Alberto
Martinelli
*a,
Dominic
Ryan
b,
Julian
Sereni
c,
Clemens
Ritter
d,
Andreas
Leineweber
e,
Ivan
Čurlík
f,
Riccardo
Freccero
g and
Mauro
Giovannini
g
aSPIN-CNR, Corso F.M. Perrone 24, 16152 Genova, Italy. E-mail: alberto.martinelli@spin.cnr.it
bPhysics Department and the Centre for the Physics of Materials, McGill University, 3600 University St., Montreal, Quebec, Canada H3A 2T8
cDepartment of Physics, CAB-CNEA, CONICET, 8400 San Carlos de Bariloche, Argentina
dInstitut Laue – Langevin, 71 Avenue des Martyrs, 38042 Grenoble Cedex 9, France
eInstitute of Materials Science, TU Bergakademie Freiberg, Gustav-Zeuner-Str. 5, D-09599 Freiberg, Germany
fFaculty of Humanities and Natural Sciences, University of Prešov, 17. novembra 1, Prešov, Slovakia
gDepartment of Chemistry and Industrial Chemistry, University of Genova, Via Dodecaneso 31, 16146 Genova, Italy
First published on 9th May 2023
The chemical bonding, structural and magnetic properties of EuPdSn2 have been investigated by DFT, synchrotron X-ray and neutron powder diffraction and 151Eu Mössbauer spectroscopy. As a result, no structural transition is observed in the thermal range of 5–290 K, whereas ferromagnetic and antiferromagnetic orderings are found to coexist below 12 K and compete in the ground state. This magnetic phase separation is likely triggered by the minimization of the global energy resulting from the coexistence of the different magnetic configurations. Chemical bonding analysis in position space reveals the presence of heteroatomic 4a- and 5a-bonds, involving each species, and two-atomic Eu–Pd polar covalent interactions building up graphite-like distorted honeycomb layers.
In the course of our systematic investigation on the Eu–Pd–Sn system several novel compounds have been discovered, namely Eu2Pd2Sn, EuPd2Sn4 and EuPdSn2; strong indications of a complex magnetism with a divalent state of Eu were found by magnetization, specific heat and resistivity measurements in all these Eu-bearing compounds.5–7 Neutron diffraction is by far the best method of determining a magnetic structure. However, in the case of europium these measurements are hampered by the rather large absorption cross section for thermal neutrons (4530 b). Recently, it was demonstrated that neutron diffraction patterns of Eu compounds can be successfully collected by using a large-area flat-plane geometry.8 In fact, this technique has already been successfully used to determine the incommensurate magnetic structure of EuPdSn.4
By magnetic susceptibility and specific heat measurements, a complex magnetic behaviour below 12.5 K was evidenced in EuPdSn2, where Eu is found in a divalent state.7 In particular, the magnetic ordering was found to evolve from a state which is not simply antiferromagnetic to a ferromagnetic state with an increasing external magnetic field. A symptom that indicated that, at zero magnetic field, the ground state is not purely antiferromagnetic, is that the transition does not shift at all strengths by applying a magnetic field between 0 and 0.45, whereas a magnetic field of 0.5 T causes a rearrangement of magnetic moments towards a ferromagnetic state.7 These results prompted us to deepen the investigation of the complex and peculiar magnetic behaviour characterizing EuPdSn2.
In this paper, we combine 151Eu Mössbauer spectroscopy and neutron powder diffraction to show that antiferromagnetic and ferromagnetic domains coexist and compete at low temperature in EuPdSn2. Our study is complemented by a position-space chemical bonding analysis and synchrotron X-ray powder diffraction measurements.
151Eu Mössbauer spectroscopy measurements were carried out using a 4 GBq 151SmF3 source, driven in the sinusoidal mode. The drive motion was calibrated using a standard 57CoRh/α-Fe foil. Isomer shifts are quoted relative to EuF3 at ambient temperature. The 21.6 keV gamma rays were recorded using a thin NaI scintillation detector. The sample was cooled in a vibration-isolated closed-cycle helium refrigerator with the sample in the helium exchange gas. The spectra were used to fit the parameters describing a sum of Lorentzian lines with the positions and intensities derived from a full solution to the nuclear Hamiltonian.9
Synchrotron X-ray powder diffraction (XRPD) analysis was carried out at the high-intensity– high-resolution ID22 beamline of ESRF, using a wavelength λ = 0.3543 Å. XRPD data were collected in the thermal range 5–290 K. A reference pattern from LaB6 powder (NIST 660a) has been recorded to assess the instrumental line broadening.
Neutron powder diffraction (NPD) analysis was performed at the Institut Laue Langevin (ILL; Grenoble – F) using the high-intensity D20 diffractometer. NPD patterns were collected using a wavelength λ = 2.414 Å in the thermal range 2.5–20 K; in particular higher statistic data were collected at 20 K (in the paramagnetic state) and at 10.7 K and 2.5 K (below the magnetic transition temperature). In order to minimize the strong neutron absorption of natural Eu (σabs = 4530 b), a large-area flat-plat geometry sample holder was used;10 for this purpose, powdered EuPdSn2 was dispersed and fixed in the single-crystal silicon flat-wafer sample holder by using an alcoholic solution (ethanol) of GE varnish.
Structural refinements were carried out according to the Rietveld method11 using the program FullProf. For XRPD data, a file describing the instrumental resolution function (obtained by analysing a standard LaB6 sample) and a Thompson–Cox–Hastings pseudo-Voigt convoluted using an axial divergence asymmetry function were used during calculations. In the final cycle, the following parameters were refined: the scale factor; the zero point of detector; the background; the unit cell parameters; the atomic site coordinates not constrained by symmetry; the atomic displacement parameters; the anisotropic microstrain broadening using parameters described in ref. 12 tracing back to an approach introduced in ref. 13. For NPD data, Rietveld refinement was carried out by fitting the difference pattern obtained by subtracting data collected at 2.5 K (where magnetic ordering is complete) minus data collected at 18.5 K, in the paramagnetic regime, that is, by fitting the difference plot constituted of purely magnetic Bragg peaks.
On the XRPD data taken at 4 K, specific checks of different versions of the microstrain broadening model already used in the Fullprof evaluations were carried out applying a parametrization described in ref. 14. These evaluations were performed using the TOPAS software15 allowing for the use of user-designed line broadening models.
![]() | ||
Fig. 2 Total and atom-projected electronic density of states for EuPdSn2vs. energy; the Fermi energy EF is indicated by a dashed line. |
The non-zero states at the Fermi level indicate that the title phase is a metal, in line with resistivity measurements.7 The DOS region below ∼−5.0 eV is mainly contributed by the Sn 5s states; the Sn 5p are located at higher energies where they energetically overlap with the Pd and Eu states hinting toward Sn–Pd/Eu bonding interactions. The Pd 4d states are the main contributors in the energy window from about −5.0 to −2.0 eV. The fact that such states lie below the Fermi energy EF, and are spread in a quite wide energy range, implies a Pd charge acceptor behaviour and active involvement in chemical bonding. It is worth noting that the Eu 4f states are localised, as shown by their reduced bandwidth, and located around −2.7 eV. The goodness of both the chosen U parameter and structural model for EuPdSn2 is confirmed when compared with the (p)DOS of CaPdSn2 (Fig. 2, inset).21
Quantum-chemical techniques in position-space were selected to undertake further investigations on the interactions taking place among the constituents. Interesting insights may be obtained from the effective charges (Qeff) and shapes of the QTAIM atomic basins (see Fig. 3).
Tin is almost neutral having a Qeff(Sn) of −0.05. Since the above-described Sn–Sn distances reveal that the formation of four homopolar bonds per Sn atom is unrealistic, the origin of such a low value deserves additional investigations. Interestingly, Eu is the only QTAIM cation (+1.02) and Pd bears a negative effective charge of −0.92, enriching the family of intermetallic compounds containing an anionic late transition metal.44–50 The Qeff(Eu) is almost half of the formal one supporting its involvement in covalent interactions. The shapes of the atomic basins are quite characteristic for transition metal rare-earth tetrelides: Eu cations are quite spherical whereas both Pd and Sn display polyhedral shapes, sharing convex surfaces with Eu and flat among them. The ELI-D field possesses three kinds of attractors occupying the 16 h (0.2798, 0.0776, 0.0390), the 4c (0, 0.1872, ¼) and the 8f (0, 0.0936, 0.6277) Wyckoff positions; the related ELI-D basins are shown in Fig. 4 to the left and are indicated with grey, greenish and purple colours, respectively.
Such a distribution confirms the absence of a covalent (4b) Sn network, in line with the aforementioned interatomic-distance considerations. In fact, there are no ELI-D maxima in the vicinity of the Sn–Sn contacts, except for the shortest d1. The valence basin close to d1 (greenish) is indeed the only one intersected by two Sn atoms whereas for the others only one contributes. In addition to the two Sn atoms, the d1-basin is intersected by two Pd and one Eu QTAIM atoms, leading to the following bond fractions: p(BSni) = 0.35, p(BPdi) = 0.14, p(BEui) = 0.02. Thus, such a basin may be described as effectively five-atomic (5a-Sn2Pd2Eu1) and not just 2a as it would have been in the case of a classical Sn–Sn homopolar interaction, with a bond fraction per Sn atom of about 0.5. The fact that europium has been included in the atomicity (number of QTAIM atoms contributing to the ELI-D basin population51) despite its low bond fraction deserves additional comments. In fact, similar values were recently published for the binary CaGe compound describing the Ca–Ge interactions as 5a-Ge1Ca4 bonds after the application of a specifically developed method, called Penultimate Shell Corrections (PSC0).27 As a result of this correction, the total bond fraction of 4Ca increased from 0.08 to 0.15. The PSC0 was conceived particularly to account for underestimated valence contributions due to considerable charge storage in the penultimate shell. Both calcium in CaGe and europium in EuPdSn2 display a core overpopulation of about 0.3 e, suggesting a similar scenario. When elements with an ambiguous oxidation state, like Pd, are involved the PSC0 cannot be employed. The remaining ELI-D valence basins are populated by 1.76 (purple) and 1.44 e (grey) and have almost identical bond fractions: 0.76 for Sn and 0.05 for 2 Eu in both cases, and 0.18 and 0.19 for Pd, respectively. Then, they should be interpreted as 4a-Sn1Pd1Eu2 polar covalent bonds, and not as Sn lone pairs due to the not negligible metal contributions, particularly for Pd.
The Pd penultimate shell basins (see Fig. 4, on the right) show three bulges that extend in the valence region pointing toward the closest Eu species, one at 3.160 Å and two at 3.457 Å (Fig. 4, on the right). In fact, they are intersected by the corresponding Eu QTAIM atoms (see green regions in Fig. 4 to the right). The same kind of feature was recently reported for some ternary La2TGe6 germanides (T = Ag, Pd)27 and for LaAuMg2,38 which is isostructural with the title compound. For these phases, bulges were interpreted as 2a polar covalent metal–metal bonds, also thanks to a complementary analysis of the ELI-D relative Laplacian. Therefore, Pd and Eu form heteropolar interactions realizing 2∞[EuPd] distorted honeycomb layers parallel to (001), analogously to Au and La within LaAuMg2. According to our knowledge, this is the first Eu–Pd bond reported so far. Finally, the origin of the almost zero charge of tin may be accounted for. Each Sn realizes four multiatomic covalent bonds practically without transferring its four valence electrons to the metal species, similarly to what happens when homopolar Sn–Sn bonds are formed. In fact, subtracting the Sn ELI-D core basin population from the Sn QTAIM one yields a valence electron number of 4.19 e, very close to the ideal 4.00 value.
![]() | ||
Fig. 5 Rietveld refinement plot for EuPdSn2 (synchrotron XRPD data collected at 4 K); the panel on the right shows an enlarged view of the fitting in the high-Q region. |
Lattice parameters (Å) | ||
---|---|---|
a | b | c |
4.4575(1) | 11.6073(1) | 7.4677(1) |
4.4480(1) | 11.5420(1) | 7.4266(1) |
Atomic positions | ||||
---|---|---|---|---|
Atom | Wyckoff site | x | y | z |
Sn | 8f | 0 | 0.1483(1) | 0.0483(1) |
0.1478(1) | 0.0474(1) | |||
Eu | 4c | 0 | 0.4339(1) | ¼ |
0.4340(1) | ||||
Pd | 4c | 0 | 0.7062(1) | ¼ |
0.7056(1) |
Agreement factors | |||
---|---|---|---|
R F-factor | 4.59 | R Bragg | 2.84 |
2.37 | 3.89 |
By using the structural data obtained by the XRPD analysis in the thermal range 4–290 K, the thermal expansion behaviour of EuPdSn2 was investigated by fitting the cell volume in the whole inspected thermal range, using a Grüneisen second-order approximation for the zero-pressure equation of state:52
![]() | (1) |
![]() | (2) |
Fig. 6 shows the resulting fitting curve: it can be seen that the Grüneisen law reasonably accounts for the observed temperature dependence of the cell volume. Since the curve describes the temperature dependence of the cell volume expected for a nonmagnetic phase, the departure observed at low temperature should account for magnetostriction in the developing magnetic phase.
![]() | ||
Fig. 6 Thermal evolution of the cell volume; the solid line shows the best fit to a second-order Grüneisen approximation. |
In a previous work on the Ce2Pd2In compound,56 it was found that magnetic properties are strongly influenced by faint variation of the chemical composition. In particular, excess of Pd favours antiferromagnetism, whereas excess of Ce induces ferromagnetism; moreover, both magnetic transitions occur at the nearly same temperature. Hence, it is fundamental to ascertain if the composition of the analysed EuPdSn2 sample is homogeneous, without significant chemical fluctuations that could induce different and co-existing magnetic orderings. At this scope, microstructural properties were accurately investigated by analysing the broadening of the diffraction lines and their temperature dependence.
The hkl dependent line widths evident in the XRPD data were calculated in the course of the FullProf evaluations by a widely used anisotropic microstrain model.12,14 The refined parameters reveal the temperature-dependent line widths along different directions, as shown in Fig. 7(a). Evidently, the most pronounced broadening occurs in the [100] direction and decreases to some degree with increasing temperature.
![]() | ||
Fig. 7 (a) Temperature dependent evolution of the width of the microstrain broadening in different directions given by hkl (FullProf evaluations). (b) Tensor surface representing the direction dependence of the microstrain broadening for the 4 K data (TOPAS evaluation) according to the more general model eqn (4) (equivalent to the FullProf evaluations) and (c) according to the more restricted model according to eqn (5). No scale is given; hence the surfaces do not depict the extent but the anisotropy. |
In view of exploring the possible origin of the microstrain broadening, its anisotropy was further analyzed by Rietveld refinements within the TOPAS software for the XRPD data recorded at 4 K. The microstrain broadening appeared to be of Lorentzian shape. Refinements were done (in view of the orthorhombic symmetry) using two different models for the anisotropy of the microstrain:
(a) The squared direction-dependent microstrain assumes the usually adopted 4th order polynomial in hkl. That model is equivalent to that used in the FullProf refinements, using, however, a parametrisation detailed in ref. 14. In that case the width of the microstrain broadening (projected on the diffraction vector during a powder-diffraction experiments), e.g. its full width of half maximum Bε amounts
![]() | (3) |
(b) The square of the direction-dependent microstrain is proportional to the square of a 2nd order polynomial. In that case
![]() | (5) |
Refinement with the 6 parameters of eqn (4) should always result in a better fitting than refinement using the special case in eqn (5) with only 3 parameters. This is indeed the case. Using the model according to eqn (5), one obtains a weighted profile R value of wRp = 0.075, whereas it decreases to 0.066 upon use of eqn (4), suggesting the significance of the additional degrees of freedom. The direction dependences from the refined parameters are depicted by the tensor surfaces shown in Fig. 7(b) [eqn (4)] and (c) [eqn (5)]. Both confirm the result of Fig. 7(a), i.e. that the largest microstrain occurs in the [100] direction. The bulges perpendicular to [100] visible in Fig. 7(b) can only be modeled using the degrees of freedom provided by eqn (4).
One possible considered origin of the peculiar two-phase magnetic structure (see what follows) was a possible inhomogeneous character of the sample. As described in ref. 14, anisotropic microstrain broadening due to composition variations should be compatible with eqn (5), which moreover would unlikely show a Lorentzian shape. Hence, we do not associate the anisotropic broadening to an inhomogeneous character of the sample.
Tracking the hyperfine field Bhf as a function of temperature (Fig. 9) shows a smooth evolution that can be fitted to a conventional mean-field Brillouin function yielding a transition temperature of 12.9(1) K. However, we had to use J = 1/2 rather than J = 7/2 expected for the Eu2+ ion, indicating that the local anisotropy at the europium site has a significant Ising-like character. In particular, a single hyperfine field is detected in the whole thermal range; this result indicates that the amplitude of the magnetic moments is rather homogenous in the whole sample, regardless of whether it belongs to a unique or different magnetic structures.
Interestingly, a divalent (and therefore moment carrying) component is still detected below ∼12 K. Since this component is associated with the Eu sub-structure, but it behaves paramagnetically (i.e. it does not interact with ordered moments), it should be sited at a zero intensity point of the global molecular field. The amount of these non-interacting ions decreases on cooling and it is definitively suppressed below ∼4 K (Fig. 10).
![]() | ||
Fig. 11 Temperature dependent evolution of the magnetic Bragg peaks in EuPdSn2 (NPD data; panels have different colors scales; red: high intensity, violet: low intensity). |
Notably, it is possible to distinguish two sets of magnetic peaks. In the first set the peaks are almost saturated at 10.7 K (arrows in Fig. 12); in the second set the intensities progressively increase only on cooling below 12 K. It is not possible to index both sets with the same magnetic propagation vector. Instead it was found that the first set follows k1 = [0,0,0] while the second set is created through k2 = [0,0,½].
Magnetic symmetry analysis57 using the programs BASIREPS58,59 and MAXMAGN60 was employed to determine the allowed magnetic structures. Tables 2 and 3 list the irreducible representations (IR) and their basis vectors (BV) of Eu on the Wyckoff position 4c for k1 = 0 and k2 = [0,0,½], respectively. All the magnetic Bragg peaks at 2.5 K are satisfactorily fitted by combining IR5 of k1 (Table 2) with either one of two different combinations of the BV of the IR2 with k2 (Table 3). As the propagation vectors associated with the magnetic structures correspond to different points in the Brillouin zone of the crystal structure, hereinafter magnetic structures with k1 and k2 are referred to as Γ and Z, respectively; the corresponding IR is indicated as subscript.
Eu Atoms | IR2 | IR3 | IR4 | IR5 | IR7 | IR8 |
---|---|---|---|---|---|---|
0 y ¼ | 0 1 0 | 0 0 1 | 1 0 0 | 0 1 0 | 1 0 0 | 0 0 1 |
0 −y ¾ | 0 −1 0 | 0 0 1 | −1 0 0 | 0 1 0 | 1 0 0 | 0 0 −1 |
Eu Atoms | IR1 | IR2 | ||||
---|---|---|---|---|---|---|
BV1 | BV2 | BV1 | BV2 | BV3 | BV4 | |
0 y ¼ | 0 0 0 | 1 0 0 | 0 1 0 | 0 0 0 | 0 0 0 | 0 0 1 |
0 −y ¾ | −1 0 0 | 0 0 0 | 0 0 0 | 0 0 1 | 0 1 0 | 0 0 0 |
As mentioned above, the microstructural analysis provides no clear evidence for significant chemical fluctuations in our sample. As a consequence, two models can be supposed in order to describe the magnetic structure: (1) a magnetic phase constituted of the superposition of the k1 and k2 structures (all reflections originate from the same magnetic domain); (2) a magnetic phase coexistence of the ferromagnetic phase following k1 with the antiferromagnetic phase created through k2 (the two sets of reflections originate from distinct magnetic domains).
On account of the strong Eu absorption reducing the coherently scattered intensity, Rietveld refinements were carried out by fitting the difference pattern obtained by subtracting the data collected in the paramagnetic regime at 18.5 K from the data collected at 2.5 K (where magnetic ordering is complete; Fig. 13, on the right). As a result, the difference plot is thus constituted of purely magnetic Bragg peaks. The scale-factor needed for the calculation of absolute magnetic moment values is determined from a refinement of the 18.5 K data using the purely nuclear phase (Fig. 13, on the left).
The Γ5 model (IR5 of k1; Table 2) corresponds to a ferromagnetic ordering belonging to the Cm′cm′ magnetic space group type (#63.464) with magnetic moments aligned along the b-axis (Fig. 14). Two magnetic structural models for Z2 are created through different combinations of the basis vectors of IR2 for the antiferromagnetic k2 coupling (Table 3); they belong to magnetic space group types Aama2 – #40.208 and Cc2/c – #15.90, respectively, and provide similar goodness of fit parameters, preventing a reliable selection based on NPD data only.
![]() | ||
Fig. 14 Magnetic moment orderings fulfilling the Γ5 and Z2 IRs and corresponding magnetic space group. |
The magnetic structure crystallizing in the Aama2 magnetic space group type is characterized by a non-collinear antiferromagnetic ordering and leads to a large difference of the magnetic moment values for different Eu positions pertaining to the same crystallographic site (Fig. 14). Conversely, in the Cc2/c magnetic structural model the magnetic moments are collinear and are of equal magnitude (Fig. 14). In model 1, the coupling between the ferromagnetic and the antiferromagnetic components leads to a unique multi-k model where both couplings are superposed within the same magnetic phase. As a result, strong variations of the total magnetic moment values are found, as the ferromagnetic component pointing in the b-direction (Fig. 14) will add to or be subtracted from the antiferromagnetic component pointing along the b-direction which is present in both (Fig. 14) of the possible k2 structures. At lower temperatures, this scenario leads to unphysical large values of the magnetic moment on half of the Eu-sites. As a consequence, this unique multi-k commensurate magnetic structure model must be rejected.
Taking into account the 151Eu Mössbauer spectroscopy outcomes, revealing that magnetic moment values must be of comparable amplitude even though they belong to different orderings, the magnetic structure crystallizing in the Cc2/c magnetic space group type (Fig. 14) should be thus preferred in the magnetic phase coexistence scenario (model 2). Then, by assuming equal magnetic moment amplitudes in both magnetic phases, it results in ∼33% of ferromagnetic phase coexisting with ∼67% of antiferromagnetic phase, with a magnetic moment of about 6.7μB at 2.5 K, in fair agreement with magnetization measurements where a saturation of approximately 6.8μB was found at 2 K.6
The magnetic behaviour of EuPdSn2 can be thus outlined as follows. The ferromagnetic Γ5 order parameter develops between 13.4 K and ∼ 10 K (Fig. 12, on the right). Conversely, the antiferromagnetic Z2 phase (Cc2/c magnetic structural model) grows at a slightly lower temperature, below 12.3 K, and the transition completes below ∼4 K (Fig. 12, on the right). It is interesting to observe that the magnetic peak intensity of the ferromagnetic phase undergoes a slight decrease as the antiferromagnetic phase develops (Fig. 12, on the right), indicating some kind of competition between these phases. The intensity decrease affecting the ferromagnetic peaks on cooling can have two different origins: (1) a decrease of the net magnetic moment of the ferromagnetic phase; and (2) a decrease of the volume percentage of the ferromagnetic phase. This second option should be preferred in our case. In fact, at around 12 K the 151Eu Mössbauer indicates that there is about 10% of paramagnetic phase (Fig. 10) and the hyperfine field is ∼14 T (Fig. 9). This implies that (i) about 90% of the Eu2+ in the sample is magnetically ordered and (ii) that the magnetic moment at this temperature is far from saturation, as the hyperfine field significantly increases on cooling down to 4 K, up to over 24 T. Taking into account the fact that the amplitudes of the magnetic moments must be comparable in both phases, the thermal dependence of the hyperfine field indicates that the net magnetic moment increases in both phase fractions, which automatically implies a decrease of the ferromagnetic phase fraction below 10.6 K.
Within this scenario, the neutron diffraction and the Mössbauer data can be consistently explained for the different temperature regions: at 12 K only ferromagnetic ordering (with ∼ 3μB) occurs in about 90% of the magnetic Eu sub-structure, the remaining 10% being still in the paramagnetic state. As the temperature is decreased down to 10.7 K, the phase percentage of this ferromagnetic phase decreases down to ∼60%, but its net moment increases up to ∼4.5μB; at this same temperature the antiferromagnetic phase percentage would amount then to about 33%, with a net magnetic moment of as well ∼4.5μB, whereas the paramagnetic phase percentage decreases down to 7%. For T < 4 K, the paramagnetic phase is no longer present and both magnetic orderings coexist within separated domains in the ground state as described above for the 2.5 K data.
The origin of the FM + AFM phases coexistence deserves further discussion. Indeed, DFT/LSDA+U calculations predict that there are no total energy differences (within the accuracy of DFT) between the FM and the two AFM orderings. On the other hand, faint deviations from the ideal chemical and structural model used for calculations could play a major role in promoting and switching the nature of the experimental magnetic ordering. In the real case, the presence of different inhomogeneities at both chemical and structural level can be hypothesized, such as Eu valence fluctuations, and local compositional or microstructural variation. Indeed, their occurrence would compromise the delicate energy balance characterizing the FM and the AFM states. As a matter of fact, all our analyses found no evidence for any kind of chemical and structural fluctuation. 151Eu Mössbauer spectroscopy indicates that all Eu is found as divalent; no evidence for Eu3+ diluted within the magnetic Eu2+ sub-structure can be gained, ruling out the possibility that magnetic interactions can be locally affected by this non-interacting ionic species. Indeed, magnetic properties could be strongly influenced by the chemical composition. In the similar intermetallic compound Ce2Pd2In two branches of solid solutions are observed, where excess of Pd favours antiferromagnetism, whereas excess of Ce induces ferromagnetism.56 In EuPdSn2 microstructural analysis detected no clear evidence for both compositional (non-stoichiometric composition) and structural (incipient symmetry breaking) fluctuations that could affect the average homogeneity of the crystal composition/symmetry even at the local scale (thus favouring different magnetic interactions). In conclusion, by considering the complementary 151Eu Mössbauer spectroscopy and neutron powder diffraction data, the peculiar magnetic phase coexistence detected in EuPdSn2 is fully consistent with the DFT/LSDA+U calculation prediction.
This journal is © The Royal Society of Chemistry 2023 |