Home Synthesis and characterization of the alkali borate-nitrates Na3–x Kx[B6O10]NO3 (x=0.5, 0.6, 0.7)
Article Publicly Available

Synthesis and characterization of the alkali borate-nitrates Na3–x Kx[B6O10]NO3 (x=0.5, 0.6, 0.7)

  • Teresa S. Ortner , Daniel Schildhammer , Martina Tribus , Bastian Joachim and Hubert Huppertz EMAIL logo
Published/Copyright: February 15, 2017
Become an author with De Gruyter Brill

Abstract

Three novel mixed alkali borate-nitrates Na3−x Kx[B6O10]NO3 (x=0.5, 0.6, 0.7) were synthesized hydrothermally; their crystal structures were determined through Rietveld analyses, and supported through EDX as well as vibrational spectroscopy. The phases represent solid solutions of the alkali borate-nitrate Na3(NO3)[B6O10], which was reported in 2002 as a “New type of boron-oxygen framework in the Na3(NO3)[B6O10] crystal structure” (O. V. Yakubovich, I. V. Perevoznikova, O. V. Dimitrova, V. S. Urusov, Dokl. Phys.2002, 47, 791). Only two of the three crystallographically independent Na+ positions in the new structures are partially substituted by K+; a pure potassium borate-nitrate was not formed until now. The cell parameters of the novel phases vary from a=1261.72(5)–1267.12(5), b=1004.19(5)–1007.96(4), c=770.55(3)–774.38(3) pm, and V=0.97630(6)–0.98905(6) nm3 in the orthorhombic space group Pnma (no. 62), in alignment with increasing K+ content.

1 Introduction

The structural possibilities in the solid state chemistry of borates are enormous, stemming from the ability of boron cations to coordinate to three or four oxide ions forming trigonal-planar or tetrahedral building blocks [1], [2], [3]. Nitrate anions are very similar to BO3 groups regarding their molecular geometry, however, showing very different chemical and electronic properties [4], [5]. Borate-nitrates combine borate and nitrate building blocks, paving the way to an even more diverse structural chemistry. To date, twelve different fully characterized borate-nitrates are known in the literature. In 1974, Bither and Young reported the nitratoboracites M3B7O13NO3 (M=Co, Ni, Cu, Zn, Cd) [6] obtained through high-pressure/high-temperature syntheses. Powder X-ray diffraction as well as IR data were obtained. However, X-ray single-crystal examinations still remain to be done. In 2003, Ce[B5O8(OH)]NO3·3 H2O [7] was reported, whereby a final structure solution, postulating one coordinating and two crystal water molecules Ce[B5O8(OH)(H2O)]NO3·2 H2O [8], was given in 2012. A defect variant, La[B5O8(OH)]NO3·2 H2O [9], was found in 2005. The hygroscopic phase served as a model system to gain further insight into the formation of fundamental borate building blocks [10]. In 2010, a potassium-neptunium borate-nitrate K2[(NpO2)3B10O16(OH)2(NO3)2] [11] was discovered with mixed-valent neptunium (Np4+, Np5+, and Np6+). In 2013, the first lead borate-nitrate [Pb3(B3O7)](NO3) [12] was discovered. The same group reported Pb2(BO3)(NO3) [13] with a honeycomb-like layer of [Pb2BO3] and a second harmonic generation (SHG) effect about nine times of the magnitude of conventionally used potassium dihydrogen phosphate (KDP), pertaining mainly to the nitrate groups and the lead cations. Also, for the latter borate-nitrate phase a direct band gap semiconducting effect was found. Three more lead borate-nitrates [Pb64-O)4(BO3)](NO3), H[Pb63-O)2(BO3)2](NO3)3, and H[Pb84-O)33-O)(BO3)2](NO3)3 with varying structures were also reported [14]. The first mixed alkali borate-nitrate K3Na[B6O9(OH)3]NO3 [15] was published in 2015. The thirdly reported rare earth structure is the lutetium borate-nitrate Lu2B2O5(NO3)2·2 H2O [16], which forms a hitherto unknown structural motif built up solely from pyroborate and nitrate anions, and water molecules. Our group was recently able to contribute to the structural type of Ce[B5O8(OH)(H2O)]NO3·2 H2O through the syntheses and characterization of two additional defect variants: Nd[B5O8(OH)(H2O)0.85]NO3·2 H2O and Sm[B5O8(OH)]NO3·2 H2O [17].

In 2002, the first alkali borate-nitrate with the composition Na3(NO3)[B6O10] was fully characterized [18]. Adding to the still very scarcely investigated field of rare earth borate-nitrates, we present three new substitutional variants Na3−x Kx[B6O10]NO3 with x=0.5, 0.6, and 0.7.

2 Experimental section

2.1 Hydrothermal syntheses

All three alkali borate-nitrates were synthesized hydrothermally by charging Teflon-lined stainless-steel autoclaves (8 mL) with Na2B4O7·10 H2O (p.a., Merck KGaA, Darmstadt, Germany) and KNO3 (≥99%, Carl Roth GmbH, Karlsruhe, Germany). The mixtures were ground together, heated to and kept at 513 K for 3 days, cooled slowly down to 330 K (2 K h−1), and finally quenched to room temperature. Colorless micro-crystals (approx. 5 μm in diameter and <1 μm thickness) were isolated after removing excess educts through washing the products with hot deionized water. The reported borate-nitrates were obtained in the Na2B4O7·10 H2O:KNO3 ratio ranging from 5:1 to 1:1 for Na2.5K0.5[B6O10]NO3, from 1:2 to 1:6 for Na2.4K0.6[B6O10]NO3, and from 1:10 to 1:50 for the phase richest in potassium Na2.3K0.7 [B6O10]NO3. A brief graphical overview of the molar educt ratios is given in Fig. 1.

Fig. 1: Na2B4O7·10 H2O to KNO3 educt ratios yielding Na2.5K0.5[B6O10]NO3 (green squares), Na2.4K0.6[B6O10]NO3 (red squares), and Na2.3K0.7[B6O10]NO3 (blue squares).
Fig. 1:

Na2B4O7·10 H2O to KNO3 educt ratios yielding Na2.5K0.5[B6O10]NO3 (green squares), Na2.4K0.6[B6O10]NO3 (red squares), and Na2.3K0.7[B6O10]NO3 (blue squares).

2.2 Crystal structure analyses

Powder X-ray diffraction patterns were collected with a Stoe Stadi P powder diffractometer with transmission geometry and MoKα1 (λ=70.93 pm) radiation with a focusing Ge(111) primary beam monochromator and a Mythen 1K detector (Dectris, Switzerland) in the 2θ range 2.0–50.0° with a step size of 0.005°. The crystal structure determinations were performed by Rietveld refinements using the program suite Topas 4.2 [19], [20]. Hardware parameters and reflection shape refinements were carried out with a LaB6 standard. The single-crystal structure data of the compound Na3(NO3)[B6O10] of Yakubovich et al. [18] served as starting parameters for the refinements.

All relevant details of the data collections and the refinements using the Rietveld method are listed in Table 1. Table 2 shows the positional parameters, occupation factors, and displacement parameters. Selected interatomic distances and bond angles are given in Tables 3 and 4 , respectively.

Table 1:

Crystal data obtained through the Rietveld refinement of powder diffraction patterns (standard deviations in parentheses).

Powder diffractometerSTOE Stadi P
Radiation; wavelength, pmMoKα1; λ=70.93
Temperature, K296(2)
θ range, deg2–50
Step width, deg0.005
Crystal systemorthorhombic
Space groupPnma (no. 62)
Formula units per cell, Z4
Empirical formulaNa2.5K0.5[B6O10]NO3Na2.4K0.6[B6O10]NO3Na2.3K0.7[B6O10]NO3
a, pm1261.72(5)1265.74(3)1267.12(5)
b, pm1004.19(5)1006.94(2)1007.96(4)
c, pm770.55(3)772.06(2)774.38(3)
V, nm30.97630(6)0.98400(3)0.98905(6)
Molar mass, g mol−1363.78364.91367.68
Calculated density, g cm−32.482.462.47
Absorption coefficient, cm−15.96.16.8
Number of reflections940947955
Rexp, %1.160.8731.136
Rwp, %5.7655.3674.972
Rp, %4.7884.2794.231
Table 2:

Atomic coordinates, according Wyckoff positions, and site occupation factors (S.O.F.) other than 1 for Na2.5K0.5[B6O10]NO3 with standard deviations in parentheses. The respective data for Na2.4K0.6[B6O10]NO3 and Na2.3K0.7[B6O10]NO3 are given in the Supporting Information Table S1.

AtomWyckoff sitexyzS.O.F.
Na14b1/200
Na24a0000.79(3)
K24a0000.21(3)
Na34c0.2916(8)1/40.871(2)0.72(2)
K34c0.2916(8)1/40.871(2)0.29(2)
B18d0.1674(8)0.126(2)0.277(2)
B24c0.311(2)1/40.439(2)
B34c–0.003(2)1/40.220(3)
B48d0.257(2)0.023(2)0.528(2)
N14c0.054(2)1/40.707(2)
O14c0.2135(9)1/40.342(2)
O28d0.1768(7)0.017(2)0.403(2)
O38d0.0539(6)0.1318(8)0.233(2)
O48d0.2183(8)0.100(2)0.117(2)
O58d0.3267(8)0.1289(8)0.544(2)
O64c–0.0232(9)1/40.609(2)
O74c0.8924(7)1/40.186(2)
O88d0.0806(8)0.1400(8)0.769(2)
Table 3:

Selected interatomic distances (pm) in Na2.5K0.5[B6O10]NO3 obtained from powder diffraction data refined with the Rietveld method (standard deviations in parentheses). The respective data for Na2.4K0.6[B6O10]NO3 and Na2.3K0.7[B6O10]NO3 are given in the Supporting Information Table S2.

Na1–O2 (2×)235.8(9)B1–O4141(2)B3–O3 (2×)135(2)
Na1–O3 (2×)254(2)B1–O3146(2)B3–O7139(2)
Na1–O6 (2×)266.2(5)B1–O2147(2)Ø=136(1)
Na1–O8 (2×)270(2)B1–O1148(2)
Ø=257(1)Ø=146(1)B4–O2139(2)
B4–O5140(2)
Na2/K2–O3 (2×)233(2)B2–O7140(2)B4–O4144(2)
Na2/K2–O8 (2×)248(2)B2–O1144(2)Ø=141(1)
Na2/K2–O5 (2×)256.4(9)B2–O5 (2×)148(2)
Ø=246(1)Ø=145(1)N–O6123(2)
N–O8 (2×)125(2)
Na3/K3–O6234(2)Ø= 124(1)
Na3/K3–O4 (2×)260(2)
Na3/K3–O2 (2×)272(2)
Na3/K3–O5 (2×)283(2)
Na3/K3–O8 (2×)299(2)
Ø=270(1)
Table 4:

Selected bond angles (deg) in Na2.5K0.5[B6O10]NO3 obtained from powder diffraction data refined with the Rietveld method (standard deviations in parentheses). The respective data for Na2.4K0.6[B6O10]NO3 and Na2.3K0.7[B6O10]NO3 are given in the Supporting Information Table S3.

O1–B1–O4106.2(9)O3–B3–O7121.0(3)
O3–B1–O4115.7(9)O3–B3–O7117.9(6)
O2–B1–O4104.8(8)O3–B3–O3121.0(3)
O3–B1–O1104.7(9)Ø=120.0(2)
O2–B1–O1112.0(9)
O2–B1–O3113.4(9)O2–B4–O5124(2)
Ø=109.5(6)O4–B4–O5117(2)
O4–B4–O2118(2)
O1–B2–O7106(2)Ø=120(1)
O5–B2–O7113.4(8)
O5–B2–O7111(2)O8–N–O6116.7(8)
O5–B2–O1106.4(8)O8–N–O6124(2)
O5–B2–O1113.4(8)O8–N–O8116.7(8)
O5–B2–O5106.4(8)Ø=119(1)
Ø=109.4(1)

Further details of the crystal structure investigations may be obtained from FIZ Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany (fax: +49-7247-808-666; e-mail: crysdata@fiz-karlsruhe.de) on quoting the deposition numbers CSD-432296, CSD-432295, and CSD-432294, for Na2.5K0.5[B6O10]NO3, Na2.4K0.6 [B6O10]NO3, and Na2.3K0.7[B6O10]NO3, respectively.

2.3 Vibrational spectroscopy

The transmission FT-IR spectra were collected in the spectral range of 600–4000 cm−1with a Bruker Vertex 70 FT-IR spectrometer (spectral resolution 4 cm−1), equipped with an MCT (Mercury Cadmium Telluride) detector, and attached to a Hyperion 3000 microscope. As mid-infrared source, a Globar (silicon carbide) rod was used, whereby 32 scans of the micro-crystals of each phase placed on a BaF2 sample holder were acquired. Primary data was recorded with support of the spectrometer software Opus [21].

Confocal Raman spectra recorded in the range of 400–4000 cm−1 with a HORIBA Jobin Yvon LabRam-HR 800 Raman micro-spectrometer. The samples were excited using a frequency doubled 100 mW Nd-YAG-laser (λ=532.22 nm) under an Olympus 100×objective lens with a numerical aperture of 0.9. The scattered light was dispersed by an optical grating with 1800 lines mm−1, and collected by a 1024×256 open electrode CCD detector. Measurement time of 100 s for each sample was chosen, whereby the intense NO3 bands around 1050 cm−1 were recorded additionally with 10 s measurement time in the range of 9002000 cm−1. All spectra were measured with a resolution higher than 2 cm−1 determined by measuring the Rayleigh line. A manual background correction was applied [22].

2.4. Elemental analyses

The atomic mass ratio of the alkali cations was determined by energy-dispersive X-ray spectrometry (EDX). The samples were prepared by placing various crystals of each of the three phases on a carbon pad, coating them with a thin carbon film in order to obtain conductivity. The equipment used was a high-resolution electron microscopy system (Jeol JSM-6010 LV) with a Bruker QUANTAX system and a Peltier-cooled Bruker XFlash 410-M silicon drift detector. Spectra acquisition was performed at 10–15 kV accelerating voltage with an output count rate of 2800 cps. The measured spot sizes were 40–50 μm². The lifetime during spectrum collection was 60 s and the dead time was less than 2%. The Bruker Esprit 1.9 software was used for standard-less spectra evaluation through the peak-to-background model and subsequent ZAF correction. Due to the standard-less method, the totals were normalized to 100%. Backscattered imaging was obtained by scanning electron microscopy (SEM).

3 Results and discussion

3.1 Synthetic conditions

The title compounds were obtained hydrothermally, as described in the experimental section. A schematic representation of the educt ratios yielding different Na2B4O7·10 H2O:KNO3 ratios of 5:1–1:1 for Na2.5K0.5[B6O10]NO3, 1:2–1:6 for Na2.4K0.6[B6O10]NO3, and 1:≥10 for Na2.3K0.7[B6O10]NO3 is given in Fig. 1. Syntheses with an even higher excess of Na2B4O7·10 H2O (>5:1) did not lead to any product formation. To promote single-crystal growth, a low cooling rate of 2 K h−1 was applied. Even slower cooling and/or temperature variations did, however, not lead to crystallite sizes larger than 5 μm.

Various syntheses were also performed using potassium borate K2B4O7·4 H2O and KNO3 which lead to the formation of the known K4B10O15(OH)4 [23]. Analogously, Na2B4O7·10 H2O and NaNO3 educt mixtures were tested, which gave the known Na3(NO3)[B6O10] as the sole reaction product. Whenever the reaction temperature was lowered to 473 K, and/or the reaction time was shortened (1–2 days instead of 3 days at maximum temperature), the known compound Na3(NO3)[B6O10] was obtained from educt mixtures of Na2B4O7·10 H2O and KNO3 ranging from 10:1 to 1:10. All further synthetic attempts including also the use of lithium borate Li2B4O7 did not yield any additional solid solutions.

3.2 Crystal structure discussion

The three alkali borate-nitrates Na3−x Kx[B6O10]NO3 (x=0.5, 0.6, 0.7) are substitution variants of the known sodium borate-nitrate Na3(NO3)[B6O10] reported by Yakubovich et al. in 2002 [18]. All phases crystallize with four formula units (Z=4) in the orthorhombic space group Pnma (no. 62). The lattice parameters of the new compound with the smallest potassium content, Na2.5K0.5[B6O10]NO3, are a=1261.72(5), b=1004.19(5), c=770.55(3) pm, and V=0.97630(6) nm3. The second phase, Na2.4K0.6[B6O10]NO3, shows unit cell parameters of a=1265.74(3), b=1006.94(2), c=772.06(2) pm, and V=0.98400(3) nm3. The borate-nitrate Na2.3K0.7[B6O10]NO3 accordingly displays the largest lattice parameters of a=1267.12(5), b=1007.96(4), c=774.38(3) pm, and V=0.98905(6) nm3. The growing cell volume of the structures in accordance with the increasing potassium content is also reflected in their X-ray powder diffraction patterns. Fig. 2 shows all three experimentally obtained patterns in direct comparison to the isotypic compound Na3(NO3)[B6O10]. A clear shift towards smaller diffraction angles compared to the pure sodium phase can be detected, agreeing with the increase in cell volume.

Fig. 2: Top to bottom: in black, the experimental X-ray powder diffraction patterns (MoKα1; λ=70.93) of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 are given. The theoretical powder pattern simulated from single-crystal data of Na3(NO3)[B6O10] by Yakubovich et al. [18] is indicated in red in each diagram for comparison.
Fig. 2:

Top to bottom: in black, the experimental X-ray powder diffraction patterns (MoKα1; λ=70.93) of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 are given. The theoretical powder pattern simulated from single-crystal data of Na3(NO3)[B6O10] by Yakubovich et al. [18] is indicated in red in each diagram for comparison.

Figure 3 shows the Rietveld plots including the difference curves obtained for the fits. All refinements showed good reliability factors as presented in Table 1. Table 2 gives the positional parameters and the site occupation factors of all positions in one of the three solid solutions based on Na3(NO3)[B6O10] obtained through the Rietveld refinement. The position Na1 is fully occupied by sodium in all synthesized phases. The original Na2 and Na3 sites of Na3(NO3)[B6O10] are mixed occupied with sodium and potassium in the title compounds with values as listed in Table 2 and Table S1 of the Supporting Information. From these site occupation factors, a sodium content of 2.51(4) Na+ per unit cell and a corresponding K+ content of 0.50(4) were refined leading to the formula Na2.5K0.5[B6O10]NO3. The refinement of the second synthesized phase yielded a Na+ content per unit cell of 2.44(1) and a respective K+ content of 0.56(1) resulting in the formula Na2.4K0.6[B6O10]NO3. For the phase with the highest amount of K+, the site occupation factors of the solid solution yielded a Na+ content per unit cell of 2.27(4) and a respective K+ content of 0.74(4) giving an idealized formula Na2.3K0.7[B6O10]NO3. These calculated values were supported by EDX analyses (vide infra).

Fig. 3: Top to bottom: In black, the experimental powder patterns of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 are given. The patterns calculated via Rietveld analyses are given in red, the difference curves in blue.
Fig. 3:

Top to bottom: In black, the experimental powder patterns of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 are given. The patterns calculated via Rietveld analyses are given in red, the difference curves in blue.

Figure 4 shows the unit cell of Na2.3K0.7[B6O10]NO3 viewed along [010], representative for all three solid solutions. The mixed occupied positions Na2/K2 and Na3/K3 are indicated with gray spheres, the unsubstituted, pure sodium positions Na1 are given as black spheres. The three different cation positions show eight-fold (Na1; distorted cubic), six-fold (Na2/K2; octahedral), and nine-fold oxygen coordination (Na3/K3; a distorted hexagonal bipyramid with one split top can be drawn as the coordination polyhedron). In direct comparison to Na3(NO3)[B6O10], Na2.5K0.5[B6O10]NO3 shows coordination distances that vary only within the estimated standard deviations. In Na2.4K0.6[B6O10]NO3, a shift towards larger bond lengths can be seen: the averaged value for Na1–O increases from 256.3 to 257.0(3) pm, the Na2–O distances rise from 250.4 to 251.3(3) pm for the equivalent Na2/K2–O coordination, and the values for Na3–O rise from 269.2 to 272.4(3) pm in the case of Na3/K3–O on average. With increasing potassium content, the cation-to-oxygen distances tend to augment: in Na2.3K0.7[B6O10]NO3, the Na1–O coordination partners are found at a mean value of 260(1) pm from each other. However, the average Na2/K2–O distance does not show a significant growth compared to the pure sodium site occupation. The average Na3/K3–O distance reaches a significantly larger value of 273(1) pm. Fig. 5 shows the three different coordination polyhedra with clear trends in bond lengths indicated in colors.

Fig. 4: Unit cell of Na2.3K0.7[B6O10]NO3 as viewed along [010]. The mixed alkali positons are given as gray spheres.
Fig. 4:

Unit cell of Na2.3K0.7[B6O10]NO3 as viewed along [010]. The mixed alkali positons are given as gray spheres.

Fig. 5: Coordination polyhedra in Na2.3K0.7[B6O10]NO3. Bonds elongated compared to Na3(NO3)[B6O10] are highlighted in red, shortened bonds are given in green.
Fig. 5:

Coordination polyhedra in Na2.3K0.7[B6O10]NO3. Bonds elongated compared to Na3(NO3)[B6O10] are highlighted in red, shortened bonds are given in green.

In analogy to the sodium borate-nitrate published by Yakubovich et al., all three solid solutions are built up from trigonal-planar as well as tetrahedral borate groups. The basic structural motif is formed by three BO4 groups sharing a common apex with the O1 position forming an OB3 group. Three additional BO3 units extend this borate to a three-dimensional network. This structural feature was first reported in the mineral tunellite [24]. The nitrate triangles are isolated and participate in the cation coordination. Fig. 6 shows the fundamental building units in Na2.3K0.7[B6O10]NO3 representatively for all three structures. Bond lengths and angles within the trigonal-planar BO3 and NO3 units, as well as of the tetrahedral borate groups are listed in Tables 3 and 4 as well as Tables S2 and S3 given in the Supporting Information. All distances and angles concur well with reported values [5], [25], [26]. Interestingly, in all substitution variants the trigonal-planar borate group centered by B4 shows equally larger overall bond lengths of 141(1), 139.9(4), and 140(1) pm, compared to 136.4 pm in Na3(NO3)[B6O10].

Fig. 6: Fundamental building unit in Na2.3K0.7[B6O10]NO3 representative for all three substitution variants.
Fig. 6:

Fundamental building unit in Na2.3K0.7[B6O10]NO3 representative for all three substitution variants.

3.3 Vibrational spectra

The IR and Raman spectra of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 are given in Figs. 7 and 8. All IR and Raman spectra display similar vibrational bands, assignments were done by comparison to literature. In general, the spectra confirm the structural findings, so bending and coupled vibrations of BO3 and BO4 units are visible between 900 and 1400 cm−1 [27]. From 850 to 1600 cm−1, stretching vibrations of B–O units are detected, whereby absorptions of BO4 tetrahedra dominate in the region of 850–1100 cm−1 [28], [29], [30] and absorptions of BO3 units between 1200 and 1450 cm−1[29], [30], [31], [32]. Bands at 1770 cm−1 and ~2430 cm−1 visible in the IR spectra are likely to stem from the vibrational absorptions of the nitrate groups [33]. No water bands were visible in the IR experiments. The sharp intense bands clearly visible in all Raman spectra at 1059 and 1057 cm−1 are characteristic symmetric ν1 stretching modes of the NO3 groups. Since the local D3h symmetry of the nitrate triangle is lowered due to the coordination to the cations, split peak shapes are observed [34], [35]. The pronounced bands at ~1380 cm−1 are likely to contain contributions of N–O stretching modes of the nitrate groups [36].

Fig. 7: Top to bottom: IR spectra of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 in the range of 600 to 3500 cm−1.
Fig. 7:

Top to bottom: IR spectra of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 in the range of 600 to 3500 cm−1.

Fig. 8: Top to bottom: Raman spectra of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 in the range of 400 to 2000 cm−1. Upper right corners: 10 s measurement time in the range of 1000–1100 cm−1 displaying the full, intense NO3 group vibrations.
Fig. 8:

Top to bottom: Raman spectra of Na2.5K0.5[B6O10]NO3, Na2.4K0.6[B6O10]NO3, and Na2.3K0.7[B6O10]NO3 in the range of 400 to 2000 cm−1. Upper right corners: 10 s measurement time in the range of 1000–1100 cm−1 displaying the full, intense NO3 group vibrations.

3.4. Elemental analyses

The sodium-to-potassium ratio obtained through the refinement of the powder diffraction data via the Rietveld method as discussed above was confirmed in the course of energy dispersive X-ray spectroscopy (EDX) within the accuracy permitted by the method. Various micro crystals of all title compounds were measured as described in detail in the experimental section. An average atomic ratio of Na+/K+=5.3(3) was obtained for Na2.5K0.5[B6O10]NO3. For the compound Na2.4K0.6[B6O10]NO3, the average atomic ratio of Na+/K+=4.6(3) was acquired and for the last phase Na2.3K0.7[B6O10]NO3 a Na+/K+ ratio of 2.8(3) was detected. In summary, the measured ratios Na+/K+ of 5.3(3), 4.6(3), and 2.8(3) correspond to the ratios of 5.0(4), 4.4(1), and 3.1(4) obtained in the course of the Rietveld refinements, respectively.

Figure 9 shows an image obtained through backscattered electron spectroscopy of the Na2.5K0.5[B6O10]NO3 plate-like crystallites produced hydrothermally, representative for all three phases, obtained with the SEM equipment.

Fig. 9: BSE SEM image of the Na2.5K0.5[B6O10]NO3 plate-like crystallites obtained hydrothermally at 450-fold magnification, representative for the habitus of all three phases.
Fig. 9:

BSE SEM image of the Na2.5K0.5[B6O10]NO3 plate-like crystallites obtained hydrothermally at 450-fold magnification, representative for the habitus of all three phases.

4 Conclusions

The new borate-nitrates Na3−x Kx[B6O10]NO3 (x=0.5, 0.6, 0.7) representing solid solutions were obtained from mild hydrothermal syntheses. The structures show increasing potassium substitution with rising K+ content of the educt mixtures, whereby no potassium content exceeding Na+:K+ of 3.3:1 could be observed. Electron dispersive spectroscopy confirmed the cation ratios obtained by the refinement of powder diffraction data employing the Rietveld method. Other substitution variants of the Na3(NO3)[B6O10] structure are the subjects of further investigations.

5 Supporting information

Additional crystallographic data for Na2.4K0.6[B6O10]NO3 and Na2.3K0.7[B6O10]NO3 (atomic coordinates, according Wyckoff positions, site occupation factors, selected interatomic distances and angles) are given as Supporting Information available online (DOI: 10.1515/znb-2016-0242).

Acknowledgement

We thank Mag. Daniela Vitzthum for collecting the FT-IR data, as well as Univ.-Prof. Dr. Roland Stalder (Institute for Mineralogy and Petrography, University of Innsbruck) for granting access to the FTIR-microscope.

References

[1] E. L. Belokoneva, Crystallogr. Rev.2005, 11, 151.10.1080/08893110500230792Search in Google Scholar

[2] H. Huppertz, Z. Naturforsch.2003, 58b, 278.10.1515/znb-2003-0406Search in Google Scholar

[3] R. S. Bubnova, S. K. Filatov, Z. Kristallogr.2013, 228, 395.10.1524/zkri.2013.1646Search in Google Scholar

[4] D. Jarosch, J. Zemann, Monatsh. Chem.1983, 114, 267.10.1007/BF00798948Search in Google Scholar

[5] A. Leclaire, J. Solid State Chem.1979, 28, 235.10.1016/0022-4596(79)90075-6Search in Google Scholar

[6] T. A. Bither, H. S. Young, J. Solid State Chem.1974, 10, 302.10.1016/0022-4596(74)90039-5Search in Google Scholar

[7] L. Li, X. Jin, G. Li, Y. Wang, F. Liao, G. Yao, J. Lin, Chem. Mater.2003, 15, 2253.10.1021/cm030004dSearch in Google Scholar

[8] W. Sun, T.-T. Zhu, B.-C. Zhao, Y.-X. Huang, J.-X. Mi, Acta Crystallogr.2012, E68, i33.10.1107/S1600536812016169Search in Google Scholar

[9] L. Y. Li, G.-B. Li, F.-H. Liao, J. H. Lin, Acta Phys. Chim. Sin.2005, 21, 769.10.3866/PKU.WHXB20050714Search in Google Scholar

[10] B.-C. Zhao, W. Sun, W.-J. Ren, Y.-X. Huang, Z. Li, Y. Pan, J.-X. Mi, J. Solid State Chem.2013, 206, 91.10.1016/j.jssc.2013.07.032Search in Google Scholar

[11] S. Wang, E. V. Alekseev, W. Depmeier, T. E. Albrecht-Schmitt, Chem. Commun.2010, 46, 3955.10.1039/c002588gSearch in Google Scholar PubMed

[12] J.-L. Song, C.-L. Hu, X. Xu, F. Kong, J.-G. Mao, Inorg. Chem.2013, 52, 8979.10.1021/ic401175rSearch in Google Scholar PubMed

[13] J.-L. Song, C.-L. Hu, X. Xu, F. Kong, J.-G. Mao, Angew. Chem. Int. Ed.2015, 54, 3679.10.1002/anie.201412344Search in Google Scholar PubMed

[14] J.-L. Song, X. Xu, C.-L. Hu, F. Kong, J.-G. Mao, CrystEngComm2015, 17, 3953.10.1039/C5CE00509DSearch in Google Scholar

[15] T. S. Ortner, K. Wurst, L. Perfler, M. Tribus, H. Huppertz, J. Solid State Chem.2015, 221, 66.10.1016/j.jssc.2014.09.016Search in Google Scholar

[16] T. S. Ortner, K. Wurst, C. Hejny, H. Huppertz, J. Solid State Chem.2016, 233, 329.10.1016/j.jssc.2015.11.005Search in Google Scholar

[17] T. S. Ortner, K. Wurst, B. Joachim, H. Huppertz, Eur. J. Inorg. Chem.2016, 3659.10.1002/ejic.201600480Search in Google Scholar

[18] O. V. Yakubovich, I. V. Perevoznikova, O. V. Dimitrova, V. S. Urusov, Dokl. Phys.2002, 47, 791.10.1134/1.1526424Search in Google Scholar

[19] S. Parsons, H. D. Flack, T. Wagner, Acta Crystallogr.2013, 69, 249.10.1107/S2052519213010014Search in Google Scholar

[20] Topas (version 4.2), Bruker AXS GmbH, Karlsruhe (Germany) 2009.Search in Google Scholar

[21] Opus (version 6.5), Bruker Optik GmbH, Ettlingen (Germany) 2008.Search in Google Scholar

[22] LabSpec (version 5), HORIBA Jobin Yvon S.A.S., Villeneuve d’Ascq (France) 2010.Search in Google Scholar

[23] H.-X. Zhang, J. Zhang, S.-T. Zheng, G.-Y. Yang, Cryst. Growth Des.2005, 5, 157.10.1021/cg049888jSearch in Google Scholar

[24] J. R. Clark, Science1963, 141, 1178.10.1126/science.141.3586.1178Search in Google Scholar PubMed

[25] E. Zobetz, Z. Kristallogr.1982, 160, 81.10.1524/zkri.1982.160.1-2.81Search in Google Scholar

[26] E. Zobetz, Z. Kristallogr.1990, 191, 45.10.1524/zkri.1990.191.1-2.45Search in Google Scholar

[27] H. R. Xia, L. X. Li, B. Teng, W. Q. Zheng, G. W. Lu, H. D. Jiang, J. Y. Wang, J. Raman Spectrosc.2002, 33, 278.10.1002/jrs.847Search in Google Scholar

[28] S. D. Ross, Spectrochim. Acta1972, 28A, 1555.10.1016/0584-8539(72)80126-0Search in Google Scholar

[29] J. P. Laperches, P. Tarte, Spectrochim. Acta1966, 22, 1201.10.1016/0371-1951(66)80023-1Search in Google Scholar

[30] K. Machida, H. Hata, K. Okuno, G. Adachi, J. Shiokawa, J. Inorg. Nucl. Chem.1979, 41, 1425.10.1016/0022-1902(79)80205-5Search in Google Scholar

[31] W. C. Steele, J. C. Decius, J. Chem. Phys.1956, 25, 1184.10.1063/1.1743175Search in Google Scholar

[32] G. D. Chryssikos, J. Raman Spectrosc.1991, 22, 645.10.1002/jrs.1250221109Search in Google Scholar

[33] F. Matossi, V. Hohler, Z. Naturforsch.1967, 22a, 1525.10.1515/zna-1967-1009Search in Google Scholar

[34] J. C. G. Buenzli, E. Moret, J. R. Yersin, Helv. Chim. Acta1978, 61, 762.10.1002/hlca.19780610224Search in Google Scholar

[35] J. R. Ferraro, J. Mol. Spectrosc.1960, 4, 99.10.1016/0022-2852(60)90071-0Search in Google Scholar

[36] S. A. Asher, D. D. Tuschel, T. A. Vargson, L. Wang, S. J. Geib, J. Phys. Chem. A2011, 115, 4279.10.1021/jp200406qSearch in Google Scholar PubMed


Supplemental Material:

The online version of this article (DOI: 10.1515/znb-2016-0242) offers supplementary material, available to authorized users.


Received: 2016-11-23
Accepted: 2016-12-16
Published Online: 2017-2-15
Published in Print: 2017-3-1

©2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 26.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/znb-2016-0242/html
Scroll to top button