Next Article in Journal
Copper Oxide/Functionalized Graphene Hybrid Nanostructures for Room Temperature Gas Sensing Applications
Next Article in Special Issue
Photoluminescence of the Eu3+-Activated YxLu1−xNbO4 (x = 0, 0.25, 0.5, 0.75, 1) Solid-Solution Phosphors
Previous Article in Journal
Straightforward One-Pot Synthesis of New 4-Phenyl-1,2,5,6-tetraazafluoranthen-3(2H)-one Derivatives: X-ray Single Crystal Structure and Hirshfeld Analyses
Previous Article in Special Issue
Thermal Stability and Radiation Tolerance of Lanthanide-Doped Cerium Oxide Nanocubes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low-Temperature Magnetic and Magnetocaloric Properties of Manganese-Substituted Gd0.5Er0.5CrO3 Orthochromites

1
Department of Physics, Central University of Rajasthan Bandarsindri, Ajmer 305817, Rajasthan, India
2
Department of Physics, Sri Venkateswara College, University of Delhi, Dhaula Kuan, New Delhi 110021, Delhi, India
3
National Research Council, Washington, DC 20001, USA
4
Materials and Manufacturing Directorate, Air Force Research Laboratory, 3005 Hobson Way, Wright Patterson Air Force Base, OH 45433, USA
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(2), 263; https://doi.org/10.3390/cryst12020263
Submission received: 29 January 2022 / Revised: 10 February 2022 / Accepted: 12 February 2022 / Published: 15 February 2022
(This article belongs to the Special Issue Optical and Spectroscopic Properties of Rare-Earth-Doped Crystals)

Abstract

:
Rare-earth chromites have been envisioned to replace gas-based refrigeration technology because of their promising magnetocaloric properties at low temperatures, especially in the liquid helium temperature range. Here, we report the low-temperature magnetic and magnetocaloric properties of Gd0.5Er0.5Cr1−xMnxO3 (x = 0, 0.1, 0.2, 0.3, 0.4 and 0.5) rare-earth orthochromites. The Néel transition temperature (TN) was suppressed from 144 K for Gd0.5Er0.5CrO3 to 66 K for the Gd0.5Er0.5Cr0.5Mn0.5O3 compound. Furthermore, magnetization reversal was observed in the magnetization versus temperature behavior of the Gd0.5Er0.5Cr0.6Mn0.4O3 and Gd0.5Er0.5Cr0.5Mn0.5O3 compounds at 100 Oe applied magnetic field. The magnetic entropy change (−∆S) value varied from 16.74 J/kg-K to 7.46 J/kg-K, whereas the relative cooling power (RCP) ranged from 375.94 J/kg to 220.22 J/kg with a Mn ion concentration at 5 T field and around 7.5 K temperature. The experimental results were substantiated by a theoretical model. The present values of the magnetocaloric effect are higher than those of many undoped chromites, manganites and molecular magnets in the liquid helium temperature range.

1. Introduction

Perovskite materials have been widely recognized for their diverse and intriguing properties, such as magnetic, electrical and photocatalytic, etc. [1,2,3,4]. As far as the magnetic properties of various perovskite materials are concerned, rare-earth orthochromites RCrO3 (R = rare-earth ion) exhibit interesting magnetic properties viz. exchange bias [5], spin reorientation [6], magnetization reversal [7,8,9] and magnetocaloric effect (MCE) [10,11,12,13]. It is now a well-established fact that there are three types of magnetic interactions, namely Cr3+-Cr3+, Cr3+-R3+ and R3+-R3+, in RCrO3 materials, which govern their distinct magnetic properties in different temperature ranges [5]. The interaction Cr3+-Cr3+ is responsible for the canted antiferromagnetic behavior below the Néel transition temperature, whereas the Cr3+-R3+ interaction gives rise to negative magnetization or magnetization reversal below a certain temperature, known as the compensation temperature, provided that the R3+ ion is magnetic. The negative magnetization occurs due to the antiparallel coupling between the Cr3+ moments and rare-earth ion moments [14]. Since the last decade, researchers have also been exploring these materials for their usefulness as a low-temperature refrigerant employing the magnetocaloric effect, which is measured in terms of magnetic entropy change and relative cooling power [10,11,12,13,15,16,17,18,19,20]. Magnetic refrigeration involving solid materials is a cooling technology that is more efficient, inexpensive, safer and more environmentally friendly than the conventional vapor-compression-based technology. It was reported that Gadolinium and Gadolinium-based compounds are important working materials for magnetic refrigeration. For example, Yin et al. reported 31.6 J/kg-K magnetic entropy change in GdCrO3 single crystal at 3 K under a 4.4 T applied magnetic field [18]. Shi et al. also reported a magnetic entropy change of 27.6 J/kg-K in Er0.33Gd0.67CrO3 at 5 K and 7 T [19]. The effect of Mn substitution at the Cr site in Gd0.5Er0.5CrO3 has not been explored yet. Therefore, it is worthwhile to investigate the magnetic and magnetocaloric properties of Mn-substituted Gd0.5Er0.5CrO3 compounds.
In the present work, we synthesized Gd0.5Er0.5Cr1−xMnxO3 ( x = 0, 0.1, 0.2, 0.3, 0.4, 0.5) compounds and investigated the effect of Mn substitution on the magnetic and magnetocaloric properties. We also compared the experimental results with theoretical models.

2. Experimental Details

Gd0.5Er0.5Cr1−xMnxO3 ( x   = 0, 0.1, 0.2, 0.3, 0.4, 0.5) (abbreviated as GECMO) orthochromites were prepared by the solid-state reaction method. Stoichiometric amounts of high-purity Gd2O3 (99.99%, Sigma Aldrich, St. Louis, MO, USA), Er2O3 (99.99%, Sigma Aldrich, St. Louis, MO, USA), Cr2O3 (99.99%, Sigma Aldrich, St. Louis, MO, USA) and MnO2 (99.99%, Sigma Aldrich, St. Louis, MO, USA) were mixed and grounded. The mixed powders were calcined for 24 h at 800 °C and 1100 °C. Finally, the calcined powders were reground and pressed into uniform cylindrical pellets, which were sintered at 1450 °C for 48 h under ambient conditions. The samples were characterized for phase identification using an X-ray diffractometer (XRD; PANalytical-Empyrean, Almelo, The Netherlands) at room temperature using Cu-Kα radiation with wavelength 1.5404 Å at a step of 0.01° per second between 20° and 80°. Magnetic measurements were performed using a physical properties measurement system (PPMS DynaCool). Magnetization temperature data were collected between 2 K and 300 K in field-cooled (FC) and zero-field-cooled (ZFC) modes. The ramp rate was 4K/minute for both the modes. The magnetization–applied field ( M H ) isotherms were measured under H = ±8 T.

3. Results and Discussion

3.1. Structural Study

Figure 1 displays the Rietveld-refined XRD patterns of the sintered GECMO orthochromites. It is clear from the XRD patterns that there are no impurity peaks in these series of samples and the diffraction peaks are well matched with the pristine GdCrO3 compound (JCPDS Card No. 25-1056). The refinement was carried out by using Fullprof software with orthorhombic structure and Pnma space group. The calculated lattice parameters, unit cell volume ( V ), reliability factors, octahedral bond length (Cr–O), octahedral tilt angles, strain factor and Goldschmidt tolerance factor ( t ) of all compounds are presented in Table 1. It can be seen that there was an increase in the lattice constants a and c , while the b value was reduced with increasing Mn3+ ion content. Their variation is shown in Figure 2a. Further, the lattice volume was also increased because of the larger ionic radius of the Mn3+ ion (0.645 Å) compared to that of the Cr3+ ion (0.615 Å) (Figure 2a). When the Mn3+ ion was substituted at the Cr3+ ion site, a strain developed in the lattice because of the different ionic radii, and, to ensure the stability of the crystal structure, the CrO6 octahedra rotated or tilted. The strain factor was calculated by S = 2( a     c )/( a + c ), which arises as a result of octahedral rotation/tilting [21]. The distortion in the perovskite structure is also represented by the octahedral bond length and tilt angles θ[101] and ϕ[010]. In the present study, the octahedral bond length Cr–O and tilt angles θ[101] and ϕ[010] were calculated by the Zhao formalism [22] as follows:
Cr O =     a b 4 c ,   θ [ 101 ] = cos 1 c a   and   ϕ [ 010 ] = cos 1     2 c b
The octahedral bond length Cr–O increased in Mn-substituted compounds because the Mn3+ ion had a larger ionic radius than the Cr3+ ion. The rotation of the CrO6 octahedron occurred in both planes, i.e., in-plane (along a and c axis) and out-of-plane (along b axis). The tilt angle θ[101] was found to increase, while the rotational angle ϕ[010] decreased with an increase in the Mn3+ ion concentration (Figure 2b). In determining the stability of the perovskite structure, the Goldschmidt tolerance factor ( t ) plays a key role. In order to calculate the tolerance factor ( t ) for the present compounds, we used the formula t = < Gd/Er–O >/ 2 < Cr/Mn–O >. The unity value of t corresponds to an ideal perovskite structure, while 0.75 ≤ t ≤ 0.9 favors a distorted orthorhombic structure [23,24]. Further, it should also be noted that the classical consideration of ionic size alone cannot explain the structural characteristics. To obtain a more accurate understanding of the structural characteristics when Mn3+ ions replace Cr3+ ions, we performed a detailed structural analysis. Figure 3a presents the crystal structure of the GECO compound generated using VESTA software. It can be noticed that the structure is distorted from the ideal cubic structure (distortion shown in Figure 3b). This results in variation in the bond lengths. Figure 3c shows the Cr–O bond length variation in the Cr–O6 octahedron as a function of the Mn concentration. The difference among the three Cr–O bond lengths in all compounds can be correlated with the Jahn–Teller (JT) distortion of Mn3+ ions, along with a contribution from intrinsic structural distortion. The in-plane (orthorhombic-like) and out-of-plane (tetragonal-like) distortions evaluated by Q2 [= l y     l x ], and Q3 [= (2 ×   l z     l x     l y )/ 3 ], respectively, are shown in Figure 3d. Here, l x , l y and l z denote the long and short Cr–O2 bond lengths mainly in the a c plane and the middle bond length Cr–O1 mainly along the b axis (out of plane), respectively. O1 and O2 stand for the apical and equatorial oxygen atoms. Here, the large and positive magnitude of Q2 is associated with the Jahn–Teller distortion and the low and negative value of Q3 indicates that an out-of-plane distortion along the c axis is competing with the JT distortion [25,26]. This implies that the lattice deformation is primarily confined to the a c plane. A higher value of Q2 in comparison to Q3 indicates a decrease in the b lattice parameter value and an increase in the a and c parameters, as shown in Figure 2a. The continuous variation in the in-plane and out-of-plane distortions with Mn substitution should have an impact on the magnetic properties.

3.2. Magnetic Study

The zero-field-cooled (ZFC) and field-cooled (FC) magnetization plots of the GECMO series as a function of temperature from 2 K to 300 K at 100 Oe applied magnetic field are shown in Figure 4a–f. Above TN (i.e., the Néel transition temperature), both the FC and ZFC curves possess almost identical magnetization, whereas the curves become bifurcated below TN. The canted antiferromagnetism of Cr3+ ions may be responsible for this bifurcation. The transition temperature for the pristine ( x = 0) sample at 100 Oe magnetic field is ~144 K, which was evaluated by the first-order derivative of the magnetization temperature curve. The transition temperature decreased with an increase in Mn3+ ion content. The decrease in TN was due to the weakening of the AFM interaction between Cr3+ ions and the development of ferromagnetic double-exchange interaction Cr3+-Mn3+ via the O2- ion. It can be seen that, below TN, magnetization in both modes (ZFC and FC) increased to a certain value with the lowering of temperature. It attained the maximum value below 10 K, corresponding to the ordering of rare-earth Gd3+ and Er3+ ions, and after this, it started decreasing. The magnetization increased with Mn3+ ion doping for the samples with x = 0, 0.1, 0.2, 0.3 in FC mode, but it decreased in the case of the samples with higher Mn content. The magnetization in both modes (FC and ZFC) passed the zero magnetization and became negative for the two samples with x = 0.4 and 0.5. The temperature where the magnetization crosses the temperature axis ( M = 0) is known as the compensation temperature (Tcomp). Further, in the case of the compound with x = 0.4, magnetization again started increasing below 10 K. This may have been due to the rotation of the net magnetic moment from one easy direction to the other. In the present study, it is likely that a higher concentration of manganese was responsible for the creation of a negative internal field that rotated the magnetization of Gd3+ and Er3+ ions in the opposite direction of the magnetization of Cr3+ ions. As we shall see later, the calculated value of the internal field was negative (i.e., opposite to magnetization of Cr3+ ions) for compounds with x = 0.4 and 0.5.
Figure 5a–d demonstrate the zero-field-cooled and field-cooled magnetization as a function of temperature from 2 K to 300 K for the compound with x = 0.5 at 500 Oe, 1 kOe, 2.5 kOe and 5 kOe applied magnetic field. Here, we observed that, for a higher magnetic field, negative magnetization disappeared. This means that the applied field of these values dominated the internal field, so the overall magnetization became positive because of the flipping of the magnetization of Gd3+/Er3+ ions in the direction of the applied field.
Further, the FC magnetization data of the compounds, in which negative magnetization appeared, were fitted using the following equation [27]:
M = M Cr   + C H + H I T θ
In Equation (1), M is the total magnetization, MCr is the magnetization of canted Cr3+ ions, C is the Curie constant, HI is the internal field due to Cr3+ ions, H is the applied field and θ is the Weiss temperature. The fitting is shown by a solid red line in Figure 6a,b. The obtained parameters are listed in Table 2. We found negative values of internal field (HI) for these compounds, which support the notion that induced magnetization occurred in the opposite direction to the applied field.
Figure 7a–f express the inverse of magnetic susceptibility ( χ 1 ) versus temperature (T) plots of GECMO compounds measured at 100 Oe applied magnetic field. In the high-temperature region, the inverse of magnetic susceptibility obeys the Curie–Weiss law χ = C/(T-θ), where C is the Curie constant and θ is the Weiss temperature. The solid red line shows the fitted curve using the Curie–Weiss law in the paramagnetic region above the transition temperature. The experimental value of the effective magnetic moment was calculated in the unit of Bohr magneton ( μ B ) using the fitted Curie constant (C) value by the following equation [28]:
μ eff e x p . = 3 K B   C N = 8 c
where N is the Avogadro number and k B is the Boltzmann constant. The theoretical effective magnetic moment value was also estimated using the free ionic moments of Gd3+ (7.9 μ B ), Er3+ (9.59   μ B ), Cr3+ (3.87 μ B ) and Mn3+ (4.90 μ B ) by the following equation:
μ eff T h . = 0.5 μ eff G d 2 + 0.5 μ eff E r 2 + 1 x μ eff C r 2 + x μ eff M n 2
The value of fitted parameters (C and θ) and effective magnetic moment (calculated and experimental) are summarized in Table 3 for all compounds. Here, a negative value of θ suggests that the interaction is antiferromagnetic [29]. It is quite satisfactory that both the values of effective magnetic moment (calculated and experimental) match fairly well with each other.
As is apparent from Figure 7a–f, the 1/χ versus T data fit well with the Curie–Weiss law above the transition temperature ( T N ) and it deviates close to T N as there is an abrupt drop in 1/χ with temperature reduction. The reason for this is that our samples were in a canted antiferromagnetic state instead of simple antiferromagnetic. Such a deviation from the Curie–Weiss behavior can be accounted for by the antisymmetric exchange (DM) interaction [30]. Thus, the susceptibility was fitted by the modified Curie–Weiss law obtained from the DM interaction:
χ = C T θ T T 0 T T N C r ,  
where T 0   and T N C r are the fitted parameters, which can be calculated by the formula given by Moriya [30]:
T 0   = 2 J e   Z S S + 1 3 k B ,
T N C r = T 0   1 + D e 2   J e 2 1 / 2
where Z is the coordination number of Cr3+ ions, S is the spin quantum number of Cr3+ ions, and Je and De are the strength of symmetric and antisymmetric exchange interactions between Cr3+ and Cr3+ ions, respectively. The data fitted using Equation (4) are shown as the insets of Figure 7. The fitted parameters C , θ , T0 and T N C r are summarized in Table 3. The non-zero values of D e imply the presence of a DM interaction in these compounds.

3.3. Magnetocaloric Study

With the aim of exploring the efficiency of the materials for magnetic refrigeration applications, the isothermal magnetization M H curves of the GECMO samples were obtained in the first quadrant with an applied field of up to 8 T. Figure 8a–f show the M H curves obtained from 5 K to 170 K with an interval of 5 K for GECMO samples. There are two criteria that are used to characterize the magnetocaloric effect: the first is the magnetic entropy change (∆S) and the other is the relative cooling power (RCP). The magnetic entropy change was calculated by using the following Maxwell’s relation [31]:
S ( T ) H = H I H F M T ,       H H d H
Figure 9a–f display the magnetic entropy change calculated from Equation (7) as a function of temperature under various applied field strengths. It can be noticed that the value of S increased rapidly with a decrease in temperature. The maximum value of S is 26.78 J Kg−1K−1 at 7.5 K and at 8 T applied field for the x = 0 compound. This is comparable to the Er0.33Gd0.67CrO3 compound [19] and higher than other compounds such as DyCrO3, HoCrO3, ErCrO3, etc., and dysprosium acetate tetrahydrate [32]. Furthermore, with the Mn doping in the pristine compound, the value of S decreased. RCP was evaluated by the following equation
RCP = T 1 T 2 Δ S T Δ H d T
where T 1 and T 2 are the lower and higher temperature of the refrigeration cycle ( T 1 = 7.5 K and T 2 = 172.5 K in this study), respectively. A greater RCP value generally denotes a better magnetocaloric compound because of its high cooling efficiency. As displayed in the inset of Figure 9a–f, the value of RCP increased with the increase in applied field. The values of S and RCP for the GECMO series are listed in Table 4 for different applied magnetic field values.
To gain a better understanding of the magnetocaloric effect, we further explored the magnetic entropy change ( S) in view of the magneto-elastic coupling and electron interaction according to the Landau theory of phase transition [33,34]. According to this, the Gibbs free energy can be defined as:
G T , M = 1 2 A M 2 + 1 4 B M 4 + 1 6 C M 6 M H
where A , B and C are the temperature-dependent parameters (Landau coefficients). At equilibrium, energy should be minimized, i.e., G M = 0, and Equation (9) gives
H   M =   A   + B M 2 + C M 4
The Landau coefficients were calculated from the Arrott’s plots ( H / M vs. M 2) by the best fit method. Magnetic entropy change was obtained by differentiating Gibbs free energy with respect to temperature:
Δ S M T , H = 1 2 A T M 2 1 4 B T M 4 1 6 C T M 6
The magnetic entropy change calculated by Landau theory was compared with that calculated by Maxwell’s thermodynamic relations under 5 T magnetic field and the comparison is shown in Figure 10a–f. The variation in the Landau coefficients is also shown in the inset of each plot. It can be seen from the plots that the magnetic entropy change versus temperature behavior is the same according to different models.

4. Conclusions

In this paper, we studied Gd0.5Er0.5Cr1−xMnxO3 (x = 0, 0.1, 0.2, 0.3, 0.4, 0.5) compounds for their low-temperature magnetic and magnetocaloric properties. The M T measurements indicate that the transition temperature for the pristine sample was ~144 K, which decreased with Mn substitution. We also observed negative magnetization at lower magnetic fields in compounds with x = 0.4 and 0.5 Mn concentrations. The existence of the magnetization reversal offers the characteristic switching of magnetization. The magnetocaloric parameters such as magnetic entropy change ( S) varied from 16.74 J/kg-K to 7.46 J/kg-K and relative cooling power (RCP) from 375.94 J/kg to 220.22 J/kg with Mn concentration under a 5 T field change. These values are significant for these materials to be useful as low-temperature magnetic refrigerants, especially below 10 K.

Author Contributions

Conceptualization, N.P.; methodology, N.P.; validation, N.P.; formal analysis, K.S.; investigation, K.S., N.P., Y.B.; resources, S.K.; writing—original draft preparation K.S., K.K., V.S.P.; original draft correct N.P., Y.B. All authors have read and agreed to the published version of the manuscript.

Funding

The research was funded by DST SERB, New Delhi, India through grant No. ECR/2017/002681.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are thankful to the low-temperature high-magnetic-field facility of CURAJ for the magnetic measurements. Neeraj Panwar would like to thank DST SERB, New Delhi, UGC DAE CSR Mumbai and IUAC New Delhi for the grants through Projects ECR/2017/002681, CRS-M-298 and UFR-62317, respectively. V.S.P. acknowledges the National Research Council (NRC) senior research associate fellowship program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ateia, E.E.; Ismail, H.; Elshimy, H.; Abdelmaksoud, M.K. Structural and Magnetic Tuning of LaFeO3 Orthoferrite Substituted Different Rare Earth Elements to Optimize Their Technological Applications. J. Inorg. Organomet. Polym. Mater. 2021, 31, 1713–1725. [Google Scholar] [CrossRef]
  2. Chakraborty, P.; Basu, S. Structural, electrical and magnetic properties of Er doped YCrO3 nanoparticles. Mater. Chem. Phys. 2021, 259, 124053. [Google Scholar] [CrossRef]
  3. Yoshii, K. Magnetic properties of perovskite GdCrO3. J. Solid State Chem. 2001, 159, 204–208. [Google Scholar] [CrossRef]
  4. Aamir, M.; Bibi, I.; Ata, S.; Majid, F.; Kamal, S.; Alwadai, N.; Sultan, M.; Iqbal, S.; Aadil, M.; Iqbal, M. Graphene oxide nanocomposite with Co and Fe doped LaCrO3 perovskite active under solar light irradiation for the enhanced degradation of crystal violet dye. J. Mol. Liq. 2021, 322, 114895. [Google Scholar] [CrossRef]
  5. Wang, L.; Wang, S.W.; Zhang, X.; Zhang, L.L.; Yao, R.; Rao, G.H. Reversals of magnetization and exchange-bias in perovskite chromite YbCrO3. J. Alloys Compd. 2016, 662, 268–271. [Google Scholar] [CrossRef]
  6. Hornreich, R.M. Magnetic interactions and weak ferromagnetism in the rare-earth orthochromites. J. Magn. Magn. Mater. 1978, 7, 280–285. [Google Scholar] [CrossRef]
  7. Su, Y.; Zhang, J.; Feng, Z.; Li, L.; Li, B.; Zhou, Y.; Chen, Z.; Cao, S. Magnetization reversal and Yb3+/Cr3+ spin ordering at low temperature for perovskite YbCrO3 chromites. J. Appl. Phys. 2010, 108, 013905. [Google Scholar] [CrossRef]
  8. Cao, Y.; Cao, S.; Ren, W.; Feng, Z.; Yuan, S.; Kang, B.; Lu, B.; Zhang, J. Magnetization switching of rare earth orthochromite CeCrO3. Appl. Phys. Lett. 2014, 104, 232405. [Google Scholar] [CrossRef]
  9. Panwar, N.; Joby, J.P.; Kumar, S.; Coondoo, I.; Vashundhara, M.; Kumar, N.; Palai, R.; Singhal, R.; Katiyar, R.S. Observation of magnetization reversal behavior in Sm0.9Gd0.1Cr0.85Mn0.15O3 orthochromites. AIP Adv. 2018, 8, 055818. [Google Scholar] [CrossRef] [Green Version]
  10. Yin, S.; Sharma, V.; McDannald, A.; Reboredo, F.A.; Jain, M. Magnetic and magnetocaloric properties of iron substituted holmium chromite and dysprosium chromite. RSC Adv. 2016, 6, 9475–9483. [Google Scholar] [CrossRef]
  11. Sharma, M.K.; Mukherjee, K. Magnetic and universal magnetocaloric behavior of rare-earth substituted DyFe0.5Cr0.5O3. J. Magn. Magn. Mater. 2017, 444, 178–183. [Google Scholar] [CrossRef] [Green Version]
  12. Yin, S.; Seehra, M.S.; Guild, C.J.; Suib, S.L.; Poudel, N.; Lorenz, B.; Jain, M. Magnetic and magnetocaloric properties of HoCrO3 tuned by selective rare-earth doping. Phys. Rev. B 2017, 95, 184421. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, H.; Qian, H.; Xie, L.; Guo, Y.; Liu, Y.; He, X. The spin reorientation and improvement of magnetocaloric effect in HoCr1−xGaxO3 (0 ≤ x ≤ 0.5). J. Alloys Compd. 2021, 885, 160863. [Google Scholar] [CrossRef]
  14. Yoshii, K. Positive exchange bias from magnetization reversal in La1−xPrxCrO3 (x ~ 0.7–0.85). Appl. Phys. Lett. 2011, 99, 142501. [Google Scholar] [CrossRef]
  15. Oliveira, G.N.P.; Machado, P.; Pires, A.L.; Pereira, A.M.; Araújo, J.P.; Lopes, A.M.L. Magnetocaloric effect and refrigerant capacity in polycrystalline YCrO3. J. Phys. Chem. Solids 2016, 91, 182–188. [Google Scholar] [CrossRef]
  16. Oliveira, G.N.P.; Pires, A.L.; Machado, P.; Pereira, A.M.; Araújo, J.P.; Lopes, A.M.L. Effect of chemical pressure on the magnetocaloric effect of perovskite-like RCrO3 (R-Yb, Er, Sm and Y). J. Alloys Compd. 2019, 797, 269–276. [Google Scholar] [CrossRef]
  17. Kumar, S.; Coondoo, I.; Vasundhara, M.; Kumar, S.; Kholkin, A.L.; Panwar, N. Structural, magnetic, magnetocaloric and specific heat investigations on Mn doped PrCrO3 orthochromites. J. Phys. Condens. Matter 2017, 29, 195802. [Google Scholar] [CrossRef] [Green Version]
  18. Yin, L.H.; Yang, J.; Kan, X.C.; Song, W.H.; Dai, J.M.; Sun, Y.P. Giant magnetocaloric effect and temperature induced magnetization jump in GdCrO3 single crystal. J. Appl. Phys. 2015, 117, 133901. [Google Scholar] [CrossRef]
  19. Shi, J.; Yin, S.; Seehra, M.S.; Jain, M. Enhancement in magnetocaloric properties of ErCrO3 via A-site Gd substitution. J. Appl. Phys. 2018, 123, 193901. [Google Scholar] [CrossRef]
  20. Kanwar, K.; Coondoo, I.; Anas, M.; Malik, V.K.; Kumar, P.; Kumar, S.; Kulriya, P.K.; Kaushik, S.D.; Panwar, N. A comparative study of the structural, optical, magnetic and magnetocaloric properties of HoCrO3 and HoCr0.85Mn0.15O3 orthochromites. Ceram. Int. 2021, 47, 7386–7397. [Google Scholar] [CrossRef]
  21. Chan, T.S.; Liu, R.S.; Yang, C.C.; Li, W.-H.; Lien, Y.H.; Huang, C.Y.; Lee, J.F. Chemical Size Effect on the Magnetic and Electrical Properties in the (Tb1−x Eux)MnO3(0 ≤ x ≤ 1.0) System. J. Phys. Chem. B 2007, 111, 2262–2267. [Google Scholar] [CrossRef] [PubMed]
  22. Zhao, Y.; Weidner, D.J.; Parise, J.B.; Cox, D.E. Thermal expansion and structural distortion of perovskite-data for NaMgF3 perovskite. Part I. Phys. Earth Planet. Inter. 1993, 76, 1–16. [Google Scholar] [CrossRef]
  23. Søndenå, R.; Stølen, S.; Ravindran, P.; Grande, T.; Allan, N.L. Corner-versus face-sharing octahedra in AMnO3 perovskites (A = Ca, Sr, and Ba). Phys. Rev. B 2007, 75, 184105. [Google Scholar] [CrossRef] [Green Version]
  24. Kim, K.; Siegel, D.J. Correlating lattice distortions, ion migration barriers, and stability in solid electrolytes. J. Mater. Chem. A 2019, 7, 3216–3227. [Google Scholar] [CrossRef]
  25. Mahana, S.; Pandey, S.K.; Rakshit, B.; Nandi, P.; Basu, R.; Dhara, S.; Turuchini, S.; Zema, N.; Manju, U.; Mahanti, S.D.; et al. Site substitution in GdMnO3: Effects on structural, electronic, and magnetic properties. Phys. Rev. B 2020, 102, 245120. [Google Scholar] [CrossRef]
  26. Chiang, F.K.; Chu, M.W.; Chou, F.C.; Jeng, H.T.; Sheu, H.S.; Chen, F.R.; Chen, C.H. Effect of Jahn-Teller distortion on magnetic ordering in Dy(Fe,Mn)O3 perovskites. Phys. Rev. B 2011, 83, 245105. [Google Scholar] [CrossRef]
  27. Cooke, A.H.; Martin, D.M.; Wells, M.R. Magnetic interactions in gadolinium orthochromite, GdCrO3. J. Phys. C Solid State Phys. 1974, 7, 3133. [Google Scholar] [CrossRef]
  28. Kumar, D.; Jena, P.; Singh, A.K. Structural, magnetic and dielectric studies on half-doped Nd0.5Ba0.5CoO3 perovskite. J. Magn. Magn. Mater. 2020, 516, 167330. [Google Scholar] [CrossRef]
  29. Deng, D.; Wang, X.; Zheng, J.; Qian, X.; Yu, D.; Sun, D.; Jing, C.; Lu, B.; Kang, B.; Cao, S.; et al. Phase separation and exchange bias effect in Ca doped EuCrO3. J. Magn. Magn. Mater. 2015, 395, 283–288. [Google Scholar] [CrossRef]
  30. Moriya, T. Anisotropic Superexchange Interaction and Weak Ferromagnetism. Phys. Rev. 1960, 120, 91–98. [Google Scholar] [CrossRef]
  31. Tishin, A.M. Magnetocaloric Effect in the Vicinity of Phase Transitions. Handb. Magn. Mater. 1999, 12, 395–524. [Google Scholar] [CrossRef]
  32. Lorusso, G.; Roubeau, O.; Evangelisti, M. Rotating Magnetocaloric Effect in an Anisotropic Molecular Dimer. Angew. Chem. Int. Ed. 2016, 55, 3360–3363. [Google Scholar] [CrossRef] [PubMed]
  33. Landau, L.D.; Lifshitz, E.M. Statistical Physics; Pergamon: New York, NY, USA, 1958. [Google Scholar]
  34. Amaral, V.S.; Amaral, J.S. Magnetoelastic coupling influence on the magnetocaloric effect in ferromagnetic materials. J. Magn. Magn. Mater. 2004, 276, 2104–2105. [Google Scholar] [CrossRef]
Figure 1. Rietveld-refined XRD patterns of the GECMO series.
Figure 1. Rietveld-refined XRD patterns of the GECMO series.
Crystals 12 00263 g001
Figure 2. (a) Lattice parameters, volume, (b) strain parameter and tilt angle variation as a function of Mn concentration.
Figure 2. (a) Lattice parameters, volume, (b) strain parameter and tilt angle variation as a function of Mn concentration.
Crystals 12 00263 g002
Figure 3. (a) Crystal structure of Gd0.5Er0.5CrO3 compound obtained from the VESTA software; l x , l y and l z denote the long and short Cr O2 bond lengths mainly in the a c plane and the middle length Cr O1 mainly along the b axis (out of plane), respectively. (b) Q2 and Q3 represent the orthogonal-like and tetragonal-like distortions, respectively. (c) Bond lengths and (d) octahedral distortion variation as a function of Mn concentration.
Figure 3. (a) Crystal structure of Gd0.5Er0.5CrO3 compound obtained from the VESTA software; l x , l y and l z denote the long and short Cr O2 bond lengths mainly in the a c plane and the middle length Cr O1 mainly along the b axis (out of plane), respectively. (b) Q2 and Q3 represent the orthogonal-like and tetragonal-like distortions, respectively. (c) Bond lengths and (d) octahedral distortion variation as a function of Mn concentration.
Crystals 12 00263 g003
Figure 4. (af): ZFC and FC magnetization curves of GECMO series measured under 100 Oe magnetic field.
Figure 4. (af): ZFC and FC magnetization curves of GECMO series measured under 100 Oe magnetic field.
Crystals 12 00263 g004
Figure 5. (ad): ZFC and FC magnetization curves of x = 0.5 compound measured under 500 Oe, 1 kOe, 2.5 kOe and 5 kOe magnetic field.
Figure 5. (ad): ZFC and FC magnetization curves of x = 0.5 compound measured under 500 Oe, 1 kOe, 2.5 kOe and 5 kOe magnetic field.
Crystals 12 00263 g005aCrystals 12 00263 g005b
Figure 6. (a,b): The FC magnetization curves fitted to Equation (1) for compounds with x = 0.4 and x = 0.5 measured under 100 Oe magnetic field.
Figure 6. (a,b): The FC magnetization curves fitted to Equation (1) for compounds with x = 0.4 and x = 0.5 measured under 100 Oe magnetic field.
Crystals 12 00263 g006
Figure 7. (af): Inverse magnetic susceptibility versus temperature plots of GECMO series measured at 100 Oe applied magnetic field. Open circles represent the experimental data points whereas solid lines of plot display the fitting to the Curie–Weiss law. Insets represent the fitting of inverse magnetic susceptibility versus temperature using Equation (4).
Figure 7. (af): Inverse magnetic susceptibility versus temperature plots of GECMO series measured at 100 Oe applied magnetic field. Open circles represent the experimental data points whereas solid lines of plot display the fitting to the Curie–Weiss law. Insets represent the fitting of inverse magnetic susceptibility versus temperature using Equation (4).
Crystals 12 00263 g007aCrystals 12 00263 g007b
Figure 8. (af): Isothermal magnetization M–H curves of the GECMO series.
Figure 8. (af): Isothermal magnetization M–H curves of the GECMO series.
Crystals 12 00263 g008aCrystals 12 00263 g008b
Figure 9. (af): The temperature dependence of magnetic entropy change ( S) and relative cooling power (inset) of GECMO series in magnetic fields up to 8 T.
Figure 9. (af): The temperature dependence of magnetic entropy change ( S) and relative cooling power (inset) of GECMO series in magnetic fields up to 8 T.
Crystals 12 00263 g009aCrystals 12 00263 g009b
Figure 10. (af): Temperature dependence of magnetic entropy change ( S) of GECMO series calculated using Landau theory and Maxwell’s relation. Insets show the variation in Landau coefficients.
Figure 10. (af): Temperature dependence of magnetic entropy change ( S) of GECMO series calculated using Landau theory and Maxwell’s relation. Insets show the variation in Landau coefficients.
Crystals 12 00263 g010aCrystals 12 00263 g010b
Table 1. Lattice parameters, unit cell volume ( V ), reliability factors, octahedral bond length, tilt angles, strain factor and Goldschmidt tolerance factor ( t ) of GECMO series.
Table 1. Lattice parameters, unit cell volume ( V ), reliability factors, octahedral bond length, tilt angles, strain factor and Goldschmidt tolerance factor ( t ) of GECMO series.
x = 0 x = 0.1 x = 0.2 x = 0.3 x = 0.4 x = 0.5
a   (Å)5.52205.53355.55305.58355.60725.6372
b (Å)7.56487.55417.54397.53767.51837.5029
c (Å)5.26915.27095.27215.27675.27405.2744
V 3)220.109220.330220.860222.081222.340223.085
χ2 (%)3.292.482.081.942.622.03
R P (%)1916.716.116.719.519.2
R WP (%)14.912.611.411.713.612.8
Cr–O (Å)1.9821.98251.98641.99251.99832.0047
θ (ο)17.4017.7118.2919.0619.8420.65
ϕ (ο)9.9229.328.758.067.226.19
S 0.04680.04860.05180.05640.06120.0665
t 0.85230.85100.84970.84840.84720.8459
Table 2. List of fitting parameters for negative magnetization of M T curves in FC mode as shown in Figure 6a,b.
Table 2. List of fitting parameters for negative magnetization of M T curves in FC mode as shown in Figure 6a,b.
CompoundExternal Field (Oe) H I
(Oe)
θ
(K)
x = 0.4100−301.81−24.43
x = 0.5100−360.70−24.59
Table 3. Magnetic parameters: Curie constant C (emu-K/Oe-mole), Weiss temperature θ (K), experimental and theoretical effective magnetic moment obtained from Curie–Weiss fitting, and T N C r (K), θ   (K), C   (emu-K/Oe-mole), fitting parameter T0 (K) and symmetric and antisymmetric exchange constant obtained from modified Curie–Weiss fitting of inverse susceptibility vs. temperature data.
Table 3. Magnetic parameters: Curie constant C (emu-K/Oe-mole), Weiss temperature θ (K), experimental and theoretical effective magnetic moment obtained from Curie–Weiss fitting, and T N C r (K), θ   (K), C   (emu-K/Oe-mole), fitting parameter T0 (K) and symmetric and antisymmetric exchange constant obtained from modified Curie–Weiss fitting of inverse susceptibility vs. temperature data.
Compound x = 0 x = 0.1 x = 0.2 x = 0.3 x = 0.4 x = 0.5
θ (K)−26.59−28.57−26.29−24.39−23.73−21.48
C(emu-K/Oe-mole)10.9911.8011.5611.5911.7611.70
μ eff e x p . (μB)9.389.729.629.639.709.68
μ eff T h . (μB)9.609.659.699. 749.799.83
T N C r (K)144.28132.77119.10103.9186.0268.01
T0 (K)143.03131.33117.13102.6884.2366.71
θ (K)−53.42−48.98−51.07−36.70−36.8−28.34
C   (emu-K/Oe-mole)11.9112.5012.4112.0012.1911.91
Je/ k B9.548.767.816.855.624.53
De/ k B2.532.602.882.132.331.80
Table 4. Magnetic entropy change ( S) and relative cooling power (RCP) at different applied magnetic fields for the Gd0.5Er0.5Cr1−xMnxO3 series.
Table 4. Magnetic entropy change ( S) and relative cooling power (RCP) at different applied magnetic fields for the Gd0.5Er0.5Cr1−xMnxO3 series.
Compound x = 0 x = 0.1 x = 0.2 x = 0.3 x = 0.4 x = 0.5
HMax.
(T)
SMax.
(J/kg-K)
RCP
(J/kg)
SMax.
(J/kg-K)
RCP
(J/kg)
SMax.
(J/kg-K)
RCP
(J/kg)
SMax.
(J/kg-K)
RCP
(J/kg)
SMax.
(J/kg-K)
RCP
(J/kg)
SMax.
(J/kg-K)
RCP
(J/kg)
1 3.3475.183.2577.192.8770.152.9465.522.1058.731.4944.04
2 6.69150.376.51154.385.74140.305.88131.044.20117.462.9888.09
3 10.04225.569.77231.578.62210.468.82196.566.31176.204.47132.13
4 13.39300.7513.02308.7611.49280.6111.76262.088.41234.935.96176.18
5 16.74375.9416.28385.9514.36350.7614.70327.6110.52293.677.46220.22
6 20.09451.1319.54463.1417.24420.9217.64393.1312.62352.408.95264.27
7 23.44526.3222.79540.3320.11491.0720.58458.6514.72411.1310.44308.32
8 26.78601.5126.05617.5222.99561.2323.53524.1716.83469.8711.93352.36
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Panwar, N.; Singh, K.; Kanwar, K.; Bitla, Y.; Kumar, S.; Puli, V.S. Low-Temperature Magnetic and Magnetocaloric Properties of Manganese-Substituted Gd0.5Er0.5CrO3 Orthochromites. Crystals 2022, 12, 263. https://doi.org/10.3390/cryst12020263

AMA Style

Panwar N, Singh K, Kanwar K, Bitla Y, Kumar S, Puli VS. Low-Temperature Magnetic and Magnetocaloric Properties of Manganese-Substituted Gd0.5Er0.5CrO3 Orthochromites. Crystals. 2022; 12(2):263. https://doi.org/10.3390/cryst12020263

Chicago/Turabian Style

Panwar, Neeraj, Kuldeep Singh, Komal Kanwar, Yugandhar Bitla, Surendra Kumar, and Venkata Sreenivas Puli. 2022. "Low-Temperature Magnetic and Magnetocaloric Properties of Manganese-Substituted Gd0.5Er0.5CrO3 Orthochromites" Crystals 12, no. 2: 263. https://doi.org/10.3390/cryst12020263

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