Skip to content
Publicly Available Published by De Gruyter January 4, 2020

The UV-phosphor strontium fluorooxoborate Sr[B5O7F3]:Eu

  • Martin J. Schäfer , Stephan G. Jantz and Henning A. Höppe EMAIL logo

Abstract

The fluorooxoborates Sr[B5O7F3] (Ccm21 (no. 36), Z = 4, a = 864.777(3) pm, b = 1001.037(4) pm, c = 810.110(5) pm, 630 refl., 68 param., RF = 0.061, RBragg = 0.051) and Sr[B5O7F3]:Eu2+ were synthesized via solid-state reactions from Sr(BF4)2 and B2O3. Doping was achieved by adding EuF3, which was reduced to its divalent state during the synthesis. Sr[B5O7F3] was investigated by infrared spectroscopy. Photoluminescence spectra of the doped compound revealed an Eu2+ 5d-4f emission in the UV regime peaking at 371 nm with a full width at half maximum (FWHM) of 21 nm upon excitation at 239 nm, which proves a weak coordination of europium by the anionic fluorooxoborate network.

1 Introduction

In the course of our systematic investigations of the broad field of silicate-analogous materials [1], [2], [3] we also came across compounds comprising BO3F tetrahedra, i.e. fluorooxoborates; systems generated through the formal addition of fluoride anions to boron compounds comprising BO3 triangles are of interest in applications as host materials for luminescence or as NLO materials [4], [5], [6]. Surprisingly, one of these, Sn[B2O3F2], represents as precursor for borates not accessible otherwise, like the first crystalline tin borate Sn3[B4O9] [2]. Such host structures should provide significantly weaker coordination strength, i.e. smaller ligand fields than the respective oxides. Accordingly, we substituted Eu2+ into our new and apparently very weakly coordinating barium fluorooxoborate Ba[B4O6F2] [6], in which europium is situated on the barium sites. Divalent europium normally causes broad-band emissions due to parity allowed 5d-4f emissions; contrarily to the 4f, the 5d states are subject to nephelauxetic and ligand field effects, and thus the recorded emission strongly depends on the nature of the ligand atoms. Therefore emissions from the deep red to the deep UV can be caused by strongly and weakly coordinating host structures, respectively. In case of the aforementioned Ba[B4O6F2]:Eu we observed a rarely seen narrow-banded 4f-4f emission (6P→8S) at 359 nm at room temperature on top of the commonly known broad 5d-4f band peaking around 366 nm; such transitions had previously only been observed for pure fluorides. Such emissions in the deep UV are of interest for water disinfection purposes. Fluorooxoborates are less reactive than metal fluorides against glass and other materials. The sensibility of divalent europium towards nephelauxetic and ligand field changes can be tuned by just replacing the alkaline earth host by a larger or smaller element allowing for less or more direct influence of the ligand atoms on the 5d states.

For the work described in this contribution we chose divalent strontium, which is often used in host materials for Eu2+ due to the same size and charge, both being beneficial for efficient doping. Our efforts yielded Sr[B5O7F3], the crystal structure of which had been published [7], [8] just before our own synthesis was successful. Here we present our own synthetic approach to this fluorooxoborate. Its crystal structure was re-investigated by means of powder X-ray diffraction and its luminescence properties after doping with Eu2+ were investigated.

2 Results and discussion

2.1 Crystal structure

Sr[B5O7F3] crystallises in the same structure type as its calcium counterpart, which was published as Ca2[B10O14F6] with a doubled sum formula but half the formula units per unit cell. The title compound adopts the non-centrosymmetric space group Ccm21. The fundamental building block (FBB) is shown in Fig. 1 and consists of two trigonal planar BO3 units and three BO3F tetrahedra; two tetrahedra and one BO3 triangle form a six-membered ring. Two of those rings are condensed via a common edge and thus the FBB can be described according to Burns et al. [9] by the descriptor 2∆3□:[Φ]<∆2□>|<∆2□>|.

Fig. 1: FBB of Sr[B5O7F3], featuring the slightly bent bicyclic structure. Boron atoms in green, oxygen atoms in red, fluorine atoms in yellow.
Fig. 1:

FBB of Sr[B5O7F3], featuring the slightly bent bicyclic structure. Boron atoms in green, oxygen atoms in red, fluorine atoms in yellow.

The FBBs are connected via common O31[2] oxygen atoms bridging two adjacent boron atoms to form layers in the (100) plane. Thus, four FBBs form an 18-membered ring, in which the strontium atoms are situated. All fluorine atoms point out of the planes (Fig. 2). The atoms O12, B2 and F2 lie on a mirror plane together with the shared edge. Table 1 summarizes selected interatomic distances and angles.

Fig. 2: Layers of Sr[B5O7F3] viewed along the [010] direction (upper picture) and the [001] direction (lower picture). Strontium atoms in grey, boron atoms in green, oxygen atoms in red, fluorine atoms in yellow.
Fig. 2:

Layers of Sr[B5O7F3] viewed along the [010] direction (upper picture) and the [001] direction (lower picture). Strontium atoms in grey, boron atoms in green, oxygen atoms in red, fluorine atoms in yellow.

Table 1:

Selected interatomic distances and angles in Sr[B5O7F3]; standard deviations in parentheses; B represents trigonal planar coordinated and B represents tetrahedrally coordinated boron atoms.

AtomsCoordination numberDistance/pmØ/pm
Sr–O6(O)+3(F)258.4(6)–261.6(4)259(2)
Sr–F239.4(5)–261.6(4)252(9)
B–O3134.6(12)–141.5(11)137(3)
B–O3(O)+1(F)138.4(11)–157.4(12)148(6)
B–F139.1(15)–141.5(10)140(1)
O12–B13142.6(17)
O12–B2157.4(12)
Angle/degØ/deg
O–B–O116.4(9)–122.0(9)119.8(3)
O–B–O102.0(8)–112.6(1.1)109.6(4)
O–B–F105.3(8)–113.5(1.1)109.2(3)
B–O12–B116.8(11)–118.7(13)118.1(9)

An unusual feature of the structure is the oxygen atom O12[3] connecting three adjacent BO3F tetrahedra within the FBB. The distances between O12[3] and the neighbouring boron atoms B1 and B2 are 157.4(17) and 142.6(17) pm, respectively, and thus considerably larger than usual B–O[2] bonds (see caption to Table 1 for definition of B and B∆) which are usually in the range from 145 to 147 pm, as observed in Sn[B2O3F2] and Ba[B4O6F2] [5], [6]. Such rarely found three-fold-coordinated oxygen atoms O[3] were reported before in borates like Ni3B18O28(OH)4·H2O or LiBa3(OH)[B9O16][B(OH)4] [10], [11], and in alumosilicates like Be3Al2Si6O18 or SiAl2O5 [12], [13], [14]. The average B–O bond lengths in the title compound of 137 pm (B∆–O) and 148 pm (B–O) agree well with the sum of the ionic radii [15]. Yet the variance of the bond lengths is slightly larger, compared with other, similar compounds. This variance, as well as the elongated bonds with the O12 atom are also reflected in the slightly larger but still uncritical deviations of the tetrahedral units from a regular tetrahedron, calculated according to Balic-Žunic and Makovicky [16], [17] to be 0.68% (B1) and 0.27% (B2).

The strontium cation is situated within the layers of the anionic substructure, in a ring built up by four B5O7F32− FBBs. Its nine-fold coordination consists of six oxygen atoms of the aforementioned ring and by three fluorine atoms of the layers above and below. The coordination polyhedron can be described as a strongly distorted triangular cupola (Johnson solid J3). The Sr–O distances range from 258 to 262 pm, which is slightly shorter than the sum of their ionic radii (267 pm) [15]. The Sr–F distances are quite different, as the different fluorine anions F1 and F2 coordinate the strontium cation from different layers. Both distances (239 pm for Sr–F1, 256 pm for Sr–F2) are shorter than the sum of their ionic radii (260 pm).

2.2 Infrared Spectroscopy

The IR spectrum of Sr[B5O7F3] was recorded in the range from 4000 to 400 cm−1 (Fig. 3). The inset shows the whole measured area with only very weak bands caused by traces of adhesive moisture beyond 3000 cm−1. The assignments listed in Table 2 are based on the infrared spectrum of Ba[B4O6F2] and confirmed by DFT calculations, as well as on other reference data [18], [19]. As expected, considering the comparable structural units (BO3 and BO3F), the spectra are similar. The bands around 1500–1275 cm−1 and the band near 960 cm−1 are caused by the asymmetric and symmetric stretching modes of the trigonal BO3 units. According to our calculations on comparable infrared spectra [5], [6] the asymmetric and symmetric stretching modes of the B–O bonds of tetrahedral BO3F units are found in the regimes 1100–980 and 960–850 cm−1, respectively. Characteristic stretching modes of B–F bonds in fluorooxoborates are found in the regimes 1000–920 and 860–730 cm−1, deformation modes of BO3F tetrahedra typically lie around 860 cm−1 for the similar Ba[B4O6F2] [6]. Based on these regimes we tentatively assigned the experimentally observed modes of the title compound according to Table 2. Aside from a slight shift in wavenumbers for most bands, the most obvious difference is an additional band around 1220 cm−1 in the spectrum of Sr[B5O7F3]. This was also reported by Mutailipu et al. [7], but was not explained further. We suspect that this band is caused by B–O stretching vibrations including the three-fold-coordinated O12[3] as it does not appear in Ba[B4O6F2]. A B–O[3] bond is presumably weaker than a B–O[2] bond causing a shift to smaller wavenumbers from the standard regime. The presence of an infrared band in this range in other borates featuring three-fold-coordinated oxygen [10], [20], [21], [22] supports our tentative assignment. A respective asymmetric mode is to be expected at lower frequencies, maybe the sharp vibration around 720 cm−1 might fit – but this is rather speculative as the spectrum is very busy in this regime.

Fig. 3: IR spectrum of Sr[B5O7F3]; the inset shows the complete region up to 4000 cm−1.
Fig. 3:

IR spectrum of Sr[B5O7F3]; the inset shows the complete region up to 4000 cm1.

Table 2:

IR bands of Sr[B5O7F3], as well as their assignments, using data from Ba[B4O6F2] [6].

IR/cm−1Assignment
1500–1275νas(BO3)
1220ν(OB3)
1063νas(B–O) in BO3F
1027νas(B–O) in BO3F
992ν(B–F) in BO3F
960νs(BO3)
880δ(BO3F), νs(B–O), ν(B–F) in BO3F (?)
848νs(B–O), ν(B–F) in BO3F

2.3 Fluorescence spectroscopy on Sr[B5O7F3]:Eu2+

The fluorescence spectrum of Sr[B5O7F3]:Eu2+ (Fig. 4) features an emission with a single 5d-4f transition peaking at 372 nm with a full width at half maximum (FWHM) of 21 nm. The excitation spectrum comprises numerous peaks assigned to 4f-5d transitions. Below 370 nm, 4f-4f transitions of Eu2+ can be excited as well [23], [24], but they are hard to identify since the resolution of the spectrum is limited by the employed deuterium lamp. The width of all excitation peaks remains equal, even at lower temperatures, as shown in Fig. 5.

Fig. 4: Photoluminescence spectrum of Sr[B5O7F3]:Eu2+; emission spectrum (violet line) measured at λexc=239 nm, excitation spectrum (black line) measured at λem=372 nm.
Fig. 4:

Photoluminescence spectrum of Sr[B5O7F3]:Eu2+; emission spectrum (violet line) measured at λexc=239 nm, excitation spectrum (black line) measured at λem=372 nm.

Fig. 5: Photoluminescence spectrum of Sr[B5O7F3]:Eu2+; emission and excitation at liquid nitrogen temperature (blue line) and room temperature (grey line); the spectra were measured with λexc=239 nm and λem=372 nm, respectively.
Fig. 5:

Photoluminescence spectrum of Sr[B5O7F3]:Eu2+; emission and excitation at liquid nitrogen temperature (blue line) and room temperature (grey line); the spectra were measured with λexc=239 nm and λem=372 nm, respectively.

Compared to the fluorescence spectrum of Ba[B4O6F2]:Eu2+, the emission spectrum of Sr[B5O7F3]:Eu2+ shows no sharp peak which could be attributed to a 4f-4f transition, but the emission peak of the 5d-4f transition shows the same FWHM; this is reasonable as both host structures are quite similar. Both structures comprise layered polymeric fluorooxoborate anions; the cations situated in the voids of these layers are weakly coordinated by oxygen and fluorine atoms with large interatomic distances due to rather large coordination numbers (Sr[B5O7F3]: 9, Ba[B4O6F2]: 13). Since these distances are significantly shorter in the title compound compared to Ba[B4O6F2], it is not surprising that the emission wavelength is slightly red-shifted from 366 to 372 nm and that the 4f-4f transition cannot be observed in Sr[B5O7F3]:Eu2+ due to efficient relaxation from the 6P (4f) to the 5d states. Nevertheless, the ligand field splitting and nephelauxetic effect of the 5d states in Eu2+ is still small enough that the emission of the d-f transition lies in the UV regime of the electromagnetic spectrum.

3 Conclusion

The effects observed for Sr[B5O7F3]:Eu2+ fit into the range of luminescence emissions which can be realised by Eu2+-doped borates. With decreasing influence of the ligands introduced in the borate framework, borates are able to generate red (e.g. Ba2Mg(BO3)2:Eu2+), green (e.g. Sr3(BO3)2:Eu2+), blue (e.g. BaNaB9O15:Eu2+), and violet (e.g. SrB4O7:Eu2+) [25] emissions. Perovskite-like alkaline earth fluorides [26] or weakly coordinating silicates and sulphates [27], [28] and more recently the condensed fluorooxoborates Sr[B5O7F3]:Eu2+ and Ba[B4O6F2]:Eu2+ [6] have extended that trend into the deep UV regime, up to the point where 4f-4f emissions are observed as well in few cases. Investigations on further known alkaline earth fluorooxoborates like Ca[B5O7F3] [29], Ba[BOF3] [30] Ba[B5O8F] ·H2O [31] and Ba[B2O3F2] [32] doped with Eu2+ are under way and will be presented in due course.

4 Experimental Section

4.1 Synthesis of Sr(BF4)2

The synthesis of the starting material Sr(BF4)2 was achieved according to the following equation:

SrCO3+2 HBF4Sr(BF4)2+H2O+CO2

Three grams of SrCO3 (20 mmol, Riedel-de Häen, 96% purity) were suspensed in 5 mL of demineralised water inside of a PVC container. 4.8 mL of HBF4 (38 mmol, Alfa Aesar, 50% solution). The excess of SrCO3 ensured the complete reaction of the tetrafluoroboric acid. The suspension was then filtered and the filtrate was slowly dried in vacuo, yielding phase-pure Sr(BF4)2.

4.2 Synthesis of Sr[B5O7F3] and Sr[B5O7F3]:Eu2+

The synthesis of Sr[B5O7F3] was achieved according to the following equation:

3 Sr(BF4)2+9 B2O33 Sr[B5O7F3]+2 B2O3+5 BF3

The pre-dried educts Sr(BF4)2 (255.0 mg, 0.9761 mmol) and B2O3 (160.0 mg, 2.298 mmol, Alfa Aesar, 99.999% purity) were weighed and ground in an agate mortar in air and transferred into a BN crucible. This mixture was dried at T=105°C overnight in a compartment drier to get rid of any remaining moisture. The powder was then compressed inside the crucible which was then screwed and sealed in a silica ampoule under argon atmosphere. The sample was heated to 500°C with a rate of 50 K h−1, held at that temperature for 500 h and then cooled to room temperature with a rate of 5 K h−1. 263.1 mg (based on Sr[B5O7F3]: 0.8469 mmol, 86.8% yield) of a slightly greyish, air and moisture insensitive powder were isolated.

For the synthesis of the doped sample, 1 mol% of the Sr(BF4)2 was replaced by EuF3 (ChemPur, 99.9% purity). The mixture of dry educts was ground and pressed into a pellet inside of a glovebox and then filled into a silver crucible and sealed in a silica glass ampule under vacuum (2.5×10−3 mbar). The sample was heated to 500°C with a rate of 50 K h−1, held at that temperature for 50 h and then cooled to 400°C in a first step and then to room temperature with the rates of 1 K h−1 and 40 K h−1, respectively. The product was used for fluorescence spectroscopy without any further purification.

4.3 Spectroscopic methods

IR spectra were recorded on a Bruker EQUINOX 55 FT-IR spectrometer equipped with a Platinum ATR unit in the range 4000–400 cm−1 with a resolution of 4 cm−1 and 32 scans. Fluorescence excitation and emission spectra were recorded on a Horiba Fluoromax-4 spectrometer, scanning a range from 200 to 800 nm.

4.4 X-ray structure determination

The structure of Sr[B5O7F3] was solved from powder X-ray diffraction data. These data were collected on a Bruker D8 Advance capillary diffractometer in the range of 5.000≤2θ≤119.966°, with steps of 0.02° and an irradiation of 36 s per step. The structure was solved by Direct Methods using the program Expo2014 [33], further refinement was conducted with the program package Fullprof [34] with the graphical interface WinPLOTR [35]. During the structure refinement, all atom positions were refined without any restraints. After some refinement cycles, the z position of the strontium atom was fixed to fix the structure along [001]. The XRD pattern including the refinement is portrayed in Fig. 6, the results of the structure refinement are listed in Tables 3 and 4.

Fig. 6: Observed (red dots) and calculated (black line) powder diffraction pattern of Sr[B5O7F3], as well as the difference plot (blue line); the green vertical lines correspond to the possible Bragg-peak positions of Sr[B5O7F3]; the only peak of a possible side phase (SrF2) is labelled with a blue arrow.
Fig. 6:

Observed (red dots) and calculated (black line) powder diffraction pattern of Sr[B5O7F3], as well as the difference plot (blue line); the green vertical lines correspond to the possible Bragg-peak positions of Sr[B5O7F3]; the only peak of a possible side phase (SrF2) is labelled with a blue arrow.

Table 3:

Parameters of the crystal structure determination of Sr[B5O7F3] (standard deviations in parentheses).

Sum formulaSr[B5O7F3]
Molar mass/g·mol−1310.67
Crystal systemOrthorhombic
Space groupCcm21
a/pm864.777(3)
b/pm1001.037(4)
c/pm810.110(5)
Volume/×106 pm3701.291(5)
Z4
ρX-ray/g cm−32.94
DiffractometerBruker D8 Advance
Radiation | λ/nmCu1,2 | 0.154056, 0.154439
Temperature/K298(2)
2θ range/deg5.000–119.966
Measured intensities630
Refined parameters68
RP0.020
RwP0.033
RF0.061
RBragg0.051
χ20.100
Table 4:

Refined atomic coordinates, Wyckoff symbols and isotropic dislocation parameters Ueq/pm2 in Sr[B5O7F3] (standard deviations in parentheses).

AtomWyckoff symbolxyzUeq
Sr14a0.25811(17)00.29644a88.7
B18b0.8525(11)0.8661(10)0.2106(12)126.7
B24a0.6740(16)00.4048(17)126.7
B38b0.7372(16)0.7669(10)0.4684(10)126.7
O124a0.8058(6)00.2990(20)126.7
O138b0.8089(4)0.7569(4)0.3196(9)126.7
O238b0.6760(8)0.8826(6)0.5223(7)126.7
O318b0.7130(7)0.6460(6)0.5551(7)126.7
F18b0.0160(4)0.8678(4)0.2037(5)126.7
F24a0.5341(6)00.3200(9)126.7
  1. aFixed parameter.

Further details of the crystal structure investigation may be obtained from Fachinformationszentrum Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany (fax: +49-7247-808-555; e-mail: crysdata@fiz-karlsruhe.de, http://www.fiz-informationsdienste.de/en/DB/icsd/depot_anforderung.html) on quoting the deposition number CSD-1962709.


Dedicated to: Professor Arndt Simon on the occasion of his 80th birthday.


References

[1] P. Netzsch, P. Gross, H. Takahashi, S. Lotfi, J. Brgoch, H. A. Höppe, Eur. J. Inorg. Chem.2019, 3975.10.1002/ejic.201900838Search in Google Scholar

[2] M. J. Schäfer, S. G. Jantz, H. A. Höppe, Dalton Trans.2019, 48, 10398.10.1039/C9DT01901DSearch in Google Scholar

[3] P. Netzsch, P. Gross, H. Takahashi, H. A. Höppe, Inorg. Chem.2018, 57, 8530.10.1021/acs.inorgchem.8b01234Search in Google Scholar PubMed

[4] S. G. Jantz, L. van Wüllen, A. Fischer, E. Libowitzky, E. J. Baran, M. Weil, H. A. Höppe, Eur. J. Inorg. Chem.2016, 1121.10.1002/ejic.201501143Search in Google Scholar

[5] S. G. Jantz, M. Dialer, L. Bayarjargal, B. Winkler, L. van Wüllen, F. Pielnhofer, J. Brgoch, R. Weihrich, H. A. Höppe, Adv. Opt. Mater.2018, 1800497.10.1002/adom.201800497Search in Google Scholar

[6] S. G. Jantz, F. Pielnhofer, L. van Wüllen, R. Weihrich, M. J. Schäfer, H. A. Höppe, Chem. Eur. J.2018, 24, 443.10.1002/chem.201704324Search in Google Scholar PubMed

[7] M. Mutailipu, M. Zhang, B. Zhang, L. Wang, Z. Yang, X. Zhou, S. Pan, Angew. Chem. Int. Ed.2018, 57, 6095.10.1002/anie.201802058Search in Google Scholar PubMed

[8] M. Luo, F. Liang, Y. Song, D. Zhao, F. Xu, N. Ye, Z. Lin, J. Am. Chem. Soc.2018, 140, 3884.10.1021/jacs.8b01263Search in Google Scholar PubMed

[9] P. C. Burns, J. D. Grice, F. C. Hawthorne, Can. Mineral.1995, 33, 1131.Search in Google Scholar

[10] M. K. Schmitt, O. Janka, R. Pöttgen, K. Wurst, H. Huppertz, Eur. J. Inorg. Chem.2017, 3508.10.1002/ejic.201700464Search in Google Scholar

[11] Q. Wei, J.-J. Wang, C. He, J.-W. Cheng, G.-Y. Yang, Chem. Eur. J.2016, 22, 10759.10.1002/chem.201602135Search in Google Scholar PubMed

[12] L. A. Blaginina, A. F. Zatsepin, I. A. Dmitriev, Fiz. Khim. Stekla1987, 19, 398.Search in Google Scholar

[13] C. W. Burnham, Z. Kristallogr.1963, 118, 337.10.1524/zkri.1963.118.5-6.337Search in Google Scholar

[14] J. B. Burt, N. L. Ross, R. J. Angel, M. Koch, Am. Mineral.2006, 91, 319.10.2138/am.2006.1875Search in Google Scholar

[15] R. D. Shannon, Acta Crystallogr.1976, A32, 751.10.1107/S0567739476001551Search in Google Scholar

[16] T. Balic-Žunic, E. Makovicky, Acta Crystallogr.1996, B52, 78.10.1107/S0108768195008251Search in Google Scholar

[17] E. Makovicky, T. Balic-Žunic, Acta Crystallogr.1998, B54, 766.10.1107/S0108768198003905Search in Google Scholar

[18] J. Weidlein, U. Müller, K. Dehnicke, Schwingungsfrequenzen I, Hauptgruppenelemente, 1st edition, Georg Thieme Verlag, Stuttgart, New York 1981.Search in Google Scholar

[19] K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination Compounds, Part A, 6th edition, John Wiley and Sons Inc., Hoboken 2009.10.1002/9780470405888Search in Google Scholar

[20] G. Sohr, V. Falkowski, M. Schauperl, K. R. Liedl, H. Huppertz, Eur. J. Inorg. Chem.2015, 527.10.1002/ejic.201403006Search in Google Scholar

[21] D. A. Köse, B. Zümreoglu-Karan, T. Hökelek, E. Sahin, Z. Anorg. Allg. Chem.2009, 635, 563.10.1002/zaac.200801288Search in Google Scholar

[22] G. Sohr, D. M. Többens, J. Schmedt auf der Günne, H. Huppertz, Chem. Eur. J.2014, 20, 17059.10.1002/chem.201404018Search in Google Scholar

[23] P. Dorenbos, J. Lumin.2003, 104, 239.10.1016/S0022-2313(03)00078-4Search in Google Scholar

[24] P. Dorenbos, J. Phys.: Condens. Matter2003, 15, 575.10.1088/0953-8984/15/3/322Search in Google Scholar

[25] A. Diaz, D. A. Keszler, Mater. Res. Bull.1996, 31, 2, 105.10.1016/0025-5408(95)00182-4Search in Google Scholar

[26] R. A. Hewes, M. V. Hoffman, J. Lumin.1971, 261.10.1016/0022-2313(71)90064-0Search in Google Scholar

[27] F. M. Ryna, W. Lehmann, D. W. Feldman, J. Murphy, J. Electrochem. Soc.1974, 121, 11, 1475.Search in Google Scholar

[28] J. M. P. J. Verstegen, J. L. Sommerdijk, J. Lumin.1974, 9, 297.10.1016/0022-2313(74)90042-8Search in Google Scholar

[29] Z. Zhang, Y. Wang, B. Zhang, Z. Yang, S. Pan, Inorg. Chem.2018, 57, 9, 4820.Search in Google Scholar

[30] D. Jaing, Y. Wang, H. Li, Z. Yang, S. Pan, Dalton Trans. 2018, 47, 5157.10.1039/C8DT00403JSearch in Google Scholar

[31] C. Huang, G. Han, H. Li, F. Zhang, Z. Yang, S. Pan, Dalton Trans.2019, 48, 6714.10.1039/C9DT00106ASearch in Google Scholar

[32] C. Huang, F. Zhang, H. Li, Z. Yang, H. Yu, S. Pan, Chem. Eur. J.2019, 25, 6693.10.1002/chem.201806350Search in Google Scholar

[33] A. Altomare, C. Cuocci, C. Giacovazzo, A. Moliterni, R. Rizzi, N. Corriero, A. Falcicchio, J. Appl. Crystallogr.2013, 46, 1231.10.1107/S0021889813013113Search in Google Scholar

[34] J. Rodríguez-Carvajal, Phys. B1993, 192, 55.10.1016/0921-4526(93)90108-ISearch in Google Scholar

[35] T. Roisnel, J. Rodríguez-Carvajal, WinPLOTR, A Windows Tool for Powder Diffraction Pattern Analysis; see: T. Roisnel, J. Rodríguez-Carvajal, Mater. Sci. Forum2001, 378–3, 118.10.4028/www.scientific.net/MSF.378-381.118Search in Google Scholar

Received: 2019-10-31
Accepted: 2019-11-19
Published Online: 2020-01-04
Published in Print: 2020-02-25

©2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2019-0174/html
Scroll to top button