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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Single crystals of SnTe3O8 in the millimetre range grown by chemical vapor transport reactions

crossmark logo

aInstitute for Chemical Technologies and Analytics, Division of Structural Chemistry, TU Wien, Getreidemarkt 9/164-SC, A-1060 Vienna, Austria
*Correspondence e-mail: matthias.weil@tuwien.ac.at

Edited by W. T. A. Harrison, University of Aberdeen, Scotland (Received 5 November 2021; accepted 8 November 2021; online 12 November 2021)

Tin(IV) trioxidotellurate(IV), SnTe3O8, is a member of the isotypic MIVTeIV3O8 (M = Ti, Zr, Hf, Sn) series crystallizing with eight formula units per unit cell in space group Ia[\overline{3}]. In comparison with the previous crystal structure model of SnTe3O8 based on powder X-ray diffraction data [Meunier & Galy (1971[Meunier, G. & Galy, J. (1971). Acta Cryst. B27, 602-608.]). Acta Cryst. B27, 602–608], the current model based on single-crystal X-ray data is improved in terms of precision and accuracy. Nearly regular [SnO6] octa­hedra (Sn site symmetry .[\overline{3}].) are situated in the voids of an oxidotellurate(IV) framework built up by corner-sharing [TeO4] bis­phenoids (Te site symmetry 2..). A qu­anti­tative structural comparison revealed a very high degree of similarity for the structures with M = Ti, Zr, Sn in the MIVTe3O8 series.

1. Chemical context

The crystal chemistry of oxidotellurates(IV) is dominated by the presence of the 5s2 electron lone pair that, in the majority of cases, is stereochemically active, thus enabling one-sided coordination spheres around the TeIV atom (Christy et al., 2016[Christy, A. G., Mills, S. J. & Kampf, A. R. (2016). Miner. Mag. 80, 415-545.]). This peculiar building block often results in compounds with non-centrosymmetric structures or structures with polar directions exhibiting inter­esting physical properties (Ra et al., 2003[Ra, H.-S., Ok, K. M. & Halasyamani, P. S. (2003). J. Am. Chem. Soc. 125, 7764-7765.]; Kim et al., 2014[Kim, Y. H., Lee, D. W. & Ok, K. M. (2014). Inorg. Chem. 53, 5240-5245.]). In this context, the microwave dielectric properties of MIVTe3O8 (M = Sn, Zr) ceramics were investigated some time ago (Subodh & Sebastian, 2008[Subodh, G. & Sebastian, M. T. (2008). Jpn. J. Appl. Phys. 47, 7943-7946.]).

The crystal structure of the isotypic series MIVTe3O8 was originally determined for M = Ti from a single crystal in space group Ia[\overline{3}] using photographic Weissenberg X-ray data, whereas for M = Sn, Zr and Hf, the crystal structures were refined from powder X-ray data (Meunier & Galy, 1971[Meunier, G. & Galy, J. (1971). Acta Cryst. B27, 602-608.]). In subsequent studies, crystal-structure refinements on the basis of single-crystal X-ray data were reported for the mineral winstanleyite with composition (Ti0.96Fe0.04)Te3O8 (Bindi & Cipriani, 2003[Bindi, L. & Cipriani, C. (2003). Can. Mineral. 41, 1469-1473.]), and for the synthetic compound ZrTe3O8 (Noguera et al., 2003[Noguera, O., Thomas, P., Masson, O. & Champarnaud-Mesjard, J. C. (2003). Z. Kristallogr. New Cryst. Struct. 218, 293-294.]; Lu et al., 2019[Lu, W., Gao, Z., Du, X., Tian, X., Wu, Q., Li, C., Sun, Y., Liu, Y. & Tao, X. (2019). Inorg. Chem. 58, 7794-7802.]). A powder X-ray study of the solid solution Sn0.59Ti0.41Te3O8 crystallizing in the MIVTe3O8 structure type has also been reported (Ben Aribia et al., 2008[Ben Aribia, W., Loukil, M., Kabadou, A. & Ben Salah, A. (2008). Powder Diffr. 23, 228-231.]).

Single-crystal growth of oxidotellurates(IV) can be accomplished through various crystallization methods including, for example, experiments under hydro­thermal conditions (Weil et al., 2018[Weil, M., Shirkhanlou, M., Füglein, E. & Libowitzky, E. (2018). Crystals, 8, 51.]), cooling from the melt (Stöger et al., 2009[Stöger, B., Weil, M., Zobetz, E. & Giester, G. (2009). Acta Cryst. B65, 167-181.]), from salt melts as fluxing agents (Weil, 2019[Weil, M. (2019). Acta Cryst. E75, 26-29.]), or from chemical vapor transport reactions (Missen et al., 2020[Missen, O. P., Weil, M., Mills, S. J. & Libowitzky, E. (2020). Acta Cryst. B76, 1-6.]). The latter method (Binnewies et al., 2012[Binnewies, M., Glaum, R., Schmidt, M. & Schmidt, P. (2012). Chemical Vapor Transport Reactions. Berlin, Boston: De Gruyter,.]) is particularly suitable for growing large crystals of high quality and was the method of choice for crystal growth of SnTe3O8 for which a more precise and accurate structure refinement appeared to be desirable.

2. Structural commentary

The asymmetric unit of SnTe3O8 comprises one SnIV atom, one TeIV atom, and two oxide anions, residing on sites 8a (site symmetry .[\overline{3}].), 24d (2..), 48e (1) and 16c (.3.), respectively. The tin atom is in an almost regular octa­hedral coordination by oxygen, with six equal Sn1—O1 distances, all trans angles equal to 180°, and cis angles ranging from 86.09 (4) to 93.91 (4)°. The Te1 site is coordinated by four O atoms in pairs of shorter (O1) and longer distances (O2) (Table 1[link]). The resulting [TeO4] coordination polyhedron is a distorted bis­phenoid. Considering the 5s2 electron lone pair at the TeIV atom, the corresponding [ΨTeO4] polyhedron has a shape inter­mediate between a square pyramid and a trigonal bipyra­mid with the non-bonding electron pair occupying an equatorial position (Fig. 1[link]). The geometry index τ5 of the [ΨTeO4] polyhedron is 0.471 (τ5 = 0 for an ideal square pyramid and τ5 = 1 for an ideal trigonal bipyramid; Addison et al., 1984[Addison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349-1356.]). The position of the electron lone pair was calculated with the LPLoc software (Hamani et al., 2020[Hamani, D., Masson, O. & Thomas, P. (2020). J. Appl. Cryst. 53, 1243-1251.]), with resulting fractional coordinates of x = 0.28655, y = 0, z = 1/4. The radius of the electron lone pair was calculated to be 1.07 Å with a distance of 0.90 Å from the Te1 position. The coordination numbers of the oxide anions are two and three: O1 coord­inates to Sn1 and Te1 at the shorter of the two Te1—O distances whereas O2 coordinates to three Te1 atoms at the longer of the two Te1—O distances.

Table 1
Selected geometric parameters (Å, °)

Sn1—O1i 2.0421 (11) Te1—O2 2.1278 (3)
Te1—O1ii 1.8800 (11)    
       
O1iii—Te1—O1ii 102.42 (8) O2iv—Te1—O2 157.05 (6)
O1iii—Te1—O2 86.60 (6) Te1—O2—Te1v 117.94 (2)
O1ii—Te1—O2 79.05 (4)    
Symmetry codes: (i) [-x+{\script{1\over 2}}, -y, z-{\script{1\over 2}}]; (ii) [-z+{\script{1\over 2}}, -x+{\script{1\over 2}}, -y+{\script{1\over 2}}]; (iii) [-z+{\script{1\over 2}}, x-{\script{1\over 2}}, y]; (iv) [x, -y, -z+{\script{1\over 2}}]; (v) z, x, y.
[Figure 1]
Figure 1
The coordination environment around Te1. Displacement ellipsoids are drawn at the 90% probability level; the electron lone pair is given as a green sphere of arbitrary radius. [Symmetry codes: (v) –z + [{1\over 2}], x–1/2, y; (vii) −z + [{1\over 2}], −x + [{1\over 2}], −y + [{1\over 2}]; (viii) x, −y, −z + [{1\over 2}].]

In the crystal structure of SnTe3O8, the [SnO6] octa­hedra are isolated from each other and arranged in rows running parallel to [100]. Each of the [TeO4] bis­phenoids shares corners (O2) with other [TeO4] bis­phenoids to form a three-dimensional oxidotellurate(IV) framework. The [SnO6] octa­hedra are situated in the voids of this framework, thereby sharing each of the six corners with an individual [TeO4] bis­phenoid. The crystal structure of SnTe3O8 is depicted in Fig. 2[link].

[Figure 2]
Figure 2
The crystal structure of SnTe3O8 in polyhedral representation, showing a projection along [[\overline{1}]00]. Displacement ellipsoids are as in Fig. 1[link]; [TeO4] polyhedra are red, [SnO6] octa­hedra are blue.

The unit-cell parameter a from the previous powder X-ray study, 11.144 (3) Å, as well as inter­atomic distances of Sn1—O1 = 2.032 Å (6×), Te1—O1 = 1.850 Å (2×), Te1—O2 = 2.124 Å (2×), and angles O1—Te1—O1′ = 102.9°, and O2—Te1—O2′ = 156.8° (Meunier & Galy, 1971[Meunier, G. & Galy, J. (1971). Acta Cryst. B27, 602-608.]) agree with the present single-crystal study (Table 1[link]), but with lower precision and accuracy. In comparison with the previous model based on powder X-ray data, the values of the bond-valence sums (Brown, 2002[Brown, I. D. (2002). The Chemical Bond in Inorganic Chemistry: The Bond Valence Model. Oxford University Press.]) using the parameters of Brese & O'Keeffe (1991[Brese, N. E. & O'Keeffe, M. (1991). Acta Cryst. B47, 192-197.]) are much closer to the expected values of 4 for Sn and Te and 2 for O on basis of the current model [previous model: Sn1 4.28 valence units (v.u.), Te1 4.10 v.u., O1 2.09 v.u., O2 2.08 v.u.; current model: Sn1: 4.14 v.u., Te1 3.93 v.u., O1 1.99 v.u., O2 2.00 v.u.].

The relation of the isotypic crystal structures of MIVTe3O8 compounds with that of the fluorite structure has been discussed previously for TiTe3O8 (Meunier & Galy, 1971[Meunier, G. & Galy, J. (1971). Acta Cryst. B27, 602-608.]; Wells, 1975[Wells, A. F. (1975). Structural Inorganic Chemistry, 4th ed, pp. 207-208. Oxford: Clarendon Press.]). The unit-cell parameter a of cubic TiTe3O8 is ∼2a of cubic CaF2, whereby the ordered distribution of the cationic sites leads to a doubling of the unit cell and also to a considerable distortion of the respective coordination environments. The original cubic coordination around the CaII cation in the fluorite structure is changed to an octa­hedral coordination of SnIV and a fourfold coordination of TeIV in the superstructure of the MIVTe3O8 compounds. Note that there are two additional O atoms at a distance of 3.2446 (19) Å around the MIV site and two pairs of additional O atoms at a distance of 2.9076 (12) and 3.3957 (13) Å around the Te1 site in SnTe3O8, completing an eightfold coordination in each case. Correspondingly, each of the two O sites has a fourfold coordination in case the much longer distances are counted.

A qu­anti­tative structural comparison of the MIVTe3O8 structures where single crystal data are available (M = Ti, Zr, Sn) was undertaken with the program compstru (de la Flor et al., 2016[Flor, G. de la, Orobengoa, D., Tasci, E., Perez-Mato, J. M. & Aroyo, M. I. (2016). J. Appl. Cryst. 49, 653-664.]) available at the Bilbao Crystallographic Server (Aroyo et al., 2006[Aroyo, M. I., Perez-Mato, J. M., Capillas, C., Kroumova, E., Ivantchev, S., Madariaga, G., Kirov, A. & Wondratschek, H. (2006). Z. Kristallogr. 221, 15-27.]). Table 2[link] lists the degree of lattice distortion (S), the maximum distance between the atomic positions of paired atoms (|u|), the arithmetic mean of all distances, and the measure of similarity (Δ) relative to SnTe3O8 as the reference structure. All these values show a very high similarity between the crystal structures in the isotypic MIVTe3O8 series.

Table 2
Atom pairs and their absolute distances |u| (Å) in the isotypic series MTe3O8 with SnTe3O8 as the reference structure, as well as degree of lattice distortion (S), arithmetic mean of the distances (dav, Å) and measure of similarity (Δ)

  TiTe3O8 a (Ti0.96Fe0.04)Te3O8 b ZrTe3O8c ZrTe3O8 d
MIV1 0 0 0 0
Te1 0.0475 0.0360 0.0065 0.0059
O1 0.1061 0.0834 0.0713 0.0694
O2 0.1374 0.0968 0.0543 0.0446
S 0.0107 0.0102 0.0076 0.0092
dav 0.0878 0.0668 0.0453 0.0436
Δ 0.011 0.008 0.006 0.006
Notes: (a) a = 10.956 (3) Å; Meunier & Galy (1971[Meunier, G. & Galy, J. (1971). Acta Cryst. B27, 602-608.]); (b) a = 10.965 (1) Å; Bindi & Cipriani (2003[Bindi, L. & Cipriani, C. (2003). Can. Mineral. 41, 1469-1473.]); (c) a = 11.308 (1) Å; Noguera et al. (2003[Noguera, O., Thomas, P., Masson, O. & Champarnaud-Mesjard, J. C. (2003). Z. Kristallogr. New Cryst. Struct. 218, 293-294.]); (d) a = 11.340 (4) Å; Lu et al. (2019[Lu, W., Gao, Z., Du, X., Tian, X., Wu, Q., Li, C., Sun, Y., Liu, Y. & Tao, X. (2019). Inorg. Chem. 58, 7794-7802.]).

3. Synthesis and crystallization

Reagent-grade chemicals were used without further purification. SnO2 (71 mg, 0.47 mmol) and TeO2 (225 mg, 1.40 mmol) were thoroughly mixed in the molar ratio 1:3 and placed in a silica tube to which 50 mg of TeCl4 were added as the transport agent. The silica ampoule was then evacuated and torch-sealed, placed in a two-zone furnace using a temperature gradient 973 K (source) → 873 K (sink) for three days. Cubic, canary-yellow crystals had formed in the millimetre size range in the colder sink region as the only product (Fig. 3[link]). Powder X-ray diffraction of the remaining material in the source region revealed SnTe3O8 as the main phase and SnO2 as a side phase. For the single-crystal diffraction study, a fragment was broken from a larger crystal.

[Figure 3]
Figure 3
Photograph of Sn3TeO8 single crystals grown by chemical vapor transport reactions.

4. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. Atomic coordinates and the labelling scheme were adapted from isotypic TiTe3O8 (Meunier & Galy, 1971[Meunier, G. & Galy, J. (1971). Acta Cryst. B27, 602-608.]).

Table 3
Experimental details

Crystal data
Chemical formula SnTe3O8
Mr 629.49
Crystal system, space group Cubic, Ia[\overline{3}]
Temperature (K) 296
a (Å) 11.1574 (4)
V3) 1388.96 (15)
Z 8
Radiation type Mo Kα
μ (mm−1) 16.04
Crystal size (mm) 0.06 × 0.06 × 0.01
 
Data collection
Diffractometer Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.])
Tmin, Tmax 0.452, 0.748
No. of measured, independent and observed [I > 2σ(I)] reflections 14087, 735, 697
Rint 0.048
(sin θ/λ)max−1) 0.907
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.014, 0.030, 1.07
No. of reflections 735
No. of parameters 21
Δρmax, Δρmin (e Å−3) 1.27, −0.86
Computer programs: APEX3 and SAINT (Bruker, 2018[Bruker (2018). APEX3 and SAINT. Bruker-AXS Inc. Madison, Wisconsin, USA.]), SHELXL (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), ATOMS (Dowty, 2006[Dowty, E. (2006). ATOMS. Shape Software, Kingsport, Tennessee, USA.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Computing details top

Data collection: APEX3 (Bruker, 2018); cell refinement: SAINT (Bruker, 2018); data reduction: SAINT (Bruker, 2018); program(s) used to solve structure: coordinates from previous refinement; program(s) used to refine structure: SHELXL (Sheldrick, 2015); molecular graphics: ATOMS (Dowty, 2006); software used to prepare material for publication: publCIF (Westrip, 2010).

Tin(IV) trioxidotellurate(IV) top
Crystal data top
SnTe3O8Mo Kα radiation, λ = 0.71073 Å
Mr = 629.49Cell parameters from 5266 reflections
Cubic, Ia3θ = 3.7–38.9°
a = 11.1574 (4) ŵ = 16.04 mm1
V = 1388.96 (15) Å3T = 296 K
Z = 8Plate, light yellow
F(000) = 21600.06 × 0.06 × 0.01 mm
Dx = 6.021 Mg m3
Data collection top
Bruker APEXII CCD
diffractometer
697 reflections with I > 2σ(I)
ω– and φ–scansRint = 0.048
Absorption correction: multi-scan
(SADABS; Krause et al., 2015)
θmax = 40.2°, θmin = 3.7°
Tmin = 0.452, Tmax = 0.748h = 1820
14087 measured reflectionsk = 2020
735 independent reflectionsl = 1920
Refinement top
Refinement on F20 restraints
Least-squares matrix: full w = 1/[σ2(Fo2) + (0.0127P)2 + 1.9293P]
where P = (Fo2 + 2Fc2)/3
R[F2 > 2σ(F2)] = 0.014(Δ/σ)max = 0.001
wR(F2) = 0.030Δρmax = 1.27 e Å3
S = 1.07Δρmin = 0.85 e Å3
735 reflectionsExtinction correction: SHELXL-2017/1 (Sheldrick 2015), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
21 parametersExtinction coefficient: 0.00046 (3)
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.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Sn10.0000000.0000000.0000000.00501 (4)
Te10.20584 (2)0.0000000.2500000.00804 (4)
O10.43242 (10)0.13738 (10)0.39972 (11)0.0129 (2)
O20.16789 (10)0.16789 (10)0.16789 (10)0.0078 (3)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Sn10.00501 (4)0.00501 (4)0.00501 (4)0.00031 (3)0.00031 (3)0.00031 (3)
Te10.00518 (5)0.01273 (6)0.00620 (5)0.0000.0000.00229 (4)
O10.0095 (4)0.0117 (4)0.0174 (5)0.0019 (3)0.0018 (4)0.0094 (4)
O20.0078 (3)0.0078 (3)0.0078 (3)0.0020 (3)0.0020 (3)0.0020 (3)
Geometric parameters (Å, º) top
Sn1—O1i2.0421 (11)Sn1—O1vi2.0421 (11)
Sn1—O1ii2.0421 (11)Te1—O1v1.8800 (11)
Sn1—O1iii2.0421 (11)Te1—O1vii1.8800 (11)
Sn1—O1iv2.0421 (11)Te1—O2viii2.1278 (3)
Sn1—O1v2.0421 (11)Te1—O22.1278 (3)
O1i—Sn1—O1ii86.09 (4)O1iv—Sn1—O1vi93.91 (4)
O1i—Sn1—O1iii93.91 (4)O1v—Sn1—O1vi86.09 (4)
O1ii—Sn1—O1iii180.00 (9)O1v—Te1—O1vii102.42 (8)
O1i—Sn1—O1iv86.09 (4)O1v—Te1—O2viii79.05 (4)
O1ii—Sn1—O1iv93.91 (4)O1vii—Te1—O2viii86.60 (6)
O1iii—Sn1—O1iv86.09 (4)O1v—Te1—O286.60 (6)
O1i—Sn1—O1v93.91 (4)O1vii—Te1—O279.05 (4)
O1ii—Sn1—O1v86.09 (4)O2viii—Te1—O2157.05 (6)
O1iii—Sn1—O1v93.91 (4)Te1ix—O1—Sn1x134.17 (6)
O1iv—Sn1—O1v180.00 (9)Te1—O2—Te1xi117.94 (2)
O1i—Sn1—O1vi180.00 (9)Te1—O2—Te1xii117.94 (2)
O1ii—Sn1—O1vi93.91 (4)Te1xi—O2—Te1xii117.94 (2)
O1iii—Sn1—O1vi86.09 (4)
Symmetry codes: (i) y, z+1/2, x1/2; (ii) x+1/2, y, z1/2; (iii) x1/2, y, z+1/2; (iv) z1/2, x+1/2, y; (v) z+1/2, x1/2, y; (vi) y, z1/2, x+1/2; (vii) z+1/2, x+1/2, y+1/2; (viii) x, y, z+1/2; (ix) y+1/2, z+1/2, x+1/2; (x) x+1/2, y, z+1/2; (xi) z, x, y; (xii) y, z, x.
Atom pairs and their absolute distances |u| (Å) in the isotypic series MTe3O8 with SnTe3O8 as the reference structure, as well as degree of lattice distortion (S), arithmetic mean of the distances (dav / Å) and measure of similarity (Δ) top
TiTe3O8 a(Ti0.96Fe0.04)Te3O8 bZrTe3O8cZrTe3O8 d
MIV10000
Te10.04750.03600.00650.0059
O10.10610.08340.07130.0694
O20.13740.09680.05430.0446
S0.01070.01020.00760.0092
dav0.08780.06680.04530.0436
Δ0.0110.0080.0060.006
Notes: (a) a = 10.956 (3) Å; Meunier & Galy (1971); (b) a = 10.965 (1) Å; Bindi & Cipriani (2003); (c) a = 11.308 (1) Å; Noguera et al. (2003); (d) a = 11.340 (4) Å; Lu et al. (2019).
 

Acknowledgements

The X-ray centre of the Vienna University of Technology is acknowledged for financial support and for providing access to the single-crystal and powder X-ray diffractometers.

References

First citationAddison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349–1356.  CSD CrossRef Web of Science Google Scholar
First citationAroyo, M. I., Perez-Mato, J. M., Capillas, C., Kroumova, E., Ivantchev, S., Madariaga, G., Kirov, A. & Wondratschek, H. (2006). Z. Kristallogr. 221, 15–27.  Web of Science CrossRef CAS Google Scholar
First citationBen Aribia, W., Loukil, M., Kabadou, A. & Ben Salah, A. (2008). Powder Diffr. 23, 228–231.  CAS Google Scholar
First citationBindi, L. & Cipriani, C. (2003). Can. Mineral. 41, 1469–1473.  Web of Science CrossRef ICSD CAS Google Scholar
First citationBinnewies, M., Glaum, R., Schmidt, M. & Schmidt, P. (2012). Chemical Vapor Transport Reactions. Berlin, Boston: De Gruyter,.  Google Scholar
First citationBrese, N. E. & O'Keeffe, M. (1991). Acta Cryst. B47, 192–197.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBrown, I. D. (2002). The Chemical Bond in Inorganic Chemistry: The Bond Valence Model. Oxford University Press.  Google Scholar
First citationBruker (2018). APEX3 and SAINT. Bruker-AXS Inc. Madison, Wisconsin, USA.  Google Scholar
First citationChristy, A. G., Mills, S. J. & Kampf, A. R. (2016). Miner. Mag. 80, 415–545.  Web of Science CrossRef CAS Google Scholar
First citationDowty, E. (2006). ATOMS. Shape Software, Kingsport, Tennessee, USA.  Google Scholar
First citationFlor, G. de la, Orobengoa, D., Tasci, E., Perez-Mato, J. M. & Aroyo, M. I. (2016). J. Appl. Cryst. 49, 653–664.  Web of Science CrossRef IUCr Journals Google Scholar
First citationHamani, D., Masson, O. & Thomas, P. (2020). J. Appl. Cryst. 53, 1243–1251.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKim, Y. H., Lee, D. W. & Ok, K. M. (2014). Inorg. Chem. 53, 5240–5245.  Web of Science CrossRef ICSD CAS PubMed Google 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 citationLu, W., Gao, Z., Du, X., Tian, X., Wu, Q., Li, C., Sun, Y., Liu, Y. & Tao, X. (2019). Inorg. Chem. 58, 7794–7802.  Web of Science CrossRef ICSD PubMed Google Scholar
First citationMeunier, G. & Galy, J. (1971). Acta Cryst. B27, 602–608.  CrossRef ICSD IUCr Journals Web of Science Google Scholar
First citationMissen, O. P., Weil, M., Mills, S. J. & Libowitzky, E. (2020). Acta Cryst. B76, 1–6.  Web of Science CrossRef ICSD IUCr Journals Google Scholar
First citationNoguera, O., Thomas, P., Masson, O. & Champarnaud-Mesjard, J. C. (2003). Z. Kristallogr. New Cryst. Struct. 218, 293–294.  CrossRef ICSD CAS Google Scholar
First citationRa, H.-S., Ok, K. M. & Halasyamani, P. S. (2003). J. Am. Chem. Soc. 125, 7764–7765.  Web of Science CrossRef ICSD PubMed CAS Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationStöger, B., Weil, M., Zobetz, E. & Giester, G. (2009). Acta Cryst. B65, 167–181.  Web of Science CrossRef ICSD IUCr Journals Google Scholar
First citationSubodh, G. & Sebastian, M. T. (2008). Jpn. J. Appl. Phys. 47, 7943–7946.  Web of Science CrossRef CAS Google Scholar
First citationWeil, M. (2019). Acta Cryst. E75, 26–29.  Web of Science CrossRef ICSD IUCr Journals Google Scholar
First citationWeil, M., Shirkhanlou, M., Füglein, E. & Libowitzky, E. (2018). Crystals, 8, 51.  Web of Science CrossRef ICSD Google Scholar
First citationWells, A. F. (1975). Structural Inorganic Chemistry, 4th ed, pp. 207–208. Oxford: Clarendon Press.  Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals 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 logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds