Lars Schumachera,
Florian Schreiner
b,
Aylin Koldemir
a,
Oliver Janka
c,
Michael Ryan Hansen
*b and
Rainer Pöttgen
*a
aInstitut für Anorganische und Analytische Chemie, Universität Münster, Corrensstraße 30, 48149 Münster, Germany. E-mail: pottgen@uni-muenster.de
bInstitut für Physikalische Chemie, Universität Münster, Corrensstrasse 30, D-48149 Münster, Germany
cUniversität des Saarlandes, Anorganische Festkörperchemie, Campus C4 1, 66123 Saarbrücken, Germany
First published on 23rd January 2025
The cadmium-rich intermetallic compounds ARh2Cd20 (A = Ca, Sr, Y, La-Nd, Sm-Lu) were synthesized from the elements in sealed tantalum tubes. The elements were reacted in an induction furnace and the samples were post-annealed to increase phase purity and crystallinity. The ARh2Cd20 phases crystallize with the cubic CeCr2Al20 type structure, space group Fdm. The polycrystalline samples were characterized by X-ray powder diffraction. The structures of SrRh2Cd20, LaRh2Cd20, CeRh2Cd20, TbRh2Cd20 and DyRh2Cd20 were refined from X-ray single crystal diffractometer data. Temperature dependent magnetic susceptibility data show diamagnetism for CaRh2Cd20, SrRh2Cd20, YRh2Cd20, LaRh2Cd20 and YbRh2Cd20, thus substantiating stable divalent ytterbium in the latter phase. The remaining phases are Curie Weiss paramagnets. EuRh2Cd20 contains stable divalent europium and orders ferromagnetically at TC = 13.8 K. Antiferromagnetic ordering was detected for SmRh2Cd20 (TN = 4.3 K), GdRh2Cd20 (TN = 9.3K), TbRh2Cd20 (TN = 6.6 K) and DyRh2Cd20 (TN = 3.9 K). TbRh2Cd20 and DyRh2Cd20 exhibit metamagnetic transitions at critical fields of 19 respectively 10 kOe. The divalent ground state in EuRh2Cd20 was also confirmed by a 151Eu Mössbauer spectrum which shows an isomer shift of δ = −10.86(1) mm s−1. Solid-state NMR spectroscopy was performed for the samples YRh2Cd20, LaRh2Cd20, and YbRh2Cd20. The 113Cd NMR spectra include three distinct Cd sites for each sample in accordance with the crystallographic data. All samples further show negative Knight shifts for 103Rh, suggesting high s–d exchange interaction at the Rh site. In the case of 139La, a residual quadrupolar coupling was observed despite cubic site symmetry, confirming the existence of local defect sites. 89Y and 171Yb NMR spectra were recorded, the latter confirming the divalent nature of Yb in YbRh2Cd20.
Meanwhile more than 200 representatives of the CeCr2Al20 type are known2 and their crystal chemistry and physical properties have recently been reviewed.3 The striking physical properties of these phases concern superconductivity (e.g., TC = 1.65 K for Ga0.2V2Al20 or TC = 1.00 K for ScV2Al20),4,5 thermoelectric materials6 and especially the heavy fermion zincides YbT2Zn20 (T = Fe, Co, Ru, Rh, Os and Ir). The outstanding compound in this structural family is YbCo2Zn20 with a γ value as large as 7900 mJ mol−1 K−2.7,8
Although a huge number of aluminides, zinc and cadmium phases has been reported, the rare earth-based series are far from been entirely investigated with respect to phase analyses and also the physical property studies are far from being complete. In the present study we complete the series of RERh2Cd20 compounds. So far only the members with RE = Ce and Pr have been studied;9–12 however, no lattice parameters were reported, and no precise structure refinements were performed. The characterization relied only on the magnetic properties. CeRh2Cd20 orders magnetically at 0.3 K.12 PrRh2Cd20 saturates above 300 kOe (without a metamagnetic transition). Besides the diffraction experiments (X-ray powder data for all samples and structure refinements for SrRh2Cd20, LaRh2Cd20, CeRh2Cd20, TbRh2Cd20 and DyRh2Cd20) we also report on the magnetic properties, 151Eu Mössbauer spectroscopy on EuRh2Cd20 and a detailed 89Y, 103Rh, 113Cd and 139La solid-state NMR spectroscopic characterization.
The samples were then placed in a water-cooled sample chamber of an induction furnace (Hüttinger Elektronik, Freiburg, Germany, type TIG 1.5/300).14 First, they were heated to 1000 K for 30 min to react the cadmium and to prevent the ampoule to burst. After that, the samples were heated twice to 1300 K for 90 s and cooled again to 1100 K. This temperature was kept for two hours, and the samples were cooled to room temperature by turning off the furnace.
The ampoules were subsequently sealed in silica tubes under vacuum (as oxidation protection) and heated to 823 K for up to 20 days in tube furnaces. The early rare earth elements formed the desired phases already after short thermal treatments while the samples with the late rare earth elements deserved longer annealing steps. All samples were silvery with a metallic luster and stable in air over month.
![]() | ||
Fig. 1 Calculated (top) and experimental (bottom) Guinier powder patterns (CuKα1 radiation) of the YbRh2Cd20 sample. |
Compound | a (pm) | V (nm3) | Ref. |
---|---|---|---|
a Single crystal data. | |||
CaRh2Cd20 | 1562.9(2) | 3.8176 | This work |
SrRh2Cd20 | 1569.1(1) | 3.8632 | This work |
SrRh2Cd20a | 1570.29(9) | 3.8720 | This work |
YRh2Cd20 | 1555.76(6) | 3.7655 | This work |
LaRh2Cd20 | 1564.8(1) | 3.8316 | This work |
LaRh2Cd20a | 1565.92(9) | 3.8398 | This work |
CeRh2Cd20 | 1562.9(1) | 3.8176 | This work |
CeRh2Cd20a | 1564.67(8) | 3.8306 | This work |
CeRh2Cd20 | 1558 | 3.7818 | 12 |
PrRh2Cd20 | 1561.43(6) | 3.8069 | This work |
NdRh2Cd20 | 1561.0(1) | 3.8037 | This work |
SmRh2Cd20 | 1559.0(1) | 3.7891 | This work |
EuRh2Cd20 | 1567.0(2) | 3.8478 | This work |
GdRh2Cd20 | 1557.3(1) | 3.7767 | This work |
TbRh2Cd20a | 1556.13(9) | 3.7682 | This work |
TbRh2Cd20 | 1556.4(1) | 3.7702 | This work |
DyRh2Cd20 | 1555.1(1) | 3.7608 | This work |
DyRh2Cd20a | 1555.38(10) | 3.7628 | This work |
HoRh2Cd20 | 1554.7(1) | 3.7579 | This work |
ErRh2Cd20 | 1554.35(5) | 3.7553 | This work |
TmRh2Cd20 | 1553.32(9) | 3.7479 | This work |
YbRh2Cd20 | 1563.0(2) | 3.8184 | This work |
LuRh2Cd20 | 1553.4(1) | 3.7484 | This work |
Additional powder X-ray diffraction (PXRD) patterns of the LaRh2Cd20 samples for Rietveld refinements were recorded at room temperature on a D8-A25-Advance diffractometer (Bruker, Karlsruhe, Germany) in Bragg–Brentano θ–θ-geometry (goniometer radius 280 mm) with CuKα-radiation (λ = 154.0596 pm). A 12 μm Ni foil working as Kβ filter and a variable divergence slit were mounted at the primary beam side. A LYNXEYE detector with 192 channels was used at the secondary beam side. Experiments were carried out in a 2θ range of 6 to 130° with a step size of 0.013° and a total scan time of 1 h.
Irregularly-shaped single crystals were selected from the carefully crushed SrRh2Cd20, LaRh2Cd20, CeRh2Cd20, TbRh2Cd20 and DyRh2Cd20 samples. The crystals were fixed to small glass fibres using beeswax, and their quality for diffractometer data collection was tested by Laue photographs on a Buerger camera with white Mo radiation. Complete data sets were collected at room temperature using a Stoe IPDS-II image plate system (graphite-monochromatized MoKα 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 refinement data are summarized in Table 2.
Refined composition | SrRh2Cd20 | LaRh2Cd20 | CeRh2Cd20 | TbRh2Cd20 | DyRh2Cd20 |
---|---|---|---|---|---|
Formula weight, g mol−1 | 2541.6 | 2592.9 | 2594.1 | 2612.9 | 2616.5 |
Lattice parameter, pm (single crystal data) | a = 1570.29(9) | a = 1565.92(9) | a = 1564.67(8) | a = 1556.13(9) | a = 1555.38(10) |
Unit cell volume, nm3 | 3.8720 | 3.8398 | 3.8306 | 3.7682 | 3.7628 |
Calculated density, g cm−3 | 8.72 | 8.97 | 9.00 | 9.21 | 9.24 |
Crystal size, μm | 20 × 55 × 60 | 20 × 20 × 80 | 20 × 30 × 210 | 10 × 30 × 35 | 20 × 60 × 160 |
Transmission (min/max) | 0.430/0.674 | 0.455/0.789 | 0.195/0.664 | 0.157/0.658 | 0.131/0.726 |
Detector distance, mm | 70 | 70 | 70 | 70 | 70 |
ω range, increment, ° | 0.0–180, 1.0 | 0.0–180, 1.0 | 0–180, 1.0 | 0–180, 1.0 | 0–180, 1.0 |
Exposure time, min | 10 | 20 | 12 | 10 | 5 |
Integr. param. (A, B, EMS) | 14.0/−1.0/0.030 | 12.7/−0.3/0.012 | 14.0/−1.0/0.030 | 12.7/−0.3/0.012 | 14.0/−1.0/0.030 |
Abs. coefficient, mm−1 | 25.8 | 25.4 | 25.6 | 27.4 | 27.6 |
F(000), e | 8704 | 8856 | 8864 | 8920 | 8928 |
θ range, ° | 2.25–33.28 | 2.25–33.39 | 2.25–33.29 | 2.27–33.33 | 2.27–33.38 |
Range in hkl | ±24/−23 to 24/±24 | −24 to 22/±24/±24 | ±23/−24 to 21/±24 | −20 to 24/−22 to 21/±24 | ±23/±23/−24 to 22 |
Total no. reflections | 11![]() |
11![]() |
11![]() |
11![]() |
11![]() |
Independent refl./Rint | 404/0.0930 | 403/0.0840 | 401/0.0420 | 392/0.0650 | 391/0.0792 |
Refl. with I ≥ 3σ(I)/Rσ | 319/0.0194 | 313/0.0161 | 365/0.0059 | 318/0.0122 | 294/0.0174 |
Data/parameters | 404/17 | 403/17 | 401/17 | 392/17 | 391/17 |
Goodness-of-fit on F2 | 1.05 | 0.98 | 1.17 | 1.00 | 0.99 |
R/wR for I > 3σ(I) | 0.0192/0.0178 | 0.0155/0.0160 | 0.0134/0.0315 | 0.0143/0.0148 | 0.0163/0.0304 |
R/wR for all data | 0.0334/0.0196 | 0.0323/0.0178 | 0.0186/0.0331 | 0.0234/0.0161 | 0.0276/0.0325 |
Extinction coefficient | 1420(40) | 330(20) | 1390(50) | 2490(50) | 2040(50) |
Largest diff. peak /hole, e Å−3 | 1.70/−2.21 | 1.17/−3.38 | 1.64/−1.65 | 0.87/−1.93 | 2.04/−1.90 |
Atom | Wyck. site | x | y | z | U11 | U22 | U33 | U12 | U13 | U23 | Ueq |
---|---|---|---|---|---|---|---|---|---|---|---|
SrRh2Cd20 | |||||||||||
Sr | 8a | 1/8 | 1/8 | 1/8 | 94(3) | U11 | U11 | 0 | 0 | 0 | 94(2) |
Rh | 16d | 1/2 | 1/2 | 1/2 | 75(2) | U11 | U11 | −11(2) | U12 | U12 | 75(1) |
Cd1 | 96g | 0.05888(2) | x | 0.32648(3) | 198(2) | U11 | 127(2) | −76(2) | −7(1) | U13 | 174(1) |
Cd2 | 48f | 0.48879(3) | 1/8 | 1/8 | 116(2) | 114(2) | U22 | 0 | 0 | −40(2) | 115(1) |
Cd3 | 16c | 0 | 0 | 0 | 266(3) | U11 | U11 | −59(2) | U12 | U12 | 266(2) |
LaRh2Cd20 | |||||||||||
La | 8a | 1/8 | 1/8 | 1/8 | 72(2) | U11 | U11 | 0 | 0 | 0 | 72(1) |
Rh | 16d | 1/2 | 1/2 | 1/2 | 76(2) | U11 | U11 | −13(2) | U12 | U12 | 76(1) |
Cd1 | 96g | 0.05921(2) | x | 0.32552(2) | 177(2) | U11 | 116(2) | −62(1) | −9(1) | U13 | 157(1) |
Cd2 | 48f | 0.48845(3) | 1/8 | 1/8 | 109(2) | 111(1) | U22 | 0 | 0 | −35(2) | 110(1) |
Cd3 | 16c | 0 | 0 | 0 | 225(2) | U11 | U11 | −48(3) | U12 | U12 | 225(1) |
CeRh2Cd20 | |||||||||||
Ce | 8a | 1/8 | 1/8 | 1/8 | 119(1) | U11 | U11 | 0 | 0 | 0 | 119(1) |
Rh | 16d | 1/2 | 1/2 | 1/2 | 94(1) | U11 | U11 | −11(1) | U12 | U12 | 94(1) |
Cd1 | 96g | 0.05905(1) | x | 0.32569(2) | 191(1) | U11 | 141(1) | −55(1) | −10(1) | U13 | 174(1) |
Cd2 | 48f | 0.48862(2) | 1/8 | 1/8 | 129(1) | 129(1) | U22 | 0 | 0 | −36(1) | 129(1) |
Cd3 | 16c | 0 | 0 | 0 | 206(1) | U11 | U11 | −35(1) | U12 | U12 | 206(1) |
TbRh2Cd20 | |||||||||||
Tb | 8a | 1/8 | 1/8 | 1/8 | 87(1) | U11 | U11 | 0 | 0 | 0 | 87(1) |
Rh | 16d | 1/2 | 1/2 | 1/2 | 76(1) | U11 | U11 | −13(1) | U12 | U12 | 76(1) |
Cd1 | 96g | 0.05972(1) | x | 0.32414(1) | 175(1) | U11 | 124(2) | −60(1) | −14(1) | U13 | 158(1) |
Cd2 | 48f | 0.48803(2) | 1/8 | 1/8 | 109(2) | 110(1) | U22 | 0 | 0 | −37(1) | 109(1) |
Cd3 | 16c | 0 | 0 | 0 | 199(2) | U11 | U11 | −26(2) | U12 | U12 | 199(1) |
DyRh2Cd20 | |||||||||||
Dy | 8a | 1/8 | 1/8 | 1/8 | 93(1) | U11 | U11 | 0 | 0 | 0 | 93(1) |
Rh | 16d | 1/2 | 1/2 | 1/2 | 79(2) | U11 | U11 | −12(2) | U12 | U12 | 79(1) |
Cd1 | 96g | 0.05978(2) | x | 0.32400(2) | 177(1) | U11 | 129(2) | −61(1) | −14(1) | U13 | 161(1) |
Cd2 | 48f | 0.48804(3) | 1/8 | 1/8 | 110(2) | 112(1) | U22 | 0 | 0 | −36(2) | 112(1) |
Cd3 | 16c | 0 | 0 | 0 | 199(1) | U11 | U11 | −20(2) | U12 | U12 | 199(1) |
Tb: | 4 | Cd3 | 336.9 | Cd2: | 2 | Rh | 275.7 |
12 | Cd1 | 341.6 | 2 | Cd1 | 292.7 | ||
Rh: | 6 | Cd2 | 275.7 | 4 | Cd2 | 301.4 | |
6 | Cd1 | 303.6 | 4 | Cd1 | 307.3 | ||
Cd1: | 1 | Cd1 | 287.3 | Cd3: | 12 | Cd1 | 331.1 |
1 | Cd2 | 292.7 | 2 | Tb | 336.9 | ||
2 | Cd1 | 294.6 | |||||
1 | Rh | 303.6 | |||||
2 | Cd2 | 307.3 | |||||
2 | Cd1 | 316.0 | |||||
2 | Cd3 | 331.1 | |||||
1 | Tb | 341.6 |
CCDC 2393492 (SrRh2Cd20), 2393494 (LaRh2Cd20), 2393488 (CeRh2Cd20), 2393482 (TbRh2Cd20) and 2393487 (DyRh2Cd20)† contain the supplementary crystallographic data for this paper.
Solid-state NMR spectroscopic experiments were conducted using the static Wideline Uniform Rate Smooth Truncation (WURST) Carr–Purcell–Meiboom–Gill (WCPMG) sequence for 89Y, 103Rh, and 171Yb,25 the WCPMG-MAS sequence for 113Cd,26 and direct acquisition for 139La. Optimal pulse power νopt for WURST pulses was calculated as νopt = 0.265/(I + 1/2)*sqrt(R) with I being the nuclear spin and R being the sweep rate of the WURST pulse, calculated as R = Ω/τ with sweep width Ω and pulse length τ. Note that low sweep rates R were generally used to avoid pulse artifacts27 and keep the power requirements low for νopt, which was particularly crucial for the static 103Rh WCPMG NMR spectroscopic experiments. All experimental parameters for the WCPMG and WCPMG-MAS experiments are summarized in Table S1.† The direct excitation 139La MAS NMR experiments were performed at a MAS frequency νMAS = 62.5 kHz using central transition selective π/8 pulses of 1.25 μs with a recycle delay of 25 ms and 524288 scans. All data was processed with a lab-written Python code and plotted using the Matplotlib package.28 Fitting of the spectra was performed using the ssNake software package29 and shifts are given in the Haeberlen-Mehring-Spiess30 convention. Since contributions from chemical, Knight, and pseudo-contact shifts to the observed shift cannot be distinguished, all further mentions will be made using shift and shift anisotropy (SA).
The plot of the lattice parameter obtained in this work (Fig. 2) shows the expected lanthanide contraction. We observe a smooth decrease of the lattice parameters from the lanthanum to the lutetium compound. The parameter for YRh2Cd20 fits in between the values for TbRh2Cd20 and DyRh2Cd20. This is also the case for several other series of rare earth-transition metal-cadmium intermetallics.32 The lattice parameters of EuRh2Cd20 and YbRh2Cd20 show strong positive deviations from the smooth plot. Both compounds show stable divalent ground states, as confirmed by magnetic susceptibility measurements as well as Mössbauer and solid-state NMR spectroscopy (vide infra). Similar valence behavior was also reported for the isotypic series of RETi2Al20 phases.33 In agreement with the course of the ionic radii, the cell parameters of CaRh2Cd20 and SrRh2Cd20 are comparable to YbRh2Cd20 and EuRh2Cd20 with divalent ytterbium and europium, respectively.
As for the LaRh2Cd20 sample, additional signals were observed in the initial NMR spectroscopic experiments (vide infra) indicating a potential formation of a side product not directly visible in the Guinier X-ray diffraction patterns. Therefore, high resolution powder X-ray diffraction patterns were recorded for Rietveld refinements. Analysis of the first recorded diffraction pattern (Fig. S17a†) clearly indicated that the sample showed the targeted phase as the main product, however, additional reflections were observed, which could be attributed to RhCd9+δ, a brass related superstructure (Zn90.6Ir11.1 type, space group F3m, Pearson code cF452).34 Due to the complex structure and the possibility of disorder alongside the quite similar atomic form factors of Rh and Cd, the structural data was used as published, only the lattice parameter was refined (Table S2†). The first measurement of the sample yielded ∼30 mass% RhCd9+δ. In order to prove the proposed instability of LaRh2Cd20, the sample was remeasured after three weeks, indicating an increase of the RhCd9+δ phase to ∼40 mass% (Fig. S17b†). The property studies were then conducted with a new, phase-pure sample (on the lever of X-ray powder diffraction). The observation of this decomposition reaction for the cadmium-based phases is a new feature in the family of CeCr2Al20 type compounds. The RE-Al and RE-Zn phase diagrams show binary phases with distinctly different crystal structures in the aluminum- and zinc-rich parts. Thus, the decomposition reaction is unique for the cadmium-based phases.
Of the whole ARh2Cd20 family, the structures of SrRh2Cd20, LaRh2Cd20, CeRh2Cd20, TbRh2Cd20 and DyRh2Cd20 were refined from single crystal X-ray diffractometer data. Herein we exemplarily discuss the structure of TbRh2Cd20 (Fig. 3). The substructure of the terbium and rhodium atoms has the topology of the cubic Laves phase MgCu2. This is also the case for binary TbRh2.35 This substructure is then strongly expanded (the lattice parameter increases from 749.1 pm for TbRh2 to 1556.4 pm for TbRh2Cd20) and the terbium and rhodium atoms are solely coordinated by cadmium. The smaller rhodium atoms have coordination number (CN) 12 in form of an icosahedron and the terbium atoms have CN16 in form of a Frank–Kasper36,37 polyhedron. The TbRh2Cd20 structure can consequently be described by a packing of these two coordination polyhedra. The cadmium coordination prevents direct Tb–Rh contacts. This crystal chemical motif is somewhat comparable to the rare earth-rich phases La15Rh5Cd2,38 Gd4RhCd,39 Er10RhCd340 and Gd23Rh7Cd4
41 where the rhodium and cadmium atoms are completely separated from each other. The basic building units in these phases are rhodium centered trigonal rare earth prisms and Cd4 tetrahedra, a rare structural motif.
The Tb–Cd distances in the Frank–Kasper polyhedron of TbRh2Cd20 are 4 × 337 and 12 × 342 pm. They compare well with the Tb@Cd12 anticuboctahedron in TbCd342 (317–351 pm Tb–Cd) and the Tb@Cd16 Frank–Kasper polyhedron in TbCd643 (326–362 pm Tb–Cd). These Tb–Cd distances are all slightly longer than the sum of the covalent radii of 300 pm for Tb + Cd.44 This is also the case for the Rh–Cd distances (6 × 276 and 6 × 304 pm) within the icosahedra (the sum of the covalent radii44 is 266 pm for Rh + Cd). Similar ranges of Rh–Cd distances were observed in the binary compounds Rh2Cd5
45 (274–300 pm) and Rh2Cd15
46 (280–298 pm).
The TbRh2Cd20 structure exhibits three crystallographically independent cadmium sites with coordination numbers 14 (Cd3) and 12 (Cd1 and Cd2) in the form of bicapped hexagonal respectively pentagonal prisms (Fig. 3). The various Cd–Cd distances range from 287 to 342 pm, comparable to hcp cadmium (6 × 298 and 6 × 329 pm).47 The different coordination numbers for the cadmium atoms influence the displacement parameters. In both TbRh2Cd20 and DyRh2Cd20 the Cd3 atoms (CN14) show enhanced displacements, a consequence of the slightly larger cage. This crystal chemical feature was observed for all other single crystal structure refinements for RET2Al20, RET2Zn20 and RET2Cd20 phases.3
![]() | ||
Fig. 4 Temperature dependence of the magnetic susceptibilities of CaRh2Cd20, SrRh2Cd20, YRh2Cd20, LaRh2Cd20 and YbRh2Cd20 measured at 10 kOe. The zero line is shown as a guide to the eye. |
Compound | TN (K) | TC (K) | μeff (μB) | μtheo (μB) | θP (K) | μsat (μB per f.u.) | gJ × J | Bcrit (kOe) | Ref. |
---|---|---|---|---|---|---|---|---|---|
a At 90 kOe.b At 90 kOe and 500 kOe.c At 90 kOe and 400 kOe. | |||||||||
CaRh2Cd20 | χ(300 K) = −6.2(1) × 10−4 emu mol−1 | ||||||||
SrRh2Cd20 | χ(300 K) = −1.7(1) × 10−3 emu mol−1 | ||||||||
YRh2Cd20 | χ(300 K) = −1.4(1) × 10−3 emu mol−1 | ||||||||
LaRh2Cd20 | χ(300 K) = −5.9(1) × 10−4 emu mol−1 | ||||||||
CeRh2Cd20 | — | — | 2.48(1) | 2.54 | 0.1(1) | 1.22(1) | 2.14 | — | This work |
CeRh2Cd20 | 0.3 | — | 2.60–2.70 | 2.54 | −2.0 | 1.2/2b | 2.14 | — | 9 and 12 |
PrRh2Cd20 | — | — | 3.71(1) | 3.58 | −1.2(1) | 2.11(1)a | 3.2 | — | This work |
PrRh2Cd20 | — | — | — | — | — | 2–2.65/2.90–3.19c | 3.2 | — | 10 |
NdRh2Cd20 | — | — | 3.91(1) | 3.62 | −6.1(1) | 2.21(1) | 3.27 | — | This work |
SmRh2Cd20 | 4.3(1) | — | 0.78(1) | 0.85 | −2.9(1) | 0.14(1) | 0.71 | — | This work |
EuRh2Cd20 | — | 13.8(1) | 7.94(1) | 7.94 | 15.7(1) | 6.61(1)a | 7 | — | This work |
GdRh2Cd20 | 9.3(1) | — | 8.25(1) | 7.94 | −4.5(1) | 6.71(1) | 7 | — | This work |
TbRh2Cd20 | 6.6(1) | — | 10.32(1) | 9.72 | −4.0(1) | 8.00(1) | 9 | 19(1) | This work |
DyRh2Cd20 | 3.9(1) | — | 10.96(1) | 10.65 | −3.9(1) | 9.09(1)a | 10 | 10(1) | This work |
HoRh2Cd20 | — | — | 10.73(1) | 10.61 | −0.5(1) | 9.82(1)a | 10 | — | This work |
ErRh2Cd20 | — | — | 9.66(1) | 9.58 | −1.7(1) | 8.42(1) | 9 | — | This work |
TmRh2Cd20 | — | — | 7.70(1) | 7.56 | −2.5(1) | 6.17(1) | 7 | — | This work |
YbRh2Cd20 | χ(300 K) = −9.4(1) × 10−4 emu mol−1 |
CaRh2Cd20, SrRh2Cd20, YRh2Cd20, LaRh2Cd20 and YbRh2Cd20 show diamagnetic behavior at 300 K (Fig. 4) with the susceptibility values listed in Table 5. Thus, for all five samples the Pauli-paramagnetic contribution is overcompensated by the core diamagnetism. This furthermore confirms the divalent ytterbium in YbRh2Cd20 with a stable [Xe]4f14 configuration, in agreement with the plot of the cell parameters (Fig. 2).
All other compounds show paramagnetic behavior. The experimental magnetic moments (Table 5) fit with the calculated values for the free RE3+ ions.48 An exception is EuRh2Cd20 which also exhibits a stable divalent ground state ([Xe]4f7 configuration). The compounds RERh2Cd20 (with RE = Ce–Nd and Ho–Tm) show paramagnetic behavior over the whole temperature region with no sign of magnetic ordering. The experimental data of NdRh2Cd20 is shown as an example in Fig. 5. The paramagnetic behavior is also supported by the course of the magnetization isotherms. For CeRh2Cd20 we find good agreement with earlier work by Takayama et al.12 These values are listed for comparison in Table 5.
EuRh2Cd20 is the only ferromagnet in the series of RERh2Cd20 intermetallics. The precise Curie temperature of 13.8(1) K (Fig. 6) was derived from dχ/dT of a field-cooling measurement with an external flux density of 100 Oe. The paramagnetic Curie temperature of EuRh2Cd20 has a positive value of 15.7(1) K, indicating ferromagnetic interactions in the paramagnetic temperature regime. Fig. 6 (bottom) shows the magnetization isotherms. In the paramagnetic range (50 and 100 K isotherms), the magnetization increases in an almost linear manner as is usual for simple paramagnets. The isotherms in the magnetically ordered regime (2 and 10 K) show a steep increase and especially at 2 K rapid saturation. The experimental saturation magnetization at 2 K and 90 kOe of 6.61μB is close to the maximum value of 7μB according to g × J.48 We observe almost no hysteresis, classifying EuRh2Cd20 as a typical soft ferromagnet.
The susceptibility data of GdRh2Cd20 is shown in Fig. 7 as an example for an antiferromagnet with TN = 9.3(1) K. The 2.5 K magnetization isotherm increases in a linear fashion and tends towards saturation near 90 kOe. Since Gd3+ and Eu2+ are isoelectronic, the maximal magnetization is also 7μB. Also, TbRh2Cd20 and DyRh2Cd20 show antiferromagnetic transitions, however, both phases exhibit metamagnetic transitions at critical fields of 19 respectively 10 kOe.
SmRh2Cd20 shows van-Vleck paramagnetism (Fig. 8), resulting from the small energy difference between the excited J = 7/2 multiplet and the J = 5/2 ground state of the Sm3+ ion.49 For a free Sm3+ ion, this energy is calculated to be ΔE = 1550 K.49 The low effective moment of Sm3+ can be described by the antiparallel coupling of the L = 5, S = 5/2 Russel Saunders states. Stewart proposed a theory to describe the magnetism of samarium intermetallics. The expression χ(T) = χ0 + D/(T − θ) considers the Van Vleck transition between multiplet levels, the Heisenberg interactions and the induced conduction-electron polarization.50 The experimental data of the inverse susceptibility of SmRh4B4 could be described by Hamaker et al.51 by the following equation (μeff is the effective moment, θP the Weiss constant, μB the Bohr magneton, NA the Avogadro number and kB the Bolzmann constant):
The value δ is defined as δ = 7ΔE/20 in units of K and describes the energy difference between the ground and excited states. The first term represents the Curie Weiss contribution of the J = 5/2 ground state, while the second term is the temperature independent Van Vleck correction caused by the accessible J = 7/2 state.51 This description does not include crystal field splitting and intermediate coupling between the lowest energy states.52,53
The susceptibility of SmRh2Cd20 was fitted with the Hamaker equation in the temperature range of 35–300 K (Fig. 8), resulting in an μeff = 0.78μB Sm per atom, θ = –2.9(1) K and δ = 260 K. This leads to ΔE = 743 K, somewhat smaller than the predicted value of 1550 K. Comparable values have been observed for isotypic SmCo2Zn20 (ΔE = 412 K), SmRu2Zn20 (ΔE = 265 K) and SmPd2Cd20 (ΔE = 1488 K).54 Further examples are SmNiAl4Ge2 (ΔE = 706 K),55 SmOs4Sb12 (ΔE = 850 K)56 and SmRh4B4 (ΔE = 1080 K).51 SmRh2Cd20 orders antiferromagnetically at TN = 4.3 K.
Finally, we draw back to the course of the Néel temperatures within the RERh2Cd20 series. The comparatively large distances between the rare earth atoms indicate Ruderman–Kittel–Kasuya–Yosida (RKKY) interactions. Therefore, the transition temperatures should be proportional to the de Gennes factor (Fig. 9).57 The Curie temperature of EuRh2Cd20 doesn't fit into the linear behavior of the other samples, because only phases with the same ordering type should be compared. DyRh2Cd20 shows a small deviation from the linear trend, most likely indicating crystal field effects.58
![]() | ||
Fig. 9 TN of RERh2Cd20 (RE = Sm, Gd–Dy) and TC of EuRh2Cd20 versus de Gennes factor (gJ− 1)2 × J(J + 1). The linear fit was only applied to the Néel temperatures. |
![]() | ||
Fig. 10 113Cd WCPMG-MAS NMR spectra of ARh2Cd20 with A = Y, La and Yb at a MAS frequency of 12.5 kHz; fits of all spectra are shown in the ESI in Fig. S25–S28.† |
![]() | ||
Fig. 11 Static 103Rh WCPMG NMR spectra of ARh2Cd20 with A = Y, La and Yb; fits of all spectra are shown in the ESI in Fig. S29–S31.† |
![]() | ||
Fig. 12 Static 89Y WCPMG NMR spectrum of YRh2Cd20; experimental spectrum (black line) and fit (red dashed line). |
![]() | ||
Fig. 14 Static 171Yb WCPMG NMR spectrum of YbRh2Cd20; experimental spectrum (black line) and fit (red dashed line). |
Compound | Nucleus | Wyckoff site | δiso/ppm | ζ/ppm | η |
---|---|---|---|---|---|
YRh2Cd20 | 113Cd | 16c | 3350 | — | — |
113Cd | 96g | 2213 | −554 | 0.5 | |
113Cd | 48f | 1387 | −1169 | 0.0 | |
89Y | 8a | 2756 | — | — | |
103Rh | 16d | −9173 | 1075 | 0.0 | |
LaRh2Cd20 | 113Cd | 16c | 4750 | — | — |
113Cd | 96g | 2247 | −778 | 0.7 | |
113Cd | 48f | 1313 | −1263 | 0.0 | |
103Rh | 16d | −8937 | 1210 | 0.0 | |
YbRh2Cd20 | 113Cd | 16c | 4775 | — | — |
113Cd | 96g | 2671 | −367 | 0.6 | |
113Cd | 48f | 1600 | −974 | 0.4 | |
171Yb | 8a | 7546 | — | — | |
103Rh | 16d | −8880 | 1870 | 0.0 | |
Compound | Nucleus | Wyckoff site | δiso/ppm | CQ/MHz | η |
LaRh2Cd20 | 139La | 8a | 4307 | 0.60 | 0.0 |
The static 103Rh WCPMG NMR spectra are shown in Fig. 11 and can be seen to be influenced by a notable SA when compared to the few organometallic samples reported so far.66 While the change in 103Rh shift is only around 300 ppm between the samples, a pronounced difference of reduced anisotropy is observed. The reduced anisotropy increases from ζ = 1075 ppm in YRh2Cd20 to 1210 ppm in LaRh2Cd20 to 1870 ppm in YbRh2Cd20. This difference might be attributed to the change in oxidation state of the A atom from +II to +III in the case of A = Yb. It should also be noted that the Knight shift, which typically is of positive sign, shows a negative sign for 103Rh in these samples. This phenomenon is typically attributed to an additional term in the Knight shift containing a non-negligeable polarization of core electrons.67 This phenomenon has been shown in e.g., platinum metal.68
The solid-state NMR spectroscopic measurements of 89Y and 171Yb in ARh2Cd20 with A = Y and Yb show sharp resonances in accordance with the expected Wyckoff site symmetry. Both nuclei with shifts of 2756 and 7546 ppm, respectively, fall in the expected range of intermetallic compounds.69,70 For 139La, instead of a singular resonance due to cubic site symmetry, a quadrupolar pattern can be observed and fitting this resonance with one component did not yield satisfactory results. Instead, a fit that includes two components of equal shift but one constrained to a pure Voigt shape was necessary. This observation can be explained through defects in the structure breaking local symmetry, resulting in a non-zero EFG for 139La (I = 7/2).71 As the line width of the observable NMR resonance for I > 1/2 in the solid state is governed by the magnitude of the quadrupolar coupling constant CQ = (eQ/h)Vzz with Vzz being the z-component of the EFG tensor, this effect gets significantly amplified by the absolute magnitude of the quadrupolar moment Q for the probed quadrupolar nucleus (Q = 20.6 fm2 for 139La).71 In contrast, since 89Y and 171Yb are both I = 1/2 nuclei, such a break of local symmetry might not suffice to manifest itself in the SA (magnitude proportional to the applied magnetic field) and therefore is likely only detectable via the more sensitive quadrupolar coupling induced by stronger EFGs.
![]() | ||
Fig. 15 Experimental (data points) and simulated (red line) 151Eu Mössbauer spectrum of EuRh2Cd20 measured at 78 K. |
It is worthwhile to compare the isomer shift of EuRh2Cd20 with the recently reported series of the aluminides EuT2Al20 with T = Ti, V, Nb, Ta, Cr, Mo and W.72 The latter series shows a small range of the isomer shifts from −8.34 to −8.98 mm s−1. The distinctly more negative isomer shift of EuRh2Cd20 points to increasing ionic bonding contributions.73,74
The crystal structure refinement data has been deposited at the Cambridge center.
All other data is documented in the main manuscript and in the ESI.†
For any further requests please contact the corresponding authors.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4dt03523b |
This journal is © The Royal Society of Chemistry 2025 |