Cadmium-rich intermetallic phases ARh2Cd20 – structure, magnetic behavior, 151Eu Mössbauer and 113Cd solid-state NMR spectroscopy

Lars Schumachera, Florian Schreinerb, Aylin Koldemira, Oliver Jankac, 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

Received 23rd December 2024 , Accepted 22nd January 2025

First published on 23rd January 2025


Abstract

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 Fd[3 with combining macron]m. 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 sd 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.


1. Introduction

The CeCr2Al20 structure1 is one of the important types within the family of so-called cage structures. Although its unit cell contains 184 atoms, its structure can easily be described as derivative of the cubic Laves phase structure of MgCu2. The substructure of the cerium and chromium atoms corresponds to the magnesium respectively copper sites of MgCu2. Both atom types have aluminum coordination in form of Frank–Kasper polyhedra, i.e., Ce@Al16 and Cr@Al12. The packing of these polyhedra avoids direct Ce–Ce, Ce–Cr and Cr–Cr contacts.

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.

2. Experimental

Synthesis

Starting materials for the synthesis of the ARh2Cd20 (A = Ca, Sr, Y, La–Nd, Sm–Lu) samples were calcium granules (Alfa Aesar, 99.5%), sublimed strontium ingots (Smart Elements, >99.9%), ingots of the rare earth elements (smart elements, >99.9%), rhodium heating tapes (Heraeus, >99.9%) and cadmium pellets (onyxmet, >99.9999%). The surface impurities of the moisture sensitive rare earth elements were mechanically removed, and the pieces were stored under dry argon (Westfalen, 99.998%, purified by using titanium sponge (T = 873 K), silica gel and molecular sieves) prior to use. The metals were weighed in a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]21 and sealed in tantalum ampoules under argon atmosphere (800 mbar) using an arc furnace.13 The slight excess of cadmium was used to compensate the loss due to evaporation of cadmium (high vapour pressure of cadmium) and recondensation at the top of the vertically positioned ampoules. The excess cadmium was found as a thin film at the top lids of the tantalum tubes.

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.

EDX data

The CeRh2Cd20, TbRh2Cd20 and DyRh2Cd20 single crystals were studied by EDX analyses using a Zeiss EVO® MA10 scanning electron microscope operated in variable pressure mode. CeO2, TbF3, DyF3, Rh and Cd were used as standards. The experimentally observed compositions (average of three point analyses on each crystal surface) of 5 ± 1 at% Ce: 8 ± 1 at% Rh: 87 ± 1 at% Cd and 4 ± 1 at% RE: 9 ± 1 at% Rh: 87 ± 1 at% Cd for the terbium and dysprosium containing crystals were close to the ideal one (4.3[thin space (1/6-em)]:[thin space (1/6-em)]8.6[thin space (1/6-em)]:[thin space (1/6-em)]87.0). No impurity elements (especially no tantalum) were found.

X-ray diffraction

The polycrystalline RERh2Cd20 samples were characterized by powder X-ray diffraction using an Enraf-Nonius FR552 Guinier camera (equipped with an imaging plate detector, coupled with a Fujifilm BAS–1800 readout system; CuKα1 radiation) and α-quartz (a = 491.30 and c = 540.46 pm) as an internal standard. The lattice parameters (Table 1) were refined by least-squares fits of the experimental 2θ values. Intensity calculations (Lazy Pulverix routine15) facilitated the correct assignment of the indices. The experimental and simulated powder patterns of the YbRh2Cd20 sample are shown as an example in Fig. 1. The patterns of the remaining samples are presented in the ESI (Fig. S1–S16). The calcium and strontium representatives contained small amounts of yet unknown by-products.
image file: d4dt03523b-f1.tif
Fig. 1 Calculated (top) and experimental (bottom) Guinier powder patterns (CuKα1 radiation) of the YbRh2Cd20 sample.
Table 1 Lattice parameters (Guinier powder and single crystal data) of ARh2Cd20 (A = Ca, Sr, Y, La–Nd, Sm–Lu), CeCr2Al20 type, space group Fd[3 with combining macron]m. Standard deviations are given in parentheses. The lattice parameter from ref. 12 was taken from a graph
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 θθ-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.

Table 2 Crystal data and structure refinement parameters for SrRh2Cd20, LaRh2Cd20, CeRh2Cd20, TbRh2Cd20 and DyRh2Cd20, CeCr2Al20-type, space group Fd[3 with combining macron]m, Z = 8, Stoe IPDS II diffractometer
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[thin space (1/6-em)]507 11[thin space (1/6-em)]891 11[thin space (1/6-em)]635 11[thin space (1/6-em)]552 11[thin space (1/6-em)]539
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


Structure refinements

The five data sets showed F centered cubic lattices and the systematic extinctions were compatible with the diamond space group Fd[3 with combining macron]m. The starting atomic parameters were obtained with the charge-flipping algorithm16 implemented in Superflip17 and the structures were refined on F2 with the JANA2020 software package18,19 with anisotropic displacement parameters for all sites. Separate refinements of the occupancy parameters gave no hints for deviations from the ideal compositions. The final difference Fourier synthesis revealed no significant residual peaks. The final atomic coordinates, displacement parameters and interatomic distances (exemplarily for TbRh2Cd20) are listed in Tables 3 and 4.
Table 3 Atomic coordinates and anisotropic displacement parameters (pm2) for SrRh2Cd20, LaRh2Cd20, CeRh2Cd20, TbRh2Cd20 and DyRh2Cd20. The anisotropic displacement factor exponent takes the form: −2π2[(ha*)2U11+…+2hka*b*U12]. Ueq is defined as one third of the trace of the orthogonalized Uij tensor
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)


Table 4 Interatomic distances (pm) in the structure of TbRh2Cd20. All distances of the first coordination spheres are listed. Standard deviations are all equal or less than 0.1 pm
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.

Physical property measurements

The ARh2Cd20 samples were ground to fine powders and filled into polypropylene capsules or glued to a quartz paddle using a low temperature adhesive (GE Varnish). The capsules were attached to a brass sample holder rod before being investigated using the VSM (vibrating sample magnetometer) option of a Physical Property Measurement System (PPMS2) or a Dynacool Physical Property Measurement System by Quantum Design. The magnetization M(H, T) was measured in the temperature range of 3.5–305 K or 2–305 K using fields up to 80 kOe or 90 kOe (depending on the machine). Fitting and plotting the data was done with OriginPro202420 and the graphical editing with the program CorelDraw2023.21

89Y, 103Rh, 113Cd and 139La solid-state NMR spectroscopy

Solid-state NMR spectroscopy was performed using a Bruker Avance III (ν0(1H) = 300.13 MHz) and Bruker NEO (ν0(1H) = 500.39 MHz) spectrometer. Chemical shifts of all nuclei are referenced relative to the 1H resonance of solid adamantane (δ(1H) = 1.85 ppm) or TMS in CDCl3 (5% solution v/v; δ(1H) = 0.00 ppm) using the unified chemical shift scale.22,23 Pulse calibration was performed on liquid state standards of saturated solutions of Y(NO3)3 with Fe(NO3)3 as relaxation assistant in D2O for 89Y (ν0(89Y) = 24.520 MHz, I = 1/2), CdBr2 in D2O for 113Cd (ν0(113Cd) = 66.608 MHz, I = 1/2), 0.1 M LaCl3 in D2O for 139La (ν0(139La) = 70.683 MHz, I = 7/2), and 0.171 M Yb(Cp*)2(THF)2 (ν0(171Yb) = 87.567 MHz, I = 1/2). For 103Rh (ν0(103Rh) = 15.945 MHz, I = 1/2) pulses were calibrated on a neat GeCl4 (ν0(73Ge) = 17.455 MHz, I = 9/2) standard.24 Experiments were performed using a 4 mm HX magic-angle spinning (MAS) probe for 113Cd, a 5 mm HX static probe for 89Y and 171Yb, as well as a 1.3 mm HFXY MAS probe for 139La. Samples were ground to a fine powder and diluted with MgO 1[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v. For NMR spectroscopic experiments using the 4 mm HX MAS probe, 4 mm zirconia rotors were packed with the sample being centered by PTFE tape. Samples for the 1.3 mm probe were packed between small amounts of MgO at either end of the rotor to center the sample and ensure a filled rotor. Samples for static NMR spectroscopic experiments using the 5 mm HX probe were packed in 5 mm o.d. quartz tubes between PTFE tape.

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).

Mössbauer spectroscopy

The 21.53 keV transition of 151Eu with an activity of 55 MBq (1% of the total activity of a 151Sm:EuF3 source) was used for the 151Eu Mössbauer spectroscopic experiment on EuRh2Cd20, which was performed in transmission geometry at 78 K. The source was kept at room temperature. The sample was mixed with α-quartz to ensure an even distribution within the sample holder (thin walled PMMA container with a diameter of 20 mm). Fitting of the data was performed by using the WinNormos for Igor7 program package.31 and the graphical editing with the program CorelDRAW2023.21

3. Results and discussion

Crystal chemistry

Our phase analytical work in the RE-Rh–Cd systems completes the series of RERh2Cd20 compounds. Also, the isotypic phases CaRh2Cd20 and SrRh2Cd20 were obtained. The lattice parameters are listed in Table 1 and a graphical presentation (Iandelli plot) is given in Fig. 2. So far, only few basic crystallographic data is available on the RERh2Cd20 phases. CeRh2Cd20 has partially been substituted by indium and lead and lattice parameters for samples of the solid solutions CeRh2Cd20−xInx and CeRh2Cd20−xPbx were presented only graphically.12 The corresponding lattice parameter for CeRh2Cd20 was estimated from the respective plot.
image file: d4dt03523b-f2.tif
Fig. 2 Course of the lattice parameter in the series of the cubic cadmium-rich phases RERh2Cd20.

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 F[4 with combining macron]3m, 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 Er10RhCd3[thin space (1/6-em)]40 and Gd23Rh7Cd4[thin space (1/6-em)]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.


image file: d4dt03523b-f3.tif
Fig. 3 The crystal structure of TbRh2Cd20. The terbium, rhodium and cadmium atoms are drawn as gray, blue and red circles, respectively. The MgCu2 related substructure of terbium and rhodium is shown in the middle. Terbium and rhodium are coordinated by cadmium as shown on the right-hand side of the figure with gray and blue shading. The three crystallographically independent cadmium sites are shown on the left-hand side along with their site symmetry.

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 TbCd6[thin space (1/6-em)]43 (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[thin space (1/6-em)]45 (274–300 pm) and Rh2Cd15[thin space (1/6-em)]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

Magnetic properties

The whole series of ARh2Cd20 (A = Ca, Sr, Y, La–Nd, Sm–Lu) phases was magnetically characterized. Only the LuRh2Cd20 sample was not obtained in X-ray pure form. The magnetic data of the samples with A = Ca, Sr, Y, La, Nd, Eu, Sm, Gd and Yb are shown in Fig. 4–8. The remaining plots are documented within the electronic ESI (Fig. S18–S24). The experimental results are summarized in Table 5.
image file: d4dt03523b-f4.tif
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.

image file: d4dt03523b-f5.tif
Fig. 5 Magnetic data of NdRh2Cd20: temperature dependence of magnetic susceptibilities and inverse susceptibilities measured at 10 kOe (top) and magnetization isotherms recorded at 2.5, 10 and 50 K (bottom).

image file: d4dt03523b-f6.tif
Fig. 6 Magnetic data of EuRh2Cd20: zero field cooled/field cooled (ZFC/FC) measurement at an applied field of 100 Oe (inset) and inverse susceptibilities measured at 10 kOe (top) and magnetization isotherms recorded at 2, 10, 50 and 100 K (bottom).

image file: d4dt03523b-f7.tif
Fig. 7 Magnetic data of GdRh2Cd20: zero field cooled/field cooled (ZFC/FC) measurements at an applied field of 100 Oe (inset), inverse susceptibilities measured at 10 kOe; the red line emphasizes the fit regime (top) and magnetization isotherms recorded at 2.5, 10, 50 and 100 K (bottom).

image file: d4dt03523b-f8.tif
Fig. 8 Magnetic data of SmRh2Cd20: temperature dependence of magnetic susceptibilities (inset) and inverse susceptibilities measured at 25 kOe; the red line emphasizes the fit regime (top) and magnetization isotherms recorded at 3, 10 and 50 K (bottom).
Table 5 Magnetic properties of ARh2Cd20: Néel temperature TN; Curie temperature TC, effective magnetic moment μeff, theoretical magnetic moment μtheo, paramagnetic Curie temperature θP, experimental saturation magnetization μsat, theoretical saturation magnetization gJ × J, critical field Bcrit[thin space (1/6-em)]48
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/(Tθ) 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):

image file: d4dt03523b-t1.tif

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 SmCo2Zn20E = 412 K), SmRu2Zn20E = 265 K) and SmPd2Cd20E = 1488 K).54 Further examples are SmNiAl4Ge2E = 706 K),55 SmOs4Sb12E = 850 K)56 and SmRh4B4E = 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


image file: d4dt03523b-f9.tif
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.

Solid-state NMR spectroscopic results

Within the large family of CeCr2Al20 type phases, the main solid state NMR spectroscopic studies focused on the RET2Al20 phases.59–65 Herein we present the first complete study of a series of isotypic cadmium compounds. The solid-state NMR spectroscopic characterization was performed for the samples ARh2Cd20 with A = Y, La and Yb using 113Cd WCPMG-MAS NMR, static 89Y and 103Rh WCPMG NMR, and 139La MAS NMR. Fig. 10–14 and Table 6 summarize the results. All spectra presented in Fig. 10 and 11 are additionally shown in Fig. S25 to S31 including partial fits. The 113Cd WCPMG-MAS NMR spectroscopic experiments show three distinct 113Cd signals (see Fig. 10) with Knight shifts ranging from 5000 to 0 ppm (equivalent to Knight shifts between 0.0 and 0.5%). Two of the three 113Cd signals are clearly influenced by SA and the corresponding fit parameters are summarized in Table 6. The 113Cd resonances without SA exhibit Knight shifts of 3350, 4750 and 4775 ppm for ARh2Cd20 with A = Y, La and Yb, respectively. These 113Cd resonances can be assigned to the Wyckoff site 16c based on the crystallographic site symmetry [3 with combining macron]m. For YRh2Cd20 in Fig. 10, the two spectrally broader 113Cd resonances could be fitted with the SA parameters δ = 2213 ppm, η = 0.48 and ζ = −554 ppm as well as δ = 1387 ppm, η = 0.0 and ζ = −1169 ppm. The presence of a 113Cd SA tensor with axial symmetry, i.e. η = 0.0, is surprising given the X-ray data (Table 3). In this structure, no Cd atom, except for the one at the Wyckoff 16c site, is located on a site with site symmetry of 3-fold or higher, which would imply η = 0.0. However, upon closer examination of the Cd coordination polyhedra (Fig. 3) a possible assignment to the 48f site can be made. Although the site symmetry of 2mm does not enforce any restrictions by itself, the distorted bicapped pentagonal prism might allow for sufficiently close axial symmetry, leading to η = 0. This is in line with the fact that this site shows η = 0.0 in the case of LaRh2Cd20 and η = 0.41 for YbRh2Cd20. The remaining 113Cd site in each spectrum in Fig. 10 can therefore be assigned to the 96g site. We note that the Knight shifts for all 113Cd sites exhibit a consistent pattern, characterized by a decrease in shift from A = Yb to La to Y. In addition, a distinct trend for reduced anisotropy is discernible, where, A = La exhibits the highest reduced anisotropy, while A = Yb shows the lowest.
image file: d4dt03523b-f10.tif
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.

image file: d4dt03523b-f11.tif
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.

image file: d4dt03523b-f12.tif
Fig. 12 Static 89Y WCPMG NMR spectrum of YRh2Cd20; experimental spectrum (black line) and fit (red dashed line).

image file: d4dt03523b-f13.tif
Fig. 13 Single-pulse 139La MAS NMR spectrum of LaRh2Cd20 recorded at a MAS frequency of 62.5 kHz; experimental spectrum (black line), sum of the fits (red dashed line) and partial fits (blue and green shadings) are shown; the spectrum is fitted with two overlapping components because side band intensities could not be reproduced with one site.

image file: d4dt03523b-f14.tif
Fig. 14 Static 171Yb WCPMG NMR spectrum of YbRh2Cd20; experimental spectrum (black line) and fit (red dashed line).
Table 6 NMR spectroscopic data for 89Y, 103Rh, 113Cd, 139La and 171Yb for RERh2Cd20 with RE = Y, La and Yb. Shifts are given including the isotropic shift δiso, reduced anisotropy ζ, and asymmetry η (top);30 Parameters for quadrupolar nuclei are given as isotropic shift, quadrupolar coupling constant CQ and asymmetry η (bottom)
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.

151Eu Mössbauer spectroscopy

The 151Eu Mössbauer spectrum of EuRh2Cd20 (78 K data) is shown in Fig. 15 along with a transmission integral fit. We observe a single signal at an isomer shift of δ = −10.86(1) mm s−1 and an experimental line width of Γ = 2.34(3) mm s−1. In agreement with the cubic site symmetry of the europium atoms no quadrupole splitting occurs. The isomer shift clearly proves the stable divalent ground state of europium, in accord with the plot of the lattice parameters and the susceptibility measurement.
image file: d4dt03523b-f15.tif
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

4. Conclusions

The series of ARh2Cd20 compounds (cubic CeCr2Al20 type) forms with A = Ca, Sr, Y, La–Nd and Sm–Lu. X-ray diffraction, magnetic susceptibility, Mössbauer and solid-state NMR spectroscopic data point to divalent europium and ytterbium, while the remaining phases show stable trivalent rare earth states. CaRh2Cd20, SrRh2Cd20, YRh2Cd20, LaRh2Cd20 and YbRh2Cd20 are diamagnets while the remaining phases are Curie–Weiss paramagnets. EuRh2Cd20 is a 13.8 K ferromagnet; 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). The 113Cd NMR spectra of the A = Y, La and Yb samples revealed the presence of three distinct Cd sites in accordance with X-ray diffraction, which via local symmetry considerations could be assigned to the specific Wyckoff positions. The solid-state NMR spectra of 139La suggest local defects in the vicinity of the 8a site although similar observation could not be found for A = Y or Yb. For 89Y and 171Yb, the NMR spectra could be recorded and fitted with pure Voigt line shapes showing no SA. A negative Knight shift was found for 103Rh in all measured samples hinting at strong s-d exchange correlation.

Author contributions

Lars Schumacher: synthesis, structure refinements, magnetic characterization. Florian Schreiner: solid state NMR spectroscopic characterization. Aylin Koldemir: 151Eu Mössbauer spectroscopic characterization. Oliver Janka: X-ray powder diffraction, writing. Michael Ryan Hansen: conceptualization, supervision, interpretation of the solid-state NMR spectra, writing. Rainer Pöttgen: conceptualization, supervision, project administration, writing – review & editing.

Data availability

All data of this contribution is available.

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.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We thank Dipl.-Ing. J. Kösters for the X-ray data collections. This research was funded by Universität Münster and Deutsche Forschungsgemeinschaft (INST 211/1034-1). Instrumentation and technical assistance for this work were provided by the Service Center X-ray Diffraction, with financial support from Saarland University and German Science Foundation (project number INST 256/349-1).

References

  1. P. I. Krypyakevych and O. S. Zarechnyuk, Dopov. Akad. Nauk Ukr. RSR, Ser. A, 1968, 364–367 Search PubMed.
  2. Pearson's Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2023/24), ed. P. Villars and K. Cenzual, ASM International®, Materials Park, Ohio, USA, 2023 Search PubMed.
  3. R. Pöttgen and O. Janka, Rev. Inorg. Chem., 2023, 43, 357–383 CrossRef.
  4. Z. Hiroi, A. Onosaka, Y. Okamoto, J. I. Yamaura and H. Harima, J. Phys. Soc. Jpn., 2012, 81, 124707 CrossRef.
  5. M. J. Winiarski, B. Wiendlocha, M. Sternik, P. Wiśniewski, J. R. O'Brien, D. Kaczorowski and T. Klimczuk, Phys. Rev. B, 2016, 93, 134507 CrossRef.
  6. E. D. Mun, S. Jia, S. L. Bud'ko and P. C. Canfield, Phys. Rev. B, 2012, 86, 115110 CrossRef.
  7. M. S. Torikachvili, S. Jia, E. D. Mun, S. T. Hannahs, R. C. Black, W. K. Neils, D. Martien, S. L. Bud'ko and P. C. Canfield, Proc. Natl. Acad. Sci. U. S. A., 2007, 104, 9960–9963 CrossRef CAS PubMed.
  8. Y. Shimura, T. Kitazawa, S. Tsuda, S. Bachus, Y. Tokiwa, P. Gegenwart, R. Yamamoto, Y. Yamane, I. Nishihara, K. Umeo, T. Onimaru, T. Takabatake, H. T. Hirose, N. Kikugawa, T. Terashima and S. Uji, Phys. Rev. B, 2020, 101, 241102 CrossRef CAS.
  9. Y. Hirose, H. Doto, T. Kawano, R. Settai, T. Takeuchi, T. Tahara, T. Kida, M. Hagiwara, H. Miyake, M. Tokunaga, R. Kulkarni, A. Thamizhavel and F. Honda, JPS Conf. Ser., 15aPS-28, 2016.
  10. Y. Hirose, H. Doto, T. Takeuchi, T. Kawano, R. Settai, A. Miyake and M. Tokunaga, JPS Conf. Ser., 22pPSA-32, 2017.
  11. R. Imai, M. Akatsu, Y. Nemoto, T. Goto, R. Kurihara, K. Mitsumoto, H. Doto, Y. Hirose and R. Settai, JPS Conf. Ser., 11pPSB-9, 2018.
  12. K. Takayama, Y. Hirose, T. Kawano, H. Doto, F. Honda, Y. Homma, A. Nakamura, D. Aoki, A. Thamizhavel and R. Settai, JPS Conf. Proc., 2020, vol. 30, pp. 011122.
  13. R. Pöttgen, Th. Gulden and A. Simon, GIT Labor-Fachz., 1999, 43, 133–136 Search PubMed.
  14. R. Pöttgen, A. Lang, R.-D. Hoffmann, B. Künnen, G. Kotzyba, R. Müllmann, B. D. Mosel and C. Rosenhahn, Z. Kristallogr., 1999, 214, 143–150 Search PubMed.
  15. K. Yvon, W. Jeitschko and E. Parthé, J. Appl. Crystallogr., 1977, 10, 73–74 CrossRef.
  16. L. Palatinus, Acta Crystallogr. B, 2013, 69, 1–16 CrossRef CAS PubMed.
  17. L. Palatinus and G. Chapuis, J. Appl. Crystallogr., 2007, 40, 786–790 CrossRef CAS.
  18. V. Petříček, M. Dušek and L. Palatinus, Z. Kristallogr., 2014, 229, 345–352 Search PubMed.
  19. V. Petříček, L. Palatinus, J. Plášil and M. Dušek, Z. Kristallogr., 2023, 238, 271–282 Search PubMed.
  20. OriginPro, Version 2024, OriginLab Corporation, Northampton, MA, USA, 2024 Search PubMed.
  21. CorelDRAW Graphics Suite 2023 (version 23.0.0.363), Corel Corporation, Ottawa, Ontario, Canada, 2023 Search PubMed.
  22. R. K. Harris, E. D. Becker, S. M. Cabral de Menezes, R. Goodfellow and P. Granger, Pure Appl. Chem., 2001, 73, 1795–1818 CrossRef CAS.
  23. S. Hayashi and K. Hayamizu, Bull. Chem. Soc. Jpn., 1991, 64, 685–687 CrossRef CAS.
  24. N. T. Duong, S. Viel, F. Ziarelli, P. Thureau and G. Mollica, J. Magn. Reson., 2024, 358, 107614 CrossRef CAS.
  25. R. W. Schurko and L. A. O'Dell, Chem. Phys. Lett., 2008, 464, 97–102 CrossRef.
  26. M. R. Hansen, J. Koppe and M. Bußkamp, J. Phys. Chem. A, 2021, 125, 5643–5649 CrossRef.
  27. M. R. Hansen and J. Koppe, J. Phys. Chem. A, 2020, 124, 4314–4321 CrossRef.
  28. J. D. Hunter, Comput. Sci. Eng., 2007, 9, 90–95 Search PubMed.
  29. S. G. J. van Meerten, W. M. J. Franssen and A. P. M. Kentgens, J. Magn. Reson., 2019, 301, 56–66 CrossRef CAS.
  30. R. K. Harris, E. D. Becker, S. M. C. de Menezes, P. Granger, R. E. Hoffman and K. W. Zilm, Pure Appl. Chem., 2008, 80, 59–84 CrossRef CAS.
  31. R. A. Brand, WinNormos for Igor7 (version for Igor 7.010 or above: 01/03/2020), Universität Duisburg, Duisburg (Germany), 2020 Search PubMed.
  32. F. Tappe and R. Pöttgen, Rev. Inorg. Chem., 2011, 31, 5–25 CAS.
  33. S. Niemann and W. Jeitschko, J. Solid State Chem., 1995, 114, 337–341 CrossRef CAS.
  34. P. P. Jana, Z. Kristallogr., 2017, 232, 611–617 CAS.
  35. A. E. Dwight, J. W. Downey and R. A. Conner Jr., Trans. Metall. Soc. AIME, 1966, 236, 1509–1510 CAS.
  36. F. C. Frank and J. S. Kasper, Acta Crystallogr., 1958, 11, 184–190 CrossRef CAS.
  37. F. C. Frank and J. S. Kasper, Acta Crystallogr., 1959, 12, 483–499 CrossRef CAS.
  38. F. Tappe, U. Ch. Rodewald, R.-D. Hoffmann and R. Pöttgen, Z. Naturforsch., 2011, 66b, 559–564 CrossRef.
  39. F. M. Schappacher and R. Pöttgen, Monatsh. Chem., 2008, 139, 1137–1142 CrossRef CAS.
  40. T. Block, S. Klenner, L. Heletta and R. Pöttgen, Z. Naturforsch., 2018, 73b, 35–42 CrossRef.
  41. F. Tappe and R. Pöttgen, Z. Naturforsch., 2009, 64b, 184–188 CrossRef.
  42. G. Bruzzone, M. L. Fornasini and F. Merlo, J. Less-Common Met., 1973, 30, 361–375 CrossRef CAS.
  43. S. Y. Piao, C. P. Gómez and S. Lidin, Z. Naturforsch., 2006, 60b, 644–649 CrossRef.
  44. J. Emsley, The Elements, Oxford University Press, Oxford, 1999 Search PubMed.
  45. B. Koley, S. Chatterjee and P. P. Jana, J. Solid State Chem., 2017, 246, 302–308 CrossRef CAS.
  46. W. Xie, M. Liu, Z. Wang, N.-P. Ong and R. J. Cava, Inorg. Chem., 2016, 55, 7605–7609 CrossRef CAS.
  47. J. Donohue, The Structures of the Elements, Wiley, New York, 1974 Search PubMed.
  48. H. Lueken, Magnetochemie, B. G. Teubner, Stuttgart, Leipzig, 1999 Search PubMed.
  49. J. H. Van Vleck, The Theory of Electric and Magnetic Susceptibilities, Oxford U. P., London, 1932 Search PubMed.
  50. A. M. Stewart, Phys. Rev. B, 1972, 6, 1985–1998 CrossRef CAS.
  51. H. C. Hamaker, L. D. Woolf, H. B. MacKay, Z. Fisk and M. B. Maple, Solid State Commun., 1979, 32, 289–294 CrossRef CAS.
  52. H. Zhou, S. E. Lambert, M. B. Maple, S. K. Malik and B. D. Dunlap, Phys. Rev. B: Condens. Matter Mater. Phys., 1987, 36, 594–598 CrossRef CAS.
  53. A. M. Stewart, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 11242–11246 CrossRef CAS PubMed.
  54. D. Yazici, B. D. White, P.-C. Ho, N. Kanchanavatee, K. Huang, A. J. Friedman, A. S. Wong, V. W. Burnett, N. R. Dilley and M. B. Maple, Phys. Rev. B, 2014, 90, 144406 CrossRef.
  55. M. Witt, J. Bönnighausen, F. Eustermann, A. Savourat, J. P. Scheifers, B. P. T. Fokwa, C. Doerenkamp, H. Eckert and O. Janka, Z. Naturforsch., 2020, 75b, 149–162 CrossRef.
  56. W. M. Yuhasz, N. A. Frederick, P. C. Ho, N. P. Butch, B. J. Taylor, T. A. Sayles, M. B. Maple, J. B. Betts, A. H. Lacerda, P. Rogl and G. Giester, Phys. Rev. B, 2005, 71, 104402 CrossRef.
  57. A. Szytuła and J. Leciejewicz, Handbook of Crystal Structures and Magnetic Properties of Rare Earth Intermetallics, CRC Press, Boca Raton, Florida, 1994 Search PubMed.
  58. I. R. Fisher, Z. Islam and P. C. Canfield, J. Magn. Magn. Mater., 1999, 202, 1–10 CrossRef CAS.
  59. Y. Tokunaga, H. Sakai, S. Kambe, A. Sakai, S. Nakatsuji and H. Harima, Phys. Rev. B, 2013, 88, 085124 CrossRef.
  60. J. Chi, Y. Li, W. Gou, V. Goruganti, K. D. D. Rathnayaka and J. H. Ross Jr., Phys. B, 2008, 403, 1426–1427 CrossRef CAS.
  61. T. Taniguchi, M. Yoshida, H. Takeda, M. Takigawa, M. Tsujimoto, A. Sakai, Y. Matsumoto and S. Nakatsuji, J. Phys.: Conf. Ser., 2016, 683, 012016 CrossRef.
  62. T. Kubo, H. Kotegawa, H. Tou, R. Higashinaka, A. Nakama, Y. Aoki and H. Sato, J. Phys. Soc. Jpn. Conf. Proc, 2014, 3, 012031 Search PubMed.
  63. T. Kubo, H. Kotegawa, H. Tou, R. Higashinaka, A. Nakama, Y. Aoki and H. Sato, J. Phys.: Conf. Ser., 2015, 592, 012093 CrossRef.
  64. T. Kubo, H. Kotegawa, H. Tou, R. Higashinaka, A. Nakama, Y. Aoki and H. Sato, J. Phys.: Conf. Ser., 2016, 683, 012015 CrossRef.
  65. T. Kubo, H. Kotegawa, H. Tou, H. Harima, R. Higashinaka, A. Nakama, Y. Aoki and H. Sato, J. Phys.: Conf. Ser., 2017, 807, 032006 CrossRef.
  66. S. T. Holmes, J. Schönzart, A. B. Philips, J. J. Kimball, S. Termos, A. R. Altenhof, Y. Xu, C. A. O'Keefe, J. Autschbach and R. W. Schurko, Chem. Sci., 2024, 15, 2181–2196 RSC.
  67. A. M. Clogston and V. Jaccarino, Phys. Rev., 1961, 121, 1357–1362 CrossRef CAS.
  68. M. Shimizu and A. Katsuki, J. Phys. Soc. Jpn., 1964, 19, 614–617 CrossRef CAS.
  69. C. Benndorf, S. Stein, L. Heletta, M. Kersting, H. Eckert and R. Pöttgen, Dalton Trans., 2017, 46, 250–259 RSC.
  70. J. M. Gerdes, L. Schumacher, M. R. Hansen and R. Pöttgen, Z. Naturforsch., 2024, 79b, 13–19 CrossRef.
  71. O. Janka, M. Radzieowski, T. Block, J. Koppe, M. R. Hansen and R. Pöttgen, J. Phys. Chem. C, 2021, 125, 1454–1466 CrossRef.
  72. A. Koldemir, S. Klenner and R. Pöttgen, Z. Kristallogr. – New Cryst. Struct., 2023, 238, 507–509 CrossRef CAS.
  73. F. Steudel, J. A. Johnson, C. E. Johnson and S. Schweizer, Materials, 2018, 11, 828 CrossRef PubMed.
  74. S. Engel, E. C. J. Gießelmann, R. Pöttgen and O. Janka, Rev. Inorg. Chem., 2023, 43, 571–582 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4dt03523b

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.