Skip to content
Publicly Available Published by De Gruyter May 1, 2015

The stannides RE3Au6Sn5 (RE = La, Ce, Pr, Nd, Sm) – synthesis, structure, magnetic properties and 119Sn Mössbauer spectroscopy

  • Thomas Fickenscher , Ute Ch. Rodewald , Oliver Niehaus , Birgit Gerke , Sandra Haverkamp , Hellmut Eckert and Rainer Pöttgen EMAIL logo

Abstract

The Ce3Pd6Sb5-type rare earth stannides RE3Au6Sn5 (RE= La, Ce, Pr, Nd, Sm) were synthesized by arc-melting of the elements and subsequent annealing in open tantalum crucibles within sealed evacuated silica ampoules. The polycrystalline samples were studied by powder X-ray diffraction. The structures of three crystals were refined from single crystal X-ray diffractometer data: Pmmn, a= 1360.3(9), b= 455.9(2), c= 1023.6(4) pm, wR2 = 0.0275, 1069 F2 values, 48 variables for Ce3Au6Sn5, a= 1352.4(4), b= 455.1(1), c= 1023.7(3) pm, wR2 = 0.0367, 1160 F2 values, 48 variables for Nd3Au6Sn5, and a= 1339.8(2), b= 452.80(7), c= 1012.4(2) pm, wR2 = 0.1204, 1040 F2 values, 49 variables for Sm3Au5.59(2)Sn5.41(2). One of the gold sites of the samarium compound shows a significant degree of Au/Sn mixing. The RE3Au6Sn5 structures are composed of three-dimensional [Au6Sn5] polyanionic networks with the two crystallographically independent rare earth atoms in larger cages, i.e., RE1@Au10Sn6 and RE2@Au8Sn8. The [Au6Sn5] network is stabilized by Au–Sn (266–320 pm), Au–Au (284–301 pm) as well as Sn–Sn (320 pm; distances given for the cerium compound) interactions. Temperature-dependent magnetic susceptibility measurements reveal an antiferromagnetic ordering only for Sm3Au6Sn5, while the other compounds exhibit Curie–Weiss paramagnetism. 119Sn Mössbauer spectroscopy shows resonances in the typical range for intermetallic tin compounds where tin takes part in the polyanionic network [isomer shifts between 1.73(1) and 2.28(1) mm·s−1]. With the help of theoretical electric field gradient calculations using the WIEN2k code it was possible to resolve the spectroscopic contributions of all three crystallographically independent atomic tin sites in the 119Sn spectra of RE3Au6Sn5 (RE= La, Ce, Pr, Nd, Sm).

1 Introduction

Many alkaline earth (AE) and rare earth (RE) transition metal (T) tetrelides AET2tetr2 and RET2tetr2 (tetr= Si, Ge, Sn) crystallize with the ThCr2Si2 or CaBe2Ge2 structures which are site occupancy or superstructure variants of the BaAl4 type [1, 2]. Especially the silicides and germanides have been studied intensely with respect to their magnetic and electrical properties. Several of the magnetic structures have been determined from single crystal and powder neutron diffraction data [3].

Compared to the huge family of silicides and germanides, the number of stannides with compositions of or close to RET2Sn2 is small [2, 4]. Only few of these phases have been structurally characterized from single crystal diffraction data and furthermore, some of these phases show defect formation on the transition metal or tin substructures. Typical examples are CeCu2Sn1.9 [5] and CePt1.88Sn1.89 [6]. An ordering of such defects along with small atomic displacements leads to superstructure formation. This has been observed for the series of RE3Co6Sn5 (RE= Y, Nd, Sm, Gd–Tm) [7–9] and Yb3Cu6Sn5 [10] stannides. Superstructures that derive from the CaBe2Ge2 type (P branch of BaAl4 superstructures) are Ce3Pd6Sb5 [11], Ce8Rh17Sb14 [12] and Yb5Cu11Sn8 [10, 13].

The stannides RE3Au6Sn5 (RE= La–Nd, Sm) [14–16] crystallize with the Ce3Pd6Sb5 type. A first hint for such phases was obtained by Boulet et al. [14]. They observed tripling of a ThCr2Si2 subcell along the b axis for a sample of the starting composition CeAu2Sn2, which apparently was Ce3Au6Sn5. Magnetic data of this sample showed a slightly reduced magnetic moment of 2.46 μB per cerium atom and no magnetic ordering down to 2 K. The praseodymium compound was discovered during a study of the isothermal section of the Pr–Au–Sn system at 870 K [15]. Its structure was refined on the basis of single crystal data. Lattice parameters of La3Au6Sn5, Nd3Au6Sn5 and Sm3Au6Sn5 were published recently [16].

In continuation of our structure–property relationship studies of the structurally closely related stannides SrAu2Sn2 [17] and EuAu2Sn2 [18], we were interested in the magnetic behavior and the 119Sn Mössbauer spectroscopic characterization of the RE3Au6Sn5 series. Herein we report on the magnetic and spectroscopic characterization of these RE3Au6Sn5 stannides and give evidence for homogeneity ranges and Au/Sn mixing within the [Au6Sn5] polyanionic networks.

2 Experimental

2.1 Synthesis

Starting materials for the syntheses of the RE3Au6Sn5 phases were pieces of sublimed rare earth elements (smart elements), gold shots (Agosi, 99.9 %) and tin granules (Merck, >99.99 %). The moisture sensitive rare earth pieces were kept under dry argon in Schlenk tubes prior to the reactions. Argon was purified over titanium sponge (900 K), silica gel and molecular sieves.

Samples with lanthanum and cerium were prepared with the ideal elemental mixture of 3RE: 6Au: 5Sn. The elements were reacted in an arc-melting furnace [19] under an argon pressure of ca. 700 mbar and re-melted three times on each side to ensure homogeneity. The product buttons were then placed in small open tantalum crucibles which were sealed in evacuated silica tubes. The ampoules were annealed at 870 K for 3 weeks followed by quenching to room temperature. The praseodymium, neodymium, and samarium containing samples were obtained from batches of the starting compositions 21RE: 44Au: 35Sn (to avoid formation of the REAuSn and/or RE2Au5Sn2 phases as by-products), arc-melted and annealed in the same sequence. The resulting samples are brittle and exhibit silvery luster while ground powders appear light grey. The samples are stable in air over months.

When searching for isotypic compounds with the heavier rare earth elements we observed the new phases RE2Au5Sn2 [20] with an ordered arrangement of the Au7Ga2 type [21].

2.2 X-ray diffraction

The polycrystalline RE3Au6Sn5 samples were studied by X-ray powder diffraction using the Guinier technique: imaging plate detector, Fujifilm BAS-1800, CuKα1 radiation and α-quartz (a = 491.30, c = 540.46 pm) as an internal standard. The orthorhombic lattice parameters (Table 1) were deduced from least-squares refinements of the Guinier powder data. Proper indexing was ensured by comparison of the experimental patterns with calculated ones [22].

Table 1

Lattice parameters (Guinier powder data) of the orthorhombic stannides RE3Au6Sn5.

Compounda (pm)b (pm)c (pm)V (nm3)References
La3Au6Sn51365(2)458.0(4)1026.6(9)0.6418this work
La3Au6Sn51371.6(3)461.8(1)1022.5(3)0.6477[16]
Ce3Au6Sn51360.3(9)455.9(2)1023.6(4)0.6348this work
Ce3Au6Sn51358.6(2)457.47(9)1021.4(2)0.6348[16]
‘CeAu2Sn21359.0(1)457.5(1)1017.6(1)0.6327[14]
Pr3Au6Sn51353.8(7)455.9(2)1021.2(5)0.6303this work
Pr3Au6Sn51353.4(3)455.7(1)1021.5(3)0.6300[16]
Pr3Au6Sn51354.8(5)455.6(1)1020.8(3)0.6301[15]
Nd3Au6Sn51352.4(4)455.1(1)1023.7(3)0.6300this work
Nd3Au6Sn51350.1(2)455.30(6)1014.6(2)0.6237[16]
Sm3Au6Sn51343.7(4)453.7(2)1015.2(4)0.6190this work
Sm3Au6Sn51338.9(5)452.5(2)1012.4(3)0.6134[16]

Standard deviations are given in parentheses.

Crystal fragments were selected from the crushed samples with cerium, neodymium, and samarium as the rare earth component. The crystals were characterized on a Buerger camera (using white Mo radiation) to check their quality. Intensity data of a Nd3Au6Sn5 and a Sm3Au5.59(2)Sn5.41(2) crystal were collected at room temperature on a Stoe IPDS-II image plate system (graphite monochromatized Mo radiation; λ = 71.073 pm) in oscillation mode. The Ce3Au6Sn5 data set was measured on a Stoe STADIVARI diffractometer equipped with a Mo micro focus source and a Pilatus detection system and scaled subsequently with reference to the Gaussian-shaped profile of the X-ray source. Numerical absorption corrections (along with scaling for the STADIVARI data set) were applied to the data sets. Details about the data collections and the crystallographic parameters are summarized in Table 2.

Table 2

Crystal data and structure refinement results for Ce3Au6Sn5, Nd3Au6Sn5 and Sm3Au5.59(2)Sn5.41(2) with Ce3Pd6Sb5 structure type; space group Pmmn, Z = 2; Pearson code oP28.

Empirical formulaCe3Au6Sn5Nd3Au6Sn5Sm3Au5.59(2)Sn5.41(2)
Formula weight, g mol−12195.612207.972193.81
Lattice parameters frompowder datapowder datasingle crystal data
a, pm1360.3(9)1352.4(4)1339.8(2)
b, pm455.9(2)455.1(1)452.80(7)
c, pm1023.6(4)1023.7(3)1012.4(2)
V, nm30.63480.63000. 6142
Calculated density, g cm−311.4911.6411.86
Absorption coefficient, mm−189.191.391.2
DiffractometerSTADIVARIIPDS2IPDS2
DetectorPilatus 100KImage PlateImage Plate
Detector distance, mm608080
Exposure time, min1205
ω range/increment, deg0–180/0.30–180/1.00–180/1.00
Integr. param. A/B/EMS6.2/−5.0/0.01313.2/3.1/0.01213.0/2.7/0.018
F(000), e179618081796
Crystal size, μm35 × 20 × 2010 × 20 × 2520 × 20 × 50
Transm. ratio (max/min)5.572.902.52
θ range, deg2–302–312–30
Range in hkl±19, ±6, ±14±19, ±6, ±14±18, ±6, ±14
Total no. reflections689171575893
Independent reflections/Rint1069/0.03171160/0.08351040/0.1136
Reflections with I ≥ 2 σ(I)/Rσ891/0.0242721/0.0800797/0.0622
Data/parameters1069/481160/481040/49
Goodness-of-fit on F20.8460.8170.998
R1/wR2 for I ≥ 2 σ(I)0.0155/0.02700.0363/0.03100.0470/0.1141
R1/wR2 for all data0.0207/0.02750.0846/0.03670.0630/0.1204
Extinction coefficient0.00080(2)0.00045(2)0.0018(2)
Largest diff. peak/hole, e Å−31.37/−1.312.95/−3.068.10/−7.04

The three diffractometer data sets showed primitive orthorhombic lattices and the systematic extinctions were compatible with space group Pmmn. Since isotypism with Ce3Pd6Sb5 [11] was already evident from the X-ray powder data, we used the atomic parameters of the antimonide as starting values. The structures were then refined with anisotropic displacement parameters for all atoms with Shelxl-97 (full-matrix least-squares on Fo2) [23, 24]. As a check for the correct site assignments and/or possible mixed occupancies, all occupancy parameters were refined in separate series of least-squares cycles. For Ce3Au6Sn5 and Nd3Au6Sn5 all sites were fully occupied within two standard deviations and the ideal compositions were assumed again in the following cycles. The samarium-based crystal showed too low scattering power for the 2b Au2 site, indicating mixing with tin. This mixed occupancy was refined as a least-squares variable in the following cycles, leading to the composition Sm3Au5.59(2)Sn5.41(2) for the investigated crystal. Final difference Fourier syntheses revealed no residual peaks. The refined atomic positions, displacement parameters, and interatomic distances (exemplarily for Ce3Au6Sn5) are given in Tables 3 and 4.

Table 3

Atomic coordinates and isotropic displacement parameters (pm2) of Ce3Au6Sn5, Nd3Au6Sn5, and Sm3Au5.59(2)Sn5.41(2).

AtomWyck.Occ.xyzUeq
Ce3Au6Sn5
 Ce12a1001/41/40.75535(7)160(1)
 Ce24f1000.56448(3)1/40.75432(6)153(1)
 Au14f1000.39412(3)1/40.02322(3)196(1)
 Au22b1001/43/40.03606(6)212(1)
 Au34f1000.58779(3)1/40.38037(4)204(1)
 Au42b1001/43/40.44580(6)273(1)
 Sn14f1000.39097(5)1/40.47599(6)182(1)
 Sn24f1000.58620(4)1/40.12090(6)167(1)
 Sn32a1001/41/40.22579(8)197(2)
Nd3Au6Sn5
 Nd12a1001/41/40.7554(2)90(2)
 Nd24f1000.56244(5)1/40.75462(18)88(2)
 Au14f1000.39273(8)1/40.02594(10)120(2)
 Au22b1001/43/40.03878(17)137(3)
 Au34f1000.59124(8)1/40.37896(10)139(2)
 Au42b1001/43/40.43640(17)185(4)
 Sn14f1000.39312(15)1/40.47566(16)111(4)
 Sn24f1000.58662(13)1/40.11986(17)104(4)
 Sn32a1001/41/40.2304(2)87(4)
Sm3Au5.59(2)Sn5.41(2)
 Sm12a1001/41/40.75342(15)107(3)
 Sm24f1000.55799(8)1/40.75619(10)104(3)
 Au14f1000.39373(6)1/40.02813(8)132(3)
 Au2/Sn42b59(2)/41(2)1/43/40.04545(16)154(6)
 Au34f1000.59505(7)1/40.37819(9)159(3)
 Au42b1001/43/40.42588(16)244(4)
 Sn14f1000.39457(11)1/40.47351(14)114(3)
 Sn24f1000.58987(13)1/40.11837(16)180(4)
 Sn32a1001/41/40.2309(2)143(5)

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

Table 4

Interatomic distances (pm), calculated with the powder lattice parameters of Ce3Au6Sn5.

Ce1:2Au1337.1Au3:1Sn2265.6
4Sn2343.02Sn1272.8
2Sn1344.31Au4283.5
4Au3346.31Sn1285.1
2Au2366.82Ce2337.5
2Au4390.32Ce1346.3
Ce2:1Au4325.01Ce2384.1
2Au1327.1Au4:2Au3283.5
1Au2331.34Sn1299.5
2Sn2332.12Sn3320.4
2Sn1333.52Ce2325.0
2Au3337.52Ce1390.3
2Sn3340.7Sn1:2Au3272.8
1Au1359.81Au3285.1
1Sn1370.02Au4299.5
1Sn2376.41Sn3319.9
1Au3384.12Ce2333.5
Au1:2Sn2272.81Ce1344.3
1Sn2279.81Ce2370.0
1Sn3285.4Sn2:1Au3265.6
2Au2300.92Au1272.8
2Ce2327.11Au2274.7
1Ce1337.11Au1279.8
1Ce2359.82Ce2332.1
Au2:2Sn2274.72Ce1343.0
2Sn3299.51Ce2376.4
4Au1300.9Sn3:2Au1285.4
2Ce2331.32Au2299.5
2Ce1366.82Sn1319.9
2Au4320.4
4Ce2340.7

All distances within the first coordination spheres are listed. Standard deviations are all equal or <0.2 pm.

Further details of the crystal structure investigation may be obtained from Fachinformationszentrum Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany (fax: +49-7247-808-666; e-mail: crysdata@fiz-karlsruhe.de, http://www.fiz-karlsruhe.de/request_for_deposited_data.html) on quoting the deposition number CSD-428665 (Ce3Au6Sn5), CSD-428666 (Nd3Au6Sn5) and CSD-428667 (Sm3Au5.59Sn5.41).

2.3 EDX data

Semiquantitative EDX analyses of the single crystals studied on the diffractometers were carried out in variable pressure mode with a Zeiss EVO® MA10 scanning electron microscope with the rare earth trifluorides, gold, and tin as standards. The experimentally observed compositions were close to the ideal ones. No impurity elements were detected.

2.4 Physical property studies

The magnetic and heat capacity measurements were carried out on a Quantum Design Physical Property Measurement System (PPMS) using dc-MS and heat capacity options, respectively. Dc magnetic measurements were performed using the VSM (Vibrating Sample Magnetometer) option. For these measurements approximately 30 mg of the crushed samples were fixed to the sample holder rod by kapton foil. For the heat capacity measurement a piece of Sm3Au6Sn5 (36.225 mg) was fixed to a pre-calibrated heat capacity puck using Apiezon N grease. Magnetic investigations were performed in the temperature range of 2.5–305 K with magnetic flux densities up to 80 kOe (1 kOe = 7.96 × 104 A m−1). Heat capacity measurements were done in the temperature range of 2.1–310 K.

2.5 119Sn Mössbauer spectroscopy

A Ca119mSnO3 source was used for the Mössbauer spectroscopic experiments. The measurements were carried out at ambient temperature, in a nitrogen bath cryostat in case of 78 K measurements and in a helium flow cryostat to reach temperatures around 5 K. The Mössbauer source was kept at room temperature. A palladium foil of 0.05 mm thickness was used to reduce the tin K X-rays concurrently emitted by this source. The samples were enclosed in small PMMA containers at a thickness corresponding to about 10 mg Sn per cm2. Fitting of the data was done by using the Normos-90 program package [25].

2.6 Field gradient calculations

Theoretical electric field gradient calculations were conducted using the WIEN2k code, a full-potential all electron method based on the LAPW+LO method [26]. SCF calculations were done with Rmt parameters of 2.38 atomic units for Sn and 2.5 atomic units for the rest. Separation energies between the core and valence states were set to −6 Ry. The plane wave cutoff parameter Rmtmin×Kmax was optimized in steps of 0.5 units within the range of 5.50–7.00. For Ce3Au6Sn5 a value of 6.50 was used and for Nd3Au6Sn5 a value of 7.00. For describing the first Brillouin zone, 3 k-points were used initially, and this was successively increased up to 400 k-points. 16 816 (Nd3Au6Sn5) and 16 930 (Ce3Au6Sn5) plane waves were used to describe the electronic state of the crystal structures. The nuclear quadrupole moment Q= 13.2(1) fm2 used to calculate the quadrupolar coupling constant CQ for the first exited state (I = 3/2) of the 119Sn nuclei was obtained by Barone et al. [27].

3 Results and discussion

3.1 Crystal chemistry

The stannides RE3Au6Sn5 (RE= La, Ce, Pr, Nd, Sm) crystallize with the orthorhombic Ce3Pd6Sb5 type [11]. The present single crystal data fully confirm the X-ray powder data published recently [14–16]. In the series of antimonides RE3Pd6Sb5 also the representatives with RE= Pr, Nd and Sm [28] exist. Their structures can be considered as CaBe2Ge2 superstructures with ordered vacancies [13].

Since our investigation mainly focuses on the magnetic and 119Sn Mössbauer spectroscopic characterization of the stannides RE3Au6Sn5 (RE= La, Ce, Pr, Nd, Sm) and since the crystal chemistry of the prototype has been discussed in detail [11, 28], herein we only briefly discuss the structural peculiarities of Ce3Au6Sn5.

A view of the Ce3Au6Sn5 structure approximately along the b axis is presented in Fig. 1. Comparison of the positional parameters of the RE3Au6Sn5 stannides with those of Ce3Pd6Sb5 shows one distinct difference. The 2a Sn3 sites of the stannides all have higher z values than the antimonides [11, 28]. This means a change in the coordination sphere of the Sn3 atoms as compared to Sb3 in Ce3Pd6Sb5 [11]. Consequently one should call these structures rather isopointal [29, 30] than isotypic. The [Au6Sn5] polyanion is three-dimensional in Ce3Au6Sn5, while in the Ce3Pd6Sb5 structure, there is no bonding of Sb3 to the adjacent layer (see Fig. 2 in [13]).

Fig. 1: View of the Ce3Au6Sn5 structure approximately along the b axis. Cerium, gold, and tin atoms are drawn as medium grey, blue, and magenta circles, respectively. The three-dimensional [Au6Sn5] network is emphasized. The two crystallographically independent cerium sites are indicated.
Fig. 1:

View of the Ce3Au6Sn5 structure approximately along the b axis. Cerium, gold, and tin atoms are drawn as medium grey, blue, and magenta circles, respectively. The three-dimensional [Au6Sn5] network is emphasized. The two crystallographically independent cerium sites are indicated.

Fig. 2: Coordination of the gold and tin atoms in Ce3Au6Sn5. Cerium, gold, and tin atoms are drawn as medium grey, blue, and magenta circles, respectively. The site symmetries are indicated.
Fig. 2:

Coordination of the gold and tin atoms in Ce3Au6Sn5. Cerium, gold, and tin atoms are drawn as medium grey, blue, and magenta circles, respectively. The site symmetries are indicated.

The shortest interatomic distances in the Ce3Au6Sn5 structure occur for Au–Sn (266–320 pm). They are even slightly shorter than the sum of the covalent radii of 274 pm [31], indicating substantial Au–Sn bonding within the polyanionic network. The range of Au–Sn distances is similar to that of other ternary gold stannides, e.g., CeAuSn (280 pm) [32], Eu2Au2Sn5 (272–301 pm) [33], EuAuSn (276–310 pm) [34] or Er2Au2Sn (307 pm) [35].

Besides Au–Sn bonding, the [Au6Sn5] polyanion is further stabilized by Au–Au bonding (aurophilic interactions). Each of the four crystallographically independent gold atoms has at least one short Au–Au distance. The various Au–Au distances range from 284 to 301 pm, close to that in fcc gold (288 pm) [36]. The importance of TT bonding in such BaAl4 related structures has been discussed in detail on the basis of electronic structure calculations [37, 38].

The strong shift (positional displacement) of the Sn3 sites (as compared to the antimonides) leads to the formation of Sn3–Sn1 bonds with a bond length of 320 pm. This Sn–Sn distance is comparable to that in the β-Sn structure (4 × 302 and 2 × 318 pm) [36] and such Sn–Sn distances typically occur in tin-rich stannides [39].

Finally we need to discuss the gold-tin coloring within the three-dimensional [Au6Sn5] networks. The crystals of the cerium and the neodymium compound revealed complete Au–Sn ordering while a substantial degree of Au/Sn mixing was observed for the 2b Au2 site of the samarium compound. Also small differences in the lattice parameters (Table 1) between our data and those published previously [14–16] most likely derive from inhomogeneities in the bulk samples.

3.2 Magnetic properties

Magnetic investigations have been performed for RE3Au6Sn5 with RE= Ce, Pr, Nd, Sm and are represented in Figs. 3 and 4. Table 5 lists the determined properties and the fitting results. Except for Sm3Au6Sn5, which exhibits the typical van Vleck paramagnetism, all compounds show Curie–Weiss behavior. Fitting the reciprocal susceptibility data leads to effective magnetic moments μeff that fit very well to the theoretical values of the respective free RE3+ ions. The small effective magnetic moment of Sm3Au6Sn5 can be explained by the van Vleck paramagnetism and will be discussed below in more detail. A negative Weiss constant, which indicates dominant antiferromagnetic interactions in the paramagnetic range, could only be obtained for Sm3Au6Sn5. The Weiss constants of Ce3Au6Sn5 and Nd3Au6Sn5 are almost zero, while a positive value (θp = 11(2) K) could be observed for Pr3Au6Sn5. An antiferromagnetic transition could only be obtained for Sm3Au6Sn5. Consequently, the Weiss constants are in accordance with the determined magnetic behavior, however, Pr3Au6Sn5 shows no ferromagnetic ordering.

Fig. 3: Magnetic properties of Ce3Au6Sn5: Temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. The inset displays the magnetization isotherms at 3, 10 and 50 K.
Fig. 3:

Magnetic properties of Ce3Au6Sn5: Temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. The inset displays the magnetization isotherms at 3, 10 and 50 K.

Fig. 4: Magnetic properties of Sm3Au6Sn5: (top) temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. The inset presents the magnetic susceptibility in zero-field-cooled/field-cooled (ZFC/FC) mode at 1000 Oe in the low-temperature range; (middle) Magnetization isotherms at 3, 50 and 150 K; (bottom) Heat capacity measured in the temperature range of 2.1–300 K without an applied field. The inset shows the low-temperature area to highlight the singularity.
Fig. 4:

Magnetic properties of Sm3Au6Sn5: (top) temperature dependence of the magnetic susceptibility (χ and χ−1 data) measured at 10 kOe. The inset presents the magnetic susceptibility in zero-field-cooled/field-cooled (ZFC/FC) mode at 1000 Oe in the low-temperature range; (middle) Magnetization isotherms at 3, 50 and 150 K; (bottom) Heat capacity measured in the temperature range of 2.1–300 K without an applied field. The inset shows the low-temperature area to highlight the singularity.

Table 5

Magnetic properties of RE3Au6Sn5 (RE = Ce, Pr, Nd, Sm).

REμexp (μB/RE)μeff (μB/RE)θP (K)μsm (μB/fu)μsm(calcd) (μB/fu)TN (K)
Ce2.42(1)2.540.6(5)0.9(1)2.14
Pr3.58(1)3.5811(2)1.9(1)3.2
Nd3.54(1)3.620.3(5)1.5(1)3.3
Sm0.70(1)0.85−26(2)0.03(5)0.79.3(5)

μexp, experimental magnetic moment; μeff, effective magnetic moment; θP, paramagnetic Curie temperature; μsm, experimental saturation magnetization; μsm(calcd), calculated saturation magnetization; fu, formula unit; TN, Néel temperature.

In Fig. 3 the magnetic properties of Ce3Au6Sn5 are presented exemplarily for the compounds that show no magnetic ordering down to 2.5 K. The temperature dependence of the magnetic susceptibility and its inverse (χ and χ−1 data) measured at 10 kOe is shown, and Curie–Weiss behavior can be clearly identified. The effective magnetic moment of μeff= 2.42(1) μB per Ce atom agrees well with the theoretical value of 2.54 μB for a free Ce3+ ion. Magnetic properties reported for the composition ‘CeAu2Sn2’ (≡CeAu2Sn1.66) show an effective magnetic moment of μeff= 2.46(1) μB per Ce [14]. Most likely in that publication the magnetic properties of Ce3Au6Sn5 with small amounts of tin have been reported [14].

To ensure that no magnetic ordering takes place down to 2.5 K, low-field measurements with an external field strength of 100 Oe were performed in a zero-field and field-cooled (ZFC/FC) mode (not shown here) for all three compounds. The inset in Fig. 3 displays the magnetization isotherms of Ce3Au6Sn5 measured at 3, 10 and 50 K. All isotherms are in line with paramagnetic character as reflected by the linear field dependence of the magnetization. At higher fields almost no tendency for saturation can be observed for the 3 K isotherm, and the magnetic moment at 3 K and 80 kOe [0.9(1) μB per Ce atom] is much lower than the expected saturation magnetization of 2.14 μB according to gJ × J. Such reduced magnetization values often occur in cerium intermetallic compounds [32, 40, 41] and can be attributed to crystal field splitting of the J= 5/2 magnetic ground state of Ce3+. Similar reduced values could be obtained for the praseodymium and neodymium compounds (see Table 5).

One particularity is the van Vleck paramagnetism typically shown by Sm compounds due to the proximity of the excited-state J= 7/2 multiplet to the ground-state J= 5/2 multiplet of the Sm3+ ions. The energy difference between these states is only about 1550 K, while the other angular momentum levels are correspondingly higher [42]. It has been proved that polycrystalline samples can be described by the equation

χM(T)=NAkB[μeff23(Tθp)+μB2δ],

where μeff is the effective magnetic moment, θp is the Weiss-constant, μB is the Bohr magneton, NA is the Avogadro number and kB is the Boltzmann constant. δ is an energy scale, which is defined as δ= 7 ΔE/20. The first term represents the Curie–Weiss susceptibility of the J= 5/2 ground state, while the second part represents the van Vleck susceptibility due to the small energy difference to the J= 7/2 multiplet [43]. A more detailed description can be found in the literature ([44] and references therein).

By using this equation the data could be described very well with the fitting parameters μeff= 0.70(1) μB, θp= −26(2) K and δ= 580(5) K (see Fig. 4). The effective magnetic moment is less than the 0.845 μB of the free ion for the J= 5/2 Hund’s rule ground state of Sm3+, and δ= 580(1) K corresponds to ΔE= 1657 K. This is somewhat higher than the 1550 K predicted by Stewart [42].

A heat capacity measurement of Sm3Au6Sn5, depicted in the bottom panel of Fig. 4, confirms the existence of the magnetic ordering around 9.0(5) K. In the inset the low-temperature area is presented and no typical λ anomaly due to a broadened signal can be observed. As described above, the samarium single crystal exhibits Au/Sn mixing on one site leading to a distribution of different domains around the ideal composition resulting in a broadened signal.

3.3 119Sn Mössbauer spectroscopy

119Sn spectra of RE3Au6Sn5 (RE = La–Nd, Sm) are presented in Fig. 5 together with transmission integral fits. The corresponding fitting parameters are listed in Table 6. The spectra taken at ambient temperature could be well reproduced by a superposition of three resonances exhibiting isomer shifts in the range typical for tin in polyanionic networks of intermetallic compounds [between 1.73(1) and 2.28(1) mm·s−1] with strongly different quadrupole splittings. In order to ensure the presence of tin sites having a distinctly different electronic environment, theoretical electric field gradient calculations were carried out for the Ce and Nd compounds (Table 7). As a result one obtains two more or less similar electronic surroundings for Sn1 and Sn3. The third tin site – Sn2 – shows a significantly weaker electric field gradient. This fact does not go along with a simple comparison of the nearest neighbor geometries (Fig. 2) and the site symmetries for the different tin atoms. Keeping this observation in mind the recorded spectra can be fitted by using three resonances by constraining the transmission integral ratios of Sn1:Sn2:Sn3 to 40:40:20 according to the multiplicities of the respective tin sites. Furthermore the calculated values for CQ and ηQ can be directly converted into quadrupole splitting values (Table 7) and give a hint to the start of the fitting procedure. Finally a fit with 85 % of the calculated value for ΔEQ led to the best agreement with the measured spectra, and in order to obtain consistent results, this variable was fixed, too. In the cases of the La, Pr and Sm compounds the values for the Ce and Nd compounds were extra- or interpolated. The calculations and the experimental fits reveal that for Sn1 and Sn2 the quadrupole splittings remain independent of the nature of the rare earth ion, whereas for Sn3 the quadrupole splitting decreases systematically with decreasing cation size. Regarding the deviation of the theoretical ΔEQ value from the experimental one it should be kept in mind that the literature value of the 119Sn excited state nuclear electric quadrupole moment – which is essential for calculating CQ – has been fluctuating during the past decades between a value of 8 fm2 [45] and the currently accepted value of 13.2(1) fm2 [27] and must be considered a quantity with an unknown degree of accuracy.

Fig. 5: Experimental (data points) and simulated (continuous lines) 119Sn Mössbauer spectra of RE3Au6Sn5 (RE = La–Sm) at ambient temperature.
Fig. 5:

Experimental (data points) and simulated (continuous lines) 119Sn Mössbauer spectra of RE3Au6Sn5 (RE = La–Sm) at ambient temperature.

Table 6

Fitting parameters of 119Sn Mössbauer spectroscopic measurements of RE3Au6Sn5 (RE = La, Ce, Pr, Nd, Sm) at 298 K.

Compoundδ (mm s−1)ΔEQ (mm s−1)Γ (mm s−1)Ratio (%)
La3Au6Sn52.00(1)1.205*0.89(1)40*
1.83(1)0.435*0.71(1)40*
2.28(1)1.04*0.63(1)20*
Ce3Au6Sn52.02(1)1.21*0.88(1)40*
1.79(1)0.43*0.86(1)40*
2.26(1)1.00*0.71(2)20*
Pr3Au6Sn52.02(2)1.215*0.86(1)40*
1.78(1)0.425*0.87(1)40*
2.22(2)0.96*0.80(2)20*
Nd3Au6Sn52.02(1)1.22*0.88(1)40*
1.73(1)0.42*0.93(1)40*
2.17(1)0.92*0.80(2)20*
Sm3Au6Sn52.00(1)1.225*0.85(1)40*
1.75(1)0.415*1.00(1)40*
2.11(2)0.88*1.02(3)20*

δ, isomer shift; ΔEQ, electric quadrupole splitting; Γ, experimental line width. Parameters marked with an asterisk were kept fixed during the fitting procedure.

Table 7

Theoretically calculated values for the nuclear electric quadrupolar coupling constant CQ and electric field gradient asymmetry parameter ηQ for the 119Sn nucleus in the RE3Au6Sn5 (RE = Ce, Nd) compounds and the corresponding quadrupole splitting.

CompoundTin siteCQtheo (MHz)ηQtheoΔEQ (mm s−1)
Ce3Au6Sn5Sn150.060.721.43
Sn216.980.650.50
Sn341.020.781.17
Nd3Au6Sn5Sn150.290.641.43
Sn216.570.650.49
Sn338.040.781.09

Although the single crystal data of the samarium compound indicated substantial Au/Sn mixing (vide supra), the 119Sn spectrum of Sm3Au6Sn5 gives no hint for substantial disorder in the bulk sample, also in comparison to the other spectra. The only detail that could be mentioned is the slightly broadened experimental line width. Similar Au/Sn mixing as found in the single crystal investigation of Sm3Au6Sn5 has also been reported for other CaBe2Ge2 related stannides like RET2Sn2 (RE= La, Ce, Sm; T= Ni, Cu) [46], SrAu2Sn2 [17] and EuT2Sn2 (T= Pd, Pt, Au) [18]. In these systems the disorder had a great influence on the transmission integral ratios and could therefore be identified to be a bulk phenomenon.

In addition to room temperature experiments, measurements at 78 and 5 K were carried out for the La compound (not shown here) in order to study temperature dependent effects. No significant differences in isomer shift, quadrupole splitting and signal ratios could be detected. Based on these results, we conclude that the three Sn sites have similar Lamb-Mössbauer factors, thus validating the above analysis procedure.


Corresponding author: Rainer Pöttgen, Institut für Anorganische und Analytische Chemie, Universität Münster, Corrensstrasse 30, 48149 Münster, Germany, e-mail:

Acknowledgments

This work was supported by the Deutsche Forschungsgemeinschaft. B.G. and O.N. are indebted to the Fonds der Chemischen Industrie and the NRW Forschungsschule Molecules and Materials – A Common Design Principle for PhD fellowships.

References

[1] G. Just, P. Paufler, J. Alloys Compd. 1996, 232, 1.Search in Google Scholar

[2] P. Villars, K. Cenzual, Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2013/14), ASM International®, Materials Park, Ohio (USA) 2013.Search in Google Scholar

[3] A. Szytuła, J. Leciejewicz, Handbook of Crystal Structures and Magnetic Properties of Rare Earth Intermetallics, CRC Press, Boca Raton, Florida, 1994.Search in Google Scholar

[4] R. V. Skolozdra, in Handbook on the Physics and Chemistry of Rare Earths, (Eds.: K. A. Gschneidner Jr., L. Eyring), Vol. 24, Elsevier Science, Amsterdam, 1997, chapter 164, p. 399.10.1016/S0168-1273(97)24009-2Search in Google Scholar

[5] G. Cordier, E. Czech, H. Schäfer, P. Woll, J. Less-Common Met. 1985, 110, 327.Search in Google Scholar

[6] H.-P. Liu, M. Colarieti-Tosti, A. Broddefalk, Y. Andersson, E. Lidström, O. Eriksson, J. Alloys Compd. 2000, 306, 30.Search in Google Scholar

[7] R. Pöttgen, Z. Naturforsch. 1995, 50b, 175.Search in Google Scholar

[8] R. Pöttgen, J. Alloys Compd. 1995, 224, 14.Search in Google Scholar

[9] D. Kaczorowski, Ya. Mudryk, P. Rogl, L. Romaka, Yu. Gorelenko, J. Phys.: Condens. Matter2003, 15, 2515.10.1088/0953-8984/15/17/307Search in Google Scholar

[10] M. L. Fornasini, P. Manfrinetti, D. Mazzone, P. Riani, G. Zanicchi, J. Solid State Chem. 2004, 177, 1919.Search in Google Scholar

[11] R. A. Gordon, F. J. DiSalvo, R. Pöttgen, J. Alloys Compd. 1995, 228, 16.Search in Google Scholar

[12] I. Schellenberg, R.-D. Hoffmann, C. Schwickert, R. Pöttgen, Solid State Sci. 2011, 13, 1740.Search in Google Scholar

[13] I. Schellenberg, R.-D. Hoffmann, S. Seidel, C. Schwickert, R. Pöttgen, Z. Naturforsch. 2011, 66b, 985.Search in Google Scholar

[14] P. Boulet, D. Mazzone, H. Noël, P. Rogl, R. Ferro, J. Alloys Compd. 2001, 317–318, 350.Search in Google Scholar

[15] D. Mazzone, R. Marazza, P. Riani, G. Zanicchi, G. Cacciamani, M. L. Fornasini, P. Manfrinetti, Calphad2009, 33, 31.10.1016/j.calphad.2008.09.017Search in Google Scholar

[16] M. L. Fornasini, D. Mazzone, A. Provino, M. Michetti, D. Paudyal, K. A. Gschneidner Jr., P. Manfrinetti, Intermetallics2014, 53, 169.10.1016/j.intermet.2014.05.001Search in Google Scholar

[17] C. Schwickert, B. Gerke, R. Pöttgen, Z. Naturforsch. 2014, 69b, 767.Search in Google Scholar

[18] C. Schwickert, F. Winter, R. Pöttgen, Z. Naturforsch. 2014, 69b, 775.Search in Google Scholar

[19] R. Pöttgen, Th. Gulden, A. Simon, GIT Labor-Fachzeitschrift1999, 43, 133.Search in Google Scholar

[20] T. Fickenscher, U. C. Rodewald, R. Pöttgen, Z. Kristallogr. 2015, 230, 117.Search in Google Scholar

[21] K. Frank, J. Less-Common Met. 1971, 23, 83.Search in Google Scholar

[22] K. Yvon, W. Jeitschko, E. Parthé, J. Appl. Crystallogr.1977, 10, 73.Search in Google Scholar

[23] G. M. Sheldrick, Shelxl-97, Program for the Refinement of Crystal Structures, University of Göttingen, Göttingen (Germany) 1997.Search in Google Scholar

[24] G. M. Sheldrick, Acta Crystallogr.2008, A64, 112.Search in Google Scholar

[25] R. A. Brand, Normos Mössbauer fitting Program, Universität Duisburg, Duisburg (Germany) 2007.Search in Google Scholar

[26] P. Blaha, K. H. Schwarz, G. K. H. Madsen, D. Kvasnicka, J. Luitz, WIEN2k, An Augmented Plane Wave + Local Orbitals Program for Calculating Crystal Properties, Technical University Vienna, Vienna (Austria) 2001; ISBN 3-9501031-2.Search in Google Scholar

[27] G. Barone, R. Mastalerz, M. Reiher, R. Lindh, J. Phys. Chem. A2008, 112, 1666.10.1021/jp710388tSearch in Google Scholar PubMed

[28] A. Imre, A. Mewis, Z. Naturforsch. 2006, 61b, 17.Search in Google Scholar

[29] L. M. Gelato, E. Parthé, J. Appl. Crystallogr. 1987, 20, 139.Search in Google Scholar

[30] E. Parthé, L. M. Gelato, Acta Crystallogr. 1984, A40, 169.Search in Google Scholar

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

[32] D. Niepmann, R. Pöttgen, K. M. Poduska, F. J. DiSalvo, H. Trill, B. D. Mosel, Z. Naturforsch. 2001, 56b, 1.Search in Google Scholar

[33] D. Kußmann, R. Pöttgen, U. Ch. Rodewald, C. Rosenhahn, B. D. Mosel, G. Kotzyba, B. Künnen, Z. Naturforsch. 1999, 54b, 1155.Search in Google Scholar

[34] R. Pöttgen, R.-D. Hoffmann, R. Müllmann, B. D. Mosel, G. Kotzyba, Chem. Eur. J. 1997, 3, 1852.Search in Google Scholar

[35] R. Pöttgen, Z. Naturforsch. 1994, 49b, 1309.Search in Google Scholar

[36] J. Donohue, The Structures of the Elements, Wiley, New York, 1974.Search in Google Scholar

[37] D. Johrendt, C. Felser, O. Jepsen, O. K. Andersen, A. Mewis, J. Rouxel, J. Solid State Chem. 1997, 130, 254.Search in Google Scholar

[38] R. Pöttgen, D. Johrendt, Intermetallics – Synthesis, Structure, Function, De Gruyter, Berlin, 2014.10.1524/9783486856187Search in Google Scholar

[39] R. Pöttgen, Z. Naturforsch. 2006, 61b, 677.Search in Google Scholar

[40] D. Johrendt, G. Kotzyba, H. Trill, B. D. Mosel, H. Eckert, Th. Fickenscher, R. Pöttgen, J. Solid State Chem. 2002, 164, 201.Search in Google Scholar

[41] D. Niepmann, R. Pöttgen, B. Künnen, G. Kotzyba, C. Rosenhahn, B. D. Mosel, Chem. Mater. 1999, 11, 1597.Search in Google Scholar

[42] A. M. Stewart, Phys. Rev. B1972, 6, 1985.10.1103/PhysRevB.6.1985Search in Google Scholar

[43] H. C. Hamaker, L. D. Woolf, H. P. MacKay, Z. Fisk, M. B. Maple, Solid State Commun.1979, 32, 289.Search in Google Scholar

[44] S. Seidel, O. Niehaus, S. F. Matar, O. Janka, B. Gerke, U. Ch. Rodewald, R. Pöttgen, Z. Naturforsch. 2014, 69b, 1105.Search in Google Scholar

[45] A. J. F. Boyle, D. S. P. Bunbury, C. Edwards, Proc. Phys. Soc. 1962, 79, 416.Search in Google Scholar

[46] E. A. Görlich, K. Łątka, J. Moser, Hyperfine Int. 1989, 50, 723.Search in Google Scholar

Received: 2015-3-18
Accepted: 2015-3-24
Published Online: 2015-5-1
Published in Print: 2015-6-1

©2015 by De Gruyter

Downloaded on 8.5.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2015-0050/html
Scroll to top button