Abstract
Orthorhombic HP-Al2B3O7(OH) was synthesized in a Walker-type multianvil apparatus under high-pressure/high-temperature conditions of 12.4 GPa and 1200 °C, respectively. Its structure is isotypic to that of Ga2B3O7(OH) and has been determined via single-crystal X-ray diffraction at room temperature. HP-Al2B3O7(OH) crystallizes in the space group Cmce (Z = 8) with the lattice parameters a = 10.3124(4), b = 7.3313(3), c = 10.4801(5) Å, and V = 792.33(6) Å3. The compound has also been characterized by IR and Raman spectroscopy.
1 Introduction
Regarding the system Al–B–O–H, the literature research yields five compounds so far. Besides the well-known mineral Jeremejevite (Al6(BO3)5(F,OH)3) [1], our working group recently prepared the compound Al5B12O25(OH) which represents a new member of the high-pressure structure type M5B12O25(OH) (M = Al, Ga, In, Ga/In) [2], [3]. Jianhua Lin et al. who are working on hydrothermally synthesized zeolite-like borate structures, published the remaining three compounds in the system Al–B–O–H. In detail, this research working group found three microporous aluminum borates, which they named PKU-1 [4], PKU-2 [5], and PKU-6 [6] (PKU stands for Peking University). PKU-1 with the sum formula HAl3B6O12(OH)4 was published in 2003. In this compound all Al atoms are octahedrally coordinated building up a porous structure with 10- and 18-membered ring windows. PKU-2 (Al2B5O9(OH)3·nH2O) is even more porous featuring 24-membered rings around channels of edge-sharing AlO6 octahedra. PKU-6 (Al2B3O7(OH) – designated as HAl2B3O8 in reference [6]) was published in 2007 and shares the sum formula and point group with our title compound but shows a completely different structure. Infinite chains of square-pyramidally coordinated Al atoms are interlinked by three corner-sharing trigonal planar BO3 groups. Thus, in the hydrothermally synthesized PKU-6, the Al atoms are pentacoordinated and the B atoms show a three-fold coordination. In our title compound HP-Al2B3O7(OH), the coordination numbers of the aluminum and boron atoms are increased by one as one would expect for a high-pressure phase due to the pressure-coordination rule in solid state chemistry. However, the condensed structure of HP-Al2B3O7(OH) is not unprecedented as it has already been described in 2015 for the gallium containing isotype Ga2B3O7(OH) [7].
The present report contains a detailed description of the synthesis and crystal structure of our title compound as well as vibrational spectroscopic data of HP-Al2B3O7(OH) single crystals.
2 Experimental section
2.1 Synthesis
The starting materials Al(NO3)3·9H2O (≥98%, Sigma-Aldrich) and H3BO3 (99.5% Carl Roth) were mixed in a stoichiometric ratio of 2:3, ground in an agate mortar and encapsulated in platinum foil (0.027 mm, 99.9%, ChemPUR). The platinum capsule was then placed into a crucible with a lid, both made of α-BN (Henze Boron Nitride Products AG) and centered in a “14/8” assembly which includes, amongst others, a graphite furnace for resistance heating. Inside a uniaxially compressed Walker-type module, six steel wedges and eight tungsten carbide cubes surround the octahedral assembly to generate a quasi-hydrostatic pressure on the sample. A detailed description of the experimental setup is given in the literature [8], [9], [10]. HP-Al2B3O7(OH) was formed under the extreme reaction conditions of 12.4 GPa and 1200 °C. The required pressure and temperature were adjusted within 6 h and 10 min, respectively. The heating power was maintained for 10 min and then gradually lowered to T = 800 °C within 30 min. Finally, the sample was quenched to room temperature and slowly decompressed to ambient conditions. After freeing the platinum capsule from its surroundings and cutting it open with a scalpel, a clean-white reaction product with colorless single crystals appeared. Compared to our starting mixture, the product quantity was obviously diminished, which we explain with the great amount of volatile parts in aluminum nitrate nonahydrate. HP-Al2B3O7(OH) did also form as a byproduct in syntheses performed with Al2O3 (99.9%, Alfa Aesar) instead of Al(NO3)3·9H2O, slightly varying ratios of Al to B, higher temperatures up to 1400 °C, or a lower pressure of 12.0 GPa.
2.2 Crystal structure determination by X-ray diffraction
The reaction product was analyzed with a STOE Stadi P powder diffractometer equipped with a Mythen 1 K detector (Dectris, Switzerland). The measurement was carried out with Ge(111)-monochromatized MoKα1 radiation (λ = 0.7093 Å) in transmission geometry across a 2θ range of 2.0–50.0°. Figure 1 shows a comparison of the experimental powder pattern with a simulation derived from single-crystal structure data. HP-Al2B3O7(OH) clearly constitutes the main phase, but nevertheless the experimental powder pattern shows additional reflections of at least one byproduct.

Experimental powder pattern (MoKα1) of the optimized synthesis leading to HP-Al2B3O7(OH) (top) compared to a simulation of the powder pattern of HP-Al2B3O7(OH) based on single-crystal data (bottom). The asterisk-marked reflections in the experimental powder pattern refer to an unidentified byproduct.
The colorless single crystals of HP-Al2B3O7(OH) were measured with a Bruker D8 Quest diffractometer equipped with a Photon 100 CMOS detector. For structure solution and parameter refinement, the software tools Shelxs-2013/1 [11], [12] and Shelxl-2013/4 [13] implemented in the program WinGx-2018.1 [14] were used. According to the systematic reflection conditions, HP-Al2B3O7(OH) was solved and refined in the space group Cmce (no. 64) and afterwards standardized with the routine Structure Tidy[15] implemented in Platon[16](version 170613). Except for the proton, all atoms could be refined anisotropically. Details of the data acquisition can be found in the synoptical Table 1. Table 2 contains the positional parameters and Wyckoff positions and Table 3 the displacement parameters of all atoms in HP-Al2B3O7(OH).
Single-crystal data and structure refinement of HP-Al2B3O7(OH).
Empirical formula | Al2B3O7(OH) |
Molar mass, g mol−1 | 215.40 |
Crystal system | orthorhombic |
Space group | Cmce (no. 64) |
Single-crystal diffractometer | Bruker D8 Quest Kappa |
Radiation/wavelength λ, pm | MoKα/71.07 |
a, Å | 10.3124(4) |
b, Å | 7.3313(3) |
c, Å | 10.4801(5) |
V, Å3 | 792.33(6) |
Formula units per cell Z | 8 |
Calculated density, g cm−3 | 3.61 |
Crystal size, mm3 | 0.030 × 0.050 × 0.090 |
Temperature, K | 183(2) |
Absorption coefficient, mm−1 | 0.7 |
F(000), e | 848 |
θ range, deg | 3.89–39.44 |
Range in hkl | ±18; ±13; ±18 |
Refl. total/independent | 8079/1236 |
Rint/Rσ | 0.0324/0.0226 |
Refl. with I > 2 σ(I) | 1095 |
Data/ref. parameters | 1236/69 |
Absorption correction | multiscan |
Final R1/wR2 (I > 2 σ(I)) | 0.0198/0.0469 |
Final R1/wR2 (all data) | 0.0250/0.0497 |
Goodness-of-fit on F2 | 0.921 |
Largest diff. peak/hole, e Å−3 | 0.38/−0.34 |
Wyckoff positions, atomic coordinates, isotropic Uiso or equivalent isotropic displacement parameters Ueq (Å2) for HP-Al2B3O7(OH). Ueq is defined as one third of the trace of the orthogonalized Uij tensor (standard deviations in parentheses).
Atom | Wyckoff position | x | y | z | Ueq (Uiso for H4) |
---|---|---|---|---|---|
Al1 | 8f | 1/2 | 0.78853(3) | 0.88079(2) | 0.00190(6) |
Al2 | 8d | 0.85164(2) | 0 | 0 | 0.00208(6) |
B1 | 16g | 0.62480(6) | 1.14999(9) | 0.87176(6) | 0.00290(9) |
B2 | 8e | 3/4 | 0.8957(2) | 3/4 | 0.0030(2) |
O1 | 16g | 0.62447(4) | 1.24742(6) | 0.98924(4) | 0.00291(7) |
O2 | 16g | 0.63173(4) | 0.78036(6) | 0.73973(4) | 0.00280(7) |
O3 | 16g | 0.73359(4) | 1.01421(6) | 0.86459(4) | 0.00263(7) |
O4 | 8f | 1/2 | 0.53723(8) | 0.88783(6) | 0.00289(9) |
O5 | 8f | 1/2 | 1.04199(8) | 0.86111(6) | 0.00277(9) |
H4 | 8f | 1/2 | 0.468(3) | 0.812(3) | 0.065(8) |
Anisotropic displacement parameters Uij (Å2) of HP-Al2B3O7(OH) (standard deviations in parentheses).
Atom | U11 | U22 | U33 | U12 | U13 | U23 |
---|---|---|---|---|---|---|
Al1 | 0.0023(2) | 0.0016(2) | 0.0019(2) | 0 | 0 | 0.00032(7) |
Al2 | 0.0023(2) | 0.0019(2) | 0.0020(2) | 0 | 0 | −0.00002(7) |
B1 | 0.0029(2) | 0.0030(2) | 0.0029(2) | −0.0003(2) | −0.0002(2) | 0.0001(2) |
B2 | 0.0032(3) | 0.0026(3) | 0.0031(3) | 0 | 0.0000(2) | 0 |
O1 | 0.0035(2) | 0.0029(2) | 0.0024(2) | −0.0003(2) | 0.0002(2) | −0.0007(2) |
O2 | 0.0027(2) | 0.0029(2) | 0.0028(2) | −0.0007(2) | 0.0003(2) | −0.0008(2) |
O3 | 0.0027(2) | 0.0028(2) | 0.0024(2) | 0.0008(2) | −0.0002(2) | −0.0006(2) |
O4 | 0.0035(2) | 0.0028(2) | 0.0024(2) | 0 | 0 | −0.0003(2) |
O5 | 0.0020(2) | 0.0023(2) | 0.0040(2) | 0 | 0 | −0.0001(2) |
CSD-2027675 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre viawww.ccdc.cam.ac.uk/data_request/cif.
2.3 Vibrational spectroscopy
The transmission FT-IR spectrum of an HP-Al2B3O7(OH) single crystal was measured in the spectral range of 600–4000 cm−1 with a Vertex 70 FT-IR spectrometer (spectral resolution 4 cm−1) equipped with a KBr beam splitter, a liquid nitrogen cooled MCT (Mercury Cadmium Telluride) detector and a Hyperion 3000 microscope (Bruker). 320 scans of the sample were acquired using a Globar (silicon carbide) rod as mid-IR source and a 15× IR objective as focus. During the measurement, the sample was positioned on a BaF2 window. Atmospheric influences were corrected with the software Opus 6.5 [17].
A single-crystal Raman spectroscopic analysis was performed using a HORIBA JOBIN-YVON LabRam-HR800 spectrometer with a focal length of 800 mm. The spectrum was recorded unpolarized at ambient conditions. The analyzed area was excited using a (green) Nd-YAG laser with an excitation wavelength of 532 nm. Scattered light was dispersed by an optical grating with 1800 grooves per mm resulting in a spectral resolution of 1.92 cm−1, and collected with a cooled Andor™ CCD collector. An Olympus 50× objective, a confocal pinhole of 1000 µm and a slit of 100 µm was used for the measurement to maximize the intensity at a lateral resolution of ∼5 µm2. Before the analyses, the spectral position of the Raman mode of a Si standard wafer was measured against the position of the Rayleigh line, resulting in the expected Raman shift of 520.7 cm−1. The spectrum was obtained in multi-window acquisition mode provided by the LabSpec (version 5.93.20) [18] software package in the frequency range of 200–1500 cm−1 and represents the average of two single measurements with an acquisition time of 250 s each. Background subtraction and the determination of band position were performed with the same software. The background was subtracted using an 8th order line-function. Band positions were obtained by fitting a Gaussian–Lorentzian function to the individual bands.
3 Results and discussion
3.1 Crystal structure
HP-Al2B3O7(OH) crystallizes in the orthorhombic space group Cmce (no. 64) with eight formula units (Z = 8) and the lattice parameters a = 10.3124(4), b = 7.3313(3), c = 10.4801(5) Å, and V = 792.33(6) Å3. Details of the single-crystal data and structure refinement can be found in Table 1. Similar to the isotypic compound Ga2B3O7(OH), HP-Al2B3O7(OH) consists of two fundamental building blocks (BB1 and BB2). BB1 represents a cluster of four edge-sharing AlO6 octahedra (Figure 2a), and BB2 is composed of 12 corner-sharing BO4 tetrahedra forming a flexed unit of two sechser rings [19] that are interlinked by two vierer rings (Figure 2b). As depicted in Figure 3, each BB1 is encircled by two BB2s. The flexed BB2 units are interconnected forming endless zig-zag chains along the crystallographic b axis. Between these zig-zag chains, the BB1 clusters fill the space in alternating orientations in relation to c (visualized in Figure 4). All oxygen atoms belong to both building blocks except O4, which connects to three Al atoms and the proton in HP-Al2B3O7(OH). While the proton position was geometrically constructed in Ga2B3O7(OH), it was possible to freely refine it in the aluminum compound. The proton is located in the middle of a BO4 sechser ring at a distance of 0.94(3) Å to O4 and forms hydrogen bonds to two O2 atoms and to O5. Figure 5 shows the hydrogen bonding situation in HP-Al2B3O7(OH). As the proton needs space, it provokes the connecting O4 atom to a position farther away resulting in diminished O–Al–O angles of 76.53(3)° for O4–Al2–O4 and 78.13(2)° for O1–Al2–O4 instead of 90° required for a regular octahedron. All Al–O distances in the range between 1.8439(6) and 2.0086(5) Å and the average values of 1.90 and 1.92 Å are in accordance with data from the literature [20], [21], [22]. Likewise, the B–O distances (Ø = 1.49 Å) and tetrahedral angles (Ø = 109.5°) are as expected [23]. All interatomic distances and angles are given in Tables 4–6. To support the crystal structure solution, the Madelung part of lattice energy (MAPLE value) [24], [25] of HP-Al2B3O7(OH) was calculated and compared to the stoichiometric sum of the compounds α-Al2O3, HP-B2O3, and H3BO3. The results showed only a minimal deviation of 0.15% (see Table 7). The bond valence sums (BVS) for HP-Al2B3O7(OH) were calculated with the bond-length/bond-strength concept [26] and the charge distribution was derived by using the charge distribution in solids (CHARDI) concept [27]. The results are listed in Table 8. Both calculations led to reasonable values with only one noticeable BVS deviation for O4 being the donor of the hydrogen bond.

Building blocks in HP-Al2B3O7(OH).
a) BB1: a cluster of four edge-sharing AlO6 octahedra; b) BB2: an interlinked ring-formation of 12 corner-sharing BO4 tetrahedra.

Two BB2 building blocks consisting of BO4 tetrahedra fully enclose one BB1 Al4O16 cluster.

Structural buildup of HP-Al2B3O7(OH): Alternatingly oriented Al4O16 clusters (BB1) are connected through endless chains of BO4 tetrahedra along c. For reasons of clarity, the protons and the interlinking BO4 tetrahedra forming the vierer rings in BB2 are omitted.

Hydrogen bonding situation in HP-Al2B3O7(OH).
Interatomic distances (Å) in HP-Al2B3O7(OH) (standard deviations in parentheses).
Al1– | O4 | 1.8439(6) | B1– | O1 | 1.4234(7) |
O5 | 1.8696(6) | O3 | 1.5017(8) | ||
O1 | 1.8900(5) 2× | O2 | 1.5112(7) | ||
O2 | 2.0086(5) 2× | O5 | 1.5152(7) | ||
Ø Al1–O | 1.92 | Ø B1–O | 1.49 | ||
Al2– | O3 | 1.8726(5) 2× | B2– | O2 | 1.4880(7) 2× |
O1 | 1.8715(5) 2× | O3 | 1.4919(7) 2× | ||
O4 | 1.9486(4) 2× | Ø B2–O | 1.49 | ||
Ø Al2–O | 1.90 |
Interatomic angles (deg) in HP-Al2B3O7(OH) (standard deviations in parentheses).
O4–Al1–O1 | 80.32(2) 2× | O4–Al2–O4 | 76.53(3) |
O2–Al1–O2 | 85.11(3) | O1–Al2–O4 | 78.13(2) 2× |
O1–Al1–O1 | 85.55(3) | O1–Al2–O4 | 89.93(2) 2× |
O5–Al1–O2 | 87.04(2) 2× | O3–Al2–O4 | 92.61(2) 2× |
O4–Al1–O2 | 89.98(2) 2× | O3–Al2–O1 | 94.37(2) 2× |
O1–Al1–O2 | 93.8(2) 2× | O3–Al2–O1 | 95.45(2) 2× |
O5–Al1–O1 | 102.6(2) 2× | O3–Al2–O3 | 98.9(3) |
Ø O–Al1–O90 | 89.9 | Ø O–Al2–O90 | 89.7 |
O1–Al1–O2 | 170.25(2) 2× | O1–Al2–O1 | 164.87(3) |
O4–Al1–O5 | 175.96(3) | O3–Al2–O4 | 167.26(2) 2× |
Ø O–Al1–O180 | 172.2 | Ø O–AL2–O180 | 166.5 |
O3–B1–O5 | 106.53(4) | O3–B2–O2 | 107.23(2) 2× |
O2–B1–O5 | 108.28(4) | O3–B2–O3 | 108.76(6) |
O1–B1–O5 | 108.9(4) | O2–B2–O2 | 110.74(6) |
O3–B1–O2 | 110.2(4) | O3–B2–O2 | 111.45(2) 2× |
O1–B1–O2 | 110.58(4) | Ø O–B3–O | 109.5 |
O1–B1–O3 | 112.19(4) | ||
Ø O–B1–O | 109.5 |
Hydrogen bonds (Å, deg) in HP-Al2B3O7(OH) (standard deviations in parentheses).
d(D–H) | d(H···A) | d(D···A) | ∠(D–H–A) | |
---|---|---|---|---|
O4–H4···O2 2× | 0.94(3) | 2.01(2) | 2.6793(7) | 127(2) |
O4–H4···O5 | 0.94(3) | 1.90(3) | 2.6091(8) | 131(2) |
Comparison of the calculated MAPLE value (Madelung Part of Lattice Energy) of HP-Al2B3O7(OH) and the MAPLE value received from α-Al2O3 [28], HP-B2O3 [29], and H3BO3 [30] according to the reaction equation: 3 Al2O3 + 4 B2O3 + H3BO3 → 3 HP-Al2B3O7(OH).
Calculated MAPLE value for HP-Al2B3O7(OH), kJ mol−1 | 161077 |
Calculated MAPLE value based on α-Al2O3, HP-B2O3, and H3BO3, kJ mol−1 | 161316 |
Difference, % | 0.15 |
Charge distribution in HP-Al2B3O7(OH), calculated with the bond-length/bond-strength (V) and CHARDI (Q) concept.
Al1 | Al2 | B1 | B2 | O1 | O2 | O3 | O4 | O5 | H4 | |
---|---|---|---|---|---|---|---|---|---|---|
V | 2.96 | 3.10 | 2.93 | 2.90 | −1.94 | −1.84 | −1.97 | −2.34 | −1.97 | 1.01 |
Q | 2.98 | 2.94 | 3.01 | 3.09 | −2.04 | −1.99 | −2.11 | −1.90 | −2.04 | 0.98 |
3.2 Vibrational spectroscopy
For the isotypic compound Ga2B3O7(OH) [7], the IR and Raman bands were assigned based on DFT calculations and therefore provide a good basis for the interpretation of the spectra of HP-Al2B3O7(OH). The FT-IR spectrum of an HP-Al2B3O7(OH) single crystal is shown in Figure 6. In the fingerprint region up to ∼1200 cm−1, at the lower wavenumbers presumably O–Al–O bending as well as Al–O stretching vibrations contribute to the spectral pattern, followed mainly by O–B–O bending and B–O stretching vibrations. As in Ga2B3O7(OH), the O–H stretching vibrations give a distinct signal at about 3500 cm−1. Unlike the Ga containing analog, the single crystal of HP-Al2B3O7(OH) does not show any contamination with organic compounds.

IR spectrum of an HP-Al2B3O7(OH) single crystal in the range of 600–4000 cm−1.
The Raman spectrum is presented in Figure 7 and shows only the frequency range between 200 and 1500 cm−1 due to strong luminescence at higher Raman shifts that becomes even more pronounced when exciting with a red (633 nm) laser, so that OH-related vibrations (expected between 3000 and 4000 cm−1) are not displayed. The DFT calculations for the isotypic compound Ga2B3O7(OH) also allow assignments for the aluminum analog. The Al3+ = Ga3+ exchange mainly results in higher vibrational frequencies and consequently increasing Raman shifts of Al-related bands if compared to the respective Ga-related counterparts. Bands in the 250–700 cm−1 frequency range can be mainly assigned to vibrations of the AlO6 units. The Al–O–H, Al–O–Al bending vibrations are expected to appear in a wide range between 250 and 1150 cm−1, while the stretching vibrations for the BO4 tetrahedra appear in the 800–1150 cm−1 region.

Raman spectrum of an HP-Al2B3O7(OH) single crystal in the range of 200–1500 cm−1. The following band positions can be assigned to the HP-Al2B3O7(OH) structure (cm−1): 251, 266, 277, 303, 316, 321, 339, 350, 364, 412, 429, 438, 492, 501, 539, 582, 607, 616, 650, 674, 683, 702, 720, 749, 772, 788, 817, 833, 871, 899, 969, 1023, 1059, 1104, 1136, 1098.
4 Conclusion
The present report presents a detailed synthesis protocol and the determination and discussion of the crystal structure of the new aluminum borate HP-Al2B3O7(OH), which is isotypic to Ga2B3O7(OH). While the proton position in the gallium compound had to be calculated geometrically, it was possible to locate it in the electron density map and refine it via difference Fourier analysis of the aluminum analog. IR data additionally confirm the presence of the hydroxyl group. Attempts to synthesize the isotypic indium compound have not been successful yet.
Dedicated to: Professor Robert Glaum on the occasion of his 60th birthday.
Acknowledgements
We thank Ass.-Prof. Dr. Klaus Wurst for the recording of the single-crystal data, Freia Ruegenberg for the IR spectrometric data, and Univ.-Prof. Dr. Roland Stalder for granting us access to the single-crystal IR spectrometer.
Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
Research funding: None declared.
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
1. Rodellas, C., Garcia-Blanco, S. Z. Kristallogr. 1983, 165, 255–260; https://doi.org/10.1524/zkri.1983.165.1-4.255.Search in Google Scholar
2. Vitzthum, D., Wurst, K., Pann, J. M., Brüggeller, P., Seibald, M., Huppertz, H. Angew. Chem. Int. Ed. 2018, 57, 11451–11455; https://doi.org/10.1002/anie.201804083.Search in Google Scholar
3. Vitzthum, D., Wimmer, D. S., Widmann, I., Huppertz, H. Z. Naturforsch. 2020, 75b, 605–613; https://doi.org/10.1515/znb-2020-0075.Search in Google Scholar
4. Ju, J., Lin, J., Li, G., Yang, T., Li, H., Liao, F., Loong, C.-K., You, L. Angew. Chem. Int. Ed. 2003, 42, 5607–5610; https://doi.org/10.1002/anie.200352263.Search in Google Scholar
5. Yang, T., Bartoszewicz, A., Ju, J., Sun, J., Liu, Z., Zou, X., Wang, Y., Li, G., Liao, F., Martín‐Matute, B., Lin, J. Angew. Chem. Int. Ed. 2011, 50, 12555–12558; https://doi.org/10.1002/anie.201106310.Search in Google Scholar
6. Yang, T., Ju, J., Li, G., Liao, F., Zou, X., Deng, F., Chen, L., Wang, Y., Lin, J. Inorg. Chem. 2007, 46, 4772–4774; https://doi.org/10.1021/ic070073k.Search in Google Scholar
7. Vitzthum, D., Schauperl, M., Strabler, C. M., Brüggeller, P., Liedl, K. R., Griesser, U. J., Huppertz, H. Inorg. Chem. 2016, 55, 676–681; https://doi.org/10.1021/acs.inorgchem.5b02027.Search in Google Scholar
8. Huppertz, H. Z. Kristallogr. 2004, 219, 330–338; https://doi.org/10.1524/zkri.219.6.330.34633.Search in Google Scholar
9. Walker, D., Carpenter, M. A., Hitch, C. M. Am. Mineral. 1990, 75, 1020–1028.Search in Google Scholar
10. Walker, D. Am. Mineral. 1991, 76, 1092–1100.10.1007/978-1-4615-3968-1_10Search in Google Scholar
11. Sheldrick, G. M. Shelxs 2013/1, Program for the Solution of Crystal Structures; University of Göttingen: Göttingen (Germany), 2013.Search in Google Scholar
12. Sheldrick, G. M. Acta Crystallogr. 2008, A64, 112–122; https://doi.org/10.1107/s0108767307043930.Search in Google Scholar
13. Sheldrick, G. M. Acta Crystallogr. 2015, C71, 3–8. https://doi.org/10.1107/S2053229614024218.Search in Google Scholar
14. Farrugia, L. J. J. Appl. Crystallogr. 2012, 45, 849–854; https://doi.org/10.1107/s0021889812029111.Search in Google Scholar
15. Gelato, L., Parthé, E. J. Appl. Crystallogr. 1987, 20, 139–143; https://doi.org/10.1107/s0021889887086965.Search in Google Scholar
16. Spek, A. J. Appl. Crystallogr. 2003, 36, 7–13; https://doi.org/10.1107/s0021889802022112.Search in Google Scholar
17. Opus (version 6.5); Bruker Optik GmbH: Ettlingen (Germany), 2007.Search in Google Scholar
18. LABSPEC (version 5.93.20); HORIBA Jobin Yvon GmbH: Bensheim (Germany), 2005.Search in Google Scholar
19. Liebau, F. Structural Chemistry of Silicates; Springer-Verlag: Berlin, 1985.10.1007/978-3-642-50076-3Search in Google Scholar
20. Fischer, R. X., Kahlenberg, V., Voll, D., MacKenzie, K. J., Smith, M. E., Schnetger, B., Brumsack, H.-J., Schneider, H. Am. Mineral. 2008, 93, 918–927; https://doi.org/10.2138/am.2008.2744.Search in Google Scholar
21. Ju, J., Yang, T., Li, G., Liao, F., Wang, Y., You, L., Lin, J. Chem. Eur J. 2004, 10, 3901–3906; https://doi.org/10.1002/chem.200400066.Search in Google Scholar
22. Vegas, A., Cano, F., García-Blanco, S. Acta Crystallogr. 1977, B33, 3607–3609; https://doi.org/10.1107/s0567740877011650.Search in Google Scholar
23. Zobetz, E. Z. Kristallogr. 1990, 191, 45–57; https://doi.org/10.1524/zkri.1990.191.1-2.45.Search in Google Scholar
24. Hübenthal, R. Maple (version 4). Program for the Calculation of MAPLE Values; University of Gießen: Gießen (Germany), 1993.Search in Google Scholar
25. Hoppe, R. Angew. Chem., Int. Ed. Engl. 1966, 5, 95–106; https://doi.org/10.1002/anie.196600951.Search in Google Scholar
26. Brown, I. D., Altermatt, D. Acta Crystallogr. 1985, B41, 244–247; https://doi.org/10.1107/s0108768185002063.Search in Google Scholar
27. Hoppe, R. Z. Kristallogr. 1979, 150, 23–52; https://doi.org/10.1524/zkri.1979.150.1-4.23.Search in Google Scholar
28. Kondo, S., Tateishi, K., Ishizawa, N. Jpn. J. Appl. Phys. 2008, 47, 616; https://doi.org/10.1143/jjap.47.616.Search in Google Scholar
29. Prewitt, C. T., Shannon, R. D. Acta Crystallogr. 1968, B24, 869–874; https://doi.org/10.1107/s0567740868003304.Search in Google Scholar
30. Zachariasen, W. H. Acta Crystallogr. 1954, 7, 305–310; https://doi.org/10.1107/s0365110x54000886.Search in Google Scholar
© 2020 Daniela Vitzthum 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