi:10.1016/

S0031-8914(41)90517-7.[26] B. Puchala, M. L. Falk, K. Garikipati, Elastic effects

on relaxation volume tensor calculations, Phys. Rev. B77 (2008) 174116. doi:10.1103/PhysRevB.77.174116.

[27] R. C. Pasianot, On the determination of defect dipolesfrom atomistic simulations using periodic boundaryconditions, Philos. Mag. Lett. 96 (2016) 447–453. doi:10.1080/09500839.2016.1250965.

[28] M. J. Gillan, The volume of formation of defects inionic crystals, Philos. Mag. A 43 (1981) 301–312. doi:10.1080/01418618108239410.

[29] M. J. Gillan, The long-range distortion caused by pointdefects, Philos. Mag. A 48 (1983) 903–919. doi:10.

1080/01418618308244326.[30] G. Subramanian, D. Perez, B. P. Uberuaga, C. N.

Tome, A. F. Voter, Method to account for arbi-

19

Page 21: Elastic modeling of point-defects and their interaction

trary strains in kinetic Monte Carlo simulations, Phys.Rev. B 87 (2013) 144107. doi:10.1103/PhysRevB.87.

144107.[31] T. Garnier, V. R. Manga, P. Bellon, D. R. Trinkle,

Diffusion of Si impurities in Ni under stress: A first-principles study, Phys. Rev. B 90 (2014) 024306. doi:10.1103/physrevb.90.024306.

[32] C. Varvenne, E. Clouet, Elastic dipoles of point defectsfrom atomistic simulations, Phys. Rev. B 96 (2017)224103. doi:10.1103/PhysRevB.96.224103.

[33] Z. Chen, N. Kioussis, N. Ghoniem, D. Seif, Strain-field effects on the formation and migration energies ofself interstitials in alpha-Fe from first principles, Phys.Rev. B 81 (2010) 094102. doi:10.1103/PhysRevB.81.

094102.[34] H. Kanzaki, Point defects in face-centred cubic lattice

— I Distortion around defects, J. Phys. Chem. Solids 2(1957) 24 – 36. doi:10.1016/0022-3697(57)90003-3.

[35] I. D. Faux, A. B. Lidiard, The volume of formationof Schottky defects in ionic crystals, Zeitschrift furNaturforschung 26a (1971) 62–68.

[36] V. K. Tewary, Green-function method for lattice stat-ics, Adv. Phys. 22 (1973) 757–810. doi:10.1080/

00018737300101389.[37] H. R. Schober, K. W. Ingle, Calculation of relaxation

volumes, dipole tensors and kanzaki forces for pointdefects, J. Phys. F Metal. Phys. 10 (1980) 575–581.doi:10.1088/0305-4608/10/4/009.

[38] A. B. Lidiard, The volume of formation of schottkydefects in ionic crystals, Philos. Mag. A 43 (1981) 291–300. doi:10.1080/01418618108239409.

[39] G. Simonelli, R. Pasianot, E. Savino, Point-defectcomputer simulation including angular forces in bcciron, Phys. Rev. B 50 (1994) 727–738. doi:10.1103/

PhysRevB.50.727.[40] C. Domain, C. S. Becquart, J. Foct, Ab initio study

of foreign interstitial atom (C, N) interactions withintrinsic point defects in alpha -Fe, Phys. Rev. B 69(2004) 144112. doi:10.1103/PhysRevB.69.144112.

[41] E. Hayward, C. Deo, B. P. Uberuaga, C. N. Tome, Theinteraction of a screw dislocation with point defectsin bcc iron, Philos. Mag. 92 (2012) 2759–2778. doi:

10.1080/14786435.2012.674646.[42] M. I. Mendelev, G. J. Ackland, Development of an in-

teratomic potential for the simulation of phase trans-formations in zirconium, Philos. Mag. Lett. 87 (2007)349–359. doi:10.1080/09500830701191393.

[43] R. Nazarov, J. S. Majevadia, M. Patel, M. R. Wen-man, D. S. Balint, J. Neugebauer, A. P. Sutton,First-principles calculation of the elastic dipole ten-sor of a point defect: Application to hydrogen inα-zirconium, Phys. Rev. B 94 (2016) 241112. doi:

10.1103/PhysRevB.94.241112.[44] G. J. Ackland, Theoretical study of the effect of point

defects on the elastic constants of copper, J. Nucl.Mater. 152 (1988) 53–63. doi:10.1016/0022-3115(88)90140-7.

[45] A. F. Bialon, T. Hammerschmidt, R. Drautz, Ab ini-tio study of boron in α-iron: Migration barriers andinteraction with point defects, Phys. Rev. B 87 (2013)104109. doi:10.1103/PhysRevB.87.104109.

[46] M. R. Fellinger, L. G. H. Jr., D. R. Trinkle, Ab initiocalculations of the lattice parameter and elastic stiff-ness coefficients of bcc Fe with solutes, Comp. Mater.Sci. 126 (2017) 503–513. doi:10.1016/j.commatsci.

2016.09.040.[47] E. Clouet, L. Ventelon, F. Willaime, Dislocation core

field. II. Screw dislocation in iron, Phys. Rev. B 84(2011) 224107. doi:10.1103/PhysRevB.84.224107.

[48] C. S. Roberts, Effect of carbon on the volume frac-tions and lattice parameters of retained austenite andmartensite, Trans. AIME 197 (1953) 203–204.

[49] L. Cheng, A. Bottger, T. H. de Keijser, E. J.Mittemeijer, Lattice parameters of iron-carbon andiron-nitrogen martensites and austenites, ScriptaMetall. Mater. 24 (1990) 509–514. doi:10.1016/

0956-716X(90)90192-J.[50] A. Nowick, W. Heller, Anelasticity and stress-induced

ordering of point defects in crystals, Adv. Phys. 12(1963) 251–298. doi:10.1080/00018736300101293.

[51] J. C. Swartz, J. W. Shilling, A. J. Schwoeble, Dipolarstrains of c and n in alpha-iron, Acta Metall. 16 (1968)1359–1364. doi:10.1016/0001-6160(68)90156-9.

[52] H. Trinkaus, On determination of the double-force ten-sor of point defects in cubic crystals by diffuse x-ray scattering, Phys. Stat. Sol. B 51 (1972) 307–319.doi:10.1002/pssb.2220510131.

[53] O. Bender, P. Ehrhart, Self-interstitial atoms, va-cancies and their agglomerates in electron-irradiatednickel investigated by diffuse scattering of X-rays, J.Phys. F: Metal. Phys. 13 (1983) 911. doi:10.1088/

0305-4608/13/5/006.[54] T. Michelitsch, A. Wunderlin, On the theory of Huang

X-ray scattering caused by point defects in hexagonalcrystals, Phys. Stat. Sol. B 198 (1996) 615–620. doi:

10.1002/pssb.2221980206.[55] E. Clouet, S. Garruchet, H. Nguyen, M. Perez, C. S.

Becquart, Dislocation interaction with C in α-Fe:A comparison between atomic simulations and elas-tic theory, Acta Mater. 56 (2008) 3450–3460. doi:

10.1016/j.actamat.2008.03.024.[56] R. G. A. Veiga, M. Perez, C. S. Becquart, E. Clouet,

C. Domain, Comparison of atomistic and elasticity ap-proaches for carbon diffusion near line defects in α-iron, Acta Mater. 59 (2011) 6963–6974. doi:10.1016/j.actamat.2011.07.048.

[57] P. H. Dederichs, K. Schroeder, Anisotropic diffusionin stress fields, Phys. Rev. B 17 (1978) 2524. doi:

10.1103/PhysRevB.17.2524.[58] A. H. Cottrell, B. A. Bilby, Dislocation theory of yield-

ing and strain ageing of iron, Proc. Phys. Soc. LondonSer. A 62 (1949) 49. doi:10.1088/0370-1298/62/1/

308.[59] C. Woo, Irradiation creep due to elastodiffusion, J.

Nucl. Mater. 120 (1984) 55 – 64. doi:10.1016/

0022-3115(84)90170-3.[60] M. J. Aziz, Thermodynamics of diffusion under pres-

sure and stress: Relation to point defect mechanisms,Appl. Phys. Lett. 70 (1997) 2810–2812. doi:10.1063/1.119066.

[61] M. S. Daw, W. Windl, N. N. Carlson, M. Laudon,M. P. Masquelier, Effect of stress on dopant and defectdiffusion in Si: A general treatment, Phys. Rev. B 64(2001) 045205. doi:10.1103/PhysRevB.64.045205.

[62] G. H. Vineyard, Frequency factors and isotope effectsin solid state rate processes, J. Phys. Chem. Solids 3(1957) 121. doi:10.1016/0022-3697(57)90059-8.

[63] R. G. A. Veiga, M. Perez, C. S. Becquart, C. Domain,S. Garruchet, Effect of the stress field of an edge dislo-cation on carbon diffusion in α-iron: Coupling molec-

20

Page 22: Elastic modeling of point-defects and their interaction

ular statics and atomistic kinetic Monte Carlo, Phys.Rev. B 82 (2010) 054103. doi:10.1103/PhysRevB.82.

054103.[64] A. Goyal, S. R. Phillpot, G. Subramanian, D. A.

Andersson, C. R. Stanek, B. P. Uberuaga, Impactof homogeneous strain on uranium vacancy diffusionin uranium dioxide, Phys. Rev. B 91 (2015) 094103.doi:10.1103/PhysRevB.91.094103.

[65] R. E. Howard, A. B. Lidiard, Matter transport insolids, Rep. Prog. Phys. 27 (1964) 161–240. doi:

10.1088/0034-4885/27/1/305.[66] A. R. Allnatt, A. B. Lidiard, Atomic Transport in

Solids, Cambridge University Press, 1993.[67] D. R. Trinkle, Diffusivity and derivatives for intersti-

tial solutes: activation energy, volume, and elastod-iffusion tensors, Philos. Mag. 96 (2016) 2714–2735.doi:10.1080/14786435.2016.1212175.

[68] R. Agarwal, D. R. Trinkle, Light-element diffusion inMg using first-principles calculations: Anisotropy andelastodiffusion, Phys. Rev. B 94 (2016) 054106. doi:

10.1103/PhysRevB.94.054106.[69] A. D. Brailsford, R. Bullough, The rate theory

of swelling due to void growth in irradiated met-als, J. Nucl. Mater. 44 (1972) 121. doi:10.1016/

0022-3115(72)90091-8.[70] H. Rouchette, L. Thuinet, A. Legris, A. Ambard,

C. Domain, Influence of shape anisotropy of self-interstitials on dislocation sink efficiencies in Zr: Mul-tiscale modeling, Phys. Rev. B 90 (2014) 014104.doi:10.1103/physrevb.90.014104.

[71] P. T. Heald, The preferential trapping of interstitialsat dislocations, Philos. Mag. 31 (1975) 551. doi:10.

1080/14786437508226537.[72] H. Rouchette, L. Thuinet, A. Legris, A. Ambard,

C. Domain, Quantitative phase field model for disloca-tion sink strength calculations, Comp. Mater. Sci. 88(2014) 50. doi:10.1016/j.commatsci.2014.02.011.

[73] H. L. Heinisch, B. N. Singh, S. I. Golubov, The effectsof one-dimensional glide on the reaction kinetics of in-terstitial clusters, J. Nucl. Mater. 283–287 (2000) 737.doi:10.1016/s0022-3115(00)00258-0.

[74] L. Malerba, C. S. Becquart, C. Domain, Object ki-netic Monte Carlo study of sink strengths, J. Nucl.Mater. 360 (2007) 159. doi:10.1016/j.jnucmat.2006.10.002.

[75] K. Schroeder, K. Dettmann, Diffusion reactions in longrange potentials, Z. Phys. B 22 (1975) 343. doi:10.

1007/bf01312804.[76] C. H. Woo, The sink strength of a dislocation loop in

the effective medium approximation, J. Nucl. Mater.98 (1981) 279. doi:10.1016/0022-3115(81)90154-9.

[77] B. C. Skinner, C. H. Woo, Shape effect in the drift dif-fusion of point defects into straight dislocations, Phys.Rev. B 30 (1984) 3084–3097. doi:10.1103/PhysRevB.

30.3084.[78] C. H. Woo, W. S. Liu, M. S. Wuschke, A finite-

difference calculation of point defect migration into adislocation loop, Tech. Rep. AECL-6441, Atomic En-ergy of Canada Limited (1979).

[79] R. Bullough, D. W. Wells, J. R. Willis, M. H. Wood,The interaction energy between interstitial atoms anddislocations and its relevance to irradiation damageprocesses, in: M. F. Ashby, R. Bullough, C. S. Hartley,J. P. Hirth (Eds.), Dislocation Modelling of PhysicalSystems, Pergamon, 1981, p. 116.

[80] V. I. Dubinko, A. S. Abyzov, A. A. Turkin, Numericalevaluation of the dislocation loop bias, J. Nucl. Mater.336 (2005) 11. doi:j.jnucmat.2004.07.034.

[81] T. Jourdan, Influence of dislocation and dislocationloop biases on microstructures simulated by rate equa-tion cluster dynamics, J. Nucl. Mater. 467 (2015) 286.doi:j.jnucmat.2015.09.046.

[82] A. B. Sivak, V. M. Chernov, V. A. Romanov, P. A.Sivak, Kinetic monte-carlo simulation of self-point de-fect diffusion in dislocation elastic fields in bcc ironand vanadium, J. Nucl. Mater. 417 (2011) 1067–1070.doi:10.1016/j.jnucmat.2010.12.176.

[83] A. Vattre, T. Jourdan, H. Ding, M.-C. Marinica, M. J.Demkowicz, Non-random walk diffusion enhances thesink strength of semicoherent interfaces, Nat. Com-mun. 7 (2016) 10424. doi:10.1038/ncomms10424.

[84] D. T. Gillespie, A general method for numericallysimulating the stochastic time evolution of coupledchemical reactions, J. Comput. Phys. 22 (1976) 403.doi:10.1016/0021-9991(76)90041-3.

[85] A. B. Bortz, M. H. Kalos, J. L. Lebowitz, A new al-gorithm for Monte-Carlo simulation of ising spin sys-tems, J. Comput. Phys. 17 (1975) 10. doi:10.1016/

0021-9991(75)90060-1.[86] A. Vattre, M. Demkowicz, Determining the Burgers

vectors and elastic strain energies of interface dislo-cation arrays using anisotropic elasticity theory, ActaMater. 61 (2013) 5172–5187. doi:10.1016/j.actamat.2013.05.006.

[87] A. J. Vattre, M. J. Demkowicz, Partitioning of elasticdistortions at a semicoherent heterophase interface be-tween anisotropic crystals, Acta Mater. 82 (2015) 234.doi:10.1016/j.actamat.2014.09.014.

[88] D. Carpentier, T. Jourdan, Y. Le Bouar, M.-C.Marinica, Effect of saddle point anisotropy of point de-fects on their absorption by dislocations and cavities,Acta Mater. 136 (2017) 323. doi:10.1016/j.actamat.2017.07.013.

[89] V. A. Borodin, A. I. Ryazanov, C. Abromeit, Voidbias factors due to the anisotropy of the point defectdiffusion, J. Nucl. Mater. 207 (1993) 242. doi:10.

1016/0022-3115(93)90266-2.[90] V. A. Borodin, A. I. Ryazanov, The effect of diffusion

anisotropy on dislocation bias and irradiation creep incubic lattice materials, J. Nucl. Mater. 210 (1994) 258.doi:10.1016/0022-3115(94)90180-5.

[91] J. E. Sinclair, P. C. Gehlen, R. G. Hoagland, J. P.Hirth, Flexible boundary conditions and nonlineargeometric effects in atomic dislocation modeling, J.Appl. Phys. 49 (1978) 3890–3897. doi:10.1063/1.

325395.[92] S. Rao, C. Hernandez, J. P. Simmons, T. A.

Parthasarathy, C. Woodward, Green’s functionboundary conditions in two-dimensional and three-dimensional atomistic simulations of dislocations, Phi-los. Mag. A 77 (1998) 231–256. doi:10.1080/

01418619808214240.[93] Y. Liu, G. Lu, Z. Chen, N. Kioussis, An improved

QM/MM approach for metals, Modelling Simul.Mater. Sci. Eng. 15 (2007) 275–284. doi:10.1088/

0965-0393/15/3/006.[94] X. Zhang, G. Lu, W. A. Curtin, Multiscale quan-

tum/atomistic coupling using constrained densityfunctional theory, Phys. Rev. B 87 (2013) 054113.doi:10.1103/PhysRevB.87.054113.

21

Page 23: Elastic modeling of point-defects and their interaction

[95] L. Huber, B. Grabowski, M. Militzer, J. Neugebauer,J. Rottler, A QM/MM approach for low-symmetry de-fects in metals, Comp. Mater. Sci. 118 (2016) 259–268.doi:10.1016/j.commatsci.2016.03.028.

[96] M. Leslie, N. J. Gillan, The energy and elastic dipoletensor of defects in ionic crystals calculated by thesupercell method, J. Phys. C Solid State 18 (1985)973–982. doi:10.1088/0022-3719/18/5/005.

[97] G. Makov, M. C. Payne, Periodic boundary conditionsin ab initio calculations, Phys. Rev. B 51 (1995) 4014–4022. doi:10.1103/PhysRevB.51.4014.

[98] S. E. Taylor, F. Bruneval, Understanding and cor-recting the spurious interactions in charged super-cells, Phys. Rev. B 84 (2011) 075155. doi:10.1103/

PhysRevB.84.075155.[99] C. Varvenne, F. Bruneval, M.-C. Marinica, E. Clouet,

Point defect modeling in materials: Coupling ab ini-tio and elasticity approaches, Phys. Rev. B 88 (2013)134102. doi:10.1103/PhysRevB.88.134102.

[100] W. Cai, V. V. Bulatov, J. Chang, J. Li, S. Yip,Periodic image effects in dislocation modelling,Philos. Mag. 83 (2003) 539–567. doi:10.1080/

0141861021000051109.[101] M.-C. Marinica, F. Willaime, J.-P. Crocombette,

Irradiation-induced formation of nanocrystallites withc15 laves phase structure in bcc iron, Phys. Rev. Lett.108 (2012) 025501. doi:10.1103/PhysRevLett.108.

025501.[102] R. C. Pasianot, Self-interstitials structure in the hcp

metals: A further perspective from first-principles cal-culations, J. Nucl. Mater. 481 (2016) 147–152. doi:

10.1016/j.jnucmat.2016.09.021.[103] M. Souissi, Y. Chen, M. H. Sluiter, H. Numakura,

Ab initio characterization of B, C, N, and O in bcciron: Solution and migration energies and elastic strainfields, Comp. Mater. Sci. 124 (2016) 249–258. doi:

10.1016/j.commatsci.2016.07.037.

22