Brought to you by:
Paper

Rare-earth tuned magnetism and magnetocaloric effects in double perovskites R2NiMnO6

, , , and

Published 16 December 2021 © 2021 IOP Publishing Ltd
, , Citation Anzar Ali et al 2022 J. Phys.: Condens. Matter 34 095803 DOI 10.1088/1361-648X/ac3e9e

0953-8984/34/9/095803

Abstract

We present a comprehensive experimental study of magnetization (2 < T < 300 K, 1 < H < 8 T) and magnetocaloric effect in double perovskite materials R2NiMnO6 with R = Pr, Nd, Sm, Gd, Tb, and Dy. While a paramagnetic to ferromagnetic transition, with TC in the range $\sim 100-200\enspace $ K, is a common feature that can be attributed to the ordering of Mn4+ and Ni2+ magnetic moments, qualitatively distinct behavior depending on the choice of R is observed at low temperatures. These low-temperature anomalies in magnetization are also manifest in the change in magnetic entropy, −ΔSM, whose sign depends on the choice of R. In order to understand these results, we present theoretical analysis based on mean-field approximation and Monte Carlo simulations on a minimal spin model. The model correctly captures the key features of the experimental observations.

Export citation and abstract BibTeX RIS

1. Introduction

Magnetic materials have undoubtedly played an important role in advancing the technology to its commonly used current form [1]. What underlies these technologies are the key concepts or phenomena in magnetism that have been harnessed to our advantage via a careful design. Some of the technologically useful phenomena displayed by magnetic materials are, giant magnetoresistance, magnetocapacitance and magnetocaloric effect (MCE) [26]. At the level of fundamental physics, a common ingredient seems to be a delicate interplay of spin, charge, and lattice degrees of freedom that leads to various magnetic phase transitions, and the sensitivity of these transitions to external fields or pressure manifests in the form of a technologically useful effect [79].

Oxides of transition metals are perhaps the most well known materials that are equally interesting from fundamental physics as well as application points of view [1012]. Double perovskites (DP) with formula R2BB'O6, where R is rare earth and B, B' are the transition metal (TM) ions, belong to this interesting class of materials [1316]. The presence of two TM ions leads to multiplet of possibilities in terms of tuning the properties in a desired manner [1719]. The coupling between Rare earth and TM network further enriches the low temperature magnetic properties [7].

MCE remains a topic of immense interest as it opens pathway toward realization of clean and energy efficient magnetic refrigeration technology with clear advantage over the conventional vapor-compression techniques that use potentially harmful chlorofluorocarbon gases [20]. Recent studies show a giant and reversible MCE in various magnetic materials such as MnFeP0.45As0.55 [21] (magnetic entropy change −ΔSM = 18.0 J kg−1 K−1 for ΔH = 0–5 T), (MnNiSi)0.56(FeNiGe)0.44 (−ΔSM = 70 J kg−1 K−1 for ΔH = 0–5 T) [22], nano-crystalline films of EuTiO3 (−ΔSM = 24 J kg−1 K−1 for ΔH = 0–2 T) [23], GdCrTeO6 (−ΔSM ≈ 42 J kg−1 K−1 for ΔH = 0–7 T) [24], Gd5(Six Ge1−x ) [25], Ni–Mn–In [26] (adiabatic temperature change ΔTad = 6.2 K), Gd2NiMnO6 [27] (−ΔSM = 35.5 J kg−1 K−1 for ΔH = 0–7 T). The materials which show MCE at low temperatures can be advantageous for cryogenic magnetic cooling to obtain a sub-kelvin temperature as an alternative option of He3/He4 liquid whose prices are constantly increasing. Observation of a significant MCE in DPs opens up possibilities for using the high degree of flexibility available in these oxides to enhance the effect further. We have recently studied the magnetic and magneto-caloric properties of Nd2NiMnO6 and proposed that an interplay of the two magnetic sub-lattices can be used as a control knob to tune the MCE properties of magnetic materials with multiple magnetic sublattices and specifically the DPs [8]. It would be instructive to test this proposal by varying the moment size on one of the magnetic sublattices to see how the magnetic and MCE properties evolve. This motivates our present study.

In this work, we present detailed experimental investigations of magnetization and magnetocaloric behavior from 2 K to 300 K and in the field range 1 < H < 8 T, on R2NiMnO6 (R = Sm, Gd, Pr, Nd, Tb, and Dy). The data for Nd2NiMnO6 is taken from reference [8] for comparison with the other DPs. All studied DPs go through a paramagnetic to ferromagnetic phase transition due to the ordering of Ni2+ and Mn4+ magnetic ions. The choice of R, in addition to affecting the value of TC, also influences the magnetization and consequently the MCE behavior as inferred from the nature of low-temperature anomalies for different R. We propose a simple explanation for this behavior in terms of a Heisenberg model that takes into account the coupling of the spin on R with the spin on Ni–Mn network, as well as the local spin–orbit coupling on R sites. We support our experimental findings via Monte Carlo simulations and mean-field analysis on the two-sublattice Heisenberg model. The remainder of the paper is organized as follows. In section 2, we present the magnetization measurements for all studied DPs. The low temperature anomalies in magnetization as well as in magnetic entropy change are presented and discussed here. In section 3, we present a two-sublattice Heisenberg model for magnetism in DPs and describe the mean-field approximation and Monte-Carlo simulation method to study the model. Further, we discuss the results obtained by these theoretical methods in this section. We summarize and conclude in section 4.

Although several previous studies have been devoted to the study of the magnetic and MCE properties of DPs. However, we believe that a detailed study of the evolution of these properties across the whole series, a tracking of these properties as a function of lattice volume and rare-earth moment size, and a theoretical estimation of the temperature dependence of these properties using a simple Heisenberg Hamiltonian which takes into account the magnetic coupling of both sub-lattices with each other, have not been attempted before. In this sense our work presents new results.

2. Experimental results

Polycrystalline samples of R2NiMnO6 where R = Pr, Nd, Sm, Gd, Tb, and Dy were synthesized using a solid state reaction method as described previously in references [28, 29]. It is known that synthesis conditions are important in controlling the oxygen stoichiometry and anti-site disorders. Off-stoichiometry or anti-site disorder have been shown to effect the magnetic properties significantly, leading to a much reduced magnetization and to glassy magnetic features at low temperatures and fields [8, 3032]. The materials used in this study do not show any glassy magnetic behaviors and the magnetization is similar to the highest values reported for various DPs [30, 31, 33]. The phase purity of all DPs was examined by powder x-ray diffraction using a Rigaku Ultima IV Diffractometer. These x-ray diffraction results confirmed that all materials were single phase and crystallized in the expected crystal structure. Figure 1 shows the temperature dependent magnetization M(T) for all synthesized DPs at different fields as indicated in the plots.

Figure 1.

Figure 1. Temperature dependence of the magnetization measured in field-cooled protocol in applied magnetic fields up to 8 T for all the synthesized R2NiMnO6 compounds.

Standard image High-resolution image

All the DPs show a rapid increase in magnetization (M) at temperatures TC which is associated with the ferromagnetic ordering of Ni2+–Mn4+ sub-lattice. For higher field this upturn in magnetization moves to higher temperature which is a common feature of a ferromagnetic transition. At lower temperatures, a second anomaly is observed for all the DPs at T2. The values of TC and T2 for all DPs which were studied, are given in table 1, which are consistent with the previous reports [29, 30, 34]. The nature of the low temperature anomaly depends on the R ion. For R = Nd, and Sm, there is a downturn in M at T2 when measured in low magnetic fields. This downturn in M can be suppressed on the application of a magnetic field and changed to an upturn at larger magnetic fields as previously reported [8]. For R = Pr this downturn is not observed, possibly because the lowest field of measurement, H = 1 T is already strong enough to suppress it. This downturn in magnetization however is not a signature of long range ordering of R ions as demonstrated for Nd2NiMnO6 for which a low temperature heat capacity study did not show any anomaly [8] and microscopic probes like XMCD did not show any ordered moment on Nd3+ ions in Nd2NiMnO6 [33]. On the other hand, for R = Gd, Tb, and Dy there is an increase in M at T2 at low fields which is enhanced in larger fields. Taken at face value the above observations seem to indicate that there is antiferromagnetic coupling between the R and Ni–Mn sublattice for R = Pr, Nd, and Sm, while this coupling is ferromagnetic for R = Gd, Tb, and Dy [3537]. However, recent DFT calculations of the exchange constants in Nd2NiMnO6 have shown that the magnetic exchange between the Nd spin and the Ni–Mn sublattice is ferromagnetic [33]. The downturn in the magnetization can then be understood by realizing that for less than half-filled 4f shells, the orbital moment of the R is oppositely aligned to it's spin moment and is larger in magnitude than the spin moment. Therefore, although the spin is coupled ferromagnetically to the Ni–Mn sublattice, the orbital moment is opposite to the Ni–Mn sublattice resulting in a downturn in the magnetization for R = Pr, Nd, and Sm. At large enough magnetic fields, it becomes energetically favorable for the net effective magnetic moment to align with the field, which leads to a switching of the magnetization anomaly at T2 from a downturn to an upturn as field is increased. For R = Gd, Tb, and Dy the orbital moment is in the same direction as the spin moment and so the ferromagnetic coupling between the spin of R and the Ni–Mn sublattice leads to an upturn at T2.

Table 1. Theoretically estimated effective magnetic moment of rare earth ions (m) [38], ferromagnetic transition temperature (TC), anomaly at lower temperature (T2), and theoretically estimated and experimentally observed saturation magnetization (MS ) for all DPs R2NiMnO6.

R2NiMnO6 m (μB)TC (K)T2 (K)MS (theoretical)M'S (observed)
Dy2NiMnO6 10.63952125.0015.90
Tb2NiMnO6 9.721131523.0015.45
Gd2NiMnO6 7.941302019.0018.90
Nd2NiMnO6 3.621955011.546.02
Pr2NiMnO6 3.582132711.405.25
Sm2NiMnO6 0.85158206.435.10

Figure 2 shows the isothermal magnetization M(H) of all R2NiMnO6 at different temperatures as indicated in plots. All of the DPs were found to show a negligible hysteresis in magnetization down to the lowest measured temperature 2 K, indicating soft ferromagnetic behavior which is a desirable feature for the magnetic refrigeration techniques. At 300 K, in the paramagnetic state, M(H) for all the DPs varies linearly. Below TC but above T2, M(H) shows tendency to saturate but with a weak linear increase at large H as expected in the ferromagnetic state of Ni–Mn, while the R ions are still paramagnetic. Below T2 the M(H) curves are complicated by the different behavior for DPs with less than (R = Pr, Nd, Sm) and more than or equal to (R = Gd, Tb, Dy) half filled 4f-shells. For R = Pr, Nd, and Sm, the M(H) at low fields is lower than for T > T2 as seen in the downturn in the M(T) below T2 due to the orbital moment being anti-aligned with the Ni–Mn sublattice. At higher fields for which the downturn in M(T) changes to an upturn due to field induced aligning of the orbital moment with Ni–Mn sublattice, the M(H) increases and shows a tendency to saturate. However, a weak linear increase in M(H) with field remains down to T = −2 K, suggesting that the full magnetic moment of R has still not aligned with the field. The values of observed saturation magnetization (M'S ) at T = 2 K, H = 9 T, and the theoretically expected saturation magnetization (MS ) if all (R, Ni, and Mn) magnetic moments have saturated are also given in table 1. The value of observed saturated magnetization for all R2NiMnO6 is lower than the theoretically predicted saturation magnetization except for Gd2NiMnO6 probably because Gd has no orbital moment (L).

Figure 2.

Figure 2. Isothermal curves of magnetization verses magnetic field for all the R2NiMnO6 at various temperatures indicated in the plots.

Standard image High-resolution image

These field and temperature dependent magnetic responses also manifest themselves in the magnetic entropy change. The isotherms of magnetization M(H) were collected over a broad temperature range from 2 K to 300 K at temperature intervals of 2 K. Using these M(H) data, the change in magnetic entropy (ΔSM ) can be calculated from Maxwell's thermodynamic relation [2]:

Equation (1)

Figure 3 show the magnetic entropy change (−ΔSM ) versus temperature at different applied magnetic fields up to 8 T for all the R2NiMnO6 compounds. All the compounds show two prominent features. The first one at the ferromagnetic phase transition at TC and second one at lower temperature. The value of −ΔSM is positive for all the DPs across TC and it's magnitude increases with field, which is consistent with ferromagnetic ordering of the Ni–Mn sublattice at TC as indicated in the magnetization versus T and H. When the temperature is lowered, the behavior of −ΔSM like M(T), is of two kinds. For R = Pr, Nd, and Sm, the −ΔSM monotonically decreases below TC and even becomes negative for R = Nd and Sm at low fields. On increasing the applied magnetic field, the negative −ΔSM is suppressed and a positive peak or anomaly appears at low temperatures and high fields. This can be understood interms of the anti-alignment of the total effective moment of the R ions with the Ni–Mn sublattice below T2 at low fields. So the magnetic entropy is high. At higher fields, the R moment switches and aligns with the Ni–Mn sublattice and also the applied magnetic field, resulting in a low magnetic entropy like in the case of the ferromagnetic transition at TC. Gd2NiMnO6, Tb2NiMnO6, and Dy2NiMnO6 show a positive anomaly in −ΔSM below T2 since the effective moment of R is already tending to align with the Ni–Mn sublattice. This alignment simply increases as higher magnetic fields are applied resulting in an enhancement in the magnitude of −ΔSM . For R = Gd, Tb, and Dy the value of −ΔSM is very large at low temperatures, particularly for R = Gd, which make them potentially useful for low temperature magnetic refrigeration.

Figure 3.

Figure 3. Temperature dependence of magnetic entropy change (−ΔSM (T)) from 2 K to 300 K at applied magnetic fields up to 8 T for all the R2NiMnO6 compounds.

Standard image High-resolution image

Another quantity which is used to quantify the usefulness of MCE materials is called RCP which is defined as:

Equation (2)

where δTFWHM is the full width at half maxima of −ΔSM for a specific value of applied magnetic field. Values of −ΔSM (at TC), δTFWHM, RCP, and −ΔS'M (at low temperature below T2) were calculated for all DPs for ΔH = 8 T and are given in table 2.

Table 2. Theoretically estimated effective magnetic moment of rare earth ions (m), −ΔSM at TC, δTFWHM, relative cooling power (RCP), −ΔS'M at T2 for all DPs R2NiMnO6 for ΔH = 8 T.

R2NiMnO6 m (μB)−ΔSM (J kg−1 K−1) δTFWHM RCP (J kg−1)−ΔS'M (J kg−1 K−1)
Dy2NiMnO6 10.635.8267389.4714.31
Tb2NiMnO6 9.725.6371400.4112.59
Gd2NiMnO6 7.945.1672371.0037.18
Nd2NiMnO6 3.622.95100295.002.50
Pr2NiMnO6 3.582.67109291.430.94
Sm2NiMnO6 0.853.7374276.621.34

Figure 4 shows the evolution of TC and −ΔSM at TC with the ionic radius of R3+ in R2NiMnO6. We find that the ferromagnetic phase transition temperature TC increases linearly and −ΔSM at TC decreases monotonically with the ionic size of R3+ for all R2NiMnO6. The decrease in TC in going from Pr to Dy is understood to arise from a decrease in the Ni–O–Mn bond angle which leads to a decrease of the super-exchange strength [39]. Additionally from table 2 we observe that the paramagnetic background of rare earth ions spreads the magnetic entropy over a large temperature range across TC.

Figure 4.

Figure 4. Ferromagnetic transition temperature (TC) and magnetic entropy change (−ΔSM (T)) as a function of the ionic radii of the rare earth ion for all the R2NiMnO6 compounds.

Standard image High-resolution image

3. Theoretical results

3.1. Heisenberg model for R2NiMnO6

The experimental data discussed above presents intriguing magnetic behavior for Ni–Mn DPs. What is particularly interesting is the dramatically different magnetic response for different R at low temperatures. In order to comprehend the experimental observations, we propose a simple Heisenberg model on a body-centered cubic (BCC) lattice that takes into account spin degrees of freedom on B sites and both spin and orbital degrees of freedom on R site. The model is specified by the Hamiltonian,

Equation (3)

In the above, Si with the appropriate superscript denotes the spin of the relevant Ni, Mn or R ions, and ${\mathbf{L}}_{i}^{R}$ denotes the orbital angular momentum on the rare earth ion. The model explicitly considers two coupled sublattices A and B. Spins on R3+ ions on sublattice A are coupled to those on sublattice B of Mn4+ and Ni2+. In the model, the first three terms have the summation over nearest neighbors (nn) as indicated by angular brackets. i ∈ A (B) represents lattice sites on A (B) sublattices. J1 is the ferromagnetic coupling between Mn4+ and Ni2+ and ${J}_{2}({J}_{2}^{\prime })$ is the ferromagnetic coupling of R3+ to Mn4+ (Ni2+). The strength of spin–orbit coupling in rare earth ions is represented by λ. H symbolizes the strength of uniform external field which couples to the z components of the spin and orbital magnetic moments. We use gs = 2 and gL = 1 as the spin and orbital g-factors. In the BCC lattice, Mn4+ $(S=\frac{3}{2})$ couples to 6 Ni2+ (S = 1) and vice versa. Ni2+ couples to 8 R3+ and similarly coordination number of Mn4+ with R3+ is 8. J1 = 1 sets the energy scale for the model. For simplicity we assume ${J}_{2}^{\prime }={J}_{2}$, and different R ions are parameterized by different choices of J2 and λ, with the choice J1J2, λ inspired by the experimental data.

3.2. Methodology

Mean-field approximation. The approximation proceeds by rewriting the model as a sum of three effective single-spin Hamiltonians as,

Equation (4)

The effective magnetic field or the molecular field for the spins on different inequivalent sites is given by,

Equation (5)

In the above, we replace each spin operator SNi(Mn) by its average value $\langle {\mathbf{S}}_{i}^{\text{Ni}(\mathrm{M}\mathrm{n})}\rangle $ in the mean-field spirit, and the sublattice-resolved magnetization is then given by, MNi(Mn) = gs SNi(Mn)⟩. The subscript in Ms is necessary for the R sites in order to differentiate between spin and orbital contributions. Similarly we find,

Equation (6)

In the above equation, we used ${M}_{L}^{R}={g}_{L}\langle {L}_{R}\rangle $. Also,

Equation (7)

Following the textbook procedure of writing statistical average ⟨S⟩ in terms of the partition function, we get magnetization on different sublattices as [38],

Equation (8)

with $x=\frac{{g}_{s}{S}^{i}{H}_{\mathrm{e}\mathrm{f}\mathrm{f}}^{{S}^{i}}}{T}$ and $y=\frac{{g}_{L}{L}^{R}{H}_{\mathrm{e}\mathrm{f}\mathrm{f}}^{{L}^{R}}}{T}$, where Si and LR take the maximum value of the respective spin and orbital angular momenta and the Brillouin function is given by,

Equation (9)

This leads to a system of coupled equations which are then solved self-consistently. The total magnetization is calculated as, $M={M}^{\mathrm{M}\mathrm{n}}+{M}^{\mathrm{N}\mathrm{i}}+{M}_{s}^{R}+{M}_{L}^{R}$. We also determine the change in entropy (ΔSM ) as a function of temperature via,

Equation (10)

ΔSM (Ti , H) is thus the change in magnetic entropy at a certain discrete value Ti of temperature. Hj arises from uniform discretization of interval [0, H] such that, H1 = 0 and Hp = H. The above equation is the discrete version of continum equation, equation (1).

Classical Monte Carlo method. Since the mean-field approximations discussed above completely ignores spatial correlations, we use Monte-Carlo simulations to ensure that the low-temperature magnetic behavior is correctly captured. We employ the standard Markov chain Monte Carlo method with the Metropolis algorithm [40]. Typical simulated lattice size is 2 × 163 sites, with periodic boundary conditions. For thermal equilibration and averaging of quantities, we use $\sim 1{0}^{5}$ Monte Carlo steps each. Alongside, the change in entropy (ΔSM ) as a function of temperature (equation (10)), the z component of total magnetization is calculated as,

Equation (11)

where N is total number of spins and the angular bracket denotes the thermal average over Monte Carlo generated equilibrium configurations.

3.3. Results and discussions

We begin with the results obtained from the classical Monte Carlo simulations using Hamiltonian equation (3). First, we discuss the variation of magnetization (Mz ) with temperature at various values of magnetic field strength (H). In figure 5 we find that for all choices of R, there is a smooth rise in Mz as the temperature is lowered. This indicates the presence of a second-order phase transition from paramagnetic to ferromagnetic state below a critical temperature (TC ≈ 4). The transition is due to ferromagnetic coupling between the magnetic moments of Ni–Mn network in the paramagnetic background of rare earth moments [33]. The interesting behavior is noted at low temperature where Mz shows an upturn or a downturn depending on the choice of R. The experimental data clearly indicates an upturn for R = Gd, Tb, Dy, and a weak downturn for R = Sm, Pr, Nd. Since Gd does not support any orbital contribution to magnetism, the upturn in magnetization for Gd confirms that the coupling J2 is ferromagnetic. The sign of the coupling between spin and orbital moments on R ions, however, depends on the filling of f-shells. Indeed, in accordance with the Hund's rule, the orbital and spin moments will antialign (align) for less (more) than half filled band. If this argument borrowed from atomic physics holds then we expect that a simple change in sign of λ should capture the experimental behavior of magnetization across the entire rare earth series. Indeed, using the magnitude of ferromagnetic J2 and λ as model parameters, our Monte Carlo results describe the experimental data very well. For R = Sm, Pr, Nd, at lower values of H, we notice a downturn in magnetization much below TC (see figures 5(a), (c) and (d)). The downturn at low temperatures indicates an antiferromagnetic correlation of rare earth moments with Ni–Mn moments. However, the origin of this distinct behavior lies in the spin–orbit coupling. Once the applied field becomes stronger than the antiferromagnetic tendency, magnetization begins to show an upturn. The simulations for R = Gd, Tb, Dy are also consistent with the experiments, and display an upturn (see figures 5(b), (e) and (f)).

Figure 5.

Figure 5. Monte Carlo results for magnetization (Mz ) as function of temperature (T) at various strengths of external field for SR and LR values corresponding to (a) R = Pr, (b) R = Nd, (c) R = Sm, (d) R = Gd, (e) R = Tb, and (f) R = Dy. We set J1 = −1.0 as the energy scale. J2 and λ are model parameters.

Standard image High-resolution image

We have seen how the Monte Carlo data resembles the experimental data for all choices of R by simply tuning the values of λ and J2. We clarify further the underlying key idea by even simpler mean-field calculations described in the methods section. In order to emphasize the essential features, we set J1 = 1 and J2 = 0.1 for the mean field calculations, and use LR = 6 corresponding to R = Nd. The low-temperature behavior of magnetization obtained within mean-field calculations (see figures 6(a) and (b)) confirms that the Hund's rule controlling the antialignment or alignment of spin and orbital moments on the rare-earth ion can explain the qualitatively distinct behavior for different R. Also from figure 6(c), we find that for all values of λ, R spin moments begin to align ferromagnetically with the Ni–Mn sublattice as the temperature is lowered starting at TC. On the other hand, the orbital moments on R remain disordered until the temperature scale becomes lower than the energy scale of the spin orbit coupling λ. For a positive sign of λ, the orbital and spin moments on R antialign (see figure 6(d)). Similar mean-field results are obtained by using the LR values corresponding to Sm and Pr (not shown). As mentioned earlier, the AFM spin–orbit interaction on R ions can be justified on the basis of Hund's rule. The R+3 ions with R = Sm, Pr, Nd have 4f subshells which are less than half-filled. According to Hund's rule, if the subshell is less than half-filled, then the stable configuration has the minimum total angular momentum (J) value. The minimum J is possible only if the orbital moment of rare-earth ion aligns oppositely to the spin magnetic moment.

Figure 6.

Figure 6. Mean-field results for LR = 6 corresponding to Nd at H = 0.02. (a) and (b) The variation of magnetization (Mz ) as function of temperature (T) for different strengths and signs of spin–orbit coupling (λ). (c) and (d) Temperature dependence of spin and orbital magnetization on R sublattice for positive λ. The results are obtained for J1 = 1 and J2 = 0.1.

Standard image High-resolution image

The low temperature downturn in Mz , for instance in Nd2NiMnO6 can be understood as follows. The orbital moment of Nd+3, L = 6, is much larger than the corresponding spin value. As the orbital moment and the spin moment of Nd+3 are oppositely aligned, the total moment is in the direction of orbital moment. Additionally, the Nd spin moments on A sublattice is ferromagnetically coupled to Ni–Mn spin moments on B sublattice. Therefore, the total magnetic moment of the Nd sublattice is antialigned to the total magnetic moment of the Ni–Mn sublattice leading to a decrease in magnetization at low temperatures. With increasing magnetic field the Hund's rule energy is compensated by the external field leading to an upturn in Mz . Similar physics persists in Sm2NiMnO6 and Pr2NiMnO6 DP. On the other hand, in DP with R = Tb, Dy we find that ferromagnetic coupling between the rare earth orbital moment and the spin moment captures the upturn in Mz in the low-temperature regime. The alignment of spin and orbital contributions to magnetization in case of Tb and Dy is again attributed to Hund's rules. As the 4f ions have more than half-filled f orbitals so the minimum energy state must have maximum J. The maximum J is obtained when the orbital moment is parallel to the spin magnetic moment of rare-earth ion. The upturn can thus be related to the large total magnetic moment of rare-earth ion on A sublattice coupling ferromagnetically to B sublattice. R = Gd represents a special case where the orbital contribution does not exist, and therefore it also serves as a checkpoint for the choice of sign of J2 used in our model. Our mean-field results on Gd2NiMnO6 are consistent with a previous mean-field study on this material [39].

From the above discussion, we conclude that the low temperature behavior in DPs with R = Sm, Nd, Gd, Tb, Dy observed in our experiments is qualitatively captured via a minimal model that explicitly considered the orbital degree of freedom on R sites. The case of R = Pr presents a slight disagreement as we do not find a downturn in Mz at low temperatures in our low-field data. One possibility for this could be a smaller spin–orbit coupling in R = Pr that can be easily overcome even by a small magnetic field.

Further, we present results of change in entropy as a function of temperature obtained via our Monte-Carlo simulations. The results are shown in figure 7. Given that ΔSM is a quantity derived from Mz (H, T), it is not surprising that except for R = Pr the results match well with the data reported in our experiments. We obtain −ΔSM > 0 near the ferromagnetic transition of Ni–Mn sublattice for all choices of R. The anti-alignment (alignment) of the total magnetic moment of Nd sublattice to the magnetic moment of Ni–Mn sublattice at low temperatures leads to −ΔSM < (>)0 as seen in figures 7(a)–(c) (figure 7(d)–(f)) [37]. In case of antialignment, external magnetic field can enforce an alignment and hence ΔSM also shows a sign reversal at sufficiently large magnetic field (see, for example, figure 7(b)). As a result, the IMCE at low temperatures changes to conventional MCE upon increasing the magnetic field strength.

Figure 7.

Figure 7. Monte Carlo simulation results for change in magnetic entropy (−ΔSM ) as function of temperature (T) at various strengths of external field for SR and LR values corresponding to (a) R = Pr, (b) R = Nd, (c) R = Sm, (d) R = Gd, (e) R = Tb, and (f) R = Dy.

Standard image High-resolution image

From figure 7 it can be seen that for R = Gd, Tb, Dy there exist two peaks with −ΔSM > 0, one at TC and other at low temperatures. Upon increasing magnetic field, both the peaks display characteristic features associated with convensional MCE, such as the broadening of temperature range and increase in the peak height. These observations are in accordance with the experimental findings.

4. Conclusions

In conclusion, we have presented a detailed experimental and theoretical study of the magnetization and MCE of rare-earth based DP R2NiMnO6 (R = Pr, Nd, Sm, Gd, Tb, and Dy) as a function of temperature (2 ⩽ T ⩽ 300 K) and magnetic field (1 ⩽ H ⩽ 8 T). In contrast to Y2NiMnO6, which shows a single anomaly at the ferromagnetic transition temperature TC caused by the ordering of Ni2+ and Mn4+ magnetic ions, all the R2NiMnO6 show anomalies in magnetization and MCE at TC, where the Ni–Mn sublattice orders ferromagnetically, as well as additional features in the magnetization and MCE at low temperature (T2) because of the coupling of R3+ ions and ordered Ni–Mn sublattice. Pr2NiMnO6, Nd2NiMnO6 and Sm2NiMnO6 show a reduction in magnetization below T2 which occurs inspite of the ferromagnetic coupling between the R spins and the Ni/Mn spins. This can be understood as the orbital moment is anti-aligned to the spin for the first half of the R elements. This decrease in magnetization below T2 can be overcome by a large field strong enough to polarize the full effective moment of the R3+ ions. This leads to a negative value of (−ΔSM ) in MCE measurements at low temperature. Just like the magnetization, the reduction of the MCE can be reversed on the application of a large enough magnetic field. Materials from the second half of the R series, R2NiMnO6 (R = Gd, Tb, and Dy) show an upturn in the magnetization below T2 because the orbital and spin moments of these R ions are aligned with the Ni–Mn sublattice. The large rare earth moments of these heavier R ions leads to a huge and positive value of (–ΔSM ) at low temperature, which may potentially be useful in magnetic refrigeration. Additionally, the paramagnetic background of rare-earth ions spreads out the magnetic entropy over a larger temperature range around TC which broadens the MCE profile at TC resulting in a large RCP. The large RCPs make these materials potentially attractive for MCE based refrigeration techniques. For the microscopic understanding of the magnetic behavior, we investigated a phenomenological two-sublattice Heisenberg model that takes into account both spin and orbital degrees of freedom. To study the model we use mean-field theory and classical Monte Carlo simulations. Both these methods reproduce the main features of the experimental observation for magnetization and for magnetic entropy change. We find that the magnetic behavior at low temperatures in the DP containing f block elements is controlled by spin–orbit coupling. The more than or less than half-filled 4f orbitals decides the nature of spin–orbit coupling to be ferromagnetic or antiferromagnetic. The weak ferromagnetic coupling J2 between the rare earth ion sublattice and Ni–Mn sublattice and antiferromagnetic spin–orbit coupling (λ) in DPs with R = Nd, Pr and Sm is responsible for the downturn in magnetization at low temperature. On the other hand, the weak ferromagnetic coupling J2 and ferromagnetic λ in DPs with R = Gd, Tb and Dy explains the upturn in magnetization at low temperatures.

Acknowledgments

We acknowledge the support of the x-ray facility at IISER Mohali for powder XRD measurements. KP thanks DST, Government of India, for the award of Inspire faculty fellowship.

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

Please wait… references are loading.
10.1088/1361-648X/ac3e9e