Home Physical Sciences Ternary rhombohedral Laves phases RE2Rh3Ga (RE = Y, La–Nd, Sm, Gd–Er)
Article Publicly Available

Ternary rhombohedral Laves phases RE2Rh3Ga (RE = Y, La–Nd, Sm, Gd–Er)

  • Stefan Seidel , Oliver Janka , Christopher Benndorf , Bernhard Mausolf , Frank Haarmann , Hellmut Eckert , Lukas Heletta and Rainer Pöttgen EMAIL logo
Published/Copyright: March 18, 2017
Become an author with De Gruyter Brill

Abstract:

The ordered Laves phases RE2Rh3Ga (RE=Y, La–Nd, Sm, Gd–Er) were synthesized by arc-melting of the elements and subsequent annealing. The samples were characterized by powder X-ray diffraction (XRD). They crystallize with the rhombohedral Mg2Ni3Si type structure, space group Rm. Three structures were refined from single crystal X-ray diffractometer data: a=557.1(1), c=1183.1(2), wR2=0.0591, 159 F2 values, 10 variables for Y2Rh3Ga, a=562.5(2), c=1194.4(2) pm, wR2=0.0519, 206 F2 values, 11 variables for Ce2Rh3Ga and a=556.7(2), c=1184.1(3) pm, wR2=0.0396, 176 F2 values, 11 variables for Tb2Rh3Ga. The Rh3Ga tetrahedra are condensed via common corners and the large cavities left by the network are filled by the rare earth atoms. The RE2Rh3Ga Laves phases crystallize with a translationengleiche subgroup of the cubic RERh2 Laves phases with MgCu2 type. Magnetic susceptibility measurements reveal Pauli paramagnetism for Y2Rh3Ga and La2Rh3Ga. Ce2Rh3Ga shows intermediate cerium valence while all other RE2Rh3Ga phases are Curie–Weiss paramagnets which order magnetically at low temperatures. The 89Y and 71Ga solid state nuclear magnetic resonance (NMR) spectra of the diamagnetic representative Y2Rh3Ga show well-defined single resonances in agreement with an ordered bulk phase. In comparison to the binary Laves phase YRh2 a strongly increased 89Y resonance frequency is observed owing to a higher s-electron spin density at the 89Y nuclei as proven by density of states (DOS) calculations.

1 Introduction

Laves phases are an important family of AB2 intermetallics which crystallize with three basic structure types: MgCu2 (space group Fdm), MgZn2 (space group P63/mmc) and MgNi2 (space group P63/mmc) [1], [2]. Further stacking sequences with a different degree of hexagonality are possible [3], [4]. These phases are formed with a broad variety of element combinations. Especially the rare earth (RE)-based Laves phases RET2 (T=3d transition metal) [5] and TT2 combinations have been intensively studied with respect to their magnetic [6] and magnetocaloric properties [7], hydrogen storage [8], [9], [10] and their catalytic behavior [11]. The material properties of many of these phases were modified by substitution experiments (change of the valence electron count) on both the A and B sites, leading to a manifold of solid solutions.

Ternary Laves phases can be formed by ordered replacements of parts of the A and B atoms. The early systematic crystal chemical work by Teslyuk [12], [13] already pointed to phases of composition A2B3B′ and AA′B4 as ordering variants for MgZn2 and MgCu2 with the prototype phases Mg2Cu3Si [14] and MgSnCu4 [15], [16]. Electronic and geometrical factors as well as magnetic ordering or magnetostriction play an important role for phase formation and meanwhile many different ordering variants have been reported [17], [18], [19], [20], [21], [22], [23], [24], [25], [26], [27], [28], [29].

An interesting case is the iron compound TbFe2 [30], the first example of a direct rhombohedral distortion of the MgCu2 type. Again, this distortion arises from magnetostriction. The closely related cerium compound CeFe2 [31] shows the same cubic-to-rhombohedral transition under high-pressure conditions. The TbFe2 type has been observed for a few Laves phases and several hydrogenated samples [32]. The symmetry reduction generates further degrees of freedom through splitting of the transition metal site and a variable c/a ratio.

The first ternary A2B3B′ example with this rhombohedral structure was the silicide Mg2Ni3Si [33]. Later on, the silicides RE2Rh3Si (RE=Pr, Er) [34], U2Ru3Si [35], Ce2Rh3+x Si1−x [36], [37], the germanides RE2Rh3Ge (RE=Y, Pr, Er) [34], Ca2Pd3Ge [38], Sm2Rh3Ge [39], U2Ru3Ge [35], and the phosphide Mg2Ni3P [40] were structurally characterized.

In the present contribution we present the first gallides with TbFe2/Mg2Ni3Si type structures. Substitution of one fourth of the rhodium atoms in the cubic Laves phases RERh2 (≡RE2Rh4) leads to the series of rhombohedral RE2Rh3Ga compounds. The synthesis, structure refinements and the magnetic properties as well as solid state nuclear magnetic resonance (NMR) spectroscopic data of these ordered Laves phases are reported herein.

2 Experimental

2.1 Synthesis

Starting materials for the synthesis of the RE2Rh3Ga samples were pieces of the sublimed rare earth elements (Y, La, Pr, Nd, Sm, Gd–Er from Smart Elements; Ce from Sigma Aldrich, Sigma Aldrich Germany, Hamburg), rhodium powder (Agosi, Pforzheim, Germany) and gallium lumps (Emmerich am Rhein, Germany), all with stated purities better than 99.9%. The moisture-sensitive early rare earth elements La–Nd and Sm were kept in Schlenk tubes under purified argon prior to the reactions. The argon was purified with a titanium sponge (873 K), silica gel and molecular sieves.

The elements were weighed in the ideal 2:3:1 atomic ratio (the rhodium powder was previously cold-pressed into pellets with a diameter of 5 mm) and arc-melted in a water-cooled copper mold of a MAM-1 (Edmund Bühler GmbH, Hechingen, Germany) arc-furnace under an argon atmosphere of 800 mbar. Remelting the obtained buttons several times from each site ensured complete homogeneity. With the exception of the cerium-based sample the synthesized compounds showed some impurity phases, especially in the form of non-ordered Laves phases with the presumed composition RERh2−x Gax [MgCu2 type, confirmed by X-ray diffraction (XRD)].

To increase the phase purity all samples were sealed in evacuated quartz ampoules and treated with different annealing procedures. The Y2Rh3Ga sample was placed in the water cooled sample chamber of an induction furnace (Hüttinger Elektronik, Freiburg, Typ TIG 5/300) [41] and rapidly heated to a temperature of approximately 1100 K. This temperature was kept for 8 h until the sample was quenched by switching off the power supply of the high-frequency furnace. Also the Tb2Rh3Ga sample was inductively heated. A piece of the crushed arc-melted regulus was sealed in an evacuated quartz tube and heated shortly below its melting point for 10 min followed by reducing the power of the furnace within 20 min to a temperature of approximately 900 K which was kept for another 8 h followed by quenching.

The remaining samples were annealed in muffle furnaces. They were heated up to 1073 K within 2 h and then kept at this temperature for 10 (RE=La–Nd, Gd), respectively 28 days (Tb, Dy). The best results for the compounds containing holmium and erbium were obtained by annealing these samples at 773 K for 6 weeks. Except for Sm2Rh3Ga, these different annealing procedures led to X-ray pure samples suitable for physical property measurements. The crushed samples showed metallic luster and were stable in air over weeks.

2.2 X-ray image plate data and data collections

The RE2Rh3Ga bulk samples were characterized by X-ray powder diffraction using a Guinier camera equipped with an image plate system (Fujifilm, BAS-1800) using Cu Kα1 radiation and α-quartz (a=491.30, c=540.46 pm) as an internal standard. The trigonal lattice parameters (Table 1) have been derived from least-squares refinements of the powder data. Comparing the experimental patterns to calculated ones [42] ensured correct indexing. As an example, we present the experimental and calculated powder pattern of Dy2Rh3Ga in Fig. 1.

Table 1:

Lattice parameters of the rhombohedral gallides RE2Rh3Ga, space group Rm.

Compounda (pm)c (pm)V (nm3)
Y2Rh3Ga557.1(1)1183.1(2)0.3180
La2Rh3Ga569.4(2)1214.6(3)0.3410
Ce2Rh3Ga562.5(2)1194.4(2)0.3273
Pr2Rh3Ga565.3(1)1194.8(2)0.3306
Nd2Rh3Ga563.8(1)1192.7(3)0.3283
Sm2Rh3Ga561.3(1)1188.9(2)0.3244
Gd2Rh3Ga558.3(1)1190.1(1)0.3213
Tb2Rh3Ga556.7(1)1184.1(3)0.3178
Dy2Rh3Ga556.6(1)1179.8(2)0.3165
Ho2Rh3Ga554.6(1)1183.2(2)0.3152
Er2Rh3Ga552.4(1)1180.6(3)0.3120
Fig. 1: Experimental (top) and calculated (bottom) Guinier powder pattern (CuKα1 radiation) of Dy2Rh3Ga.
Fig. 1:

Experimental (top) and calculated (bottom) Guinier powder pattern (CuKα1 radiation) of Dy2Rh3Ga.

Mechanical fragmentation of the Y2Rh3Ga, Ce2Rh3Ga and Tb2Rh3Ga (annealed in the induction furnace) samples resulted in irregularly shaped single crystal fragments. These were glued to thin quartz fibers using beeswax. The crystal quality was subsequently tested by Laue photographs on a Burger camera (white molybdenum radiation, image plate technique, Fujifilm, BAS-1800). The intensitiy data sets were collected at room temperature using a Stoe StadiVari single crystal diffractometer equipped with a Mo micro focus source and a Pilatus detection system. The Gaussian-shaped profile of the X-ray source required scaling along with numerical absorption corrections. All relevant crystallographic data and details of the data collections and evaluations are listed in Table 2.

Table 2:

Crystallographic data and structure refinement for Ce2Rh3Ga, Tb2Rh3Ga and Y2Rh3Ga, space group Rm,Z=3, Mg2Ni3Si type.

FormulaCe2Rh3GaTb2Rh3GaY2Rh3Ga
Molar mass, g mol−1658.7696.3556.2
Lattice parameters, pma=562.5(2)a=556.7(1)a=557.1(1)
Guinier powder datac=1194.4(2)c=1184.1(3)c=1183.1(2)
Cell volume, nm3V=0.3273V=0.3178V=0.3180
Density calc., g cm−310.0310.918.71
Crystal size, μm20×30×6010×20×5010×20×50
Detector distance, mm404040
Exposure time, s123560
Integr. param. A, B, EMS3.5; −0.1; 0.0127.0; −6.0; 0.0307.0; −6.0; 0.030
Range in hkl±9, ±9, ±19±8, ±8, ±18±8, ±8, ±17
θmin/θmax, deg4.5/35.44.6/33.24.6/31.9
Linear absorption coeff., mm−137.450.444.5
No. of reflections357434833156
Rint0.07610.03590.0600
No. of independent reflections206176159
Reflections used [I>3 σ(I)]187156127
F(000), e846888732
R1/wR2 for I>3 σ(I)0.0212/0.05110.0173/0.03830.0281/0.0575
R1/wR2 for all data0.0232/0.05190.0208/0.03960.0418/0.0591
Data/parameters206/11176/11159/10
Goodness-of-fit on F21.881.431.46
Extinction coefficient440(30)10(8)
Diff. Fourier residues (min/max), e Å−3−1.61/1.31−1.40/0.84−1.20/2.02

2.3 EDX data

The Y2Rh3Ga, Ce2Rh3Ga and Tb2Rh3Ga crystals measured on the diffractometer were analyzed semi-quantitatively using a Zeiss EVO MA10 scanning electron microscope with CeO2, TbF3, Rh and GaP as standards. No impurity elements heavier than sodium (detection limit of the instrument) were observed. The experimentally determined element ratios (34±2 at.% Y:50±2 at.% Rh:16±2 at.% Ga, 33±2 at.% Ce:52±2 at.% Rh:15±2 at.% Ga and 32±2 at.% Tb:53±2 at.% Rh:15±2 at.% Ga) were in close agreement with the ideal compositions (33.3:50:16.6). The variations of ±2 at.% result from the irregular shape of the crystal surfaces (conchoidal fracture).

2.4 Physical property measurements

Fragments of the buttons of the X-ray pure phases were attached to the sample holder rod of a vibrating sample magnetometer unit (VSM) using Kapton foil for measuring the magnetization M(T, H) in a quantum design physical property measurement system (PPMS). The samples were investigated in the temperature range of 2.5–300 K with magnetic flux densities up to 80 kOe (1 kOe=7.96×104 A m−1). For the Nd2Rh3Ga and Dy2Rh3Ga samples additionally heat capacity (HC) measurements were conducted. A piece of the sample was fixed to a pre-calibrated heat capacity puck using Apiezon N grease and measured in the temperature range of 2 to 20 K.

2.5 Solid state NMR spectroscopy

The 89Y and 71Ga solid state NMR spectra of the diamagnetic representatives Y2Rh3Ga and YRh2 were measured on Bruker DSX 500 (B0=11.7 T), Bruker DSX 400 (B0=9.4 T) and Bruker Avance III (B0=7.05 T) NMR spectrometers at resonance frequencies of 24.496 MHz, 19.597 MHz (both 89Y) and 91.486 MHz (71Ga). The spectra were referenced to aqueous solutions of Y(NO3)3 (8 mol L−1 with 0.25 mol L−1 Fe(NO3)3; δiso=−22.8 ppm) and Ga(NO3)3 (1 mol L−1; δiso=0 ppm) at ambient temperature. The finely powdered sample was mixed with dry potassium chloride in an approximate volume ratio of 1:2 to reduce the electrical conductivity and density and was filled in conventional Si3N4 or ZrO2 MAS rotors with 7 mm (Si3N4) and 4 mm (ZrO2) diameter, respectively. The 89Y spectra were recorded by using conventional rotor-synchronized π/2–τ–π–τ-spin-echo sequences with typical π/2-pulse lengths of 20–20.5 μs, relaxation delays of 0.5 s and MAS spinning frequencies of 4–7 kHz. Static 71Ga investigations were carried out by using the wideband uniform-rate smooth truncation (WURST) QCPMG sequence [43] with the WURST-80 pulse shape and 8-step cycling version. All spectra were recorded using the Bruker Topspin software [44] and analyzed with the Dmfit software [45]. The experimental data are summarized in Table 7.

2.6 Theoretical investigations

Quantum mechanical calculation of the electric field gradient (EFG) and the density of states (DOS) for YRh2 and Y2Rh3Ga were performed on the basis of density functional theory. The experimental structures were optimized with the VASP package which uses the plane augmented waves (PAW) [46] method based on the local density approximation (LDA) [47]. The cubic YRh2 was sampled with a 14×14×14 k-grid and the hexagonal Y2Rh3Ga with a 16×16×8 k-grid. Convergence was with respect to the k-meshes and was checked carefully, and calculations were performed with a cut-off energy of 500 eV.

The full potential linear augmented plane wave (FP-LAPW) method as implemented in the WIEN2k code [48] was used to calculate the electric field gradients main principal axis VZZ(Ga) and the asymmetry parameter ηQ(Ga). The QTL and TETRA programs included in the code were applied to calculate the s-DOS of Y in both compounds. To avoid core charge leakage and to ensure that the calculations run with appropriate efficiency the atomic-sphere radii (RMT) were chosen as large as possible (Y: 2.5–2.45 a.u., Rh: 2.5–2.41 a.u. and Ga: 2.29 a.u.). The default value for the separation energies of −6.0 Ry was chosen to separate core and valence states. The implemented GGA PBE exchange-correlation functional was chosen [49]. For the plane-wave cut-off, defined by RMT×Kmax, we applied the default value of 7. The k-meshes for Wien2k calculations were 10×10×10 for YRh2 and the rhombohedral representation of Y2Rh3Ga.

3 Results and discussion

3.1 Structure refinements

Careful analyses of the three intensity data sets revealed R-centered lattices. The absence of further systematic extinctions led to the possible space groups Rm, R3m, R32, R3̅, and R3, of which the centrosymmetric group Rm was found to be correct during structure refinement. Isotypism of the RE2Rh3Ga compounds with the Mg2Ni3Si type structure [33] was already evident from the Guinier powder patterns and from a systematic check of the Pearson database [32]. Subsequently the atomic parameters of Mg2Ni3Si were used as starting values and the structures were refined on F2 with anisotropic displacement parameters for all atoms using the Jana2006 [50], [51] routine. As a check for the correct composition and site assignment, the occupancy parameters were refined in a separate series of least-squares cycles. All sites were fully occupied within three standard deviations. No significant residual peaks were evident in the final difference Fourier syntheses. At the end, the positional parameters were transformed to the setting required for the group–subgroup scheme discussed below. The final positional parameters and interatomic distances are listed in Tables 35.

Table 3:

Atom positions and equivalent isotropic displacement parameters (pm2) for Ce2Rh3Ga, Tb2Rh3Ga and Y2Rh3Ga.

AtomWyckoff positionxyzUeq
Ce2Rh3Ga
 Ce6c000.37198(5)171(2)
 Rh9d1/201/2148(2)
 Ga3a000134(3)
Tb2Rh3Ga
 Tb6c000.37090(4)144(2)
 Rh9d1/201/2127(2)
 Ga3a000109(3)
Y2Rh3Ga
 Y6c000.36984(16)179(4)
 Rh9d1/201/2155(3)
 Ga3a000144(6)

Ueq is defined as one third of the trace of the orthogonalized Uij tensor.

Table 4:

Anisotropic displacement parameters (pm2) for Ce2Rh3Ga, Tb2Rh3Ga and Y2Rh3Ga.

AtomU11U22U33U12U13U23
Ce2Rh3Ga
 Ce162(2)U11189(3)81(1)0U13
 Rh151(2)124(3)159(3)62(1)7(1)15(2)
 Ga142(4)U11120(5)71(2)0U13
Tb2Rh3Ga
 Tb138(2)U11157(2)69(1)0U13
 Rh126(2)99(3)146(3)50(1)7(1)13(2)
 Ga116(4)U1195(5)58(2)0U13
Y2Rh3Ga
 Y172(5)U11193(8)86(2)0U13
 Rh161(4)128(5)166(4)64(2)5(2)10(4)
 Ga150(7)U11132(11)75(4)0U13

Coefficients Uij of the anisotropic displacement factor tensor of the atoms are defined by: −2π2[(ha*)2U11+…+2hka*b*U12].

Table 5:

Interatomic distances (pm) for Ce2Rh3Ga, Tb2Rh3Ga and Y2Rh3Ga. All distances of the first coordination spheres are listed.

Ce2Rh3GaTb2Rh3GaY2Rh3Ga
Ce:3Rh294.1Tb:3Rh290.4Y:3Rh289.2
1Ce305.81Tb305.71Y308.0
6Rh320.16Rh317.66Rh318.3
3Ga328.03Ga324.53Ga324.5
3Ce337.63Tb333.53Tb333.0
Rh:2Ga256.9Rh:2Ga254.5Rh:2Ga254.4
4Rh281.34Rh278.34Rh278.6
2Ce294.12Tb290.42Y289.2
4Ce320.14Tb317.64Y318.3
Ga:6Rh256.9Ga:6Rh254.5Ga:6Rh254.4
6Ce328.06Tb324.56Y324.5

All standard uncertainties were less than 0.2 pm.

Further details of the structure refinements may be obtained from FIZ Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany (fax: +49-7247-808-666; e-mail: crysdata@fiz-karlsruhe.de) on quoting the deposition numbers CSD-432339 (Y2Rh3Ga), CSD-432341 (Ce2Rh3Ga) and CSD-432340 (Tb2Rh3Ga).

3.2 Crystal chemistry

The gallides RE2Rh3Ga (RE=Y, La–Nd, Sm, Gd–Er) are new representatives of the Mg2Ni3Si type structure [33], space group Rm, Pearson symbol hR18 and Wyckoff sequence dca. The cell volumes (Table 1 and Fig. 2) decrease from the lanthanum to the erbium compound as expected from the lanthanide contraction. Ce2Rh3Ga shows a drastic deviation from this smooth behavior. The significantly smaller cell volume (even smaller than Nd2Rh3Ga) is a clear hint for an intermediate cerium valence. This is discussed in more detail below along with the magnetic properties. A very similar trend of the cell volume is observed for the cubic Laves phases RERh2 (Fig. 2) [32]. The cell volumes per formula unit (Z) of the gallides are slightly larger than those of the RERh2 phases. This trend is also observed in the sequence GdRh2 (MgCu2 type, V/Z=0.0530 nm3 [52])→GdRhGa (TiNiSi type, V/Z=0.0573 nm3 [53])→GdGa2 (AlB2 type, V/Z=0.0639 nm3 [54]), indicating generally stronger RE–Rh vs. weaker RE–Ga bonding. Small anomalies in the c lattice parameter of the RE2Rh3Ga samples might be indicative of small homogeneity ranges.

Fig. 2: Plot of the cell volumes of the RE2Rh3Ga and RERh2 [32] Laves phases as a function of the rare earth element.
Fig. 2:

Plot of the cell volumes of the RE2Rh3Ga and RERh2 [32] Laves phases as a function of the rare earth element.

The crystal structure of Ce2Rh3Ga is exemplarily presented in Fig. 3 and discussed in the following. The network of condensed tetrahedra shows full rhodium-gallium ordering. All Rh3Ga tetrahedra exclusively share corners, but in a well-defined motif. The Kagomé networks which extend in the ab plane are formed only by the rhodium atoms, and the gallium atoms connect adjacent networks. This ordering pattern leads to substantial differences in the geometry of the tetrahedra. Each rhodium atom has two gallium neighbors at 257 pm and four rhodium neighbors at the much longer distance of 281 pm.

Fig. 3: The crystal structure of Ce2Rh3Ga. Cerium, rhodium, and gallium atoms are drawn as medium gray, black filled and open circles, respectively. The network of condensed Rh3Ga tetrahedra and the diamond-related cerium substructure (magenta lines for the shorter and gray lines for the longer Ce–Ce distances) are emphasized. Interatomic distances are given in pm.
Fig. 3:

The crystal structure of Ce2Rh3Ga. Cerium, rhodium, and gallium atoms are drawn as medium gray, black filled and open circles, respectively. The network of condensed Rh3Ga tetrahedra and the diamond-related cerium substructure (magenta lines for the shorter and gray lines for the longer Ce–Ce distances) are emphasized. Interatomic distances are given in pm.

The Rh–Ga distance of 257 pm is indicative of substantial Rh–Ga bonding. It is only slightly longer than the sum of the covalent radii for Rh+Ga of 250 pm [55]. The [RhGa] (265–268 pm Rh–Ga) and [Rh3Ga9] (259–263 pm Rh–Ga) networks of CeRhGa [56] and Eu2Rh3Ga9 [57] show even longer Rh–Ga distances.

The cerium atoms fill the large cavities formed by the network of tetrahedra. The topology reminds of the diamond structure. The substantial Rh–Ga interlayer bonding (between the Kagomé networks) leads to a strong decrease of the c/a ratio (√2√3=2.45 for CeRh2 [52] in a related setting vs. 2.12 for Ce2Rh3Ga) and this directly influences the Ce–Ce distances: 4×327 pm in CeRh2vs. 1×306 and 3×338 pm in Ce2Rh3Ga. All of these Ce–Ce distances are well below the Hill limit [58] for f electron localization (340 pm). This is in full agreement with the intermediate cerium valence (vide infra) and the absence of any magnetic ordering.

The structures of CeRh2(≡Ce2Rh4) and Ce2Rh3Ga are strictly related by a group–subgroup scheme, which is presented in the Bärnighausen formalism [59], [60], [61] in Fig. 4. The translationengleiche symmetry reduction of index 4 (t4) from Fdm to Rm leads to a splitting of the 16c rhodium site, enabling the 3:1 Rh/Ga ordering. Furthermore, we obtain a free c/a ratio and a free z parameter for the cerium atoms. These parameters allow the optimization of chemical bonding forced by the rhodium/gallium coloring. A 7:1 ordering on the tetrahedral network has recently been reported for orthorhombic Cd4Cu7As [28].

Fig. 4: Group–subgroup scheme in the Bärnighausen formalism [59], [60], [61], [62] for the structures of CeRh2 [52] and Ce2Rh3Ga. The index for the translationengleiche (t) symmetry reduction, the unit cell transformation, and the evolution of the atomic parameters are given.
Fig. 4:

Group–subgroup scheme in the Bärnighausen formalism [59], [60], [61], [62] for the structures of CeRh2 [52] and Ce2Rh3Ga. The index for the translationengleiche (t) symmetry reduction, the unit cell transformation, and the evolution of the atomic parameters are given.

Finally we focus on the valence electron count (VEC) of RERh2 and RE2Rh3Ga (≡RERh1.5Ga0.5). The formation of a cubic or a hexagonal Laves phase depends on VEC. Johnston and Hoffmann [62] calculated the variation of the relative energy of the hexagonal and cubic tetrahedral network of Laves phases as a function of VEC. Cubic RERh2 (VEC=21) and hexagonal RERh1.5Ga0.5 (VEC=18) perfectly match the predictions. Even the two cerium compounds with slightly higher VEC (a consequence of intermediate cerium valence; vide infra) fall in the same range.

3.3 Magnetic properties

Magnetic susceptibility data has been obtained only for the X-ray pure RE2Rh3Ga samples with RE=Y, La–Nd, and Gd–Er. The basic magnetic data that has been derived from these measurements are listed in Table 6. The results are discussed in the following paragraphs.

Table 6:

Magnetic properties of the gallides RE2Rh3Ga.

TN (K)TC (K)μeffB)μcalcB)θp (K)Hcrit (kOe)μsatB)gJ×JB)
Pr2Rh3Ga6.3(1)3.54(1)3.583.8(1)1.6(1)3.20
Nd2Rh3Ga4.8(1)3.58(1)3.626.4(1)1.7(1)3.27
Gd2Rh3Ga43.0(1)8.03(1)7.9451.7(1)6.69(5)7.0(1)7
Tb2Rh3Ga33.4(1)9.47(1)9.7237.5(1)4.49(5)7.0(1)9
Dy2Rh3Ga10.9(1)11.01(1)10.6511.7(1)8.7(1)10
Ho2Rh3Ga10.9(1)10.56(1)10.6120.9(1)8.3(1)10
Er2Rh3Ga13.1(1)9.70(1)9.585.0(1)5.1(1)9

TN, Néel temperature; TC, Curie temperature; μeff, effective magnetic moment; μcalc, calculated magnetic moment; θp, paramagnetic Curie temperature; Hcrit, critical field for the spin reorientation, saturation moment; μsat and saturation according to gJ×J. The experimental saturation magnetizations were obtained at 3 K and 80 kOe.

3.3.1 Y2Rh3Ga and La2Rh3Ga

The temperature dependences of the magnetic susceptibility of the yttrium and lanthanum compound are presented in Fig. 5. Both compounds are Pauli paramagnets with room temperature susceptibilities of χ=1.85(1)×10−4 emu mol−1 and χ=1.30(1)×10−3 emu mol−1, respectively. The weak upturns at lower temperature arise from tiny amounts of paramagnetic impurities. The present data clearly proves the absence of local moments on the rhodium atoms. Thus, the magnetic properties of the remaining phases arise from the rare earth elements only.

Fig. 5: Temperature dependence of the magnetic susceptibility (χ data) of Y2Rh3Ga (red) and La2Rh3Ga (black) measured at 10 kOe.
Fig. 5:

Temperature dependence of the magnetic susceptibility (χ data) of Y2Rh3Ga (red) and La2Rh3Ga (black) measured at 10 kOe.

3.3.2 Ce2Rh3Ga

The susceptibility and inverse susceptibility data of Ce2Rh3Ga are shown in Fig. 6 (top), indicating values in the order of 10−3 emu mol−1 without any sign of magnetic ordering down to low temperatures. The χ−1 data is curved, consequently no Curie–Weiss behavior is observed. The shape of the χ−1 curves with the weak temperature dependence is typical for Ce-based intermetallics which exhibit valence fluctuations. The magnetic susceptibility of valence fluctuating compounds can be described with the so called interconfiguration fluctuation (ICF) model proposed by Hirst [63], which was later applied by Sales and Wohlleben in order to explain valence fluctuation behavior in intermetallic Yb compounds [64]. The ICF model can be used for systems were two distinct states of the rare earth atom, here Ce3+ and Ce4+, exist. The temperature dependence of the magnetic susceptibility is described as:

Fig. 6: Magnetic properties of Ce2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe; the fit according to Sales and Wohlleben is depicted in red. Bottom: Magnetization isotherms at 3, 10, and 50 K along with the calculated temperature dependence of the cerium valence, depicted in red.
Fig. 6:

Magnetic properties of Ce2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe; the fit according to Sales and Wohlleben is depicted in red. Bottom: Magnetization isotherms at 3, 10, and 50 K along with the calculated temperature dependence of the cerium valence, depicted in red.

(1)χ(T)=(NA3kB)[μn2ν(T)+μn12{1ν(T)}T+Tsf]+nCT+χ0

with NA being the Avogardo’s number, kB the Boltzmann constant, μ the magnetic moment of the respective Ce ion and the spin fluctuation temperature Tsf. χ0 is a temperature independent term, C the Curie constant (for free Ce3+, C=0.807 emu mol−1 K−1), and n the fraction of stable Ce3+. The overall susceptibility is described by three terms: (i) a valence fluctuation part, (ii) the contribution of the stable Ce3+ ions and (iii) a temperature independent part. The valence fluctuations are described by the pseudo-Boltzmann statistic shown in equation (2) which contains the spin fluctuation temperature Tsf and Eex describing the energy difference between the two cerium ground states according to Eex=E(Ce3+)–E(Ce4+).

(2)ν(T)=(2Jn+1){(2Jn+1)+(2Jn1+1)exp[EexkB(T+Tsf)]}

For the Ce cations the fluctuations take place between the 4f0 and 4f1 configurations. With J1=μeff,1=0, J2=5/2 and μeff,2=2.54 μB the magnetic data can be fitted and Tsf=1599(15) K, Eex=3072(27) K, n=0.0268 and χ0=1.6×10−3 emu mol−1 were extracted. In order to prevent over-determination, the temperature independent part χ0 and n were fitted first and fixed for the rest of the procedure, hence no standard deviations are given for these two parameters. The extracted values for Tsf and Eex are larger compared to, e.g. valence fluctuating CeMo2Si2C [65], CeRhSi2 [66] or Ce2Rh3Si5 [66], but comparable to the ones found for the solid solutions CeRu1−x Nix Al [67] or CeRu1−x Pdx Al [68]. In addition to parameters extracted from the fitted data, the average cerium valence can be obtained using eq. 2 to calculate ν(T). The change of the cerium valence vs. temperature is plotted in Fig. 6 (bottom, red), showing a temperature-dependent shift of the cerium valence. The magnetization isotherms exhibit highly reduced magnetic moments with saturation magnetizations of μsat=0.05(1) μB per formula unit compared to the expected magnetization of μsat,theo=2.14 according to gJ×J.

Finally it is interesting to note that binary CeRh2 also shows intermediate cerium valence. This was substantiated for polycrystalline material as well as for Czochralski grown single crystals via resistivity measurements, evaluation of the de Haas–van Alphen effect and LIII absorption studies [69], [70].

3.3.3 Pr2Rh3Ga–Er2Rh3Ga

For the remaining RE2Rh3Ga series with RE=Pr, Nd, and Gd–Er conventional magnetism with cooperative ordering phenomena is observed. The magnetic measurements are depicted in Figs. 713. The top panels show the magnetic susceptibility (χ and χ−1 data), measured at 10 kOe. The paramagnetic region is fitted using the Curie-Weiss law to extract the effective magnetic moment μeff and the Weiss constant θP (listed in Table 6). For all compounds the effective magnetic moments are in good agreement with the calculated theoretical moments μcalc. The middle panels depict the zero-field cooling (ZFC) and field-cooling (FC) measurements at an external field of 100 Oe. The bottom panels display the magnetization isotherms recorded at different temperatures. The isotherms way above the ordering temperatures show linear temperature dependencies, as expected for paramagnetic materials. The saturation magnetization (μsat) was determined from the 3 K isotherm at 80 kOe. In the following paragraphs the measurements of the individual compounds are briefly discussed and interesting features are outlined.

Fig. 7: Magnetic properties of Pr2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe; the inset shows the derivative dχ/dT vs. T used to determine the Curie-temperature. Bottom: Magnetization isotherms at 3, 10, 25, 75, and 100 K.
Fig. 7:

Magnetic properties of Pr2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe; the inset shows the derivative dχ/dT vs. T used to determine the Curie-temperature. Bottom: Magnetization isotherms at 3, 10, 25, 75, and 100 K.

Fig. 8: Magnetic properties of Nd2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe; the heat capacity measurement is depicted in red. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.
Fig. 8:

Magnetic properties of Nd2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe; the heat capacity measurement is depicted in red. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.

Fig. 9: Magnetic properties of Gd2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 30, 50, and 80 K.
Fig. 9:

Magnetic properties of Gd2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 30, 50, and 80 K.

Fig. 10: Magnetic properties of Tb2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.
Fig. 10:

Magnetic properties of Tb2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.

Fig. 11: Magnetic properties of Dy2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.
Fig. 11:

Magnetic properties of Dy2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.

Fig. 12: Magnetic properties of Ho2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.
Fig. 12:

Magnetic properties of Ho2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.

Fig. 13: Magnetic properties of Er2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.
Fig. 13:

Magnetic properties of Er2Rh3Ga. Top: temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. Middle: temperature dependence of the magnetic susceptibility in zero-field cooling and field cooling mode measured at 100 Oe. Bottom: magnetization isotherms at 3, 10, 50, and 100 K.

The ZFC/FC measurements of Pr2Rh3Ga (Fig. 7) show an anomaly characteristic for ferromagnetic ordering at TC=6.3(1) K. The transition temperature was determined from the derivative dχ/dT vs. T curve (Fig. 7, middle, inset). The magnetization isotherm recorded at 3 K confirms the ferromagnetic ground state of Pr2Rh3Ga.

Two anomalies occur in the ZFC/FC measurements of Nd2Rh3Ga (Fig. 8). This is unusual since the crystal structure contains only one rare earth site, therefore additional heat capacity measurements were conducted (Fig. 8, middle, red). One observes a sharp λ-shaped anomaly at T=4.8(1) K along with a small shoulder. The shoulder can be attributed to the impurity of the sample (although pure on the level of X-ray powder diffraction) while the main anomaly corresponds to the intrinsic magnetic transition. The ferromagnetic character of the transition is again evident from the magnetization isotherm recorded at 3 K.

The susceptibility measurements of Gd2Rh3Ga (Fig. 9) at 10 kOe reveal a positive Weiss constant θP=51.7(1) K, suggesting ferromagnetic interactions in the paramagnetic temperature region. The ZFC/FC measurement in contrast shows an anomaly characteristic for antiferromagnetic ordering at a Neél temperature of TN=43.0(1) K. The antiferromagnetic character of the transition is also evident from the magnetization isotherms recorded at 3, 10 and 25 K, all showing a so called meta-magnetic step (spin-reorientation from an antiparallel to a parallel alignment). The corresponding critical field is below 10 kOe, therefore the high-field measurement suggests ferromagnetic ordering.

Similar behavior is observed for Tb2Rh3Ga (Fig. 10) with θP=37.5(1) K. The ZFC/FC measurement shows two anomalies, one characteristic for antiferromagnetic, the other characteristic for ferromagnetic ordering. Since the order of magnitude for the susceptibility of ferromagnetic materials is much larger compared to antiferromagnetic materials, the weak FM transition most likely originates from a tiny amount of an unknown impurity (although the sample is pure on the level of X-ray powder diffraction). For the AFM transition a Neél temperature of TN=33.4(1) K was determined. The antiferromagnetic character of the transition is also evident from the magnetization isotherms recorded at 3 and 10 K, which show a meta-magnetic step, at a critical field below 10 kOe.

The ZFC/FC measurement of Dy2Rh3Ga (Fig. 11) shows two anomalies with characteristics of ferromagnetic ordering at TC,1=10.9(1) K and TC,2=13.3(1) K. To investigate which one is intrinsic, a heat capacity measurement was conducted (Fig. 11, middle, inset). Two lambda anomalies were observed at T=10.4(1) K and T=13.0(1) K. Since the area under the signal corresponds to the amount of the respective phase present, the anomaly at T=10.4(1) K has to be the intrinsic one. The magnetization isotherms recorded at 3 and 10 K, below the ordering temperature, confirm the ferromagnetic ground state of the compound and show weak hystereses.

Ho2Rh3Ga (Fig. 12) and Er2Rh3Ga (Fig. 13) order ferromagnetically at TC=10.9(1) and TC=13.1(1) K, respectively. The ferromagnetic ground state of both gallides is also evident from the magnetization isotherms recorded at 3 and 10 K.

3.4 Solid state NMR spectroscopy

The 89Y and 71Ga solid state NMR spectra of the compound Y2Rh3Ga and, for comparison, the binary Laves phase YRh2 are shown in Figs. 14 and 15. In agreement with their crystal structures, each spectrum exhibits only one single signal characterized by a very high resonance shift, which exceeds the chemical shift range of typical diamagnetic yttrium compounds (100–400 ppm) significantly [71]. Based on the theoretical calculations discussed below we attribute the high resonance frequency to a strong Knight shift contribution originating from the interaction of the magnetic moments of the 89Y and 71Ga nuclei with the spin density of the conduction electrons near the Fermi edge.

Fig. 14: Experimental and simulated (red lines) 89Y MAS NMR spectra of YRh2 (top; B0=11.7 T; νrot=4 kHz) and Y2Rh3Ga (middle and bottom; B0=11.7 T and 9.4 T; νrot=7 kHz). The signal of the ternary phase exhibits anisotropically broadened spinning sideband manifolds due to a strong Knight shift anisotropy.
Fig. 14:

Experimental and simulated (red lines) 89Y MAS NMR spectra of YRh2 (top; B0=11.7 T; νrot=4 kHz) and Y2Rh3Ga (middle and bottom; B0=11.7 T and 9.4 T; νrot=7 kHz). The signal of the ternary phase exhibits anisotropically broadened spinning sideband manifolds due to a strong Knight shift anisotropy.

Fig. 15: Experimental 71Ga NMR spectrum of Y2Rh3Ga with simulation (red line) recorded at an external magnetic flux density of B0=7.05 T under static conditions.
Fig. 15:

Experimental 71Ga NMR spectrum of Y2Rh3Ga with simulation (red line) recorded at an external magnetic flux density of B0=7.05 T under static conditions.

The ternary Laves phase Y2Rh3Ga shows a significantly higher resonance frequency, suggesting a larger Knight shift contribution than observed for the binary compound YRh2. In addition, a sizeable resonance shift anisotropy is observed, manifesting itself in a spinning sideband manifold. The difference of s-electron density at the Fermi-edge, also determined by total-DOS calculations (vide infra), suggests a partial compensation of the electron withdrawing effect of the highly electronegative Rh (EN=2.28 [55]) atoms by the additional electron density of the Ga atoms (EN=1.81 [55]). The observed shift anisotropy (Δσ=−560 ppm, ησ =0 (Table 7), observed in the field-dependent NMR spectra) is consistent with the axially symmetric local environment of the Y atoms (crystallographic 6c site, symmetry 3m).

Table 7:

89Y and 71Ga NMR spectroscopic parameters of Y2Rh3Ga and YRh2: resonance shift δ (±1)/ppm, full width at half maximum Δ/kHz and degree of Gaussian (G) vs. Lorentzian (L) character of the signal, shift anisotropy and asymmetry parameter Δσ and ησ experimental and calculated quadrupolar coupling constant CQ, exp and CQ, calc/MHz, magnetic flux density of the external field B0/T, MAS spinning frequency νrot/kHz.

CompoundδΔG/LΔσησCQ, expaCQ, calcaB0νrot
Y2Rh3Ga
89Y45454.920.20–560011.77
45443.920.61–56009.47
71Ga184223.2423.537.05
YRh289Y 1342 4.7 0.87 11.7 4

aηQ, calc=ηQ, exp=0.

The lineshape of the static 71Ga NMR spectrum is dominated by strong second-order quadrupolar perturbations originating from the interaction of the I=3/2 nucleus with a strong electric field gradient produced by the non-cubic coordination sphere of the Ga atoms. The experimental quadrupole coupling parameters (CQ, exp=23.24 MHz, ηQ, exp=0) are found in excellent agreement with the calculated values (CQ, calc=23.53 MHz, ηQ, calc=0), revealing an axially symmetric electric field gradient as expected from the symmetry (3̅m) of the Wyckoff site 3a. Overall, the extremely well-defined 71Ga NMR spectrum characterized by a singular value of CQ implies an essentially perfectly ordered structure and the absence of disordering and/or mixed site occupancy. The 71Ga resonance shift is comparable to values found for other intermetallic gallides [72], [73], [74].

3.5 Electronic structure and NMR parameters

To understand the experimentally observed differences of the 89Y NMR signal shifts, electronic structure calculations of YRh2 and Y2Rh3Ga were performed. The total DOS of both compounds is presented in Fig. 16. In a first approximation, the NMR signal shift of metallic materials is frequently dominated by the s-electrons at the Fermi level (s-DOSEF), since they have a nonzero probability at the nuclear site [75], [76]. This correlation was verified for the di- and tetragallides of the alkaline earth metals and NaGa4 [73], [77]. Recently, the validity of this approximation has been confirmed by a careful analysis of the shift contributions in these materials [78]. Consistent with this simple correlation, Y2Rh3Ga possesses a significantly higher calculated s-DOSEF than YRh2 as indicated in Fig. 17.

Fig. 16: Total density of states (DOS) for Y2Rh3Ga (red line) and YRh2 (black line).
Fig. 16:

Total density of states (DOS) for Y2Rh3Ga (red line) and YRh2 (black line).

Fig. 17: s-Contributions of the angular momentum decomposed electronic density of states (s-DOS) for Y in YRh2 (black) and Y2Rh3Ga (red).
Fig. 17:

s-Contributions of the angular momentum decomposed electronic density of states (s-DOS) for Y in YRh2 (black) and Y2Rh3Ga (red).

Beside this, the calculations of the EFGs main principal axis VZZ and the asymmetry parameter ηQ of the Ga atoms are in excellent agreement with the NMR experiments (see Section 3.4 and Table 7). The highly symmetric Y2Rh3Ga possesses a relatively large and negative value of VZZ=−9.092 1021 Vm−2. This indicates a very anisotropic prolate charge distribution of the Ga atoms compared to other Ga compounds [72], [73], [74]. The asymmetry parameter ηQ=0 is in agreement with the experimental value and the site symmetry (3̅m). Considering this site symmetry, VZZ has to be aligned parallel to the c axis of the unit cell. The charge distributions point towards the Rh triangles above and below the Ga atoms with short Rh–Ga distances. This may indicate bonding interactions of these two types of atoms.


Dedicated to: Professor Walter Frank on the Occasion of his 60th birthday.


Acknowledgments

We thank Dipl.-Ing. U. Ch. Rodewald for collection of the single crystal diffractometer data.

References

[1] F. Laves, Theory of Alloy Phases, Am. Soc. Met., Cleveland, 1956.Search in Google Scholar

[2] W. B. Pearson, Acta Crystallogr. B1968, 24, 7.10.1107/S0567740868001627Search in Google Scholar

[3] F. Stein, M. Palm, G. Sauthoff, Intermetallics, 2004, 12, 713.10.1016/j.intermet.2004.02.010Search in Google Scholar

[4] F. Stein, M. Palm, G. Sauthoff, Intermetallics, 2005, 13, 1056.10.1016/j.intermet.2004.11.001Search in Google Scholar

[5] K. A. Gschneidner Jr., V. K. Pecharsky, Z. Kristallogr. 2006, 221, 375.10.1524/zkri.2006.221.5-7.375Search in Google Scholar

[6] O. Eriksson, B. Johansson, M. S. S. Brooks, H. L. Skriver, Phys. Rev. B1989, 40, 9519.10.1103/PhysRevB.40.9519Search in Google Scholar

[7] K. A. Gschneidner Jr., V. K. Pecharsky, A. O. Tsokol, Rep. Prog. Phys. 2005, 68, 1479.10.1088/0034-4885/68/6/R04Search in Google Scholar

[8] D. Shaltiel, J. Less-Common Met. 1978, 62, 407.10.1016/0022-5088(78)90055-3Search in Google Scholar

[9] S. Hong, C. L. Fu, Phys. Rev. B2002, 66, 094109.10.1103/PhysRevB.66.094109Search in Google Scholar

[10] H. Kohlmann, J. Phys. Chem. C2010, 114, 13153.10.1021/jp104156gSearch in Google Scholar

[11] E. Wu, W. Li, J. Li, Int. J. Hydrogen Energy2012, 37, 1509.10.1016/j.ijhydene.2011.10.016Search in Google Scholar

[12] M. Y. Teslyuk, G. I. Oleksiv, Dopov. Akad. Nauk. SSSR, Ser. B1965, 10, 1329.Search in Google Scholar

[13] M. Y. Teslyuk, Metallic Compounds with Laves Type Structure, Nauka, Moscow, 1969.Search in Google Scholar

[14] H. Witte, Metallwirtsch. Metallwiss. Metalltech. 1939, 18, 459.10.1093/jn/18.5.459Search in Google Scholar

[15] E. I. Gladyshevskii, P. I. Krypyakevich, M. Y. Teslyuk, Dopov. Akad. Nauk. SSSR. 1952, 85, 81.Search in Google Scholar

[16] K. Osamura, Y. Murakami, J. Less-Common Met. 1978, 60, 311.10.1016/0022-5088(78)90185-6Search in Google Scholar

[17] H. Arnfelt, A. Westgren, Jernkontorets Ann. 1935, 118, 185.Search in Google Scholar

[18] A. Simon, W. Brämer, B. Hillenkötter, H.-J. Kullmann, Z. Anorg. Allg. Chem. 1976, 419, 253.10.1002/zaac.19764190306Search in Google Scholar

[19] R. Horyń, J. Less-Common Met. 1977, 56, 103.10.1016/0022-5088(77)90223-5Search in Google Scholar

[20] A. S. Markosyan, Sov. Phys. Solid State1981, 23, 965.Search in Google Scholar

[21] E. Gratz, Solid State Commun. 1983, 48, 825.10.1016/0038-1098(83)91027-XSearch in Google Scholar

[22] A. C. Lawson, J. L. Smith, J. O. Willis, J. A. O’Rourke, J. Faber, R. L. Hitterman, J. Less-Common Met. 1985, 107, 243.10.1016/0022-5088(85)90083-9Search in Google Scholar

[23] R. Z. Levitin, A. S. Markosyan, J. Magn. Magn. Mater. 1990, 84, 247.10.1016/0304-8853(90)90102-VSearch in Google Scholar

[24] R. Cywinski, S. H. Kilcoyne, C. A. Scott, J. Phys.: Condens. Matter1991, 3, 6473.10.1088/0953-8984/3/33/023Search in Google Scholar

[25] Y. Zhao, F. Chu, R. B. Von Dreele, Q. Zhu, Acta Crystallogr. B2000, 56, 601.10.1107/S0108768100003633Search in Google Scholar

[26] X. Yan, A. Grytsiv, P. Rogl, H. Schmidt, G. Giester, A. Saccone, X.-Q. Chen, Intermetallics2009, 17, 336.10.1016/j.intermet.2008.11.006Search in Google Scholar

[27] V. Pavlyuk, G. Dmytriv, I. Tarasiuk, I. Chumak, H. Ehrenberg, Acta Crystallogr. C2011, 67, i59.10.1107/S0108270111048566Search in Google Scholar PubMed

[28] O. Osters, T. Nilges, M. Schöneich, P. Schmidt, J. Rothballer, F. Pielnhofer, R. Weihrich, Inorg. Chem. 2012, 51, 8119.10.1021/ic3005213Search in Google Scholar

[29] Y. Mydryk, D. Paudyal, A. K. Pathak, V. K. Pecharsky, K. A. Gschneidner, Jr. J. Mater. Chem. C2016, 4, 4521.10.1039/C6TC00867DSearch in Google Scholar

[30] A. E. Dwight, C. W. Kimball, Acta Crystallogr. B1974, 30, 2791.10.1107/S0567740874008156Search in Google Scholar

[31] J. Wang, Y. Feng, R. Jaramillo, J. van Wezel, P. C. Canfield, T. F. Rosenbaum, Phys. Rev. B2012, 86, 014422.10.1103/PhysRevB.86.014422Search in Google Scholar

[32] P. Villars, K. Cenzual, Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2015/16), ASM International®, Materials Park, Ohio (USA) 2015.Search in Google Scholar

[33] D. Noréus, L. Eriksson, L. Göthe, P.-E. Werner, J. Less-Common Met. 1985, 107, 345.10.1016/0022-5088(85)90093-1Search in Google Scholar

[34] K. Cenzual, B. Chabot, E. Parthé, J. Solid State Chem. 1987, 70, 229.10.1016/0022-4596(87)90061-2Search in Google Scholar

[35] A. Vernière, P. Lejay, P. Bordet, J. Chenavas, J. P. Brison, P. Haen, J. X. Boucherle, J. Alloys Compd. 1994, 209, 251.10.1016/0925-8388(94)91108-8Search in Google Scholar

[36] A. Lipatov, A. Gribanov, A. Grytsiv, S. Safronov, P. Rogl, J. Rousnyak, Y. Seropegin, G. Giester, J. Solid State Chem. 2010, 183, 829.10.1016/j.jssc.2010.01.029Search in Google Scholar

[37] D. Kaczorowski, A. Lipatov, A. Gribanov, Yu. Seropegin, J. Alloys Compd. 2011, 509, 6518.10.1016/j.jallcom.2011.03.144Search in Google Scholar

[38] I. Doverbratt, S. Ponou, S. Lidin, J. Solid State Chem. 2013, 197, 312.10.1016/j.jssc.2012.09.003Search in Google Scholar

[39] A. V. Morozkin, J. Alloys Compd. 2004, 385, L1.10.1016/j.jallcom.2004.05.008Search in Google Scholar

[40] V. Keimes, A. Mewis, Z. Naturforsch. 1992, 47b, 1351.10.1515/znb-1992-1002Search in Google Scholar

[41] R. Pöttgen, A. Lang, R.-D. Hoffmann, B. Künnen, G. Kotzyba, R. Müllmann, B. D. Mosel, C. Rosenhahn, Z. Kristallogr. 1999, 214, 143.10.1524/zkri.1999.214.3.143Search in Google Scholar

[42] K. Yvon, W. Jeitschko, E. Parthé, J. Appl. Crystallogr. 1977, 10, 73.10.1107/S0021889877012898Search in Google Scholar

[43] L. A. O’Dell, R. W. Schurko, Chem. Phys. Lett. 2008, 464, 97.10.1016/j.cplett.2008.08.095Search in Google Scholar

[44] Bruker Corp., Topspin (Version 2.1), Karlsruhe, Germany, 2008.Search in Google Scholar

[45] D. Massiot, F. Fayon, M. Capron, I. King, S. Le Calvé, B. Alonso, J.-O. Durand, B. Bujoli, Z. Gan, G. Hoatson, Magn. Reson. Chem. 2002, 40, 70.10.1002/mrc.984Search in Google Scholar

[46] H. M. Petrilli, P. E. Blöchl, P. Blaha, K. Schwarz, Phys. Rev. B 1998, 57, 14690.10.1103/PhysRevB.57.14690Search in Google Scholar

[47] G. Kresse, J. Furthmüller, Phys. Rev. B, 1996, 54, 11169.10.1103/PhysRevB.54.11169Search in Google Scholar

[48] P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, J. Luitz, WIEN2k, Vienna University of Technology, Vienna, Austria, 2001.Search in Google Scholar

[49] J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett.1996, 77, 3865.10.1103/PhysRevLett.77.3865Search in Google Scholar PubMed

[50] V. Petříček, M. Dušek, L. Palatinus, Jana2006, The Crystallographic Computing System, Institute of Physics, Academy of Sciences of the Czech Republic, Prague (Czech Republic) 2006.Search in Google Scholar

[51] V. Petříček, M. Dušek, L. Palatinus, Z. Kristallogr.2014, 229, 345.10.1515/zkri-2014-1737Search in Google Scholar

[52] V. B. Compton, B. T. Matthias, Acta Crystallogr. 1959, 12, 651.10.1107/S0365110X59001918Search in Google Scholar

[53] F. Hulliger, J. Alloys Compd. 1996, 239, 131.10.1016/0925-8388(96)02262-1Search in Google Scholar

[54] N. C. Baenziger, J. L. Moriarty Jr., Acta Crystallogr. 1961, 14, 948.10.1107/S0365110X6100276XSearch in Google Scholar

[55] J. Emsley, The Elements, Oxford University Press, Oxford 1999.Search in Google Scholar

[56] B. Chevalier, B. Heying, U. Ch. Rodewald, C. Peter Sebastian, E. Bauer, R. Pöttgen, Chem. Mater. 2007, 19, 3052.10.1021/cm0705338Search in Google Scholar

[57] O. Sichevych, W. Schnelle, Yu. Prots, U. Burkhardt, Yu. Grin, Z. Naturforsch. 2006, 61b, 904.10.1515/znb-2006-0719Search in Google Scholar

[58] H. H. Hill, in Plutonium and Other Actinides, Nuclear Materials Series, AIME, (Ed.: W. N. Mines), 1970, 17, 2.Search in Google Scholar

[59] H. Bärnighausen, Commun. Math. Chem. 1980, 9, 139.Search in Google Scholar

[60] U. Müller, Z. Anorg. Allg. Chem. 2004, 630, 1519.10.1002/zaac.200400250Search in Google Scholar

[61] U. Müller, Symmetriebeziehungen zwischen verwandten Kristallstrukturen, Vieweg, Teubner Verlag, Wiesbaden, 2012.10.1007/978-3-8348-8342-1Search in Google Scholar

[62] R. L. Johnston, R. Hoffmann, Z. Anorg. Allg. Chem. 1992, 616, 105.10.1002/zaac.19926161017Search in Google Scholar

[63] L. L. Hirst, Phys. Kondens. Materie1970, 11, 255.10.1007/BF02422643Search in Google Scholar

[64] B. C. Sales, D. K. Wohlleben, Phys. Rev. Lett.1975, 35, 1240.10.1103/PhysRevLett.35.1240Search in Google Scholar

[65] U. B. Paramanik, Anupam, U. Burkhardt, R. Prasad, C. Geibel, Z. Hossain, J. Alloys Compd.2013, 580, 435.10.1016/j.jallcom.2013.05.169Search in Google Scholar

[66] D. Kaczorowski, A. P. Pikul, U. Burkhardt, M. Schmidt, A. Ślebarski, A. Szajek, M. Werwiński, Y. Grin, J. Phys.: Condens. Matter2010, 22, 215601.10.1088/0953-8984/22/21/215601Search in Google Scholar

[67] O. Niehaus, U. C. Rodewald, P. M. Abdala, R. S. Touzani, B. P. T. Fokwa, O. Janka, Inorg. Chem.2014, 53, 2471.10.1021/ic402414fSearch in Google Scholar

[68] O. Niehaus, P. M. Abdala, R. Pöttgen, Z. Naturforsch.2015, 70b, 253.10.1515/znb-2015-0003Search in Google Scholar

[69] H. Sugawara, T. Yamazaki, J. Itoh, M. Takashita, T. Ebihara, N. Kimura, P. Svoboda, R. Settai, Y. Ōnuki, H. Aoki, S. Uji, Physica B1994, 199 & 200, 570.10.1016/0921-4526(94)91908-9Search in Google Scholar

[70] T. Mihalisin, A. Harrus, S. Raaen, R. D. Parks, J. Appl. Phys. 1984, 55, 1966.10.1063/1.333534Search in Google Scholar

[71] J. Mason, Multinuclear NMR, Kluwer Academic/Plenum Publishers, Buckinghamshire, 1987.10.1007/978-1-4613-1783-8Search in Google Scholar

[72] F. Haarmann, K. Koch, D. Grüner, W. Schnelle, O. Pecher, R. Cardoso-Gil, H. Borrmann, H. Rosner, Yu. Grin, Chem. Eur. J. 2009, 15, 1673.10.1002/chem.200801131Search in Google Scholar PubMed

[73] F. Haarmann, K. Koch, P. Jeglič, O. Pecher, H. Rosner, Yu. Grin, Chem. Eur. J. 2011, 17, 7560.10.1002/chem.201003486Search in Google Scholar PubMed

[74] L. Vasylechko, U. Burkhardt, W. Schnelle, H. Borrmann, F. Haarmann, A. Senyshyn, D. Trots, K. Hiebl, Yu. Grin, Solid State Sci. 2012, 14, 746.10.1016/j.solidstatesciences.2012.03.029Search in Google Scholar

[75] A. Abragam, Principles of Nuclear Magnetism, Oxford University Press, Oxford, New York, 1961.10.1063/1.3057238Search in Google Scholar

[76] C. P. Slichter, Principles of Magnetic Resonance, 3rd ed., Springer-Verlag, Berlin, Heidelberg, New York, 1990.10.1007/978-3-662-09441-9Search in Google Scholar

[77] F. Haarmann, Quadrupolar NMR of Intermetallic Compounds, in Enzyclopedia of Magnetic Resonance, (Eds.: R. K. Harris, R. E. Wasylishen), John Wiley & Sons, Ltd, Chichester, 2011.10.1002/9780470034590.emrstm1225Search in Google Scholar

[78] R. Laskowski, K. H. Khoo, F. Haarmann, P. Blaha, J. Phys. Chem. C2017, 121, 753.10.1021/acs.jpcc.6b11210Search in Google Scholar

Received: 2016-12-20
Accepted: 2017-1-24
Published Online: 2017-3-18
Published in Print: 2017-4-1

©2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 19.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/znb-2016-0265/html
Scroll to top button