Skip to content
Publicly Available Published by De Gruyter March 30, 2021

Coloring and distortion variants of the bcc packing and for the aristotypes BaAl4 and CeMg2Si2

  • Stefan Seidel and Rainer Pöttgen EMAIL logo

Abstract

The huge number of intermetallic structure types with many representatives calls for structural systemization. The combination of crystal chemistry with group theory is an efficient tool for such systemization and can be displayed in a concise and compact way via group-subgroup schemes. The present overview deals with such group-subgroup schemes (Bärnighausen trees) for coloring and distortion variants of the bcc packing as well as superstructures that derive from the aristotypes BaAl4 and CeMg2Si2.

1 Introduction

In the early times of X-ray structure determination, names for structure types were assigned on the basis of the Strukturbericht designation: A for elements, B for AB compounds, … D for AmBn compounds, … L for alloys, O for organic structures etc. [1], [2], [3], [4], [5], [6]. With the broad use of X-ray diffraction and the increasing number of compounds with complex compositions it soon became clear that this notation quickly reached its limits.

The Pearson data base [7], the TYPIX compilation [8] and The AFLOW Library of Crystallographic Prototypes [9], [10] use a pure crystallographic fingerprint; the space group symbol, the Pearson code (which includes the number of atoms within the unit cell) and the Wyckoff sequence. With this fingerprint it is possible to find all known compounds which have that given structural arrangement. Nevertheless, also this fingerprint has its limitations when it comes to the crystal-chemical interpretation. To give an example, CsCl and ErMg both crystallize in space group Pm3m, the Pearson code cP2 and the Wyckoff sequence ba. The crystal-chemical difference concerns the chemical bonding, i.e. salt-like CsCl versus intermetallic ErMg. Consequently, both compounds are rather isopointal [11], [12] than isotypic.

A further crystal-chemical feature concerns superstructures and distortion variants that derive from a structure type of higher symmetry. Drawing back to CsCl and ErMg as examples, these are the many bcc superstructures [13], [14]. Such superstructure and distortion variants are related by group-subgroup schemes. The compact and concise Bärnighausen formalism [15], [16], [17], [18] facilitates the crystal-chemical understanding of larger structural families. So-called Bärnighausen trees have been reported for the families of perovskite [19], [20] and rutile [21], [22] related structures in the field of ionic solids as well as for the intermetallic families of the aristotypes AlB2 [23] and BaAl4 [24], [25], [26], [27], [28], covering several thousand compounds that are crystal chemically related. Especially in the field of intermetallics [14, 29], these group-subgroup schemes are helpful for the understanding of coloring problems (introduced by the Miller group [30], [31], [32]; which Wyckoff position is occupied by which element?).

Herein we summarize some new crystal-chemical data that relate the structures of some inorganic (mainly intermetallic) compounds by group-subgroup schemes. These compounds are derived from the bcc packing or from the aristotypes BaAl4 and CeMg2Si2.

2 Ordering variants of the bcc packing

The first examples stem from the bcc packing, one of the basic metal structure types. The simplest ordering variant is the CsCl type. The Cs+/Cl ordering induces a klassengleiche symmetry reduction from Im3m to Pm3m and this is detectable in the diffraction patterns through superstructure reflections. This is the textbook example [13, 29] for bcc superstructures.

Herein we concentrate on klassengleiche bcc superstructures of index 3 (k3). This ordering is possible for a combination of two elements (typical AB2 compounds) that are different in size. The space group symmetry is then I4/mmm, with the Pearson code tI6 and the Wyckoff sequence ea. The Pearson data base [7] lists almost 200 entries for this crystallographic fingerprint; however, these compounds show distinctly different bonding peculiarities as a consequence of the size and the chemical nature of the atoms. A reasonable classification is possible via the c/a ratio of the unit cell. This was already reported by Parthé and coworkers in the early paper on the Structure Tidy program [12] and in a compilation on Elements of Inorganic Structural Chemistry [33].

From the 200 entries we present some examples in Figure 1. The respective structures are arranged with increasing c/a ratio. We start with the intermetallic compounds MoSi2 [34], AlCr2 [35], CuZr2 [36], BePd2 [37], and CdTi2 [38]. Their c/a ratio increases from 2.46 for MoSi2 to 4.68 for CdTi2. The structural flexibility for these phases is achieved mainly by changes in the lattice parameters. For most phases (where diffraction data is available) the z parameters of the 4e site do not substantially deviate from the subcell value of 1/3. The drastic changes of the c/a ratios then influence the bonding peculiarities. To account for these differences, these AB2 phases were devided into the MoSi2 (c/a ratios <3) and CuZr2 (c/a ratios >3) branches. Although these phases all have the same crystallographic fingerprint, those of the different branches are only isopointal [11], [12]. It is interesting to note that also LaI2 [39] adopts this fingerprint. The excess electron of trivalent lanthanum leads to metallic behavior for LaI2 [40].

Figure 1: Projections of several crystal structures with the crystallographic fingerprint of space group type I4/mmm, Pearson code tI6 and Wyckoff sequence ea. The structures are presented with increasing c/a ratio. For details see text.
Figure 1:

Projections of several crystal structures with the crystallographic fingerprint of space group type I4/mmm, Pearson code tI6 and Wyckoff sequence ea. The structures are presented with increasing c/a ratio. For details see text.

We now turn to the structure of calcium acetylide CaC2 [41]. This salt-like carbide has the same crystallographic fingerprint, but a drastically different bonding situation. The carbon atoms in the C22 acetylide unit have a C–C distance of 120 pm and this leads to a collapse of the c axis and thus a drastic decrease of the c/a ratio to 1.64. The whole series of AC2, AEC2, REC2 (A = alkali cation, AE = alkaline earth cation, RE = rare earth cation) acetylides, the AO2 superoxides, the AEO2 peroxides and the AEN2 diazenides [7] are isotypic (and only isopointal with the compounds discussed above!) with CaC2 and form the third branch within the family of bcc superstructures.

A unique case is the XeF2 structure [42], [43]. In contrast to the intermetallic compounds and the ionic compound CaC2 discussed above, XeF2 is a molecular compound with covalent intramolecular and van der Waals intermolecular bonding. The shift of the fluorine atoms in XeF2 is opposite to the shift of the carbon atoms in CaC2. This also leads to a collapse of the c axis and a small c/a ratio and to the formation of linear XeF2 units. The low-temperature modification of KrF2 [44] is isotypic with XeF2.

The packing motif of the XeF2 molecules is preserved in several ternary compounds which can be considered as ‘filled’ variants. This has first been observed for the oxidomercurate Na2[HgO2] [45] and later for other oxides, nitrides, hydrides and carbides [7]. As an example we present the packing of the linear polyanionic units [BC2]5−, [PdD2]2− and [HgO2]2− in Sc2BC2 [46], Na2PdD2 [47] and Na2HgO2 [45], respectively, in Figure 2. The c/a ratio of these examples is again dominated by the nature of the atoms forming the respective structure. It is worthwhile to note that the structures of the cyanamides Na2CN2 [48] and K2CN2 [49] are monoclinic distortion variants of the Na2HgO2 type. The complete Bärnighausen tree is presented for Na2CN2 in ref. [48].

Figure 2: The crystal structures of Sc2BC2 [46], Na2PdD2 [47] and Na2HgO2 [45]. The linear polyanionic units [BC2]5−, [PdD2]2− and [HgO2]2− are emphasized.
Figure 2:

The crystal structures of Sc2BC2 [46], Na2PdD2 [47] and Na2HgO2 [45]. The linear polyanionic units [BC2]5−, [PdD2]2− and [HgO2]2− are emphasized.

The description of the ‘filled’ variants is only intuitive and simply guided by comparison with the XeF2 molecules. A closer look at the unit cells of Sc2BC2, Na2PdD2 and Na2HgO2 (Figure 2) readily reveals that the substructures [Sc2B], [Na2Pd] and [Na2Hg] correspond to a MoSi2-type arrangement with the carbon, deuterium and oxygen atoms filling distorted C@Sc5B, D@Na5Pd and O@Na5Hg octahedral voids in an ordered manner. The corresponding group-subgroup scheme for the isotypic carbide Th2NiC2 [50] is presented in Figure 3.

Figure 3: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of Th2NiC2 [50] and YCoC [51] along with the evolution of the atomic parameters for the octahedral voids (VAC denotes vacancies). In each carbide only one void (not shaded) is occupied.
Figure 3:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of Th2NiC2 [50] and YCoC [51] along with the evolution of the atomic parameters for the octahedral voids (VAC denotes vacancies). In each carbide only one void (not shaded) is occupied.

The tungsten structure (the aristotype) exhibits six crystallographically equivalent distorted octahedral voids per unit cell (Wyckoff site 6b) which correspond to the middle of the edges and the faces. For steric reasons, only partial occupancy of the voids is possible and this can only be achieved through a symmetry reduction. In going to the protactinium structure (the compressed version of the tungsten type) with a translationengleiche symmetry reduction of index 3 (t3) to I4/mmm we get a 4 + 2 splitting of the 6b site. The subsequent tripling of the unit cell along the c axis (isomorphic transition of index 3) leads to a further 4 + 2 splitting of all three sites, and in total we obtain the two sites for thorium and nickel and four possible octahedral voids of which only the Wyckoff site 4e is occupied by carbon. This leads to the linear carbidonickelate units presented in Figure 4.

Figure 4: The crystals structures of Th2NiC2 [50] and YCoC [51]. Thorium (yttrium), nickel (cobalt) and carbon atoms are drawn as light gray, blue and black circles, respectively. The [NiC2] and [CoC] substructures are emphasized.
Figure 4:

The crystals structures of Th2NiC2 [50] and YCoC [51]. Thorium (yttrium), nickel (cobalt) and carbon atoms are drawn as light gray, blue and black circles, respectively. The [NiC2] and [CoC] substructures are emphasized.

A similar ordered occupancy of a bcc derived metal substructure has been observed for YCoC [51], SrLiN [52] and CaNiN [53]. The symmetry reduction is explained for the prototype YCoC in Figure 3. The first step is the decentering of the lattice in going to the CsCl structure which leads to a 3 + 3 splitting of the 6b site. This is the decisive difference to the body-centered branch discussed before. The subsequent translationengleiche step of index 3 (t3) leads to the tetragonal CuAu structure (P4/mmm) and in the final step the unit cell is doubled along the c axis via a klassengleiche symmetry reduction of index 2 (k2) to space group P42/mmc. This model leaves five crystallographically independent octahedral voids of which only the Wyckoff site 2c is filled with carbon atoms. This leads to linear [CoC] chains which change their orientation by 90° in every other layer (Figure 4) as a consequence of the 42 screw axis. The isotypic nitrides contain [LiN], respectively [NiN] chains.

Especially the [CoC] and [NiN] chains have attracted considerable interest by theoreticians [54], [55], [56] with respect to a possible Peierls distortion. Keeping the [NiC2] units in Th2NiC2 and the infinite [CoC] chains in YCoC in mind, the most plausible Peierls distortion would by a pairwise shift of the carbon atoms towards the cobalt atoms as emphasized by red arrows in Figure 5. This ordering leaves two crystallographically independent cobalt sites, emphasized by blue and reddish color in Figure 5. The resulting structural model for a Peierls-distorted YCoC variant would crystallize in space group P42/mnm (Figure 6). The klassengleiche symmetry reduction of index 2 (k2) leads to the splitting of the cobalt site and leaves a free x parameter for the carbon atoms, allowing the formation of [CoC2] pairs, similar to [NiC2] in Th2NiC2.

Figure 5: Cutouts of the structures of YCoC [51] and a possible Peierls-distorted variant (see the Bärnighausen tree in Figure 6). Only one layer is shown. The red arrows indicate the shift of the carbon atoms during Peierls distortion. The two crystallographically independent cobalt atoms in the distorted model are shaded in the same manner as in the Bärnighausen tree.
Figure 5:

Cutouts of the structures of YCoC [51] and a possible Peierls-distorted variant (see the Bärnighausen tree in Figure 6). Only one layer is shown. The red arrows indicate the shift of the carbon atoms during Peierls distortion. The two crystallographically independent cobalt atoms in the distorted model are shaded in the same manner as in the Bärnighausen tree.

Figure 6: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for a possible Peierls distorted variant of YCoC [51] and the nitridometalate Sr2[CoLiN2] [57], [58]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given. The differently shaded cobalt positions correspond to the ordering presented in this Figure.
Figure 6:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for a possible Peierls distorted variant of YCoC [51] and the nitridometalate Sr2[CoLiN2] [57], [58]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given. The differently shaded cobalt positions correspond to the ordering presented in this Figure.

The plausibility of this model is nicely underlined by the structure of isotypic nitrodometalate Sr2[CoLiN2] [57], [58], where ordered chains –N–Co–N–Li– are embedded in the strontium matrix. It is remarkable that the Co+ cations show ferromagnetic ordering at a comparatively high Curie temperature of 44 K [58]. Ca2[Li1.25Cu0.75N2], Ca2[Li0.93Cu1.07N2] [59] and Ca2[Li1.18Fe0.82N2] [60] are further representatives of this structure type.

A look into the Pearson data base [7] for the fingerprint #136, tP12, gdba yields further representatives. The structure of Li3[BN2] [61], [62] shows linear [BN2] nitridoborate units that are linked via one kind of lithium cations to linear chains. This is similar to the oxidonickelates(I) A3[NiO2] with A = K, Rb and Cs [63], [64], where one of the alkali cations forms linear chains with the [NiO2] units. It is interesting to note that K3[NiO2] exhibits a phase transition at around 423 K. α-K3[NiO2] crystallizes in the polar space group P41212. The complete Bärnighausen tree is presented in ref. [64].

Keeping the temperature-induced phase transition of the nickelate in mind we redraw to the xenon difluoride structure which shows two different pressure-induced phase transitions [65]. The polymorphs HP1-XeF2 (Immm) and HP2-XeF2 (Pnma) are derived from the normal-pressure modification NP-XeF2 (I4/mmm) via two different symmetry reductions (Figure 7); I4/mmm is the common supergroup of Immm and Pnma. HP1-XeF2 shows just an orthorhombic distortion of the unit cell (loss of the four-fold axis) and this proceeds via a translationengleiche transition of index 2 (t2) from I4/mmm to Immm along with a splitting of the axis to a = 329.5 and b = 325.7 pm. The XeF2 units remain in parallel alignment in the HP1-XeF2 polymorph (Figure 8). The intermetallic compounds ReSi2 [66] and TiPd2 [67] are isopointal [11], [12] with HP1-XeF2; however, as discussed above, with larger c/a and c/b ratios.

Figure 7: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of the normal-pressure (NP) and high-pressure (HP) polymorphs of XeF2 [65]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given. Note that the relationship between the XeF2 modifications and the intermetallic phases is solely isopointal [11], [12]. For further details see text.
Figure 7:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of the normal-pressure (NP) and high-pressure (HP) polymorphs of XeF2 [65]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given. Note that the relationship between the XeF2 modifications and the intermetallic phases is solely isopointal [11], [12]. For further details see text.

Figure 8: Arrangement of the molecules in the normal-pressure (NP) and high-pressure (HP) polymorphs of XeF2 [65]. Xenon and fluorine atoms are drawn as blue and magenta circles.
Figure 8:

Arrangement of the molecules in the normal-pressure (NP) and high-pressure (HP) polymorphs of XeF2 [65]. Xenon and fluorine atoms are drawn as blue and magenta circles.

The HP2-XeF2 structure (Figure 8) shows substantial tilting of the XeF2 units. The symmetry reduction proceeds in three steps (Figure 7) via the space groups Fmmm and Cmcm. The essential step is the unit cell enlargement during the transition from I4/mmm to Fmmm. A search in the Pearson data base [7] for the intermediate space groups Fmmm and Cmcm revealed the representatives TiSi2 (observed in thin films) [68] and VAu2 [62, 69]. Again, these two intermetallics are only isopointal [11], [12], since the bonding peculiarities are different.

We now turn to the equiatomic compounds ReGaSi and the recently reported isotypic compound ReAlSi [70], [71]. These silicides attracted attention with respect to their low valence electron count of 14 and the occurrence of superconductivity in the solid solution ReAl1+xSi1−x (TC = 3.5 K for ReAl1.2Si0.8). Samples of that solid solution with a slightly higher aluminum content showed structural disorder [72]. A structure refinement of ReAl1.28Si0.72 revealed the MoSi2-type structure with a mixed occupancy (64% Al + 36% Si) for the 4e site (Figure 9). A 1:1 ordering of aluminum and silicon atoms in ReAlSi and of gallium and silicon atoms in ReGaSi requires a symmetry reduction and a splitting of the 4e subcell site into two two-fold sites. A look into the International Table A1 [73] for space group I4/mmm reveals two possibilities for klassengleiche transitions of index 2 which allow that kind of splitting. The first possibility is the transition to the primitive space group P4/mmm. Here, also the 2a site splits into two one-fold sites, and this superstructure model is realized for the series RE2RuMg3 and RE3Ru2Mg [74]. For ReAlSi and ReGaSi, the second possibility is realized. The space group symmetry is reduced to P4/nmm and the 4e site splits into two sites 2c (Figure 10). The 1:1 ordering (the ReAlSi structure is presented as an example in Figure 9) was unambiguously determined from neutron diffraction data (aluminum and silicon have too close scattering factors for a reliable refinement from X-ray diffraction data).

Figure 9: The crystal structures of ReAl1.28Si0.72 and ReAlSi [70], [71]. Rhenium, aluminum and silicon atoms are drawn as light gray, blue and magenta circles, respectively. The mixed occupied site in ReAl1.28Si0.72 is emphasized by segments. The heights of the differently compressed cubes are given in units of pm.
Figure 9:

The crystal structures of ReAl1.28Si0.72 and ReAlSi [70], [71]. Rhenium, aluminum and silicon atoms are drawn as light gray, blue and magenta circles, respectively. The mixed occupied site in ReAl1.28Si0.72 is emphasized by segments. The heights of the differently compressed cubes are given in units of pm.

Figure 10: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of ReAl1.28Si0.72 and ReAlSi [70], [71]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given. Note that the setting is for origin choice 1.
Figure 10:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of ReAl1.28Si0.72 and ReAlSi [70], [71]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given. Note that the setting is for origin choice 1.

The ordering in space group P4/nmm allows the stacking of three different cubes (Figure 9) with edge lengths of 262.7, 257.1 and 283.9 pm. The a lattice parameter of ReAlSi is 317.7 pm, thus all cubes are compressed. In the P4/mmm superstructure of RE2RuMg3 (RE = Dy, Ho, Er, Tm, Lu) and RE3Ru2Mg (RE = Sc, Er, Lu) [74] the mirror planes are kept perpendicular to the c axis leaving only two crystallographically independent cubes; a decisive difference to the P4/nmm superstructure.

ReAlSi and ReGaSi are simple ternary stacking sequences of three ordered and distorted cubes. They belong to a larger family of bcc superstructures which show a stacking in c direction with up to 13 cubes. Selected representatives of this family are summarized in Table 1. For an overview of more complex bcc superstructures (which are also enlarged in the a and b directions) we refer to the textbook by Müller [13] and two review articles [14, 82].

Table 1:

Further examples for bcc superstructures with subcell stacking in c direction.

CompoundNSGPCWSRef.
Sc2RuMg2P4/mmmtP4hba[75]
ReAlSi3P4/nmmtP6c3[70], [71]
MoSi23I4/mmmtI6ea[34]
Er2RuMg33P4/mmmtP6hgda[74]
Sc3Ru2Mg3P4/mmmtP6hgda[74]
Os2Al35I4/mmmtI10e2a[76]
Er2RuMg25I4/mmmtI10e2a[77]
Ti3Cu47I4/mmmtI14e3a[78]
Lu3Ru2Mg27I4/mmmtI14e3a[79]
Nb1.72Ta3.28S27I4/mmmtI14e3a[80]
Ta9(S, Se)413I4/mmmtI26e6a[81]
  1. The number of stacked cubes (N), the space group symbols (SG), the Pearson code (PC) and the Wyckoff sequence (WS) are listed. The complete group-subgroup schemes are presented in the respective references.

Finally we draw back to the crystallographic fingerprint (space group P4/nmm, Pearson code oP6 and Wyckoff sequence c3) of ReGaSi. In Figure 11 we present projections of the ReGaSi [70], Ta2Se [83], EuSnP [84], NbCrN [85], and CaGaN [86], [87] structures along the a axes. As discussed above, the ReGaSi structure shows a stacking of three slightly compressed cubes and is classified as an intermetallic phase, similar to the intermetallics discussed in Figure 1. The three cubes are emphasized by differently colored shadings. The Ta2Se structure shows one significantly stretched cube which contains the selenium atoms (van der Waals bonding between adjacent selenium layers). This is also the case for the isotypic members of the solid solution Ta2−xNbxS [88], [89] and Ta2Se0.66Te0.34 [81]. The remaining three compounds presented in Figure 11 show distinctly different chemical bonding and are only isopointal with ReGaSi. The differences are briefly discussed. The superstructure model in space group P4/nmm leaves free z parameters for all atoms, and in the cases of EuSnP (SrSnP [90]), NbCrN (NbMoN [91], TaMoN [92]), and CaGaN we observe substantial shifts of the phosphorus and nitrogen atoms off the sub-cube centers towards the sub-cube faces. These positions now correspond to a distorted octahedral coordination: P@Eu5Sn, N@Nb5Cr and N@Ca5Ga. The comparison of these five structures nicely underlines, that we need to divide the P4/nmm, tP6, c3 phases into three branches: (i) ReGaSi, (ii) Ta2Se and (iii) EuSnP, NbCrN and CaGaN, in analogy to the I4/mmm, tI6, ea phases discussed above. These branches account for the differences in chemical bonding and also underline the limits of structure comparison based on group-subgroup schemes.

Figure 11: Projection of the crystal structures of ReGaSi [70], [71], Ta2Se [83], EuSnP [84], NbCrN [85] and CaGaN [86], [87] which all have the crystallographic fingerprint of space group type P4/nmm, Pearson code oP6 and Wyckoff sequence c3. For details see text.
Figure 11:

Projection of the crystal structures of ReGaSi [70], [71], Ta2Se [83], EuSnP [84], NbCrN [85] and CaGaN [86], [87] which all have the crystallographic fingerprint of space group type P4/nmm, Pearson code oP6 and Wyckoff sequence c3. For details see text.

3 Ordering variants of BaAl4

The structures of more than four thousand intermetallic compounds are derived from the BaAl4 type [93]. This simple binary structure, space group I4/mmm, tI10 and Wyckoff sequence eda allows for a coloring on the e and d sites, leading to a manifold of ternary representatives. These ordering variants are listed in the Pearson data base [7] under the prototypes CeAl2Ga2 [94]/ThCr2Si2 [95] (3445 entries) and TlCu2Se2 [96] (627 entries). Herein we discuss some coloring variants from the so-called 1:2:2 pnictide family.

Depending on the size of the atoms and the valence electron count (VEC), the pnictogen atoms form homoatomic bonds or one observes pronounced two-dimensional substructures. These structural features have repeatedly been discussed [8, 24], [25], [26], [27], [28, 97], [98]. As an example we present the EuFe2As2 structure [99], [100] in Figure 12. The electron-precise description Eu2+(Fe2+)2(As3−)2 is compatible with the As–As distance of 333.4 pm, which is much longer than the As–As single bond of 252 pm in α-arsenic [101], thus underlining isolated arsenide anions.

Figure 12: The crystal structures of EuFe2As2 [100] and EuRbFe4As4 [109]. Europium, rubidium, iron and arsenic atoms are drawn as light gray, green, blue and magenta circles, respectively. The tetrahedral [FeAs] substructures are emphasized and relevant interatomic distances are given in units of pm.
Figure 12:

The crystal structures of EuFe2As2 [100] and EuRbFe4As4 [109]. Europium, rubidium, iron and arsenic atoms are drawn as light gray, green, blue and magenta circles, respectively. The tetrahedral [FeAs] substructures are emphasized and relevant interatomic distances are given in units of pm.

The field of BaAl4 related materials rapidly grew especially in the last 10 years in the context of pnictide based superconductors [102], [103], [104], [105], [106], [107], [108]. In the present subchapter we focus on some new superstructure variants which extend the previously described Bärnighausen trees [24], [25], [26], [27], [28]. The first ordering variant concerns the quaternary phases EuRbFe4As4 (Figure 12) [109], CaAFe4As4 (A = K, Rb, Cs) and SrAFe4As4 (A = Rb, Cs) [110], [111], [112] which contain a 1:1 distribution of mono- and divalent cations. The ordering is only possible through a loss of the body-centering. The corresponding group-subgroup relation is shown in Figure 13 for the pair EuFe2As2 and EuRbFe4As4. Due to the klassengleiche transition of index 2 (k2) from I4/mmm to P4/mmm (Figure 13), the X-ray powder patterns reveal superstructure reflections (so-called primitive reflections) which indicate the ordering. The 2a subcell site splits into two one-fold sites for europium and rubidium. The complete ordering is a consequence of the distinctly different ionic radii [113] of smaller Eu2+ (125 pm for coordination number 8) and larger Rb+ (161 pm for coordination number 8). The unit cell drastically expands from EuFe2As2 (c = 1212.47 pm) to EuRbFe4As4 (c = 1331.09 pm) and this significantly enlarges the As–As distances around the rubidium atoms (Figure 12). The lowering of the space group symmetry also affects the iron and arsenic atoms. The iron atoms lose their subcell mirror plane and are shifted to z = 0.2351 while the 4e arsenic site splits into two two-fold arsenic sites which independently coordinate to europium and rubidium.

Figure 13: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of EuFe2As2 [100] and EuRbFe4As4 [109]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given.
Figure 13:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of EuFe2As2 [100] and EuRbFe4As4 [109]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given.

The incorporation of a monovalent cation has an influence on the electronic structure of the quaternary arsenides. In order to keep an electron precise description, one iron cation per formula unit needs to be oxidized, leading to an overall oxidation state of +2.25 for iron. Magnetic susceptibility and resistivity measurements indicated superconductivity for these quaternary phases with transition temperatures in the range of 31–36 K [110]. Keeping the huge number of CeAl2Ga2/ThCr2Si2 intermetallics in mind, the 1:1 ordering on the cation site should be possible for many other element combinations, allowing for a number of new phases with modified magnetic properties.

At this point we should note that P4/mmm-type ordering was also considered for the uranium compounds UPd2Si2, UPd2Ge2, URh2Si2 and URh2Ge2 [114], [115]; however, the primitive reflections result rather from a CaBe2Ge2-type arrangement, and these tetrelides exhibit space group P4/nmm.

We now turn to a stacking of BaAl4 related cells along the c axis. First examples were the structures of BaMg2Sn2 (i2 of P4/nmm) [116] with a doubling the c axis, and the structure of BaCu2Sb2 (i3 of I4/mmm) [117] with a tripling of the c axis (Figures 14 and 15). These two superstructure variants were already discussed in the first Bärnighausen tree of BaAl4 superstructures in 1999 [24]. An X-ray-pure sample of BaCu2Sb2 was obtained by a reaction of BaSb3, Ba5Sb3 and Cu2Sb in a NaCl/KCl flux at 900 °C. BaCu2Sb2 has recently been reinvestigated with respect to its physical properties [118], [119]. A lead flux synthesis with a final annealing temperature of 550 °C led to platelets of α-BaCu2Sb2. Single crystal X-ray diffraction data proved the CaBe2Ge2-type structure, space group P4/nmm [123]. Reinvestigation of the high-temperature phase β-BaCu2Sb2 indicated lowering of the space group symmetry to Immm, C2/m or probably even to P21/c [124]. This situation might be similar to that observed for SrRh2As2 [120]; however, the quality of the diffraction data was not sufficient. The β-BaCu2Sb2 structure deserves further investigations in order to clarify possible phase transitions along with symmetry reductions in going to lower temperature.

Figure 14: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of BaFe2As2 [121]/α-BaCu2As2 [117], α-BaCu2Sb2 [118], β-BaCu2Sb2 [117]/Sr3Cd8Ge4 [122], and β-BaCu2As2 [119, 123]. For details see text.
Figure 14:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of BaFe2As2 [121]/α-BaCu2As2 [117], α-BaCu2Sb2 [118], β-BaCu2Sb2 [117]/Sr3Cd8Ge4 [122], and β-BaCu2As2 [119, 123]. For details see text.

Figure 15: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of BaFe2As2 [121], β-BaCu2Sb2 [117] and Sr3Cd8Ge4 [122]. The index for the isomorphic (i) symmetry reduction and the evolution of the atomic parameters are given. The shaded areas emphasize the change in coloring between copper and antimony in β-BaCu2Sb2, respectively cadmium and germanium in Sr3Cd8Ge4.
Figure 15:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of BaFe2As2 [121], β-BaCu2Sb2 [117] and Sr3Cd8Ge4 [122]. The index for the isomorphic (i) symmetry reduction and the evolution of the atomic parameters are given. The shaded areas emphasize the change in coloring between copper and antimony in β-BaCu2Sb2, respectively cadmium and germanium in Sr3Cd8Ge4.

We can thus conclude that α- and β-BaCu2Sb2 have the common supergroup I4/mmm (ThCr2Si2 type) and the phase transition is of a reconstructive type. There is no direct group-subgroup relation between the space group types of α- and β-BaCu2Sb2 (Figure 14). We start the crystal-chemical discussion of the congener BaCu2As2 [117] with the α modification which adopts the ThCr2Si2 type, space group I4/mmm, similar to the well know spin density wave material BaFe2As2 [121]. The copper atoms have tetrahedral arsenic coordination (Figure 16) and adjacent layers of condensed tetrahedra are related by mirror planes. Thus, these layers are crystallographically identical and we call them I layers in Figure 16. Due to the mirror symmetry, the arsenic atoms of adjacent layers point towards each other.

Figure 16: The crystal structures of α- and β-BaCu2As2 [117, 119, 123] α- and β-BaCu2Sb2 [117], [118] and Sr3Cd8Ge4 [122]. Strontium (barium), copper (cadmium) and arsenic (antimony, germanium) atoms are drawn as medium gray, blue and magenta circles, respectively. The tetrahedral substructures are emphasized. I and P denote the slabs that are derived from α-BaCu2As2 (ThCr2Si2 type, I4/mmm) and α-BaCu2Sb2 (CaBe2Ge2 type, P4/nmm). For details see text.
Figure 16:

The crystal structures of α- and β-BaCu2As2 [117, 119, 123] α- and β-BaCu2Sb2 [117], [118] and Sr3Cd8Ge4 [122]. Strontium (barium), copper (cadmium) and arsenic (antimony, germanium) atoms are drawn as medium gray, blue and magenta circles, respectively. The tetrahedral substructures are emphasized. I and P denote the slabs that are derived from α-BaCu2As2 (ThCr2Si2 type, I4/mmm) and α-BaCu2Sb2 (CaBe2Ge2 type, P4/nmm). For details see text.

In the structure of α-BaCu2Sb2 (CaBe2Ge2 type, space group P4/nmm), the coloring within each other [Cu2Sb2] layer is inverted to Sb@Cu4 tetrahedra and we now observe Cu–Sb (i.e. heteroatomic) bonding between adjacent layers. These layers are called P layers in Figure 16. The β-polymorph of BaCu2Sb2 shows a tripling of the unit cell (isomorphic transition of index 3). The decisive point is the inverted Cu/Sb coloring of the two Wyckoff sites shaded in Figure 15. This allows for the different stacking of P and I slabs (4 + 2 per unit cell), and one observes homoatomic (Sb–Sb) as well as heteratomic (Cu–Sb) interactions between the slabs. In other words, the β-BaCu2Sb2 structure is an intergrowth variant of ThCr2Si2 and CaBe2Ge2 related slabs.

The longest stacking period of a BaAl4/ThCr2Si2 superstructure has been observed for β-BaCu2As2 [119]. The original work on BaCu2As2 powders and single crystals (now called α-BaCu2As2) [117, 119] revealed the ThCr2Si2 type, space group I4/mmm. X-ray-pure samples of BaCu2As2 were obtained from an adequate mixture of Ba + CuCl + As. The resulting BaCl2 was extracted with diluted acetic acid and water.

Recent flux growth experiments [119, 123] with tin, lead, copper arsenide and gallium as flux agents at different temperatures resulted in a new superstructure variant. Only the Sn flux with a final annealing temperature of 500 or 550 °C led to crystals of β-BaCu2As2, which is an i5 superstructure of α-BaCu2As2. Similar to β-BaCu2Sb2 described above, we observe a switch in coloring (blue shading in Figure 17) with a different stacking of P and I slabs (4 + 6 per unit cell). The quintupled unit cell was unambiguously determined from a combination of single crystal X-ray diffraction data and high resolution electron microscopy.

Figure 17: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of α-BaCu2As2 and β-BaCu2As2 [117, 119, 123]. The index for the isomorphic (i) symmetry reduction and the evolution of the atomic parameters are given. The shaded areas emphasize the change in coloring between copper and arsenic.
Figure 17:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of α-BaCu2As2 and β-BaCu2As2 [117, 119, 123]. The index for the isomorphic (i) symmetry reduction and the evolution of the atomic parameters are given. The shaded areas emphasize the change in coloring between copper and arsenic.

An extension of the i3 superstructure variant was reported for the germanides Sr3Cd8Ge4 and Eu3Cd8Ge4 [122]. The Wyckoff sequence (Figure 15) is the same as for β-BaCu2Sb2; however, the coloring is different. Thus, Sr3Cd8Ge4 and β-BaCu2Sb2 are isopointal [11], [12], not isotypic. The occupancy of three Wyckoff positions with cadmium leads to a pronounced cadmium substructure; exemplarily shown for Sr3Cd8Ge4 in Figure 16. The two crystallographically independent germanium atoms are isolated (Ge2) and paired (251 pm Ge1–Ge1). For a detailed discussion of chemical bonding in these ternary germanides we refer to the original work.

Further compounds with the i3 superstructure variant are SrPt2Ge2 and BaPt2Ge2 [125], [126], [127], and the germanium-rich phases SrAu1.65Ge2.35 and SrAu1.60Ge2.40 [128]. So far only powder diffraction data is available for these three germanides. For SrPt2Ge2 and BaPt2Ge2 an inverse coloring has been reported and furthermore, monoclinic distortions [126], [127] might be possible. The structures of these three phases need to be studied on the basis of precise single crystal diffraction data in the future.

Further examples for BaAl4 superstructures are summarized in Table 2. The underlying group-subgroup relations are documented in the respective references.

Table 2:

Further examples for BaAl4 superstructures.

CompoundSGPCWSRef.
LaAg0.6Al3.4P4/nmmtP20jfc3b[129]
La2NiAl7I4mmtI40c2b3a6[130]
YbCu0.15Ga3.85C2/mmS10i2a[27]
BaGa1.9Hg2.1I41/amdtI40fe3[131]
BaNi2Ge2PnmaoP20d2c[28]
SrPdGa3CmcmoS20ec3[132]
γ-SrRh2As2P21/cmP20e5[120]
CePt2Al2CmmeoS20g3ba[133], [134]
LaPt2Al2CmmeoS20g3ba[135]
  1. The space group symbols (SG), the Pearson code (PC) and the Wyckoff sequence (WS) are listed. The complete group-subgroup schemes are presented in the respective references.

4 Ordering variants that are derived from the CeMg2Si2 type

The third group of compounds is derived from the CeMg2Si2-type structure [136], [137], [138], [139], [140], space group P4/mmm, Pearson code tP5 and Wyckoff sequence hea. The CeMg2Si2 type is realized for only few of the so-called 122 phases. Besides isotypic LaMg2Si2 [141], the pnictides BaPd2P2 [142] and BaPd2As2 [142], [143], [144] have been reported.

The relatively simple CeMg2Si2 structure is presented in Figure 18. The cerium and magnesium atoms form a three-dimensional network of corner-sharing Ce2Mg4 octahedra. The large cavities within this network (formed by eight triangular faces of the octahedra) are filled by Si2 dumbbells with 257 pm Si–Si. The chemical bonding in CeMg2Si2 has been studied on the basis of extended Hückel calculations along with a discussion on the valence electron concentration with respect to the competing structure types BaAl4/ThCr2Si2 [138].

Figure 18: The crystal structures of CeMg2Si2 [136], SrAu3Ge [146] and LaRu2Al2B [156]. Cerium (strontium, lanthanum), magnesium (gold, ruthenium) and silicon (germanium, aluminum) atoms are drawn as medium gray, blue and magenta circles, respectively. The octahedral building units Ce2Mg4, Sr2Au4 and B@La2Ru4 and the Si2, Au–Ge and Al2 dumbbells are emphasized. For details see text.
Figure 18:

The crystal structures of CeMg2Si2 [136], SrAu3Ge [146] and LaRu2Al2B [156]. Cerium (strontium, lanthanum), magnesium (gold, ruthenium) and silicon (germanium, aluminum) atoms are drawn as medium gray, blue and magenta circles, respectively. The octahedral building units Ce2Mg4, Sr2Au4 and B@La2Ru4 and the Si2, Au–Ge and Al2 dumbbells are emphasized. For details see text.

Superstructure formation was first reported for BaAu3Ge [145] and later also for SrAu3Ge [146], BaAu3+xSi1−x [147] and SrPt3P [148], [149], [150], and recently for CePt3P [151]. We exemplarily discuss SrAu3Ge [146]. The superstructure formation concerns coloring within the Si2 dumbbells sites of the aristotype CeMg2Si2. This deserves a symmetry reduction and a splitting of the 2h site. The corresponding group-subgroup scheme is shown in Figure 19. In the original version [146], the group-subgroup scheme has inconsistencies with respect to the space group symbols and the Wyckoff sites. The splitting is realized through a klassengleiche transition of index 2 (k2) from P4/mmm to P4/nmm (√2a × √2a × c cell). The Au–Ge dumbbells in SrAu3Ge (Figure 18) point alternatingly in +c and −c direction.

Figure 19: Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of CeMg2Si2 [136] and SrAu3Ge [146]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given.
Figure 19:

Group-subgroup scheme in the Bärnighausen formalism [15], [16], [17], [18] for the structures of CeMg2Si2 [136] and SrAu3Ge [146]. The index for the klassengleiche (k) symmetry reduction and the evolution of the atomic parameters are given.

The network of corner-sharing octahedra reminds of the Cu3Au and perovskite structures; however, the Ce2Mg4 octahedra in CeMg2Si2 are empty. It is interesting to note, that these octahedral voids can indeed be filled with small atoms, leading to quaternary phases. This was first observed for CeCr2Si2C [152] and then for other RECr2Si2C (RE = Y, La–Nd, Sm, Gd–Ho) silicide carbides [153], [154], [155]. Later, also borides, e.g. LaRu2Al2B [156] and CeRu2Ga2B [157], as well as the oxides BaTi2As2O [158] and BaTi2Sb2O [159] with this structure type have been discovered. The structure of LaRu2Al2B is presented in Figure 18. Such phases have repeatedly been discussed as anti-perovskite variants. The ‘LaRu2B’ substructure of LaRu2Al2B corresponds to the ‘TiO3’ network of perovskite. The cavities are filled by the calcium cations in perovskite, and by Al2 dumbbells aligned parallel to c in LaRu2Al2B, causing a tetragonal distortion.

5 Conclusion

The present overview discussed coloring and distortion variants of the bcc packing of the aristotypes BaAl4 and CeMg2Si2 on the basis of group-subgroup schemes. The many ordering variants include examples from intermetallics, ionic compounds and also some molecules. The Bärnighausen trees are a compact tool for structure systemization and nicely merge crystal chemistry and group theory. However, the many ordering variants discussed in the present overview also manifest the limits of comparability. The coloring with different atom types and/or the resulting distortions often drastically modify the bonding pattern. Thus, coloring variants are often only isopointal and not isotypic.


Corresponding author: Rainer Pöttgen, Institut für Anorganische und Analytische Chemie, Universität Münster, Corrensstrasse 30, 48149Münster, Germany, E-mail:

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: None declared.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

1. Ewald, P. P., Hermann, C., Eds. Strukturbericht 1913–1928; Akademische Verlagsgesellschaft M. B. H.: Leipzig, 1931.Search in Google Scholar

2. Wells, A. F. Structural Inorganic Chemistry, 5th ed.; Clarendon Press: Oxford, 1984.Search in Google Scholar

3. Kleber, W., Bautsch, H.-J., Bohm, J., Klimm, D. Einführung in die Kristallographie, 19. Aufl.; Oldenbourg Verlag: München, 2010.10.1515/9783486598858Search in Google Scholar

4. Ferro, R., Saccone, A. Intermetallic Chemistry; Elsevier: Amsterdam, 2008.Search in Google Scholar

5. Steurer, W., Dshemuchadse, J. Intermetallics: Structures, Properties, and Statistics, IUCr Monographs on Crystallography; Oxford University Press: New York, Vol. 26, 2016, ISBN-10: 0198714556.10.1093/acprof:oso/9780198714552.001.0001Search in Google Scholar

6. Mehl, M. J. J. Phys.: Conf. Ser. 2019, 1290, 012016; https://doi.org/10.1088/1742-6596/1290/1/012016.Search in Google Scholar

7. Villars, P., Cenzual, K., Eds. Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2020/21); ASM International®: Materials Park, Ohio (USA), 2020.Search in Google Scholar

8. Parthé, E., Gelato, L., Chabot, B., Penzo, M., Cenzual, K., Gladyshevskii, R. TYPIX-Standardized Data and Crystal Chemical Characterization of Inorganic Structure Types. Gmelin Handbook of Inorganic and Organometallic Chemistry, 8th ed.; Springer: Berlin (Germany), 1993.10.1007/978-3-662-10641-9Search in Google Scholar

9. Mehl, M. J., Hicks, D., Toher, C., Levy, O., Hanson, R. M., Hart, G., Curtarolo, S. Comput. Mater. Sci. 2017, 136, S1–S828; https://doi.org/10.1016/j.commatsci.2017.01.017.Search in Google Scholar

10. Hicks, D., Mehl, M. J., Gossett, E., Toher, C., Levy, O., Hanson, R. M., Hart, G., Curtarolo, S. Comput. Mater. Sci. 2019, 161, S1–S1011; https://doi.org/10.1016/j.commatsci.2018.10.043.Search in Google Scholar

11. Parthé, E., Gelato, L. M. Acta Crystallogr. 1984, A40, 169–183; https://doi.org/10.1107/s0108767384000416.Search in Google Scholar

12. Gelato, L. M., Parthé, E. J. Appl. Crystallogr. 1987, 20, 139–143; https://doi.org/10.1107/s0021889887086965.Search in Google Scholar

13. Müller, U. Inorganic Structural Chemistry, 2nd ed.; Wiley: Chichester, 2007.Search in Google Scholar

14. Pöttgen, R. Z. Anorg. Allg. Chem. 2014, 640, 869–891; https://doi.org/10.1002/zaac.201400023.Search in Google Scholar

15. Bärnighausen, H. Commun. Math. Chem. 1980, 9, 139–175.10.1007/BF01674443Search in Google Scholar

16. Müller, U. Z. Anorg. Allg. Chem. 2004, 630, 1519–1537; https://doi.org/10.1002/zaac.200400250.Search in Google Scholar

17. Müller, U. Relating crystal structures by group-subgroup relations. In International Tables for Crystallography; Wondratschek, H., Müller, U., Eds.; Symmetry Relations Between Space Groups, 2nd ed.; John Wiley & Sons, Ltd: Chichester, Vol. A1, 2010; pp. 44–56.10.1107/97809553602060000795Search in Google Scholar

18. Müller, U. Symmetriebeziehungen zwischen verwandten Kristallstrukturen; Vieweg + Teubner Verlag: Wiesbaden, 2012.10.1007/978-3-8348-8342-1Search in Google Scholar

19. Bärnighausen, H. Acta Crystallogr. 1975, A31, part S3, 01. 1–9.Search in Google Scholar

20. Bock, O., Müller, U. Acta Crystallogr. 2002, B58, 594–606; https://doi.org/10.1107/s0108768102001490.Search in Google Scholar

21. Baur, W. H. Z. Kristallogr. 1994, 209, 143–150; https://doi.org/10.1524/zkri.1994.209.2.143.Search in Google Scholar

22. Baur, W. H. Crystallogr. Rev. 2007, 13, 65–113; https://doi.org/10.1080/08893110701433435.Search in Google Scholar

23. Hoffmann, R.-D., Pöttgen, R. Z. Kristallogr. 2001, 216, 127–145; https://doi.org/10.1007/978-3-662-10258-9_5.Search in Google Scholar

24. Kußmann, D., Pöttgen, R., Rodewald, U. Ch., Rosenhahn, C., Mosel, B. D., Kotzyba, G., Künnen, B. Z. Naturforsch. 1999, 54b, 1155–1164; https://doi.org/10.1515/znb-1999-0911.Search in Google Scholar

25. Johrendt, D., Hosono, H., Hoffmann, R.-D., Pöttgen, R. Z. Kristallogr. 2011, 226, 435–446; https://doi.org/10.1524/zkri.2011.1363.Search in Google Scholar

26. Kneidinger, F., Salamakha, L., Bauer, E., Zeiringer, I., Rogl, P., Blaas-Schenner, C., Reith, D., Podloucky, R. Phys. Rev. B 2014, 90, 024504; https://doi.org/10.1103/physrevb.90.024504.Search in Google Scholar

27. Verbovytskyy, Yu. Visn. Lviv Univ. Ser. Chem. 2015, 56, 80–86.Search in Google Scholar

28. Hlukhyy, V., Trots, D., Fässler, T. F. Inorg. Chem. 2017, 56, 1173–1185; https://doi.org/10.1021/acs.inorgchem.6b02190.Search in Google Scholar

29. Pöttgen, R., Johrendt, D. Intermetallics, 2nd ed.; De Gruyter: Berlin, 2019.10.1515/9783110636727Search in Google Scholar

30. Miller, G. J. Eur. J. Inorg. Chem. 1998, 523–536; https://doi.org/10.1002/(sici)1099-0682(199805)1998:5<523::aid-ejic523>3.0.co;2-l.10.1002/(SICI)1099-0682(199805)1998:5<523::AID-EJIC523>3.0.CO;2-LSearch in Google Scholar

31. Miller, G. J. Z. Anorg. Allg. Chem. 2006, 632, 2078; https://doi.org/10.1002/zaac.200670006.Search in Google Scholar

32. Han, M. K., Miller, G. J. Inorg. Chem. 2008, 47, 515–528; https://doi.org/10.1021/ic701311b.Search in Google Scholar

33. Parthé, E. Elements of Inorganic Structural Chemistry; Pöge: Leipzig, 1990.Search in Google Scholar

34. Zachariasen, W. H. Z. Phys. Chem. Stoechiom. Verwandtschaftsl. 1927, 128, 39–48.10.1515/zpch-1927-12804Search in Google Scholar

35. Kallel, A. C. R. Séances Acad. Sci., Ser. B 1969, 268, 455–458.Search in Google Scholar

36. Nevitt, M. V., Downey, J. W. Trans. Metall. Soc. AIME 1962, 224, 195–196.Search in Google Scholar

37. Wannek, C., Harbrecht, B. Z. Anorg. Allg. Chem. 2002, 628, 1597–1601; https://doi.org/10.1002/1521-3749(200207)628:7<1597::aid-zaac1597>3.0.co;2-a.10.1002/1521-3749(200207)628:7<1597::AID-ZAAC1597>3.0.CO;2-ASearch in Google Scholar

38. Schablaske, R. V., Tani, B. S., Chasanov, M. G. Trans. Metall. Soc. AIME 1962, 224, 867–868.Search in Google Scholar

39. Warkentin, E., Bärnighausen, H. Abstr. 3rd Eur. Crystallogr. Meet., Zürich, Switzerland, 1976, 354–356.Search in Google Scholar

40. Simon, A., Mattausch, H. In Comprehensive Inorganic Chemistry II: From Elements to Applications; Reedijk, J., Poeppelmeier, K., Eds. Elsevier: Amsterdam, Vol. 2, 2013; pp. 535–581.10.1016/B978-0-08-097774-4.00224-2Search in Google Scholar

41. Stackelberg, M. V. Z. Phys. Chem. B 1930, 9, 437–475.10.1515/zpch-1930-0929Search in Google Scholar

42. Elliott, H. St. A., Lehmann, J. F., Mercier, H. P. A., Jenkins, H. D. B., Schrobilgen, G. J. Inorg. Chem. 2010, 49, 8504–8523; https://doi.org/10.1021/ic101152x.Search in Google Scholar

43. Siegel, S., Gebert, E. J. Am. Chem. Soc. 1963, 85, 240; https://doi.org/10.1021/ja00885a035.Search in Google Scholar

44. Lehmann, J. F., Dixon, D. A., Schrobilgen, G. J. Inorg. Chem. 2001, 40, 3002–3017; https://doi.org/10.1021/ic001167w.Search in Google Scholar

45. Hoppe, R., Röhrborn, H. J. Naturwissenschaften 1962, 49, 419–420; https://doi.org/10.1007/bf00624409.Search in Google Scholar

46. Shi, Y., Leithe-Jasper, A., Tanaka, T. J. Solid State Chem. 1999, 148, 250–259; https://doi.org/10.1006/jssc.1999.8446.Search in Google Scholar

47. Noréus, D., Törnroos, K. W., Börje, A., Szabò, T., Bronger, W., Spittank, H., Auffermann, G., Müller, P. J. Less Common Met. 1988, 139, 233–239; https://doi.org/10.1016/0022-5088(88)90004-5.Search in Google Scholar

48. Becker, M., Nuss, J., Jansen, M. Z. Anorg. Allg. Chem. 2000, 626, 2505–2508; https://doi.org/10.1002/1521-3749(200012)626:12<2505::aid-zaac2505>3.0.co;2-#.10.1002/1521-3749(200012)626:12<2505::AID-ZAAC2505>3.0.CO;2-#Search in Google Scholar

49. Zhang, S., Zherebtsov, D. A., DiSalvo, F., Niewa, R. Z. Anorg. Allg. Chem. 2012, 638, 2111–2116; https://doi.org/10.1002/zaac.201100413.Search in Google Scholar

50. Moss, M. A., Jeitschko, W. Z. Anorg. Allg. Chem. 1991, 603, 57–67; https://doi.org/10.1002/zaac.19916030109.Search in Google Scholar

51. Gerss, M. H., Jeitschko, W. Z. Naturforsch. 1986, 41b, 946–950; https://doi.org/10.1515/znb-1986-0804.Search in Google Scholar

52. Cordier, G., Gudat, A., Kniep, R., Rabenau, A. Angew. Chem., Int. Ed. Engl. 1989, 28, 201–202; https://doi.org/10.1002/anie.198902011.Search in Google Scholar

53. Chern, M. Y., DiSalvo, F. J. J. Solid State Chem. 1990, 88, 459–464; https://doi.org/10.1016/0022-4596(90)90242-p.Search in Google Scholar

54. Hoffmann, R., Li, J., Wheeler, R. A. J. Am. Chem. Soc. 1987, 109, 6600–6602; https://doi.org/10.1021/ja00256a005.Search in Google Scholar

55. Green, M. T., Hughbanks, T. Inorg. Chem. 1993, 32, 5611–5615; https://doi.org/10.1021/ic00076a031.Search in Google Scholar

56. Singh, D. J. Phys. Rev. B 2002, 66, 132414; https://doi.org/10.1103/physrevb.66.014415.Search in Google Scholar

57. Höhn, P., Kniep, R. Z. Naturforsch. 1992, 47b, 434–436; https://doi.org/10.1515/znb-1992-0322.Search in Google Scholar

58. Ballé, T. J., Zangeneh, Z., Hozoi, L., Jesche, A., Höhn, P. Phys. Rev. B 2019, 99, 094422; https://doi.org/10.1103/physrevb.99.094422.Search in Google Scholar

59. Jäger, J., Kniep, R. Z. Naturforsch. 1992, 47b, 1290–1296; https://doi.org/10.1515/znb-1992-0913.Search in Google Scholar

60. Klatyk, J., Kniep, R. Z. Kristallogr. NCS 1999, 214, 451–452; https://doi.org/10.1515/ncrs-1999-0425.Search in Google Scholar

61. Yamane, H., Kikkawa, S., Koizumi, M. J. Solid State Chem. 1987, 71, 1–11; https://doi.org/10.1016/0022-4596(87)90135-6.Search in Google Scholar

62. Cenzual, K., Gelato, L. M., Penzo, M., Parthé, E. Acta Crystallogr. 1991, B47, 433–439; https://doi.org/10.1107/s0108768191000903.Search in Google Scholar

63. Đuriš, K., Magdysyuk, O. V., Jansen, M. Solid State Sci. 2012, 14, 1399–1404.10.1016/j.solidstatesciences.2012.06.006Search in Google Scholar

64. Đuriš, K., Müller, U., Jansen, M. Z. Anorg. Allg. Chem. 2012, 638, 737–743.10.1002/zaac.201100511Search in Google Scholar

65. Wu, G., Huang, X., Huang, Y., Pan, L., Li, F., Li, X., Liu, M., Liu, B., Cui, T. J. Phys. Chem. C 2017, 121, 6264–6271; https://doi.org/10.1021/acs.jpcc.6b11558.Search in Google Scholar

66. Siegrist, T., Hulliger, F., Travaglini, G. J. Less Common Met. 1983, 92, 119–129; https://doi.org/10.1016/0022-5088(83)90233-3.Search in Google Scholar

67. Krautwasser, P., Bhan, S., Schubert, K. Z. Metallkd. 1968, 59, 724–729.10.1515/ijmr-1968-590907Search in Google Scholar

68. Smith, P. L., Ortega, R., Brennan, B. Mater. Res. Soc. Symp. Proc. 1998, 481, 605–610.10.1557/PROC-481-605Search in Google Scholar

69. Stolz, E., Schubert, K. Z. Metallkd. 1962, 53, 433–444.10.1086/349630Search in Google Scholar

70. Decocq, V., Gui, X., Neeson, A., Xie, W., Heitmann, T., Wang, F. Inorg. Chem. 2020, 59, 17310–17319; https://doi.org/10.1021/acs.inorgchem.0c02578.Search in Google Scholar

71. Xie, W., Gustin, L., Bian, G. Inorg. Chem. 2017, 56, 5165–5172; https://doi.org/10.1021/acs.inorgchem.7b00309.Search in Google Scholar

72. Kuz’ma, Y. B., Milyan, V. V. Inorg. Mater. 1977, 13, 760–761.Search in Google Scholar

73. Wondratschek, H., Müller, U., Eds. International Tables for Crystallography; Symmetry Relations Between Space Groups, 2nd ed.; John Wiley & Sons, Ltd: Chichester, Vol. A1, 2010.10.1107/97809553602060000110Search in Google Scholar

74. Kersting, M., Niehaus, O., Hoffmann, R.-D., Rodewald, U. Ch., Pöttgen, R. Z. Kristallogr. 2014, 229, 285–294; https://doi.org/10.1515/zkri-2013-1717.Search in Google Scholar

75. Kersting, M., Rodewald, U. Ch., Pöttgen, R. Z. Kristallogr. 2015, 230, 151–155; https://doi.org/10.1515/zkri-2014-1831.Search in Google Scholar

76. Edshammar, L.-E. Acta Chem. Scand. 1965, 19, 871–874; https://doi.org/10.3891/acta.chem.scand.19-0871.Search in Google Scholar

77. Kersting, M., Niehaus, O., Hoffmann, R.-D., Pöttgen, R. Z. Kristallogr. 2013, 228, 643–649; https://doi.org/10.1524/zkri.2013.1692.Search in Google Scholar

78. Schubert, K. Z. Metallkd. 1965, 56, 197–199.10.1515/ijmr-1965-560308Search in Google Scholar

79. Stein, S., Kersting, M., Heletta, L., Pöttgen, R. Z. Naturforsch. 2017, 72b, 447–455; https://doi.org/10.1515/znb-2017-0048.Search in Google Scholar

80. Yao, X., Franzen, H. F. J. Am. Chem. Soc. 1991, 113, 1426–1427; https://doi.org/10.1021/ja00004a061.Search in Google Scholar

81. Harbrecht, B., Degen, T., Conrad, M. J. Alloys Compd. 1997, 246, 37–50; https://doi.org/10.1016/s0925-8388(96)02475-9.Search in Google Scholar

82. Sluiter, M. H. F. Phase Trans. 2007, 80, 299–309; https://doi.org/10.1080/01411590701228562.Search in Google Scholar

83. Harbrecht, B. Angew. Chem., Int. Ed. Engl. 1989, 28, 1660–1662; https://doi.org/10.1002/anie.198916601.Search in Google Scholar

84. Payne, A. C., Sprauve, A. E., Holm, A. P., Olmstead, M. M., Kauzlarich, S. M., Klavins, P. J. Alloys Compd. 2002, 338, 229–234; https://doi.org/10.1016/s0925-8388(02)00235-9.Search in Google Scholar

85. Jack, D. H., Jack, K. H. J. Iron Steel Inst., London 1972, 210, 790–792.Search in Google Scholar

86. Verdier, P., Marchand, R., Lang, J. C. R. Acad. Sci. Paris, Sér. C 1973, 276, 607–609.Search in Google Scholar

87. Verdier, P., L’Haridon, P., Maunaye, M., Marchand, R. Acta Crystallogr. 1974, B30, 226–228; https://doi.org/10.1107/s0567740874002500.Search in Google Scholar

88. Yao, X., Miller, G. J., Franzen, H. F. J. Alloys Compd. 1992, 183, 7–17; https://doi.org/10.1016/0925-8388(92)90725-o.Search in Google Scholar

89. Nanjundaswamy, K. S., Hughbanks, T. J. Solid State Chem. 1992, 98, 278–290; https://doi.org/10.1016/s0022-4596(05)80236-1.Search in Google Scholar

90. Eisenmann, B., Jordan, H., Schäfer, H. J. Less Common Met. 1986, 116, 251–258; https://doi.org/10.1016/0022-5088(86)90233-x.Search in Google Scholar

91. Vendl, A. Monatsh. Chem. 1979, 110, 103–108; https://doi.org/10.1007/bf00903751.Search in Google Scholar

92. Vendl, A. Monatsh. Chem. 1978, 109, 1001–1004; https://doi.org/10.1007/bf00907321.Search in Google Scholar

93. Andress, K. R., Alberti, E. Z. Metallkd. 1935, 27, 126–128.10.1163/187124035X00090Search in Google Scholar

94. Zarechnyuk, O. S., Krypyakevych, P. I., Gladyshevskii, E. I. Sov. Phys. Crystallogr. 1965, 9, 706–708.Search in Google Scholar

95. Ban, Z., Sikirica, M. Acta Crystallogr. 1965, 18, 594–599; https://doi.org/10.1107/s0365110x6500141x.Search in Google Scholar

96. Avilov, A. S., Imamov, R. M., Pinsker, Z. G. Sov. Phys. Crystallogr. 1971, 16, 542–544.Search in Google Scholar

97. Mewis, A. Z. Naturforsch. 1980, 35b, 141–145; https://doi.org/10.1515/znb-1980-0205.Search in Google Scholar

98. Hoffmann, R., Zheng, C. J. Phys. Chem. 1985, 89, 4175–4181; https://doi.org/10.1021/j100266a007.Search in Google Scholar

99. Marchand, R., Jeitschko, W. J. Solid State Chem. 1978, 24, 351–357; https://doi.org/10.1016/0022-4596(78)90026-9.Search in Google Scholar

100. Tegel, M., Rotter, M., Weiss, V., Schappacher, F. M., Pöttgen, R., Johrendt, D. J. Phys. Condens. Matter 2008, 20, 452201; https://doi.org/10.1088/0953-8984/20/45/452201.Search in Google Scholar

101. Donohue, J. The Structures of the Elements; Wiley: New York, 1974.Search in Google Scholar

102. Just, G., Paufler, P. J. Alloys Compd. 1996, 232, 1–25; https://doi.org/10.1016/0925-8388(95)01939-1.Search in Google Scholar

103. Shatruk, M. J. Solid State Chem. 2019, 272, 198–209; https://doi.org/10.1016/j.jssc.2019.02.012.Search in Google Scholar

104. Sadovskii, M. V. Phys. Usp. 2008, 51, 1201–1228; https://doi.org/10.1070/pu2008v051n12abeh006820.Search in Google Scholar

105. Ivanovski, A. L. Phys. Usp. 2008, 51, 1229–1260.10.1070/PU2008v051n12ABEH006703Search in Google Scholar

106. Izyumov, Yu. A., Kurmaev, E. Z. Phys. Usp. 2008, 51, 1261–1286; https://doi.org/10.1070/pu2008v051n12abeh006733.Search in Google Scholar

107. Johnston, D. C. Adv. Phys. 2010, 59, 803–1061; https://doi.org/10.1080/00018732.2010.513480.Search in Google Scholar

108. Johrendt, D. J. Mater. Chem. 2011, 21, 13726–13736; https://doi.org/10.1039/c1jm12318a.Search in Google Scholar

109. Liu, Y., Liu, Y.-B., Tang, Z.-T., Jiang, H., Wang, Z.-C., Ablimit, A., Jiao, W.-H., Tao, Q., Feng, C.-M., Xu, Z.-A., Cao, G.-H. Phys. Rev. B 2016, 93, 214503; https://doi.org/10.1103/physrevb.93.214503.Search in Google Scholar

110. Iyo, A., Kawashima, K., Kinjo, T., Nishio, T., Ishida, S., Fujihisa, H., Gotoh, Y., Kihou, K., Eisaki, H., Yoshida, Y. J. Am. Chem. Soc. 2016, 138, 3410–3415; https://doi.org/10.1021/jacs.5b12571.Search in Google Scholar

111. Mou, D., Kong, T., Meier, W. R., Lochner, F., Wang, L.-L., Lin, Q., Wu, Y., Bud’ko, S. L., Eremin, I., Johnson, D. D., Canfield, P. C., Kaminski, A. Phys. Rev. Lett. 2016, 117, 277001; https://doi.org/10.1103/physrevlett.117.277001.Search in Google Scholar

112. Meier, W. R., Kong, T., Kaluarachchi, U. S., Taufour, V., Jo, N. H., Drachuck, G., Böhmer, A. E., Saunders, S. M., Sapkota, A., Kreyssig, A., Tanatar, M. A., Prozorov, R., Goldman, A. I., Balakirev, F. F., Gurevich, A., Bud’ko, S. L., Canfield, P. C. Phys. Rev. B 2016, 94, 064501; https://doi.org/10.1103/physrevb.94.064501.Search in Google Scholar

113. Shannon, R. D. Acta Crystallogr. 1976, A32, 751–767; https://doi.org/10.1107/s0567739476001551.Search in Google Scholar

114. Ptasiewicz-Bąk, H., Leciejewicz, J., Zygmunt, A. J. Phys. F Met. Phys. 1981, 11, 1225–1235; https://doi.org/10.1088/0305-4608/11/6/009.Search in Google Scholar

115. Ptasiewicz-Bąk, H., Leciejewicz, J., Zygmunt, A. Solid State Commun. 1985, 55, 601–604.10.1016/0038-1098(85)90820-8Search in Google Scholar

116. Eisenmann, B., Schäfer, H. Z. Anorg. Allg. Chem. 1974, 403, 163–172; https://doi.org/10.1002/zaac.19744030210.Search in Google Scholar

117. Dünner, J., Mewis, A., Roepke, M., Michels, G. Z. Anorg. Allg. Chem. 1995, 621, 1523–1530.10.1002/zaac.19956210915Search in Google Scholar

118. Saparov, B., Sefat, A. S. J. Solid State Chem. 2012, 191, 213–219; https://doi.org/10.1016/j.jssc.2012.03.036.Search in Google Scholar

119. Wu, S. F., Richard, P., van Roekeghem, A., Nie, S. M., Miao, H., Xu, N., Qian, T., Saparov, B., Fang, Z., Biermann, S., Sefat, A. S., Ding, H. Phys. Rev. B 2015, 91, 235109; https://doi.org/10.1103/physrevb.91.235109.Search in Google Scholar

120. Zinth, V., Petricek, V., Dusek, M., Johrendt, D. Phys. Rev. B 2012, 85, 014109; https://doi.org/10.1103/physrevb.85.014109.Search in Google Scholar

121. Rotter, M., Tegel, M., Johrendt, D., Schellenberg, I., Hermes, W., Pöttgen, R. Phys. Rev. B 2008, 78, 020503; https://doi.org/10.1103/physrevb.78.020503.Search in Google Scholar

122. Suen, N.-T., Huang, L., Meyers, J. J., Bobev, S. Inorg. Chem. 2018, 57, 833–842; https://doi.org/10.1021/acs.inorgchem.7b02781.Search in Google Scholar

123. Wu, H., Li, S., Wu, Z., Wang, X., Ofenstein, G. A., Kwon, S., Kim, M. J., Chu, P. C. W., Lv, B. Cryst. Growth Des. 2020, 20, 5922–5930; https://doi.org/10.1021/acs.cgd.0c00614.Search in Google Scholar

124. Anand, V. K., Perera, P. K., Pandey, A., Goetsch, R. J., Kreyssig, A., Johnston, D. C. Phys. Rev. B 2012, 85, 214523; https://doi.org/10.1103/physrevb.85.214523.Search in Google Scholar

125. Kessens, R., Jung, W. Z. Anorg. Allg. Chem. 2002, 628, 2175; https://doi.org/10.1002/1521-3749(200209)628:9/10<2175::aid-zaac11112175>3.0.co;2-j.10.1002/1521-3749(200209)628:9/10<2175::AID-ZAAC11112175>3.0.CO;2-JSearch in Google Scholar

126. Ku, H. C., Chen, I. A., Huang, C. H., Chen, C. W., You, Y. B., Tai, M. F., Hsu, Y. Y. Physica C 2013, 493, 93–95; https://doi.org/10.1016/j.physc.2013.03.036.Search in Google Scholar

127. Chen, I. A., Huang, C. H., Chen, C. W., You, Y. B., Tai, M. F., Ku, H. C., Hsu, Y. Y. J. Appl. Phys. 2013, 113, 213902; https://doi.org/10.1063/1.4808287.Search in Google Scholar

128. Zeiringer, I., Grytsiv, A., Bauer, E., Giester, G., Rogl, P. Z. Anorg. Allg. Chem. 2015, 641, 1404–1421; https://doi.org/10.1002/zaac.201500045.Search in Google Scholar

129. Cordier, G., Dietrich, C., Doersam, G. Z. Kristallogr. NCS 1997, 212, 288; https://doi.org/10.1524/ncrs.1997.212.1.288.Search in Google Scholar

130. Gout, D., Barker, T. J., Gourdon, O., Miller, G. J. Chem. Mater. 2005, 17, 3661–3667; https://doi.org/10.1021/cm050513a.Search in Google Scholar

131. Wendorff, M., Röhr, C. Z. Naturforsch. 2013, 68b, 307–322; https://doi.org/10.5560/znb.2013-3084.Search in Google Scholar

132. Seidel, S., Hoffmann, R.-D., Pöttgen, R. Z. Kristallogr. 2014, 229, 421–426; https://doi.org/10.1515/zkri-2014-1742.Search in Google Scholar

133. Doležal, P., Duverger-Nédellec, E., Mičková, Z., Proschek, P., Illková, K., Daniš, S., Bartha, K., Javorský, P. Inorg. Chem. 2020, 59, 12263–12275.10.1021/acs.inorgchem.0c01292Search in Google Scholar

134. Murashova, E., Morozova, Y., Dunaev, S., Kurenbaeva, Z., Tursina, A. Crystals 2020, 10, 465 (15 pages); https://doi.org/10.3390/cryst10060465.Search in Google Scholar

135. Doležal, P., Diviš, M., Bartha, K., Javorský, P. J. Alloys Compd. 2020, 848, 5, 156360.10.1016/j.jallcom.2020.156360Search in Google Scholar

136. Zmii, O. F., Gladyshevskii, E. I. Sov. Phys. Crystallogr. 1971, 15, 817–819.Search in Google Scholar

137. Zmii, O. F., Gladyshevskii, E. I. Visn. Lviv Derzh. Univ., Ser. Khim. 1974, 15, 24–26.Search in Google Scholar

138. Zheng, C., Hoffmann, R. Z. Naturforsch. 1986, 41b, 292–320; https://doi.org/10.1515/znb-1986-0304.Search in Google Scholar

139. Dhar, S. K., Manfrinetti, P., Palenzona, A. J. Alloys Compd. 1997, 252, 24–27; https://doi.org/10.1016/s0925-8388(96)02706-5.Search in Google Scholar

140. Han, W.-D., Li, K., Hu, F., Li, X.-D., Dai, J. Mater. Res. Express 2019, 6, 12, 1165f8; https://doi.org/10.1088/2053-1591/ab3687.Search in Google Scholar

141. De Negri, S., Saccone, A., Delfino, S. Calphad 2009, 33, 44–49; https://doi.org/10.1016/j.calphad.2008.07.012.Search in Google Scholar

142. Mewis, A. Z. Naturforsch. 1984, 39b, 713–720; https://doi.org/10.1515/znb-1984-0603.Search in Google Scholar

143. Anand, V. K., Kim, H., Tanatar, M. A., Prozorov, R., Johnston, D. C. Phys. Rev. B 2013, 87, 224510; https://doi.org/10.1103/physrevb.87.224510.Search in Google Scholar

144. Kudo, K., Yamada, Y., Takeuchi, T., Kimura, T., Ioka, S., Matsuo, G., Kitahama, Y., Nohara, M. J. Phys. Soc. Jpn. 2017, 86, 4, 063704; https://doi.org/10.7566/jpsj.86.063704.Search in Google Scholar

145. Zeiringer, I., Melnychenko-Koblyuk, N., Grytsiv, A., Bauer, E., Giester, G., Rogl, P. J. Phase Equilibria Diffus. 2011, 32, 115–127; https://doi.org/10.1007/s11669-011-9852-7.Search in Google Scholar

146. Lin, Q., Corbett, J. D. J. Am. Chem. Soc. 2012, 134, 4877–4884; https://doi.org/10.1021/ja211625w.Search in Google Scholar

147. Zeiringer, I., Chen, M. X., Grytsiv, A., Bauer, E., Podloucky, R., Effenberger, H., Rogl, P. Acta Mater. 2012, 60, 2324–2336; https://doi.org/10.1016/j.actamat.2011.12.040.Search in Google Scholar

148. Takayama, T., Kuwano, K., Hirai, D., Katsura, Y., Yamamoto, A., Takagi, H. Phys. Rev. Lett. 2012, 108, 237001; https://doi.org/10.1103/physrevlett.108.237001.Search in Google Scholar

149. Hu, K., Gao, B., Ji, Q., Ma, Y., Li, W., Xu, X., Zhang, H., Mu, G., Huang, F., Cai, C., Xie, X., Jiang, M. Phys. Rev. B 2016, 93, 214510; https://doi.org/10.1103/physrevb.93.214510.Search in Google Scholar

150. Zhigadlo, N. D. J. Cryst. Growth 2016, 455, 94–98; https://doi.org/10.1016/j.jcrysgro.2016.10.003.Search in Google Scholar

151. Chen, J., Wang, Z., Zheng, S., Feng, C., Dai, J., Xu, Z. Sci. Rep. 2017, 7, 41853; https://doi.org/10.1038/srep41853.Search in Google Scholar

152. Tang, C., Fan, S., Zhu, M. J. Alloys Compd. 2000, 299, 1–4; https://doi.org/10.1016/s0925-8388(99)00631-3.Search in Google Scholar

153. Pohlkamp, M. W., Jeitschko, W. Z. Naturforsch. 2001, 56b, 1143–1148; https://doi.org/10.1515/znb-2001-1108.Search in Google Scholar

154. Klosek, V., Vernière, A., Malaman, B., Tobola, J., Kaprzyk, S. Phys. Rev. B 2008, 78, 104419; https://doi.org/10.1103/physrevb.78.104419.Search in Google Scholar

155. Chen, G., Du, Q., Yang, W., Lai, Y., Yang, J., Wang, C., Han, J., Liu, S., Zhang, Y., Du, H., Zhang, C., Xia, Y., Zhou, D., Yang, Y. J. Phys. D Appl. Phys. 2016, 49, 025001; https://doi.org/10.1088/0022-3727/49/2/025001.Search in Google Scholar

156. Zaikina, J. V., Jo, Y.-J., Latturner, S. E. Inorg. Chem. 2010, 49, 2773–2781; https://doi.org/10.1021/ic902151d.Search in Google Scholar

157. Baumbach, R. E., Lu, X., Ronning, F., Thompson, J. D., Bauer, E. D. J. Phys. : Condens. Matter 2012, 24, 325601; https://doi.org/10.1088/0953-8984/24/32/325601.Search in Google Scholar

158. Wang, X. F., Yan, Y. J., Ying, J. J., Li, Q. J., Zhang, M., Xu, N., Chen, X. H. J. Phys. : Condens. Matter 2010, 22, 075702; https://doi.org/10.1088/0953-8984/22/7/075702.Search in Google Scholar

159. Yajima, T., Nakano, K., Takeiri, F., Ono, T., Hosokoshi, Y., Matsushita, Y., Hester, J. R., Kobayashi, Y., Kageyama, H. J. Phys. Soc. Jpn. 2012, 81, 103706; https://doi.org/10.1143/jpsj.81.103706.Search in Google Scholar

Received: 2021-03-04
Accepted: 2021-03-13
Published Online: 2021-03-30
Published in Print: 2021-05-26

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2021-0022/html
Scroll to top button