Skip to content
Publicly Available Published by De Gruyter May 11, 2017

The nitridoborate nitrides Mg3[BN2]N and Ca3[BN2]N – electronic structure and chemical bonding

  • Samir F. Matar , Adel F. Al Alam and Rainer Pöttgen EMAIL logo

Abstract

The nitridoborates Mg3[BN2]N (P63/mmc) and Ca3[BN2]N (P4/mmm) are electron-precise compounds with discrete linear [BN2]3− and isolated N3− anions. Electronic structure calculations reveal pronounced B–N bonding within the [BN2]3− units with more covalent Mg–N vs. Ca–N bonding. Total energy calculation for hexagonal normal-pressure Mg3[BN2]N, orthorhombic high-pressure Mg3[BN2]N and a hypothetical Ca3[BN2]N-type tetragonal Mg3[BN2]N modification revealed that the hexagonal modification is the ground state structure. The band structure for orthorhombic high-pressure Mg3[BN2]N indicates a substantial metallization (delocalization in the high-pressure regime). This peculiar result calls for a reinvestigation of high-pressure Mg3[BN2]N under different high-pressure high-temperature conditions along with physical property studies.

1 Introduction

Nitridoborates form with the alkali, alkaline earth and rare earth metals [1], [2], [3], [4]. They can be synthesized from the binary nitrides and hexagonal boron nitride [5] or via metathesis reactions [6], [7]. Among all ternary and quaternary nitridoborates, the linear or slightly bent [BN2]3− anion is the most frequent anionic motif. The structure-property relationships of these materials have recently been reviewed [8].

Among the 40 different [BN2]3− containing phases [8] only few nitridoborate nitrides have been reported: two modifications of Mg3[BN2]N [9], [10], [11], [12], Mg6Al2[BN2]2N4 and Mg6Ga2[BN2]2N4 [13], [14] and Ca3[BN2]N [15]. These phases are electron-precise and show complete anion ordering between [BN2]3− and N3−. The profound interest in these nitridoborate nitrides lies in the field of catalysis for the high-pressure high-temperature transformation of hexagonal to cubic boron nitride. Especially Mg3[BN2]N has repeatedly been discussed as an important intermediate product [16], [17], [18], [19], [20], [21], [22], [23], although the complete mechanism in not yet fully understood. The second latest research topic concerns luminescent materials. Various Eu2+-doped nitridoborate nitrides are promising candidates for red-emitting phosphors [12], [13], [14].

The dimorphism and the electronic structure of Mg3[BN2]N have not yet been studied in detail. Only the density of states of the normal-pressure modification has been reported [24], substantiating a direct band gap of ca. 0.88 eV at the Γ point. The theoretically obtained transformation pressure of ca. 2 GPa is close to the experimentally used pressure of 4 GPa at 1573 K [10].

Herein the theoretical study is extended to include both modifications of Mg3[BN2]N along with data for tetragonal Ca3[BN2]N [15], presenting energy-volume curves, density of states and the band structures. Also, an analysis of chemical bonding on the basis of overlap populations is performed.

2 Crystal chemistry

The crystal structures of the normal-pressure (NP) and high-pressure (HP) modifications of Mg3[BN2]N are presented in Fig. 1. The nitride anions in hexagonal NP-Mg3[BN2]N, space group P63/mmc, have trigonal bi-pyramidal magnesium coordination. These N@Mg5 units are condensed via common corners, leading to layers, which extend in the ab plane. The layers are stacked in ABAB sequence (a consequence of the 63 axis) and separated by the [BN2]3− units. The latter are coordinated by six magnesium cations in a strongly distorted octahedral manner.

Fig. 1: The crystal structures of NP- and HP-Mg3[BN2]N. Magnesium, boron and nitrogen atoms are drawn as blue, green and medium grey circles, respectively. The magnesium coordination of the isolated nitride anions and the B–N distances (Å) are emphasized.
Fig. 1:

The crystal structures of NP- and HP-Mg3[BN2]N. Magnesium, boron and nitrogen atoms are drawn as blue, green and medium grey circles, respectively. The magnesium coordination of the isolated nitride anions and the B–N distances (Å) are emphasized.

The transformation to HP-Mg3[BN2]N is a reconstructive phase transition and one observes different coordination modes. For the nitride anions the coordination number increases from five to six and the resulting octahedra are condensed to layers via common edges. According to the pressure-distance paradoxon, the high-pressure phase exhibits longer Mg–N distances (2.056–2.202 Å) as compared to NP-Mg3[BN2]N (2.040–2.046 Å). In the orthorhombic high-pressure phase we observe an increasing coordination number for the [BN2]3− units: 10 Mg2+ in an orthorhombically distorted bi-capped, square prismatic arrangement. The corresponding interatomic distances are listed in Table 1.

Table 1:

Interatomic distances (Å) in Mg3[BN2]N (Pmmm) and Ca3[BN2]N (P4/mmm).

Mg3[BN2]NCa3[BN2]N
Mg/Ca1:4N12.2022.510
2N22.5122.755
Mg/Ca2:1N12.0562.282
4N22.2482.554
B:2N21.3381.352

At this point it is interesting to compare the structure of HP-Mg3[BN2]N with that of the higher congener Ca3[BN2]N. Both compounds follow the empirical pressure-homologue rule: HP-Mg3[BN2]N crystallizes approximately with the structure of the higher congener. The striking difference is the space group symmetry, Pmmm for HP-Mg3[BN2]N (the orthorhombic distortion has experimentally clearly been proven on the basis of powder X-ray diffraction in [10]) and P4/mmm for Ca3[BN2]N. Thus, HP-Mg3[BN2]N crystallizes with a translationengleiche subgroup of P4/mmm. The corresponding group-subgroup scheme is presented in the Bärnighausen formalism [26], [27] in Fig. 2. From this scheme it is readily evident that the distortion only concerns the course of the lattice parameters: a=b=3.5494 Å for tetragonal Ca3[BN2]N as compared to a=3.0933 and b=3.1336 Å for HP-Mg3[BN2]N. There is no need for other free positional parameters. These structural peculiarities are addressed based on ab initio calculations in the following.

Fig. 2: Group-subgroup scheme in the Bärnighausen formalism [26, 27] for the structures of Ca3[BN2]N [15] and HP-Mg3[BN2]N [9]. The index for the translationengleiche symmetry reduction (t) and the evolution of the atomic parameters is given.
Fig. 2:

Group-subgroup scheme in the Bärnighausen formalism [26, 27] for the structures of Ca3[BN2]N [15] and HP-Mg3[BN2]N [9]. The index for the translationengleiche symmetry reduction (t) and the evolution of the atomic parameters is given.

3 Computational details

Within the accurate quantum mechanical density functional theory DFT [28], [29] two methods have been used complementarily.

The first calculations are performed using the Vienna ab initio simulation package (VASP) code [30], [31] which allows geometry optimization, total energy calculations, and establishing the energy-volume equations of state (EOS). This method is based on the projector augmented wave (PAW) [31], [32] which is executed for atomic potentials built within the generalized gradient approximation (GGA) scheme following Perdew, Burke and Ernzerhof (PBE) [33]. This exchange-correlation (XC) scheme is preferred to the local density approximation LDA [34] known to underestimate interatomic distances and energy band gaps. Relaxing atoms of the different crystal setups is based on the conjugate-gradient algorithm [35] and the tetrahedron method with Blöchl corrections [36] as well as a Methfessel-Paxton [37] scheme are applied for both geometry relaxation and total energy calculations. Brillouin zone (BZ) integrals are approximated using a special k-point sampling of Monkhorst and Pack [38]. Optimization of structural parameters has been performed until the forces on the atoms were less than 0.02 eV Å−1 and all stress components less than 0.003 eV Å−3. Calculations are converged at an energy cut-off of 350 eV for the plane-wave basis set with respect to the k-point integration with a starting mesh of 6×6×6 up to 12×12×12 for best convergence and relaxation to zero strains.

The second method fully accounts for the electronic structure, electronic band structure, site projected density of states (PDOS) and properties of chemical bonding based on the overlap matrix (Sij) with the COOP criterion [39]. It is based on the full potential augmented spherical wave (ASW) method [40], [41] and the generalized gradient approximation GGA [33] for exchange-correlation effects. In the minimal ASW basis set, outermost shells are used for representation of valence states and the matrix elements are constructed using partial waves up to lmax+1=3 for Ca, Mg and lmax+1=2 for B and N. Self-consistency is achieved when charge transfers and energy changes between two successive cycles are ΔQ<10−8 and ΔE<10−6 eV, respectively. The BZ integrations are performed using the linear tetrahedron method within the irreducible hexagonal and orthorhombic wedges following Blöchl [36].

4 Electronic structure and chemical bonding

4.1 Energy-volume curves

Table 2 shows the starting experimental and calculated atomic positions and structure parameters of the two compounds. There is a reasonably good agreement between experiment and calculations as well as with the computations by Medvedeva et al. [24] and Wang et al. [25]. Further, following the introduction of the orthorhombic form of Mg3[BN2]N crystallizing with a translationengleiche subgroup of the tetragonal group P4/mmm of Ca3[BN2]N with group-subgroup correspondence on one hand and on the other hand in view of the computational results by Wang et al. [25] where Mg3[BN2]N is computed in the tetragonal Ca3[BN2]N type structure, additional calculations were carried out in order to clearly evaluate the energetic consequences of the closeness of the two structures (orthorhombic versus tetragonal). Such energy discriminations between the three forms of Mg3[BN2]N can be assessed from the energy (E)-volume (V) values calculated for the optimized structure parameters (Table 2). The E(V) data, which then present a quadratic evolution, are fitted with a Birch EOS [42] up to the third order following:

Table 2:

Experimental and (calculated) crystal data.

CompoundSpace groupa, b, c (Å)z (Ca/Mg); z (N1)
Ca3[BN2]NP4/mmm3.5494 (3.52); 8.2136 (8.31)0.2778 (0.28); 0.3354 (0.33)
Mg3[BN2]NPmmm3.0933 (3.10); 3.1336 (3.12); 7.7005 (7.84)0.2670 (0.25); 0.3262 (0.32)
Mg3[BN2]NP63/mmc3.5445 (3.52); 16.0354 (16.2)0.1228 (0.13); 0.0851 (0.09)
Mg3[BN2]NP4/mmm(3.122); (7.91)(0.258); (0.33)
  1. Orthorhombic/tetragonal: Ca1/Mg1(0 0 0), Ca2/Mg2(1/2 1/2 z), N1(0 0 z), N2(1/2 1/2 0), B(0 0 1/2); hexagonal: Mg1(0 0 1/4), Mg2(1/3 2/3 z), N1(1/3 2/3 1/4), N2(0 0 z), B(0 0 0).

E(V)=E0(V0)+98V0B0[(V0/V)2/31]2+916B0(B4)V0[(V0/V)2/31]3

The latter provides equilibrium parameters for the energy, the volume, the zero pressure bulk modulus and its pressure derivative: E0, V0, B0 and B′, respectively. The obtained values with accurate goodness of fit χ2 ~10−5/10−4 magnitudes are displayed in the insert of Fig. 3 for tetragonal Ca3[BN2]N and the three forms of Mg3[BN2]N. For the former the volume comes close to the experiment. This is opposed to the smaller volume of the Mg compounds which come close to ~76 Å3 per formula unit (FU) for the orthorhombic and tetragonal forms and larger (V~88 Å3 per FU) for the hexagonal normal-pressure modification. The consequence is a large change of the magnitudes of the corresponding bulk moduli, i.e. the smallest one occurs for hexagonal Mg3[BN2]N: this follows from the fact that the larger the volume the larger the compressibility or the smaller the bulk module. The relevant result for the Mg3[BN2]N structural models is the proposition of the hexagonal form as the ground state structure which exhibits by far the lowest energy. The orthorhombic modification is very close to the tetragonal one, although the latter is found at slightly lower energy and smaller volume (i.e. a slightly larger bulk modulus of 140 GPa). One might suggest that the orthorhombic and tetragonal modification form under different high-pressure high-temperature conditions.

Fig. 3: Energy-volume curves and fit values from Birch EOS calculations for Ca3[BN2]N and the NP, HP and hypothetic Ca3[BN2]N-type modifications of Mg3[BN2]N.
Fig. 3:

Energy-volume curves and fit values from Birch EOS calculations for Ca3[BN2]N and the NP, HP and hypothetic Ca3[BN2]N-type modifications of Mg3[BN2]N.

4.2 Electronic band structure and density of states

Figure 4 provides the band structure and the site projected density of states (PDOS) of Ca3[BN2]N. The compound is characterized by a semi-conducting behavior with a direct gap at the Γ point, i.e. at the Brillouin zone (BZ) center with ΓV→ΓC of ~1 eV magnitude. The zero energy is with respect to the top of the valence band (VB). At the VB bottom two distinct types of bands are identified corresponding to s-like states whereas p-like states occupy the energy range from –6 eV to –EV. The site projected DOS clearly show different (chemical) characters of N1 versus N2, i.e. with less electronegative behavior for the latter and the presence of its p-PDOS at the top of the valence band (VB). Note the energetically low-lying s states of N1 at –15 eV whereas the s-N2 states are at approximately –10 eV.

Fig. 4: Ca3[BN2]N: Electronic band structure along major lines of the tetragonal Brillouin zone (top) and site-projected density-of-states (DOS) (bottom).
Fig. 4:

Ca3[BN2]N: Electronic band structure along major lines of the tetragonal Brillouin zone (top) and site-projected density-of-states (DOS) (bottom).

Similar features are observed for hexagonal Mg3[BN2]N (normal-pressure structure) in Fig. 5 showing the band structure, with a smaller, direct ΓV→ΓC band gap with albeit less ionic character, i.e. a broader VB with p states over ~7 eV due to the presence of Mg. The top of the VB is with respect to EV. More drastic changes are observed for the orthorhombic form which exhibits metallic behavior with flat localized bands (bottom panel) still with N-p states now at EF, and the VB is largely broadened for all states. Then the transition from NP-Mg3[BN2]N to HP-Mg3[BN2]N involves metallization (higher degree of delocalization in the high-pressure phase).

Fig. 5: Band structures of NP-Mg3[BN2]N (top) and HP-Mg3[BN2]N (bottom) along the major lines of the irreducible wedge of the hexagonal (top) and orthorhombic (bottom) Brillouin zones.
Fig. 5:

Band structures of NP-Mg3[BN2]N (top) and HP-Mg3[BN2]N (bottom) along the major lines of the irreducible wedge of the hexagonal (top) and orthorhombic (bottom) Brillouin zones.

The PDOS given in Fig. 6 mirrors the band structure observations for the transition from NP-Mg3[BN2]N to HP-Mg3[BN2]N. In the middle panel (hypothetical tetragonal Mg3[BN2]N), besides similar features to the actual compound as for instance the relative positions of the s and p PDOS, a band gap is observed, similar to Ca3[BN2]N with, however, a smaller magnitude due to the less ionic character of Mg as compared to Ca.

Fig. 6: Site-projected density-of-states (PDOS) for NP-Mg3[BN2]N (top) hypothetic tetragonal Mg3[BN2]N (middle) and HP-Mg3[BN2]N (bottom).
Fig. 6:

Site-projected density-of-states (PDOS) for NP-Mg3[BN2]N (top) hypothetic tetragonal Mg3[BN2]N (middle) and HP-Mg3[BN2]N (bottom).

Insofar as both experimental orthorhombic HP-Mg3[BN2]N and hypothetical tetragonal Mg3[BN2]N are very close in energy (only 0.02 eV difference) with slightly more stabilization for the latter (cf. Fig. 3/EOS), and in view of the unexpected metallization in such a compound, it becomes questionable as to whether the stable high-pressure modification of Mg3[BN2]N should not be tetragonal. This inquiry needs further experimental investigation.

4.3 Crystal orbital overlap population analyses

Bonding is mainly operating in the p states-dominated VB part, i.e. extending over 6 eV below EV. The plot for Ca3[BN2]N (Fig. 7) reveals interesting features whereby B–N1 covalent bonding within the [BN2]3− units dominates over the lower energy part of VB from –6 to –3 eV, whereas the ionic Ca–N2 bonding shows COOP at the higher energy part extending over 2 eV below EV; two different features appear for the anions [BN2]3− and N3−.

Fig. 7: Crystal orbital overlap population analyses for Ca3[BN2]N.
Fig. 7:

Crystal orbital overlap population analyses for Ca3[BN2]N.

For Mg3[BN2]N (Fig. 8), in spite of features similar to Ca3[BN2]N, less energy differentiation between the B–N and Mg–N COOP due to the decrease of iconicity can be observed. These features stress further the more covalent character of the Mg–N bonds versus Ca–N.

Fig. 8: Crystal orbital overlap population analyses for NP-Mg3[BN2]N (top) and HP-Mg3[BN2]N (bottom).
Fig. 8:

Crystal orbital overlap population analyses for NP-Mg3[BN2]N (top) and HP-Mg3[BN2]N (bottom).

4.4 Bader analyses

We further assess these results by analyzing the charge density issued from the self-consistent calculations using the AIM (atoms in molecules theory) approach [43] developed by Bader who proposed an intuitive way of splitting molecules into atoms as based on electronic charge density. The charge density reaches a minimum between atoms and this is a natural region to separate them from each other. Such an analysis can be useful when trends between similar compounds are examined as it is done herein; they do not constitute a tool for evaluating absolute ionizations. Bader’s analysis is done using a fast algorithm operating on a charge density grid. In order to include core electrons within the PAW method in VASP, use is made of the “LAECHG=.TRUE.” parameter in the ‘INCAR’ control file followed by operating a summation of core and valence charge densities using a simple perl program. The results of computed charge changes (ΔQ) are such that they lead to neutrality when the respective multiplicities are accounted for; the obtained values are:

Ca3[BN2]NQ(Ca1)=+0.98; Q(Ca2)=+1.09; Q(N2)=–1.20; Q(N1)=–2.40; Q(B)=+2.84

For one formula unit: Q(Ca1)+2×Q(Ca1)+Q(N2) +2×Q(N1)+Q(B)=0 (this holds also for the remaining phases)

Hexagonal NP-Mg3[BN2]NQ(Mg1)=+1.03; Q(Mg2)=+1.12; Q(N2)=–1.43; Q(N1)=–2.40; Q(B)=+2.96

Orthorhombic HP-Mg3[BN2]NQ(Mg1)=+1.10; Q(Mg2)=+1.13; Q(N2)=–1.45;Q(N1)=–2.42; Q(B)=+2.93

Hypothetic tetragonal Mg3[BN2]NQ(Mg1)=+1.11; Q(Mg2)=+1.14; Q(N2)=–1.45; Q(N1)=–2.42; Q(B)=+2.90

The resulting charges fully underline the different chemical bonding behavior of the nitridoborate and nitride anions and support further the DOS and COOP observations.

5 Conclusion

Electronic structure calculations quantify the different bonding situation of the covalently bonded [BN2]3− and isolated N3− anions in the nitridoborate nitrides Ca3[BN2]N and Mg3[BN2]N. Total energy calculations reveal the hexagonal normal-pressure Mg3[BN2]N structure as the ground state structure. The high-pressure phase of Mg3[BN2]N is reported in space group Pmmm, an orthorhombically distorted variant of the Ca3[BN2]N type. A comparative study of orthorhombic Mg3[BN2]N and a hypothetic tetragonal, Ca3[BN2]N-type high-pressure phase show substantial metallization of the orthorhombic variant. These unusual bonding characteristics call for a reinvestigation of the HP-Mg3[BN2]N structure along with property studies.

Acknowledgement

Part of the calculations has been performed on the computer clusters of MCIA/University of Bordeaux and workstations of CSR-USEK.

References

[1] J. Goubeau, W. Anselment, Z. Anorg. Allg. Chem. 1961, 310, 248.10.1002/zaac.19613100410Search in Google Scholar

[2] M. Somer, U. Herterich, J. Curda, W. Carillo-Cabrera, K. Peters, H. G. von Schnering, Z. Anorg. Allg. Chem. 1997, 623, 18.10.1002/zaac.19976230104Search in Google Scholar

[3] M. Somer, U. Herterich, J. Čurda, W. Carillo-Cabrera, A. Zürn, K. Peters, H. G. von Schnering, Z. Anorg. Allg. Chem. 2000, 626, 625.10.1002/(SICI)1521-3749(200003)626:3<625::AID-ZAAC625>3.0.CO;2-4Search in Google Scholar

[4] B. Blaschkowski, H. Jing, H.-J. Meyer, Angew. Chem. Int. Ed. 2002, 41, 3322.10.1002/1521-3773(20020916)41:18<3322::AID-ANIE3322>3.0.CO;2-8Search in Google Scholar

[5] M. Häberlen, J. Glaser, H.-J. Meyer, J. Solid State Chem. 2005, 178, 1478.10.1016/j.jssc.2005.02.008Search in Google Scholar

[6] H. Jing, B. Blaschkowski, H.-J. Meyer, Z. Anorg. Allg. Chem. 2002, 628, 1955.10.1002/1521-3749(200209)628:9/10<1955::AID-ZAAC1955>3.0.CO;2-MSearch in Google Scholar

[7] H.-J. Meyer, Dalton Trans. 2010, 39, 5973.10.1039/c001031fSearch in Google Scholar

[8] R. Pöttgen, O. Reckeweg, Z. Kristallogr. 2017, 232, in press. DOI: 10.1515/zkri-2017-2043.10.1515/zkri-2017-2043Search in Google Scholar

[9] H. Hiraguchi, H. Hashizume, O. Fukunga, A. Takenaka, M. Sakata, J. Appl. Crystallogr. 1991, 24, 286.10.1107/S0021889891001334Search in Google Scholar

[10] H. Hiraguchi, H. Hashizume, S. Sasaki, S. Nakano, O. Fukunaga, Acta Crystallogr. 1993, B49, 478.10.1107/S0108768192013533Search in Google Scholar

[11] O. Reckeweg, Carbido- und Nitridoborate der Erdalkali- und Seltenerdmetalle: Synthese, Kristallstrukturen und Eigenschaften, Dissertation, Universität Tübingen, Tübingen (Germany) 1998.Search in Google Scholar

[12] J. Schölch, T. Dierkes, D. Enseling, M. Ströbele, T. Jüstel, H.-J. Meyer, Z. Anorg. Allg. Chem. 2015, 641, 803.10.1002/zaac.201400487Search in Google Scholar

[13] M. Ströbele, K. Dolabdjian, D. Enseling, D. Dutczak, B. Mihailova, T. Jüstel, H.-J. Meyer, Eur. J. Inorg. Chem. 2015, 1716.10.1002/ejic.201403116Search in Google Scholar

[14] D. Dutczak, K. M. Wurst, M. Ströbele, D. Enseling, T. Jüstel, H.-J. Meyer, Eur. J. Inorg. Chem. 2016, 861.10.1002/ejic.201501090Search in Google Scholar

[15] M. Häberlen, J. Glaser, H.-J. Meyer, Z. Anorg. Allg. Chem. 2002, 628, 1959.10.1002/1521-3749(200209)628:9/10<1959::AID-ZAAC1959>3.0.CO;2-ZSearch in Google Scholar

[16] T. Endo, O. Fukunga, M. Iwata, J. Mater. Sci. 1979, 14, 1676.10.1007/BF00569290Search in Google Scholar

[17] V. P. Elyutin, N. I. Polushin, K. P. Burdina, V. P. Polyakov, Ya. A. Kalashnikov, K. N. Semenenko, Yu. A. Lavlov, Dokl. Akad. Nauk SSSR1981, 259, 112.Search in Google Scholar

[18] C. Hohlfeld, J. Mater. Sci. Lett.1989, 8, 1082.10.1007/BF01730494Search in Google Scholar

[19] S. Nakano, H. Ikawa, O. Fukunaga, J. Am. Ceram. Soc. 1992, 75, 240.10.1111/j.1151-2916.1992.tb05477.xSearch in Google Scholar

[20] G. Bocquillon, C. Loriers-Susse, J. Loriers, C. R. Acad. Sci. Paris, Sér. II1992, 315, 1069.Search in Google Scholar

[21] A. N. Zhukov, K. P. Burdina, K. N. Semenenko, Russ. J. Gen. Chem. 1994, 64, 1117.Search in Google Scholar

[22] H. Lorenz, I. Orgzall, E. Hinze, Diamond Rel. Mater. 1995, 4, 1050.10.1016/0925-9635(95)00276-6Search in Google Scholar

[23] A. N. Zhukov, K. P. Burdina, K. N. Semenenko, Russ. J. Gen. Chem. 1996, 66, 1043.Search in Google Scholar

[24] N. I. Medvedeva, Yu. E. Medvedeva, A. L. Ivanovskii, Dokl. Phys. Chem. 2001, 379, 168.10.1023/A:1019277711896Search in Google Scholar

[25] J. Wang, Y. Zhou, Z. Lin, T. Liao, J. Solid State Chem. 2006, 179, 2739.10.1016/j.jssc.2006.05.016Search in Google Scholar

[26] H. Bärnighausen, Commun. Math. Chem. 1980, 9, 139.Search in Google Scholar

[27] U. Müller, Z. Anorg. Allg. Chem. 2004, 630, 1519.10.1002/zaac.200400250Search in Google Scholar

[28] P. Hohenberg, W. Kohn, Phys. Rev.1964, 136, B864.10.1103/PhysRev.136.B864Search in Google Scholar

[29] W. Kohn, L. J. Sham, Phys. Rev.1965, 140, A1133.10.1103/PhysRev.140.A1133Search in Google Scholar

[30] G. Kresse, J. Furthmüller, Phys. Rev. B1996, 54, 11169.10.1103/PhysRevB.54.11169Search in Google Scholar PubMed

[31] G. Kresse, J. Joubert, Phys. Rev. B1999, 59, 1758.10.1103/PhysRevB.59.1758Search in Google Scholar

[32] P. E. Blöchl, Phys. Rev. B1994, 50, 17953.10.1103/PhysRevB.50.17953Search in Google Scholar

[33] J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 1996, 77, 3865.10.1103/PhysRevLett.77.3865Search in Google Scholar PubMed

[34] D. M. Ceperley, B. J. Alder, Phys. Rev. Lett. 1980, 45, 566.10.1103/PhysRevLett.45.566Search in Google Scholar

[35] W. H. Press, B. P. Flannery, S. A. Teukolsky, W. T. Vetterling, Numerical Recipes, Cambridge University Press, New York, 1986.Search in Google Scholar

[36] P. E. Blöchl, O. Jepsen, O. K. Andersen, Phys. Rev. B1994, 49, 16223.10.1103/PhysRevB.49.16223Search in Google Scholar

[37] M. Methfessel, A. T. Paxton, Phys. Rev. B1989, 40, 3616.10.1103/PhysRevB.40.3616Search in Google Scholar

[38] H. J. Monkhorst, J. D. Pack, Phys. Rev. B1976, 13, 5188.10.1103/PhysRevB.13.5188Search in Google Scholar

[39] R. Hoffmann, Angew. Chem. Int. Ed. Engl. 1987, 26, 846.10.1002/anie.198708461Search in Google Scholar

[40] A. R. Williams, J. Kübler, C. D. Gelatt, Phys. Rev. B1979, 19, 6094.10.1103/PhysRevB.19.6094Search in Google Scholar

[41] V. Eyert, The Augmented Spherical Wave Method – A Comprehensive Treatment, Lecture Notes in Physics, Springer, Heidelberg, 2007, chapter 719.Search in Google Scholar

[42] F. Birch, J. Geophys. Res. B31978, 83, 1257.10.1029/JB083iB03p01257Search in Google Scholar

[43] R. F. W. Bader, Chem. Rev. 1991, 91, 893.10.1021/cr00005a013Search in Google Scholar

Received: 2017-2-26
Accepted: 2017-3-31
Published Online: 2017-5-11
Published in Print: 2017-5-24

©2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 17.5.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2017-0035/html
Scroll to top button