on the accuracy of the non-self-consistent calculation of the electronic structure of solids with...

4
Physics Letters A 376 (2012) 879–882 Contents lists available at SciVerse ScienceDirect Physics Letters A www.elsevier.com/locate/pla On the accuracy of the non-self-consistent calculation of the electronic structure of solids with hybrid functionals Fabien Tran Institute of Materials Chemistry, Vienna University of Technology, Getreidemarkt 9/165-TC, A-1060 Vienna, Austria article info abstract Article history: Received 23 November 2011 Accepted 16 January 2012 Available online 18 January 2012 Communicated by V.M. Agranovich Keywords: Density functional theory Hybrid functionals Band gap A simple approximation within the framework of the hybrid methods for the calculation of the electronic structure of solids is presented. By considering only the diagonal elements of the matrix of the perturba- tion operator (Hartree–Fock exchange minus semilocal exchange) calculated in the basis of the semilocal orbitals, the computational time is drastically reduced, while keeping very well in most studied cases the accuracy of the results obtained with hybrid functionals when applied without any approximations. © 2012 Elsevier B.V. All rights reserved. 1. Introduction Quantum calculations for solids are commonly done with the local density approximation (LDA) or the generalized gradient ap- proximation (GGA) of density-functional theory (DFT) [1,2]. While the properties calculated using the total energy (e.g., lattice con- stant) are fairly well described by LDA or GGA, the electronic band structure cannot be described correctly in many cases. For instance, it is well known that the semilocal approximations (LDA and GGA) lead to band gaps in semiconductors and insulators which are too small with respect to experiment (see, e.g., Ref. [3]). More ad- vanced theories can provide more accurate band structures [3–8]. The best known are the GW approximation to the self-energy Σ xc within the many-body perturbation theory (Ref. [9]) and the hy- brid functionals [10,11] which consist of a mixing of semilocal and Hartree–Fock (i.e., exact) exchange. The GW method represents the state-of-the-art [4], and leads very often to very accurate results in particular if it is applied self- consistently [12–14] and with vertex corrections [14]. However, due to the huge computational effort required by a GW calculation, most of the GW results from the literature were obtained non-self- consistently (i.e., one-shot G 0 W 0 ). Within the G 0 W 0 method, the quasiparticle energies G 0 W 0 nk are obtained as solutions of the non- linear equation G 0 W 0 nk = SL nk + ψ SL nk Σ xc ( G 0 W 0 nk ) v SL xc ψ SL nk , (1) E-mail address: [email protected]. where ψ SL nk and SL nk are the orbitals and the corresponding ener- gies obtained from a previous DFT calculation with a semilocal (SL) functional E SL xc [in Eq. (1), v SL xc = δ E SL xc ρ ]. The nondiagonal terms of the matrix of Σ xc v SL xc are neglected [15]. The quasi- particle energies G 0 W 0 nk are most of the time close to experiment, but there are known cases (e.g., NiO [16]) were self-consistency is really needed. Beside the heavy cost of a GW calculation (G 0 W 0 can still be considered as a very expensive method), a drawback is that the convergence of the results with respect to the number of unoccupied states can be extremely slow as, for example, for ZnO [17,18]. The hybrid functionals are becoming more and more popular for solids (see, e.g., Ref. [7]). On average, the accuracy of the elec- tronic band structure they provide is quite similar to G 0 W 0 . In hybrid functionals, a fraction α x of semilocal exchange is replaced by the Hartree–Fock (HF) exchange: E hybrid xc = E SL xc + α x ( E HF x E SL x ) . (2) The value of α x which leads to the best agreement with experi- ment depends on (a) the system under study, (b) the considered property, and (c) the underlying semilocal functional E SL xc . Most of the time, the value of α x lies in the range 0–0.5. Among the best- known hybrid functionals, there is the so-called PBE0 [19,20] (the functional considered in the present work), where the semilocal functional in Eq. (2) is the GGA of Perdew, Burke, and Ernzerhof [21] (PBE) and the amount α x of Hartree–Fock exchange is set to 0.25 (Ref. [22]). Another way of constructing hybrid function- als consists of replacing only the short-range part of the semilocal exchange by short-range Hartree–Fock as proposed in Ref. [23] 0375-9601/$ – see front matter © 2012 Elsevier B.V. All rights reserved. doi:10.1016/j.physleta.2012.01.022

Upload: fabien-tran

Post on 11-Sep-2016

214 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: On the accuracy of the non-self-consistent calculation of the electronic structure of solids with hybrid functionals

Physics Letters A 376 (2012) 879–882

Contents lists available at SciVerse ScienceDirect

Physics Letters A

www.elsevier.com/locate/pla

On the accuracy of the non-self-consistent calculation of the electronic structureof solids with hybrid functionals

Fabien Tran

Institute of Materials Chemistry, Vienna University of Technology, Getreidemarkt 9/165-TC, A-1060 Vienna, Austria

a r t i c l e i n f o a b s t r a c t

Article history:Received 23 November 2011Accepted 16 January 2012Available online 18 January 2012Communicated by V.M. Agranovich

Keywords:Density functional theoryHybrid functionalsBand gap

A simple approximation within the framework of the hybrid methods for the calculation of the electronicstructure of solids is presented. By considering only the diagonal elements of the matrix of the perturba-tion operator (Hartree–Fock exchange minus semilocal exchange) calculated in the basis of the semilocalorbitals, the computational time is drastically reduced, while keeping very well in most studied cases theaccuracy of the results obtained with hybrid functionals when applied without any approximations.

© 2012 Elsevier B.V. All rights reserved.

1. Introduction

Quantum calculations for solids are commonly done with thelocal density approximation (LDA) or the generalized gradient ap-proximation (GGA) of density-functional theory (DFT) [1,2]. Whilethe properties calculated using the total energy (e.g., lattice con-stant) are fairly well described by LDA or GGA, the electronic bandstructure cannot be described correctly in many cases. For instance,it is well known that the semilocal approximations (LDA and GGA)lead to band gaps in semiconductors and insulators which are toosmall with respect to experiment (see, e.g., Ref. [3]). More ad-vanced theories can provide more accurate band structures [3–8].The best known are the GW approximation to the self-energy Σxcwithin the many-body perturbation theory (Ref. [9]) and the hy-brid functionals [10,11] which consist of a mixing of semilocal andHartree–Fock (i.e., exact) exchange.

The GW method represents the state-of-the-art [4], and leadsvery often to very accurate results in particular if it is applied self-consistently [12–14] and with vertex corrections [14]. However,due to the huge computational effort required by a GW calculation,most of the GW results from the literature were obtained non-self-consistently (i.e., one-shot G0W0). Within the G0W0 method, thequasiparticle energies ε

G0 W0nk are obtained as solutions of the non-

linear equation

εG0 W0nk = εSL

nk + ⟨ψSL

nk

∣∣Σxc(ε

G0 W0nk

) − vSLxc

∣∣ψSLnk

⟩, (1)

E-mail address: [email protected].

0375-9601/$ – see front matter © 2012 Elsevier B.V. All rights reserved.doi:10.1016/j.physleta.2012.01.022

where ψSLnk and εSL

nk are the orbitals and the corresponding ener-gies obtained from a previous DFT calculation with a semilocal(SL) functional ESL

xc [in Eq. (1), vSLxc = δESL

xc/δρ]. The nondiagonalterms of the matrix of Σxc − vSL

xc are neglected [15]. The quasi-

particle energies εG0 W0nk are most of the time close to experiment,

but there are known cases (e.g., NiO [16]) were self-consistency isreally needed. Beside the heavy cost of a GW calculation (G0W0can still be considered as a very expensive method), a drawbackis that the convergence of the results with respect to the numberof unoccupied states can be extremely slow as, for example, forZnO [17,18].

The hybrid functionals are becoming more and more popularfor solids (see, e.g., Ref. [7]). On average, the accuracy of the elec-tronic band structure they provide is quite similar to G0W0. Inhybrid functionals, a fraction αx of semilocal exchange is replacedby the Hartree–Fock (HF) exchange:

Ehybridxc = ESL

xc + αx(

EHFx − ESL

x

). (2)

The value of αx which leads to the best agreement with experi-ment depends on (a) the system under study, (b) the consideredproperty, and (c) the underlying semilocal functional ESL

xc . Most ofthe time, the value of αx lies in the range 0–0.5. Among the best-known hybrid functionals, there is the so-called PBE0 [19,20] (thefunctional considered in the present work), where the semilocalfunctional in Eq. (2) is the GGA of Perdew, Burke, and Ernzerhof[21] (PBE) and the amount αx of Hartree–Fock exchange is setto 0.25 (Ref. [22]). Another way of constructing hybrid function-als consists of replacing only the short-range part of the semilocalexchange by short-range Hartree–Fock as proposed in Ref. [23]

Page 2: On the accuracy of the non-self-consistent calculation of the electronic structure of solids with hybrid functionals

880 F. Tran / Physics Letters A 376 (2012) 879–882

for the HSE functional (also based on PBE), where the short- andlong-range parts of exchange are defined by splitting the Coulomboperator with the error function. Not considering the long-rangeHartree–Fock exchange avoids technical problems and makes thecalculations faster. Note that the idea of screening the Hartree–Fock exchange was already used by Bylander and Kleinman fortheir screened-exchange LDA functional (sX-LDA) [6], which canbe considered as a hybrid functional with 100% (αx = 1) of short-range Hartree–Fock exchange (see Ref. [24] for recent sX-LDA cal-culations). However, even without long-range Hartree–Fock, theuse of hybrid functionals for solids leads to calculations whichare one or two orders of magnitude more expensive than withsemilocal functionals. The tendency of the PBE0 functional is tooverestimate small (< 3 eV) band gaps and to underestimate large(> 10 eV) band gaps (see, e.g., Ref. [25]), and due to the neglect ofthe long-range Hartree–Fock in HSE, the HSE band gaps are smallerthan the PBE0 band gaps [25–27]. Note that in Ref. [25] it wasshown that the performance of PBE0 and HSE could be improvedby making the fraction αx of Hartree–Fock exchange dependent oneither the average of |∇ρ|/ρ in the unit cell (as done in Ref. [8] forthe modified Becke–Johnson potential [28]) or the static dielectricconstant.

In this work, a fast way of getting the energies of orbitalsfrom hybrid functionals is proposed. In Ref. [29], the implemen-tation of hybrid functionals (screened and unscreened) into theWIEN2k code [30], which is based on the full-potential linearizedaugmented plane-wave plus local orbitals method [31,32] to solvethe Kohn–Sham equations, was reported. The Hartree–Fock methodwas implemented following the method of Massidda, Posternak,and Baldereschi [33], which is based on the pseudocharge methodto solve the Poisson equation [34].

2. Theory and results

The calculation of the matrix of the nonlocal Hartree–Fock op-erator v̂HF

x is done in a second variational procedure, i.e., the oper-ator αx(v̂HF

x − vSLx ) is considered as a perturbation and the semilo-

cal orbitals are used as basis functions:⟨ψSL

nk

∣∣αx(

v̂HFx − vSL

x

)∣∣ψSLn′k

⟩. (3)

The second variational procedure, which was also adopted forthe implementation of the Hartree–Fock equations in other LAPWcodes [33,35,36] leads to cheaper calculations, since in practicethe number of orbitals ψSL

nk which are used for the constructionof Eq. (3) can be chosen to be much smaller than the numberof LAPW basis functions. Since the shape of the orbitals obtainedfrom the semilocal and hybrid functionals can be expected to berather similar, the most important matrix elements of Eq. (3) arethe diagonal ones. This leads to the simple proposition which con-sists of calculating the orbital energies from hybrid functional thefollowing way:

ε̃hybridnk = εSL

nk + ⟨ψSL

nk

∣∣αx(

v̂HFx − vSL

x

)∣∣ψSLnk

⟩, (4)

i.e., the nondiagonal terms are neglected, similarly as done inEq. (1) for the G0W0 method [15]. Within this approximationthe orbitals are not updated since the matrix of the operatorαx(v̂HF

x − vSLx ) [Eq. (3)] is diagonal, which means that the ψSL

nk arealready the eigenvectors [and used to calculate the Hartree–Fockpotential v̂HF

x in Eq. (4)]. In the following, the results obtainedwith Eq. (4) will be named PBE00 (i.e., one-shot PBE0) in analogywith G0W0. Compared to a self-consistent calculation, using Eq. (4)leads to calculations of the band structure which are at least two or-ders of magnitude faster (neglect of the nondiagonal terms and onlyone iteration). Naturally, the PBE orbitals were used as the semilo-cal orbitals in Eq. (4). Actually, Eq. (4) is also very closely related

Table 1Transition energies (in eV) obtained with the PBE, PBE0, and PBE00 methods. Theexperimental value for Cu2O is from Ref. [41]. See Table I of Ref. [36] for the othersolids.

Solid Transition PBE PBE0 PBE00 Expt.

Ar Γ → Γ 8.69 11.09 11.11 14.2C Γ → Γ 5.59 7.69 7.64 7.3

Γ → X 4.76 6.64 6.59Γ → L 8.46 10.76 10.73

Si Γ → Γ 2.56 3.95 3.90 3.4Γ → X 0.71 1.91 1.87Γ → L 1.53 2.86 2.80 2.4

GaAs Γ → Γ 0.53 1.99 1.87 1.63Γ → X 1.46 2.66 2.65 2.18, 2.01Γ → L 1.01 2.35 2.29 1.84, 1.85

MgO Γ → Γ 4.79 7.23 7.23 7.7Γ → X 9.16 11.58 11.65Γ → L 7.95 10.43 10.48

NaCl Γ → Γ 5.22 7.29 7.28 8.5Γ → X 7.59 9.80 9.80Γ → L 7.33 9.40 9.40

Cu2O Γ → Γ 0.53 2.77 2.68 2.17MnO Γ → Γ 1.47 4.23 4.04NiO Γ → Γ 2.41 6.07 5.91

to the way the exchange part of the derivative discontinuity is cal-culated:

x = 〈ψn′k′ |v̂HFx − vEXX

x |ψn′k′ 〉 − 〈ψnk|v̂HFx − vEXX

x |ψnk〉, (5)

where ψnk and ψn′k′ are the orbitals at the valence band maxi-mum and conduction band minimum, respectively, and vEXX

x is themultiplicative exact exchange (EXX) potential obtained with theoptimized effective potential method [see Ref. [3] for more dis-cussion on Eq. (5)].

In order to test the accuracy of Eq. (4) for the calculation of theorbital energies, we have considered the following semiconductorsand insulators (the structure and cubic lattice constant are given inparenthesis): Ar (fcc, 5.260 Å), C (diamond, 3.567 Å), Si (diamond,5.430 Å), GaAs (zinc blende, 5.648 Å), MgO (rocksalt, 4.207 Å), NaCl(rocksalt, 5.595 Å), Cu2O (Pn3m, 4.27 Å), MnO (Fm3m, 4.445 Å),and NiO (Fm3m, 4.171 Å). The unit cell of Cu2O contains six atoms.Formally, Cu has a valency of +1 and therefore the Cu-3d shell isfull, which means that the correlation effects in the Cu-3d shellshould not play an important role as it is the case for CuO [37].MnO and NiO have a cubic symmetry, but by taking into accountthe antiferromagnetic phase (along the [111] direction of the cubiccell), the symmetry is reduced to a rhombohedral one (four atomsin the unit cell). MnO and NiO are two of the most studied Mottinsulators, for which the semilocal functionals are very inaccuratedue to the strongly localized character of the 3d electrons [38].Detailed descriptions of the structures of Cu2O and MnO/NiO canbe found in Refs. [39] and [40], respectively.

The calculated transition energies, along with experimental val-ues, are shown in Table 1. As expected, the PBE transition energiesare by far too small compared to the experiment, while the PBE0values are much larger and closer to the experiment. However,as already mentioned above, PBE0 overestimates small transitionenergies (e.g., GaAs) and underestimates large transition energies(e.g., Ar). Actually, the important observation of the present workis that the PBE0 transition energies calculated with the approx-imation given by Eq. (4) (PBE00 in Table 1) are very similar tothe self-consistent PBE0 results. Indeed, the largest difference isfor MnO (∼ 0.2 eV), which, anyway, represents less than 10% ofthe difference between PBE (1.47 eV) and PBE0 (4.23 eV). For theΓ → Γ transition in GaAs and Cu2O, the difference between PBE0and PBE00 is ∼ 0.1 eV and less than 0.1 eV in all other cases. Inthe case of MgO and NaCl, the accuracy of Eq. (4) is impressive.

Figs. 1, 2, and 3 show the density of states (DOS) of GaAs, MnO,and NiO, respectively. In the case of GaAs, the PBE0 and PBE00

Page 3: On the accuracy of the non-self-consistent calculation of the electronic structure of solids with hybrid functionals

F. Tran / Physics Letters A 376 (2012) 879–882 881

Fig. 1. (Color online.) Density of states of GaAs calculated with the PBE, PBE0, andPBE00 methods. The Fermi energy is set at zero.

DOSs are indistinguishable, which can be explained by the fact thatalready the PBE and PBE0 DOSs are very similar and differ only bythe shift of the unoccupied bands. For MnO as well, the agreementbetween PBE0 and PBE00 is very good, albeit small differences canbe seen. In the case of NiO (the most difficult case considered inthis work), more visible differences between PBE0 and PBE00 canbe observed. For instance, the PBE00 DOS in the energy range −7and −3 eV looks intermediate between the PBE and PBE0 DOSs.There is less O-p states in the latter case. Also, the unoccupiedNi-d peaks are higher in energy by ∼ 0.8 eV with PBE0 than withPBE00.

In Refs. [16] and [12], band gaps of 1.1 and 4.8 eV for NiO cal-culated with non-self-consistent [Eq. (1)] and self-consistent GWcalculations, respectively, were reported. The importance of self-consistency in this case is rather extreme. Such a huge differenceis not observed for NiO with hybrid functionals when using Eq. (4)instead of doing a self-consistent PBE0 calculation. The explana-tion lies in the fact that in the case of Eq. (4), only the semilocalorbitals are used for the construction of the nonlocal Hartree–Fockoperator v̂HF

x , while for G0W0 [Eq. (1)] both the orbitals and theirenergies are used for the calculation of the self-energy Σxc. In thecase of Cu2O it was shown in Ref. [13], that with respect to non-self-consistent G0W0, updating only the orbital energies in Σxcalready increases the band by 0.46 eV (from 1.34 to 1.80 eV), whileupdating also the orbitals increases further the band gap, but by amuch smaller value (0.17 eV). Other such examples can be found inRef. [42]. However, the effect of self-consistency in the GW methodcan be reduced if the input orbitals and energies are calculatedfrom the hybrid (see Ref. [4]) or LDA + U (see Ref. [43]) methods.

3. Summary

In summary, it has been shown that for calculations with hybridfunctionals, neglecting the nondiagonal terms in the constructionof the Hartree–Fock Hamiltonian (within the second variationalprocedure) leads to a very good approximation. By doing so, theorbital energies [Eq. (4)] are very close to the energies obtainedfrom a self-consistent calculation done without any approxima-

Fig. 2. (Color online.) Density of states of one spin component of MnO calculatedwith the PBE, PBE0, and PBE00 methods. The Fermi energy is set at zero.

Fig. 3. (Color online.) Density of states of one spin component of NiO calculatedwith the PBE, PBE0, and PBE00 methods. The Fermi energy is set at zero.

tions. Several types of semiconductors and insulators, includingthe more difficult cases of antiferromagnetic MnO and NiO, havebeen considered and the accuracy of the approximation has beenshown to be (very) good in most cases. For NiO, the results areless impressive than in the other cases. The approximation leads tocalculations of the band structure of solids with hybrid functionalswhich are at least two orders of magnitude faster than the self-consistent calculations. This approximation allows the calculationof the electronic band structure of solids with hybrid functionalson much larger systems.

Page 4: On the accuracy of the non-self-consistent calculation of the electronic structure of solids with hybrid functionals

882 F. Tran / Physics Letters A 376 (2012) 879–882

Acknowledgements

Helpful discussions with Peter Blaha are greatly acknowledged.This work was supported by Projects SFB-F41 (ViCoM) and P20271-N17 of the Austrian Science Fund.

References

[1] P. Hohenberg, W. Kohn, Phys. Rev. 136 (1964) B864.[2] W. Kohn, L.J. Sham, Phys. Rev. 140 (1965) A1133.[3] M. Städele, J.A. Majewski, P. Vogl, A. Görling, Phys. Rev. Lett. 79 (1997) 2089.[4] F. Bechstedt, F. Fuchs, G. Kresse, Phys. Status Solidi B 246 (2009) 1877.[5] V.I. Anisimov, J. Zaanen, O.K. Andersen, Phys. Rev. B 44 (1991) 943.[6] D.M. Bylander, L. Kleinman, Phys. Rev. B 41 (1990) 7868.[7] F. Corà, M. Alfredsson, G. Mallia, D.S. Middlemiss, W.C. Mackrodt, R. Dovesi, R.

Orlando, Struct. Bonding (Berlin) 113 (2004) 171.[8] F. Tran, P. Blaha, Phys. Rev. Lett. 102 (2009) 226401.[9] L. Hedin, Phys. Rev. 139 (1965) A796.

[10] A.D. Becke, J. Chem. Phys. 98 (1993) 1372.[11] A.D. Becke, J. Chem. Phys. 98 (1993) 5648.[12] S.V. Faleev, M. van Schilfgaarde, T. Kotani, Phys. Rev. Lett. 93 (2004) 126406.[13] F. Bruneval, N. Vast, L. Reining, M. Izquierdo, F. Sirotti, N. Barrett, Phys. Rev.

Lett. 97 (2006) 267601.[14] M. Shishkin, M. Marsman, G. Kresse, Phys. Rev. Lett. 99 (2007) 246403.[15] L. Hedin, Int. J. Quantum Chem. 56 (1995) 445.[16] M. van Schilfgaarde, T. Kotani, S.V. Faleev, Phys. Rev. B 74 (2006) 245125.[17] B.-C. Shih, Y. Xue, P. Zhang, M.L. Cohen, S.G. Louie, Phys. Rev. Lett. 105 (2010)

146401.[18] C. Friedrich, M.C. Müller, S. Blügel, Phys. Rev. B 83 (2011) 081101(R).[19] M. Ernzerhof, G.E. Scuseria, J. Chem. Phys. 110 (1999) 5029.[20] C. Adamo, V. Barone, J. Chem. Phys. 110 (1999) 6158.

[21] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865;J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 78 (1997) 1396.

[22] J.P. Perdew, M. Ernzerhof, K. Burke, J. Chem. Phys. 105 (1996) 9982.[23] J. Heyd, G.E. Scuseria, M. Ernzerhof, J. Chem. Phys. 118 (2003) 8207;

J. Heyd, G.E. Scuseria, M. Ernzerhof, J. Chem. Phys. 124 (2006) 219906.[24] S.J. Clark, J. Robertson, Phys. Status Solidi B 248 (2011) 537.[25] M.A.L. Marques, J. Vidal, M.J.T. Oliveira, L. Reining, S. Botti, Phys. Rev. B 83

(2011) 035119.[26] J. Heyd, J.E. Peralta, G.E. Scuseria, R.L. Martin, J. Chem. Phys. 123 (2005) 174101.[27] J. Paier, M. Marsman, K. Hummer, G. Kresse, I.C. Gerber, J.G. Ángyán, J. Chem.

Phys. 124 (2006) 154709;J. Paier, M. Marsman, K. Hummer, G. Kresse, I.C. Gerber, J.G. Ángyán, J. Chem.Phys. 125 (2006) 249901.

[28] A.D. Becke, E.R. Johnson, J. Chem. Phys. 124 (2006) 221101.[29] F. Tran, P. Blaha, Phys. Rev. B 83 (2011) 235118.[30] P. Blaha, K. Schwarz, G.K.H. Madsen, D. Kvasnicka, J. Luitz, in: K. Schwarz (Ed.),

WIEN2k: An Augmented Plane Wave plus Local Orbitals Program for Calculat-ing Crystal Properties, Vienna University of Technology, Austria, 2001.

[31] O.K. Andersen, Phys. Rev. B 12 (1975) 3060.[32] E. Sjöstedt, L. Nordström, D.J. Singh, Solid State Commun. 114 (2000) 15.[33] S. Massidda, M. Posternak, A. Baldereschi, Phys. Rev. B 48 (1993) 5058.[34] M. Weinert, J. Math. Phys. 22 (1981) 2433.[35] R. Asahi, W. Mannstadt, A.J. Freeman, Phys. Rev. B 59 (1999) 7486.[36] M. Betzinger, C. Friedrich, S. Blügel, Phys. Rev. B 81 (2010) 195117.[37] J. Ghijsen, L.H. Tjeng, J. van Elp, H. Eskes, J. Westerink, G.A. Sawatzky, M.T.

Czyzyk, Phys. Rev. B 38 (1988) 11322.[38] K. Terakura, T. Oguchi, A.R. Williams, J. Kübler, Phys. Rev. B 30 (1984) 4734.[39] P. Marksteiner, P. Blaha, K. Schwarz, Z. Phys. B: Condens. Matter 64 (1986) 119.[40] M. Cococcioni, S. de Gironcoli, Phys. Rev. B 71 (2005) 035105.[41] P.W. Baumeister, Phys. Rev. 121 (1961) 359.[42] M. Shishkin, G. Kresse, Phys. Rev. B 75 (2007) 235102.[43] H. Jiang, R.I. Gomez-Abal, P. Rinke, M. Scheffler, Phys. Rev. B 82 (2010) 045108.