In this report, the structural and magnetic properties of GdCrTiO5 nanoparticles were explored, which were synthesized through the sol-gel technique and subsequently calcined (at 800 °C). X-ray diffraction (XRD) studies revealed the orthorhombic crystal structure of synthesized GdCrTiO5 nanoparticles with space group Pbam. The transmission electron microscopy (TEM) images, with selected area electron diffraction (SAED) pattern, showed the particle size as 38.0 ± 0.4 nm and single crystalline nature of the sample. The temperature dependence of dc magnetization, M(T), was measured for GdCrTiO5 nanoparticles, and different magnetic transitions were confirmed, including the spin reorientation (TSR), Néel temperature (TN) and compensation temperatures (Tcomp1 and Tcomp2) in the material. Irreversibility appeared in field-cool-cooling (FCC), and field-cool-warming (FCW) curves at low temperatures, indicating a ferromagnetic-antiferromagnetic (FM-AFM) transition. Earlier, this FM-AFM transition and TSR, Tcomp was not observed in bulk GdCrTiO5. Both the FM nature and exchange bias (EB) effect are further established from the field-dependent magnetization measurements. Furthermore, a change in isothermal magnetic entropy (−ΔSm) of 22 ± 3 J.kg-1.K-1 is found below 10 K, for a 7 T difference in the field. The obtained magnetic properties in this report are discussed in terms of exchange frustration originating from the competing interactions of the magnetic sublattices of the Cr3+ and Gd3+ in the GdCrTiO5 nanoparticles.

The interest in materials with sizable magnetocaloric effect (MCE) has grown significantly because of the possibilities of their future uses as solid refrigerants in magnetic cooling systems.1 Magnetic cooling has been recognized as a promising alternative to conventional vapour-cycle refrigeration because of its low power consumption, improved efficiency, and environmental friendliness.1–4 Magnetic refrigeration is an entropically-driven thermodynamic phenomenon, which couples magnetic entropy with lattice thermal energy and permits cooling from room temperature to the sub-Kelvin temperature region.5 

Multiferroic materials consist of multiple ferroic states,6,7 and it is envisioned that the co-dependence of the order parameters will contribute towards spintronics and innovative memory devices.8–10 In the RCrTiO5 series, one can expect to find exceptional and fascinating magnetic behavior because of the co-existence of two magnetic sublattices, Cr3+ and R3+.11–14 There are limited reports on this group of materials, mostly on basic magnetic properties in polycrystalline bulk materials.6–13 Few reports systematically studied magnetic properties of bulk GdCrTiO5.9–11 Recently, a study demonstrated magnetic transitions such as spin reorientation (SR) and compensation in bulk DyCrTiO5, as well as in HoCrTiO5, and also observed exchange bias (EB) effects.13 Due to this exchange coupling, a shift in the hysteresis loop across the applied field axis originates in ferromagnetic-antiferromagnetic (FM-AFM) systems.15 The EB effect due to exchange coupling among different atoms of an oxide compound having a rare-earth constituent depicts interesting magnetic behavior.13–15 Nowadays, the application of nanostructured materials is enhanced due to their improved properties, which is a consequence of the increase in the surface-to-volume ratio and uncompensated surface spins present on the grain surfaces.16,17 The uncompensated surface spins present at the surface, significantly modify magnetic properties and other physical properties.16,17 The change in the structural properties subsequently brings about concomitant changes in the magnetic nature of these materials.17–22 Zheng et al.,23 observed the EB effect in YMnO3, which is ascribed to the uncompensated surface spins of the nanoparticles. Previously, the MCE has been studied in bulk GdCrTiO5 compound.6 In this contribution, GdCrTiO5 nanoparticles were synthesized, and their magnetic properties were studied.

The GdCrTiO5 specimen was synthesized by a cost-effective sol-gel technique.14,16,17 Stochiometric amounts of Gd(NO3)3, Cr(NO3)3 and Ti(OC3H7)4 were added to absolute ethanol. The prepared solution was stirred continuously by adding a few drops of HNO3. After that, 50 mg of cetyltrimethylammoniumbromide (CTAB) was added to stirring solution, subsequently 10 ml of distilled water was poured into the solution while stirring was continued for another 30 minutes. The prepared reaction mixture was allowed to age for a day. The aged solution was dried, and then the obtained powder was consequently crushed and calcined at 800 °C for three hours in a box furnace. The structural and morphological features of the calcined powders were explored by x-ray diffraction (XRD) technique using Cu-Kα radiation (λ = 1.5406 Å), and transmission electron microscope (TEM). A vibrating sample magnetometer (VSM) using a Cryogenic, Cryogen Free Physical Properties Measurement System, with a vibrating sample magnetometer insert was used in order to study the magnetic behaviour as a function of temperature and applied magnetic field.

XRD pattern of GdCrTiO5 sample along with the profile fitting using FULLPROF24 are plotted in Fig. 1(a). The simulated XRD pattern is obtained by considering the orthorhombic crystal structure with a Pbam space group. The cell parameters a, b and c are found to be 7.381 ± 0.002, 8.679 ± 0.002 and 5.863 ± 0.002 Å, respectively, while the cell volume found from the refinement parameters is 375.6 ± 0.2 Å3. An earlier report on a polycrystalline bulk GdCrTiO5 sample, which has been prepared by the solid-state reaction method, shows an orthorhombic phase with the cell parameters a, b and c given as 7.4385, 8.5814, and 5.7782 Å, respectively.6 In a comparison of the lattice parameters of the synthesized powder with that of the bulk sample,6 it is observed that the lattice size is increased, which might be the consequence of the decreased size of the particles.25 It is observed that as the size of the crystallites is reduced, the lattice parameter increases.26 Minor peak shifting has appeared in a few peak positions as a result of internal stresses in nano-dimensional material.27 Broadening of the XRD peaks is observed when the size of the particle is reduced.28 In the present study peaks are broadened compared to the bulk,6 because of the reduction in size of the particle to nanoscale as confirmed by the TEM. Certain XRD peaks are merged when samples are in the nano form.14 As a result of peak shifting and merging of XRD reflections, a difference is observed between the experimental and calculated data is obtained in the present case.27,28 Furthermore, the crystallite size of the sample was calculated using Scherrer’s formula17 and found to be 33.0 ± 0.2 nm.

FIG. 1.

(a) The Le-Bail profile fit to the XRD pattern of the GdCrTiO5 nanoparticles using the FULLPROF program.31 After fitting, the observed pattern and simulated data are shown as red dots and a continuous black line, respectively. In contrast, the difference between observed to simulated data is shown in a blue-colored line. The green lines situated in the middle of the fitted pattern indicate Bragg peaks’ location and the difference data is indicated in blue. (b) The TEM image of GdCrTiO5 nanoparticles, with the corresponding (c) particle size distribution using log-normal fit, (d) SAED pattern, and (e) energy-dispersive x-ray spectroscopy.

FIG. 1.

(a) The Le-Bail profile fit to the XRD pattern of the GdCrTiO5 nanoparticles using the FULLPROF program.31 After fitting, the observed pattern and simulated data are shown as red dots and a continuous black line, respectively. In contrast, the difference between observed to simulated data is shown in a blue-colored line. The green lines situated in the middle of the fitted pattern indicate Bragg peaks’ location and the difference data is indicated in blue. (b) The TEM image of GdCrTiO5 nanoparticles, with the corresponding (c) particle size distribution using log-normal fit, (d) SAED pattern, and (e) energy-dispersive x-ray spectroscopy.

Close modal

In Fig. 1(b) the TEM micrograph for GdCrTiO5 is shown. The morphology indicates that the particles are nearly spherical in shape. The particle size histogram with a log-normal fit29 to the histogram is shown in Fig. 1(c). The average particle size for the synthesized sample is found to be 38.0 ± 0.4 nm, with particle size ranging from 15 to 90 nm. The particle size and crystallite size are nearly equal, which indicates single crystalline nature of the particles. In this contribution, the nanoparticles of GdCrTiO5 were obtained after three hours of calcination at a temperature of 800 °C, while the bulk GdCrTiO5 sample was prepared at a sintering temperature of 1400 °C over several days.6 The selected area electron diffraction (SAED) pattern is depicted in Fig. 1(d). The dotted pattern indicates the crystalline sample with single crystal nature.17 The elements present in the sample are confirmed using energy-dispersive x-ray spectroscopy (EDS). The EDS spectrum is shown in Fig. 1(e), indicating that the sample consists of the constituent elements Gd, Cr, Ti and O. The presence of C and Cu in the spectra (Fig. 1(e)) is from the carbon-coated Cu grid used for the microscopy.

To understand the magnetic properties of the synthesized GdCrTiO5 sample, magnetic moment as a function of temperature, M(T), was measured in zero-field-cooling (ZFCW), field-cool-cooling (FCC) and field-cool-warming (FCW) cycles at the applied magnetic field, 0.01, 0.05 and 0.1 T.30–38 The M(T) curves as a function of the indicated applied magnetic fields for the GdCrTiO5 sample are shown in Fig. 2(a)-(c). The M(T) curves diverge on reducing the temperature from 150 to 140 K. This type of behaviour is attributed to the competing magnetic interactions in the material.14,39 The temperature range 0 to 200 K is shown in Fig. 2(a)-(c), as the sample remains paramagnetic beyond 200 K. The nature of the M(T) curves shows the irreversible nature in all three measurement cycles at low field. The complicated magnetic properties in this system are expected because of the competing Cr3+-Cr3+, Gd3+-Gd3+ and Gd3+-Cr3+ exchange interactions.29 From Fig. 2(a), a magnetic transition associated with the Néel temperature, TN = 190 ± 2 K, is observed. In the temperature range T < TN, the magnetization curves show very complex and interesting features, as the magnetization drastically changes in this region. The antiferromagnetic (AFM) moment is slightly canted from their arrangement and a ferromagnetic (FM) component is emerges because of the anti-symmetric exchange interaction of the Cr3+-Cr3+ ions, called as Dzyaloshinskii-Moriya (DM) interaction.13–15 In addition, at low temperatures, an anisotropic exchange develops between the Gd3+ and Cr3+ spins, which generates an effective field at the Gd3+ site, showing a moment opposite to Cr3+.13,15 With an increase in the temperature from 2 K, the magnetization decreases rapidly due to the faster decrease of Gd3+ moment.13 In FCC and FCW cycles, all the local moments of Gd3+ and Cr3+, cancel each other at 9 K. As a result, the net magnetic moment becomes zero, termed as the compensation temperature, referred to as Tcomp1.29 Above Tcomp1, the resultant magnetization vector orients in the reverse direction to the external applied magnetic field with increasing temperature. As a result, negative magnetization is observed, and this extends up to a minimum value at TSR of about 20 K. At this temperature, a new magnetic phase transition occurs because of spin reorientation of Cr3+ ion.13 Increasing the temperature further, the net magnetization enhances and again compensates at a temperature of 115 K, referred to as Tcomp2.29 In between Tcomp1 and Tcomp2 the sample shows negative magnetization.

FIG. 2.

M(T) data for GdCrTiO5 nanoparticles with probing field (a) 0.01 T, (b) 0.05 T, and (c) 0.1 T.

FIG. 2.

M(T) data for GdCrTiO5 nanoparticles with probing field (a) 0.01 T, (b) 0.05 T, and (c) 0.1 T.

Close modal

For T > Tcomp2 the sample shows positive magnetization. Another important feature is that higher M(ZFC) values are observed compared to the M(FC) values, in this case at 0.01 and 0.05 T applied field. A previous report stated that various interactions like FM, AFM, and canted spin structure in the material result in higher magnetization in the ZFC cycle than FC.40 The magnetization as a function of temperature was previously studied in bulk GdCrTiO5 at 0.005 T, showing a decrease in magnetization with increasing temperature, but TSR and Tcomp has not observed.6 Therefore, for the first time, the spin reorientation and negative magnetization features are reported here for the nanoparticles of GdCrTiO5. Furthermore, a loop is observed in FCC and FCW cycles below the temperature 60 K (Fig. 2(a)-(c)). This nature of FCC and FCW cycles was previously observed in some manganites and metallic alloy systems.30–38 However, this behavior is not observed in the GdCrTiO5 system.6–13 The obtained feature indicates an FM-AFM phase transition at low temperature.36,37 The observed loop has been suppressed with the increase in probing magnetic field, which is completely suppressed at the higher fields.30,32 This phenomenon is termed kinetic arrest.30–38 In the present case, a similar suppression of the loop between FCC and FCW magnetization data has been observed (Fig. 2(c)) when the probing field becomes 0.1 T.30,32

A further effect of the magnetic field on the magnetization of the material is explored. Fig. 3(a) shows the five-segment Mμ0H loop of GdCrTiO5, recorded at 2, 10 and 50 K. Considering the low field region shown in Fig. 3(b), all three measurement cycles show small coercivity, which indicates the FM nature of the material due to DM interaction in Cr3+ moment.15 The obtained loops are asymmetric in nature, referred to as the EB effect.13–15 The EB field is evaluated by using the relation, Hex = H1+H2/2, where H1 and H2 are the positive and negative intercepts of the Mμ0H curve on the field axis.14 The obtained value of Hex for the nanoparticles GdCrTiO5 is 0.008 T. To examine the magnetocaloric nature of the material; isothermal magnetic measurement is carried out from 0 to 7 T in the temperature range 2 to 50 K, shown in Fig. 3(c). At low temperatures, the Mμ0H loops are nonlinear but do not attain saturation, whereas, at high temperatures, these curves become linear. This change in the behaviour of the Mμ0H loops is observed because of the broken exchange interactions within Gd3+ ions.4 

FIG. 3.

(a) The magnetization as a function of applied field hysteresis loops measured at 2, 10, and 50 K, for GdCrTiO5 nanoparticles; with (b) showing the zoomed view of the curves for the low field regions. (c) Magnetization as a function of applied fields between 0 to 7 T, measured at fixed temperatures in the range from 2 to 50 K. (d) Arrott plots and (e) temperature dependence of isothermal entropy change for fixed magnetic fields indicated in the legend.

FIG. 3.

(a) The magnetization as a function of applied field hysteresis loops measured at 2, 10, and 50 K, for GdCrTiO5 nanoparticles; with (b) showing the zoomed view of the curves for the low field regions. (c) Magnetization as a function of applied fields between 0 to 7 T, measured at fixed temperatures in the range from 2 to 50 K. (d) Arrott plots and (e) temperature dependence of isothermal entropy change for fixed magnetic fields indicated in the legend.

Close modal

In order to better understand the critical behaviour of spins in a magnetic phase transition, Arrott plots are plotted at different temperatures, shown in Fig. 3(d). According to the literature,4,39 Arrott plots are broadly used to establish Curie or Néel temperature in magnetic materials and order of the magnetic transition. If the transition is first-order, then the curve’s slope is negative, and for the second-order transition, a positive slope is expected.4,39,41 In the present case, all Arrott plots exhibit a positive slope, indicating the second-order nature of the material.4 The magnetocaloric effect (MCE) of the material is examined by utilizing the isothermal magnetization data. In this report, the magnetization is measured at different intervals of temperature and field. Therefore, the change in isothermal magnetic entropy ΔSm is numerically calculated with the following equation,6 

ΔSm=iMi+1MiTi+1TiΔHi,

where, Mi+1 and Mi is the magnetization at temperature Ti+1 and Ti for the change in the magnetic field ΔHi. The temperature-dependent ΔSm is calculated using the expression, and data is plotted in Fig. 3(e). A large negative value of ΔSm is obtained for this system similar to that observed in other Gd-based systems.2–5 The maximum value of the change in isothermal magnetic entropy (−ΔSm) of 22 ± 3 J.kg-1.K-1 is found below 10 K, for a 7 T change in the field. The MCE in these GdCrTiO5 nanoparticles is associated with degenerated frustrated magnetic states due to the competition between DM interaction and spatial anisotropy.6 When a magnetic field is applied, the degeneracy is removed and frustration in magnetic moment appears in the material; consequently a change in magnetic entropy occurs.

This is the first report on GdCrTiO5 nanoparticles, synthesized using a simple sol-gel technique to explore the role of particle size on structural and magnetic properties. The sample was stabilized with an orthorhombic crystal structure (space group Pbam). The single crystalline nature of the particles with a size of 38.0 ± 0.4 nm is evidenced from the TEM analysis. From the temperature-dependent magnetization measurements, different magnetic transitions, including TSR and TN, are observed. Magnetization as a function of field measurement showed the good MCE (−ΔSm = 22 ± 3 J.kg-1.K-1) in the synthesized sample. The observed magnetic properties of the material are attributed to the coupling between Cr3+ and Gd3+ ions.

Financial aid from the SANRF (Grant No. 120856), as well as the University of Johannesburg (UJ) URC and FRC (Grant Nos. 282810, 083135), as well as Spectrum is acknowledged. The use of the NEP Cryogenic Physical Properties Measurement System at UJ, obtained with the financial support from the SANRF (Grant No. 88080) and UJ, RSA, is acknowledged. The use of the Spectrum facility in the FoS at UJ is acknowledged.

The authors have no conflicts to disclose.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
M.
Balli
,
S.
Jandl
,
P.
Fournier
, and
D. Z.
Dimitrov
,
Appl. Phys. Letters
108
,
102401
(
2016
).
2.
K.
Dey
,
A.
Indra
,
S.
Majumdar
, and
S.
Giri
,
J. Mater. Chem. C
5
,
1646
(
2017
).
3.
J. Y.
Zhang
,
J.
Luo
,
J. B.
Li
,
J. K.
Liang
,
Y. C.
Wang
,
L. N.
Ji
,
Y. H.
Liu
, and
G. H.
Rao
,
J. Alloy and Compounds
469
,
15
(
2009
).
4.
D. N.
Petrov
,
V.
Lovchinov
,
B.
The Huy
,
P.
The Long
,
N. T.
Dang
, and
D. S.
Yang
,
J. Appl. Phys.
126
,
135101
(
2019
).
5.
Z.
Yang
,
J.-Yi
Ge
,
S.
Ruan
,
H.
Cui
, and
Yu-J.
Zeng
,
J. Mater. Chem. C
9
,
6754
(
2021
).
6.
M.
Das
,
S.
Roy
,
N.
Khan
, and
P.
Mandal
,
Phys. Rev. B
98
,
104420
(
2018
).
7.
J.
Hwang
,
E. S.
Choi
,
H. D.
Zhou
,
J.
Lu
, and
P.
Schlottmann
,
Phys. Rev. B
85
,
024415
(
2012
).
8.
J.
Saha
,
G.
Sharma
, and
S.
Patnaik
,
J. Magn. Magn. Mater.
360
,
34
(
2014
).
9.
T.
Basu
,
D. T.
Adroja
,
F.
Kolb
,
H.-A.
Krug von Nidda
,
A.
Ruff
,
M.
Hemmida
,
A. D.
Hillier
,
M.
Telling
,
E. V.
Sampathkumaran
,
A.
Loidl
, and
S.
Krohns
,
Phys. Rev. B
96
,
184431
(
2017
).
10.
E. B.
Guedes
,
H. P.
Martins
,
M.
Abbate
,
S. H.
Masunaga
,
F. E. N.
Ramirez
,
R. F.
Jardim
, and
R. J. O.
Mossanek
,
J. Alloy and Compounds
724
,
67
(
2017
).
11.
T.
Basu
,
K.
Singh
,
S.
Gohil
,
S.
Ghosh
, and
E. V.
Sampathkumaran
,
J. Appl. Phys.
118
,
234103
(
2015
).
12.
M.
Greenblatt
,
R. M.
Hornreich
, and
B.
Sharon
,
J Solid State Chemistry
10
,
371
(
1974
).
13.
M.
Das
,
S.
Roy
,
K.
Mahalingam
,
V.
Ganesan
, and
P.
Mandal
,
J. Phys.: Condens. Matter
32
,
035802
(
2020
).
14.
B.
Bharati
,
P.
Mohanty
,
C. J.
Sheppard
, and
A. R. E.
Prinsloo
,
J. Magn. Magn. Mater.
546
,
168862
(
2022
).
15.
D. A.
Kumar
and
M.
Yusuf
,
J. Appl. Phys.
127
,
213903
(
2020
).
16.
B.
Bharati
,
A. K.
Sonkar
,
N.
Singh
,
D.
Dash
, and
C.
Rath
,
Mater. Res. Express
4
,
085503
(
2017
).
17.
B.
Bharati
,
N. C.
Mishra
,
A. S. K.
Sinha
, and
C.
Rath
,
Mater. Res. Bulletin
123
,
110710
(
2020
).
18.
B.
Bharati
,
N. C.
Mishra
,
D.
Kanjilal
, and
C.
Rath
,
Applied Surface Science
428
,
723
(
2018
).
19.
B.
Bharati
,
N. C.
Mishra
, and
C.
Rath
,
Applied Surface Science
455
,
717
(
2018
).
20.
P.
Mohanty
,
V. P.
Singh
,
N. C.
Mishra
,
S.
Ojha
,
D.
Kanjilal
, and
C.
Rath
,
J. Phys. D: Appl. Phys.
47
,
315001
(
2014
).
21.
B.
Bharati
and
C.
Rath
,
AIP Advances
11
,
035110
(
2021
).
22.
P.
Mohanty
,
D.
Kabiraj
,
R. K.
Mandal
,
P. K.
Kulriya
,
A. S. K.
Sinha
, and
C.
Rath
,
J. Magn. Magn. Mater.
355
,
240
(
2014
).
23.
H. W.
Zheng
,
Y. F.
Liu
,
W. Y.
Zhang
,
S. J.
Liu
,
H. R.
Zhang
, and
K. F.
Wang
,
J. Appl. Phys.
107
,
053901
(
2010
).
24.
J.
Rodríguez-Carvajal
,
Physica B
192
,
55
(
1993
).
25.
P. M.
Diehm
,
P.
Ágoston
, and
K.
Albe
,
Chem Phys Chem
13
,
2443
(
2012
).
26.
D.
Prieur
,
W.
Bonani
,
K.
Popa
,
O.
Walter
,
K. W.
Kriegsman
,
M. H.
Engelhard
,
X.
Guo
,
R.
Eloirdi
,
T.
Gouder
,
A.
Beck
,
T.
Vitova
,
A. C.
Scheinost
,
K.
Kvashnina
, and
P.
Martin
,
Inorg. Chem.
59
,
5760
(
2020
).
27.
S.
Sen
,
S. K.
Halder
, and
S. P. S.
Gupta
,
J. Phys. D: Appl. Phys.
6
,
1978
(
1973
).
28.
L.
Kumar
,
P.
Kumar
,
A.
Narayan
, and
M.
Kar
,
International Nano Letters
3
,
8
(
2013
).
29.
J.
Shi
,
T.
Sauyet
,
Y.
Dang
,
S. L.
Suib
,
M. S.
Seehra
, and
M.
Jain
,
Condens. Matter.
33
,
205801
(
2021
).
30.
A.
Lakhani
,
A.
Banerjee
,
P.
Chaddah
,
X.
Chen
, and
R. V.
Ramanujan
,
J. Phys.: Condens. Matter
24
,
386004
(
2012
).
31.
A. K.
Pramanik
and
A.
Banerjee
,
Phys. Rev. B
79
,
214426
(
2009
).
32.
A.
Lakhani
,
S.
Dash
,
A.
Banerjee
,
P.
Chaddah
,
X.
Chen
, and
R. V.
Ramanujan
,
Appl. Phys. Letters
99
,
242503
(
2011
).
33.
A. K.
Nayak
,
M.
Nicklas
,
C.
Shekhar
, and
C.
Felser
,
J. Appl. Phys.
113
,
17E308
(
2013
).
34.
V.
Siruguri
,
P. D.
Babu
,
S. D.
Kaushik
,
A.
Biswas
,
S. K.
Sarkar
,
M.
Krishnan
, and
P.
Chaddah
,
J. Phys.: Condens. Matter
25
,
496011
(
2013
).
35.
S. B.
Roy
and
M. K.
Chattopadhyay
,
Phys. Rev. B
79
,
052407
(
2009
).
36.
X. F.
Miao
,
Y.
Mitsui
,
A. I.
Dugulan
,
L.
Caron
,
N. V.
Thang
,
P.
Manuel
,
K.
Koyama
,
K.
Takahashi
,
N. H.
vanDijk
, and
E.
Bruck
,
Phys. Rev. B
94
,
094426
(
2016
).
37.
A.
Banerjee
,
P.
Chaddah
,
S.
Dash
,
K.
Kumar
, and
A.
Lakhani
,
Phys. Rev. B
84
,
214420
(
2011
).
38.
X.-L.
Qian
,
J.
Kang
,
B.
Lu
,
S.-X.
Cao
, and
J.-C.
Zhang
,
RSC Adv
6
,
10677
(
2016
).
39.
K.
Gautam
,
A.
Ahad
,
S. S.
Majid
,
A.
Sagdeo
,
S.
Francoual
,
R. J.
Choudhary
, and
D. K.
Shukla
,
J. Magn. Magn. Mater.
478
,
260
(
2019
).
40.
P.
Kollu
,
S.
Prathapani
,
E. K.
Varaprasadarao
,
C.
Santosh
,
S.
Mallick
,
A. N.
Grace
, and
D.
Bahadur
,
Appl. Phys. Letter
105
,
052412
(
2014
).
41.
S.
Gupta
and
K. G.
Suresh
,
Materials Letters
113
,
195
(
2013
).