Skip to content
Publicly Available Published by De Gruyter April 18, 2019

Magnetic hyperfine field splitting in the Zintl phase Eu2Mg4Si3

  • Theresa Block , Ryosuke Numakura , Masashi Kosaka EMAIL logo , Shinji Michimura and Rainer Pöttgen EMAIL logo

Abstract

Eu2Mg4Si3 ≡ (2Eu2+)(4Mg2+)(3Si4−) is an electron-precise Zintl phase. Its Hf2Co4P3-type structure contains three crystallographically independent europium sites. The divalent state of europium was manifested through 151Eu Mössbauer spectroscopy. In the paramagnetic regime (T = 78 K) the isomer shifts range from −9.16 to −11.29 mm s−1. Eu2Mg4Si3 shows complex magnetic hyperfine field splitting at T = 5.7 K with a superposition of three subspectra with magnetic hyperfine fields of 5.4 (Eu2), 20.4 (Eu1) and 22.4 (Eu3) T.

1 Introduction

The alkaline earth metals form a large variety of binary electron-precise Zintl phases with the tetrel elements [1], [2], [3]. The typical textbook examples are Ca2Si, CaSi and CaSi2 with isolated Si4− ions, Si2− zig-zag chains and Si arsenic-related puckered hexagons, respectively. Due to the close proximity of the ionic radii [4] of the pairs Ca2+ (100 pm)/Yb2+ (102 pm) and Sr2+ (118 pm)/Eu2+ (117 pm), such Zintl phases are also formed with the divalent rare earth elements. Especially the europium compounds of this family are of special interest since the half-filled 4f shell of Eu2+ leads to interesting magnetic properties. A striking example is Eu3Si4 which orders ferromagnetically at T=117 K [5].

An extension of this family of binary Zintl phases is easily possible through a combination of large and small cations, offering a large structural variety. Broad phase analytical studies were conducted in the Eu-Mg-Si and Eu-Mg-Ge phase diagrams. The distinctly different radii [4] of Eu2+ (117 pm) and Mg2+ (57 pm) allow for an ordering on different Wyckoff sites. The simplest examples are the equiatomic compounds EuMgSi [6], [7] and EuMgGe [6], [8] with TiNiSi-type structure. The europium and magnesium atoms deliver four valence electrons and these tetrelides contain isolated Si4−, respectively Ge4− Zintl anions. 151Eu Mössbauer spectra have confirmed the presence of stable divalent europium in EuMgSi and EuMgGe. Magnetic susceptibility measurements showed magnetic ordering at 14 (EuMgSi), respectively 16 K (EuMgGe). Neutron powder diffraction data revealed an incommensurate magnetic structure for the silicide.

Isolated Si4− and Ge4− Zintl anions and Si38− and Ge38− units – isoelectronic with Cl2O – are found in Eu3Mg5Si5≡ [Eu2+]3[Mg2+]5[Si4−]2[Si38−] and Eu3Mg5Ge5≡[Eu2+]3[Mg2+]5 [Ge4−]2[Ge38−] [9]. The silicide shows the slightly higher magnetic ordering temperature of 19.1 K as compared to the germanide (14.2 K). Metallic Eu2−xMg2−yGe3 (x=0.1 and y=0.5) contains isolated Ge4− Zintl anions besides chains of oligomeric Ge2− ions and shows a small homogeneity range [10]. The europium magnetic moments order antiferromagnetically below TN=5 K.

The tetrelides Eu5+xMg18−xSi13 (x=2.2) [11], Eu5+xMg18−xGe13 (x=0.1) and Eu8Mg16Ge12≡Eu2Mg4Ge3 [12] are isopointal with the metal-rich phosphide Hf2Co4P3 [13] and show substantial homogeneity ranges because of mixed occupied Eu/Mg sites. The three compounds show stable divalent europium; however, the properties strongly depend on the europium concentration, e.g. a field-induced inversion of the magnetoresistive effect. The recently reported silicide Eu2Mg4Si3 [14] belongs to this family of compounds, i.e. it is isotypic with Eu2Mg4Ge3 [12] with Eu/Mg ordering. Eu2Mg4Si3 single crystals with typical dimensions of 0.2×0.2×3 mm3 were grown from a magnesium self-flux technique allowing orientation dependent magnetic susceptibility and resistivity measurements. Eu2Mg4Si3 becomes ordered antiferromagnetically at TN=9.6 K followed by a reorientation at TN=8.4 K. A metamagnetic transition is observed at a critical field of ca. 5 kOe (1 kOe=7.96×104 A m−1) in the magnetization isotherm at 1.8 K, manifesting the antiferromagnetic ground state.

To complete the property studies, herein we report on a 151Eu Mössbauer spectroscopic characterization of a polycrystalline sample of Eu2Mg4Si3.

2 Experimental

2.1 Synthesis

Single-crystalline samples of Eu2Mg4Si3 were synthesized by the Mg self-flux technique. Europium ingots (99.9%, Rare Metallic Co.), magnesium shots (99.99%, Rare Metallic Co.), and a silicon wafer (99.999%, Rare Metallic Co.) were used as starting materials. The mixture of Eu, Mg, and Si in the molar ratio of 2:30:3 was placed in a molybdenum crucible and sealed by arc-welding under a pure argon atmosphere in order to prevent evaporation of Eu and Mg. The crucible was encapsulated in an evacuated quartz ampoule. At first, the ampoule was slowly heated up to T=1470 K and subsequently cooled to 970 K with a rate of 5 K h−1 in an electric furnace. Finally, the excess Mg flux was removed by spinning the ampoule in a centrifuge. Needle-shaped single crystals of Eu2Mg4Si3 were obtained.

2.2 Mössbauer spectroscopy

For the 151Eu Mößbauer spectroscopic measurement of Eu2Mg4Si3 the 21.53 keV transition of 151Eu of a 151Sm:EuF3 source with an activity of 50 MBq was used (0.91% of the total activity; I=7/2 to I=5/2 transition). The sample was prepared in a thin-walled PMMA container (2 cm diameter) with an optimized thickness according to Long et al. [15]. The measurement was conducted in a continuous flow cryostat system (Janis Research Co LLC) at 5.7, 9 and 78 K while the source was kept at room temperature. The program WinNormos for Igor was used to fit the spectrum [16].

3 Results and discussion

3.1 Crystal chemistry

For the interpretation of the Mössbauer spectroscopic data we need a short crystal chemical description of Eu2Mg4Si3. This silicide is isotypic with the metal-rich phosphide Hf2Co4P3, space group P6̅2m [13]. The Pearson data base [3] lists 26 entries for this structure type. Besides several other metal-rich phosphides, also the non-magnetic (compounds without a permanent magnetic moment) silicides Sr2Mg4Si3 and Ba2Mg4Si3 are isotypic with Eu2Mg4Si3. Keeping the electronegativity differences in mind it is interesting to note that also the aurides Sr2In4Au3 and Eu2In4Au3 [17] and the arsenide Yb2Cu4As3 [18] adopt the same structure type.

The Eu2Mg4Si3 structure contains three crystallographically independent europium sites. Their Mg/Si coordination is presented in Fig. 1 along with the site symmetry. The Eu1 atoms have coordination number 14 (site symmetry m2m) which consists of a hexagonal prismatic coordination by two Mg3Si3 hexagons and two additional Mg1 atoms in-between. The highest coordination number (CN 15) occurs for the Eu2 atoms with the same hexagonal prismatic coordination by two Mg3Si3 hexagons and three additional Mg2 atoms. This results in 6̅.. site symmetry. The Eu3 atoms have the most irregular coordination by two Mg3Si2 pentagons with two Mg2 and one Si1 atom in-between. We will address these coordination modes again in the 151Eu Mössbauer spectroscopic section.

Fig. 1: Coordination of the europium atoms in the Eu2Mg4Si3 structure. Europium, magnesium and silicon atoms are drawn as medium grey, blue and magenta circles, respectively. Atom designations and site symmetries are indicated.
Fig. 1:

Coordination of the europium atoms in the Eu2Mg4Si3 structure. Europium, magnesium and silicon atoms are drawn as medium grey, blue and magenta circles, respectively. Atom designations and site symmetries are indicated.

Another structural feature concerns the europium substructure. The structural arrangement of the europium atoms determines the magnetic coupling. It is readily evident from Fig. 2 that several triangular arrangements appear in this peculiar europium substructure. The Eu1 atoms form triangles with 451 pm edge length. These triangles are shifted with respect to the Eu2–Eu3 substructure (437 pm Eu2–Eu3) by half the lattice parameter c. Such triangular arrangements typically lead to frustration in the magnetically ordered regimes, similar to the many Fe2P/ZrNiAl-type phases [19] or the recently reported rare earth (RE)-rich phases RE10TX3 (X=Al, Mg, Cd) [20], [21].

Fig. 2: Projection of the europium layers of the Eu2Mg4Si3 structure onto the ab plane. The three crystallographically independent europium sites at z=0 and z=1/2 are indicated. The in-plane distances are Eu1–Eu1 of 451 pm and Eu2–Eu3 of 437 pm. The shortest Eu–Eu distance in the c direction corresponds to the lattice parameter (443 pm).
Fig. 2:

Projection of the europium layers of the Eu2Mg4Si3 structure onto the ab plane. The three crystallographically independent europium sites at z=0 and z=1/2 are indicated. The in-plane distances are Eu1–Eu1 of 451 pm and Eu2–Eu3 of 437 pm. The shortest Eu–Eu distance in the c direction corresponds to the lattice parameter (443 pm).

3.2 151Eu Mössbauer spectroscopy

Figure 3 shows the 151Eu Mössbauer spectra of the Eu2Mg4Si3 sample at 78, 9 and 5.7 K along with transmission integral fits. The corresponding fitting parameters are summarized in Table 1.

Fig. 3: Experimental and simulated 151Eu Mössbauer spectra of Eu2Mg4Si3 at T=78 (top), 9 (middle) and 5.7 (bottom) K.
Fig. 3:

Experimental and simulated 151Eu Mössbauer spectra of Eu2Mg4Si3 at T=78 (top), 9 (middle) and 5.7 (bottom) K.

Table 1:

Fitting parameters of 151Eu Mössbauer-spectroscopic measurements of Eu2Mg4Si3 at T=78, 9 and 5.7 K.

T (K)Siteδ (mm·s−1)ΔEQ (mm·s−1)Γ (mm·s−1)BHf (T)Ratio (%)
78Eu1+Eu2−11.29(4)−2.6(3)2.73(8)62.5*
Eu3−9.16(2)−2.1(4)2.4(2)37.5*
9Eu1+Eu2+Eu3−10.72(1)−4.4(2)5.21(8)100.0
5.7Eu2−11.13(6)−2.1(5)2.7*5.4(1)25.0*
Eu1−10.87(5)−2.6(2)2.7*20.4(1)37.5*
Eu3−8.60(5)−2.0(2)2.7*22.4(1)37.5*
  1. δ=isomer shift, ΔEQ=electric quadrupole splitting, Γ=experimental line width, BHf=hyperfine field splitting. Parameters marked with an asterisk were kept fixed during the fitting procedure.

The 78 K spectrum looks like a singlet at first sight; however, the experimental line width parameter was too large to account for a single europium site. A reasonable fit of the spectrum was obtained with a superposition of two signals. This fit relies on the analyses of the individual europium coordination polyhedra (Fig. 1). The larger signal (62.5% intensity) with an isomer shift of −11.29 mm s−1 covers the Eu1 and Eu2 sites. The latter have six silicon neighbors and both polyhedra are very similar. In contrast we observe only five silicon neighbors for the Eu3 atoms, and their signal appears at δ=−9.16 mm s−1. The lower number of silicon neighbors removes less electron density from the Eu3 atoms. We can thus conclude that Eu1 and Eu2 show a slightly higher iconicity as compared to Eu3. In the paramagnetic range Eu1 and Eu2 are very similar, and their individual signals cannot be resolved in the 151Eu Mössbauer spectrum. The quadrupole splitting parameters (which account for the lower site symmetry) and the experimental line width parameters could be refined without constraints.

Now we directly turn to the spectrum at T=5.7 K, collected in the magnetically ordered regime. We observe a broad signal with some structure, readily indicating a complex superposition of three sub-signals with individual hyperfine fields. The spectrum was well reproduced with the following model: Eu1 and Eu3 with higher hyperfine fields of 20.4 and 22.4 T, respectively, and a much smaller hyperfine field of 5.4 T for Eu2, responsible for the large hump of the spectrum. The higher fields for Eu1 and Eu3 indicate almost full hyperfine field splitting. Similar fields have been observed in a variety of magnetically ordered europium intermetallics [22]. Due to the comparatively low resolution of the spectrum (caused by the superpositions) and correlations between the fitting parameters, we fixed the line width parameter of all three sub-signals to 2.7 mm s−1. These results are similar to those of isotypic Eu2In4Au3 [17]. The magnetically different behavior of the distinct europium sites is certainly due to magnetic frustration within the triangular arrangement of the europium atoms (Fig. 2).

Finally we discuss the spectrum at T=9 K, which was taken in between the two Néel temperatures of TN1=9.6 and TN2=8.4 K [14]. The signal is almost symmetric but significantly broadened with respect to the 78 K spectrum. Attempts to fit the spectrum with a superposition of two or three signals with different small hyperfine fields failed; a consequence of the vicinity to the magnetic ordering temperature. Usually in such temperature ranges one observes large correlation of the parameters. The 9 K signal thus corresponds to the (not resolved) envelope of the three sub-spectra and the enhanced line width and quadrupole parameter account for the distribution of small hyperfine fields on the europium sites. Again, this is similar to the 10 K spectrum of Eu2In4Au3 [17].

Acknowledgements

This work was partially supported by a Grant-in-Aid for Research (No. 24540359 and 18K03504) from the Ministry of Education, Culture, Sports, Science and Technology, Japan.

References

[1] H. Schäfer, B. Eisenmann, W. Müller, Angew. Chem.1973, 85, 742.10.1002/ange.19730851704Search in Google Scholar

[2] R. Pöttgen, D. Johrendt, Intermetallics, De Gruyter, Berlin, 2014.10.1524/9783486856187Search in Google Scholar

[3] P. Villars, K. Cenzual, Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2018/19), ASM International®, Materials Park, Ohio (USA) 2018.Search in Google Scholar

[4] R. D. Shannon, Acta Crystallogr.1976, A32, 751.10.1107/S0567739476001551Search in Google Scholar

[5] F. Weitzer, Yu. Prots, W. Schnelle, K. Hiebl, Yu. Grin, J. Solid State Chem.2004, 177, 2115.10.1016/j.jssc.2004.02.013Search in Google Scholar

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

[7] O. Niehaus, D. H. Ryan, R. Flacau, P. Lemoine, D. Chernyshov, V. Svitlyk, E. Cuervo-Reyes, A. Slabon, R. Nesper, I. Schellenberg, R. Pöttgen, J. Mater. Chem. C2015, 3, 7203.10.1039/C5TC01017ASearch in Google Scholar

[8] A. Slabon, E. Cuervo-Reyes, O. Niehaus, F. Winter, C. Mensing, R. Pöttgen, R. Nesper, Z. Anorg. Allg. Chem.2014, 640, 1861.10.1002/zaac.201400176Search in Google Scholar

[9] I. Schellenberg, M. Eul, C. Schwickert, C. M. Kubata, E. Cuervo Reyes, R. Nesper, U. Ch. Rodewald, R. Pöttgen, Z. Anorg. Allg. Chem.2012, 638, 1976.10.1002/zaac.201200114Search in Google Scholar

[10] A. Slabon, E. Cuervo-Reyes, C. Kubata, M. Wörle, C. Mensing, R. Nesper, Z. Anorg. Allg. Chem.2012, 638, 1417.10.1002/zaac.201200132Search in Google Scholar

[11] A. Slabon, C. Mensing, C. Kubata, E. Cuervo-Reyes, R. Nesper, Angew. Chem. Int. Ed.2013, 52, 2122.10.1002/anie.201209036Search in Google Scholar PubMed

[12] A. Slabon, E. Cuervo-Reyes, C. Kubata, C. Mensing, R. Nesper, Z. Anorg. Allg. Chem.2012, 638, 2020.10.1002/zaac.201200058Search in Google Scholar

[13] E. Ganglberger, Monatsh. Chem.1968, 99, 566.10.1007/BF00901205Search in Google Scholar

[14] R. Numakura, R. Iizuka, S. Michimura, S. Katano, M. Kosaka, Physica B2018, 536, 262.10.1016/j.physb.2017.11.030Search in Google Scholar

[15] G. J. Long, T. E. Cranshaw, G. Longworth, Moessbauer Eff. Ref. Data J.1983, 6, 42.Search in Google Scholar

[16] R. A. Brand, WinNormos for Igor6 (version for Igor 6.2 or above: 22/02/2017), Universität Duisburg, Duisburg (Germany) 2017.Search in Google Scholar

[17] R.-D. Hoffmann, R. Pöttgen, C. Rosenhahn, B. D. Mosel, B. Künnen, G. Kotzyba, J. Solid State Chem.1999, 145, 283.10.1006/jssc.1999.8265Search in Google Scholar

[18] R. O. Demchyna, Yu. B. Kuz’ma, U. Burkhardt, W. Schnelle, Yu. Grin, Inorg. Mater.2003, 39, 1112.10.1023/A:1027328904725Search in Google Scholar

[19] 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

[20] O. Niehaus, M. Johnscher, T. Block, B. Gerke, R. Pöttgen, Z. Naturforsch.2016, 71b, 57.10.1515/znb-2015-0145Search in Google Scholar

[21] C. Benndorf, H. Eckert, O. Janka, Dalton Trans.2017, 48, 1083.10.1039/C6DT04314CSearch in Google Scholar

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

Received: 2019-03-28
Accepted: 2019-04-10
Published Online: 2019-04-18
Published in Print: 2019-05-27

©2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2019-0055/html
Scroll to top button