Abstract
Several samples of the solid solutions CeRu1–xPdxSn and CeRh1–xPdxSn have been synthesized by arc-melting and characterized by X-ray powder diffraction. Guinier powder patterns prove that the ZrNiAl-type structure is the dominating one, besides the CeRuSn and TiNiSi type structures. The structures of CeRu0.28Pd0.72Sn (ZrNiAl type, P6̅2m, a = 751.95(3), c = 418.70(2) pm, wR2 = 0.0274, 332 F2 values, 14 variables) and CeRh0.66Pd0.34Sn (ZrNiAl type, P6̅2m, a = 750.26(3), c = 411.59(2) pm, wR2 = 0.0533, 358 F2 values, 14 variables) were refined from single crystal diffractometer data. Magnetic measurements in combination with XANES (X-Ray Absorption Near Edge Structure) clearly proved intermediate cerium valencies for most compounds and revealed the best fitting parameters for those with the ICF model (Interconfiguration fluctuation). The electrical resistivity is also influenced by the substitutions. At low and high valence electron counts (VECs) metallic character is present, while around the VEC of CeRhSn the typical resistivity behavior for valence fluctuating compounds is observed.
1 Introduction
The equiatomic cerium transition metal (T) stannides CeTSn (T = Ni, Cu, Zn, Ru, Rh, Pd, Ag, Ir, Pt, Au) have thoroughly been studied in the last twenty years with respect to their broadly varying magnetic properties. Their crystal structure depends on the valence electron count (VEC). CeCuSn [1], CeAgSn [2], CeAuSn [2–4], and CeZnSn [5] crystallize with hexagonal ordering variants of the AlB2 type [6]. The transition metal (T) and tin atoms form puckered T3X3 networks which are separated by the cerium atoms. These four stannides contain stable trivalent cerium and exhibit magnetic ordering at low temperatures.
Puckered T3X3 networks also occur in the structures of CeNiSn, CePdSn [7, 8] and CePtSn [9]; however, with strong orthorhombic distortions. These stannides crystallize with the TiNiSi type, space group Pnma. CePdSn (TN = 7 K) [10] and CePtSn (TN = 7.5 K) [9, 11] order antiferromagnetically while CeNiSn is an intermediate-valent compound [12]. Application of hydrostatic pressure leads to reconstructive phase transitions [13–15]. The high-pressure modifications crystallize with the hexagonal ZrNiAl type and one observes changes in the magnetic behavior. The Néel temperature for HP-CePdSn drops to 5 K, and HP-CePtSn shows no magnetic ordering down to 2 K. Changes in the magnetic behavior can also be induced by hydrogenation. CeNiSnH1.8 [16] is a 7 K ferromagnet and the Néel temperature is reduced from 7.5 to 5.0 K on going from CePdSn to CePdSnH [17]. CeRhSn and CeIrSn crystallize with the ZrNiAl type under ambient pressure conditions. Both stannides are valence-fluctuating compounds [18, 19]. Hydrogenation changes the magnetic ground state from an intermediate valent to a nearly trivalent state [20].
The most interesting candidate among the CeTSn stannides is CeRuSn [21–23], a mixed-valent compound. Ordering of the trivalent and intermediate valent cerium leads to extremely short Ce–Ru distances for the intermediate-valence site [24]. CeRuSn shows valence modulations over the whole temperature range [25].
Since one observes a variety of different magnetic ground states as a function of the VEC and the crystal structure, several solid solutions CeT1–xT′xSn with two different transition metals have been studied with respect to changes in the magnetic properties. Prominent examples are CePt1–xNixSn with a transition from an antiferromagnetic Kondo lattice to a mixed valence semiconductor [26], CeNi1–xRhxSn with an interplay between spin-glass-like and non-Fermi-liquid behavior [27] or CeNi1–xRhxSn with first-order valence phase transitions [28]. By substitution of CeNiSn with Cu and Pt the development of magnetism as a function of two parameters, the volume and the electron density, has also been investigated [29].
During our systematic study of cerium-ruthenium intermetallics with short Ce–Ru distances [24, 30] we have investigated several solid solutions of CeRuSn, i. e. CeRu1–xRhxSn [31], Ce1–xLaxRuSn [32], and CeRu1–xNixSn [33]. While CeRuSn is almost intolerant for substitution, the ZrNiAl type becomes the dominant structure in these series offering the opportunity to investigate the physical properties in dependence on the VEC. An important feature is the cerium valence which has been investigated by magnetic and XANES spectroscopic data. Similar developments have also been reported for CeRu1–xNixAl [34]. Herein we discuss, referring to the solid solutions CeRu1–xPdxSn and CeRh1–xPdxSn, in more detail which fitting procedure might be the best for explaining the results of the susceptibility measurements. The Curie-Weiss law, a modified variety of it and the ICF model (Interconfiguration fluctuation) [35] are taken into account. Furthermore the influence of the substitution on the crystal structure and on the electrical resistivity is discussed.
2 Experimental
2.1 Synthesis
Starting materials for the syntheses of the CeRu1–xPdxSn and CeRh1–xPdxSn samples were a cerium ingot (Sigma Aldrich), ruthenium and rhodium powder (Allgemeine Pforzheim; Merck), a palladium sheet (Allgussa) and tin granules (Merck), all with a stated purity better than 99.9%. Small pieces of cerium were cut under dried (sodium wire) paraffin oil, washed with dried (sodium wire) n-hexane, and kept under argon in Schlenk tubes prior to the reactions. At first small cold-pressed pellets of ruthenium, respectively rhodium powder were prepared to prohibit splashing during the arc-melting procedure. The remaining elements were then weighed in the ideal stoichiometric ratios in relation to the compositions of these pellets and arc-melted [36] under argon (ca. 800 mbar). The argon was purified over molecular sieves, silica gel and a titanium sponge (900 K). The molten buttons were re-melted three times to ensure homogeneity. Bulk samples have metallic luster while ground polycrystalline powders are dark grey. The samples are air-stable.
2.2 X-ray diffraction
The polycrystalline samples of both solid solutions were characterized by Guinier powder patterns (imaging plate technique, Fujifilm BAS-1800) using CuKα1 radiation and α-quartz (a = 491.30, c = 540.46 pm) as an internal standard. Standard least-squares refinements led to the hexagonal and orthorhombic lattice parameters listed in Table 1. Comparison of the experimental patterns with calculated [37] ones assured correct indexing.
Refined lattice parameters (Guinier powder data) of different samples of the solid solutions CeRu1–xPdxSn and CeRh1–xPdxSn.
| Compound | a (pm) | b (pm) | c (pm) | V (nm3) |
|---|---|---|---|---|
| CeRu0.9Pd0.1Sn | 741.5(3) | – | 409.6(1) | 0.1950 |
| CeRu0.8Pd0.2Sn | 743.2(2) | – | 411.7(1) | 0.1969 |
| CeRu0.75Pd0.25Sn | 744.2(1) | – | 412.6(1) | 0.1978 |
| CeRu0.7Pd0.3Sn | 745.1(3) | – | 412.5(2) | 0.1983 |
| CeRu0.6Pd0.4Sn | 747.6(1) | – | 414.0(1) | 0.2004 |
| CeRu0.5Pd0.5Sn | 749.4(1) | – | 414.8(1) | 0.2017 |
| CeRu0.4Pd0.6Sn | 750.9(1) | – | 415.0(1) | 0.2026 |
| CeRu0.3Pd0.7Sn | 752.2(1) | – | 414.9(1) | 0.2033 |
| CeRu0.2Pd0.8Sn | 754.7(4) | – | 415.4(2) | 0.2049 |
| CeRu0.2Pd0.8Sn | 748.0(2) | 469.8(1) | 794.3(2) | 0.2791 |
| CeRu0.1Pd0.9Sn | 752.3(1) | 470.9(1) | 795.8(1) | 0.2819 |
| CeRhSn [20] | 744.8(2) | – | 408.0(1) | 0.1960 |
| CeRhSn | 744.7(1) | – | 408.0(1) | 0.1960 |
| CeRh0.9Pd0.1Sn | 746.4(1) | – | 409.5(1) | 0.1976 |
| CeRh0.8Pd0.2Sn | 748.2(1) | – | 410.5(1) | 0.1990 |
| CeRh0.7Pd0.3Sn | 749.7(1) | – | 411.3(1) | 0.2002 |
| CeRh0.6Pd0.4Sn | 751.1(1) | – | 411.9(1) | 0.2012 |
| CeRh0.5Pd0.5Sn | 752.2(2) | – | 412.1(1) | 0.2019 |
| CeRh0.4Pd0.6Sn | 753.9(2) | – | 412.7(1) | 0.2031 |
| CeRh0.3Pd0.7Sn | 754.7(1) | – | 413.3(1) | 0.2039 |
| CeRh0.2Pd0.8Sn | 757.4(2) | – | 413.7(1) | 0.2055 |
| CeRh0.1Pd0.9Sn | 751.5(1) | 469.6(1) | 795.9(1) | 0.2809 |
| CePdSn [17] | 753.0(2) | 469.8(1) | 795.7(2) | 0.2815 |
Standard deviations are given in parentheses.
Irregularly shaped crystals were selected from the CeRu0.28Pd0.72Sn and CeRh0.66Pd0.34Sn samples by mechanical fragmentation. The crystals were glued to quartz fibers using beeswax and investigated by Laue photographs on a Buerger camera (white molybdenum radiation, image plate technique, Fujifilm, BAS-1800) in order to check crystal quality and suitability for intensity data collection. Single-crystal diffraction intensities were collected at room temperature on a Stoe IPDS-II image plate system (graphite monochromatized Mo radiation; λ = 71.073 pm) in oscillation mode. Numerical absorption corrections were applied to the data sets. Details of the data collections and the structure refinements are listed in Table 2.
Crystal data and structure refinement for CeRu0.28Pd0.72Sn and CeRh0.66Pd0.34Sn with ZrNiAl type structure (P6̅2m, Z = 3).
| Empirical formula | CeRu0.28Pd0.72Sn | CeRh0.66Pd0.34Sn |
| Formula weight, g mol–1 | 363.7 | 361.9 |
| Unit cell dimensions, pm | a = 751.95(3) | a = 750.26(3) |
| c = 418.70(2) | c = 411.59(2) | |
| Unit cell volume, nm3 | 0.2050 | 0.2006 |
| Pearson code | hP9 | hP9 |
| Diffractometer type | ipds-II | ipds-II |
| Calculated density, g cm–3 | 8.84 | 8.98 |
| Crystal size, μm3 | 40 × 50 × 60 | 20 × 30 × 80 |
| Transm. ratio (min / max) | 0.161 / 0.243 | 0.264 / 0.718 |
| Detector distance, mm | 70 | 60 |
| Exposure time, min | 2 | 5 |
| ω range / increment, deg | 0–180 / 1.0 | 0–180 / 1.0 |
| Integr. param. A / B / EMS) | 10.9 / –1.7 / 0.012 | 12.4 / 1.9 / 0.012 |
| Absorption coefficient, mm–1 | 31.9 | 31.8 |
| F(000), e | 460 | 460 |
| θ range for data collection, deg | 3.1–33.4 | 3.1–34.8 |
| Range in hkl | ±11, ±11, ±6 | ±12, ±12, ±6 |
| Total no. reflections | 2965 | 3229 |
| Independent reflections / Rint | 332 / 0.0227 | 358 / 0.0146 |
| Reflections with I ≥ 2 σ(I) / Rσ | 327 / 0.0062 | 351 / 0.0053 |
| Data / ref. parameters | 332 / 14 | 358 / 14 |
| Goodness-of-fit on F2 | 1.15 | 2.30 |
| R1 / wR2 for I > 2σ(I)) | 0.0099 / 0.0273 | 0.0166 / 0.0533 |
| R1 / wR2 (all data) | 0.0104 / 0.0274 | 0.0171 / 0.0533 |
| Flack parameter | –0.02(2) | –0.01(4) |
| Extinction coefficient | 0.00233(9) | 0.00161(16) |
| Largest diff. peak / hole, e Å–3 | 1.63 / –1.09 | 1.21 / –1.27 |
2.3 EDX data
The CeRu0.28Pd0.72Sn and CeRh0.66Pd0.34Sn crystals studied on the diffractometer were analyzed by semiquantitative EDX analysis using a Zeiss EVO MA10 scanning electron microscope with CeO2, ruthenium, rhodium, palladium, and tin as standards. No impurity elements heavier than sodium (detection limit of the instrument) were observed. The obtained ratios of the transition metals are with respect to standard deviations of the EDX technique in good agreement to the refined ratios.
2.4 Structure refinements
The two data sets showed hexagonal lattices with high Laue symmetry and no further systematic extinctions. Since isotypism with CeRhSn (space group P6̅2m) was already evident from the Guinier powder patterns, we used the atomic parameters of the rhodium compound [38] as starting values. Both structures were refined using jana2006 [39] with anisotropic atomic displacement parameters for all atoms. A severe problem concerns the small difference in scattering power between ruthenium and palladium (2 electrons) and rhodium and palladium (1 electron). In the first cycles we refined the T1 (2d) and T2 (1a) sites exclusively with ruthenium, respectively rhodium and then refined their occupancy parameters. The 2d site of the ruthenium containing crystal showed full palladium occupancy within two standard deviations and the 1a site was refined with Ru/Pd mixing, leading to the data listed in Table 3. In contrast, we obtained no stable refinement with free rhodium/palladium mixing for the rhodium containing crystal. In agreement with the starting composition and the results from EDX analyses of the crystal, we fixed the Rh/Pd ratio at 66/34 in the final cycles. Refinement of the correct absolute structures was ensured through calculation of the Flack parameter [40, 41]. The final difference Fourier syntheses were flat (Table 2). The positional parameters and interatomic distances (exemplarily for CeRh0.66Pd0.34Sn in comparison with CeRhSn [38]) are listed in Tables 3 and 4. Further information on the structure refinements is available.[1]
Atomic coordinates and equivalent isotropic displacement parameters (pm2) of CeRu0.28Pd0.72Sn and CeRh0.66Pd0.34Sn.
| Atom | Wyckoff position | x | y | z | Ueq | Occ. |
|---|---|---|---|---|---|---|
| CeRu0.28Pd0.72Sn | ||||||
| Ce | 3f | 0.58927(5) | 0 | 0 | 117(1) | |
| Pd1 | 2d | 2/3 | 1/3 | 1/2 | 103(1) | |
| Ru/Pd2 | 1a | 0 | 0 | 0 | 88(2) | 84/16 |
| Sn | 3g | 0.24669(5) | 0 | 1/2 | 87(1) | |
| CeRh0.66Pd0.34Sn | ||||||
| Ce | 3f | 0.58627(8) | 0 | 0 | 86(2) | |
| Rh1/Pd1 | 2d | 2/3 | 1/3 | 1/2 | 97(2) | 66/34 |
| Rh2/Pd2 | 1a | 0 | 0 | 0 | 85(3) | 66/34 |
| Sn | 3g | 0.24827(9) | 0 | 1/2 | 75(2) |
Ueq is defined as one third of the trace of the orthogonalized Uij tensor.
Interatomic distances (pm) of CeRhSn (data taken from ref. [38]) and CeRh0.66Pd0.34Sn.
| CeRhSn [38] | CeRh0.66Pd0.34Sn | ||||||
| Ce | 4 | Rh1 | 303.6 | Ce | 4 | T1 | 305.7 |
| 1 | Rh2 | 309.1 | 1 | T2 | 310.4 | ||
| 2 | Sn | 323.1 | 2 | Sn | 326.6 | ||
| 4 | Sn | 338.2 | 4 | Sn | 340.0 | ||
| 4 | Ce | 388.9 | 4 | Ce | 391.5 | ||
| 2 | Ce | 408.6 | 2 | Ce | 411.6 | ||
| Rh1 | 3 | Sn | 284.9 | T1 | 3 | Sn | 287.4 |
| 6 | Ce | 303.6 | 6 | Ce | 305.7 | ||
| Rh2 | 6 | Sn | 276.5 | T2 | 6 | Sn | 277.6 |
| 3 | Ce | 309.1 | 3 | Ce | 310.4 | ||
| Sn | 2 | Rh2 | 276.5 | Sn | 2 | T2 | 277.6 |
| 2 | Rh1 | 284.9 | 2 | T1 | 287.4 | ||
| 2 | Sn | 322.8 | 2 | Sn | 322.6 | ||
| 2 | Ce | 323.1 | 2 | Ce | 326.6 | ||
| 4 | Ce | 338.2 | 4 | Ce | 340.0 | ||
All distances within the first coordination spheres are listed. Standard deviations are equal or less than 0.1 pm.
2.5 X-Ray Absorption Near Edge Structure (XANES)
A few samples of the solid solutions CeRu1–xPdxSn and CeRh1–xPdxSn have been investigated by X-ray absorption spectra at the CeLIII edge. The data were recorded in transmission mode at the BM01B station (Swiss Norwegian Beamlines, SNBL), ESRF, Grenoble, France. The measurements were performed using a Si(111) double crystal monochromator. The second crystal of the monochromator was detuned by 60% in order to suppress higher harmonic radiation from the synchrotron. The intensities of the incident and transmitted X-rays were monitored with ionization chambers (nitrogen and helium gas filled). All spectra were acquired in a continuous scanning mode from 5680 to 5900 eV, with energy steps of 0.5 eV and 1 s integration time. Each scan was completed after about 7.5 min, 4 scans were collected for each sample and merged.
The samples were ground in an agate mortar under cyclohexane in order to avoid oxidation and then homogeneously mixed with small amounts of cellulose (Sigma-Aldrich) and pressed to pellets. The amount of sample was optimized for transmission measurements. The experiments were performed at ambient conditions. Energy calibration was based on the energy of the first peak of CeO2 (reference compound) at 5731 eV. The athena software was used for data normalization [42].
2.6 Magnetic measurements
The magnetic susceptibility measurements were carried out on a Quantum Design Physical Property Measurement System using the VSM or the AC-Transport option, respectively. For the magnetic measurements approximately 40 mg pieces of the crushed samples were fixed with kapton foil and attached to the sample holder rod. Magnetic investigations were performed in the temperature range of 2.5–350 K with magnetic flux densities up to 80 kOe (1 kOe = 7.96 × 104 A m–1).
For the resistivity measurements two different procedures were pursued. Arc-melted buttons were either grinded and pressed to pellets or directly sintered at 500 K. Subsequently, the annealed buttons were polished in a parallel arrangement. To achieve a parallel geometry, the buttons were embedded in a polymethylmethacrylate (PMMA) matrix. After the first polishing step, a self-built device allowed parallel polishing of the second side. The samples were polished until an approximately 1 mm thick specimen remained inside the polymer matrix. The disc shaped samples were removed by dissolving the PMMA matrix in acetone. The resistivity measurements were carried out by the van-der-Pauw technique [43] in AC transport mode. The ACT puck was modified by a van-der-Pauw press contact assembly purchased from Wimbush Science & Technology. The probes are spring contacts, gold plated over nickel and the distances between the pins were set to 2 mm. The resistivity was measured between 2.5 and 300 K and the recorded data of channel 1 and 2 were converted according to the van-der-Pauw equation given in the Quantum Design Application Note 1076-304.
3 Discussion
3.1 Crystal chemistry
Depending on the valence electron count (VEC), the stannides of the solid solutions CeT1–xT′xSn crystallize with different structure types (see Introduction). The main question concerns a possible change of the structure type through stepwise substitution of the transition metal. Within the solid solution CePt1–xNixSn [26] the VEC remains constant and all members of the solid solution crystallize with the orthorhombic TiNiSi type [44]. This structure type also occurs for the solid solution CeNi1–xCoxSn [28] up to x = 0.5, indicating a certain electronic flexibility. A change in structure type arises for CeNi1–xRhxSn [27]. The orthorhombic TiNiSi type is retained up to x ≈ 0.5 and then switches to the hexagonal ZrNiAl type [45–47], similar to pure CeRhSn. A double switch in structure type occurs for the recently reported solid solution CeRu1–xNixSn [33]. The monoclinic CeRuSn type allows only for tiny ruthenium-nickel substitution and one readily observes the ZrNiAl type up to x ≈ 0.5. In this range the VEC is close to isotypic CeRhSn. With increasing nickel content, the orthorhombic TiNiSi type forms and the cerium ground state switches from intermediate valent to stable trivalent cerium.
In the present work we studied the Ru/Pd and Rh/Pd substitutions in CeRu1–xPdxSn and CeRh1–xPdxSn. Again, monoclinic CeRuSn allows almost no ruthenium substitution, and the structure type readily changes to ZrNiAl. In the entire region up to x ≈ 0.8, the cerium atoms remain in an intermediate valence state, followed by a switch to the orthorhombic TiNiSi type and trivalent cerium. The rhodium-based system shows similar behavior (Fig. 1). Each switch in structure type produces a two-phase region, probably due to small immiscibility gaps.

Development of the unit cell volume per formula unit in the solid solutions CeRu1–xPdxSn and CeRh1–xPdxSn as a function of the palladium content. The development of the unit cell volumes of the phases with ZrNiAl type structure is presented by a line of the best fit for Vegard type behavior. Data points due to powder or single crystal diffraction and literature are drawn as black squares, red circles and blue triangles, respectively.
In the ZrNiAl type regime, substitution of ruthenium (124 pm) and rhodium (125 pm) by the slightly larger palladium (128 pm) atoms [48] leads to an increase of both the a and c lattice parameters (Fig. 2). A small excess increase of the unit cell volume results from the Ru/Pd and Rh/Pd disorder. This results in different bond lengths and hampers perfect long-range order. Table 4 lists the interatomic distances of ternary CeRhSn along with data for the palladium substituted stannide CeRh0.66Pd0.34Sn, showing an almost isotropic increase of the bond lengths.

Development of the a and c lattice parameters in the solid solutions CeRu1–xPdxSn (black) and CeRh1–xPdxSn (red) with ZrNiAl type structure as a function of the palladium content.
Finally we point to the coordination of the transition metal atoms (Fig. 3) in the different structure types. The coordination is always derived from a trigonal prism. These prisms are ideal in the hexagonal ZrNiAl type phases while small distortions occur in the orthorhombic TiNiSi and in the monoclinic CeRuSn type. Only CeRuSn exhibits additional weak Ru–Ru bonding, while there are no close T–T contacts in the orthorhombic and hexagonal phases.

The near neighbor coordination of the transition metal atoms in the ZrNiAl type (top), CeRuSn type (bottom, left and middle) and TiNiSi type (bottom, right) phases CeTSn. Cerium, transition metal and tin atoms are drawn as medium grey, black filled and open circles, respectively.
The structure types TiNiSi and ZrNiAl are well known basic structure types for intermetallic phases, both with several hundred representatives [49]. For further crystal chemical details we refer to review articles [6, 50–52]. The main topic of the present contribution concerns the course of the cerium valence as a function of the transition metal substitution, studied spectroscopically as well as by temperature dependent magnetic susceptibility measurements (vide infra).
3.2 X-Ray Absorption Near Edge Structure (XANES)
Figures 4 and 5 display the results of XANES measurements of the CeLIII edge absorption spectra of the solid solutions CeRu1–xPdxSn and CeRh1–xPdxSn, respectively. Except for CeRu0.1Pd0.9Sn, which crystallizes in the TiNiSi type structure (dashed line), all investigated compounds exhibit the ZrNiAl type structure (solid lines). All spectra have in common a main white line at about 5726 eV that can be attributed to the configuration of Ce3+. Compounds with high ruthenium or rhodium contents exhibit an additional peak approximately 10 eV above the main white line that can be ascribed to the Ce4+ electronic configuration. This characteristic double peak feature reflects a mixed valence state in these compounds as already described in the literature [31–34]. In addition, a systematic development of the relative intensity of the Ce3+ and Ce4+ lines as a function of the VEC can be observed. In general, a higher VEC leads to a decrease of the Ce4+ peak.

Normalized Ce LIII spectra of samples from the solid solution CeRu1–xPdxSn for different palladium contents. The dashed line belongs to the compound that exhibits the TiNiSi type, whereas the solid lines illustrate samples with ZrNiAl type structure. The arrows serve as a guide to the eye to underline the development of the cerium valences.

Normalized Ce LIII spectra of samples from the solid solution CeRh1–xPdxSn for different palladium contents. The arrows serve as a guide to the eye to underline the development of the cerium valences.
For an estimation of the cerium valence of the mixed valence compounds, fittings of the experimental data were performed with a combination of Lorentzians and arctangent functions. The Athena software was used for these purposes [42]. The ratio of the fractions of Ce3+ and Ce4+ ions was estimated from the ratio of the areas of the two Lorentzian curves. This procedure has been applied in previous works [31–34]. The results of this analysis are reported in Table 5 for both solid solutions and are in accordance with the tendencies described above. No second peak that would indicate Ce4+ was observed for CeRu0.1Pd0.9Sn, which exhibits the TiNiSi type, and for CeRh1–xPdxSn with x ≥ 0.5. Thus fitting of the magnetic susceptibility with a Curie-Weiss law should at least be possible for these samples, as discussed in more detail in the next part.
Estimated average cerium valence for the stannides CeRu1–xPdxSn and CeRh1–xPdxSn determined by fitting of the XANES spectra.
| Compound | Average Ce valence |
|---|---|
| CeRu0.9Pd0.1Sn | 3.20(2) |
| CeRu0.8Pd0.2Sn | 3.21(2) |
| CeRu0.7Pd0.3Sn | 3.18(2) |
| CeRu0.6Pd0.4Sn | 3.16(2) |
| CeRu0.5Pd0.5Sn | 3.10(2) |
| CeRu0.4Pd0.6Sn | 3.09(2) |
| CeRu0.1Pd0.9Sn | 3.0(1) |
| CeRh0.9Pd0.1Sn | 3.13(2) |
| CeRh0.8Pd0.2Sn | 3.13(2) |
| CeRh0.7Pd0.3Sn | 3.11(2) |
| CeRh0.5Pd0.5Sn | 3.0(1) |
| CeRh0.3Pd0.7Sn | 3.0(1) |
3.3 Magnetic properties
Magnetic measurements were performed for all phase pure samples that could be obtained with the ZrNiAl type structure in order to analyze the influence of the VEC. Additionally, for CeRu0.1Pd0.9Sn (TiNiSi type structure) no magnetic ordering down to 2.5 K could be observed and the fitting results obtained with the Curie-Weiss law are in good accordance with a trivalent state of cerium (Table 6). The temperature dependence of the reciprocal magnetic susceptibility (χ–1 data) of CeRu1–xPdxSn and CeRh1–xPdxSn is depicted in Figs. 6 and 7, respectively. All samples were investigated in the temperature range of 2.5–350 K with a magnetic field of 10 kOe. No magnetic ordering down to 2.5 K could be observed for any of these compounds in low-field studies. Remarkable is the clear tendency of decreasing χ–1 values with increasing VEC. For CeRu0.1Pd0.9Sn a steep increase of χ–1 can be observed until a negative curvature leads to an almost temperature independent region above 200 K. However, the samples with a higher VEC exhibit a more and more linear increase like it is expected for Curie paramagnetism, which is caused by trivalent cerium. Higher values of the reciprocal susceptibility can only be explained by partially tetravalent cerium.
Fitting parameters of the magnetic measurements (χ–1 data) with the Curie-Weiss law above Tfit.
| Compound | μexp (μB / Ce) | θP (K) | Tfit (K) | Valence |
|---|---|---|---|---|
| CeRu0.4Pd0.6Sn | 2.49(1) | –109 | 130 | 3.02 |
| CeRu0.1Pd0.9Sn | 2.56(1) | –46 | 100 | 3.0 |
| CeRhSn | 2.34(1) | –209 | 120 | 3.04 |
| CeRh0.9Pd0.1Sn | 2.41(1) | –170 | 175 | 3.05 |
| CeRh0.8Pd0.2Sn | 2.47(1) | –138 | 150 | 3.03 |
| CeRh0.7Pd0.3Sn | 2.40(1) | –79 | 130 | 3.05 |
| CeRh0.6Pd0.4Sn | 2.36(1) | –35 | 90 | 3.07 |
| CeRh0.5Pd0.5Sn | 2.40(1) | –29 | 60 | 3.06 |
| CeRh0.4Pd0.6Sn | 2.39(1) | –7 | 60 | 3.06 |
| CeRh0.3Pd0.7Sn | 2.50(1) | –22 | 60 | 3.02 |
| CeRh0.1Pd0.9Sn | 2.56(1) | –47 | 100 | 3.0 |

Temperature dependence of the reciprocal magnetic susceptibility χ–1 of the solid solution CeRu1–xPdxSn with 0.1 ≤ x ≤ 0.6 measured with a magnetic field of 10 kOe.

Temperature dependence of the reciprocal magnetic susceptibility χ–1 of the solid solution CeRh1–xPdxSn with 0 ≤ x ≤ 0.7 measured with a magnetic field of 10 kOe.
In both solid solutions an increase of the amount of trivalent cerium with the increasing valence electron count can be observed. Consequently, the key question is which theoretical model is the best to reasonably describe the data. The most common way to fit the magnetic susceptibility χ in dependence of the temperature T is the Curie-Weiss law.

C is the Curie constant and θp the Weiss constant. This equation neglects the Pauli-paramagnetic contribution of the conduction electrons, which is considered in a modified version by a temperature independent coefficient χ0.

Due to this additional coefficient it is possible to fit a slightly curved development of the reciprocal susceptibility. However, both Curie-Weiss laws do not consider the temperature dependency of the effective magnetic moment, which is the only possibility to explain any decreasing χ–1 value. Such a case can only be described by a higher population of the Ce3+ state at high temperatures in comparison to lower ones. Sales and Wohlleben developed a model (Interconfiguration fluctuation model, ICF) addressing a mixed-valent behavior of rare earth atoms [35].


In these expressions: (i) Eex = E(Ce3+) – E(Ce4+) is the energy gap between the ground states of each cerium configuration; (ii) Tsf the quantum spin fluctuation temperature; (iii) ν(T) the fractional occupancy of the Ce4+ state (valence = 3 + ν(T)); (iv) J1, J2, μ1 and μ2, respectively, the quantum numbers and effective magnetic moments corresponding to the two levels (J1 = μ1 = 0 for Ce4+; J2 = 5/2 and μ2 = 2.54 μB for Ce3+); (v) n the proportion of stable Ce3+ (C = 0.807 emu K mol–1) impurity, and (vi) χ0 the temperature independent part of magnetism. Characteristic features of samples with such a behavior are small susceptibility values (of the order of 10–3 emu per mol Ce atom) and even more important is a characteristic broad peak [53]. Such a peak is caused by a faster population of the state with higher energy (in this case Ce3+) in comparison to the decrease of the susceptibility. Consequently, the absence of a broad peak, as observed in the investigated compounds, does not clearly exclude the usability of the ICF model. Simply a smaller population of the Ce3+ state has to be fulfilled and a situation like this should possibly be described as intermediate-valent instead of interconfiguration fluctuating. Nevertheless the ICF model is the only one of the three ones mentioned above that considers Ce3+ as well as Ce4+ states. A more detailed description of this model can be found in the literature [35, 53, 54].
Referring to the results from XANES it is known that the ruthenium and rhodium rich compounds show significant amounts of tetravalent cerium, while in the palladium rich ones the cerium atoms are almost or totally trivalent. As a consequence, fittings with all mentioned laws have been obtained. The results achieved with the ICF model are summarized in Table 7, while results of the Curie-Weiss fits are listed in Table 6. No parameters obtained from fitting with the modified Curie-Weiss law are mentioned because no physically meaningful results could be achieved. All attempts have led to very small effective magnetic moments, e.g., for CeRu0.5Pd0.5Sn a cerium valence of 3.44 could be calculated, which is not compatible with the XANES and susceptibility results, though very good R2 values have been obtained for the fitting procedure. For four compounds both fitting parameters are listed in Tables 6 and 7 (CeRu0.4Pd0.6Sn and CeRh1–xPdxSn with x = 0–0.2). Cerium valencies calculated with the ICF model are in the range of 3.11–3.13 for these four compounds, while Curie-Weiss fitting leads to values between 3.02 and 3.05. In comparison to the XANES results that show cerium valences in the range of 3.09–3.11, clearly the ICF model shows the better consensus. For CeRh0.7Pd0.3Sn a cerium valence of 3.11 and 3.05 was determined by XANES and Curie-Weiss fitting, respectively. Unfortunately, no fitting with the ICF model was possible without obtaining a negative value for Eex which is synonymous with the physically impossible situation that the Ce3+ has a lower energy than the Ce4+ state. Nevertheless, the results obtained with the Curie-Weiss fit are only slightly deviating and are in even better compliance for higher rhodium contents. For comparison of the susceptibility fitting and the XANES results the cerium valences at 300 K are plotted in Fig. 8 for CeRu1–xPdxSn and in Fig. 9 for CeRh1–xPdxSn. One important remark for the comparison of cerium valencies detected by susceptibility measurements and XANES is the maximum value of about 3.5 in CeO2 determined by XANES [55], while magnetic measurements clearly prove diamagnetism.
Fitting parameters of the magnetic measurements (χ–1 data) with the ICF model.
| Compound | Eex/kB (K) | Tsf (K) | n | χ0 (emu mol–1) | Valence at 300 K |
|---|---|---|---|---|---|
| CeRu0.9Pd0.1Sn | 1494(1) | 951(1) | 0.0334* | 3.2E–04* | 3.34 |
| CeRu0.8Pd0.2Sn | 1100(1) | 687(1) | 0.0333* | 4.4E–04* | 3.33 |
| CeRu0.75Pd0.25Sn | 781(1) | 496(1) | 0.0575* | 3.6E–04* | 3.29 |
| CeRu0.7Pd0.3Sn | 638(1) | 432(1) | 0.0689* | 3.4E–04* | 3.27 |
| CeRu0.6Pd0.4Sn | 396(1) | 319(1) | 0.0518* | 3.2E–04* | 3.23 |
| CeRu0.5Pd0.5Sn | 245(2) | 271(1) | 0.1756* | 1.7E–04* | 3.17 |
| CeRu0.4Pd0.6Sn | 13(3) | 271* | 0.2523(4) | 4.1E–05* | 3.11 |
| CeRhSn | 26(1) | 310* | 0.0836(2) | 3.3E–05* | 3.13 |
| CeRh0.9Pd0.1Sn | 64(2) | 415* | 0.2128(2) | 4.5E–05* | 3.12 |
| CeRh0.8Ru0.2Sn | 9(2) | 278* | 0.2072(3) | 2.6E–05* | 3.11 |
Parameters marked with an asterisk were kept fixed during the last fitting step.

The cerium valence of CeRu1–xPdxSn at 300 K in dependence of the palladium content. The data calculated by the ICF model are compared with those determined by XANES.

The cerium valence of CeRh1–xPdxSn at 300 K in dependence of the palladium content. The data calculated by the ICF model are compared with those determined by XANES.
In Fig. 10 the calculated cerium valencies with the ICF model are plotted in dependence of the temperature. These results are achieved by insertion of the fitting parameters in eq. 4. It is obvious that the temperature dependency of the cerium valence becomes less with increasing palladium content. CeRu0.9Pd0.1Sn exhibits a change of the valence of approximately 0.1 in the investigated temperature range, while it is constant for CeRu0.4Pd0.6Sn. This might be important evidence that Curie-Weiss fitting also becomes useful for these compounds because no change in the cerium valence leads to a linear increase of the inverse susceptibility. The remaining tetravalent character should result in a reduced effective magnetic moment. However, the fitting with the ICF model seems to be the better way for this compound as discussed above. Two general points concerning the fitting with the ICF model should be finally discussed. The first one concerns the high values for n, the proportion of stable Ce3+, which was originally introduced as “impurity term”. As discussed above, less valence change with temperature is observed and the “impurity term” is just a simplified form of the Curie-Weiss law. Consequently, an increasing n underlines again the more sensible applicability of the Curie-Weiss law. Consequently, n cannot anymore be neglected for the calculation of the cerium valence and is considered for all compounds. The second point concerns the fixed fitting parameters during the last refinement cycles. In general, the temperature independent part χ0 and the amount of stable Ce3+n were fixed to prevent over-determination. Due to the increasing importance of n, this parameter was refined, while the quantum spin fluctuation temperature Tsf was fixed.
![Fig. 10: Calculated cerium valence for the solid solution CeRu1–xPdxSn with 0.1 ≤ x ≤ 0.6 in the temperature range from 5 to 350 K by applying equation 2 of the ICF model [35].](/document/doi/10.1515/znb-2015-0003/asset/graphic/znb-2015-0003_fig10.jpg)
Calculated cerium valence for the solid solution CeRu1–xPdxSn with 0.1 ≤ x ≤ 0.6 in the temperature range from 5 to 350 K by applying equation 2 of the ICF model [35].
3.4 Electrical properties
In order to avoid data overlap, only selected resistivity data of both solid solutions are shown in Figs. 11 and 12. All remaining compounds are totally in line with the behavior and the tendencies that will be discussed in this part. As described in the experimental section, two different ways for the sample preparation were chosen. To achieve a better comparability and to neglect the influence of grain boundaries, relative resistivities and no absolute values are discussed here. For the solid solution CeRu1–xPdxSn the sample preparation via sintered tablets was favored, while for the CeRh1–xPdxSn one polished arc-melted buttons turned out to be the best opportunity.

Temperature dependence of the reduced electrical resistivity of samples from the solid solution CeRu1–xPdxSn with x = 0.1, 0.3, 0.5 and 0.7.

Temperature dependence of the reduced electrical resistivity of CeRhSn and samples from the solid solution CeRh1–xPdxSn with x = 0.1, 0.3, 0.5 and 0.7.
Starting with the sample with the lowest VEC (CeRu0.9Pd0.1Sn), metallic behavior is observed. At low temperatures saturation due to the residual resistivity, and a linear increase of ρ(T) can be observed with higher temperatures. Nevertheless, two small anomalies should be discussed. A slightly curved character instead of a totally linear increase and a weak increase instead of linear saturation are observed. Both characteristics are more evident for the samples with higher palladium contents. The slightly curved character turns into a weak and broad maximum around 75–100 K. This would be even more pronounced if the lattice contribution could be subtracted as often described in literature [56–58]. In literature such a behavior is often attributed to crystal field (CF) effects, as for example reported for CeAl2 and CeAl3 ([57, 58] and references therein).
While all results shown for this solid solution exhibit an increase of ρ(T) with increasing temperature, for CeRu0.5Pd0.5Sn it decreases logarithmically. This is an interesting parallel to CeRhSn, which exhibits the same VEC. This decrease at high temperatures is a characteristic of the Kondo effect. Nevertheless, the most prominent feature of a Kondo system, the rapid decrease at low temperatures due to the formation of a spin-compensated state (Kondo state) [59, 60], is not observed for any of the reported compounds. The electrical resistivity of CeRhSn has already been discussed in detail in the literature and is in accordance with the observed result. It is described as Kondo-intermediate valence behavior [18, 56] due to the features discussed above. At temperatures above 50 K an almost identical development of CeRu0.5Pd0.5Sn and CeRhSn is observed, whereas below 50 K both compounds behave totally different. While for CeRhSn a linear and steep decrease is obtained, for CeRu0.5Pd0.5Sn there is an increase. Such an increase has for example been reported for CeNiSn below 6 K and was related to the opening of a pseudo gap [61]. Nevertheless, further investigations on different single crystals proved that the increase of the resistivity at lower temperatures is caused by impurities, and CeNiSn exhibits metallic behavior [62]. According to the investigations on CeNiSn the observed increase below 25 K is most likely due to intrinsic Ce3+ impurities. These are caused by the increasing palladium content and find also expression in higher amounts of stable Ce3+ (n) in the ICF fitting (vide supra).
One particularity is the lower ρ(T)/ρ(270 K) value at low temperatures (e.g., 10 K) of CeRu0.3Pd0.7Sn in comparison with the remaining members of the solid solution. In general, an opposite trend can be observed. Taking into account that the solid solution contains rhodium instead of ruthenium, this feature seems to be consistent. Except for CeRhSn, CeRh0.9Pd0.1Sn exhibits the highest ρ(T)/ρ(270 K) value at 10 K, and a continuous decrease with higher palladium content is observed indicating an increasing metallic character.
4 Conclusion
As recently reported for the solid solution CeRu1–xNixSn, three different structure types have also been proven for the solid solution CeRu1–xPdxSn. Due to the absence of substitution within the CeRuSn type the phase diagrams for both solid solutions are almost identical. The ZrNiAl type is the dominating one and only at high VECs the TiNiSi type is established. This enables the investigation of physical properties in dependence of a broad valence electron count. Magnetic measurements and XANES spectra prove a higher amount of tetravalent cerium with higher ruthenium/rhodium contents or in general higher VECs. From a comparison of both methods it appears that the ICF model is the best fitting procedure for compounds with significant population of the Ce4+ state, at least higher than 10%. However, some deviations from the original published model had to be introduced. The most important one is the different interpretation of the “impurity term” as intrinsic stable Ce3+ due to the increasing palladium content. The increasing amount of trivalent cerium influences also the electrical resistivity. For the lowest and the highest VECs almost ideal metallic behavior can be predicted, while in between typical valence fluctuating behavior with characteristics of a Kondo system is the best description. Very interesting is the almost identical development of CeRhSn and CeRu0.5Pd0.5Sn above 50 K and the opposite one below this temperature. The deviation below 50 K can also be related to the Ce–Pd hybridization.
Acknowledgments
We thank Dipl.-Ing. U. Ch. Rodewald for collection of the single crystal diffraction data. This work was supported by the Deutsche Forschungsgemeinschaft. O. N. is indebted to the NRW Forschungsschule Molecules and Materials – A Common Design Principle for a PhD fellowship.
References
[1] C. P. Sebastian, S. Rayaprol, R.-D. Hoffmann, U. Ch. Rodewald, T. Pape, R. Pöttgen, Z. Naturforsch. 2007, 62b, 647.Search in Google Scholar
[2] M. Lenkewitz, S. Corsépius, G. R. Stewart, J. Alloys Compd. 1996, 241, 121.Search in Google Scholar
[3] D. Niepmann, R. Pöttgen, K. M. Poduska, F. J. DiSalvo, H. Trill, B. D. Mosel, Z. Naturforsch. 2001, 56b, 1.Search in Google Scholar
[4] K. Ļątka, W. Chajec, K. Kmieć, A. W. J. Pacyna, J. Magn. Magn. Mater. 2001, 224, 241.Search in Google Scholar
[5] W. Hermes, B. Chevalier, U. Ch. Rodewald, S. F. Matar, F. Weill, I. Schellenberg, R. Pöttgen, H. Lueken, M. Speldrich, Chem. Mater. 2011, 23, 1096.Search in Google Scholar
[6] R.-D. Hoffmann, R. Pöttgen, Z. Kristallogr. 2001, 216, 127.Search in Google Scholar
[7] D. Rossi, V. Contardi, R. Marazza, D. Mazzone, G. Zanicchi, Atti XVI Congr. Naz. di Chimica Inorganica, Ferrara, September 12–16, Università degli Studi di Ferrara, Ferrara (Italy) 1983, p. 240.Search in Google Scholar
[8] D. Rossi, R. Marazza, R. Ferro, J. Less-Common Met.1985, 107, 99.Search in Google Scholar
[9] J. Sakurai, Y. Yamaguchi, S. Nishigori, T. Suzuki, T. Fujita, J. Magn. Magn. Mater.1990, 90&91, 422.Search in Google Scholar
[10] D. T. Adroja, S. K. Malik, B. D. Padalia, R. Vijayaraghavan, Solid State Commun.1988, 66, 1201.Search in Google Scholar
[11] M. Kyogaku, Y. Kitaoka, H. Nakamura, K. Asayama, F. Teshima, T. Takabatake, H. Fujii, Y. Yamaguchi, J. Sakurai, J. Magn. Magn. Mater. 1990, 90&91, 487.Search in Google Scholar
[12] T. Takabatake, Y. Nakazawa, M. Ishikawa, Jpn. J. Appl. Phys. Suppl. 31987, 26, 547.Search in Google Scholar
[13] J. F. Riecken, G. Heymann, T. Soltner, R.-D. Hoffmann, H. Huppertz, D. Johrendt, R. Pöttgen, Z. Naturforsch. 2005, 60b, 821.Search in Google Scholar
[14] G. Heymann, J. F. Riecken, S. Rayaprol, S. Christian, R. Pöttgen, H. Huppertz, Z. Anorg. Allg. Chem. 2007, 633, 77.Search in Google Scholar
[15] J. F. Riecken, G. Heymann, W. Hermes, U. Ch. Rodewald, R. D. Hoffmann, H. Huppertz, R. Pöttgen, Z. Naturforsch. 2008, 63b, 695.Search in Google Scholar
[16] B. Chevalier, J.-L. Bobet, M. Pasturel, E. Bauer, F. Weill, R. Decourt, J. Etourneau, Chem.Mater.2003, 15, 2181.Search in Google Scholar
[17] B. Chevalier, A. Wattiaux, J.-L. Bobet, J. Phys.: Condens. Matter2006, 18, 1743.10.1088/0953-8984/18/5/026Search in Google Scholar
[18] Y. Bando, T. Suemitsu, K. Takagi, H. Tokushima, Y. Echizen, K. Katoh, K. Umeo, Y. Maeda, T. Takabatake, J. Alloys Compd. 2000, 313, 1.Search in Google Scholar
[19] A. Ślebarski, J. Low Temp. Phys. 2007, 147, 147.Search in Google Scholar
[20] B. Chevalier, C. P. Sebastian, R. Pöttgen, Solid State Sci. 2006, 8, 1000.Search in Google Scholar
[21] J. F. Riecken, W. Hermes, B. Chevalier, R.-D. Hoffmann, F. M. Schappacher, R. Pöttgen, Z. Anorg. Allg. Chem. 2007, 633, 1094.Search in Google Scholar
[22] J. Mydosh, A. M. Strydom, M. Baenitz, B. Chevalier, W. Hermes, R. Pöttgen, Phys. Rev. B2011, 83, 054411.10.1103/PhysRevB.83.054411Search in Google Scholar
[23] K. Prokeš, V. Petříček, E. Ressouche, S. Hartwig, B. Ouladdiaf, J. A. Mydosh, R.-D. Hoffmann, Y. K. Huang, R. Pöttgen, J. Phys.: Condens. Matter2014, 26, 122201.10.1088/0953-8984/26/12/122201Search in Google Scholar
[24] W. Hermes, S. F. Matar, R. Pöttgen, Z. Naturforsch. 2009, 64b, 901.Search in Google Scholar
[25] R. Feyerherm, E. Dudzik, K. Prokeš, J. A. Mydosh, Y. K. Huang, R. Pöttgen, Phys. Rev. B2014, 90, 041104.10.1103/PhysRevB.90.041104Search in Google Scholar
[26] J. Sakurai, R. Kawamura, T. Taniguchi, S. Nishigori, S. Ikeda, H. Goshima, T. Suzuki, T. Fujita, J. Magn. Magn. Mater. 1992, 104–107, 1415.Search in Google Scholar
[27] A. Ślebarski, M. B. Maple, R. E. Baumbach, T. A. Sayles, Phys. Rev. B2008, 77, 245133.10.1103/PhysRevB.77.245133Search in Google Scholar
[28] D. T. Adroja, B. D. Rainford, J. M. de Teresa, A. del Moral, M. R. Ibarra, K. S. Knight, Phys. Rev. B1995, 52, 12790.10.1103/PhysRevB.52.12790Search in Google Scholar
[29] G. M. Kalvius, A. Kratzer, G. Grosse, D. R. Noakes, R.Wäppling, H. V. Löhneysen, T. Takabatake, Y. Echizen, Physica B2000, 289–290, 256.10.1016/S0921-4526(00)00387-2Search in Google Scholar
[30] T. Mishra, R.-D. Hoffmann, C. Schwickert, R. Pöttgen, Z. Naturforsch. 2011, 66b, 771.Search in Google Scholar
[31] O. Niehaus, P. M. Abdala, J. F. Riecken, F. Winter, B. Chevalier, R. Pöttgen, Z. Naturforsch. 2013, 68b, 960.Search in Google Scholar
[32] O. Niehaus, B. Chevalier, P. M. Abdala, F. Winter, R. Pöttgen, Z. Naturforsch. 2013, 68b, 1279.Search in Google Scholar
[33] O. Niehaus, P. M. Abdala, R. S. Touzani, B. P. T. Fokwa, R. Pöttgen, Solid State Sci. 2015, 40, 36.Search in Google Scholar
[34] O. Niehaus, U. Ch. Rodewald, P. M. Abdala, R. S. Touzani, B. P. T. Fokwa, O. Janka, Inorg. Chem.2014, 53, 2471.Search in Google Scholar
[35] B. C. Sales, D. K. Wohlleben, Phys. Rev. Lett. 1975, 35, 1240.Search in Google Scholar
[36] R. Pöttgen, Th. Gulden, A. Simon, GIT Labor-Fachzeitschrift1999, 43, 133.Search in Google Scholar
[37] K. Yvon, W. Jeitschko, E. Parthé, J. Appl. Crystallogr. 1977, 10, 73.Search in Google Scholar
[38] T. Schmidt, D. Johrendt, C. P. Sebastian, R. Pöttgen, K. Łątka, R. Kmieć, Z. Naturforsch. 2005, 60b, 1036.Search in Google Scholar
[39] V. Petříček, M. Dušek, L. Palatinus, Z. Kristallogr.2014, 229, 345.Search in Google Scholar
[40] H. D. Flack, G. Bernadinelli, Acta Crystallogr. 1999, A55, 908.Search in Google Scholar
[41] H. D. Flack, G. Bernadinelli, J. Appl. Crystallogr. 2000, 33, 1143.Search in Google Scholar
[42] B. Ravel, M. Newville, J. Synchrotron Rad.2005, 12, 537.Search in Google Scholar
[43] L. J. van der Pauw, Philips Res. Rep.1958, 13, 1.Search in Google Scholar
[44] C. B. Shoemaker, D. P. Shoemaker, Acta Crystallogr. 1965, 18, 900.Search in Google Scholar
[45] P. I. Krypyakevich, V. Ya. Markiv, E. V. Melnyk, Dopov. Akad. Nauk. Ukr. RSR, Ser. A, 1967, 750.Search in Google Scholar
[46] A. E. Dwight, M. H. Mueller, R. A. Conner, Jr., J. W. Downey, H. Knott, Trans. Met. Soc. AIME1968, 242, 2075.Search in Google Scholar
[47] M. F. Zumdick, R.-D. Hoffmann, R. Pöttgen, Z. Naturforsch. 1999, 54b, 45.Search in Google Scholar
[48] J. Emsley, The Elements, Oxford University Press, Oxford (UK) 1999.Search in Google Scholar
[49] P. Villars, K. Cenzual, Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2014/15), ASM International®, Materials Park, Ohio (USA) 2014.Search in Google Scholar
[50] E. Parthé, L. Gelato, B. Chabot, M. Penzo, K. Cenzual, R. Gladyshevskii, TYPIX-Standardized Data and Crystal Chemical Characterization of Inorganic Structure Types, Gmelin Handbook of Inorganic and Organometallic Chemistry, 8th edition, Springer, Berlin (Germany) 1993.10.1007/978-3-662-10641-9Search in Google Scholar
[51] G. Nuspl, K. Polborn, J. Evers, G. A. Landrum, R. Hoffmann, Inorg. Chem. 1996, 35, 6922.Search in Google Scholar
[52] M. F. Zumdick, R. Pöttgen, Z. Kristallogr. 1999, 214, 90.Search in Google Scholar
[53] D. Kaczorowski, P. Rogl, K. Hiebl, Phys. Rev. B1996, 54, 9891.10.1103/PhysRevB.54.9891Search in Google Scholar
[54] B. Buffat, B. Chevalier, M. H. Tuilier, B. Lloret, J. Etourneau, Solid State Commun. 1986, 59, 17.Search in Google Scholar
[55] R. Niewa, Z. Hu, C. Grazioli, U. Rößler, M. S. Golden, M. Knupfer, J. Fink, H. Giefers, G. Wortmann, F. M. F. de Groot, F. J. DiSalvo, J. Alloys Compd. 2002, 346, 129.Search in Google Scholar
[56] A. Ślebarski, M. B. Maple, E. J. Freeman, C. Sirvent, M. Radłowska, A. Jezierski, E. Granado, Q. Huang, J. W. Lynn, Phil. Mag. B2002, 82, 943.10.1080/13642810110112955Search in Google Scholar
[57] B. Cornut, B. Coqblin, Phys. Rev. B1972, 5, 4541.10.1103/PhysRevB.5.4541Search in Google Scholar
[58] V. U. S. Rao, W. E. Wallace, Phys. Rev. B1970, 2, 4613.10.1103/PhysRevB.2.4613Search in Google Scholar
[59] J. Kondo, Prog. Theor. Phys.1964, 32, 37.Search in Google Scholar
[60] K. H. J. Buschow, H. J. Van Daal, Phys. Rev. Lett. 1969, 23, 408.Search in Google Scholar
[61] T. Takabatake, F. Teshima, H. Fujii, S. Nishigori, T. Suzuki, T. Fujita, Y. Yamaguchi, J. Sakurai, D. Jaccard, Phys. Rev. B1989, 41, 9607.10.1103/PhysRevB.41.9607Search in Google Scholar
[62] T. Takabatake, G. Nakamoto, T. Yoshino, H. Fujii, K. Izawa, S. Nishigori, H. Goshima, T. Suzuki, T. Fujita, K. Maezawa, T. Hiraoko, Y. Okayama, I. Oguro, A. A. Menovsky, K. Neumaier, A. Brückl, K. Andres, Physica B1996, 223–224, 413.10.1016/0921-4526(96)00137-8Search in Google Scholar
©2015 by De Gruyter
Articles in the same Issue
- Frontmatter
- In this Issue
- High-pressure syntheses and crystal structures of orthorhombic DyGaO3 and trigonal GaBO3
- Crystal structure, magnetic, fluorescent, electrochemical properties and thermal stability of a new copper(II) coordination polymer [Cu2(C5H4NCOO)2(C7H5N4)2]n
- Efficient synthesis of 2,3-dimethoxy-5-methyl-6-morpholinomethyl-1,4-benzoquinone hydrochloride
- Orthoamides and iminium salts, LXXXIX. Reactions of N,N,N′,N′,N″,N″,N′″,N′″-octamethyl-acetylene-bis(carboxamidinium) tetrafluoroborate with nucleophilic reagents – new methods for the preparation of amidinium salts and ketene aminalsa
- Synthesis of bis-thiazolidin-4-ones from N,N,N″-(1,ω-alkanediyl)bis(N″-organylthiourea) derivatives
- Tl2C2O4·H2C2O4: a new crystalline form of thallium(I) oxalate
- The solid solutions CeRu1–xPdxSn and CeRh1–xPdxSn – Applicability of the ICF model to determine intermediate cerium valencies by comparison with XANES data
- Sr(Hg1–xSnx)4: variations of the EuIn4-type structure
- A diethylhydroxylaminate based mixed lithium/beryllium aggregate
- Note
- The crystal structure of Sc5Co2In4
Articles in the same Issue
- Frontmatter
- In this Issue
- High-pressure syntheses and crystal structures of orthorhombic DyGaO3 and trigonal GaBO3
- Crystal structure, magnetic, fluorescent, electrochemical properties and thermal stability of a new copper(II) coordination polymer [Cu2(C5H4NCOO)2(C7H5N4)2]n
- Efficient synthesis of 2,3-dimethoxy-5-methyl-6-morpholinomethyl-1,4-benzoquinone hydrochloride
- Orthoamides and iminium salts, LXXXIX. Reactions of N,N,N′,N′,N″,N″,N′″,N′″-octamethyl-acetylene-bis(carboxamidinium) tetrafluoroborate with nucleophilic reagents – new methods for the preparation of amidinium salts and ketene aminalsa
- Synthesis of bis-thiazolidin-4-ones from N,N,N″-(1,ω-alkanediyl)bis(N″-organylthiourea) derivatives
- Tl2C2O4·H2C2O4: a new crystalline form of thallium(I) oxalate
- The solid solutions CeRu1–xPdxSn and CeRh1–xPdxSn – Applicability of the ICF model to determine intermediate cerium valencies by comparison with XANES data
- Sr(Hg1–xSnx)4: variations of the EuIn4-type structure
- A diethylhydroxylaminate based mixed lithium/beryllium aggregate
- Note
- The crystal structure of Sc5Co2In4