Abstract
GdCuMg has been synthesized by induction-melting of the elements in a sealed niobium ampoule followed by annealing in a muffle furnace. The sample was studied by powder and single crystal X-ray diffraction: ZrNiAl type, P6̅2m (a=749.2(4), c=403.3(1) pm), wR2=0.0242, 315 F2 values and 15 variables. Temperature dependent magnetic susceptibility measurements have revealed an experimental magnetic moment of 8.54(1) μB per Gd atom. GdCuMg orders ferromagnetically below TC=82.2(5) K and based on the magnetization isotherms it can be classified as a soft ferromagnet.
1 Introduction
Intermetallic gadolinium compounds are in the focus with respect to their promising magnetocaloric properties, especially when searching for materials for magnetic refrigeration near room temperature. The key feature for characterizing such materials is phase analyses along with a study of the basic magnetic and magnetocaloric data [1], [2], [3], [4], [5], [6], [7], [8], [9]. Another important feature concerns the electron density at the gadolinium nuclei which is readily recordable via 155Gd Mössbauer spectroscopy [10], [11], [12]; the isomer shift is related to the local electron densities, mainly of s character, giving a finger-print characteristic of the respective phase.
Besides many binary gadolinium intermetallics [9], a huge number of ternary compounds of the general compositions GdxTyXz (T=transition metal; X=element of the 3rd, 4th or 5th main group) has been characterized with respect to the basic magnetic properties in order to select suitable compounds for further magnetocaloric studies. We have added a number of compounds with X=Mg, Zn and Cd [13], [14], [15], [16] to this family of materials. Magnesium, zinc and cadmium are fully embedded within the three-dimensional polyanionic networks, underlining their covalent bonding character. The important feature for property studies, however, concerns the reduction of the electron count (with respect to group III, IV or V elements), leading to changes of the magnetic ground state and thus the magnetic ordering temperature. To give an example, substitution of indium in GdRhIn by zinc leads to a drastic reduction of the magnetic ordering temperature from TC=34 K for GdRhIn [17] to TN=12.7 K for GdRhZn [15], along with a change of the spin alignment in the magnetically ordered state.
Of the magnesium-based ternary systems we already studied the magnetic behavior of the equiatomic compounds GdTMg for T=Pd, Ag, Pt and Au [18], [19], which show comparatively high magnetic ordering temperatures (Table 1). GdAgMg shows the by far lowest Curie temperature of 39.3 K, a consequence of different chemical bonding. We have now closed the gap for this series of compounds and present the synthesis, structural, and magnetic characterization of GdCuMg herein.
Lattice parameters and magnetic ordering temperatures TM of several equiatomic GdTMg compounds with ZrNiAl type structure.
Compound | a, pm | c, pm | V, nm3 | TM, K | Reference |
---|---|---|---|---|---|
GdPdMg | 750.1(1) | 404.10(4) | 0.1969 | TC=95.7 | [18] |
GdPtMg | 738.0(1) | 409.02(5) | 0.1929 | TC=97.6 | [18] |
GdCuMg | 749.2(4) | 403.3(1) | 0.1960 | TC=82.2 | This work |
GdAgMg | 768.0(2) | 419.92(9) | 0.2145 | TC=39.3 | [18] |
GdAuMg | 756.3(1) | 412.71(7) | 0.2044 | TN=81.1 | [19] |
TN=Néel temperature; TC=Curie temperature.
2 Experimental
2.1 Synthesis
GdCuMg was synthesized directly from the elements (1 mmol quantity). Gadolinium pieces (Smart Elements, 99.99%), copper pieces (Alfa Aesar, 99.999%) and turnings of a magnesium rod (Alfa Aesar, 99.8%) were mixed in the 1:1:1 atomic ratio and arc-welded [20] in a niobum ampoule under an argon pressure of about 700 mbar. The argon (Westfalen, 99.998%) was purified over titanium sponge, silica gel and molecular sieves. The reaction container was placed into the water cooled quartz sample chamber of a high frequency furnace (TIG 2.5/300, Hüttinger Elektronik) [21]. The ampoule was heated to 1300 K and kept at that temperature for 5 min. Subsequently, the temperature was rapidly increased to 1800 K, kept for 2 min at this maximum and then lowered rapidly to 1350 K. After 120 min at 1350 K the temperature was reduced to 850 K within 90 min and the sample was annealed at that temperature for another 180 min, followed by quenching. The temperature was controlled through a Methis MS09 pyrometer (Sensortherm, Sulzbach, Germany) with an accuracy of ±50 K. This inductively heated niobium container was then sealed in an evacuated quartz tube and annealed at 773 K for 14 days within a conventional muffle furnace. The polycrystalline product showed no reaction with the container material. GdCuMg is air-stable.
2.2 X-ray diffraction
The GdCuMg cell parameters were refined from powder X-ray diffraction data: Guinier camera (Enraf-Nonius, type FR 552), imaging plate technique (Fuji Film, BAS-READER 1800), CuKα1 radiation and α-quartz (Sigma-Aldrich, a=491.30 and c=540.46 pm) as an internal reference. Correct indexing was ensured through an intensity calculation [22].
An irregulary shaped GdCuMg single-crystal was isolated from the bulk sample with the help of a light microscope and fixing it to a silica fiber with beeswax. The crystal was investigated on a Buerger precession camera (white Mo radiation and a Fuji film imaging plate) to check its quality. An intensity data set was collected at ambient temperature by using an IPDS-II image plate system (STOE; graphite-monochromatized MoKα radiation, oscillation mode). A numerical absorption correction was applied to the data set. Details on the crystallographic data and the structure refinement are given in Table 2.
Crystal data and structure refinement for GdCuMg, space group P6̅2m, Z=3.
Empirical formula | GdCuMg |
Formula weight, g mol−1 | 245.1 |
Unit cell dimensions, pm | a=749.2(4) |
(powder data) | c=403.3(1) |
Unit cell volume, nm³ | 0.1960 |
Calculated density, g cm−3 | 6.23 |
Crystal size, μm3 | 20×30×30 |
Detector distance, mm | 70 |
Exposure time, s | 360 |
Wavelength, pm | 71.359 (MoKα) |
ω-range/increment, deg | 0–180/1.0 |
Integr. param. (A/B/EMS) | 13.0/3.0/0.015 |
Abs. coefficient, mm−1 | 33.7 |
Transm. ratio (max/min) | 0.505/0.320 |
F(000), e | 315 |
θ range, deg | 3.1–33.5 |
Range in hkl | ±11; ±11; ±6 |
Total no. reflections | 7851 |
Independent refl. /Rint | 315/0.0739 |
Refl. with I>3σ(I) | 302/0.0120 |
Data/ref. parameters | 315/15 |
Goodness-of-fit on F2 | 0.86 |
R1/wR2 for I>3 σ(I)) | 0.0106/0.0236 |
R1/wR2 (all data) | 0.0128/0.0243 |
Extinction coefficient | 1290(80) |
Flack parameter | −0.01(3) |
Largest diff. peak/hole, e Å−3 | 0.33/−0.37 |
2.3 Structure refinement
Isotypism of GdCuMg with the GdTMg (T=Pd, Ag, Pt, Au) [18], [19] phases was already evident from the Guinier powder pattern. The data set showed the corresponding hexagonal lattice and no further systematic extinctions in agreement with space group P6̅2m. The atomic parameters of GdAuMg [19] were taken as starting values and the structure was refined with the Jana2006 package (full-matrix least-squares on Fo2) [23] with anisotropic displacement parameters for all atoms. Refinement of the occupancy parameters indicated no deviation from the ideal composition. The final difference Fourier syntheses revealed no significant residual peaks. The refined atomic positions, displacement parameters, and interatomic distances are given in Tables 3 and 4 .
Atomic coordinates and displacement parameters (pm2) for GdCuMg.
Atom | Wyck. | x | y | z | U11 | U22 | U33 | U12 | Ueq |
---|---|---|---|---|---|---|---|---|---|
Gd | 3f | 0.41449(3) | 0 | 0 | 123(1) | 131(1) | 129(1) | 66(1) | 127(1) |
Cu1 | 2d | 1/3 | 2/3 | 1/2 | 123(2) | U11 | 142(4) | 62(1) | 129(2) |
Cu2 | 1a | 0 | 0 | 0 | 155(3) | U11 | 109(6) | 78(2) | 140(3) |
Mg | 3g | 0.7577(2) | 0 | 1/2 | 102(7) | 105(7) | 143(9) | 53(4) | 117(6) |
The isotropic displacement parameter Ueq is defined as: Ueq=1/3 (U11+ U22+ U33); U13=U23=0. Standard deviations are given in parentheses.
Interatomic distances (pm) in the gadolinium coordination sphere of the GdTMg compounds.
T= | Pd | Pt | Cu | Ag | Au | ||
---|---|---|---|---|---|---|---|
4 | T1 | 303.2 | 302.5 | 302.6 | 312.4 | 307.6 | |
1 | T2 | 310.3 | 303.6 | 310.5 | 317.9 | 312.0 | |
2 | Mg | 327.6 | 326.0 | 326.8 | 336.0 | 330.6 | |
4 | Mg | 337.3 | 334.3 | 337.2 | 347.4 | 341.3 | |
4 | Gd | 391.5 | 386.0 | 390.7 | 400.7 | 395.1 | |
2 | Gd | 404.1 | 409.0 | 403.3 | 419.9 | 412.7 |
Standard deviations are equal or smaller than 0.3 pm.
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-432999.
2.4 EDX data
The GdCuMg single-crystal studied on the diffractometer was semiquantitatively investigated by an EDX analysis by use of a Zeiss EVO® MA10 scanning electron microscope equipped with a LaB6 cathode and using GdF3, Cu and MgO as internal standards. The experimentally observed composition was close to the ideal one with an accuracy of ±5%, a consequence of the conchoidal fracture of the irregular crystal surface. No impurity elements heavier than sodium were observed, which in particular rules out the presence of the container material niobium. The investigation was accomplished in variable pressure mode with an internal chamber pressure of 60 Pa.
2.5 Magnetic characterization
The magnetic susceptibility measurements were performed with a Quantum Design physical property measurement system (PPMS) using the vibrating sample magnetometer (VSM). A polypropylene capsule was filled with 14.318 mg of the powdered GdCuMg sample and attached to the sample holder rod of the VSM option. GdCuMg was investigated in the temperature range of 2.5–300 K with magnetic fields of up to 80 kOe (1 kOe=7.96×104 A m−1).
3 Crystal chemistry
GdCuMg crystallizes with the ZrNiAl type structure, space group P6̅2m with the lattice parameters a=749.2(4) and c=403.3(1) pm. So far, only few systematic phase analytical studies have been performed for the RE-Cu-Mg systems, i.e. Y-Cu-Mg [24], La-Cu-Mg [25], Tb-Cu-Mg [26], and Yb-Cu-Mg [27]. Solely the equiatomic compounds YCuMg (a=744.5, c=399.5 pm) [24], LaCuMg (a=773.2, c=417.9 pm) [25] and CeCuMg (a=766.8, c=416.4 pm) [28] have been reported. Already the Tb-Cu-Mg system [26] shows no ZrNiAl type phase. Thus GdCuMg is the last existing phase in the RECuMg series. The smaller rare earth elements do not allow formation of a RECuMg phase. High-pressure high-temperature studies will be performed in order to test if the series can be extended under such conditions, similar to the related series of rare earth stannides [29], [30]. In going from LaCuMg to GdCuMg, the unit cell decreases more or less isotropically as expected from the lanthanide contraction (3.2% decrease in a and 3.5% decrease in c).
The crystal chemistry of ZrNiAl type intermetallics has repeatedly been reviewed, e.g. in [1], [31], [32] and also discussed for equiatomic RETMg intermetallics [13], [16], [19]. We therefore only briefly discuss the GdCuMg structure. A projection of the latter is presented in Fig. 1. Both crystallographically independent copper sites show trigonal prismatic coordination, Cu1@Gd6 and Cu2@Mg2. The Cu1@Gd6 prisms show hexameric units and include the Cu2@Mg2 prisms. Both prism types are shifted with respect to each other by half the translation period c. Especially the triangular arrangement of the gadolinium atoms calls for frustration of the magnetic moments in the magnetically ordered state (vide infra).

Projection of the GdCuMg structure onto the xy plane. Gadolinium, copper and magnesium atoms are drawn as medium grey, blue, and magenta circles, respectively. All atoms lie on mirror planes at z=0 (thin lines) and z=1/2 (thick lines). The trigonal prismatic coordination around the two crystallographically independent copper sites is emphasized.
The shortest interatomic distances in the GdCuMg structure occur between the copper and magnesium atoms, ranging from 271 to 290 pm, slightly longer than the sum of the covalent radii of Cu+Mg of 253 pm [33]. We can therefore assume weaker Cu–Mg bonding. Together, the copper and magnesium atoms build up a three-dimensional network which incorporates the gadolinium atoms. Bonding of gadolinium to the [CuMg] network proceeds through Gd–Cu contacts at 303 and 311 pm.
At this point it is interesting to compare GdCuMg with the isotypic indide GdCuIn (a=746.3, c=399.6 pm) [34]. Although indium is larger than magnesium, the indide has smaller lattice parameters and thus also shows stronger Cu–In vs Cu–Mg bonding, similar to the series of RE2Cu2In vs RE2Cu2Mg compounds [35]. These compounds crystallize with ternary ordered versions of the U3Si2 type, again with trigonal prismatic rare earth coordination for the copper atoms.
Finally, in Table 4 we compare the relevant interatomic distances for the whole series of GdTMg (T=Cu, Pd, Ag, Pt, Au) phases. It is readily evident, that GdAgMg shows a pronounced anomaly. Although silver and gold have comparable covalent radii [33], the lattice parameters of the silver compound are distinctly larger than those of GdAuMg. This leads to much weaker Ag–Mg and Gd–Ag bonding and thus strongly influences the gadolinium magnetic ground state (vide infra). Although all of the listed GdTMg phases (Table 1) crystallize with the hexagonal ZrNiAl type, there are distinct differences in chemical bonding and a simple application of a rigid band model for understanding of the structure-property relations is critical.
4 Magnetic properties
The temperature dependence of the magnetic susceptibility (χ and χ−1 data) of GdCuMg measured in zero-field cooled (ZFC) mode with an applied field of 10 kOe is displayed in the top panel of Fig. 2. The inverse magnetic susceptibility was fitted using the Curie-Weiss law in the temperature range of 125–300 K revealing a paramagnetic Curie temperature of θP=82.0(1) K and an effective magnetic moment of μexp=8.54(1) μB per Gd atom, slightly higher than the free ion value of Gd3+ (7.94 μB). The observed excess magnetic moment can at least partially be ascribed to the gadolinium 5d electrons induced via 4f-5d exchange interactions, also observed for GdAuMg [19] and GdPtMg [18] or the solid solution GdxLa1−x Ni5 [36].

(Top) Temperature dependence of the magnetic susceptibility of GdCuMg measured with an applied field of 10 kOe. The inset shows the ZFC and FC data measured at 100 Oe. (Bottom) Magnetization isotherms measured at 3, 50, 100 and 150 K with a magnetic field strength of up to 80 kOe.
The inset of Fig. 2 depicts the zero-field cooled and field cooled (FC) mode measurement of GdCuMg at 100 Oe. The large bifurcation between the ZFC and the FC curves indicates dominant ferromagnetic interactions. The Curie temperature of TC=82.2(5) K was determined from the minimum of the derivative dχ/dT of the FC curve. The magnetization isotherms of GdCuMg measured at 3, 50, 100 and 150 K in a field range of 0–80 kOe are shown in the bottom panel of Fig. 2. The isotherms recorded above the Curie temperature exhibit the expected almost linear field dependence of a paramagnetic compound. Below the magnetic ordering temperature, a sharp increase of the magnetization is evident which is due to saturation effects. The observed full saturation value at 3 K and 80 kOe of μsat=7.3(1) μB per Gd atom is slightly higher than the theoretical value of 7 μB per gadolinium atom (according to g×J). The steep increase of the magnetization and the almost negligible hysteresis classify GdCuMg as a soft ferromagnet.
Finally it is interesting to compare GdCuMg with the other isotypic GdTMg (T=Pd, Pt, Ag, Au) [18], [19] phases. The Curie temperature of 82.2 K is comparable to the ordering temperatures published for GdPdMg (TC=95.7 K), GdPtMg (TC=97.6 K) [18] and GdAuMg (TN=81.1 K) [19]. GdAgMg has the much lower Curie temperature of TC=39.3 K [18]. This is in line with the bonding characteristics discussed above. In the series of GdTMg compounds GdAgMg shows the largest lattice parameters (Table 1) and consequently also larger interatomic distances (Table 4), of which Gd–Ag and Gd–Gd are the important ones with respect to magnetic coupling, resulting in weaker Gd(4f)-Ag(4d) exchange and weaker direct Gd–Gd coupling, leading to the much lower ordering temperature.
A simple correlation of the magnetic ordering temperature with the valence electron count (VEC) of several GdCuX phases is not possible: TC=82.2 K (GdCuMg; VEC=16), TC=82 K (GdCuAl [37]; VEC=17), TN=20 K (GdCuIn [38]; VEC=17) and TN=14 K (GdCuGe [39]; VEC=18). The magnetic ordering temperature and the type of spin alignment show large differences, indicating that the magnetic ground state is largely influenced by small changes in chemical bonding as discussed above for GdAgMg within the series of GdTMg compounds.
Acknowledgements
We thank Dipl.-Ing. U. Ch. Rodewald for the collection of the single crystal diffractometer data.
References
[1] A. Szytuła, J. Leciejewicz, Handbook of Crystal Structures and Magnetic Properties of Rare Earth Intermetallics, CRC Press, Boca Raton, 1994.Suche in Google Scholar
[2] V. K. Pecharsky, K. A. Gschneidner, Jr., Appl. Phys. Lett. 1997, 70, 3299.10.1063/1.119206Suche in Google Scholar
[3] V. K. Pecharsky, K. A. Gschneidner, Jr., Phys. Rev. Lett. 1997, 78, 4494.10.1103/PhysRevLett.78.4494Suche in Google Scholar
[4] V. K. Pecharsky, K. A. Gschneidner, Jr., Cryocoolers1999, 10, 629.Suche in Google Scholar
[5] E. V. Sampathkumaran, R. Mallik, NATO Sci. Ser. 2003, 110, 353.10.1007/978-94-010-0213-4_35Suche in Google Scholar
[6] K. A. Gschneidner, Jr., V. K. Pecharsky, A. O. Tsokol, Rep. Prog. Phys. 2005, 68, 1479.10.1088/0034-4885/68/6/R04Suche in Google Scholar
[7] O. Tegus, B. Li-Hong, S. Lin, Chin. Phys. B2013, 22, 037506.10.1088/1674-1056/22/3/037506Suche in Google Scholar
[8] L. Petit, D. Paudyal, Y. Mudryk, K. A. Gschneidner, Jr., V. K. Pecharsky, M. Lüders, Z. Szotek, R. Banerjee, J. B. Staunton, Phys. Rev. Lett. 2015, 115, 207201.10.1103/PhysRevLett.115.207201Suche in Google Scholar PubMed
[9] P. Villars, K. Cenzual, Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2016/17), ASM International®, Materials Park, Ohio (USA), 2016.Suche in Google Scholar
[10] G. Czjzek, in Mössbauer Spectroscopy of New Materials Containing Gadolinium (Eds.: G. J. Long, F. Grandjean), Plenum Press, New York, 1993, chapter 9, pp. 373.10.1007/978-1-4899-2409-4_9Suche in Google Scholar
[11] R. Pöttgen, K. Łątka, Z. Anorg. Allg. Chem. 2010, 636, 2244.10.1002/zaac.201000171Suche in Google Scholar
[12] K. Łątka, Nukleonika2013, 58, 23.Suche in Google Scholar
[13] U. Ch. Rodewald, B. Chevalier, R. Pöttgen, J. Solid State Chem. 2007, 180, 1720.10.1016/j.jssc.2007.03.007Suche in Google Scholar
[14] F. Tappe, R. Pöttgen, Rev. Inorg. Chem. 2011, 31, 5.10.1515/revic.2011.007Suche in Google Scholar
[15] W. Hermes, R. Pöttgen, Solid State Sci. 2009, 11, 706.10.1016/j.solidstatesciences.2008.11.003Suche in Google Scholar
[16] T. Mishra, G. Heymann, H. Huppertz, R. Pöttgen, Intermetallics2012, 20, 110.10.1016/j.intermet.2011.08.012Suche in Google Scholar
[17] S. B. Gupta, K. G. Suresh, AIP Conf. Proc. 2013, 1512, 1090.10.1063/1.4791425Suche in Google Scholar
[18] K. Łątka, Z. Tomkowicz, R. Kmieć, A. W. Pacyna, R. Mishra, T. Fickenscher, R.-D. Hoffmann, R. Pöttgen, H. Piotrowski, J. Solid State Chem.2002, 168, 331.10.1006/jssc.2002.9718Suche in Google Scholar
[19] K. Łątka, R. Kmieć, A. W. Pacyna, Th. Fickenscher, R.-D. Hoffmann, R. Pöttgen, Solid State Sci. 2004, 6, 301.10.1016/j.solidstatesciences.2004.01.006Suche in Google Scholar
[20] R. Pöttgen, Th. Gulden, A. Simon, GIT Labor-Fachzeitschrift1999, 43, 133.Suche in Google Scholar
[21] 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;2-0Suche in Google Scholar
[22] K. Yvon, W. Jeitschko, E. Parthé, J. Appl. Crystallogr.1977, 10, 73.10.1107/S0021889877012898Suche in Google Scholar
[23] V. Petříček, M. Dušek, L. Palatinus, Z. Kristallogr. 2014, 229, 345.10.1515/zkri-2014-1737Suche in Google Scholar
[24] S. De Negri, P. Solokha, A. Saccone, V. Pavlyuk, Intermetallics2009, 17, 614.10.1016/j.intermet.2009.02.001Suche in Google Scholar
[25] S. De Negri, M. Giovannini, A. Saccone, J. Alloys Compd.2007, 427, 134.10.1016/j.jallcom.2006.02.062Suche in Google Scholar
[26] P. Solokha, V. Pavlyuk, S. De Negri, A. Saccone, O. Zelinska, Xth International Conference on Crystal Chemistry of Intermetallic Compounds, September 17–20, 2007, Lviv (Ukraine), Poster O3.Suche in Google Scholar
[27] S. De Negri, A. Saccone, P. Rogl, G. Giester, Intermetallics2008, 16, 1285.10.1016/j.intermet.2008.08.004Suche in Google Scholar
[28] V. V. Kinzhibalo, A. T. Tyvanchuk, E. V. Mel’nyk, R. M. Rykhal, Visn. Lviv Derzh. Univ., Ser. Khim. 1988, 29, 17.Suche in Google Scholar
[29] C. P. Sebastian, G. Heymann, B. Heying, U. C. Rodewald, H. Huppertz, R. Pöttgen, Z. Anorg. Allg. Chem. 2007, 633, 1551.10.1002/zaac.200700011Suche in Google Scholar
[30] G. Heymann, B. Heying, U. Ch. Rodewald, O. Janka, H. Huppertz, R. Pöttgen, J. Solid State Chem. 2016, 236, 138.10.1016/j.jssc.2015.06.044Suche in Google Scholar
[31] 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, 1993.10.1007/978-3-662-10641-9Suche in Google Scholar
[32] R. Pöttgen, B. Chevalier, Z. Naturforsch. 2015, 70b, 289.10.1515/znb-2015-0018Suche in Google Scholar
[33] J. Emsley, The Elements, Oxford University Press, Oxford, 1999.Suche in Google Scholar
[34] L. V. Sysa, V. I. Zaremba, Ya. M. Kalychak, V. M. Baranyak, Visn. Lviv Derzh. Univ., Ser. Khim. 1988, 29, 32.Suche in Google Scholar
[35] R. Mishra, R.-D. Hoffmann, R. Pöttgen, Z. Naturforsch.2001, 56b, 239.10.1515/znb-2001-0304Suche in Google Scholar
[36] E. Burzo, L. Chioncel, I. Costina, S. G. Chiuzbăian, J. Phys.: Condens. Matter2006, 18, 4861.10.1088/0953-8984/18/20/011Suche in Google Scholar
[37] P. Javorský, L. Havela, V. Sechovský, H. Michor, K. Jurek, J. Alloys Compd. 1998, 264, 38.10.1016/S0925-8388(97)00198-9Suche in Google Scholar
[38] J. W. C. De Vries, R. C. Thiel, K. H. J. Buschow, J. Less-Common Met. 1985, 111, 313.10.1016/0022-5088(85)90203-6Suche in Google Scholar
[39] S. Rayaprol, C. P. Sebastian, R. Pöttgen, J. Solid State Chem. 2006, 179, 2041.10.1016/j.jssc.2006.04.005Suche in Google Scholar
©2017 Walter de Gruyter GmbH, Berlin/Boston
Artikel in diesem Heft
- Frontmatter
- In this Issue
- Synthesis and structure of large 24-mer and 36-mer oxamate-based macrocycles
- A [Mo2O2S2]-based ring system incorporating tartrate as the bridging ligand: synthesis, structure and catalytic activity of Cs4[Mo2O2(μ-S)2]2(μ4-tart)2 (tart=[C4H2O6]4−)
- Two new pyrrolo[2,3-d]pyrimidines (7-deazapurines): ultrasonic-assisted synthesis, experimental and theoretical characterizations as well as antibacterial evaluation
- One-pot desilylation-Sonogashira coupling
- Synthesis of benzodiazepines catalyzed by CoFe2O4@SiO2-PrNH2 nanoparticles as a reusable catalyst
- Synthesis, structural characterization, and hydrogen bonds of Co9(OH)14[SO4]2
- GdCuMg with ZrNiAl-type structure – an 82.2 K ferromagnet
- Notes
- Complete X-ray single-crystal structure determination and Raman spectrum of NH4[C(CN)3]
- Synthesis and crystal structure of a new homoleptic tetraarylruthenium(IV) complex Ru(2,4,5-Me3C6H2)4
- Book Review
- Lead: Its Effects on Environment and Health
Artikel in diesem Heft
- Frontmatter
- In this Issue
- Synthesis and structure of large 24-mer and 36-mer oxamate-based macrocycles
- A [Mo2O2S2]-based ring system incorporating tartrate as the bridging ligand: synthesis, structure and catalytic activity of Cs4[Mo2O2(μ-S)2]2(μ4-tart)2 (tart=[C4H2O6]4−)
- Two new pyrrolo[2,3-d]pyrimidines (7-deazapurines): ultrasonic-assisted synthesis, experimental and theoretical characterizations as well as antibacterial evaluation
- One-pot desilylation-Sonogashira coupling
- Synthesis of benzodiazepines catalyzed by CoFe2O4@SiO2-PrNH2 nanoparticles as a reusable catalyst
- Synthesis, structural characterization, and hydrogen bonds of Co9(OH)14[SO4]2
- GdCuMg with ZrNiAl-type structure – an 82.2 K ferromagnet
- Notes
- Complete X-ray single-crystal structure determination and Raman spectrum of NH4[C(CN)3]
- Synthesis and crystal structure of a new homoleptic tetraarylruthenium(IV) complex Ru(2,4,5-Me3C6H2)4
- Book Review
- Lead: Its Effects on Environment and Health