Magnetic and magneto-caloric properties of polycrystalline powder samples of HoCrO3 with four different particle sizes are reported here. The samples were prepared by citrate method and were annealed at 700, 900, 1100, and 1300 °C to yield average particle sizes of 60 nm, 190 nm, 320 nm, and 425 nm, respectively, as determined by the analysis of X-ray diffraction patterns and images obtained by scanning electron microscopy. Additional structural characterization was done using Raman spectroscopy. Measurements of the magnetization of the samples were done from 5 K to 300 K in magnetic fields up to 70 kOe. Analysis of the temperature dependence of the paramagnetic susceptibility in terms of the modified Curie-Weiss law, including the Dzyaloshinsky-Moriya (DM) interaction, show small but systematic changes in the Néel temperature T N C r of Cr3+ ions, exchange constant J, and the DM interaction with variation in particle size. However, below T N C r the largest size-dependent effects are observed at 5 K, and the measured magnitudes of coercivity field HC are 1930, 2500, 4660, and 7790 Oe for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively, which can be interpreted by a single domain model. Enhancement of the particle size gives about a fourfold increase in the magnitude of the energy product, HC * MS, where MS is the saturation magnetization. However, as the particle size rises, an opposite trend is observed in the max magnetic entropy (ΔSM = 8.73, 7.22, 7.77, and 6.70 J/kg K) and the refrigerant capacity (RC = 388, 354, 330, and 310 J/kg) for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively. These results suggest ways to optimize the properties of HoCrO3 for applications in magnetic storage and magnetic refrigeration.

As a safe, efficient, and environmentally friendly technology, solid-state refrigeration is an alternative solution to conventional vapor compression technology.1–4 Solid-sate refrigeration is based on three kinds of caloric effects: baro-caloric,1 electro-caloric,2 and magneto-caloric effect (MCE),3,4 which are, respectively, driven by the application of isotropic stress, electric field, and magnetic field.5–7 Nowadays, the interest has been extended to multi-caloric materials, which include more than one of these caloric effects. Multicaloric materials have been regarded as useful for improving the efficiency of refrigeration devices,8–11 and strong candidates among them are multiferroic materials with two or more ferroic phases (ferroelectricity, ferromagnetism, and ferroelasticity). This is because cross-coupling effects between these ferroic phases lead to large entropy change.12–17 Recently, rare-earth chromites (RCrO3, R = Ho, Dy, Er, etc.) were proposed to be a new family of multiferroics showing large MCE,18–20 thereby becoming suitable candidates for application in magnetic refrigeration (MR).21–25 Among RCrO3, HoCrO3 with orthorhombically distorted perovskite structure has been a of great interest because of its multiferroic and MCE properties.22,26–30 It has been reported that HoCrO3 polycrystalline powder has MCE value of 7.2 J/kg K and refrigerant capacity (RC) ∼ 408 J/kg at 20 K and 7 T (70 kOe), and therefore deemed suitable for MR applications in low temperature range from 5 to 30 K.22 Tiwari et al. reported that, and Cr3+ ions in CrO6 octahedra in HoCrO3 shows splitting in spin-allowed d-d transition, induced by local magnetic field.27 Ghosh et al., reported a high electric polarization ∼0.32 μCcm−2 at 10 K, an ordering temperature of Cr3+ ( T N C r ) at 142 K in HoCrO3 polycrystalline sample, and polar ordering at a much higher temperature of 240 K indicating atypical multiferroic order.26 

In spite of the above-mentioned studies, an essential understanding of the size effect on the magnetic properties of HoCrO3 polycrystalline samples is still lacking. This is an important issue to investigate in HoCrO3 since prior studies in other perovskites have shown that particle size considerably affects the magnetic properties of polycrystalline powders.31–34 For example, Wu et al. synthesized uniform ErMn2O5 nanorods of different lengths by a surfactant-templated hydrothermal process and the magnetization of these nanorods increases with their length, reaching a maximum value at the length of 175 nm because of the competition between surface strain and uncompensated spin.31 Park et al. synthesized BiFeO3 nanoparticles of particle size ranging from 15 nm to 100 nm by controlling the annealing temperatures, and demonstrated that the nanoparticles with size smaller than 62 nm exhibited strong size-dependent magnetic properties.32 In the work of Cai et al., a modified polyacrylamide sol-gel method was employed to prepare DyMnO3 nanoparticles of various sizes from 51 nm to 106 nm. With the reduction in particle size, the critical field where DyMnO3 nanoparticles went through a field-induced metamagnetic transition at 4 K increases monotonically, whereas all the lattice constants and the antiferromagnetic interaction of Mn3+ ions decreased, because of reduction in Mn-O-Mn bond angle.33 Likewise, the dependence of magnetic properties on particle size was also reported in DyCrO3 nano-platelets synthesized using citric acid and oxalic acid as chelating agents. The average particle size were ∼90 nm and ∼50 nm, respectively, leading to different Néel temperatures (144 K and 146 K).34 

Particle size also influences the magnetic hysteresis behavior of a material.35–37 This is important since in the refrigeration process, larger magnetic hysteresis leads to more energy loss, and thus abates MCE.38 Since HoCrO3 polycrystalline sample displays large magnetic hysteresis,22 variation of the particle size is expected to impact its MCE properties. As an example, the particle size of Gd5Si2Ge2 alloys powders prepared via a milled processing condition in the micrometer range remarkably affected the MCE properties and powders with particle size 38.5–50 μm exhibited maximum entropy change.39 Also, Pękała et al. revealed that MCE properties of La0.7Ca0.3MnO3 nanocrystalline spreads over a broader temperature interval than its polycrystalline counterpart.40 In addition, MCE properties of Ni51Mn34In14Si1 alloy of particle size ranging from 20 to 1000 μm were studied and samples with sizes 600–710 μm had high values of entropy change and RC.41 Similar result was reported by Toliński and Synoradzki in DyCo3B2 system.42 Likewise, the electro-caloric effect of BaTiO3 nano-ceramics also depends on the particle size.43 

The particle size dependence of the MCE and electro-caloric effects observed in many systems noted above motivated us to undertake similar investigations in polycrystalline HoCrO3 powder samples. The size-dependent structural, magnetic, and magneto-caloric properties for the 60 nm, 190 nm, 320 nm, and 425 nm size particles are reported in this work.

Powder samples of HoCrO3 were synthesized using the citrate method. In the first step, equimolar quantities of Ho(NO3)3 and Cr(NO3)3 precursors were dissolved in water separately and then mixed together, followed by the addition of citric acid as the chelating agent. The mixed solution was then heated and dried on a hot plate followed by annealing at 700, 900, 1100, and 1300 °C, respectively for 2 h in an O2 atmosphere to acquire the four HoCrO3 powder samples with different particle sizes.

The structural properties and phase purity of the samples were determined by X-ray diffraction (XRD) measurements (using Bruker D2 Phaser X-ray diffractometer with Cu-Kα radiation λ = 1.54 Å). The scan mode was 2°/min with 0.02°/step from 20° to 90°. The experimental XRD data were Rietveld-refined via Fullprof Suite software in order to obtain the lattice constants of the samples. The Williamson-Hall analysis was adopted to estimate the particle size by Person 7 peak fitting of the XRD data (Fityk Software). Samples were imaged by field-emission scanning electron microscope (FE-SEM) to ascertain the microstructure and particle sizes, which were further confirmed by plotting the histogram of the measured particle diameters. Room temperature micro-Raman spectra were recorded using a 514 nm Argon laser in the Renishaw System 2000. The temperature dependent dc magnetization data {measured in zero-field cooled (ZFC) and field-cooled (FC) modes} and isothermal dc magnetization versus magnetic-field curves at diverse temperatures were evaluated using a vibrating sample magnetometer system attached to an Evercool Physical Property Measurement System from Quantum Design Inc.

The room temperature XRD patterns of the four prepared samples of HoCrO3 are presented in Fig. 1(a), where the samples are labeled by their particle sizes obtained in the following discussion. The experimental data were fitted by Rietveld refinement that indicates that all the observed diffraction peaks in the four samples match with the orthorhombically distorted perovskite structure with space group Pbnm. Fig. 1(b) shows the representative refinement result of the HoCrO3 sample annealed at 700 °C (labeled as 60 nm). It can be seen that no extra peak was identified, suggesting that the samples were phase pure. The lattice parameters (a, b, and c), Cr-O-Cr bond angles, and Cr-O bond lengths acquired from the Rietveld refinements are summarized in Table I. The values of the orthorhombic distortions from the cubic structure, defined by the orthorhombic strain factor S = 2(b−a)/(b+a),33 are also summarized in Table I. These parameters for the HoCrO3 samples annealed at the four different temperatures were found to be only slightly different from each other.

FIG. 1.

(a) Room temperature X-ray diffraction patterns for the four HoCrO3 samples;. (b) Representative X-ray diffraction data and Rietveld refinement results of the 60 nm HoCrO3 sample were compared.

FIG. 1.

(a) Room temperature X-ray diffraction patterns for the four HoCrO3 samples;. (b) Representative X-ray diffraction data and Rietveld refinement results of the 60 nm HoCrO3 sample were compared.

Close modal
TABLE I.

The annealing temperature (T), lattice parameters, Cr-O-Cr bond angles, and Cr-O bond length acquired by Rietveld refinement of the experimental XRD data, crystallite size (D) and strain (η) obtained by Williamson-Hall analysis of the XRD data of the present HoCrO3 samples, and the particle size (D′) obtained from the histogram of SEM images. The orthorhombic strain factor S is calculated using the lattice parameters.

T ( °C) 700 °C 900 °C 1100 °C 1300 °C
a (Å)  5.250  5.248  5.249  5.247 
b (Å)  5.526  5.525  5.527  5.524 
c (Å)  7.548  7.545  7.546  7.543 
V3 218.97  218.79  218.93  218.59 
S  0.05136  0.05142  0.05152  0.05142 
Cr1-O1-Cr1 (deg)  147.92  146.64  147.49  146.51 
Cr1-O2-Cr1 (deg)  147.57  145.34  146.58  146.54 
Cr1-O1 (Å)  1.963  1.975  1.965  1.969 
Cr1-O2 (Å)  2.013  2.010  2.006  2.000 
Cr1-O2 (Å)  1.956  1.966  1.973  1.977 
D (nm)  88 ± 9  182 ± 61  250 ± 60  487 ± 200 
η (%)  0.202 ± 0.049  0.134 ± 0.075  0.194 ± 0.049  0.121 ± 0.041 
D (nm)  60 ± 18  190 ± 60  320 ± 80  425 ± 100 
T ( °C) 700 °C 900 °C 1100 °C 1300 °C
a (Å)  5.250  5.248  5.249  5.247 
b (Å)  5.526  5.525  5.527  5.524 
c (Å)  7.548  7.545  7.546  7.543 
V3 218.97  218.79  218.93  218.59 
S  0.05136  0.05142  0.05152  0.05142 
Cr1-O1-Cr1 (deg)  147.92  146.64  147.49  146.51 
Cr1-O2-Cr1 (deg)  147.57  145.34  146.58  146.54 
Cr1-O1 (Å)  1.963  1.975  1.965  1.969 
Cr1-O2 (Å)  2.013  2.010  2.006  2.000 
Cr1-O2 (Å)  1.956  1.966  1.973  1.977 
D (nm)  88 ± 9  182 ± 61  250 ± 60  487 ± 200 
η (%)  0.202 ± 0.049  0.134 ± 0.075  0.194 ± 0.049  0.121 ± 0.041 
D (nm)  60 ± 18  190 ± 60  320 ± 80  425 ± 100 

The crystallite sizes (D) and strain (η) of the HoCrO3 samples were calculated using the Williamson-Hall analysis44 

(1)

where K is the shape parameter (0.89 was used in this study in accordance with earlier reports of synthesis of similar materials),44 λ = 0.154 (nm) is the wavelength of the X-ray beam, β is the full width at half maximum of the Bragg diffraction peak, and η is the strain in the sample. For particle sizes above 50 nm, the instrumental width has to be considered by replacing β in Eq. (1) with β = β 2 b 2 , where b is the instrumental width (0.042° for the Bruker D2 Phaser system adopted here).

As listed in Table I, D was calculated to be 88, 182, 250, and 487 nm with the uncertainties using Eq. (1) for the HoCrO3 samples annealed at 700, 900, 1100, and 1300 °C, respectively. Notably, D increases with an increase in the annealing temperature, because larger particles grow at the expense of smaller particles. This phenomenon is also known as Oswald ripening.45 In the report of Ghosh et al., HoCrO3 polycrystalline sample annealed at 1400 °C for 12 h by solid-state reaction method was examined to have even larger D of ∼1 μm,26 which is consistent with our results. Similarly, Park et al.32 and Cai et al.33 reported BiFeO3 and DyMnO3 nanoparticles synthesized with various D (14–245 nm and 51–106 nm, respectively) by tuning the annealing temperatures. Therefore, it can be inferred that the crystallite size of HoCrO3 powders is tunable by simply controlling the annealing temperature (in addition to the sample synthesis method).

In Figs. 2(a)–2(d), the particle size (D) of the HoCrO3 samples was further explored by using the FESEM images, along with the particle diameter histograms accounting for about 100 particles (fitted to log-normal distribution). From these histograms, D are estimated to be 60 nm, 190 nm, 320 nm, and 425 nm, respectively, which are very close to the crystallite sizes (D) determined from the Williamson-Hall analysis and are within the experimental uncertainties (see Table I). This shows that the particle size rises with increase in the annealing temperature. In general, the crystallite size determined by XRD may be smaller than the particle size determined by SEM technique since a particle may contain more than one crystallite. However, as shown above, in the present case, the sizes determined by XRD and SEM are nearly equal within experimental uncertainties of the two methods. Therefore, in the rest of this paper, the average particle sizes obtained from the histograms as listed above have been used to label the HoCrO3 samples. The shape of the particles was also measured in terms of the aspect ratio of the longest dimension to the shortest dimension, which were determined to be 1.45, 1.45, 1.33, and 1.21 for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively. Therefore, the aspect ratio decreases as the annealing temperature goes up.

FIG. 2.

Scanning electron microscopy images of the four HoCrO3 samples with the average particle sizes of 60 nm (a), 190 nm (b), 320 nm (c), and 425 nm (d), as determined here. The particle diameter histogram acquired by measuring sizes of about 100 particles in each case are shown on the right side and fitted to log normal distribution (black curves).

FIG. 2.

Scanning electron microscopy images of the four HoCrO3 samples with the average particle sizes of 60 nm (a), 190 nm (b), 320 nm (c), and 425 nm (d), as determined here. The particle diameter histogram acquired by measuring sizes of about 100 particles in each case are shown on the right side and fitted to log normal distribution (black curves).

Close modal

Room temperature Raman spectra of the samples are presented in Fig. 3. Theoretical studies show that the orthorhombic Pbnm perovskite structure has 24 Raman active modes: 7Ag+5B1g+7B2g+5B3g, 12 of which occur in the range of wave-number 100 cm−1 to 600 cm−1. These modes were assigned by following the recent room temperature Raman study on RCrO3 system.21,46 All of the vibration modes observed here are indicative of phase purity of the samples. There is slight shift in the Raman modes between these HoCrO3 samples, attributed to the difference in the strain and lattice parameters (see Table I). It is noted that the width of the Raman modes becomes narrower with the rise in the particle size; this is interpreted in terms of the variation of lattice distortion, which is consistent with the trend observed in DyMnO3 nanoparticles.33 Because neither XRD nor Raman measurements show any impurity peaks (or modes) for all samples, it is inferred that even after annealing these samples at higher temperatures (1300 °C), no impurity phase was detected in the present samples.

FIG. 3.

Room-temperature Raman spectra of the four HoCrO3 samples with the major peaks marked with identified modes in the text.

FIG. 3.

Room-temperature Raman spectra of the four HoCrO3 samples with the major peaks marked with identified modes in the text.

Close modal

The temperature dependent dc magnetization (M vs. T) data for the four HoCrO3 samples under the field-cooled (FC) and zero-field-cooled (ZFC) protocols with applied field of H = 50 Oe are presented in Fig. 4. The FC data bifurcates from the ZFC data below ∼ 140 K which represents magnetic ordering as discussed below. Note that below 140 K, there is no evidence of a peak in the M vs. T data for the ZFC case. This implies that there is no blocking phenomenon characteristics of nanoparticle systems and that the sizes of these particles are larger than the superparamagnetic limit. This information is used later in the discussion of the size dependence of coercivity in these samples.

FIG. 4.

The temperature dependent dc magnetization (M) data for the four HoCrO3 samples in the (black square) zero-field-cooled (ZFC) and (red cycle) field-cooled (FC) modes with applied field H = 50 Oe.

FIG. 4.

The temperature dependent dc magnetization (M) data for the four HoCrO3 samples in the (black square) zero-field-cooled (ZFC) and (red cycle) field-cooled (FC) modes with applied field H = 50 Oe.

Close modal

For the determination of the Néel temperature ( T N C r , where Cr3+ ions order), d(χT)/dT data (χ = M/H is susceptibility using dc field cooled magnetization data) are plotted as a function of temperature as shown in Fig. 5, revealing T N C r of 139.1, 139.6, 140.1, and 139.7 K for the 60 nm, 190 nm, 320 nm, and 425 nm size particle samples, respectively. It is noted that the FC magnetization data were recorded with an interval of 0.5 K near T N C r for better resolution. Theoretically and experimentally, it is well established that in antiferromagnets, the position of TN from the temperature dependent magnetic susceptibility data is best determined by the peak position in d(χT)/dT vs T plot.47,48 In the present samples with diverse particle size, there is only minor variations (<1 K) in the observed T N C r values. However, in the report of Gupta et al., DyCrO3 nano-platelets prepared by citric acid and oxalic acid methods with an average particle size of 90 nm and 50 nm, yielded slightly different T N C r values (144 K and 146 K, respectively).34 In addition to T N C r , a second transition at ∼8 K, associated with the ordering of Ho3+ ions ( T N H o ), was observed in the present samples as shown in the inset of Fig. 5.

FIG. 5.

Derivative of the product of temperature and susceptibility with respect to temperature (d(χT)/dT) under an applied field of 50 Oe of the four HoCrO3 samples. The inset shows derivative of the magnetization with respect to temperature (dχ/dT) from 5 K to 50 K, in order to reveal the Ho3+ ordering temperature.

FIG. 5.

Derivative of the product of temperature and susceptibility with respect to temperature (d(χT)/dT) under an applied field of 50 Oe of the four HoCrO3 samples. The inset shows derivative of the magnetization with respect to temperature (dχ/dT) from 5 K to 50 K, in order to reveal the Ho3+ ordering temperature.

Close modal

Below 140 K ( T N C r ), the magnetization of the present HoCrO3 samples rises gradually with reduction in temperature. This is not typical of fully compensated collinear antiferromagnets, so that some canting of the moments is likely present in HoCrO3. This is also verified by the temperature dependence of magnetization just above T N C r in Sec. III B. In addition, there is clearly a size effect on the magnetization of HoCrO3 samples measured with H = 50 Oe below T N C r for both the ZFC and FC cases, as revealed in Fig. 4. The FC magnetization increases with an increase in particle size, whereas an opposite trend is evident for the ZFC magnetization. It is noted that such an increase of FC magnetization value with increasing particle size was also reported in multiferroic TbMnO3 nanoparticles;49 however, no such trend was observed for multiferroic DyMnO3 nanoparticles in the size range of 51 to 106 nm.33 

In the paramagnetic region above 140 K, magnetization of the four HoCrO3 samples with different particle sizes slowly decreases with increase in temperature, due to paramagnetism of Cr3+ and Ho3+ ions. Hence, the data of χ vs. T for T T N C r were first fitted to the Curie-Weiss law

(2)

Here, C is the Curie constant and θ is the Weiss temperature. The fitting results of the present samples are shown in Figures 6(a)–6(d) and the fitted parameters (C and θ) are summarized in Table II. With the fitted C value, the effective magnetic moment μ e f f was then calculated using

(3)

where kB is Boltzmann constant and N is Avogadro constant.50 The effective magnetic moment μeffa was also calculated using the free ionic moments of Cr3+ (3.8 μB) and Ho3+ ions (10.4 μB)22 according to μ e f f a = μ H o 2 + μ C r 2 . These parameters are also summarized in Table II.

FIG. 6.

Temperature dependence of the inverse magnetic susceptibility (1/χ) measured with H = 50 Oe in the field-cooled mode (represented by the black dots) were fitted to the Curie–Weiss law of Eq. (2) (the red line stands for the fitting), with the magnitudes of parameters listed in Table II.

FIG. 6.

Temperature dependence of the inverse magnetic susceptibility (1/χ) measured with H = 50 Oe in the field-cooled mode (represented by the black dots) were fitted to the Curie–Weiss law of Eq. (2) (the red line stands for the fitting), with the magnitudes of parameters listed in Table II.

Close modal
TABLE II.

The particle sizes (D′), and the magnetic parameters: Cr3+ ordering temperature T N Cr (K), the Weiss temperature θ (K), Curie constant C (emu K/(Oe mol)), and effective magnetic moment μeffB) gained by Curie-Weiss fit of the dc susceptibility data, and T N Cr (K), θ (K), C′ (emu K/(Oe mol)), the fitting parameter T0 (K), the symmetric exchange constant Je (K), and the antisymmetric exchange constant De resulting from the modified Curie-Weiss fitting of the dc susceptibility data. μeffa is calculated by free ionic moments of Cr3+ and Ho3+ ions.

D′ 60 nm 190 nm 320 nm 425 nm
T N C r (K)  139.1  139.6  140.1  139.7 
T N C r (K)  140.99 ± 0.05  141.87 ± 0.05  140.48 ± 0.04  141.35 ± 0.06 
T0  140.79 ± 0.06  141.4 ± 0.1  139.6 ± 0.1  140.8 ± 0.1 
Je  9.386 ± 0.004  9.429 ± 0.007  9.306 ± 0.007  9.390 ± 0.007 
De  1.0 ± 0.2  1.5 ± 0.2  2.1 ± 0.1  1.6 ± 0.2 
θ (K)  −20.12 ± 0.68  −23.55 ± 0.66  −22.92 ± 0.45  −29.16 ± 0.49 
θ (K)  −54 ± 1.4  −41 ± 7  −49 ± 8  −48 ± 7 
C (emu K/(Oe mol))  15.43 ± 0.06  16.01 ± 0.04  15.77 ± 0.03  16.13 ± 0.04 
C (emu K/(Oe mol))  15.91 ± 0.07  17.0 ± 0.4  17.1 ± 0.5  17.2 ± 0.4 
μeffB 11.10 ± 0.02  11.31 ± 0.02  11.23 ± 0.01  11.36 ± 0.01 
μeffB 11.28 ± 0.03  11.6 ± 0.1  11.7 ± 0.2  11.7 ± 0.1 
μeffaB 11.23 
D′ 60 nm 190 nm 320 nm 425 nm
T N C r (K)  139.1  139.6  140.1  139.7 
T N C r (K)  140.99 ± 0.05  141.87 ± 0.05  140.48 ± 0.04  141.35 ± 0.06 
T0  140.79 ± 0.06  141.4 ± 0.1  139.6 ± 0.1  140.8 ± 0.1 
Je  9.386 ± 0.004  9.429 ± 0.007  9.306 ± 0.007  9.390 ± 0.007 
De  1.0 ± 0.2  1.5 ± 0.2  2.1 ± 0.1  1.6 ± 0.2 
θ (K)  −20.12 ± 0.68  −23.55 ± 0.66  −22.92 ± 0.45  −29.16 ± 0.49 
θ (K)  −54 ± 1.4  −41 ± 7  −49 ± 8  −48 ± 7 
C (emu K/(Oe mol))  15.43 ± 0.06  16.01 ± 0.04  15.77 ± 0.03  16.13 ± 0.04 
C (emu K/(Oe mol))  15.91 ± 0.07  17.0 ± 0.4  17.1 ± 0.5  17.2 ± 0.4 
μeffB 11.10 ± 0.02  11.31 ± 0.02  11.23 ± 0.01  11.36 ± 0.01 
μeffB 11.28 ± 0.03  11.6 ± 0.1  11.7 ± 0.2  11.7 ± 0.1 
μeffaB 11.23 

As evident in Figures 6(a)–6(d), the χ vs. T data just above T N C r does not quite fit to the Curie-Weiss law of Eq. (2) due to the sharp drop in 1/χ with reduction in temperature. This is because HoCrO3 is not a simple antiferromagnet, but a canted antiferromagnet giving rise to weak ferromagnetism, whose paramagnetic susceptibility versus temperature behavior was modeled by Moriya by including the antisymmetric exchange interaction, known as the Dzyaloshinsky-Moriya (DM) interaction.51 Only the susceptibility in the direction parallel to the easy magnetization axis of the crystal follows the Curie-Weiss law, while the susceptibility in the direction perpendicular to the easy axis must take the DM interaction into consideration. For the present polycrystalline powder samples, the parallel and perpendicular susceptibility could not be evaluated independently, and the contribution from the perpendicular part plays the dominant role. Thus, the susceptibility were fitted to the equation51 

(4)

where T0 and T N C r are fitted parameters, expressed by Moriya51 as

(5)
(6)

where Z = 6 is the coordination of Cr3+ with respect to other Cr3+, S = 3/2 is the spin quantum number of Cr3+, Je and De are the strength of the symmetric and antisymmetric Cr3+-Cr3+ exchange interactions, respectively. Equations (5) and (6) provide a semi-quantitative analysis of Je and De.

The fitting of the data according to Eqs. (4)–(6) in Fig. 7 shows major improvements over the fitting to the Curie-Weiss law of Eq. (2). The fitted parameters, including C′, T0, θ, and T N C r , are also given in Table II. Note that the fitted values of T N C r agree well with those obtained by the peak in d(χT)/dT. The Cr3+-Cr3+ exchange interactions have been extracted from the above parameters, since other weaker exchange interactions (like Ho3+-Ho3+ and Ho3+-Cr3+ interactions) are ignorable at temperature well above the Ho3+ ordering temperature (∼8 K). Note that for the HoCrO3 samples, De is an order of magnitude smaller than Je as expected in most cases.53 However, as evident from Table II, De increases with growth in particle size whereas Je is essentially the same for the four present samples. Since the canting angle and hence magnetization is proportional to the ratio (De/Je), the rise in magnetization (FC mode) with increase in particle size may be interpreted by this mechanism.

FIG. 7.

Same as Fig. 6 except that the data points are only for temperature above 140 K and the red solid lines are fits to the modified Curie-Weiss law of Eq. (4) with the magnitudes of the fitted parameters listed in Table II.

FIG. 7.

Same as Fig. 6 except that the data points are only for temperature above 140 K and the red solid lines are fits to the modified Curie-Weiss law of Eq. (4) with the magnitudes of the fitted parameters listed in Table II.

Close modal

Fig. 8 shows M vs. H curve of the four HoCrO3 samples measured at 5 K with H up to 40 kOe. As H is increased, M increases, but the low H variation of M is quite different among the four samples. As discussed later, this is related to significantly different coercivity (HC) of the four samples. At H = 40 kOe, magnitudes of M for the 60 nm, 190 nm, 320 nm, and 425 nm particles are 80.1, 79.8, 77.4, and 75.8 emu/g, respectively. However, M does not saturate until applied field of 40 kOe for any of the four samples.

FIG. 8.

Magnetic field dependence of magnetization (M) data at 5 K for the four HoCrO3 samples.

FIG. 8.

Magnetic field dependence of magnetization (M) data at 5 K for the four HoCrO3 samples.

Close modal

The isothermal hysteresis behavior of M vs. H in the four HoCrO3 samples with H up to 40 kOe was recorded from 5 K to 160 K with the representative hysteresis loops at 5 K, 50 K, 100 K, and 145 K shown in Fig. 9. This M vs. H variation is interpreted in terms of two components: one is a ferromagnetic component MFM resulting from canting of spins and responsible for the hysteresis, and the other is an antiferromagnetic component varying linearly with H or mathematically M = MFM + χAF*H. Both MFM and χAF are temperature dependent, and MFM = 0 at temperature well above T N C r , so that χAF changes to the paramagnetic susceptibility. The plots of the hysteresis loop in terms of MFM = MχAF*H at 5 K are shown in Fig. 10. From such plots MFM vs. H at each temperature, coercivity HC, remanence MR, saturation magnetization MS, and linear susceptibility χAF were determined and plotted in Fig. 11. As the temperature is increased, MR, χAF and MS decrease quite rapidly reaching zero at T N C r except for χAF that remains non-zero at T N C r as expected and noted above. In contrast, HC shows insignificant variation versus temperature between 5 K and 120 K, except for the largest size 425 nm particles for which there is a steady decrease in HC with increasing temperature. Above about 120 K, HC for all four sizes reduces quite rapidly with the increase of temperature and approaches zero at T N C r  ∼ 140 K, because the system is in the paramagnetic state for T >  T N C r .

FIG. 9.

Hysteresis loop measurements of magnetization M vs applied magnetic field H for the four HoCrO3 samples at four temperatures listed in the panel (a).

FIG. 9.

Hysteresis loop measurements of magnetization M vs applied magnetic field H for the four HoCrO3 samples at four temperatures listed in the panel (a).

Close modal
FIG. 10.

Magnetization M vs. H data of Fig. 8 at 5 K (black square) is subtracted using the linear term χAF·H from M to yield the hysteresis loops of MFM vs. H (red circle) for the four HoCrO3 samples.

FIG. 10.

Magnetization M vs. H data of Fig. 8 at 5 K (black square) is subtracted using the linear term χAF·H from M to yield the hysteresis loops of MFM vs. H (red circle) for the four HoCrO3 samples.

Close modal
FIG. 11.

The temperature dependence of (a) coercive field (HC), (b) remnant magnetization (MR), (c) the saturation magnetization (Ms), and (d) the linear susceptibility (χAF) of the four HoCrO3 samples. The inset of the top panel (a) shows the plot of Hc vs. particle size at 5 K with the data of 1000 nm sample taken from Ref. 26.

FIG. 11.

The temperature dependence of (a) coercive field (HC), (b) remnant magnetization (MR), (c) the saturation magnetization (Ms), and (d) the linear susceptibility (χAF) of the four HoCrO3 samples. The inset of the top panel (a) shows the plot of Hc vs. particle size at 5 K with the data of 1000 nm sample taken from Ref. 26.

Close modal

According to the plots in Fig. 11, although all these four parameters (MR, χAF, MS and HC) exhibit some particle size dependence especially at the lower temperatures, the size dependence of HC is strongest that shows several fold increase with growth in the particle size. As plotted in the inset of Fig. 11(a), HC measured at 5 K is 1930 Oe, 2500 Oe, 4660 Oe, and 7790 Oe at 5 K for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively, including a data point on the 1 μm size particles of HoCrO3 polycrystalline powder showing HC = 15 kOe from prior report.26 It is noted that the magnitude of HC rises almost linearly with increment in particle size from 60 nm to about 1000 nm, suggesting a simple way to tune the magnetic hysteresis of HoCrO3 samples. Such an increment in HC at 5 K with growing particle size from 45 nm to 60 nm was also observed in multiferroic TbMnO3 nanoparticles.49 This near-linear relationship between HC and the particle size is important for applications in magnetic storage, since increasing the coercivity of the magnetic materials utilized in write-heads is a key issue.52 Therefore, this study could prospectively strengthen the efforts towards synthesis and designing of new magnetic storage medium.

Another quantity of interest in connection with magnetic materials is the energy product MS·HC, which is often adopted to compare the strength of magnetic materials. The particle size dependence of this energy product for the four HoCrO3 particles at 5 K is plotted in Fig. 12. The largest 425 nm particle-size sample has the largest magnitude of the energy product.

FIG. 12.

The particle size dependence of the energy product of coercive field (HC) and saturated magnetization (MS) for the four HoCrO3 samples.

FIG. 12.

The particle size dependence of the energy product of coercive field (HC) and saturated magnetization (MS) for the four HoCrO3 samples.

Close modal

In an effort to understand the source of HC and its particle size dependence in HoCrO3, coercivity at 5 K was also measured in the FC mode by cooling the samples from room temperature to 5 K in H = 10 kOe, followed by collecting the hysteresis loop data. This comparison of observed hysteresis loop at 5 K between the ZFC and FC cases is shown in Fig. 13. Also, the measured HC and the loop asymmetry - exchange bias (EB) between the FC and ZFC cases for the four particles are compared and plotted in Fig. 14, showing insignificant differences. The magnitude of EB is also quite small, perhaps well within the uncertainties of measuring large HC values, thereby signifying the absence of interface effects between the ferromagnetic and antiferromagnetic components of the magnetic phases in these samples. This EB is much smaller than that reported for Dy1-xNdxCrO3 solid solution.53 

FIG. 13.

Isothermal hysteresis loop data of magnetization M vs. applied magnetic field at 5 K of the four HoCrO3 samples are shown for the case of zero-field cooled (red curves) and for the case of field cooled in H = 10 kOe (black curves).

FIG. 13.

Isothermal hysteresis loop data of magnetization M vs. applied magnetic field at 5 K of the four HoCrO3 samples are shown for the case of zero-field cooled (red curves) and for the case of field cooled in H = 10 kOe (black curves).

Close modal
FIG. 14.

Particle size dependence of the (a) coercive field (HC) and (b) exchange bias (EB) at 5 K for the zero-field cooled (black squares) and field cooled modes (red circles) of the four HoCrO3 samples. Lines connecting the data points are visual guides.

FIG. 14.

Particle size dependence of the (a) coercive field (HC) and (b) exchange bias (EB) at 5 K for the zero-field cooled (black squares) and field cooled modes (red circles) of the four HoCrO3 samples. Lines connecting the data points are visual guides.

Close modal

As noted earlier, the absence of the blocking temperature in the M vs. T data of Fig. 4 shows that the particle sizes of HoCrO3 investigated here are above the threshold of superparamagnetic limit. Nevertheless, it is still worthwhile to note that the observed size dependence of HC in HoCrO3 is quite different from that reported in nanoparticles of antiferromagnetic oxides such as CuO,54 in which case HC and EB of nanoparticles with sizes below about 50 nm show quite different particle size and cooling field dependence. In CuO nanoparticles, HC increases with decrease in particle size which, along with the observed EB, is interpreted in terms of the interface effects between the disordered surface spins that yield the ferromagnetic component and the antiferromagnetically ordered spins in the core of particles. Therefore, the basic mechanism for the large HC in HoCrO3 and its particle size dependence in the range of 60–1000 nm reported here has to be quite different.

The weak ferromagnetism reported here in HoCrO3 and interpreted in terms of the spin canting induced by the DM interaction is very similar to that known in hematite (α-Fe2O3), which is a well-recognized weak ferromagnet at room temperature due to canting produced by the DM interaction. In a recent paper, Ozdemir and Dunlop55 have presented a thorough review of the particle size dependence of the magnetic parameters including HC in hematite covering the particle size range of about 30 nm to 1000 μm. As the particle size rises, HC increases until a maximum around Do = 2 μm, and then decreases thereafter. It is argued that for D < Do the particles have single domain (SD), whereas for D > Do the particles have multiple-domains (MD) with domain walls. In the SD state, the spin rotation with changing H must overcome exchange forces, making HC grow with increasing particle size. On the other hand, in the MD state with domain walls present, relatively lower H is required to overcome the domain wall energy, making HC decrease with increase in particle size. For the present HoCrO3 samples with sizes in the range between 60 nm and 1 μm, HC rises with increasing particle size. Therefore, these particles are in the SD state and the size Do required for the SD to MD transition must be larger than 1 μm. It is noted that in a strong ferromagnet with large MS such as Fe and Co, the SD to MD transition occurs at a much lower Do ∼ 30 nm.56 Since the temperature dependence of HC is considerably more sluggish than that for MR and MS (Fig. 11), magnetocrystalline anisotropy alone cannot explain the large magnitudes of HC in HoCrO3. Because of the weak ferromagnetism and relatively small magnitude of MS, demagnetization field and shape anisotropy are also less important. Other possible sources of HC are magnetoelastic anisotropy and defects.57,58 However, at present, a quantitative interpretation of HC in HoCrO3 samples is still lacking.

In order to further evaluate these samples for their applications in magnetic refrigeration, the MCE properties of the samples were acquired by collecting the isothermal magnetization data up to field of 7 T. Two parameters—magnetic entropy change ΔSM(T,H) and refrigerant capacity (RC)—were used to characterize the MCE properties and are given by4,38

(7)
(8)

where T1 and T2 are the low temperature and high temperature of the refrigeration circle, respectively. Fig. 15 shows the dependence of ΔSM on temperature of the present HoCrO3 samples. For all the four samples, ΔSM increases from 1 T (10 kOe) to 7 T (70 kOe) because larger magnetic field generates larger magnetization, yielding larger ΔSM according to Eq. (7). As temperature grows, ΔSM rises from 5 to 15 K and declines thereafter. The maximum ΔSM at ∼15 K is related to the ordering of Ho3+ ions and determined to be 8.73, 7.22, 7.77, and 6.70 J/kg K for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively. It can be seen that the HoCrO3 sample with smallest particle size in this work shows the largest value of ΔSM. The ΔSM of HoCrO3 samples is smaller than that of DyCrO3 polycrystalline sample (max 5.13 J/kg K compared to 8.4 J/kg K at 4 T),21 but much larger than those of YCrO3 (0.36 J/kg K at 5 T),59 and LaCrO3 (0.11 J/kg K at 8 T).60 

FIG. 15.

Temperature dependence of entropy change (ΔSM) for the four HoCrO3 samples in seven different magnetic fields from 1 T (10 kOe) to 7 T (70 kOe) as listed in the inset of panel (a).

FIG. 15.

Temperature dependence of entropy change (ΔSM) for the four HoCrO3 samples in seven different magnetic fields from 1 T (10 kOe) to 7 T (70 kOe) as listed in the inset of panel (a).

Close modal

The field-dependent RC value of the present samples calculated by Eq. (8) (setting T1 = 5 K, T2 = 100 K) was shown in Fig. 16. With increasing magnetic field, RC value raises, because larger field results in larger ΔSM and thus larger integration of ΔSM. At 7 T, RC values were found to be 388, 354, 330, and 310 J/kg for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively. Alternatively, the refrigeration capacity can be characterized by the product of ΔSM and the full width at half maximum δTFWHM in the ΔSM(T) curve, known as relative cooling power (RCP = −δTFWHM × ΔSM) and calculated to be 307, 265, 258, and 252 J/kg, respectively. Evidently, RC value of HoCrO3 sample increases with the reduction of particle size. This, on the one hand, might be explained by the larger surface-to-volume ratio of the smaller particle size, which enhances the tangible contribution to the overall magnetization by the uncompensated surface spins.32 On the other hand, the HoCrO3 sample with smaller particle size has smaller magnetic hysteresis, generating less energy lost in the thermal process, and thus the RC value is larger. As in the report of Phan and Yu, materials with nearly zero magnetic hysteresis were considered suitable for the application in MR because of better working efficiency.38 In spite of larger RC value, the sample with 190 nm particle size has smaller ΔSM than the sample with 320 nm particle size, because RC is the integration of sample's ΔSM and it also depends on the width of ΔSM vs. T curve. In addition, the RC values of the present samples are larger than those of polycrystalline YCrO3 (7.1 J/kg at 5 T),59 and Y-doped DyCrO3, i.e., Dy0.7Y0.3CrO3 (150 J/kg at 4 T),24 but smaller than those of pure DyCrO3, Dy0.7Ho0.3CrO3, and Dy0.7Er0.3CrO3.24 It should be clarified here that some rare-earth based alloys and oxides are also promising candidates for low temperature MR application, and exhibit even larger MCE in the similar temperature range.61–64 For example, Li et al. reported large ΔSM = 17.1 J/kg K and 9.3 J/kg K in the temperature of 4–20 K and low field of 2 T for ErNiBC and GdNiBC compounds, respectively.61 A TmZn compound was also reported to have much larger ΔSM = 26.9 J/kg K at 8.6 K and field of 7 T, but smaller RC value of 214 J/kg than the present HoCrO3 samples.65 Since RC value is the figure of merit for MCE property, the work presented here reveals important insights on a promising way to optimize the MCE property of an oxide material system via the control of particle size, such as by tuning the annealing temperature in the solution synthesis method.

FIG. 16.

Magnetic field dependence of the refrigerant capacity (RC) in the refrigeration cycle from 5 K to 100 K for the four HoCrO3 samples. The line connecting the data points are visual guides.

FIG. 16.

Magnetic field dependence of the refrigerant capacity (RC) in the refrigeration cycle from 5 K to 100 K for the four HoCrO3 samples. The line connecting the data points are visual guides.

Close modal

In this paper, polycrystalline HoCrO3 samples of various particle sizes were synthesized via a solution method and annealed at temperature from 700 °C to 1300 °C. Williamson-Hall analysis of the experimental XRD data and scanning electron microscopy images demonstrate that the samples are phase pure and have average particle size ∼60 nm, 190 nm, 320 nm, and 425 nm, respectively. Raman scattering data further certify the absence of impurity and there are frequency shift in the phonon modes between the HoCrO3 samples with different sizes. The dc magnetization measurements indicate that the Néel temperature of Cr3+ ions are 139.1 K, 139.6 K, 140.1 K, and 139.7 K, and the max magnetization at 50 Oe are 19.73, 30.76, 30.85, and 33.14 emu/g for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively. The temperature dependent magnetization data was fitted to the Curie-Weiss law, and also the modified Curie-Weiss law including a correction from Dzyaloshinsky-Moriya interaction. The field dependent magnetization of the HoCrO3 samples were collected at 5 K and the HoCrO3 sample with smaller particle size show higher magnetization at the same field. Isothermal magnetization curves reveal that the HoCrO3 sample with larger particle size has larger magnetic hysteresis loop that are characterized by the dependence of the coercive field and remnant magnetization on temperature. The max coercive field is 1930, 2500, 4660, and 7790 Oe for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively, which can be interpreted by a single domain model, but their remnant magnetization is very similar. The max magnetic entropy changes (ΔSM) are 8.73, 7.22, 7.77, and 6.70 J/kg K and the refrigerant capacity values are 388, 354, 330, and 310 J/kg for the 60 nm, 190 nm, 320 nm, and 425 nm size particles, respectively. Therefore, the particle size of the HoCrO3 samples affects the magnetic and magneto-caloric properties considerably and prospectively the MCE property can be optimized by controlling the particle size during processing of the materials.

This work was funded by the National Science Foundation Grant No. DMR-1310149. M.S.S. thanks D. J. Dunlop for bringing Ref. 57 to his attention.

1.
L.
Mañosa
,
D.
González-Alonso
,
A.
Planes
,
E.
Bonnot
,
M.
Barrio
,
J.-L.
Tamarit
,
S.
Aksoy
, and
M.
Acet
,
Nat. Mater.
9
,
478
(
2010
).
2.
A. S.
Mischenko
,
Q.
Zhang
,
J. F.
Scott
,
R. W.
Whatmore
, and
N. D.
Mathur
,
Science
311
,
1270
(
2006
).
3.
V.
Franco
,
J.
Blázquez
,
B.
Ingale
, and
A.
Conde
,
Annu. Rev. Mater. Res.
42
,
305
(
2012
).
4.
5.
A.
Planes
,
T.
Castán
, and
A.
Saxena
,
Philos. Mag.
94
,
1893
(
2014
).
6.
X.
Moya
,
S.
Kar-Narayan
, and
N. D.
Mathur
,
Nat. Mater.
13
,
439
(
2014
).
7.
M. M.
Vopson
,
J. Phys. D: Appl. Phys.
46
,
345304
(
2013
).
8.
A.
Planes
,
L.
Mañosa
, and
M.
Acet
,
J. Phys. Condens. Matter
21
,
233201
(
2009
).
9.
P. O.
Castillo-Villa
,
D. E.
Soto-Parra
,
J. A.
Matutes-Aquino
,
R. A.
Ochoa-Gamboa
,
A.
Planes
,
L.
Mañosa
,
D.
González-Alonso
,
M.
Stipcich
,
R.
Romero
,
D.
Ríos-Jara
, and
H.
Flores-Zúñiga
,
Phys. Rev. B
83
,
174109
(
2011
).
10.
T.
Castán
,
A.
Planes
, and
A.
Saxena
,
Phys. Rev. B
85
,
144429
(
2012
).
11.
S.
Lisenkov
,
B. K.
Mani
,
C. M.
Chang
,
J.
Almand
, and
I.
Ponomareva
,
Phys. Rev. B
87
,
224101
(
2013
).
12.
M. M.
Vopson
,
Solid State Commun.
152
,
2067
(
2012
).
13.
N. A.
Spaldin
,
S.
Cheong
, and
R.
Ramesh
,
Phys. Today
63
(
10
),
38
(
2010
).
14.
T.
Kimura
,
T.
Goto
,
H.
Shintani
,
K.
Ishizaka
,
T.
Arima
, and
Y.
Tokura
,
Nature
426
,
55
(
2003
).
15.
N. A.
Spaldin
and
M.
Fiebig
,
Science
309
,
391
(
2005
).
16.
M.
Staruch
,
G.
Lawes
,
A.
Kumarasiri
,
L. F.
Cotica
, and
M.
Jain
,
Appl. Phys. Lett.
102
,
062908
(
2013
).
17.
A.
McDannald
,
M.
Staruch
,
G.
Sreenivasulu
,
C.
Cantoni
,
G.
Srinivasan
, and
M.
Jain
,
Appl. Phys. Lett.
102
,
122905
(
2013
).
18.
J. R.
Sahu
,
C. R.
Serrao
,
N.
Ray
,
U. V.
Waghmare
, and
C. N. R.
Rao
,
J. Mater. Chem.
17
,
42
(
2007
).
19.
Z. X.
Cheng
,
X. L.
Wang
,
S. X.
Dou
,
H.
Kimura
, and
K.
Ozawa
,
J. Appl. Phys.
107
,
09D905
(
2010
).
20.
S.
Yin
,
T.
Sauyet
,
C.
Guild
,
S. L.
Suib
, and
M.
Jain
,
J. Magn. Magn. Mater.
428
,
313
(
2017
).
21.
A.
McDannald
,
L.
Kuna
, and
M.
Jain
,
J. Appl. Phys.
114
,
113904
(
2013
).
22.
S.
Yin
,
V.
Sharma
,
A.
Mcdannald
,
F. A.
Reboredo
, and
M.
Jain
,
RSC Adv.
6
,
9475
(
2016
).
23.
L. H.
Yin
,
J.
Yang
,
R. R.
Zhang
,
J. M.
Dai
,
W. H.
Song
, and
Y. P.
Sun
,
Appl. Phys. Lett.
104
,
032904
(
2014
).
24.
A.
McDannald
and
M.
Jain
,
J. Appl. Phys.
118
,
043904
(
2015
).
25.
S.
Yin
and
M.
Jain
,
J. Appl. Phys.
120
,
43906
(
2016
).
26.
A.
Ghosh
,
A.
Pal
,
K.
Dey
,
S.
Majumdar
, and
S.
Giri
,
J. Mater. Chem. C
3
,
4162
(
2015
).
27.
B.
Tiwari
,
M. K.
Surendra
, and
M. S. R.
Rao
,
J. Phys. Condens. Matter
25
,
216004
(
2013
).
28.
Y.
Su
,
J.
Zhang
,
Z.
Feng
,
Z.
Li
,
Y.
Shen
, and
S.
Cao
,
J. Rare Earths
29
,
1060
(
2011
).
29.
Y. L.
Su
,
J. C.
Zhang
,
L.
Li
,
Z. J.
Feng
,
B. Z.
Li
,
Y.
Zhou
, and
S. X.
Cao
,
Ferroelectrics
410
,
102
(
2010
).
30.
S.
Lei
,
L.
Liu
,
C. Y.
Wang
,
C. N.
Wang
,
D.
Guo
,
S.
Zeng
,
B.
Cheng
,
Y.
Xiao
, and
L.
Zhou
,
J. Mater. Chem. A
1
,
11982
(
2013
).
31.
S.
Wu
,
Y.
Lv
,
M.
Lu
, and
Z.
Lin
,
J. Mater. Chem. C
3
,
3121
(
2015
).
32.
T. J.
Park
,
G. C.
Papaefthymiou
,
A. J.
Viescas
,
A. R.
Moodenbaugh
, and
S. S.
Wong
,
Nano Lett.
7
,
766
(
2007
).
33.
X.
Cai
,
L.
Shi
,
S.
Zhou
,
J.
Zhao
,
Y.
Guo
, and
C.
Wang
,
J. Appl. Phys.
116
,
103903
(
2014
).
34.
P.
Gupta
,
R.
Bhargava
,
R.
Das
, and
P.
Poddar
,
RSC Adv.
3
,
26427
(
2013
).
35.
G.
Herzer
,
IEEE Trans. Magn.
26
,
1397
(
1990
).
36.
D.
AlMawlawi
,
N.
Coombs
, and
M.
Moskovits
,
J. Appl. Phys.
70
,
4421
(
1991
).
37.
E. F.
Kneller
and
F. E.
Luborsky
,
J. Appl. Phys.
34
,
656
(
1963
).
38.
M. H.
Phan
and
S. C.
Yu
,
J. Magn. Magn. Mater.
308
,
325
(
2007
).
39.
L.
Zhang
and
H.
Zeng
,
Int. J. Microstruct. Mater. Prop.
8
,
325
(
2013
).
40.
M.
Pękała
,
V.
Drozd
,
J. F.
Fagnard
,
P.
Vanderbemden
, and
M.
Ausloos
,
Appl. Phys. A
90
,
237
(
2008
).
41.
R.
Das
,
A.
Perumal
, and
A.
Srinivasan
,
J. Alloys Compd.
572
,
192
(
2013
).
42.
T.
Toliński
and
K.
Synoradzki
,
Acta Phys. Pol., A
126
,
160
(
2014
).
43.
J. H.
Qiu
and
Q.
Jiang
,
J. Appl. Phys.
105
,
034110
(
2009
).
44.
A.
McDannald
,
L.
Kuna
,
M. S.
Seehra
, and
M.
Jain
,
Phys. Rev. B
91
,
224415
(
2015
).
45.
A.
Baldan
,
J. Mater. Sci.
37
,
2171
(
2002
).
46.
M. C.
Weber
,
J.
Kreisel
,
P. A.
Thomas
,
M.
Newton
,
K.
Sardar
, and
R. I.
Walton
,
Phys. Rev. B
85
,
054303
(
2012
).
47.
P.
Dutta
,
M. S.
Seehra
,
S.
Thota
, and
J.
Kumar
,
J. Phys. Condens. Matter
20
,
015218
(
2008
).
48.
E. E.
Bragg
and
M. S.
Seehra
,
Phys. Rev. B
7
,
4197
(
1973
).
49.
R.
Dhinesh Kumar
,
M.
Subramanian
,
M.
Tanemura
, and
R.
Jayavel
,
J. Nanoparticle Res.
16
,
2501
(
2014
).
50.
J. M. D.
Coey
,
Magnetism and Magnetic Materials
(
Cambridge University Press
,
Cambridge
,
2009
).
51.
52.
R. L.
Comstock
,
J. Mater. Sci. Mater. Electron.
13
,
509
(
2002
).
53.
A.
Mcdannald
,
C. R.
Cruz
,
M. S.
Seehra
, and
M.
Jain
,
Phys. Rev. B
93
,
184430
(
2016
).
54.
M. S.
Seehra
and
A.
Punnoose
,
Solid State Commun.
128
,
299
(
2003
).
55.
O.
Ozdemir
and
D. J.
Dunlop
,
J. Geophys. Res. Solid Earth
119
,
2582
, doi: (
2014
).
56.
W.
Williams
and
D.
Dunlop
,
J. Geophys. Res.
100
,
3859
, doi: (
1995
).
57.
C.
Rath
,
K. K.
Sahu
,
S. D.
Kulkarni
,
S.
Anand
,
S. K.
Date
,
R. P.
Das
, and
N. C.
Mishra
,
Appl. Phys. Lett.
75
,
4171
(
1999
).
58.
D. L.
Leslie-Pelecky
and
R. D.
Rieke
,
Chem. Mater.
8
,
1770
(
1996
).
59.
G. N. P.
Oliveira
,
P.
MacHado
,
A. L.
Pires
,
A. M.
Pereira
,
J. P.
Araújo
, and
A. M. L.
Lopes
,
J. Phys. Chem. Solids
91
,
182
(
2016
).
60.
B.
Tiwari
,
A.
Dixit
,
R.
Naik
,
G.
Lawes
, and
M. S. R.
Rao
,
Mater. Res. Express
2
,
026103
(
2015
).
61.
L.
Li
,
M.
Kadonaga
,
D.
Huo
,
Z.
Qian
,
T.
Namiki
, and
K.
Nishimura
,
Appl. Phys. Lett.
101
,
122401
(
2012
).
62.
L.
Li
,
O.
Niehaus
,
M.
Kersting
, and
R.
Pöttgen
,
Appl. Phys. Lett.
104
,
092416
(
2014
).
63.
Y.
Zhang
,
X.
Xu
,
Y.
Yang
,
L.
Hou
,
Z.
Ren
,
X.
Li
, and
G.
Wilde
,
J. Alloys Compd.
667
,
130
(
2016
).
64.
Y.
Yang
,
Y.
Zhang
,
X.
Xu
,
S.
Geng
,
L.
Hou
,
X.
Li
,
Z.
Ren
, and
G.
Wilde
,
J. Alloys Compd.
692
,
665
(
2017
).
65.
L.
Li
,
Y.
Yuan
,
Y.
Zhang
,
T.
Namiki
,
K.
Nishimura
,
R.
Pöttgen
, and
S.
Zhou
,
Appl. Phys. Lett.
107
,
132401
(
2015
).