Abstract
GdPtMg and YPtMg (both crystallize with the ZrNiAl-type structure) form a complete solid solution Gd1–xY x PtMg. Samples in x = 0.1 steps were synthesized from the elements in sealed tantalum ampoules in an induction furnace and characterized by Guinier powder patterns. The structures of four members of the solid solution were refined from single-crystal X-ray diffractometer data, confirming the mixed occupation of the Gd/Y site; however, without any indication for Gd/Y ordering. Temperature dependent magnetic susceptibility measurements reveal Curie-Weiss behavior for all samples and ferromagnetic ordering in the low-temperature regime. The Curie temperature drops linearly from 97.6 K for GdPtMg to 3.7 K for Gd0.1Y0.9PtMg. All samples are soft ferromagnets. The Gd/Y substitution is a suitable tool for adjusting magnetic ordering temperatures of gadolinium intermetallics over a broad temperature range.
1 Introduction
The magnetic ordering in rare earth-based intermetallic compounds can effectively be influenced through the formation of diverse solid solutions [1]. In a general ternary compound RE x T y X z with RE = rare earth element, T = electron-rich transition metal, X = element of the third, fourth or fifth main group, all three sites can be substituted, effecting the magnetic long-range ordering of the rare earth substructure.
To give some examples, substitution of the T element by another one (i) modifies the RE–T hybridization and/or (ii) the electron count. An interesting case concerns the solid solution CeRh1−xIr x Ge [2, 3], where a change from antiferromagnetic CeRhGe to intermediate-valent CeIrGe sets in at 0.28 ≤ x ≤ 0.36. Nickel substitution by cobalt decreases the electron count in the solid solution CeNi1−xCo x Sn [4]. A first order valence phase transition takes place in the x range from 0.35 to 0.40. As a function of the cobalt concentration, the transition temperature changes from 40 to 80 K [4]. CeRu1−xPd x Sn and CeRh1−xPd x Sn are examples for solid solutions where the valence change of cerium is accompanied also by a structural phase transition [5].
The X site substitution was e.g. tested for the solid solution CeRhSb1−xSn x [6]. The TiNiSi-type structure is stable up to x < 0.2. The Kondo gap disappears by tin substitution and the solid solution switches to a non-Fermi liquid system. A remarkable example with an X site substitution is the solid solution EuAu4Cd2–xMg x . These samples show ferromagnetic ordering at around 16 K and the structural Cd/Mg disorder has no major influence on the ferromagnetic ground state [7].
Substitution of the rare earth site by replacing the paramagnetic rare earth element by a diamagnetic one allows a study of the spin dilution. Typical pairs are Ce3+/La3+, Eu2+/Sr2+, Gd3+/Y3+ and Yb3+/Lu3+. Especially the equiatomic CeTX compounds were studied with respect to solid solutions Ce1–xLa x TX [8], [9], [10]. Examples for the Eu2+/Sr2+ and Yb3+/Lu3+substitutions are the ferromagnetic solid solution Eu1–xSr x PtIn2 [11] and the segmented Heisenberg quantum spin chains in Yb4–xLu x As3 [12]. Many other intermetallic and oxydic materials have been studied in this direction.
From the point of view of the very close covalent radii, the Gd3+ (161 pm)/Y3+ (162 pm) pair is an obvious case [13]. Nevertheless, only a few studies of such solid solutions have been reported, e.g. Gd1–xY x Co2 [14], Gd1–xY x Mn2Ge2 [15], Gd1–xY x TiGe [16], and Gd1–xY x Co2B2 [17]. Recent investigations showed that the intermetallic compounds GdAuMg and YAuMg form a continuous solid solution Gd1–xY x AuMg. The gadolinium substitution leads to a strong decrease of the Néel temperature from TN = 81.1(1) K for GdAuMg to 3.5(1) K for Gd0.1Y0.9AuMg [18]. Dilution of the gadolinium substructure with yttrium drives the magnetization isotherms to nearly Brillouin-type behavior as is known for the free Gd3+ ion.
In continuation of these substitution experiments we herein report on the solid solution Gd1–xY x PtMg which starts from ferromagnetic GdPtMg with TC = 97.6 K [19].
2 Experimental
2.1 Synthesis
The samples of the solid solution Gd1–xY x PtMg were synthesized directly from the elements in sealed tantalum ampoules. Starting materials were gadolinium ingots (Smart Elements, 99.99%), yttrium ingots (Smart Elements, 99.99%), platinum powder (Agosi, 99.9%) and magnesium shavings (Alfa Aesar, 99.8%). First, the gadolinium and yttrium ingots were cut into smaller pieces and arc melted to buttons in an atmosphere of ca. 800 mbar argon [20]. The argon (Westfalen, 99.998%) was purified over titanium sponge (870 K), silica gel, and molecular sieves. The samples were synthesized by weighing all elements in the ideal stoichiometric ratio into tantalum ampoules which were subsequently arc-welded [20] under a reduced pressure of 800 mbar. The tantalum ampoules were then positioned in a water-cooled sample chamber [21] of a high-frequency furnace (Hüttinger Elektronik, Freiburg, Typ TIG 1.5/300), first heated to 1570 K and kept at that temperature for 1 min, followed by a rapid temperature reduction to 1270 K. The samples were cooled to 900 K within 30 min and finally annealed for another 2 h, followed by quenching. The temperature was controlled through a radiation pyrometer (Metis MS09, Sensortherm) with an accuracy of ±50 K. The polycrystalline samples have metallic luster, while ground powders are dark gray. No reaction with the crucible material was evident, all samples were separated mechanically from the tantalum ampoules. The samples are stable in air over several weeks.
2.2 X-ray diffraction
The polycrystalline Gd1–xY x PtMg samples were studied through Guinier powder patterns (Enraf-Nonius FR 552 camera): CuKα1 radiation, imaging plate detector (Fuji film, BAS-READER 1800) and α-quartz (a = 491.30 and c = 540.46 pm) as an internal standard. The hexagonal lattice parameters (Table 1) were deduced from least-squares refinements of the experimental 2θ values. Correct indexing was ensured by comparison with calculated patterns (Lazy Pulverix routine [22]).
Refined lattice parameters for the samples of the solid solution Gd1–xY x PtMg based on Guinier powder data. Standard deviations are given in parentheses.
Compound | a (pm) | c (pm) | V (nm3) | Reference |
---|---|---|---|---|
GdPtMga | 738.0(1) | 409.02(5) | 0.1929 | [19] |
GdPtMg | 737.5(1) | 408.32(7) | 0.1923 | This work |
Gd0.9Y0.1PtMg | 736.2(1) | 408.61(6) | 0.1918 | This work |
Gd0.8Y0.2PtMg | 735.7(1) | 408.34(8) | 0.1914 | This work |
Gd0.7Y0.3PtMg | 735.3(1) | 408.50(8) | 0.1913 | This work |
Gd0.692(1)Y0.308(1)PtMga | 733.50(7) | 406.99(4) | 0.1896 | This work |
Gd0.691(2)Y0.309(2)PtMga | 732.99(7) | 407.45(4) | 0.1896 | This work |
Gd0.6Y0.4PtMg | 735.1(1) | 408.35(8) | 0.1911 | This work |
Gd0.5Y0.5PtMg | 734.5(1) | 408.67(9) | 0.1909 | This work |
Gd0.485(7)Y0.515(7)PtMga | 734.30(7) | 408.26(4) | 0.1906 | This work |
Gd0.4Y0.6PtMg | 733.8(1) | 408.30(7) | 0.1904 | This work |
Gd0.3Y0.7PtMg | 733.2(2) | 408.3(1) | 0.1901 | This work |
Gd0.298(5)Y0.702(5)PtMga | 731.75(5) | 408.15(3) | 0.1893 | This work |
Gd0.2Y0.8PtMg | 732.8(1) | 408.27(9) | 0.1899 | This work |
Gd0.1Y0.9PtMg | 731.5(1) | 407.81(6) | 0.1890 | This work |
YPtMg | 730.6(2) | 407.6(1) | 0.1884 | This work |
YPtMg | 733.1(2) | 408.83(8) | 0.1903 | [23] |
-
aSingle-crystal data.
Irregularly shaped crystals were mechanically separated from the crushed Gd1–xY x PtMg samples with x = 0.3, 0.5 and 0.7. The crystals were glued to quartz fibers with beeswax and their quality was tested by Laue photographs on a Buerger camera (white molybdenum radiation, image plate technique, Fujifilm, BAS-1800). Complete data sets were measured at room temperature on a STOE StadiVari (Mo micro focus source and a Pilatus detection system) single-crystal diffractometer. The Gaussian-shaped profile of the micro focus X-ray source required scaling along with numerical absorption corrections. One data set was collected on a STOE IPDS-II diffractometer (graphite-monochromatized MoKα radiation; oscillation mode). A numerical absorption correction was applied. Details about the data collections and the structure refinements are listed in Table 2.
Crystal data and structure refinement parameters for Gd0.692(1)Y0.308(1)PtMg, Gd0.691(2)Y0.309(2)PtMg, Gd0.485(7)Y0.515(7)PtMg and Gd0.298(5)Y0.702(5)PtMg; ZrNiAl type, space group P
Empirical formula | Gd0.692(1)Y0.308(1)PtMg | Gd0.691(2)Y0.309(2)PtMg | Gd0.485(7)Y0.515(7)PtMg | Gd0.298(5)Y0.702(5)PtMg |
---|---|---|---|---|
Formula weight, g mol−1 | 355.6 | 355.5 | 341.4 | 328.7 |
Lattice parameters (single-crystal data) | ||||
a, pm | 733.50(7) | 732.99(7) | 734.30(7) | 731.75(5) |
c, pm | 406.99(4) | 407.45(4) | 408.26(4) | 408.15(3) |
Cell volume, nm3 | 0.1896 | 0.1896 | 0.1906 | 0.1893 |
Calculated density, g cm−3 | 9.34 | 9.34 | 8.92 | 8.65 |
Crystal size, µm3 | 10 × 30 × 50 | 10 × 30 × 40 | 20 × 30 × 50 | 10 × 30 × 50 |
Transmission (min/max) | 0.068/0.355 | 0.087/0.410 | 0.081/0.456 | 0.092/0.420 |
Absorption coefficient, mm−1 | 80.1 | 80.1 | 79.1 | 79.0 |
Detector distance, mm | 40 | 40 | 70 | 40 |
Exposure time, s | 60 | 35 | 900 | 46 |
ω range/increment, deg | −58.0–25.5/0.5 | −53.0–25.5/0.5 | 0–180/1 | −58.0–25.5/0.3 |
Integration param. (A/B/EMS) | 7/−6.0/0.030 | 7/−6.0/0.030 | 14.0/−1.0/0.030 | 7/−6.0/0.030 |
F(000), e | 439 | 439 | 423 | 409 |
θ range, deg | 3.2–33.6 | 3.2–33.6 | 3.2–33.4 | 3.2–34.0 |
Range in hkl | ±11/±11/±6 | ±11/±11/±6 | ±11/±11/±6 | ±11/±11/±6 |
Total no. reflections | 4454 | 3981 | 2277 | 6157 |
Independent reflections/Rint | 314/0.0451 | 315/0.0546 | 308/0.0237 | 321/0.0409 |
Reflections with I ≥ 3 σ(I)/R σ | 311/0.0076 | 310/0.0110 | 301/0.0078 | 314/0.0054 |
Data/parameters | 314/16 | 315/16 | 308/16 | 321/16 |
Goodness-of-fit on F2 | 1.72 | 2.62 | 1.00 | 1.17 |
R1/wR2 for I ≥ 3 σ(I) | 0.0184/0.0436 | 0.0272/0.0727 | 0.0108/0.0250 | 0.0113/0.0279 |
R1/wR2 for all data | 0.0184/0.0436 | 0.0275/0.0728 | 0.0110/0.0250 | 0.0118/0.0281 |
Flack parameter | 0.00(3) | 0 | 0.01(2) | 0.03(2) |
Extinction coefficient | 41(10) | 88(2) | 107(7) | 123(9) |
Largest diff. peak/hole, e Å−3 | +1.15/−1.44 | +1.49/−1.19 | +0.67/−0.71 | +0.58/−0.74 |
2.3 EDX analysis
The single crystals studied on the diffractometer were semiquantitatively analyzed by EDX using a Zeiss EVO® MA10 scanning electron microscope (tungsten cathode) in variable pressure mode (60 Pa) with GdF3, Y, Pt and MgO as standards. The experimentally obtained compositions were close to the compositions refined from the X-ray data (Table 2). The crystals were especially analyzed with respect to potential impurities resulting from the tantalum containers. No such impurity elements were detected.
2.4 Physical property studies
The magnetic properties of the solid solution Gd1–xY x PtMg were investigated on the X-ray pure samples by magnetic susceptibility and magnetization experiments. The samples were ground to fine powders, filled into polypropylene capsules and then attached to the brass sample holder rod of a vibrating sample magnetometer (VSM) of a Quantum Design Physical Property Measurement System (PPMS). The experiments were conducted in the temperature range of 2.5–300 K with applied external magnetic fields of up to 80 kOe (1 kOe = 7.96 × 104 A m−1). Fitting and plotting of the data was done with OriginPro 2016G [24] and the graphical editing with the program CorelDRAW2017 [25].
3 Structure refinements
The four data sets showed hexagonal lattices with high Laue symmetry without further systematic extinctions. Space group P
Atomic coordinates, anisotropic and equivalent isotropic displacement parameters (pm2) for Gd0.692(1)Y0.308(1)PtMg, Gd0.691(2)Y0.309(2)PtMg, Gd0.485(7)Y0.515(7)PtMg and Gd0.298(5)Y0.702(5)PtMg (ZrNiAl type, space group P
Atom | Wyck. | x | y | z | U 11 | U 22 | U 33 | U 12 | U eq |
---|---|---|---|---|---|---|---|---|---|
Gd 0.691(2) Y 0.309(2) PtMg | |||||||||
Gd/Y | 3f | 0.41115(2) | 0 | 0 | 159(4) | 161(6) | 178(6) | 80(3) | 166(4) |
Pt1 | 2d | 2/3 | 1/3 | 1/2 | 160(3) | U 11 | 177(4) | 80(2) | 166(2) |
Pt2 | 1a | 0 | 0 | 0 | 173(4) | U 11 | 156(5) | 87(2) | 168(3) |
Mg | 3g | 0.7555(10) | 0 | 1/2 | 130(20) | 150(30) | 170(30) | 77(15) | 150(20) |
Gd 0.692(1) Y 0.308(1) PtMg | |||||||||
Gd/Y | 3f | 0.58874(9) | 0 | 0 | 221(3) | 227(3) | 201(3) | 113(2) | 216(2) |
Pt1 | 2d | 1/3 | 2/3 | 1/2 | 219(2) | U 11 | 202(2) | 109(1) | 213(1) |
Pt2 | 1a | 0 | 0 | 0 | 234(2) | U 11 | 181(3) | 117(1) | 216(2) |
Mg | 3g | 0.2453(5) | 0 | 1/2 | 193(12) | 196(16) | 208(16) | 98(8) | 199(11) |
Gd 0.485(7) Y 0.515(7) PtMg | |||||||||
Gd/Y | 3f | 0.58884(6) | 0 | 0 | 120(2) | 127(2) | 128(2) | 64(1) | 124(2) |
Pt1 | 2d | 1/3 | 2/3 | 1/2 | 127(1) | U 11 | 131(1) | 63(1) | 128(1) |
Pt2 | 1a | 0 | 0 | 0 | 139(2) | U 11 | 111(2) | 70(1) | 130(1) |
Mg | 3g | 0.2450(3) | 0 | 1/2 | 115(8) | 124(11) | 138(10) | 62(5) | 125(7) |
Gd 0.298(5) Y 0.702(5) PtMg | |||||||||
Gd/Y | 3f | 0.41094(6) | 0 | 0 | 172(2) | 178(2) | 182(2) | 89(1) | 177(2) |
Pt1 | 2d | 2/3 | 1/3 | 1/2 | 174(1) | U 11 | 180(1) | 87(1) | 176(1) |
Pt2 | 1a | 0 | 0 | 0 | 186(1) | U 11 | 160(2) | 93(1) | 177(1) |
Mg | 3g | 0.7550(3) | 0 | 1/2 | 153(6) | 172(9) | 187(9) | 86(5) | 169(6) |
Interatomic distances (pm) for Gd0.692(1)Y0.308(1)PtMg, Gd0.691(2)Y0.309(2)PtMg, Gd0.485(7)Y0.515(7)PtMg and Gd0.298(5)Y0.702(5)PtMg. Standard deviations are equal or smaller than 0.1 pm. All distances of the first coordination spheres are listed.
Gd0.692Y0.308PtMg | Gd0.691Y0.309PtMg | Gd0.485Y0.515PtMg | Gd0.298Y0.702PtMg | |||
---|---|---|---|---|---|---|
Gd/Y: | 4 | Pt1 | 300.8 | 300.9 | 301.4 | 300.9 |
1 | Pt2 | 301.7 | 301.4 | 301.9 | 300.7 | |
2 | Mg | 323.8 | 324.4 | 324.7 | 324.1 | |
4 | Mg | 332.4 | 332.3 | 333.0 | 332.1 | |
4 | Gd/Y | 383.7 | 383.5 | 384.1 | 382.9 | |
Pt1: | 3 | Mg | 282.4 | 282.6 | 282.8 | 281.9 |
6 | Gd/Y | 300.8 | 300.9 | 301.4 | 300.9 | |
Pt2: | 6 | Mg | 271.6 | 271.3 | 272.1 | 271.6 |
3 | Gd/Y | 301.7 | 301.4 | 301.9 | 300.7 | |
Mg: | 2 | Pt2 | 271.6 | 271.3 | 272.1 | 271.6 |
2 | Pt1 | 282.4 | 282.6 | 282.8 | 281.9 | |
2 | Mg | 311.6 | 310.4 | 311.6 | 310.5 | |
2 | Gd/Y | 323.8 | 324.4 | 324.7 | 324.1 | |
4 | Gd/Y | 332.4 | 332.3 | 333.0 | 332.1 |
CCDC 2102223 (Gd0.692Y0.308PtMg), 2102225 (Gd0.691Y0.309PtMg), 2102220 (Gd0.485Y0.515PtMg) and 2102224 (Gd0.298Y0.702PtMg) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.
4 Crystal chemistry
The intermetallic phases GdPtMg [19] and YPtMg [23] form a complete solid solution Gd1–xY x PtMg. Samples in steps of x = 0.1 were synthesized and characterized through their X-ray powder patterns along with their refined lattice parameters (Table 1). The solid solution Gd1–xY x PtMg shows small differences when compared with the isotypic series Gd1–xY x AuMg [18]. While both the a and c lattice parameters weakly decrease from the gadolinium to the yttrium compound for Gd1–xY x AuMg [18], for Gd1–xY x PtMg the c parameter remains more or less constant and only the a lattice parameter decreases by less than 1%. In total we observe a small decrease of the cell volume (Figure 1) in going to the pure yttrium compound. The single-crystal data is not plotted in Figure 1. Differences in the lattice parameters between powder and single-crystal data are mainly due to the use of different techniques (no internal standard for the single-crystal diffractometer).

Course of the cell volumes of the samples of the solid solution Gd1–xY x PtMg.
GdPtMg and YPtMg crystallize with the ZrNiAl-type structure [30], [31], [32], one of the basic structure types for ternary intermetallic compounds with more than 1600 entries in the Pearson data base [33]. The various crystal chemical facets of the ZrNiAl type have repeatedly been reviewed and we refer to review articles [34, 35] for further details. Only some important aspects for the Gd1–xY x PtMg series are discussed herein.
As an example we present a projection of the Gd0.485Y0.515PtMg structure in Figure 2. The two basic structural motifs are platinum centered trigonal prisms formed by the rare earth (Gd/Y mixing) and magnesium atoms. The two prism types are shifted by half the c parameter with respect to each other. The Pt1@(Gd/Y)6 prisms are condensed via common edges leading to six-membered rings. The resulting channels are filled by columns (sharing triangular faces in c direction) of the Pt2@Mg6 prisms. The (Gd/Y)3 triangles of the Pt1@(Gd/Y)6 prisms are a striking structural motif with respect to the magnetic behavior, since such triangles lead to geometrical frustration, often observed in such kagome like networks [36], [37], [38].

Projection of the Gd0.485Y0.515PtMg structure onto the ab plane. Gadolinium, yttrium, platinum and magnesium atoms are drawn as medium gray, green, blue and magenta circles. The Gd/Y mixing is emphasized by segments. The trigonal prismatic coordination around the two crystallographically independent platinum sites is highlighted. The two different prismatic building units are shifted by half the c translation period, emphasized by thin and thick lines.
Similar to the solid solution Gd1–xY x AuMg [18], we have also carefully checked the data sets from the Gd1–xY x PtMg series with respect to possible Gd/Y ordering. Again, no hints for Gd/Y ordering were observed. Most likely the very similar size of gadolinium and yttrium avoids any ordering. In that view we draw to the data sets of Gd0.692(1)Y0.308(1)PtMg and Gd0.691(2)Y0.309(2)PtMg with similar compositions but slightly differing lattice parameters. This might be indicative of different clustering in the rare earth substructure.
5 Magnetic properties
GdPtMg [19] becomes ordered ferromagnetically at TC = 97.6 K. Yttrium-substituted samples in x = 0.1 steps were measured with respect to the temperature dependence of the magnetic susceptibility. The magnetic data of the Gd0.6Y0.4PtMg sample is presented in Figure 3. The sample shows Curie-Weiss behavior above 100 K with an experimental magnetic moment of 8.35(3) µB per gadolinium atom and a Weiss constant of 59.0(4) K. The magnetic moment is slightly higher than the free ion value [39] of 7.94 µB per gadolinium atom for Gd3+. This is due to the gadolinium 5d electrons that induce 4f–5d exchange interactions [40, 41]. The Gd1–xY x AuMg series [18] also shows this behavior. The precise Curie temperature of 55.3(1) K of Gd0.6Y0.4PtMg was determined from a low-field measurement in zero-field-cooled/field-cooled mode (Figure 4).

(Top) Temperature dependence (χ and χ−1 data) of the magnetic susceptibility of Gd0.6Y0.4PtMg measured at 10 kOe. (Bottom) Magnetization isotherms of Gd0.6Y0.4PtMg measured at 3, 10, 50 and 100 K.

Temperature dependence of the magnetic susceptibility of the members of the solid solution Gd1–xY x PtMg measured at 100 Oe (zero field-cooling (dots) and field-cooling (thin lines) curves).
The magnetization isotherms of Gd0.6Y0.4PtMg are shown at the bottom of Figure 3. At 100 K, well above the Curie temperature, we observe an almost linear increase of the magnetization as expected for a paramagnetic material. At 50 K, only slightly below the Curie temperature, the magnetization rapidly rises and becomes saturated around 5 µB per gadolinium atom at 80 kOe. The steep increase and the almost negligible hysteresis classifies Gd0.6Y0.4PtMg as a soft ferromagnet. The isotherms at 3 and 10 K are congruent. They show full saturation at 80 kOe (7.17(1) µB per Gd atom).
The remaining samples all show very similar behavior. The basic magnetic data derived from the susceptibility and magnetization measurements is summarized in Table 5. The results of all zero-field-cooled/field-cooled measurements are plotted in Figure 4. It is remarkable that the ferromagnetic ground state is retained down to the Gd0.1Y0.9PtMg sample, however, with a drastically decreased Curie temperature of 3.7(1) K. The course of TC as a function of x is presented in Figure 5.
Magnetic properties of the samples of the solid solution Gd1–xY x PtMg, with TC, Curie temperature, μeff, effective magnetic moment, μcalc, calculated magnetic moment, θp, paramagnetic Curie temperature, μsat saturation moment and calculated saturation magnetization according to gJ × J.
Compound | TC (K) | µeff (µB) | µeff, theo (µB) | θP (K) | µsat (µB) | gJ × J (µB) |
---|---|---|---|---|---|---|
GdPtMg [19] | 97.6(1) | 8.27 | 7.94 | 90.9(5) | 7.22 | 7 |
Gd0.9Y0.1PtMg | 86.6(1) | 8.28(1) | 7.94 | 86.1(1) | 7.09(1) | 7 |
Gd0.8Y0.2PtMg | 76.8(1) | 8.34(1) | 7.94 | 75.9(1) | 7.16(1) | 7 |
Gd0.7Y0.3PtMg | 68.2(1) | 8.37(1) | 7.94 | 68.8(1) | 7.11(1) | 7 |
Gd0.6Y0.4PtMg | 55.3(1) | 8.35(3) | 7.94 | 59.0(4) | 7.17(1) | 7 |
Gd0.5Y0.5PtMg | 49.5(1) | 8.35(1) | 7.94 | 54.2(1) | 7.20(1) | 7 |
Gd0.4Y0.6PtMg | 38.4(1) | 8.47(1) | 7.94 | 41.5(1) | 7.17(1) | 7 |
Gd0.3Y0.7PtMg | 26.6(1) | 8.39(1) | 7.94 | 32.5(1) | 7.11(1) | 7 |
Gd0.2Y0.8PtMg | 16.2(1) | 8.45(1) | 7.94 | 22.5(1) | 7.15(1) | 7 |
Gd0.1Y0.9PtMg | 3.7(1) | 8.43(1) | 7.94 | 10.7(1) | 7.13(1) | 7 |

Course of the Curie temperature of the solid solution Gd1–xY x PtMg.
Summing up, the results of the studies of the solid solution presented here nicely underpin that gadolinium spin dilution by diamagnetic yttrium is a suitable tool to adjust magnetic ordering temperatures of gadolinium intermetallics over a wide temperature range.
Acknowledgments
We thank Dipl.-Ing. J. Kösters for collecting the single-crystal X-ray data.
-
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: None declared.
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
1. Szytuła, A., Leciejewicz, J. Handbook of Crystal Structures and Magnetic Properties of Rare Earth Intermetallics; CRC Press: Boca Raton, 1994.Search in Google Scholar
2. Chevalier, B., Rogl, P., Hlil, E. K., Tuilier, M. H., Dordor, P., Etourneau, J. Z. Phys. B Condens. Matter 1991, 84, 205; https://doi.org/10.1007/bf01313538.Search in Google Scholar
3. Gaudin, E., Chevalier, B., Heying, B., Rodewald, U. Ch., Pöttgen, R. Chem. Mater. 2005, 17, 2693; https://doi.org/10.1021/cm040203c.Search in Google Scholar
4. Adroja, D. T., Rainford, B. D., de Teresa, J. M., del Moral, A., Ibarra, M. R., Knight, K. S. Phys. Rev. B 1995, 52, 12790; https://doi.org/10.1103/physrevb.52.12790.Search in Google Scholar PubMed
5. Niehaus, O., Abdala, P. M., Pöttgen, R. Z. Naturforsch. 2015, 70b, 253; https://doi.org/10.1515/znb-2015-0003.Search in Google Scholar
6. Ślebarski, A. J. Low Temp. Phys. 2007, 147, 147.10.1007/s10909-007-9306-0Search in Google Scholar
7. Reimann, M. K., Bönnighausen, J., Klenner, S., Pöttgen, R. Z. Naturforsch. 2020, 75b, 643; https://doi.org/10.1515/znb-2020-0066.Search in Google Scholar
8. Pöttgen, R., Chevalier, B. Z. Naturforsch. 2015, 70b, 289; https://doi.org/10.1515/znb-2015-0018.Search in Google Scholar
9. Pöttgen, R., Janka, O., Chevalier, B. Z. Naturforsch. 2016, 71b, 165; https://doi.org/10.1515/znb-2016-0013.Search in Google Scholar
10. Janka, O., Niehaus, O., Pöttgen, R., Chevalier, B. Z. Naturforsch. 2016, 71b, 737; https://doi.org/10.1515/znb-2016-0101.Search in Google Scholar
11. Klenner, S., Heletta, L., Pöttgen, R. Dalton Trans. 2019, 48, 3648; https://doi.org/10.1039/c9dt00035f.Search in Google Scholar PubMed
12. Matysiak, R., Gegenwart, P., Ochiai, A., Antkowiak, M., Kamieniarz, G., Steglich, F. Phys. Rev. B 2013, 88, 224414; https://doi.org/10.1103/physrevb.88.224414.Search in Google Scholar
13. Emsley, J. The Elements; Oxford University Press: Oxford, 1999.Search in Google Scholar
14. Murata, K., Fukamichi, K., Fujinaga, Y., Syono, Y., Goto, T. J. Magn. Magn. Mater. 1995, 140–144, 833; https://doi.org/10.1016/0304-8853(94)00827-2.Search in Google Scholar
15. Granovsky, S., Gaidukova, I., Ritter, C. EPJ Web Conf. 2018, 185, 04009; https://doi.org/10.1051/epjconf/201818504009.Search in Google Scholar
16. Morozkin, A. V., Viting, L. M., Sviridov, I. A., Tskhadadze, I. A. J. Alloys Compd. 2000, 297, 168; https://doi.org/10.1016/s0925-8388(99)00583-6.Search in Google Scholar
17. Ballou, R., Burzo, E., Mincic, A., Pop, V. J. Magn. Magn. Mater. 1993, 118, L285; https://doi.org/10.1016/0304-8853(93)90427-4.Search in Google Scholar
18. Reimann, M. K., Pöttgen, R. Z. Anorg. Allg. Chem. 2021, 647; https://doi.org/10.1002/zaac.202100168.Search in Google Scholar
19. Łątka, K., Tomkowicz, T., Kmieć, R., Pacyna, A. W., Mishra, R., Fickenscher, T., Hoffmann, R.-D., Pöttgen, R., Piotrowski, H. J. Solid State Chem. 2002, 168, 331.10.1006/jssc.2002.9718Search in Google Scholar
20. Pöttgen, R., Gulden, Th., Simon, A. GIT Labor-Fachzeitschrift 1999, 43, 133.Search in Google Scholar
21. Pöttgen, R., Lang, A., Hoffmann, R.-D., Künnen, B., Kotzyba, G., Müllmann, R., Mosel, B. D., Rosenhahn, C. Z. Kristallogr. 1999, 214, 143; https://doi.org/10.1524/zkri.1999.214.3.143.Search in Google Scholar
22. Yvon, K., Jeitschko, W., Parthé, E. J. Appl. Crystallogr. 1977, 10, 73; https://doi.org/10.1107/s0021889877012898.Search in Google Scholar
23. Stein, S., Heletta, L., Block, T., Gerke, B., Pöttgen, R. Solid State Sci. 2017, 67, 64; https://doi.org/10.1016/j.solidstatesciences.2017.03.006.Search in Google Scholar
24. OriginPro 2016G (version 9.3.2.303); OriginLab Corporation: Northampton, Massachusetts (USA), 2016.Search in Google Scholar
25. CorelDRAW Graphics Suite 2017 (version 19.0.0.328); Corel Corporation: Ottawa, Ontario (Canada), 2017.Search in Google Scholar
26. Petříček, V., Dušek, M., Palatinus, L. Z. Kristallogr. 2014, 229, 345.10.1515/zkri-2014-1737Search in Google Scholar
27. Flack, H. D., Bernadinelli, G. Acta Crystallogr. A 1999, 55, 908; https://doi.org/10.1107/s0108767399004262.Search in Google Scholar PubMed
28. Flack, H. D., Bernadinelli, G. J. Appl. Crystallogr. 2000, 33, 1143; https://doi.org/10.1107/s0021889800007184.Search in Google Scholar
29. Parsons, S., Flack, H. D., Wagner, T. Acta Crystallogr. B 2013, 69, 249; https://doi.org/10.1107/s2052519213010014.Search in Google Scholar PubMed PubMed Central
30. Krypyakevich, P. I., Markiv, V. Ya., Melnyk, E. V. Dopov. Akad. Nauk. Ukr. RSR, Ser. A 1967, 750.Search in Google Scholar
31. Dwight, A. E., Mueller, M. H., Conner, R. A.Jr., Downey, J. W., Knott, H. Trans. Met. Soc. AIME 1968, 242, 2075.Search in Google Scholar
32. Zumdick, M. F., Hoffmann, R.-D., Pöttgen, R. Z. Naturforsch. 1999, 54b, 45; https://doi.org/10.1515/znb-1999-0111.Search in Google Scholar
33. Villars, P., Cenzual, K., Eds. Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds (release 2020/21); ASM International®: Materials Park, Ohio (USA), 2020.Search in Google Scholar
34. Parthé, E., Gelato, L., Chabot, B., Penzo, M., Cenzual, K., Gladyshevskii, R. TYPIX-standardized data and crystal chemical characterization of inorganic structure types. In Gmelin Handbook of Inorganic and Organometallic Chemistry, 8th ed.; Springer: Berlin, 1993.10.1007/978-3-662-02909-1Search in Google Scholar
35. Zumdick, M. F., Pöttgen, R. Z. Kristallogr. 1999, 214, 90; https://doi.org/10.1524/zkri.1999.214.2.90.Search in Google Scholar
36. Ramirez, A. P. Annu. Rev. Mater. Sci. 1994, 24, 453; https://doi.org/10.1146/annurev.ms.24.080194.002321.Search in Google Scholar
37. Collins, M. F., Petrenko, O. A. Can. J. Phys. 1997, 75, 605; https://doi.org/10.1139/p97-007.Search in Google Scholar
38. Klenner, S., Reimann, M. K., Pöttgen, R. Z. Kristallogr. 2021, 236; https://doi.org/10.1515/zkri-2021-2026.Search in Google Scholar
39. Lueken, H. Magnetochemie; Teubner: Stuttgart, 1999.10.1007/978-3-322-80118-0Search in Google Scholar
40. Łątka, K. In Magnetism and Hyperfine Interactions in GdT2Si2 Systems, Report No1443/PS; Institute of Nuclear Physics: Cracow, 1989; p. 68. Chapt. 5.Search in Google Scholar
41. Czjzek, G., Oestreich, V., Schmidt, H., Łątka, K., Tomala, K. J. Magn. Magn. Mater. 1989, 79, 42; https://doi.org/10.1016/0304-8853(89)90290-4.Search in Google Scholar
© 2021 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- In this issue
- Research Articles
- Antimycobacterial cycloartane derivatives from the roots of Trichilia welwistchii C. DC (Meliaceae)
- Three metal(II) complexes constructed using the 2-(1H-benzo[d]imidazol-2-yl)quinoline ligand
- Low-perchlorate blue-flame pyrotechnic compositions
- A convenient synthesis of trifluoromethyl-substituted quinolino[8,7-h]quinolines and quinolino[7,8-h]quinolines
- Curie temperature adjustment in the solid solution Gd1–xY x PtMg
- Synthesis of oligotetramethylene oxides with terminal amino groups as curing agents for an epoxyurethane oligomer
- Synthesis, characterization and optoelectronic properties of 2D hybrid RPbX4 semiconductors based on an isomer mixture of hexanediamine-based dications
- Electrochemical sensing of hydrogen peroxide on a carbon paste electrode modified by a silver complex based on the 1,3-bis(1H-benzimidazole-2-yl)propane ligand
Articles in the same Issue
- Frontmatter
- In this issue
- Research Articles
- Antimycobacterial cycloartane derivatives from the roots of Trichilia welwistchii C. DC (Meliaceae)
- Three metal(II) complexes constructed using the 2-(1H-benzo[d]imidazol-2-yl)quinoline ligand
- Low-perchlorate blue-flame pyrotechnic compositions
- A convenient synthesis of trifluoromethyl-substituted quinolino[8,7-h]quinolines and quinolino[7,8-h]quinolines
- Curie temperature adjustment in the solid solution Gd1–xY x PtMg
- Synthesis of oligotetramethylene oxides with terminal amino groups as curing agents for an epoxyurethane oligomer
- Synthesis, characterization and optoelectronic properties of 2D hybrid RPbX4 semiconductors based on an isomer mixture of hexanediamine-based dications
- Electrochemical sensing of hydrogen peroxide on a carbon paste electrode modified by a silver complex based on the 1,3-bis(1H-benzimidazole-2-yl)propane ligand