• Aucun résultat trouvé

Effect of interface dislocation Burgers vectors on elastic fields in anisotropic bicrystals

N/A
N/A
Protected

Academic year: 2021

Partager "Effect of interface dislocation Burgers vectors on elastic fields in anisotropic bicrystals"

Copied!
12
0
0

Texte intégral

(1)

Effect of interface dislocation Burgers vectors

on elastic fields in anisotropic bicrystals

The MIT Faculty has made this article openly available.

Please share

how this access benefits you. Your story matters.

Citation

Vattré, A.J., and M.J. Demkowicz. “Effect of Interface Dislocation

Burgers Vectors on Elastic Fields in Anisotropic Bicrystals.”

Computational Materials Science 88 (June 2014): 110–115.

As Published

http://dx.doi.org/10.1016/j.commatsci.2014.02.044

Publisher

Elsevier

Version

Author's final manuscript

Citable link

http://hdl.handle.net/1721.1/103929

Terms of Use

Creative Commons Attribution-NonCommercial-NoDerivs License

(2)

Effect of interface dislocation Burgers vectors on elastic fields in

anisotropic bicrystals

A.J. Vattr´ea,∗, M.J. Demkowiczb

aCEA, DAM, DIF, F-91297 Arpajon, France

bMIT, Department of Materials Science and Engineering, Cambridge MA, 02139

Abstract

A recent anisotropic elasticity formalism for quantifying interface dislocation arrays is used to compute the variationsin long-range elastic fields and short-range strain energiesof misfit dislocations due to changes in the Burgers vectors of the interface dislocations. The importance of selecting proper reference states for tilt and twist grain boundaries as well as a heterophase interface formed by tetragonal crystals is discussed in terms of partitioning of strain and rotation fields. For constrained interfaces consistent with the Frank-Bilby equation, the present work shows that examining the strain energies of different admissible dislocation configurations may be used as a criterion for predicting the most favorable structures.

Keywords: Grain boundary, semi-coherent interface, anisotropic elasticity theory, dislocation arrays, interface energy

1. Introduction

Grain boundaries and heterophase interfaces in polycrystalline solids are barriers to dislocation motion [1] as well as sinks [2] and sources for point defects [3]. Experimental [4], theoretical [5] and modeling [6] investigations have demonstrated that these processes depend on the interface crystallographic character and, in the case of semicoherent interfaces, on their dislocation structures. There has been extensive work on predicting dislocation structures of general semicoherent interfaces formed by joining two neighboring crystals [7–9]. One widely used approach describes interface dislocation structures by applying the closely related Frank-Bilby [10, 11] and O-lattice [12, 13] techniques. Both procedures require the selection of a ”reference state”, within which the Burgers vectors of individual dislocations are defined. Partitioning of elastic distortions between the adjoining crystals at the interface is a critical part of finding the proper reference state. The importance of partitioning of elastic fields and selecting proper reference states have been recently discussed using elasticity theory in isotropic [14–16] and anisotropic bicrystals [16].

The purpose of the present work is to investigate, using heterogeneous anisotropic elasticity theory, variations inlong-range elastic fields and short-range strain energiesof misfit dislocations due to changes in the Burgers vectors of the interface dislocations.The reference state of the median lattice suggested by Frank [10] has been successfully applied to grain boundaries, symmetrically disposed between the two rotated crystals [1, 9]. This reference state would therefore apply to pure symmetric tilt or pure symmetric twist boundaries, with equal partitioning of the rotations. However, it would not apply in general: for instance,

(3)

F A F B Reference state Coherency stresses Volterra dislocations

with far-field stresses

+

Natural state

Interface dislocations free of far-field stresses

=

Figure 1: Mapping from a coherent reference state to the natural state using displacement gradientsAF andBF. Volterra dislocations

introduced into the reference state remove coherency stresses and may change the misorientation of the neighboring crystals [16].

for asymmetrical boundaries or heterophase interfaces, for which early interpretations of the Frank-Bilby equation claimed considerable freedom in selecting the Burgers vectors of interface misfit dislocations, even

going so far as to suggest that they are arbitrary [12, 17]. The present work illustrates the effect of improper

selection of interface dislocation Burgers vectors on grain boundaries in Cu. Furthermore, some interfaces may have multiple admissible misfit dislocation configurations that meet constraints of crystallography and vanishing far-field stresses. For such interfaces, examining the strain energies of the different structures may be used as a criterion for predicting the most likely one. We illustrate this case on an interface formed by tetragonal L10-ordered FePd bicrystals. Assuming zero far-field stresses, the complete elastic field solutions

are used to compute the strain energies associated with equilibrium dislocation structure for this interface.

2. Theory of interface dislocations

We follow the procedure for computing far-field elastic distortions described in [16]. Starting from a coherent reference state, materials A and B are mapped separately into new configurations that yield an interface with a prescribed crystallography, as illustrated in Fig. (1). For a grain boundary, the maps applied to materials A and B are proper rotations while for a pure misfit interface they are pure strains. To account for general heterophase interfaces, whose misorientations may include both rotations and strains, the maps are described by homogeneous displacement gradientsAF andBF from the reference state into

the ”natural state”. Interface dislocations may be modeled as Volterra dislocations that are introduced in the reference state for constrained interfaces [16]: these dislocations are created by the Volterra process in which the uniform displacement across a cut is a translation Burgers vector of the reference state. Such arrays of Volterra dislocations (sometimes so-called ”extrinsic” [18], ”coherency” [19] or ”stress-generator” [9] dislocations) produce a staircase disregistry [20]. In our previous work [16], the Volterra dislocations have

been distinguished from ”interface dislocations” that generate a sawtooth disregistry. In general, arrays of

Volterra dislocations have non-zero far-field strains, rotations, or both. For equilibrium arrays, it is therefore required that the far-field stresses due to these dislocationsAσ∞dis andBσ∞dis are equal and opposite to the

coherency stressesAσcandBσcin the reference state [15, 16]. This requirement leads to the removal of all

far-field stresses in the natural state:

Aσc+Aσ

dis= 0 and, Bσc+Bσ

dis= 0 . (1)

Although a variety of shapes of interface dislocation networks have been observed [21], the present work is restricted to arrays that may be represented by j ≤ 2 arrays of parallel dislocations with Burgers vectors bi, line directions ξi, and inter-dislocation spacings di. An interface containing only one array of straight

(4)

types are distinguished directly by the number of sets of dislocations (e.g. stated from experimental inves-tigations or atomistic calculations) and the rank of the matrix T (in eq. (2)) [9, 22]. However, topological dislocation changes that might result in recombination or zipping the dislocation segments into a junction are not able to handle: this may be managed by using nodal dislocation dynamics simulations dedicated to interface dislocations, for which local rules for approximating dislocation reactions are necessary to take into account [23].

Following previous investigators [9–11], the quantities {bi, ξi, di} may be related to the resultant Burgers

vector of admissible Volterra dislocations B in the reference state and interface crystallography as B = j

i=1  n × ξi di · p  bi= AF −1 BF −1 | {z } T  p = (ADc−BDc) p , (2)

where n is a unit vector normal to the interface and the so-called probe vector p is any vector contained within the interface plane. Eq. (2) is known as the quantized Frank-Bilby equation [9, 22], where Dcare regarded

as the coherency distortion fields, that are needed to maintain the (coherent) reference state [15, 16]. The distortions presented in eq. (2) are purely elastic, so that the interface coherency strain relaxation that might be induced by plastic deformation or phase transformation (i.e. inelastic strains) is not included in the present formalism. However, the relaxation processes might be accomplished by incorporating kinetic treatments of dislocations and interfaces [24]. Here, the elastic coherency strains that align crystals A and B in the reference state (see Fig. 1) are generated in the whole bicrystal to produce homogeneous biaxial distortions parallel to the interface of interest. This operation leads to a mechanical force equilibrium in the coherent bicrystal that is not explicitly taken into account in the Olson-Cohen-Bonnet approaches [19, 25, 26], for which the coherency strains are concentrated in the vicinity of the interfaces.

The complete elastic distortion field Dtotmay be written as the superposition of the uniform coherency

and the Volterra dislocation distortions, denoted by Ddis. Due to the periodicity of the interface dislocation

structures, described by the two O-lattice vectors po16= po

2in a Cartesian coordinate system with basis vectors

(x1, x2k n, x3) and with the interface located at x2= 0, the complete distortion field may be expressed,

outside of dislocation cores, as the biperiodic Fourier series at any position x [16, 25, 26], i.e. Dtot(x) = Dc+ Ddis(x) = Dc+

k 6= 0

ei2πk · rDk(x2) , (3)

where i =√−1 and the sum spans over all non-zero wavevectors k. The Fourier amplitudes of the complete distortion waves Dk(x2) are required to converge (not necessary to zero) in the far-field, i.e. x2→ ±∞. The

components k1and k3of the wavevector k are related to the geometry of the interface dislocation structures

[16] and satisfy: k · r = k1x1+ k3x3=  n csc φ | po 1| −m ctg φ | po 2|  x1+ m | po 2| x3, (4)

with φ the oriented angle between ξ2and ξ1. From eq. (3), the far-field distortion D∞

disproduced by Volterra

dislocations is obtained by adding the individual contributions of each dislocation sets with x2→ ±∞ and

then computing the sum over n and m from −∞ to ∞. The distortion may be written as follows [16] D∞ dis= −sgn (x2) Re 2

i=1 di−1 3

α = 1 ¯ λαii + ¯ζiαGαi∗, (5)

where Re stands for the real part and∗indicates the complex conjugates of quantities indexed by α = 1, 2, 3

[16, 27]. The remaining unknowns { ¯λα

i, ¯ζiα} and {G α

i, G

α

i∗} are determined by solving the boundary

(5)

Crystal Lattice ( ˚A) Elastic constant (GPa)

a c c11 c33 c12 c13 c44 c66

Cu 3.615 - 168.4 - 121.4 - 75.4

-FePd 3.847 3.726 231.6 226.5 143.0 143.0 91.5 92.7

Table I: Material properties for Cu [1], and a γ1-FePd single crystal [30]

rotations produced by the interface dislocation arrays are then deduced from the symmetric and antisymmet-ric parts of eq. (5), i.e. E∞

dis= sym D∞disand Ω∞dis= skew D∞dis, respectively. Thus, the requirement in eq. (1)

may be written in terms of strains: lim

x2→±∞

Etot(x) = E∞tot= Ec+ E∞dis= 0 , (6)

for which the total elastic strain fields Etot in each material must decay to zero at long range. In the fol-lowing, when eqs. (1) or (6) are not satisfied, the associated dislocation structures will be designated as non-equilibrium structures. In theses circumstances, a residual long-range stress field and therefore a con-stant non-zero elastic energy density far from the interface due to an improper selection of the reference state might remain in the bicrystals [15, 16], meaning that the stored elastic energy in an infinite bicrystal is infinite. In turn, the zero far-field stresses and strains guarantee a finite stored elastic energy per unit area at the interfaces when the equilibrium condition is satisfied. An additional consideration arises when mul-tiple dislocation configurations satisfy both the Frank-Bilby equation as well as the equilibrium condition addressed above. In that case, an energy minimization operation may be performed to find which of the possible interface dislocation structures have lowest elastic energy, as it will become clearer in section 3.2.

3. Model applications

We apply the above model to describe variations in long-range elastic fields of misfit dislocations due to the changes in the Burgers vectors of interface dislocations in tilt and twist grain boundaries, as well as the

(short-range) strain energies of interfaces formed by tetragonal materials. The materials properties used in

these examples are listed in Table I. 3.1. Tilt and twist grain boundaries

Pure tilt boundaries that contain one set of straight parallel dislocations have been discussed extensively [1, 9, 31]. A symmetrical tilt boundary with [001] tilt axis and tilt angle θ = 2◦may be modeled as an edge dislocation array with b1= aCu[010] k n, for which eq. (2) yields ξ1= [001] and d1= 10.3567 nm [16]. Such

a grain boundary is free of long-range stresses. We consider the effect of converting the above dislocation array into another one by continuously varying the Burgers vectors from their ideal equilibrium values to two alternative end states: a pure edge array with b1k x1= [100] (Fig. 2a) and a pure screw array with b1k ξ1

(Fig. 2b). Since the elastic fields in all cases studied are symmetric, only solutions in the upper crystal are plotted in Fig. (2). The coherency strains and stresses remain zero for all cases: Ec= 0 ⇔ σc= 0.

As expected, in the ideal equilibrium structure, there are no far-field stresses, but there are far-field rota-tions Ω∞

12dis= −Ω

21dis= 0.0175, corresponding to a net rigid body rotation vector about the tilt axis given

by $ = − [0, 0, 0.03490] = −x1× b1/d1. In contrast, all other families of dislocation networks produce

non-zero far-field stresses. For example, the edge dislocation array with b1k x1= [100] produces

long-range stress components σ∞

11dis= 1.41 GPa and σ ∞

33dis= 0.59 GPa, but no rotations. The screw dislocation

array with b1k ξ1 exhibits long-range stress component σ13∞dis= −1.32 GPa as well as rotation

compo-nent Ω∞

(6)

superposition of additional elastic sources, e.g. external fields or additional set of dislocations, that may produce a rotation field with the same magnitude but opposite signs than the dependence of the component

values shown in Fig. (2b), all boundaries with non zero Ω∞

13disshould not be considered as tilt boundaries

but, strictly speaking, as single dislocation arrays which do not satisfy the Frank-Bilby equation. Such

boundaries are not proper grain boundaries but are better thought of as dislocation pile-ups. For all these

(non-equilibrium) grain boundaries that possess non-zero far-field stresses, however, equilibrium solutions may be reached by rotating the (010) interface habit plane [15]. Fig. (2c) illustrates the conversion of an array of screw dislocations, associated with a θ = 2◦Cu twist boundary on a (010) plane to an array of edge dislocations that remain orthogonal to each other with respect to θ⊥. Stresses vary approximately logarith-mically with θ⊥while the rotation component varies inversely to the stress. With φ = π/2, this boundary contains a square grid of screw dislocations with line directions ξ1= 1/√2¯101 and ξ2= 1/√2 [101] and spacings d1= d2= 7.3233 nm. If converted to the pure edge array, as shown in Fig. (2c), the dislocations

produce no rotation and long-range stress fields σ∞

11dis= σ33∞dis= 2.01 GPa in the upper crystal.Interestingly, Fig. (2c) illustrates that the introduction of a second array of screw dislocations to the case of a single screw dislocation array (shown in Fig. 2b) gives an equilibrium dislocation structure that is free of long-range stresses.

For tilt and twist grain boundaries that meet the condition of vanishing far-field strain and prescribed misorientations, the incomplete cancellation of the coherency and Volterra fields near the boundaries gives rise to short-range stresses and strains. These fields may also be used to compute the elastic strain energies consistent with the Frank-Bilby equation [16], such that the stored energies are obtained by using judiciously the divergence theorem in the linear elastic limit [28] with the aid of the following conditions: − the elastic fields are assumed to follow the periodicity of the two-dimensional interface dislocation structures, − the elastic fields are partitioned between the crystals in a manner that depends on the equilibrium equations of elasticity, − and the stress and strain fields decay to zero at long range. Following standard practice, because the short-range stress fields at dislocation cores diverge, a pre-determined cutoff distance is introduced in the calculations of elastic energies [1]. Fig. (3) shows the elastic energies computed using two core cutoff

parameters for both grain boundaries in Cu as a function of the tilt and twist angle up to 10◦. Experimental

values are plotted as triangles [29]. Qualitatively, both theoretical and experimental results illustrate that energies increase with the increasing rotation angles and the tilt energy is slightly higher than the twist

energy. Our calculations show that the core cutoff parameter r0= b/2 fits better the experiments up to ∼ 6◦,

while a higher value of r0might be more appropriated in the range of ∼ 6 − 10◦.

3.2. Misfit interface in tetragonal bicrystals

The final example applies to a pair of L10-ordered FePd crystals joined on (010) faces and rotated

with respect to one another by π/2, as illustrated in Fig. (4), for which epitaxial FePd thin films exhibit excellent mechanical strength for advanced structural and functional applications. FePd is tetragonal with a c/a ratio smaller than unity and a moderately high anisotropy ACu= 2c44/(c11− c12) = 2.63, as listed in

Table I. Following the methodology described in [16], the total far-field strains (and, hence, also stresses) are removed if the long-range strains produced by the dislocation arrays equal:

AE ∞ dis=    0.01598 0 −0 0 0 −0 0 0 −0.01598   = −BE ∞ dis, (7) withAΩ ∞ dis=BΩ ∞

dis= 0. Because eqs. (6) are satisfied, the uniform coherency matrices are equivalently

deduced from eqs. (7) by:AEc= −AE

disandBEc= −BE

(7)

in the plane of the interface and the corresponding principal coherency strains (that may be related to a pure shear produced by a screw dislocation array, see solution 2, below) reveal thatAFePd is compressed in the

[100] direction and extended in the [001] direction. To ensure perfect registry across the interface,BFePd

is extended in the [100] direction and compressed in the [001] direction. The resulting elastic strain is equally partitioned, as expected, so that the lattice parameter ac in the coherent reference state is given by

ac= 2ac/(a + c) = 0.37855 nm. Hence, one possible solution (solution 1) is the orthogonal array of edge

dislocations [24, 32, 33], as shown in Fig. (4b). These have Burgers vectors b⊥1 = ac[100] and b⊥2 = ac[001],

where the dislocation spacings of the former set of edges are d1⊥= d2⊥= 11.8462 nm.

Another possibility (solution 2) is the array of screw dislocations in Fig. (4c) with Burgers vectors b 1 = ac/2 [101] and b 2 = ac/2

¯101. The spacings are given by d

i = d

i /

2 = 8.3765 nm, with i = 1, 2. Both solutions (solution 1 and 2) relieve misfit and both have no far-field rotations.

In fact, an infinite number of admissible solutions may be expressed by continuously interpolating bi

between b⊥i and b i, such that bi= (1 − δ) b⊥i + δ b

i with 0 ≤ δ ≤ 1, for which intermediate dislocation

structures have mixed characters. Because the misorientation remains constant and the far-field stresses are zero along this entire path, the divergence theorem may be used to determine the elastic strain energy of equilibrium interface dislocations by converting the volume integral to the following surface integral. The definition of the short-range strain energy is therefore confined to the misfitting interface and to the discontinuous quantities across it, i.e.

Ee(r0) = 1 2 A Z Z A σtot(x1, 0, x3) n · ∆ udis(x1, x3) | {z } W dS, (8)

where A = d1d2 is the area of the orthogonal unit cells and the dislocation cores are excluded from the

domain of integration. In eq. (8), σtot(x1, 0, x3) n is the traction vector produced at the interface and

∆ udis(x1, x3) is the disregistry produced by the network of the two sets of interface dislocations, given

in terms of relative displacements between neighboring atomic planes.

Local contributions to the interface energies W (values of the integrand in eq. (8)) for the two above solutions are shown in Fig. (4b−c) and the elastic energy Ee as a function of δ is plotted in Fig. (5) for

three cutoff parameters: r0 = b/2, r0 = b/3, and r0 = b/4, with b the magnitude of bi. For instance,

Ee = 0.24381 J.m−2for the dislocation structure of solution 1 with r0= b/4, whereas Ee⊥= 0.50375 J.m−2

for solution 2. The ratio of energies for the two endpoint structuresis Ee E⊥ e = 0.48399 ∼(1 − ν)√ 2 with, b ⊥ = √ 2 b , (9)

where the isotropic approximation with Poisson’s ratio ν = 0.33 of γ1-FePd has been used [30]. Eq. (9)

shows that the anisotropic linear elasticity and isotropic approximation[15, 32], derived by using the van der

Merwe expressions for energy of dislocation arrays [34],give the same trend in terms of ratios of energies,

for which the character factor (1 − ν) and the b /b⊥Burgers vectors ratio both favor the screw dislocation array energetically.

In general, one may expect dislocation arrays with screw characters to have larger cores [1], so that the cutoff parameters may not alter the qualitative results in Fig. (5): the elastic energy decreases monotonically as a function of δ, showing that the array of screw dislocations is favored energetically. Thus, Eq. (8) may be used as a criterion to find the most likely dislocation structures among all solutions of the Frank-Bilby equation and find metastable configurations of equilibrium dislocations along admissible paths. The minimum of elastic energy given in Fig. (5) corresponds to a local minimum, strictly associated to the choice of our specified path. Other paths may certainly be defined, for which the Frank-Bilby equation is always satisfied and a global minimum may exist.

(8)

4. Concluding remarks

We have examined the effect of one type of variation in interface dislocation network configurations, namely changes in Burgers vectors, using heterogeneous anisotropic elasticity theory. As expected, incorrect reference states lead to non-zero far field stresses, incorrect far-field rotations, or both. In real tilt and twist boundaries, equilibrium is achieved when the rotational distortion fields are symmetrically partitioned with respect to the interface plane. Correct reference states for tilt and twist grain boundaries, for which elastic fields meet the condition of vanishing far-field strains and prescribed misorientations at long range, are uniquely defined.

In reality, interface dislocation structures may deviate from their ideal equilibrium state in other ways be-sides those considered here, for instance due to changes in misfit dislocation orientations during propagation and migration of interfaces, relaxation and reconstruction of dislocation networks formed by intersecting dislocation arrays, displacive phase transformations, or epitaxial growth process. Some of these phenomena may give rise to deviations of far-field elastic fields and interface strain energies from the ideal case. The formalism proposed here may be extended to address these cases as well.

The present model also resolves the ambiguity arising from the infinite number of reference states avail-able when interface dislocation structure is analyzed using the Frank-Bilby equation alone (i.e., without considering elastic fields). It is shown that an infinite of reference states are available for a pair of tetragonal (or orthorhombic) FePd crystals, rotated with respect to one another by π/2. When multiple dislocation structures are consistent with the Frank-Bilby equation, the elastic strain energies may be used to find the most likely one, i.e. the one with least strain energy. Applications of this energy criterion to the analy-sis of complex interfaces for determining a unique solution of the Frank-Bilby equation and investigating (short-range) elastic strain relaxations will be presented in follow-on studies.

Modeling of core energies is a possible extension of the present formalism, which is purely linear elastic. However, highly nonlinear fields may be expected in dislocation core regions, defined here by a core radius. These nonlinear effects become especially important for small dislocation spacings. To model them, a Peierls-Nabarro [35, 36] or van der Merwe [34] -type approach may be combined smoothly with linear elasticity by modeling continuous distributions of disregistry, rather than discrete Burgers vectors.

5. Acknowledgements

MJD acknowledges support from the National Science Foundation under Grant No. 1150862.

References

[1] Hirth JP, Lothe J. Theory of Dislocations. 2nd ed. Melbourne: Kriger; 1992. [2] Siegel RW, Chang SM, Balluffi RW. Acta Metall 1980; 28:249.

[3] Hirth JP, Pieraggi B, Rapp RA. Acta Metall Mater 1995; 43:1065.

[4] Aaronson HI. In: Zackay VF, Aaronson HI, editors. Decomposition of austenite by diffusional pro-cesses. New York: Wiley Interscience; 1962.

[5] Deymier P, Janit L, Li J, Dobrzynski L. Phys Rev B 1989; 39:1512.

[6] Demkowicz MJ, Wang J, Hoagland RG. Interfaces between dissimilar crystalline solids. Dislocations in Solids, edited by Hirth JP. Amsterdam: Elsevier, vol. 14, p. 141; 2008.

(9)

[7] Hwang M, Laughlin DE, Bernstein IM. Acta Metall 1979; 28:621. [8] Hall MG, Rigsbee JM, Aaronson HI. Acta Metall 1986; 34:1419.

[9] Sutton AP, Balluffi RW. Interfaces in Crystalline Materials. Oxford: Oxford University Press; 1995. [10] Frank FC. Acta Metall 1953; 1:15.

[11] Bilby BA, Bullough R, Smith E. Proc Roy Soc A (London) 1955; 231:263.

[12] Bollmann W. Crystal defects and crystalline interfaces. Berlin: Springer-Verlag; 1970. [13] Zhang WZ, Purdy. Acta Metall Mater 1993; 41:543;

[14] Hirth JP, Pond RC. Philos Mag Let 2010; 23:3129;

[15] Hirth JP, Pond RC, Hoagland RG, Liu XY, Wang J. Prog Mater Sci 2013; 58:749. [16] Vattr´e AJ, Demkowicz MJ. Acta Mat 2013; 61:5172.

[17] Christian JW. Trans JIM 1976;Suppl. 17:211. [18] Hirth JP, Balluffi RW. Acta Metall 1973; 21:973. [19] Olson GB, Cohen M. Acta Metall 1979; 27:1907. [20] Bollmann W. Acta Cryst 1977; 33:730.

[21] Amelinckx S. The direct observation of dislocations. New-York: Academic Press; 1964. [22] Yang JB, Nagai Y, Yang ZG, Hasegawa M. Acta Mater 2009; 57:4874.

[23] Bulatov V, Cai W. Computer simulations of dislocations Oxford Series on Materials Modelling. New York: Oxford University Press; 2006.

[24] Howe JM, Pond RC, Hirth JP. Prog Mater Sci 2009; 54:792. [25] Bonnet R. Acta Metall 1981; 29:437.

[26] Bonnet R. Philos Mag A 1981; 43:1165. [27] Stroh AN. Philos Mag 1958; 3:625.

[28] Willis JR, Jain SC, Bullough R. Phil Mag A 1990; 62:115. [29] Gjostein NA, Rhines FN. Acta Metall 1959; 7:319.

[30] Al-Ghaferi A, M¨ullner, Heinrich H, Kostorz G, Wiezorek JMK. Acta Mat 2006; 54:881. [31] Hirth JP, Barnett DM, Lothe J. Philos Mag A 1979; 40:39.

[32] Matthews JW. Phil Mag 1974; 29:797. [33] Hirth JP. J Mater Res 1993; 8:1572.

[34] Van der Merwe JH. Proc Phys Soc A 1950; 63:616. [35] Peierls R. Proc Roy Soc A (London) 1940; 52:34. [36] Nabarro FRN. Proc Roy Soc A (London) 1947; 59:256.

(10)

θ (°)

θ (°)

θ (°)

Long-range stress fields GPa

Long-range rotation field

−1.5 −1.0 −0.5 0. 0 30 60 90 0. 0.006 0.012 0.018 b x1 x2 x3 b θ σ Ω 33 12

a.

σ118 8 8

Edge dislocation array Edge dislocation array

Log-range stress field GPa

Long-range rotation fields

−1.5 −1.0 −0.5 0. 0 30 60 90 0. 0.006 0.012 −bb θ x1 x2 x3 σ13 Ω12 Ω13 8 8 8

Edge dislocation array Screw dislocation array

Long-range stress fields GPa

Long-range rotation field

0. 0 30 60 90 −0.005 x1 x2 x3 Ω13 σ 118 = 8

Screw dislocation array Edge dislocation array

0.5 1.0 1.5 2.0 2.5 0. −0.010 −0.015 −0.020 σ 338 b2 b1 b2 b

b.

c.

case 1 case 2 case 3 1 1 1 b1 0.018 a ⊥ ◦ 9

(11)

0 2 4 6 8 10 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 Ee (J.m ) −2 θ (°) Exp. data Exp. data Tilt r = b / 3 0 Tilt r = b / 2 0 Twist r = b / 3 0 Twist r = b / 2 0

Figure 3:Elastic energies computed using two different core cutoff parameters for [001] tilt (full lines) and (010) twist (dotted lines)

grain boundaries in Cu as a function of the rotation angle θ. Experimental values are shown with triangles [29].

solution 2 solution 1 1. material B material A coherent state a c set 2 set 1 ac

a.

b.

c.

b1 b2 b 1 b2 001 100 0.5 0.3 0.05 0.2 0.02 0.1

Figure 4: (a) Interface between two tetragonal L10crystals with the same c/a ratios rotated by π/2 with respect to one another

(differences in atom positions have been exaggerated). The reference state, for which the strain is equally partitioned, is illustrated in

black. The coherency strains may be canceled by arrays of either(b) edge or (c) screw dislocations. Local elastic energy densities W

(12)

0.2 r = b / 30 r = b / 20 r = b / 40 E (J.m ) −2 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.1 0.2 0.3 0.4 0.5 e δ 0.35 0.25 0.15 0.6 0.8 1.0

Edge dislocation array

(solution 1) Screw dislocation array(solution 2)

Figure

Figure 1: Mapping from a coherent reference state to the natural state using displacement gradients A F and B F
Table I: Material properties for Cu [1], and a γ 1 -FePd single crystal [30]
Figure 4: (a) Interface between two tetragonal L1 0 crystals with the same c/a ratios rotated by π/2 with respect to one another (differences in atom positions have been exaggerated)
Figure 5: Dependence on the elastic energies per unit area E e on δ in γ 1 -FePd for three core cutoff parameters r 0 .

Références

Documents relatifs

These dislocations having a Burgers vector b = 1/3 [11-20],. The interface is defined by its planeparallel to the dislocation line and disorientation varies between 0they are located

The results show that dislocation motion under the image force effect depends on the elastic and crystallographic properties of the materials constituting the bicrystals

Thus a multiband structure, consisting of one anisotropic part and one other part having larger warping and larger Fermi velocities, leads to an anomalous

Because the singularity at the center of the bottom surface (lower left corner in Fig. 3) looks more like the radial singularity of figure 2a, one might anticipate that

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

We have used a cross correlation based analysis of EBSD patterns to map the variation of the elastic strain tensor (and hence stress tensor) and small lattice rotations in the

In the Appendix we show that the formal perturbation expansion around the infinite temperature dynamics reproduces the strictly local high temperature expansion in the case of

(1) Polycrystalline aggregates containing 85 grains in average with different grain shapes and crystal orientations, except at one free surface where the 2D grain shape and