Next Article in Journal
A Review on the Applications of Natural Biodegradable Nano Polymers in Cardiac Tissue Engineering
Next Article in Special Issue
The Adsorption Effect of Methane Gas Molecules on Monolayer PbSe with and without Vacancy Defects: A First-Principles Study
Previous Article in Journal
Regulating the Electron Depletion Layer of Au/V2O5/Ag Thin Film Sensor for Breath Acetone as Potential Volatile Biomarker
Previous Article in Special Issue
Facile Fabrication of Highly Active CeO2@ZnO Nanoheterojunction Photocatalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Properties and Magnetocaloric Effect of Polycrystalline and Nano-Manganites Pr0.65Sr(0.35−x)CaxMnO3 (x ≤ 0.3)

1
Faculty of Physics, Babes-Bolyai University, Str. Kogalniceanu 1, 400084 Cluj-Napoca, Romania
2
National Institute for Research and Development of Isotopic and Molecular Technologies, Str. Donath 67-103, 400293 Cluj-Napoca, Romania
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(8), 1373; https://doi.org/10.3390/nano13081373
Submission received: 19 March 2023 / Revised: 10 April 2023 / Accepted: 11 April 2023 / Published: 14 April 2023

Abstract

:
Here we report investigations of bulk and nano-sized Pr0.65Sr(0.35−x)CaxMnO3 compounds (x ≤ 0.3). Solid-state reaction was implemented for polycrystalline compounds and a modified sol–gel method was used for nanocrystalline compounds. X-ray diffraction disclosed diminishing cell volume with increasing Ca substitution in Pbnm space group for all samples. Optical microscopy was used for bulk surface morphology and transmission electron microscopy was utilized for nano-sized samples. Iodometric titration showed oxygen deficiency for bulk compounds and oxygen excess for nano-sized particles. Measurements of resistivity of bulk samples revealed features at temperatures associated with grain boundary condition and with ferromagnetic (FM)/paramagnetic (PM) transition. All samples exhibited negative magnetoresistivity. Magnetic critical behavior analysis suggested the polycrystalline samples are governed by a tricritical mean field model while nanocrystalline samples are governed by a mean field model. Curie temperatures values lower with increasing Ca substitution from 295 K for the parent compound to 201 K for x = 0.2. Bulk compounds exhibit high entropy change, with the highest value of 9.21 J/kgK for x = 0.2. Magnetocaloric effect and the possibility of tuning the Curie temperature by Ca substitution of Sr make the investigated bulk polycrystalline compounds promising for application in magnetic refrigeration. Nano-sized samples possess wider effective entropy change temperature (ΔTfwhm) and lower entropy changes of around 4 J/kgK which, however, puts in doubt their straightforward potential for applications as magnetocaloric materials.

1. Introduction

Organic life is possible in a certain range of temperature because chemical exchange is destroyed above and below that range [1]. Throughout history, while artists were driven by the need to stay “cool”, the engineering part of the human brain kept searching for more practical solutions [2]. It is obvious that the best method for cooling our environment is the one that does not harm it. As such, magnetocaloric materials present a viable alternative to harmful gasses [3,4].
Magnetocaloric effect was reported in 1881 [5], but only recently has its potential for everyday use been discussed. Gadolinium Gd has been a forerunner in excellent magnetocaloric effect for a while, until its alloys were found to have higher values of entropy change [3]. They are expensive and require fields of over 5 T for operational use [4], and so the search for even more effective compounds has begun.
Compounds of the type A1−xBxMnO3 (where A is a trivalent rare earth cation and B is a divalent alkaline earth cation [6] (pp. 1–153) have a perovskite structure. They allow for Mn3+-O-Mn4+ interaction where an electron from Mn3+ can hop to Mn4+, thus aligning the magnetic moments. Such interaction is named “double exchange” and is the main reason for their electric and magnetic properties [7] (pp. 18–21), [8] (pp. 167–293).
Recent work has shown that compounds such as La1−xSrxMnO3 and Pr1−xBaxMnO3 [9,10] exhibit large magnetic entropy change. Structurally, a divalent Sr2+ and Ba2+ are doped in place of trivalent La3+ and Pr3+. The strongest “double exchange” interaction results are achieved at the doping level of about x = 0.3 [8]. Further introduction of divalent elements leads to antiferromagnetic (AFM) arrangement and can cause localization of charge—the so-called charge ordered (CO) state [11]. Because of this, further substitution of trivalent or divalent ions with different size ions can change the structure, volume and the magnetic properties.
In this work, we have prepared polycrystalline and nanocrystalline samples of Pr0.65Sr0.35−xCaxMnO3 (x = 0.02, 0.05, 0.1, 0.2, 0.3). Magnetic properties of the parent bulk compound Pr0.65Sr0.35MnO3 have been reported previously in the literature [12]. Ca2+ ions act as the substitute for divalent Sr2+ ions in order to bring TC down from 295 K of the parent compound Pr0.65Sr0.35MnO3 [12] in bulk and TC = 257 K of nano-sized compounds. Systems were crystallographically and morphologically investigated by X-ray diffraction (XRD), optical microscopy and transmission electron microscopy (TEM), and Rietlveld refinement analysis of XRD. Stoichiometry [13] and structure [8,9] of the samples greatly affects properties. Preparation method can affect stoichiometry of the compounds, sometimes causing accidental vacancies in Pr3+ ions. Some changes in magnetic and electrical behavior, therefore, could be observed, but are unquantifiable by this experiment. Additionally, oxygenation plays an important role in final measurements. Oxygen deficiency and excess change stoichiometry of the compounds by changing Mn3+/Mn4+ ratio electrical conductivity and overall magnetization. In our work, oxygen content was investigated by chemical analysis of Iodometry. Critical behavior was analyzed by construction of modified Arrott plots (MAP), which was also confirmed by the Kouvel–Fisher (KF) method. Nanocrystalline compounds have lower maximum entropy change values but much wider effective temperature range. Electrical measurements, in the bulk samples, revealed colossal magnetoresistance behavior [11,14,15].
This paper is organized as follows: in Section 2 we describe the preparation methods, as well as all the characterization methods we used in structural, morphological, oxygen stoichiometric, electrical, and magnetic investigation. In Section 3 we present the results of our investigation, analysis of data, and we discuss the magnetic critical behavior, electrical and magnetic properties of our samples. Finally, we summarize our results in Section 4.

2. Materials and Methods

Polycrystalline samples were prepared by conventional solid-state reaction. High purity oxides of principal elements Pr6O11, SrO, MnO2, and carbonate CaCO3 were purchased from Alfa Aesar (Heysham, UK). After being mixed by hand in a mortar for 3 h, the powders were calcinated at 1100 °C for 24 h in air. Later, the powder was pressed into a pellet at 3 tons and sintered in air at 1350 °C for 30 h.
Nanocrystalline samples were prepared by sucrose sol-gel method. Nitrates Pr(NO3)3 · 6H2O, Sr(NO3)2, Ca(NO3)2 · 4H2O, and Mn(NO3)2 · 6H2O were dissolved in pure water (18.2 MΩ × cm at 25 °C) for up to 1 h at 60 °C after which, 10 g of sucrose was added. This enables positive ions to attach themselves to OH hubs of the sucrose chain. After stirring for 45–60 min temperature was turn down to room temperature. Two grams of pectin was added to expand the xero-gel and mixed for further 20 min. Solutions were dried at 200 °C for 24 h or until all water evaporated. Finally, gels were burned at 1000 °C for 2 h to obtain the nano-sized particles. If the reaction does not happen, as was the case in this experiment, some small amount of Acetic acid can be added to the mix at the last stage before drying in order to bring the chains closer together. However, such action usually results in somewhat bigger particles.
X-ray diffraction (XRD) was implemented for structural categorization of all samples. Optical microscopy was used on bulk samples for grain size determination and surface defects. Transmission electron microscopy (TEM) allowed for nanocrystalline particle size determination. Analysis of XRD data was performed by Rietveld refinement method, as well as Williamson–Hall (W-H) method.
Iodometric analytical titration was applied in order to detect deficiency or excess of oxygen in the samples. A small amount of a sample was placed in a closed vessel containing hydrochloric acid (HCl) where positive Mn+ ions react with Cl to produce Cl2. An inert gas pushes Cl2 into another vessel containing potassium iodine (KI) where it reacts to produce I2. This mixture was titrated with sodium thiosulfate and the ratio of Mn3+ to Mn4+ was calculated by the stoichiometry of balanced chemical equations [16].
Electrical properties were found using the four-point technique in a cryogen-free superconducting setup. Four-point chips, measuring voltage and current separately were placed in applied magnetic fields of up to 5 T with a varied temperature between 10 K and 300 K. Resistance of the samples was recorded, resistivity was calculated using sample dimensions.
Magnetic measurements were made using a Vibrating Sample Magnetometer (VSM) in the temperature range of 4–300 K in external magnetic fields of up to 4 T.

3. Results

3.1. Structural Analysis

Visual inspection of stack XRD patterns, as shown in Figure 1, suggests possible decrease in cell dimensions due to the shift to the right of patterns with increasing Ca content. All samples are single phase with less than 5% levels of impurities. Wider peaks for nanocrystalline samples are indicative of smaller crystallites size in the compounds [17] which is further confirmed by Rietveld refinement analysis and Williamson–Hall method in Table 1 and Table 2.
Investigation of XRD data by Rietveld refinement analysis established orthorhombic (Pbnm) space group no. 62 for all samples, with decreasing cell dimensions and volume for each subsequent substitution of Sr. The results are illustrated in Figure 2. Stability of the structure in orthorhombic perovskite materials is assessed by the Goldschmidt tolerance factor. It was calculated using following relation [18,19]:
t = R A + R 0 2 R B + R 0 ,
where RA is the radius of A cation, RB is the radius of B cation, and R0 is the radius of the anion. All samples fall within orthorhombic/rhombohedral tolerance range of 0.7–1 [20] (pp. 707–714).
An addition of smaller crystal radius Ca2+ (1.32 Å) atoms in place of Sr2+ (1.45 Å) into the structure increases disorder [8,19] and decreases average Mn–O bond length, as seen in Table 1 and Table 2, which plays a crucial role in “double exchange” interaction [8]. The average angle between Mn–O–Mn stays relatively the same throughout the range of substitutions at 157.72(3)°.
The Williamson–Hall (W-H) method is advantageous in that it takes into account strain between crystallites, as opposed to the Scherrer method, which does not [21]. Thus, results form W-H analysis are expected to be closer to the real values. A comparison between Rietveld refinement analysis and Williamson–Hall method for approximation of crystallite size can be performed by contrasting them with TEM and optical microscopy results. As can be observed in Table 1 for bulk samples, W-H and Rietveld results are similar, but are two orders of magnitude different from real values. This can be attributed to each grain containing many crystallites. In Table 2, for nano-sized samples, it can be seen that values from W-H calculations are on average larger than those performed by Rietveld analysis and closer to values from TEM. Real sizes for nanocrystalline particles are in the upper limit of nano dimensions at approximately 70–80 nm. Selected samples of optical microscope pictures are presented in Figure 3 and selected TEM pictures in Figure 4.

3.2. Oxygen Content

Implementation of iodometric titration analysis for samples is a reliable method for determining deficiency or excess of oxygen in manganites [13,16]. All results are presented in Table 3.
Titration of bulk compounds revealed oxygen deficiency for all samples. An average of O2.93±0.02 for x = 0.02, similar to other samples, shows larger than expected deficiency of approximately O2.98 [22]. This is attributed to preparation methods and difficulty in oxygenating the sample during calcination and sintering. Lower levels of oxygen would affect magnetic and electrical properties as it affects the stoichiometry and changes Mn3+/Mn4+ ratio which would lower the amount of “double exchange” interaction [13].
All nanocrystalline compounds exhibited small excess of oxygen stoichiometry. The level of excess did not exceed O3.02±0.01 as is with x = 0.02 sample. We suggest that the culprit for such result should be found in the high surface to volume ratio of nano-sized particle. Broken bonds on the surface will increase the ratio of Mn4+ to Mn3+ by attracting oxygen in the air. This, in turn, can create vacancies and affect its magnetic and electrical properties.
Finally, it can be observed that standard deviation, or deviation from the “mean”, is larger for bulk samples than the nano-sized samples. This is explained by the variation of oxygenation within the bulk. Nevertheless, all relative standard deviations do not exceed 3% and can be considered reliable.

3.3. Electrical Measurements

An investigation of electrical behavior for polycrystalline samples x = 0.02; 0.05; 0.1 and 0.2 was performed using the four-point probe method in external fields of μ0H = 0 T; 1 T; 2 T in temperature range of 10–290 K. For the sample with x = 0.3 fields of up to 5 T were used. Data were used to estimate metal–insulator transition temperature Tp and magnetoresistivity MR according to the following equation [23]:
MR% = [(ρ(H) − ρ(0))/ρ(0)] × 100,
The first observation to be made is that all samples, except x = 0.3, exhibit a transition temperature Tp1 typically associated with grain boundary conditions [23,24,25]. Visual inspection shows that it is a wide and smooth transition, different from the usual sharp peaks associated with ferromagnetic–paramagnetic transition (and consequently metal-insulator) at Tp2 (inside of the grains). The addition of Ca ions increases disorder, which tends to spread itself toward the edges of the grains in order to conserve energy [26,27]. Boundaries act as insulators or semi-conductors [27,28]. As can be seen in Figure 5, the application of external field shifts Tp to the right. This is caused by lowering spin fluctuations and delocalization of charge carriers [19].
A separation between Tp1 and Tp2 (which is associated with TC) is reported when grain boundary conditions are strong enough due to mismatch in ionic radius at the A-site [26]. Samples x = 0.02 and 0.05 do not show a peak Tp2 because it is beyond the temperature range. Starting with x = 0.1 a sharp peak appears which lies higher in temperature than TC value. With application of a field, Tp2 tends to smooth out, as can be seen with the sample x = 0.2 in Figure 5b.
Special attention should be paid to the sample with x = 0.3. Its graph is presented in Figure 5c. Magnetic fields of up to 5 T were applied. For μ0H = 0 T; 1 T the sample conductivity is below the detection level for temperatures below 50 K. With the application of 2 T resistivity drops significantly. Its curve can be seen in the main portion of the figure. Curves for fields of 3–5 T can be seen in the inset and show further drop in resistivity and a shift to higher temperatures for the peak. This suggests that higher field overcomes intergrain resistance and further increase in the field induces a parallel alignment of manganese ions. In higher applied magnetic fields, the magnetic moments of the neighboring grains tend to be parallel, enhancing the tunneling of the conduction electrons similarly with a giant magnetoresistance effect (GMR) [26].
Magnetoresistivity is negative in all samples. Table 4 shows an increasing value of MR with increasing level of Ca substitution. The highest MRMax (2 T) is 51.96% and 29.08% for 1 T for x = 0.2 at 217 K. Sample with x = 0.02 exhibits the lowest MRMax of 11.9% and 4.45% for 2 T and 1 T at 289 K, respectively. A sample with x = 0.3 has MR of 77.62% between 3 T and 4 T at 118 K and 99.99% for 2–3 T at 130 K (shown in brackets in Table 4). In order to calculate MR, both the peak value of resistivity from the lower field data and an isothermal value from the higher field data are taken and then inserted into Equation (2). The highest peak resistivity ρpeak of about 3 Ωcm at 275 K is observed in sample x = 0.05. The lowest ρpeak for x = 0.3 at 5 T is 73 Ωcm (in brackets) at 174 K. All relevant data, including TC and Tp, are presented in Table 4.

3.4. Magnetic Properties

Investigation of magnetization vs. temperature (M vs. T), with the samples cooled in zero and in an applied magnetic field of μ0H = 0.05 T (ZFC-FC) was carried out in a vibrating sample magnetometer (VSM). All samples except x = 0.3 exhibit strong ferromagnetic behavior. Samples with x = 0.3 for both systems exhibit antiferromagnetic charge-exchange-type (CE-AFM) and charged ordered (CO) state. In CE-AFM structure, Mn3+ and Mn4+ are arranged like a checkerboard in the ab-plane (Mn-O2); exchange interaction causes Mn3+ eg electron to occupy either d3×2−r2 or d3y2−r2 orbital. As a result, there are FM zig-zag chain arrangements which are AFM to each other and also stack antiferromagnetically in the c-plane. Figure 6 presents selected graphs of M vs. T with insets representing a derivative dM/dT of the plots for which the minimum is associated with TC. The addition of Ca2+ ions causes smaller cell dimensions; increased disorder and it lowers the Curie temperature [17]. The increase in Ca substitution lowers TC further. Nano-sized particles show lower values of TC than their bulk counterpart due to their size and surface effect, including broken bonds, canted spins, and reduced magnetization [29,30]. In addition, the curves for nano-sized samples are “smoother”, covering a wider temperature range in their ferromagnetic–paramagnetic change but have lower maximum value of magnetization M(T).
An observation of ZFC-FC curves, especially for bulk samples, reveals an upturn in magnetization at lower temperatures at around 70–100 K, an example of which is in Figure 6a. This can be attributed to the magnetization of praseodymium Pr ions [10]. Samples with x = 0.3 also exhibit increased magnetization at low temperatures starting at around 100 K, seen in Figure 6c,d. Maximum value is not as high as for the rest of the samples, approximately 0.04 μB/f.u. for x = 0.3 vs. 1 μB/f.u. for x = 0.1. Other CO compounds, such as La0.4Ca0.6Mn0.9Ga0.1O3 [11] and Pr0.75Na0.25MnO3 [14] have been reported with similar increase in magnetization due to suppression of CO state. The difference between La3+ and Pr3+ is their crystal ionic size (1.356 Å for La3+ and 1.319 Å for Pr3+ [20]) and their electron configuration: La3+ has no 4f electrons while Pr3+ has the configuration = [Xe]4f2. The size difference affects the bandwidth, with Pr-based manganites being referred to as narrow bandwidth manganites. Appearance of the charged ordered state is influenced by Mn3+/Mn4+ ratio as well as the bandwidth [31], therefore Pr based manganites enter CO state more easily. The low temperature increase in magnetization is associated with phase separation (PS), i.e., existence of FM clusters, forming among AFM matrix [11]. The graphs for x = 0.3 at higher temperatures show behavior typical of charged ordered state. A small feature at around 170 K represents Neél temperature, TN (most visible in the bulk sample), and the peak at around 210 K representing the onset of the charge ordering phase, TCO. Such a behavior was revealed in several manganites systems such as La0.250Pr0.375Ca0.375MnO3 [32], Pr0.7Ca0.3MnO3 [33], La0.3Ca0.7Mn0.8Cr0.2O3 [34], Pr0.57Ca0.41Ba0.02MnO3 [35], Nd0.5Ca0.5MnO3 [36], and La0.4Ca0.6MnO3 [37], and it arises as a result of doping and/or due to the reducing of the sizes of the particles to the range of nanometers.
The feature from 45 K, in bulk sample, could be the signature of the blocking of isolated spins between FM clusters as seen in spin glass materials [38,39].
According to Landau’s mean field theory [40], Gibbs free energy of the system around a critical point can be expanded in a Taylor series as:
G (T,M) = G_0 + MH + aM2 + bM4 + …,
where coefficients a and b depend on temperature. A derivative of the energy with respect to magnetization, to find the minima, results in an expression:
H/M = 2a + 4bM2,
An easy way to confirm the correctness of the mean field theory approach for given samples is the Arrott plot, i.e., M2 vs. H/M [41]. If isotherms are straight and parallel to each other around TC then the assumption is correct. Observation of Arrott plots for our samples reveals curved, non-parallel lines for bulk compounds (Figure 7a,c) and almost straight lines for nano-sized compounds (Figure 7b,d). Furthermore, Banerjee criterion is a useful tool for determining the order of the phase transition [42]. According to the criterion, positive slope represents second order phase transition and negative slope corresponds to first order transition. Samples with x = 0, 0.02, 0.05, 0.1 show only positive slope. As can be seen in Figure 7c,d, samples with x = 0.2 show negative slope in the first stage of the magnetization at temperatures above TC suggesting first order transition. Samples with x = 0.3, not pictured, show mostly negative slope plots.
Inexactness of the exponents in the Equation (4) can be solved by constructing an Arrott–Noakes plot, i.e., M1/β vs. μ0H/M1/γ [43] with proper exponents β and γ. The exponent β relates to the spontaneous magnetization below TC and γ relates to the inverse susceptibility above TC [44] (pp. 7–10). An additional exponent δ relates magnetization and external field at TC. In mean field model: β = 0.5 and γ = 1, but there exist more models with different exponent values. The value of the exponents relates to the system dimensionality, spin, and range of the interaction J(r). In renormalization group theory [45], exchange interaction is defined as J(r) = 1/rd+σ where d—dimensionality of the system and σ—range of interaction. For σ greater than 2: β = 0.355, γ = 1.366 and δ = 4.8—is the case of the 3D Heisenberg model. If σ < 3/2, long-range interactions occur, according to mean field theory. For the tricritical point, the critical exponents are: β = 0.25, γ = 1, and δ = 5 [46]. The tricritical point sets a boundary between two different ranges of order phase transitions (first order and second order). These exponents can be generalized into following equations [43]:
MS(T) = M0 (−ε)β, T < TC,
χ 1 T = h 0 M 0 ε γ ,   T   >   T C ,
M = D0H1), T = TC,
where ε is the reduced temperature (TTC)/TC and M0, h0/M0, and D are critical amplitudes.
To find the proper exponents, we constructed Arrott–Noakes plots by implementing the Modified Arrott plot (MAP) method [43]. It is an iterative method. It includes, at first, construction of Arrott–Noakes plots by choosing exponents which produce straight parallel lines, with the line at TC crossing at the origin. Secondly, we find intercepts of lines around TC; on the abscissa, above Curie temperature, to obtain the values of χ0−1 and on the ordinate, below TC, to find the values of Ms. Finally, these values are plugged into the Equations (5) and (6) to obtain new values of β and γ. This process is repeated until results stabilize. The exponent δ is found using the Widom relation: β + γ = β δ [10]. Selected MAP graphs are shown in Figure 8. Full results for MAP are presented in Table 5.
Remarkably, polycrystalline samples are more closely governed by tricritical mean field model (β = 0.212, γ = 1.057, δ = 5.986 for x = 0.02), rather than 3D Heisenberg model as reported for other manganites in the literature [10,47]. Sample x = 0.3 does not exhibit ferromagnetic behavior in the high temperature range, only showing small magnetization at low temperatures due to FM clusters and Praseodymium ions, therefore, no critical values are presented in this work. Alternatively, all critical exponents for nano-sized particles fall within the mean field model values (β = 0.541, γ = 1.01, δ = 2.867 for x = 0.02) which is comparable to reported values in La0.7Ba0.3−xCaxMnO3 compounds [22]. Similarly, to the bulk, critical values for nano-sized sample x = 0.3 are not presented.
The Kouvel–Fisher (KF) method for determining critical exponents is a widely implemented tool for ferromagnetic materials [47,48,49]. We used KF to confirm the results from MAP as the two methods are regarded to be very accurate and reliable [47]. Akin to MAP, it is also an iterative method. It requires the construction of Arrott–Noakes plot, finding the intercepts on ordinate and abscissa and fitting these values in the equations [47,49]:
Ms {dMs/dT}−1 = (TTC)/β
χ0−1 {d χ0−1/dT}−1 = (TTC)/γ
Ideally, plot of Ms {dMs/dT}−1 vs. T is a straight line with slope = 1/β and the intercept giving TC/ β. Same logic applies to the plot of χ0−1 {d χ0−1/dT}−1 vs. T which gives slope equal to 1/γ. We present, in Figure 9, selected examples of results from KF calculations compared to results from MAP for the same compounds. It is evident, that the lines in KF plots are close to being straight and their slope values result in β and γ being close to results from MAP.
Magnetic entropy change was calculated using the following formula [50,51]:
Δ S m ( T , H 0 ) = S m ( T , H 0 ) S m ( T , 0 ) = 1 Δ T 0 H 0 M ( T + Δ T , H ) M ( T , H ) d H ,
where magnetization M0H) is taken from isothermal data at fields of 1–4 T.
Plots of −ΔSM vs. T (temperature) were constructed in order to better estimate magnetocaloric effect of the compounds, including calculations of relative cooling power RCP [51,52]:
R C P S = Δ S m T , H × δ T F W H M
where δ T F W H M is the range of temperature at full width half maximum.
Selected graphs of samples’ entropy change are presented in Figure 10. It is noteworthy that, generally, polycrystalline samples exhibit higher maximum entropy change −ΔSM compared to their nano-sized counterparts, as seen in Figure 10a vs. 10b. Bulk compound with x = 0.05 shows −ΔSM (max) = 5.56 J/kgK at 4 T and nano-sized sample shows −ΔSM (max) = 3.25 J/kgK at 4 T. Bulk samples with x = 0.1 and x = 0.2 exhibit high maximum entropy change at 6.9 J/kgK and 9.2 J/kgK respectively. For all samples, maximum entropy change occurs at around their Curie temperature TC. It is also important to note that nanocrystalline compounds exhibit wider range of effective temperature δ T F W H M at around 40–70 K while bulk samples show range of 20–30 K.
Construction of cooling equipment should not be “blindly” reliant on reported values of RCP [53]; nevertheless, for a long time it has been a useful tool in estimating the applicability of the magnetic material. Although it is the width of the temperature change in nano-sized compounds that is criticized as unreliable, in this work, nanocrystalline compounds deserve attention for their satisfactory level of entropy change, greater or close to 4 J/(kgK) for Δμ0H = 4 T. Both nano-sized and bulk systems exhibit values of entropy change and RCP comparable with other manganites reported in the literature [10,52,54,55].
Delving deeper, unfortunately, RCP tends to overestimate the merits of the materials which have a large temperature range of magnetocaloric effect (δTFWHM) but small entropy changes [55,56] in no small part because materials with same relative cooling power can behave differently in a magnetic cooling simulation [57]. Moreover, one of the last printed books about magnetic cooling ignores this figure of merit [58].
In order to be of use in magnetic refrigeration applications the chosen materials have to show a large magnetocaloric effect—a large magnetic entropy change. Besides this, there is a list of requirements related with heat transfer, heat capacity, heat conductivity, chemical and mechanical stability, hysteresis losses, eddy currents losses, etc. Therefore, magnetic measurements can be a trusted guide in deciding if a material deserves the effort of complete characterization. Some compounds of the type A1−xBxMnO3 can be considered to be of interest for magnetic refrigeration because some of these compounds show −ΔSm values (−ΔSm ≈ 4–6 J·kg−1·K−1 for ΔH = 5 T) comparable to intermetallic alloys [56]. This is the case of our investigated compounds.
The best way to decide the practicality of a material in magnetic cooling applications is by testing directly in a magnetic refrigerator, working on the principle of the AMRR (Active Magnetic Regenerative Refrigeration) cycle [12]. Recent investigations confirmed the magnetocaloric performance and the potential in magnetic refrigeration of the perovskite oxide Pr0.65Sr0.35MnO3 [12,59], which has the stoichiometry of our parent compound. When Ca substitutes for Sr in this compound, the magnetic parameters and magnetocaloric effect can be tuned for various temperature applications without significant changes in the magnitude of magnetic entropy change. Our investigations support the bulk polycrystalline Pr0.65Sr(0.35−x)CaxMnO3 compounds to be promising for magnetic cooling technology, while lower magnitude of magnetocaloric effect in nanocrystalline samples, unfortunately, seems to somewhat hinder their direct practical applications [53,56,57].
Special mention should be made of some feature for samples with x = 0.3. Both bulk (Figure 10c) and nano-sized (not pictured) compounds exhibit positive entropy change at temperatures associated with antiferromagnetic (CO)/ferromagnetic/paramagnetic phase transition. This phase transition can be observed in the ZFC-FC plots in Figure 6 at around 200–210 K for both samples. As can be observed in Figure 10c, bulk sample exhibits inverse (materials cool down when a magnetic field is adiabatically applied) to normal MCE change suggesting AFM/FM transition, but it seems the changes in the lattice parameters can also modify the magnetic exchange interaction to give rise to a such behavior [60,61]. Additionally, Figure 10d shows a negative entropy change of 5.1 J/kgK at 4 T for nano-sized samples at 75 K. Similar behavior and even higher entropy change is exhibited by the bulk sample at around 40 K (not pictured) at the suppression of the AFM state. These materials can be used to produce both cooling and heating when they are adiabatically demagnetized. Unfortunately, the large values of the magnetic entropy changes can be reach only at low temperatures, far from the range of domestic refrigeration, in the case of these samples.
Low coercivity is of utmost importance in application of cooling materials [51]. All samples exhibit low coercive fields as measured in a hysteresis loop at external fields of up to 4 T at 4 K. Largest coercive field exhibited by a bulk sample is of 190 Oe for x = 0.2; with the smallest being 130 Oe for x = 0.05. Nanocrystalline samples carry larger coercive field compared to polycrystalline samples. It increases with decreasing size of the particles until single domain size. Particles become superparamagnetic with further decrease in size [62]. Largest coercive field is produced by the parent nano-sized with x = 0 at 810 Oe. All values for coercive field and magnetic saturation Ms are presented in Table 6 and Table 7. Nanocrystalline sample x = 0.3 does not reach saturation at 4 T.

4. Conclusions

Two sample systems of Pr0.65Sr0.35−xCaxMnO3 (x = 0.02, 0.05, 0.1, 0.2, 0.3) were prepared. Solid state reaction as a method of preparation of polycrystalline compounds resulted in samples with an average 12,000 nm grain size. Sucrose based sol–gel method for production of nanocrystalline samples resulted in particles of an average size of approximately 70 nm. X-ray diffraction measurements revealed a single crystallographic phase for all samples. Rietveld refinement analysis confirmed single phase, orthorhombic (Pbnm) symmetry, diminishing cell volume and Mn–O bond length with increasing Ca substitution. Iodometric titration was implemented on both systems to determine their oxygen content. All bulk samples show oxygen deficiency close to O2.93±0.02 attributed to the preparation method. All nano-sized samples exhibit small oxygen excess of O3.02±0.01 or less, credited to the effects of the surface/volume ratio of the particles. Bulk compounds were measured for their electrical properties and revealed a feature at a temperature Tp1 associated with high disorder at grain boundaries due to Ca substitution. Additionally, samples with Ca levels x = 0.1, 0.2 exhibit a peak at a temperature Tp2 associated with the ferromagnetic-paramagnetic transition. With the increase in applied magnetic field, Tp2 tends to become higher and smooth out. The sample with x = 0.3 possesses “infinite” resistance for μ0H = 0 T and 1 T at around 50 K. Further increase in applied field up to 5 T results in overcoming intergrain resistance and shift in the resistance peak. Negative magnetoresistivity was observed for all bulk samples. The smallest value of MR was 11.99% for x = 0.02 at 2 T at Tp1, while the largest value of MR for x = 0.2 of 51.96% for Tp2. Zero-field cooled–field cooled graphs reveal ferromagnetic behavior for x = 0.02, 0.05, 0.01, 0.2 samples. Both x = 0.3 show ferromagnetic-like behavior at low temperatures and antiferromagnetic behavior at higher temperatures than 80 K, with transition to ferromagnetic to paramagnetic behavior at 210 K. Curie temperature TC lowers with each Ca substitution: 273 K for bulk x = 0.02 compared to 295 K for parent compound; 252 K for nano-sized x = 0.02 vs. 257 K for parent compound. Low coercivity was found for all samples. Nanocrystalline samples exhibiting larger coercivity compared to their bulk counterparts: 720 Oe vs. 180 Oe for x = 0.02. Arrott plots confirm second order phase transition for all samples except for x = 0.3. Modified Arrott plot (MAP) analysis disclosed critical behavior. Critical exponents for polycrystalline samples belong to tricritical mean field model while exponents for nano-sized particles belong to mean field model. The Kouvel–Fisher (KF) method for analyzing critical exponents confirmed MAP results. Bulk compounds reveal higher magnetic saturation Ms and maximum entropy change ΔSM than nano-sized compounds. Bulk sample with x = 0.2 exhibits highest entropy change ΔSM = 9.21 J/kgK (4 T) at 204 K. Relative cooling powers (RCP) for equivalent bulk and nano-sized compounds are comparable in value: 166 J/kg vs. 175 J/kg (4 T) for bulk and nano-sized with x = 0.02 This is due to the wide effective temperature range of entropy change δTFWHM in nanocrystalline samples. All values of RCP are high and comparable with relevant compounds reported in the literature. Bulk samples with x = 0.3 exhibits positive entropy change at antiferromagnetic-ferromagnetic transition temperature Magnetic entropy changes and magnetic parameters of the bulk polycrystalline samples indicate them as potential candidates for magnetocaloric materials. Possibly, they can be combined in construction of multistep refrigeration processes to increase their temperature range and effectiveness. The nanocrystalline samples, which show comparatively lower maximum entropy change are moved to the immediate background of the picture of usable magnetocaloric compounds but could, possibly, be still useful in future developments.

Author Contributions

R.A., conceptualization, investigation, methodology, writing—original draft, writing—review and editing, visualization, supervision; D.A., R.H., R.B., G.S., A.S. and L.B.-T., methodology, investigation, review and editing; I.G.D., conceptualization, investigation, methodology, visualization, supervision, review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data presented in this study are available in this article.

Acknowledgments

The authors acknowledge the Institute of Physics “Ioan Ursu” of Faculty of Physics from Babes Bolyai University for assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mckay, C.P. Requirements and limits for life in the context of exoplanets. Proc. Natl. Acad. Sci. USA 2014, 111, 12628–12633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Briley, G.C. A History of Refrigeration. Ashrae J. 2004, 46, S31–S34. [Google Scholar]
  3. Gschneidner, K.A.; Pecharsky, V.K. Thirty Years of near Room Temperature Magnetic Cooling: Where We Are Today and Future Prospects. Int. J. Refrig. 2008, 31, 945–961. [Google Scholar] [CrossRef] [Green Version]
  4. Salazar-Munoz, V.E.; Lobo Guerrero, A.; Palomares-Sanchez, S.A. Review of magnetocaloric properties in lanthanum manganites. J. Magn. Magn. Mater. 2022, 562, 169787. [Google Scholar] [CrossRef]
  5. Smith, A. Who discovered the magnetocaloric effect? EPJH 2013, 38, 507–517. [Google Scholar] [CrossRef]
  6. Dagotto, E.; Hotta, T.; Moreo, A. Colossal Magnetoresistant Materials: The Key Role of Phase Separation. Phys. Rep. 2001, 344, 1–153. [Google Scholar] [CrossRef] [Green Version]
  7. Pavarini, E.; Koch, E.; Anders, F.; Jarrell, M. Correlated Electrons: From Models to Materials Modeling and Simulation; Forschungszentrum Julich: Jülich, Germany, 2012; Chapter 7; Volume 2, pp. 18–21. ISBN 978-3-89336-796-2. [Google Scholar]
  8. Coey, J.M.D.; Viret, M.; Von Molnár, S. Mixed-Valence Manganites. Adv. Phys. 1999, 48, 167–293. [Google Scholar] [CrossRef]
  9. Rostamnejadi, A.; Venkatesan, M.; Alaria, J.; Boese, M.; Kameli, P.; Salamati, H.; Coey, J.M.D. Conventional and Inverse Magnetocaloric Effects in La0.45Sr 0.55MnO3 Nanoparticles. J. Appl. Phys. 2011, 110, 043905. [Google Scholar] [CrossRef] [Green Version]
  10. Varvescu, A.; Deac, I.G. Critical Magnetic Behavior and Large Magnetocaloric Effect in Pr0.67Ba0.33MnO3 Perovskite Manganite. Phys. B Condens. Matter 2015, 470–471, 96–101. [Google Scholar] [CrossRef]
  11. Badea, C.; Tetean, R.; Deac, I.G. Suppression of Charge and Antiferromagnetic Ordering in Ga-doped La0.4Ca0.6MnO3. Rom. J. Phys. 2018, 63, 604. [Google Scholar]
  12. Guillou, F.; Legait, U.; Kedous-Lebouc, A.; Hardy, V. Development of a new magnetocaloric material used in a magnetic refrigeration device. EPJ Web Conf. 2012, 29, 21. [Google Scholar] [CrossRef] [Green Version]
  13. Licci, F.; Turilli, G.; Ferro, P. Determination of Manganese Valence in Complex La-Mn Perovskites. J. Magn. Magn. Mater. 1996, 164, L268–L272. [Google Scholar] [CrossRef]
  14. Zhang, X.H.; Li, Z.Q.; Song, W.; Du, X.W.; Wu, P.; Bai, H.L.; Jiang, H.Y. Magnetic properties and charge ordering in Pr0.75Na0.25MnO3 manganite. Solid State Commun. 2005, 135, 356. [Google Scholar] [CrossRef]
  15. Rao, C.N.R. Charge, Spin, and Orbital Ordering in the Perovskite Manganates, Ln1−xAxMnO3 (Ln = Rare Earth, A = Ca or Sr). J. Phys. Chem. B 2000, 104, 5877–5889. [Google Scholar] [CrossRef]
  16. Tali, R. Determination of Average Oxidation State of Mn in ScMnO3 and CaMnO3 by Using Iodometric Titration. Damascus Univ. J. Basic Sci. 2007, 23, 9–19. [Google Scholar]
  17. Ehsani, M.H.; Kameli, P.; Ghazi, M.E.; Razavi, F.S.; Taheri, M. Tunable magnetic and magnetocaloric properties of La0.6Sr0.4MnO3 nanoparticles. J. Appl. Phys. 2013, 114, 223907. [Google Scholar] [CrossRef]
  18. Dagotto, E. Nanoscale Phase Separation and Colossal Magnetoresistance, 1st ed.; Springer Science & Business Media: New York, NY, USA, 2002; pp. 271–284. [Google Scholar]
  19. Raju, K.; Manjunathrao, S.; Venugopal Reddy, P. Correlation between Charge, Spin and Lattice in La-Eu-Sr Manganites. J. Low Temp. Phys. 2012, 168, 334–349. [Google Scholar] [CrossRef]
  20. Rao, C.N.R. Perovskites. In Encyclopedia of Physical Science and Technology; Elsevier: Amsterdam, The Netherlands, 2003; pp. 707–714. [Google Scholar]
  21. Nath, D.; Singh, F.; Das, R. X-ray Diffraction Analysis by Williamson-Hall, Halder-Wagner and Size-Strain Plot Methods of CdSe Nanoparticles—A Comparative Study. Mater. Chem. Phys. 2020, 239, 2764–2772. [Google Scholar] [CrossRef]
  22. Atanasov, R.; Bortnic, R.; Hirian, R.; Covaci, E.; Frentiu, T.; Popa, F.; Iosif Grigore Deac, I.G. Magnetic and Magnetocaloric Properties of Nano- and Polycrystalline Manganites La(0.7−x)EuxBa0.3MnO3. Materials 2022, 15, 7645. [Google Scholar] [CrossRef]
  23. Saw, A.K.; Channagoudra, G.; Hunagund, S.; Hadimani, R.L.; Dayal, V. Study of transport, magnetic and magnetocaloric properties in Sr2+ substituted praseodymium manganite. Mater. Res. Express 2020, 7, 016105. [Google Scholar] [CrossRef]
  24. Deac, I.G.; Tetean, R.; Burzo, E. Phase Separation, Transport and Magnetic Properties of La2/3A1/3Mn1−XCoxO3, A = Ca, Sr (0.5 ≤ x ≤ 1). Phys. B Condens. Matter 2008, 403, 1622–1624. [Google Scholar] [CrossRef]
  25. Panwar, N.; Pandya, D.K.; Agarwal, S.K. Magneto-Transport and Magnetization Studies of Pr2/3Ba 1/3MnO3:Ag2O Composite Manganites. J. Phys. Condens. Matter 2007, 19, 456224. [Google Scholar] [CrossRef]
  26. Panwar, N.; Pandya, D.K.; Rao, A.; Wu, K.K.; Kaurav, N.; Kuo, Y.K.; Agarwal, S.K. Electrical and Thermal Properties of Pr 2/3(Ba1−XCsx)1/3MnO 3 Manganites. Eur. Phys. J. B 2008, 65, 179–186. [Google Scholar] [CrossRef]
  27. Ibrahim, N.; Rusop, N.A.M.; Rozilah, R.; Asmira, N.; Yahya, A.K. Effect of grain modification on electrical transport properties and electroresistance behavior of Sm0.55Sr0.45MnO3. Int. J. Eng. Technol. 2018, 7, 113–117. [Google Scholar]
  28. Kubo, K.; Ohata, N. A Quantum Theory of Double Exchange. J. Phys. Soc. Jpn. 1972, 33, 21–32. [Google Scholar] [CrossRef]
  29. Arun, B.; Suneesh, M.V.; Vasundhara, M. Comparative Study of Magnetic Ordering and Electrical Transport in Bulk and Nano-Grained Nd0.67Sr0.33MnO3 Manganites. J. Magn. Magn. Mater. 2016, 418, 265–272. [Google Scholar] [CrossRef]
  30. Peters, J.A. Relaxivity of Manganese Ferrite Nanoparticles. Prog. Nucl. Magn. Reson. Spectrosc. 2020, 120, 72–94. [Google Scholar] [CrossRef] [PubMed]
  31. Maheswar Repaka, D.V. Magnetic Control of Spin Entropy, Thermoelectricity and Electrical Resistivity in Selected Manganites. Ph.D. Thesis, National University of Singapore, Singapore, 2014. [Google Scholar]
  32. Deac, I.G.; Diaz, S.V.; Kim, B.G.; Cheong, S.-W.; Schiffer, P. Magnetic relaxation in La0.250Pr0.375Ca0.375MnO3 with varying phase separation. Phys. Rev. B 2002, 64, 174426. [Google Scholar] [CrossRef] [Green Version]
  33. Deac, I.G.; Mitchell, J.; Schiffer, P. Phase Separation and the Low-Field Bulk Magnetic Properties of Pr0.7Ca0.3MnO3. Phys. Rev. B 2001, 63, 172408. [Google Scholar] [CrossRef] [Green Version]
  34. Sudyoadsuka, T.; Suryanarayananb, R.; Winotaia, P.; Wengerc, L.E. Suppression of charge-ordering and appearance of magnetoresistance in a spin-cluster glass manganite La0.3Ca0.7Mn0.8Cr0.2O3. J. Magn. Magn. Mater. 2004, 278, 96–106. [Google Scholar] [CrossRef] [Green Version]
  35. Anuradha, K.N.; Rao, S.S.; Bhat, S.V. Complete ‘melting’ of charge order in hydrothermally grown Pr0.57Ca0.41Ba0.02MnO3 nanowires. J. Nanosci. Nanotechnol. 2007, 7, 1775–1778. [Google Scholar] [CrossRef] [Green Version]
  36. Rao, S.S.; Tripathi, S.; Pandey, D.; Bhat, S.V. Suppression of charge order, disappearance of antiferromagnetism, and emergence of ferromagnetism in Nd0.5Ca0.5MnO3 nanoparticles. Phys. Rev. B 2006, 74, 144416. [Google Scholar] [CrossRef]
  37. Lu, C.L.; Dong, S.; Wang, K.F.; Gao, F.; Li, P.L.; Lv, L.Y.; Liu, J.M. Charge-order breaking and ferromagnetism in La0.4Ca0.6MnO3 nanoparticles. Appl. Phys. Lett. 2007, 91, 032502. [Google Scholar] [CrossRef] [Green Version]
  38. Hüser, D.; Wenger, L.E.; van Duynevelt, A.J.; Mydosh, J.A. Dynamical behavior of the susceptibility around the freezing temperature in (Eu,Sr)S. Phys. Rev. B 1983, 27, 3100. [Google Scholar] [CrossRef]
  39. Cao, G.; Zhang, J.; Wang, S.; Yu, J.; Jing, C.; Cao, S.; Shen, X. Reentrant spin glass behavior in CE-type AFM Pr0.5Ca0.5MnO3 manganite. J. Magn. Magn. Mater. 2007, 310, 169. [Google Scholar] [CrossRef]
  40. Jeddi, M.; Gharsallah, H.; Bejar, M.; Bekri, M.; Dhahri, E.; Hlil, E.K. Magnetocaloric Study, Critical Behavior and Spontaneous Magnetization Estimation in La0.6Ca0.3Sr0.1MnO3 Perovskite. RSC Adv. 2018, 8, 9430–9439. [Google Scholar] [CrossRef] [Green Version]
  41. Arrott, A. Criterion for Ferromagnetism from Observations of Magnetic Isotherms. Phys. Rev. 1957, 108, 1394–1396. [Google Scholar] [CrossRef]
  42. Banerjee, B.K. On a Generalised Approach to First and Second Order Magnetic Transitions. Phys. Lett. 1964, 12, 16–17. [Google Scholar] [CrossRef]
  43. Arrott, A.; Noakes, J.E. Approximate Equation of State for Nickel Near Its Critical Temperature. Phys. Rev. Lett. 1967, 19, 786–789. [Google Scholar] [CrossRef]
  44. Stanley, H.E. Introduction to Phase Transitions and Critical Phenomena; Oxford University Press: Oxford, UK, 1987; pp. 7–10. [Google Scholar]
  45. Fisher, M.E.; Ma, S.K.; Nickel, B.G. Critical Exponents for Long-Range Interactions. Phys. Rev. Lett. 1972, 29, 917. [Google Scholar] [CrossRef] [Green Version]
  46. Pathria, R.K.; Beale, P.D. Phase Transitions: Criticality, Universality, and Scaling. In Statistical Mechanics; Elsevier: Amsterdam, The Netherlands, 2022; pp. 417–486. [Google Scholar]
  47. Vadnala, S.; Asthana, S. Magnetocaloric effect and critical field analysis in Eu substituted La0.7−xEuxSr0.3MnO3 (x = 0.0, 0.1, 0.2, 0.3) manganites. J. Magn. Magn. Mater. 2018, 446, 68–79. [Google Scholar] [CrossRef]
  48. Kim, D.; Revaz, B.; Zink, B.L.; Hellman, F.; Rhyne, J.J.; Mitchell, J.F. Tri-critical Point and the Doping Dependence of the Order of the Ferromagnetic Phase Transition of La1−xCaxMnO3. Phys. Rev. Lett. 2002, 89, 227202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Pelka, R.; Konieczny, P.; Fitta, M.; Czapla, M.; Zielinski, P.M.; Balanda, M.; Wasiutynski, T.; Miyazaki, Y.; Inaba, A.; Pinkowicz, D.; et al. Magnetic systems at criticality: Different signatures of scaling. Acta Phys. Pol. 2013, 124, 977. [Google Scholar] [CrossRef]
  50. Souca, G.; Iamandi, S.; Mazilu, C.; Dudric, R.; Tetean, R. Magnetocaloric Effect and Magnetic Properties of Pr1−XCexCo3 Compounds. Stud. Univ. Babeș-Bolyai Phys. 2018, 63, 9–18. [Google Scholar] [CrossRef]
  51. Zverev, V.; Tishin, A.M. Magnetocaloric Effect: From Theory to Practice. In Reference Module in Materials Science and Material Engineering; Elsevier: Amsterdam, The Netherlands, 2016; pp. 5035–5041. [Google Scholar] [CrossRef]
  52. Deac, I.G.; Vladescu, A. Magnetic and magnetocaloric properties of Pr1−xSrxCoO3 cobaltites. J. Magn. Magn. Mater. 2014, 365, 1–7. [Google Scholar] [CrossRef]
  53. Griffith, L.D.; Mudryk, Y.; Slaughter, J.; Pecharsky, V.K. Material-based figure of merit for caloric materials. J. Appl. Phys. 2018, 123, 034902. [Google Scholar] [CrossRef]
  54. Naik, V.B.; Barik, S.K.; Mahendiran, R.; Raveau, B. Magnetic and Calorimetric Investigations of Inverse Magnetocaloric Effect in Pr0.46Sr0.54MnO3. Appl. Phys. Lett. 2011, 98, 112506. [Google Scholar] [CrossRef]
  55. Gong, Z.; Xu, W.; Liedienov, N.A.; Butenko, D.S.; Zatovsky, I.V.; Gural’skiy, I.A.; Wei, Z.; Li, Q.; Liu, B.; Batman, Y.A.; et al. Expansion of the multifunctionality in off-stoichiometric manganites using post-annealing and high pressure: Physical and electrochemical studies. Phys. Chem. Chem. Phys. 2002, 24, 21872–21885. [Google Scholar] [CrossRef]
  56. Sandeman, K.G. Magnetocaloric materials: The search for new systems. Scr. Mater. 2012, 67, 566–571. [Google Scholar] [CrossRef] [Green Version]
  57. Smith, A.; Bahl, A.C.R.H.; Bjørk, R.; Engelbrecht, K.; Kaspar, K.; Nielsen, K.K.; Pryds, N. Materials challenges for high performance magnetocaloric refrigeration devices. Adv. Energy Mater. 2012, 2, 1288–1318. [Google Scholar] [CrossRef]
  58. Kitanovski, A.; Tusek, J.; Tomc, U.; Plaznik, U.; Ozbolt, M.; Poredos, A. Magentocaloric Energy Conversion: From Theory to Applications, 1st ed.; Springer: Cham, Switzerland, 2015. [Google Scholar] [CrossRef]
  59. Legait, U.; Guillou, F.; Kedous-Lebouc, A.; Hardy, A.V.; Almanza, M. An experimental comparison of four magnetocaloric regenerators using three different materials. Int. J. Refrig. 2014, 37, 147–155. [Google Scholar] [CrossRef]
  60. Krishnamoorthi, C.; Barik, S.K.; Siu, Z.; Mahendiran, R. Normal and inverse magnetocaloric effects in La0.5Ca0.5Mn1−xNixO3. Solid State Commun. 2010, 150, 1670–1673. [Google Scholar] [CrossRef]
  61. von Ranke, P.J.; de Oliveira, N.A.; Alho, B.P.; Plaza, E.J.R.; de Sousa, V.S.R.; Caron, L.; Reis, M.S. Understanding the inverse magnetocaloriceffect in antiferro- and ferrimagnetic arrangements. J. Phys. Condens. Matter 2009, 21, 056004. [Google Scholar] [CrossRef] [PubMed]
  62. Majumder, D.D.; Majumder, D.D.; Karan, S. Magnetic Properties of Ceramic Nanocomposites. In Ceramic Nanocomposites; Banerjee, R., Manna, I., Eds.; Woodhead Publishing Series in Composites Science and Engineering; Elsevier B.V.: Amsterdam, The Netherlands, 2013; pp. 51–91. [Google Scholar] [CrossRef]
  63. Hueso, L.E.; Sande, P.; Miguéns, D.R.; Rivas, J.; Rivadulla, F.; López-Quintela, M.A. Tuning of the magnetocaloric effect in nanoparticles synthesized by sol–gel techniques. J. Appl. Phys. 2002, 91, 9943–9947. [Google Scholar] [CrossRef]
  64. Anis, B.; Tapas, S.; Banerjee, S.; Das, I. Magnetocaloric properties of nanocrystalline Pr0.65(Ca0.6Sr0.4)0.35MnO3. J. Appl. Phys. 2008, 103, 013912. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns for (a) Pr0.65Sr0.35−xCaxMnO3 polycrystalline bulk samples and (b) Pr0.65Sr0.35−xCaxMnO3 nano-sized samples.
Figure 1. X-ray diffraction patterns for (a) Pr0.65Sr0.35−xCaxMnO3 polycrystalline bulk samples and (b) Pr0.65Sr0.35−xCaxMnO3 nano-sized samples.
Nanomaterials 13 01373 g001
Figure 2. Plot of cell volume change for polycrystalline and nanocrystalline samples.
Figure 2. Plot of cell volume change for polycrystalline and nanocrystalline samples.
Nanomaterials 13 01373 g002
Figure 3. Selected optical microscope pictures for Pr0.7Sr0.3−xCaxMnO3 bulk samples (a) for x = 0.02, (b) x = 0.05.
Figure 3. Selected optical microscope pictures for Pr0.7Sr0.3−xCaxMnO3 bulk samples (a) for x = 0.02, (b) x = 0.05.
Nanomaterials 13 01373 g003
Figure 4. Selected TEM pictures for Pr0.7Sr0.3−xCaxMnO3 nano-sized samples for x = 0.02 (a), x = 0.01 (b).
Figure 4. Selected TEM pictures for Pr0.7Sr0.3−xCaxMnO3 nano-sized samples for x = 0.02 (a), x = 0.01 (b).
Nanomaterials 13 01373 g004
Figure 5. Graphs of resistivity vs. temperature for bulk samples: (a) x = 0.02, (b) x = 0.2, (c) x = 0.3; The inset shows plot of resistivity at higher fields which are not visible in the main plot.
Figure 5. Graphs of resistivity vs. temperature for bulk samples: (a) x = 0.02, (b) x = 0.2, (c) x = 0.3; The inset shows plot of resistivity at higher fields which are not visible in the main plot.
Nanomaterials 13 01373 g005
Figure 6. ZFC-FC plots in external field of 0.05 T (a) ZFC-FC curves and derivative of magnetization (inset) for the bulk sample Pr0.65Sr0.25Ca0.1MnO3; (b) ZFC-FC curves and derivative (inset) for nanocrystalline sample Pr0.65Sr0.25Ca0.1MnO3. (c) ZFC-FC curves and derivative shown as inset, for polycrystalline sample Pr0.65Sr0.05Ca0.3MnO3; (d) ZFC-FC curves and derivative shown in the inset for nanocrystalline sample Pr0.65Sr0.05Ca0.3MnO3.
Figure 6. ZFC-FC plots in external field of 0.05 T (a) ZFC-FC curves and derivative of magnetization (inset) for the bulk sample Pr0.65Sr0.25Ca0.1MnO3; (b) ZFC-FC curves and derivative (inset) for nanocrystalline sample Pr0.65Sr0.25Ca0.1MnO3. (c) ZFC-FC curves and derivative shown as inset, for polycrystalline sample Pr0.65Sr0.05Ca0.3MnO3; (d) ZFC-FC curves and derivative shown in the inset for nanocrystalline sample Pr0.65Sr0.05Ca0.3MnO3.
Nanomaterials 13 01373 g006
Figure 7. Arrott plot (M2 vs. H/M) for (a) the bulk sample with x = 0.02; (b) the nanocrystalline sample with x = 0.02; (c) bulk sample with x = 0.2; (d) nanocrystalline sample x = 0.2.
Figure 7. Arrott plot (M2 vs. H/M) for (a) the bulk sample with x = 0.02; (b) the nanocrystalline sample with x = 0.02; (c) bulk sample with x = 0.2; (d) nanocrystalline sample x = 0.2.
Nanomaterials 13 01373 g007
Figure 8. Modified Arrott plots for (a) the bulk sample x = 0.02 with β = 0.212, γ = 1.057, δ = 5.986 and for (b) the nanocrystalline sample x = 0.02 with β = 0.541, γ = 1.01, δ = 2.867.
Figure 8. Modified Arrott plots for (a) the bulk sample x = 0.02 with β = 0.212, γ = 1.057, δ = 5.986 and for (b) the nanocrystalline sample x = 0.02 with β = 0.541, γ = 1.01, δ = 2.867.
Nanomaterials 13 01373 g008
Figure 9. Calculated values for critical exponents for (a) MAP analysis for the bulk sample with x = 0.02 (b) KF analysis for the bulk sample with x = 0.02. (c) MAP analysis for the nanocrystalline sample x = 0.02 (d) KF analysis for nanocrystalline sample x = 0.02.
Figure 9. Calculated values for critical exponents for (a) MAP analysis for the bulk sample with x = 0.02 (b) KF analysis for the bulk sample with x = 0.02. (c) MAP analysis for the nanocrystalline sample x = 0.02 (d) KF analysis for nanocrystalline sample x = 0.02.
Nanomaterials 13 01373 g009
Figure 10. Magnetic entropy change vs. temperature for selected samples: (a) x = 0.05 bulk; (b) x = 0.05 nanocrystalline sample; (c) x = 0.3 bulk at AFM-FM; (d) x = 0.3 nanocrystalline at low temperature.
Figure 10. Magnetic entropy change vs. temperature for selected samples: (a) x = 0.05 bulk; (b) x = 0.05 nanocrystalline sample; (c) x = 0.3 bulk at AFM-FM; (d) x = 0.3 nanocrystalline at low temperature.
Nanomaterials 13 01373 g010
Table 1. Calculated tolerance factors, Mn-O lengths, and crystallite sizes for polycrystalline samples using the Williamson–Hall and Rietveld methods and including particle diameters by optical microscopy.
Table 1. Calculated tolerance factors, Mn-O lengths, and crystallite sizes for polycrystalline samples using the Williamson–Hall and Rietveld methods and including particle diameters by optical microscopy.
Ca Content (Bulk)t (Tolerance Factor)Mn-O (Å)Mn-O-MnAverage Particle Diameter (μm)Williamson–Hall Size (nm)Average Rietveld Size (nm)Strain
(°)
x = 0.020.9251.964157.7315111.3396.670.0022
x = 0.050.9241.962157.7411156.63113.680.0019
x = 0.10.9211.961157.7313132.88192.590.0024
x = 0.20.9171.956157.6910144.85125.640.0023
x = 0.30.9121.954157.7211123.5686.790.0021
Table 2. Calculated Mn-O lengths and crystallite sizes for nanocrystalline samples using the Williamson–Hall and Rietveld methods and including grain diameters from TEM.
Table 2. Calculated Mn-O lengths and crystallite sizes for nanocrystalline samples using the Williamson–Hall and Rietveld methods and including grain diameters from TEM.
Ca Content (Nano)Mn-O (Å)Mn-O-MnAverage Particle Diameter (nm)Williamson–Hall Size (nm)Average Rietveld Size (nm)Strain
(°)
x = 0 1.972157.7468.172.4554.330.0018
x = 0.021.964157.7564.871.1545.640.0019
x = 0.051.958157.6978.284.5957.210.0016
x = 0.11.957157.6987.582.9539.750.0019
x = 0.21.955157.771.566.7345.620.0017
x = 0.31.951157.7165.769.6951.110.002
Table 3. Average oxygen content calculated using iodometry for bulk and nanocrystalline samples.
Table 3. Average oxygen content calculated using iodometry for bulk and nanocrystalline samples.
Ca ContentAverage Mn3+/Mn4+ RatioStandard DeviationRelative Standard Deviation (%)Average Oxygen Content
x = 0.02 bulk0.7880.01672.12O2.93±0.02
x = 0.05 bulk0.7830.01211.55O2.93±0.01
x = 0.1 bulk0.7780.01391.79O2.94±0.02
x = 0.2 bulk0.7810.01822.33O2.94±0.02
x = 0.3 bulk0.7650.02262.95O2.94±0.02
x = 0 nano 0.6240.0081.28O3.01±0.01
x = 0.02 nano0.6120.0091.47O3.02±0.01
x = 0.05 nano0.6220.0111.33O3.01±0.01
x = 0.1 nano0.6150.0091.77O3.02±0.01
x = 0.2 nano0.6330.0081.26O3.01±0.01
x = 0.3 nano0.6380.0121.88O3.01±0.01
Table 4. Experimental values for Pr0.65Sr0.35−xCaxMnO3 bulk materials: electrical properties.
Table 4. Experimental values for Pr0.65Sr0.35−xCaxMnO3 bulk materials: electrical properties.
Compound (Bulk)TC (K)Tp1 (K)
(Tp2 (K))
ρpeak (Ωcm)
in 0 T
MRMax (%)
(1 T)
MRMax (%)
(2 T)
Pr0.65Sr0.33Ca0.02MnO32732740.2134.4511.99
Pr0.65Sr0.3Ca0.05MnO32612733.09412.2822.75
Pr0.65Sr0.25Ca0.1MnO3244256 (258)1.60223.2833.27
Pr0.65Sr0.15Ca0.2MnO3201210 (217)0.34129.0851.96
Pr0.65Sr0.05Ca0.3MnO3 ->100 × 108 (72.985 in 5 T)77.62 (between 3 and 4 T)99.99 (between 2 and 3 T)
Table 5. Critical exponent values for all samples from modified Arott plot method.
Table 5. Critical exponent values for all samples from modified Arott plot method.
CompoundγβδTC (K)
x = 0.02bulk1.0570.2125.986273
x = 0.05bulk0.9810.2325.228261
x = 0.1bulk0.9850.254.94244
x = 0.2bulk1.0360.2175.774201
x = 0.3bulk----
x = 0nano1.0230.5522.853257
x= 0.02nano1.010.5412.867252
x = 0.05nano0.9860.5082.941249
x = 0.1nano0.9770.5122.908239
x = 0.2nano0.9670.5312.821191
x = 0.3nano----
Mean field model10.53
3D Heisenberg model1.3660.3554.8
Ising model1.240.3254.82
Tricritical mean field model10.255
Table 6. Experimental values for Pr0.65Sr0.35−xCaxMnO3 bulk materials: magnetic measurements.
Table 6. Experimental values for Pr0.65Sr0.35−xCaxMnO3 bulk materials: magnetic measurements.
Compound (Bulk)TC (K)MsB/f.u.)Hci (Oe)SM|
(J/kgK)
μ0ΔH = 1 T
SM|
(J/kgK)
μ0ΔH = 4 T
RCP (S)
(J/kg)
μ0ΔH = 1 T
RCP (S)
(J/kg)
μ0ΔH = 4 T
Refs.
Pr0.65Sr0.35MnO3295 2.3 [12]
Pr0.65Sr0.33Ca0.02MnO32733.711802.045.5334166This work
Pr0.65Sr0.3Ca0.05MnO32613.781702.245.5644167This work
Pr0.65Sr0.25Ca0.1MnO32443.941903.036.9145186This work
Pr0.65Sr0.15Ca0.2MnO32013.971604.489.2160270This work
Pr0.65Sr0.05Ca0.3MnO3 3.811704.6 (40 K)15.3 (40 K)92 (40 K)380 (40 K)This work
La0.7Ca0.3MnO3256 1.38 41 [10]
La0.7Sr0.3MnO3365 -4.44 (5 T) 128 (5 T)[10]
La0.6Nd0.1Ca0.3MnO3233 1.95 37 [10]
Gd5Si2Ge2276 -18 (5 T)-535 (5 T)[10]
Gd293 2.8 35 [10]
Table 7. Experimental values for Pr0.65Sr0.35−xCaxMnO3 nano materials.
Table 7. Experimental values for Pr0.65Sr0.35−xCaxMnO3 nano materials.
Compound (Nano)Tc
(K)
Ms
B/f.u.)
Hci
(Oe)
SM|
(J/kgK)
μ0ΔH = 1 T
SM|
(J/kgK)
μ0ΔH = 4 T
RCP(S)
(J/kg)
μ0ΔH = 1 T
RCP(S)
(J/kg)
μ0ΔH = 4 T
Refs.
Pr0.65Sr0.35MnO3257 3.68101.624.5554263This work
Pr0.65Sr0.33Ca0.02MnO32523.087200.692.538175This work
Pr0.65Sr0.3Ca0.05MnO32493.165400.993.2548178This work
Pr0.65Sr0.25Ca0.1MnO32393.495101.414.3739215This work
Pr0.65Sr0.15Ca0.2MnO31913.396201.083.941185This work
Pr0.65Sr0.05Ca0.3MnO3 -6001.23(75 K)5.12(75 K)48(75 K)204(75 K)This work
La0.67Ca0.33MnO3260 0.97 (5 T) 27 (5 T)[63]
Pr0.65(Ca0.6Sr0.4)0.35MnO3220 0.75 21.8 [64]
La0.6Sr0.4MnO3365 1.5 66 [17]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Atanasov, R.; Ailenei, D.; Bortnic, R.; Hirian, R.; Souca, G.; Szatmari, A.; Barbu-Tudoran, L.; Deac, I.G. Magnetic Properties and Magnetocaloric Effect of Polycrystalline and Nano-Manganites Pr0.65Sr(0.35−x)CaxMnO3 (x ≤ 0.3). Nanomaterials 2023, 13, 1373. https://doi.org/10.3390/nano13081373

AMA Style

Atanasov R, Ailenei D, Bortnic R, Hirian R, Souca G, Szatmari A, Barbu-Tudoran L, Deac IG. Magnetic Properties and Magnetocaloric Effect of Polycrystalline and Nano-Manganites Pr0.65Sr(0.35−x)CaxMnO3 (x ≤ 0.3). Nanomaterials. 2023; 13(8):1373. https://doi.org/10.3390/nano13081373

Chicago/Turabian Style

Atanasov, Roman, Dorin Ailenei, Rares Bortnic, Razvan Hirian, Gabriela Souca, Adam Szatmari, Lucian Barbu-Tudoran, and Iosif Grigore Deac. 2023. "Magnetic Properties and Magnetocaloric Effect of Polycrystalline and Nano-Manganites Pr0.65Sr(0.35−x)CaxMnO3 (x ≤ 0.3)" Nanomaterials 13, no. 8: 1373. https://doi.org/10.3390/nano13081373

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop