Home Mechanochemical synthesis and structural evaluation of a metastable polymorph of Ti3Sn
Article Publicly Available

Mechanochemical synthesis and structural evaluation of a metastable polymorph of Ti3Sn

  • Eva M. Heppke EMAIL logo , Aylin Koldemir , Rainer Pöttgen , Thomas Bredow and Martin Lerch EMAIL logo
Published/Copyright: March 13, 2023
Become an author with De Gruyter Brill

Abstract

A metastable polymorph of Ti3Sn (called c-Ti3Sn) exhibiting the cubic Cr3Si-type structure was prepared by a mechanochemical route. At temperatures of about 450 °C, it transforms to the well-known hexagonal phase (Ni3Sn type, called h-Ti3Sn). 0.5% of iron was incorporated into the material originating from the steel beaker and the steel balls. However, quantum-chemical calculations show that this should not lead to a stabilization of the Cr3Si type. c-Ti3Sn shows a single signal at an isomer shift of δ = 1.70(1) mm s−1 in its 119Sn Mössbauer spectrum at 78 K.

1 Introduction

In the course of preparing new catalyst materials, we intended to synthesize an intermetallic phase of the type A3B with composition Ti3Sn which may be used as a precursor for further transformations. h-Ti3Sn as an intermetallic compound has been known since the 1950s and exhibits an “unusually high damping capacity” [1] which makes it interesting for application in various composite materials. It is also of interest as an anode material for Li-ion batteries [2, 3] as Sn-based materials appear to have high storage capacities [4], [5], [6], [7], [8].

h-Ti3Sn is known to adopt the Ni3Sn-type structure (space group P63/mmc) [9]. This crystal structure type is derived from a hexagonal closest packing (hcp) of the atoms and represents an ordered variant with layers exhibiting a 3:1 occupation of Ti and Sn. An investigation of the solidification process of undercooled Ti3Sn melts revealed two metastable martensitic Ti3Sn phases (an orthorhombic and a monoclinic phase) [10]. A reversible phase transition (starting at around 260 K) to an orthorhombic phase (o-Ti3Sn) with space group Cmcm was observed by Du et al. upon cooling [11]. Also, for non-stoichiometric Ti3Sn, a phase transition to an orthorhombic phase with space group Cmcm has been mentioned [12]. A high-pressure phase of Ti3Sn is also known, crystallizing in the Cu3Au type [13].

In this contribution, we report on the preparation of c-Ti3Sn by ball milling, an economical and environmentally friendly synthesis method that is now used in inorganic as well as organic synthesis, including catalysts [14], [15], [16]. In this process, the mechanical energy applied induces a chemical reaction of the starting materials used. Various parameters such as milling time and atmosphere, rotational speed, material and size of the grinding balls and beakers, and also the ball-to-powder weight or volume/surface ratio can be varied in order to optimize the milling process and thus ensure a successful outcome of an experiment [17, 18]. It should be mentioned that this procedure may be also of interest for the preparation of metastable compounds [17, 19]. Here we present the synthesis of a new metastable polymorph of Ti3Sn and its structural evaluation (on an experimental but also theoretical level).

2 Results and discussion

As described in the experimental section, the sample we report here was prepared by ball milling (steel beaker) with a subsequent annealing step. The powder exhibits a dark grey color and appears to be stable in air. According to energy dispersive X-ray spectroscopy (EDX), the chemical composition is in good agreement with its theoretical one showing a Ti/Sn ratio of 3:1. In addition, an average of 0.5 at% iron was detected which is most likely due to abrasion of the used steel beaker (mean values for Ti: 74.7(7) at%, Sn: 24.9(6) at%, Fe: 0.5(2) at%).

Due to the low annealing temperature (450 °C) maintained for the preparation of our sample, its crystallinity is rather poor. However, reflections are clearly visible in the diffraction pattern and, consequently, a structural evaluation is still possible. It should be stated that the observed reflections in the diffraction pattern do not match h-Ti3Sn with Ni3Sn-type structure or any other previously described polymorph of Ti3Sn. Applying a slightly higher crystallization temperature (550 °C) led to the formation of h-Ti3Sn. With longer holding times at 450 °C, the hexagonal phase also began to form and a two-phase-mixture was obtained. The crystal structure of our sample has been determined using X-ray powder diffraction. The diffraction pattern was first examined using the software X’pert HighScore Plus (PANalytical, Almelo, The Netherlands) referring to the Inorganic Crystal Structure Database ICSD (FIZ Karlsruhe). The results suggested the Cr3Si-type structure with space group Pm 3 n of cubic symmetry. Interestingly, c-Ti3Sn with Cr3Si-type structure has already been mentioned in the literature [20, 21]. However, a detailed description of the synthesis or a structural evaluation cannot be found. Unfortunately, access to presumably available, more detailed information was not possible [22].

Consequently, Rietveld refinements [23] using the Cr3Si-type structure model were performed. A refinement using a single-phase model led to a mediocre residual value of Rwp = 0.0392. The reflection near 2 θ  = 40° appears to be slightly broader than the other ones and cannot be fitted into the single-phase model (see Figure 1a). This indicates the presence of an additional phase. Based on diffraction pattern analysis, possible side phases such as α -Ti or Ti2Sn could be identified, both crystallizing in the hexagonal crystal system with space group P63/mmc [2425]. It should be noted that the strongest reflection of these phases appears at around 2 θ  = 40°. Due to the fact that the secondary phase could not be clearly assigned, only a Le-Bail refinement of the side phase was performed. Finally, an improved residual value of Rwp = 0.0233 could be reached. From the refined lattice parameters, the presence of titanium metal seemed to be a reasonable assumption. As a result, the broadened reflection could be refined thoroughly and the bump in the difference plot finally was smoothed out (see Figure 1b). Interestingly, the formation of α -Ti as a minor phase using arc melting for the preparation of Ti3Sn samples was observed as well [26]. The X-ray powder diffraction pattern with the results of the Rietveld refinements are shown in Figure 1a and b. Refined structural and atomic parameters are summarized in Tables 1 and 2.

Figure 1: 
(a) Powder X-ray diffraction pattern of the mechanochemically prepared c-Ti3Sn phase with the results of the Rietveld refinement using the Cr3Si-type model (space group Pm




3
‾



$\overline{3}$


n). Red: measured, black: calculated, blue: measured 



−


${-}$


 calculated. (b) Results of the Rietveld refinement considering a secondary phase in space group P63/mmc.
Figure 1:

(a) Powder X-ray diffraction pattern of the mechanochemically prepared c-Ti3Sn phase with the results of the Rietveld refinement using the Cr3Si-type model (space group Pm 3 n). Red: measured, black: calculated, blue: measured calculated. (b) Results of the Rietveld refinement considering a secondary phase in space group P63/mmc.

Table 1:

Results of the Rietveld refinement for mechanochemically prepared Ti3Sn (standard deviations in parenthesis).

Empirical formula Ti3Sn
Color Dark grey
Structure type Cr3Si
Space group Pm 3 n
Crystal system Cubic
Z 2
a, Å 5.240(2)
V, Å3 143.86(11)
Calculated density, g/cm3 6.06
Diffractometer PANalytical X’Pert MDP Pro
Radiation Cu radiation
Wavelength, Å λ1 = 1.54056, λ2 = 1.54439
R p 0.0170
R wp 0.0233
R exp 0.0256
R Bragg 0.0139
S 0.91
Table 2:

Atomic and displacement parameters for Ti3Sn (standard deviations in parenthesis).

Atom Wyckoff x y z s.o.f Biso2
Sn 2a 0 0 0 1 1.32(18)
Ti 6c 1/4 0 1/2 1 1.6(2)

The Cr3Si type belongs to one of the most dense crystal structures [27]. The Sn atoms form a bcc-related substructure where the six faces of the unit cell are occupied by pairs of Ti atoms. These units are assembled into chains in three perpendicular directions through the crystal structure. The metal–metal distances are rather short which formally results in a coordination number of 14 for the Ti atoms, which are surrounded by ten Ti and four Sn atoms (Figure 2). The Ti–Ti distances in c-Ti3Sn correlate well with those in other Cr3Si-type phases (e.g. Ti3Sb, Ti3Au calculated from [28, 29]) and vary from 2.6207(11) Å (2×) to 3.2097(10) Å (8×). In comparison, elemental titanium (space group P63/mmc) exhibits Ti–Ti distances of 2.8953(1) Å (calculated from [30]). Comparable titanium chains have also been observed in the structures of Ti2In5 (300 pm Ti–Ti) [31] and Ti3Rh2In3 (307 pm Ti–Ti) [32]. The Sn atom is surrounded by 12 Ti atoms forming an icosahedron (Figure 2). The Ti–Sn distances exhibit an average value of 2.9300(10) Å and are in good agreement with those observed for h-Ti3Sn [33] and the cubic Ti3Sn high-pressure phase [13].

Figure 2: 
Coordination polyhedra around Sn (left) and Ti (right) in the Cr3Si-type structure of c-Ti3Sn. Orange: Sn, blue: Ti, unit cell in black.
Figure 2:

Coordination polyhedra around Sn (left) and Ti (right) in the Cr3Si-type structure of c-Ti3Sn. Orange: Sn, blue: Ti, unit cell in black.

In order to investigate if the incorporation of Fe leads to a stabilization of the Cr3Si-type structure with respect to the Ni3Sn structure, we performed quantum-chemical calculations at DFT level with the dispersion-corrected meta-GGA functional r2SCAN. We also considered other structure types such as the orthorhombic structure with space group Cmcm [11], the Cu3Au structure type [13], and the Li2AgSb structure type as observed for TiCuHg2 [34]. Here the 4c Wyckoff site was occupied with Sn, the 4a, 4b and 4d sites with Ti. The primitive cells of the Cmcm structure of Ni3Sn, and Cr3Si contain 6 Ti atoms and 2 Sn atoms. For the other two polymorphs supercells of corresponding size were constructed. One of the six Ti atoms was replaced by an Fe atom. In cases where the Ti sites are not symmetry-equivalent, the most stable substitution position was considered. In Table 3 the relative stabilities per formula unit are given for the five polymorphs. For comparison, the corresponding Ti3Sn energies are also given. The optimized lattice parameters are shown in Table 4. After Ti/Fe substitution, the symmetry was lowered. Therefore, some lattice parameters were averaged.

Table 3:

Relative energies per formula unit (eV) calculated with r2SCAN-D4.

Structure type/space group Ti3Sn Ti2.5Fe0.5Sn
o-Ti3Sn/Cmcm 0.00 0.00
Ni3Sn/P63/mmc 0.02 0.02
Cr3Si/Pm 3 n 0.08 0.11
Cu3Au/Pm 3 m 0.17 0.04
Li2AgSb/F 4 3m 0.59 0.11
Table 4:

Optimized lattice vectors (Å) calculated with r2SCAN-D4. For the Cmcm structure, the primitive cell vectors are given.

Structure type/space group Ti3Sn Ti2.5Fe0.5Sn
a c a c
o-Ti3Sn/Cmcm 5.75 4.73 5.74 4.70
Ni3Sn/P63/mmc 5.87 4.74 5.85* 4.70
Cr3Si/Pm 3 n 5.22 5.16*
Cu3Au/Pm 3 m 5.85 5.83*
Li2AgSb/F 4 3m 8.02 7.94
  1. *Averaged value.

For Ti3Sn, the Cmcm structure is slightly more stable than the experimentally observed Ni3Sn-type structure. This is in agreement with previous DFT calculations [11]. Also, the Cr3Si structure is only 0.08 eV less stable, followed by Cu3Au (0.17 eV) and Li2AgSb (0.59 eV). The Ti/Fe substituted structures are energetically much closer (within 0.11 eV) than the Ti3Sn structures. In particular the Cu3Au and Li2AgSb structures are strongly stabilized, so that the Cr3Si-type structure of Ti2.5Fe0.5Sn is now the least stable among the considered structure types, although its relative stability with respect to the Ni3Sn-type structure is not significantly different.

The optimized lattice parameter of the Cr3Si-type structure (a = 5.22 Å) is in good agreement with the measured value (a = 5.24 Å, Table 1). It decreases to 5.16 Å when 16.6% Ti atoms are substituted with Fe. It can be expected that the structural effect is much smaller if only 0.5% Ti is replaced by Fe. For the other structures (except Li2AgSb) the reduction of the lattice parameters is much smaller.

From the experimental and theoretical results presented above, the mechanochemically prepared cubic phase has to be considered as metastable. At about 450 °C, the transformation to the well-known hexagonal phase (Ni3Sn type) was observed. All this is in accordance with Ostwald’s step rule, which predicts the crystallization of a metastable polymorph out of a mainly amorphous material. Our calculations point to the fact that the incorporation of iron due to the use of a steel beaker does not stabilize this particular cubic structure type. Interestingly, it should be mentioned that other milling materials such as silicon nitride or zirconia directly led to the formation of the hexagonal Ti3Sn phase using identical crystallizing conditions. Apparently, the local structure of the as-milled materials seems to be an important determinant for the formed polymorph.

2.1 119Sn Mössbauer spectroscopy

Two different Ti3Sn:Fe samples were analyzed through their 119Sn Mössbauer spectra at 78 K. The spectrum of sample A is shown as an example in Figure 3. It was well reproduced with a single signal at an isomer shift of δ = 1.70(1) mm s−1. Due to the cubic site symmetry of the tin atoms (2a, m 3 .) no quadrupole splitting parameter was introduced. The experimentally determined line width parameter of Γ = 1.19(1) mm s−1 is in the usual range observed for intermetallic tin compounds. The 119Sn Mössbauer spectrum thus confirms the single crystallographic tin site. Sample B is essentially identical to sample A; however, a very weak shoulder (<5% of the intensity) was observed in the spectrum at a slightly higher δ value. This signal could not be attributed to any known phase.

Figure 3: 
Experimental (data points) and simulated (colored line) 119Sn Mössbauer spectrum of the Ti3Sn sample measured at 78 K.
Figure 3:

Experimental (data points) and simulated (colored line) 119Sn Mössbauer spectrum of the Ti3Sn sample measured at 78 K.

The isomer shift value of the present sample is slightly lower than the one reported for an arc-melted h-Ti3Sn sample which was additionally annealed at 600 °C and cooled to room temperature at a rate of 4 K h−1 [35]: δ = 1.80(1) mm s−1, ΔEQ = 0.98(1) mm s−1 and Γ = 0.50(1) mm s−1. The lower isomer shift indicated slightly lower s electron density in the present sample [36]. The quadrupole splitting parameter reflects the non-cubic site symmetry ( 6 m2) in the hexagonal modification.

2.2 Magnetic properties

The Fe-substituted structures have an open-shell ground state. This was checked by comparing the total energies of closed and open shell solutions. In all cases the magnetic moment of the Sn atoms is negligible, while it is between 0.1 and 1.0 for Ti and higher than 2 for Fe (see Table 5).

Table 5:

Atomic magnetizations m s (a.u.) and averaged hyperfine coupling parameters a (MHz) calculated with r2SCAN-D4.

Structure type/space group Ti2.5Fe0.5Sn
m s a
o-Ti3Sn/Cmcm 2.7 36.8
Ni3Sn/P63/mmc 2.7 43.0
Cr3Si/Pm 3 n 2.8 36.0
Cu3Au/Pm 3 m 2.5 639.8
Li2AgSb/F 4 3m 2.5 25.2

3 Experimental section

3.1 Synthesis

All samples were prepared by ball milling using a high energy mill (Pulverisette 7, Fritsch, Idar-Oberstein, Germany) followed by an annealing step in N2 atmosphere. Ti (Koch-Light, 99.5%) and Sn (Merck, 99.9%) were mixed in the stoichiometric atomic ratio and milled at a rotational speed of 700 rpm for 6 h (30 min milling and 30 min break per milling step) using a 12 mL stainless steel grinding beaker with six stainless steel balls with a diameter of 10 mm. To prevent oxidation of the metallic powders, the grinding beaker was filled under Argon atmosphere. In order to improve the crystallinity, a subsequent annealing step at 450 °C for 1 h in N2 atmosphere (12.5 L h−1) was set.

3.2 Structural and chemical characterization

All powders were structurally characterized by X-ray powder diffraction using a PANalytical X’Pert MDP Pro diffractometer (CuK α radiation). The diffraction data were obtained in a Bragg-Brentano arrangement ( θ / θ ). Rietveld refinements [23] were performed using the program package FullProf Suite [37], where a pseudo-Voigt function was applied and the background was fitted using a linear interpolation between a set of various background points with refineable heights.

The bulk samples were semiquantitatively analysed by EDX, using a Zeiss EVO® MA10 scanning electron microscope, which was operated in variable pressure mode (60 Pa). Elemental titanium, iron and tin were used as standards.

3.3 119Sn Mössbauer spectroscopy

A Ca119mSnO3 source was used for the Mössbauer spectroscopic experiments on the Ti3Sn samples. The measurements were carried out in a standard liquid nitrogen bath cryostat at 78 K. The source was kept at room temperature. The samples were mixed with alpha-quartz or glucose and enclosed in small PMMA containers with an optimized thickness [38]. Data was fitted using the WinNormos for Igor6 program package [39]. One sample was also tested by 57Fe Mössbauer spectroscopy using a 57Co/Rh source. As a consequence of the low iron content, no signal was detectable.

3.4 Computational details

Structure optimizations and energy calculations of the polymorphs were performed with the plane-wave program package VASP version 6.3.2 [40]. A high kinetic energy cutoff E c  = 900 eV was set to define the plane-wave valence basis. Core electrons were described with semi-valence projector augmented waves [41]. The meta-GGA density functional r2SCAN [42] was applied together with the D4 dispersion correction [43]. This combination is considered as highly accurate for the calculation of structures and energies [44]. The Monkhorst-Pack k-point mesh was set to 12 × 12 × 12. The Ti2.5Fe0.5Sn structures were calculated with unrestricted Kohn-Sham theory (ISPIN = 2). The lowest energies were obtained when initial magnetic moments of Ti atoms were set to small negative values (−0.7), and to +2.0 for Fe atoms.


Dedicated to Professor Gerhard Müller on the occasion of his 70th birthday.



Corresponding authors: Eva M. Heppke and Martin Lerch, Institut für Chemie, Technische Universität Berlin, Straße des 17.Juni 135, 10623 Berlin, Germany, E-mail: ,

Acknowledgment

We thank Dr. S. Seidel for the EDX analyses. The authors (E.H., M.L.) further thank the Federal Ministry of Education and Research, Germany (Bundesministerium für Bildung und Forschung, BMBF, Verbundvorhaben TransHyDE Forschungsverbund AmmoRef, supportcode: 03HY203C), for funding.

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

  2. Research funding: None declared.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

1. Vdovychenko, O. V., Bulanova, M. V., Fartushna, Y. V., Shcheretsky, A. A. Scripta Mater. 2010, 62, 758–761; https://doi.org/10.1016/j.scriptamat.2010.01.036.Search in Google Scholar

2. Jacobs, J. K., Dasgupta, S. US Patent No. 6,007,945, 1999.Search in Google Scholar

3. Todd, A. D. W., Mar, R. E., Dahn, J. R. J. Electrochem. Soc. 2006, 153, A1998–A2005; https://doi.org/10.1149/1.2257985.Search in Google Scholar

4. Derrien, G., Hassoun, J., Panero, S., Scrosati, B. Adv. Mater. 2007, 19, 2336–2340; https://doi.org/10.1002/adma.200700748.Search in Google Scholar

5. Guo, H., Zhao, H., Jia, X., Li, X., Qiu, W. Electrochim. Acta 2007, 52, 4853–4857; https://doi.org/10.1016/j.electacta.2007.01.058.Search in Google Scholar

6. Wachtler, M., Winter, M., Besenhard, J. O. J. Power Sources 2002, 105, 151–160; https://doi.org/10.1016/s0378-7753(01)00934-x.Search in Google Scholar

7. Wolfenstine, J., Campos, S., Foster, D., Read, J., Behl, W. K. J. Power Sources 2002, 109, 230–233; https://doi.org/10.1016/s0378-7753(02)00061-7.Search in Google Scholar

8. Yang, R., Gu, Y., Li, Y., Zheng, J., Li, X. Acta Mater. 2010, 58, 866–874; https://doi.org/10.1016/j.actamat.2009.10.001.Search in Google Scholar

9. Worner, H. W. J. Inst. Met. 1953, 81, 521–528.Search in Google Scholar

10. McCullough, C., Miller, D. S., Levi, C. G. Metall. Mater. Trans. A 1993, 24, 1481–1496; https://doi.org/10.1007/bf02646589.Search in Google Scholar

11. Du, M., Cui, L., Liu, F. Mater. 2019, 12, 2484; https://doi.org/10.3390/ma12152484.Search in Google Scholar PubMed PubMed Central

12. Ivanova, O., Karpets, M., Yavari, A. R., Georgarakis, K., Podrezov, Y. J. Alloys Compd. 2014, 582, 360–363; https://doi.org/10.1016/j.jallcom.2013.07.198.Search in Google Scholar

13. Cannon, J. F., Homan, C., MacCrone, R. K., Whalley, E. Mater. Res. Soc. Symp. Proc. 1984, 22, 113–116.Search in Google Scholar

14. James, S. L., Adams, C. J., Bolm, C., Braga, D., Collier, P., Friščić, T., Grepioni, F., Harris, K. D. M., Hyett, G., Jones, W., Krebs, A., Mack, J., Maini, L., Orpen, A. G., Parkin, I. P., Shearouse, W. C., Steed, J. W., Waddell, D. C. Chem. Soc. Rev. 2012, 41, 413–447; https://doi.org/10.1039/c1cs15171a.Search in Google Scholar PubMed

15. Howard, J. L., Cao, Q., Browne, D. L. Chem. Sci. 2018, 9, 3080–3094; https://doi.org/10.1039/c7sc05371a.Search in Google Scholar PubMed PubMed Central

16. Friščić, T., Mottillo, C., Titi, H. M. Angew. Chem. 2020, 132, 1030–1041; https://doi.org/10.1002/ange.201906755.Search in Google Scholar

17. Suryanarayana, C. Prog. Mater. Sci. 2001, 46, 1–184; https://doi.org/10.1016/s0079-6425(99)00010-9.Search in Google Scholar

18. Burmeister, C. F., Kwade, A. Chem. Soc. Rev. 2013, 42, 7660–7667; https://doi.org/10.1039/c3cs35455e.Search in Google Scholar PubMed

19. Koch, C. C. Mater. Sci. Forum 1992, 88–90, 243–262; https://doi.org/10.4028/www.scientific.net/msf.88-90.243.Search in Google Scholar

20. Tarutani, Y., Kudo, M. J. Less-Common Met. 1977, 55, 221–229; https://doi.org/10.1016/0022-5088(77)90196-5.Search in Google Scholar

21. Tütüncü, H. M., Bağcı, S., Srivastava, G. P. Phys. Rev. B 2010, 82, 214510.10.1103/PhysRevB.82.214510Search in Google Scholar

22. Savitskii, E. M., Baron, V. V., Efimo, Y. V., Bychkova, M. I., Myzenkova, L. F. Superconducting Materials; Plenum Press: New York, 1973.10.1007/978-1-4615-8672-2Search in Google Scholar

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

24. Clark, H. T. J. Met. 1949, 1, 588–589; https://doi.org/10.1007/bf03398899.Search in Google Scholar

25. Pietrokowsky, P., Frink, E. P. Trans. Am. Soc. Met. 1957, 49, 339–342.Search in Google Scholar

26. Kordan, V., Zelinska, O., Pavlyuk, V., Oshchapovsky, I., Serkiz, R. Chem. Met. Alloys 2016, 9, 84–91; https://doi.org/10.30970/cma9.0327.Search in Google Scholar

27. Andersson, S. J. Solid State Chem. 1978, 23, 191–204; https://doi.org/10.1016/0022-4596(78)90065-8.Search in Google Scholar

28. Kjekshus, A., Grønvold, F., Thorbjørnsen, J., Briggs, T., Haslewood, G. A. D., Flood, H. Acta Chem. Scand. 1962, 16, 1493–1510; https://doi.org/10.3891/acta.chem.scand.16-1493.Search in Google Scholar

29. Reddy, M. S., Sadanandam, J., Suryanarayana, S. V. J. Mater. Sci. Lett. 1983, 2, 166–168; https://doi.org/10.1007/bf00735567.Search in Google Scholar

30. Sabeena, M., Murugesan, S., Anees, P., Mohandas, E., Vijayalakshmi, M. J. Alloys Compd. 2017, 705, 769–781; https://doi.org/10.1016/j.jallcom.2016.12.155.Search in Google Scholar

31. Pöttgen, R. Z. Naturforsch. 1995, 50b, 1505–1509.10.1515/znb-1995-1011Search in Google Scholar

32. Zumdick, M. F., Landrum, G. A., Dronskowski, R., Hoffmann, R. D., Pöttgen, R. J. Solid State Chem. 2000, 150, 19–30; https://doi.org/10.1006/jssc.1999.8541.Search in Google Scholar

33. Du, M., Cui, L., Ren, Y., Liu, F. Mater. Lett. 2019, 252, 161–164; https://doi.org/10.1016/j.matlet.2019.05.121.Search in Google Scholar

34. Puselj, M., Ban, Z. Croat. Chem. Acta 1969, 41, 79–83.Search in Google Scholar

35. O’Brien, J. W., Dunlap, R. A., Dahn, J. R. J. Alloys Compd. 2003, 353, 60–64; https://doi.org/10.1016/s0925-8388(02)01304-x.Search in Google Scholar

36. Lippens, P. E. Phys. Rev. B 1999, 60, 4576–4586; https://doi.org/10.1103/physrevb.60.4576.Search in Google Scholar

37. Rodríguez-Carvajal, R. Satellite Meeting on Powder Diffraction of the 15th International Congress of the IUCr: Toulouse, France, 1990; p. 127.Search in Google Scholar

38. Long, G. J., Cranshaw, T. E., Longworth, G. Moss. Eff. Ref. Data J. 1983, 6, 42–49.Search in Google Scholar

39. Brand, R. A. WinNormos for Igor6 (Version for Igor 6.2 or Above: 22/02/2017); Universität Duisburg: Duisburg, Germany, 2017.Search in Google Scholar

40. Kresse, G., Furthmüller, J. Phys. Rev. B 1996, 54, 11169–11186; https://doi.org/10.1103/physrevb.54.11169.Search in Google Scholar PubMed

41. Kresse, G., Joubert, D. Phys. Rev. B 1999, 59, 1758–1775; https://doi.org/10.1103/physrevb.59.1758.Search in Google Scholar

42. Furness, J. W., Kaplan, A. D., Ning, J., Perdew, J. P., Sun, J. J. Phys. Chem. Lett. 2020, 11, 8208–8215; https://doi.org/10.1021/acs.jpclett.0c02405.Search in Google Scholar PubMed

43. Caldeweyher, E., Bannwarth, C., Grimme, S. J. Chem. Phys. 2017, 147, 34112; https://doi.org/10.1063/1.4993215.Search in Google Scholar PubMed

44. Ehlert, S., Huniar, U., Ning, J., Furness, J. W., Sun, J., Kaplan, A. D., Perdew, J. P., Brandenburg, J. G. J. Chem. Phys. 2021, 154, 61101; https://doi.org/10.1063/5.0041008.Search in Google Scholar PubMed

Received: 2023-02-20
Accepted: 2023-02-23
Published Online: 2023-03-13
Published in Print: 2023-03-28

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Articles in the same Issue

  1. Frontmatter
  2. In this issue
  3. Preface
  4. Professor Dr. Gerhard Müller. Editor-in-Chief der Zeitschrift für Naturforschung BChemical Sciences. zum 70. Geburtstag
  5. Research Articles
  6. Ferrocenylmethylation of theophylline
  7. Electron density of a cyclic tetrasaccharide composed of benzoylated galactose units
  8. Orthoamide und Iminiumsalze, CIX. Umsetzungen von Orthoamiden der Alkincarbonsäuren mit Diolen, Ethandithiol und CH-aciden Nitroverbindungen
  9. 1,4-Divinylphenylene-bridged diruthenium complexes with 2-hydroxypyridine- and 2- or 8-hydroxyquinoline-olate ligands
  10. The calcium oxidotellurates Ca2(TeIVTeVIO7), Ca2(TeIVO3)Cl2 and Ca5(TeIVO3)4Cl2 obtained from salt melts
  11. N-heterocyclic carbene-mediated oxidation of copper(I) in an imidazolium ionic liquid
  12. Synthesis, crystal structure, thermal and spectroscopic properties of ZnX2-2-methylpyrazine (X = Cl, Br, I) coordination compounds
  13. Solid-state molecular structures of Se(IV) and Te(IV) dihalides X2Se(CH3)(C6F5) and the gas-phase structure of Se(CH3)(C6F5)
  14. Ein neuartiger T-förmiger 14-Elektronen-Iridium(I)-Komplex stabilisiert durch eine agostische Ir–H-Wechselwirkung
  15. Exploring dicyanamides with two different alkali-metal cations: phase separations, solid solutions and the new compound Rb1.667Cs0.333[N(CN)2]2
  16. Eu4Al13Pt9 – a coloring variant of the Ho4Ir13Ge9 type structure
  17. Decoration of the [Nb6O19]8– cluster shell with six Cu2+-centred complexes generates the [(Cu(cyclen))6Nb6O19]4+ moiety: room temperature synthesis, crystal structure and selected properties
  18. Structure and spectroscopic properties of etherates of the beryllium halides
  19. The palladium-rich silicides RE3Pd20Si6 (RE = Sc, Y and Lu)
  20. Azido and desamino analogs of the marine natural product oroidin
  21. High-pressure high-temperature preparation of CeGe3
  22. On the synthesis and crystal structure of praseodymium(III) metaborate molybdate(VI) – PrBO2MoO4
  23. A third polymorph of the zwitterionic complex trichlorido-((dimethylphosphoryl)methanaminium-κO)zinc(II)
  24. Mechanochemical synthesis and structural evaluation of a metastable polymorph of Ti3Sn
  25. Synthesis and application of calcium silicate hydrate (C-S-H) nanoparticles for early strength enhancement by eco-friendly low carbon binders
  26. Sterically crowded di-indazolyl-pyridines: Iron(II) complexation studies
Downloaded on 24.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/znb-2023-0309/html
Scroll to top button