Home Magnetic properties of RE10TCd3 (RE = Ho, Er, Tm, Lu; T = Fe, Co, Ni, Ru) and 57Fe Mössbauer spectroscopic data of Y10FeCd3
Article Publicly Available

Magnetic properties of RE10TCd3 (RE = Ho, Er, Tm, Lu; T = Fe, Co, Ni, Ru) and 57Fe Mössbauer spectroscopic data of Y10FeCd3

  • Oliver Niehaus , Michael Johnscher , Theresa Block , Birgit Gerke and Rainer Pöttgen EMAIL logo
Published/Copyright: December 7, 2015
Become an author with De Gruyter Brill

Abstract

Fourteen X-ray-pure intermetallic compounds RE10TCd3 (RE = Ho, Er, Tm, Lu; T = Fe, Co, Ni, Ru) and Y10FeCd3 were obtained through high-frequency melting of the elements in sealed niobium tubes and subsequent annealing in a muffle furnace. They adopt the Er10FeCd3 structure, a ternary ordered version of the Co2Al5 type. Temperature-dependent magnetic susceptibility measurements show Pauli paramagnetism for the lutetium compounds Lu10FeCd3, Lu10CoCd3, and Lu10RuCd3. The RE10TCd3 phases with holmium, erbium, and thulium show Curie–Weiss paramagnetism and the experimental magnetic moments match with the free ion values of RE3+. All these compounds order antiferromagnetically. The highest Néel temperature was observed for the holmium compounds, e.g. 46.5 K for Ho10RuCd3. Some of the RE10TCd3 phases show field-induced spin reorientations. A 57Fe Mössbauer spectrum of Y10FeCd3 confirms the single crystallographic iron site.

1 Introduction

Rare earth-rich (rare earth: RE) intermetallic compounds that form in the ternary systems RE-T-X (T = electron-rich transition metal; X = element of the 3rd, 4th, or 5th main group, Mg, Zn, or Cd) are interesting candidates for magnetic and magnetocaloric materials, especially for the rare earth elements with high effective magnetic moments. Typical compositions for those materials are RE4TX and RE23T7X4 with X = Mg, Zn, Cd, In [1–4], RE15Rh5Cd2 [5], RE14T3In3 [6], RE6TX2 (X = Sb, Bi) [7] or RE6T2X (X = Ga) [8]. The common structural motifs of these phases are transition metal centered trigonal rare earth prisms that show different connectivity patterns. Some of these compounds show high magnetic ordering temperatures, e.g. TN = 92 K for Gd4NiMg [9] or TC = 150.1 K for Eu4PdMg [10].

Recently, phase analytical studies in the RET–Cd systems revealed the new family of rare earth-rich intermetallics RE10TCd3, which are formed with a large variety of rare earth and transition metals: RE = Y, Tb–Tm, Lu; T = Fe, Co, Ni, Ru, Rh, Pd [11]. These phases are ordered ternary versions of the Co2Al5 type and exhibit the same structural motif as the phases described above, i.e. a trigonal prismatic coordination of the transition metal atoms by rare earth atoms. The recently reported aluminide Er10Co1.3Al2.7 [12] is isotypic with the cadmium phases and orders antiferromagnetically at TN = 12.7 K. First magnetic data for the RE10TCd3 series are reported herein.

2 Experimental

2.1 Synthesis

Starting materials (all with stated purities better than 99.5 %) for the synthesis of the RE10TCd3 samples were pieces of the rare earth elements (Smart Elements), ruthenium (Allgemeine Gold- und Silberscheideanstalt, Pforzheim), nickel powder (Merck), cobalt and iron pieces (Alfa Aesar), and a cadmium rod (Sigma-Aldrich). The elements were weighed in the ideal atomic ratios and sealed in niobium ampoules [13] under an argon pressure of ca. 700 mbar. The argon was purified with titanium sponge (870 K), silica gel and molecular sieves. The ampoules were placed in a water-cooled sample chamber of a high-frequency furnace (Hüttinger Elektronik, Freiburg, type TIG 1.5/300) [14]. They were rapidly heated to ca. 1500 K and kept at that temperature for 2 min. The temperature was then directly reduced to 1300 K and kept for another 45 min, followed by cooling to 900 K within 30 min and further annealing at that temperature for 2 h. The temperature was controlled through a Sensor Therm Methis MS09 pyrometer with an accuracy of ±30 K. The niobium ampoules were then sealed in evacuated silica tubes and annealed at 873 K for 5 d in a muffle furnace to achieve homogeneity. Finally, the samples were separated from the niobium tubes by mechanical fragmentation. The RE10TCd3 phases are stable at air over month.

2.2 X-ray diffraction

As a purity check, all polycrystalline RE10TCd3 samples were characterized through Guinier powder patterns (Enraf-Nonius camera, type FR 552): imaging plate detector, Fujifilm BAS-1800, CuKα1 radiation and α-quartz (a = 491.30, c = 540.46 pm) as an internal standard. The experimental patterns matched with simulated ones [15], taking the positional parameters of the previously refined structures. The samples studied herein were pure on the level of powder X-ray diffraction.

2.3 Physical property measurements

Magnetic susceptibility and heat capacity investigations were performed with a Quantum Design Physical-Property-Measurement-System (PPMS) using the VSM (Vibrating Sample Magnetometer), ACMS and heat capacity options, respectively.

For the VSM measurements approximately 20 mg of the powdered samples was packed in polypropylene capsules and attached to the sample holder rod. The sample was investigated in the temperature range of 2.5–305 K and with magnetic flux densities up to 80 kOe (1 kOe = 7.96 × 104 A m–1).

For ac-MS measurements 67.754 mg of the powdered Ho10RuCd3 sample was enclosed in a thin-walled gelatine capsule. These investigations were performed in the temperature range of 5–100 K with internal frequencies from 41 to 9999 Hz and different magnetic field strengths up to 20 kOe.

The heat capacity of Ho10RuCd3 was measured with a 2.350 mg piece of the sintered tablet, which was fixed to a pre-calibrated heat capacity puck using Apiezon N grease and investigated between 2.5 and 300 K.

2.4 57Fe Mössbauer spectroscopy

A 57Co/Rh source was available for the 57Fe Mössbauer spectroscopic investigation. The Y10FeCd3 sample was placed in a PMMA container and the measurement was run in the usual transmission geometry at ambient temperature. Fitting of the spectrum was performed with the normos-90 program system [16].

3 Results and discussion

3.1 Crystal chemistry

The intermetallic cadmium compounds RE10TCd3 [11] crystallize with a ternary ordered version of the Co2Al5 type [17]. The three crystallographically independent rare earth atoms occupy the aluminum sites, while the transition metal and cadmium atoms reside on the Co2 and Co1 sites, respectively. The crystal chemical details have been discussed in [11]. Herein we briefly focus on the structural details that are relevant with respect to the magnetic characterization.

The RE10TCd3 structures contain three different polyhedral building units (Fig. 1). The RE3 atoms have slightly distorted icosahedral coordination by six cadmium and six RE1 atoms. These RE3@Cd6RE16 icosahedra are condensed via common triangular faces to infinite polyhedral chains along c, resulting in the motif of a rod packing. The space between these rods is filled by a second rod type which consists of condensed empty RE6 octahedra and transition metal filled T@RE6 trigonal prisms. For a detailed discussion of the interatomic distances we refer to [11].

Fig. 1: The crystal structure of the RE10TCd3 phases in polyhedral presentation.
Fig. 1:

The crystal structure of the RE10TCd3 phases in polyhedral presentation.

Figure 2 shows the two different layers of rare earth atoms as a projection onto the ab plane. As a consequence of the hexagonal space group symmetry, the rare earth substructure shows several equilateral, triangular units that might cause frustration in magnetic ordering. The layer around z ≈ 0 of the RE3 and RE1 atoms is slightly puckered, while the RE2 atoms are located on the mirror plane at z = 1/4. Due to the 63 screw axes, these motifs repeat around z ≈ 1/2 and z = 3/4.

Fig. 2: Projection of the rare earth layers of the RE10TCd3 structures onto the ab plane. The heights of the atoms are indicated. Both layer types repeat at z ≈ 1/2 and z = 3/4 as a consequence of the 63 screw axis.
Fig. 2:

Projection of the rare earth layers of the RE10TCd3 structures onto the ab plane. The heights of the atoms are indicated. Both layer types repeat at z ≈ 1/2 and z = 3/4 as a consequence of the 63 screw axis.

3.2 Magnetic properties

At first we discuss the magnetic properties of the three lutetium compounds Lu10TCd3 (T = Fe, Co, Ru). The temperature dependence of the magnetic susceptibility measured at 10 kOe is shown in Fig. 3. The overall susceptibility values are small. The susceptibilities of Lu10CoCd3 and Lu10RuCd3 are almost independent of temperature down to about 100 K. This is compatible with Pauli paramagnetism originating from the conduction electrons. The room temperature values are 5.6(2) × 10–4 and 4.6(2) × 10–4 emu mol–1 for Lu10CoCd3 and Lu10RuCd3, respectively. The upturns at low temperatures are due to tiny amounts of paramagnetic impurities. Also for Lu10FeCd3 we can conclude Pauli paramagnetic behavior, however, with higher susceptibility values and stronger temperature dependence (Fig. 3). This can be ascribed to iron containing impurity phases at the grain boundaries. Nevertheless, the absolute susceptibility values are two orders of magnitude lower than those of the paramagnetic holmium, erbium and thulium compounds discussed below. We can thus conclude that the transition metal atoms in the RE10TCd3 phases carry no magnetic moment. Similar behavior has been reported for the aristotype Co2Al5 [18].

Fig. 3: Temperature dependence of the magnetic susceptibility of Lu10TCd3 (T = Fe, Co, Ru) measured at 10 kOe.
Fig. 3:

Temperature dependence of the magnetic susceptibility of Lu10TCd3 (T = Fe, Co, Ru) measured at 10 kOe.

The high-temperature parts of the susceptibility data of all remaining compounds could be fitted with the Curie-Weiss law. The resulting experimental magnetic moments and Weiss constants are listed in Table 1. Most experimental moments are close to the theoretical values of the RE3+ free ion values, indicating stable trivalent rare earth elements. The enhanced value observed for Ho10FeCd3 is again a consequence of iron-containing impurity phases and/or elemental iron at the grain boundaries. Such a behavior is frequently observed for iron containing intermetallics.

Table 1

Magnetic properties of various RE10TCd3 compounds.

CompoundμexpB/RE)μeffB/RE)θP (K)μSMB/RE)μSM(calc)B/RE)HC (3 K) (kOe)TN (K)
Ho10FeCd311.16(5)10.6119(2)5.4(2)1062(5)43(1)
Ho10CoCd310.59(5)10.618.8(5)5.4(2)1053(5)38(1)
Ho10NiCd310.33(5)10.6119(2)5.4(2)1033(5)29(3)
Ho10RuCd310.72(5)10.612.2(5)3.7(2)1046.5(5)
Er10CoCd39.53(1)9.583.2(5)6.0(1)917.7(5)
Er10RuCd39.60(1)9.580.9(5)6.0(1)917(2)
Tm10FeCd37.46(5)7.56–0.4(5)3.6(2)717(4)9.2(5)
Tm10CoCd37.67(1)7.56–6.7(5)3.8(1)75.8(5)
Tm10NiCd37.46(5)7.56–4.8(5)3.5(1)73.9(5)
Tm10RuCd37.66(1)7.56–4.8(5)3.9(1)725(3)9.8(5)

μexp, Experimental magnetic moment; μeff, effective magnetic moment; θP, paramagnetic Curie temperature; μsm, experimental saturation magnetization; μsm(calc), calculated saturation magnetization; HC, critical field at 3 K; TN, Néel temperature.

Except for Ho10FeCd3 and Ho10NiCd3, all Weiss constants have low values close to zero. This is somehow in contrast to the magnetic ordering temperatures and is indicative of compensating magnetic interactions in the paramagnetic regimes in the case of the Ho and Er compounds. For instance, compensating antiferromagnetic and ferromagnetic interactions can result in small values for the Weiss constant. Nevertheless we can exclude magnetic frustration, since this would cause higher Weiss constants. According to Schiffer and Ramirez, geometrical frustration is evident if the frustration index (the quotient of the amount of the Weiss constant and the Néel temperature) is higher than 10 [19]. Our compounds have much smaller indices.

Exemplarily we describe the magnetic properties of Tm10RuCd3 and Ho10RuCd3 herein in detail, whereas the important parameters of the remaining compounds are listed in Table 1. The temperature dependence of the magnetic and inverse magnetic susceptibility (χ and χ–1 data) of Tm10RuCd3 are displayed in the top panel of Fig. 4 (10 kOe data). The Néel temperature of 9.8(5) K was deduced from a zero-field-cooled (ZFC)/field-cooled (FC) measurement at 100 Oe (inset of Fig. 4). This is in line with the Weiss constant of –4.8(5) K. Similar behavior has been observed for Tm10TCd3 with T = Fe, Co and Ni (Table 1).

Fig. 4: Magnetic properties of Tm10RuCd3: (top) Temperature dependence of the magnetic susceptibility and its reciprocal (χ and χ–1 data) measured at 10 kOe. The inset displays the magnetic susceptibility in ZFC/FC mode at 100 Oe; (bottom) Magnetization isotherms of Tm10RuCd3 at 3, 8, 20 and 50 K.
Fig. 4:

Magnetic properties of Tm10RuCd3: (top) Temperature dependence of the magnetic susceptibility and its reciprocal (χ and χ–1 data) measured at 10 kOe. The inset displays the magnetic susceptibility in ZFC/FC mode at 100 Oe; (bottom) Magnetization isotherms of Tm10RuCd3 at 3, 8, 20 and 50 K.

Magnetization isotherms of Tm10RuCd3 taken at 3, 8, 20 and 50 K are presented in the bottom panel of Fig. 4. At 20 and 50 K, well above the magnetic ordering temperature we observe linear field dependence as expected for a paramagnet. An s-shaped field dependence is observed for the 3 K isotherm. The spin reorientation (antiparallel to parallel realignment) takes place at a critical field strength of 23(3) kOe, underlining the antiferromagnetic ground state. The saturation moment at 3 K at the highest achievable field of 80 kOe is 3.9(1) μB per thulium atom, significantly lower than the theoretical value of 7 μB. The other three thulium compounds (Table 1) show comparable values for the saturation magnetization.

Data of the magnetic measurements of Ho10RuCd3 are shown in Fig. 5. The ZFC measurement (top panel of Fig. 5), carried out at an external field strength of 10 kOe, manifests antiferromagnetic ordering below approximately 50 K. The Weiss constant of 2.2(5) K was deduced from this measurement. ZFC/FC data at 50, 100 and 500 Oe are shown in the middle panel of that figure. We observe a first broad transition at TN = 46.5(5) K and a second anomaly below 20 K. The curves show a steep increase in this temperature range and the splitting between the ZFC/FC measurements is smaller with increasing field strength. This is indicative of weak intrinsic magnetic interactions or due to tiny amounts of impurities.

Fig. 5: Magnetic properties of Ho10RuCd3: (top) Temperature dependence of the magnetic susceptibility and its reciprocal (χ and χ–1 data) measured at 10 kOe; (middle) magnetic susceptibility in zero-field-cooled/field-cooled (ZFC/FC) mode at 50, 100 and 500 Oe; (bottom) magnetization isotherms of Ho10RuCd3 at 3, 40, 60 and 100 K.
Fig. 5:

Magnetic properties of Ho10RuCd3: (top) Temperature dependence of the magnetic susceptibility and its reciprocal (χ and χ–1 data) measured at 10 kOe; (middle) magnetic susceptibility in zero-field-cooled/field-cooled (ZFC/FC) mode at 50, 100 and 500 Oe; (bottom) magnetization isotherms of Ho10RuCd3 at 3, 40, 60 and 100 K.

A pronounced splitting of ZFC/FC curves is characteristic for ferromagnetic compounds or for short-range interactions (spin-glass phenomenon), which might be accompanied by magnetic frustration. This is in contrast to the antiferromagnetic transition at 46.5 K, which becomes more pronounced at higher field strengths, underlining the intrinsic nature. The magnetization isotherms of Ho10RuCd3 at 3, 40, 60 and 100 K up to maximum field strength of 80 kOe are presented in the bottom panel of Fig. 5. The isotherms in the paramagnetic range (60 and 100 K data) show a linear increase. The 3 K isotherm shows a distinct upturn above approximately 45 kOe, which can be attributed to a field induced spin-reorientation confirming the antiferromagnetic ground state, similar to Pr2Pd2Mg [20]. However, most metamagnetic materials show weak ferromagnetism in the high-field range and almost no hysteresis. The maximum magnetization of Ho10RuCd3 at 3 K and 80 kOe is 3.7(1) μB per holmium atom, significantly smaller than the theoretical value of 10 μB. All other holmium compounds of this series show higher saturation magnetization. A further metamagnetic step for Ho10RuCd3 at higher field strengths cannot be excluded.

In view of the two magnetic transitions observed for Ho10RuCd3 in the low-temperature regime, we also measured the temperature dependence of the specific heat from 2 to 300 K without an applied magnetic field (Fig. 6). The measurement shows a pronounced transition at 45.4(5) K, in good agreement with TN = 46.5(5) K determined from the susceptibility data. No further anomalies are evident towards lower temperatures, excluding a second intrinsic long-range ordering at lower temperatures. This is in line with the suppression of the 20 K anomaly by higher field strengths.

Fig. 6: Heat capacity of Ho10RuCd3 in the temperature range of 2.0–300 K without an applied field. The inset shows the magnified low-temperature area to highlight the magnetic ordering at 45.5 K.
Fig. 6:

Heat capacity of Ho10RuCd3 in the temperature range of 2.0–300 K without an applied field. The inset shows the magnified low-temperature area to highlight the magnetic ordering at 45.5 K.

The complex magnetic behavior of Ho10RuCd3 was additionally studied by ac-susceptibility data. The ZFC χ′ (ω, T) susceptibilities at frequencies ranging from 41 to 9999 Hz and an amplitude of 8 Oe are plotted in Fig. 7. In agreement with the dc-susceptibility data we observe two ordering phenomena at 45 and 15 K. The absence of a frequency dependence excludes dynamic processes. This is in line with the χ″ (ω, T) susceptibilities (Fig. 7, middle). The maximum value of χ″ (ω, T) corresponds to approximately one sixtieth of χ′ (ω, T). For common spin glasses a ratio of one tenth is usually observed [21]. We can thus conclude simple antiferromagnetic behavior for the first magnetic transition, whereby one possible reason for the broadening of these peaks is a weak homogeneity range of the sample. The anomaly below 20 K can most likely be attributed to a tiny impurity contribution. A wide-range magnetic phase transition can be excluded due to the absence of an anomaly in the heat capacity. Short-range magnetic interactions usually accompany dynamic processes, which are almost not evident in the ac-susceptibility measurements.

Fig. 7: ac-Susceptibility of Ho10RuCd3: (top) Temperature dependence of the magnetic susceptibility χ′ (ω) recorded after ZFC with internal frequencies ranging from 41 to 9999 Hz and an amplitude of 8 Oe; (middle) Magnified temperature area around the magnetic ordering temperatures; (bottom) Temperature dependence of the out-of-phase susceptibility χ″ (ω) around the second anomaly.
Fig. 7:

ac-Susceptibility of Ho10RuCd3: (top) Temperature dependence of the magnetic susceptibility χ′ (ω) recorded after ZFC with internal frequencies ranging from 41 to 9999 Hz and an amplitude of 8 Oe; (middle) Magnified temperature area around the magnetic ordering temperatures; (bottom) Temperature dependence of the out-of-phase susceptibility χ″ (ω) around the second anomaly.

3.3 57Fe Mössbauer spectroscopy

The room temperature 57Fe Mössbauer spectrum of Y10FeCd3 along with a transmission integral fit is presented in Fig. 8. In agreement with the crystallographic data the spectrum could be well reproduced with a single iron site at an isomer shift of δ = –0.38(1) mm s–1, subjected to quadrupole splitting of ΔEQ von 0.82(1) mm s–1 (a consequence of the trigonal prismatic yttrium coordination). The isomer shift value is within the expected range for iron in intermetallic compounds [22]. The experimental line width is Γ = 0.29(1) mm s–1.

Fig. 8: 57Fe Mössbauer spectrum of Y10FeCd3 at room temperature.
Fig. 8:

57Fe Mössbauer spectrum of Y10FeCd3 at room temperature.


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

Acknowledgments

O. N. and B. G. are indebted to the NRW Forschungsschule Molecules and Materials – A Common Design Principle for a PhD fellowship.

References

[1] U. Ch. Rodewald, B. Chevalier, R. Pöttgen, J. Solid State Chem. 2007, 180, 1720.10.1016/j.jssc.2007.03.007Search in Google Scholar

[2] F. Tappe, R. Pöttgen, Rev. Inorg. Chem. 2011, 31, 5.10.1515/revic.2011.007Search in Google Scholar

[3] R. Zaremba, U. Ch. Rodewald, R.-D. Hoffmann, R. Pöttgen, Monatsh. Chem. 2007, 138, 523.10.1007/s00706-007-0663-9Search in Google Scholar

[4] O. Niehaus, T. Bartsch, R. Pöttgen, Solid State Sci. 2015, 46, 95.10.1016/j.solidstatesciences.2015.06.008Search in Google Scholar

[5] F. Tappe, U. Ch. Rodewald, R.-D. Hoffmann, R. Pöttgen, Z. Naturforsch. 2011, 66b, 559.10.1515/znb-2011-0602Search in Google Scholar

[6] R. Zaremba, U. Ch. Rodewald, R. Pöttgen, Z. Naturforsch. 2007, 62b, 1574.10.1515/znb-2007-1216Search in Google Scholar

[7] A. V. Morozkin, R. Nirmala, S. K. Malik, J. Alloys Compd. 2005, 394, 75.10.1016/j.jallcom.2004.10.042Search in Google Scholar

[8] R. E. Gladyshevskii, Y. Grin, Y. P. Yarmolyuk, Dopov. Akad. Nauk Ukr. RSR, Ser. A1983, 2, 67.Search in Google Scholar

[9] S. Tuncel, J. G. Roquefère, C. Stan, J.-L. Bobet, B. Chevalier, E. Gaudin, R.-D. Hoffmann, U. Ch. Rodewald, R. Pöttgen, J. Solid State Chem. 2009, 182, 229.10.1016/j.jssc.2008.10.026Search in Google Scholar

[10] M. Kersting, S. F. Matar, C. Schwickert, R. Pöttgen, Z. Naturforsch. 2012, 67b, 61.10.1515/znb-2012-0111Search in Google Scholar

[11] M. Johnscher, T. Block, R. Pöttgen, Z. Anorg. Allg. Chem. 2015, 641, 369.10.1002/zaac.201400475Search in Google Scholar

[12] N. Nasri, M. Pasturel, V. Dorcet, B. Belgacem, R. Ben Hassen, O. Tougait, J. Alloys Compd. 2015, 650, 528.10.1016/j.jallcom.2015.07.277Search in Google Scholar

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

[14] D. Kußmann, R.-D. Hoffmann, R. Pöttgen, Z. Anorg. Allg. Chem. 1998, 624, 1727.10.1002/(SICI)1521-3749(1998110)624:11<1727::AID-ZAAC1727>3.0.CO;2-0Search in Google Scholar

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

[16] R. A. Brand, Normos Mössbauer Fitting Program, University of Duisburg, Duisburg (Germany) 2002.Search in Google Scholar

[17] J. B. Newkirk, P. J. Black, A. Damjanovic, Acta Crystallogr. 1961, 14, 532.10.1107/S0365110X61001637Search in Google Scholar

[18] G. Foex, J. Wucher, J. Phys. Radium1956, 17, 454.10.1051/jphysrad:01956001705045401Search in Google Scholar

[19] P. Schiffer, A. P. Ramirez, Comments Condens. Matter Phys. 1996, 18, 21.Search in Google Scholar

[20] R. Kraft, T. Fickenscher, G. Kotzyba, R.-D. Hoffmann, R. Pöttgen, Intermetallics2003, 11, 111.10.1016/S0966-9795(02)00189-9Search in Google Scholar

[21] K. Binder, A. P. Young, Rev. Mod. Phys.1986, 58, 801.10.1103/RevModPhys.58.801Search in Google Scholar

[22] G. K. Shenoy, F. E. Wagner, Mössbauer Isomer Shifts, North-Holland Publishing Company, Amsterdam, 1978.Search in Google Scholar

Received: 2015-9-16
Accepted: 2015-10-20
Published Online: 2015-12-7
Published in Print: 2016-1-1

©2016 by De Gruyter

Articles in the same Issue

  1. Frontmatter
  2. In this Issue
  3. Gas electron diffraction of increased performance through optimization of nozzle, system design and digital control
  4. Syntheses and crystal structures of two new sodium borates [Na2(H2O)3][B5O8(OH)2] and Na[enH2][B7O10(OH)4]
  5. Three new prenylflavonol glycosides from heat-processed Epimediumkoreanum
  6. Nano-SiO2: a heterogeneous and reusable catalyst for the one-pot synthesis of symmetrical and unsymmetrical 3,3-di(aryl)indolin-2-ones under solvent-free conditions
  7. Heterocycles [h]-fused to 4-oxoquinoline-3-carboxylic acid. Part XI: Synthesis and antibacterial activity of 4-fluoro-6-oxoimidazo[4,5-h]quinoline-7-carboxylic acids
  8. Synthesis, structure and magnetic properties of a binuclear copper(II) complex constructed by a new coordination mode of the tetrachlorophthalate ligand
  9. Structural and IR-spectroscopic characterization of magnesium acesulfamate
  10. Magnetic properties of RE10TCd3 (RE = Ho, Er, Tm, Lu; T = Fe, Co, Ni, Ru) and 57Fe Mössbauer spectroscopic data of Y10FeCd3
  11. Synthesis and characterization of the novel rare earth orthophosphates Y0.5Er0.5PO4 and Y0.5Yb0.5PO4
  12. Glutamyl-glutamate – a tailor-made chelating ligand for the [Be4O]6+ core in basic beryllium complexes and implications on investigations on the origins of chronic beryllium disease
  13. Notes
  14. Improved synthesis and crystal structure of the parent 1,3,5-trisilacyclohexane
  15. 1,3,5-Tris[(trimethylstannyl)ethynyl]- 1,3,5-trimethyl-1,3,5-trisilacyclohexane
  16. Corrigendum
  17. Corrigendum to: Ionic binuclear ferrocenyl compounds containing 1,1,3,3-tetracyanopropenide anions – synthesis, structural characterization and catalytic effects on thermal decomposition of main components of solid propellants
Downloaded on 26.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/znb-2015-0145/html
Scroll to top button