research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296

LaTe1.82(1): modulated crystal structure and chemical bonding of a chalcogen-deficient rare earth metal polytelluride1

CROSSMARK_Color_square_no_text.svg

aInorganic Chemistry, Technische Universität Dresden, Bergstrasse 66, Dresden 01069, Germany, and bTheoretical Chemistry, Technische Universität Dresden, Bergstrasse 66c, Dresden 01069, Germany
*Correspondence e-mail: thomas.doert@tu-dresden.de

Edited by A. L. Spek, Utrecht University, The Netherlands (Received 7 February 2020; accepted 10 April 2020; online 6 May 2020)

Crystals of the rare earth metal polytelluride LaTe1.82(1), namely, lanthanum telluride (1/1.8), have been grown by molten alkali halide flux reactions and vapour-assisted crystallization with iodine. The two-dimensionally incommensurately modulated crystal structure has been investigated by X-ray diffraction experiments. In contrast to the tetra­gonal average structure with unit-cell dimensions of a = 4.4996 (5) and c = 9.179 (1) Å at 296 (1) K, which was solved and refined in the space group P4/nmm (No. 129), the satellite reflections are not compatible with a tetra­gonal symmetry but enforce a symmetry reduction. Possible space groups have been derived by group–subgroup relationships and by consideration of previous reports on similar rare earth metal polychalcogenide structures. Two structural models in the ortho­rhom­bic superspace group, i.e. Pmmn(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 (No. 59.2.51.39) and Pm21n(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 (No. 31.2.51.35), with modulation wave vectors q1 = αa* + βb* + [1 \over 2]c* and q2 = −αa* + βb* + [1 \over 2]c* [α = 0.272 (1) and β = 0.314 (1)], have been established and evaluated against each other. The modulation describes the distribution of defects in the planar [Te] layer, coupled to a displacive modulation due to the formation of different Te anions. The bonding situation in the planar [Te] layer and the different Te anion species have been investigated by density functional theory (DFT) methods and an electron localizability indicator (ELI-D)-based bonding analysis on three different approximants. The temperature-dependent electrical resistance revealed a semiconducting behaviour with an estimated band gap of 0.17 eV.

1. Introduction

The structural chemistry of the rare earth metal chalcogenides REX2–δ (RE = Y, La–Nd, Sm, Gd–Lu; X = S, Se or Te; 0 ≤ δ ≤ 0.2) with trivalent RE metals has attracted attention because of their structural variety in a quite small compositional range. The sulfides and selenides of this class of compounds have been intensively investigated, illuminating several different (super)structures due to different amounts of defects δ and the formation and arrangements of chalcogenide X2− and polychalcogenide Xn2− anions for charge balancing. A comprehensive overview discussing these aspects can be found in Doert & Müller (2016[Doert, T. & Müller, C. J. (2016). Binary Polysulfides and Polyselenides of Trivalent Rare-Earth Metals, in Reference Module in Chemistry, Molecular Sciences and Chemical Engineering, edited by J. Reedijk. Waltham, MA: Elsevier.]). At the beginning of the 21st century, the structures of four rare earth metal polytellurides RETe2–δ (RE = La–Nd) (Stöwe, 2000a[Stöwe, K. (2000a). J. Solid State Chem. 149, 155-166.],b[Stöwe, K. (2000b). J. Alloys Compd. 307, 101-110.],c[Stöwe, K. (2000c). Z. Anorg. Allg. Chem. 626, 803-811.], 2001[Stöwe, K. (2001). Z. Kristallogr. Cryst. Mater. 216, 215-224.]), were thoroughly (re)investigated, revealing considerable differences to the sulfides and selenides of analogous compositions, while still maintaining the same general structural motif. This general structural motif of all binary polychalcogenides REX2–δ of trivalent rare earth metals is closely related to the structure of ZrSSi (space group P4/nmm, No. 129; a0 ≃ 3.80 and c0 ≃ 8.00 Å) (Onken et al., 1964[Onken, H., Vierheilig, K. & Hahn, H. (1964). Z. Anorg. Allg. Chem. 333, 267-279.]; Klein Haneveld & Jellinek, 1964[Klein Haneveld, A. & Jellinek, F. (1964). Recl. Trav. Chim. Pays-Bas, 83, 776-783.]), which shows an alternating stacking of puckered [ZrS] slabs and square-planar [Si] layers along [001]. The binary rare earth metal chalcogenides REX2–δ (RE = Y, La–Nd, Sm, Gd–Lu; X = S, Se or Te; 0 ≤ δ ≤ 0.2) comprise puckered [REX]+ slabs and planar chalcogenide layers, which can formally be stated as [X] (Doert & Müller, 2016[Doert, T. & Müller, C. J. (2016). Binary Polysulfides and Polyselenides of Trivalent Rare-Earth Metals, in Reference Module in Chemistry, Molecular Sciences and Chemical Engineering, edited by J. Reedijk. Waltham, MA: Elsevier.]). For electronic reasons, the chalcogenide layers of the stoichiometric REX2 compounds, especially the di­sulfides RES2 and diselenides RESe2, feature only X22− dianions, resulting in a distortion from an idealized square-planar net towards a planar herringbone pattern; for ditellurides RETe2, the structural situation is not that uniform (Stöwe, 2000a[Stöwe, K. (2000a). J. Solid State Chem. 149, 155-166.],b[Stöwe, K. (2000b). J. Alloys Compd. 307, 101-110.],c[Stöwe, K. (2000c). Z. Anorg. Allg. Chem. 626, 803-811.]). Going to the off-stoichiometric REX2–δ (0 < δ ≤ 0.2) compounds, vacancies in the planar chalcogenide layers are observed, together with X2− to maintain the overall net charge [X] for the layer. This structural change is obvious for the CeSe1.9 structure type, but can also be seen for the related, intrinsically disordered Gd8Se15-type structures (Doert et al., 2012[Doert, T., Graf, C., Vasilyeva, I. G. & Schnelle, W. (2012). Inorg. Chem. 51, 282-289.]; Doert, Dashjav et al., 2007[Doert, T., Dashjav, E. & Fokwa, B. P. T. (2007). Z. Anorg. Allg. Chem. 633, 261-273.]). Hence, the two most important factors accounting for structural differences are the amount of vacancies in the chalcogenide layer and the ionic radii of the trivalent rare earth metal cations, as they largely determine the Coulomb repulsion between the anions in the [X] layers in this series. In addition, in accordance with the Zintl-like electron localization, the polysulfides RES2–δ and polyselenides RESe2–δ are semiconductors.

To distinguish between different anionic fragments in the chalcogenide layer, classical electron counting has been proven a simple but powerful way to describe these structures, as briefly explained for the CeSe1.9 structure type; the CeSe1.9 type is a [\sqrt 5] × [\sqrt 5] × 2 superstructure of the basic ZrSSi unit cell and crystallizes in the space group P42/n (No. 86) (Plambeck-Fischer et al., 1989[Plambeck-Fischer, P., Abriel, W. & Urland, W. (1989). J. Solid State Chem. 78, 164-169.]). The planar [Se] layer of this compound is built up by four dinuclear Se22− anions, forming a pinwheel-like arrangement around a vacancy. The complementary isolated Se2− anion is surrounded by four Se22− anions in a spoke-like manner (Lee & Foran, 1994[Lee, S. & Foran, B. (1994). J. Am. Chem. Soc. 116, 154-161.]). Assuming only trivalent rare earth metal cations, ten positive charges per [REX]+ layer and unit cell need to be balanced by nine atoms of the planar [X] layer. This is achieved by four Se22− anions and one isolated Se2− anion. This kind of charge-ordered superstructure has only been reported once for a rare earth metal telluride, namely for CeTe1.9 (Ijjaali & Ibers, 2006[Ijjaali, I. & Ibers, J. A. (2006). J. Solid State Chem. 179, 3456-3460.]), whereas many examples are known for the rare earth metal polysulfides and polyselenides (Doert, Graf, Lauxmann et al., 2007[Doert, T., Graf, C., Lauxmann, P. & Schleid, T. (2007). Z. Anorg. Allg. Chem. 633, 2719-2724.]; Grupe & Urland, 1991[Grupe, M. & Urland, W. (1991). J. Less-Common Met. 170, 271-275.]; Plambeck-Fischer et al., 1989[Plambeck-Fischer, P., Abriel, W. & Urland, W. (1989). J. Solid State Chem. 78, 164-169.]; Urland et al., 1989[Urland, W., Plambeck-Fischer, P. & Grupe, M. (1989). Z. Naturforsch. Teil B, 44, 261-264.]; Dashjav et al., 2000[Dashjav, E., Doert, T., Böttcher, P., Mattausch, H. & Oeckler, O. (2000). Z. Kristallogr. New Cryst. Struct. 215, 337-338.]; Müller et al., 2012[Müller, C. J., Schwarz, U. & Doert, T. (2012). Z. Anorg. Allg. Chem. 638, 2477-2484.]).

An unusual case of charge balancing for the deficient REX2–δ compounds has been reported for structures with a composition of RETe1.8 (Sm, Gd–Dy) by forming larger anionic fragments (Ijjaali & Ibers, 2006[Ijjaali, I. & Ibers, J. A. (2006). J. Solid State Chem. 179, 3456-3460.]; Wu et al., 2002[Wu, Y., Doert, T. & Böttcher, P. (2002). Z. Anorg. Allg. Chem. 628, 2216.]; Gulay et al., 2007[Gulay, L. D., Stępień-Damm, J., Daszkiewicz, M. & Pietraszko, A. (2007). J. Alloys Compd. 427, 166-170.]; Poddig et al., 2018[Poddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886-1896.]). Here, a similar enlargement of the basic lattice parameters of [\sqrt 5] × [\sqrt 5] × 2 is observed, and the compounds crystallize in a 10-fold superstructure of the aristotype in P4/n (No. 85). In contrast to the respective sulfides and selenides, a motif of statistically disordered Te22− anions and linear Te34− anions are found here. Linear Te34− anions have rarely been reported in REX2–δ compounds, although the presence of an Se34− anion was discussed for DySe1.84, but neglected after computational studies (van der Lee et al., 1997[Lee, A. van der, Hoistad, L. M., Evain, M., Foran, B. J. & Lee, S. (1997). Chem. Mater. 9, 218-226.]). The bonding situation in such linear trinu­clear anions, like Te34− or I3, requires the occupation of nonbonding states, similar to the situation of the prominent XeF2 mol­ecule. Within the concept of mol­ecular orbital (MO) theory, this situation can be described as a 3c–4e bond (Rundle, 1963[Rundle, R. E. (1963). J. Am. Chem. Soc. 85, 112-113.]; Assoud et al., 2007[Assoud, A., Xu, J. & Kleinke, H. (2007). Inorg. Chem. 46, 9906-9911.]). A density functional theory (DFT)-based study clearly evidenced such a linear Te3 unit in GdTe1.8 (Poddig et al., 2018[Poddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886-1896.]) and confirmed an alternative method of electron localization for this composition of REX1.8: ten positive charges of one puckered [REX]+ layer per unit cell are balanced by two Te34− anions and one Te22− anion.

Starting from the results of the RETe1.8 (RE = Sm, Gd–Dy) compounds, we were inter­ested in identifying the structural motifs of the early rare earth metal tellurides RETe2–δ with δ > 0. This was especially motivated by the reported differences between the structures of LaTe2, CeTe2 and PrTe2 (Stöwe, 2000a[Stöwe, K. (2000a). J. Solid State Chem. 149, 155-166.],b[Stöwe, K. (2000b). J. Alloys Compd. 307, 101-110.],c[Stöwe, K. (2000c). Z. Anorg. Allg. Chem. 626, 803-811.]), and the corresponding sulfides and selenides. Structural data on tellurides RETe2–δ with a comparable low chalcogen content have rarely been reported; RETe1.8 (RE = Sm, Gd–Dy) are a few examples (Ijjaali & Ibers, 2006[Ijjaali, I. & Ibers, J. A. (2006). J. Solid State Chem. 179, 3456-3460.]; Wu et al., 2002[Wu, Y., Doert, T. & Böttcher, P. (2002). Z. Anorg. Allg. Chem. 628, 2216.]; Gulay et al., 2007[Gulay, L. D., Stępień-Damm, J., Daszkiewicz, M. & Pietraszko, A. (2007). J. Alloys Compd. 427, 166-170.]; Poddig et al., 2018[Poddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886-1896.]). The first results on the lanthanum compound LaTe1.82(1) are presented in the following.

2. Experimental

2.1. Synthesis

All preparation steps were carried out in an argon-filled (5.0, Praxair) glove-box (MBraun, Garching, Germany). Crystals were grown by the addition of a small amount of I2 to the reaction mixture in closed silica ampoules. In a standard synthesis, 500 mg of a stoichiometric mixture of lanthanum (99.9%, Edelmetall Recycling m&k GmbH) and tellurium (Merck, >99.9%, reduced in a H2 stream at 673 K) were placed in a quartz tube and flame sealed under dynamic vacuum (p ≤ 1 × 10−3 mbar). The ampoule was heated slowly with a ramp of 2 K min−1 to 1073 K. The reaction takes place in a gradient from 1123 to 1073 K with I2 (Roth, >99.8%, purified by sublimating twice prior to use) as transporting agent. After 7 d, the ampoule was cooled to room temperature. As we observed a slow degrading of the compounds under atmos­pheric conditions, the samples were stored under argon.

2.2. Powder diffraction

Data collection was performed at 296 (1) K with an Empyrean diffractometer (PANalytical) equipped with a curved Ge(111) monochromator using Cu Kα1 radiation (λ = 1.54056 Å). The scans covered the angular range from 5 to 90° 2θ. Rietveld refinement using the fundamental parameter approach was performed with TOPAS (Version 5; Coelho, 2018[Coelho, A. A. (2018). J. Appl. Cryst. 51, 210-218.]).

2.3. Single-crystal diffraction

Crystal data, data collection and structure refinement details are summarized in Table 1[link]. Data for the modulated structure were integrated and corrected for Lorentz and polarization factors, before applying a numerical absorption correction with the program JANA2006 (Petříček et al., 2014[Petříček, V., Dušek, M. & Palatinus, L. (2014). Z. Kristallogr. Cryst. Mater. 229, 345-352.]). The structure was solved using the charge-flipping method of the program SUPERFLIP (Palatinus & Chapuis, 2007[Palatinus, L. & Chapuis, G. (2007). J. Appl. Cryst. 40, 786-790.]) implemented in the JANA2006 software; the atomic positions were synchronized with those of the average structure. Structure refinement was performed with JANA2006 against F2 including anisotropic displacement parameters for all atoms. Second-order satellites were neglected because of their low intensity (about 99% of these reflections were found with intensities below 3σ) and two harmonic waves have been used for the fit of the atomic modulation functions.

Table 1
Crystallographic data and refinement details for LaTe1.82(1)

  Model Pmmn       Model Pm21n      
Refined composition LaTe1.811 (4)       LaTe1.825 (3)      
Formula weight (g mol−1); F(000) 370.8; 303       371.8; 304      
Crystal size (mm−3) 0.0588 × 0.0472 × 0.0094
Diffractometer, radiation Bruker Kappa APEX II, Mo Kα (0.71073 Å)
Temperature 296 (1) K
Lattice parameters (Å) a = 4.5020 (5), b = 4.4985 (5), c = 9.181 (1)
  α = β = γ = 90°
Modulation vectors q1 = αa* + βb* + γc*
  q2 = −αa* + βb* + γc*
  α = 0.272 (1), β = 0.314 (1), γ = [1 \over 2]
Index range measured −7≤h≤8; −8≤k≤8; −16≤l≤9; −1≤m,n≤1
  2.18≤θ≤38.36
Measured reflections 30619
Absorption coefficient μ (mm) 25.201
Tmin, Tmax 0.3929, 0.4983
Extinction parameter (Becker & Coppens, 1974[Becker, P. J. & Coppens, P. (1974). Acta Cryst. A30, 148-153.]) 0.151       0.153      
Independent reflections 1757, 738 > 3σ(I)       6103, 1386 > 3σ(I)      
Main reflection 415, 323 > 3σ(I)       1080, 719 > 3σ(I)      
First-order satellites 1342, 415 > 3σ(I)       3946, 664 > 3σ(I)      
Rint; Rσ 0.0556, 0.0461 for I > 3σ(I) 0.0469, 0.0600 for I > 3σ(I)        
  0.1563, 0.1934 for all 0.1435, 0.2810 for all
Superspace group, Z Pmmn(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 (No. 59.2.51.39), 2 Pm21n(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 (No. 31.2.51.35), 2
Refinement method JANA2006, full-matrix against F2
Restrictions/parameters 0/33       0/66      
  R1 [3σ(I)] R1 (all) wR2 [3σ(I)] wR2 (all) R1 [3σ(I)] R1 (all) wR2 [3σ(I)] wR2 (all)
All reflections 0.0495 0.1300 0.0872 0.1111 0.0506 0.2131 0.0883 0.1287
Main reflections 0.0247 0.0385 0.0509 0.0542 0.0326 0.0554 0.0575 0.0610
First-order satellites 0.1079 0.2470 0.1891 0.2479 0.1000 0.3333 0.1949 0.2928
Goodness-of-fit [3σ(I)/all] 1.53/1.26       1.30/0.90      
Largest difference peak/hole (e Å−3) 15.11/−17.25       13.21/−14.87      
Computer programs: APEX2 (Bruker, 2010[Bruker (2010). APEX2. Bruker AXS Inc., Madison, Wisconsin, USA.]), SAINT-Plus (Bruker, 2017[Bruker (2017). SAINT-Plus. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT2014 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SUPERFLIP (Palatinus & Chapuis, 2007[Palatinus, L. & Chapuis, G. (2007). J. Appl. Cryst. 40, 786-790.]), SHELXL2019 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), JANA2006 (Petříček et al., 2014[Petříček, V., Dušek, M. & Palatinus, L. (2014). Z. Kristallogr. Cryst. Mater. 229, 345-352.]), OLEX2 (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]), SADABS (Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.]) and DIAMOND (Brandenburg, 2019[Brandenburg, K. (2019). DIAMOND. Crystal Impact GbR, Bonn, Germany.]).

2.4. Scanning electron microscopy (SEM) and EDS

SEM was performed with an SU8020 (Hitachi) with a triple-detector system for secondary and low-energy backscattered electrons (Ua = 5 kV). The composition of selected single crystals was determined by semiqu­anti­tative energy dispersive X-ray analysis (Ua = 20 kV) with a Silicon Drift Detector (SDD) X-MaxN (Oxford).

2.5. Computational methods

Solid-state calculations were performed with the all-electron code FHI-aims (Blum et al., 2009[Blum, V., Gehrke, R., Hanke, F., Havu, P., Havu, V., Ren, X., Reuter, K. & Scheffler, M. (2009). Comput. Phys. Commun. 180, 2175-2196.]) for three ordered structural models of LaTe1.82(1). The FHI-aims calculations for subsequent real-space analysis were performed with a (2 × 2 × 2) k-point grid (model in P1) and a (3 × 2 × 2) k-point grid (model in Amm2 and A2) using the zeroth-order scalar relativistic zora correction, collinear spins, the numerical atom-centred basis of light level and the PBE functional (Perdew et al., 1996[Perdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865-3868.]). Real-space properties were evaluated subsequently with the help of the program package DGrid (Kohout, 2016[Kohout, M. (2016). DGrid. Version 5.0. https://www2.cpfs.mpg.de/kohout/dgrid.html.]). Hereby, the electron localizability indicator (ELI-D) was calculated on a grid with a 0.1 Bohr mesh size.

2.6. Temperature-dependent electrical resistance

The electrical resistance of LaTe1.82(1) was measured between 7 and 345 K with a mini-CFMS (Cryogenic Ltd, London). Four gold contacts were attached to the surface of a single crystal in a linear set-up with a carbon conductive composite 7105 (DuPont) to establish the electrical contact between the crystal and the gold wires.

3. Results and discussion

3.1. Synthesis

Black plate-like single crystals of previously unreported LaTe1.82(1) were obtained starting from the elements by alkali halide flux reactions or solid-state reactions with a small amount of I2 for mineralization in fused-silica ampoules. Temperatures above 1173 K in the presence of I2 lead to an attack on the ampoule wall, whereas temperatures of about 1073 to 1123 K are well suited for crystal growth, without noticeable side reactions with the ampoule material. The best results were achieved in a small gradient from 1123 to 1073 K, where crystals of about 0.3 mm in length were grown. The amount of added I2 needs to be compensated by excess La when preparing the experiment due to the formation of LaI3 during and after the experiment, which in turn alters the composition slightly. Solid LaI3 was mainly found at the sink of the ampoule, whereas the desired product was found at the source of the ampoule. A similar procedure at higher tem­per­atures was chosen, e.g. for RETe1.8 (RE = Gd, Tb or Dy) (Poddig et al., 2018[Poddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886-1896.]).

3.2. Diffraction image

The strong reflections of the powder pattern can be indexed with a tetra­gonal unit cell with a = b ≃ 4.50 Å and c ≃ 9.17 Å. As a starting point for the Rietveld refinement, the space group (P4/nmm, No. 129) and the atom sites of the aristotype, the ZrSSi type, were chosen. The fit shows some additional unindexed reflections, which are not compatible with the space group P4/nmm and its most prominent lower symmetric subgroups (Fig. 1[link]).

[Figure 1]
Figure 1
Rietveld refinement of LaTe1.82(1). The space group of the aristotype ZrSSi (P4/nmm), with corresponding atom sites as the starting point for structure analyses, has been chosen. Satellite reflections are marked with an asterisk.

The diffraction image of a single crystal at ambient temperature reveals, as already indicated by powder diffraction, additional weak reflections in the layers (hkn) with n = 0, ±1, ±2,… (Fig. S1 in the supporting information). Slightly stronger reflections can be observed in the layers (hkn) with n = ±0.5, ±1.5,… (Fig. S1). These additional reflections can­not be indexed with a commensurate superstructure of the basic unit cell and were treated as satellites. Moreover, the distribution of the satellites with respect to the main reflections suggest a two-dimensional modulation. The whole diffraction image can then be indexed with five indices hklm1m2 according to:

[H_i = ha^* + kb^* + lc^* + m_1q_1 + m_2q_2]

with

[q_1 = \alpha a^* + \beta b^* + {{1}\over{2}}c^*]

and

[q_2 = -\alpha a^* + \beta b^* + {{1}\over{2}} c^*.]

The modulation wave vector components α and β were determined to be 0.272 (1) and 0.314 (1), respectively. A schematic image of the relative positions of the satellite reflections with respect to the main reflections is displayed in Fig. 2[link] and reconstructed precession images are shown in Fig. S1 (see supporting information). The schematic figure illustrates also that two different q vectors are necessary to index the complete diffraction image. Equivalent satellites are linked by a twofold rotational axis (Fig. S1), pointing towards ortho­rhom­bic (or lower) symmetry for the modulated structure. Furthermore, the weak additional reflections in the (hkn) (n = ±1, ±2,…) plane correspond to the linear combinations of q1 and q2, which cannot be explained by twinning and are, thus, evidencing a true [3 + 2]-dimensional modulated structure.

[Figure 2]
Figure 2
Projection of the relative position of the main reflections and first-order satellites along the [001] (left) and [010] (right) directions. The dotted line indicates the deviation of the satellite positions from the [110] mirror plane in Laue class 4/mmm.

3.3. Average crystal structure

Single-crystal data collected at ambient temperature indicated the same lattice parameters for a and b within standard deviations so that a tetra­gonal unit cell of a = 4.4996 (5) and c = 9.179 (1) Å was chosen. The main reflections are clearly compatible with high tetra­gonal Laue symmetry and the space group P4/nmm (No. 129) was chosen for structure refinement of the average structure according to the reflection condition h+k = 2n. The refinement resulted in a reasonable structural model (Table S1 in the supporting information), with a reduced occupancy factor of the Te site in the Te layer of about 81.1 (4)%. The composition derived from semiqu­anti­tative EDS (energy-dispersive X-ray spectroscopy) measurements points towards a composition of LaTe1.79 (1). Throughout the article, we will refer to this compound as LaTe1.82(1), based on the refined composition of the modulated structure.

The average structure of LaTe1.82(1) can be described with puckered [LaTe] layers sandwiched by square-planar [Te] layers. The partial occupation of the Te position in the [Te] layer, together with its large oblate anisotropic displacement parameters (ADPs) in the ab plane and prolate ADPs of the La atom along the [001] direction already give hints towards the modulation (Fig. 3[link]).

[Figure 3]
Figure 3
Average crystal structure of LaTe1.82(1). Displacement ellipsoids are drawn with a probability level of 99.9%.

The La atoms are regularly surrounded by five Te atoms of the puckered layer [4 × 3.3006 (4) and 1 × 3.324 (1) Å] and four Te atoms of the [Te] layer [4 × 3.3563 (7) Å], forming a regular tricapped trigonal prism. The planar [Te] layer is a perfect square net of Te atoms, with a Te—Te distance of 3.1821 (4) Å, which is significantly longer than a single Te—Te bond with about 2.80 Å [e.g. 2.78 Å in Rb2Te2, 2.86 Å in α-K2Te2 or 2.790 (1) Å in β-K2Te2; Böttcher et al., 1993[Böttcher, P., Getzschmann, J. & Keller, R. (1993). Z. Anorg. Allg. Chem. 619, 476-488.]]. The large ADPs indicate a considerably reduced electron density and a positional shift of the Te atoms due to the formation of anionic Te units. Similar observations have been made for the incommensurable modulated selenides RESe1.85 (RE = La–Nd or Sm) (Doert, Graf, Schmidt et al., 2007[Doert, T., Graf, C., Schmidt, P., Vasilieva, I. G., Simon, P. & Carrillo-Cabrera, W. (2007). J. Solid State Chem. 180, 496-509.]; Graf & Doert, 2009[Graf, C. & Doert, T. (2009). Z. Kristallogr. Cryst. Mater. 224, 568-579.]).

3.4. Refinement of the modulated structure

The tetra­gonal symmetry of the average structure discussed above is violated by the modulation vectors, as mentioned before. The observed satellite positions are incompatible with a fourfold rotational axis (Fig. 2[link]), resulting in a lower symmetric Laue class. To establish a suitable basic structure as starting model, ortho­rhom­bic and monoclinic subgroups of the space groups of the average structure were considered. The highest possible ortho­rhom­bic space group would then be Pmmn, which is a translationengleiche subgroup of the index 2 (t2) of P4/nmm, as displayed in Fig. 4[link]. However, a very similar basic structure has been used for DySe1.84, with similar modulation wave vectors q1 = αa* + βb* + [1 \over 2]c* and q2 = αa*βb* + [1 \over 2]c*, with α = 0.333 and β = 0.273. The structure model obtained in superspace group Pmmn(α,β,[1 \over 2])000(α,−β,[1 \over 2])000 [No. 59.2.51.39 according to Stokes et al. (2011[Stokes, H. T., Campbell, B. J. & van Smaalen, S. (2011). Acta Cryst. A67, 45-55.])] contained linear Se34− units besides Se22− and Se2− anions. The existence of linear Se34− fragments, however, was ex­cluded due to ener­getic consideration based on the μ2-Hückel method (van der Lee et al., 1997[Lee, A. van der, Hoistad, L. M., Evain, M., Foran, B. J. & Lee, S. (1997). Chem. Mater. 9, 218-226.]) and the final model for DySe1.84 was established in the noncentrosymmetric space group Pm21n(α,β,[1 \over 2])000(α,−β,[1 \over 2])000 (No. 31.2.51.35). We there­fore decided to establish a second structure model for LaTe1.82(1) in Pm21n(α,β,[1 \over 2])000(−α,β,[1 \over 2])000, too, and evaluate it against the centrosymmetric one in Pmmn(α,β,[1 \over 2])000(−α,β,[1 \over 2])000. The group–subgroup relationship between these two space groups of the basic structures is displayed in Fig. 4[link], indicating a t2 group–subgroup relationship between them. The limits of both models shall be discussed in the following and the crystallographic details of the refinements are given in Table 1[link].

[Figure 4]
Figure 4
Group–subgroup relationship between the space group of the aristotpye ZrSSi (P4/nmm) and the chosen ortho­rhom­bic space groups of the basic structures.

The main reflections meet the conditions for a primitive tetra­gonal lattice nearly perfectly, the data derived by powder diffraction and single-crystal diffraction give no hint of an ortho­rhom­bic or even monoclinic distortion of the lattice within standard uncertainties. However, taking the symmetry of the satellite reflections into account, the final unit-cell parameters were restrained to the conditions of an ortho­rhom­bic unit cell and were used for the integration of the intensities, as well as for structure refinements. According to the two modulation vectors and reflection conditions, the proper superspace group is Pmmn(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 (No. 59.2.51.39). The modulated structure clearly contradicts twinning: the result of a fourfold rotational axis as twin element would result in four additional satellite reflections around the position of the main reflections in layers hkl, l = ±0.5, ±1.5,… (Fig. S2 in the supporting information). The reconstructed precession images, however, reveal only the four expected satellite reflections corresponding to ±q1 and ±q2.

A second model in superspace group Pm21n(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 (No. 31.2.51.35; we keep the nonstandard setting for a concise structure description) has been evaluated against the centrosymmetric model to check if there are also differences in the structural model, as in the case of DySe1.84. For simplicity, we will use the three-dimensional space-group symbols during the structure descriptions to distinguish between the two different modulated structures in the following.

The refinement in Pm21n has been adjusted by considering inversion twinning. The refinement converged with a twin volume fraction of about 40% for the second domain.

The atomic modulation functions (amf) were stepwise included in both refinements, by first modelling the positional displacement of all atoms, before including an additional occupational modulation for the Te2 atom. As already expected from the average structure, the La1 and Te1 position show mainly shifts in the c direction, whereas the Te2 atom in the [Te] layer shows a strong displacement in the ab plane, as displayed in the two t plots in Fig. 5[link]. Note, that there is a small difference for the positional modulation along [001] for the Te2 atom between the models in Pmmn and Pm21n, as a result of the higher degree of freedom in the noncentrosymmetric space group. In Pm21n, the t plot shows a slightly sinusoidal curve in the c direction with a very small amplitude.

[Figure 5]
Figure 5
t-plots (a) of the displacement of La1 and Te1 in the [LaTe] layer and (b) of Te2 in the [Te] layer.

The displacement in the [LaTe] layer along c can be explained as a reaction to the modulation in the [Te] layer; the La atom aims to compensate the missing Te atoms in the coordination sphere by getting closer to the [Te] layer. Consequently, the Te1 atom reacts accordingly to the La1 dislocation by adjusting its position along c as well. The displacement of the Te2 atom in the [Te] layer is slightly more pronounced (Fig. 5[link]), due to vacancy formation and the creation of different Te anions. This holds for both models, as mentioned before. As a second step in the refinement, the occupational modulation in the [Te] layer was introduced by adding two harmonic functions. This improved the structural model in Pmmn and Pm21n considerably and the areas of low electron density at certain points in the Fourier map around Te2 are now also covered by the atomic modulation function (Fig. 6[link]).

[Figure 6]
Figure 6
Fourier sections around Te2. The electron density is drawn in contour line steps of 30 e Å−3. The central thick black line indicates the refined modulation function.

The Fourier sections in Fig. 6[link] also reveal a partially discontinuous behaviour of the electron density around Te2, although it is not very pronounced (see also Fig. S3 in the supporting information for a two-dimensional plot of the Te occupancy). The drawback in modelling this with harmonic functions are some overshooting and truncation effects in the final structure model, which we assume to be one major reason for the large residual electron-density maxima (see Table 1[link]). The inter­atomic distances are, nevertheless, in good agreement with previously reported distances for Te anions and the refined composition matches that of the semiqu­anti­tative EDS analysis. In the second evaluated model in Pm21n, harmonic functions, as well as crenel functions, have been utilized. The refinement with harmonic waves in Pm21n converged with slightly better R values and a similar residual electron density compared to the refinement in Pmmn, mainly due to the greater number of independently refined parameters. The use of discontinuous functions, such as crenel or sawtooth functions (Petříček et al., 2016[Petříček, V., Eigner, V., Dušek, M. & Čejchan, A. (2016). Z. Kristallogr. Cryst. Mater. 231, 301-312.]), failed in Pmmn but refined stably in Pm21n, although they did not improve the structural model. The comparison between both types of functions suggests that treating the occupational and positional modulation of Te2 by harmonic functions is suitable. The large residual peaks in the difference Fourier (FoFc) maps decrease considerably if two modulation functions are applied to the ADPs of Te2 as well. However, this leads to nonpositive-definite values for Umin at some values of t and has hence been rejected for the final structure model.

3.5. Discussion of the modulated crystal structure of LaTe1.82(1)

The structure model derived in Pmmn(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 for LaTe1.82(1) is used in the following paragraph for the discussion of the structural features as the structural differences between both models are negligible, as stated before.

The general feature of an alternating stacking of a puckered [LaTe] layer and [Te] layer is easily seen from Fig. S4 (see supporting information), which additionally shows the displacement of the La atoms along the c direction, as expected from the average structure. The motif of the puckered layer is very stable and does not show large deviations between different REX2–δ compounds, whereas the planar [X] layer is the more inter­esting structural feature and will be discussed in the following.

The change of the occupation and the variations of the Te—Te distances for the model in Pmmn are shown in Fig. 7[link]. The t-plot for the occupational modulation displays a static behaviour along t, which is shifted for different u values. The changes of the Te—Te distances in the modulated [Te] layer are displayed in the second t plot (Fig. 7[link]). Short distances with a lower limit of 2.801 (4) Å correspond to a Te—Te single bond (see above) and medium distances up to 2.959 (1) Å are in good agreement with distances reported for a linear Te34− anion (see, for example: Poddig et al., 2018[Poddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886-1896.]), as well as the often observed Te22− anions (see, for example: Stöwe, 2000a[Stöwe, K. (2000a). J. Solid State Chem. 149, 155-166.]). Larger distances of 2.984 (1) to 3.563 (4) Å are mainly considered as nonbonding inter­actions between Te entities. Compared to the known rare earth metal polysulfides and selenides, the inter­pretation of the Te—Te distances from a purely crystallographic viewpoint is more difficult as we face a much larger variety of distances, and boundaries between bonding and nonbonding inter­actions in the RETe2–δ system are floating. Reported Te—Te distances in dinuclear Te22− anions ranging from 2.868 (1) Å in GdTe1.8 (Poddig et al., 2018[Poddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886-1896.]) to 3.036 (2) Å in LaTe2 (Stöwe, 2000a[Stöwe, K. (2000a). J. Solid State Chem. 149, 155-166.]) have been inter­preted as single bonds. Nevertheless, the observed distances in the modulated structure of LaTe1.82(1) are in good agreement with the distances found in comparable com­pounds.

[Figure 7]
Figure 7
t-plots of (a) the occupation and (b) the distances shown for LaTe1.82(1).

A section of the modulated [Te] layer is depicted in Fig. 8[link], with Te positions displayed for a refined occupancy factor of 0.5 and greater. Solid lines indicate distances from 2.801 (4) up to 2.981 (2) Å, which should illustrate probable bonding situations; dashed lines are drawn up to 3.282 (2) Å for a better visualization of possible larger fragments, which have been reported for other polychalcogenides. Considering only these distances, three different motifs can be distinguished: a Te8 unit, probably consisting of smaller anions, like Te2−, Te22−, bent Te32− and linear Te34− anions, as well as isolated Te2− anions surrounded by different anionic Te entities. Eight-membered rings of chalcogen atoms and isolated X2− anions are, as already pointed out, common motifs in the crystal structures of the rare earth metal sulfides and selenides with compositions of RES1.9 (RE = La–Nd, Sm, Gd–Tm) (Doert, Graf, Lauxmann et al., 2007[Doert, T., Graf, C., Lauxmann, P. & Schleid, T. (2007). Z. Anorg. Allg. Chem. 633, 2719-2724.]; Tamazyan et al., 2000[Tamazyan, R., Arnold, H., Molchanov, V., Kuzmicheva, G. & Vasileva, I. (2000). Z. Kristallogr. Cryst. Mater. 215, 272-277.]; Müller et al., 2012[Müller, C. J., Schwarz, U. & Doert, T. (2012). Z. Anorg. Allg. Chem. 638, 2477-2484.]), RESe1.9 (RE = La–Nd, Sm, Gd–Tm) (Grupe & Urland, 1991[Grupe, M. & Urland, W. (1991). J. Less-Common Met. 170, 271-275.]; Plambeck-Fischer et al., 1989[Plambeck-Fischer, P., Abriel, W. & Urland, W. (1989). J. Solid State Chem. 78, 164-169.]; Urland et al., 1989[Urland, W., Plambeck-Fischer, P. & Grupe, M. (1989). Z. Naturforsch. Teil B, 44, 261-264.]; Müller et al., 2012[Müller, C. J., Schwarz, U. & Doert, T. (2012). Z. Anorg. Allg. Chem. 638, 2477-2484.]), RE8S15–δ (RE = Y, Tb–Ho) (Doert et al., 2012[Doert, T., Graf, C., Vasilyeva, I. G. & Schnelle, W. (2012). Inorg. Chem. 51, 282-289.]), RE8Se15–δ (RE = Y, Gd–Er; δ = 0 ≤ δ ≤ 0.3) (Doert, Dashjav et al., 2007[Doert, T., Dashjav, E. & Fokwa, B. P. T. (2007). Z. Anorg. Allg. Chem. 633, 261-273.]) and RESe1.85 (RE = La–Nd or Sm) (Doert, Graf, Schmidt et al., 2007[Doert, T., Graf, C., Schmidt, P., Vasilieva, I. G., Simon, P. & Carrillo-Cabrera, W. (2007). J. Solid State Chem. 180, 496-509.]; Graf & Doert, 2009[Graf, C. & Doert, T. (2009). Z. Kristallogr. Cryst. Mater. 224, 568-579.]).

[Figure 8]
Figure 8
Section of the modulated Te layer, with a cut-off occupancy at 0.5. Black-framed squares emphasize the voids in the layer. Solid lines are drawn from 2.801 to 2.981 Å and dashed lines are drawn for distances between 2.984 to 3.282 Å.

In the modulated [Te] layer of LaTe1.82(1), a Te4 square is apparent additionally when choosing small cut-off values for the approximant crystal structure for visualization (see, for example: Fig. S5 in the supporting information). As all four atoms in these fragments have an occupation value of about 0.52 (5), the presence of all four at the same time seems unrealistic. Instead, an unresolved superposition of a dinu­clear Te22− anion with two adjacent vacancies is the most likely explanation. Moreover, there is no evidence for anionic [Te4] squares with Te—Te distances of 2.80 Å in the literature. Cationic Te42+ and the corresponding Se42+ entities, on the other hand, are well known (see, for example: Barr et al., 1968[Barr, J., Gillespie, R. J., Kapoor, R. & Pez, G. P. (1968). J. Am. Chem. Soc. 90, 6855-6856.]; Beck et al., 1997[Beck, J., Kasper, M. & Stankowski, A. (1997). Chem. Ber. 130, 1189-1192.]; Forge & Beck, 2018[Forge, F. & Beck, J. (2018). Z. Anorg. Allg. Chem. 644, 1525-1531.]; Ruck & Locherer, 2015[Ruck, M. & Locherer, F. (2015). Coord. Chem. Rev. 285, 1-10.]) and their bonding situation has been investigated by computational methods in 1980 already (Rothman et al., 1980[Rothman, M. J., Bartell, L. S., Ewig, C. S. & Wazer, J. R. V. (1980). J. Comput. Chem. 1, 64-68.]). The typical Te—Te distance in Te42+ is about 2.65 to 2.70 Å (Ruck & Locherer, 2015[Ruck, M. & Locherer, F. (2015). Coord. Chem. Rev. 285, 1-10.]) and the Te—Te—Te angles are often close to 90°, which results in a slight distortion from idealized D4h symmetry. Taking the EDS results and the site-occupation factor from the average structure (both resulting in a composition of about LaTe1.8) into account, a substantial number of voids in the [Te] layer is expected, which also supports the idea of a disordered motif of dinuclear Te22− anions and adjacent vacancies instead.

A similar discussion of apparent structure motifs and possible (super)positions due to unresolved disorder shall be deduced for the apparent Te8 rings (Fig. 8[link]). Arrangements of disordered X22− anions have been identified as the constituents of eight-membered ring-like motifs in the sulfides RE8S15–δ (RE = Y, Tb–Ho; Doert et al., 2012[Doert, T., Graf, C., Vasilyeva, I. G. & Schnelle, W. (2012). Inorg. Chem. 51, 282-289.]) and selenides RE8Se15–δ (RE = Y, Gd–Er; δ = 0 ≤ δ ≤ 0.3; Doert, Dashjav et al., 2007[Doert, T., Dashjav, E. & Fokwa, B. P. T. (2007). Z. Anorg. Allg. Chem. 633, 261-273.]). In LaTe1.82(1), these apparent eight-membered rings may also consist of different disordered constituents, like Te32− and Te34− anions, along with the more common Te2− and Te22− motifs, around central vacancies. However, this is hard to resolve solely from the diffraction data. To gain more insight into this structural motif and the chemical bonding situation in the modulated [Te] layers in general, chemical bonding analyses were performed for three different com­mensurate approximants.

3.6. Bonding analysis

Quantum mechanical calculations based on density functional theory (DFT) and bonding analyses with the electron localizability indicator (ELI-D) (Kohout, 2004[Kohout, M. (2004). Int. J. Quantum Chem. 97, 651-658.], 2006[Kohout, M. (2006). Faraday Discuss. 135, 43-54.]; Pendás et al., 2012[Pendás, A. M., Kohout, M., Blanco, M. A. & Francisco, E. (2012). Modern Charge-Density Analysis, pp. 303-358. Dordrecht, Heidelberg, London, New York: Springer.]) have been performed for three approximant structures of LaTe1.82(1) in order to provide additional information on the bonding situation in the Te layers, especially regarding the (presumably disordered) Te4 and Te8 entities. As these calculations require three-dimensional commensurate structure models as bases, a suitable commensurate ortho­rhom­bic B-centred 4 × 3 × 2 supercell of the basic ZrSSi-type structure was chosen by approximating the q vector components α and β by [1 \over 4] and [1 \over 3], respectively. The respective three-dimensional space groups and the atomic positions were derived by the JANA2006 software package (Petříček et al., 2014[Petříček, V., Dušek, M. & Palatinus, L. (2014). Z. Kristallogr. Cryst. Mater. 229, 345-352.]) by enabling the commensurate option, after the final refinement of the modulated structure. According to the previously reported structures of RESe1.875–δ compounds (Doert, Dashjav et al., 2007[Doert, T., Dashjav, E. & Fokwa, B. P. T. (2007). Z. Anorg. Allg. Chem. 633, 261-273.]; Stöwe, 2001[Stöwe, K. (2001). Z. Kristallogr. Cryst. Mater. 216, 215-224.]), this cell was transformed into an A-centred setting for a better comparison, resulting in a 3 × 4 × 2 supercell with unit-cell dimensions of a = 13.4859 (4), b = 17.9812 (4) and c = 18.3446 (8) Å. As the highest possible symmetry, space group Amm2 (No. 38), the space group of the RE8Se15–δ compounds, was chosen for one approximant. The respective Amm2 structure model exhibits bent Te3 units in the Te8 ring, as well as a Te4 square, which cannot be resolved due to the symmetry restrictions in this space group (Fig. S6 in the supporting information). A second model has been established in the space group A2 (No. 5), i.e. removing the two perpendicular mirror planes. This space group has also been used to describe the disorder in the structures of the compounds RE8S15–δ. Here, only Te22− anions with alternating short (bonding) and longer (nonbonding) distances are considered as building units of the Te8 rings (Fig. S7 in the sup­porting information), enabling a direct comparison between the bent X3 fragments in Amm2, and the X22− patterns known from different REX1.9 and RE8X15–δ structures (cf. above). A third model in the space group P1 (No. 1) was developed starting from the previous model in A2 to lift the symmetry restrictions completely. The disorder of the apparent Te4 unit can then be resolved by assuming two vacancies and a single Te22− anion (Fig. S8 in the supporting information). Bear in mind that energetic comparisons are only possible between models with the same number of atoms. This is the case for the models in Amm2 and A2, but not for P1, due to the additional vacancies when taking the occupational disorder of the Te4 square into account. This means that an identification of the favoured structure is not possible based on the computed net energies only.

The calculated band gaps for all models are finite, but small, e.g. 0.04 eV for the model in A2, so that semiconducting electronic properties are expected (see below). The corresponding stoichiometric LaTe2 was reported as metallic (Stöwe, 2000a[Stöwe, K. (2000a). J. Solid State Chem. 149, 155-166.]).

Regarding the large Te8 entities, the (disordered) structure model in Amm2 shows a lower energy than the corresponding (ordered) model in A2 (ΔE = 0.16 eV). The Amm2 structure would imply the existence of bent Te3 entities with bond angles of about 90.0 (1)° in the planar Te layer, as mentioned before. Angular Te32− anions are well known, e.g. from the binary dialkali metal tritellurides A2Te3 with A = K, Rb or Cs (Eisenmann & Schäfer, 1978[Eisenmann, B. & Schäfer, H. (1978). Angew. Chem. 90, 731.]; Böttcher, 1980[Böttcher, P. (1980). J. Less-Common Met. 70, 263-271.]) and have Te—Te distances of about 2.80 Å, but significantly larger bond angles of about 100°. These Te32− anions are, however, more or less isolated in the structures and no inter­actions amongst them or with other anionic fragments are expected. Bent Te3 entities with Te—Te—Te angles close to 90° were described as parts of the anionic substructure in disordered polytellurides like KRE3Te8 (Stöwe et al., 2003[Stöwe, K., Napoli, C. & Appel, S. (2003). Z. Anorg. Allg. Chem. 629, 1925-1928.]; Patschke et al., 1998[Patschke, R., Heising, J., Schindler, J., Kannewurf, C. R. & Kanatzidis, M. (1998). J. Solid State Chem. 135, 111-115.]) and RbUSb0.33Te6 (Choi & Kanatzidis, 2001[Choi, K.-S. & Kanatzidis, M. G. (2001). J. Solid State Chem. 161, 17-22.]), and in the modulated structures of K1/3Ba2/3AgTe2 (Gourdon et al., 2000[Gourdon, O., Hanko, J., Boucher, F., Petricek, V., Whangbo, M.-H., Kanatzidis, M. G. & Evain, M. (2000). Inorg. Chem. 39, 1398-1409.]), LnTe3 (Malliakas et al., 2005[Malliakas, C., Billinge, S. J. L., Kim, H. J. & Kanatzidis, M. G. (2005). J. Am. Chem. Soc. 127, 6510-6511.]) and RESeTe2 (Fokwa et al., 2002[Fokwa, B. P. T., Doert, T., Simon, P., Söhnel, T. & Böttcher, P. (2002). Z. Anorg. Allg. Chem. 628, 2612-2616.]; Fokwa Tsinde & Doert, 2005[Fokwa Tsinde, B. P. & Doert, T. (2005). Solid State Sci. 7, 573-587.]), for example.

Orthoslices of the ELI-D within the [Te] layers of the Amm2 and the P1 approximant are shown in Fig. 9[link]. The isolines of both ELI-D images within the [Te] plane discriminates most of the observed Te atoms in three groups: isolated Te2−, dinuclear Te22− and bent trinuclear Te32− anions. The ELI-D in the P1 model suggest a slightly more pronounced tendency to form Te22− dumbbells as main polynuclear building units, in accordance with the reported structures of rare earth metal sulfides and selenides (Doert & Müller, 2016[Doert, T. & Müller, C. J. (2016). Binary Polysulfides and Polyselenides of Trivalent Rare-Earth Metals, in Reference Module in Chemistry, Molecular Sciences and Chemical Engineering, edited by J. Reedijk. Waltham, MA: Elsevier.]) and with theoretical considerations for the rare earth metal selenides (Lee & Foran, 1994[Lee, S. & Foran, B. (1994). J. Am. Chem. Soc. 116, 154-161.]). As discussed above, these entities are expected to represent the dominant bonding inter­actions in the planar layer, but the local bonding situation and the stability of the corresponding fragment are also influenced by inter­actions with other telluride anions in the [Te] layers, as well as by the surrounding La atoms in the layers below and above. Indeed, substantial inter­actions between these small anionic fragments in the [Te] layer have to be considered based on relatively high isovalues of the ELI-D between the dominating species in all models (Fig. 9[link]). For nonbonding or anti­bonding inter­actions, deep valleys (depicted in blue) would be expected, like, for example, the dark-blue regions between strongly localized lone-pair regions in P1.

[Figure 9]
Figure 9
Orthoslices of ELI-D of the Te layer of LaTe1.82(1) with isocontour lines. (a) The largely unordered model in Amm2, with bent Te3 units in the Te8 entities and a Te4 square. (b) The essentially ordered model in P1, exhibiting linear Te22− units in the Te8 entities and with only one Te22− anion instead of a Te4 square. Both pictures hint towards the inter­actions between different Te fragments in the planar layer.

The ELI-D slices in Fig. 9[link] show some additional inter­esting features. Significant localized lone-pair regions are found for those Te atoms located directly adjacent to vacancies. This may be taken as evidence for the anionic character of the [Te] layers. The respective lone pairs are localized in the structural voids with no hint of bonding inter­actions between Te fragments encasing the voids. The additional vacancies of the P1 model (Fig. 9[link]b) seem to be used to accommodate the lone pairs of different anionic Te fragments, again supporting the ionic description of the [Te] layer and in accordance with the calculated band gap and the measured semiconducting properties of LaTe1.81 (2) (see below).

Note, that the evaluated approximant structures indicate compositions of about LaTe1.95 (Amm2 and A2) and LaTe1.875 (P1), i.e. a higher tellurium content as compared to the actual composition LaTe1.82(1). Thus, additional vacancies would be necessary to get a more realistic image of the Te substructure. The bonding features between different constituents should nevertheless be comparable.

3.7. Electrical resistance of LaTe1.82(1)

The temperature-dependent electrical resistance of LaTe1.82(1) has been recorded by a four-point measurement between 7 and 345 K. The observed temperature dependence of the resistance of LaTe1.82(1) is characteristic for a semiconductor (Fig. 10[link]). The band gap, Eg, can be estimated from the highest temperature ρ values using a fit of the form ρ = ρ0exp(Eg/2kBT), where kB is the Boltzmann constant and T is the absolute temperature. The estimated Eg value is 0.17 eV for LaTe1.82(1), which is slightly larger than the calculated value for the model in A2 (cf. above). However, comparable com­pounds like NdTe1.89 (1), GdTe1.8 and SmTe1.84 show similar band gaps of 0.14 (Stöwe, 2001[Stöwe, K. (2001). Z. Kristallogr. Cryst. Mater. 216, 215-224.]), 0.19 (Poddig et al., 2018[Poddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886-1896.]) and 0.04 eV (Park et al., 1998[Park, S.-M., Park, S.-J. & Kim, S.-J. (1998). J. Solid State Chem. 140, 300-306.]), respectively, in contrast to LaTe2, which was reported to be metallic (Stöwe, 2000a[Stöwe, K. (2000a). J. Solid State Chem. 149, 155-166.]).

[Figure 10]
Figure 10
Temperature-dependent electrical resistance of LaTe1.82(1) between 7 and 345 K, showing a semiconducting behaviour.

4. Conclusions

The modulated structure of the rare earth metal polytelluride LaTe1.82(1) has been solved and refined using the superspace approach. The diffraction pattern evidences that the tetra­gonal symmetry of the average structure is not preserved in the modulation. Two different models evaluated in superspace groups Pmmn(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 (No. 59.2.51.39) and Pm21n(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 (No. 31.2.51.35) show only slightly different results, suggesting that the highest possible superspace group Pmmn(α,β,[1 \over 2])000(−α,β,[1 \over 2])000 should describe the structure accordingly. In the regime of the different REX2–δ compounds, LaTe1.82(1) may be best described as a depleted REX1.9 or REX1.875 structure. For the latter two structure types, it is possible to accommodate the respective anionic vacancies structurally isolated, i.e. separated between different (poly)telluride anions and in a commensurate superstructure. In the title compound, 18% of the Te2 positions are unoccupied, which leads to two obvious consequences: a commensurate ordering of vacancies and remaining anions is not possible anymore, and a considerable number of adjacent anion defects occur. In other words, LaTe1.82(1) exhibits a higher propensity for missing Te22− dianions. This description deviates significantly from the structures described for RETe1.8 (RE = Sm, Gd–Dy; Ijjaali & Ibers, 2006[Ijjaali, I. & Ibers, J. A. (2006). J. Solid State Chem. 179, 3456-3460.]; Wu et al., 2002[Wu, Y., Doert, T. & Böttcher, P. (2002). Z. Anorg. Allg. Chem. 628, 2216.]; Gulay et al., 2007[Gulay, L. D., Stępień-Damm, J., Daszkiewicz, M. & Pietraszko, A. (2007). J. Alloys Compd. 427, 166-170.]; Poddig et al., 2018[Poddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886-1896.]). However, this different structure fits well with the overall trend for the rare earth metal tellurides RETe2–δ, where different structures have been observed for a similar composition, as pointed out for the stoichiometric RETe2 (RE = La, Ce or Pr) compounds. Quantum mechanical calculations based on DFT with subsequent ELI-D-based bonding analysis for the ionic Te layer reveal Te22− units as dominant species, however, with significant long-range inter­actions amongst them. Temperature-dependent resistance measurements suggest a semiconducting behaviour with a band gap of about 0.17 eV, which is in good agreement with comparable rare earth metal compounds.

Supporting information


Computing details top

For both structures, data collection: APEX2 (Bruker, 2010). Cell refinement: APEX2 (Bruker, 2010) for (I); SAINT-Plus (Bruker, 2017) for (II). For both structures, data reduction: SAINT-Plus (Bruker, 2017). Program(s) used to solve structure: SHELXT2014 (Sheldrick, 2015a) for (I); SUPERFLIP (Palatinus & Chapuis, 2007) for (II). Program(s) used to refine structure: SHELXL2019 (Sheldrick, 2015b) for (I); JANA2006 (Petricek et al., 2014) for (II). For both structures, molecular graphics: OLEX2 (Dolomanov et al., 2009) and DIAMOND (Brandenburg, 2019). Software used to prepare material for publication: OLEX2 (Dolomanov et al., 2009) for (I); JANA2006 (Petricek et al., 2014) for (II).

Lanthanum telluride (1/1.8) (I) top
Crystal data top
LaTe1.81Dx = 6.621 Mg m3
Mr = 370.50Mo Kα radiation, λ = 0.71073 Å
Tetragonal, P4/nmm:1Cell parameters from 872 reflections
a = 4.4996 (5) Åθ = 4.4–34.7°
c = 9.1794 (12) ŵ = 25.18 mm1
V = 185.85 (3) Å3T = 296 K
Z = 2Block, metallic black
F(000) = 3030.03 × 0.02 × 0.02 mm
Data collection top
Bruker APEXII CCD
diffractometer
289 reflections with I > 2σ(I)
Radiation source: sealed X-ray tubeRint = 0.046
profile data from CCD detector scansθmax = 38.5°, θmin = 2.2°
Absorption correction: multi-scan
(SADABS; Krause et al., 2015)
h = 77
Tmin = 0.641, Tmax = 0.748k = 77
3402 measured reflectionsl = 815
351 independent reflections
Refinement top
Refinement on F2Secondary atom site location: difference Fourier map
Least-squares matrix: full w = 1/[σ2(Fo2) + (0.0248P)2 + 0.9467P]
where P = (Fo2 + 2Fc2)/3
R[F2 > 2σ(F2)] = 0.024(Δ/σ)max < 0.001
wR(F2) = 0.057Δρmax = 3.15 e Å3
S = 1.07Δρmin = 2.35 e Å3
351 reflectionsExtinction correction: SHELXL2019 (Sheldrick, 2015b), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
11 parametersExtinction coefficient: 0.0266 (18)
0 restraintsAbsolute structure: No quotients, so Flack parameter determined by classical intensity fit
Primary atom site location: dual
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Refinement. Single-crystal X-ray diffraction was performed with the four-circle diffractometer Kappa Apex2 (Bruker) equipped with a CCD-detector using graphite-monochromated Mo-Kα radiation (λ= 0.71073 Å) at 296 (1) K. Data for the average structure was corrected for Lorentz and polarization factors, and multi-scan absorption corrections was applied (Krause et al., 2015). The structure was solved using the dual space approach of the program package SHELXT (Sheldrick, 2015a). Structure refinement was performed with the program package SHELXL against F2 including anisotropic displacement parameters for all atoms (Sheldrick, 2015b).

Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3–10.

Sheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.

Sheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
La10.0000000.5000000.27121 (7)0.01498 (15)
Te10.0000000.5000000.63329 (7)0.01164 (15)
Te20.0000000.0000000.0000000.0408 (4)0.813 (4)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
La10.00971 (16)0.00971 (16)0.0255 (3)0.0000.0000.000
Te10.00909 (17)0.00909 (17)0.0167 (3)0.0000.0000.000
Te20.0568 (6)0.0568 (6)0.0088 (4)0.0000.0000.000
Geometric parameters (Å, º) top
La1—Te1i3.3002 (3)La1—Te2vi3.3555 (6)
La1—Te1ii3.3002 (3)La1—Te2vii3.3555 (5)
La1—Te1iii3.3002 (3)Te2—Te2vi3.1817 (2)
La1—Te1iv3.3002 (3)Te2—Te2viii3.1817 (2)
La1—Te13.3237 (10)Te2—Te2v3.1817 (2)
La1—Te2v3.3555 (6)Te2—Te2ix3.1817 (2)
La1—Te23.3555 (5)
Te1i—La1—Te1ii149.19 (3)La1i—Te1—La1iii85.954 (10)
Te1i—La1—Te1iii85.954 (10)La1ii—Te1—La1iii85.954 (9)
Te1ii—La1—Te1iii85.954 (10)La1i—Te1—La1iv85.954 (9)
Te1i—La1—Te1iv85.954 (9)La1ii—Te1—La1iv85.954 (10)
Te1ii—La1—Te1iv85.954 (10)La1iii—Te1—La1iv149.19 (3)
Te1iii—La1—Te1iv149.19 (3)La1i—Te1—La1105.404 (15)
Te1i—La1—Te174.597 (15)La1ii—Te1—La1105.403 (15)
Te1ii—La1—Te174.597 (15)La1iii—Te1—La1105.404 (15)
Te1iii—La1—Te174.597 (15)La1iv—Te1—La1105.404 (15)
Te1iv—La1—Te174.597 (15)Te2vi—Te2—Te2viii180.0
Te1i—La1—Te2v130.854 (14)Te2vi—Te2—Te2v90.000 (6)
Te1ii—La1—Te2v74.930 (12)Te2viii—Te2—Te2v90.000 (6)
Te1iii—La1—Te2v74.930 (12)Te2vi—Te2—Te2ix90.000 (6)
Te1iv—La1—Te2v130.854 (14)Te2viii—Te2—Te2ix90.000 (6)
Te1—La1—Te2v137.896 (9)Te2v—Te2—Te2ix180.0
Te1i—La1—Te2130.854 (14)Te2vi—Te2—La1v118.301 (6)
Te1ii—La1—Te274.930 (11)Te2viii—Te2—La1v61.699 (6)
Te1iii—La1—Te2130.854 (14)Te2v—Te2—La1v61.699 (6)
Te1iv—La1—Te274.930 (11)Te2ix—Te2—La1v118.301 (6)
Te1—La1—Te2137.896 (8)Te2vi—Te2—La1x118.301 (5)
Te2v—La1—Te256.601 (10)Te2viii—Te2—La1x61.699 (5)
Te1i—La1—Te2vi74.930 (12)Te2v—Te2—La1x118.301 (5)
Te1ii—La1—Te2vi130.854 (14)Te2ix—Te2—La1x61.699 (5)
Te1iii—La1—Te2vi130.854 (14)La1v—Te2—La1x123.398 (10)
Te1iv—La1—Te2vi74.930 (12)Te2vi—Te2—La1vi61.699 (6)
Te1—La1—Te2vi137.896 (9)Te2viii—Te2—La1vi118.301 (6)
Te2v—La1—Te2vi84.208 (18)Te2v—Te2—La1vi118.301 (6)
Te2—La1—Te2vi56.601 (10)Te2ix—Te2—La1vi61.699 (6)
Te1i—La1—Te2vii74.930 (11)La1v—Te2—La1vi84.208 (17)
Te1ii—La1—Te2vii130.854 (14)La1x—Te2—La1vi123.398 (10)
Te1iii—La1—Te2vii74.930 (11)Te2vi—Te2—La161.699 (5)
Te1iv—La1—Te2vii130.854 (14)Te2viii—Te2—La1118.301 (5)
Te1—La1—Te2vii137.896 (8)Te2v—Te2—La161.699 (5)
Te2v—La1—Te2vii56.601 (10)Te2ix—Te2—La1118.301 (5)
Te2—La1—Te2vii84.208 (16)La1v—Te2—La1123.399 (10)
Te2vi—La1—Te2vii56.601 (10)La1x—Te2—La184.208 (16)
La1i—Te1—La1ii149.19 (3)La1vi—Te2—La1123.398 (10)
Symmetry codes: (i) x+1/2, y+3/2, z+1; (ii) x1/2, y+1/2, z+1; (iii) x1/2, y+3/2, z+1; (iv) x+1/2, y+1/2, z+1; (v) x1/2, y+1/2, z; (vi) x+1/2, y+1/2, z; (vii) x, y+1, z; (viii) x1/2, y1/2, z; (ix) x+1/2, y1/2, z; (x) x, y1, z.
(II) top
Crystal data top
LaTe1.811(4)F(000) = 303
Mr = 370Dx = 6.626 Mg m3
Orthorhombic, PmmnMo Kα radiation, λ = 0.71073 Å
q1 = 0.275080a* + 0.309805b* + 0.500000c*; q2 = -0.275080a* + 0.309805b* + 0.500000c*Cell parameters from 872 reflections
a = 4.4861 (5) Åθ = 4.4–34.7°
b = 4.5030 (5) ŵ = 25.20 mm1
c = 9.1806 (11) ÅT = 296 K
V = 185.46 (4) Å3Block, black
Z = 20.06 × 0.05 × 0.01 mm
† Symmetry operations: (1) x1, x2, x3, x4, x5; (2) −x1, −x2, x3, x3x4, x3x5; (3) −x1+1/2, x2+1/2, −x3, −x3+x5, −x3+x4; (4) x1+1/2, −x2+1/2, −x3, −x5, −x4; (5) −x1+1/2, −x2+1/2, −x3, −x4, −x5; (6) x1+1/2, x2+1/2, −x3, −x3+x4, −x3+x5; (7) x1, −x2, x3, x3x5, x3x4; (8) −x1, x2, x3, x5, x4.

Data collection top
Bruker APEXII CCD
diffractometer
738 reflections with I > 3σ(I)
Radiation source: sealed X-ray tubeRint = 0.156
profile data from CCD detector scansθmax = 38.4°, θmin = 2.2°
Absorption correction: gaussian
Jana2006
h = 78
Tmin = 0.393, Tmax = 0.498k = 88
30619 measured reflectionsl = 169
1757 independent reflections
Refinement top
Refinement on F21 constraint
R[F > 3σ(F)] = 0.050Weighting scheme based on measured s.u.'s w = 1/(σ2(I) + 0.0004I2)
wR(F) = 0.111(Δ/σ)max = 0.045
S = 1.26Δρmax = 15.11 e Å3
1757 reflectionsΔρmin = 17.25 e Å3
24 parametersExtinction correction: B-C type 1 Gaussian isotropic (Becker & Coppens, 1974)
0 restraintsExtinction coefficient: 1500 (80)
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
La100.50.27147 (8)0.01052 (18)
Te100.50.63322 (8)0.01043 (19)
Te20000.0204 (4)0.811 (4)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
La10.0099 (3)0.0090 (3)0.0127 (4)000
Te10.0096 (3)0.0095 (3)0.0121 (4)000
Te20.0258 (7)0.0266 (8)0.0088 (5)000
Geometric parameters (Å, º) top
AverageMinimumMaximum
La1—Te13.319 (2)3.234 (2)3.408 (2)
La1—Te1i3.296 (3)3.245 (3)3.353 (3)
La1—Te1ii3.297 (3)3.245 (3)3.354 (3)
La1—Te1iii3.296 (3)3.245 (3)3.353 (3)
La1—Te1iv3.297 (3)3.245 (3)3.354 (3)
La1—Te23.371 (2)3.253 (3)3.475 (3)
La1—Te2v3.363 (2)3.253 (3)3.475 (3)
La1—Te2vi3.361 (3)3.248 (4)3.463 (4)
La1—Te2vii3.361 (3)3.248 (4)3.463 (4)
Te2—Te2viii3.167 (3)2.799 (5)3.558 (5)
Te2—Te2vi3.167 (3)2.799 (5)3.558 (5)
Te2—Te2ix3.167 (3)2.799 (5)3.558 (5)
Te2—Te2vii3.167 (3)2.799 (5)3.558 (5)
Te1—La1—Te1i74.66 (5)71.62 (5)77.67 (5)
Te1—La1—Te1ii74.73 (5)71.63 (5)77.66 (5)
Te1—La1—Te1iii74.66 (5)71.62 (5)77.67 (5)
Te1—La1—Te1iv74.73 (5)71.63 (5)77.66 (5)
Te1—La1—Te2138.00 (7)131.71 (7)144.29 (8)
Te1—La1—Te2v138.09 (7)131.70 (7)144.25 (8)
Te1—La1—Te2vi138.12 (7)132.30 (6)143.81 (7)
Te1—La1—Te2vii138.12 (7)132.30 (6)143.81 (7)
Te1i—La1—Te1ii86.19 (5)84.52 (5)87.67 (5)
Te1i—La1—Te1iii85.78 (5)84.22 (5)87.09 (5)
Te1i—La1—Te1iv149.39 (6)143.50 (6)155.11 (7)
Te1i—La1—Te274.94 (7)72.45 (7)77.30 (7)
Te1i—La1—Te2v130.63 (8)125.30 (8)136.53 (8)
Te1i—La1—Te2vi75.07 (7)72.70 (7)77.50 (7)
Te1i—La1—Te2vii130.42 (8)125.13 (8)136.45 (8)
Te1ii—La1—Te1iii149.39 (6)143.50 (6)155.11 (7)
Te1ii—La1—Te1iv85.78 (5)84.22 (5)87.08 (5)
Te1ii—La1—Te2130.63 (8)125.33 (8)136.51 (8)
Te1ii—La1—Te2v74.95 (7)72.45 (7)77.29 (7)
Te1ii—La1—Te2vi75.16 (7)72.70 (7)77.51 (7)
Te1ii—La1—Te2vii130.54 (8)125.09 (8)136.47 (8)
Te1iii—La1—Te1iv86.19 (5)84.52 (5)87.67 (5)
Te1iii—La1—Te274.94 (7)72.45 (7)77.30 (7)
Te1iii—La1—Te2v130.63 (8)125.30 (8)136.53 (8)
Te1iii—La1—Te2vi130.42 (8)125.13 (8)136.45 (8)
Te1iii—La1—Te2vii75.07 (7)72.70 (7)77.50 (7)
Te1iv—La1—Te2130.63 (8)125.33 (8)136.51 (8)
Te1iv—La1—Te2v74.95 (7)72.45 (7)77.29 (7)
Te1iv—La1—Te2vi130.54 (8)125.09 (8)136.47 (8)
Te1iv—La1—Te2vii75.16 (7)72.70 (7)77.51 (7)
Te2—La1—Te2v83.88 (7)71.67 (7)96.54 (6)
Te2—La1—Te2vi56.18 (7)48.42 (7)64.86 (7)
Te2—La1—Te2vii56.18 (7)48.42 (7)64.86 (7)
Te2v—La1—Te2vi56.28 (7)48.45 (7)64.83 (7)
Te2v—La1—Te2vii56.28 (7)48.45 (7)64.83 (7)
Te2vi—La1—Te2vii83.69 (6)72.62 (6)95.33 (6)
La1—Te1—La1i105.41 (6)102.47 (5)108.25 (5)
La1—Te1—La1ii105.30 (6)102.48 (5)108.24 (5)
La1—Te1—La1iii105.41 (6)102.47 (5)108.25 (5)
La1—Te1—La1iv105.30 (6)102.48 (5)108.24 (5)
La1i—Te1—La1ii86.21 (5)85.55 (5)86.71 (5)
La1i—Te1—La1iii85.79 (5)85.18 (5)86.27 (5)
La1i—Te1—La1iv149.29 (6)144.20 (7)154.25 (7)
La1ii—Te1—La1iii149.29 (6)144.20 (7)154.25 (7)
La1ii—Te1—La1iv85.80 (5)85.18 (5)86.27 (5)
La1iii—Te1—La1iv86.21 (5)85.55 (5)86.71 (5)
La1x—Te2—La183.87 (5)80.45 (7)88.10 (7)
La1x—Te2—La1viii123.37 (7)114.57 (9)132.12 (6)
La1x—Te2—La1ix123.37 (7)114.57 (9)132.12 (6)
La1x—Te2—Te2viii61.76 (6)57.09 (5)66.19 (6)
La1x—Te2—Te2vi118.22 (7)113.84 (6)122.96 (7)
La1x—Te2—Te2ix61.76 (6)57.09 (5)66.19 (6)
La1x—Te2—Te2vii118.22 (7)113.84 (6)122.96 (7)
La1—Te2—La1viii123.37 (7)114.57 (9)132.12 (6)
La1—Te2—La1ix123.37 (7)114.57 (9)132.12 (6)
La1—Te2—Te2viii118.22 (7)113.84 (6)122.96 (7)
La1—Te2—Te2vi61.76 (6)57.09 (5)66.19 (6)
La1—Te2—Te2ix118.22 (7)113.84 (6)122.96 (7)
La1—Te2—Te2vii61.76 (6)57.09 (5)66.19 (6)
La1viii—Te2—La1ix83.64 (5)79.93 (7)88.22 (7)
La1viii—Te2—Te2viii61.92 (6)57.32 (6)66.24 (7)
La1viii—Te2—Te2vi61.92 (6)57.32 (6)66.24 (7)
La1viii—Te2—Te2ix118.09 (7)113.58 (9)122.75 (8)
La1viii—Te2—Te2vii118.09 (7)113.58 (9)122.75 (8)
La1ix—Te2—Te2viii118.09 (7)113.58 (9)122.75 (8)
La1ix—Te2—Te2vi118.09 (7)113.58 (9)122.75 (8)
La1ix—Te2—Te2ix61.92 (6)57.32 (6)66.24 (7)
La1ix—Te2—Te2vii61.92 (6)57.32 (6)66.24 (7)
Te2viii—Te2—Te2vi90.14 (9)88.89 (9)91.31 (7)
Te2viii—Te2—Te2ix89.86 (9)88.67 (7)91.14 (8)
Te2viii—Te2—Te2vii179.76 (9)179.54 (9)180
Te2vi—Te2—Te2ix179.76 (9)179.54 (9)180
Te2vi—Te2—Te2vii89.86 (9)88.67 (7)91.14 (8)
Te2ix—Te2—Te2vii90.14 (9)88.89 (9)91.31 (7)
Symmetry codes: (i) x11/2, x21/2, x3+1, x3+x5, x3+x4; (ii) x11/2, x2+1/2, x3+1, x3+x5, x3+x4; (iii) x1+1/2, x21/2, x3+1, x3+x5, x3+x4; (iv) x1+1/2, x2+1/2, x3+1, x3+x5, x3+x4; (v) x1, x2+1, x3, x4, x5; (vi) x11/2, x2+1/2, x3, x3+x5, x3+x4; (vii) x1+1/2, x2+1/2, x3, x3+x5, x3+x4; (viii) x11/2, x21/2, x3, x3+x5, x3+x4; (ix) x1+1/2, x21/2, x3, x3+x5, x3+x4; (x) x1, x21, x3, x4, x5.
Crystallographic data and refinement details for LaTe1.82 (1) top
Model PmmnModel Pm21n
Refined compositionLaTe1.811 (4)LaTe1.825 (3)
Formula weight (g mol-1); F(000)370.8; 303371.8; 304
Crystal size (mm-3)0.0588 × 0.0472 × 0.0094
Diffractometer, radiationBruker Kappa APEXII, Mo Kα (0.71073 Å)
Temperature296 (1) K
Lattice parameters (Å)a = 4.5020 (5), b = 4.4985 (5), c = 9.181 (1)
α = β = γ = 90°
Modulation vectorsq1 = αa* + βb* + γc*
q2 = -αa* + βb* + γc*
α = 0.272 (1), β = 0.314 (1), γ = 1/2
Index range measured-7h8; -8k8; -16l9; -1m,n1
2.18θ38.36
Measured reflections30619
Abs. coefficient µ (mm)25.201
Tmin, Tmax0.3929, 0.4983
Extinction parameter (Becker & Coppens, 1974)0.1510.153
Independet reflections1757, 738 > 3σ(I)6103, 1386 > 3σ(I)
Main reflection415, 323 > 3σ(I)1080, 719 > 3σ(I)
First-order satellites1342, 415 > 3σ(I)3946, 664 > 3σ(I)
Rint; Rsigma0.0556, 0.0461 for I > 3σ(I)0.1563, 0.1934 for all
0.0469, 0.0600 for I > 3σ(I)0.1435, 0.2810 for all
Superspace group, ZPmmn(α,β,1/2)000(-α,β,1/2)000 (No. 59.2.51.39), 2Pm21n(α,β,1/2)000(α,β,)000 (No. 31.2.51.35), 2
Refinement methodJANA2006, full-matrix against F2
Restrictions/parameters0/330/66
R1 [3σ(I)]R1 (all)wR2 [3σ(I)]wR2 (all)R1 [3σ(I)]R1 (all)wR2 [3σ(I)]wR2 (all)
All reflections0.04950.13000.08720.11110.05060.21310.08830.1287
Main reflections0.02470.03850.05090.05420.03260.05540.05750.0610
First-order satellites0.10790.24700.18910.24790.10000.33330.19490.2928
Goodness-of-fit (3σ(I)/all)1.53/1.261.30/0.90
Largest difference peak/hole (e Å-3)15.11/-17.2513.21/-14.87
Computer programs: APEX (Bruker, 2010), APEX2 (Bruker, 2010), SAINT-Plus (Bruker, 2017), SHELXT2014 (Sheldrick, 2015a), SUPERFLIP (Palatinus & Chapuis, 2007), SHELXL2019 (Sheldrick, 2015b), JANA2006 (Petricek & Dusek, 2014), OLEX2 (Dolomanov et al., 2009) and DIAMOND (Brandenburg, 2019).
 

Footnotes

1Dedicated to Professor Stephen Lee on the occasion of his 65th birthday.

Acknowledgements

KF acknowledges the Technische Universität Dresden for funding within the framework of a Maria Reiche fellowship.

Funding information

Funding for this research was provided by: Deutsche Forschungsgemeinschaft (grant No. Do 590/6 to TD).

References

First citationAssoud, A., Xu, J. & Kleinke, H. (2007). Inorg. Chem. 46, 9906–9911.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBarr, J., Gillespie, R. J., Kapoor, R. & Pez, G. P. (1968). J. Am. Chem. Soc. 90, 6855–6856.  CrossRef CAS Web of Science Google Scholar
First citationBeck, J., Kasper, M. & Stankowski, A. (1997). Chem. Ber. 130, 1189–1192.  CrossRef ICSD CAS Web of Science Google Scholar
First citationBecker, P. J. & Coppens, P. (1974). Acta Cryst. A30, 148–153.  CrossRef IUCr Journals Web of Science Google Scholar
First citationBlum, V., Gehrke, R., Hanke, F., Havu, P., Havu, V., Ren, X., Reuter, K. & Scheffler, M. (2009). Comput. Phys. Commun. 180, 2175–2196.  Web of Science CrossRef CAS Google Scholar
First citationBöttcher, P. (1980). J. Less-Common Met. 70, 263–271.  Google Scholar
First citationBöttcher, P., Getzschmann, J. & Keller, R. (1993). Z. Anorg. Allg. Chem. 619, 476–488.  Google Scholar
First citationBrandenburg, K. (2019). DIAMOND. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationBruker (2010). APEX2. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBruker (2017). SAINT-Plus. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationChoi, K.-S. & Kanatzidis, M. G. (2001). J. Solid State Chem. 161, 17–22.  Web of Science CrossRef ICSD CAS Google Scholar
First citationCoelho, A. A. (2018). J. Appl. Cryst. 51, 210–218.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationDashjav, E., Doert, T., Böttcher, P., Mattausch, H. & Oeckler, O. (2000). Z. Kristallogr. New Cryst. Struct. 215, 337–338.  CAS Google Scholar
First citationDoert, T., Dashjav, E. & Fokwa, B. P. T. (2007). Z. Anorg. Allg. Chem. 633, 261–273.  Web of Science CrossRef ICSD CAS Google Scholar
First citationDoert, T., Graf, C., Lauxmann, P. & Schleid, T. (2007). Z. Anorg. Allg. Chem. 633, 2719–2724.  Web of Science CrossRef ICSD CAS Google Scholar
First citationDoert, T., Graf, C., Schmidt, P., Vasilieva, I. G., Simon, P. & Carrillo-Cabrera, W. (2007). J. Solid State Chem. 180, 496–509.  Web of Science CrossRef CAS Google Scholar
First citationDoert, T., Graf, C., Vasilyeva, I. G. & Schnelle, W. (2012). Inorg. Chem. 51, 282–289.  Web of Science CrossRef ICSD CAS PubMed Google Scholar
First citationDoert, T. & Müller, C. J. (2016). Binary Polysulfides and Polyselenides of Trivalent Rare-Earth Metals, in Reference Module in Chemistry, Molecular Sciences and Chemical Engineering, edited by J. Reedijk. Waltham, MA: Elsevier.  Google Scholar
First citationDolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339–341.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEisenmann, B. & Schäfer, H. (1978). Angew. Chem. 90, 731.  CrossRef ICSD Google Scholar
First citationFokwa, B. P. T., Doert, T., Simon, P., Söhnel, T. & Böttcher, P. (2002). Z. Anorg. Allg. Chem. 628, 2612–2616.  CrossRef CAS Google Scholar
First citationFokwa Tsinde, B. P. & Doert, T. (2005). Solid State Sci. 7, 573–587.  Web of Science CrossRef ICSD CAS Google Scholar
First citationForge, F. & Beck, J. (2018). Z. Anorg. Allg. Chem. 644, 1525–1531.  Web of Science CrossRef ICSD CAS Google Scholar
First citationGourdon, O., Hanko, J., Boucher, F., Petricek, V., Whangbo, M.-H., Kanatzidis, M. G. & Evain, M. (2000). Inorg. Chem. 39, 1398–1409.  Web of Science CrossRef ICSD PubMed CAS Google Scholar
First citationGraf, C. & Doert, T. (2009). Z. Kristallogr. Cryst. Mater. 224, 568–579.  CAS Google Scholar
First citationGrupe, M. & Urland, W. (1991). J. Less-Common Met. 170, 271–275.  CrossRef ICSD CAS Web of Science Google Scholar
First citationGulay, L. D., Stępień-Damm, J., Daszkiewicz, M. & Pietraszko, A. (2007). J. Alloys Compd. 427, 166–170.  Web of Science CrossRef ICSD CAS Google Scholar
First citationIjjaali, I. & Ibers, J. A. (2006). J. Solid State Chem. 179, 3456–3460.  Web of Science CrossRef ICSD CAS Google Scholar
First citationKlein Haneveld, A. & Jellinek, F. (1964). Recl. Trav. Chim. Pays-Bas, 83, 776–783.  CrossRef ICSD CAS Google Scholar
First citationKohout, M. (2004). Int. J. Quantum Chem. 97, 651–658.  Web of Science CrossRef CAS Google Scholar
First citationKohout, M. (2006). Faraday Discuss. 135, 43–54.  Web of Science CrossRef Google Scholar
First citationKohout, M. (2016). DGrid. Version 5.0. https://www2.cpfs.mpg.de/kohout/dgrid.htmlGoogle Scholar
First citationKrause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3–10.  Web of Science CSD CrossRef ICSD CAS IUCr Journals Google Scholar
First citationLee, S. & Foran, B. (1994). J. Am. Chem. Soc. 116, 154–161.  CrossRef CAS Web of Science Google Scholar
First citationLee, A. van der, Hoistad, L. M., Evain, M., Foran, B. J. & Lee, S. (1997). Chem. Mater. 9, 218–226.  CrossRef Web of Science Google Scholar
First citationMalliakas, C., Billinge, S. J. L., Kim, H. J. & Kanatzidis, M. G. (2005). J. Am. Chem. Soc. 127, 6510–6511.  Web of Science CrossRef ICSD PubMed CAS Google Scholar
First citationMüller, C. J., Schwarz, U. & Doert, T. (2012). Z. Anorg. Allg. Chem. 638, 2477–2484.  Google Scholar
First citationOnken, H., Vierheilig, K. & Hahn, H. (1964). Z. Anorg. Allg. Chem. 333, 267–279.  CrossRef ICSD CAS Web of Science Google Scholar
First citationPalatinus, L. & Chapuis, G. (2007). J. Appl. Cryst. 40, 786–790.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPark, S.-M., Park, S.-J. & Kim, S.-J. (1998). J. Solid State Chem. 140, 300–306.  Web of Science CrossRef ICSD CAS Google Scholar
First citationPatschke, R., Heising, J., Schindler, J., Kannewurf, C. R. & Kanatzidis, M. (1998). J. Solid State Chem. 135, 111–115.  Web of Science CrossRef ICSD CAS Google Scholar
First citationPendás, A. M., Kohout, M., Blanco, M. A. & Francisco, E. (2012). Modern Charge-Density Analysis, pp. 303–358. Dordrecht, Heidelberg, London, New York: Springer.  Google Scholar
First citationPerdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865–3868.  CrossRef PubMed CAS Web of Science Google Scholar
First citationPetříček, V., Dušek, M. & Palatinus, L. (2014). Z. Kristallogr. Cryst. Mater. 229, 345–352.  Google Scholar
First citationPetříček, V., Eigner, V., Dušek, M. & Čejchan, A. (2016). Z. Kristallogr. Cryst. Mater. 231, 301–312.  Google Scholar
First citationPlambeck-Fischer, P., Abriel, W. & Urland, W. (1989). J. Solid State Chem. 78, 164–169.  CAS Google Scholar
First citationPoddig, H., Donath, T., Gebauer, P., Finzel, K., Kohout, M., Wu, Y., Schmidt, P. & Doert, T. (2018). Z. Anorg. Allg. Chem. 644, 1886–1896.  Web of Science CrossRef ICSD CAS Google Scholar
First citationRothman, M. J., Bartell, L. S., Ewig, C. S. & Wazer, J. R. V. (1980). J. Comput. Chem. 1, 64–68.  CrossRef CAS Web of Science Google Scholar
First citationRuck, M. & Locherer, F. (2015). Coord. Chem. Rev. 285, 1–10.  Web of Science CrossRef CAS Google Scholar
First citationRundle, R. E. (1963). J. Am. Chem. Soc. 85, 112–113.  CrossRef CAS Web of Science Google Scholar
First citationSheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationStokes, H. T., Campbell, B. J. & van Smaalen, S. (2011). Acta Cryst. A67, 45–55.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationStöwe, K. (2000a). J. Solid State Chem. 149, 155–166.  Google Scholar
First citationStöwe, K. (2000b). J. Alloys Compd. 307, 101–110.  Google Scholar
First citationStöwe, K. (2000c). Z. Anorg. Allg. Chem. 626, 803–811.  Google Scholar
First citationStöwe, K. (2001). Z. Kristallogr. Cryst. Mater. 216, 215–224.  Google Scholar
First citationStöwe, K., Napoli, C. & Appel, S. (2003). Z. Anorg. Allg. Chem. 629, 1925–1928.  Web of Science CrossRef ICSD CAS Google Scholar
First citationTamazyan, R., Arnold, H., Molchanov, V., Kuzmicheva, G. & Vasileva, I. (2000). Z. Kristallogr. Cryst. Mater. 215, 272–277.  CAS Google Scholar
First citationUrland, W., Plambeck-Fischer, P. & Grupe, M. (1989). Z. Naturforsch. Teil B, 44, 261–264.  CrossRef ICSD CAS Google Scholar
First citationWu, Y., Doert, T. & Böttcher, P. (2002). Z. Anorg. Allg. Chem. 628, 2216.  Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296
Follow Acta Cryst. C
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds