You are on page 1of 5

MATERIALS FORUM VOLUME 31 - 2007 Edited by J.M. Cairney and S.P.

Ringer Institute of Materials Engineering Australasia

FIRST PRINCIPLES CALCULATIONS OF STACKING FAULT ENERGIES FOR MAGNESIUM AND TITANIUM
A. E. Smith School of Physics, Monash University, Victoria, 3800 ARC Centre of Excellence for Design in Light Metals, Monash University, Victoria 3800 ABSTRACT Properties of magnesium and titanium have been calculated for bulk, surface and stacking faults, regarded as homogeneous interfaces, using density functional theory within a plane wave pseudopotential technique. Using a supercell approach, stacking fault energies have been determined for intrinsic, extrinsic and twin-like structures for the basal plane together with generalised stacking fault energies on basal, pyramidal and prismatic planes.

1. INTRODUCTION Light metal alloys are increasingly being used because of their light weight and high strength, however they can have low plastic formability. Accordingly, there is considerable interest in the deformation mechanism and comparison of slip on different crystallographic planes in the corresponding elemental metals. Stacking faults, which correspond to mismatches of close packed planes, provide a measure of the ease of slip. They can be regarded as homogeneous material interfaces and for hexagonally close packed metals, such as magnesium and titanium are observed on basal, prismatic or pyramidal planes. The preponderance of observed faults on particular planes for different metals has been ascribed to different c/a lattice constant ratios, but this simplistic picture based on a geometrical description of ball stacking does not fit all experimental observations.1 A more fundamental computational approach is to calculate the energy required to create a stacking fault on a certain surface in comparison with others.2,3 Indeed it is possible to employ computed stacking fault energies to model dislocation nucleation as a measure of intrinsic ductility or brittleness.4,5 This paper carries out first principle electronic calculations of the energy by using the density functional ABINIT package.6,7 It extends energy computations for magnesium previously performed by Chetty and Weinert2 using a Local Density Approximation (LDA) potential for stacking fault and twin-like structures together with those by Uesugi et al. 3 using a Generalised Gradient Approximation (GGA) potential for the generalised stacking fault energy, i.e. the energy obtained for fractional faulted displacements. It also reports initial results for titanium and compares them with results obtained from a non first principles approach which is based on the use of EAM (Embedded Atom Model) potentials. 8

2. COMPUTATIONAL PARAMETERS The present calculations were performed using the ABINIT package6,7, employing pseudopotentials and planewaves and so reliant on an efficient Fast Fourier Transform algorithm9 for conversions between real and reciprocal space. It is based on the adaptation to a fixed potential of the band-by-band conjugate gradient method10 and on a potential-based conjugate-gradient algorithm for the determination of the self-consistent potential.11The present work uses fully separable non-local Trouiller-Martins type pseudopotentials12 generated by the FHI98PP code,13 with exchange correlation calculated within the generalised gradient approximation using the Perdew Burke Emzerhof (PBE) form. 14 Calculations have been performed for different magnesium and titanium structures: bulk, surface and homogeneous interface (stacking faults), with the latter two found through the use of supercells. Numerical convergence needed to be monitored more strictly for the stacking fault calculations as the reported values of stacking fault energies, though somewhat uncertain, are only in the range of 10 150 meV 8,15-22 compared with surface and binding energies in the range of several eV. 23 Again regarding stacking faults as homogeneous material interfaces, this range should be compared with the much larger cleavage- bonding energies for heterogeneous interfaces. 24,25 Accordingly, monitoring of the numerical convergence of the stacking fault calculations was aimed at achieving as far as possible convergence per atom ~ 0.1meV (4x10-6 H) for the different aspects of the calculation. For instance the electron wave functions then need to be expanded in a plane wave basis with a kinetic energy cut-off greater than ~ 30 H for the case of magnesium but with a much larger value of ~ 55 H for titanium (see Figure 1).

71

1600 1400

Relative Total Energy (meV)

1200 1000 800

high power of (see Figure 2). 27,28,29 Accordingly, adequate convergence can then be achieved with a choice of parameter = 1x10-2 eV, which agrees with less rigorous determinations. 29,30
1.6 1.4

Relative Total Energy (meV)

600 400 200 0 -200 10 20 30 40 50 Ecut (Hartree) 60 70

1.2 1 0.8 0.6 0.4 0.2 0 -0.2 0 0.005 0.01 0.015 0.02 Smearing width (eV) 0.025 0.03

Figure 1 Total energy convergence of optimised bulk hcp Ti lattice for increasing kinetic energy cut-off (Ecut). Calculations performed with 352 k points in the Irreducible Brillouin Zone (IBZ) and smearing width parameter set at 0.01 eV. Integrations in k space were performed according to the Monkhorst Pack scheme together with group symmetry reduction.26 The number of special k points in the irreducible wedge of the Brillouin zone (IBZ) required depends on the size of the periodic cell / supercell and Figure 2 shows the convergence of the total energy as a function of number of k points for the titanium bulk hexagonal close packed structure. On the other hand, most of the present modelling of stacking faults in this paper is based on the use of twelve close packed planes and, for instance, the prismatic stacking fault calculation employed 462 special points.
1.6 1.4

Figure 3 Total energy convergence of optimised bulk hcp Mg lattice as function of smearing width parameter . Calculations performed with Ecut set at 30 H and nkpt set at 352. As a means of cross checking, results of calculations were compared between the bulk calculations and the supercell calculations. Whilst the appropriate 12 layer super cell calculations rapidly converged to the face centred cubic (fcc) bulk calculations with full symmetrised applied within ABINIT, it was necessary to turn off the full symmetrisation for the equivalent hcp calculations. This is indicative of the lower symmetry of the hexagonal system where there are two available parameters for variation.

Relative Total Energy (meV)

1.2 1 0.8 0.6 0.4 0.2 0 -0.2 0 100 200 300 400 Number of k points 500 600 700

3. RESULTS OF BULK AND SURFACE CALCULATIONS Bulk and surface properties were determined from the structural optimisations that employed the Broyden-Fletcher-Goldfarb-Shano (BFGS) method.31,32 For instance, to begin this study the minimum-energy lattice parameters with GGA for bulk magnesium in the simplest close packed structures were found to be: a = 3.18 and c = 5.19 (with c/a = 1.630) for the hcp (ideal c/a = 1.633) and a = 3.19 for the fcc structure. In particular this hcp structure is lower in energy than this fcc structure by 0.44 mH (12.0 meV) per atom. Table 1 compares these results with those previously reported using LDA and GGA approximations. 2,23,33

Figure 2 Total energy convergence of optimised bulk hcp Ti lattice as function of number of k points (nkpt) in IBZ. Calculations performed with Ecut set at 55 H and set at 0.01 eV. As magnesium and titanium are metallic systems, the quality of integration can be improved by use of cold Fermi surface smearing with width parameter 27,28 which should also be checked for convergence properties (see Figure 3). This is not an independent parameter and it is also necessary to choose a sufficiently large number of k points in order to observe the necessary requirement of total energy convergence proportional to a reasonably 72

Table 1 Comparison of calculated lattice constants for magnesium with experimental values hcp: a () 3.18 3.12 3.13 3.18 3.20 3.21 hcp: c/a 1.630 1.624 1.616 1.615 1.660 1.624 fcc: a () 3.19 3.12

This work GGA Weinert and Chetty LDA2 Wachowicz and Kiejna LDA23 Wachowicz and Kiejna GGA23 Fuchs et al. GGA 33 Experiment23

However in the present work the use of the dipole correction is avoided by adopting a sufficiently large vacuum space of eleven unoccupied layers between the eleven occupied layers of the slab. For instance, with a slab consisting of eleven occupied layers the total energy with eleven vacuum layers had converged to closer than 10-6 H of the value with twenty one vacuum layers.

4. CALCULATION OF INTERFACE PROPERTIES AND STACKING SEQUENCE ENERGIES To determine stacking sequence energies it is necessary to employ a basal lattice parameter and using a LDA potential Chetty and Weinert2 carried out stacking sequence energy calculations for magnesium based on the hcp lattice parameter. For bulk calculations (Table 1) the present GGA calculations for magnesium agrees with their results that the hcp structure is of lower energy than the fcc structure, but go further in determining the lattice parameter a to be closer to experimental values33 and larger for the fcc than the hcp structure. This increases the relative size of the fcc to the hcp structure caused by the c/a ratio being less than ideal for the magnesium hcp structure and so adds to the stability 2 . In practice this is only a small difference in energy value of 0.22 meV between the fcc structures with lattice constants a = 3.18 and 3.19. Stacking sequence energies have been determined for magnesium and titanium and compared for the hexagonal (hcp) , face centred cubic (fcc), two low energy intrinsic stacking faults (I1 and I2), extrinsic stacking fault (E) and a low energy twin like structure (T2), following previous work2,8 and using standard notation. 34,35 The sequences are denoted by ABC notation in Table 4 where for instance I1 is formed by removing a plane followed by a shear; I2 is formed by a shear; whilst E is formed by inserting a plane. The present GGA results shown in Table 4 for magnesium are in general agreement with the previous LDA results,2 though showing somewhat lesser energy differences. Table 4 stacking sequence energies for magnesium with number of faults denoted for different low energy structures Atoms Supercell Energy (meV) (GGA) 0 12 10 20 32 22 Energy (meV) (LDA)

In a similar way the BFGS method was used to determine the minimum-energy lattice parameters with GGA For bulk titanium magnesium in the hcp structure, this results in a = 2.959 with c/a = 1.618. This hcp structure is lower in energy than the corresponding fcc structure by 15 meV per atom. Table 2 compares these results with parameters previously employed to construct an EAM potential for titanium based on experimental results. 8 Table 2 Comparison of calculated lattice constants for magnesium with EAM fitted and experimental values hcp: a () This work GGA EAM fitted parameters8 Experiment8 2.96 2.95 1.95 hcp: c/a 1.618 1.585 1.588

Similar structural optimisation calculations were carried out to determine the surface properties using the slab method for the hcp structure. Table 3 shows the results for magnesium in comparison with those previously reported using LDA and GGA approximations. 23 Table 3 Comparison of calculated surface properties for magnesium with experimental values This work GGA a() c/a cohesive energy (eV) surface energy (eV) work function (eV) 3.18 1.6307 1.51 0.31 3.71 Wachowicz and Kiejna GGA 23 3.18 1.615 1.50 0.30 3.76 Wachowicz and Kiejna LDA 23 3.13 1.616 1.78 0.35 3.88 Expt.23

3.21 1.624 1.51 0.28 3.84 hcp fcc I1 I2 E T2

12 12 12 12 13 12

ABABABABABAB ABCABCABCABC ABABABCBCBCB ABABABCACACB ABABABCABABAB ABABABCBABAB

2 2 1 1

0 15 11 23 36 27

Table 5 shows the corresponding results for titanium in comparison with previous EAM results. 8 The sequences 73

are denoted by the same ABC notation as in Table 4. There are considerable differences between the different schemes especially for the low energy structures. In particular results from the use of the stated computational parameters in the present method is not able to distinguish a significant energy difference between the hcp and I1 structures. However, an experimental value for the stacking fault energy of the I1 structure is not available in the literature.8 Furthermore Ti has a complex phase diagram with several high pressure phases, as might be expected for a material with a number of low energy phases close to the hcp structure.36,37 For instance this has required a posteori adjustments of parameters by ~ 10 meV for non-first principle tight binding methods to achieve good descriptions. 37 In general these preliminary GGA results do support previously proposed relatively small values of stacking fault energies in comparison with existing experimental determinations. 8 Table 5 stacking sequence energies for titanium denoted for different low energy structures
Atoms Supercell Energy (meV) (GGA) 0 15 0 14 40 16 Energy (meV) (EAM)
8

Table 6 Stacking fault energies for magnesium along basal (0001) , prismatic (1010) and pyramidal

(10 1 0) planes.
Stacking fault energies (mJ/m2) Basal Prismatic Pyramidal This work 36 265 344 Chetty and Weinert (LDA) 2 44 Uesegi et al. (GGA) 3 32 255 Experiment
21,16

<50; 78

This has been termed the stable stacking fault energy in comparison with the unstable stacking fault energy corresponding to the maximum of this curve. 38 Whilst the corresponding positions are determined by crystal symmetry, relative stability or instability in different directions can be dependent on the details of the crystal potential. 39 Table 6 also shows results of stacking fault displacements for magnesium along the prismatic (1010) and pyramidal (10 1 0) planes, both of which are unstable, i.e. the end points corresponding to Figure 4 are local maxima, and which are compared with previous results for the basal and prismatic planes. 3 Also noted in Table 6 are the experimental values of 78 mJ/m2 3,16 and <50 mJ/m2 , 21 which are somewhat smaller than older less directly measure values (> 100 mJ/m2) . 22,35

hcp fcc I1 I2 E T2

12 12 12 12 13 12

ABABABABABAB ABCABCABCABC ABABABCBCBCB ABABABCACACB ABABABCABABAB ABABABCBABAB

2 2 1 1

0 11 14 26 39 -

5. CONCLUSION The results of first-principles calculations of stacking fault energies in magnesium and titanium using GGA potentials are in broad agreement with previous calculations. In particular the computed values are in general agreement with modern directly measured experimental results. The small differences in energy values from previous calculations for magnesium can be ascribed to choice of density functional theory approximations, increased distance between stacking faults in periodic cells and different energy minimisation relaxation algorithms. The unstable stacking fault energy of the prismatic plane is 30% greater than that on the pyramidal plane. The GGA stacking fault energies calculated in this work are consistently slightly smaller (several %) than corresponding LDA values. The preliminary results for titanium indicate smaller stacking fault energies than those presently derived in the experimental literature. In addition it is of interest to extend the present calculations for titanium to obtain stacking fault energies on the prismatic and pyramidal faces in addition to the basal face.

Figure4. shows the results of carrying out a series of stacking fault calculations for magnesium where the displacement along the basal plane is a fraction of the displacement for the final stacking fault. 3 The energy difference between the initial and final positions corresponds to the stacking fault value of Table 6.

Generalised Stacking Fault Energy (meV)

120 100 80 60 40 20 0 0 0.2 0.4 0.6 0.8 Displacement along 1/3<10-10> 1

Figure 4 The generalised stacking fault energy for the basal plane along the 1 3 10 1 0 direction.

74

References 1. 2. 3. 4. 5. 6. A. Kelly and G.W. Groves, Crystallography and Crystal Defects (Longmans, Harlow, 1970) N. Chetty and M. Weinert, Phys. Rev. B 56 (1997) 10844. T. Uesugi, M. Kohyama, M. Kohzu and K. Higashi, Mat. Sci. Forum 419-422 (2003) 225 Y. Juan and E. Kaxiras, Phil. Mag. 74 (1996) 1367-1384. J.R. Rice, J. Mech. Phys. Solids 40 (1992) 239-271. X. Gonze, J.-M. Beuken, R. Caracas, F. Detraux, M. Fuchs, G.-M. Rignanese, L. Sindic, M. Verstraete, G. Zerah, F. Jollet, M. Torrent, A. Roy, M. Mikami, Ph. Ghosez, J.-Y. Raty, D.C. Allan, Comput. Mat. Sci. 25 (2002) 478. X. Gonze, G.-M. Rignanese, M. Verstraete, J.-M. Beuken, Y. Pouillon, R. Caracas, F. Jollet, M. Torrent, G. Zerah, M. Mikami, P. Ghosez, M. Veithen, J.-Y. Raty, V. Olevano, F. Bruneval, L. Reining, R. Godby, G. Onida, D.R. Hamann and D.C. Allan. Z. Kristallogr. 220 (2005) 558. The ABINIT code is a common project of the Universit Catholique de Louvain, Corning Incorporated, and other contributors (URL http://www.abinit.org). R.R. Zope and Y. Mishin, Phys. Rev. B 68 (2003) 024102. S. Goedecker, SIAM J. on Scientific Computing 18 (1997) 1605. M.C. Payne, M.P. Teter, D.C. Allan, T.A. Arias and J.D. Joannopoulos, Rev. Mod. Phys. 64 (1992) 1045. X. Gonze, Phys. Rev. B 54 (1996) 4383. N. Troullier and J.L. Martins, Phys. Rev. B 43 (1991) 1993. M. Fuchs and M. Scheffler, Comput. Phys. Commun. 119 (1999) 67. J.P. Perdew, K. Burke and M. Emzerhof, Phys. Rev. Lett. 77 ( 1996) 3865. A. Seeger, S. Mader and H. Kronmller, Electron Microcopy and Strength of Crystals, (Interscience, New York,1963) p 665. D.K Sastry, Y.V.R.K. Prasad and K.I. Vasu, Scripta

7.

8. 9. 10. 11. 12. 13. 14. 15. 16.

Met. 3 (1969) 927. 17. P.S. Dobson and R.E. Smallman, Scripta Met. 4 (1970) 345. 18. D.K Sastry, Y.V.R.K. Prasad and K.I. Vasu, Scripta Met. 4 (1970) 347. 19. R.E. Smallman and P.S. Dobson, Metal. Trans.1 (1970) 2383. 20. J.F. Devlin, J. Phys. F: Metal Phys. 4 (1974) 1865. 21. A. Couret and D. Caillard, Acta Metall. 33 (1985) 1455 22. R.L. Fleischer, Scripta Met. 20 (1986) 223. 23. E. Wachowicz and A. Kiejna, J. Phys.: Condens. Matter, 13 ( 2001) 10767. 24. I.G. Batirev, A. Alavi, M.W. Finnis and T. Deutsch, Phys. Rev. Lett. 82 (1999) 1510. 25. L.M. Liu, S.Q. Wang and H.Q. Ye, Surf. Sci. 550 (2004) 46. 26. H. Monkhorst and J. Pack, Phys. Rev. B 13 (1976) 5188. 27. M. Verstraete and X. Gonze, Comput. Mat. Sci. 30 (2004) 27. 28. N. Mazari, Ph.D. thesis, University of Cambridge, 1996, http://alfaromeo.princeton.edu/marzari/preprints/. 29. S. de Gironcoli, Phys. Rev. B 51 (1995) 6773. 30. E.G. Moroni, G. Kresse, J. Hafner and J. Furthmller, Phys. Rev. B 56 (1997) 15629. 31. C. Broyden, Math. Comput. 19 (1965) 577. 32. D.F. Shano, Math. Op. Res. 3 (1978) 244. 33. M. Fuchs, M. Bockstedte, E. Pehlke and M. Scheffler, Phys. Rev. B 57 (1998) 2134. 34. F.C. Franck, Phil. Mag. 42 (1951) 809. 35. J.P. Hirth and J. Lothe, Theory of dislocations, 2nd ed. (Wiley, New York, 1982) 36. S.P. Rudin, M.D. Albers and R.C. Albers, Phys. Rev. B 69 (2004) 094117. 37. M.J. Mehl and D.A. Papaconstantopoulos, Europhys. Lett. 60 (2002) 248-254. 38. H.J. Gotsis, D.A. Papaconstantopoulos and M.J. Mehl, Phys. Rev. B 65 (2002) 1344101. 39. V. Vitek and M. Igarashi, Phil. Mag. A 63 (1991) 1059.

75

You might also like