Skip to content
BY 4.0 license Open Access Published by De Gruyter (O) November 3, 2022

Mixed-valent 1:1 oxidotellurates(IV/VI) of Na, K and Rb: superstructure and three-dimensional disorder

  • Hermann Völkl , Felix Eder , Berthold Stöger ORCID logo EMAIL logo and Matthias Weil ORCID logo

Abstract

Single crystals of three mixed-valent oxidotellurates(IV/VI) with general formula M2TeIVTeVIO6 (H2O) x (M = Na, K, Rb) were grown under hydrothermal conditions at 210 °C from M2CO3, TeO2 and Te(OH)6 in a 4:1:1 molar ratio. Na2Te2O6·11/8H2O (where TeIVTeVI is abbreviated as Te2) crystallizes as a monoclinic eightfold superstructure with occupational modulation of the water molecules. The structure is built of diperiodic corrugated [Te2O6]2− networks interleaved by the Na atoms and H2O molecules. The hexagonal crystal structures of M2Te2O4(OH)4 (M = K, Rb; x = 2) with P63/mmc symmetry are made up of [Te2O4(OH)4]2− rods extending in the [001] direction. The rods are connected by the M atoms. The [TeIVO2(OH)2] units are disordered about the threefold axis, resulting in pronounced three-dimensional diffuse scattering.

1 Introduction

Tetravalent oxidotellurates feature a rich crystal chemistry owing to the stereochemically active 5s2 electron lone pair and coordination numbers generally ranging from three to five (in extreme cases also two and six) [1]. Within this class of compounds, often modular structures are formed, because the oxidotellurate(IV) anions can condense into rods, layers or three-dimensional networks, where the lone pairs protrude into the empty space between these modules. Moreover, charge-balancing may lead to superstructures with occupational modulation of the O atoms, such as in Ba6Te10O25Br2 [2] or KCa3Te5O12Cl3 [3].

The structural complexity [4] of oxidotellurates(IV) may be increased by combining them with oxidotellurates(VI), which usually feature sixfold coordination for TeVI, forming close to regular [TeVIO6] octahedra. The resulting mixed-valent oxidotellurates(IV/VI) are not only interesting from a structural point of view, but also may have potential for various applications. In general, mixed-valent compounds are interesting due to electronic coupling between the sites with formal charges, associated with the process of intervalence charge transfer (IVCT; [56]) which makes them promising materials as controlled valency semiconductors or as internal redox couples [7, 8]. Only a small number of oxidotellurates(IV/VI) have been reported with 34 unique compositions deposited with the Inorganic Crystal Structure Database (ICSD).

As has already been noted by [9], mixed-valent oxidotellurates(IV/VI) seem to favor a higher coordination number of the TeIV atom. For example, the rare [TeIVO5] coordination has been reported for NH4TeIVTeVIO5(OH) [10] and the very unusual octahedral [TeIVO6] coordination for Rb2TeIVTe3VIO12 [11] and Cs2TeIVTe3VIO12 [12], though the latter compound is prone to losing oxygen [13]. Counter-examples are provided by Cd2Te2IVTeVIO9 and Cd2TeIVTeVIO7 [14], which feature [TeIVO3] trigonal pyramids in the oxidotellurate(IV/VI) frameworks.

Only few hydrous phases or hydrates of mixed-valent oxidotellurates(IV/VI) have been reported, such as the acid H2TeIVTeVIO6 [15] or the hydrate Na2TeIVTeVIO6·3/2H2O [16]. During a systematic study on the formation conditions of such compounds with alkali metals as cations, we obtained three hydrous oxidotellurate(IV/VI) phases, which can be formally written as M2TeIVTeVIO6(H2O) x with M = Na, K, Rb. The Na compound corresponds to the Na2TeIVTeVIO6·3/2H2O phase [16] mentioned above. However, the volume of the primitive unit cell is increased by a factor of eight and the water content of the new Na2TeIVTeVIO6·xH2O phase was determined as x = 11/8 = 1.375.

The K and Rb compounds (x = 2) are isotypic and can be formulated as M2TeIVTeVIO4(OH)4, since no water of crystallization exists in these phases. Henceforth, for brevity the TeIVTeVI pair will always be written as ‘Te2’ in sum formulas.

Attempts to form an analogous Cs compound resulted in the oxidotellurate(VI) CsTeVIO3(OH), which features an order-disorder phase transition and will be described elsewhere.

2 Experimental

2.1 Synthesis

The title compounds were obtained from experiments under hydrothermal conditions. M2CO3 (M = Na, K, Rb), TeO2 and Te(OH)6 in a 4:1:1 molar ratio with a combined mass of 700 mg were thoroughly mixed and introduced into Teflon containers with ca. 3 ml inner volume. The containers were filled to three quarters with deionized water and placed in steel autoclaves. The samples were heated to 210 °C and kept under autogenous pressure at that temperature for one week. After cooling to room temperature within 3 h, the solids were separated, washed with deionized water, ethanol and acetone and dried at ambient conditions. For the Na batch, Na2Te2O6·11/8H2O was obtained as a single phase, whereas for K and Rb additionally colorless needles of the oxidotellurates(VI) KTeO3(OH) [17] and RbTeO3(OH) were obtained. These phases are non-isotypic but closely related to CsTeO3(OH) and will be discussed in a subsequent publication.

2.2 Data collection and refinement

The colourless crystals were preselected under a polarizing microscope. Na2Te2O6·11/8H2O crystallizes as small tabular crystals, M2Te2O4(OH)4 (M = K, Rb) as large hexagonal pillars. The crystals were cut into fragments with a size suitable for single crystal diffraction.

Data were collected under a dry stream of nitrogen on a Bruker KAPPA ApexII diffractometer system using MoK α radiation and fine-sliced ω- and φ-scans. Data were reduced to intensity values with Saint-Plus [18] and a correction for absorption effects applied using the multi-scan approach implemented in Sadabs [18]. The structures were solved with the dual-space approach implemented in Shelxt [19] and refined against F2 with Shelxl [20]. Details on data collection and refinement are summarized in Table 1. Additional crystallographic information on the title structures were deposited in the ICSD database of the FIZ Karlsruhe with the CSD deposition numbers listed at the end of Table 1.

Table 1:

Data collection and structure refinement details.

Na2Te2O6·11/8H2O K2Te2O4(OH)4 Rb2Te2O4(OH)4
Crystal data
Chemical formula H2.75Na2O7.375Te2 H4K2O8Te2 H4O8Rb2Te2
M r 421.95 465.43 558.17
Temperature (K) 100 100 137
Crystal system, space group Monoclinic, P21/n Hexagonal, P63/mmc Hexagonal, P63/mmc
a, b, c (Å) 35.893(3), 14.3864(15), 12.5546(13) 6.2036(8), 6.2036(8), 13.0114(17) 6.4077(9), 6.4077(9), 13.0352(18)
β (°) 54.664(5) 90 90
V3) 5288.5(9) 433.65(13) 463.50(15)
Z 32 2 2
Radiation type MoK α MoK α MoK α
μ (mm−1) 8.951 7.694 16.735
Crystal size (mm) 0.23 × 0.15 × 0.05 0.27 × 0.21 × 0.10 0.25 × 0.20 × 0.02
Data collection
Diffractometer Bruker KAPPA Apex II CCD Bruker KAPPA Apex II CCD Bruker KAPPA Apex II CCD
Absorption correction Multi-scan (Sadabs) Multi-scan (Sadabs) Multi-scan (Sadabs)
T min , T max 0.233, 0.663 0.230, 0.513 0.134, 0.279
No. of measured, independent and observed [I > 2σ(I)] reflections 95135, 19646, 15996 14816, 574, 573 8692, 592, 536
R int 0.025 0.038 0.041
Refinement
(sinθ/λ) max −1) 0.764 0.915 0.904
R[F2>2σ(F2)], wR(F2)

0.024, 0.059


0.026, 0.068


0.026, 0.055
S 1.06 1.35 1.01
Δρ max , Δρ min (e Å−3) 1.80, −1.55 1.13, −0.78 1.91, −1.31
CSD number 2152744 2152745 2152746

Localization and refinement of the H positions as well as refinement of the displacement parameters of disordered O atoms was only possible after removing erroneous reflection intensities (for example owing to shadowing by the beamstop) and optimizing the weighting of reflections.

H atoms of Na2Te2O6·11/8H2O were located from difference Fourier maps and the O–H distance restrained to 0.87(2) Å. For K2Te2O4(OH)4 and Rb2Te2O4(OH)4, the single H site features a 2/3 occupancy and therefore was difficult to locate. Ultimately, a crystal-chemically reasonable position could be located and refined.

K2Te2O4(OH)4 and Rb2Te2O4(OH)4 crystallize in the space group P63/mmc and feature disorder about the threefold rotation axis. All attempts to resolve the disorder by refinement of models in subgroups of P63/mmc, with the lost operation as twin operations, failed. Indeed, distinct three-dimensional diffuse scattering, lack of superstructure reflections, and excellent reliability factors all favor the P63/mmc model. Moreover, the platy crystals featured a hexagonal shape and showed no birefringence when viewed perpendicular to the main plane, as expected for trigonal or hexagonal symmetry.

3 Results and discussion

According to the Robin-Day classification of mixed-valent compounds [6], all title compounds adopt class I, i.e. different sites with specific valences (here +IV and +VI for Te) are developed.

3.1 Na2Te2O6·11/8H2O

Na2Te2O6·11/8H2O [P21/n, V = 5288.5(9) Å3, Z = 32] crystallizes as an eightfold superstructure. Atoms were labeled according to their position in the basic structure with an added letter (e.g. Te1A, Te1B, etc.). The set of atoms that correspond to a single position in the basic structure will be indicated by an appended ‘x’ (e.g. Te1x stands for Te1A, Te1B, etc.).

The basic structure [C2/c, V = 1322.1(2) Å3, Z = 8] corresponds, except for the water molecules, to the Na2Te2O6·3/2H2O structure reported by [16]. The eightfold superstructure is expressed by a quadrupling of the cell volume and the loss of the C-centring. The unit cells are related by ( a , b , c ) = ( 4 a b , b b , a b + c b ) , where a subscript ‘b’ marks basis vectors of the basic structure. Equivalently, the structure can be considered as a commensurately modulated structure with the modulation wave vector q = a b */4−c b */4 (Figure 1).

Figure 1: 
Reconstructed hk2 section of reciprocal space of Na2Te2O6·11/8H2O. The reciprocal basis of the basic structure and the q-vector are indicated.
Figure 1:

Reconstructed hk2 section of reciprocal space of Na2Te2O6·11/8H2O. The reciprocal basis of the basic structure and the q-vector are indicated.

In a crystal-chemical sense, Na2Te2O6·11/8H2O can be described as an alternation of corrugated [Te2O6]2− layers and [Na2·11/8H2O]2+ layers extending parallel to (010) (Figure 2).

Figure 2: 
The crystal structure of Na2Te2O6·11/8H2O viewed down [100]. Te (yellow), Na (pink), O (red) and H (white) atoms are represented by spheres of arbitrary radii.
Figure 2:

The crystal structure of Na2Te2O6·11/8H2O viewed down [100]. Te (yellow), Na (pink), O (red) and H (white) atoms are represented by spheres of arbitrary radii.

The [Te2O6]2− layers, shown including the Na1x atoms in Figure 3(a), feature a diperiodic network, which has already been described by [16]. [TeIVO4+1] units (with TeIV in distorted square-pyramidal coordination) are connected by edges to [Te2IVO8] groups (Figure 4). The [TeVIO6] octahedra are likewise connected via edges to [Te2VIO10] groups (Figure 4).

Figure 3: 
Layers in Na2Te2O6·11/8H2O projected on the layer plane (010): (a) [Te2O6]2− including the Na1x atoms and (b) [Na2·11/8H2O]2+ excluding the Na1x atoms, but including the O1 atoms of the [TeVIO6] octahedra. Atom color codes are as in Figure 2. Super and basic structure cells are indicated in red and black, respectively. Atoms are labeled according to the basic structure for clarity. The symmetry elements of the layers in the basic structure are indicated in one cell using the standard graphical symbols [21].
Figure 3:

Layers in Na2Te2O6·11/8H2O projected on the layer plane (010): (a) [Te2O6]2− including the Na1x atoms and (b) [Na2·11/8H2O]2+ excluding the Na1x atoms, but including the O1 atoms of the [TeVIO6] octahedra. Atom color codes are as in Figure 2. Super and basic structure cells are indicated in red and black, respectively. Atoms are labeled according to the basic structure for clarity. The symmetry elements of the layers in the basic structure are indicated in one cell using the standard graphical symbols [21].

Figure 4: 
One of the eight [Te2VIO10] (Te1A, Te1B) and [Te2IVO8] (Te2A, Te2B) groups in Na2Te2O6·11/8H2O. For clarity, O atoms are only labeled according to the basic structure. Te and O atoms are represented by yellow and red ellipsoids drawn at the 75% probability level.
Figure 4:

One of the eight [Te2VIO10] (Te1A, Te1B) and [Te2IVO8] (Te2A, Te2B) groups in Na2Te2O6·11/8H2O. For clarity, O atoms are only labeled according to the basic structure. Te and O atoms are represented by yellow and red ellipsoids drawn at the 75% probability level.

In the basic structure, both kinds of dimers ([Te2IVO8] and [TeVIO6]) are located on centres of inversion. The dimers are connected via corner-sharing to a checkerboard pattern. The individual atoms are all located on the general position in the basic structure and therefore also in the superstructure, since the site symmetry cannot increase. This means that each position in the basic structure is split into eight positions in the superstructure.

The [Te2O6]2− layers show a moderate modulation with respect to the basic structure, as compiled in Table 2. The largest modulation amplitude is observed for the Te2G (0.237 Å) and the O1B (0.403 Å) atoms, respectively. In most cases, the atoms are displaced by less than 0.2 Å. Owing to the modulation, the description of the [Te2IVO8] dimers is ambiguous. The bridging oxygen atoms (O6x) partake in the shortest [1.8496(18)–1.8779(17) Å] and longest [2.2557(18)–2.5763(18) Å] bonds (Figure 4). Whereas the shorter of these long bonds clearly constitute 4 + 1-coordination, in some cases, where d(Te—O) > 2.5 Å, the [TeIVO4+1] units could be better described as two isolated [TeIVO4] units. As is common for oxidotellurates(IV) [1], a clear distinction cannot be made.

Table 2:

Distances d of the Te and O atoms (excluding water) in Na2Te2O6·11/8H2O to the corresponding positions in the average basic structure. To determine the latter, the atoms were transformed into the basic cell and the C2/c symmetry was applied. The positions of close atoms were averaged. All atoms are located on the general position in the basic and superstructures.

Atom Atom d (Å) Atom Atom d (Å) Atom Atom d (Å) Atom Atom d (Å)
Te1 Te1A 0.072 Te2 Te2A 0.172 O1 O1A 0.062 O2 O2A 0.173
Te1B 0.175 Te2B 0.100 O1B 0.403 O2B 0.075
Te1C 0.081 Te2C 0.150 O1C 0.205 O2C 0.099
Te1D 0.104 Te2D 0.060 O1D 0.198 O2D 0.092
Te1E 0.106 Te2E 0.155 O1E 0.109 O2E 0.064
Te1F 0.096 Te2F 0.071 O1F 0.190 O2F 0.128
Te1G 0.151 Te2G 0.237 O1G 0.379 O2G 0.100
Te1H 0.120 Te2H 0.110 O1H 0.210 O2H 0.199
O3 O3A 0.179 O4 O4A 0.084 O5 O5A 0.124 O6 O6A 0.183
O3B 0.170 O4B 0.278 O5B 0.230 O6B 0.031
O3C 0.123 O4C 0.100 O5C 0.122 O6C 0.178
O3D 0.113 O4D 0.162 O5D 0.192 O6D 0.110
O3E 0.133 O4E 0.086 O5E 0.064 O6E 0.107
O3F 0.014 O4F 0.184 O5F 0.193 O6F 0.115
O3G 0.267 O4G 0.151 O5G 0.131 O6G 0.298
O3H 0.020 O4H 0.263 O5H 0.176 O6H 0.291

The [Na2·11/8H2O]2+ layers feature three Na positions in the basic structure. Na1 [Figure 3(a)] is located on the twofold axis and Na2 [Figure 3(b)] on a general position. In both cases, the superstructure atoms are located on general positions, leading to four (Na1x) and eight (Na2x) independent positions, which all feature mild modulation (max. 0.311 Å, Table 3). The Na3 atom is located on a centre of inversion in the basic structure, which is retained for two positions in the superstructure. These atoms therefore feature no positional modulation. The three remaining Na3x atoms are located on general positions and exhibit distinctly stronger modulation than the previously discussed atoms: 0.473–0.720 Å.

Table 3:

Distances d of the Na atoms and water molecules in Na2Te2O6·11/8H2O to the corresponding positions in the basic structure. The basic structure positions were determined as in Table 2, atoms close to a special position were moved onto that position. Site symmetries are given in parentheses.

Atom Atom d (Å) Atom Atom d (Å)
Na1 (2) Na1A (1) 0.311 OW1 (2) OW1A (1) 0.458
Na1B (1) 0.155 OW1B (1) 0.822
Na1C (1) 0.175 OW1C (1) 0.628
Na1D (1) 0.181 OW1D (1) 0.595
Na2 (1) Na2A (1) 0.272 OW2 (2) OW2A (1) 0.241
Na2B (1) 0.263 OW3 (1) OW3A (1) 0.638
Na2C (1) 0.183 OW3B (1) 0.388
Na2D (1) 0.165 OW3C (1) 0.201
Na2E (1) 0.114 OW3D (1) 0.149
Na2F (1) 0.288 OW3E (1) 0.333
Na2G (1) 0.130 OW3F (1) 0.235
Na2H (1) 0.204
Na3 ( 1 ) Na3A ( 1 ) 0
Na3B (1) 0.473
Na3C ( 1 ) 0
Na3F (1) 0.550
Na3E (1) 0.720

This can be explained by the connectivity to the water molecules, which can be considered as the source of the modulation [Figure 3(b)]. In the basic structure, the OW1 atom is located on the twofold axis, however the four corresponding atoms on the general position of the superstructure are strongly modulated (0.458–0.822 Å, Table 3). For the remaining water molecules, two cases can be distinguished: OW2 is realized in one fourth of the cases and located on the twofold rotation axis of the basic structure, from which it deviates slightly in the single superstructure position (0.241 Å). In the remaining three quarters of the cases, OW3 in the basic structure is located on a general position around the twofold axis. There are six independent positions in the superstructure, with modulation amplitudes of 0.149–0.638 Å.

In total, there are eleven (four OW1, one OW2 and six OW3) water positions in the superstructure resulting in the peculiar composition of Na2Te2O6·11/8H2O. Thus, the water molecules can be considered as occupationally modulated. If only OW2 or only OW3 were realized, the composition would read as Na2Te2O6·H2O or Na2Te2O6·3/2H2O, respectively. As far as we can tell from the refined H-positions, the water molecules only donate hydrogen bonds to O atoms of the [Te2O6]2− layers and not to other water molecules. However, in some cases the situation is not entirely clear.

The strong positional modulation of the Na3x atoms is due to their direct contact to the water molecules. One of the positions in the superstructure (Na3B) is disordered in a 78.7:21.3(3) manner, whereby the second position is close to the centre of inversion in the basic structure. This could be interpreted as another Na3A- or Na3C-type position, which are located on an inversion centre of the superstructure. An Na3A-type position results in a reduced number of water molecules, because it leads to the coordination of a single OW2 position rather than two OW3 positions. However, with an Na3C-like position the composition remains unchanged, as an OW3-like coordination is established, where the Na atom coordinates to two water molecules. Thus, for now the water content is given as 11/8 molecules per formula unit.

3.2 M2Te2O4(OH)4 (M = K, Rb)

K2Te2O4(OH)4 and Rb2Te2O4(OH)4 crystallize in isotypic disordered crystal structures in the hexagonal space group P63/mmc. The main building blocks of the structures are [Te2O4(OH)4]2− rods with 63/mmc symmetry [Figure 5(a) and (b)]. The rods are connected by M atoms located on a site with 3m symmetry.

Figure 5: 
The crystal structure of M2Te2O4(OH)4 (M = K, Rb) viewed down (a) [001] and (b) [110]; (c) a [Te2O4(OH)4]2− rod. M is represented by pink spheres, color codes of the other atoms are as in Figure 2. H atoms were omitted in (a) and (b) for clarity and in (b) the O atoms disordered about the .2. position were merged to a single position. In (c) only one of the three possible positions of the [TeIVO4] groups is shown in each case.
Figure 5:

The crystal structure of M2Te2O4(OH)4 (M = K, Rb) viewed down (a) [001] and (b) [110]; (c) a [Te2O4(OH)4]2− rod. M is represented by pink spheres, color codes of the other atoms are as in Figure 2. H atoms were omitted in (a) and (b) for clarity and in (b) the O atoms disordered about the .2. position were merged to a single position. In (c) only one of the three possible positions of the [TeIVO4] groups is shown in each case.

The [Te2O4(OH)4]2− rods are built of an alternation of [(Te1)VIO2(OH)4] and [(Te2)IVO4] groups [Figure 5(c)]. The TeVI atom is located on the 3 m position and coordinated by six O1 atoms forming a close to regular octahedron [six equal distances: 1.926(2) Å (M = K), 1.9239(18) Å (M = Rb); O–Te–O angles deviating by less than 2° from the ideal angles]. The TeIV atom is located on an mm2 position and bonded to four O atoms within an disphenoidal coordination with short Te–O distances [1.879(7) Å (M = K) and 1.862(5) Å (M = Rb)] for the two terminal O2 atoms and longer distances [2.181(2) Å (M = K) and 2.1843(18) Å (M = Rb)] for the two bridging O1 atoms, which connect the two kinds of oxidotellurate groups.

The [TeIVO4] units are disordered around the threefold axis of the rod group. Hence, only one out of three positions is realized. This leaves the other two O1 atoms that are not connected to the [TeIVO4] units and correspond to the OH groups. The latter form strong hydrogen bonds to the O2 atoms of the [TeIVO4] units [O ⋯ O: 2.551(5) Å (M = K) and 2.562(3) Å (M = Rb)] [Figure 5(c)].

The disorder of the rods can be explained by application of the OD theory [22], which is usually applied to structures built of layers. Here, the rods are decomposed into two kinds of blocks. The [TeVIO6] octahedra (excluding H atoms) feature 3 m symmetry and will be called A1. The [TeIVO4] units and OH groups form the second block and feature mm2 symmetry and will be called A2. The O1 atoms are located at the interface and attributed to both adjacent blocks. Both blocks are non-polar in the sense that they are symmetric by an operation that inverts the rod directions ( 3 , .2. and m.., ..2, respectively). In analogy to structures of layers, a rod structure of two non-polar blocks can be considered as a category IV structure of two kinds of blocks [23].

By application of the threefold rotation of the A1 block onto the A2 block, it becomes evident that given an A1 block, the adjacent A2 block can be placed in three ways that result in congruent (and therefore energetically equivalent) pairs of adjacent blocks. This equivalence of pairs of modules is the crucial aspect of OD structures, since it means that all possible arrangements are locally equivalent. However, triples of adjacent objects may differ. Indeed, there are two kinds of A2A1A2 triples, namely those where the two A2 blocks are related by a twofold screw rotation and those where they are related by a sixfold screw rotation in clockwise (65) or counterclockwise (61) direction (Figure 6). An example of triples with a 61 screw is given in Figure 5(c). The triples possess the overall symmetry .2/m. and .2., respectively.

Figure 6: 
Schematic representation of the two types of A2A1A2 triples in M2Te2O4(OH)4: (a) .2/m. and (b, c) two orientations of .2.. [TeVIO6] units are represented by a green octahedron, the direction of the lone pair of the [TeIVO4] units by a yellow arrow. Symmetry elements are indicated in black.
Figure 6:

Schematic representation of the two types of A2A1A2 triples in M2Te2O4(OH)4: (a) .2/m. and (b, c) two orientations of .2.. [TeVIO6] units are represented by a green octahedron, the direction of the lone pair of the [TeIVO4] units by a yellow arrow. Symmetry elements are indicated in black.

Polytypes of a maximum degree of order (MDO) are those that cannot be decomposed into simpler polytypes [24, 25], that is into polytypes made up of only one subset of the triples, quadruples, etc. of adjacent modules. Applied to [Te2O4(OH)4]2−, there are four MDO rods, which are schematized in Figure 7. MDO1 (mmc, c = 2c0) is composed only of the .2/m. triple and the generating operation is a 21.. operation. MDO2 (6122, c = 6c0) and MDO2 (6522, c = 6c0) are composed of either of the two orientations of the .2. triple and the generating operation is a 61.. or 65.. screw rotation. Finally, in MDO3 (211, c = 2c0) the two orientations of the 0.2. triple alternate and the generating operation is .2..

Figure 7: 
Schematic representation of the MDO rods in M2Te2O4(OH)4. The electron lone pairs of [TeIVO4] units are represented by yellow arrows. (a) MDO1, (b) MDO2, (c) MDO′2, (d) MDO3
Figure 7:

Schematic representation of the MDO rods in M2Te2O4(OH)4. The electron lone pairs of [TeIVO4] units are represented by yellow arrows. (a) MDO1, (b) MDO2, (c) MDO′2, (d) MDO3

The M atoms (site symmetry 3m.) connect the [Te2O4(OH)4]2− rods [Figure 5(a)] Their coordination polyhedron can be derived from an anticuboctahedron [Figure 8]. Six O1 atoms (which make up the [TeVIO6] octahedra) form a central, slightly distorted hexagon [M–O distances: 3.1115(5) Å (M = K), 3.2155(5) Å (M = Rb); O–M–O angles: 66.94(10), 52.89(10)° (M = K) and 69.04(7), 50.89(7)° (M = Rb)]. Three more O1 atoms form an equilateral triangular base [M–O distances: 2.808(3) Å (M = K) and 2.938(2) Å (M = Rb)]. The second triangular base is formed by O2 atoms of the partially occupied [TeIVO4] units [M–O distances: 3.103(6) Å (M = K) and 3.146(4) Å (M = Rb)]. However, on average only two out of three vertices (each consisting of two possible O2 sites) are occupied and thus the M site can be considered as eleven-coordinated.

Figure 8: 
M/O-network in the crystal structure of M2Te2O4(OH)4 (M = K, Rb). Atoms as in Figure 5. Non-labeled oxygen atoms are O1.
Figure 8:

M/O-network in the crystal structure of M2Te2O4(OH)4 (M = K, Rb). Atoms as in Figure 5. Non-labeled oxygen atoms are O1.

Figure 9(a) shows a section of the M2Te2O4(OH)4 crystal structure with fixed z. As stated above, the lone pairs of the TeIV atoms point towards the threefold axis, on which the M atoms are located at a different z value. It is reasonable to assume that only one TeIV for a given z-coordinate can point into a given channel. Thus, the distribution of the [TeIVO4] units in a fixed z section is restricted. In particular, one can describe such a section as a bipartite honeycomb net, where the nodes stand for the centres of the [Te2O4(OH)4] rods and the channels of empty space (represented by yellow and blue disks in Figure 9).

Figure 9: 
(a) Slice parallel to (001) through the TeIVO4 units of M2Te2O4(OH)4 (M = K, Rb). Atom colours are as in Figure 4. Only one out of three TeIVO4 units is realized. Centres of the Te2O4(OH)4 rods and the empty space, towards which the lone pairs point are indicated, by yellow and blue circles, respectively. (b, c) Examples of dimer partitions of a honeycomb net corresponding to a valid M2Te2O4(OH)4 (M = K, Rb) structure. Arrows indicate lone pairs protruding into the channels of empty space. A grey backdrop indicates a unit cell.
Figure 9:

(a) Slice parallel to (001) through the TeIVO4 units of M2Te2O4(OH)4 (M = K, Rb). Atom colours are as in Figure 4. Only one out of three TeIVO4 units is realized. Centres of the Te2O4(OH)4 rods and the empty space, towards which the lone pairs point are indicated, by yellow and blue circles, respectively. (b, c) Examples of dimer partitions of a honeycomb net corresponding to a valid M2Te2O4(OH)4 (M = K, Rb) structure. Arrows indicate lone pairs protruding into the channels of empty space. A grey backdrop indicates a unit cell.

Finding a valid distribution of the [TeIVO4] units then is equivalent to finding a dimer covering of a honeycomb net [26]. Two examples of such nets are given in Figure 9(b) and (c). In particular, these can be considered as the two ‘MDO layers’: In Figure 9(b), all lone pairs point in the same direction (layer symmetry pm2m), whereas in Figure 9(c) adjacent lone pairs always point in different directions [layer symmetry p 6 2 m with lattice basis (2a + b, −a + b)]. These two configurations have been named ‘ladder’ and ‘columnar’, respectively [26]. In the former, all M atoms are eleven-coordinated. In the latter, a third of the M atoms are nine- and the other two thirds are twelve-coorindated. Thus, from the point of view of the M atoms, these arrangements are not equivalent. The M atoms might compensate the change in coordination number by displacement in the [001] direction.

The description in terms of disordered blocks is proved by the diffraction pattern of the M2Te2O4(OH)4 crystals, which feature distinct three-dimensional diffuse scattering (Figure 10). In the c* direction, the diffuse scattering is entirely unstructured, which means that the orientation of succesive [TeIVO4] units along c is uncorrelated.

Figure 10: 
Reconstructed sections of reciprocal space of M2Te2O4(OH)4. Spurious sharp diffraction spots, in particular for the K sample, are due to microcrystalline contaminations on the crystal surface. (a) K2Te2O4(OH)4, (hk0)*; (b) K2Te2O4(OH)4, (h2/3l)*; (c) Rb2Te2O4(OH)4, (hk0)*; (d) Rb2Te2O4(OH)4, (h2/3l)*.
Figure 10:

Reconstructed sections of reciprocal space of M2Te2O4(OH)4. Spurious sharp diffraction spots, in particular for the K sample, are due to microcrystalline contaminations on the crystal surface. (a) K2Te2O4(OH)4, (hk0)*; (b) K2Te2O4(OH)4, (h2/3l)*; (c) Rb2Te2O4(OH)4, (hk0)*; (d) Rb2Te2O4(OH)4, (h2/3l)*.

In planes parallel to (001)* (with constant l), the Rb analogue features distinctly less pronounced diffuse scattering than the K analogue. Thus, the distribution of the [TeIVO4] units is rather well ordered in the (001) plane of Rb2Te2O4(OH)4. The observed intensities at (h+1/3)a* + (k + 1/3)b* and (h + 2/3)a* + (k + 2/3)b* suggest a layer with tripled unit cell volume and the layer lattice basis (2a + b, −a + b). This cell is consistent with the ‘columnar’ arrangement [Figure 9(c)].

The layers in the K analogue are distinctly less ordered. Diffuse scattering still centres around positions suggesting a ‘columnar’ arrangement. Distinct elongation in the a* + b* direction suggests an increased ordering in rods running along 1 1 0 with respect to the basic structure, which is equivalent to the 〈100〉 directions with respect to the basis vectors of the ’columnar’ arrangement [Figure 9(b) and (c)].

4 Conclusion and outlook

Again, oxidotellurates have turned out to be a treasure trove for the study of modular crystal structures, with the anionic networks forming distinct layers or rods. During the current study of mixed-valent alkali oxidotellurates(IV/VI), we could confirm that highly condensed anionic networks with high coordination numbers of the TeIV atoms (4 + 1 and 4) are formed.

The crystal structure of Na2Te2O6·11/8H2O is remarkable, since the eightfold superstructure is not due to an occupational modulation of the O atoms of the Te/O network, but rather an occupational modulation of the water molecules. More synthetic and structural studies will be necessary to confirm the unmodulated structure proposed by [16] or perhaps even other phases with modulated structures and/or other water contents.

The distribution of the TeIV atoms in the M2Te2O4(OH)4 (M = K, Rb) family of structures is related to the dimer partitioning of a honeycomb net, which in turn is equivalent to the lozenge partitioning of a plane or box-partitions. These are well-known, yet still actively researched topics of algebraic combinatorics [27]. Modeling of the disorder with simulation of the diffuse scattering obvserved in the diffraction pattern of M2Te2O4(OH)4 phases is planned in the future.


Corresponding author: Berthold Stöger, X-Ray Centre, TU Wien, Getreidemarkt 9, 1060 Vienna, Austria, E-mail:

Acknowledgements

The authors acknowledge TU Wien Bibliothek for financial support through its Open Access Funding Program.

  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. Christy, A. G., Mills, S. J., Kampf, A. R. A review of the structural architecture of tellurium oxycompounds. Mineral. Mag. 2016, 80, 415–545; https://doi.org/10.1180/minmag.2016.080.093.Search in Google Scholar

2. Stöger, B., Weil, M. The barium oxotellurate(IV) bromides Ba6Te10O25Br2 and Ba3Te3O8Br2 with channel structures. Z. Anorg. Allg. Chem. 2012, 638, 2150–2157.10.1002/zaac.201200082Search in Google Scholar

3. Larvor, C., Stöger, B., Weil, M. The allotwinning of KCa3Te5O12Cl3: an OD interpretation. Z. Kristallogr. 2018, 233, 849–859.10.1515/zkri-2018-2071Search in Google Scholar

4. Krivovichev, S. V. Structural complexity and configurational entropy of crystals. Acta Crystallogr. B 2016, 72, 274–276; https://doi.org/10.1107/s205252061501906x.Search in Google Scholar

5. Murase, R., Leong, C. F., D’Allesandro, D. M. Mixed valency as a strategy for achieving charge delocalization in semiconducting and conducting framework materials. Inorg. Chem. 2017, 56, 14373–14382; https://doi.org/10.1021/acs.inorgchem.7b02090.Search in Google Scholar PubMed

6. Robin, M. B., Day, P. Mixed valent chemistry. Adv. Inorg. Chem. Radiochem. 1968, 10, 247–422.10.1016/S0065-2792(08)60179-XSearch in Google Scholar

7. Day, P. Mixed valent chemistry: a survey of 10 Years progress. Int. Rev. Phys. Chem. 1981, 1, 149–193; https://doi.org/10.1080/01442358109353319.Search in Google Scholar

8. Prassides, K., Ed. Mixed valency systems: applications in chemistry, physics and biology, volume 343 of NATO ASI series; Springer: Dordrecht, 1991.10.1007/978-94-011-3606-8Search in Google Scholar

9. Loeksmanto, W., Moret, J., Maurin, M., Philippot, E. Etude cristallochimique, comparée et conductivité électrique de deux tellurates mixtes: AgxNa2-xTe2IVTe3VIO14 (x = 0,4) et K2TeIVTe3VIO12. J. Solid State Chem. 1980, 33, 209–217.10.1016/0022-4596(80)90122-XSearch in Google Scholar

10. Philippot, E., Benmiloud, L., Maurin, M., Moret, J. Pentacoordination de l’atome de tellure(IV) par les atomes d’oxygène. Etude cristallochimique d’un oxotellurate mixte: NH4[TeIVTeVIO5(OH)]. Acta Crystallogr. 1979, B35, 1986–1989.10.1107/S0567740879008311Search in Google Scholar

11. Meier, S. F., Schleid, Th. RbTe2O6: ein gemischtvalentes rubidiumoxotellurat(IV,VI). Z. Anorg. Allg. Chem. 2002, 628, 2198.10.1002/1521-3749(200209)628:9/10<2198::AID-ZAAC11112198>3.0.CO;2-YSearch in Google Scholar

12. Loopstra, B. O., Goubitz, K. The structures of four caesium tellurates. Acta Crystallogr. 1986, C42, 520–523; https://doi.org/10.1107/s0108270186095537.Search in Google Scholar

13. Siritanon, T., Li, J., Stalick, J. K., Macaluso, R. T., Sleight, A. W., Subramanian, M. A. CsTe2O6-x: novel mixed-valence tellurium oxides with framework-deficient pyrochlore-related structure. Inorg. Chem. 2011, 50, 8494–8501.10.1021/ic2010375Search in Google Scholar

14. Weil, M. New phases in the systems Ca-Te-O and Cd-Te-O: the calcium tellurite(IV) Ca4Te5O14, and the cadmium oxocompounds Cd2Te3O9 and Cd3Te2O7 with mixed-valent oxotellurium(IV/VI) anions. Solid State Sci. 2004, 6, 29–37.10.1016/j.solidstatesciences.2003.11.003Search in Google Scholar

15. Lindqvist, O., Moret, J. The crystal structure of a tellurium(IV,VI) oxyhydroxide, H2Te2O6. Acta Crystallogr. 1973, B29, 956–963.10.1107/S0567740873003687Search in Google Scholar

16. Lee, D. W., Ok, K. M. New polymorphs of ternary sodium tellurium oxides: hydrotherm synthesis, structure determination, and characterization of β-Na2Te4O9 and Na2Te2O6·1.5H2O. Inorg. Chem. 2014, 53, 10642–10648.10.1021/ic501739uSearch in Google Scholar

17. Lindqvist, O. A redetermination of the crystal structure of KTeO3(OH). Acta Chem. Scand. 1972, 26, 4107–4120.10.3891/acta.chem.scand.26-4107Search in Google Scholar

18. Bruker. Apexii, Rlatt, Saint, Sadabs and Twinabs, 2020.Search in Google Scholar

19. Sheldrick, G. M. Shelxt – integrated space-group and crystal-structure determination. Acta Crystallogr. 2015b, A71, 3–8; https://doi.org/10.1107/s2053273314026370.Search in Google Scholar

20. Sheldrick, G. M. Crystal structure refinement with Shelxl. Acta Crystallogr. 2015a, C64, 3–8; https://doi.org/10.1107/s2053229614024218.Search in Google Scholar

21. Hahn, Th., Aroyo, M. I. Symbols of symmetry elements. In Space-group symmetry, volume A of international tables for crystallography; IUCr: Chester, 2016, pp. 144–148. Chapter 2.1.2.10.1107/97809553602060000114Search in Google Scholar

22. Dornberger-Schiff, K., Grell-Niemann, H. On the theory of order-disorder (OD) structures. Acta Crystallogr. 1961, 14, 167–177; https://doi.org/10.1107/s0365110x61000607.Search in Google Scholar

23. 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. 1982, A38, 49–54; https://doi.org/10.1107/s0567739482000096.Search in Google Scholar

24. Dornberger-Schiff, K. Geometrical properties of MDO polytypes and procedures for their derivation. I. General concept and applications to polytype families consisting of OD layers all of the same kind. Acta Crystallogr. 1982, A38, 483–491; https://doi.org/10.1107/s0567739482001041.Search in Google Scholar

25. Dornberger-Schiff, K., Grell, H. Geometrical properties of MDO polytypes and procedures for their derivation. II. OD families containing OD layers of kinds and their MDO polytypes. Acta Crystallogr. 1982, A38, 491–498.10.1107/S0567739482001053Search in Google Scholar

26. Yamamoto, H., Kamiyama, S., Yamada, I., Kimura, H. Cation dimerization in a 3d1 honeycomb lattice system. J. Am. Chem. Soc. 2021, 144, 1082–1086.10.1021/jacs.1c10977Search in Google Scholar PubMed

27. Yeliussizov, D. Random plane partitions and corner distributions. Algebraic Combinatorics 2021, 4, 599–617; https://doi.org/10.5802/alco.171.Search in Google Scholar

Received: 2022-06-02
Accepted: 2022-10-13
Published Online: 2022-11-03
Published in Print: 2023-01-27

© 2022 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 28.4.2024 from https://www.degruyter.com/document/doi/10.1515/zkri-2022-0036/html
Scroll to top button