Home Antiferromagnetic ordering in the plumbide EuPdPb
Article Publicly Available

Antiferromagnetic ordering in the plumbide EuPdPb

  • Lukas Heletta , Steffen Klenner , Theresa Block and Rainer Pöttgen EMAIL logo
Published/Copyright: November 18, 2017
Become an author with De Gruyter Brill

Abstract

The plumbide EuPdPb was synthesized in polycrystalline form by reaction of the elements in a sealed niobium ampoule in a muffle furnace. The structure was refined from single-crystal X-ray diffractometer data: TiNiSi type, Pnma, a=752.4(2), b=476.0(2), c=826.8(2) pm, wR2=0.0485, 704 F2 values and 20 variables. The europium atoms are coordinated by two tilted and puckered Pd3Pb3 hexagons (280–289 pm Pd–Pb) with pronounced Eu–Pd bonding (312–339 pm). Temperature-dependent magnetic susceptibility measurements show Curie-Weiss behaviour and an experimental magnetic moment of 7.35(1) μB per Eu atom. EuPdPb orders antiferromagnetically at TN=13.8(5) K and shows a metamagnetic transition at a critical field of 15 kOe. 151Eu Mössbauer spectra confirm divalent europium (δ=–10.04(1) mm s−1) and show full magnetic hyperfine field splitting (Bhf=21.1(1) T) at 6 K.

1 Introduction

Among the huge number of ternary intermetallic rare earth (RE) compounds [1], the equiatomic RETX phases (T=transition metal; X=element of the 3rd, 4th, or 5th main group) have deeply been investigated and are the subject of several hundred publications. Numerous structural aspects and the broadly varying physical properties have been reviewed [2], [3], [4], [5], [6]. Especially the cerium [7], [8], [9], europium [10], and ytterbium [11] containing members have attracted the interest of solid state chemists and physicists when searching for materials with valence instabilities, i.e. CeIII/CeIV, EuII/EuIII and YbII/YbIII. The synthesis conditions for the RETX compounds depend on the boiling temperature of the elements. If elements with high and similar boiling temperatures are used, simple arc-melting leads to larger sample quantities. Europium and ytterbium samples cannot be synthesized in such a quasi-open setup, since substantial evaporation changes the starting composition, leading to multi-phased samples. The same holds true for magnesium, zinc, cadmium, and lead as X components. Such samples need to be prepared in sealed high-melting metal tubes.

Besides the restrictions in synthesis techniques, especially the RETPb plumbides [12] are sensitive to moisture and need to be handled under inert condition. For these reasons, only few equiatomic plumbides have been characterized with respect to their magnetic and transport properties [12]. In extension of our recent studies on the equiatomic EuTMg [13], EuTZn [14] and EuTCd [15] compounds, we have now focused on the EuTPb plumbides. So far, X-ray powder data have been reported for EuTPb with T=Mg, Zn, Pd, Ag, Cd, Au and Hg [16], [17], [18], [19], [20]. Ordering of the T and X atoms was postulated for EuMgPb and EuPdPb (TiNiSi type) as well as EuCdPb and EuHgPb (LiGaGe type), while the remaining plumbides were described with T/X statistics (KHg2 type).

Small single crystals of EuPdPb were now obtained from an induction-melted sample. Herein we report on the structure refinement and a combined study by magnetic susceptibility and 151Eu Mössbauer spectroscopic measurements.

2 Experimental

2.1 Synthesis

Starting materials for the syntheses of the EuPdPb samples were europium pieces (American Elements), palladium sheets (Agosi AG) and lead granules (ABCR). All elements had stated purities better than 99.9%. Pieces of the elements were weighed in the ideal 1:1:1 atomic ratio and arc-welded [21] in a small niobium ampoule under an argon pressure of 700 mbar. The argon was purified over titanium sponge (870 K), silica gel, and molecular sieves. The sealed ampoule was placed in a water-cooled sample chamber [22] of a high-frequency furnace (type TIG 1.5/300, Hüttinger Elektronik, Freiburg, Germany) under flowing argon and first annealed at 1670 K for about 5 min, followed by rapid cooling to 1070 K and a subsequent annealing at this temperature for 5 h. The temperature was controlled through a Sensor Therm Methis MS09 pyrometer with an accuracy of ±30 K. The polycrystalline sample could easily be separated from the metal tube. EuPdPb is slightly sensitive to moisture and was kept under dry argon in a Schlenk tube.

2.2 X-ray diffraction

The EuPdPb sample was characterized by a Guinier pattern (Enraf-Nonius FR552 camera, imaging plate detector, Fujifilm BAS-1800) with 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 a standard least-squares procedure. Correct indexing of the pattern was ensured through an intensity calculation [23]. The present data are in good agreement with the lattice parameters obtained by Sendlinger from a Rietveld refinement (a=751.9(1), b=475.8(1), c=827.1(1) pm) [18].

Table 1:

Crystal data and structure refinement for EuPdPb, TiNiSi type, space group Pnma, Z=4.

Empirical formulaEuPdPb
Formula weight, g mol−1465.6
Unit cell dimension (Guinier powder data)
a, pm752.4(2)
b, pm476.0(2)
c, pm826.8(2)
Cell volume, nm3V=0.2961
Calculated density, g cm−310.44
Crystal size, μm320×20×20
Transm. ratio (min/max)0.034/0.093
Absorption coefficient, mm−183.3
Detector distance, mm60
Exposure time, min10
ω Range/increment, deg0–180/1
Integr. param. A/B/EMS14.0/–4.0/0.03
F(000), e764
θ Range, deg3–35
Range in hkl±12, ±7, ±13
Total no. of reflections4729
Independent reflections/Rint704/0.0714
Reflections with I>2 σ(I)/Rσ629/0.0099
Data/parameters704/20
Goodness-of-fit on F21.65
R1/wR2 for I>2 σ(I)0.0214/0.0478
R1/wR2 for all data0.0251/0.0485
Extinction coefficient1710(80)
Largest diff. peak/hole, e Å−32.30/–1.58

Single-crystal fragments were selected from the crushed annealed EuPdPb sample, glued to quartz fibers using beeswax and studied on a Buerger camera (using white Mo radiation) to check their quality. Since this plumbide is slightly moisture sensitive, the crystals were additionally coated with Paratone-N® oil. An intensity data set was collected on a Stoe IPDS-II diffractometer (graphite monochromatized MoKα radiation; oscillation mode) and a numerical absorption correction was applied. Details about the data collection and the crystallographic parameters are summarized in Table 1.

2.3 Structure refinement

The EuPdPb data set showed a primitive orthorhombic lattice and the systematic extinctions were compatible with space group Pnma, confirming isotypy with the TiNiSi-type EuTX intermetallics [10]. The starting atomic parameters were deduced by the charge-flipping algorithm [24] of Superflip [25] and the structure was refined with anisotropic displacement parameters for all atoms using the Jana2006 package (full-matrix least-squares on Fo2) [26]. Separate refinement of the occupancy parameters of all sites revealed full occupancy. The final difference Fourier synthesis revealed no residual peaks. The refined atomic positions, displacement parameters, and interatomic distances are given in Tables 2 and 3 .

Table 2:

Atomic coordinates and displacement parameters (pm2) for EuPdPb.

AtomxzU11U22U33U13Ueq
Eu0.02351(5)0.68792(5)148(2)151(2)167(2)–10(1)156(1)
Pd0.28388(8)0.39395(9)188(3)145(2)175(3)6(2)169(2)
Pb0.16571(4)0.07290(4)167(1)129(1)145(1)11(1)147(1)
  1. All atoms lie on Wyckoff site 4c (x 1/4 z). Ueq is defined as one third of the trace of the orthogonalized Uij tensor; U12=U23=0.

Table 3:

Interatomic distances (pm) for EuPdPb. Standard deviations are equal or smaller than 0.2 pm.

Eu:1Pd312.2Pd:1Pb279.9Pb:1Pd279.9
2Pd326.62Pb282.82Pd282.8
1Pb335.81Pb288.61Pd288.6
2Pd338.71Eu312.21Eu335.8
2Pb340.62Eu326.62Eu340.6
1Pb344.92Eu338.71Eu344.9
2Pb346.91Eu389.92Eu346.9
1Pd389.92Pb365.2
2Eu390.0
2Eu393.0
  1. All distances of the first coordination spheres are listed.

Further details of the crystal structure investigation may be obtained from FIZ Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany (fax: +49-7247-808-666; e-mail: crysdata@fiz-karlsruhe.de) on quoting the deposition number CSD-433678.

2.4 EDX data

The EuPdPb crystal studied on the diffractometer was semiquantitatively analysed by EDX in variable pressure mode with a Zeiss EVO® MA10 scanning electron microscope. EuF3, Pd and PbF2 were used as standards. The measurement (29±3 at.% Eu:39±3 at.% Pd:32±3 at.% Pb) confirmed the ideal composition. The elevated standard deviations result from the irregular crystal surface (conchoidal fracture) and the coating with beeswax and Paratone-N® oil. No impurity elements (especially from the container material) were observed.

2.5 Magnetic characterization

The temperature dependence of the magnetic susceptibility was measured using the vibrating sample magnetometer (VSM) option of a quantum design physical property measurement system (PPMS). About 30 mg of the powdered sample was filled into a polypropylene capsule and attached to the brass sample holder. The magnetic properties were detected from temperatures of 2.5 to 300 K with magnetic field strengths of up to 80 kOe (1 kOe=7.96×104 A m−1).

2.6 Mössbauer spectroscopy

The 21.53 keV transition of 151Eu with an activity of 130 MBq (2% of the total activity of a 151Sm:EuF3 source) was used for the Mössbauer spectroscopic characterization of EuPdPb. The measurement was performed in a continuous flow cryostat system (Janis Research Co LLC) at 6 and 78 K. The source was kept at room temperature. The temperature was controlled by a resistance thermometer (±0.5 K accuracy). The sample was placed in a thin-walled PMMA container with an optimized absorber thickness according to Long et al. [27]. Fitting of the spectra was performed with the Normos-90 program system [28].

3 Crystal chemistry

The single crystal structure refinement of EuPdPb fully confirms the Rietveld powder data reported by Sendlinger [18]; however, the single crystal data are by two orders of magnitude more precise. EuPdPb crystallizes with the orthorhombic TiNiSi type structure, space group Pnma with complete palladium-lead ordering. Each palladium atom has four lead neighbors in strongly distorted tetrahedral coordination with Pd–Pb distances ranging from 280–289 pm, close to the sum of the covalent radii [29] for Pd+Pb of 282 pm. This is consistent with covalent Pd–Pb bonding within the three-dimensional [PdPb] network. A cutout of this network is presented in Fig. 1, highlighting the europium near-neighbor environment. Each europium atom is coordinated by two tilted and puckered Pd3Pb3 hexagons, where the inter-layer Pd–Pb distance of 289 pm is only slightly longer than the intra-layer distances of 280 and 283 pm, underlining the three-dimensional character of the [PdPb] network. The puckering leads to a distorted europium coordination. Of the 12 atoms from the two coordinating Pd3Pb3 hexagons, one palladium atom is significantly shifted off the first coordination sphere (390 pm Eu–Pd) as compared to the Eu–Pd distances of 312–339 pm for the five closer neighbors. Each europium atom has two europium neighbors above and below the hexagons at Eu–Eu distances of 393 pm and two further neighbors at 390 pm at the open side of the [PdPb] network. Besides intra-layer Pd–Pb bonding we also observe weaker Pb–Pb contacts. The Pb–Pb distance of 365 pm within the Pd2Pb2 rhombs is close to that in fcc lead [30] with 350 pm Pb–Pb. These bonding characteristics fit into the whole family of TiNiSi type intermetallic compounds. For further details we refer to review articles [2], [9], [31], [32], [33], [34], [35].

Fig. 1: (left) View of the EuPdPb structure approximately along the crystallographic b axes. Europium, palladium and lead atoms are drawn as medium grey, blue and red circles, respectively. (right) Coordination of the europium atoms in EuPdPb. Relevant interatomic distances are given in pm.
Fig. 1:

(left) View of the EuPdPb structure approximately along the crystallographic b axes. Europium, palladium and lead atoms are drawn as medium grey, blue and red circles, respectively. (right) Coordination of the europium atoms in EuPdPb. Relevant interatomic distances are given in pm.

Finally we compare EuPdPb with the complete series of REPdPb plumbides [36], [37], [38]. The phases with RE=Y, La–Nd, Sm and Gd–Yb crystallize with the hexagonal ZrNiAl type. Except YbPdPb, the cell volumes follow the lanthanide contraction. The positive deviation for YbPdPb is pronounced (the cell volume is even larger than that of GdPdPb) pointing towards intermediate-valent or divalent ytterbium. EuPdPb shows a stable divalent ground state (vide infra) and a switch in structure type with a cell volume per formula unit larger than that for LaPdPb. This is typically observed also in other series of equiatomic RETX intermetallics [1], [6].

For the closely related plumbides EuAgPb [19] and EuAuPb [20] only subcell data (KHg2 type) with Ag/Pb and Au/Pb statistics has been reported on the basis of powder X-ray diffraction and single-counter diffractometer data. In view of the well-resolved palladium-lead ordering observed in the present study, we have reinvestigated the EuAgPb and EuAuPb structures and indeed observed transition-metal lead ordering. The complicated modulated EuAgPb and EuAuPb structures will be reported in a forthcoming publication.

4 Magnetic properties

Figure 2 (top) shows the susceptibility (χ and χ−1 data) measured in zero-field-cooled mode of a polycrystalline EuPdPb sample at an external magnetic flux density of 10 kOe in the temperature range of 3–300 K. EuPdPb exhibits Curie Weiss behavior above ca. 100 K. Fitting the reciprocal susceptibility in the temperature range of 120–300 K leads to an experimental magnetic moment of 7.35(1) μB/Eu which is lower than the theoretical value of the free Eu2+ ion of 7.94 μB/Eu [39]. The paramagnetic Curie temperature obtained from the Curie Weiss fit is θP=23(1) K, indicating ferromagnetic correlations in the paramagnetic range.

Fig. 2: Magnetic properties of EuPdPb. (top) Temperature dependence of the magnetic susceptibility (χ and χ−1 data) at 10 kOe; (middle) zero-field/field-cooled data at 100 Oe in the low-temperature regime and (bottom) magnetization isotherms at T=3, 10, 50 and 100 K.
Fig. 2:

Magnetic properties of EuPdPb. (top) Temperature dependence of the magnetic susceptibility (χ and χ−1 data) at 10 kOe; (middle) zero-field/field-cooled data at 100 Oe in the low-temperature regime and (bottom) magnetization isotherms at T=3, 10, 50 and 100 K.

In addition, we performed a zero-field-cooled/field-cooled measurement with an applied magnetic field of 100 Oe in the temperature range from 2.5 to 100 K (Fig. 2, middle). The pronounced anomaly is the antiferromagnetic ground state of EuPdPb below a Neél temperature of TN=13.8(5) K. Trace amounts of EuO were ascertained by its characteristic Curie temperature of TC=71 K [40], [41], [42]. This anomaly is already seen in the high-field measurement (Fig. 2; top).

Magnetization isotherms (Fig. 2, bottom) were measured at 3, 10, 50 and 100 K with magnetic field strengths of up to 80 kOe. At 50 and 100 K, well above the Néel temperature of EuPdPb, we observe a linear increase of the magnetization, as expected for a paramagnetic material. The 10 K isotherm (slightly below the Néel temperature) shows a steeper increase. At 3 K the magnetization isotherm shows a clear metamagnetic transition at a critical field of 15 kOe, fully underlining the antiferromagnetic ground state. The highest observed magnetization at 3 K and 80 kOe is 6.4(1) μB/Eu, slightly smaller the theoretical value of g×J=7 μB for Eu2+ [39]. Similar saturation magnetization values were observed for many other equiatomic EuTX compounds [10].

5 151Eu Mössbauer spectroscopy

The 151Eu Mössbauer spectra of EuPdPb at 78 and 6 K are presented in Fig. 3. The corresponding fitting parameters are listed in Table 4. The 78 K spectrum shows a main spectral component at an isomer shift of δ=–10.04(1) mm s−1, subjected to quadrupole splitting of ΔEQ=3.56(9) mm s−1, a consequence of the non-cubic site symmetry around the europium atoms. The isomer shift is consistent with divalent europium. The small signal at δ=0.6(1) mm s−1 indicates an Eu(III) contribution, most likely arising from surface oxidation. This signal was included as a simple Lorentzian within the fitting procedure. The Eu(II) signal shows full magnetic hyperfine field splitting at 6 K with a hyperfine field of 21.1(1) T. Almost similar hyperfine fields of 22.2 T were observed for EuPdIn [43] and EuPdSn [44].

Fig. 3: Experimental and simulated 151Eu Mössbauer spectra of EuPdPb at 78 (top) and 6 K (bottom).
Fig. 3:

Experimental and simulated 151Eu Mössbauer spectra of EuPdPb at 78 (top) and 6 K (bottom).

Table 4:

Fitting parameters of 151Eu Mössbauer spectroscopic measurements of EuPdPb at 6 and 78 K: δ=isomer shift, ΔEQ=electric quadrupole splitting, Γ=experimental line width.

T (K)Siteδ (mm s−1)ΔEQ (mm s−1)Γ (mm s−1)Area (%)
78Eu(II)−10.04(1)3.56(9)2.28(6)92(2)
Eu(III)0.6(1)0*2.5*8(2)
6Eu(II)−9.85(2)1.1(1)2.97(4)89(2)
Eu(III)0.81(4)0*2.5(1)11(2)
  1. Parameters marked with an asterisk were kept fixed during the fitting procedure.

EuPdPb belongs to the large family of EuTX intermetallic compounds [10]. The 151Eu Mössbauer spectroscopic data have been relationized. The isomer shifts δ correlate with the valence electron count (VEC) [13], [44]. EuPdPb with VEC=16 and δ=–10.04(1) mm s−1 perfectly matches this δ vs VEC plot [13].

Acknowledgements

We thank Dr. R.-D. Hoffmann for collection of the single crystal diffractometer data.

References

[1] P. Villars, K. Cenzual, Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2017/18), ASM International®, Materials Park, Ohio (USA), 2017.Search in Google Scholar

[2] E. Parthé, L. Gelato, B. Chabot, M. Penzo, K. Cenzual, R. Gladyshevskii, TYPIX–Standardized Data and Crystal Chemical Characterization of Inorganic Structure Types, Gmelin Handbook of Inorganic and Organometallic Chemistry, 8th edition, Springer, Berlin, 1993.10.1007/978-3-662-10641-9Search in Google Scholar

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

[4] M. L. Fornasini, F. Merlo, J. Alloys Compd. 1995, 219, 63.10.1016/0925-8388(94)05010-4Search in Google Scholar

[5] A. Szytuła, Croatica Chem. Acta1999, 72, 171.Search in Google Scholar

[6] S. Gupta, K. G. Suresh, J. Alloys Compd. 2015, 618, 562.10.1016/j.jallcom.2014.08.079Search in Google Scholar

[7] R. Pöttgen, B. Chevalier, Z. Naturforsch. 2015, 70b, 289.10.1515/znb-2015-0018Search in Google Scholar

[8] R. Pöttgen, O. Janka, B. Chevalier, Z. Naturforsch. 2016, 71b, 165.10.1515/znb-2016-0013Search in Google Scholar

[9] O. Janka, O. Niehaus, R. Pöttgen, B. Chevalier, Z. Naturforsch. 2016, 71b, 737.10.1515/znb-2016-0101Search in Google Scholar

[10] R. Pöttgen, D. Johrendt, Chem. Mater. 2000, 12, 875.10.1021/cm991183vSearch in Google Scholar

[11] R. Pöttgen, D. Johrendt, D. Kußmann, in Handbook on the Physics and Chemistry of Rare Earths, Vol. 32 (Eds.: K. A. Gschneidner Jr., L. Eyring, G. H. Lander) North-Holland/Elsevier, Amsterdam, 2001, chapter 207, pp. 453–513.10.1016/S0168-1273(01)32006-8Search in Google Scholar

[12] R. Pöttgen, U. Ch. Rodewald, in Handbook on the Physics and Chemistry of Rare Earths, Vol. 38 (Eds.: K. A. Gschneider Jr., V. K. Pecharsky, J.-C. Bünzli), Elsevier, Amsterdam, 2008, chapter 237, pp. 55–103.10.1016/S0168-1273(07)38002-1Search in Google Scholar

[13] S. Stein, L. Heletta, T. Block, B. Gerke, R. Pöttgen, Solid State Sci. 2017, 67, 64.10.1016/j.solidstatesciences.2017.03.006Search in Google Scholar

[14] T. Mishra, W. Hermes, T. Harmening, M. Eul, R. Pöttgen, J. Solid State Chem. 2009, 182, 2417.10.1016/j.jssc.2009.06.034Search in Google Scholar

[15] R. Mishra, R. Pöttgen, R.-D. Hoffmann, D. Kaczorowski, H. Piotrowski, P. Mayer, C. Rosenhahn, B. D. Mosel, Z. Anorg. Allg. Chem. 2001, 627, 1283.10.1002/1521-3749(200106)627:6<1283::AID-ZAAC1283>3.0.CO;2-LSearch in Google Scholar

[16] F. Merlo, M. Pani, M. L. Fornasini, J. Less-Common Met.1991, 171, 329.10.1016/0022-5088(91)90155-WSearch in Google Scholar

[17] F. Merlo, M. Pani, M. L. Fornasini, J. Alloys Compd. 1993, 196, 145.10.1016/0925-8388(93)90585-BSearch in Google Scholar

[18] B. Sendlinger, Hochdruck-Untersuchungen an den ambivalenten Verbindungen MTX (M=Yb, Ca, Eu, Sr, Ba; T=Pd, Pt; X=Si, Ge, Sn, Pb), PhD Thesis, Universität München, München, 1993.Search in Google Scholar

[19] F. Merlo, M. Pani, M. L. Fornasini, J. Alloys Compd. 1996, 232, 289.10.1016/0925-8388(95)01952-9Search in Google Scholar

[20] P. E. Arpe, Synthese und strukturchemische Untersuchungen an ternären Plumbiden des Ytterbiums. Staatsexamensarbeit, Universität Münster, Münster, 1998.Search in Google Scholar

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

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

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

[24] L. Palatinus, Acta Crystallogr. B2013, 69, 1.10.1107/S2052519212051366Search in Google Scholar

[25] L. Palatinus, G. Chapuis, J. Appl. Crystallogr. 2007, 40, 786.10.1107/S0021889807029238Search in Google Scholar

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

[27] G. J. Long, T. E. Cranshaw, G. Longworth, Mössbauer Effect Ref. Data J. 1983, 6, 42.Search in Google Scholar

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

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

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

[31] G. Nuspl, K. Polborn, J. Evers, G. A. Landrum, R. Hoffmann, Inorg. Chem. 1996, 35, 6922.10.1021/ic9602557Search in Google Scholar

[32] G. A. Landrum, R. Hoffmann, J. Evers, H. Boysen, Inorg. Chem. 1998, 37, 5754.10.1021/ic980223eSearch in Google Scholar

[33] R.-D. Hoffmann, R. Pöttgen, Z. Kristallogr. 2001, 216, 127.10.1524/zkri.216.3.127.20327Search in Google Scholar

[34] M. D. Bojin, R. Hoffmann, Helv. Chim. Acta2003, 86, 1653.10.1002/hlca.200390140Search in Google Scholar

[35] M. D. Bojin, R. Hoffmann, Helv. Chim. Acta2003, 86, 1683.10.1002/hlca.200390141Search in Google Scholar

[36] A. Iandelli, J. Alloys Compd. 1994, 203, 137.10.1016/0925-8388(94)90724-2Search in Google Scholar

[37] R. Marazza, D. Mazzone, P. Riani, G. Zanicchi, J. Alloys Compd. 1995, 220, 241.10.1016/0925-8388(94)06022-3Search in Google Scholar

[38] W. Hermes, S. Rayaprol, R. Pöttgen, Z. Naturforsch. 2007, 62b, 901.10.1515/znb-2007-0705Search in Google Scholar

[39] H. Lueken, Magnetochemie, B. G. Teubner Stuttgart, Leipzig 1999.10.1007/978-3-322-80118-0Search in Google Scholar

[40] T. R. McGuire, M. W. Shafer, J. Appl. Phys.1964, 35, 984.10.1063/1.1713568Search in Google Scholar

[41] B. D. McWhan, P. C. Souers, G. Jura, Phys. Rev. 1966, 143, 385.10.1103/PhysRev.143.385Search in Google Scholar

[42] B. Stroka, J. Wosnitza, E. Scheer, H. von Löhneysen, W. Park, K. Fischer, Z. Phys. Condens. Matter1992, 89, 39.10.1007/BF01320827Search in Google Scholar

[43] R. Müllmann, B. D. Mosel, H. Eckert, G. Kotzyba, R. Pöttgen, J. Solid State Chem. 1998, 137, 174.10.1006/jssc.1998.7750Search in Google Scholar

[44] R. Müllmann, U. Ernet, B. D. Mosel, H. Eckert, R. K. Kremer, R.-D. Hoffmann, R. Pöttgen, J. Mater. Chem. 2001, 11, 1133.10.1039/b100055lSearch in Google Scholar

Received: 2017-10-13
Accepted: 2017-10-25
Published Online: 2017-11-18
Published in Print: 2017-12-20

©2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 22.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/znb-2017-0166/html
Scroll to top button