MATERIALS TRANSACTIONS
Online ISSN : 1347-5320
Print ISSN : 1345-9678
ISSN-L : 1345-9678
Engineering Materials and Their Applications
Magnetic and Magnetocaloric Properties of SrRu1−xFexO3
T.A. HoP.D. ThangN.T. DangW.B.K. PutriT.L. PhanS.C. Yu
Author information
JOURNAL FREE ACCESS FULL-TEXT HTML

2022 Volume 63 Issue 10 Pages 1489-1495

Details
Abstract

We have systematically investigated the crystalline and electronic structures, magnetic and magnetocaloric properties of polycrystalline SrRu1−xFexO3 (x = 0 and 0.1) samples fabricated by the solid-state reaction method. The X-ray diffraction analyses of the samples indicated single phase orthorhombic perovskite structure. A thorough analysis of X-ray-absorption-based electronic structure revealed that both Fe2+ and Fe3+ ions are present in x = 0.1, in which concentration of Fe3+ ions is higher than that of Fe2+ ions. More careful analyses on the isothermal magnetization data derived from the Banerjee’s criterion demonstrated that the ferromagnetic-paramagnetic phase transition in all samples belongs to the second-order phase transition type. These results were also confirmed by a recently proposed quantitative criterion, which considered an exponent n from the magnetic field and temperature dependences of the magnetic entropy change (ΔSm). Especially, around TC, we have found that the ΔSm reaches the maximum values of 1.65 and 1.32 J.kg−1.K−1 for x = 0 and 0.1, respectively, for a field change of ΔH = 50 kOe. Magnetic-field dependences of the maximum magnetic-entropy change (ΔSmax) obey a power law of ΔSmax(H) ∝ Hn, where the values of n = 0.92–0.94 are far from the mean-field-theory value (2/3), indicating short-range magnetic order existing in the samples.

Fig. 4 (a), (b) M(H) data, (c), (d) Arrott plots of M2 vs. H/M, and (e), (f) −ΔSm(T) curves at various ΔH values for SrRu1−xFexO3. Arrow lines indicate the direction of increasing H value.

1. Introduction

Currently, divalent alkaline-earth ruthenates with the general formula of ARuO3 (A = Ca, Sr, Ba) are still sparking rigorous spotlight from the solid-state-physics community because of their peculiar electronic and magnetic properties relying upon the radius of the alkali metal ion.18) For BaRuO3, its ferromagnetic order is suppressed under pressure of 8 GPa, and fluctuations in the vicinity of the Curie temperature (TC) increase as changing pressure.2) While CaRuO3 is a paramagnetic conductor, SrRuO3 is a 4d ferromagnetic conductor with a Curie temperature TC = 160 K, despite the fact that both compounds crystallize in the orthorhombic structure (Pbnm space group).3) The magnetic-order difference in these compounds is ascribed to the effect of A2+ ion on the relative populations of the spin-up and spin-down bands and on the widths of the 4d bands of (RuO3)2−.9) The 4d bands in CaRuO3 is narrower than those in SrRuO3 because there are more covalent characters in the Ca–O bond than in the Sr–O bond. With a high chemical stability and rich variety of electronic and magnetic features, SrRuO3 has been vastly explored fundamentally and practically.1,3,4,1013) Furthermore, it is previously pinpointed that at 125 kOe and 4.2 K, the magnetic moment of SrRuO3 is about 1.55 μB/mol. The magnetization of this compound seems unsaturated even at high fields, indicating the very narrow 4d band; specifically when it is being compared to a 3d band of iron3,9) and to a large magnetocrystalline anisotropy which was caused by a strong spin-orbit interaction.1) Ferromagnetism in SrRuO3 is still regarded to have an obscure origin since the saturation moment is less than the magnetic moment (2 μB/mol) presumed for the localized t2g low-spin configuration of Ru4+.14,15) Earlier reports suggested that the ferromagnetism of SrRuO3 is vagrant,10) and also the compound itself can be viewed as a bad metal.16) Moreover, SrRuO3 thin film’s magnetic critical behavior is affiliated to the 3D Ising model, and it is in accordance with experimentally examined high uniaxial anisotropy which lingers above TC.11,17) The transport and magnetic properties of SrRuO3 can be adjusted via partial substitutions at Sr and Ru sites with various cations. Interestingly, Mieville et al.18) exposed that the chemical substitution on the two separate cation sites in SrRuO3 could produce various results. The ferromagnetic behavior prevails with up to 65% substitution of Ca2+ for Sr2+ while this behavior was destroyed with 20% substitution of Ti4+ for Ru4+. Pi et al.19) revealed that when Zn2+, Ni2+, Co2+ and Mn3+ were doped into the Ru site of SrRuO3, TC was notably decreased from 166 K down to 105 K. On the contrary, TC was increased up to 188 K with the Cr3+ doping. Additionally, the resistivity of this group of oxides is altered due to the deviation of the oxygen stoichiometric ratio at the grain boundary. Biswas et al.10) inspected the magnetic and electronic behaviors of SrRu1−xIrxO3 (0 ≤ x ≤ 0.25) and found that all samples showed ferromagnetic order and a fully-metallic behavior, despite the increased resistivity and the decreased TC. They concluded that the sample with x = 0.25 had gone through a metal-to-insulator transition at 75 K and a spin-glass-like behavior at 45 K along with the omission of ferromagnetic long-range ordering.

The Fe-doping effect on the transport and magnetic behaviors of SrRu1−xFexO3 was first reported by Gibb et al.20) For x values up to 0.2, a less significant influence on the magnetic and electronic properties was detected, despite the increasing level of oxygen deficiency. With further increasing of x up to 0.3, the structural transformation can be identified by the X-ray diffraction study. For x = 0.3, the sample is at an intermediate between a localized-electron system and a collective-electron system. The conflicting exchange mechanism in the system created intricate magnetic interactions. As x > 0.3, the decreased oxygen deficiency was generated from the oxidation of Ru4+ to Ru5+ states. Bansal et al.21) also has thoroughly inspected the alteration in the transport and magnetic behaviors of (SrFe)1+x(RuFe)1+xO3+δ by substituting Fe at both alkaline earth and transition-metal sites to reduce oxygen deficiency. They discovered that the conduction behavior had transformed from the metallic behavior (x ≤ 0.1) to the non-metallic one (x > 0.26). The minimum resistivity was observed as a function of temperature for 0.1 < x < 0.26. A transition of the ferromagnetic-to-paramagnetic state was observed as x ≥ 0.3. These results could be associated with the initial small mean free path of the system of this bad metal, and are also likely connected to the substitution of Fe3+ for Ru4+.

Apart from studying the magnetic and transport properties of 4d transition ferromagnets of pure and doped SrRuO3, a very few reports on their magnetocaloric (MC) properties have also been found, with some anomalous characters. Zhang et al.12) analyzed the MC effect of SrRuO3 under an applied field of 65 kOe and acquired the magnetic entropy change (ΔSm) of 2.5 J/kg K and the adiabatic temperature change (ΔTad) of 3.1 K at 160 K. Conversely, Sarkar et al.1) attained both normal and inverse MC effects in SrRuO3 co-doped with Ba and Ti. They concluded that the paramagnetic-ferromagnetic transition along with the magnetization reversal are the reasons leading to this phenomenon. Furthermore, the large magnetocrystalline anisotropy, the coupling between local and itinerant moments, and the random orientation of easy axis are perceived as the key factors to inquire some of the intriguing magnetic and MC properties of this system. To gain more knowledge about this material system, in the present work, we show a systematical study on the structure, oxidation state of Fe, magnetization, and MC effect of polycrystalline SrRu1−xFexO3 (x = 0 and 0.1) compounds.

2. Experimental Details

Polycrystalline SrRu1−xFexO3 (x = 0 and 0.1) perovskite-type samples were prepared by the ceramic technique in air, using Sigma-Aldrich chemicals of SrCO3, RuO2 and Fe2O3 (purity: 99.9%) as starting materials. These well-mixed powders according to nominal masses of SrRu1−xFexO3 were pre-annealed at 850°C for 12 h. Following several times of intermediate grinding alongside the pre-annealing at temperatures below 1000°C, the pre-annealed samples were pressed into pellets and sintered at 1120°C for 12 h. After annealing, and cooling down to room temperature, the crystalline structure of the acquired samples was examined by an X-ray diffractometer (D8 Discover, Bruker) equipped with a Cu-Kα radiation source. The Rietveld refinement was also employed to analyze collected X-ray diffraction data. The oxidation number of Fe was analyzed by using X-ray absorption near edge structure (XANES) at the K-edge. The magnetic and MC properties were studied by means of a superconducting quantum interference device (SQUID) working in the temperature and magnetic-field ranges of T = 15–250 K and H = 0–50 kOe, respectively. Here, the samples in powder were used for magnetization measurements.

3. Results and Discussion

Figures 1(a), (b) present powder X-ray diffraction (XRD) patterns of SrRu1−xFexO3 (x = 0 and 0.1). The data analysis has revealed that both samples are single-phase and have crystalized in an orthorhombic perovskite-type structure (the space-group Pnma) with the unit-cell parameters of ac$a_{p}\sqrt{2} $ and b ≈ 2ap (ap is defined as the lattice parameter of a pseudo-cubic perovskite unit cell), as represented in Fig. 1(c). The refined values of the lattice parameters and unit-cell volume are a = 5.564(4) Å, b = 7.840(8) Å, c = 5.526(4) Å and V = 241.113 Å3 for x = 0, and a = 5.571(5) Å, b = 7.848(8) Å, c = 5.532(6) Å and V = 241.90 Å3 for x = 0.1. The observation of the Fe-doping induced lattice expansion of SrRuO3 contrasts with the shrinkage of the lattice at the same Fe doping level previously reported by Fan et al.11) This discrepancy can be associated with different tendencies towards the occupation of the Fe ion. Indeed, Felner et al.22) observed that the Fe ions exhibit a preference to occupy in the Sr site instead of in the Ru site, resulting in the reduced lattice parameters because of the smaller size of Fe ion compared to the Sr2+ one. Therefore, in our case, the lattice expansion indicates the preferred occupation of Fe ions at the Ru site.

Fig. 1

XRD patterns of SrRu1−xFexO3 with (a) x = 0 and (b) x = 0.1 analyzed using the Rietveld refinement. Experimental points (red symbols) and refined curves (blue lines) are shown. Vertical blue ticks denote the diffraction reflections of the orthorhombic phase. A representation of the orthorhombic Pnma structure is graphed in (c).

Together with the XRD analysis, we have also based on the XANES spectra to identify the oxidation number of Fe in SrRu1−xFexO3 (x = 0.1). According to Teo,23) the absorption-edge position of an investigating sample in comparison with that of standard samples indicates the information about the oxidation number of an absorbing atom. Therefore, the XANES spectra of FeCl2 and CaFeO2.5 containing Fe2+ and Fe3+, respectively, have been recorded as the standard samples. The XANES spectra of all investigating samples shown in Fig. 2(a) indicate a rapid increase in the intensity at 7120∼7130 eV, which is associated with the Fe K-edge absorption. The middle points (Emid) of the absorption edges are the maxima of the first derivative curves of the corresponding samples. Figure 2(b) shows that the Emid values are about 7122.7, 7127.5 and 7128.8 eV for FeCl2, SrRu1−xFexO3 and CaFeO2.5, respectively. The Emid value of SrRu1−xFexO3 occupies between the values of FeCl2 and CaFeO2.5, but close toward that of CaFeO2.5. This proves that both Fe2+ and Fe3+ ions are present in SrRu1−xFexO3 (x = 0.1), and concentration of Fe3+ is higher than that of Fe2+ ions with an estimated Fe3+/Fe2+ value of 8/2. It is known that parent SrRuO3 contains Ru4+ ions and the partial replacement of Fe3+ and Fe2+ for Ru4+ could lead to the appearance of Ru5+ due to charge disproportionation of Fe3+ + Ru4+ → Fe2+ + Ru5+. This observation is well consistent with a recent adjustment by Lei and co-workers24) that the combination of Fe3+ and Ru5+ ions is energetically favored over that of Fe4+ and Ru4+ ions. Such mixed valence state in SrRu1−xFexO3 influences magnetic and MC behaviors, as being shown below.

Fig. 2

(a) Fe K-edge XANES spectra and (b) their derivative curves of SrRu1−xFexO3 (x = 0.1) and standard samples FeCl2 (Fe2+) and CaFeO2.5 (Fe3+) in the energy range E = 7110∼7180 eV.

To investigate the magnetic properties of SrRu1−xFexO3, temperature dependences of zero-field-cooled (ZFC) and field-cooled (FC) magnetization, MZFC/FC(T), at an applied field of 100 Oe were measured, in which the entire measurements were carried out according to the heating direction. The results are displayed in Figs. 3(a), (b). For the sample with x = 0, its MZFC(T) curve exhibits a sharp cusp at T = 160 K, and below 160 K, the magnetization is slowly decreased with decreasing T. This cusp could be connected to cluster spin-glass behavior at temperatures below 160 K, and the cusp peak is considered as the freezing temperature. Below this temperature, spins freeze in one of metastable states located in the system.25) This behavior was also observed in (Ba, Ca)xSr1−xFe0.5Ru0.5O3,25) La0.7Ca0.3Mn1−xScxO3,26) and La0.7Ca0.3Mn1−xZnO3.27) Above 160 K, the magnetization is abruptly reduced to zero because the sample experiences a transition from the ferromagnetic (FM) to paramagnetic (PM) state. This phase transition occurs within a narrow temperature range. In contrast, for x = 0.1, in its MZFC(T) curve, no cusp was detected, and the M value decreases with increasing temperature, and the broadened FM-PM transition is observed. Notably, at T = 15 K, the MZFC(T) and MFC(T) curves display a substantial divergence of about 10 and 7 emu/g for x = 0 and 0.1, respectively. This feature has also been found in many perovskite-type compounds, which is related to the irreversible magnetization contribution,1,28) or to a large magnetic anisotropy induced by magnetic domains in the samples.11)

Fig. 3

(a), (b) MZFC(T) and MFC(T) curves of SrRu1−xFexO3 for H = 100 Oe; the insets in (a), (b) show the first derivative of dMFC(T)/dT and (c) χ−1(T) plots (open symbols) in the PM region; the solid lines are fitting curves to the Curie-Weiss law.

The value of TC obtained from the maximum of the first order deviation with respect to temperature of an MFC(T) curve (as shown in the inset of Figs. 3(a), (b)) are about 165 and 150 K for x = 0 and 0.1, respectively, indicating a decrease of TC with increasing x. TC of the un-doped sample is slightly higher than the TC values reported by Fan et al.11) and Zhang et al.,12) plausibly due to the discrepancies in sample preparation conditions, which could influence the oxidation number of Fe and oxygen deficiency. According to our XANES analysis, when Ru4+ in SrRuO3 was partially replaced by Fe3+, some Ru4+ ions were altered into Ru5+ ions and a small amount of Fe2+ also appeared to balance the valence in the system. Super-exchange interactions between these ions strongly compete with the pre-existing FM coupling of Ru4+–Ru4+.11) Consequently, the FM-PM transition temperature has been reduced as observed.

Temperature dependences of the inverse magnetic susceptibility, χ−1(T), are shown in Fig. 3(c) for all the investigated samples. The straight line in the PM part of each curve is fitted by the Curie-Weiss (CW) law of $\chi (T) = \frac{C}{T - \theta }$, where $\text{C} = \frac{N_{A}\mu_{\textit{eff}}^{2}}{3k_{B}}$ is the Curie constant, NA is Avogadro’s constant, kB is the Boltzmann constant, μeff is the effective magnetic moment, and θ is the CW temperature. The fit values obtained for $\mu_{\textit{eff}}^{\textit{exp}}$ and θ are 2.85μB and 164.7 K, respectively, for x = 0, and 4.29μB and 125 K, respectively, for x = 0.1. Additionally, the theoretically effective magnetic moment, $\mu_{\textit{eff}}^{\textit{cal}}$, of SrRu1−xFexO3 could be expressed as11) $\mu_{\textit{eff}}^{\textit{cal}} = \sqrt{ {(1 - 2x)\mu_{\textit{Ru}4 + }^{2} + x(0.8\mu_{\textit{Fe}3 + }^{2} + 0.2\mu_{\textit{Fe}2 + }^{2} - \mu_{\textit{Ru}5 + }^{2})}\mathstrut} $. Using the effective magnetic moments of Fe3+ (5.9μB), Fe2+ (4.9μB), Ru4+ (2.83μB) and Ru5+ (3.87μB), the calculated effective magnetic moment for x = 0 and 0.1 are 2.83μB and 2.86μB, respectively. It can be seen that TC ≈ θ for x = 0 indicating that this sample obeys the Curie-Weiss law well. A large discrepancy between TC and θ, $\mu_{\textit{eff}}^{\textit{exp}}$ and $\mu_{\textit{eff}}^{\textit{cal}}$ observed for x = 0.1 could be due to FM clusters in the PM region. Interestingly, for x = 0.1, there is a downturn deviation from the CW law in a temperature range from TC to the Griffiths temperature TG. As indicated by the arrow in Fig. 3(c), TG is defined as the temperature that χ−1(T) begins to drift from the linear behavior in the high-temperature PM state. This downturn in χ−1(T) is characteristic of the Griffiths phase,29) and described by an expression $\chi^{ - 1} \propto (T - T_{C}^{\textit{rand}})^{1 - \lambda }$, where $T_{C}^{\textit{rand}}$ is the random transition temperature and λ is a non-universal exponent with 0 < λ < 1. Fitting the χ−1(T) data above TC to this equation, the acquired value of λ is 0.39 and $ T_{C}^{\textit{rand}}$ is about 156 K. In the temperature regime between $ T_{C}^{\textit{rand}}$ and TG, the interaction between the FM and PM phases is dominant. This indicates that the system is not a perfect FM ordering or a complete PM behavior. Concerning the Griffiths behavior, its origin has been suggested to be from a quenched disorder owing to a structural distortion30) and/or arising from a mismatch of radius and valence states of doping ion to the A/Ru – site as a random dilution and/or a formation of FM clusters in the PM phase matrix.29) In this Fe doped sample, compared to x = 0 sample, there is a coexistence of FM and PM phase above TC. This may be related to the interaction of Ru–O–Ru, Ru–O–Fe, and Fe–O–Fe due to complicated exchange interactions like double-exchange and super-exchange interactions. Sarkar et al.31) also observed the Griffith phase in Sr1−xBaxRu1−xTixO3 as x ≥ 0.2, and assigned it to be associated with diamagnetic BaTiO3. Meanwhile, Gupta et al.32) found the Griffith phase in SrRu1−xTixO3 as x > 0.3. They suggested a Griffith singularity occurring as a result of site dilution, and depicted the appearance of small-sized FM clusters above TC. The cluster size decreases with increasing Ti content. The Griffith-phase signature has also been attested in Sr1−xCaxRuO3.33) The authors explained that the Ca substitution suppressed the Ru–O–Ru angle bond, and diluted the FM interaction between Ru ions, leading to the Griffith-phase-like behavior.

The nature of FM-PM transition and the MC effect of SrRu1−xFexO3 have been studied from a set of isothermal magnetization curves, M(H), that were measured at temperatures near TC with a 2 K interval. For all samples, typical curves are displayed in Figs. 4(a), (b). Prior to every magnetization measurement, the samples were initially heated to a room temperature and then cooled down to the targeted temperature without any applied field. This step is to ensure a perfect demagnetization. Further, each of the measured sample shows that the value of M monotonically decreases with increasing temperature at a given H; while at a given temperature, M increases with H. Notably, for all samples, even at H = 50 kOe, the magnetization saturation is still out of reach. This is because the lowest temperature has a value near the TC and because of the coexistence of FM and PM phases. As T increases, the nonlinear dependences of M(H) in the FM region transform into the linear ones corresponding to the PM state. Through Arrott plot (i.e., the M2 vs. H/M curves), we could assess the characteristic of the FM-PM phase transition,34) as presented in Figs. 4(c), (d). At high H values, the plots near TC are not a series of parallel straight lines, and there is no straight line at T = TC passing through the origin. These traits confirm short-range magnetic order, meaning that the magnetic interactions in the samples do not obey the mean-field theory proposed for long-range order. Based on Banerjee’s criteria,35) the particular traits of the FM-PM phase transition could be taken into account from the slope sign of the M2 vs. H/M curves. The magnetic transition is a second-order phase transition (SOPT) when the entire slope of M2 vs. H/M curves near TC exudes a positive slope. Inversely, the magnetic transition is a first-order phase transition (FOPT) when the slopes show a negative sign. The results shown in Figs. 4(c), (d) indicate the Arrott-plot curves having positive slopes. This proves a SOPT taking place in the samples.

Fig. 4

(a), (b) M(H) data, (c), (d) Arrott plots of M2 vs. H/M, and (e), (f) −ΔSm(T) curves at various ΔH values for SrRu1−xFexO3. Arrow lines indicate the direction of increasing H value.

The MC effect in SrRu1−xFexO3 has been studied through magnetic entropy change (ΔSm), which was assessed up to 50 kOe. The |ΔSm| was determined from an indirect calculation using the thermodynamic Maxwell’s relation (with zero initial magnetic field).36)   

\begin{equation} |\Delta S_{m}(T,H)| = \int_{0}^{H\textit{max}}\left(\frac{\partial M}{\partial T}\right)_{H}dH \end{equation} (1)
For point-to-point measurements, the above equation can be written as a summation of   
\begin{equation} |\Delta S_{m}| = \sum \frac{M_{i} - M_{i + 1}}{T_{i + 1} - T_{i}}H_{i} \end{equation} (2)
where Mi and Mi+1 are the M values measured at Ti and Ti+1 in a magnetic-field sweep of Hi, respectively. According to eq. (1), |ΔSm| value is a $\frac{\partial M}{\partial T}$ slope-dependent, inferring that during an abrupt magnetic transition, a large |ΔSm| could be obtained. Also, |ΔSm| is maximized around TC, where the rapid decays of the magnetization mostly occur.

The magnetic entropy change as a function of temperature for various applied field values can be seen in Figs. 4(e), (f). Because only one magnetic phase (PM-FM) transition is recorded in our samples, normal MC effect would be observed. Sarkar et al.1) reported that in SrRuO3 codoped with Ba and Ti, both normal and inverse MC effects coexist, due to the two magnetic phase transitions, namely a second-order PM-FM transition and a first-order FM-antiferromagnetic transition. As shown in Figs. 4(e), (f), |ΔSm| increases with increasing H and achieves the highest values near TC, as presumed from eq. (1). The maximum |ΔSm| values (denoted as |ΔSmax|) obtained for ΔH = 50 kOe are about 1.65 and 1.32 J.kg−1.K−1 for x = 0, and 0.1, respectively. This result indicates the suppressed the MC effect by Fe dopants, which is related to the dilution of Ru concentration. Here, the |ΔSmax| value of x = 0 is fairly consistent with previous report on the same compound,12) implying that the current sample is of eminent quality. Specially, Fig. 6 shows broad |ΔSm(T)| curves that results in a large relative cooling power (RCP), which is known as a one of the important parameters to evaluate the MC effect.22) It can be calculated by using an expression of RCP = |ΔSm| × δT, where δT is the full-width-at-half-maximum of a |ΔSm| curve.22) The obtained values are 114 J.kg−1 for x = 0 and 150 J.kg−1 for x = 0.1 under an applied field of ΔH = 50 kOe. Although the |ΔSmax| magnitude slightly decreased with increasing x, this disadvantage is offset by an increase in RCP values, which are increased by Fe doping. Having reviewed previous reports on typical (Sr, Ca)RuO3-based materials, as shown in Table 1, one can see that the |ΔSmax| values obtained in this work are higher than those obtained from (Ba, Ti)-codoped SrRuO3 and comparable to those from Bi0.4Ca0.6Mn0.8Ru0.2O3.

Table 1 Experiment values of TC, and MC-related parameters (|ΔSmax|, and RCP) for SrRu1−xFexO3 compounds compared with those for (Sr, Ca)RuO3-based materials.

As considering magnetic-field dependences of |ΔSmax|, we have found these data follow a power function of |ΔSmax| ∝ Hn, where H varies from 5 to 50 kOe, and n is an exponent affiliated to magnetic order, which is dependent on T and H.36) Fitting |ΔSmax| data to the power law, as shown in Fig. 5, yielded n values of approximately 0.97 and 0.92 for x = 0, and 0.1, respectively.

Fig. 5

Field dependences of ΔSmax for SrRu1−xFexO3. The data are fitted to the power law |ΔSmax| ∝ Hn and the fittings are indicated by the solid lines.

A recent publication by Law et al.36) described a new quantitative criterion to determine the order of magnetic phase transition. This is done by computing n(T, H) associated with the magnetic entropy change as follows:   

\begin{equation} n(T,H) = \frac{d\,\mathit{ln}|S_{m}|}{d\,\mathit{ln}\,H} \end{equation} (3)
Using the mentioned criteria for magnetic materials, at temperatures well below TC, n is supposed to be in the value tendency of 1, whereas n has the tendency of 2 when T has much higher value than TC. Near the TC, for materials with a FOPT, the peak is n > 2; while for materials with SOPT, these large values will be absent.37) These values of n should be acknowledged because the samples are in a multi-domain state in a small field. Thus, by employing eq. (3), the exponent n at an applied field above 10 kOe could be estimated, as seen in Fig. 6. Evidently, one can see a variation trend of 1 → minimum → 2 in the n(T) curves for all samples. The absent of large values n > 2 is the criterion to ascertain that the FM-PM transition in the samples is a SOPT type (according to the above quantitative criterion), in accordance with those obtained from Banerjee’s criteria. This corroborates the benefit of applying the quantitative criterion besides Banerjee’s criterion in determining the order of a phase transition. Moreover, the minimum n values at TC under an applied field of H = 50 kOe are 0.61 and 0.54 for x = 0 and 0.1, respectively. These values are contrasting with the mean-field-theory value of n = 2/3. This difference is due to the absence of long-range magnetic order, meaning the presence of magnetic inhomogeneity in the samples. It is well known that the replacement of Ru by Fe leads to the presence of Ru–O and Fe–O bond lengths together with three types of network arrangements of Ru–O–Ru, Ru–O–Fe, and Fe–O–Fe in a crystal lattice. Long-range magnetic Ru–O–Ru and Fe–O–Fe networks are partially destroyed by short-range magnetic Ru–O–Fe. Also, the substitution of Fe3+ and Fe2+ for Ru4+ created Ru5+ and led to an addition of interactions between Fe3+, Fe2+, Ru4+ and Ru5+ ions. These interactions compete with the intrinsic FM coupling of Ru4+–Ru4+. Thus, it is possible to present some antiferromagnetism11) which results in the decrease of the TC, and |ΔSmax| as increasing Fe doping concentration in SrRu1−xFexO3.

Fig. 6

n(T, H) curves of the SrRu1−xFexO3 samples for H = 10 and 50 kOe.

4. Conclusions

We studied the structural characterization, oxidation state of Fe, the magnetic and magnetocaloric properties of SrRu1−xFexO3 (x = 0, and 0.1). The Fe-doping induced lattice expansion of SrRuO3 that is assigned to the appearance of Fe3+ and Fe2+ with estimated value of Fe3+/Fe2+ is about 8/2 replacing partially for Ru4+ ions, leading to the appearance of Ru5+. The anti-FM interactions associated with Fe3+/Fe2+–Ru4+/Ru5+ pairs compete with FM interactions associated with Ru4+–Ru4+ pairs cause short-range magnetic order, and hence reduce the TC as increasing Fe doping concentration in SrRu1−xFexO3. Detailed inspections in connection to the CW law and Griffiths phase for the M(T) data disclosed the coexistence of FM, anti-ferromagnetic, and PM interactions at temperatures TC < T < TG in the Fe-doped sample. The analyses of the M(H) data using the Arrott plots and the quantitative criterion n(T) indicated that the SOPT occurs in the both samples. Investigations into the magnetocaloric effect indicated the Fe-doping caused the slightly reduced |ΔSmax|.

Acknowledgement

This research is supported by Vietnam National Foundation for Science and Technology Development (NAFOSTED) under grant number 103.02-2019.335, and the research at Korea was supported by the National Research Foundation of Korea Grant No. 2020R1A2C1008115.

REFERENCES
 
© 2022 The Japan Institute of Metals and Materials
feedback
Top