Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Physics: Condensed Matter

J. Phys.: Condens. Matter 32 (2020) 035810 (10pp) https://doi.org/10.1088/1361-648X/ab4845

Structural, optical and magnetic properties


of Sm3+ doped yttrium orthoferrite (YFeO3)
obtained by sol–gel synthesis route
P S J Bharadwaj1, Swarup Kundu1, Vijay Sai Kollipara1
and Kalidindi B R Varma1,2
1
  Department of Physics, Sri Sathya Sai Institute of Higher Learning, Vidyagiri, Prasanthi Nilayam,
Andhra Pradesh, India
2
  Materials Research Centre, Indian Institute of Science, Bengaluru, India

E-mail: kvijaysai@sssihl.edu.in

Received 19 July 2019, revised 13 September 2019


Accepted for publication 26 September 2019
Published 23 October 2019

Abstract
Fine powders of Y1−xSmxFeO3 (x  =  0, 0.05, 0.10, 0.15) were synthesized via citrate based
sol–gel route. While the as synthesized powders were amorphous, the calcined (900 °C/8 h)
powders were confirmed to be polycrystalline by x-ray powder diffraction (XRD) studies. The
calcined powders were found to crystallize in an orthorhombic structure associated with the
lattice parameters a  =  5.59 Å, b  =  7.60 Å, c  =  5.28 Å. These lattice parameters increased with
the increase in Sm3+ content at yttrium sites. The strain that was obtained by the Williamson–
Hall method increased with the increase in dopant (Sm3+) concentration vis-à-vis a decrease in
crystallite size. Diffuse reflectance spectroscopic studies suggest an increase in band gap as Sm
doping level increased. Significant enhancement in magnetization associated with a decrease
in coercive field accompanied by a transition from anti-ferromagnetic to soft ferromagnetic
behaviour in Sm doped YFeO3 were encountered. It is hoped that these materials with the
enhanced magnetic properties could be of potential use for multifarious applications.

Keywords: rare-earth orthoferrites, multiferroics, photoluminescence, superexchange


interaction

(Some figures may appear in colour only in the online journal)

Introduction canting of Fe3+ moments, as Y3+ being diamagnetic, cannot


contribute to the magnetization [6]. Further, the magnetic
Multifunctional rare-earth orthoferrites (RFeO3, where properties of the material are characterized by the extent of
R  =  Nd3+, Gd3+, Sm3+ etc) were extensively studied, pri- the octahedral tilting of FeO6 [7, 8]. YFeO3 is reported to pos-
marily owing to a variety of physical properties that they sess moderately high resistivity (109 Ω m) and also, a very
exhibit. These qualify orthoferrites for multifarious applica- high Neel temperature (TN  =  645 K) [9, 10]. YFeO3 possess
tions that include capacitors, memory devices, magneto-elec- an orthorhombic structure associated with slight distortion,
tric sensors in transducers, spintronic devices and microwave where the FeO6 octahedron is tilted towards Y3+ ion at the
electronic devices [1–4]. centre while maintaining the Y3+–O2− bonding. The unit
YFeO3 has a perovskite structure belonging to the generic lattice structure of YFeO3 is depicted in figure  1 employing
formula ABO3. The Y3+ cations (A-site) occupy the body VESTA software.
centre which co-ordinate with 12 oxygen anions. Fe3+ cat- Earlier studies demonstrate that the introduction of magn­
ions (B-site), occupy the corner positions of the cube and co- etic or non-magnetic ions can effectively tweak the structural
ordinate with the six oxygens, as an octahedron [5]. YFeO3 and physical properties of rare-earth orthoferrites [11–13]. To
exhibits its net magnetization, originating entirely from the the best of our knowledge, sufficient attention was not paid to

1361-648X/ 20 /035810+10$33.00 1 © 2019 IOP Publishing Ltd  Printed in the UK


J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

Figure 1.  Crystal structure of YFeO3.

the magnetic-ion substitution at Y-site in YFeO3. It was felt


that the introduction of magnetic Sm3+ ion at diamagnetic Y3+
site in YFeO3 may cause crystal structure distortion as Sm3+
ionic radius is slightly larger than that of Y3+ and in turn influ-
ence its physical properties.
The synthesis of YFeO3 was undertaken using several
routes including solid-state, hydrothermal, microwave, polyol
and sol–gel citrate methods [7, 14–17]. The wet chemical syn-
thesis enables the control of the stoichiometry, particle size
and homogeneity, albeit with the disadvantages of long reac- Figure 2.  Schematic for the synthesis of Y1−xSmxFeO3 (X  =  0,
tion time and expensive solvent medium. 0.05, 0.10, 0.15) polycrystalline powders by citrate based sol–gel
In this article, we present our findings pertaining to the synthesis route.
effect of Samarium (Sm3+) doping on the structural, optical Y1−xSmxFeO3 (X  =  0, 0.05, 0.10, 0.15). The process of this
and magnetic properties of an antiferromagnetic polycrystal- citrate based sol–gel synthesis of these phases is schemati-
line YFeO3. cally illustrated in figure 2.

Materials and methods Characterization

Sol-gel technique was used for synthesizing fine powders of The crystallographic information for all the calcined samples
Y1−xSmxFeO3(X  =  0, 0.05, 0.10, 0.15). For this, the precursors was obtained using x-ray powder Diffractometer (X-pert
used were Y2O3 (AR grade, purity  >  99%), Fe(NO3)3·9H2O Pro, Panalytical) with Cu-Kα  =  1.5408 Å. The average
(AR grade, purity   >  98%) and Sm2O3 (Sigma-Aldrich, crystallite size and lattice strain as a function of dopant
purity  >  99%). concentration were calculated using Scherrer formula and
Y2O3 and Fe(NO3)3·9H2O were dissolved in de-ionized Williamson–Hall plot respectively. The stability of the phase
water as per the stoichiometric requirements. Sm2O3 was was confirmed by invoking Goldschmidt tolerance factor.
separately dissolved in dilute nitric acid. These two separate The Raman spectra were recorded for the powder samples
solutions were mixed under vigorous stirring using a magn­ using Raman microscope, Thermo Scientific (laser excitation
etic stirrer. Citric acid which acts as chelating agent was source λ  =  780 nm). The FTIR studies were carried out on
added in the molar ratio of 1:1 with respect to the inorganics powder samples of all the compositions under study using
to enhance the complex formation of metal ions and to pro- FTIR spectrometer (Cary 630, Agilent).The microstructure
mote the polymerization of the mixture. After the addition of associated with elemental analysis studies were carried out
citric acid, the solution was continuously stirred for 30 min, using scanning electron microscope (SEM), Jeol IT300. The
to which ammonia solution (25%) was added drop-wise to optical studies in the wavelength range of 200–800 nm were
maintain the optimum pH level of 7–8. This solution was vig- done using UV–vis (Perkin-Elmer) spectrophotometer. The
orously stirred at room temperature for 3–4 h. Subsequently, it photoluminescence studies were carried out using photolumi-
was heated for 12 h at 120 °C to obtain the gel. Thus formed nescence spectrometer (Perkin Elmer LS 55). VSM (Model
gel was heated to 280 °C, which caused combustion of excess EZ9, MicroSense, USA) was employed to study the magn­
citrate to yield brown residue. This was finely ground and cal- etic properties at room temperature for all the synthesized
cined at 900 °C for 8 h to obtain polycrystalline powders of powders.

2
J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

Figure 3.  (a) XRD Patterns recorded for Y1−xSmxFeO3 (X  =  0, 0.05, 0.10, 0.15). (b) Shift in the peak positions.

Table 1.  Refinement and Lattice parameters Y1−xSmxFeO3(X  =  0, 0.05, 0.10, 0.15).

Y1−xSmxFeO3 Rwp (%) RP (%) GOF(χ2) a (Å) b (Å) c (Å)

X  =  0.00 5.64 4.43 1.20 5.588 7.600 5.278


X  =  0.05 4.93 3.83 1.05 5.593 7.607 5.285
X  =  0.10 5.12 4.00 1.03 5.595 7.613 5.291
X  =  0.15 5.26 4.11 1.01 5.596 7.618 5.297

Results and discussion Table 2.  Goldschmidt tolerance factor, crystallite size, lattice strain
as function of Y1−xSmxFeO3 (X  =  0, 0.05, 0.10, 0.15).
The x-ray diffractograms obtained in the 2θ range of 20°–60° Tolerance Crystallite Lattice
for all the samples calcined at 900 °C/8 h, confirmed to be Y1−xSmxFeO3 factor size (nm) strain
monophasic (figure 3) with no detectable impurity phases.
X  =  0 0.8516 176.4 0.183
However, there was a considerable shift of x-ray peak posi-
tions towards lower angles (at 2θ  =  33.2°) on increasing X  =  0.05 0.8529 133.7 0.210
the Sm3+ content in YFeO3 implying an increase in lattice X  =  0.10 0.8565 120.8 0.460
parameters. X  =  0.15 0.8560 106.3 0.488
The lattice parameter data along with refinement are illus-
trated in table 1. The increase in lattice parameters a, b and c
is ascribed to the larger ionic radius of Sm3+ (1.24 Å) than that Kλ
L= .
of Y3+ (1.06 Å) [18]. βCosθ
The x-ray diffraction peaks of all the compositions in the The stability of orthorhombic phase was calculated using
present investigation could be indexed to the bulk YFeO3 Goldschmidth tolerance factor [19],
(ICSD 98-008-0866) associated with orthorhombic crystal
Ra + Rb
structure with ‘Pmna’ space group of the point group ‘mmm’. T=√ .
Scherrer formula given below was used to calculate the 2(Rb + RO )
average crystallite size. Here L is average crystallite size, K is Where Ra is radius of A-site ion, Rb is radius of B-site ion, and
shape factor, λ is wavelength of the x-ray used, β is the value Ro is radius of O in ABO3 perovskite structure. The typical
of full width at half maximum (FWHM) and θ is the Bragg value for stable orthorhombic phase is 0.8 to 1 [20]. We have
angle. found the Goldschmidth tolerance factor for Y1−XSmxFeO3 to

3
J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

Figure 4.  Rietveld refined XRD Patterns for Y1−xSmxFeO3 (a) x  =  0, (b) 0.05, (c) 0.10, (d) 0.15.

be around 0.85 for different concentrations of Sm there by residual factor without considering the background and GOF
confirming the stability of the orthorhombic phase for dif- (Goodness of Fit, χ2) which is the measure of (Rwp)2/(Rp)2
ferent levels of doping. The crystallite sizes along with lattice were found to be low, which suggest the goodness of refine-
strain and tolerance factors obtained for the samples ranging ment. The corresponding refined lattice parameters are sum-
from x  =  0–0.15 are depicted in table 2. The average crystal- marised in table 1. There is an increase in lattice parameters
lite size decreases with the increase in dopant concentration with the increase in dopant concentration with a concomitant
while strain increases. increase in strain.
The lattice strain was calculated using the Williamson– We have calculated octahedral tilts of FeO6 and Fe–O–Fe
Hall plot and the observations are summarized in table 2. An bond angles, as the magnetic properties of orthoferrites are
increase in lattice strain was observed with the increase of strongly dependent on these parameters. There were two
dopant (Sm3+) concentration at Y3+ site in YFeO3. bond angles θ1 and θ2 owing to two kinds of Fe–O–Fe super
The lattice parameters were obtained by rietveld refinement exchange bonds. The formation of these two super exchange
using High score plus software. The refined plots are captured bonds are attributed to the sharing of O2− ions between two
in figure 4 depicting the experimental data as open circles and octahedral structures of FeO6 in YFeO3 orthorhombic crystal
calculated intensities as solid lines. The vertical lines at the structure.
bottom are the allowed Bragg positions for Pmna space group. O’Keefe geometrical approximations [21] were used to
The solid line in figure  4 represents the difference between determine the extent of distortion of the FeO6 octahedra in
measured and calculated intensities. In the refinement pro- terms of bond and octahedral tilt angles;
cess, we have taken oxygen positions as free parameters. The  
 2 − 5cos2 ϕ1 
other atomic positions were considered to be fixed. The other θ1 = cos−1  
parameters like lattice constants, shape parameters, scale fac- 2 + cos2 ϕ1 
tors, isothermal parameters and occupancies were taken to
 
be fixed. The pseudo-voigt function has been used to correct  1 − 4cos2 ϕ2 
θ2 = cos −1  .
the background. The values of Rwp, weighted profile residual  3 
factor which takes background into consideration, Rp, profile

4
J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

Table 3.  Bond angles Fe–O–Fe (θ1, θ2), Bond lengths (Fe–O) and octahedral tilt angles (ϕ1, ϕ2) of Y1−xSmxFeO3.

Bond angles Tilt angles


Bond length (Å)
Y1−xSmxFeO3 θ1 θ2 ϕ1 ϕ2 (Fe–O)

x  =  0 151.450 137.393 17.307 26.426 1.926


x  =  0.05 144.063 143.637 20.088 22.467 2.045
x  =  0.10 144.0625 143.624 21.663 22.474 2.046
x  =  0.15 144.0605 143.615 21.664 22.480 2.047

Figure 5.  Raman spectra for Y1−xSmxFeO3 (x  =  0, 0.05, 0.10, 0.15) powders.

Where θ1, θ2 and ϕ1, ϕ2 are bond angles and octahedral tilt
angles respectively.
In order to obtain these bond angles and octahedral tilt
angles we have used crystallographic data obtained from
reitveld refinement and represented in VESTA software. The
values of these bond angles and tilt angles for Y1−xSmxFeO3
are elucidated in table 3 which suggest that there is a change
of tilt angles which cause the distortion in crystal lattice with
increase in Sm3+ doping concentration.
The Raman spectra for Y1−xSmxFeO3 (x  =  0, 0.05, 0.10,
0.15) is presented in the figure 5.The modes assigned to the
samples matches quite well with those reported previously
[22]. The peaks at 149 cm−1 and 180 cm−1 are attributed to
the vibrations of yttrium ion. The peak around 220 cm−1 is
associated with Fe3+ ion vibration. The other relatively low
intense peaks around 281, 340, 430, 498 cm−1 are due to the
excitation of the Fe3+ magnetic ions within the realm of octa­
hedral oxygen [23]. The small systematic shift towards lower
wave numbers, observed with the increase in doping concen- Figure 6.  FT-IR spectra for Y1−xSmxFeO3 (x  =  0, 0.05, 0.10, 0.15)
tration of Sm is rationalized using the relation ω ~ (k/M)1/2, powders.
where ω is the frequency, k is the spring constant and M is The microstructural features of all the calcined samples
atomic mass [24]. under investigation were studied using SEM and the data are
The IR spectra of Y1−xSmxFeO3 shown in figure 6 reveal depicted in figure 7. It is observed that there is a reduction in
the characteristic peaks of YFeO3. The peaks assigned around the crystallite size with increasing dopant (Sm3+) concentra-
3435 cm−1 and 1635 cm−1 are assigned to ν (O–H) and δ tion. These results are consistent with those obtained by x-ray
(H–O–H) respectively [25]. The strong peaks located around studies (Scherrer analysis). The crystallite size distributions
575 cm−1 and 485 cm−1 are ascribed to Fe–O stretching vibra- are shown as insets in each micrograph. This decrease in crys-
tion modes in the orthoferrite structure [26]. tallite size by doping samarium(Sm) at A-site is ascribed to

5
J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

Figure 7.  SEM images of Y1−xSmxFeO3 (X  =  0, 0.05, 0.10, 0.15), the histograms in the insets of each micrograph are the corresponding
crystallite size distributions.

Figure 8.  EDS spectra for Y1−xSmxFeO3 (X  =  0, 0.05, 0.10, 0.15).

6
J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

Figure 9.  (a) Absorption spectra for Y1−xSmxFeO3(x  =  0, 0.05, 0.10, 0.15). (b) Tauc plot for Y1−xSmxFeO3 (x  =  0, 0.05, 0.10, 0.15).

Table 4.  Optical bandgap for differernt Sm3+ contents in YFeO3.

Y1−xSmxFeO3 Band gap (eV)

x  =  0 2.1121
x  =  0.05 2.1328
x  =  0.10 2.1542
x  =  0.15 2.1553

the ionic radii difference between the dopant (Sm3+) and the
host (Y3+) which inhibits the grain boundary mobility due to
the lattice distortion and consequently impede nucleation and
hinder propagation of nucleation and the growth of crystal-
lites. The observed peaks in the EDS (figure 8) spectra are
assigned to Y, Fe, O, and Sm. The mole ratios of Y, Sm, Fe,
O are consistent with the expected stoichiometry within the
experimental errors.
Figure 10.  PL spectra for Y1−xSmxFeO3 (x  =  0.00, 0.05, 0.10, 0.
15) samples.
Optical band gap studies
(0 0 2)) suggesting an increase in lattice spacing on increasing
Diffuse reflectance studies carried out for all the synthesized Sm3+ content in YFeO3 lattice. Yet another reason for an
samples in the wavelength range of 200–800 nm were used enhancement in the bandgap for the present compound could
to detemine the band gaps associated with the samples under be due to the presence of microstrain [28] originating from
study. We have observed a strong transition in the 500–600 nm slight difference that exists between the ionic sizes of Sm3+
range in the absorbance plots, which correspond to electronic and Y3+ apart from the defect states arising out of prevalence
transitions from 2p valence states of oxygen (O) to 3d states of of oxygen vacancies. These oxygen vacancies lower the adja-
iron (Fe) conduction band. There is a downward shift (towards cent Fe 3d levels, which result in sub-bandgap defect states.
lower wavelength) in the absorption edge with an increase in The sub-band gap states are likely to lower the overall energy
Sm3+ in YFeO3(figure 9(a)) implying that there is an increase band gap [29]. Further, the increase in bandgap is attributed to
in bandgap. The bandgap Eg for Y1−xSmxFeO3(x  =  0, 0.05, the change in bond length (Fe–O) and bond angle (Fe–O–Fe)
0.10, 0.15) was calculated using the Tauc-plot (Kubelka– with Sm3+ doping as these are related to the bandgap (Eg);
Monk versu Photon energy) and the results are illustrated
in figure  9(b). The bandgap for Y1−xSmxFeO3 (x  =  0) was Eg = ∆ − W
calculated to be 2.11 eV and is consistent with that reported where Δ is charge transfer energy and W is one-electron
in the literature [27]. The bandgap values for Y1−xSmxFeO3 band width which is given by
(x  =  0, 0.05, 0.10, 0.15) lie in the range of 2.11 eV to 2.15 eV cos ω
as depicted in table 4. This increase in bandgap with increase W ∝ 3.5
dFe–O
in Sm content may be attributed to more than one mechanism.
There is an expansion in the lattice vis-à-vis a decrease in where ω is 12 [π − (Fe–O–Fe)] and dFe–O
3.5
is Fe–O bond length.
crystallite size as evidenced by our XRD (figure 3) followed As the Fe–O bond length is found to increase on increasing
by SEM (figure 7) studies could be one of the reasons. This is Sm3+ content (table 3), one would expect a decrease in W and
substantiated by a downward shift in the x-ray peaks ((1 2 1), hence an increase in Eg [30].

7
J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

Figure 11. M–H loop of Y1−xSmxFeO3 (a) x  =  0.00, (b) 0.05, (c) 0.10, (d) 0.15.

Photoluminiscence (PL) studies orbitals and Fe3+ ions have partially filled orbitals. Therefore,
the magnetic properties arise only from Fe3+–Fe3+ interac-
The emission spectra of Y1−xSmxFeO3 (x  =  0, 0.05, 0.10,
tions. The super exchange interaction is an oxygen mediated
0.15) polycrystalline samples have a broad peak around
exchange between transition metal ions based on virtual hop-
600 nm when excited at 450 nm, corresponding to photon
ping processes of the oxygen 2p electrons. The O2− ions have
energy of 2.7 eV are depicted in figure 10. The emission lines
fully occupied triple degenerate (2px, 2py, 2pz) orbitals with six
around 600nm are attributed to yttrium. The oxygen vacan-
electrons. In Fe3+, the 5d orbitals are not degenerate and they
cies and defects on the surface of Y1−xSmxFeO3 polycrystal-
split as triple degenerate t2g (dxy, dyz, dxz) and doubly degen-
line powders make the electrons to bind to form excitons.
erate eg (dx2 −y2 , dz2 ). These eg orbitals which point along the
The formation of excitons leads to the introduction of exciton
crystal axes, overlap with the 2p orbitals of oxygen leading to
energy level near the bottom of conduction band. Therefore,
the super exchange interaction of Fe3+–Fe3+ through O2− ions
higher the oxygen vacancies and defects, stronger will be the
at 180 °C as predicted by the GKA (Goodenough, Kanamori
PL spectrum [31]. The reduction in intensity of the emission
and Anderson) rules which give the super exchange interac-
peaks at 600 nm was observed with an increase in the dopant
tion [33] and the anti-ferromagnetic nature of YFeO3. We have
(Sm3+) concentration(x  =  0, 0.05, 0.10, 0.15). This can be
observed that at room temperature, there was an enhancement
ascribed to the decrease of surface oxygen vacancies and
in magnetization with the increase in doping concentration of
defects with increase in Sm3+ dopant concentration.
Sm, owing to magnetic contribution of Sm which is stronger
than super exchange of Fe–O–Fe as shown in figure  12.
However, the coercive field was noticed to decrease drasti-
Magnetic properties
cally even with the x  =  0.05 (figure 13), which is attributed
The origin of magnetic properties in rare-earth ortho-ferrites to the influence of Sm3+–Fe3+ interaction on magnetic aniso­
is due to super-exchange interaction of Fe3+–Fe3+, R3+–R3+, tropy [34].
R3+–Fe3+ through O2− ion [32]. The M–H hysteresis loop The linear increase in magnetization was observed with
shown in figure 11(a) suggests the antiferromagnetic behav- the increase in doping concentration as shown in figure  13.
iour of YFeO3. The magnetic interactions between Y3+ and One possibility of reduction in antiferromagnetic nature is due
Fe3+ are absent as Y3+ is diamagnetic as it has fully filled to the unstable hybridization of Y3+ and O2− and Fe3+ and

8
J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

transfer, magnetic field screening, electric motors, electro-


magnetic pole-pieces etc [38].

Conclusions

Samarium doped yttrium orthoferrite Y1−xSmxFeO3 (x  =  0,


x  =  0.05, x  =  0.10, x  =  0.15) powders were successfully syn-
thesized using the sol–gel technique. The synthesized ceramic
powders were demonstrated to possess orthorhombic crystal
structure. The value obtained for Goldschmidt tolerance factor
confirms the phase stability. A slight increase in optical band
gap was observed with an increase in dopant (Sm3+) concen-
tration. The obtained magnetization versus magnetic field
hysteresis loops at room temperature stipulate that there is
a transition from antiferromagnetic to a soft ferromagnetic
behaviour in Sm doped YFeO3 polycrystalline powders.

Acknowledgments
Figure 12.  Comparison of M–H loops of Y1−xSmxFeO3 (x  =  0.00,
0.05, 0.10, 0.15). The Authors dedicate the work to the founder chancellor
Bhagawan Sri Sathya Sai Baba, and thank the Sri Sathya Sai
Central Trust for providing characterization facilities at Cen-
tral Research Instruments Facility (SSSIHL-CRIF).

ORCID iDs

Vijay Sai Kollipara https://orcid.org/0000-0003-3674-0011

References

[1] Schmid H 2008 Some symmetry aspects of ferroics and single


phase multiferroics J. Phys.: Condens. Matter 20 434201
[2] Nan C W, Bichurin M I, Dong S, Viehland D and Srinivasan G
2008 Multiferroic magnetoelectric composites: historical
perspective, status, and future directions J. Appl. Phys.
Figure 13.  Magnetization and Coercive field as function of Sm 103 031101
doping concentration. [3] Mitoseriu L 2005 Magnetoelectric phenomena in single-phase
and composite systems Bol. Soc. Esp. Ceram. V 44 177–84
[4] Spaldin N A 2017 Multiferroics: past, present, and future MRS
Bull. 42 385–9
O2− as compared to more stable Fe3+–O2−–Fe3+ hybridiza- [5] King G and Woodward P M 2010 Cation ordering in
tion being formed. The other possibility could be due to the perovskites J. Mater. Chem. 20 5785–96
distortion in crystal structure. With the introduction of Sm [6] Yuan X, Sun Y and Xu M 2012 Effect of Gd substitution on
in the crystal lattice, distortion takes place as evident from the structure and magnetic properties of YFeO3 ceramics
an increase in strain (table 2). This distortion consequently J. Solid State Chem. 196 362–6
[7] Zhou Z, Guo L, Yang H, Liu Q and Ye F 2014 Hydrothermal
leads to a decrease in Fe3+–O2−–Fe3+ bond angle as evi- synthesis and magnetic properties of multiferroic rare-earth
denced from O’Keefe approximation and therefore a reduc- orthoferrites J. Alloys Compd. 583 21–31
tion in magnetic super exchange effect. Further, this results in [8] Lee J H, Jeong Y K, Park J H, Oak M A, Jang H M, Son J Y
exchange interaction between Fe3+–O2−–Sm3+ which signifi- and Scott J F 2011 Spin-canting-induced improper
cantly influences the spin re-orientation of Fe3+ ions leading ferroelectricity and spontaneous magnetization reversal in
SmFeO3 Phys. Rev. Lett. 107 1–5
to weak-ferromagnetic behaviour. In addition, owing to the [9] Shang M, Zhang C, Zhang T, Yuan L, Ge L, Yuan H and
Dzyaloshinsky–Moriya anti-symmetric exchange, the magn­ Feng S 2013 The multiferroic perovskite YFeO3 Appl. Phys.
etic moments are not completely anti-parallel and they are Lett. 102 1–4
slightly canted with a small angle, resulting in a weak fer- [10] Zhang R L, Chen C Le, Jin K X, Niu L W, Xing H and
romagnetic behaviour [35–37]. Though it is weak, these mat­ Luo B C 2014 Dielectric behavior of hexagonal and
orthorhombic YFeO3 prepared by modified sol–gel method
erials may be exploited for various applications that include J. Electroceram. 32 187–91
magnetic shielding, power supply transformers, DC–DC [11] Dahmani A, Taibi M, Nogues M, Aride J, Loudghiri E and
converters, electrical to mechanical power converters, signal Belayachi A 2003 Magnetic properties of the perovskite

9
J. Phys.: Condens. Matter 32 (2020) 035810 P S J Bharadwaj et al

compounds YFe1  −  xCrxO3 (0.5  ⩽  x  ⩽  1) Mater. Chem. Alvarez-Bada J R 2018 Mechanochemical synthesis of


Phys. 77 912–7 YFeO3 nanoparticles: optical and electrical properties of
[12] Ma Y, Wu Y J, Lin Y Q and Chen X M 2010 Microstructures thin films J. Clust. Sci. 29 225–33
and multiferroic properties of YFe1  −  xMnxO3 ceramics [26] Zhang W, Fang C, Yin W and Zeng Y 2013 One-step synthesis
prepared by spark plasma sintering J. Mater. Sci. Mater. of yttrium orthoferrite nanocrystals via sol–gel auto-
Electron. 21 838–43 combustion and their structural and magnetic characteristics
[13] Madolappa S, Ponraj B, Bhimireddi R and Varma K B R 2017 Mater. Chem. Phys. 137 877–83
Enhanced magnetic and dielectric properties of Ti-doped [27] Díez-García M I, Celorrio V, Calvillo L, Tiwari D, Gómez R
YFeO3 ceramics J. Am. Ceram. Soc. 100 2641–50 and Fermín D J 2017 YFeO3 Photocathodes for Hydrogen
[14] Carotta M C, Martinelli G, Sadaoka Y, Nunziante P and Evolution Electrochim. Acta 246 365–71
Traversa E 2002 Gas-sensitive electrical properties of [28] Mocherla P S V, Karthik C, Ubic R, Rao M S R and
perovskite-type SmFeO3 thick films Sensors Actuators B Sudakar C 2013 Tunable bandgap in BiFeO3 nanoparticles:
48 270–6 the role of microstrain and oxygen defects Appl. Phys. Lett.
[15] Ding J, Lü X, Shu H, Xie J and Zhang H 2010 Microwave- 103 022910
assisted synthesis of perovskite ReFeO3 (Re: La, Sm, Eu, [29] Hauser A J, Zhang J, Mier L, Ricciardo R A, Woodward P M,
Gd) photocatalyst Mater. Sci. Eng. B 171 31–4 Gustafson T L, Brillson L J and Yang F Y 2008
[16] Niu X, Li H and Liu G 2005 Preparation, characterization and Characterization of electronic structure and defect states
photocatalytic properties of REFeO3 (RE  =  Sm, Eu, Gd) of thin epitaxial films by UV-visible absorption and
J. Mol. Catal. A 232 89–93 cathodoluminescence spectroscopies Appl. Phys. Lett.
[17] Siemons M, Weirich T, Mayer J and Simon U 2004 92 222901
Preparation of nanosized perovskite-type oxides via polyol [30] Hasan M, Basith M A, Zubair M A, Hossain S, Mahbub R,
method Z. Anorg. Allg. Chem. 630 2083–9 Hakim M A and Islam F 2016 Saturation magnetization and
[18] Shannon R D 1976 Revised effective ionic radii in halides and band gap tuning in BiFeO3 nanoparticles via co-substitution
chalcogenides Acta Crystallogr. A 32 751 of Gd and Mn J. Alloys Compd. 687 701–06
[19] Wang M, Wang T, Song S and Tan M 2017 Structure- [31] Jing L, Xin B, Yuan F, Xue L, Wang B and Fu H 2006
controllable synthesis of multiferroic YFeO3 nanopowders Effects of surface oxygen vacancies on photophysical and
and their optical and magnetic properties Materials 10 626 photochemical processes of Zn-doped TiO2 nanoparticles
[20] Kosuke N and Kauzuhiko K 2002 Metastable phase formation and their relationships J. Phys. Chem. B 110 17860–5
from an undercooled rare-earth orthoferrite melt J. Am. [32] Jaiswal A, Das R, Adyanthaya S and Poddar P 2011 Surface
Ceram. Soc. 56 2550–6 effects on morin transition, exchange bias, and enchanced
[21] O’Keeffe M and Hyde B G 1977 Some Structures spin reorientation in chemically synthesized DyFeO3
topologically related to cubic perovskites(E21),ReO3(DO9) nanoparticles J. Phys. Chem. C 115 2954–60
and Cu3Au (L12) Acta.Cryst. B33 3802–13 [33] Anderson P W 1959 New approach to the theory of
[22] Venugopalan S, Dutta M, Ramdas A K and Remeika J P superexchange interactions Phys. Rev. 115 2–13
1985 Magnetic and vibrational excitations in rare-earth [34] Chen L, Li T, Cao S, Yuan S, Hong F and Zhang J 2012
orthoferrites: a Raman scattering study Phys. Rev. B The role of 4f-electron on spin reorientation transition of
31 1490–7 NdFeO3: a first principle study J. Appl. Phys. 111 103905
[23] Coutinho P V, Cunha F and Barrozo P 2017 Structural, [35] Moriya T 1960 Anisotropic superexchange interaction and
vibrational and magnetic properties of the orthoferrites weak ferromagnetism Phys. Rev. 120 91–8
LaFeO3 and YFeO3 : a comparative study Solid State [36] Dzyaloshinsky I 1958 A thermodynamic theory of ‘weak’
Commun. 252 59–63 ferromagnetism of antiferromagnetics J. Phys. Chem.
[24] Stojadinovic B, Dohcevic-Mitrovic Z, Paunovic N, Ilic N, Solids. 4 241–55
Tasic N, Petronijevic I, Popovic D and Stojanovic B 2015 [37] Yuvaraj S, Layek S, Vidyavathy S M, Yuvaraj S, Meyrick D
Comparative study of structural and electrical properties of and Selvan R K 2015 Electrical and magnetic properties
Pr and Ce doped BiFeO3 ceramics synthesized by auto- of spherical SmFeO3 synthesized by aspartic acid assisted
combustion method J. Alloys Compd. 657 866–72 combustion method Mater. Res. Bull. 72 77–82
[25] Vázquez-Olmos A R, Sánchez-Vergara M E, Fernández- [38] Goldman A 1999 Handbook of Modern Ferromagnetic
Osorio A L, Hernández-García A, Sato-Berrú R Y and Materials (London: Kluwer Academic)

10

You might also like