Next Article in Journal
Anti-Melanogenic Effects of Umbelliferone: In Vitro and Clinical Studies
Previous Article in Journal
Improved Sensitivity of a Taste Sensor Composed of Trimellitic Acids for Sweetness
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of the Binary Nitrides YN, LaN and LuN by Solid-State NMR Spectroscopy

Department of Chemistry, University of Munich (LMU), Butenandtstr. 5-13, 81377 Munich, Germany
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(23), 5572; https://doi.org/10.3390/molecules29235572
Submission received: 24 October 2024 / Revised: 19 November 2024 / Accepted: 22 November 2024 / Published: 25 November 2024
(This article belongs to the Section Inorganic Chemistry)

Abstract

:
Based on their various and outstanding properties, binary nitrides are used as (synthesis) materials in industry and research. Hence, their comprehensive characterization by analytical methods is of particular interest. Since Nuclear Magnetic Resonance (NMR) spectroscopy is very sensitive to the symmetry of the electronic density distribution, it is a suitable tool for the investigation of rock-salt structure types and, especially, for those with known stoichiometry issues. Here, we report on magic-angle spinning NMR spectra of the nuclides 89Y ( I = 1 2 ), 139La ( I = 7 2 ) and 14N ( I = 1 ) in polycrystalline samples of YN, LaN and LuN. Due to the high symmetry of their crystal structures, the spectra of all nuclides do not exhibit anisotropic effects of significant magnitude. The resulting isotropic chemical shift values are δ i s o (89Y) = 516 ppm for YN, δ i s o (139La) = 1294 ppm for LaN, and δ i s o (14N) = 457 ppm (YN), 788 ppm (LaN) and 384 ppm (LuN). The newly determined δ i s o (14N) values for these three binary nitrides fit well into the previously reported linear correlation between nitrogen distance to the nearest cation and isotropic chemical shift, leading to a better correlation coefficient and reduced error margins for the fit parameters.

Graphical Abstract

1. Introduction

Binary nitrides are widely applied in industry and daily-life products covering a broad field of various applications. For example, the hexagonal form of boron nitride is an excellent lubricant in a wide temperature range [1]. AlN possesses both high thermal conductivity and piezoelectric properties and can be found, as wide-bandgap material combined with gallium or indium to III-nitride semiconductors, in light-emitting diodes [2,3]. Furthermore, 3d-metal nitrides like TiN are used in microelectronics [4], and VN with its large electrochemical potential window is of interest for supercapacitors [5]. Also, yttrium and lanthanum nitride might be of use in electronic and/or optoelectronic applications [6]. However, for cations beyond the main-group elements, as for example transition metals, the synthesis of compounds with defined stoichiometry becomes a challenging task. The deviation from the 1:1 ratio leads to nitrogen vacancies that might actually contribute to the stabilisation of the rock-salt structure types [7]. This defect formation also produces free charge carriers and strongly affects the material properties, which has been investigated mainly theoretically by means of quantum mechanical calculations [8,9,10,11]. There have been efforts to realize samples with ideal 1:1 stoichiometry, using optimized conditions or special approaches, including laser ablation [12], ion beam sputtered coatings [13] or supercritical nitrogen fluids in laser-heated diamond anvil cells [14]. There have been attempts to quantify the nitrogen content in the products by different analytical methods like the ones of Dumas [15] or Khjeldal [16], with synthesis of single crystals also reported [17,18]. Despite all these efforts, very few clear and reliable synthesis instructions can be found in the literature, especially for bulk samples. But while it is desirable to obtain stoichiometric products for the investigation of nanoparticles, thin films, powder samples and single crystals, defects and vacancies can actually be advantageous for applications like the catalytic effects of surface nitrogen-vacancies in lanthanum nitride to enable N2 activation for ammonia synthesis [19].
Regardless of the desired stoichiometry of the product, nuclear magnetic resonance (NMR) spectroscopy is a sensitive and at the same time non-destructive technique which is helpful for the analysis of binary nitrides. Since NMR directly probes the local electronic environment of a nucleus, valuable additional information complementing X-ray analysis may be obtained. Due to its sensitivity to small deviations in the local geometric and electronic environment of the nucleus of interest, it is a suitable method to detect even small amounts of vacancies [20,21] and impurities [22,23]. While NMR data are already available on several binary nitrides [24,25,26,27,28], the current work will extend the series to include three more representatives, namely with the 4d- and 5d-metals yttrium and lanthanum as well as the rare-earth element lutetium as cations. These elements were deliberately chosen so that in theory, only closed-shell states should be present for ideal 1:1 compositions. For other samples, i.e., TiN and VN, the valence electrons of the cations lead to the presence of paramagnetic properties [29,30], which made it necessary to dilute them in the NMR rotor for magic angle spinning (MAS) experiments [28,31]. For the binary nitrides investigated in this work, we report NMR spectra of the nuclides 89Y and 139La, and those of 14N. In the following, their evaluation as well as the linear correlation between the obtained isotropic chemical shift values for 14N and the shortest distances of the nitrogen atoms to the next cation in the respective structures will be described in detail.

2. Results and Discussion

The investigated compounds can be denoted as 89Y14N, 139La14N and Lu14N in good approximation due to the natural abundance of these nuclides, i.e., 100% for 89Y, 99.91 % for 139La and 99.64 % for 14N, indicating the most suitable choices for NMR experiments. In a strong external magnetic field, the various interactions each nucleus experiences at its position in the crystal structure can be considered as perturbations to the Zeeman transition energies. The influence of these interactions on the resonance frequencies may be expressed by the following equation:
ν = ν 0 + ν C S + ν I S + ν k ( 1 ) + ν k 2 ( 2 ) +
For nuclei with spin I = 1 2 , like the 89Y isotope in this work, the respective resonance frequency ν is predominantly determined by their Larmor frequency ν 0 , which is slightly modified by the chemical shift ν C S and dipolar couplings ν I S . For quadrupolar nuclei, the interaction of their non-spherical charge distributions with the electric field gradient (EFG) of the surrounding must be taken into account, usually by means of perturbation theory to first and second order [32]. These contributions are expressed by ν k ( 1 ) and ν k 2 ( 2 ) , respectively, in the above equation. Here, the parameter k denotes the relevant transition, being connected to the familiar magnetic quantum number m for the transition | m | m ± 1 by [33,34]:
k = m ± 1 2
For purpose of quantifying the quadrupolar interaction, the coupling constant χ (or C Q ) in units of [Hz] and the dimensionless asymmetry parameter η Q are commonly used. From the definition of χ (given below) it may be seen that this quantity scales directly with the quadrupole moment Q [35] of the respective nucleus and the largest eigenvalue V 33 of the EFG tensor, whereas η Q is defined as a ratio of V-tensor elements (with | V 11 | | V 22 | | V 33 | ):
χ = e Q h V 33 η Q = V 11 V 22 V 33
Considering the cubic symmetry of the investigated nitrides which all crystallize in space group F m 3 ¯ m (no. 225), both cations and anions are located at the centers of regular octahedra of the respective counter ion. Furthermore, the criterion of an highly isotropic environment is fulfilled not only in this first coordination sphere but also in the more distant ones. Hence, in the ideal rock-salt structure type, the quadrupolar coupling constants χ for nuclides with spin I > 1 2 are expected to be zero. Even for small structural distortions, which would lead to non-zero χ ’s, the asymmetry parameter η Q should be zero due to the occupation of Wyckoff positions with three-fold rotational symmetry by all atoms inducing cylindrical symmetry.
Regarding dipolar couplings, their effect on an isolated, heteronuclear spin pair scales directly with the gyromagnetic ratios γ of the two spins I and S and inversely with their cubed distance r I S to each other. As a measure for the interaction strength, the so-called dipolar coupling constant b I S = ( μ 0 γ I γ S ) / ( 8 π 2 r I S 3 ) is commonly used. Calculating b I S for the possible spin combinations in the three investigated binary nitrides, the resulting absolute values do not exceed 73.5 Hz, which is the largest value found for the heteronuclear case of 175Lu–14N. Hence, under MAS conditions with a spinning rate of 10 kHz, such small contributions should be completely averaged out, and therefore show no net effect on our spectra.
For LuN, the isotope 175Lu with its natural abundance of 97.41 % appears to be also a suitable candidate for NMR experiments. However, compared to the already large quadrupolar moment Q = 20 fm2 of 139La [36], this quantity is more than 15 times larger for 175Lu. Being one of the largest known for non-radioactive isotopes it has a value of Q = 349 fm2 [37]. Therefore, vacancies or impurities in the crystal structure of LuN are expected to have an enormous influence on the spectral appearance of the resonance line of 175Lu. Due to the more ionic bonding character between lutetium and nitrogen, the formation of such defects is more probable than for the heavier homologues of binary Lu pnictides which have been already investigated by means of NMR spectroscopy [38]. Under these unfavorable conditions we did not succeed in obtaining a spectrum of 175Lu.

2.1. NMR of Spin-1: 14N

For a spin I = 1 nucleus like 14N, there exist two single-quantum transitions with k = ± 1 2 , see Equation (2). The two significant contributions influencing the position and the shape of the resonance line are the chemical shift and the quadrupolar interaction. The effects of the latter on our samples are described by:
ν k ( 1 ) = I = 1 k · 3 χ 2 · 3 cos 2 β 1 2
For the rock-salt structure the quadrupolar interaction should vanish completely, but the 14N spectra of all the nitrides investigated here (see Figure 1) show a substantial pattern of spinning sidebands (SSBs). These SSBs are caused by the presence of anisotropic interactions, whose source can be quadrupolar coupling effects but might as well be chemical shift anisotropies (CSA). The breakdown of the ideal symmetry of the surroundings is most likely caused by nitrogen vacancies or oxygen impurities in the crystal structure. As outlined in the Introduction, achieving a perfect 1:1 stoichiometry and phase purity is a very challenging task for this compound class. The distribution of vacancies and/or impurities throughout the crystal structure results in additional line broadening, as can be seen especially in the spectrum of La14N. While these effects are smaller in the case of Y14N and Lu14N, a certain distribution of NMR parameters is also present in their corresponding spectra. This is indicated by the characteristic distribution tailing of each resonance line in the spectrum towards lower frequencies [39]. Overall, however, the envelopes of the SSB patterns do not show the characteristic lineshape with two pronounced maxima (belonging to the two k = ± 1 2 transitions) as it would be expected for static 14N powder spectra under the influence of first-order quadrupolar interactions (FOQI) [40]. This effect persists experimentally even for relatively small quadrupolar coupling constants of about 140 kHz at an MAS rate of 12 kHz, as has been observed for h-BN [24]. Consequently, possible FOQI effects have to be significantly smaller for the compounds investigated here, and second-order effects will be practically absent. Under the assumption that the observed SSB patterns are caused chiefly by the quadrupolar interaction, the 14N spectra were simulated using the Simpson [41] package (see Supplementary Material). Comparing the results with the obtained spectra in Figure 1, an upper limit for χ of about 43 kHz was determined for LaN, since for larger values of χ , the isotropic peak is not the one with the highest intensity anymore. For LuN and, especially, for YN with a significantly less intense sideband pattern, χ has to be even smaller.
The position of the resonances of highest intensity in the spectra can be identified with the isotropic chemical shift of the 14N isotope, with the assignment being verified by additional measurements with a spinning rate of 8 kHz. This leads to δ i s o values of 788 ppm for LaN, 457 ppm for YN and 384 ppm for LuN.

2.2. 14N Chemical Shift Relation to Crystal Structure

Correlation effects between NMR interaction parameters and crystal structure are of particular interest as a possible means to aid structure elucidation by NMR spectroscopy. One such possibility is to correlate the values of the isotropic chemical shift with the average distance to neighbouring atoms in the first coordination sphere for the observed nuclide [42,43,44]. To avoid the fraught debate about the constituting atoms of a coordination polyhedron (i.e., which atoms in the structure are actually considered to be neighbours), it has been suggested to just use the distance to the nearest atom, which is also an indirect measure of the packing density of the structure. This approach has been successfully applied to examples with 207 PbO x , 23 NaO x , 27 AlO x and 43 CaO x polyhedra, regardless of the compound class [45,46] and has been carried over to the class of binary nitrides with good results [28].
In this work, three new entries are added to this correlation (see Table 1). In fact, to look for the 14N signals of the new nitrides we successfully used the previously published correlation [28] to estimate the required spectral range.With Y being the first 4d metal and, La and Lu situated even in the next period of the periodic table, the 14N chemical shift of their binary nitrides extended the existing graph to larger atomic distances. In comparison to the previously published plot [28], the new values not only put this linear correlation on a broader base of data points, but they also reduce the errors of the fit parameters, while slightly improving the Pearson correlation coefficient. In Figure 2, all values of δ i s o (14N), as measured by us or obtained from the literature, are plotted against the shortest distance to the next cation. The corresponding crystal structures were taken from the Inorganic Crystal Structure Database (ICSD, provided by FIZ Karlsruhe GmbH), choosing the one closest to the average of frequently published values.

2.3. NMR of Cations

In the case of 139La, primarily the quadrupolar interaction as well as the chemical shift have to be considered. The spectra of nuclei with spin I = 7 2 consist of a central transition (CT) with k = 0 , symmetrically flanked by satellite transitions (ST) with k = ± 1 ,   ± 2 ,   ± 3 . While the position of the CT is independent of first-order effects, the spacing of the STs is directly proportional to χ as it can be seen from the following equation (for η = 0 ):
ν k ( 1 ) = I = 7 2 k · χ 14 · 3 cos 2 β 1 2
Figure 3 shows the 139La spectrum of the LaN sample. Due to the distribution of vacancies and/or impurities now already occurring in the first coordination sphere of the investigated nucleus, the obtained resonance line suffers from severe line broadening and small spinning sidebands are present. This SSB pattern is again caused by residual anisotropic interactions. Nevertheless, the intense central transition is visible as a single, symmetric resonance indicating the absence of substantial second-order effects. Neglecting CSA and only assuming first-order effects, a maximum value for the qudrupolar coupling constant can be estimated from the width of the SSB pattern. Its outermost limits belonging to the STs with k = ± 3 span a range of approximately 300 kHz. For an angle of β = 0 a value for χ m a x = ( 150 kHz · 14 ) / 3 = 700 kHz can be calculated as an upper limit from Equation (5). For a quadrupolar coupling constant of this magnitude, the corresponding SSB pattern calculated with the Simpson program for a single crystallographic site was of higher intensity and the shape of the envelope deviated from the measured one (see Supplementary Material). This reflects the fact that a substantial part of the experimentally observed broadening is due to the presence of a distribution. However, even for the overestimated value of χ = 700 kHz the second-order quadrupolar-induced shift (QIS) on the signal position of the CT would be just of about 0.25 ppm [61]:
ν Q I S ( 2 ) = I = 7 2 1 392 · χ 2 ν 0
Hence, it may be assumed that the center band is mainly affected by the chemical shift and its position at 1294 ppm can be identified in good approximation as δ i s o ( 139 La) in the LaN structure.
Yttrium offers the advantage of being a mononuclidic element. Unfortunately, NMR measurements of this isotope are complicated by its very low gyromagnetic ratio resulting in a degraded sensitivity. In addition, long relaxation times are often present for yttrium in inorganic compounds, for example about 4 h for Y2O3 or even 6.5 h for Y2O2S [62]. Therefore, only a moderate amount of yttrium-containing compounds has been studied by means of 89Y NMR spectroscopy so far. Due to its spin I = 1 2 , quadrupolar coupling effects for the isotope 89Y are not existent and the signal position is only affected by the chemical shift. However, with its 36 electrons, Y3+ reacts sensitively to changes in the surrounding electron density distribution and thus, it spans a relatively wide range of chemical shift values. In addition, defects like vacancies and impurities enlarge the CSA significantly. In our case, this leads, even under MAS conditions, to a broad signal at about 516 ppm (see Figure 4). Nevertheless, the signal position is shifted to higher frequencies compared to the ones of yttrium oxide [63] as is expected for a nucleus in octahedral nitrogen coordination instead of oxygen.

3. Materials and Methods

The three investigated binary nitrides were synthesized by direct nitridation of metal shavings in a high-temperature approach using a radio-frequency furnace (detailed synthesis conditions as well as corresponding Rietveld analyses of the products can be found in the Supplementary Materials).
All shown NMR spectra were measured at a MAS frequency of 10 kHz and were, except for 89Y, carried out on a Avance-III 500 WB spectrometer (Bruker, Karlsruhe, Germany) at LMU Munich with the polycrystalline samples filled into a 4 mm rotor. Single-pulse experiments were used for the spectra of the cations while for 14N a spin-echo pulse-sequence was applied to minimize baseline roll and background effects [64]. The respective Larmor frequencies are ν 0 ( 139 La ) = 70.647 MHz and ν 0 ( 14 N ) = 36.141 MHz. Further experimental details are summarized in Table 2. Yttrium measurements were done on a Neo 600 WB instrument (Bruker, Karlsruhe, Germany) at MPI-FKF Stuttgart, at ν 0 ( 89 Y ) = 29.408 MHz. The shown spectra were all processed in magnitude (non-phase sensitive) mode and without any baseline correction. The 14N spectra were referenced against solid NH4Cl, the other nuclides against the secondary reference of the 1H resonance of 1 % Si(CH3)4 in CDCl3.

4. Conclusions

Samples of the binary nitrides YN, LaN and LuN were synthesised and the nuclides 89Y, 139La and 14N investigated by means of MAS NMR spectroscopy. For the isotopes of yttrium and lanthanum, these results expand the comparatively scarce knowledge on NMR data of nitrogen-containing compounds. The spectra presented here also provide reference values for future NMR investigations especially for multinary nitride systems containing those rare-earth elements. In theory, quadrupolar coupling effects or chemical shift anisotropies should be completely absent in those cubic, rock-salt type structures leading to narrow, symmetric lineshapes. In practice, the appearance of the spectra shows the influence of nitrogen vacancies and/or oxygen impurities, which are known difficulties in the synthesis of these samples. However, these defects actually demonstrate the value of NMR as an analytical method: As long as their concentration is not high enough to affect the lattice parameters or even the atomic arrangement itself, their presence is hardly or, at least, very difficult to detect by means of PXRD experiments. This was confirmed by our experiments since the exact NMR signal positions and lineshapes vary slightly according to minor changes in the synthesis conditions, even for phase pure samples according to PXRD. The structural distortions and the distribution of NMR parameters are reflected in broad lineshapes or in a characteristic tailing of the resonance signals, respectively. This is yet another demonstration of the well-known fact that solid-state NMR of quadrupolar nuclei is very sensitive to deviations from perfect symmetry [65]. Nowadays, NMR investigations are frequently augmented by density functional theory (DFT) calculations. However, applying DFT methods to structures with imperfections is a challenging task. Commonly used program packages do not allow the straightforward calculation of the electron density distribution in combination with statistically distributed defects. This difficult and complex approach is still a topic of current research among DFT specialists [66,67,68]. Nevertheless, the results of the present work provide initial data points for possible synthesis optimization or further studies which might then be appropriately supplemented by quantum chemical calculations. Moreover, the respective 14N isotropic chemical shift values determined for YN, LaN and LuN fitted nicely into the previously suggested correlation of nitrogen distance to the nearest cation in each structure.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29235572/s1, Table S1: Crystallographic data of YN, LaN and LuN from Rietveld refinement. Figures S1–S3: Rietveld refinement plots of the three synthesized binary nitrides. Figures S4 and S5: Simulations for 14N and 139La NMR spectra. The authors have cited additional references within the Supporting Materials [69,70,71,72,73,74,75].

Author Contributions

Conceptualization, J.S., G.K. and T.B.; methodology, J.S., G.K. and T.B.; validation, J.S., G.K. and T.B.; formal analysis, J.S., G.K. and T.B.; investigation, J.S., G.K. and T.B.; resources, W.S.; data curation, J.S. and G.K.; writing—original draft preparation, J.S. and T.B.; writing—review and editing, J.S., G.K., W.S. and T.B.; visualization, J.S. and G.K.; supervision, W.S. and T.B.; funding acquisition, W.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Acknowledgments

The authors would like to thank Igor L. Moudrakovski (MPI-FKF, Stuttgart) for providing measurement time to acquire the 89Y spectrum and for helpful discussions relating to that spectrum.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Podgornik, B.; Kosec, T.; Kocijan, A.; Donik, Č. Tribological behaviour and lubrication performance of hexagonal boron nitride (h-BN) as a replacement for graphite in aluminium forming. Tribol. Int. 2015, 81, 267–275. [Google Scholar] [CrossRef]
  2. Dridi, Z.; Bouhafs, B.; Ruterana, P. Pressure dependence of energy band gaps for AlxGa1−xN, InxGa1−xN und InxAl1−xN. New J. Phys. 2002, 4, 94. [Google Scholar] [CrossRef]
  3. Polster, T.; Hoffmann, M. Aluminum nitride based 3D, piezoelectric, tactile sensor. Procedia Chem. 2009, 1, 144–147. [Google Scholar] [CrossRef]
  4. Lima, L.; Diniz, J.; Doi, I.; Godoy Fo, J. Titanium nitride as electrode for MOS technology and Schottky diode: Alternative extraction method of titanium nitride work function. Microelectron. Eng. 2012, 92, 86–90. [Google Scholar] [CrossRef]
  5. Liu, Y.; Wu, Q.; Liu, L.; Manasa, P.; Kang, L.; Ran, F. Vanadium nitride for aqueous supercapacitors: A topic review. J. Mater. Chem. A 2020, 8, 8218–8233. [Google Scholar] [CrossRef]
  6. Soto, G.; Moreno-Armenta, M.; Reyes-Serrato, A. First principles study on the formation of yttrium nitride in cubic and hexagonal phases. Comp. Mater. Sci. 2008, 42, 8–13. [Google Scholar] [CrossRef]
  7. Ravi, C.; Sahu, H.; Valsakumar, M.; van de Walle, A. Cluster expansion Monte Carlo study of phase stability of vanadium nitrides. Phys. Rev. B 2010, 81, 104111. [Google Scholar] [CrossRef]
  8. Porte, L. Stoichiometric ScN and nitrogen deficient scandium nitride layers studied by photoelectron spectroscopy. J. Phys. C. Solid State Phys. 1985, 18, 6701–6709. [Google Scholar] [CrossRef]
  9. Čapková, P.; Smrčok, V.; Šlma, V.; Valvoda, V.; Chalupa, P.; Ettmayer, B. Thermal Vibrations in Substoichiometric Vanadium Nitrides. Dynamical Deformation Effect. Phys. Stat. Sol. 1987, 143, 471–476. [Google Scholar] [CrossRef]
  10. Guemmaz, M.; Mosser, A.; Ahuja, R.; Parlebas, J. Theoretical and experimental investigations on elastic properties of substoichiometric titanium nitrides: Influence of lattice vacancies. Int. J. Inorg. Mater. 2001, 3, 1319–1321. [Google Scholar] [CrossRef]
  11. Deng, Z.; Kioupakis, E. Semiconducting character of LaN: Magnitude of the bandgap and origin of the electrical conductivity. AIP Adv. 2021, 11, 065312. [Google Scholar] [CrossRef]
  12. De La Cruz, W.; Díaz, J.; Mancera, L.; Takeuchi, N.; Soto, G. Yttrium nitride thin films grown by reactive laser ablation. J. Phys. Chem. Solids 2003, 64, 2273–2279. [Google Scholar] [CrossRef]
  13. Larijani, M.M.; Kiani, M.; Tanhayi, M.; Majdabadi, A. Characterization of ion beam sputtered ZrN coatings prepared at different substrate temperatures. Cryst. Res. Technol. 2011, 46, 351–356. [Google Scholar] [CrossRef]
  14. Hasegawa, M.; Yagi, T. Systematic study of formation and crystal structure of 3d-transition metal nitrides synthesized in a supercritical nitrogen fluid under 10 GPa and 1800 K using diamond anvil cell and YAG laser heating. J. Alloy. Compd. 2005, 403, 131–142. [Google Scholar] [CrossRef]
  15. Dumas, J.-B.A. Recherches de Chimie organique. Ann. Chim. Phys. 1833, 53, 164–181. [Google Scholar]
  16. Kjeldahl, J. Neue Methode zur Bestimmung des Stickstoffs in organischen Körpern. Z. Anal. Chem. 1883, 22, 366–382. [Google Scholar] [CrossRef]
  17. Motojima, S.; Baba, K.; Kitatani, K.; Takahashi, Y.; Sugiyama, K. Single crystal growth of titanium nitride by chemical vapour deposition and measurement of the linear growth rate on a (100) plane. J. Cryst. Growth 1976, 32, 141–148. [Google Scholar] [CrossRef]
  18. Wachter, P. Physical Properties of Some Stoichiometric Rare Earth Nitride Single Crystals. Adv. Mater. Phys. Chem. 2015, 5, 96–131. [Google Scholar] [CrossRef]
  19. Ye, T.-N.; Park, S.-W.; Lu, Y.; Li, J.; Sasase, M.; Kitano, M.; Tada, T.; Hosono, H. Vacancy-enabled N2 activation for ammonia synthesis on an Ni-loaded catalyst. Nature 2020, 583, 391–395. [Google Scholar] [CrossRef]
  20. Kim, N.; Stebbins, J.F. Vacancy and Cation Distribution in Yttria-Doped Ceria: An 89Y and 17O MAS NMR Study. Chem. Mater. 2007, 19, 5742–5747. [Google Scholar] [CrossRef]
  21. Wang, Q.; Li, W.; Hung, I.; Mentink-Vigier, F.; Wang, X.; Qi, G.; Wang, X.; Gan, Z.; Xu, J.; Deng, F. Mapping the oxygen structure of γ-Al2O3 by high-field solid-state NMR spectroscopy. Nat. Comm. 2020, 11, 3620. [Google Scholar] [CrossRef] [PubMed]
  22. Eckert, H.; Deo, G.; Wachs, I.; Hirt, A. Solid state 51V NMR structural studies of vanadium(V) oxide catalysts supported on TiO2(anatase) and TiO2(rutile). The influence of surface impurities on the vanadium(V) coordination. Colloids Surf. 1990, 45, 347–359. [Google Scholar] [CrossRef]
  23. Bräuniger, T.; Kempgens, P.; Harris, R.K.; Howes, A.P.; Liddell, K.; Thompson, D.P. A combined 14N/27Al nuclear magnetic resonance and powder X-ray diffraction study of impurity phases in β-sialon ceramics. Solid State Nucl. Magn. Reson. 2003, 23, 62–76. [Google Scholar] [CrossRef] [PubMed]
  24. Jeschke, G.; Hoffbauer, W.; Jansen, M. A comprehensive NMR study of cubic and hexagonal boron nitride. Solid State Nucl. Magn. Reson. 1998, 12, 1–7. [Google Scholar] [CrossRef]
  25. Bräuniger, T.; Müller, T.; Pampel, A.; Abicht, H.-P. Study of Oxygen–Nitrogen Replacement in BaTiO3 by 14N Solid-State Nuclear Magnetic Resonance. Chem. Mater. 2005, 17, 4114–4117. [Google Scholar] [CrossRef]
  26. Kempgens, P.; Britton, J. Powder-XRD and 14N magic angle-spinning solid-state NMR spectroscopy of some metal nitrides. Magn. Reson. Chem. 2016, 54, 371–376. [Google Scholar] [CrossRef]
  27. Zeman, O.E.O.; Moudrakovski, I.L.; Hartmann, C.; Indris, S.; Bräuniger, T. Local Electronic Structure in AlN Studied by Single-Crystal 27Al and 14N NMR and DFT Calculations. Molecules 2020, 25, 469. [Google Scholar] [CrossRef]
  28. Steinadler, J.; Eisenburger, L.; Bräuniger, T. Characterization of the Binary Nitrides VN and ScN by Solid-State NMR Spectroscopy. Z. Anorg. Allg. Chem. 2022, 648, e202200201. [Google Scholar] [CrossRef]
  29. Zhao, B.; Chen, L.; Luo, H. Superconducting and normal-state properties of vanadium nitride. Phys. Rev. B 1984, 29, 6198–6202. [Google Scholar] [CrossRef]
  30. Gupta, S.; Moatti, A.; Bhaumik, A.; Sachan, R.; Narayan, J. Room-temperature ferromagnetism in epitaxial titanium nitride thin films. Acta Mater. 2019, 166, 221–230. [Google Scholar] [CrossRef]
  31. MacKenzie, K.J.D.; Meinhold, R.H.; McGavin, D.G.; Ripmeester, J.A.; Moudrakovski, I.L. Titanium carbide, nitride and carbonitrides: A 13C, 14N, 15N and 47,49Ti solid-state nuclear magnetic resonance study. Solid State Nucl. Magn. Reson. 1995, 4, 193–201. [Google Scholar] [CrossRef] [PubMed]
  32. Cohen, M.H.; Reif, F. Quadrupole Effects in Nuclear Magnetic Resonance Studies of Solids. Solid State Phys. 1957, 5, 321–438. [Google Scholar] [CrossRef]
  33. Zeman, O.E.O.; Bräuniger, T. Quantifying the quadrupolar interaction by 45Sc-NMR spectroscopy of single crystals. Solid State Nucl. Magn. Reson. 2022, 117, 101775. [Google Scholar] [CrossRef]
  34. Bräuniger, T. High-Precision Determination of NMR Interaction Parameters by Measurement of Single Crystals: A Review of Classical and Advanced Methods. Molecules 2024, 29, 4148. [Google Scholar] [CrossRef] [PubMed]
  35. Pyykkö, P. Year-2017 nuclear quadrupole moments. Mol. Phys. 2018, 116, 1328–1338. [Google Scholar] [CrossRef]
  36. Jacob, C.R.; Visscher, L.; Thierfelder, C.; Schwerdtfeger, P. Nuclear quadrupole moment of 139La from relativistic electronic structure calculations of the electric field gradients in LaF, LaCl, LaBr, and LaI. J. Chem. Phys. 2007, 127, 204303. [Google Scholar] [CrossRef]
  37. Dey, W.; Ebersold, P.; Leisi, H.J.; Scheck, F.; Walter, H.K.; Zehnder, A. Nuclear spectroscopic ground-state quadrupole moments from muonic atoms. The quadrupole moment of 175Lu. Nucl. Phys. A 1979, 326, 418–444. [Google Scholar] [CrossRef]
  38. Benndorf, C.; de Oliveira Junior, M.; Bradtmüller, H.; Stegemann, F.; Pöttgen, R.; Eckert, H. Rare-earth solid-state NMR spectroscopy of intermetallic compounds: The case of the 175Lu isotope. Solid State Nucl. Magn. Reson. 2019, 101, 63–67. [Google Scholar] [CrossRef] [PubMed]
  39. d’Espinose de Lacaillerie, J.-B.; Fretigny, C.; Massiot, D. MAS NMR spectra of quadrupolar nuclei in disordered solids: The Czjzek model. J. Magn. Reson. 2008, 192, 244–251. [Google Scholar] [CrossRef]
  40. O’Dell, L.A. Direct detection of nitrogen-14 in solid-state NMR spectroscopy. Prog. Nucl. Magn. Reson. Spectrosc. 2011, 59, 295–318. [Google Scholar] [CrossRef]
  41. Bak, M.; Rasmussen, J.T.; Nielsen, N.C. SIMPSON: A general simulation program for solid-state NMR spectroscopy. J. Magn. Reson. 2011, 213, 366–400. [Google Scholar] [CrossRef]
  42. Grimmer, A.-R.; Radeglia, R. Correlation between the isotropic 29Si chemical shifts. Chem. Phys. Lett. 1984, 106, 262–265. [Google Scholar] [CrossRef]
  43. Dupree, R.; Howes, A.P.; Kohn, S.C. Natural abundance solid state 43Ca NMR. Chem. Phys. Lett. 1997, 276, 399–404. [Google Scholar] [CrossRef]
  44. Stebbins, J.F. Cation sites in mixed-alkali oxide glasses: Correlations of NMR chemical shift data with site size and bond distance. Solid State Ionics 1998, 112, 137–141. [Google Scholar] [CrossRef]
  45. Zeman, O.E.O.; Steinadler, J.; Hochleitner, R.; Bräuniger, T. Determination of the Full 207Pb Chemical Shift Tensor of Anglesite, PbSO4, and Correlation of the Isotropic Shift to Lead–Oxygen Distance in Natural Minerals. Crystals 2019, 9, 43. [Google Scholar] [CrossRef]
  46. Steinadler, J.; Zeman, O.E.O.; Bräuniger, T. Correlation of the Isotropic NMR Chemical Shift with Oxygen Coordination Distances in Periodic Solids. Oxygen 2022, 2, 327–336. [Google Scholar] [CrossRef]
  47. Eichhorn, K.; Kirfel, A.; Grochowski, J.; Serda, P. Accurate structure analysis with synchrotron radiation. An application to borazone, cubic BN. Acta Cryst. 1991, B47, 843–848. [Google Scholar] [CrossRef]
  48. Budak, E.; Bozkurt, C. Synthesis of hexagonal boron nitride with the presence of representative metals. Physics B 2010, 405, 4702–4705. [Google Scholar] [CrossRef]
  49. Wang, J.; Zhao, M.; Jin, S.F.; Li, D.D.; Yang, J.W.; Hu, W.J.; Wang, W.J. Debye temperature of wurtzite AlN determined by X-ray powder diffraction. Powder Diffr. 2014, 29, 352–355. [Google Scholar] [CrossRef]
  50. Xu, Y.-N.; Ching, W.Y. Electronic, optical, and structural properties of some wurtzite crystals. Phys. Rev. B 1993, 48, 4335–4351. [Google Scholar] [CrossRef]
  51. Paszkowicz, W.; Černý, R.; Krukowski, S. Rietveld refinement for indium nitride in the 105–295 K range. Powder Diffr. 2003, 18, 114–121. [Google Scholar] [CrossRef]
  52. Inamura, S.; Nobugai, K.; Kanamaru, F. The preparation of NaCl-type Ti1−xAlxN solid solution. J. Solid State Chem. 1987, 68, 124–127. [Google Scholar] [CrossRef]
  53. Christensen, A.N. A Neutron Diffraction Investigation on Single Crystals of Titanium Carbide, Titanium Nitride, and Zirconium Nitride. Acta Chem. Scand. A 1975, 29, 563–564. [Google Scholar] [CrossRef]
  54. Spear, K.E.; Leitnaker, J.M. Equilibrium investigations of carbon-rich V(C,N) solutions. High Temp. Sci. 1969, 1, 401–411. [Google Scholar]
  55. Klesnar, H.P.; Rogl, P. Phase relations in the ternary systems rare-earth metal(RE)-boron-nitrogen, where RE = Tb, Dy, Ho, Er, Tm, Lu, Sc and Y. High Temp. High Press. 1990, 22, 453–457. [Google Scholar]
  56. Schuster, J.C.; Bauer, J. The ternary systems Sc–Al–N and Y–Al–N. J. Less-Common Met. 1985, 109, 345–350. [Google Scholar] [CrossRef]
  57. Holleck, H.; Smailos, E. Mischnitride von Thorium mit Seltenen Erden. J. Nucl. Mater. 1980, 91, 237–239. [Google Scholar] [CrossRef]
  58. Vendl, A. Über die gegenseitige Mischbarkeit von LaN, CeN, NdN und GdN. J. Nucl. Mater. 1979, 79, 246–248. [Google Scholar] [CrossRef]
  59. Khitrin, A.K.; Fung, B.M. 14N nuclear magnetic resonance of polycrystalline solids with fast spinning at or very near the magic angle. J. Chem. Phys. 1999, 111, 8963–8969. [Google Scholar] [CrossRef]
  60. Healy, M.A.; Morris, A. Reference compounds for 14N nuclear magnetic resonance; the relative chemical shifts of aqueous nitrate ion and nitromethane. Spectrochim. Acta A 1975, 31, 1695–1697. [Google Scholar] [CrossRef]
  61. Samoson, A. Satellite transition high-resolution NMR of quadrupolar nuclei in powders. Chem. Phys. Lett. 1985, 119, 29–32. [Google Scholar] [CrossRef]
  62. Harazono, T.; Watanabe, T. Chemical Shift, Chemical Shift Anisotropy, and Spin-Lattice Relaxation Time in 89Y-MAS and-Static NMR of Yttrium Compounds. Bull. Chem. Soc. Jpn. 1997, 70, 2383–2388. [Google Scholar] [CrossRef]
  63. Florian, P.; Gervais, M.; Douy, A.; Massiot, D.; Coutures, J.-P. A Multi-nuclear Multiple-Field Nuclear Magnetic Resonance Study of the Y2O3–Al2O3 Phase Diagram. J. Phys. Chem. B 2001, 105, 379–391. [Google Scholar] [CrossRef]
  64. Kunwar, A.C.; Turner, G.L.; Oldfield, E. Solid-state spin-echo Fourier transform NMR of 39K and 67Zn salts at high field. J. Magn. Reson. 1986, 69, 124–127. [Google Scholar] [CrossRef]
  65. Koller, H.; Engelhardt, G.; Kentgens, A.P.M.; Sauer, J. 23Na NMR Spectroscopy of Solids: Interpretation of Quadrupole Interaction Parameters and Chemical Shifts. J. Phys. Chem. 1994, 98, 1544–1551. [Google Scholar] [CrossRef]
  66. Ebert, H.; Ködderitzsch, D. Minár, J. Calculating condensed matter properties using the KKR-Green’s function method—Recent developments and applications. Rep. Prog. Phys. 2011, 74, 096501. [Google Scholar] [CrossRef]
  67. Freysoldt, C.; Grabowski, B.; Hickel, T.; Neugebauer, J.; Kresse, G.; Janotti, A.; Van de Walle, C.G. First-principles calculations for point defects in solids. Rev. Mod. Phys. 2014, 86, 253–305. [Google Scholar] [CrossRef]
  68. Niklaus, R.; Minár, J.; Häusler, J.; Schick, W. First-principles and experimental characterization of the electronic properties of CaGaSiN3 and CaAlSiN3: The impact of chemical disorder. Phys. Chem. Chem. Phys. 2017, 19, 9292–9299. [Google Scholar] [CrossRef] [PubMed]
  69. Schnick, W.; Huppertz, H.; Lauterbach, R. High temperature syntheses of novel nitrido- and oxonitrido-silicates and sialons using rf furnaces. J. Mater. Chem. 1999, 9, 289–296. [Google Scholar] [CrossRef]
  70. Rietveld, H.M. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
  71. Schütze, M. Feinreinigung von Gasen mit einem hochaktiven Kupferkatalysator. Angew. Chem. 1958, 70, 697–699. [Google Scholar] [CrossRef]
  72. Coelho, A.A. TOPAS-Academic, Version 6; Coelho Software: Brisbane, Australia, 2016.
  73. Cheary, R.W.; Coelho, A. A Fundamental Parameters Approach to X-ray Line-Profile Fitting. J. Appl. Crystallogr. 1992, 25, 109–121. [Google Scholar] [CrossRef]
  74. Cheary, R.W.; Coelho, A.A.; Cline, J.P. Fundamental Parameters Line Profile Fitting in Laboratory Diffractometers. J. Res. Natl. Inst. Stand. Technol. 2004, 109, 1–25. [Google Scholar] [CrossRef] [PubMed]
  75. OriginPro, Version 2019b; OriginLab Corporation: Northhampton, MA, USA, 2019.
Figure 1. 14N-NMR MAS spectra at 10 kHz spinning speed for polycrystalline samples of LaN, YN and LuN with the vertical lines indicating the respective isotropic chemical shift.
Figure 1. 14N-NMR MAS spectra at 10 kHz spinning speed for polycrystalline samples of LaN, YN and LuN with the vertical lines indicating the respective isotropic chemical shift.
Molecules 29 05572 g001
Figure 2. Isotropic chemical shift values δ i s o , plotted against the shortest existing X–N distance in the respective crystal structures. Actual data and references are listed in Table 1. The dotted line represents the fit line according to δ i s o ( 14 N ) = a + b · d m i n with the obtained fit parameters being a = 867 ± 180 and b = 5642 ± 900 , with a Pearson correlation coefficient r of 0.91.
Figure 2. Isotropic chemical shift values δ i s o , plotted against the shortest existing X–N distance in the respective crystal structures. Actual data and references are listed in Table 1. The dotted line represents the fit line according to δ i s o ( 14 N ) = a + b · d m i n with the obtained fit parameters being a = 867 ± 180 and b = 5642 ± 900 , with a Pearson correlation coefficient r of 0.91.
Molecules 29 05572 g002
Figure 3. 139La-NMR MAS spectrum at 10 kHz spinning rate for the polycrystalline LaN sample with the center band located at 1294 ppm.
Figure 3. 139La-NMR MAS spectrum at 10 kHz spinning rate for the polycrystalline LaN sample with the center band located at 1294 ppm.
Molecules 29 05572 g003
Figure 4. 89Y-NMR MAS spectrum at 10 kHz spinning rate for the polycrystalline YN sample with the broad resonance line located at about 516 ppm.
Figure 4. 89Y-NMR MAS spectrum at 10 kHz spinning rate for the polycrystalline YN sample with the broad resonance line located at about 516 ppm.
Molecules 29 05572 g004
Table 1. 14N isotropic chemical shifts, referenced against solid NH4Cl and listed in ascending order, for the investigated and selected binary nitrides.
Table 1. 14N isotropic chemical shifts, referenced against solid NH4Cl and listed in ascending order, for the investigated and selected binary nitrides.
CompoundStructure TypeCrystallogr. Data Ref. δ iso /ppmNMR Ref.
c-BNsphalerite[47] 17.6 [24]
h-BNgraphite-related[48] 61.2 [24]
AlN *wurtzite[49] 62.4 [27]
GaN *wurtzite[50] 270.4 [26]
InN *wurtzite[51] 270.9 [26]
TiNrock-salt[52]359[25]
ZrN *rock-salt[53] 361.4 [26]
VNrock-salt[54]378[28]
LuNrock-salt[55]384this work
ScNrock-salt[56]442[28]
YNrock-salt[57]457this work
LaNrock-salt[58]788this work
* Values for δ i s o re-referenced against NH4+ shifted by 342.2 ppm relative to liquid nitromethane [59] and by 355 ppm to a NO3-solution [60].
Table 2. Selected measurement parameters for the shown NMR spectra with their respective recycle delays (d1), the accumulated number of scans (ns), the applied line-broadening (lb) factor and the resulting width of the isotropic peak at half the maximum height (fwhm).
Table 2. Selected measurement parameters for the shown NMR spectra with their respective recycle delays (d1), the accumulated number of scans (ns), the applied line-broadening (lb) factor and the resulting width of the isotropic peak at half the maximum height (fwhm).
89YN139LaNY14NLa14NLu14N
d1/s240118015240
ns51210009606256320
lb/Hz50500000
fwhm/kHz2.215.51.22.01.2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Steinadler, J.; Krach, G.; Schnick, W.; Bräuniger, T. Investigation of the Binary Nitrides YN, LaN and LuN by Solid-State NMR Spectroscopy. Molecules 2024, 29, 5572. https://doi.org/10.3390/molecules29235572

AMA Style

Steinadler J, Krach G, Schnick W, Bräuniger T. Investigation of the Binary Nitrides YN, LaN and LuN by Solid-State NMR Spectroscopy. Molecules. 2024; 29(23):5572. https://doi.org/10.3390/molecules29235572

Chicago/Turabian Style

Steinadler, Jennifer, Georg Krach, Wolfgang Schnick, and Thomas Bräuniger. 2024. "Investigation of the Binary Nitrides YN, LaN and LuN by Solid-State NMR Spectroscopy" Molecules 29, no. 23: 5572. https://doi.org/10.3390/molecules29235572

APA Style

Steinadler, J., Krach, G., Schnick, W., & Bräuniger, T. (2024). Investigation of the Binary Nitrides YN, LaN and LuN by Solid-State NMR Spectroscopy. Molecules, 29(23), 5572. https://doi.org/10.3390/molecules29235572

Article Metrics

Back to TopTop