Skip to content
Publicly Available Published by De Gruyter April 5, 2024

Ni2Mo3N: crystal structure, thermal properties, and catalytic activity for ammonia decomposition

  • Sophie Hund , Oscar Gómez-Cápiro , Holger Ruland , Eva M. Heppke and Martin Lerch EMAIL logo

Abstract

Ni2Mo3N was synthesized by ammonolysis of NiMoO4, prepared by a sol-gel-based modified Pechini route. X-ray powder diffraction measurements confirmed that Ni2Mo3N crystallizes in a filled β-Mn type (cubic space group P4132) with a lattice parameter of a = 6.6338 Å. Group theoretical methods were applied to elucidate the relation between the crystal structure of Ni2Mo3N and that of the rock salt type. The high-temperature behavior was investigated in-situ by X-ray diffraction measurements in flowing ammonia gas at temperatures up to 875 °C. Ni2Mo3N exhibits significant catalytical activity for ammonia decomposition, which is critically discussed in comparison to literature.

1 Introduction

Nitrides are materials with many interesting properties and applications. Some metal nitrides such as TiN, ZrN, and CrN [1, 2] are very hard materials which are used as protective coatings against wear and corrosion; especially TiN coatings are deposited on turbine blades [3, 4]. Si3N4 ceramics, however, are suitable for high-temperature applications due to their remarkable temperature resistance [5]. Various nitrides are also used for photocatalytic water splitting, such as γ-C3N4-based materials [6] and Ta3N5 [7] offering new routes to generate hydrogen as an energy carrier and energy storage material [8], [9], [10]. However, water is not the only potential hydrogen carrier; given that ammonia has higher hydrogen content of 17.6 wt% than water (11.1 wt%). The well-established infrastructure for synthesis and transportation of ammonia on an industrial scale offers clear benefits. Some nitrides have already been proven to be suitable catalysts for ammonia decomposition, e. g. γ-Mo2N [11, 12] and VN [13]. Ternary nitrides, for instance Fe3Mo3N and Co3Mo3N, exhibit a higher catalytic activity than the mentioned binary nitrides. Srifa et al. [14] presented an ammonia conversion of ∼63 % at 550 °C for Mo2N while Co3Mo3N reached a value of ∼94 % at the same temperature. Furthermore, the authors reported a nickel molybdenum nitride with the composition Ni3Mo3N. Having a closer look at their XRD pattern, it is clear that they synthesized Ni2Mo3N with nickel as by-product. They measured the catalytic activity for ammonia decomposition, but it is unknown whether the activity came from nickel metal or the nitride. In contrast, measurements on Ni2Mo3N, carried out by Zaman et al. [15] proved that Ni2Mo3N is a promising material with an ammonia conversion of more than 90 % at 600 °C. The crystal structure of Ni2Mo3N has been described by Herle et al. [16] and Weil et al. [17] as filled β-Mn-type structure (space group P4132). An overview of phases crystallizing in this structure type was given by Jeitschko et al. [18]. In our work we take a closer look at the crystal structure, also using group theoretical concepts, the high-temperature behavior, and the catalytic activity of Ni2Mo3N for ammonia decomposition.

2 Results and discussion

2.1 Chemical characterization

Ni2Mo3N was prepared as phase-pure powder of dark gray color with metallic sheen. The expected cation ratio for nickel molybdenum nitride of Ni:Mo = 2:3 was confirmed by energy dispersive X-ray spectroscopy (EDX) with 39.7 at% for nickel and 60.3 at% for molybdenum, normalized to the cations. The anion contents were determined by means of hot gas extraction: 3.41 wt% nitrogen and 0.49 wt% oxygen was measured (calculated: 3.34 wt% nitrogen, 0 wt% oxygen). The standard deviations for EDX and hot gas extraction are estimated to be 2 %. In our previous work on Fe3Mo3N [19] we showed that the oxygen content is caused by an oxygen-rich surface layer (scanning transmission electron microscopy and electron energy loss spectroscopy (STEM-EELS)). No oxygen was found in the bulk of the crystals. Consequently, the oxygen content was neglected for the Rietveld [20] refinement. We assume that the composition is comparable for the nickel molybdenum nitride presented here and formulated as Ni2Mo3N.

2.2 Crystal structure

The powder X-ray diffraction (XRD) pattern with the results of the Rietveld refinement is presented in Figure 1. The structural model reported by Weil et al. [17] was used as starting input.

Figure 1: 
Powder X-ray diffraction pattern of Ni2Mo3N with the results of the Rietveld refinement. The measured pattern is given in red, the calculated one in black, while the reflection positions are presented in green and the difference plot in blue color.
Figure 1:

Powder X-ray diffraction pattern of Ni2Mo3N with the results of the Rietveld refinement. The measured pattern is given in red, the calculated one in black, while the reflection positions are presented in green and the difference plot in blue color.

The background was fitted by linear interpolation between 27 points. Table 1 lists the Wyckoff positions, atomic coordinates, and Debye-Waller factors for Ni2Mo3N. Table 2 sums up the results of the Rietveld refinement.

Table 1:

Refined atomic parameters for Ni2Mo3N (standard deviations in parentheses).

Atom Wyckoff x y z Biso2)
Ni 8c 0.06697(19) x x 0.61(6)
Mo 12d 1/8 0.20157(9) 0.45157(9) 0.41(4)
N 4a 3/8 3/8 3/8 1a
  1. aNot refined.

Table 2:

Results of the Rietveld refinement for Ni2Mo3N (standard deviations in parentheses).

Ni2Mo3N
Space group P4132
Structure type Filled β-Mn type
a, Å 6.63384(9)
V, ų 291.941(6)
Z 4
Dcalc, g cm−3 9.54
Diffractometer PANalytical X’Pert Pro
Radiation CuKα
Wavelengths, Å 1.54056, 1.54439
R p 0.0102
R wp 0.0136
R exp 0.0110
R Bragg 0.0152
S 1.24

Ni2Mo3N crystallizes in the filled β-Mn type (space group P4132). The cubic unit cell is shown in Figure 2. In the β-Mn type, the atoms occupy the 8c and 12d sites, the filled β-Mn type contains additional atoms on the 4a site (here: nitrogen). The lattice parameter of β-Mn is a = 6.315 Å, which is slightly smaller than that of Ni2Mo3N (a = 6.634 Å) [21]. The (NMo6) octahedra, depicted in Figure 2, are corner-shared, forming a three-dimensional network similar to that in Fe3Mo3N. The screw axis provides rotation of the polyhedra. Weil et al. [17] described the relation between the filled β-Mn type and the η-carbide type (e.g., adopted by Fe3Mo3N). Besides the obvious fact that the unit cell of the η-carbide type is much larger (for Fe3Mo3N: a = 11.0777 Å [19]) than that of the filled β-Mn type, the main difference seems to be the space available for the metal atoms that are located between the octahedra. In the η-carbide type, the octahedra form rings with a hexagonal shaped void, but in the β-Mn type the void inside the ring has a pentagonal geometry (see Figure 3).

Figure 2: 
Crystal structure of Ni2Mo3N (Ni green, Mo violet, N blue, and (NMo6) octahedra light blue colored), the unit cell is presented in black.
Figure 2:

Crystal structure of Ni2Mo3N (Ni green, Mo violet, N blue, and (NMo6) octahedra light blue colored), the unit cell is presented in black.

Figure 3: 
Rings of (NMo6) octahedra in (a) Fe3Mo3N with a hexagonal shaped void and (b) in Ni2Mo3N with a pentagonal shaped void.
Figure 3:

Rings of (NMo6) octahedra in (a) Fe3Mo3N with a hexagonal shaped void and (b) in Ni2Mo3N with a pentagonal shaped void.

The coordination spheres of all atom types in Ni2Mo3N are illustrated in Figure 4. The polyhedra around the nickel atoms may be described as strongly distorted icosahedra consisting of three nickel and nine molybdenum atoms. In the crystal structure of Fe3Mo3N [19], distorted icosahedra are also present – but with a different composition of (Fe[Fe6Mo6]). As described above, (NMo6) octahedra can be found in both compounds. Interestingly, the coordination spheres around the molybdenum atoms are different: An icosahedron-like environment was found for Fe3Mo3N, while Ni2Mo3N exhibits a complex environment with a coordination number of 14. A list of selected interatomic distances is given in Table 3. The lengths are in accordance with values in the literature [17] with exception of the Mo–Mo distances. In Ni2Mo3N, there are four Mo–Mo distances of 2.775 Å and two of 2.818 Å – which are reversed in the reference.

Figure 4: 
Coordination spheres for Ni2Mo3N with nickel in green, molybdenum in violet, and nitrogen in blue. The faces of the polyhedra have the colors of the central atoms.
Figure 4:

Coordination spheres for Ni2Mo3N with nickel in green, molybdenum in violet, and nitrogen in blue. The faces of the polyhedra have the colors of the central atoms.

Table 3:

Selected interatomic distances for Ni2Mo3N (in Å) compared to the values reported by Weil et al. [17] (standard deviations in parentheses).

Ni2Mo3N Ni2Mo3N [17]
Ni–Ni 2.4686(19) [3×] Ni–Ni 2.468(9) [3×]
Ni–Mo 2.8169(15) [3×] Ni–Mo 2.817(3) [3×]
Ni–Mo 2.7409(15) [3×] Ni–Mo 2.739(8) [3×]
Ni–Mo 2.7304(15) [3×] Ni–Mo 2.732(2) [3×]
Mo–Mo 2.7752(9) [4×] Mo–Mo 2.815(9) [4×]
Mo–Mo 2.8182(6) [2×] Mo–Mo 2.776(3) [2×]
Mo–Ni 2.8169(15) [2×] Mo–Ni 2.817(3) [2×]
Mo–Ni 2.7409(15) [2×] Mo–Ni 2.739(8) [2×]
Mo–Ni 2.7304(15) [2×] Mo–Ni 2.732(2) [2×]
Mo–N 2.0814(6) [2×] Mo–N 2.081(2) [2×]
N–Mo 2.0814(6) [6×] N–Mo 2.081(2) [6×]

Using group-theoretical concepts [22], [23], [24], the crystal structure of Ni2Mo3N can be derived from the rock salt type. Figure 5 shows the group-subgroup relations (Bärnighausen formalism) for one possible pathway: we begin with space group F m 3 m for the rock salt type, where the atoms occupy the Wyckoff positions 4a and 4b. The first step is a translationengleiche transition of index 2 (t2), resulting in space group F432. The following decentering of the lattice by a klassengleiche transition of index 4 (k4) leads to space group P4232. In the latter step a screw axis is created and the 4a and 4b site become 4c and 4b, respectively. Next, space group I4132 is reached by doubling the lattice parameters and a klassengleiche transition of index 4 (k4) takes place again. The 4b site splits to 8a and a 24h, 4c splits to an 8b and a 24g site. The final symmetry reduction leads to space group P4132, adopted by Ni2Mo3N, by a klassengleiche transition of index 2 (k2), splitting the 8b site to 4a, occupied by nitrogen, and 4b. The 24g site splits to two non-equivalent 12d sites, one of them occupied by molybdenum. The 8a site becomes an 8c site, occupied by nickel, and the 24h site is transformed to a 24e site. Interestingly, in a previous work by Stegemann et al. [25], an alternative Bärnighausen tree starting from the alloy NaAu2, adopting the MgCu2 type, and ending with Na2Au3Al has been shown.

Figure 5: 
Group-subgroup relations (Bärnighausen formalism) between the rock salt type and the crystal structure of Ni2Mo3N (filled β-Mn type).
Figure 5:

Group-subgroup relations (Bärnighausen formalism) between the rock salt type and the crystal structure of Ni2Mo3N (filled β-Mn type).

The high-temperature behavior of Ni2Mo3N was investigated by in-situ XRD measurements by heating the powder in flowing ammonia in steps of 50 K to 875 °C. Interestingly, and in clear contrast to Fe3Mo3N [19], the material was stable over the whole temperature range.

2.3 Catalytic measurements

The results of the decomposition of NH3 to N2 and H2 in the range of 400–600 °C for the newly prepared Ni2Mo3N are given in Figure 6 and Table 4. For comparison, an industrial Ni-based reference, and the inert material (SiC) used as reactor filler are included, the latter being chosen in order to prove that it is practically inactive. Although the industrial reference is more active, Ni2Mo3N shows a significant conversion and might be an interesting starting point for further investigations. Especially a comparison of the conversion between the heating and cooling cycle is indicating a slight activation for both catalysts. A direct comparison of the conversion for a similar nitride catalyst reported in the literature is not possible due to the different test conditions applied, and missing information [15]. There appears to be a typing error in the given units as the hydrogen productivities presented are utopian high and do not fit to the gas hourly space velocity (GHSV) applied. A comparison of the hydrogen productivities reveals that with 129 mmol H2 gcat−1 h−1 a similar productivity at 500 °C for the Ni2Mo3N catalyst investigated in this work and with 122 mmol H2 gcat−1 h−1 reported by Zaman et al. [15] there is a similar performance. At higher temperatures the values obtained in this work are far higher than the ones reported by Zaman et al. (335 vs. 170 mmol H2 gcat−1 h−1 at 550 °C and 684 vs. 206 mmol H2 gcat−1 h−1 at 600 °C), most likely due to thermodynamic or transport limitations set under the different test conditions applied. The difficulties in comparing catalysts from both works highlight the need to apply standardized test conditions and to report the data obtained appropriately to enable a reliable comparison of literature studies and identify suitable catalysts for industrial application [26, 27].

Figure 6: 
Conversion of the Ni2Mo3N and an industrial Ni-based catalyst as reference for the decomposition of ammonia in a temperature range between 400 and 600 °C. Reaction conditions: p = 1 atm, flow rate = 0.6 mL mgcat−1 min−1, rate of temperature changes between the steps = 2 K min−1.
Figure 6:

Conversion of the Ni2Mo3N and an industrial Ni-based catalyst as reference for the decomposition of ammonia in a temperature range between 400 and 600 °C. Reaction conditions: p = 1 atm, flow rate = 0.6 mL mgcat−1 min−1, rate of temperature changes between the steps = 2 K min−1.

Table 4:

Summary of the results of the ammonia decomposition tests for the Ni2Mo3N samples presented here and an industrial Ni-based catalyst as reference in the temperature range between 400 and 600 °C. Reaction conditions: p = 1 atm, flow rate = 0.6 mL mgcat−1 min−1, rate of temperature changes between the steps = 2 K min−1.

Catalyst Parameter Temperature (°C)
400 450 500 550 600a
Ni 2 Mo 3 N XNH3 heating (%) 1.1 ± 0.1 2.4 ± 0.1 5.4 ± 0.2 12.7 ± 0.5 28.4 ± 1.1
XNH3 cooling (%) 1.0 ± 0.1 2.1 ± 0.1 5.3 ± 0.3 14.0 ± 0.4
H2 prod. heat. (mmol gcat−1 h−1) 26.8 ± 1.8 58.4 ± 2.2 128.8 ± 3.7 305.5 ± 11.3 683.6 ± 25.9
H2 prod. cool. (mmol gcat−1 h−1) 25.2 ± 0.6 49.4 ± 2.6 128.6 ± 6.1 336.0 ± 8.9
Industrial Ni reference XNH3 heating (%) 7.3 ± 0.1 18.9 ± 0.7 39.4 ± 1.1 70.6 ± 1.2 97.2 ± 0.3
XNH3 cooling (%) 8.9 ± 0.4 21.0 ± 0.7 41.9 ± 1.1 71.4 ± 0.8
H2 prod. heat. (mmol gcat−1 h−1) 176.7 ± 2.2 456.4 ± 17.9 947.4 ± 25.9 1699.4 ± 30.6 2340.7 ± 9.1
H2 prod. cool. (mmol gcat−1 h−1) 214.2 ± 9.3 506.1 ± 17.8 1008.7 ± 27.1 1717.4 ± 20.3
  1. aThe values for the segments heating and cooling are the same at 600 °C according to the temperature program.

3 Conclusions

Ni2Mo3N was prepared and characterized by EDX and powder XRD measurements. It crystallizes in a filled β-Mn type (cubic space group P4132), which was confirmed by powder XRD measurements and Rietveld refinement. A possible group-subgroup relation between the crystal structures of Ni2Mo3N and the rock salt type is presented. XRD measurements carried out in-situ under industrial relevant conditions as applied for ammonia decomposition revealed the structural stability of the material. Ni2Mo3N has been shown to have high catalytic activity for ammonia decomposition. Although it exhibits a lower performance than the industrial reference, it seems to be a promising candidate for optimization: further work, especially the addition of promotors may lead to a significant increase of the catalytic activity. From the difficulties met when comparing our results with data reported in the literature it is evident that there is a strong need for a standardization of the test conditions. This will be elaborated and specified in detail in forthcoming contributions.

4 Experimental section

4.1 Synthesis

In this work, Ni2Mo3N was synthesized by ammonolysis of an oxidic precursor (NiMoO4) which was prepared via a modified Pechini [28] sol-gel route. The modified Pechini route, in contrast to the original method, starts from aqueous solutions to reduce the amount of organics [29]. Nickel(II) chloride (3.1 g) and a 12-fold molar excess of citric acid (55.2 g) were dissolved in 250 mL of a mixture of equal parts absolute ethanol and distilled water to receive a nickel citrate solution. For the molybdenum citrate solution, molybdenum(V)-chloride (4.8 g) and a 12-fold molar excess of citric acid (40.6 g) were employed, and absolute ethanol (250 mL) was added dropwise. The exact metal ion concentrations of the citrate solutions were determined gravimetrically. Both citrate solutions were mixed in the desired metal ratio (Ni:Mo = 2:3) and an at least 17-fold molar excess of ethylene glycol was added. Heating in steps of 50 K to 350 °C resulted in a gel that was transformed through overnight calcination at 450 °C to a green-brown powder: the oxidic precursor NiMoO4. The solvents and additives evaporated completely during the heating process. The last step was the nitridation via ammonolysis: the sample was annealed in a tube furnace at 875 °C for 20 h with a heating rate of 400 K h−1 and an ammonia flow rate of 9 L h−1.

4.2 Chemical and structural characterization

For energy-dispersive X-ray spectroscopy (EDX) a DSM 982 GEMINI spectrometer (Carl Zeiss AG, Oberkochen, Germany) equipped with a XFlash 6 | 60 detector (Bruker, Billerica, USA) was used. The standard deviation of this method is estimated to be 2 %. A LECO TC-300/EF-300 N/O analyzer was used for hot gas extraction (ZrO2 and steel as standard materials). Hereby, the standard deviation is estimated to be 2 %. XRD was performed using a PANalytical X’Pert Pro powder diffractometer operating in a Bragg-Brentano set-up. Nickel-filtered CuKα radiation with wavelengths of 1.54056 and 1.54439 Å was used for the experiments. The 2θ measuring range was set to 10°–120° with a step size of 0.026°. The program package FullProf Suite 2021 [30] was used to execute the Rietveld refinements. In-situ powder X-ray diffraction measurements were performed with a Rigaku SmartLab 3 kW system (CuKα radiation) equipped with Reaktor X.

Further details of the crystal structure investigation may be obtained from FIZ Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany (Fax: +49 7247 808 666; E-mail: ) on quoting the deposition number CSD-2309672.

4.3 Activity test on ammonia decomposition

The catalytic activity for ammonia decomposition of Ni2Mo3N was tested in a fixed-bed reactor. Two hundred milligrams of the catalyst (particle size 200–300 µm) was diluted in a mass ratio of 1:2 with SiC (particle size 300–400 µm). The use of different particle sizes enables a later separation of the spent catalyst from the dilution material. The test procedure started with heating the sample with a rate of 1 K min−1 in nitrogen (quality 5.0) to 400 °C applying a flow rate of 1 mL mgcat−1 min−1. After reaching the desired temperature, the gas supply was switched to ammonia (quality 5.0) with a flow rate of 0.6 mL mgcat−1 min−1 and the temperature was kept constant for 60 min. Afterwards, the temperature was increased in steps of 50 K to 600 °C and back to the initial temperature applying a heating rate of 2 K min−1. For each step the temperature was kept constant for 60 min. For comparison, an industrial Ni-based catalyst was tested under similar conditions, but with a reductive pretreatment with 10 % H2 in N2. In addition, a blank test was made filling the reactor only with SiC using the same procedure as described above. Product gas analysis was performed via IR detectors for quantification of NH3 and H2O as well as a TCD for H2 (Emerson XStream XEGP). A detailed description of the experimental setup and test conditions can be found in ref. [26].


Dedicated to Professor Thomas Bredow of the University of Bonn on the occasion of his 60th birthday.



Corresponding author: Martin Lerch, Institut für Chemie, Technische Universität Berlin, Straße des 17. Juni 135, 10623 Berlin, Germany, E-mail:

Funding source: Deutsche Forschungsgemeinschaft

Award Identifier / Grant number: (INST 131/734-1 FUGG)

Acknowledgment

Many thanks go to the Federal Ministry of Education and Research, Germany (Bundesministerium für Bildung und Forschung, BMBF, Verbundvorhaben TransHyDE Forschungsverbund AmmoRef, supportcodes: 03HY203C, 03HY203A), for funding. The industrial reference catalyst used in this study was kindly provided by our project partner Clariant. Furthermore, we thank Dr. Stefan Berendts for hot gas extraction analyses and in-situ powder XRD measurements. Christoph Fahrenson from the Zentraleinrichtung Elektronenmikroskopie (Zelmi) at the TU Berlin carried out the EDX measurements.

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

  2. Research funding: This work was also supported by the Deutsche Forschungsgemeinschaft (INST 131/734-1 FUGG).

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

  4. Data availability: The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

1. Milošev, I., Strehblow, H.-H., Navinšek, B. Thin Solid Films 1997, 303, 246–254; https://doi.org/10.1016/s0040-6090(97)00069-2.Search in Google Scholar

2. Dinesh Kumar, D., Kumar, N., Kalaiselvam, S., Dash, S., Jayavel, R. Surface. Interfac. 2017, 7, 74–82; https://doi.org/10.1016/j.surfin.2017.03.001.Search in Google Scholar

3. Slot, H. M., Gelinck, E. R. M., Rentrop, C., van der Heide, E. Renewable Energy 2015, 80, 837–848; https://doi.org/10.1016/j.renene.2015.02.036.Search in Google Scholar

4. Rajendran, R. Eng. Fail. Anal. 2012, 26, 355–369; https://doi.org/10.1016/j.engfailanal.2012.07.007.Search in Google Scholar

5. Klemm, H. J. Am. Ceram. Soc. 2010, 93, 1501–1522.10.1111/j.1551-2916.2010.03839.xSearch in Google Scholar

6. Wen, J., Xie, J., Chen, X., Li, X. Appl. Surf. Sci. 2017, 391, 72–123; https://doi.org/10.1016/j.apsusc.2016.07.030.Search in Google Scholar

7. Nurlaela, E., Ziani, A., Takanabe, K. Mater. Renew. Sustain. Energy 2016, 5, 18 (21 pages); https://doi.org/10.1007/s40243-016-0083-z.Search in Google Scholar

8. Ng, K. H., Lai, S. Y., Cheng, C. K., Cheng, Y. W., Chong, C. C. Chem. Eng. J. 2021, 417, 128847; https://doi.org/10.1016/j.cej.2021.128847.Search in Google Scholar

9. Jafari, T., Moharreri, E., Amin, A. S., Miao, R., Song, W., Suib, S. L. Molecules 2016, 21, 900 (29 pages); https://doi.org/10.3390/molecules21070900.Search in Google Scholar PubMed PubMed Central

10. Acar, C., Dincer, I., Naterer, G. F. Int. J. Energy Res. 2016, 40, 1449–1473; https://doi.org/10.1002/er.3549.Search in Google Scholar

11. Wise, R. S., Markel, E. J. J. Catal. 1994, 145, 333–343.10.2307/3978050Search in Google Scholar

12. Lee, H. J., Choi, J.-G., Colling, C. W., Mudholkar, M. S., Thompson, L. T. Appl. Surf. Sci. 1995, 89, 121–130; https://doi.org/10.1016/0169-4332(95)00025-9.Search in Google Scholar

13. Oyama, S. T. J. Catal. 1992, 133, 358–369; https://doi.org/10.1016/0021-9517(92)90246-e.Search in Google Scholar

14. Srifa, A., Okura, K., Okanishi, T., Muroyama, H., Matsui, T., Eguchi, K. Catal. Sci. Technol. 2016, 6, 7495–7504; https://doi.org/10.1039/c6cy01566b.Search in Google Scholar

15. Zaman, S. F., Jolaoso, L. A., Podila, S., Al-Zahrani, A. A., Alhamed, Y. A., Driss, H., Daous, M. M., Petrov, L. Int. J. Hydrogen Energy 2018, 43, 17252–17258; https://doi.org/10.1016/j.ijhydene.2018.07.085.Search in Google Scholar

16. Herle, P. S., Hegde, M. S., Sooryanarayana, K., Row, T. N. G., Subbanna, G. B. Inorg. Chem. 1998, 37, 4128–4130; https://doi.org/10.1021/ic980211t.Search in Google Scholar PubMed

17. Weil, K. S., Kumta, P. N., Grins, J. J. Solid State Chem. 1999, 146, 22–35; https://doi.org/10.1006/jssc.1999.8296.Search in Google Scholar

18. Jeitschko, W., Nowotny, H., Benesovsky, F. Monatsh. Chem. 1964, 95, 1212–1218; https://doi.org/10.1007/bf00904716.Search in Google Scholar

19. Hund, S., Gómez-Cápiro, O., Dembélé, K., Berendts, S., Lunkenbein, T., Ruland, H., Heppke, E. M., Lerch, M. Z. Anorg. Allg. Chem. 2023, 649, 2023001522-1–e202300152-6; https://doi.org/10.1002/zaac.202300152.Search in Google Scholar

20. Rietveld, H. M. J. Appl. Crystallogr. 1969, 2, 65–71; https://doi.org/10.1107/s0021889869006558.Search in Google Scholar

21. Shoemaker, C. B., Shoemaker, D. P., Hopkins, T. E., Yindepit, S. Acta Crystallogr. 1978, B34, 3573–3576.10.1107/S0567740878011620Search in Google Scholar

22. Bärnighausen, H. MATCH, Commun. Math. Comput. Chem. 1980, 9, 139–175.10.1007/BF01674443Search in Google Scholar

23. Aroyo, M. I., Perez-Mato, J. M., Capillas, C., Kroumova, E., Ivantchev, S., Madariaga, G., Kirov, A., Wondratschek, H. Z. Kristallogr. 2006, 221, 15–27; https://doi.org/10.1524/zkri.2006.221.1.15.Search in Google Scholar

24. Müller, U. Z. Anorg. Allg. Chem. 2004, 630, 1519–1537; https://doi.org/10.1002/zaac.200400250.Search in Google Scholar

25. Stegemann, F., Benndorf, C., Zhang, Y., Bartsch, M., Zacharias, H., Fokwa, B. P. T., Eckert, H., Janka, O. Inorg. Chem. 2017, 56, 1919–1931; https://doi.org/10.1021/acs.inorgchem.6b02480.Search in Google Scholar PubMed

26. Ristig, S., Poschmann, M., Folke, J., Gómez-Cápiro, O., Chen, Z., Sanchez-Bastardo, N., Schlögl, R., Heumann, S., Ruland, H. Chem. Ing. Tech. 2022, 94, 1413–1425; https://doi.org/10.1002/cite.202200003.Search in Google Scholar

27. Gómez-Cápiro, O., Ristig, S., Folke, J., Ruland, H. Energy Technol. 2023, accepted; https://doi.org/10.1002/ente.202300996.Search in Google Scholar

28. Pechini, M. P. Method of preparing lead and alkaline earth titanates and niobates and coating method using the same to form a capacitor. US3,330,697, July 11, 1967.Search in Google Scholar

29. Sunde, T. O. L., Grande, T., Einarsrud, M.-A. Modified Pechini synthesis of oxide powders and thin films. In Handbook of Sol-Gel Science and Technology; Klein, L., Aparicio, M., Jitianu, A., Eds. Springer: Cham (Switzerland), 2016; pp. 1–30.10.1007/978-3-319-19454-7_130-1Search in Google Scholar

30. Rodríguez-Carvajal, J. FullProf2000, a program for Rietveld refinement and pattern matching analysis. In Satellite Meeting on Powder Diffraction of the 15th International Congress of the IUCr; IUCr: Toulouse, France, 1990; p. 127.Search in Google Scholar

Received: 2023-08-28
Accepted: 2023-10-31
Published Online: 2024-04-05
Published in Print: 2024-04-25

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 30.4.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2023-0071/html
Scroll to top button