Next Article in Journal
Design, Synthesis, and Investigation of Cytotoxic Activity of cis-Vinylamide-Linked Combretastatin Analogues as Potential Anticancer Agents
Next Article in Special Issue
Novel Red-Emitting BaBi2B4O10:Eu3+ Phosphors: Synthesis, Crystal Structure and Luminescence
Previous Article in Journal
On Three-Rainbow Dominationof Generalized Petersen Graphs P(ck,k)
Previous Article in Special Issue
Complexity Parameters for Molecular Solids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Symmetry Analysis of the Complex Polytypism of Layered Rare-Earth Tellurites and Related Selenites: The Case of Introducing Transition Metals

by
Dmitri O. Charkin
1,2,
Valeri A. Dolgikh
1,
Timofey A. Omelchenko
1,
Yulia A. Vaitieva
2,
Sergey N. Volkov
2,3,
Dina V. Deyneko
1,2 and
Sergey M. Aksenov
2,4,*
1
Department of Chemistry, Lomonosov Moscow State University, GSP-1, 1-3 Leninskiye Gory, 119991 Moscow, Russia
2
Laboratory of Arctic Mineralogy and Material Sciences, Kola Science Centre, Russian Academy of Sciences, Fersmana Str. 14, 184209 Apatity, Russia
3
Grebenshchikov Institute of Silicate Chemistry, Makarov Emb, 199053 St. Petersburg, Russia
4
Geological Institute, Kola Science Centre, Russian Academy of Sciences, 14 Fersman Street, 184209 Apatity, Russia
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(10), 2087; https://doi.org/10.3390/sym14102087
Submission received: 4 September 2022 / Revised: 22 September 2022 / Accepted: 27 September 2022 / Published: 7 October 2022
(This article belongs to the Special Issue Symmetry in Inorganic Crystallography and Mineralogy)

Abstract

:
Our systematic explorations of the complex rare earth tellurite halide family have added several new [Ln12(TeO3)12][M6X24] (M = Cd, Mn, Co) representatives containing strongly deficient and disordered metal-halide layers based on transition metal cations. The degree of disorder increases sharply with decrease of M2+ radius and the size disagreements between the cationic [Ln12(TeO3)12]+12 and anionic [M6Cl24]−12 layers. From the crystal chemical viewpoint, this indicates that the families of both rare-earth selenites and tellurites can be further extended; one can expect formation of some more complex structure types, particularly among selenites. Analysis of the polytypism of compounds have been performed using the approach of OD (“order–disorder”) theory.

1. Introduction

The use of species with “one-sided” coordination which terminate propagation of the net of chemical bonds and contribute to interfaces with non-covalent interactions is a common and a powerful instrument in the construction of low-dimensional, porous, and open-framework inorganic structures for various purposes [1,2,3,4,5,6]. This approach employs either organic species with hydrophilic parts contributing to the frameworks and hydrophobic parts forming the interfaces, or cations with the stereoactive lone pair of electrons such as Tl+, Sn2+, Pb2+, As3+, Sb3+, Bi3+, Se4+, Te4+, or I5+ [2,5,6,7,8,9]. The interiors of the covalent frameworks may be filled by small anions, most commonly halides, or much larger ensembles, e.g., salt-inclusion solids [10]. Combinations of “lone-pair” and “regular-coordination” cations, particularly rare-earth and halide anions, often result in layered arrangements comprised of alternating covalent (metal-oxide) and ionic (metal-halide) 2D infinite building blocks bearing opposite charges [1]. They are more frequently observed among compounds containing rare-earth cations with isotropic coordination and SeIV or TeIV which form the interfaces of the cationic metal-oxide parts.
By now, the majority of their structures contain tetragonal or pseudo-tetragonal [Ln11Mn(ChO3)12] layers (Ch = Se or Te, M = Ln or some other uni- or divalent cations) sandwiched between either single halide sheets or more complex metal-halide slabs derived from tetragonal FeS (based on tetrahedrally coordinated Cu+ or Zn2+ and Cl- [11]), CsCl (based on MX8 cubes M = K, Rb and Cs; X = Cl or Br [10,12,13,14,15]) or NaCl (based on CdCl6 or CdBr6 octahedra [16,17]) as shown schematically in Figure 1. Octahedral coordination in halides is common also for some other transition metal dications including Mn2+, Co2+, and Ni2+. Layered rare-earth-chalcogen (IV) oxyhalides containing magnetically active cations have been reported only for the [Ln10M(SeO3)12][Cl8] composition with low amounts of M2+ at very large separations [11]. The possibility of constructing transition metal-rich building blocks within this family has not been explored to date; the only hint to their possible existence is given in [11] where a [La11(SeO3)12][Co7.5Cl24] composition is once mentioned without comments. This formula corresponds to that of the Cd compounds, [Ln12(TeO3)12][Cd6X24], X = Cl, Br [16,17] (Figure 1d); the ordered arrangement of the transition metal (Co2+ in this case) cations only within the metal-halide slabs is expected to result in formation of relatively dense planar or quasi-planar magnetic sublattices situated at relatively large (~13Å) separations.
For the modular compounds with the layered structures (with the weak interactions between the adjacent layers), the presence of the stacking disorder is very common [18]. The slight shifts can form different polytypes and OD (“order–disorder”) structures [19,20,21]. The symmetry analysis of such disordered compounds is based on the concepts of OD theory [22,23] and the symmetry of the OD structures is described by groupoids (instead of crystallographic space groups), which can contain the non-crystallographic symmetry operations [24,25,26]. This approach allows to determine the symmetry relations for the whole family of known polytypes and predict the hypothetical ones [27], which is important for the searching of new materials [28,29,30,31,32,33,34]. The OD structures with the different polytypes have been also described for the tellurium compounds, in particular, for the hydrous magnesium orthotellurate (VI) Mg(H2O)2[TeO2(OH)4] [35], copper-zinc oxotellurite (IV) [Cu2ZnTeO4][SO4 H2O] [36], and alkali metal zinc oxidotellurites (IV), Rb2Zn(TeO3)(CO3)·H2O and Na2Zn2Te4O11 [37].
The presence of stacking disorders for the family of compounds containing tetragonal or pseudo-tetragonal [Ln11Mn(ChO3)12] layers have been recently reported based on the TEM images [38]. It was also mentioned that for the accurate description of the polytypes, the corresponding OD analysis is required [38]. Hereby, we report the results of our attempts to prepare such compounds; contrary to the Cd2+ based prototypes, they exhibit gross disorder in the targeted M2+ sublattices, which could be plausibly modeled in just a few cases. A new representative was also obtained for the family of Ln-Cd tellurite halides; its structure was refined to quite reasonable values and will be described here as a “best ordered” reference structure. The accurate symmetry analysis for these compounds using the approach of OD theory is given and polytypic relations have been determined.

2. Materials and Methods

2.1. Synthesis

As in the previous reports [10,15,16,17], the starting compounds were LnOX, TeO2, and MX2 (M = Mn, Co, Ni). Anhydrous Co and Mn compounds obtained from MnX24H2O and CoX26H2O via dehydration by gentle heating (up to ~100 °C) and subsequent cooling in vacuo (down to ca. 50–70 Pa). Anhydrous nickel halides were prepared in the same way starting from NiX26NH3. The anhydrous halides were pink (MnX2), deep blue (CoCl2), deep green (CoBr2), yellow (NiCl2) and brown (NiBr2). Due to high hygroscopicity of CoX2, all operations with these compounds were conducted in an argon-filled glovebox. The halides of Mn and Ni permit handling in air for a short time. Mixtures of LnOX, MX2, and TeO2 in 1:7:1 ratio were ground, placed in silica capsules, gently heated in a dynamic vacuum until the pressure dropped to ca. 50 Pa, sealed, and annealed according to the following scheme: heating to 600 °C within 12 h, plateau 12 h; heating to 850 °C within 12 h, plateau 120 h; cooling to 650 °C within 120 h. The salt fluxes were dissolved in water (Ni and Mn halides) or 96% ethanol (Co halides). The insoluble residues consisted of small colorless, pinkish (Mn) or deep blue (Co) crystals. Suitable crystals were not produced using NiX2. The novel Ce-Cd-Te oxychloride was obtained as a somewhat unexpected product from a mixture of SrTeO3, CeOCl, TeO2, SrCl2, and CdCl2 (1:1:1:3:4) targeted at a different proposed composition. It is reported here as it belongs to the same structural family.

2.2. Diagnostics

Single crystals of the Co compounds were characterized by energy-dispersive X-ray spectroscopy using a Leo Supra 50 VP electron microscope utilizing 15 kV accelerating voltage and an INCA analyzer. The results indicate that two kinds of crystals were observed in the Eu and Gd-containing samples: those containing only Ln, Te, and Cl, as well as those also containing Co. In the bromide sample, only blue crystals have been studied which were shown to contain La, Co, Te, and Br.

2.3. Single Crystal X-ray Analysis

Though formation of relatively large platelets or respective colors was observed in a variety of samples, even examination in an optical microscope showed most of them then to be piles of very thin foil-like sheets, sometimes with uneven coloring. Acceptable quality crystals were observed for M2+ = Cd, X = Cl, Ln = Ce (1), M2+ = Mn, X = Cl, Ln = Nd (2), M2+ = Mn, X = Br, Ln = La (3), M2+ = Co, X = Cl, Ln = Eu (4), and M2+ = Co, X = Cl, Ln = Gd (5). Single crystals found in the La-Co-Te-O-Br sample were of too low quality and were not studied. The others were selected under a polarizing microscope for single-crystal X-ray diffraction. The single-crystal X-ray data were collected on a Rigaku XtaLAB Synergy diffractometer (HyPix detector, MoKα radiation, λ = 0.71073 Å; ω−θ-scanning mode) (1), Bruker SMART APEX II diffractometer (CCD detector; MoKα radiation, λ = 0.71073 Å; ω−θ-scanning mode) (2,3), and Xcalibur S Oxford Diffraction diffractometer (CCD detector; MoKα radiation, λ = 0.71073 Å; ω−θ-scanning mode) (4,5). Raw data were integrated and then scaled, merged, and corrected for Lorentz-polarization effects using the CrysAlis [39] and Apex2 [40] programs. A semiempirical absorption correction based upon the intensities of equivalent reflections was applied using SADABS [41].
The observed single-crystal diffraction patterns indicated the bad quality of studied crystals which was typically manifested by essential streaking (indicating stacking disorder) and poorly resolved reflections. In accordance with the analysis of systematic absence of reflections, the space group P4/nbm (No. 125) for all compounds was chosen. The experimental details of the data collection and refinement results are also listed in Table 1.
Structure models were determined by the “charge flipping” method using the SUPERFLIP computer program [42] and were refined using the JANA2006 computer program [43]. Illustrations were produced with the JANA2006 program package in combination with the DIAMOND program [44]. Atomic scattering factors for neutral atoms together with anomalous dispersion corrections were taken from International Tables for Crystallography [45]. These essential values of Δρmin and Δρmax are most likely the result of stacking faults common for the layered structures and indicating their possible OD character that was also manifested by streaks in the diffraction patterns. Similar problems, though slightly less pronounced, were encountered in some previous studies [15,16,17]. Attempts to split the corresponding sites were not successful, leading mostly to non-positive definite sites, and were finally abandoned. Table S1 list fractional site coordinates and equivalent displacement parameters for 1 to 5. Selected bond distances are given in Table S2. CCDC 1939965, 1939966, 1974365, 2202775, 2202776 contain the supplementary crystallographic data for these compounds. The data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/structures, accessed on 17 June 2022.

3. Results

3.1. Crystal Structure Description

The crystal structures of the studied compounds 15 are similar to the previously studied ones with the space group P4/nbm (Figure 1d) Their crystal structures contain two blocks which alternate along c parameter. The first block is represented by a complex slab formed by edge-shared Lnφn-polyhedra (n = 8–12) with the general formula [Ln12(TeO3)12] (type B slab [17]) (Figure 2), where Ln = La, Ce, Nd, Eu, and Gd. The second slab is represented by three-layered close-packing formed by the halide X anions (X = Cl or Br) with octahedral sites occupied by the M2+ cations (M = Mn, Co, Cd).
In the crystal structures of compounds 15, the [Ln12–x(TeO3)12] type B slabs are nearly identical, with differences in bond lengths due to the varied size of Ln3+ cations. In the [MII6+yX24] blocks, the M2+ cations fill the octahedral voids; however, refinement of the corresponding site occupancies and analysis of the electron density maps for 13 indicate presence of maxima which we interpret as the additional weakly occupied M sites (close to the main ones) (Figure 3a). The statistical distributions of M2+ cations within the octahedral cavities lead to the formation of two types of the layers characterized by the different orientations of octahedra (Figure 3b,c).
The final refinement cycles converged to reasonable values for 1 and 2; however, an ordered model could be refined only for 1. In the structure of 2 and particularly of 35, occupancies of octahedrally coordinated M2+ positions (M2+ = Mn, Co) were evidently below unity; in addition, relatively high electron density peaks were found in different Fourier maps. Some of these could be tentatively assigned to additional weakly occupied M2+ sites. This is another clear manifestation of the OD character of the structures.

3.2. Polytypism of Tetragonal Halides Containing the [Ln12(TO3)12] Slabs

The diffraction patterns of tetragonal layered rare-earth tellurite halogenides [16,17] are typically characterized by a remarkably low quality of the single crystal X-ray diffraction data which indicates the presence of various stacking faults and possibly some other 2D defects such as single halide sheets. Within the family of related compounds (Figure 1), there are two types of the structures with the following unit cell parameters: a~15.77–16.44 Å, c~12.92–13.77 Å (sp. gr. P4/nbm, Z = 2) and a~15.80–16.12 Å, c~24.82–26.41 Å (sp. gr. I4/mcm, Z = 4). The doubling of the c parameter of the unit cell is observed predominantly for compounds with MI = alkali cation (K, Rb, Cs) with the [MI6X16] composition of the layer (the only exception is compound [Gd12(TeO3)12][Cd6Cl24] [16]).
Taking into account the layered type of the structures as well as different possible ways of their linkage, the observed structural variants can be considered as polytypes [18,46]. The presence of the twinning indicates the possible stacking disorder and OD (“order–disorder”) [47,48] character of the structures. The crystal structures of compounds with the general formulas [Ln12(TeO3)12][MII6X24] and [(Ln,MI)1-xLn11(TeO3)12][MI6X16+y] [15] can be described using the same OD groupoid family, more precisely, a family of OD structures built up by two kinds of non-polar layers (category IV) [49]. The layers are the following (Figure 4a,b):
(i)
L2n type with layer symmetry p4/nmm [or P(4/n)mm in the terms of OD notation] is formed by [M6X24], [M6X16] or [X8] layer (Figure 4b).
(ii)
L2n+1 type with layer symmetry p4/nbm11[or P(4/n)bm in terms of the OD notation, where braces in the third position indicate the direction of missing periodicity [50]] is formed by a [Ln12(TeO3)12] type B slab (Figure 4a).
The layers of both types (L2n and L2n+1) alternate along the c direction and have common translation vectors a and b, with c0, the distance between the two nearest equivalent layers, corresponding to (c/2)~6.6 Å. The ordered or disordered alternation of the two kinds of layers gives rise to a whole family of ordered polytypes or disordered sequences, which can be obtained through the action of the following symmetry operators that may be active in the L2n-type of layer: the 2 and 21 axes parallel to a ([2 1 1] or [21 1 1]) and/or b ([1 2 1] or [1 21 1]). The symmetry relation common to all polytypes of this family are described by the OD groupoid family symbol:
P ( 4 m ) m m P ( 4 n ) b m [ r , s ]
where the first line contains the layer-group symbols of the two constituting layers, while the second line indicates the positional relations between the adjacent layers [51].
In all the polytypes, as well as in the disordered sequences, pairs of adjacent layers are geometrically equivalent (according to the general principle of OD structures). Polytypes presenting the smallest possible number of different triples of layers are called MDO (Maximum Degree of Order) polytypes (the principle of MDO structures). The first MDO structure (MDO1-polytype) can be obtained when the 2x and 2y axes are active in the L2n-type layers (Figure 4c), giving the tetragonal structure with a~16.0 Å, c~13.2 Å and the space group P4/nbm. The second MDO structure (MDO2-polytype) can be obtained when the 21x and 21y axes are active in the L2n-type layers: a~16.0 Å, c~25.5 Å and the space group I4/mcm.
The similar stacking disorder with the formation of different types of polytypes have been also previously observed for the related rare-earth selenites [12,13,38,52,53,54,55]. The crystal structures of [LiPr11(SeO3)12][Rb6Cl16] [13] and [LiNd11(SeO3)12][Rb6Cl16] [12] are crystal chemical isotypical [56] to tellurites with the general formula [(Ln,MI)1-xLn11(TeO3)12][MI6X16+y] [15] and are represented by the MDO1-polytype with the space group I4/mcm.
In the crystal structures of [MIINd10(SeO3)12][Cl8] (MII = Ca, Sr; Figure 1a) [52] and [Cs0.5Sm10.5(SeO3)12][Br8] [55] the type B slabs are linked directly via a single-layered [X8] block (X = Cl, Br), forming the heteropolyhedral framework. Despite the reduced interstitial [X8] block represented only by a single layer with the one atom width the polytypic character remains common for the whole family of compounds. Compounds [MNd10(SeO3)12][Cl8] (M = Ca, Sr) [52] and [Cs0.5Sm10.5(SeO3)12][Br8] [55] are crystal chemical isotypical and represent the third type of MDO structure (MDO3-polytype). This polytype can be obtained when the 21x and 2y axes are active in the L2n-type layers (Figure 5a) giving the orthorhombic (pseudo-tetragonal) structure with a~15.7 Å, a~15.7 Å c~17.9 Å and the space group Bbab (non-standard setting of the space group Ccca).
Compounds [Sm11(SeO3)12][K7Cl16] [54] and [Nd11(SeO3)12][Cs7Cl16] [38] are characterized by another MDO structure (MDO4-polytype) which can be obtained when the 2y axes are active in the L2n-type layers (Figure 5b) giving the orthorhombic (pseudo-tetragonal) structure with a~15.7 Å, a~15.7 Å c~17.9 Å and the space group Pnan. For the compound [Nd11(SeO3)12][Cs7Cl16] [38], symmetry lowering was confirmed by a second harmonic generation test, and the space group Pna21 (which is a subgroup of the space group Pnan) was established. It needs be noted that among selenites, one site in the type B layer is mostly empty while on the CsCl layer, one additional cubic void is occupied instead by a voluminous alkali cation.
When different sequences of operators are active in the L2n-type layers, multilayered non-MDO polytypic structures with increased c occur. The simplest ones are those in which the two types of operators in L2n layers regularly alternate. The crystal structures of [Pr11(SeO3)12][Cs7Cl16] [54] and [Nd11(SeO3)12][Cs7Cl16] [38] with the orthorhombic (pseudo-tetragonal) unit cell parameters a = 15.9, b = 15.9, c = 51.6 Å and the space groups Cmce or Pban are unknown, but the TEM images of [Nd11(SeO3)12][Cs7Cl16] [38] show the regular stacking disorder and different kinds of symmetrical operation active in L2n-type layers.
Except for different MDO and non-MDO structures, the modular structures can also be formed when the different types of the interstitial block between adjacent [Ln12 (TeO3)12] type B slab are present. As a result, the different types of alternation can form a merotypic modular series [18,33,46,57,58,59]. In the crystal structure of [La11(SeO3)12]2[Cs6Cl16][Cl8] [53], there are four type B slabs, two [MI6X16] blocks and two [X8] blocks, and the crystal chemical formula can be written as {[La11(SeO3)12]+9[Cs6Cl16]−10}−1{[La11(SeO3)12]+9[Cl8]−8}+1. Such complex structure is relatively strictly ordered as only the whole layer sequence is electroneutral. The same applies to the compounds {[La11(SeO3)12]+9[Cu6Cl16]−10}−1{[La11(SeO3)12]+9[Cl8]−8}+1 and {[Nd11(SeO3)12]+9[Zn3Cl16]−10}−1{[Nd11(SeO3)12]+9[Cl8]−8}+1 [11]. Such mixed-layer structures are as yet observed only for selenites; note the compositional shift of the CsCl slabs from [MI7Cl16]−9 in “bilayer” structures (Figure 1b) to [MI6X16]−10 in their mixed-layer derivatives [11,53]. In the structures of tellurites, simple “bi-layer” sequences are charge balanced (e.g., [MILn11(TeO3)12]+10[MI6X16]−10 [10]) and formation of mixed-layer structures is not favored. A very similar pattern was observed in our earlier works on complex layered perovskites [60]. Therefore, if the [Ln11(SeO3)12]+9[MII7.5X24]−9 selenite analogs of the compounds 1–5 discussed here do actually exist, one can foresee also existence of their more complex derivatives such as {[Ln11(SeO3)12]+9[MII7X24]−10}−1{[Ln11(SeO3)12]+9[X8]−8}+1, etc.
The comparative data on unit cell parameters, space groups, and symmetry relation between different layers in compounds with the general formulas [Ln12(TeO3)12][MII6X24], [(Ln,MI)1-xLn11(ChO3)12][MI6+zX16+y] (Ch = Se or Te), and [MIILn10(SeO3)12][X8] are summarized in Table 2. Other polytypes can be easily obtained using the different combinations of the symmetry operations active in L2n layer.

4. Discussion

4.1. Synthesis

In line with our expectations, the family of layered rare-earth chalcogenite halides could be more or less successfully extended into the realm of transition-metal compounds. As yet, the generally employed self-flux synthetic pathway (the use of metal halides as both reactants and fluxing agents) is quite effective in producing single crystals of selenites but essentially less so in the case of tellurites; we estimate the probability of preparing suitable single crystals from a single synthetic run to be as low as 20–30%. The possible explanation is that the targeted tellurites are less soluble in the molten fluxes compared to the selenites. The quality of the produced tellurite crystals is also essentially lower in contrast to selenites. This is the reason why most isostructural series are only sparsely characterized, i.e., structures have been determined for certain members only. This approach also does not provide phase-pure samples. The most common byproduct is microcrystalline TeO2 which, in the case of alkali fluxes, is generated due to a side reaction yielding hexahalotellurites and their subsequent hydrolysis upon flux leaching. An alternative chemical vapor approach, which works well for structurally related bismuth selenite halides [61], does not work for rare earths due to much lower volatility of their compounds. Yet, the use of reactive MnX2 and CoX2 fluxes permitted to prepare the first representatives of two new series of rare-earth-manganese and cobalt tellurite halides.

4.2. Structural Trends

All the compounds reported here adopt the arrangement reported previously by us for the [Ln12(TeO3)12][Cd6X24] compounds (X = Cl, Br) [16,17]. This is not surprising, as the MnX2 and CoX2 halides adopt related layered structure type with octahedral coordination of the dication [62,63,64,65,66]. No analogs of the [MLn10(SeO3)12][X8] [52] or [Nd11(SeO3)12]2[Cl8][Zn3Cl16] [11] are as yet known among tellurite halides. In the meantime, selenite analogs of the compounds reported here are also unknown. No mixed-layer structures, known among selenites [11], have been reported among tellurites, as these mixed-layer structures contain the [(M,Ln)11+x(ChO3)12][X8] (Ch = chalcogen) layer sequence. By now, the only “intersection point” between the tetragonal (or pseudotetragonal) tellurite and selenite halides is the structure depicted in Figure 1b, which contains CsCl-derived metal-halide slabs.
The structure of the metal-tellurite [Ln12(TeO3)12]12+ slabs remains nearly the same in all structures reported herein and in [10,11,12,13,14,15,16,17] and are composed by three types of the sheets: the LnO8 and LnO10 polyhedra share edges and vertices to form an “inner” metal-oxide sheet decorated by the TeO3 squashed tetrahedra and LnO4X4 tetragonal antiprisms (“outer” sheet; Figure 2). Refinement of these parts of the structures proceeds relatively smoothly; they exhibit the best ordering and can be considered structure-directing. The structure of 1 is analogous to those of the other early rare-earth-cadmium compounds and adopts the same P4/nbm space group with c~13 Å and one cationic and one anionic slab per unit cell (Figure 1d). As we noted earlier in [16], there exists a considerable mismatch between the size of the layer comprised from regular CdCl6 octahedra (estimated subcell parameter of ~3.75 Å) and the cerium-tellurite slab (a/4 = 4.08Å), which can somewhat be compensated by deformations of the octahedral layer; yet some disorder is already invoked. This discrepancy increases sharply when passing from Cd2+ ([6]r = 0.95 Å) to Mn2+ ([6]r = 0.83 Å) and further to Co2+ ([6]r = 0.75 Å) [67]. In these cases, the distortions are already unable to compensate the adjustment of the metal-halide to the metal-oxide slab. Local deviations from the ideal structure which permit either to preserve the acceptable M–X distances or to adapt some M2+ cations in tetrahedral coordination are very likely to be present. Unfortunately, only an averaged and statistical picture can be extracted from the single-crystal X-ray data; the observed disorder can otherwise reflect just the overlapping contributions from various “islands” with different M–X arrangements. One may suggest that these discrepancies might decrease when passing from MCl6 to larger MBr6 octahedra; however, it is likely to be overruled by the swelling of the unit cell in the ab plane. It is also worth noting that while a polytypic-like transition from P4/nbm to I4/mcm occurs in the [Ln12(TeO3)12][Cd6Cl24] series with the borderline between Eu and Gd compounds, no such transition occurs between 4 and 5. It is, therefore, possible that among compounds of Mn and Co, this transition occurs smoothly and the OD character of the structure is represented by more or less thick lamellae of both P4/nbm and I4/mcm layer sequences. It is also rather likely that the compound mentioned in [11] as [La11(SeO3)12][Co7.5Cl24] also exhibited strong disorder in the metal-halide part of the structure, which could not be plausibly modeled and was thus abandoned. Overall, it should be noted that while metal-halide layers containing magnetically active transition metal cations can indeed be incorporated into the structures of layered rare-earth chalcogenite halides, they form strongly disordered sublattices which are unlikely to result in magnetic ordering.
We also note that by now, the [Ln12(TeO3)12][M6X24] compounds have been obtained for the M2+ cations characterized by zero (Mn2+, Cd2+) or relatively low (Co2+) values of crystal field stabilization energy (CFSE) which permit essential distortions of their octahedral environment. One can suggest that it is also possible to introduce non-magnetic Mg2+, of the size similar to Co2+ ([6]r = 0.72 Å [67]) also characterized by CFSE = 0 (as there are no d orbitals), into the structures reported here.

5. Conclusions

To summarize, we succeeded in further developing the layered rare-earth tellurite-halide family by introducing, for the first time, magnetically active transition metal cations into the halide blocks. Despite essential variation of their size, these cations tend to keep the octahedral environment of halide anions; the geometrical mismatch between the cationic and anionic layer can be, more or less effectively, compensated by the increasing degree of disorder in the transition metal sublattice. This disorder increases essentially also when passing from chlorides to bromides, which is in line with the fact that no tellurite iodide featuring the [(M,Ln)11(TeO3)12] layers has been reported so far. It is more or less evident from this study that further directions of developing this family are possible; however, there is yet little hope in the pursuit of magnetic properties. Other properties related to the type of polytypic structures, including ionic exchange and trapping or soft-chemistry transformations (under mild conditions), might, however, be of interest.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym14102087/s1, Table S1: Fractional site coordinates (xyz), site multiplicities (Q), equivalent displacement parameters (Ueq, Å2), and site occupancy factor for four compounds [Cd6Cl24][Ce12(TeO)12] (1), [Mn6Cl24][Nd12(TeO)12] (2), [Mn6Br24][La12(TeO)12] (3), [Co6Cl24][Eu12(TeO)12] (4) and [Co6Cl24][Gd12(TeO)12] (5); Table S2: Selected interatomic distances (Å) for compounds 1–5.

Author Contributions

Conceptualization, D.O.C. and S.M.A.; methodology, V.A.D.; formal analysis, T.A.O., Y.A.V., S.N.V. and D.V.D.; writing—original draft preparation, D.O.C. and S.M.A.; writing—review and editing, D.O.C. and S.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 20–77-10065 (in the part of the analysis of the polytypism and OD relationship) and the state task of Russian Federation, state registration number [122011300125–2].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zimmermann, I.; Johnsson, M. A Synthetic Route toward Layered Materials: Introducing Stereochemically Active Lone-Pairs into Transition Metal Oxohalides. Cryst. Growth Des. 2014, 14, 5252–5259. [Google Scholar] [CrossRef]
  2. Johnsson, M.; Lidin, S.; Törnroos, K.W.; Bürgi, H.-B.; Millet, P. Host–Guest Compounds in the Family of Tellurium–Nickel Oxohalogenides. Angew. Chem. Int. Ed. 2004, 43, 4292–4295. [Google Scholar] [CrossRef] [PubMed]
  3. Ok, K.M. Toward the Rational Design of Novel Noncentrosymmetric Materials: Factors Influencing the Framework Structures. Acc. Chem. Res. 2016, 49, 2774–2785. [Google Scholar] [CrossRef]
  4. Mayerová, Z.; Johnsson, M.; Lidin, S. Lone-Pair Interfaces That Divide Inorganic Materials into Ionic and Covalent Parts. Angew. Chem. Int. Ed. 2006, 45, 5602–5606. [Google Scholar] [CrossRef]
  5. Zhang, S.-Y.; Hu, C.-L.; Li, P.-X.; Jiang, H.-L.; Mao, J.-G. Syntheses, crystal structures and properties of new lead(ii) or bismuth(iii) selenites and tellurite. Dalt. Trans. 2012, 41, 9532. [Google Scholar] [CrossRef]
  6. Sun, C.; Yang, B.; Mao, J. Structures and properties of functional metal iodates. Sci. China Chem. 2011, 54, 911–922. [Google Scholar] [CrossRef]
  7. Zheng, T.; Wang, Q.; Ren, J.; Cao, L.; Huang, L.; Gao, D.; Bi, J.; Zou, G. Halogen regulation triggers structural transformation from centrosymmetric to noncentrosymmetric switches in tin phosphate halides Sn2PO4X (X = F, Cl). Inorg. Chem. Front. 2022, 9, 4705–4713. [Google Scholar] [CrossRef]
  8. Goerigk, F.C.; Schander, S.; Hamida, M.B.; Kang, D.-H.; Ledderboge, F.; Wickleder, M.S.; Schleid, T. Die monoklinen Seltenerdmetall(III)-Chlorid-Oxidoarsenate(III) mit der Zusammensetzung SE5Cl3[AsO3]4 (SE =La–Nd, Sm). Z. Naturforsch. B 2019, 74, 497–506. [Google Scholar] [CrossRef]
  9. Ahlers, R.; Ruschewitz, U. Non-centrosymmetric coordination polymers based on thallium and acetylenedicarboxylate. Solid State Sci. 2009, 11, 1058–1064. [Google Scholar] [CrossRef] [Green Version]
  10. Charkin, D.O.; Black, C.; Downie, L.J.; Sklovsky, D.E.; Berdonosov, P.S.; Olenev, A.V.; Zhou, W.; Lightfoot, P.; Dolgikh, V.A. Cs7Sm11[TeO3]12Cl16 and Rb7Nd11[TeO3]12Br16, the new tellurite halides of the tetragonal Rb6LiNd11[SeO3]12Cl16 structure type. J. Solid State Chem. 2015, 232, 56–61. [Google Scholar] [CrossRef]
  11. Hamida, M.B. Oxo-Selenate(IV) und Oxo-Arsenate(III) der Selten-Erd-Metalle und ihre Derivate. Ph.D. Thesis, Universität Oldenburg, Oldenburg, Germany, 2007. [Google Scholar]
  12. Lipp, C.; Schleid, T. Rb6LiNd11Cl16[SeO3]12: Ein durch zwei chloridische Flussmittel derivatisiertes Neodym(III)-Oxoselenat(IV). Z. Fuer Krist. 2005, 22, 165. [Google Scholar]
  13. Lipp, C.; Schleid, T. Rb6LiPr11Cl16[SeO3]12: Ein chloridisch derivatisiertes Rubidium-Lithium-Praseodym(III)-Oxoselenat(IV). Z. Für Anorg. Und Allg. Chem. 2006, 632, 2226–2231. [Google Scholar] [CrossRef]
  14. Zitzer, S.; Lipp, C.; Schleid, T. Two isostructural alkali-metal europium chloride oxochalcogenates: Rb6LiEu11Cl16[SeO3]12 versus Cs7Eu11Cl16[TeO3]12. In Proceedings of the European Conference on Solid State Chemistry (ECSCC), Lund, Sweden, 25–28 September 2011; p. 71. [Google Scholar]
  15. Charkin, D.O.; Volkov, S.N.; Dolgikh, V.A.; Aksenov, S.M. Potassium rare-earth tellurite chlorides: A new branch from the old root. Solid State Sci. 2022, 129, 106895. [Google Scholar] [CrossRef]
  16. Kharitonov, I.D.; Charkin, D.O.; Berdonosov, P.S.; Black, C.; Downie, L.J.; Lightfoot, P.; Dolgikh, V.A. Rare-Earth Cadmium Tellurite Chlorides with a Structural Type Exhibiting [Ln12(TeO3)12] Slabs Alternating with CdCl6 Octahedral Layers. Eur. J. Inorg. Chem. 2014, 2014, 3140–3146. [Google Scholar] [CrossRef]
  17. Charkin, D.O.; Grishaev, V.Y.; Volkov, S.N.; Dolgikh, V.A. Synthesis and Structure of New Rare Earth Cadmium Tellurite Halides. Russ. J. Inorg. Chem. 2020, 65, 466–471. [Google Scholar] [CrossRef]
  18. Ferraris, G.; Makovicky, E.; Merlino, S. Crystallography of Modular Materials; Oxford University Press: Oxford, UK, 2008. [Google Scholar]
  19. Zvyagin, B.B. Polytypism of crystal structures. Comput. Math. Appl. 1988, 16, 569–591. [Google Scholar] [CrossRef] [Green Version]
  20. Fichtner, K. Order-disorder structures. Comput. Math. Appl. 1988, 16, 469–477. [Google Scholar] [CrossRef] [Green Version]
  21. Ďurovič, S.; Hybler, J. OD structures in crystallography—basic concepts and suggestions for practice. Z. Krist. Cryst. Mater. 2006, 221, 63–76. [Google Scholar] [CrossRef]
  22. Dornberger-Schiff, K.; Fichtner, K. On the Symmetry of OD-Structures Consisting of Equivalent Layers. Krist. Technol. 1972, 7, 1035–1056. [Google Scholar] [CrossRef]
  23. Dornberger-Schiff, K. Grundzüge einer Theorie der OD-Strukturen aus Schichten. Abhandlungen der Dtsch. Akad. Der Wiss. Zu Berlin. Kl. Chem. Geol. Biol. 1964, 3, 107. [Google Scholar]
  24. Fichtner, K. On the Description of Symmetry of OD Structures (I) OD Groupoid Family, Parameters, Stacking. Krist. Technol. 1979, 14, 1073–1078. [Google Scholar] [CrossRef]
  25. Grell, J. Symmetry Description of OD Crystal Structures in Group Theoretical Terms. Acta Appl. Math. 1998, 52, 261–269. [Google Scholar] [CrossRef]
  26. Stöger, B. Non-Crystallographic Layer Lattice Restrictions in Order-Disorder (OD) Structures. Symmetry 2014, 6, 589–621. [Google Scholar] [CrossRef] [Green Version]
  27. Nespolo, M.; Durovic, S. Crystallographic basis of Polytypism and Twinning in Micas. Rev. Mineral. Geochem. 2002, 46, 155–279. [Google Scholar] [CrossRef]
  28. Topnikova, A.; Belokoneva, E.; Dimitrova, O.; Volkov, A.; Deyneko, D. Rb1.66Cs1.34Tb[Si5.43Ge0.57O15]·H2O, a New Member of the OD-Family of Natural and Synthetic Layered Silicates: Topology-Symmetry Analysis and Structure Prediction. Minerals 2021, 11, 395. [Google Scholar] [CrossRef]
  29. Belokoneva, E.L.; Topnikova, A.P.; Aksenov, S.M. Topolology-symmetry law of structure of natural titanosilicate micas and related heterophyllosilicates based on the extended OD theory: Structure prediction. Crystallogr. Rep. 2015, 60, 1–15. [Google Scholar] [CrossRef]
  30. Belokoneva, E.L. Borate crystal chemistry in terms of the extended OD theory: Topology and symmetry analysis. Crystallogr. Rev. 2005, 11, 151–198. [Google Scholar] [CrossRef]
  31. Reutova, O.; Belokoneva, E.; Volkov, A.; Dimitrova, O. Structure-Properties Relations in Two Iodate Families Studied by Topology-Symmetry Analysis of OD Theory. Symmetry 2021, 13, 1477. [Google Scholar] [CrossRef]
  32. Reutova, O.; Belokoneva, E.; Volkov, A.; Dimitrova, O.; Stefanovich, S. Two New Rb3Sc(IO3)6 Polytypes in Proposed Nonlinear Optical Family A3M(IO3)6 (A = K, Rb; M = Sc, In): Topology–Symmetry Analysis, Order–Disorder and Structure–Properties Relation. Symmetry 2022, 14, 1699. [Google Scholar] [CrossRef]
  33. Aksenov, S.; Antonov, A.; Deyneko, D.; Krivovichev, S.; Merlino, S. Polymorphism, polytypism and modular aspect of compounds with the general formula A2M3(TO4)4 (A = Na, Rb, Cs, Ca; M = Mg, Mn, Fe3+, Cu2+; T = S6+, P5+): Order–disorder, topological description and DFT calculations. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2022, 78, 61–69. [Google Scholar] [CrossRef] [PubMed]
  34. Aksenov, S.M.; Kuznetsov, A.N.; Antonov, A.A.; Yamnova, N.A.; Krivovichev, S.V.; Merlino, S. Polytypism of Compounds with the General Formula Cs{Al2[TP6O20]} (T = B, Al): OD (Order-Disorder) Description, Topological Features, and DFT-Calculations. Minerals 2021, 11, 708. [Google Scholar] [CrossRef]
  35. Stöger, B.; Krüger, H.; Weil, M. Mg(H2O)2 [TeO2(OH)4]: A polytypic structure with a two-mode disordered stacking arrangement. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2021, 77, 605–623. [Google Scholar] [CrossRef]
  36. Stöger, B.; Weil, M.; Missen, O.P.; Mills, S.J. The Order-Disorder (OD) Polytypism of [Cu2ZnTeO4]2+[SO4 ·H2O]2−. Cryst. Res. Technol. 2020, 55, 1900182. [Google Scholar] [CrossRef]
  37. Eder, F.; Stöger, B.; Weil, M. Order-disorder (OD) structures of Rb2Zn(TeO3)(CO3)·H2O and Na2Zn2Te4O11. Z. Krist. Cryst. Mater. 2022, 237, 329–341. [Google Scholar] [CrossRef]
  38. Berdonosov, P.S.; Akselrud, L.; Prots, Y.; Abakumov, A.M.; Smet, P.F.; Poelman, D.; Van Tendeloo, G.; Dolgikh, V.A. Cs7Nd11(SeO3)12Cl16: First Noncentrosymmetric Structure among Alkaline-Metal Lanthanide Selenite Halides. Inorg. Chem. 2013, 52, 3611–3619. [Google Scholar] [CrossRef]
  39. Oxford Diffraction. CrysAlisPro; Oxford Diffraction Ltd: Abingdon, UK, 2009. [Google Scholar]
  40. Bruker. SAINT; Bruker AXS Inc.: Madison, WI, USA, 2012. [Google Scholar]
  41. Bruker. SADABS; Bruker AXS Inc.: Madison, WI, USA, 2001. [Google Scholar]
  42. Palatinus, L.; Chapuis, G. Superflip—A computer program for the solution of crystal structures by charge flipping in arbitrary dimensions. J. Appl. Crystallogr. 2007, 40, 786–790. [Google Scholar] [CrossRef] [Green Version]
  43. Petricek, V.; Dusek, M.; Palatinus, L. Crystallographic Computing System JANA2006: General features. Z. Fuer Krist. 2014, 229, 345–352. [Google Scholar]
  44. Brandenburg, K.; Putz, H. Diamond Version 3; Crystal Impact GbR.: Bonn, Germany, 2005. [Google Scholar]
  45. Prince, E. (Ed.) International Tables for Crystallography, Mathematical, Physical and Chemical Tables, 3rd ed.; Wiley: Hoboken, NJ, USA, 2004; Volume C. [Google Scholar]
  46. Merlino, S. (Ed.) . EMU Notes in Mineralogy. Modular Aspects of Minerals; Eötvös University Press: Budapest, Hungary, 1997; Volume 1. [Google Scholar]
  47. Dornberger-Schiff, K. On order–disorder structures (OD-structures). Acta Crystallogr. 1956, 9, 593–601. [Google Scholar] [CrossRef]
  48. Dornberger-Schiff, K. OD Structures—a Game and a Bit More. Krist. Technol. 1979, 14, 1027–1045. [Google Scholar] [CrossRef]
  49. Dornberger-Schiff, K.; Grell, H. Geometrical properties of MDO polytypes and procedures for their derivation. II. OD families containing OD layers of M > 1 kinds and their MDO polytypes. Acta Crystallogr. Sect. A 1982, 38, 491–498. [Google Scholar] [CrossRef]
  50. Dornberger-Schiff, K. On the nomenclature of the 80 plane groups in three dimensions. Acta Crystallogr. 1959, 12, 173. [Google Scholar] [CrossRef]
  51. Grell, H.; Dornberger-Schiff, K. Symbols for OD groupoid families referring to OD structures (polytypes) consisting of more than one kind of layer. Acta Crystallogr. Sect. A 1982, 38, 49–54. [Google Scholar] [CrossRef]
  52. Berdonosov, P.S.; Olenev, A.V.; Dolgikh, V.A.; Lightfoot, P. The synthesis and crystal structures of the first rare-earth alkaline-earth selenite chlorides MNd10(SeO3)12Cl8 (M=Ca and Sr). J. Solid State Chem. 2007, 180, 3019–3025. [Google Scholar] [CrossRef]
  53. Berdonosov, P.S.; Oleneva, O.S.; Dolgikh, V.A. Tricaesium undecalanthanum dodecaselenate(IV) dodecachloride. Acta Crystallogr. Sect. E Struct. Rep. Online 2006, 62, i29–i31. [Google Scholar] [CrossRef]
  54. Berdonosov, P.S.; Dolgikh, V.A.; Schmidt, P.; Ruck, M. Complex chloride-selenites of rare earthes—a family of new phases. In Proceedings of the IV National Crystal Chemical Conference, Chernogolovka, Russia, 26–30 June 2006; pp. 192–193. [Google Scholar]
  55. Ruck, M.; Schmidt, P. Synthesen und Kristallstrukturen der homöotypen Selenitbromide Bi8(SeO3)9Br6 und CsSm21(SeO3)24Br16. Z. Anorg. Allg. Chem. 2003, 629, 2133–2143. [Google Scholar] [CrossRef]
  56. Lima-de-Faria, J.; Hellner, E.; Liebau, F.; Makovicky, E.; Parthé, E. Nomenclature of inorganic structure types. Report of the International Union of Crystallography Commission on Crystallographic Nomenclature Subcommittee on the Nomenclature of Inorganic Structure Types. Acta Crystallogr. Sect. A Found. Crystallogr. 1990, 46, 1–11. [Google Scholar] [CrossRef] [Green Version]
  57. Veblen, D.R. Polysomatism and polysomatic series: A review and applications. Am. Mineral. 1991, 76, 801–826. [Google Scholar]
  58. Ferraris, G.; Gula, A. Polysomatic Aspects of Microporous Minerals—Heterophyllosilicates, Palysepioles and Rhodesite-Related Structures. Rev. Mineral. Geochem. 2005, 57, 69–104. [Google Scholar] [CrossRef]
  59. Charkin, D.O. Modular approach as applied to the description, prediction, and targeted synthesis of bismuth oxohalides with layered structures. Russ. J. Inorg. Chem. 2008, 53, 1977–1996. [Google Scholar] [CrossRef]
  60. Charkin, D.O.; Akinfiev, V.S.; Alekseeva, A.M.; Batuk, M.; Abakumov, A.M.; Kazakov, S.M. Synthesis and cation distribution in the new bismuth oxyhalides with the Sillén–Aurivillius intergrowth structures. Dalt. Trans. 2015, 44, 20568–20576. [Google Scholar] [CrossRef]
  61. Aliev, A.; Kovrugin, V.M.; Colmont, M.; Terryn, C.; Huvé, M.; Siidra, O.I.; Krivovichev, S.V.; Mentré, O. Revised Bismuth Chloroselenite System: Evidence of a Noncentrosymmmetric Structure with a Giant Unit Cell. Cryst. Growth Des. 2014, 14, 3026–3034. [Google Scholar] [CrossRef]
  62. Tornero, J.D.; Fayos, J. Single crystal structure refinement of MnCl2. Z. Krist. Cryst. Mater. 1990, 192, 147–148. [Google Scholar] [CrossRef]
  63. Wollan, E.O.; Koehler, W.C.; Wilkinson, M.K. Neutron Diffraction Study of the Magnetic Properties of MnBr2. Phys. Rev. 1958, 110, 638–646. [Google Scholar] [CrossRef]
  64. Ferrari, A.; Braibanti, A.; Bigliardi, G. Refinement of the crystal structure of NiCl2 and of unit-cell parameters of some anhydrous chlorides of divalent metals. Acta Crystallogr. 1963, 16, 846–847. [Google Scholar] [CrossRef]
  65. Wilkinson, M.K.; Cable, J.W.; Wollan, E.O.; Koehler, W.C. Neutron Diffraction Investigations of the Magnetic Ordering in FeBr2, CoBr2, FeCl2, and CoCl2. Phys. Rev. 1959, 113, 497–507. [Google Scholar] [CrossRef]
  66. Ketelaar, J.A.A. Die Kristallstruktur des Nickelbromids und -jodids. Z. Krist. Cryst. Mater. 1934, 88, 26–34. [Google Scholar] [CrossRef]
  67. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
Figure 1. Different possible types of the alternations of [(Ln, M)12(TO3)12]n+ slabs with the [X8] slab (a), [MIX16+y] slabs (b), both [X8] and [MIX16+y] slabs (c) and [MII6X24] slabs (d).
Figure 1. Different possible types of the alternations of [(Ln, M)12(TO3)12]n+ slabs with the [X8] slab (a), [MIX16+y] slabs (b), both [X8] and [MIX16+y] slabs (c) and [MII6X24] slabs (d).
Symmetry 14 02087 g001
Figure 2. The general views of the [Ln12–x(TeO3)12] type slab projected on (001) (a) and the same slab with the “outer” Lnn polyhedra (b) in the crystal structures of compounds with the general formula [Ln12(TO3)12][M6X24], where M = Cd, X = Cl, Ln = Ce (compound 1), M= Mn, X = Cl, Ln = Nd (2), M = Mn, X = Br, Ln = La (3), M = Co, X = Cl, Ln = Eu (4), and M = Co, X = Cl, Ln = Gd (5).
Figure 2. The general views of the [Ln12–x(TeO3)12] type slab projected on (001) (a) and the same slab with the “outer” Lnn polyhedra (b) in the crystal structures of compounds with the general formula [Ln12(TO3)12][M6X24], where M = Cd, X = Cl, Ln = Ce (compound 1), M= Mn, X = Cl, Ln = Nd (2), M = Mn, X = Br, Ln = La (3), M = Co, X = Cl, Ln = Eu (4), and M = Co, X = Cl, Ln = Gd (5).
Symmetry 14 02087 g002
Figure 3. Different types of distribution of M cations within the [M6X24] layer: the statistical disordered arrangement (a) and ordered types (b,c).
Figure 3. Different types of distribution of M cations within the [M6X24] layer: the statistical disordered arrangement (a) and ordered types (b,c).
Symmetry 14 02087 g003
Figure 4. The general views and the symmetry of the L2n+1 (a) and L2n (b) OD-layers crystal structure of compounds with the general formula [Ln12(TeO3)12][M6X24] and [(Ln,MI)1-xLn11(TeO3)12][MI6X16+y]. The different types of the arrangements of Ln4-polyhedra of L2n+1 layers across the L2n layer and the symmetry of the triplets L2n–1,L2n,L2n+1 observed in the crystal structures of MDO1-polytype (with the space group P4/nbm) (c) and MDO2-(d) polytype (with the space group I4/mcm) are shown.
Figure 4. The general views and the symmetry of the L2n+1 (a) and L2n (b) OD-layers crystal structure of compounds with the general formula [Ln12(TeO3)12][M6X24] and [(Ln,MI)1-xLn11(TeO3)12][MI6X16+y]. The different types of the arrangements of Ln4-polyhedra of L2n+1 layers across the L2n layer and the symmetry of the triplets L2n–1,L2n,L2n+1 observed in the crystal structures of MDO1-polytype (with the space group P4/nbm) (c) and MDO2-(d) polytype (with the space group I4/mcm) are shown.
Symmetry 14 02087 g004
Figure 5. The different types of the arrangements of Ln4-polyhedra of L2n+1 layers across the L2n layer and the symmetry of the triplets L2n–1,L2n,L2n+1 observed in the crystal structures of MDO3-polytype (with the space group Bbab) (a) and MDO4-polytype (with the space group Pnan) (b).
Figure 5. The different types of the arrangements of Ln4-polyhedra of L2n+1 layers across the L2n layer and the symmetry of the triplets L2n–1,L2n,L2n+1 observed in the crystal structures of MDO3-polytype (with the space group Bbab) (a) and MDO4-polytype (with the space group Pnan) (b).
Symmetry 14 02087 g005
Table 1. Crystal data, data collection, and refinement of grown crystals.
Table 1. Crystal data, data collection, and refinement of grown crystals.
Sample Number[Ce12(TeO3)12]
[Gd6Cl24] (1)
[Nd12(TeO3)12]
[Mn6Cl24] (2)
[La12(TeO3)12]
[Mn6Br24] (3)
[Eu12(TeO3)12]
[Co6Cl24] (4)
[Gd12(TeO3)12]
[Co6Cl24] (5)
Molecular weight (g)5321.55014.35978.45135.45198.4
Temperature (K)293
Cell settingTetragonal
Space groupP4/nbm (125)
a (Å)16.3262 (1)16.0692 (11)16.4400 (49)15.8928 (6)15.8323 (7)
c (Å)13.0257 (1)12.6796 (9)13.5077 (40)12.4614 (13)12.4729 (1)
V3)3471.93 (4)3274.12 (4)3650.77 (3)3147.5 (4)3126.46 (3)
Z2
Calculated density,
Dx (g cm−3)
5.09025.0865.4385.4185.522
Crystal size
(mm)
0.02 × 0.06 × 0.080.01 × 0.03 × 0.060.03 × 0.04 × 0.070.03 × 0.05 × 0.080.08 × 0.10 × 0.12
Crystal formAnhedral grainAnhedral grainAnhedral grainAnhedral grainAnhedral grain
Crystal coloryellowishlilacpinkdeep bluedeep blue
Data collection
DiffractometerRigaku XtaLAB Synergy
(HyPix detector)
Bruker Apex II
(CCD detector)
Xcalibur S Oxford Diffraction
(CCD detector)
Radiation; λMoKα; 0.71073
Absorption
coefficient, μ (mm−1)
15.36816.69625.49619.83920.664
F (000)46154377513444764500
Data range θ(°);
h, k, l
3.2–33.39;
−18 < h < 24,
−25 < k < 24,
−19 < l < 20
2.41–30.25;
−22 < h < 22,
−22 < k < 13,
−16 < l < 17
2.31–30.32;
−23 < h < 22,
−23< k < 23,
−18 < l < 19
3.30–35.37;
−17 < h < 25,
−25 < k < 24,
−19 < l < 19
3.31–35.39;
−24 < h < 24,
−24 < k < 15,
−19 < l < 14
No. of measured
reflections
46,38816,88134,40332,93132,476
Total reflections (N2)
/observed (N1)
3194/29572314/19372625/12043432/10933411/1052
Criterion for
observed reflections
I > 3σ (I)
Rσ/Rint (%)0.96/3.42.95/4.689.77/ 2.9812.24/22.1311.50/21.46
Refinement
Refinement onFull-matrix
least squares on F
Full-matrix
least squares on F
Full-matrix
least squares on F
Full-matrix
least squares on F2
Full-matrix
least squares on F2
Weight scheme1/(σ2|F| + 0.0004F2)1/(σ2|F| + 0.0004F2)1/(σ2|F| + 0.0004F2)1/(σ2|I| + 0.0036I2)1/(σ2|I| + 0.0064I2)
R1, wR2 (all reflection)
(%)
3.65/6.014.03/5.646.29/8.275.74/20.586.26/24.03
GOF (Goodness of fit)
(%)
2.471.751.141.021.04
Max./min. residual e density, (eÅ−3)9.58/−5.003.99/−5.2210.27/−5.516.97/−6.449.15/−5.15
CCDC Number22027761974365220277519399661939965
Table 2. The comparative crystal chemical data on compounds with the general formulas [Ln12(TeO3)12][MII6X24], [(Ln,MI)1-xLn11(ChO3)12][MI6+zX16+y] (Ch = Se or Te), and [MIILn10(SeO3)12][X8] and the character of polytypic relations.
Table 2. The comparative crystal chemical data on compounds with the general formulas [Ln12(TeO3)12][MII6X24], [(Ln,MI)1-xLn11(ChO3)12][MI6+zX16+y] (Ch = Se or Te), and [MIILn10(SeO3)12][X8] and the character of polytypic relations.
CompoundSp. Gr.Symmetry of
the Triplet
L2n–1,L2n,L2n + 1
Symmetry Operation Active in
L2n Layer
Unit Cell ParametersV, Å3Ref.
a, Åb, Åc, Å
MDO1-polytype
[La12(TeO3)12][Cd6Cl24]P4/nbmP(4/n)bm2x, 2y16.40116.40112.9183474.77[17]
[Sm12(TeO3)12][Cd6Cl24]P4/nbmP(4/n)bm2x, 2y15.84215.84213.0793282.36[16]
[Eu12(TeO3)12][Cd6Cl24]P4/nbmP(4/n)bm2x, 2y15.77415.77413.1133262.94[16]
[La12(TeO3)12][Cd6Br24]P4/nbmP(4/n)bm2x, 2y16.43916.43913.7133705.66[17]
[Pr12(TeO3)12][Cd6Br24]P4/nbmP(4/n)bm2x, 2y16.23116.23113.7713627.91[17]
MDO2-polytype
[CsSm11(TeO3)12][Cs6Cl16]I4/mcmP(4/m)bm21x, 21y15.88815.88825.7376496.45[10]
[RbNd11(TeO3)12][Rb6Br16]I4/mcmP(4/m)bm21x, 21y16.11816.11825.9356737.51[10]
[Sm11.397(TeO3)12][K5.809Cl16]I4/mcmP(4/m)bm21x, 21y15.89415.89424.8856287[15]
[Gd11.86(TeO3)12][K4.42Cl16]I4/mcmP(4/m)bm21x, 21y15.80415.80424.9616234[15]
[Ho11.697(TeO3)12][K4.908Cl16]I4/mcmP(4/m)bm21x, 21y15.82615.82624.8176216[15]
[Pr11.781(TeO3)12][K5.657Cl17]I4/mcmP(4/m)bm21x, 21y16.07116.07124.8626421[15]
[Gd12(TeO3)12][Cd6Cl24]I4/mcmP(4/m)bm21x, 21y15.76315.76326.4106561.82[15]
[LiPr11(SeO3)12][Rb6Cl16]I4/mcmP(4/m)bm21x, 21y15.90615.90624.7906271.89[13]
[LiNd11(SeO3)12][Rb6Cl16]I4/mcmP(4/m)bm21x, 21y15.81715.81724.7706196.90[12]
MDO3-polytype
[CaNd10(SeO3)12][Cl8] *BbabPba(b)21x, 2y15.68115.58817.4194257.82[52]
[SrNd10(SeO3)12][Cl8] *BbabPba(b)21x, 2y15.77315.83617.6224401.64[52]
[Cs0.5Nd10.5(SeO3)12][Br8]BbabPba(b)21x, 2y15.79715.79717.6934482.58[55]
MDO4-polytype
[Nd11(SeO3)12][Cs7Cl16]Pna21P12/a(1)2y15.91115.95125.8606563.17[38]
[Sm11(SeO3)12][K7Cl16]PnanP12/a(1)2y15.63315.66425.0756140.25[54]
Non-MDO polytypes and
modular structures
[CsPr11(SeO3)12][Cs7Cl16] *Aeamn.d.n.d.15.99915.98651.69813222.28[54]
[Nd11(SeO3)12][Cs7Cl16]Pbann.d.n.d.15.94115.95451.65613137.29[38]
{[La11(SeO3)12][Cs6Cl16]}{[La11(SeO3)12][Cl8]} *AeamP(4/m)bm21x, 21y16.07316.03743.17611129.16[53]
Pba(b)21x, 2y
* The initial unit cell parameters and the space groups have been transformed to preserve the orientation and stacking direction of the OD layers and modules; n.d., no data.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Charkin, D.O.; Dolgikh, V.A.; Omelchenko, T.A.; Vaitieva, Y.A.; Volkov, S.N.; Deyneko, D.V.; Aksenov, S.M. Symmetry Analysis of the Complex Polytypism of Layered Rare-Earth Tellurites and Related Selenites: The Case of Introducing Transition Metals. Symmetry 2022, 14, 2087. https://doi.org/10.3390/sym14102087

AMA Style

Charkin DO, Dolgikh VA, Omelchenko TA, Vaitieva YA, Volkov SN, Deyneko DV, Aksenov SM. Symmetry Analysis of the Complex Polytypism of Layered Rare-Earth Tellurites and Related Selenites: The Case of Introducing Transition Metals. Symmetry. 2022; 14(10):2087. https://doi.org/10.3390/sym14102087

Chicago/Turabian Style

Charkin, Dmitri O., Valeri A. Dolgikh, Timofey A. Omelchenko, Yulia A. Vaitieva, Sergey N. Volkov, Dina V. Deyneko, and Sergey M. Aksenov. 2022. "Symmetry Analysis of the Complex Polytypism of Layered Rare-Earth Tellurites and Related Selenites: The Case of Introducing Transition Metals" Symmetry 14, no. 10: 2087. https://doi.org/10.3390/sym14102087

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop