Skip to content
BY 4.0 license Open Access Published by De Gruyter (O) August 10, 2020

High-pressure synthesis of SmGe3

  • Julia-Maria Hübner ORCID logo , Wilder Carrillo-Cabrera , Raul Cardoso-Gil , Primož Koželj , Ulrich Burkhardt , Martin Etter , Lev Akselrud , Yuri Grin and Ulrich Schwarz ORCID logo EMAIL logo

Abstract

The new samarium germanide SmGe3 is obtained by high-pressure high-temperature synthesis of pre-reacted mixtures of samarium and germanium at a pressure of 9.5 GPa and temperatures between 1073 and 1273 K. SmGe3 decomposes at 470(5) K into SmGe2, α-Sm3Ge5 and a hitherto unknown phase. SmGe3 exhibits a superstructure of the cubic Cu3Au-type. Transmission electron microscopy measurements of crystalline particles and prepared lamellae indicate a high density of defects on the nanoscale. Selected area electron diffraction and elaborate X-ray powder diffraction measurements consistently indicate a 2a0 × 2a0 × 2a0 superstructure adopting space group Fm3¯m with a = 8.6719(2) Å.

1 Introduction

The elemental semiconductors silicon and germanium form a rich variety of binary phases with electropositive partners of the alkaline, alkaline-earth or rare-earth metal groups, and their chemical bonding as well as their electron count is often within the scope of the Zintl-Klemm concept [1], [2]. The tetrel atoms form one-, two- or three-dimensional partial structures with two-center two-electron bonds, frequently yielding electron-precise electron balances. The application of high-pressure techniques has proven to be a productive strategy to grant access to structural patterns that violate classical electron counting rules.

Systematic studies of tetrel connectivities in MT3 compounds (M: alkaline earth or rare-earth metal; T: Si, Ge) [3], [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18], [19], [20], [21], [22], [23] disclose a number of structural motifs that go beyond the scope of the 8-N rule. Moreover, the phases formed with diamagnetic metal ions repeatedly exhibit superconductivity with strong electron-phonon coupling. Within the set of manufactured phases, the absence of a corresponding samarium compound is striking. Earlier density functional theory calculations [24] predict an atomic arrangement with space group P63/mmc for the compound SmGe3. However, by high-pressure high-temperature synthesis, we obtain a Cu3Au-like SmGe3 phase. The finding of superstructure reflections motivated further investigations by energy dispersive X-ray spectroscopy, Transmission electron microscopy (TEM) and extensive X-ray powder diffraction experiments.

2 Experimental section

2.1 Synthesis

SmGe3 was synthesized under high-pressure high-temperature conditions. All sample handling, except for high-pressure synthesis itself, was performed in argon-filled glove boxes (MBraun, H2O and O2 < 0.1 ppm). The precursor mixture was prepared by arc melting of samarium (Lamprecht, 99.9%) and germanium (Chempur, 99.9999+%) in the ratio 1:3 plus 6% excess of samarium for the compensation of evaporation losses during heating. The material was ground and filled into BN crucibles before being transferred into MgO octahedra with an edge length of 18 mm. High-pressure high-temperature syntheses were conducted in a Walker-type module [25] for 30 min to 5 h at pressures between 9(1) to 9.5(10) GPa and temperatures from 1073(107) and 1273(127) K before quenching under load. The calibration of pressure and temperature has been realized prior to the experiments by the observation of resistance changes of bismuth [26] and thermocouple-calibrated runs, respectively.

2.2 Sample characterization

For metallographic analysis, samples were prepared by polishing with diamond powder disks (grain size 6, 3 and 0.25 μm) after embedding in paraffin. The investigation was realized with a Philips XL 30 scanning electron microscope (SEM) (LaB6 cathode), comprising an EDAX Si(Li) detector for energy-dispersive X-ray spectroscopy (EDXS).

Differential scanning calorimetry (DSC) experiments were performed in a Netzsch DSC 404 C device (Netzsch-Gerätebau GmbH, Selb, Germany) operated with heating and cooling rates of 10 K/min under argon atmosphere using corundum crucibles.

Phase designation was conducted by X-ray powder diffraction experiments with a Huber Image Plate Guinier Camera G670 (Huber Diffraktionstechnik GmbH & Co. KG, Rimsting, Germany), using CuKα1 radiation (λ = 1.54056 Å). X-ray diffraction experiments for structure refinement were realized with synchrotron radiation (λ = 0.20709 Å) at DESY Group PETRA III using beamline P02.1, and with a Stoe Stadi MP in Bragg-Brentano geometry. The diffractometer was equipped with a DECTRIS MYTHEN2 1K silicon strip detector and operated with Cu-Kα1 radiation (λ = 1.54056 Å, curved germanium (111) Johann-type Monochromator). The sample with cylindrical shape was mounted on a zero background sample holder with one of the cross sectional area at its end oriented to the incident X-ray beam. For better data point statistics, the final powder pattern results from the sum of three individual intensity data sets, collected in the angular range 5.00° ≤ 2θ ≤ 120° (scan step = 0.06°, time pro step 40 s, ttotal ≈ 24 h).

All crystallographic calculations including determination of diffraction peak positions as well as lattice parameter and structure refinements on basis of full diffraction profiles (Rietveld technique) were performed with the WinCSD program package [27].

Thin samples for the TEM study were prepared by the focused-ion-beam (FIB) lift-out technique. Thin cross-sections of micro-crystalline grains were extracted from a broken bulk fragment. Defined crystallographic oriented cross-sections were prepared perpendicular to the staking-faults visible at the fracture surface. The FIB lift-out technique was performed on a FEI Quanta 200 3D ion/electron dual beam device (FEI Company, Eindhoven, the Netherlands) equipped with an Omniprobe micro-manipulator (tungsten needle), and can be used both as a SEM and a scanning ion microscope (SIM). First, protecting Pt layers (24 × 2 µm thickness, 2 µm high) were deposited on selected parts (parallel and perpendicular to the c axis of microcrystalline arrays) using an acceleration voltage of 30 kV and a current of 0.1 nA. Each cross section (2 μm thickness) was prepared by applying a Ga-ion beam using an acceleration voltage of 30 kV and a current of 1–0.5 nA. The manufactured cut was transferred onto a copper Omniprobe TEM holder using the in-situ lift-out technique [3]. Finally, the cross section was thinned in several discrete steps down to a thickness of about 40 nm by applying an acceleration voltage of 30 kV with currents of 0.5–0.01 nA of the Ga-ion beam.

Magnetic susceptibility was measured using a polycrystalline sample of cylindrical shape (3.0 × 2.5 × 3.4 mm) on an SQUID magnetometer (MPMS XL‐7, Quantum Design) between 1.8 and 300 K in external fields of 0.01 to 3.5 T. Electrical resistivity measurements were conducted in a temperature range from 1.8 to 300 K by using the four-point probes method with a Keithley DC current source and a Hewlett Packard nanovoltmeter inside a helium flow cryostat at zero magnetic field. The contacting of the sample with Cu wire was done inside the glove box with Ag paint.

3 Results

3.1 Composition and properties

The new samarium germanide is synthesized by high-pressure high-temperature treatment of pre-reacted Sm25.4Ge74.6 mixtures. The average chemical composition of the product as determined by energy dispersive X-ray spectroscopy amounts to Sm25.5(5)Ge74.5(5) or SmGe2.92(8).

Differential scanning calorimetry measurements (Figure 1) of SmGe3 reveal a decomposition into SmGe2 [28], Sm2Ge5 [29] and a hitherto unknown phase at 470(5) K. The feature at 737(5) K is attributed to a reaction of the decomposition products into SmGe2 [28] and Ge [30]. In full agreement with the ambient-pressure phase diagram [31], effects at 1030(5) K and 1103(5) K correspond to the peritectoid decomposition of SmGe2 into Sm2Ge3 and Ge as well as to the melting of the resulting eutectic, respectively.

Figure 1: Differential scanning calorimetry (DSC) curve of SmGe3 taken on heating (red curve) and cooling (blue curve) in the temperature range from 300 to 1175 K with a heating rate of 10 K min−1 at ambient pressure under argon atmosphere.
Figure 1:

Differential scanning calorimetry (DSC) curve of SmGe3 taken on heating (red curve) and cooling (blue curve) in the temperature range from 300 to 1175 K with a heating rate of 10 K min−1 at ambient pressure under argon atmosphere.

The magnetic susceptibility χ measured in an external field of μ0H = 3.5 T between 1.8 and 300 K (Figure 2) indicates van Vleck-type paramagnetic behavior, denoting the influence of the external field on the wave function and the transition towards excited states. Under consideration of the slope of the experimental data, the typical minimum of χ(T) for Sm3+ compounds is expected to occur around 400 K. The electrical resistivity ρ at zero-field between 1.8 and 300 K (Figure 2, inset) denotes metallic behavior with a room temperature value ρ(300 K) of 122 µΩcm and an inflection, which is attributed to the reduced scattering of charge carriers in the magnetically ordered phase [32]. The ordered state is antiferromagnetic as indicated by the cusp at 23 K in χ(T).

Figure 2: Magnetic susceptibility χ of SmGe3 for μ0H = 3.5 T at temperatures between 1.8 and 300 K. Inset: Electrical resistivity of SmGe3 at zero-field in the same temperature range.
Figure 2:

Magnetic susceptibility χ of SmGe3 for μ0H = 3.5 T at temperatures between 1.8 and 300 K. Inset: Electrical resistivity of SmGe3 at zero-field in the same temperature range.

3.2 Substructure

X-ray powder diffraction patterns of SmGe3 evidence that the strongest reflections indicate a Cu3Au-type structure motif [33]. A similar atomic arrangement is also reported for analog phases MGe3−x (x ≈ 0.15: M = Tb, Dy, Yb [34], [35]; x = 0: M = Ce [14], Np [36], Pu [37] and U [38], as well as for SmSn3 [39]), the corresponding tin compound of samarium. The crystal structure represents an ordered variety of an fcc lattice, with germanium atoms occupying the centers of cube faces and samarium being positioned on the vertices.

The refinement of a Cu3Au-type model using full diffraction profiles (Figure 3) results in a0 = 4.337(3) Å with R(P) = 0.043 and wR(P) = 0.063. As some other isostructural phases exhibit germanium deficiencies, partial occupation of the Ge site has been tested. The refinement proceeds without improvement of the residuals so that the models for phases SmGe3 and SmGe2.84(2) yield R(P) = 0.0429 and wR(P) = 0.0629 in the substructure.

Figure 3: Synchrotron powder XRD pattern of SmGe3−x (see main text) and the result of refining a Cu3Au-type substructure using full diffraction profiles (Rietveld method).
Figure 3:

Synchrotron powder XRD pattern of SmGe3−x (see main text) and the result of refining a Cu3Au-type substructure using full diffraction profiles (Rietveld method).

3.3 Superstructure

Closer inspection of the substructure refinement reveals significant extra reflections not being accounted for by the Cu3Au-aristotype, e.g., that at 2.36°. The finding points at a larger unit cell for SmGe3 and motivated additional investigations of the crystal structure. For this purpose, annealed samples (55 h at 823 K and 9.5 GPa) are selected and cleaved in order to avoid peak broadening because of grinding. Nevertheless, transmission electron microscopy investigations of a selected lamella prepared by the focused ion-beam method still reveal the presence of extended structural defects in the microstructure (Figure 4a).

Figure 4: a) TEM image of a FIB SmGe3 lamella oriented along [213] direction. The inserted [213]* Selected area (electron) diffraction (SAED) pattern is taken from an area free of {111} stacking-faults, therefore, the diffuse lines along the [111]* direction are absent. In the TEM image, the [11-1] direction is indicated by the red arrow. Dark bands are inclined {111} stacking faults; SAED diffraction images for SmGe3 phase (observed in almost all the FIB lamellar area) along the zone directions, b) [110]*, c) [112]* and d) [102]*. The observed reflection conditions are in agreement with the reflection conditions for a face-centered lattice. The diffuse lines in b) and c) originate from {111} stacking-faults/twin-boundaries.
Figure 4:

a) TEM image of a FIB SmGe3 lamella oriented along [213] direction. The inserted [213]* Selected area (electron) diffraction (SAED) pattern is taken from an area free of {111} stacking-faults, therefore, the diffuse lines along the [111]* direction are absent. In the TEM image, the [11-1] direction is indicated by the red arrow. Dark bands are inclined {111} stacking faults; SAED diffraction images for SmGe3 phase (observed in almost all the FIB lamellar area) along the zone directions, b) [110]*, c) [112]* and d) [102]*. The observed reflection conditions are in agreement with the reflection conditions for a face-centered lattice. The diffuse lines in b) and c) originate from {111} stacking-faults/twin-boundaries.

SAED (Figure 4b–d) and X-ray powder diffraction patterns consistently evidence a 2a0 × 2a0 × 2a0 superstructure (a0 unit cell of the primitive cubic phase). Systematic absences are compatible with a face-centered lattice. As the synchrotron X-ray diffraction experiment evidences significant broadening of the diffraction profiles upon grinding of the material, subsequent X-ray measurements are performed with as-grown ingots at the cost of reduced powder average. Thus, diffraction data of several samples had to be collected in order to check for reproducibility of the intensity information. The best result has been achieved for a sample having been manufactured at 9.5 GPa by heating the starting mixture to 1273 K for 30 min before annealing at 823 K for 5 h to increase the size of the crystal domains. The corresponding diffraction patterns evidence a number of weak reflections, which are absent in the non-annealed samples. In order to test if the extra lines have to be attributed to (potentially incommensurate) modulations of the superstructure, metallographic investigations have been performed. The measurements reveal the presence of a small amount of a side product with composition Sm:Ge ≈ 1:2. As the diffraction peak positions of the previously reported compounds with similar composition do not match the extra lines, a reference sample was synthesized at the same conditions as the 1:3 phase. The X-ray powder diffraction diagram of this component is compatible with a Pu3Pd5-type [40] crystal structure and fits the extra lines in the pattern of SmGe3. With these pieces of information at hand, a model for the superstructure of SmGe3 was developed in space group Fm3¯m and refined (Tables 1 and 2 and Figure 5). Decrease to tetragonal symmetry (sg I4/mmm, at = √2 a0, ct = 2 a0) does not result in improvements of the residuals of the least squares refinements. Further details on an ordered variety with presumably monoclinic symmetry and on the Pu3Pd5-type phase will be reported elsewhere.

Table 1:

Information on data collection, structure refinement and crystallographic information crystal structure of SmGe3. CCDC 2006311 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

CompositionSmGe3
Space group, Pearson symbol, Structure typeFm3¯m, cF32, SmGe3
Lattice parameters (with ABCR LaB6)
a8.6719(2)
V3652.14(4)
Formula units, Z4
Density/g cm−37.49(1)
Measurement range5.015 ≤ 2θ ≤ 120.005

0 ≤ h ≤ 5, 0 ≤ k ≤ 6, 1 ≤ l ≤ 9
Measd points/reflections7667/40
R(P)/wR(P)0.0211/0.0363
Table 2:

Atom type, Wyckoff position, site occupancy factor (S.O.F.), relative atomic coordinates x, y, z, isotropic displacement parameter Biso and site occupancy factors for the crystal structure of SmGe3.

AtomSitex/ay/bz/cBisoS.O.F.
Sm8c¼¼¼0.27(1)1.0
Ge24e0.2366(2)000.63(1)1.0
Figure 5: X-ray powder diffraction diagram of SmGe3 measured in reflection alignment. Blue indices refer to the Cu3Au-type substructure; green ones to the 2 × 2 × 2 superstructure. The tick marks below the experimental pattern and the difference curve indicate line positions for the main (black) and the side phase (ocean blue).
Figure 5:

X-ray powder diffraction diagram of SmGe3 measured in reflection alignment. Blue indices refer to the Cu3Au-type substructure; green ones to the 2 × 2 × 2 superstructure. The tick marks below the experimental pattern and the difference curve indicate line positions for the main (black) and the side phase (ocean blue).

The atomic arrangement of the 2a0 × 2a0 × 2a0 superstructure represents a new variety of the Cu3Au-type. The decrease of symmetry with respect to the lt-Cu3Au type [33] adds a degree of freedom to the positions of the germanium atoms. While germanium is coordinated by eight germanium and four metal atoms at the same distance in the undistorted lt-Cu3Au arrangement (corresponding to 12 distances of a0/√2 = 3.067(2) Å for the substructure), the distances amount to d1(Ge–Ge) = 2.903(2) Å and d2(Ge–Ge) = 3.229(2) Å, while the contacts d(Sm–Ge) = 3.0682(1) Å remain essentially unchanged in the superstructure of SmGe3 (Figure 6).

Figure 6: Crystal structure of SmGe3. (left) Segment visualizing the distortion of the germanium substructure; (right) coordination sphere of the germanium atom.
Figure 6:

Crystal structure of SmGe3. (left) Segment visualizing the distortion of the germanium substructure; (right) coordination sphere of the germanium atom.

The symmetry relation between the sub- and the superstructure is concisely described by a group subgroup relationship [41], [42] evidencing a klassengleiche transition of order 2. The relationship considering also some other fcc varieties is summarized in form of a small family tree (Figure 7) and the atomic arrangements are shown (Figure 8).

Figure 7: Group-subgroup schemes for the structures of ht- and lt-Cu3Au [43], [33], CuPt7 [44], [45], SrPb3 [46] and SmGe3. Indices for klassengleiche (k) and translationengleiche (t) symmetry reductions, unit cell transformations, and atomic positions are given.
Figure 7:

Group-subgroup schemes for the structures of ht- and lt-Cu3Au [43], [33], CuPt7 [44], [45], SrPb3 [46] and SmGe3. Indices for klassengleiche (k) and translationengleiche (t) symmetry reductions, unit cell transformations, and atomic positions are given.

Figure 8: Segments of the crystal structures of (left) lt-Cu3Au [33] and (right) the 2 × 2 × 2 superstructure of SmGe3. The unit cells are indicated by black lines. The atomic arrangement of SrPb3 [46] corresponds to that of lt-Cu3Au with a subtle tetragonal distortion (c/a = 1.014).
Figure 8:

Segments of the crystal structures of (left) lt-Cu3Au [33] and (right) the 2 × 2 × 2 superstructure of SmGe3. The unit cells are indicated by black lines. The atomic arrangement of SrPb3 [46] corresponds to that of lt-Cu3Au with a subtle tetragonal distortion (c/a = 1.014).

Within the series of rare earth metal trigermanides MGe3 (M = La, Ce, Eu, Gd, Tb, Dy, Ho, Er, Tm, Yb, Lu) [12], [13], [14], [15], [16], [17], [18], [19], [20], [21], [22] and their variants MGe3−x (M = Tb, Dy, Yb) [34], [35] and MGe3+x (M = Nd, Pr) [21], different structural motifs are observed. The overview (Figure 9) illustrates that the selected parameters do not allow for a clear separation, but a general tendency comes to light. The extended stability field of the DyGe3 type includes the heavier rare-earth metal trigermanides (M = Tb, Dy, Ho, Er; Tm and Lu, but except Yb). The compounds of the lighter rare-earth metals among these form at ambient pressure, only the Tm and Lu compound require high-pressure synthesis. Light rare-earth metals (M = Ce, Tb, Dy; but also Yb) form dense-packed lt-Cu3Au-type arrangements at elevated pressures.

Figure 9: Average interatomic distances d(Ge–Ge) of selected compounds MGe3 (M = La; rare-earth metal except Pm) [12], [13], [14], [15], [16], [17], [18], [19], [20], [21], [22], [35]. Only bond distances resulting from crystal structure refinements (except lt-Cu3Au) are considered. Filled symbols denote compounds synthesized by high-pressure methods, empty ones phases synthesized at ambient pressure. The grey field indicates distances d(Ge–Ge) of compounds being assigned to the lt-Cu3Au type. References to the crystal structures are given in form of structure types.
Figure 9:

Average interatomic distances d(Ge–Ge) of selected compounds MGe3 (M = La; rare-earth metal except Pm) [12], [13], [14], [15], [16], [17], [18], [19], [20], [21], [22], [35]. Only bond distances resulting from crystal structure refinements (except lt-Cu3Au) are considered. Filled symbols denote compounds synthesized by high-pressure methods, empty ones phases synthesized at ambient pressure. The grey field indicates distances d(Ge–Ge) of compounds being assigned to the lt-Cu3Au type. References to the crystal structures are given in form of structure types.

The interatomic distances d(Ge–Ge) of all structure types except Cu3Au fall into the range from 2.4 to roughly 3 Å (Figure 9). Within the series of isotypic DyGe3-type compounds, the distances d(Ge–Ge) remain more or less constant (distances d(M–Ge) decrease with increasing atomic number in accordance with the lanthanide contraction, not shown). Compounds, which are reported to adopt the lt-Cu3Au-type structure, exhibit unusually long interatomic distances d(Ge–Ge). This abnormality is attenuated by the symmetry breaking in the new SmGe3-type arrangement. The distortion induces a subdivision of the Ge–Ge distances into two groups. The shorter contacts, representing half of the distances d(Ge–Ge), fall now into the upper part of the range observed for DyGe3-type compounds and phases REGe3+x (RE = Pr, Nd). Whether the underlying reason for the distortion is the formation of Ge–Ge bonds or the stereochemical activity of lone pairs located at the Ge species (assuming the electron balance Sm3+[Ge]3) will be the topic of future investigations by quantum chemical methods.


Dedicated to Professor Dr. Ulrich Müller on the occasion of his 80th birthday.



Corresponding author: Ulrich Schwarz, Max-Planck-Institut für Chemische Physik fester Stoffe, Nöthnitzer Straße 40, 01187 Dresden, Germany, E-mail:

Funding source: International Max Planck Research School for Chemistry and Physics of Quantum Materials (IMPRS-CPQM)

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

  2. Research funding: Funded by International Max Planck Research School for Chemistry and Physics of Quantum Materials (IMPRS-CPQM).

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

References

1. Zintl, E., Brauer, G. Z. Phys. Chem. B 1933, 20, 245; https://doi.org/10.1515/zpch-1933-2023.Search in Google Scholar

2. Zintl, E. Angew. Chem. 1939, 52, 1; https://doi.org/10.1002/ange.19390520102.Search in Google Scholar

3. Schwarz, U., Wosylus, A., Rosner, H., Schnelle, W., Ormeci, A., Meier, K., Baranov, A., Nicklas, M., Leipe, S., Müller, C. J., Grin, Yu. J. Am. Chem. Soc. 2012, 134, 13558; https://doi.org/10.1021/ja3055194.Search in Google Scholar PubMed

4. Hübner, J.-M., Akselrud, L., Schnelle, W., Burkhardt, U., Bobnar, M., Prots, Yu., Grin, Yu., Schwarz, U. Materials 2019, 12, 145; https://doi.org/10.3390/ma12010145.Search in Google Scholar PubMed PubMed Central

5. Meier, K., Cardoso-Gil, R., Schwarz, U. Z. Kristallogr. NCS 2011, 226, 297; https://doi.org/10.1524/ncrs.2011.0131.Search in Google Scholar

6. Wosylus, A., Prots, Yu., Schwarz, U. Z. Kristallogr. NCS 2011, 226, 295; https://doi.org/10.1524/ncrs.2011.0130.Search in Google Scholar

7. Schnelle, W., Ormeci, A., Wosylus, A., Meier, K., Grin, Yu., Schwarz, U. Inorg. Chem. 2012, 51, 5509; https://doi.org/10.1021/ic300576a.Search in Google Scholar PubMed

8. Nishikawa, T., Fukuoka, H., Inumaru, K. Inorg. Chem. 2015, 54, 7433; https://doi.org/10.1021/acs.inorgchem.5b00989.Search in Google Scholar PubMed

9. Castillo, R., Schnelle, W., Baranov, A., Burkhardt, U., Bobnar, M., Cardoso-Gil, R., Schwarz, U., Grin, Yu. Z. Naturforsch. B 2016, 71, 585; https://doi.org/10.1515/znb-2016-0047.Search in Google Scholar

10. Castillo, R., Baranov, A., Burkhardt, U., Cardoso-Gil, R., Schnelle, W., Bobnar, M., Schwarz, U. Inorg. Chem. 2016, 55, 4498; https://doi.org/10.1021/acs.inorgchem.6b00299.Search in Google Scholar PubMed

11. Fukuoka, H., Tomomitsu, Y., Inumaru, K. Inorg. Chem. 2011, 50, 6372; https://doi.org/10.1021/ic200826d.Search in Google Scholar

12. Belyavina, N. M., Markiv, V. Y., Speka, M. V. J. Alloys Compd. 1999, 283, 162; https://doi.org/10.1016/s0925-8388(98)00858-5.Search in Google Scholar

13. Fukuoka, H., Suekuni, K., Onimaru, T., Inumaru, K. Inorg. Chem. 2011, 50, 3901; https://doi.org/10.1021/ic102077k.Search in Google Scholar

14. Fukuoka, H., Yamanaka, S. Chem. Lett. 2004, 33, 1334; https://doi.org/10.1246/cl.2004.1334.Search in Google Scholar

15. Castillo, R., Baranov, A., Burkhardt, U., Grin, Yu., Schwarz, U. Z. Anorg. Allg. Chem. 2015, 641, 355; https://doi.org/10.1002/zaac.201400585.Search in Google Scholar

16. Savysyuk, I. A., Gladyshevskii, E. I., Gladyshevskii, R. E. In: Proceedings of the 7th International Conference on Crystal Chemistry Intermetallic Compounds, Lviv, Ukraine, 25–28 September 1999; pp. PB17.Search in Google Scholar

17. Schobinger-Papamantellos, P., Rodriguez Carvajal, J., Buschow, K. H. J. J. Phys. Condens. Matter 2007, 19, 236201; https://doi.org/10.1088/0953-8984/19/23/236201.Search in Google Scholar

18. Schobinger-Papamantellos, P., de Mooij, D. B., Buschow, K. H. J. J. Alloys Compd. 1992, 183, 181; https://doi.org/10.1016/0925-8388(92)90743-s.Search in Google Scholar

19. Schobinger-Papamantellos, P., Rodriguez Carvajal, J., Tung, L. D., Ritter, C., Buschow, K. H. J. J. Phys. Condens. Matter 2008, 20, 195201; https://doi.org/10.1088/0953-8984/20/19/195201.Search in Google Scholar

20. Eremenko, V. N., Obushenko, I. M. Sov. Non-Ferrous Met. Res. 1981, 9, 216.Search in Google Scholar

21. Fukuoka, H., Yoshikawa, M., Baba, K., Yamanaka, S. Bull. Chem. Soc. Jpn. 2010, 83, 323; https://doi.org/10.1246/bcsj.20090310.Search in Google Scholar

22. Harada, M., Fukuoka, H., Matsumura, D., Inumaru, K. J. Phys. Chem. C 2012, 116, 2153; https://doi.org/10.1021/jp206207c.Search in Google Scholar

23. Hübner, J.-M., Bobnar, M., Akselrud, L., Prots, Yu., Grin, Yu., Schwarz, U. Inorg. Chem. 2018, 57, 10295; https://doi.org/10.1021/acs.inorgchem.8b01510.Search in Google Scholar PubMed

24. Persson, K. Materials data on SmGe3 (SG:194) by materials project by LBNL materials project; Berkeley, CA (United States): Lawrence Berkeley National Laboratory (LBNL); 2016, https://doi.org/10.17188/1316073.Search in Google Scholar

25. Walker, D., Carpenter, M. A., Hitch, C. M. Am. Mineral 1990, 75, 1020.Search in Google Scholar

26. Young, D. A. Phase diagrams of the elements; Berkeley, CA, USA: UC Press, 1991.10.1525/9780520911482Search in Google Scholar

27. Akselrud, L., Grin, Yu. J. Appl. Crystallogr. 2014, 47, 803; https://doi.org/10.1107/s1600576714001058.Search in Google Scholar

28. Mayer, I., Eshdat, Y. Inorg. Chem. 1968, 7, 1904; https://doi.org/10.1021/ic50067a042.Search in Google Scholar

29. Tobash, P. H., Lins, D., Bobev, S., Hur, N., Thompson, J. D., Sarrao, J. L. Inorg. Chem. 2006, 45, 7286; https://doi.org/10.1021/ic060913f.Search in Google Scholar PubMed

30. Straumanis, M. E., Aka, E. Z. J. Appl. Phys. 1952, 23, 330; https://doi.org/10.1063/1.1702202.Search in Google Scholar

31. Gokhale, A. B., Abbaschian, G. J. In Binary Alloy Phase Diagrams; Massalski, T. B. Ed. ASM International: Ohio, 1996.Search in Google Scholar

32. Meier, K., Cardoso-Gil, R., Schnelle, W., Rosner, H., Burkhardt, U., Schwarz, U. Z. Anorg. Allg. Chem. 2010, 636, 1466; https://doi.org/10.1002/zaac.201000065.Search in Google Scholar

33. Johansson, C. H., Linde, J. O. Ann. Phys. 1925, 78, 439; https://doi.org/10.1002/andp.19253832104.Search in Google Scholar

34. Fukuoka, H., Yamanaka, S. Chem. Lett. 2004, 33, 1334; https://doi.org/10.1246/cl.2004.1334.Search in Google Scholar

35. Tsvyashchenko, A. V., Spasskiy, A.V., Velichkov, A. I., Salamatin, A. V., Fomicheva, L. N., Salamatin, D. A., Ryasny, G. K., Nikolaev, A. V., Budzynski, M., Sadykov, R. A. J. Alloys Compd. 2013, 552, 190; https://doi.org/10.1016/j.jallcom.2012.10.025.Search in Google Scholar

36. Gal, J., Hadari, Z., Bauminger, E. R., Nowik, I., Ofer, S. Solid State Commun. 1973, 13, 647; https://doi.org/10.1016/0038-1098(73)90450-x.Search in Google Scholar

37. Coffinberry, A. S., Ellinger, F. H. In Proceedings of the United Nations International Conference on the Peaceful Uses of Atomic Energy, 1958, 8, 138.Search in Google Scholar

38. Iandelli, A., Ferro, R. Ann. Chim. 1952, 42, 598.Search in Google Scholar

39. Harris, I. R., Raynor, G. V. J. Less-Common Met. 1965, 9, 7; https://doi.org/10.1016/0022-5088(65)90029-9.Search in Google Scholar

40. Cromer, D. T. Acta Crystalllogr. B 1976, 32, 1930; https://doi.org/10.1107/s0567740876006778.Search in Google Scholar

41. Billiet, Y., Aroyo, M. I., Wondratschek, H. Tables of maximal subgroups of the space groups. In International tables for crystallography. A1. Symmetry relations between space groups; Wondratschek, H., Müller, U. Eds. Wiley: Chichester UK, 2010.10.1107/97809553602060000110Search in Google Scholar

42. Müller, U. Relations between the Wyckoff positions. In International tables for crystallography. A1. Symmetry relations between space groups; Wondratschek, H., Müller, U. Eds. Wiley: Chichester, UK, 2010.10.1107/97809553602060000794Search in Google Scholar

43. Owen, E. A., Liu, Y. H. Philos. Mag. 1947, 38, 354; https://doi.org/10.1080/14786444708521607.Search in Google Scholar

44. Schneider, A., Esch, U. Z. Elektrochem. Angew. Phys. Chem. 1944, 50, 290. https://doi.org/10.1002/bbpc.19440501107.Search in Google Scholar

45. Pöttgen, R., Z. Anorg. Allg. Chem. 2014, 640, 869. https://doi.org/10.1002/zaac.201400023.Search in Google Scholar

46. Zintl, E., Neumayr, S., Z. Elektrochem. Angew. Phys. Chem. 1933, 39, 86. https://doi.org/10.1002/bbpc.19330390209.Search in Google Scholar

Received: 2020-05-20
Accepted: 2020-06-11
Published Online: 2020-08-10
Published in Print: 2020-09-25

© 2020 Julia-Maria Hübner et al., published by De Gruyter, Berlin/Boston.

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

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/zkri-2020-0058/html
Scroll to top button