Abstract
Colorless crystals of the barium rare earth hydrogarnets Ba3[RE(OH)6]2 (RE = Sc, Y, Ho–Lu) were synthesized under hydroflux conditions with KOH at about T = 200 °C starting from the respective RE2O3 and Ba (NO3)2. Single-crystal X-ray diffraction analysis on these distorted rhombic dodecahedra revealed the cubic space group Ia
1 Introduction
Garnets, a class of nesosilicates, are described by the general formula
In the group of barium hydrogarnets, Ba3[Al(OH)6]2, Ba3[In(OH)6]2 and Ba3[Sc(OH)6]2 are the only known representatives [5], [6]. These compounds were synthesized by using the hydrothermal method yielding fine powders. Experiments to synthesize other barium rare earth hydrogarnets failed, so that these three compounds were assumed to be the only barium containing hydrogarnets existing [6]. A similar observation was made for strontium rare earth hydrogarnets, where a hydrothermal synthesis of hydrogarnets with rare earth metals larger than thulium was described as impossible [7].
We succeeded in synthesizing a variety of hydrogarnets using the novel hydroflux method as highly crystalline samples at comparatively low temperatures [8], [9]. Among them are strontium rare earth hydrogarnets Sr3[RE(OH)6]2 (RE = Sc, Y, Ho–Lu), of which we also obtained large and well-shaped crystals [4]. The reaction medium of the hydroflux method consists of an approximately equimolar mixture of alkali-metal hydroxide and water. The autogenous pressure is much lower than in classical hydrothermal processes, making the use of autoclaves dispensable. In addition, the hydroflux medium has a good solubility of oxides and hydroxides, which allows, among other things, the direct use of rare earth oxides instead of more expensive nitrates or sulfates. We now employed this approach to explore the apparently barren field of barium rare earth hydrogarnets. The seven hydrogarnets Ba3[RE(OH)6]2 (RE = Sc, Y, Ho–Lu) presented here considerably expand the knowledge about this compound class and provide the first single-crystal structures of barium hydroxometalates with the respective elements [10]. In addition, we provide single-crystal structure data of the strontium indium hydrogarnet Sr3[In(OH)6]2.
2 Results and discussion
2.1 Synthesis
An approximately equimolar mixture of water and diverse alkali-metal hydroxides can be utilized as a reaction medium in the hydroflux synthesis, with sodium and potassium hydroxide being the most common. In our experiments, no significant influence of the nature of these hydroxides was observed on the formation of the barium rare earth hydrogarnets. The base concentration of the hydroflux represents one of the most important reaction parameters [11]. By lowering the base concentration, the size and the quality of the crystals decreased until only a powder was obtained, which is in well accordance with the results of the hydrothermal experiments. Additionally, the product formation is notably slower when using the hydrothermal method, like in the synthesis of Ba3[In(OH)6]2 and Ba3[Sc(OH)6]2 where only small amounts of these hydrogarnets were obtained after 90 h at T = 200 °C and quantitative yields required 95 h at 400 °C [6]. In contrast, a typical hydroflux synthesis is completed after 10 h at 200 °C or even faster [12].
The syntheses of the here described barium rare earth hydrogarnets were performed in a potassium hydroxide hydroflux with a water-base molar ratio of n(H2O):n(KOH) = 1.6 in exception of Ba3[Sc(OH)6]2, where a ratio of 3 was used. Ba(NO3)2 and RE2O3 were added to the hydroflux in the stoichiometric ratio of the target compound. As both binary compounds are known to form hydroxides in alkaline aqueous solution [13], [14], the following metathesis reaction can be formulated:
The experiments were performed in a stainless steel autoclave with PTFE inlet considering the high corrosivity and to maintain a constant base concentration. The autoclaves were heated up to 200 °C for 10 h before slowly cooling down to room temperature. All barium hydrogarnets show signs of decomposition on their surfaces upon washing with water. Consequently, the products were washed with methanol and colorless distorted rhombic dodecahedra in single-crystal quality were obtained (Figure 1). In the case of the synthesis of Ba3[Ho(OH)6]2, large amounts of Ho(OH)3 and BaCO3 were obtained, showing parallels to the synthesis of Sr3[Ho(OH)6]2 [4]. In contrast, the remaining barium rare earth hydrogarnets were received phase pure (Figures S1 and S2, Supporting Information available online). Experiments with the remaining lanthanide oxides excluding promethium resulted in the formation of the respective hydroxides or oxide hydroxides as well as barium carbonate. In a synthesis with the same reaction parameters but starting from Sr(NO3)2 and In2O3, water-stable colorless rhombic dodecahedral crystals of Sr3[In(OH)6]2 were obtained (Figure S3, Information). Syntheses and powder X-ray analyses have already been reported for this compound [15]. Since no single-crystal diffraction data of this compound were published [10] we would like to fill also this gap.
![Figure 1: Single crystals of Ba3[Lu(OH)6]2 with the typical morphology of hydrogarnets obtained from hydroflux.](/document/doi/10.1515/znb-2020-0147/asset/graphic/j_znb-2020-0147_fig_001.jpg)
Single crystals of Ba3[Lu(OH)6]2 with the typical morphology of hydrogarnets obtained from hydroflux.
The thermal decomposition of Ba3[Y(OH)6]2 was investigated by thermogravimetry (TG) up to T = 1000 °C in a stream of CO2-free synthetic air (Figure S4, Supporting Information). In the thermogram, the first decomposition signal appears at about 250 °C, where three equivalents of water are quickly released. Subsequent heating to 900 °C leads in several steps to complete dehydration. The observed total mass loss of 13.9% matches the theoretical value (13.6%) for the elimination of six water molecules per formula unit. In addition, a sample of Ba3[Y(OH)6]2 was annealed at 1000 °C in air for 12 h. The powder X-ray diffraction (PXRD) pattern of the product (Figure S5, Supporting Information) revealed BaY2O4 and BaCO3. Thermal stability and decomposition products of Ba3[Y(OH)6]2 resemble those of Sr3[RE(OH)6]2 (RE = Sc, Y, Ho–Lu) [4].
2.2 Crystal structures
Typical hydrogarnets crystallize in the centrosymmetric cubic space group Ia
![Figure 2: Reconstructed 0kl planes of Ba3[Sc(OH)6]2 and Ba3[Tm(OH)6]2 hydrogarnets showing different reflection densities. Note the small reflections with odd k and l indices for the thullium hydrogarnet, which do not fulfill the reflection conditions of Ia3‾$‾{3}$d.](/document/doi/10.1515/znb-2020-0147/asset/graphic/j_znb-2020-0147_fig_002.jpg)
Reconstructed 0kl planes of Ba3[Sc(OH)6]2 and Ba3[Tm(OH)6]2 hydrogarnets showing different reflection densities. Note the small reflections with odd k and l indices for the thullium hydrogarnet, which do not fulfill the reflection conditions of Ia
![Figure 3: Ratio of the X-ray diffraction intensity and the standard deviation of selected reflection sets of Ba3[Sc(OH)6]2 and Ba3[Yb(OH)6]2 single crystals. For the ytterbium hydrogarnet, a small fraction of the reflections of a set have an unusually large I/σ ratio caused by the Renninger effect [19].](/document/doi/10.1515/znb-2020-0147/asset/graphic/j_znb-2020-0147_fig_003.jpg)
Ratio of the X-ray diffraction intensity and the standard deviation of selected reflection sets of Ba3[Sc(OH)6]2 and Ba3[Yb(OH)6]2 single crystals. For the ytterbium hydrogarnet, a small fraction of the reflections of a set have an unusually large I/σ ratio caused by the Renninger effect [19].
Upon close inspection of the data we are convinced that all violating reflections can be explained either by Renninger or λ/2 effects. If several reflections meet the Laue condition simultaneously, i.e., if two or more reciprocal lattice points concurrently lie on the Ewald sphere, this can lead to multiple reflection (Umweganregung, four-beam case), which enhances the intensities of weak reflections and lowers those of strong reflections significantly [19]. Such Renninger reflections never occur outside the Bravais lattice, but can violate zonal or serial reflection conditions. Renninger reflections vanish under a slightly different measuring position, i.e., the azimuthal angle ψ, of the same reflection. The effect is typically strong in crystals of high quality (low mosaic spread).
The here investigated cubic hydrogarnets are a showcase for multiple reflection. The quality of the crystals was excellent and extinction correction was necessary. Moreover, cooling to 100 (1) K reduced the lattice vibrations. The data sets (Rσ < 1%) have a high redundancy, with one and the same reflection being measured under several ψ angles. For example, the data of Ba3[Yb(OH)6]2 include the reflection 0 1 5 and its symmetry equivalents (multiplicity 24). These reflections should be systematically absent, because of the a glide planes of Ia
Some of the reflections which do not fulfill the reflection conditions of Ia
Hence, all barium rare earth hydrogarnets Ba3[RE(OH)6]2 with RE = Sc, Y, Ho–Lu crystallize with the centrosymmetric katoite structure in the space group Ia
![Figure 4: Left: Crystal structure of Ba3[Ho(OH)6]2 projected along [111], as an example of a hydrogarnet. Ellipsoids comprise 99.99% probability densities of the atoms at 100(1) K. Hydrogen atoms are omitted for clarity. Right: [RE(OH)6]3– octahedron and (O4H4)4– tetrahedra in Ba3[Ho(OH)6]2.](/document/doi/10.1515/znb-2020-0147/asset/graphic/j_znb-2020-0147_fig_004.jpg)
Left: Crystal structure of Ba3[Ho(OH)6]2 projected along [111], as an example of a hydrogarnet. Ellipsoids comprise 99.99% probability densities of the atoms at 100(1) K. Hydrogen atoms are omitted for clarity. Right: [RE(OH)6]3– octahedron and (O4H4)4– tetrahedra in Ba3[Ho(OH)6]2.
In all hydrogarnets, four adjacent hydroxide groups form an empty (O4H4)4– tetrahedron (Figure 4). The hydrogen atoms are located near faces of the polyhedron with the O-H bonds roughly bisecting the angle of the face [20]. The O (H)···O distances in the (O4H4)4– tetrahedron are long (>360 pm) and thus hydrogen bonds are supposed to be weak and not structure-directing. In the IR spectrum, the hydrogarnets Ba3[RE(OH)6]2 (RE = Sc, Y, Ho–Lu) show single infrared absorption bands in the narrow interval between 3632 and 3634 cm−1 (Figure S9, Supporting Information), which is related to the valence vibration of their hydroxide groups. Compared to the strontium rare earth hydrogarnets, the O-H valence vibrations in the barium compounds occur at higher energies, which fits well with the longer distances to the acceptor of the hydrogen bonds OD-H···OA (D: donor, A: acceptor atom of the H bridge).
2.3 Correlation between adopted space group and the ratio of interatomic distances d(M–O)/d(A–O)
Since hydrogen bonds seem to be not important for the structure, the space group of a specific hydrogarnet A3[M(OH)6]2 might depend on the size of its cations. Using the effective ionic radii tabulated by Shannon [21] did not lead to a meaningful correlation. Thus, we tested the structure immanent ratio of (average) interatomic distances d(M–O)/2d(A–O). The double weight for d(A–O) takes the occurrence of the respective interatomic distances into account (2 × [MO6] vs. 3 × [AO8] polyhedra). Table 1 lists most of the known hydrogarnets together with their space groups in the order of increasing distance ratio. Obviously, for small values of this ratio, i.e., small M3+ and large A2+ cations, the centrosymmetric space group Ia
Ratio of interatomic distances d(M–O)/2d(A–O) in strontium and barium hydrogarnets A3[M(OH)6]2. Unmarked compounds crystallize in the centrosymmetric katoite type (space group Ia
A3[M(OH)6]2 | d(M–O)/2d(A–O) | A3[M(OH)6]2 | d(M–O)/2d(A–O) |
---|---|---|---|
Sr3[Al(OH)6]2 [22] | 0.366 | Sr3[Sc(OH)6]2 [4] | 0.401 |
Ba3[Al(OH)6]2 [5] | 0.378 | Ba3[Er(OH)6]2 | 0.402 |
Sr3[Ga(OH)6]2 [23] | 0.381 | Ba3[Y(OH)6]2 | 0.403 |
Ba3[Sc(OH)6]2 | 0.381 | Ba3[Ho(OH)6]2 | 0.404 |
Sr3[Cr(OH)6]2 [9] | 0.382 | Sr3[In(OH)6]2 | 0.410 |
Ca3[Al(OH)6]2 [16] | 0.384 | Ca3[Rh(OH)6]2 [2] | 0.415 |
Sr3[Fe(OH)6]2 (own data) | 0.388 | Sr3[Lu(OH)6]2 [4] | 0.417 |
Sr3[Rh(OH)6]2 [9] | 0.392 | Sr3[Yb(OH)6]2 [4] | 0.419 |
Ba3[Lu(OH)6]2 | 0.397 | Sr3[Tm(OH)6]2 [4] | 0.420 |
Ba3[Yb(OH)6]2 | 0.399 | Sr3[Er(OH)6]2 [4] | 0.422 |
Ba3[Tm(OH)6]2 | 0.400 | Sr3[Ho(OH)6]2 [4] | 0.424 |
Ca3[Cr(OH)6]2 [9] | 0.400 | Sr3[Y(OH)6]2 [4] | 0.425 |
The distance ratio 0.410 of the strontium hydrogarnet Sr3[In(OH)6]2 falls into the undefined interval. Thus the structure, which had previously been described in Ia
3 Conclusions
Six new barium rare earth hydrogarnets Ba3[RE(OH)6]2 (RE = Y, Ho–Lu) were synthesized by using the hydroflux method starting from the respective oxide RE2O3 and Ba(NO3)2. A KOH hydroflux was used as reaction medium resulting in large colorless single crystals. Additionally, we were able to obtain single crystals of Ba3[Sc(OH)6]2 and Sr3[In(OH)6]2, whereas syntheses in literature had yielded only microcrystalline powders. X-ray diffraction analyses on single crystals of these barium rare earth hydrogarnets revealed the space group Ia
4 Experimental
4.1 Synthesis
The hydrogarnets were synthesized in a potassium hydroxide hydroflux. The reactions were carried out in a PTFE-lined 50 mL Berghof type DAB-2 autoclave starting from 0.3 mmol Ba(NO3)2 (VEB Laborchemie Apolda, 99%) and 0.1 mmol of RE2O3 [RE = Sc (ChemPur, 99.9%), Y (Fluka, 99.98%), Lu (Fluka, 99.99%), Yb (abcr, 99.9%), Tm (Riedel-de Haën, pure), Er (abcr, 99.9%), Ho (abcr, 99.995%)]. Water and potassium hydroxide (86%, Fisher Scientific) were added in a molar ratio of 1.6:1.0 to these compounds with the exception of Ba3[Sc(OH)6]2, where a ratio of 3 was used. The autoclaves were sealed and heated to 200 °C with 2 K min−1, held for 10 h before cooling down with a rate of 0.5 K min−1 to room temperature. The crystalline products were isolated by washing with methanol.
Sr3[In(OH)6]2 was synthesized under the same conditions as the lanthanide hydrogarnets but starting from 0.3 mmol Sr(NO3)2 and 0.1 mmol In2O3. The reaction product consists of water-stable colorless rhombic dodecahedral single crystals.
4.2 X-ray crystal structure determination
Intensity data was collected at T = 100(1) K with a four-circle diffractometer Kappa Apex2 (Bruker) equipped with a CCD detector using graphite-monochromated MoKα radiation (λ = 71.073 pm). The raw data were corrected for background, Lorentz and polarization factors [24], and multi-scan absorption correction was applied [25]. The structures were solved using Shelxl [26]. Structure refinement against Fo2 with Shelxl [27] included anisotropic displacement parameters for all non-hydrogen atoms (Tables S4–S12 of the Supporting Information). The hydrogen atom position was found in the difference Fourier map, and its isotropic displacement parameter was coupled to the respective oxygen atom. Graphics of the structure were developed with the program Diamond [28].
Further details of the crystal structure determinations are available from the Fachinformationszentrum Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany, crysdata@fiz-karlsruhe.de, on quoting the depository numbers listed in Table S2 (Supporting Information).
4.3 Powder X-ray diffraction
X-ray powder diffraction for phase identification and Rietveld refinement were determined at room temperature on a Stadi P diffractometer (STOE & Cie.) equipped with a curved Ge monochromator using CuKα1 radiation (λ = 154.056 pm) and with a Dectris Mythen 1K detector.
4.4 Infrared spectroscopy
IR spectra were recorded using a Vertex 70 FT-IR (Bruker) spectrometer. The device was operated in the ATR mode (diamond for a measuring range of 4000–600 cm−1). The software used to evaluate the spectra was Opus 6.5 [29].
4.5 Thermal analysis
The behavior of Ba3[Y(OH)6]2 upon heating in a synthetic air flow (CO2 free) was explored between T = 25 and 1000 °C with a heating rate of 5 K min−1 using a simultaneous thermal analyzer STA 409C/CD (Netzsch). In addition, a sample of Ba3[Y(OH)6]2 was thermally decomposed at 1000 °C in a chamber furnace for 12 h.
5 Supporting information
Additional crystallographic data, powder diffraction patterns, IR spectra, group-subgroup relationships, and structure graphics are given as online supplementary material.
Dedicated to: Professor Robert Glaum on the occasion of his 60th birthday.
Funding source: Deutsche Forschungsgemeinschaft
Award Identifier / Grant number: 501100001659
Acknowledgments
This work was financially supported by the Deutsche Forschungsgemeinschaft (project-id: 438795198).
Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
Research funding: This work was financially supported by the Deutsche Forschungsgemeinschaft (project-id: 438795198).
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
1. Zhang, H., Wang, T., Chen, X., Zhu, W. Particuology 2016, 24, 210–215; https://doi.org/10.1016/j.partic.2015.08.002.Search in Google Scholar
2. Cook, D. S., Clarkson, G. J., Dawson, D. M., Ashbrook, S. E., Fisher, J. M., Thompsett, D., Pickup, D. M., Chadwick, A. V., Walton, R. I. Inorg. Chem. 2018, 57, 11217–11224; https://doi.org/10.1021/acs.inorgchem.8b01797.Search in Google Scholar
3. Kozawa, T., Suzuki, Y., Naito, M. Mater. Chem. Phys. 2018, 212, 245–251; https://doi.org/10.1016/j.matchemphys.2018.03.060.Search in Google Scholar
4. Albrecht, R., Doert, T., Ruck, M. Z. Anorg. Allg. Chem. 2020, in press; https://doi.org/10.1002/zaac.202000031.Search in Google Scholar
5. Ahmed, A. H. M., Glasser, L. S. D. Acta Crystallogr. 1969, B25, 2169–2170; https://doi.org/10.1107/s0567740869005358.Search in Google Scholar
6. Kwestroo, W., van Gerven, H. C. A., van Hal, H. A. M. Mater. Res. Bull. 1977, 12, 161–164; https://doi.org/10.1016/0025-5408(77)90159-3.Search in Google Scholar
7. Ito, J. Mater. Res. Bull. 1968, 3, 495–500; https://doi.org/10.1016/0025-5408(68)90073-1.Search in Google Scholar
8. Bugaris, D. E., Smith, M. D., Zur Loye, H.-C. Inorg. Chem. 2013, 52, 3836–3844; https://doi.org/10.1021/ic302439b.Search in Google Scholar PubMed
9. Albrecht, R., Hunger, J., Doert, T., Ruck, M. Eur. J. Inorg. Chem. 2019, 10, 1398–1405; https://doi.org/10.1002/ejic.201801406.Search in Google Scholar
10. Hellenbrandt, M. Crystallogr. Rev. 2004, 10, 17–22; https://doi.org/10.1080/08893110410001664882.Search in Google Scholar
11. Albrecht, R., Hunger, J., Hölzel, M., Block, T., Pöttgen, R., Doert, T., Ruck, M. Chem. Open 2019, 8, 1399–1406; https://doi.org/10.1002/open.201900287.Search in Google Scholar
12. Albrecht, R., Hunger, J., Block, T., Pöttgen, R., Senyshyn, A., Doert, T., Ruck, M. Chem. Open 2019, 8, 74–83; https://doi.org/10.1002/open.201800229.Search in Google Scholar
13. Kragten, J., Decnop-Weever, L. G. Talanta 1982, 29, 219–222; https://doi.org/10.1016/0039-9140(82)80097-0.Search in Google Scholar
14. Kragten, J., Decnop-Weever, L. G. Talanta 1983, 30, 131–133; https://doi.org/10.1016/0039-9140(83)80033-2.Search in Google Scholar
15. Ito, J. Am. Mineral. 1968, 53, 1663–1673. http://www.minsocam.org/msa/collectors_corner/amtoc/toc1968.htm.Search in Google Scholar
16. Lager, G. A., Armbruster, T. Am. Mineral. 72 1987, 756. http://www.minsocam.org/msa/collectors_corner/amtoc/toc1987.htm. erwandten Kristallstrukturen: Anwendungen Der Kristallographischen Gruppenthe.Search in Google Scholar
17. Müller, U. Z. Anorg. Allg. Chem. 2004, 630, 1519–1537; https://doi.org/10.1002/zaac.200400250.Search in Google Scholar
18. Müller, U. Symmetriebeziehungen zwischen verwandten Kristallstrukturen: Anwendungen der kristallographischen Gruppentheorie in der Kristallchemie. Vieweg + Teubner Verlag: Wiesbaden, 2012.10.1007/978-3-8348-8342-1Search in Google Scholar
19. Renninger, M. Z. Physics 1937, 106, 141–176; https://doi.org/10.1007/bf01340315.Search in Google Scholar
20. Lager, G. A., Downs, R. T., Origlieri, M., Garoutte, R. Am. Mineral. 2002, 87, 642–647; https://doi.org/10.2138/am-2002-5-606.Search in Google Scholar
21. Shannon, R. D. Acta Crystallogr. 1976, A32, 751–767; https://doi.org/10.1107/s0567739476001551.Search in Google Scholar
22. Nevskii, N. N., Ivanov-Ehmin, B. N., Nevskaya, N. A., Kaziev, G. Z., Belov, N. V Dokl. Akad. Nauk SSSR 1982, 264, 857–858. http://mi.mathnet.ru/eng/dan45347.Search in Google Scholar
23. Kim, M. K., Jo, V., Shim, I.-W., Ok, K. M. Inorg. Chem. 2009, 48, 1275–1277; https://doi.org/10.1021/ic802187w.Search in Google Scholar PubMed
24. Apex2Data Reduction and Frame Integration Program for the CCD Area-Detector System; Bruker AXS Inc.: Madison, Wisconsin (USA), 2014.Search in Google Scholar
25. Sheldrick, G. M. Sadabs, Area-Detector Absorption Correction; Bruker AXS Inc.: Madison, Wisconsin (USA), 2014.Search in Google Scholar
26. Sheldrick, G. M. Acta Crystallogr. 2015, A71, 3–8; https://doi.org/10.1107/s2053273314026370.Search in Google Scholar
27. Sheldrick, G. M. Acta Crystallogr. 2015, C71, 3–8 https://doi.org/10.1107/S2053229614024218.Search in Google Scholar PubMed PubMed Central
28. Brandenburg, K. Diamond 4, Crystal and Molecular Structure Visualization; Crystal Impact GbR: Bonn (Germany), 2017.Search in Google Scholar
29. Opus (version 6.5); Bruker Optik GmbH: Ettlingen (Germany), 2007.Search in Google Scholar
Supplementary Material
The online version of this article offers supplementary material (https://doi.org/10.1515/znb-2020-0147).
© 2020 Ralf Albrecht et al., published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- In this issue
- Editorial
- Robert Glaum zum 60. Geburtstag gewidmet
- Research Articles
- Structural transition and antiferromagnetic ordering in the solid solution CePd1−xAuxAl (x = 0.1–0.9)
- Ternary plumbides ATPb2 (A = Ca, Sr, Ba, Eu; T = Rh, Pd, Pt) with distorted, lonsdaleite-related substructures of tetrahedrally connected lead atoms
- Intergrowth of niobium tungsten oxides of the tetragonal tungsten bronze type
- FeBiS2Cl – A new iron-containing member of the MPnQ2X family
- Thallium diphosphates
- Behavior of beryllium halides and triflate in acetonitrile solutions
- Hydroflux syntheses and crystal structures of hydrogarnets Ba3[RE(OH)6]2 (RE = Sc, Y, Ho–Lu)
- Three of a kind? The non-isotypic triple CsCe[P2Se6], CsSm[P2Se6] and CsEr[P2Se6]
- The crystal structure of ZrCr2D≈4 at 50 K ≤ T ≤ 200 K
- High-pressure synthesis and crystal structure of HP-Al2B3O7(OH)
- New layered supertetrahedral compounds T2-MSiAs2, T3-MGaSiAs3 and polytypic T4-M4Ga5SiAs9 (M = Sr, Eu)
- Comparative photophysical study of Pt(II) complex-nanoclay hybrid materials as dry powders and hydrogels
- Carbon subsulfide C3S2 – synthesis by flash vacuum pyrolysis and crystal structure determination
Articles in the same Issue
- Frontmatter
- In this issue
- Editorial
- Robert Glaum zum 60. Geburtstag gewidmet
- Research Articles
- Structural transition and antiferromagnetic ordering in the solid solution CePd1−xAuxAl (x = 0.1–0.9)
- Ternary plumbides ATPb2 (A = Ca, Sr, Ba, Eu; T = Rh, Pd, Pt) with distorted, lonsdaleite-related substructures of tetrahedrally connected lead atoms
- Intergrowth of niobium tungsten oxides of the tetragonal tungsten bronze type
- FeBiS2Cl – A new iron-containing member of the MPnQ2X family
- Thallium diphosphates
- Behavior of beryllium halides and triflate in acetonitrile solutions
- Hydroflux syntheses and crystal structures of hydrogarnets Ba3[RE(OH)6]2 (RE = Sc, Y, Ho–Lu)
- Three of a kind? The non-isotypic triple CsCe[P2Se6], CsSm[P2Se6] and CsEr[P2Se6]
- The crystal structure of ZrCr2D≈4 at 50 K ≤ T ≤ 200 K
- High-pressure synthesis and crystal structure of HP-Al2B3O7(OH)
- New layered supertetrahedral compounds T2-MSiAs2, T3-MGaSiAs3 and polytypic T4-M4Ga5SiAs9 (M = Sr, Eu)
- Comparative photophysical study of Pt(II) complex-nanoclay hybrid materials as dry powders and hydrogels
- Carbon subsulfide C3S2 – synthesis by flash vacuum pyrolysis and crystal structure determination