9
Full length article The electric charge and climb of edge dislocations in perovskite oxides: The case of high-pressure MgSiO 3 bridgmanite P. Hirel a, * , P. Carrez a , E. Clouet b , P. Cordier a a Unit e Mat eriaux Et Transformations, B^ at. C6, Univ. Lille 1, 59655 Villeneuve d'Ascq, France b CEA, DEN, Service de Recherches de M etallurgie Physique, UPSay, F-91191 Gif-sur-Yvette, France article info Article history: Received 10 November 2015 Received in revised form 4 January 2016 Accepted 8 January 2016 Available online 26 January 2016 Keywords: Computer simulation Perovskites Dislocations Vacancies Climb abstract Dislocation climb is expected to play a major role in high-temperature creep of complex ionic oxides, however the fundamental mechanisms for climb are poorly understood in this class of materials. In this work we investigate the interaction of vacancies with an edge dislocation by means of atomic-scale simulations in high-pressure MgSiO 3 bridgmanite, a mineral whose natural conditions of deformation are favorable to climb. The evaluation of the interaction with oxygen vacancies reveals that the dislo- cation favors a non-stoichiometric, oxygen-poor conguration, associated with a positive electric charge of maximum value þ 9.17 10 11 Cm 1 . This result is in qualitative agreement with experimental ob- servations in related perovskite oxides. Subsequently, the interactions between dislocations and va- cancies are dominated by electrostatics, with binding enthalpies of several eV in the vicinity of the dislocation core, superseding elastic effects. These results shed a new light on dislocation-vacancy in- teractions and dislocation climb in this class of materials. © 2016 Acta Materialia Inc. Published by Elsevier Ltd. This is an open access article under the CC BY-NC- ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). 1. Introduction Dislocation climb is a non-conservative process that occurs when a dislocation absorbs or emits vacancies. This mechanism allows for dislocations with edge components to move out of their glide plane. Favoring unpinning, climb plays an essential role in high-temperature creep. Although theoretical models for climb were formulated early in the development of the theory of dislo- cations [1e4], actual experimental observation of climb has been very scarce and remains very challenging even today. Edelin and Poirier were among the rst to obtain a climb deformation rate by deforming single crystals of hexagonal close-packed magnesium and beryllium-copper alloys at high temperature, along the c axis to inhibit dislocation glide [5e7]. Recently climb has been demon- strated to be the prevalent deformation mechanism in icosahedral quasicrystals in response to the high lattice friction of highly corrugated glide plane in those structures [8,9]. Climb thus appears as an alternative to dislocation glide when this mechanism is inhibited. Since the seminal work of Eshelby, Newey, Pratt and Lidiard [10], it is admitted that dislocations in ionic materials should carry an electric charge, because of the difference in formation energy of anion and cation vacancies. Charged dislocations are then sur- rounded by a cloud of charge-compensating defects, called DebyeeHückel cloud. Direct evidence of charged dislocations were obtained in some materials with the rocksalt structure, e.g. LiF [11], NaCl [12], or doped KCl [13,14]. In perovskite-type materials, recent electron energy loss spectrum (EELS) measurements as well as atomic-scale simulations have revealed that dislocations tend to be oxygen-decient, e.g. in SrTiO 3 [15e19] or in BaTiO 3 [20], however their electric charge was not determined. Yet one can expect elec- trically charged dislocations to interact strongly with charged point defects, therefore the determination of the charge of dislocations is of primary importance to the understanding of the processes related to climb. Because climb is a slow process that depends on the mobility of vacancies, molecular dynamics simulations are usually not suited to investigate it because of the short time-scales accessible to this method. However static calculations can be utilized to quantify the interactions between vacancies and dislocations, and bring un- derstanding of the elementary processes of climb [21e23]. In this work we investigate dislocation-vacancy interactions, the elementary mechanism of dislocation climb, in bridgmanite, the main constituent of the Earth's lower mantle. The natural * Corresponding author. E-mail address: [email protected] (P. Hirel). Contents lists available at ScienceDirect Acta Materialia journal homepage: www.elsevier.com/locate/actamat http://dx.doi.org/10.1016/j.actamat.2016.01.019 1359-6454/© 2016 Acta Materialia Inc. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc- nd/4.0/). Acta Materialia 106 (2016) 313e321

The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

lable at ScienceDirect

Acta Materialia 106 (2016) 313e321

Contents lists avai

Acta Materialia

journal homepage: www.elsevier .com/locate/actamat

Full length article

The electric charge and climb of edge dislocations in perovskiteoxides: The case of high-pressure MgSiO3 bridgmanite

P. Hirel a, *, P. Carrez a, E. Clouet b, P. Cordier a

a Unit�e Mat�eriaux Et Transformations, Bat. C6, Univ. Lille 1, 59655 Villeneuve d'Ascq, Franceb CEA, DEN, Service de Recherches de M�etallurgie Physique, UPSay, F-91191 Gif-sur-Yvette, France

a r t i c l e i n f o

Article history:Received 10 November 2015Received in revised form4 January 2016Accepted 8 January 2016Available online 26 January 2016

Keywords:Computer simulationPerovskitesDislocationsVacanciesClimb

* Corresponding author.E-mail address: [email protected] (P. Hirel)

http://dx.doi.org/10.1016/j.actamat.2016.01.0191359-6454/© 2016 Acta Materialia Inc. Published by End/4.0/).

a b s t r a c t

Dislocation climb is expected to play a major role in high-temperature creep of complex ionic oxides,however the fundamental mechanisms for climb are poorly understood in this class of materials. In thiswork we investigate the interaction of vacancies with an edge dislocation by means of atomic-scalesimulations in high-pressure MgSiO3 bridgmanite, a mineral whose natural conditions of deformationare favorable to climb. The evaluation of the interaction with oxygen vacancies reveals that the dislo-cation favors a non-stoichiometric, oxygen-poor configuration, associated with a positive electric chargeof maximum value þ 9.1710�11 C m�1. This result is in qualitative agreement with experimental ob-servations in related perovskite oxides. Subsequently, the interactions between dislocations and va-cancies are dominated by electrostatics, with binding enthalpies of several eV in the vicinity of thedislocation core, superseding elastic effects. These results shed a new light on dislocation-vacancy in-teractions and dislocation climb in this class of materials.© 2016 Acta Materialia Inc. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-

ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

Dislocation climb is a non-conservative process that occurswhen a dislocation absorbs or emits vacancies. This mechanismallows for dislocations with edge components to move out of theirglide plane. Favoring unpinning, climb plays an essential role inhigh-temperature creep. Although theoretical models for climbwere formulated early in the development of the theory of dislo-cations [1e4], actual experimental observation of climb has beenvery scarce and remains very challenging even today. Edelin andPoirier were among the first to obtain a climb deformation rate bydeforming single crystals of hexagonal close-packed magnesiumand beryllium-copper alloys at high temperature, along the c axis toinhibit dislocation glide [5e7]. Recently climb has been demon-strated to be the prevalent deformation mechanism in icosahedralquasicrystals in response to the high lattice friction of highlycorrugated glide plane in those structures [8,9]. Climb thus appearsas an alternative to dislocation glide when this mechanism isinhibited.

Since the seminal work of Eshelby, Newey, Pratt and Lidiard [10],

.

lsevier Ltd. This is an open access

it is admitted that dislocations in ionic materials should carry anelectric charge, because of the difference in formation energy ofanion and cation vacancies. Charged dislocations are then sur-rounded by a cloud of charge-compensating defects, calledDebyeeHückel cloud. Direct evidence of charged dislocations wereobtained in some materials with the rocksalt structure, e.g. LiF [11],NaCl [12], or doped KCl [13,14]. In perovskite-type materials, recentelectron energy loss spectrum (EELS) measurements as well asatomic-scale simulations have revealed that dislocations tend to beoxygen-deficient, e.g. in SrTiO3 [15e19] or in BaTiO3 [20], howevertheir electric charge was not determined. Yet one can expect elec-trically charged dislocations to interact strongly with charged pointdefects, therefore the determination of the charge of dislocations isof primary importance to the understanding of the processesrelated to climb.

Because climb is a slow process that depends on the mobility ofvacancies, molecular dynamics simulations are usually not suited toinvestigate it because of the short time-scales accessible to thismethod. However static calculations can be utilized to quantify theinteractions between vacancies and dislocations, and bring un-derstanding of the elementary processes of climb [21e23]. In thiswork we investigate dislocation-vacancy interactions, theelementary mechanism of dislocation climb, in bridgmanite, themain constituent of the Earth's lower mantle. The natural

article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-

Page 2: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

P. Hirel et al. / Acta Materialia 106 (2016) 313e321314

conditions of deformation of this mineral are favorable to disloca-tion climb: grain size larger than 1 mm [24], high temperature(T¼ 1500�2500 K) and extremely slow strain rates(_ε ¼ 10�12 � 10�16 s�1). The study focuses on the [100](010) edgedislocation, which belongs to one of themost favorable slip systemsin bridgmanite and whose glide properties were already investi-gated [25e28]. It is shown that at finite temperature this disloca-tion favors an oxygen-deficient configuration, and carries anintrinsic positive electric charge. As a result the vacancies, whichare also electrically charged, interact strongly with the dislocationthrough the Coulomb interaction. Based on these results we pro-pose a scenario for dislocation climb in bridgmanite.

2. Methods and models

Bridgmanite is a high-pressure magnesium silicate, stable onlyin the pressure range 26e130 GPa, that crystallizes in a distortedperovskite structure with an orthorhombic lattice (Fig. 1a). Thechemical composition of natural bridgmanite is (Mg,Al,Fe)(Si,Al)O3,and in this work we focus on the pure magnesium-rich compound,MgSiO3. We describe it within the Pbnm space group, the occupiedWyckoff positions being (4c) for Mg ions, (4b) for Si ions, and (4c)and (8d) for O ions. All calculations presented in this work areperformed using a hydrostatic pressure 30 GPa corresponding tothe uppermost lower mantle (about 670e700 km below the Earth'ssurface).

Atomic systems are constructed with our home-made softwareAtomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS package [30]. Interatomic interactionsare described by a rigid-ion pair potential function where each ionhas a fixed electric charge, and ions interact through the Coulombforce and a short-range Buckingham potential. The parameteriza-tion by Alfredsson and coworkers is used, which produces accuratelattice constants, elastic constants, and surface energies, in thepressure range 30e140 GPa [31]. We choose this potential becausethe ion charges it uses follow the relationshipqMg¼�qO¼ 0.5qSi¼ 1.672e (where e is the elementary electriccharge), which ensures that the charge of dislocations does notarise from an incomplete compensation of the charges of ions. Weverified that this potential reproduces the generalized stacking

Fig. 1. (a) The unit cell of MgSiO3 bridgmanite. Mg ions are displayed as large orange sphereshown in transparent blue. In the [001] direction normal to the figure, two layers of Mg and Sof stoichiometry MgSiO, and another one of stoichiometry O2. (b) Sketch of the atomic syscations have different core structures, noted P and N. (c) Close-up of the atomic configuratioboundary (APB) inside the dislocation core is overlaid with a thick dashed line. On each sidedensity map around the P dislocation, as calculated with our atomic-to-continuum method (is the surface of an element of the grid. Areas where the charge density vanishes appear in grdensity appear in red, due to the missing oxygen ions in the incomplete octahedra. (d) Closecompared to the bulk. This dislocation contains four oxygen ions in excess (arrows) comparshows two zones of negative charge density in blue, due to the excess oxygen ions. (For inteweb version of this article.)

fault (GSF) energy with accuracy, and yields the same dislocationcore structures as the potential used in a previous work [27].Atomic systems are optimized by means of molecular statics sim-ulations employing a conjugate-gradient algorithm until reaching aresidual force on each atom smaller than 10�11 eV/Å.

Dislocations are constructed following a procedure similar tothe one proposed by Marrocchelli et al. in SrTiO3 [19], as illustratedin Fig. 1. The system is 3-D periodic and has a square shape, of sizeabout 150� 150� 6.7 Å3. Two dislocations of opposite Burgersvectors b¼±[100] are introduced so that the line joining themforms an angle of about 45+ with their glide planes to minimizetheir elastic interaction [32]. Because the two atomic planes acrossa (010) cut plane do not have the same stoichiometry, the termi-nation of the extra half-plane is not the same for the [100] and forthe ½100� dislocation. In the example shown in Fig. 1, the dislocationof positive Burgers vector has a termination MgSiO, i.e. it is oxygen-deficient compared to the bulk, and in the followingwill be referredto as the P variant. The dislocation of negative Burgers vector has atermination O2, i.e. enriched in oxygen compared to the bulk, andwill be referred to as the N variant. The core structures of thesedislocations are described in more details in the Results section.

In the framework of the chosen rigid-ion potential, a vacancycan only be created by removing an ion with its full charge, e.g.removing an oxygen ion will create an oxygen vacancy ofcharge þ1.672e. This approach is justified because ab initio calcu-lations have determined that the most energetically favorable statefor vacancies is to carry an uncompensated electric charge in otherperovskite materials, e.g. SrTiO3 [33], PbZrO3 [34], BaZrO3 [35],CaTiO3 [36]. By analogy we assume that the charged state is themost favorable in bridgmanite as well. The Coulomb interaction iscomputed with the Ewald summation method, which includes aself-interaction correction compensating for the interactions be-tween periodic replicas of a point charge [37e39], allowing toperform calculations with supercells that are not charge-neutral.

In order to evaluate the effective charge of dislocations, we usean atomic-to-continuum approach inspired by the Ewald summa-tion method. Atoms are projected in a plane normal to the dislo-cation line, and replaced by 2-D Gaussian functions over a finite-element grid. The magnitude of each Gaussian function is equalto the electric charge of the ion it replaces, and their variance s2¼ 9

s, Si ions as medium blue spheres, and oxygen as small red spheres. SiO6 octahedra areiO6 octahedra are superimposed. A (010) plane (dashed line) separates an atomic planetem constructed, containing a dipole of ±[100](010) edge dislocations. The two dislo-n of the P dislocation, which is oxygen-deficient compared to the bulk. The anti-phaseof the APB, incomplete, truncated SiO5 octahedra are recognizable. The electric chargesee text), is reported in the right-hand side. The charge density is in units of e/Swhere Seen, those where it is positive in red, and negative in blue. Two zones of positive charge-up view of the atomic configuration of the N dislocation, which is enriched in oxygened to the P dislocation. The electric charge density map (same color conventions as c)rpretation of the references to colour in this figure legend, the reader is referred to the

Page 3: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

P. Hirel et al. / Acta Materialia 106 (2016) 313e321 315

is chosen so that the Gaussian functions overlap and yield acontinuous zero charge density in the bulk material. WhereGaussian functions only partially overlap, the resulting net chargedensity is non-zero. When the supercell has a non-vanishing totalelectric charge, a continuous neutralizing background is introducedby shifting all values of the density, similarly to what is done in theEwald summation method. Then, the integration of the chargedensity around a defect provides an estimation of its charge. Thefinal result does not depend on small variations of the variance or ofthe size of elements of the grid. Moreover we verified that applyingthis approach to a single vacancy produces a net charge opposite tothat of the removed ion, e.g. removing a Mg ion results in a netcharge of �1.672e, which validates this approach.

The energy of interaction of a vacancy with the dislocations isdefined as:

DiE ¼ Edv � Ed � Ev � E0� �

(1)

where Edv is the total energy of the system containing the dislo-cations and the vacancy, Ed the energy of the system containingonly the dislocations, Ev the energy of the system containing onlythe vacancy, and E0 the energy of the ideal bulk system with nodefect. We verified that the system is large enough so that, whenthe vacancy is close to a dislocation, the effect of the seconddislocation is negligible. Note that DiE can also be interpreted as thevariation of the formation energy of the vacancy with respect to thebulk (defect-free) material. In order to account for the effects ofpressure, the enthalpy of interaction is defined as:

DiH ¼ DiE þ V0dP (2)

where V0 is the volume of the supercell and dP the variation of thepressure due to the vacancy.

3. Results

3.1. Atomic structures of the [100](010) dislocation

The system containing the dislocation dipole was optimized,and the atomic configurations of the P and N dislocations are re-ported in Fig. 1c and d, respectively. Both dislocations contain acharacteristic anti-phase boundary (APB), recognizable by the pairsof SiO6 octahedra connected by their edges in the glide plane(dashed lines), as opposed to the bulk where octahedra are con-nected by their corners. Despite their difference in compositionboth cores are compact and have a similar spreading of about 13 Å.On one hand, the P dislocation is characterized by incomplete,truncated octahedra on each side of the APB (Fig. 1c), due to thedeficiency of oxygen ions. On the other hand, the N dislocation isbordered by complete SiO6 octahedra, marked by thick arrows inFig. 1d. There are two octahedra per unit length of dislocation oneach side of the APB, therefore the N dislocation has four extraoxygen ions per unit line length compared to the P core.

We used our atomistic-to-continuum approach to estimate thenet electric charge of these dislocations. The electric charge den-sities are reported in Fig. 1c and d. As expected, the surroundingbulk material has a continuous zero charge density. The P disloca-tion (Fig. 1c) is characterized by two maxima of positive chargelocalized on the incomplete SiO5 octahedra. This reflects the factthat the positive charge of cations is not compensated because ofmissing oxygen ions. The integration of this charge density yields anet charge of þ0.384 e/L, where L is the unit length of the dislo-cation (equal to the lattice parameter c0¼ 6.7063 Å),i.e.þ9.17 10�11 C.m�1. Similarly theN dislocation is characterized by

two maxima of negative charge localized on the SiO6 polyhedra,obviously due to the extra oxygen ions. The integration yields a netcharge of �0.381 e/L, or �9.10 10�11 C.m�1, i.e. almost exactlyopposite to the charge of the P core, as one would expect from theterminations of the inserted half-planes.

3.2. Construction of charge-neutral dislocations

The N variant of the dislocation has four oxygen ions in excesswith respect to the P variant. Therefore, in order to construct twoequivalent, charge-neutral dislocation cores, two oxygen ions haveto be displaced from the N to the P dislocation. Among the possibleresulting configurations, the one with the lowest energy wasselected. The displacement of the two oxygen ions lowers the totalenergy of the system by about 17.75 eV, confirming that it is muchmore favorable. The two dislocations have then identical atomicconfigurations, represented in Fig. 2a and referred to as the Ovariant. Inside this dislocation, on each side of the APB there is onecomplete SiO6 and one incomplete SiO5 octahedra (arrows inFig. 2a).

The density of electric charge in the O dislocation is representedin Fig. 2b. It reveals a small distribution of negative charge at thecenter, surrounded by three regions of positive charge. These re-gions of non-vanishing electric charge are due to the displacementfield of the dislocation, which prevent the ions from overlappinglike in the bulk material. However the integration of the chargedensity is zero, and overall this dislocation is indeed charge-neutral.

3.3. Optimal stoichiometry and charge of the dislocation

At any finite temperature, vacancies naturally form in the ma-terial. Typically in ionic materials the anions and cations have verydifferent formation energies and mobilities. In bridgmanite thevacancies with the lowest formation and migration energies areoxygen vacancies [40,41], therefore they are the most mobile ones.When vacancies of all types exist in the material, oxygen vacancieswill segregate the most quickly into the dislocations.

We wanted to determine how many oxygen vacancies per unitline length the charge-neutral O dislocation can absorb, in order todetermine the optimal oxygen vacancy concentration, i.e. theoptimal atomic configuration, of the [100](010) edge dislocation.Starting from the O variant, we removed half a column of oxygenions (i.e. one oxygen ion out of two along the [001] direction),relaxed the system, and computed the corresponding energy ofinteraction using Eq. (2). By repeating this step for several positionsof the vacancy column, we determined if it was more favorable foroxygen vacancies to be in the bulk or in the vicinity of the dislo-cation. Then the configuration of lowest energy was selected (i.e.the one with the oxygen vacancies at the most favorable site), andthe procedure was repeated for a second oxygen vacancy. Note thatsince the two O dislocations in our setup are identical, the sameresults are obtained from both of them, but only one is presentedbelow.

The results are summarized in Fig. 2. Since the initial O dislo-cation is charge-neutral (Fig. 2b) one would expect the interactionto be driven by elastic effects, i.e. different interactions above andbelow the glide plane. Surprinzingly the interaction map in Fig. 2cshows that the dislocation is unequivocally attractive to oxygenvacancies. The interaction is quite short-ranged and vanishesquickly outside the dislocation core. It is more favorable for oxygenvacancies to be inside the dislocation core rather than in the sur-rounding bulk material, by �5.26 eV. The effect responsible for thispeculiar interaction is illustrated in Fig. 3. The (charged) row ofvacancies is responsible for a displacement of neighboring ions:

Page 4: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

Fig. 2. (a) Atomic structure of the O dislocation, obtained by removing two oxygen ions from the initial N dislocation (Fig. 1c), and placing them into the P dislocation (Fig. 1d). (b)The electric charge density around the O dislocation is characterized by a central region of negative charge (in blue) and three surrounding regions of positive charge (in red). Theintegration results in an overall charge-neutral dislocation. (c) Map of the enthalpy of interaction of a single oxygen vacancy with the dislocation of O type. Vanishing values appearin green, positive ones in red, and negative ones in blue. Only oxygen sites are represented (Mg and Si ions are omitted), and some oxygen octahedra are overlaid with black lines.(d) The reduced O dislocation, after it has absorbed one oxygen vacancy per unit line length. Because of the oxygen vacancy, one octahedron is incomplete on the left side of the APB.(e) The electric charge density of the reduced O dislocation. The oxygen vacancy is responsible for a zone of positive charge (in red). (f) Map of the enthalpy of interaction of thereduced O vacancy with a second oxygen vacancy (same color convention as c). Although the dislocation is positively-charged, negative binding enthalpies are still found in thevicinity of the core. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

P. Hirel et al. / Acta Materialia 106 (2016) 313e321316

oxygen ions are attracted to the row of vacancies, while cations arerepelled from it. In other words the vacancies are responsible for anionic polarization of the surrounding material. When the oxygenvacancies approach the dislocation, they induce displacements ofions in the dislocation core that modify the balance of the electriccharges. The effect is quite small in magnitude and can barely beseen with the naked eye. To better quantify this effect, wecomputed a weighted average of the positions of positive andnegative charges, using the density of electric charge as weights.Even so, the changes in polarization are small and are best observedwhen the vacancy is close to the dislocation. The Fig. 3 shows thepositions of positive and negative centers of charge of the dislo-cation for various locations of the row of oxygen vacancies. Nomatter the position of the vacancies, the dislocation always polar-izes with its negative center of charge oriented towards the va-cancy. This happens because the dislocation, although charge-neutral, contains regions of positive and negative charges that donot perfectly overlap, hence acting like a dipole. The vacancies areresponsible for an induced polarization of the dislocation, which isthe reason for the attraction of the vacancy to the dislocation.

Fig. 2d shows the atomic configuration of the dislocation afterremoval of the first oxygen row. This reduced dislocation hasgained an incomplete SiO5 octahedron (on the left side of the APB),responsible for a zone of positive electric charge. The total charge ofthe dislocation becomes positive, about þ3.82 10�11 C.m�1. In thisconfiguration, zones of positive and negative charges do not coin-cide, as shown in Fig. 2e (zones in red and blue, respectively), sothat the dislocation has the aspect of a line of electric dipoles.

Despite its positive charge, this dislocation is still found to beattractive to oxygen vacancies, as illustrated in Fig. 2f. Thepositively-charged oxygen vacancies are mainly attracted to thenegatively-charged zone of the dislocation. Inside the core thebinding enthalpy is �3.1 eV, i.e. weaker than in the case of the Odislocation.

Finally, after the removal of two columns of oxygen ions thedislocation is in the P configuration, as shown in Fig. 1c, and itselectric charge is 9.17 10�11 C.m�1. Because of this important posi-tive electric charge, the P dislocation is then completely repulsive tooxygen vacancies, as will be presented in the next section.

In conclusion, our results indicate that it is favorable for the[100](010) dislocation to be deficient in oxygen vacancies, hence tobear a positive electric charge. In the limit of a super-saturation ofoxygen vacancies, the dislocation can absorb up to two oxygenvacancies per unit line length compared to the charge-neutral Ocore. These results are consistent with studies found in the litera-ture showing that dislocations tend to be oxygen-poor in otherperovskite materials [15e18,20,42,19].

3.4. Interaction of the P dislocation with vacancies

In equilibrium with a population of oxygen vacancies, theoxygen-deficient configuration of the [100](010) edge dislocationappears to be the most favorable. However this dislocation iselectrically charged and, as presented in the introduction, it is ex-pected to be surrounded by a cloud of charge-compensating de-fects. In bridgmanite the positive charge of the P dislocation can be

Page 5: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

Fig. 3. Electric charge density when an oxygen vacancy is at various positions aroundthe O dislocation. The dotted circle indicates the position of the vacancy row, and the

P. Hirel et al. / Acta Materialia 106 (2016) 313e321 317

compensated by negatively-charged Mg and Si vacancies, andnaturally these charged defects will interact with the dislocation. Inorder to quantify this interaction, the system is duplicated along thedirection of the dislocation lines. A size corresponding to foursupercells along the dislocation line (24 Å) is found to give a goodconvergence of the interaction energies. A vacancy (Mg or Si) isintroduced, the atom positions are optimized with a conjugate-gradient algorithm, and the interaction with the dislocation isevaluated (Eq. (2)). The procedure is repeated for vacancy siteswithin a radius of 30 Å around the dislocation.

The maps of interaction of the P dislocation with vacancies arereported in Fig. 4. As stated earlier this dislocation is entirelyrepulsive to oxygen vacancies, since it is already saturated in oxy-gen vacancies. The interaction is radial, and much more long-rangethan before, because of the electric charge carried by thedislocation.

By opposition, the P dislocation is strongly attractive to cationvacancies. Again the interaction is radial and long-range because ofits electrostatic nature. A Mg vacancy has a binding enthalpy of11 eV when it is at the most favorable site inside the dislocationcore, while the Si vacancy has binding enthalpy of about 22 eV at itsmost favorable position. The factor of 2 between the two defectsreflects quite well the ratio of their charges q(VSi)/q(VMg)¼2, againdue to the electrostatic nature of the interaction.

Fig. 4b, e and h show the elastic contribution of pressure, i.e. theterm VdP in Eq. (2). When the vacancy is far from the dislocation, itsformation enthalpy has a value close to that of the bulk, thereforeVdP¼ 0. When it approaches the dislocation, the vacancy isresponsible for small variations of the enthalpy term, of the order of0.2 eV. It can be seen that the pressure in the system does notdependmuch on the position of the vacancies, and the contributionof this term is quite small compared to the electrostatic effects.

Finally, Fig. 5 shows the atomic structure of the P dislocationafter it has absorbed cation vacancies. After absorbing one Mg va-cancy the dislocation core becomes narrower, and the octahedraare not connected by their edges anymore. After absorbing a Sivacancy, the atomic configuration of the dislocation changes evenmore significantly: the stacking fault is spread in the (100) climbplane, as illustrated in Fig. 5b. This stacking fault is an APB, asevidenced by the SiO6 octahedra connected by their edges in the(100) climb plane. This configuration looks very similar to that ofanother dislocation, the [010](100) edge dislocation that also dis-sociates by climb when it is submitted to pressures above 80 GPa[27]. These results show that the absorption of cation vacanciesallow the dislocation to move out of its glide plane, and initiate theprocess of climb. They also show that the dislocation adoptsdifferent atomic configurations as it absorbs vacancies during itsclimb.

4. Discussion

4.1. The charge of the edge dislocation

According to Eshelby, in ionic materials at thermal equilibriumdislocations are electrically charged [10], because the formationenergies of cation and anion vacancies are different. These argu-ments were mainly based on the atomic structure and mechanicalproperties of binary ionic materials with rocksalt structure, like

4 and. signs mark the positions of the centers of charge of the dislocation. The colorconvention is the same as in Fig. 1. (a) Vacancy at the far top left of the dislocation. (b)Vacancy on the left side of the dislocation. (c) Vacancy at the right of the dislocation.(d) Vacancy below the dislocation. (For interpretation of the references to colour in thisfigure legend, the reader is referred to the web version of this article.)

Page 6: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

Fig. 4. Interaction of the P dislocation with (a,b,c) a single oxygen vacancy, (d,e,f) a Mg vacancy, (g,h,i) a Si vacancy, visualized with Atomeye [59]. The crystal orientation is identicalto the one in Fig. 1. The (010) glide plane of the dislocation is shownwith a horizontal dashed line, and a plane normal to the glide plane, used in the bottom graphs, is shownwith avertical dotted line. (a,d,g) Maps of the enthalpy of interaction (Eq. (2)). Only the possible lattice sites for the relevant vacancy are represented (other ions are omitted), and ions arecolored according to their binding enthalpy DiH. The interaction enthalpy is always positive for the oxygen vacancy, and always negative for the cation vacancies. (b,e,h) Contri-bution of the work of pressure. Ions are colored according to the variation VdP induced by their removal. (c,f,i) Graphs showing the enthalpy of interaction DiH when the vacancybelongs to the atomic plane above the glide plane (filled triangles), below the glide plane (empty triangles), and to the plane normal to the glide plane (circles). The thick, dashedblack lines are the results of the fits of Eq. (3).

Fig. 5. Atomic configuration of the P dislocation after it has absorbed (a) one Mg va-cancy; (b) one Si vacancy. The dashed line shows the plane of the stacking fault.

P. Hirel et al. / Acta Materialia 106 (2016) 313e321318

NaCl or MgO. In the easiest 1=2½110�ð110Þ slip system of thesematerials, pure edge dislocations lie along a [001] axis, i.e. thedislocation line is along a row of alternating positive and negative

ions [32]. Therefore by construction, in a charge-neutral dislocationthe electric charges compensate perfectly and the density of chargeis zero everywhere. Consequently, the Coulomb interaction van-ishes and the interaction with vacancies is driven by elastic effects,as it was shown in MgO [43e47], NaCl [48], or NiO [49]. Thesesimulations were performed on dislocations that were constructedto be charge-neutral, without taking temperature into account. Inthe rocksalt structure the absorption of one vacancy produces a pairof jogs along the dislocation line, and because of the ionic nature ofthe material the jogs are electrically charged [32]. For instance theremoval of one Naþ ion along a 1=2½110�ð110Þ dislocation in NaClproduces a pair of jogs of negative sign. In Eshelby's model, at finitetemperature the difference in the formation energy of anions andcations results in the dislocation to have more jogs of one sign thanthe other, naturally causing the dislocation to acquire an electriccharge.

In a ternary perovskite oxide such as bridgmanite, the atomicstructure is much more complex. Even when the dislocation ischarge-neutral the electric charges do not compensate perfectly,and zones of positive and negative electric charges still exist asillustrated in Fig. 2b. Electrostatic effects are the dominant term inthe interactionwith vacancies (Fig. 3). In bridgmanite there seem tobe no configuration where the charge-neutral dislocation wouldinteract only elastically with charged vacancies. Because of thisinteraction, at any finite temperature the dislocation attracts oxy-gen vacancies and becomes positively charged. This situation isdistinct from the one described by Eshelby: in materials with therocksalt structure the absorption of one vacancy is sufficient tomove the dislocation out of its glide plane, i.e. to create a pair of jogsand make it climb. On the contrary, in bridgmanite (and otherperovskite materials) the absorption of one or even two oxygenvacancies does not form jogs, and does notmove the dislocation outof its glide plane. Indeed the dislocation spreads in the same planewhether it is enriched or depleted in oxygen, i.e. whether it is in the

Page 7: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

P. Hirel et al. / Acta Materialia 106 (2016) 313e321 319

N, O, or P configuration, as shown in Figs. 1c,d and 2a. The charge ofthe dislocation is not due to the formation of jogs like in the rocksaltstructure, but only to the concentration of oxygen vacancies insidethe dislocation.

Our results indicate that the oxygen-deficient, positively-charged dislocation core is themost favorable one in bridgmanite. Ifthe material is saturated with oxygen vacancies the dislocationcharge can reach a maximum of about þ9.17 10�11 C.m�1. Experi-mentally, we propose that it may be possible to determine thecharge of edge dislocations in bridgmanite (or in an analogueperovskite) by measuring their velocity under an applied electricfield, like it was done in other ionic materials [13,14].

4.2. Interaction with vacancies

The positively-charged dislocation is attractive to cation va-cancies, which will slowly migrate towards the dislocation core,driven by electrostatic interaction. If we approximate the disloca-tion as a straight line of lineic charge density l and the vacancy as apoint charge q, their interaction is defined as the sum of elastic andelectrostatic contributions [10,32]:

Eint ¼ Eelas þ ECoul ¼mb3p

1þ n

1� n

y2

x2 þ y2dV þ�2ql

ε

ln rð Þ þ cst:

(3)

where m is the shearmodulus of thematerial, n its Poisson ratio, ε itsrelative permittivity, b the magnitude of the Burgers vector of thedislocation, x and y the position of the vacancy (i.e. considering thatthe dislocation is located at x¼ 0,y¼ 0), and r ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðx2 þ y2Þ

pthe

dislocation-vacancy distance.In order to confirm the value of the dislocation charge given by

our atomic-to-continuum approach, we fitted Eq. (3) to the inter-action energy profiles obtained from atomistic calculations for theMg and Si vacancies. The only fit parameter was the dislocationcharge l, and the resulting fitted functions are reported as thickdashed lines in Fig. 4c, f and i. The obtained values arel¼3.37 10�10 C.m�1 when fitting to the interaction enthalpies ofthe oxygen vacancy (q(VO)¼þ1.672e), l¼6.68 10�10 C.m�1 using thedata for the Mg vacancy (q(VMg)¼�1.672e), andl¼5.01 10�10 C.m�1 using the data for the Si vacancy (q(VSi)¼�3.344e). These values are of the same order of magnitude as ouratomic-to-continuummethod, the best match being obtained fromthe fit to data for the oxygen vacancy. The discrepancy can beattributed to the divergence of the model when the dislocation-vacancy distance r/0, while in atomistic simulations the energyof the system is always finite. As a result the obtained value of ldepends critically on the data points included in the fitting pro-cedure. Including or excluding data points at distances smaller than5 Å can result in large variations of the estimated dislocationcharge. In addition, the Eq. (3) does not account for second-ordereffects like dipole-point charge interactions, which may play animportant role as in the case of the O dislocation (Fig. 3). For thesereasons we estimate that the dislocation charge obtained with ouratomic-to-continuum approach is more reliable than the one ob-tained by fitting the simple Eq. (3). Nonetheless the latter enhancesthe confidence on the order of magnitude of the dislocation charge,both qualitatively and quantitatively.

4.3. Consequences for dislocation climb

As introduced by Eshelby, a charged dislocation in an ionicmaterial is surrounded by an atmosphere of vacancies of oppositecharge, the DebyeeHückel cloud [10,32]. Our results show that in

bridgmanite the dislocation is positively charged, therefore theDebyeeHückel cloud is composed of negatively-charged Mg and Sivacancies. These cation vacancies are attracted to the chargeddislocation, and eventually they also segregate into the dislocationcore. This has two effects. First, the absorption of negatively-charged vacancies modifies the electric charge of the dislocation,making it charge-neutral or negatively charged. Our results showthat in both cases the dislocation becomes attractive to oxygenvacancies again. By sequentially absorbing oxygen and cation va-cancies the dislocation alternates between positive and negativecharge. The second effect of the absorption of cation vacancies is tomodify the plane inwhich the dislocation spreads. In other words itallows the dislocation to climb. After absorbing a complete formulaunit MgSiO3, a segment of the dislocation can climb by one latticevector.

To summarize, we propose that dislocation climb in bridg-manite, and possibly in other perovskite materials, occurs by thesequential absorption of cation and oxygen vacancies, as illustratedin Fig. 6. The absorption of oxygen vacancies has the role ofmaintaining the positive charge of the dislocation: indeed if thedislocation is charge-neutral or negatively charged it attracts oxy-gen vacancies, however their absorption does not change the planewhere the dislocation spreads. Subsequently, the positively-charged dislocation attracts cation vacancies, effectively movingthe dislocation out of its glide plane and changing its electriccharge. The dislocation then becomes attractive to oxygen va-cancies again. The cyclic repetition of this sequence allows thedislocation to climb.

Furthermore, the predominance of the Coulomb interaction hasa direct consequence on the magnitude of the dislocation-vacancyinteractions. In materials where this interaction is governed byelastic effects, the interaction energies are of the order of 0.1e0.5 eV[50,51]. By contrast, in bridgmanite we find energies of the order ofseveral eV to more than 10 eV, i.e. up to two orders of magnitudelarger. As a result, the kinetics of dislocation climb in bridgmanitemay also be significantly faster than it is in metallic systems.

Previous atomic-scale calculations have shown that bridgman-ite is a high lattice friction material under high-pressure, makingdislocation glide difficult [27,28]. Moreover the DebyeeHückelcloud is responsible for a back-stress acting on the dislocation, sothat the minimum stress required to move the dislocation is largerthan the Peierls stress predicted by considering an isolated dislo-cation [10].We expect this to be true also in bridgmanite, where theminimum stress required to move a dislocation (in Ref. [27]) wasprobably underestimated because it did not take into account theback stress due to a cloud. In the natural conditions of the Earth'slower mantle the deformation occurs under extremely slow strainrates (_ε ¼ 10�12 � 10�16 s�1), i.e. the applied stresses are extremelysmall. Also, the Earth's lower mantle is expected to be an oxygen-reducing environment [52], thus favoring the existence ofoxygen-poor dislocations rather than oxygen-enriched ones. Insuch conditions one can expect dislocation glide to be a rare event,and dislocations to stay immobile for very long times. This givestime for the cation vacancies to migrate towards the dislocationcores, despite their low diffusivity [41]. In such extreme high-temperature creep conditions dislocation climb may play animportant role.

It remains difficult to give an estimate of the climb velocity withthe available data. Other studies however, seem to corroborate thefact that dislocation climb can occur easily in perovskite materials.High-temperature creep experiments in BaTiO3 [53] and CaTiO3[54,55] revealed the formation of features named “scallops” alongdislocation lines, which were interpreted as climb-dissociateddislocation loops. The activation of climb at so short time scalessuggests an efficient mechanism for the absorption of vacancies by

Page 8: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

Fig. 6. Scenario for dislocation climb. (a) A charge-neutral dislocation surrounded by vacancies of positive and negative charges. The positively-charged oxygen vacancies migratefaster than cation vacancies towards the dislocation. (b) After absorbing neighboring oxygen vacancies the dislocation is positively charged (red) and surrounded by a space-chargeregion (dashed line), where only cation vacancies remain and form the DebyeeHückel cloud. Cation vacancies migrate slowly towards the dislocation core. (c) After absorbing cationvacancies the dislocation climbs, and becomes attractive to oxygen vacancies again. (For interpretation of the references to colour in this figure legend, the reader is referred to theweb version of this article.)

P. Hirel et al. / Acta Materialia 106 (2016) 313e321320

the dislocations, which could be explained by electrostatic in-teractions. Another supporting experimental fact is the ductile tobrittle transition around 1050 K in bulk strontium titanate SrTiO3,attributed to a change in the core structure of the ⟨110⟩ dislocations[56,57], which are the structural equivalent to the [100] disloca-tions in bridgmanite. The sudden loss of this slip system at 1050 Kcontrasts with the fact that these dislocations dissociate in theirglide plane and are responsible for the ductility at lower temper-ature [58]. Since the increase of temperature is accompanied withan increase in the concentration and mobility of vacancies, bothenhancing dislocation climb, it is possible that this dislocation ischarged and absorbs vacancies. Indeed, recent atomistic simula-tions have pointed to the predominance of electrostatic effects indislocation-vacancy interactions in SrTiO3 [19], supporting theconclusions of the present work.

5. Conclusion

Atomic-scale calculations were utilized to determine the electriccharge of the [100](010) edge dislocation in MgSiO3 perovskite at30 GPa, and quantify its interaction with intrinsic vacancies. Weshow that [100](010) edge dislocations in bridgmanite favor anoxygen-deficient composition and carry a positive electric charge.As a result the interaction with vacancies is dominated by theCoulomb interaction rather than by elastic effects. Because of itselectric charge, the dislocation alternatively attracts defects ofnegative and positive electric charges, allowing it to climb.

These findings modify significantly the understanding of dislo-cation climb in bridgmanite, and in perovskite materials in general.The climb velocity of dislocations in this class of materials may bemuch faster than in non-ionic systems. It appears necessary topursue this work in order to obtain estimates of climb velocities.

Acknowledgements

This work was supported by funding from the EuropeanResearch Council under the Seventh Framework Programme (FP7),ERC grant N.290424 - RheoMan.

References

[1] N. Mott, The mechanical properties of metals, Proc. Phys. Soc. B 64 (1951) 729.[2] J. Weertman, Theory of steady-state creep based on dislocation climb, J. Appl.

Phys. 26 (1955) 1213e1217.[3] J. Lothe, Theory of dislocation climb in metals, J. Appl. Phys. 31 (1960)

1077e1087.[4] F. Nabarro, Steady-state diffusional creep, Philos, Mag 16 (1967) 231e237.[5] G. Edelin, J. Poirier, Etude de la mont�ee des dislocations au moyen

d'exp�eriences de fluage par diffusion dans le magn�esium - I. M�ecanisme ded�eformation, Philos. Mag. 28 (1973) 1203e1210.

[6] G. Edelin, J. Poirier, Etude de la mont�ee des dislocations au moyen

d'exp�eriences de fluage par diffusion dans le magn�esium - II. Mesure de lavitesse de mont�ee, Philos. Mag. 28 (1973) 1211e1223.

[7] R. Le Hazif, G. Edelin, J.-M. Dupouy, Diffusion creep by dislocation climb inbeyllium and Be-Cu single crystals, Metall. Trans. 4 (1973) 1275e1281.

[8] F. Mompiou, D. Caillard, Dislocation-climb plasticity: modelling and compar-ison with the mechanical properties of icosahedral AlPdMn, Acta Mater 56(2008) 2262e2271.

[9] F. Mompiou, D. Caillard, Dislocations and mechanical properties of icosahedralquasicrystals, C.R. Phys. 15 (2014) 82e89.

[10] J. Eshelby, C. Newey, P. Pratt, A. Lidiard, Charged dislocations and the strengthof ionic crystals, Philos. Mag. 3 (1958) 75.

[11] R. Sproull, Charged dislocations in lithium fluoride, Philos. Mag. 5 (1960)815e831.

[12] R. Davidge, The sign of charged dislocations in NaCl, Philos. Mag. 8 (1963)1369e1377.

[13] L. Colombo, T. Kataoka, J. Li, Movement of edge dislocations in KCl by largeelectric fields, Philos. Mag. A 46 (1982) 211e215.

[14] T. Kataoka, L. Colombo, J. Li, Direct measurements of dislocation charges inCa2þ-doped KCl by using large electric fields, Philos. Mag. A 49 (1984)395e407.

[15] Z. Zhang, W. Sigle, W. Kurtz, M. Rühle, Electronic and atomic structure of adissociated dislocation in SrTiO3, Phys. Rev. B 66 (2002) 214112.

[16] Z. Zhang, W. Sigle, M. Rühle, Atomic and electronic characterization of the a[100] dislocation core in SrTiO3, Phys. Rev. B 66 (2002) 094108.

[17] Z. Zhang, W. Sigle, W. Kurtz, HRTEM and EELS study of screw dislocation coresin SrTiO3, Phys. Rev. B 69 (2004) 144103.

[18] C. Jia, A. Thust, K. Urban, Atomic-scale analysis of the oxygen configuration ata SrTiO3 dislocation core, Phys. Rev. Lett. 95 (2005) 225506.

[19] D. Marrocchelli, L. Sun, B. Yildiz, Dislocations in SrTiO3: easy to reduce but notso fast for oxygen transport, J. Am. Chem. Soc. 137 (2015) 4735e4748.

[20] H. Kurata, S. Isojima, M. Kawai, Y. Shimakawa, S. Isoda, Local analysis of theedge dislocation core in BaTiO3 thin film by STEM-EELS, J. Microsc. 236 (2009)128e131.

[21] T. Lau, X. Lin, S. Yip, K. Van Vliet, Atomistic examination of the unit processesand vacancy-dislocation interaction in dislocation climb, Scr. Mater 60 (2009)399e402.

[22] M. Kabir, T. Lau, D. Rodney, S. Yip, K. Van Vliet, Predicting dislocation climband creep from explicit atomistic details, Phys. Rev. Lett. 105 (2010) 095501.

[23] E. Clouet, Predicting dislocation climb: classical modeling versus atomisticsimulations, Phys. Rev. B 84 (2011) 092106.

[24] V. Solomatov, C. Reese, Grain size variations in the Earth's mantle and theevolution of primordial chemical heterogeneities, J. Geophys. Res. 113 (2008)B07408.

[25] P. Cordier, T. Ung�ar, L. Zsoldos, G. Tichy, Dislocation creep in MgSiO3 perov-skite at conditions of the Earth's uppermost lower mantle, Nature 428 (2004)837e840.

[26] D. Mainprice, A. Tommasi, D. Ferr�e, P. Carrez, P. Cordier, Predicted glide sys-tems and crystal preferred orientations of polycrystalline silicate Mg-Perovskite at high pressure: Implications for the seismic anisotropy in thelower mantle, Earth Planet. Sci. Lett. 271 (2008) 135e144.

[27] P. Hirel, A. Kraych, P. Carrez, P. Cordier, Atomic core structure and mobility of[100](010) and [010](100) dislocations in MgSiO3 perovskite, Acta Mater 79(2014) 117e125.

[28] A. Kraych, Ph. Carrez, P. Hirel, E. Clouet, and P. Cordier, Phys. Rev. B, https://journals.aps.org/prb/accepted/26073O06T6818c3ed21800f5ada7b75c4d40f35af (accepted).

[29] P. Hirel, Atomsk: a code for manipulating and converting atomic data files,Comput. Phys. Comm. 197 (2015) 212e219.

[30] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics,J. Comp. Phys. 117 (1995) 1e19. URL, http://lammps.sandia.gov.

[31] M. Alfredsson, J. Brodholt, D. Dobson, A. Oganov, C. Catlow, S. Parker, G. Price,Crystal morphology and surface structures of orthorhombic MgSiO3 perov-skite, Phys. Chem. Miner. 31 (2005) 671e682.

[32] J. Hirth, J. Lothe, Theory of Dislocations, second ed., McGraw-Hill Publ. Co.,

Page 9: The electric charge and climb of edge dislocations in perovskite oxides… · 2017. 3. 1. · Atomsk [29], and classical molecular statics simulations are per-formed with the LAMMPS

P. Hirel et al. / Acta Materialia 106 (2016) 313e321 321

New York, 1982.[33] J. Buban, H. Iddir, S. €O�güt, Structural and electronic properties of oxygen va-

cancies in cubic and antiferrodistortive phases of SrTiO3, Phys. Rev. B 69(2004) 180102.

[34] Z. Zhang, P. Wu, L. Lu, C. Shu, Study on vacancy formation in ferroelectricPbTiO3 from ab initio, Appl. Phys. Lett. 88 (2006) 142902.

[35] P. Sundell, M. Bj€orketun, G. Wahnstr€om, Thermodynamics of doping and va-cancy formation in BaZrO3 perovskite oxide from density functional calcula-tions, Phys. Rev. B 73 (2006) 104112.

[36] H. Lee, T. Mizoguchi, T. Yamamoto, Y. Ikuhara, First principles study onintrinsic vacancies in cubic and orthorhombic CaTiO3, Mater. Trans. 50 (2009)977e983.

[37] P. Ewald, Die Berechnung optischer und elektrostatischer Gitterpotentiale,Ann. Phys. 64 (1921) 253.

[38] D. Adams, G. Dubey, Taming the Ewald sum in the computer simulation ofcharged systems, J. Comp. Phys. 72 (1987) 156e176.

[39] A. Toukmaji, J. Board Jr., Ewald summation techniques in perspective: a sur-vey, Comp. Phys. Comm. 95 (1996) 73e92.

[40] D. Dobson, R. Dohmen, M. Wiedenbeck, Self-diffusion of oxygen and silicon inMgSiO3 perovskite, Earth Planet. Sci. Lett. 270 (2008) 125e129.

[41] M. Ammann, J. Brodholt, D. Dobson, DFT study of migration enthalpies inMgSiO3 perovskite, Phys. Chem. Miner. 36 (2009) 151e158.

[42] V. Metlenko, A. Ramadan, F. Gunkel, H. Du, H. Schraknepper, S. HoffmannEifert, R. Dittmann, R. Waser, A. De Souza, Do dislocations act as atomic au-tobahns for oxygen in the perovskite oxide SrTiO3? Nanoscale 6 (2014)12864e12876.

[43] C. Woo, M. Puls, M. Norgett, Energies of formation of point defects in perfectand dislocated ionic crystals, J. Phys. 12 (1976) 557.

[44] C. Woo, M. Puls, Atomistic breathing shell model calculations of dislocationcore configurations in ionic crystals, Philos. Mag. 35 (1977) 727e756.

[45] M. Puls, C. So, The core structure of an edge dislocation in NaCl,, Phys. Stat. Sol.98 (1980) 87e96.

[46] J. Rabier, M. Puls, Atomistic model calculations of pipe-diffusion mechanisms

in MgO, Philos. Mag. 52 (1985) 461e473.[47] F. Zhang, A. Walker, K. Wright, J. Gale, Defects and dislocations in MgO: atomic

scale models of impurity segregation and fast pipe diffusion, J. Mater. Chem.20 (2010) 10445e10451.

[48] J. Rabier, M. Puls, Atomistic calculations of point-defect interaction andmigration energies in the core of an edge dislocation in NaCl, Philos. Mag. A 59(1989) 533e546.

[49] J. Rabier, J. Soullard, M. Puls, Atomistic calculations of point-defect in-teractions with an edge dislocation in NiO, Philos. Mag. A 61 (1990) 99e108.

[50] E. Clouet, The vacancy-edge dislocation interaction in fcc metals: a compari-son between atomic simulations and elasticity theory, Acta Mater 54 (2006)3543e3552.

[51] Z. Chen, M. Mrovec, P. Gumbsch, Dislocation-vacancy interactions in tungsten,Modelling Simul, Mater. Sci. Eng. 19 (2011) 074002.

[52] D. Frost, C. McCammon, The redox state of Earth's mantle,, Annu. Rev. EarthPlanet. Sci. 36 (2008) 389e420.

[53] N. Doukhan, J. Doukhan, Dislocations in perovskites BaTiO3 and CaTiO3, Phys.Chem. Miner. 13 (1986) 403e410.

[54] K. Wright, G. Price, J.-P. Poirier, High-temperature creep of the perovskitesCaTiO3 and NaNbO3, Phys. Earth Planet. Int. 74 (1992) 9e22.

[55] P. Besson, J. Poirier, G. Price, Dislocations in CaTiO3 perovskite deformed athigh-temperature: a transmission electron microscopy study, Phys. Chem.Miner. 23 (1996) 337e344.

[56] P. Gumbsch, S. Taeri Baghbadrani, D. Brunner, W. Sigle, M. Rühle, Plasticityand an inverse brittle-to-ductile transition in strontium titanate, Phys. Rev.Lett. 87 (2001) 085505.

[57] W. Sigle, C. Sarbu, D. Brunner, M. Rühle, Dislocations in plastically deformedSrTiO3, Philos. Mag. 86 (2006) 4809e4821.

[58] P. Hirel, M. Mrovec, C. Els€asser, Atomistic study of ⟨110⟩ dislocations instrontium titanate, Acta Mater 60 (2012) 329e338.

[59] J. Li, Atomeye: an efficient atomistic configuration viewer, Modelling Simul.Mater. Sci. Eng. 11 (2003) 173.