Pramod R.
Nadig
a,
Murari
M. S.
b and
Mamatha D.
Daivajna
*a
aDepartment of Physics, Manipal Institute of Technology (MIT), Manipal Academy of Higher Education, Manipal, Karnataka 576104, India. E-mail: mamatha.daijna@manipal.edu; Fax: +91 9449331022
bDST PURSE Program, Mangalore University, Mangalagangotri, Mangalore, Karnataka 574199, India
First published on 23rd January 2024
The effect of heat treatments on bulk poly-crystalline La0.67Ca0.33MnO3 perovskite manganite is presented, to explore the possible enhancement in magnetocaloric performance. Samples were prepared via conventional solid-state reaction route with annealing and sintering at various temperatures. Detailed measurements of temperature-dependent and field-dependent magnetization were carried out to estimate the Curie point and order of magnetic transition. The increased sintering temperature results in a steep transition near the TC, and establishes the magnetic sensitivity as well as the active zone for substantial magnetocaloric performance, at about 168.2% for the LCM9 (sintered at 900 °C) sample. The cause for the significant improvement in the magnetic and magnetocaloric response is brought to light using detailed X-ray photoelectron spectroscopy (XPS) analysis, highlighting the role of oxygen in modifying the Mn3+/Mn4+ charge ratio. The maximum value of the isothermal magnetic entropy change for the optimized sample is found to be 6.4 J kg−1 K−1, achieved at 269 K, while temperature-averaged entropy change (TEC) values, TEC(ΔTH–C = 3 K) and TEC(ΔTH–C = 5 K), of 6 J kg−1 K−1 and 5.2 J kg−1 K−1, respectively, were obtained with a low magnetic field change of 20 kOe. The obtained isothermal entropy change at low field for the optimized La0.67Ca0.33MnO3 sample is higher than that of pure Gd and most oxide-based materials. The relative cooling power (RCP) value is around 93 J kg−1 (ΔH = 20 kOe). The order of the phase transition is examined with universal scaling; the scaled entropy change curves confirm the collapse onto a single curve for LCM9, asserting second-order character, whereas the breakdown of the curve with a dispersion relation (d) of 101.1% at Θ = −5 confirms the onset of intrinsic first-order nature in the case of the high-temperature-sintered samples. Calorimetry measurements show thermal hysteresis of 2.4 K and 7.1 K for LCM11 (sintered at 1100 °C) at ramp rates of 5 K min−1 and 10 K min−1, respectively, confirming the first-order nature of the magnetic transition.
Among perovskite manganites, lanthanum-based mixed-valence manganites with the general formula La1−xAxMnO3 (A is mostly a divalent ion, Ca, Ba or Sr) show a variety of fascinating properties, like colossal magnetoresistance (CMR),20–22 charge ordering (CO),23,24 phase separation (PS),25,26 the magnetocaloric effect (MCE),1,27–30etc. For (La,Ca)MnO3, both end members LaMnO3 and CaMnO3 are typical antiferromagnets with TN close to 140 K and 130 K, respectively.31 Upon substitution of a divalent cation at the A-site, the system becomes mixed valent [La1−xAx(Mn1−x3+Mnx4+)O3] with generation of Mn3+/Mn4+ charged species.32 Similar charge control can be achieved via monovalent substitution [La1−xAx (Mn1−2x3+Mn2x4+)O3] at the La site.33 The mixed valence of Mn3+ (t32g e1g) and Mn4+ (t32g e0g) results in strong hybridization with oxygen 2p states, delocalizing the eg electrons, which mediate the ferromagnetic interactions between the localized t2g spins. As a result of these competitive interactions, various intriguing properties have been observed in mixed-valence manganites.
The polycrystalline La1−xCaxMnO3 represent a class of materials that emerged as suitable candidates for magnetic refrigeration technology, as they exhibit a sharp drop in magnetization across the magnetic transition region.34 The nature of the ferromagnetic (FM)–paramagnetic (PM) transition is most interesting for the La1−xCaxMnO3 system. Only a narrow range of substitution between x ≈ 0.2–0.4 will result in a first-order phase transition (FOPT), whereas x ≤ 0.2 results in a continuous second-order phase transition (SOPT) and x ≥ 0.4 results in a continuous transition with a tricritical point. The sharp drop in magnetization at low magnetizing intensities near the transition region is a more important concern for obtaining a large magnetocaloric effect. In the present investigation, a typical value of x = 0.33 is chosen (La0.67Ca0.33MnO3) because of the elusive nature of the FM–PM transition, with claims of it being first-order, with a strong volume anomaly near the transition region,35 small thermal hysteresis in temperature-dependent magnetization and negative slopes observed in isothermal magnetization plots,34 and shifts in heat capacity/asymmetric growth of the magnetic entropy change with field,36 or being second order.37 However, finite size effects,38 external pressure39,40 and a high field41 can also cause crossover in the nature of the transition. For instance, previous work by other groups38,42 investigated the effect of size reduction on x = 0.33 composition; the intrinsic FOPT becomes continuous second-order upon size reduction, with the size threshold lying between 50 and 100 nm.
Thereby, it is hypothesized that keeping all intrinsic and extrinsic parameters unaltered, the nature of the phase transition shall differ with annealing temperature, which controls the grain size with simultaneous variation in the Mn ratio (Mn3+/Mn4+) via oxygen non-stoichiometry. Also, this study realizes the optimized conditions to obtain a sharp drop in magnetization at the Curie point, associated with the intrinsic first-order nature. Most importantly, the magnetocaloric phenomena linked to the nature of the phase transition are explored and eventually the optimum magnetocaloric performance, especially at low fields with La0.67Ca0.33MnO3 perovskite manganites, is presented.
The structure, phase purity, unit-cell parameters, bond length and bond angles were determined with the help of the Rietveld refinement method using FullProf Suite.44 The initialization and steps followed in the refinement can be found in the ESI.† The refined patterns are shown in Fig. 1a, overlaying the observed (Yobs) data on calculated (Ycalc) data. The agreement in Yobs and Ycalc is established based on the criteria of R-factors (Rp, Rwp, Rexp and Bragg-R) and χ2 values. The samples were found to have an orthorhombic structure with the Pnma space group without any impurity phase. The obtained parameters are tabulated in Table 1. The unit-cell parameters change significantly when TS is changed from 900 °C to 1000 °C. The lattice parameters a and b are unchanged, whereas c is found to decrease, reflecting similar changes in the unit-cell volume. Keeping the elemental composition intact throughout the sintering temperature range, this perhaps indicates a change in the Mn3+/Mn4+ charge ratio in the given perovskite due to changes in oxygen stoichiometry. The volume contraction could be a result of an increase in Mn4+, which has a smaller size compared to its counterpart Mn3+ ions. However, changes in unit-cell volume, MnO6 octahedral tilt, average 〈Mn–O–Mn〉 bond angle and average 〈Mn–O〉 are not observed after 1000 °C. Since the bond length and bond angles are the crucial factors in controlling the exchange interaction in perovskite manganites, similar variations shall be reflected in physical properties such as magnetism.
Sample | La0.67Ca0.33MnO3 | |||
---|---|---|---|---|
Sample code | LCM9 | LCM10 | LCM11 | LCM12 |
Structure | Orthorhombic | |||
Space group | Pnma | |||
a (Å) | 5.459 | 5.456 | 5.455 | 5.454 |
b (Å) | 7.709 | 7.708 | 7.707 | 7.707 |
c (Å) | 5.568 | 5.469 | 5.469 | 5.471 |
Unit-cell volume (Å)3 | 230.150 | 229.998 | 229.926 | 229.967 |
〈Mn–O–Mn〉 (°) | 160.37 | 161.18 | 161.74 | 160.89 |
〈Mn–O〉 (Å) | 1.955 | 1.956 | 1.956 | 1.958 |
MnO6 tilt (°) | 10.169 | 9.31 | 9.63 | 10.31 |
R p | 9.86 | 8.63 | 9.33 | 9.45 |
R wp | 13.7 | 12.9 | 13.6 | 13.7 |
R exp | 11.84 | 11.96 | 12 | 12.53 |
Bragg R-factor | 3.9 | 3.02 | 3.52 | 4.25 |
χ 2 | 1.34 | 1.17 | 1.28 | 1.19 |
T C (K) | 270.9 | 270.4 | 268.1 | 267.5 |
θ p (K) | 269.6 | 271.9 | 268.4 | 268.2 |
Curie constant C | 4.275 | 4.305 | 4.830 | 5.822 |
μ expeff (μB) (μtheff = 4.85 μB) | 5.84 | 5.86 | 6.21 | 6.85 |
Fig. 2 shows FESEM micrographs of the prepared samples. The average grain size is found to increase with TS, with an apparent reduction in the grain boundaries, and the grain distribution has become compact. The average grain size has improved from 0.03 μm2 to 6.02 μm2. The sintered samples showed an exponential grain growth with increasing sintering temperature (Fig. 1d). This observation agrees well with the mechanism of grain growth depending on the reaction rate, which is an exponential function of the sintering temperature.45 The grain size of the prepared specimens shows a log-normal distribution, given by the probability distribution function
![]() | ||
Fig. 2 The FESEM images of LCM9, LCM10, LCM11 and LCM12 samples at a scale of 1 μm. (Insets: The histograms show log-normal distributions.) |
![]() | ||
Fig. 3 [a] The FTIR absorption spectra and [b] the variation in bending mode νb with TS for LCM9, LCM10, LCM11 and LCM12 samples. |
The core O 1s spectra (Fig. 5a) exhibit an asymmetric peak with a shoulder in LCM9, LCM10, and LCM12. A nearly symmetric peak is observed in LCM11. The O 1s spectrum was fitted using three distinct components that were found in the ranges of 528.38–528.84 eV, 530.52–531.40 eV and, 532.62–533.42 eV after the Shirley background was removed. The characteristics of the O 1s spectra show chemical shifts in the oxygen core levels brought on by various chemical environments (i.e., oxygen bonding with Mn, La/Ca and other functional groups). The O 1s spectral interpretation is bit challenging due to the presence of several overlapping components. Early on, similar spectra were seen for other perovskite-type oxides, although the interpretations were ambiguous and contradictory.48–52 Some of the authors refer to oxygen bonding with each of the other elements of the compound or defects. Since the lower-binding-energy feature most closely approaches the binding energy of the O 1s core level in MnO, which is between 528.38–528.84 eV, we may infer that this peak is associated with strong covalent Mn–O bonding. On the other hand, the peak between 530.52–531.40 eV can be attributed to a layer comprising La/Ca–O or a combination of both, since ionic bonds (La/Ca–O) have a higher binding energy than covalent Mn–O bonds. According to Chu et al.,53 further deconvolution of component La/Ca–O (between 530.52–531.40 eV) into two sub-components, individually corresponding to La–O and Ca–O, is not conceivable due to the spectrometer's limited energy resolution mixed with sample effects. The fact that La/Ca–O contributes significantly to LCM12 compared to the other components shows that the surface is made up of a La/Ca–O bonding environment. Thus, the intensity of Mn–O in LCM12 is comparatively lower than that in other specimens, indicating that the sample contains an abundance of oxygen and is dominated by the surface effects.54 The higher binding energy peak (marked as O#) between 532.62–533.42 eV might be the result of surface species like carbonyl groups or crystal lattice defects. It is argued that elements located at the outermost surface have larger binding energies than their bulk-bound counterparts, which means that surface sites often display less stability than bulk sites due to the abrupt termination of the bulk lattice structure.55
![]() | ||
Fig. 5 [a] Core level oxygen O1s, [b] carbon C 1s and [c] Ca 2p spectra for LCM9, LCM10, LCM11, and LCM12 samples. |
Furthermore, the C 1s core spectra are examined (Fig. 5b), which deconvolute into three distinct components for LCM9 to LCM11 and four for LCM12. The major peak of C 1s can be assigned to adventitious carbon at 284.6 eV (CC) and serves as a reference for the correction of spectra against charging effects. The peak distribution near 285.96–286.98 eV can be assigned to C–O bonds. Finally, surface carbonates are responsible for a third component between 288.10–288.47 eV (C
O), which is in support of the O 1s third component peak between 532.62–533.42 eV. The surface carbonates (purple line) are most intense for LCM12 when compared to other specimens. Evidently, another peak at 290.82 eV (orange line; marked with ♦) is observed in LCM12 and may indicate the presence of another polymorph of carbonates.55
The Ca 2p core level spectra (Fig. 5c) display spin–orbit doublets with 2p3/2 and 2p1/2 peaks. The spectra are well fitted by four components: two major peaks (blue lines), corresponding to the Ca 2p3/2 and Ca 2p1/2 levels of the solid phase; and two other additional peaks (olive-colored lines), which may come from a small fraction of surface oxide and carbonate species.55 In comparison to the other specimens, the LCM11 has the least contribution from surface oxide and carbonate species. However, it is clear that the primary lines in LCM12 have been shifted towards higher binding energies. This might be brought about by the segregation of Ca ions at the oxide surface, which could lead to the formation of CaCO3 and/or CaO. Upon high temperature sintering, small quantities of extra oxygen are introduced into the system, where this appears to be accommodated by introducing A/B site vacancies in the structure.56 Thus, it is more evident that LCM12 contains an abundance of oxygen clusters, due to which some Ca has segregated to the surface of the sample. These findings from the O 1s and C 1s core level spectra are consistent. The existence of carbonyl group contributions and their bonding with CO (referred to as surface carbonates) were evidenced by the O 1s and C 1s core level spectra. Furthermore, in the case of LCM12, the carbonate groups – in particular, CaCO3 – provide strong evidence that Ca segregated from the A site of the perovskite. Overall, it has been shown that differences in the electron densities surrounding the oxygen anions in the Mn–O bonds have an impact on the binding energies of the Mn 2p3/2 and Mn 2p1/2 orbitals in the perovskites.
The Mn 2p XPS spectra (Fig. 6) show a spin–orbit splitting between the Mn 2p3/2 and Mn 2p1/2 peaks, with an energy separation (given in Table 2). The energy separation in the metallic state of Mn is typically 11.05 eV,57 whereas a higher energy separation value implies the occurrence of a Mn oxidative state. In addition, the difference between the binding energies of Mn 2p3/2 and O 1s is found to be 111.38 eV (for LCM9) to 113.21 eV (for LCM12). This is direct evidence of the evolution of Mn4+ ion species over Mn3+.58 The core spectra of Mn 2p3/2 and Mn 2p1/2 are deconvoluted into two peaks, unveiling the existence of mixed-valence Mn states (Mn3+ and Mn4+), which appear after the divalent substitution at the A-site of perovskite manganites. The corresponding binding energies for the said two peaks are tabulated in Table 2. The relative fractions of Mn3+ and Mn4+ ions are calculated by using the areas under the corresponding peaks. Ideally, for an x = 0.33 system, the theoretical Mn3+/Mn4+ ratio is 2.03 (Mn3+:
Mn4+ = 67
:
33). However, the values obtained are found to be 2.26, 2.11, 2.04, and 1.84 for LCM9 to LCM12, respectively. This indicates that for LCM9, the specimen is Mn3+ rich, whereas for LCM12, the specimen finds Mn4+ abundance, and for LCM11, the specimen is nearly stoichiometric with a 67
:
33 charge ratio. The larger Mn3+/Mn4+ and Mn4+/Mn3+ ratios at the extreme ends of the series compared to the stoichiometric compound (Mn3+/Mn4+ = 2.030) are indicative of oxygen off-stoichiometry (perhaps a deficiency or excess). The oxygen off-stoichiometry at low/high TS is supported by the observed asymmetric/shoulder peak in the oxygen O 1s spectra. The rise in Mn3+ and Mn4+ above a nominal value originates from oxygen vacancies/excess,59 which can alter the octahedral symmetry.60 The overall change in this ratio severely influences the double-exchange interaction via Mn3+–O2−–Mn4+ bridges, which further modifies the transport and magnetic properties.
Sample | Core spectra | Spin–orbit split | Peak position eV | Difference eV | Mn states | Peak position eV | Area under curve | Mn3+/Mn4+ (nominal 2.03) |
---|---|---|---|---|---|---|---|---|
LCM9 | Mn 2p | Mn 2p3/2 | 641.98 | 11.58 | Mn3+ | 641.02 | 33278.3 | 2.26 |
Mn4+ | 643.09 | 14702.7 | ||||||
Mn 2p1/2 | 653.56 | Mn3+ | 652.53 | 14506.02 | ||||
Mn4+ | 654.66 | 6416.5 | ||||||
LCM10 | Mn 2p | Mn 2p3/2 | 642.13 | 11.65 | Mn3+ | 641.23 | 46583.2 | 2.11 |
Mn4+ | 643.24 | 22031.7 | ||||||
Mn 2p1/2 | 653.78 | Mn3+ | 652.87 | 24642.2 | ||||
Mn4+ | 654.84 | 11634.8 | ||||||
LCM11 | Mn 2p | Mn 2p3/2 | 642.21 | 11.68 | Mn3+ | 641.3 | 43177.4 | 2.04 |
Mn4+ | 643.46 | 21134.4 | ||||||
Mn 2p1/2 | 653.89 | Mn3+ | 652.96 | 22174.3 | ||||
Mn4+ | 655.13 | 10856.2 | ||||||
LCM12 | Mn 2p | Mn 2p3/2 | 642.56 | 11.75 | Mn3+ | 641.43 | 41486.8 | 1.84 |
Mn4+ | 643.52 | 22514.6 | ||||||
Mn 2p1/2 | 654.31 | Mn3+ | 653.02 | 21337.5 | ||||
Mn4+ | 655.27 | 11542.1 |
![]() | ||
Fig. 7 The temperature-dependent magnetization (M–T) curves under 100 Oe magnetic field in ZFCW, FCC and FCW modes. The insets show the Curie point (dM/dT vs. T). |
Bifurcation between ZFCW and FCC magnetization below TC is seen with further decrease in temperature due to magnetic inhomogeneity. This is a characteristic property observed in manganites, where the magnitude of magnetization in ZFCW is always less than in FCC, attributed to the gradual freezing of magnetic moments in randomly distributed clusters, resulting in incomplete magnetization rather than long-range FM coupling.62 The M–T curve of sample LCM9 has concomitant transition broadening (FWHM ∼ 15.5 K) when compared with the other samples. This is due to enhanced grain boundary density at low sintering temperature, where each grain surface has different oxygen stoichiometry causing magnetic disorder, or in other words, the disorder can arise in perovskite manganites due to the surface termination of the crystal structure.63 This fact is established through XPS characterization, which is a surface-sensitive technique. Sample LCM9 has shown a deviation from oxygen stoichiometry, as indicated by an asymmetric peak in the O 1s spectra and consequent variation in the Mn3+/Mn4+ ratio. The onset of disappearance of these disorders is believed to happen upon annealing, where oxygen vacancies at grain boundaries are suspended, resulting in enhancement of the grain size.64 This is clearly established through XPS analysis on high-temperature-sintered samples. The oxygen vacancies have significantly reduced, turning LCM11 into a stoichiometric specimen. Interestingly, the presence of thermal hysteresis (<2 K) between FCC and FCW cycles is observed prominently in higher-temperature-sintered samples, which appears to be a signature of a weak FOPT.
To confirm the magnetic ground state and hence to estimate the effective paramagnetic moments and Curie–Weiss temperature (θP), the high temperature paramagnetic regimes after TC are fitted with the Curie–Weiss (CW) law, given by χ = C/(T − θP) where C is the Curie constant, defined as C = NAμB2 (μeff)2/(3KB), where NA = 6.023 × 1023 mol−1, μB = 9.274 × 1021 emu, KB = 1.38 × 1016 erg K−1 and μeff is the effective paramagnetic moment. The linear fit to the inverse susceptibility curve is shown in Fig. 8. The slope gives C and θP is obtained from the x intercept. The positive θP obtained from the fit concludes a dominant FM correlation between spins. Based on the assumption that the orbital moment is fully quenched ( = 0) by an internal crystal field effect,65 we have calculated the theoretical value of the effective paramagnetic moment μeff by considering spin momentum (
≠ 0). The μeff is given by
, where the Lande g-factor g = 2, S = 3/2 for Mn4+ (t32g e0g), S = 2 for Mn3+ (t32g e1g) and μeff (Mn3+) = 4.90 μB and μeff (Mn4+) = 3.87 μB.
For the La0.67Ca0.33MnO3 system, the total contribution from Mn3+ and Mn4+ ions is given by (μtheff)2 = 0.67μeff2 (Mn3+) + 0.33μeff2 (Mn4+) = 4.85 μB. The obtained values of the effective paramagnetic moments are higher than 4.85 μB, as tabulated in Table 1. This could be due to the contribution from FM clusters containing more than one Mn ion above the transition temperature. The spontaneous formation of magnetic clusters above TC was detected in a small-angle neutron scattering experiment conducted on La0.70Ca0.30MnO3 samples by De Teresa.66 Thereby, these kinds of FM clusters in the PM matrix have larger effective spins, which results in higher values of μexpeff than μtheff. A downturn in the inverse susceptibility curve is noticed for sample LCM9 (before approaching TC from high temperature) along with divergence from CW law, suggesting the presence of short-range FM interactions and thus a Griffiths phase (GP)-like singularity67 in specimen LCM9. The downturn occurring at ∼285 K is referred to as the Griffiths temperature (TG). Above TG, χ−1 exhibits pure PM behavior. The GP is formed at a characteristic freezing temperature where a short-range FM correlation exists within the cluster, which sets the PM regime and contributes to the increase in magnetic moments in the PM region. The GP singularity present between TC and TG is characterized by a power law, given by χ−1 ∝ (T − TRC)(1−λ), where λ is the magnetic susceptibility exponent and TRC is the critical temperature of random ferromagnetic clusters where susceptibility tends to diverge (χ → ∞). The choice of TRC is very crucial and is obtained using a method followed by Jiang et al.,68 in which TRC is recognized as the temperature for which fitting data in the PM regime yields λPM ∼ 0.
In order to verify the presence of a GP in LCM9, the temperature-dependent magnetization at 500 Oe is measured. Fig. 8 illustrates the clear suppression of χ−1 deviation at H = 500 Oe. This can be ascribed to the polarization of spins occurring beyond the clusters69 and, therefore, indicates the presence of a GP-like singularity. Furthermore, a plot of log(χ−1) versus log(T/TRC − 1) resembles a straight-line fit and is displayed in the inset of LCM9 in Fig. 8. The obtained values of λPM = 0.001 (for 100 Oe) and 0.004 (for 500 Oe) in the PM regime are close to ideal value of ‘0’, whereas in the GP region, the λGP values are 0.92 and 0.87 for 100 and 500 Oe, respectively. The obtained values of λGP agree well with the expected range of 0 < λ < 1. Hence, the value of λGP and its suppression with an increase in field validates the GP-like singularity present in LCM9. Consequently, LCM9 has no long-range ordering due to disordered spins at the surface and random magnetic interactions due to uneven distribution of grains. However, the downturn behavior in the inverse susceptibility was not seen for the rest of the samples, even though they have higher μexpeff values. But the difference between TC and θP decreased with TS, which is a clue that the magnetic inhomogeneity has decreased when compared with the LCM9 sample.
To get deeper insights into the magnetic transition and variation of the magnetocaloric properties, magnetic isotherms were collected in the vicinity of the magnetic transition region for all the samples and are displayed in Fig. 9. The isotherms were collected in 2 K intervals across the transition region and 4 K away from it. The isotherms below TC display an abrupt change in magnetization, showing a tendency towards saturation that confirms the FM behavior. Above TC, the magnetization changes are gradual, showing linearity far away from TC that confirms a typical PM behavior. Further, the magnitude of magnetization has increased with TS. The findings can be rationalised by the argument that, with an increase in TS, the oxygen off-stoichiometry at the grain surface and boundary vanishes, resulting in a homogenous, stoichiometric La0.67Ca0.33MnO3 compound. Among the La1−xCaxMnO3 compounds, the 67:
33 specimen has an optimal Mn3+/Mn4+ ratio to have the maximum TC in this family, or in other words, the said concentration would facilitate Zener double-exchange easily compared to the other concentrations. Variation in oxygen content would cause deviation of this ratio, resulting in a larger number of Mn3+ ions which are Jahn–Teller active. The oxygen vacancies are proven to break the octahedral symmetry, thereby reducing the energy of the eg orbital and causing a strong electron–lattice interaction. This localizes the eg electrons available for hopping, eventually suppressing ferromagnetic double-exchange (DE) interactions. Reduced DE interactions are responsible for large reductions in magnetic moment and hence the magnetization.60 This is established clearly through the oxygen vacancies/excess observed in XPS studies. We infer that the effects of oxygen off-stoichiometry are more pronounced in LCM9 (oxygen-deficient) and LCM12 (oxygen-excess).
![]() | ||
Fig. 9 The field-dependent isothermal magnetization curves (M vs. H) of samples LCM9, LCM10, LCM11 and LCM12 in the vicinity of the magnetic transition. |
Apart from the said features, interestingly, prominent S-shaped M(H) curves are observed for the high-temperature-sintered samples. This feature is observed over a certain temperature range above TC, indicative of a metamagnetic phase transition in the PM state of the sample. It has been proven that the magnetic transition at low field is accompanied by an abrupt decrease in unit-cell volume without a change in the crystal symmetry.35,70 Furthermore, the observation of a change in unit-cell volume in response to an applied magnetic field may suggest the presence of magnetic polarons. These magnetic polarons have been confirmed to exist in the PM regime (T = 1.8TC) by a small-angle neutron scattering experiment.66 At H = 0, these are non-interacting; magnetic polarons begin to interact as H grows at a fixed temperature (close to but above TC). At a critical value of magnetic field, the magnetic polarons grow and combine to form larger ferromagnetic clusters.71 Therefore, a gradual increase in magnetization is observed at lower fields, followed by a rapid increase above a certain critical magnetic field, reaching saturation at higher fields.
To determine the nature of the transition with an increase in TS, Arrott plots were constructed for all isotherms and are shown in Fig. 10. According to the Banerjee criterion, the slope of M2vs. H/M distinguishes the order of the transition. From a thermodynamic point of view, a positive slope represents a second-order transition and a negative slope represents a FOPT. The slopes of the M2 (H/M) curves of LCM9 at temperatures T > 277 K (T < 277 K) are negative (positive) at low fields; these features indicate a combination of first- and second-order characters. For the higher-temperature-sintered samples, significant negative slopes are observed in the M2 (H/M) curves, which are consistent with previous observations for bulk La0.67Ca0.33MnO3.34 Bui et al.72 investigated the coexistence of several isostructural oxygen-deficient phases and its impact on the nature of the magnetic transition of the La0.67Ca0.33MnO3 compound. They found that oxygen-deficiency led to magnetic inhomogeneity, resulting in the appearance of a Griffiths phase and causing a first- to second-order phase transformation. Earlier, work by Park et al.73 found that the magnetism in perovskite manganites at the surface boundary is significantly different from that of the bulk. Thus, the outer surface layer has a probability to suppress the bulk properties and is likely to undergo an SOPT or weak FOPT. These studies are in line with the effects of oxygen off-stoichiometry causing a partial second-order nature in a typical first-order material, La0.67Ca0.33MnO3 (LCM9).
The magnitude of the magnetic entropy change exhibits a sharp peak near the magnetic transition temperature due to a considerable change in magnetization and the change in −ΔSM with the field has a linear relationship. The change in isothermal entropy is always negative for a material that shows a peak at the TC. Fig. 11 shows −ΔSM for field variations of 10 to 50 kOe with an increment of 10 kOe. The maximum −ΔSM is observed for LCM11 over the other samples, with peak values of 4.77 J kg−1 K−1, 6.39 J kg−1 K−1, 7.36 J kg−1 K−1, 8.16 J kg−1 K−1 and 8.83 J kg−1 K−1 for field changes of 10 kOe, 20 kOe, 30 kOe, 40 kOe and 50 kOe, respectively. For a low field change of 20 kOe, the peak −ΔSM values are 3.80 J kg−1 K−1, 5.45 J kg−1 K−1, 6.39 J kg−1 K−1, and 5.46 J kg−1 K−1 for LCM9, LCM10, LCM11 and LCM12 samples, respectively. The −ΔSM value is found to increase from sample LCM9 to LCM11 and then decreases at LCM12, having similar values to LCM10. We assert that the role of oxygen non-stoichiometry at the grain surface in modifying the magnetic properties will also be reflected in the MCE properties in an analogous way.
![]() | ||
Fig. 11 The isothermal magnetic entropy change vs. temperature under various magnetic field changes for samples LCM9, LCM10, LCM11 and LCM12, obtained from Maxwell's equation. |
Though −ΔSmaxM is a physical parameter which characterizes the magnetocaloric attributes of a specimen, nevertheless, the value of adiabatic temperature change (ΔTad) is meaningful quantity to quote the applicability of a material as a magnetic refrigerant.75
In addition, the relative cooling power (RCP) is commonly used as a criterion for assessing a magnetocaloric material in terms of its efficiency, as a product of the peak entropy change and the temperature range in which the value of the peak entropy change becomes half of the maximum, also called the full width at half maximum (δTFWHM) of ΔSmaxM.110 It is expressed by
RCP = |ΔSmaxM| × δTFWHM |
Type | Composition | T C/Tpeak (K) | ΔH (kOe) | −ΔSmaxM (J kg−1 K−1) | −ΔSmaxM/ΔH (J kg−1 K−1 kOe) | RCP (J kg−1) | Ref. |
---|---|---|---|---|---|---|---|
a Approximate values. | |||||||
Prototypes | Gd | 294 | 20 | 5.3 | 26.5 | 212a | 76 |
50 | 10.2 | 20.4 | 410 | ||||
Gd5Ge2Si2 | 275 | 50 | 18.5 | 37 | 535 | 77 | |
Fe49Rh51 | 316.58 | 20 | −13.2 | 66 | 153.4 | 8 | |
La(Fe0.9Si0.1)13 | 184 | 20 | 28 | 140 | 532a | 10 | |
MnAs | 318 | 20 | 31 | 155 | 124 | 12 | |
50 | 38 | 76 | 418 | ||||
LaFe11.2Co0.7Si1.1 | 274 | 20 | 12 | 60 | 216a | 9 | |
50 | 20.3 | 40.6 | 585a | ||||
(La, Ca)MnO3 | La0.5Ca0.5MnO3 | 220 | 20 | 1.13 | 5.65 | 67.8 | 78 |
La0.55Ca0.45MnO3 | 238 | 15 | 1.9 | 12.66 | 68 | 17 | |
La0.60Ca0.40MnO3 | 263 | 30 | 5 | 16.66 | 135 | 79 | |
La0.65Ca0.35MnO3 | 235 | 20 | 1.06 | 5.3 | 52 | 80 | |
La0.70Ca0.30MnO3 | 216 | 20 | 7.3 | 36.5 | 124.1 | 81 | |
La0.70Ca0.30MnO3 | 227 | 10 | 1.95 | 19.5 | 49 | 82 | |
La0.70Ca0.30MnO3 | 251 | 10 | 5.71 | 57.1 | 40 | 83 | |
50 | 7.52 | 15.04 | 218 | ||||
La0.70Ca0.30MnO3 | 272 | 10 | 3.36 | 33.6 | 33.6 | 84 | |
La0.72Ca0.28MnO3 | 268 | 20 | 5.5 | 27.5 | 88 | 85 | |
La0.75Ca0.25MnO3 | 224 | 15 | 4.70 | 31.33 | 99 | 86 | |
La0.75Ca0.25MnO3 | 268 | 50 | 6.25 | 12.5 | 198 | 87 | |
La0.80Ca0.20MnO3 | 230 | 15 | 5.50 | 36.66 | 72 | 17 | |
(La, Sr)MnO3 | La0.5Sr0.5MnO3 | 340 | 30 | 0.63 | 2.1 | — | 88 |
La0.60Ca0.40MnO3 | 365 | 20 | 0.97 | 4.85 | 98 | 89 | |
50 | 2.14 | 4.28 | 264 | ||||
La0.70Sr0.30MnO3 | 370 | 20 | 1.10 | 5.5 | 49 | 90 | |
La0.75Sr0.25MnO3 | 340 | 15 | 1.5 | 10 | 65 | 91 | |
LaLa0.80Sr0.20MnO3 | 309 | 10 | 1.42 | 14.2 | 52.7 | 92 | |
La0.70Sr0.30MnO3 | 369 | 10 | 1.14 | 11.4 | 46.4 | ||
La0.80Sr0.20MnO3 | 330 | 20 | 1.65 | 8.25 | 103.2 | 93 | |
La0.65Sr0.35MnO3 | 305 | 10 | 2.12 | 21.2 | 106 | 94 | |
(La, Ba)MnO3 | La0.6Ba0.4MnO3 | 333 | 25 | 1.19 | 4.76 | 79.3 | 95 |
La0.7Ba0.3MnO3 | 342 | 25 | 2.06 | 8.24 | 124 | ||
La0.7Ba0.3MnO3 | 332 | 20 | 3.08 | 15.4 | 46.2a | 96 | |
La0.7Ba0.3MnO3 | 322 | 18 | 2.1 | 11.66 | 40.2a | 97 | |
ABO3 (67![]() ![]() |
La0.67Ca0.33MnO3 | 267 | 15 | 4.3 | 28.66 | 47 | 17 |
La0.67Ca0.33MnO3 | 260 | 15 | 3.7 | 24.66 | 46.25a | 98 | |
La0.67Ca0.33MnO3 | 260 | 10 | 5 | 50 | 35 | 42 | |
La0.67Ca0.33MnO3 | 260 | 10 | 1.10 | 11 | 36.3a | 99 | |
La0.67Ca0.33MnO3 | 278 | 20 | 4.33 | 21.65 | 86.6a | 100 | |
La0.67Ca0.33MnO3 | 252 | 10 | 2.3 | 23 | 49 | 101 | |
20 | 4.1 | 20.5 | 91 | ||||
50 | 6.8 | 13.6 | 247 | ||||
La0.67Sr0.33MnO3 | 370 | 20 | 2.68 | 13.4 | 85 | 102 | |
50 | 5.15 | 10.3 | 252 | ||||
La0.67Sr0.33MnO3 | 377 | 20 | 2.02 | 10.1 | 92.9a | 100 | |
La0.67Sr0.33MnO3 | 354 | 50 | 2.49 | 4.98 | 225 | 103 | |
La0.67Ba0.33MnO3 | 292 | 50 | 1.48 | 2.96 | 161 | 104 | |
La0.67Ba0.33MnO3 | 337 | 10 | 2.7 | 27 | 68 | 105 | |
La0.67Ba0.33MnO3 | 350 | 20 | 1.72 | 8.6 | 92.8a | 100 | |
Pr0.67Sr0.33MnO3 | 290 | 20 | 3.5 | 17.5 | 79.2 | 106 | |
Pr0.67Pb0.33MnO3 | 360 | 20 | 2.33 | 11.65 | 106.2 | 107 | |
La0.67Pb0.33MnO3 | 358 | 20 | 2.54 | 12.7 | 134.6 | 108 | |
Pr0.67Ba0.33MnO3 | 188 | 10 | 2.32 | 23.2 | 49 | 109 | |
40 | 550 | 13.75 | 225 | ||||
Prepared samples | La0.67Ca0.33MnO3 (LCM9) | 268 | 20 | 3.80 | 19 | 75 | Our results |
50 | 6.45 | 12.9 | 210 | ||||
La0.67Ca0.33MnO3 (LCM10) | 271 | 20 | 5.45 | 27.25 | 82 | ||
50 | 7.51 | 15.02 | 225 | ||||
La0.67Ca0.33MnO3 (LCM11) | 269 | 10 | 4.77 | 47.7 | 42 | ||
15 | 5.72 | 38.13 | 67 | ||||
18 | 6.14 | 34.11 | 80 | ||||
20 | 6.39 | 31.95 | 93 | ||||
50 | 8.83 | 17.66 | 276 | ||||
La0.67Ca0.33MnO3 (LCM12) | 269 | 20 | 5.46 | 27.3 | 84 | ||
50 | 7.71 | 15.42 | 254 |
Refrigerant capacity (RC) is used as another parameter to assess magnetocaloric properties. It describes the thermal efficiency of a magnetocaloric material in terms of energy transfer between two reservoirs in one cycle. It is given by following relation:
So far, −ΔSM was discussed here when the field changes from 0 → Hmax. Since we have confirmed the FOPT features and large MCE of our optimal LCM11, it was further examined for its MCE under a reduction in field from Hmax → 0 to ensure the repeatability. The value of −ΔSM is changed from 5.7 J kg−1 K−1 (0 → 15 kOe) to 5.3 J kg−1 K−1 (15 → 0 kOe), as shown in the Fig. 13a for comparison. This small discrepancy further confirms the magnetic hysteresis present. But authors have pointed out that such hysteresis could also be a field-sweep-dependent case37 and using an infinitely slow sweep rate during isothermal measurements resulted in no intrinsic hysteresis. Moore112 demonstrated that heat exchange dynamics may lead to a significantly sweep-rate-dependent magnetic hysteresis. However, during isothermal measurements, the stability of temperature during the magnetic field sweep was maintained up to the first decimal digit. So, we claim that hysteresis may be arising from the intrinsic first-order nature rather than the uncertainties in the field sweep.
![]() | ||
Fig. 12 The magnetic field dependence of the temperature-averaged entropy change at ΔTH–C = 3 K, 5 K, 10 K, 15 K and 20 K temperature increases for samples LCM9, LCM10, LCM11 and LCM12. |
Material | TEC(3) | TEC(10) | Ref. |
---|---|---|---|
Gd | 3.05 | 2.91 | 111 |
Gd5Ge2Si2 | 9.91 | 6.89 | |
La(Fe0.88Si0.12)13 | 17.36 | 9.11 | |
LaFe11Co0.8Si1.2 | 3.37 | 3.24 | |
FeRh | 11.09 | 9.08 | |
La0.813K0.16Mn0.987O3 | 1.49 | 1.47 | |
La0.67Ca0.33MnO3 | 4.45 | 3.51 | Our result |
Like the TEC, the normalized refrigerant capacity (NRC) is another tool to identify a material's utility for cooling performance. It is defined as a function of ΔTH–C = THot − TCold:
Fig. 14 shows the variation of the NRC as a function of temperature increase. The results show that the NRC increases with increasing ΔTH–C, which is an attractive feature of our material for the applicability in the field of magnetic refrigeration technology.
![]() | ||
Fig. 14 The normalized refrigerant capacity (NRC) as a function of temperature increase (ΔTH–C) for samples LCM9, LCM10, LCM11 and LCM12. |
Further, to validate the second-order character of LCM9 and first-order features of LCM11, the dispersion parameter, d (%) is estimated at an arbitrary Θ below TC. It is given by the expression
Irrespective of the negative slope observed in Arrott plots, the asymmetry observed in the −ΔSM curve, the hysteresis observed in the isothermal magnetic entropy change and the breakdown of universality at a typical temperature of 1100 °C (LCM11) doesn't experimentally substantiate the FOPT. DSC is one of the direct measurements that validate a FOPT.
The transition latent heats are calculated using the area under the curve (heat flow vs. time, figure not shown here) which are 4.74 J g−1 and −4.80 J g−1, respectively, for the warming and cooling curves. Interestingly, with a 10 K min−1 ramp rate, the hysteresis between the peaks was found to be 7.1 K. The FOPT is bound to display a systematic hysteresis upon an increase in the ramp rate due to a pronounced shift of the cooling curve to the lower temperature. The significant increase in the hysteresis upon a change in ramp rate, observed for LCM11, validates the FOPT.
Thus, the large magnetization with an improvement in magnetic sensitivity, the metamagnetic transition due the first-order crossover of the magnetic ground state, the consequent sizable entropy change, and the excellent figure-of-merit (TEC) seen for LCM11 (due to its apt charge ratio, Mn3+/Mn4+ = 2.04) makes it an optimal candidate amongst the other three specimens for magnetic refrigeration applications. The entropy changes can be realized even at low external fields, and those being attainable using a permanent magnet makes this candidate attractive towards cooling technology.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cp04185a |
This journal is © the Owner Societies 2024 |