Skip to content
BY 4.0 license Open Access Published by De Gruyter September 21, 2023

Crystal structures and luminescence properties of Li6MN4:Ce3+ (M = Mo, W)

  • Martin Klotz , Markus Seibald ORCID logo , Dominik Baumann , Klaus Wurst , Gunter Heymann ORCID logo and Hubert Huppertz EMAIL logo

Abstract

The compounds Li6MoN4 and Li6WN4 were prepared from the reactants M (M = Mo, W) and Li3N in a radiofrequency furnace at 1000 °C. The lithonitridometallates crystallize in the tetragonal system with the lattice parameters: a = 6.6844(1), c = 4.9294(1) Å for Li6WN4 based on single-crystal X-ray diffraction data and a = 6.6611(3), c = 4.9338(3) Å for Li6MoN4 taken from powder X-ray diffraction data. Colorless to slightly reddish single-crystals of the tungsten compound were isolated and the crystal structure was refined in the space group P42/nmc (no. 137) with Z = 2 and the powder X-ray data of the molybdenum compound were analyzed by a Rietveld refinement. Both structures belong to the Li6[ZnO4] type published by Hoppe et al. in 1987 (Untenecker H., Hoppe R. Z. Anorg. Allg. Chem. 1987, 551, 147–150) and could be doped with Ce3+ for the first time. The investigated compounds show a reddish color impression upon UV to blue irradiation and exhibit a broad emission band with a maximum at λmax = 693 nm (fwhm 97 nm) for Li6MoN4 and at λmax = 653 nm (fwhm 133 nm) for Li6WN4.

1 Introduction

In 1961, Juza et al. first reported the preparation of ternary nitrides of lithium with the transition metals Cr, Mo and W. They succeeded by reacting Li3N with the pure metals in high-temperature autoclaves, obtaining compounds with the composition Li9MN5 (M = Cr, Mo, W) [1]. Gudat et al. synthesized the compounds Li6MN4 (M = Cr, Mo, W) and Li15Cr2N9 for the first time in 1990 and described them as fluorite-related superstructures [2]. In the 1990s, lithium nitride was reacted with a larger selection of metals and a wide variety of lithonitridometallates were obtained [3], [4], [5], [6]. In 2005, Yuan et al. synthesized the compound Li6WN4 by high-temperature synthesis with N2 overpressure and analyzed it using powder X-ray diffraction data [7].

The field of nitride chemistry in connection with the element lithium has made great progress in recent decades. Not only because it had received little attention in the past compared to lithooxides, but also because the use of these compounds in batteries or as ion conductors has recently become increasingly important [5, 8, 9]. In the corrosion of Cr-containing steels, the formation of lithium nitridometallates is also observed [10, 11]. It was found that the extremely reactive lithium nitride melts attack metal crucibles forming a passivating layer. Specifically, when tungsten crucibles are used, a layer with the composition Li6WN4 is formed and thus prevents ongoing progressive attack [12, 13].

Another area of application of nitride compounds is as phosphors in so-called phosphor-converted LEDs (pc-LEDs). Here, nitridosilicates such as the compounds M2Si5N8 (M = Ca, Sr, Ba), RESi3N5 or RE3Si6N11 (RE = rare earth element) are mainly used [14], [15], [16], [17]. In addition to the nitridosilicates, the groups of lithium oxonitridosilicates/oxonitridoaluminates and alkalilithosilicates are also of great interest. The compounds LiSiON, Li3SiNO2, SLA (Sr[LiAl3N4]) and SALON (Sr[Li2Al2O2N2]) play a major role, especially in the phosphor sector [18], [19], [20], [21], [22], [23], [24], [25], [26], [27]. Among the compounds of these series, LiSiON is structurally most similar to the title compounds, since the crystals are built exclusively by a network of tetrahedra without any units with higher coordination numbers [27]. It should be noted that in the compound class of pure lithonitridometallates, no compounds are known as solid-state phosphors in the literature.

In this contribution, we report on the normal-pressure syntheses and crystal structures of Li6MN4 (M = Mo, W). For Li6WN4, we present single-crystal data for the first time. The structure model of the title compounds is compared and discussed with previously published compounds that crystallize in the Li6[ZnO4] structure type [28]. In addition, both title compounds were doped with the activator ion Ce3+ and their optical properties were investigated.

2 Results and discussion

2.1 Crystal structure

The lithonitridotungstate crystallizes in the tetragonal crystal system in the space group P42/nmc (no. 137). For the crystal structure of Li6WN4, the lattice parameters a = 6.6844(1) and c = 4.9294(1) Å and a cell volume of V = 220.252(8) Å3 were obtained from single-crystal X-ray diffraction data. The details of the structure refinement are given in Table 1 and the atomic coordinates, displacement parameters, interatomic distances and angles are given in Tables 25.

Table 1:

Single-crystal data and structure refinement of tetragonal Li6WN4.

Empirical formula Li6WN4
Molar mass/g·mol−1 281.53
Crystal system Tetragonal
Space group P42/nmc (no. 137)
Single-crystal diffractometer Bruker D8 Quest PHOTON III C14
Radiation; λ Mo; 0.71073
a 6.6844(2)
c 4.9294(2)
V3 220.25(1)
Formula units per cell Z 2
Calculated density/g·cm−3 4.25
Crystal size/mm3 0.090 × 0.090 × 0.090
Temperature/K 293(2)
Detector distance/mm 40
Frame width/deg 0.5
Exposure time per frame/s 40
Absorption coefficient/mm−1 26.1
F(000)/e 240
θ range/deg 4.31–40.90
Range in hkl ±11; ±11; ±8
Reflections total; independent 16,146; 334
R int 0.0406
Reflections with I > 2σ(I) 291
Data; ref. parameters 334; 19
Absorption correction Multi-scan (Sadabs-2016/2)
Final R1; wR2 [I ≥ 2σ(I)] 0.0070; 0.0184
Final R1; wR2 (all data) 0.0085; 0.0190
Goodness of fit on F2 1.136
Largest diff. peak; hole/e·Å−3 1.22; −0.79
Table 2:

Wyckoff positions, atomic coordinates, and equivalent isotropic displacement parameters Ueq (/Å2) of Li6WN4 based on single-crystal data (standard deviations in parentheses).

Atom Wyckoff position x y z U eq
W 2b 3/4 1/4 1/4 0.00315(7)
Li1 4d 1/4 1/4 0.3337(9) 0.0137(7)
Li2 8f 0.4620(5) x 1/4 0.0097(6)
N 8g 1/4 0.50989(2) 0.5510(2) 0.0066(2)
Table 3:

Anisotropic displacement parameters Uij (/Å2) of Li6WN4 (standard deviations in parentheses).

Atom U 11 U 22 U 33 U 23 U 13 U 12
W 0.00325(7) U 11 0.00294(8) 0 0 0
Li1 0.018(2) 0.009(2) 0.015(2) 0 0 0
Li2 0.0103(9) U 11 0.009(2) 0.0015(5) 0.0015(5) −0.001(2)
N 0.0060(5) 0.0071(5) 0.0066(4) −0.0010(4) 0 0
Table 4:

Wyckoff positions and atomic coordinates of Li6MoN4 based on the Rietveld refinement of the powder data (the Li positions were not refined) (standard deviations in parentheses).

Atom Wyckoff position x y z
Mo 2b 3/4 1/4 1/4
Li1 4d 1/4 1/4 0.3337
Li2 8f 0.462 x 1/4
N 8g 1/4 0.504(2) 0.546(3)
Table 5:

Interatomic distances and angles in Li6MN4 (M = Mo, W) of the single-crystal structure refinement and the Rietveld refinements (standard deviations in parentheses).

Distance/Å M = W SCXRD M = W PXRD M = Mo PXRD Angle/deg M = W SCXRD M = W PXRD M = Mo PXRD
M–N (4×) 1.882(2) 1.88(2) 1.93(2) N–M–N 105.78(3) 104.7(3) 106.0(4)
Ø M–N 1.882(2) 1.88(2) 1.93(2) N–M–N 117.13(7) 119.4(7) 116.8(8)
N–Li1–N 102.5(2) 103.4(6) 100.0(7)
Li1–N (2×) 2.227(3) 2.20(2) 2.22(2) N–Li1–N 109.17(3) 109.6(2) 109.8(2)
Li1–N (2×) 2.041(2) 2.05(2) 1.99(2) N–Li1–N 116.7(2) 114.4(6) 116.6(7)
Ø Li1–N 2.134(2) 2.13(2) 2.11(2) N–Li2–N 86.8(2) 86.6(6) 89.3(6)
N–Li2–N 105.63(5) 105.7(8) 105.3(7)
Li2–N (2×) 2.184(3) 2.16(6) 2.19(9) N–Li2–N 110.10(7) 109.1(5) 108.8(6)
Li2–N (2×) 2.060(2) 2.09(1) 2.05(9) N–Li2–N 130.1(2) 131.5(5) 131.6(6)
Ø Li2–N 2.122(2) 2.13(1) 2.12(9)
  1. Mean values are written in bold letters.

The tungsten and lithium atoms are tetrahedrally coordinated by nitrogen atoms (Figure 1). The W–N distances of 1.882(2) Å are in the range of the values known in the literature. For example, in the compound Ba3[WN4] of Rabenau et al., in which the tungsten atom is also tetrahedrally coordinated by nitrogen atoms, the W–N distances have a mean value of 1.902 Å [2, 29, 30]. Likewise, the Li–N distances with a range of 2.041(2)–2.227(3) Å (see Figure 1) agree with published values for other lithium nitridometallates of different transition metals. E.g., Ma et al. give a Li–N distance of 2.252 Å for the compound LiSiON, in which lithium is tetrahedrally coordinated by three oxygen atoms and one nitrogen atom [2, 23, 31].

Figure 1: 
Coordination polyhedra of the tungsten site and the two lithium sites in Li6WN4 with the interatomic distances in Å (standard deviations in parentheses). Nitrogen atoms are drawn as red spheres.
Figure 1:

Coordination polyhedra of the tungsten site and the two lithium sites in Li6WN4 with the interatomic distances in Å (standard deviations in parentheses). Nitrogen atoms are drawn as red spheres.

Regarding the coordination polyhedra, the structure is built up from a three-dimensional network of edge- and corner-linked LiN4 tetrahedra. Along [001], channels are formed by an alternating arrangement of eight edge-linked LiN4 tetrahedra of the Li1 site, with the Li2 site forming a square. The channels consist of alternating WN4 tetrahedra and vacancies (see Figure 2).

Figure 2: 
Crystal structure of Li6WN4 in the viewing directions [001] (top left), [100] (top middle) and [010] (bottom left). The unit cell is illustrated with black edges. Nitrogen atoms are shown in red and tungsten atoms in turquoise. The differently colored LiN4 tetrahedra (Li1: green and Li2: orange) represent the two different Li positions. The squarish channels are marked in yellow.
Figure 2:

Crystal structure of Li6WN4 in the viewing directions [001] (top left), [100] (top middle) and [010] (bottom left). The unit cell is illustrated with black edges. Nitrogen atoms are shown in red and tungsten atoms in turquoise. The differently colored LiN4 tetrahedra (Li1: green and Li2: orange) represent the two different Li positions. The squarish channels are marked in yellow.

In the crystal structure of Li6WN4, the positions of the nitrogen atoms are close to a ccp (cubic closest packing) arrangement where seven out of eight tetrahedral voids are occupied by W and Li (see Figure 3).

Figure 3: 
Left: edge-linked Li1 tetrahedra (green) and a W tetrahedron (turquoise) linked by corners to the Li1 channel. Middle: strands of edge-linked Li2 tetrahedra (orange) which are linked by corners. Right: the resulting distorted ccp of nitrogen atoms with the seven out of eight occupied tetrahedra voids. The distorted ccp arrangement (black lines) of the N atoms (red spheres) is drawn in all three representations for better understanding.
Figure 3:

Left: edge-linked Li1 tetrahedra (green) and a W tetrahedron (turquoise) linked by corners to the Li1 channel. Middle: strands of edge-linked Li2 tetrahedra (orange) which are linked by corners. Right: the resulting distorted ccp of nitrogen atoms with the seven out of eight occupied tetrahedra voids. The distorted ccp arrangement (black lines) of the N atoms (red spheres) is drawn in all three representations for better understanding.

2.2 MAPLE, CHARDI and BLBS calculations

Madelung Part of Lattice Energy (MAPLE) [32, 33], bond-length/bond-strength (BLBS) [34, 35] and Charge Distribution (CHARDI) [36, 37] calculations were performed to prove the electrostatic consistency of the crystal structure model. The MAPLE sums of Li6WN4 are listed in Table 6. The results of the CHARDI and BLBS calculations are shown in Table 7.

Table 6:

Comparison between the calculated MAPLE values for Li6WN4 and SiO2 and the corresponding theoretical starting materials WO3, Li3N and Si3N4 according to the following equation: 2 WO3 + 4 Li3N + Si3N4 → 2 Li6WN4 + 3 SiO2.

Educts MAPLE values/kJ mol−1 Products MAPLE values/kJ mol−1
WO3 [38] (2×) 51,934 Li 6 WN 4 (2×) 245,450
Li3N [39] (4×) 25,336 SiO 2 [40] (3×) 8960
Si3N4 [41]
53,211


Σ
130,481
Σ
130,819
Difference abs. in kJ mol−1 338
Difference in % 0.26
  1. Bold values just to emphasize the values.

Table 7:

Charge distribution in Li6WN4 calculated via BLBS (ΣV) and CHARDI (ΣQ).

W Li1 Li2 N
ΣV +6.47 +1.00 +1.02 −3.14
ΣQ +6.00 +1.00 +1.00 −3.00

Comparing the MAPLE sum based on WO3, Li3N and Si3N4 with the calculated MAPLE sum of Li6WN4 and SiO2, it was found that the values differ by only 0.26 %. The results of the BLBS method show the expected charges for the atoms, which are in good agreement with the oxidation states according to an electro-neutral sum formula. The result obtained using the BLBS method is supported by the CHARDI calculations, which is also in good agreement.

The validity of the structural model was confirmed for the powder phases by Rietveld refinement (a phase-pure sample obtained via the described synthesis pathway for the tungsten variant, and a high-quality powder sample with a minor amount of an unknown side phase for the molybdenum variant), as shown in Figures 4 and 5. The results of the Rietveld analysis for the W and Mo phases are shown in Table 8.

Figure 4: 
Rietveld fit of the measured powder diffraction data (black), the calculated pattern based on single-crystal data of Li6WN4 (red), and the difference plot (blue). Tick marks illustrate the theoretical reflection positions of Li6WN4 (green).
Figure 4:

Rietveld fit of the measured powder diffraction data (black), the calculated pattern based on single-crystal data of Li6WN4 (red), and the difference plot (blue). Tick marks illustrate the theoretical reflection positions of Li6WN4 (green).

Figure 5: 
Rietveld fit of the measured powder diffraction data (black), the calculated pattern (developed on Li6WN4 single-crystal data by substituting W for Mo (red)), and the difference plot (blue). Tick marks illustrate the theoretical reflection positions of Li6MoN4 (green). The asterisks denote reflections belonging to an unknown side phase.
Figure 5:

Rietveld fit of the measured powder diffraction data (black), the calculated pattern (developed on Li6WN4 single-crystal data by substituting W for Mo (red)), and the difference plot (blue). Tick marks illustrate the theoretical reflection positions of Li6MoN4 (green). The asterisks denote reflections belonging to an unknown side phase.

Table 8:

Crystallographic data for Rietveld refinement of Li6MN4 (M = Mo, W).

Empirical formula Li6WN4 Li6MoN4
Powder diffractometer STOE Stadi P
Radiation; λ Mo1; 0.7093
Space group P42/nmc (no. 137)
a 6.6766(2) 6.6611(3)
c 4.9328(2) 4.9338(3)
V3 219.89(2) 218.91(2)
2θ range/deg 2–50 2–50
2θ step size/deg 0.015 0.015
Rexp/% 3.53 0.71
Rwp/% 5.30 19.92
Rp/% 4.17 12.84
RBragg/% 1.699 5.361

2.3 Luminescence

The two title compounds could be successfully doped with the activator ion Ce3+ and the luminescence properties of Li6WN4:Ce3+ single crystals and Li6MoN4:Ce3+ single particles were investigated. The compound Li6WN4:Ce3+ with a nominal activator concentration of 1 mol% Ce3+ can be excited with UV-near to blue light. The single crystals, also taken for the structure determination, excited at λ = 448 nm show an emission maximum at λmax = 693 nm with a full width at half maximum (fwhm) of 97 nm (0.269 eV, 2166 cm−1). In comparison, single particles taken from a powder sample of the isotypic compound Li6MoN4:Ce3+ with a nominal activator concentration of 1 mol% Ce3+ show an emission maximum at λmax = 653 nm with a full width at half maximum (fwhm) of 133 nm (0.385 eV, 3108 cm−1). In the CIE-xy color space, this corresponds to x = 0.707(1) and y = 0.293(1) for Li6WN4:Ce3+ and x = 0.598(1) and y = 0.400(1) for Li6MoN4:Ce3+ (see Figure 6). The single crystals of the compound Li6WN4:Ce3+ have a slightly reddish body color in daylight and are stable for several months under atmospheric conditions.

Figure 6: 
Left: the blue line represents the luminescence spectrum of single particles taken from a powder sample of Li6MoN4:Ce3+ recorded at an excitation wavelength of 448 nm. The red line represents the luminescence spectrum of a Li6WN4:Ce3+ crystal recorded at an excitation wavelength of 448 nm. The three gray dashed lines signal detector artefacts that occur at wavelengths 688, 703 and 733 nm. Right: color points of the emission spectra of Li6MoN4:Ce3+ particles and Li6WN4:Ce3+ crystals in the CIE-xy color space.
Figure 6:

Left: the blue line represents the luminescence spectrum of single particles taken from a powder sample of Li6MoN4:Ce3+ recorded at an excitation wavelength of 448 nm. The red line represents the luminescence spectrum of a Li6WN4:Ce3+ crystal recorded at an excitation wavelength of 448 nm. The three gray dashed lines signal detector artefacts that occur at wavelengths 688, 703 and 733 nm. Right: color points of the emission spectra of Li6MoN4:Ce3+ particles and Li6WN4:Ce3+ crystals in the CIE-xy color space.

In the compounds Li6MN4:Ce3+ (M = Mo, W), two possible positions occur that can be occupied by the activator ion Ce3+, although a precise assignment is not possible (see Figure 7). Comparing the distances in the MN4 tetrahedra of the two Rietveld refinements (Table 5) and the ionic radii of tungsten (0.56 Å) [42] and molybdenum (0.55 Å) [42], the molybdenum-centered tetrahedra have a larger distance between the nitrogen and the metal atoms than the tungsten-centered tetrahedra (the ionic radii applied in each case are those of the oxidation states 6+ and the coordination numbers 4). This difference leads to a smaller splitting of the d orbitals and thus results in a shift of the bands (see Figure 6 left).

Figure 7: 
Left: the two possible coordination environments of the activator ion Ce3+ in the tungsten variant including the distances to the N3− anions (red spheres) (distances taken from the powder data for both compounds). Right: the two possible coordination environments of the activator ion Ce3+ in the molybdenum variant including the distances to the N3− anions (red spheres).
Figure 7:

Left: the two possible coordination environments of the activator ion Ce3+ in the tungsten variant including the distances to the N3− anions (red spheres) (distances taken from the powder data for both compounds). Right: the two possible coordination environments of the activator ion Ce3+ in the molybdenum variant including the distances to the N3− anions (red spheres).

For the Ce3+ dopant, an octahedral or even larger coordination is more likely than the tetrahedral coordination (see Figure 7) owing to the comparatively large ionic radius of cerium (r = 1.15 Å) [42]. Two different possible positions of the octahedrally coordinated activator ion Eu2+ were described for the compound Li2CaSi2N4:Eu2+ [43]. Therein, the octahedral coordination polyhedra are occupied by Ca and N atoms, with Ca–N distances ranging from 2.489(3) to 2.509(3) Å for Ca1 and 2.586(3) Å for Ca2. Compared to the distances between the center of the octahedron to its nitrogen atoms in Li6MN4:Ce3+ (M = Mo, W) (Mo: Ø center–N = 2.41(2) Å and W: Ø center–N = 2.42(2) Å), the distances found in Li2CaSi2N4:Eu2+ are slightly larger. Considering that Eu2+ (r = 1.31 Å) has a slightly larger ionic radius than Ce3+ (r = 1.15 Å), the occupation of an octahedrally coordinated center in the structure of the title compounds by Ce3+ would also be possible [42].

A spectrum unfolding was performed for the emission spectrum of the single-crystal Li6WN4:Ce3+ (see Figure 8). This unfolding shows two bands, of which one band has a maximum at λmaxP1 = 1.799 eV (14,507 cm−1) with a full width at half maximum (fwhm) of 0.218 eV (1757 cm−1) and the second band shows a maximum at λmaxP2 = 1.718 eV (13,855 cm−1) with a full width at half maximum (fwhm) of 0.301 eV (2432 cm−1). This double band is a result of spin–orbit coupling of the two sublevels 2F5/2 and 2F7/2 of the 4f1 ground state configuration of Ce3+ [44].

Figure 8: 
Left: the black line represents the luminescence spectrum of a Li6WN4:Ce3+ crystal recorded at an excitation wavelength of 448 nm. The green dashed line represents an ideal Gaussian fit of the experimental emission band. The bimodal Gaussian fit is represented by the blue and red curves. Right: optical appearance of the 1 mol% Ce3+-doped compound Li6WN4 under UV light (365 nm) at the bottom of a tungsten crucible and black particles due to corrosion of the tungsten crucible by lithium.
Figure 8:

Left: the black line represents the luminescence spectrum of a Li6WN4:Ce3+ crystal recorded at an excitation wavelength of 448 nm. The green dashed line represents an ideal Gaussian fit of the experimental emission band. The bimodal Gaussian fit is represented by the blue and red curves. Right: optical appearance of the 1 mol% Ce3+-doped compound Li6WN4 under UV light (365 nm) at the bottom of a tungsten crucible and black particles due to corrosion of the tungsten crucible by lithium.

In order to exclude defect luminescence and host structure emission, several experiments were carried out. For this purpose, the compounds Li6MN4 (M = Mo, W) were prepared without the activator ion Ce3+, for which no luminescence could be detected. Likewise, the possibility of falsification of the emissions due to impurities in the crucibles was eliminated.

3 Conclusions

The two lithonitridometallates described here were prepared in a solid-state reaction. They crystallize in the Li6[ZnO4] structure type described by Hoppe et al. [28]. The quality of the crystals of the tungsten compound could be improved by adjusting the reactor of the radiofrequency furnace. Compared to the synthesis routes of Gudat et al. [2] and Yuan et al. [7], a gas flow was used in a radiofrequency furnace and the reaction was thus carried out at normal pressure. The plausibility of the structure model was checked by single-crystal data, Rietveld refinement, MAPLE, CHARDI and BLBS calculations. The results of the structure models agree with those of other variants known in the literature.

Doping of the compounds with Ce3+ afforded single crystals of Li6WN4:Ce3+, which show an emission maximum at λmax = 693 nm with a full width at half maximum (fwhm) of 97 nm upon excitation with UV near to blue light. Single particles taken from a powder sample of the isotypic compound Li6MoN4:Ce3+ show an emission maximum at λmax = 653 nm with a full width at half maximum (fwhm) of 133 nm. The two phosphors are stable to humidity and atmospheric conditions.

In several experiments in which attempts were made to synthesize the Cr phase isotypic to the title compound, the products showed an emission with a maximum at lower wavelengths than the two reported compounds. This observation is consistent with the hypothesis of shifting the emission maximum to lower wavelengths using a smaller central atom (W, Mo vs. Cr). Due to the aggressiveness of the lithium melt during the syntheses, we were not able to produce the pure Cr phase, despite the use of different crucible materials (tungsten, tantalum, corundum) and crucible inlays (molybdenum, titanium). However, there is reason to assume that the observed emission originates from the compound Li6CrN4:Ce3+.

4 Experimental section

4.1 Synthesis

Because the syntheses involve reactants that are unstable or sensitive to hydrolysis in air, the preparations were carried out in a glovebox filled with inert gas (Ar 5.0, Messer Austria GmbH). In an agate mortar, the starting materials M (M: W – Plansee, 99.9 %; Mo – Plansee, 99.9 %) and Li3N (Alfa Aesar, 99.4 %) were mixed in a stoichiometric ratio of 1:4 together with 1 mol% Ce (Sigma Aldrich, 99.9 %) based on the metal contents. The ratio 1:4 has been found to be the best from empirical studies, although traces of corrosion on the tungsten crucible and sublimation of reactant and subsequent deposition on the cooler reaction apparatus were always observed during the syntheses. The reactants were then poured into a tungsten crucible and transferred to the radio-frequency furnace system (type TrueHeat HF 5010; Hüttinger Elektronik, Freiburg, Germany) under an inert gas atmosphere. To prevent a reaction of the lithium nitride melt with the tungsten crucible during the production of the Mo variant, the reactants were packed in a capsule made of Mo foil (Alfa Aesar, 99.95 %, 0.025 mm thick) and then placed in the tungsten crucible. During the entire annealing process, forming gas (forming gas H10 5.0, Messner Austria GmbH), which was regulated to 10 mL min−1, flowed into the reaction apparatus to ensure a reducing reaction atmosphere. The annealing process consisted of a heating ramp of 200 K h−1 to the final temperature of 1000 °C, a holding time of 12 h at this temperature, a cooling ramp of 0.15 K min−1 to 600 °C and finally quenching to room temperature. The products obtained, which had a slightly reddish body color, could be further investigated and characterized under atmospheric conditions and showed no decomposition due to humidity or atmospheric conditions for several days.

4.2 Single-crystal X-ray diffraction

Single-crystals were isolated from the obtained products using a needle selected under a polarization microscope and analyzed with the Bruker D8 Quest single-crystal diffractometer (Bruker, Billerica, USA) equipped with a PHOTON III C14 detector and an Incoatec Microfocus diffraction source (Incoatec, Geesthacht, Germany) with MoKα radiation (λ = 0.71073 Å). The programs Saint [45] and Sadabs [46] were used for data processing and multi-scan absorption correction. Olex2. solve 1.5 [47] using Direct Methods and Shelxl [48] using the least-squares method, both implemented in the program Olex2 1.5 [49], were used for structure solution and parameter refinement, respectively. The structural data were standardized using Structure Tidy [50] implemented in Platon (version 170613) [51].

Further details of the crystal structure investigation on the single-crystal of Li6WN4 can be obtained from the Cambridge Crystallographic Data Center at www.ccdc.cam.ac.uk/structures/ using the deposition number 2282797 (Li6WN4).

4.3 Powder X-ray diffraction

Powder X-ray diffraction (PXRD) was used for phase analysis, using a STOE Stadi P powder diffractometer (STOE & Cie., Darmstadt, Germany) with Ge(111)-monochromatized MoKα1 (λ = 0.7093 Å) radiation and a Mythen 1K detector in modified Debye-Scherrer geometry over a 2θ range of 2–50°. The experimental powder pattern was compared with the calculated powder pattern derived from single-crystal data of Li6WN4 on the basis of Rietveld analysis. Secondary-phase identification was performed with the ICDD PDF-2 database and the Topas 4.2 program [52] was used for Rietveld refinements.

4.4 Luminescence

The emission spectra of the compounds Li6MN4:Ce3+ (M = Mo, W) were recorded with a setup consisting of a blue laser diode (λ = 448 nm, THORLABS, Ostfildern, Germany) in combination with a CCD-based spectrometer (AVA AvaSpec 2048, AVANTES, Apeldoorn, The Netherlands). A tungsten-halogen calibration lamp (HL3-PLUS CAL, Ocean Insight, Ostfildern, Germany) was used to calibrate the spectral radiance of the system for the first time. Ava Avasoft software (version 7) was used for data preparation.


Corresponding author: Hubert Huppertz, Department of General, Inorganic and Theoretical Chemistry, University of Innsbruck, Innrain 80–82, A-6020 Innsbruck, Austria, E-mail:

  1. Research ethics: Not applicable.

  2. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  3. Competing interests: The authors declare no conflict of interest regarding this article.

  4. Research funding: None declared.

  5. Data availability: The raw data can be obtained on request from the corresponding author.

References

1. Juza, R., Haug, J. Z. Anorg. Allg. Chem. 1961, 309, 276–282; https://doi.org/10.1002/zaac.19613090506.Search in Google Scholar

2. Gudat, A., Haag, S., Kniep, R., Rabenau, A. Z. Naturforsch. 1990, 45b, 111–120; https://doi.org/10.1515/znb-1990-0203.Search in Google Scholar

3. Kniep, R. Pure Appl. Chem. 1997, 69, 185–192; https://doi.org/10.1351/pac199769010185.Search in Google Scholar

4. Niewa, R., DiSalvo, F. Chem. Mater. 1998, 10, 2733–2752; https://doi.org/10.1021/cm980137c.Search in Google Scholar

5. Gregory, D. H. J. Chem. Soc. Dalton Trans. 1999, 259–270; https://doi.org/10.1039/a807732k.Search in Google Scholar

6. Marchand, R., Tessier, F., DiSalvo, F. J. J. Mater. Chem. 1999, 9, 297–304; https://doi.org/10.1039/a805315d.Search in Google Scholar

7. Yuan, W., Hu, J., Song, Y., Wang, W., Xu, Y. Powder Diffr. 2005, 20, 18–21; https://doi.org/10.1154/1.1855325.Search in Google Scholar

8. Tapia-Ruiz, N., Segalés, M., Gregory, D. H. Coord. Chem. Rev. 2013, 257, 1978–2014; https://doi.org/10.1016/j.ccr.2012.11.008.Search in Google Scholar

9. Alpen, U. v. J. Solid State Chem. 1979, 29, 379–392; https://doi.org/10.1016/0022-4596(79)90195-6.Search in Google Scholar

10. Ruedl, E., Sasaki, T. J. Nucl. Mater. 1983, 116, 112–122; https://doi.org/10.1016/0022-3115(83)90299-4.Search in Google Scholar

11. Barker, M., Frankham, S. J. Nucl. Mater. 1982, 107, 218–221; https://doi.org/10.1016/0022-3115(82)90422-6.Search in Google Scholar

12. Rabenau, A. Festkörperprobleme 1978, XVIII, 77–108.10.1093/oxfordjournals.afraf.a096919Search in Google Scholar

13. Rabenau, A. Solid State Ionics 1982, 6, 277–293; https://doi.org/10.1016/0167-2738(82)90012-1.Search in Google Scholar

14. Xie, R.-J., Li, Y. Q., Hirosaki, N., Yamamoto, H. Nitride Phosphors and Solid-State Lighting; CRC Press: Boca Raton, FL, 2016.10.1201/b10939Search in Google Scholar

15. Huppertz, H. Strukturelle Erweiterung der Nitridosilicate, Dissertation; Universität Bayreuth: Bayreuth, 1997.Search in Google Scholar

16. Inoue, Z., Mitomo, M., Ii, N. J. Mater. Sci. 1980, 15, 2915–2920; https://doi.org/10.1007/bf00550562.Search in Google Scholar

17. Gaudé, J., Lang, J., Louër, D. Rev. Chim. Minér. 1983, 20, 523–527.Search in Google Scholar

18. Dutzler, D., Seibald, M., Baumann, D., Huppertz, H. Angew. Chem. Int. Ed. 2018, 57, 13676–13680; https://doi.org/10.1002/anie.201808332.Search in Google Scholar PubMed PubMed Central

19. Wimmer, D. S., Seibald, M., Baumann, D., Peschke, S., Wurst, K., Heymann, G., Dutzler, D., Garcia-Fuente, A., Urland, W., Huppertz, H. Eur. J. Inorg. Chem. 2021, 2021, 4470–4481; https://doi.org/10.1002/ejic.202100550.Search in Google Scholar

20. Ruegenberg, F., Seibald, M., Baumann, D., Peschke, S., Schmid, P. C., Huppertz, H. Chem. Eur. J. 2020, 26, 2204–2210; https://doi.org/10.1002/chem.201904526.Search in Google Scholar PubMed PubMed Central

21. Fang, M.-H., Mariano, C. O. M., Chen, P.-Y., Hu, S.-F., Liu, R.-S. Chem. Mater. 2020, 32, 1748–1759; https://doi.org/10.1021/acs.chemmater.9b04861.Search in Google Scholar

22. Zhao, M., Zhang, Q., Xia, Z. Mater. Today 2020, 40, 246–265; https://doi.org/10.1016/j.mattod.2020.04.032.Search in Google Scholar

23. Ma, Y. Y., Xiao, F., Ye, S., Zhang, Q. Y., Jiang, Z. H., Qian, Y. ECS J. Solid State Sci. Technol. 2012, 1, R1; https://doi.org/10.1149/2.003201jss.Search in Google Scholar

24. Wimmer, D. S., Rießbeck, K. M., Seibald, M., Baumann, D., Wurst, K., Heymann, G., Huppertz, H. Dalton Trans. 2022, 51, 16465–16478; https://doi.org/10.1039/d2dt03064k.Search in Google Scholar PubMed

25. Pust, P., Weiler, V., Hecht, C., Tücks, A., Wochnik, A. S., Henß, A.-K., Wiechert, D., Scheu, C., Schmidt, P. J., Schnick, W. Nat. Mater. 2014, 13, 891–896; https://doi.org/10.1038/nmat4012.Search in Google Scholar PubMed

26. Hoerder, G. J., Seibald, M., Baumann, D., Schröder, T., Peschke, S., Schmid, P. C., Tyborski, T., Pust, P., Stoll, I., Bergler, M., Patzig, C., Reißaus, S., Krause, M., Berthold, L., Höche, T., Johrendt, D., Huppertz, H. Nat. Commun. 2019, 10, 1824; https://doi.org/10.1038/s41467-019-09632-w.Search in Google Scholar PubMed PubMed Central

27. Laurent, Y., Guyader, J., Roult, G. Acta Crystallogr. 1981, B37, 911–913; https://doi.org/10.1107/s0567740881004561.Search in Google Scholar

28. Untenecker, H., Hoppe, R. Z. Anorg. Allg. Chem. 1987, 551, 147–150; https://doi.org/10.1002/zaac.19875510816.Search in Google Scholar

29. Shannon, R. D. Acta Crystallogr. 1976, A32, 751–767; https://doi.org/10.1107/s0567739476001551.Search in Google Scholar

30. Gudat, A., Höhn, P., Kniep, R., Rabenau, A. Z. Naturforsch. 1991, 46b, 566–572; https://doi.org/10.1515/znb-1991-0502.Search in Google Scholar

31. Juza, R., Langer, K., Von Benda, K. Angew. Chem. Int. Ed. Engl. 1968, 7, 360–370; https://doi.org/10.1002/anie.196803601.Search in Google Scholar

32. Hoppe, R. Angew. Chem. Int. Ed. Engl. 1966, 5, 95–106; https://doi.org/10.1002/anie.196600951.Search in Google Scholar

33. Hübenthal, R. Maple (version 4); University of Gießen: Gießen (Germany), 1993.Search in Google Scholar

34. Brown, I., Altermatt, D. Acta Crystallogr. 1985, B41, 244–247; https://doi.org/10.1107/s0108768185002063.Search in Google Scholar

35. Brese, N. E., O’Keeffe, M. Acta Crystallogr. 1991, B47, 192–197; https://doi.org/10.1107/s0108768190011041.Search in Google Scholar

36. Hoppe, R. Z. Kristallogr. 1979, 150, 23–52; https://doi.org/10.1524/zkri.1979.150.14.23.Search in Google Scholar

37. Hoppe, R., Voigt, S., Glaum, H., Kissel, J., Müller, H. P., Bernet, K. J. Less-Common Met. 1989, 156, 105–122; https://doi.org/10.1016/0022-5088(89)90411-6.Search in Google Scholar

38. Salje, E. Acta Crystallogr. 1977, B33, 574–577; https://doi.org/10.1107/s0567740877004130.Search in Google Scholar

39. Rabenau, A., Schulz, H. J. Less-Common Met. 1976, 50, 155–159; https://doi.org/10.1016/0022-5088(76)90263-0.Search in Google Scholar

40. d’Amour, H., Denner, W., Schulz, H. Acta Crystallogr. 1979, B35, 550–555; https://doi.org/10.1107/s056774087900412x.Search in Google Scholar

41. Hardi, D., Jack, K. Nature 1957, 180, 322–332.10.1038/180332a0Search in Google Scholar

42. Holleman, A. F., Wiberg, E., Wiberg, N. Lehrbuch der Anorganischen Chemie, Vol. 102; W. de Gruyter: Berlin, New York, 2007.10.1515/9783110206845Search in Google Scholar

43. Zeuner, M., Pagano, S., Hug, S., Pust, P., Schmiechen, S., Scheu, C., Schnick, W. Eur. J. Inorg. Chem. 2010, 2010, 4945–4951; https://doi.org/10.1002/ejic.201000671.Search in Google Scholar

44. Holloway, W. W., Kestigian, M. J. Opt. Soc. Am. 1969, 59, 60–63; https://doi.org/10.1364/josa.59.000060.Search in Google Scholar

45. Saint (version 8.34a); Bruker AXS Inc.: Madison, Wisconsin (USA), 2014.Search in Google Scholar

46. Sadabs (version 2016/2); Bruker AXS Inc.: Madison, Wisconsin (USA), 2016.Search in Google Scholar

47. Bourhis, L. J., Dolomanov, O. V., Gildea, R. J., Howard, J. A., Puschmann, H. Acta Crystallogr. 2015, A71, 59–75; https://doi.org/10.1107/s2053273314022207.Search in Google Scholar

48. Sheldrick, G. M. Acta Crystallogr. 2015, C71, 3–8; https://doi.org/10.1107/S2053229614024218.Search in Google Scholar PubMed PubMed Central

49. Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K., Puschmann, H. J. Appl. Crystallogr. 2009, 42, 339–341; https://doi.org/10.1107/s0021889808042726.Search in Google Scholar

50. Gelato, L., Parthé, E. J. Appl. Crystallogr. 1987, 20, 139–143; https://doi.org/10.1107/s0021889887086965.Search in Google Scholar

51. Spek, A. L. Acta Crystallogr. 2009, D65, 148–155; https://doi.org/10.1107/s090744490804362x.Search in Google Scholar

52. Topas (version 4.2); Bruker AXS Inc.: Madison, Wisconsin (USA), 2009.Search in Google Scholar

Received: 2023-08-02
Accepted: 2023-08-20
Published Online: 2023-09-21
Published in Print: 2023-11-27

© 2023 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 30.4.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2023-0057/html
Scroll to top button