Magnetic characteristics, magnetocaloric effect, and critical behavior of Nd1−xSrxMnO3 compounds by Sr doping (x = 0.2, 0.3, 0.4, 0.5) were studied. All samples maintained orthorhombic structures, but the space group changed from Pnma (No. 62) for x = 0.2, 0.3 to Imma (No. 74) for x = 0.4, 0.5. As Sr doping increased, the Curie temperature (TC), Curie–Weiss temperature (TCW), and magnetization increased, attributed to the double exchange (DE) interaction. A discrepancy between TCW and TC was observed due to the competition between polarons and DE interaction. The critical behavior was investigated systematically using the self-consistent (modified Arrott plots, MAP) method and the Kouvel–Fisher (KF) relation. The KF relation was suitable for the samples with x = 0.2 and 0.5, while the MAP method was suitable for the samples with x = 0.3 and 0.4. Among the Ising, XY, Heisenberg, and mean-field models, the samples with x = 0.2, 0.3, and 0.4 aligned more closely with the mean-field model, except for the x = 0.5 sample. Entropy change (−ΔSM) of Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) increased with the applied field, with the maximum value observed around TC. For the sample with x = 0.3, (−ΔSM) reached 4.315 J/kg K at μ0ΔH = 50 kOe, corresponding to a relative cooling power (RCP) of 280.48 J/kg. Remarkably, the x = 0.4 sample displayed (−ΔSM) of 3.298 J/kg K at μ0ΔH = 50 kOe near room temperature, with the RCP of 283.64 J/kg. These findings underscore the role of Sr doping in tuning the magnetic properties, critical behavior, and magnetocaloric effect of NdMnO3.
I. INTRODUCTION
Perovskite manganites exhibit a diverse range of physical properties, including spin and orbital ordering,1 piezoelectric properties,2 ferroelectric behavior,3 ionic conduction,4 superconductivity,5 magnetoresistance (MR) effects,6 and photovoltaic effects.7 These properties are significant for both practical applications and fundamental science. In rare earth manganites RMnO3 (R representing trivalent rare earth elements), the magnetic and electrical properties are mainly influenced by the transition metal ions and their oxidation states. This includes phenomena such as charge-orbital ordering,8 spin-polarized transport,9 half-metallic ferromagnetism,10 and multiferroic behavior.11 Among RMnO3 compounds, NdMnO3 stands out as a unique material. Below its Curie temperature (TC) of 75 K, Nd and Mn magnetic moments exhibit ferromagnetic coupling.12,13 Both stoichiometric and non-stoichiometric NdMnO3 show various interesting physical and chemical properties, such as phase separation,14–20 positive and negative exchange bias,21 magnetoelastic effects,22 and negative magnetization with magnetization reversal.15–13,21–25 Sudakshina et al. found that the Néel temperature (TN) of NdMnO3 is 73.51 K, with negative magnetization observed below 2 K.23 Research by Hong et al. revealed that the magnetic moment in NdMnO3 becomes negative below 13 K, with the material exhibiting A-type antiferromagnetic (AFM) ordering and a ferromagnetic (FM) ordering temperature (TC) of 79 K.21 Kumar et al. observed Nd ordering at 15 K and A-type AFM ordering around 75 K in NdMnO3.24 Additionally, NdMnO3 displays a significant magnetocaloric effect, with a magnetic entropy change (−ΔSM) of 4.4 J/kg Sudakshina K near 15 K under an external field of μ0ΔH = 50 kOe.25 Al-Yahmadi et al. reported a maximum magnetic entropy change of 1.61 J/kg K at 62 K under the same field.26 Saravanan et al. discovered that under an external magnetic field of μ0ΔH = 50 kOe, the transition from paramagnetic (PM) to antiferromagnetic (AFM) occurs at 56 K, with magnetic entropy changes of 1.85 J/kg K at 60 K, increasing to 2.5 J/kg K at 15 K.27 Kumar and colleagues found that NdMnO3 undergoes transitions at 15 and 75 K. Under an applied magnetic field of μ0ΔH = 10 kOe at 15 K, the magnetic entropy change is −2 J/kg K, while under μ0ΔH = 20 kOe at 73 K, the magnetic entropy change is 0.7 J/kg K.24
Most studies on NdMnO3 have identified two transition temperatures related to magnetic entropy. However, our previous research on the magnetic entropy and critical behavior of NdMnO3 revealed three distinct maximum magnetic entropy change values, including both positive and negative changes.28 A maximum (−ΔSM) value of 3.82 J/kg⋅K was observed at μ0ΔH = 50 kOe and 10–15 K. In the 80–85 K temperature range, a smaller (−ΔSM) value of 1.22 J/kg K was obtained under an applied magnetic field of μ0ΔH = 50 kOe. Additionally, at μ0ΔH = 5 kOe and around 8 K, a maximum negative (−ΔSM) of −0.557 J/kg K was observed, attributed to the alignment of Mn moments and the ordering of Nd sublattices. Analysis of the critical behavior confirmed the reliability of the KF relation, with an estimated magnetic exchange distance J(r) of approximately 4.563, placing it between the three-dimensional Heisenberg model29 and the mean-field model.30 This behavior can be explained by the antiferromagnetic coupling between the ferromagnetic Mn components at low temperatures around 12 K, which induces the ordering of Nd. The ordered Nd moments cause a spin reorientation of Mn, resulting in magnetization reversal and negative magnetization. At around 53 K, the high-spin state of Mn involved in antiferromagnetic interactions becomes significant, while at around 82 K, the ferromagnetic Mn components exhibit ordering in the a–b plane of NdMnO3.
For RMnO3 compounds, doping at the R-site often leads to various intriguing physical properties in manganites.31–35 Currently, much attention has been focused on A-site substitution, where divalent cations replace trivalent rare-earth ions (R3+), generating Mn4+ ions due to charge compensation. Generally, RMnO3 exhibits antiferromagnetism and significant Jahn–Teller (J-T) distortion. However, R-site doping can induce double exchange (DE) interactions, reducing the J-T distortion and simultaneously introducing metallic and ferromagnetic characteristics.36 Near the PM insulator to FM metal transition, giant negative magnetoresistance (MR) and magnetic entropy changes are often observed. Doping and temperature variations alter the Mn3+/Mn4+ ion distribution, leading to changes in Mn—O bond lengths and causing structural disorder. This disorder is closely related to the static or dynamic properties of the lattice, orbital degrees of freedom, and charge, which remain topics of ongoing debate. Researchers are particularly interested in understanding which universal principles govern magnetic phase transitions and the mechanisms that normalize interactions near critical points. The critical phenomena of the FM to PM transition in DE systems can be studied using the mean-field method,30 the Ising model,26 and the 3D Heisenberg model.29 Among RMnO3 oxides, studies on Nd-based compounds are less extensive. The smaller Nd3+ ion leads to larger lattice distortions and weaker DE interactions, making Nd-based perovskite manganites particularly intriguing.37 In doped systems like Nd1−x(Ca, Sr)xMnO3, compared to La1−x(Ca, Sr)xMnO3, there is a heightened competition among electron–phonon, electron–electron, and DE interactions. This competition often accompanies pervasive instabilities such as antiferromagnetic superexchange, orbital ordering, and charge ordering.38 Therefore, doping at the Nd site in NdMnO3 can induce a richer array of magnetic properties, offering fertile ground for further investigation.
In our previous work,28 we investigated the spin reorientation and critical behavior of NdMnO3. In the current study, we examine the effect of Nd-site doping on its magnetic and magnetocaloric properties and further explore its critical behavior. Sr was chosen as the dopant at the Nd site due to the promising MR and magnetocaloric effects exhibited by Nd1−xSrxMnO3. Al-Yahmadi et al.26 found a maximum magnetic entropy change (−ΔSM) value of 2.78 J/kg K at 127 K with an applied field of μ0ΔH = 50 kOe and a relative cooling power (RCP) value of 155.5 J/kg in Nd0.8Sr0.2MnO3. Fkhar et al.39 reported a maximum (−ΔSM) of 3.12 J/kg K at 257.5 K and μ0ΔH = 10 kOe in Nd0.67Sr0.33MnO3. Similarly, Xu et al.40 observed a maximum (−ΔSM) of 3.12 J/kg K under μ0ΔH = 30 kOe at 274 K, with an RCP value of 71 J/kg in Nd0.55Sr0.45MnO3. Ahmed et al.41 studied the MR of Nd0.6Sr0.4MnO3 and found that MR could reach 12% at 273 K under a magnetic field of 0.6 T. These studies highlight the excellent potential of the Nd1−xSrxMnO3 system for magnetic sensor and refrigeration.
Therefore, we systematically study the doping effect of polycrystalline Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) with NdMnO3 as the parent compound, investigating the critical behavior through analysis of its magnetic and magnetocaloric effects. Our findings demonstrate that Sr doping in NdMnO3 (Nd1−xSrxMnO3) induces a ferromagnetic (FM) transition, resulting in significant magnetocaloric effects. A discrepancy between TCW and TC can be observed due to the competition between polarons and DE interaction, which is different from previous similar reports on NdMnO3 by Sr doping.26,40 The critical behavior analysis via using the self-consistent (modified Arrott plots, MAP) method and the Kouvel–Fisher (KF) relation reveals that Nd1−xSrxMnO3 samples with doping concentrations of x = 0.2, 0.3, and 0.4 adhere to the mean-field model. However, the sample with a doping concentration of x = 0.5 (Nd0.5Sr0.5MnO3) does not conform to any classical model due to the presence of a charge-ordered antiferromagnetic (AFM) phase. For the sample with x = 0.3, (−ΔSM) reached a maximum of 4.315 J/kg K at μ0ΔH = 50 kOe, corresponding to a relative cooling power (RCP) of 280.48 J/kg. Remarkably, the x = 0.4 sample displayed (−ΔSM) of 3.298 J/kg K at μ0ΔH = 50 kOe near room temperature, with the RCP of 283.64 J/kg. These findings underscore their potential for magnetic refrigeration applications, meanwhile highlight the important role of Sr doping in tuning the magnetic properties, critical behavior, and magnetocaloric effect of NdMnO3.
II. EXPERIMENTAL DETAILS
Polycrystalline Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) samples were prepared using the high-temperature solid-state reaction method. Initially, stoichiometric amounts of Nd2O3, SrCO3, and MnO2 powders (Aladdin) with a purity of 99.99% were thoroughly mixed and ground. The mixtures were then calcined at 1100 °C for 24 h. After furnace cooling, the obtained powders were ground again and subjected to a second calcination at 1200 °C for 24 h. The resultant powders were pressed into disc-shaped pellets with a diameter of approximately 1 cm and sintered continuously at 1300 °C for 24 h. All the samples were obtained after furnace cooling.
The crystal structure of the samples was analyzed using x-ray diffraction (XRD) with Cu-Kα radiation (wavelength: 1.5405 Å). The surface morphology of the samples was examined using a scanning electron microscope (SEM), and the elemental composition was analyzed with an energy dispersive spectrometer (EDS). For magnetic analysis, a physical property measurement system (PPMS) incorporating a vibrating sample magnetometer (VSM) was utilized. The relationship between magnetization and temperature (M-T) was examined using different procedures: field-cooled-warming (FCW), field-cooled-cooling (FCC), and zero-field-cooled (ZFC). These procedures were carried out under four distinct magnetic field strengths: 100, 500, 2000, and 10 000 Oe. Magnetization hysteresis loops (M-H) were recorded in the range from −50 to 50 kOe at various temperatures. Additionally, isothermal magnetization curves were measured across different temperature ranges to calculate the corresponding magnetic entropy change for each sample. Electrical transport measurements were performed using the PPMS. The temperature dependence of resistivity (R-T) was measured using the four-probe method under magnetic fields of 0, 1, 2, 4, and 6 T.
III. RESULTS AND DISCUSSION
A. XRD structural characterization
The room temperature XRD measurements and Rietveld refinement results for the Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) samples are shown in Figs. 1 and 2(a)–2(h). Figures 2(e)–2(h) display the crystal structures obtained after refinement, viewed from different orientations. The positions of XRD diffraction peaks for the x = 0.2 and x = 0.3 samples are almost consistent with the reference Nd0.7Sr0.3MnO3 (PDF No. 04-008-4443), confirming a single orthorhombic phase with the space group Pnma (No. 62). Similarly, the diffraction peak positions for the x = 0.4 and x = 0.5 samples align with the reference Nd0.5Sr0.5MnO3 (PDF No. 04-011-4383), confirming a single orthorhombic phase with the space group Imma (No. 74). Table I presents the structural parameters obtained from Rietveld refinement. The goodness-of-fit values (χ2) are below 1.5, indicating the reliability of the refinement results.
Parameters . | x = 0.2 . | x = 0.3 . | x = 0.4 . | x = 0.5 . |
---|---|---|---|---|
Symmetry | Orthorhombic | Orthorhombic | Orthorhombic | Orthorhombic |
Space group | Pnma | Pnma | Imma | Imma |
a (Å) | 5.502 086 | 5.451 941 | 5.432 781 | 5.432 627 |
b (Å) | 7.710 249 | 7.712 472 | 7.668 573 | 7.622 804 |
c (Å) | 5.458 090 | 5.457 115 | 5.467 898 | 5.465 116 |
V (Å3) | 231.546 | 229.460 | 227.802 | 226.321 |
Nd | (0.033, 0.25, −0.009) | (0.024, 0.25, −0.016) | (0.00, 0.25, 0.002) | (0.00, 0.25, 0.008) |
Sr | (4.556, 0.25, −8.465) | (0.032, 0.25, 0.011) | (0.00, 0.25, −0.005) | (0.00, 0.25, −0.021) |
Mn | (0.00, 0.00, 0.50) | (0.00, 0.00, 0.50) | (0.00, 0.00, 0.50) | (0.00, 0.00, 0.50) |
O1 | (0.467, 0.25, 0.097) | (0.499, 0.25, 0.013) | (0.00, 0.25, 0.44) | (0.00, 0.25, 0.418) |
O2 | (0.287, 0.03, −0.287) | (0.303, −0.01, −0.204) | (0.75, −0.015, 0.25) | (0.75, −0.016, 0.25) |
Mn-O1 (Å) | 2.007(4) | 1.92937(4) | 1.9450(15) | 1.958(5) |
Mn-O2 (Å) | 1.974(12) | 2.31260(4) | 1.9302(11) | 1.9305(7) |
1.970(12) | 1.54721(2) | − | − | |
θ(Mn-O1-Mn) | 147.7(8) | 175.879(0) | 160.6(5) | 153.5(11) |
θ(Mn-O2-Mn) | 158.5(6) | 104.028(1) | 173.4(12) | 172.6(6) |
Rwp | 0.0864 | 0.0695 | 0.0724 | 0.0780 |
Rp | 0.0625 | 0.0525 | 0.0557 | 0.0588 |
χ2 | 1.380 | 1.124 | 1.088 | 1.160 |
Parameters . | x = 0.2 . | x = 0.3 . | x = 0.4 . | x = 0.5 . |
---|---|---|---|---|
Symmetry | Orthorhombic | Orthorhombic | Orthorhombic | Orthorhombic |
Space group | Pnma | Pnma | Imma | Imma |
a (Å) | 5.502 086 | 5.451 941 | 5.432 781 | 5.432 627 |
b (Å) | 7.710 249 | 7.712 472 | 7.668 573 | 7.622 804 |
c (Å) | 5.458 090 | 5.457 115 | 5.467 898 | 5.465 116 |
V (Å3) | 231.546 | 229.460 | 227.802 | 226.321 |
Nd | (0.033, 0.25, −0.009) | (0.024, 0.25, −0.016) | (0.00, 0.25, 0.002) | (0.00, 0.25, 0.008) |
Sr | (4.556, 0.25, −8.465) | (0.032, 0.25, 0.011) | (0.00, 0.25, −0.005) | (0.00, 0.25, −0.021) |
Mn | (0.00, 0.00, 0.50) | (0.00, 0.00, 0.50) | (0.00, 0.00, 0.50) | (0.00, 0.00, 0.50) |
O1 | (0.467, 0.25, 0.097) | (0.499, 0.25, 0.013) | (0.00, 0.25, 0.44) | (0.00, 0.25, 0.418) |
O2 | (0.287, 0.03, −0.287) | (0.303, −0.01, −0.204) | (0.75, −0.015, 0.25) | (0.75, −0.016, 0.25) |
Mn-O1 (Å) | 2.007(4) | 1.92937(4) | 1.9450(15) | 1.958(5) |
Mn-O2 (Å) | 1.974(12) | 2.31260(4) | 1.9302(11) | 1.9305(7) |
1.970(12) | 1.54721(2) | − | − | |
θ(Mn-O1-Mn) | 147.7(8) | 175.879(0) | 160.6(5) | 153.5(11) |
θ(Mn-O2-Mn) | 158.5(6) | 104.028(1) | 173.4(12) | 172.6(6) |
Rwp | 0.0864 | 0.0695 | 0.0724 | 0.0780 |
Rp | 0.0625 | 0.0525 | 0.0557 | 0.0588 |
χ2 | 1.380 | 1.124 | 1.088 | 1.160 |
B. Morphological characterization and elemental analysis
Figures 3(a)–3(d) show the surface morphology of the Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) perovskites observed using SEM. The agglomerated particles are relatively small, densely distributed, and irregularly oriented. The insets in the top right corners display the average grain sizes, calculated to be 1.90 μm for the x = 0.2 sample, 1.52 μm for the x = 0.3 sample, 1.43 μm for the x = 0.4 sample, and 2.11 μm for the x = 0.5 sample. The left insets in Figs. 3(a)–3(d) reveal a uniform elemental distribution. The EDS spectra of the Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) samples were measured to study their compositions and the compositions are analyzed in Table II. The differences between the experimental and nominal elemental compositions are very small and can be neglected.
Element . | x = 0.2 . | x = 0.3 . | x = 0.4 . | x = 0.5 . |
---|---|---|---|---|
% Nd Exp | 16.35 ± 0.21 | 11.50 ± 0.18 | 10.04 ± 0.11 | 11.55 ± 0.15 |
% Sr Exp | 4.35 ± 0.17 | 5.04 ± 0.20 | 6.27 ± 0.14 | 9.16 ± 0.21 |
% Mn Exp | 22.31 ± 0.33 | 16.65 ± 0.29 | 19.61 ± 0.20 | 22.65 ± 0.29 |
% O Exp | 56.99 ± 0.73 | 66.82 ± 0.84 | 64.08 ± 0.56 | 56.64 ± 0.71 |
Total | 100.00 | 100.00 | 100.00 | 100.00 |
Element . | x = 0.2 . | x = 0.3 . | x = 0.4 . | x = 0.5 . |
---|---|---|---|---|
% Nd Exp | 16.35 ± 0.21 | 11.50 ± 0.18 | 10.04 ± 0.11 | 11.55 ± 0.15 |
% Sr Exp | 4.35 ± 0.17 | 5.04 ± 0.20 | 6.27 ± 0.14 | 9.16 ± 0.21 |
% Mn Exp | 22.31 ± 0.33 | 16.65 ± 0.29 | 19.61 ± 0.20 | 22.65 ± 0.29 |
% O Exp | 56.99 ± 0.73 | 66.82 ± 0.84 | 64.08 ± 0.56 | 56.64 ± 0.71 |
Total | 100.00 | 100.00 | 100.00 | 100.00 |
C. Magnetic properties
X . | C (emu K/mol Oe) . | TC (K) . | TCW (K) . | . | . |
---|---|---|---|---|---|
0.2 | 5.498 | 152 | 178 | 6.636 | 4.712 |
0.3 | 5.359 | 217 | 226 | 6.551 | 4.615 |
0.4 | 4.72 | 285 | 286 | 6.148 | 4.517 |
0.5 | 3.541 | 290 | 285 | 5.325 | 4.416 |
X . | C (emu K/mol Oe) . | TC (K) . | TCW (K) . | . | . |
---|---|---|---|---|---|
0.2 | 5.498 | 152 | 178 | 6.636 | 4.712 |
0.3 | 5.359 | 217 | 226 | 6.551 | 4.615 |
0.4 | 4.72 | 285 | 286 | 6.148 | 4.517 |
0.5 | 3.541 | 290 | 285 | 5.325 | 4.416 |
In the relation, g is the Landé g-factor. It is noteworthy that transition elements undergo orbital “freezing” when forming metals and compounds, meaning that the orbital magnetic moment is quenched by the crystal field and does not contribute to magnetism. In this case, L = 0, J = S, gJ = 2. Therefore, for Mn3+: J = S = 2, gJ = 2, and for Mn4+: J = S = 1.5, gJ = 2. The calculated values are listed in Table III.
In the undoped sample (x = 0), distinct antiferromagnetic properties are observed.28 Introducing Sr2 + alters the material's characteristics: as x increases, changes occur in the Curie temperature, net magnetization, and spontaneous magnetization, likely influenced by the evolution of double exchange (DE) interactions. With partial substitution of Nd by Sr (0 < x < 0.1), a canted spin structure emerges. As x slightly increases further, the system transitions to a ferromagnetic state while remaining insulating. For compositions where 0.2 ≤ x ≤ 0.48, the material exhibits a ferromagnetic metallic phase below TC and transforms into a paramagnetic insulating phase above this temperature.43 Optical studies on Nd0.7Sr0.3MnO3 thin films have provided evidence of dynamic Jahn–Teller effects and the presence of magnetic polarons.44,45 Additionally, chemical potential investigations by Ebata et al. have confirmed the existence of these polarons.46 Below TC, DE interactions dominate, whereas above TC, polarons become significant, contributing to the complex magnetic behaviors observed.
Figures 4(a)–4(d) depict the magnetization curves of Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) samples in the temperature range of 10–350 K under an external magnetic field of 100 Oe. At low temperatures, the divergence between ZFC and FC magnetization-temperature (M-T) curves suggests a spin glass state5,28 or other forms of magnetic inhomogeneity. As temperature decreases, a transition from PM to FM state becomes evident. The insets in Figs. 4(a)–4(d) show peaks in dM/dT under the FC mode around 152 K, indicative of ferromagnetic ordering induced by double exchange (Curie temperature, TC). Notably, Fig. 4(a) reveals a significant deviation of approximately 26 K between TC derived from the magnetization curve (dM/dT) and the Curie–Weiss temperature (TCW) obtained from fitting, possibly attributed to polaron effects above TC. A Griffith-like phase, observed in the x = 0.2 sample,5 designates the temperature as the Griffith temperature (TG). Figure 4(b) displays the magnetization curve for the x = 0.3 sample, showing a PM to FM transition with TC around 217 K. The inset of Fig. 4(b) confirms a peak in dM/dT under the FC mode at this temperature. The higher doping level (increased Mn4+ content) promotes double-exchange interactions between Mn3+-O-Mn4+, resulting in elevated TC and spontaneous magnetization compared to the x = 0.2 sample. Similar magnetic behavior persists, including spin glass characteristics at low temperatures, with TC differing from TCW by approximately 9 K. The Griffith-like phase is also present in the x = 0.3 sample. Figure 4(c) exhibits the magnetization curve for the x = 0.4 sample, demonstrating increased spontaneous magnetization and a higher TC. In its inset, dM/dT under the FC mode shows a peak around 285 K, approaching room temperature. According to Al-Yahmadi, substituting 40% of Nd3+ with Sr2+ establishes strong double-exchange interactions via Mn3+—O—Mn4+ bonds, achieving a nearly equal distribution of Mn3+ (δ3+) and Mn4+ (δ4+), resulting in a high Mn4+/Mn3+ ratio of 97%.26 Consequently, the x = 0.4 sample exhibits the highest spontaneous magnetization among the concentrations studied, with TC and TCW nearly identical, highlighting pronounced double-exchange interactions.
With Sr2+ doping reaching 50%, a decrease in net magnetization and TC is observed, attributed to the increase in antiferromagnetic superexchange interactions. Figure 4(d) shows the magnetization curve for the x = 0.5 sample, with the inset displaying dM/dT in the FC mode. A PM to FM transition occurs around 290 K. At the half-doped concentration of x = 0.5, as the temperature decreases from the PM to FM state and further, a charge-exchange-type antiferromagnetic charge-ordered phase emerges, reducing spontaneous magnetization.46
D. Magnetocaloric effect and critical behavior
Figures 5(a)–5(d) present the isothermal magnetization curves for Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) in the external magnetic field range of 0–50 000 Oe. The x = 0.2 sample is measured for magnetization at different temperatures from 100 to 250 K, the x = 0.3 sample from 175 to 300 K, and the x = 0.4 and x = 0.5 samples from 200 to 320 K, to evaluate potential magnetic entropy. Figures 6(a)–6(d) display Arrott diagrams showing the relationship between the square of the magnetization (M2) and H/M, providing insight into the nature of the phase transition. In the M2 range, a positive slope indicates a second-order phase transition under the applied magnetic field, according to the Banerjee criteria.47
Figures 8(a)–8(h) depict the scaled magnetization intensity and magnetic field data separated into two universal curves above and below TC. The results indicate that the x = 0.5 sample cannot be collapsed into two universal curves, possibly due to the presence of its charge-ordered superexchange antiferromagnetic phase. In contrast, other doping ratios show well-scaled curves. Specifically, the x = 0.2 sample demonstrates better scaling and a clearer collapse into two universal curves using critical exponents calculated via the Kouvel–Fisher relation. For the x = 0.3 and x = 0.4 samples, the critical exponents obtained through the self-consistent method (MAP) provide better scaling of the universal curves. These observations provide initial insights into the more suitable method, and these parameters will be further validated in subsequent steps.
Figure 9 presents the isothermal magnetization curves at temperatures corresponding to the fitted TC values: 150 K for the x = 0.2 sample in Fig. 9(a), 220 K for the x = 0.3 sample in Fig. 9(b), 280 K for the x = 0.4 sample in Fig. 9(c), and 265 K for the x = 0.5 sample in Fig. 9(d). The insets show slopes obtained from ln–ln scaled critical isotherms, specifically at TC, as per Eq. (6). Reciprocals of these slopes yield the δ values. The critical exponent δ must satisfy the Widom scaling relation .28 Different δ values can be derived using the MAP method and the KF relation, listed in Table IV. Following comparison, the fitting method that aligns more closely with the expected relation is selected, and the parameters calculated via this method are chosen.
Doping concentration . | MAP . | KF . | DH1/δ . |
---|---|---|---|
δ . | δ . | δ . | |
x = 0.2 | 3.123 | 3.228 | 3.2 |
x = 0.3 | 3.645 | 3.738 | 3.604 |
x = 0.4 | 3.14 | 3.351 | 3.122 |
x = 0.5 | 3.555 | 3.405 | 2.923 |
Doping concentration . | MAP . | KF . | DH1/δ . |
---|---|---|---|
δ . | δ . | δ . | |
x = 0.2 | 3.123 | 3.228 | 3.2 |
x = 0.3 | 3.645 | 3.738 | 3.604 |
x = 0.4 | 3.14 | 3.351 | 3.122 |
x = 0.5 | 3.555 | 3.405 | 2.923 |
For the x = 0.2 sample, the δ value calculated from critical exponents obtained using the KF relation closely matches the δ value fitted from the critical isotherm, consistent with the universal curve. Therefore, the critical parameters derived from the KF relation are chosen. Similarly, for the x = 0.3 and x = 0.4 samples, the δ values calculated from critical exponents obtained using the MAP align better with the δ values fitted from the critical isotherm, consistent with the universal curve. Hence, the critical parameters obtained from the self-consistent method are selected. Regarding the x = 0.5 sample, where the critical exponents are less reliable, a δ value is still fitted from the critical isotherm to maintain data integrity and verify the reliability of the calculated critical exponents. A significant difference of approximately 0.5 is observed compared to the δ values of the other ferromagnetic phases. The KF relation, which produces a smaller difference, is chosen for further verification.
Apart from the x = 0.5 sample, which has coexisting antiferromagnetic phases, the reliable calculation methods for the x = 0.2, 0.3, and 0.4 samples were selected based on the universal curves and δ comparisons. The critical parameters obtained for these samples were then compared with standard models. The results indicate that these parameters are closer to the mean-field model, as shown in Table V.
Sample and models . | β . | γ . | δ . | References . | |
---|---|---|---|---|---|
x = 0.2 | KF | 0.435 ± 0.015 | 0.969 ± 0.013 | 3.228 | This work |
x = 0.3 | MAP | 0.366 ± 0.01 | 0.968 ± 0.015 | 3.645 | This work |
x = 0.4 | MAP | 0.407 ± 0.01 | 0.871 ± 0.013 | 3.14 | This work |
x = 0.5 | KF | 0.449 ± 0.009 | 1.08 ± 0.005 | 3.405 | This work |
3D Ising model | 0.3258 ± 0.094 | 1.24 | 4.82 ± 0.01 | 26 | |
3D XY model | 0.346 ± 0.009 | 1.316 ± 0.009 | 4.81 ± 0.08 | 29 | |
3D Heisenberg model | 0.3645 ± 0.0025 | 1.39 | 4.80 ± 0.04 | 29 | |
Mean field model | 0.5 | 1 | 3 | 29 |
Sample and models . | β . | γ . | δ . | References . | |
---|---|---|---|---|---|
x = 0.2 | KF | 0.435 ± 0.015 | 0.969 ± 0.013 | 3.228 | This work |
x = 0.3 | MAP | 0.366 ± 0.01 | 0.968 ± 0.015 | 3.645 | This work |
x = 0.4 | MAP | 0.407 ± 0.01 | 0.871 ± 0.013 | 3.14 | This work |
x = 0.5 | KF | 0.449 ± 0.009 | 1.08 ± 0.005 | 3.405 | This work |
3D Ising model | 0.3258 ± 0.094 | 1.24 | 4.82 ± 0.01 | 26 | |
3D XY model | 0.346 ± 0.009 | 1.316 ± 0.009 | 4.81 ± 0.08 | 29 | |
3D Heisenberg model | 0.3645 ± 0.0025 | 1.39 | 4.80 ± 0.04 | 29 | |
Mean field model | 0.5 | 1 | 3 | 29 |
Figure 10 illustrates the magnetocaloric properties of Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) by depicting the relationship between magnetic entropy change and temperature. The magnetic entropy change (−ΔSM) represents the temperature variation in a material during magnetization and demagnetization processes under varying magnetic fields. According to the Maxwell relations, the magnetic entropy values (−ΔSM) are calculated from the M-H curves using
, where are the magnetization rates at different temperatures and at the same magnetic field, respectively.
Figures 10(a)–10(d) illustrate the relationship between magnetic entropy change (−ΔSM) and temperature under various applied magnetic fields. As expected, the peak values of (−ΔSM) occur near the transition region (TC) and increase with stronger magnetic fields. Under an applied field of μ0ΔH = 50 kOe, the maximum magnetic entropy change (−ΔSM) varies across the different compositions of Nd1−xSrxMnO3. For the x = 0.2 sample, the maximum (−ΔSM) reaches 3.108 J/kg K at 160 K. The x = 0.3 sample exhibits the highest (−ΔSM) value of 4.315 J/kg K at 235 K. As the Sr content increases, the x = 0.4 sample shows a (−ΔSM) of 3.298 J/kg K at 290 K. The x = 0.5 sample demonstrates the lowest (−ΔSM) value of 2.433 J/kg K at 280 K among the studied samples. The three-dimensional (3D) magnetic entropy plots are shown in Figs. 11(a)–11(d).
Using the above equation, we can fit the data for magnetic entropy change and RCP, as shown in Figs. 12(a)–12(d). The obtained values of n and δ are listed in Table VI. Comparing n vs δ obtained from the critical parameters, the samples with x = 0.2, 0.3, and 0.4 all match well, while the sample with x = 0.5 shows a significant difference. This discrepancy may be related to its charge-ordered antiferromagnetic phase, consistent with the previous analysis, thus proving the reliability of these results.
. | Calculation of critical parameters . | Entropy and RCP fitting . | ||
---|---|---|---|---|
n . | δ . | n . | δ . | |
x = 0.2 | 0.5976 | 3.228 | 0.5995 | 3.321 |
x = 0.3 | 0.525 | 3.645 | 0.5742 | 3.752 |
x = 0.4 | 0.536 | 3.14 | 0.5415 | 3.204 |
x = 0.5 | 0.6396 | 3.405 | 0.8244 | 3.47 |
. | Calculation of critical parameters . | Entropy and RCP fitting . | ||
---|---|---|---|---|
n . | δ . | n . | δ . | |
x = 0.2 | 0.5976 | 3.228 | 0.5995 | 3.321 |
x = 0.3 | 0.525 | 3.645 | 0.5742 | 3.752 |
x = 0.4 | 0.536 | 3.14 | 0.5415 | 3.204 |
x = 0.5 | 0.6396 | 3.405 | 0.8244 | 3.47 |
The relationship between ΔCP,H and temperature for μ0ΔH in the range of 1–50 kOe is shown in Figs. 13(a)–13(d). As the temperature increases, the heat capacity undergoes a regular transition from negative to positive values, corresponding to the Curie temperature TC transition point. Below TC, the heat capacity exhibits negative values, while above TC, the heat capacity exhibits positive values, reflecting a change in the magnetic phase.
IV. CONCLUSION
The structure, magnetic characteristics, critical behavior, and MCE of Nd1−xSrxMnO3 (0.2 ≤ x ≤ 0.5) were systematically studied using XRD, SEM, and magnetization techniques. All samples exhibit a single orthorhombic phase, but with different space groups: Pnma (No. 62) for x = 0.2 and 0.3 samples, and Imma (No. 74) for x = 0.4 and 0.5 samples. TC of these samples are 152, 217, 283, and 291 K, respectively, while TCW are 178, 226, 285, and 285 K, respectively, showing a deviation from TC. The increase in TC and spontaneous magnetization is attributed to the development of DE interaction, facilitated by the substitution of Sr2+, which introduces mobile Mn4+ bonds forming Mn3+—O—Mn4+ bonds and influencing Mn positioning. The difference between TCW and TC arises from the competition between polarons and double-exchange, with DE dominating below TC and polarons becoming significant above TC, resulting in complex magnetic behaviors. Samples with doping concentrations x = 0.2 and 0.3 exhibit a Griffiths-like phase. The slight variation in TC and reduced spontaneous magnetization observed in the x = 0.5 sample is due to the presence of a charge-ordered superexchange antiferromagnetic interaction. All samples demonstrate significant MCE, with the x = 0.3 sample showing a maximum (−ΔSM) of 4.315 J/kg K and a corresponding RCP value of 280.48 J/kg under an applied field of μ0ΔH = 50 kOe. Near room temperature (290 K), the x = 0.4 sample exhibits a (−ΔSM) of 3.298 J/kg K and an RCP value of 283.64 J/kg under similar conditions. The KF relation is suitable for the samples with x = 0.2 and 0.5, while the MAP method is suitable for the samples with x = 0.3 and 0.4. Except for the x = 0.5 sample, where the presence of the charge-ordered phase complicates critical behavior analysis, the samples with other doping concentrations generally adhere to the mean-field model.
This study on the doping effects of Nd1−xSrxMnO3 provides new insights into the magnetic properties, critical behavior, and MCE of these important manganites. The observed magnetic behaviors and MCE are particularly significant for the potential application of perovskite manganites in magnetic refrigeration. This research is expected to stimulate renewed interest in manganites, prompting further investigation into how doping and structural modifications influence their diverse physical characteristics, including magnetoresistance, MCE, thermoelectric effects, dielectric properties, and optical properties.
ACKNOWLEDGMENTS
This work was supported by the National Natural Science Foundation of China (Nos. 11604067 and U1832143) and the Scientific Research Project of the Department of Education of Zhejiang Province (No. Y202352427). We want to thank colleagues from Beijing Synchrotron Radiation Facility and Shanghai Synchrotron Radiation Facility for their great support.
AUTHOR DECLARATIONS
Conflict of Interest
The authors declare no conflict of interest.
Author Contributions
Haiou Wang: Data curation (supporting); Formal analysis (equal); Investigation (lead); Project administration (equal); Writing – original draft (lead); Writing – review – editing (lead). Fuxiao Dong: Data curation (equal); Formal analysis (equal); Investigation (equal). Haochen Wang: Data curation (equal); Writing – original draft (supporting). Bojun Zhao: Data curation (equal); Investigation (supporting). Yan Wang: Data curation (equal); Writing – original draft (supporting). Weishi Tan: Formal analysis (equal); Investigation (supporting).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.