Structural, magnetic and magnetocaloric properties of hexagonal multiferroic Yb1−xScxMnO3 (x = 0.1 and 0.2)

Bhumireddi Sattibabu*a, A. K. Bhatnagar*ab, K. Vinodc, Awadhesh Manic and D. Dasa
aSchool of Engineering Sciences and Technology, University of Hyderabad, Hyderabad 500046, India. E-mail: bsb.satti@gmail.com; anilb42@gmail.com
bSchool of Physics, University of Hyderabad, Hyderabad 500046, India
cCondensed Matter Physics Division, Materials Science Group, IGCAR, Kalpakkam, 603102, India

Received 13th August 2015 , Accepted 6th October 2015

First published on 6th October 2015


Abstract

We have studied the effect of Sc doping on the structural, magnetic and magnetocaloric properties of multiferroic Yb1−xScxMnO3 (x = 0.1 and 0.2). X-ray powder diffraction shows that both samples crystallize in the hexagonal phase with P63cm space group. The structural analysis shows a decrease in the lattice parameter a, a decrease in the cell volume of the hexagonal unit cell and a decrease in the average bond length between Mn–O, with Sc substitution. Magnetic measurements show that the Néel temperature (TN) increases from 90 K for x = 0.1 to 94 K for x = 0.2 samples. Isothermal magnetic curves show that the field variation in magnetization generates a metamagnetic transition. The maximum entropy change −ΔSmaxM and the relative cooling power (RCP) of Yb1−xScxMnO3 are found to be 2.46 ± 0.40 J mol−1 K−1 and 38.5 ± 9 J mol−1 for x = 0.1 and 1.87 ± 0.31 J mol−1 K−1 and, 30.1 ± 8 J mol−1 for x = 0.2 with ΔH = 10 T. The rescaled magnetic entropy change curves for different applied fields collapse onto a single curve for materials with second-order phase transitions.


1. Introduction

Materials with the simultaneous presence of more than one ferroic property (multiferroics) with a strong coupling between them have been the subject of tremendous research activity in recent years. Control of electric polarization by the application of a magnetic field and inducing magnetic ordering by the application of an electric field in these materials are expected to lead to next generation multifunctional devices for applications in information storage processes, spintronics, multiple-state memories, magnetoelectric sensors, etc.1–4 One typical example of multiferroic materials is hexagonal manganites RMnO3 with space group P63cm, for elements at the rare-earth R site with a relatively small ionic radius, e.g., Ho, Er, Tm, Yb, Lu, Y, and Sc which exhibit a strong coupling between electric and magnetic dipoles.5–7 In these hexagonal manganites, Mn ions form a natural two dimensional edge-sharing triangular network, which becomes magnetically frustrated with an antiferromagnetic nearest–neighbor interaction.8,9

Among rare earth hexagonal manganites YbMnO3 has been studied only occasionally. The crystal structure of YbMnO3 comprises layers of corner-sharing MnO5 trigonal bipyramids with two apical (O1, O2) oxygen atoms and a triangular base of nonequivalent O3 and O4 oxygen atoms. Yb atom occupies two crystallographic sites (in Wyckoff notations), 2a and 4b of space group P63cm. The Yb–O displacements give rise to a ferroelectric moment along the c-axis (TC = 990 K), the Mn3+ moments order at the Neel temperature (TN = 85 K) and below TC ∼ 5 K, the magnetic moments of the Yb3+ ions are completely ordered.10,11 Most of the studies in such multiferroic manganites focus on their magnetic and ferroelectric behavior, but less work has been undertaken on exploring their magnetocaloric effect (MCE) properties.

MCE describes the reversible change in temperature of a material under adiabatic condition produced by the magnetic entropy change ΔSM due to the variation in applied magnetic field.12,13 The main aim in this field is to search for new materials that exhibit a large MCE and are capable of operating at different temperature ranges, depending on the intended applications. Large MCE close to room temperature would be useful for domestic and several technological applications while large MCE in the low-temperature region is important for specific technological applications such as space science and liquefaction of hydrogen in fuel industry.12,14 The guidelines for the choice of an appropriate material are that it should have low heat capacity and exhibit a large entropy change at the ferromagnetic (FM) to paramagnetic (PM) transition or field-induced metamagnetic transition from antiferromagnetic (AFM) to FM states with a minimal hysteresis.

In the present report, effect of Sc substitution on structural, magnetic and magnetocaloric properties in the Yb1−xScxMnO3 (x = 0.1 and 0.2) system is studied. In order to successfully use multiferroic materials in practical applications, the most important criterion is that the coupling between the ferroelectric and magnetic ordering should occur close to room temperature. Thus, it is necessary to enhance the low antiferromagnetic ordering TN of YbMnO3 through chemical doping. In our previous paper10 on divalent Mg ion doped YbMnO3, it is shown that TN increases marginally with doping. The increase in TN can be explained on the basis of smaller cell volume (Mg has small ionic radii), which may lead to strong exchange interactions and therefore higher ordering temperatures. Hence, it is of significance to search for other effective dopants. Hexagonal ScMnO3 exhibits an antiferromagnetic (AFM) transition at 139 K (ref. 15) and Sc3+ has small ionic radius compare to that of Yb3+. The large field induced magnetization observed in Yb1−xScxMnO3 has motivated us to investigate the magnetocaloric behavior in this system.

2. Experimental details

Polycrystalline samples of Yb1−xScxMnO3 (x = 0.1 and 0.2) were synthesized by the conventional solid state reaction method. High-purity (purity better than 99.9%) Yb2O3, Sc2O3 and MnCO3, powders are obtained from M/s Sigma-Aldrich. The precursors Yb2O3 and Sc2O3 powders are preheated at 500 K for 5 h to remove any absorbed moisture. Stoichiometric proportions of Yb2O3, Sc2O3 and MnCO3 powders are thoroughly mixed, and then calcined in platinum crucibles at 1150 °C in air for 24 h with an intermediate grinding for homogenization. The calcined mixture is cold pressed into pellets at approximately 5 × 107 Pa pressure and then sintered at 1350 °C in air for 20 hours. Finally, all samples are slowly cooled to room temperature for sufficient oxygenation. The phase purity of each sample is checked by powder X-ray diffraction (XRD) using a Bruker D8 Advance X-ray powder diffractometer operating with the Cu-Kα radiation. The magnetization measurements, performed in a Cryogenic Inc. (UK) make vibrating sample magnetometer operating at 20.4 Hz. The temperature dependence of magnetic moment was measured for zero-field cooled (ZFC) and field cooled (FC) conditions at a magnetic field of 1000 Oe. The magnetization isotherms in fields up to 10 T were measured, at different temperatures in the vicinity of low temperature ordering transition.

3. Results and discussion

3.1. Structural characterization

The X-ray diffraction patterns and the Rietveld refinement of polycrystalline samples Yb1−xScxMnO3 (x = 0.1 and 0.2) are shown in Fig. 1. Both samples are in single phase and the measured patterns can be indexed to the hexagonal phase with P63cm space group (JCPDS no. 38-1246) in agreement with a previous report.10 Refined values of lattice parameters and discrepancy factors for Yb1−xScxMnO3 (x = 0.1 and 0.2) are shown in Table 1 along with corresponding values for YbMnO3. It is clearly seen that due to Sc substitution, the relative cell parameter c/a increases. Overall the cell volume decreases. The decrease in lattice constant a and cell volume is due to the smaller ionic radius of Sc3+ [Shannon radius = 0.87 Å for coordination number (CN) = 8] than Yb3+ (Shannon radius = 0.985 Å for CN = 8). As the average A-site (i.e., Yb site) radius changes with Sc content, it is expected that the tolerance factor will also change.
image file: c5ra16357a-f1.tif
Fig. 1 Rietveld refinement of room temperature XRD pattern of Yb1−xScxMnO3 (x = 0.1 and 0.2) samples indexed in space group P63cm.
Table 1 Refined crystallographic parameters and reliability factors of Rietveld refinement for Yb1−xScxMnO3 (x = 0.1 & 0.2) samples at room temperaturea
Yb1−xScxMnO3 x = 0.1 x = 0.2 Reported values for YbMnO3 (ref. 10 and 16)
a t’ is the tolerance factor, given as: t = ((rYb+Sc + ro)/√2(rMn + ro)).
a (Å) 6.0577 (4) 6.046 (4) 6.0671 (4)
c (Å) 11.362 (4) 11.3575 (4) 11.3519 (4)
V3) 361.07 (2) 359.542 (2) 361.886 (2)
c/a 1.8756 1.8785 1.8710
χ2 4.12 5.97
RP (%) 6.55 6.67 5.06
Rwp (%) 8.28 8.86 6.82
RB (%) 4.05 5.34 4.50
t (Å) 0.846 0.842 0.850


Some selected bond distances and bond angles of Yb1−xScxMnO3 (x = 0.1 & 0.2) as well as those of YbMnO3 are given in Table 2. The value of a decreases in Sc doped sample when compared to that in YbMnO3 sample. This change is ascribed to the decrease in the average ab-plane i.e., Mn–O3, Mn–O4 is on Sc doping. The Mn–O1 and Mn–O2 bond lengths are also decrease along the c axis on Sc doping The average Mn–O distances in MnO5 units are significantly shorter in the doped samples compared to pure YbMnO3. In the hexagonal phase the Mn3+ ion is fivefold coordinated, forming a trigonal bipyramid polyhedral environment. Polyhedral distortions Δ are calculated using the following formula,17

 
image file: c5ra16357a-t1.tif(1)
where N is the coordination number (N = 5), dn is the individual distance between Mn and the nth nearest oxygen neighbor, and 〈d〉 is the average distance value.

Table 2 Main bond distances (Å) and angles for MnO5 polyhedra in Yb1−xScxMnO3 (x = 0.1 and 0.2)a
Parameter x = 0.1 x = 0.2 Reported values for YbMnO3 (ref. 7)
a Δ is the distortion of MnO5 polyhedra.
Mn–O1 1.8442 1.8434 1.8532
Mn–O2 1.8019 1.8011 1.8366
〈Mn–O1, O2〉 1.823 1.822 1.845
Mn–O3 1.9104 1.9067 1.9033
Mn–O4 2.0931 2.0891 2.1017
〈Mn–O3, O4〉 2.002 1.9979 2.0025
〈Mn–O〉 1.9125 1.9099 1.9237
∑Radii 2.00 2.00 2.00
Yb1–O1(X3) 2.3781 2.3745
Yb1–O2(X3) 2.4719 2.4676
Yb1–O3 2.3251 2.3242
Yb2–O1(X3) 2.1853 2.1816
Yb2–O2(X3) 2.2765 2.2733
Yb2–O4 2.4439 2.4729
〈Yb–O〉 2.3468 2.349
∑Radii 2.405 2.405
Mn–O3–Mn (°) 118.644 (3) 118.64 (6) 119.5 (7)
Mn–O4–Mn (°) 119.04 (3) 119.038 (5) 118.6 (2)
Δ(10−4) 27 26.5 24.14


The distortion parameter is nearly constant with the Sc content in the hexagonal phase. In general, the Sc doping does not notably affect the structure of the MnO5 polyhedron, in agreement with the weak changes of the trigonal bipyramidal crystal field, which is proved by the fact that the Mn3+ ions (3d4) remain in the high spin state (S = 2) throughout the whole doping range. The Mn–O bond lengths are in agreement with the sum of the ionic radii.15 Moreover, in the YbO7 polyhedron, the Yb–O distances are larger in x = 0.1 sample than in x = 0.2 sample, as expected due to the larger ionic radius of Yb3+ when compared to Sc3+. Also, the average Yb–O distances are close to the sum of the ionic radii.

3.2. Magnetic behaviors

Fig. 2 shows the temperature dependence of the susceptibility χ = M/H of Yb1−xScxMnO3 (x = 0.1 and 0.2) measured at a magnetic field H = 1000 Oe. The susceptibility strongly increases with the decrease in the temperature at about 5 K. In the upper inset, we have shown zero field cooled (ZFC) and field cooled (FC) curves in the low-temperature region below 10 K. Near about 5.5 K both the curves show an anomaly with a sudden increase in magnetization with decreasing T. This abrupt change corresponds to the ferromagnetic (FM) ordering of Yb3+ moments at the 2a crystallographic sites through Yb–Yb interactions.18 A second anomaly (a kink) is observed at 90 K for x = 0.1 and 94 K for x = 0.2, as shown in the lower inset of Fig. 2, which correspond to the AFM ordering of the Mn3+ moments and represent Neel temperatures (TN) for these samples. Further, from the Rietveld refinement study it is observed that the average bond distance between the manganese and oxygen atoms decreases with increases in Sc content. This increases the covalence of Mn–O bonds and results stronger exchange interaction. It may be noted that lattice parameter a decreases with the Sc content and the magnetic interaction occurs in the ab plane of the hexagonal manganites. Therefore, the increase in covalence of Mn–O accounts for both the increase in magnetic ordering temperature and the decrease in the lattice parameter a. The increase in TN from 85 K for YbMnO3 (ref. 16) to 94 K for Yb0.8Sc0.2MnO3 sample also can be explained on the basis of smaller cell volume, which might lead to strong exchange interactions and therefore higher ordering temperatures.
image file: c5ra16357a-f2.tif
Fig. 2 Temperature dependence of 1/χ with Curie–Weiss fit and χ (H = 1000 Oe) of Yb1−xScxMnO3 (x = 0.1 and 0.2): Upper inset: ZFC and FC curves of T near the FM transition of Yb3+. Lower inset: ZFC and FC curves of T near the AFM transition of Mn3+.

Temperature variation of inverse magnetic susceptibility is plotted in Fig. 2 which shows that the Curie–Weiss (CW) law is well obeyed in temperature interval between 200 and 300 K in paramagnetic phase. It fits well to the Curie–Weiss law χ = C/(TθCW). The values of paramagnetic Curie temperature (θCW) and effective magnetic moment (μeff) are calculated from this Curie–Weiss fit which are listed in Table 3. The negative Weiss temperature indicates presence of the antiferromagnetic interaction in these compounds. The experimental effective magnetic moments of Yb1−xScxMnO3 (x = 0.1 and 0.2) samples are obtained from the relation μeff = (7.99C)0.5, where C is the Curie–Weiss constant.10 It is seen that the effective moment (μeff) decreases with increases in nonmagnetic Sc content. The experimental values of μeff are in reasonable agreement with the calculated values using the following formula: μeff = ((1 − x)μYb2 + μMn2)1/2, where x is the Sc concentration while μYb and μMn are the effective magnetic moment of Yb3+ (μYb = 4.53 μB), and Mn3+ (μMn = 4.9 μB), respectively. The ratio f = |θCW/TN|, which is a measure of geometric frustration, is around 2.05. It is smaller for Sc doped samples when compared to that of YbMnO3 which indicates that the magnetic coupling between the Yb and the Mn moments in presence of Sc relieves the frustration effect as the magnitude of the magnetically coupling between Yb and Mn ions is the most important factors for the geometrical frustration.

Table 3 Values of magnetic parameter of Yb1−xScxMnO3 (x = 0.1 and 0.2). Also listed are magnetocaloric effect (MCE) parameters: ΔSmaxM and RCP with reported values
  From Curie–Wiess fit MCE
μexpeffB) μtheoreffB) θCW (K) TN (K) f = |θCW/TN| ΔH(T) ΔSmaxM (J mol−1 K−1) RCP (J mol−1)
x = 0.1 6.1 6.5 −182 90 2.02 5 1.32 13.8
10 2.43 38.5
x = 0.2 5.6 6.3 −193 94 2.05 5 1.01 10.2
10 1.88 30.1
[thin space (1/6-em)]
Reported values for YbMnO3
Ref. 10 5.91 6.68 −219 85 2.57 8 2.3 26 Ref. 20
Ref. 18 6.1 6.68 −220 85 2.58 10 2 26 Ref. 16
Ref. 19 6.40 6.68 −166 82 2.02  


The isothermal magnetization (MH) curves of Yb1−xScxMnO3 (x = 0.1 and 0.2) measured at 2.5 K are shown in Fig. 3. Both the samples show a small magnetic hysteresis loops in low fields indicating a weak ferromagnetic (FM) behavior due to Yb3+.18,20 At 2.5 K, the ferromagnetic (FM) behavior is observed with corresponding coercivity (Hc) around 500 Oe. A change in Sc concentration slightly changes coercivity but retentivity decreases with increase in the Sc content as shown in the inset Fig. 3. The isothermal magnetization (MH) at 2.5 K also shows a noticeable sudden increase in magnetization around 30 kOe, similar results are observed in RMnO3 in general and is attributed due to a magnetically induced phase transition where the Mn3+ ions ordering changes from AFM to FM along the c-axis while along the ab-plane the ordering remains unchanged to AFM.20–23 Further, M does not saturate up to 100 kOe. As for Yb0.9Sc0.1MnO3 and Yb0.8Sc0.2MnO3, the magnetization at 100 kOe is 1.6 μB and 1.3 μB, respectively. The magnetization of these samples decreases with increasing Sc concentration, which is nonmagnetic.


image file: c5ra16357a-f3.tif
Fig. 3 The hysteresis loops of Yb1−xScxMnO3 (x = 0.1 and 0.2). Insert shows up to 1 tesla.

To explain the role of applied magnetic field on metamagnetic transition, the isothermal magnetization curves as a function of magnetic field for Yb1−xScxMnO3 (x = 0.1 and 0.2) were measured in applied fields of up to 10 T in the temperature range of 2–40 K. Fig. 4 shows the typical isothermal magnetization curves of Yb1−xScxMnO3 (x = 0.1 and 0.2), which upholds a field-induced metamagnetic transition.24 The isotherms vary almost linearly in the low-field region and depending upon the temperature the slope changes at a critical field Hc without any indication of saturation. From these plots, below TC, M increases slowly with H in the low field region the slope changes at a critical field Hc and then increases slowly with further increase of H.


image file: c5ra16357a-f4.tif
Fig. 4 Field dependence of isothermal magnetization for Yb1−xScxMnO3 (x = 0.1 and 0.2) at some selective temperatures.

To understand the nature of the metamagnetic phase transitions, we have transformed the M(H) data into Arrott plots as shown in Fig. 5.25 Banerjee26 has given an experimental criterion, which allows the determination of the nature of the magnetic transition (first or second order). It consists in observing the slope of the isotherms plots M2 versus H/M. Applying a regular approach, the straight line was constructed simply by extrapolating the high magnetization parts of the curves for each studied temperature. The negative slope of the Arrott plot indicates a first order nature of the transition, while the positive slope implies a second order transition. It is seen from Fig. 5 that the Arrott plots have positive slopes above and below TC for Yb1−xScxMnO3 (x = 0.1 and 0.2) samples in the complete M2 ranges, indicating that the system exhibits a second order AFM to FM phase transition.


image file: c5ra16357a-f5.tif
Fig. 5 The Arrott plots of M2 vs. H/M at various temperatures for Yb1−xScxMnO3 (x = 0.1 and 0.2) samples.

3.3. Magnetocaloric behaviors

In the isothermal process of magnetization, the MCE of the materials can be derived from Maxwell's thermodynamic relationship:27
 
image file: c5ra16357a-t2.tif(2)

The magnetic-entropy change ΔSM, which results from the spin ordering (i.e. ferromagnetic ordering) and is induced by the variation of the applied magnetic field from 0 to H is given by

 
image file: c5ra16357a-t3.tif(3)
for magnetization measured at discrete field and temperature intervals, the magnetic entropy change defined in eqn (3) can be approximated by the eqn (4):28
 
image file: c5ra16357a-t4.tif(4)
with the uncertainty in ΔSM(Tav)ΔH as
image file: c5ra16357a-t5.tif
where, δT = TuTl is the temperature difference between the two isotherms measured at Tu and Tl with the magnetic field varying from H1 to Hn in constant steps of δH. δMk = M(Tu)kM(Tl)k is the difference in the magnetization at Tu and Tl for each magnetic field Hk. For the calculation of relative error in the entropy change, following,28 the accuracy of magnetization measurements are taken as 0.5% and the accuracy of the magnetic field as 0.1%. The manufacturer quoted temperature stability of 0.25% is used as the error for temperatures. In general, the relative error in the calculated entropy change is 10–20% except at very high fields and low temperatures (H ∼ 10 T, T < 4 K).

The magnetic entropy changes, −ΔSM of Yb1−xScxMnO3 (x = 0.1 and 0.2) samples, associated with the magnetic field variation (1 to 10 T) were calculated using eqn (4) and the data for selected fields are displayed in Fig. 6. These curves present a characteristic shape with a broad maximum in the vicinity of the FM transition of Yb moment. The magnitude of the peak increases with increasing the value of ΔH for each composition and the position of the maximum shifts from 5 to 7.5 K when the magnetic field change increases from 1 to 10 T. The maximum entropy change, −ΔSmaxM, corresponding to a magnetic field variation of 10 T is found to be 2.46 ± 0.40 J mol−1 K−1 and 1.87 ± 0.31 J mol−1 K−1 for x = 0.1 and 0.2, respectively. The magnitude of |ΔSmaxM| increases linearly with increasing magnetic field. The field induced metamagnetic transition contributes to the enhancement of ΔSM.29 These values are in agreement with the reported values in the literature which are ∼2.3 J mol−1 K−1 for H = 8 T for single crystal YbMnO3.24 For comparison, we list in data of various magnetic materials in Table 3 which could be used as magnetic refrigerants.


image file: c5ra16357a-f6.tif
Fig. 6 Temperature variation of magnetic entropy change for selected field change for Yb1−xScxMnO3 (x = 0.1 and 0.2) samples.

When comparing different magneto-caloric materials, it is useful to calculate their relative cooling power (RCP) based on the magnetic entropy change. The relative cooling power is evaluated by considering the magnitude of ΔSM and its full width at half-maximum δTFWHM was expressed as follows:30

 
RCP = |ΔSmaxM| × |δTFWHM|, (5)
it is a measure of the quantity of heat transferred by the magnetic refrigerant between hot and cold sinks. The results of these calculations are shown in Fig. 7(b). We estimate a relative error of 15–30% for the RCP values (as field increases from low: 1 T to high: 10 T) considering the relative error of 10–20% in the value of |ΔSmaxM| and considering a 5–10% error in assigning the δTFWHM values. The RCP values show increase with increasing field for both compounds. RCP values are 38.5 ± 9 J mol−1, and 30.1 ± 8 J mol−1 with ΔH = 10 T for samples with x = 0.1 and 0.2, respectively. These values are higher than those for single crystal YbMnO3 (RCP = 26 J mol−1 with ΔH = 8 T).24 Thus, RCP values of these compounds indicate that these are potential candidates for applications at low temperatures.


image file: c5ra16357a-f7.tif
Fig. 7 (a) Temperature dependence of magnetic entropy change ΔSmaxM versus h2/3 (b) relative cooling power as a function of field for Yb1−xScxMnO3 (x = 0.1 and 0.2) samples.

Fig. 7(a) shows the dependence of the magnetic entropy change on the parameter h2/3 (ref. 31 and 32) where h is the reduced field just around TC and is given by [h = (μBH)/(KBTC)]. The mean-field theory predicts that in the vicinity of second-order phase transitions, ΔSmaxM = −kMS(0)h2/3S(0,0), here k is a constant, MS(0) is the saturation magnetization at low temperatures and S(0, 0) is the reference parameter, which may not be equal to zero.33,34 Fig. 7(a) shows the linear dependence of ΔSmaxM versus h2/3 which implies the second order transition for Yb1−xScxMnO3 (x = 0.1 and 0.2). The fact that ΔSmaxM is estimated around TC and in fields larger than the critical field required for the metamagnetic transition, the conclusion about the second order transition is justified.

3.4. Universal curve

The construction of the phenomenological universal curve is based on the collapse of the ΔSM(T, H) points into one single point in the new curve corresponding to equivalent states of the system. Those equivalent states have the same height, in the (−ΔSMSmaxM) curves. The collapse of the normalized entropy change curves can be then obtained by defining a new variable for the temperature axis, θ, given by the expression
 
image file: c5ra16357a-t6.tif(6a)
 
image file: c5ra16357a-t7.tif(6b)
where Tr1 and Tr2 are the temperature of the two reference points that, for the present study, have been selected as those corresponding to ΔSM(Tr1,2) = 1/2ΔSmaxM.35 Fig. 8 shows the dependence of ΔS* (−ΔSMSmaxM) for Yb1−xScxMnO3 (x = 0.1 and 0.2) for typical field changes. It can be clearly seen that the experimental points of the samples distribute on one universal curve of the magnetic entropy change (ranging from 4 T up to 10 T). The universal curve can be well fitted by a Lorentz function35
 
image file: c5ra16357a-t8.tif(7)
where a, b, and c are the free parameters. A fit to this relation gives a = 1.0246, b = 1.01, and c = −0.05. According to eqn (7), only the position and magnitude of the peak, namely, TC and ΔSmaxM, and two reference temperatures Tr1 and Tr2, are needed to characterize the entropy change, where Tr1 < TC and Tr2 > TC. That is to say, to translate S into the “real” ΔSM(T), one needs only these values that are determined by the properties of the materials. Thus, incomplete ΔSM(T) curves, which are experimentally determined from a small temperature span in the vicinity of TC for the isothermal magnetization measurements, can be easily transformed into the complete curves, which is a helpful tool for the evaluation of material properties such as the refrigerant capacity RC.

image file: c5ra16357a-f8.tif
Fig. 8 The universal curve behavior of the curves as a function of the rescaled temperature for different magnetic field for Yb1−xScxMnO3 (x = 0.1 and 0.2).

4. Conclusion

In conclusion, we report detailed investigations of structural, magnetic and magnetocaloric properties of Yb1−xScxMnO3 (x = 0.1 and 0.2) revealing the effect of Sc doping in YbMnO3 on the structure (bond length, bond angles, unit cell volume, tolerance factor, distortion of MnO5 polyhedra) and magnetism (increase in TN, change in frustration parameter, f). From analysis of experimental data, we deduced RCP of Yb1−xScxMnO3, which are found to be 38.5 ± 9 J mol−1, and 30.1 ± 8 J mol−1 with ΔH = 10 T for x = 0.1 and 0.2 respectively. Thus, multiferroic manganites seem to be potential materials for magnetic refrigeration in the low temperature region. The behavior of ΔSmaxM vs. h2/3 curve confirm that present materials exhibit a second order transition. The phenomenological construction of the universal curve for the studied Yb1−xScxMnO3 (x = 0.1 and 0.2) with a Lorentz function is a helpful tool for the evaluation of material properties such as the refrigerant capacity RC.

Acknowledgements

This work has been supported by UGC-DAE Consortium for Scientific Research, Mumbai Centre, India in the form of a collaborative research scheme (CRS) through project number CRS-M-199. BSB acknowledges UGC-DAE CSR, Mumbai Centre for project fellowship. AKB is thankful to the National Academy of Sciences, India for their support to this work through Senior Scientist Platinum Jubilee Fellowship Scheme. Thanks to Dr K Gireesan, IGCAR, Kalpakkam for providing help in part of this work and also thanks to Dr B. Koteswara Rao, DST-INSPIRE Faculty for fruitful discussions.

References

  1. M. Fiebig, J. Phys. D: Appl. Phys., 2005, 38, R123–R152 CrossRef CAS.
  2. W. Prellier, M. P. Singh and P. Murugavel, J. Phys.: Condens. Matter, 2005, 17, R803–R832 CrossRef CAS.
  3. N. A. Spaldin and M. Fiebig, Science, 2005, 309, 391–392 CrossRef CAS PubMed.
  4. X. Qi, J. Dho, R. Tomov, M. G. Blamire and J. L. MacManus-Driscoll, Appl. Phys. Lett., 2005, 86, 062903 CrossRef PubMed.
  5. Z. J. Huang, Y. Cao, Y. Y. Sun, Y. Y. Xue and C. W. Chu, Phys. Rev. B: Condens. Matter Mater. Phys., 1997, 56, 2623–2626 CrossRef CAS.
  6. T. Katsufuji, S. Mori, M. Masaki, Y. Moritomo, N. Yamamoto and H. Takagi, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 64, 104419 CrossRef.
  7. M. Fiebig, T. Lottermoser, D. Frohlich, A. V. Goltsev and R. V. Pisarev, Nature, 2002, 419, 818–820 CrossRef CAS PubMed.
  8. T. Katsufuji, M. Masaki, A. Machida, M. Moritomo, K. Kato, E. Nishibori, M. Takata, M. Sakata, K. Ohoyama, K. Kitazawa and H. Takagi, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 66, 134434 CrossRef.
  9. P. Schiffer and A. P. Ramirez, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 18, 21 CAS.
  10. B. Sattibabu, A. K. Bhatnagar, S. Rayaprol, D. Mohan, D. Das, M. Sundararaman and V. Siruguri, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 448, 210 CrossRef CAS PubMed.
  11. T. Lonkai, D. G. Tomuta, U. Amann, J. Ihringer, R. W. A. Hendrikx, D. M. Többens and J. A. Mydosh, Phys. Rev. Lett., 2004, 69, 134108 Search PubMed.
  12. K. A. Gschneidner Jr, V. K. Pecharsky and A. O. Tsokol, Rep. Prog. Phys., 2005, 68, 1479 CrossRef and references therein.
  13. A. M. Tishin, in Handbook of Magnetic Materials, ed. K. H. Buschow, Elsevier Science B. V., New York, 1999, vol. 12, p. 395 Search PubMed.
  14. V. Provenzano, J. Li, T. King, E. Canavan, P. Shirron, M. DiPirro and R. D. Shull, J. Magn. Magn. Mater., 2003, 266, 185 CrossRef CAS.
  15. A. Muñoz, J. A. Alonso, M. J. Martínez-Lope, M. T. Casáis, J. L. Martínez and M. T. Fernández-Díaz, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 62, 9498 CrossRef.
  16. B. Sattibabu, A. K. Bhatnagar, S. S Samatham, D. Singh, S. Rayaprol, D. Das, V. Siruguri and V. Ganesan, J. Alloys Compd., 2015, 644, 830–835 CrossRef CAS PubMed.
  17. J. Rodríguez-Carvajal, M. Hennion, F. Moussa and A. H. Moudden, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 57, R3189–R3192 CrossRef.
  18. X. Fabrèges, I. Mirebeau, P. Bonville, S. Petit, G. Lebras-Jasmin, A. Forget, G. Andre and S. Pailhes, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 214422 CrossRef.
  19. S. L. Samal, T. Magdaleno, K. V. Ramanujachary, S. E. Lofland and A. K. Ganguli, J. Solid State Chem., 2010, 183, 643–648 CrossRef CAS PubMed.
  20. H. Sugie, N. Iwata and K. Kohn, J. Phys. Soc. Jpn., 2002, 71, 1558 CrossRef CAS.
  21. M. Fiebig, C. Degenhardt and R. V. Pisarev, Phys. Rev. Lett., 2002, 88, 027203 CrossRef CAS.
  22. M. Fiebig, T. Lottermoser and R. V. Pisarev, J. Appl. Phys., 2003, 93, 8194 CrossRef CAS PubMed.
  23. N. Abramov, S. Lofland and Y. Mukovskii, Phys. Status Solidi B, 2012, 1–3, 48318 Search PubMed.
  24. A. Midya, S. N. Das, P. Mandal, S. Pandya and V. Ganesan, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 235127 CrossRef.
  25. A. Arrott and J. Noakes, Phys. Rev. Lett., 1967, 19, 786 CrossRef CAS.
  26. S. K. Banerjee, Phys. Lett., 1964, 12, 16 CrossRef.
  27. A. H. Morrish, The Physical Principles of Magnetism Wiley, New York, 1965, ch. 3 Search PubMed.
  28. V. V. Pecharsky and K. A. Gschneidner, J. Appl. Phys., 1999, 86, 565 CrossRef CAS PubMed.
  29. A. Midya, P. Mandal, S. Das, S. Banerjee, L. S. Sharath Chandra, V. Ganesan and S. Roy Barman, Appl. Phys. Lett., 2010, 96, 142514 CrossRef PubMed.
  30. K. A. Gschneidner and V. K. Pecharsky, Annu. Rev. Mater. Sci., 2000, 30, 387 CrossRef CAS.
  31. J. L. Wang, S. J. Campbell, R. Zeng, C. K. Poh and S. X. Dou, J. Appl. Phys., 2009, 105, 07A909 Search PubMed.
  32. H. Oesterreicher and F. T. Parker, J. Appl. Phys., 1984, 55, 4334 CrossRef CAS PubMed.
  33. V. Franco, A. Conde, V. K. Pecharsky and K. A. Gschneidner, EPL, 2007, 79, 47009 CrossRef.
  34. N. Zaidi, S. Mnefgui, J. Dhahriaand and E. K. Hlil, RSC Adv., 2015, 5, 31901 RSC.
  35. Q. Y. Dong, H. W. Zhang, J. R. Sun, B. G. Shen and V. Franco, J. Appl. Phys., 2008, 104, 116101 CrossRef PubMed.

This journal is © The Royal Society of Chemistry 2015