1 s2.0 S092145262200878X Main
1 s2.0 S092145262200878X Main
1 s2.0 S092145262200878X Main
A R T I C L E I N F O A B S T R A C T
Keywords: The various physical properties with their intercorrelations in polycrystalline samples of Sm0.5Y0.5
Antiferromagnetism Fe0.58Mn0.42O3 [SYFM (58–42)] were investigated. The dc magnetization measurements revealed a weak
Spin reorientation ferromagnetic (WFM) transition at TÑ361 K that is followed by an incomplete spin reorientation (SR) transition
Magnetoelectricity
at TSR1~ 348 K. A first order magnetic transition (FOMT) around TSR2 ~ 292 K completes the spin reorientation
Spin-phonon coupling
transition and the material enters into a nearly collinear antiferromagnetic (AFM) state. Robust magnetodi
electric (MD) is found to be present in material at room temperature. True ferroelectric transition with TFE = 108
K and having a value of saturation polarization (~0.06 μC/cm2 at 15 K) have been found in the specimen. This
spin phonon coupling (SPC) stabilizes the ferroelectric state and responsible for magnetoelectric coupling (ME).
We argue that wave vector (q) dependence of the spin-pair correlation affects the electronic and magnetodi
electric properties in a similar way.
1. Introduction canting causing a net moment (F) [1,2]. From the symmetry consider
ations and the antiferromagnetic nature of the coupling between the
Temperature induced spin reorientation transitions (SR) is a phe magnetic ions, three types of magnetic structures are allowed for TM
nomenon often encountered in the Rare-earth Orthoferrites (RFeO3) sublattice in orthoferrites, Γ 4, Γ 2 and Γ 1. Below TN the allowed spin
with magnetic R3+ ions [1,2]. In orthoferrites, the interaction between structure Γ4 has the G-type moment directed along c-axis, while the F
two metallic ions sublattices namely Fe3+-Fe3+, R3+-Fe3+ and R3+-R3+ and A components are directed along the b and a axis of the Pnma
gives rise to a series of magnetic transitions. The dominant among these crystal, giving rise to (G z, F y, A x) type structure. The Γ4 configuration
three magnetic interactions is the antiferromagnetic Fe3+ - Fe3+ inter can be rotated such that the net moment aligns along the c axis and
action which drives the transition metal sublattice into a G-type anti major G-type antiferromagnetic moments lines along the b axis, then we
ferromagnetic state below TFe N (650–700 K). However, the iron spin
obtain the Γ 2 configuration (G y, C x, F z). In the Γ 1 configuration there
directions are not completely collinear but are slightly canted with is no net magnetization F along any direction and the major G-type
respect to one another. The spin canting is of two types viz. Hidden antiferromagnetic vector points along the b axis (A z, G x, C y).
canting causing a C-type or A-type antiferromagnetic, and the other overt Apart from the SR transitions, the anisotropic symmetric and anti
* Corresponding author.
E-mail address: [email protected] (S. Raut).
https://doi.org/10.1016/j.physb.2022.414593
Received 5 August 2022; Received in revised form 24 November 2022; Accepted 13 December 2022
Available online 22 December 2022
0921-4526/© 2022 Elsevier B.V. All rights reserved.
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
symmetric exchange interactions between the R3+-Fe3+ ions can give using Cu Kα radiation. Further confirmation have been done by the
rise to the net magnetization reversal phenomena known also as nega synchrotron x-ray diffraction studies measured at the INDUS 2, BL-11,
tive magnetization (NM), within the weak ferromagnetic (WFM) order Raja Ramanna Centre for Advanced Technology (RRCAT), Indore,
at TFe
N , that is not related to usual superconducting state. It rather implies
India. The synchrotron powder diffraction data was analyzed using
that the net moment to be opposite to the applied field [3,4]. Several of Rietveld refinement [16] available FULLPROF software.
the orthoferrites and its doped samples have shown striking properties of The magnetic measurements were performed in a Superconducting
SR transitions, NM, magnetodielectricity (MD) and magnetoelectric Quantum Interference Device (SQUID) magnetometer (Model: MPMS 3,
(ME) coupling, negative thermal expansion (NTE) [3–8] which renders Quantum Design make) installed in UGC DAE CSR, Kalapakkam Node,
them in a wide category of applications including sensors, thermo between 2 and 400 K. The sintered pellets were used for dielectric and
magnetic switches, thermally assisted magnetic random access mem impedance spectroscopic measurements in between 90 and 400 K using
ories, and other multifunctional devices. a homemade insert coupled with a Keysight E4980A LCR-meter oper
Thus, a single phase material exhibiting the abovementioned prop ating at frequency range f = 20 Hz–2 MHz. The complex dielectric
erties are very rare but can be proved to be very useful for making measurements with Magnetic field and at variable temperatures be
multifunctional devices. However, in the technological point of view, it tween 5 and 300 K were performed in an Alpha-A broadband impedance
is plausible that the material become useful if the functional properties analyser from Novo Control using an Oxford Nano systems Integra 9 T
in the material occur close to room temperature or above. Several of magnet-cryostat. The pyroelectric current (Ip) was recorded at a con
these properties exhibited by RFeO3 and their doped systems are found stant temperature sweep rate (5.0 K/min) in a PPMS II system (Quantum
to be exhibited at temperatures very less than liquid N2. Such as the NTE design) using a Keithley electrometer (model 6517 B) and integrated
in RFe0.5Cr0⋅5O3 (R = Yb, Tm) [9] and RFe0.5Cr0⋅5O3 (R = Tb, Tm) [7] with time for obtaining electric polarization (P). A poling field of 5 kV/
have been observed below the SR transitions that are <50 K. Similarly, cm was applied during cooling and short-circuited before the measure
the ferroelectricity evolved by the application of the magnetic field in ment of Ip in the warming mode for the polarization measurement.
DyFeO3 [10] only in the coexisting short range Dy3+ order with the
WFM ordering of the Fe3+ ions. Very few of the orthoferrites such as 3. Results and discussions
SmFeO3 [11] and modified orthoferrites such as YFe0⋅6Mn0⋅4O3 [5,12]
exhibits magnetodielectricity and/ferroelectricity at RT. Recently 3.1. X-ray diffraction studies
ferroelectricity is reported also in HoFeO3 [13] below TFE ≪ TN that is
invoked by a structural phase transition to a polar Pna21 space group Fig. 1(A) displays the satisfactory fit of Rietveld Refinement of the
from the high temperature centrosymmetric space group Pnma. SXRD pattern at room temperature from specimen. It shows that SYFM
These results motivated us to investigate the physical properties of a (58–42) crystalizes with an orthorhombic symmetry Pnma spacegroup,
codoped sample Sm0.5Y0⋅5Fe0⋅58Mn0⋅42O3. The Fe and Mn percentages similar to orthoferrites and Mn doped YFeO3 systems [5,14,17]. Further
have been optimized at 58% and 42% respectively so that we can get the no apparent presence of impurity phases can be observed in the mate
TN and TSR close to each other and also near room temperature as rial. In this crystal structure, as displayed in Fig. 1(B), both the R atoms
observed earlier [5,14], where we seek to find the magnetodielectric occupy randomly the crystallographic site 4c (x, 0.25, z), while the
effects. The half doping of Sm in place of Y is to ensure marked effects of Fe/Mn atoms are distributed randomly at the special Wyckoff position
Sm3+ on the magnetic properties of the system. In our previous studies 4b (0, 0, 0.5), The oxygen atoms are distributed in two inequivalent sites
on the co-doped system Sm0.5Y0⋅5Fe0⋅58Mn0⋅42O3, we reported about the as O1 in crystallographic site 4c (x, 0.25, z) and as O2 in the 8 d site (x, y,
occurrence low T linear variation of the specific heat [Cp (T) vs T] plot z). In the Pnma setting the O1 atom is the apical oxygen while O2 lies in
along with a broad Schottky anomaly which has been attributed to the the basal plane of the MO6 Octahedra. The Y/Sm ions are surrounded by
inhomogeneity of the molecular field in the re-entrant spinglass (SG) oxygen ions in a RO12 dodecahedral configuration.
state that leads to variable splitting of the Kramer’s ground-state dou
blets in Sm 3+ ions for T < 20 K [15]. The linear temperature de 3.2. Dc magnetization studies
pendency of the Cp (T) at low temperatures is observed due to the
re-entrant spinglass like state with TSG = 67.3 K. The dc magnetization measurement of SYFM (58–42), as a function
In this article we report the effect of magnetic phase coexistency on of temperature and under different external fields, have been performed
various physical properties of Sm0.5Y0⋅5Fe0⋅58Mn0⋅42O3, which is done in zero field cooled (ZFC), field-cooled-cooling (FCC) and field-cooled-
through investigation of magnetic, dielectric, electronic properties. The warming (FCW) protocols. Fig. 2(a) illustrates the M (T) curves recor
material is found to reveal a spontaneously polarized electrical state ded under 100 Oe applied field. The ZFC, FCC and FCW, M (T) curves
within the AFM state with substantial linear magnetoelectric coupling. rises sharply with decreasing T below TN ~361 K demonstrating
The temperature dependent Raman spectroscopic study revealed wave occurrence of weak ferromagnetic transition similar to earlier reports on
vector q dependent spin–phonon coupling strength at the magnetic and YFe(1-x)MnxO3 (x = 0.4, 0.45) [5,14]. However below TN several inter
ferroelectric ordering temperatures. Overall, we argue here that the q esting features can be observed in SYFM (58–42) that are similar to that
dependent spin-pair correlation has marked effect on the measured observed in YFe0⋅6Mn0⋅4O3 but several additional features observed in
physical properties in the material. SYFM (58–42) haven’t been seen before in RFeO3 and RMnO3 systems.
The following important features have been observed from the dc
2. Experimental procedure magnetization measurements:
Polycrystalline Sm0.5Y0⋅5Fe0⋅58Mn0⋅42O3 is prepared via. Solid-state 1) With decreasing temperature, recorded M (T) curves in all the pro
reaction route using high purity oxides R2O3 (Sigma Aldrich, 99.99%), tocols undergoes a sharp decrease around TSR2 ~ 292 K which is a
Fe2O3 (Sigma Aldrich, 99.9%) and Mn2O3 (Alfa Aesar, 99.9%). Oxides convincing signature of the spin reorientation transition. The huge
R2O3 were preheated at 950 K for 7 h. The stoichiometric quantities of irreversibility between the FCC and FCW magnetization states con
the binary oxides are then intimately mixed and heated at 1000 ◦ C for 24 firms the metastable nature or first order nature of the spin reor
Hrs. The final calcination and sintering were done at 1350 ◦ C for 16 Hrs ientation transition similar to that observed in DyFeO3 and YFe(1-
several times with intermittent grinding. The single phase chemical x)MnxO3 [5,14,10,18] below TN., The merging of the FC curves below
composition is confirmed by the x-ray diffraction studies at room tem 360 K implies the second order nature of the magnetic transition at
perature recorded in a Rigaku x-ray diffractometer (Model: Ultima IV) TN. Comparing with the M (T) data of the present material with
earlier reports on YFe0⋅6Mn0⋅4O3 [5,14], it can be concluded that
2
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 1. (A) Shows the Rietveld Refinement of the SXRD diffractogram of the SYFM (58–42) at 298 K. (B) Shows the schematic representation of the unit cell of
SYFM (58–42).
SYFM (58–42) undergoes a temperature induced SR transition into a At further low temperatures FC curves exhibited an upturn towards
nearly collinear (Γ1) AFM state below TSR2. M = 0 axes indicating occurrence of another compensation point for T <
2) It is important to note that an anomalous T variation of M (T) occurs 11 K. The increase in the M (T) towards positive field direction can be
at TSR1 ~ 348 K as indicated in Fig. 2(a) and (b), which is more due to ordering of Sm3+ sublattice that may occur for T < 2 K in SYFM
prominent in ZFC mode than in the FC modes. This anomaly can be (58–42). This can be verified by measuring magnetization/specific heat
regarded as an incomplete spin reorientation transition probably of as a function of temperature below 2 K, which is beyond the scope of this
the second order type that can be triggered by the Sm3+ -Fe3+ work.
anisotropic interaction below TN. In YFe (1-x) MnxO3 (x < 0.45), high In order to verify the role of paramagnetic Sm3+ ions polarized by
anisotropic character of the Mn3+ ions causes the TM spin structure to TM sublattice, in inducing the observed NM below Tcomp, we have
abruptly change from Γ4 → Γ1 configuration just below TN for x = 0.4 employed a quantitative approach using the formulation derived in
[5,14]. Also the anisotropy interaction energy of Sm3+ -Fe3+ inter Cooke et al. [19]. The FCW magnetization within T range of 11–250 K
action being comparable to Fe3+-Fe3+ interaction, also causes a and under different fields had been modelled separately using the
gradual (second order) changes spin configuration Γ4 → Γ1 in following expression [4,19]:
SmFeO3 at elevated temperature (TSR ~ 480 K) [3]. In Dy (1-x) Sm
C(H + HI )
xFeO3 single crystals [18], it has been observed that the magnitude of MNet = MFe/ + (1)
the Sm3+ -Fe3+ anisotropy interaction energy becomes dominant than
Mn (T − θ)
the Dy3+ -Fe3+ interaction energy for x > 0.2, because of which the Here, MNet, M Fe/Mn, C, H, HI and θ represents the net magnetization,
exhibited second order spin reorientation transition for T > 50 K. magnetization due to the canted M-O-M (M = Fe/Cr), a Curie constant,
an applied field, an internal field on Sm3+ ions, and a Weiss temperature,
Hence, it is plausible to consider that strong Sm3+ -Fe3+ anisotropy respectively. Fig. 2 (e)–(g) shows the satisfactory fitting of FCW M(T)
interaction energy is prevalent below TN in the present system, which with Eq. (1) (demonstrated as the solid line) for the applied field Hext. Of
induces the temperature dependent second order transition from Γ4 → Γ1 100 Oe, 800 Oe and 50 kOe respectively. The fitting results are listed in
phase below TSR1 ~ 348 K. However from the view point of high Table-1.
anisotropy character possessed by Mn3+ ions in the TM sublattice, As seen from Table 1, under 100 Oe, the internal field Hint. On
continuous spin rotation process changes into an abrupt one at TSR2, Sm3+is negative. However it being slightly greater than Hext., the FC
driving the system into the purely antiferromagnetic Γ1 phase [5,14]. induced exchange anisotropy energy of Sm3+ ions from TM molecular
Thus, the strong competition between the anisotropy energy of R3+-Fe3+ field forces them to get aligned in the direction opposite to Hext. And also
interaction and anisotropic nature of Mn3+ ions governs the SR transi to the Fe–O–Mn antiferromagnetic sublattice that is oriented in the
tions in these co-doped systems. applied field direction. With decreasing T, the polarized Sm3+ moments
increase and becomes comparable with the AFM moments of the TM
3) With decreasing temperature, FC curve exhibits negative values of sublattice at Tcomp., below which the net magnetization of the sample
magnetization below the compensation temperature Tcomp ~92 K. becomes negative as seen in FC M (T) datas measured at 100 Oe (Fig. 2
This indicates that a ferrimagnetic ground state occurs at low tem (a)).
peratures that can be attributed to the FC induced anisotropy of the For Hext. = 800 Oe, the Hint. Assumes negative values but slightly
polarized Sm3+ ions in the AFM state as discussed latter. On contrary, lesser than Hext. Thus, the Sm3+ ions gets aligned towards Hext. Due to
the ZFC magnetization remained positive down to the low measured enhanced Zeeman energy over the exchange anisotropy energy on the
temperature. The specimen is in demagnetized state within the ZFC Sm3+ ions [18]. It must be noted from Fig. 2(c) that the FC-M (T) curves
protocol. So ZFC M(T) contains nearly orthogonal AFM vectors with lie below the ZFC curve for T < 120 K. Under high magnetic fields i. e
weak ferromagnetic components (FM) randomly distributed Hext = 50 kOe, Hint. Becomes positive but less than that applied field. As
throughout the material. Thus, the FM components are compensated a result, large moment appeared for T < 120 K as seen from Fig. 2(d). In
and only the AFM moments contribute the ZFC magnetization [1]. addition, the first order transition is fully suppressed in higher magnetic
fields (Fig. 2(d)). The relative orientations of different moments relative
3
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 2. Panel (a), (c) & (d) shows thermal variation of the ZFC, FCC and FCW magnetization between 2 and 400 K under applied field of 100 Oe, 800 Oe and 50 kOe
respectively. Panel (b) T variation of the dMZFC/dT showing the minimum at TN and two maxima corresponding to the inflection points at TSR1 and TSR2. Panel (e)–(g)
shows the FCW magnetization at 100 Oe, 800 Oe and 50 kOe along with the fit with Eq (1) (solid lines) between 15 and 230 K.
4
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 3. The schematic representation of the relative orientation of the Fe/Mn sublattice and the Sm3+ions with the external (Hext.) and internal fields (Hint.) at
different field strengths in the field cooled protocols [15].
(H) is clearly seen, which can be understood as a complex behaviour TmFe0⋅5Cr0⋅5O3 system [9].
arising from the AFM (or WFM) order of the Fe/Mn sublattice, super In this report, we have further established the ZFC memory effects
imposed to this signal, the paramagnetic response of the rare earth, occurring within the reentrant spinglass-like states in the material
which is not linear for paramagnetic Sm3+ ions. Such nonlinearity of M (coexisting with the long range magnetically ordered state) at low
− H loops below the NM has also been observed previously in the temperatures [15]. Earlier in SYFM (58–42), we have observed a
5
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
reentrant SG state with the SG transition temperature TSG = 67.5 K temperatures approximately 74 K, 62 K, 52 K, 43 K, 31 K and 11 K that
through ac magnetic susceptibility studies while a contribution of linear are indicated as TW0, TW1, TW2, TW3, TW4 and TW5 respectively in the
temperature dependency of Cp (T) has been detected at low T along with inset. Prominent dips can be noticed at each TWi (i = 1–5) i. e for Twait <
a broad Schottky anomaly for T < 15 K. In this protocol, the sample had TSG. The occurrence of the SG transition in the present material couldn’t
been cooled from temperatures much above TSG, first in the absence of be detected in the (M (T) vs T measurements. This can be due to strong
any magnetic field and without any stop up to 2 K. The field of 100 Oe is background contribution from the long range ordered sublattice
then applied during warming the material from 2 to 200 K, thereby together with the increasing paramagnetic contribution of the Sm3+ ions
constituting a reference curve. The sample again cooled from 200 K in the [20]. This observation is similar to that observed in MnCr2O4 [21] where
absence of the magnetic field but with several stops with wait time the occurrence of the coexisting FIM and SG phases also hinders the
twait=5000 s at each stop. The waiting temperature Tw had been selected frequency dependency features in ac susceptibility. In that case, only the
such that twait was imposed on the system beginning from temperatures appearance of memory effect solely confirmed the occurrence of SG
just above TSG to temperature far below it. The result of the ZFC-memory transition in the spinel.
effect measurements on SYFM (58–42) is displayed in Fig. 4(B). Hence, the appearance of the ZFC memory effects in the material
Inset of Fig. 4(B) shows the difference curve ΔM = MNo wait -MWait, below SG transition confirms the freezing of spins rather than a super
spanning between 2 and 200 K. Twait were introduced successively at paramagnetic state [22]. The occurrence of magnetic glassy phases can
Fig. 5. The thermal (T) variation of the real part [εr′ (f, T)] of dielectric constant and tanδ are displayed in (A) and (B) respectively at different frequencies. The insets
of figures (A) and (B) displays the respective T variation of the derivatives w. r.t temperatures. The set of anomalies I and II in the dielectric spectrum are indicated by
the arrows in (A) and (B). Panels (C) and (D) displays the dependency of the relaxation times on the peak temperatures obtained from the dεr′ /d T vs. T curves (inset
of (A)) for anomaly I and II respectively. The solid lines in (C) and (D) are fit with the V–F law described in the text. (E) The T variation of the εr′ (T) at 0.5 kHz in
heating and cooling cycle between 90 and 400 K showing significant irreversibility across the FOMT envelope centered at TSR2.
6
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
be attributed to the intrinsic disorderness created from the random oc second order nature.
cupancy of the magnetic ions at same crystallographic site and frustration In order to seek the contributions of the various microstructural re
of magnetic interaction namely AFM and FM between the magnetic gions to the measured dielectric constant of the material, the dielectric
species due to the mixed valency of Fe and Mn ions. The site and bond and impedance spectroscopy had been conducted at various tempera
(FM and AFM) disorderness in the material can create a freezing in of the tures between 90 and 400 K and within the frequency interval of 100 Hz-
spins as observed earlier by us in YFe0⋅9Cr0⋅1O3 [23]. These glassy 2 MHz. Fig. 6(A) and (B) displays the ε′ r (f, T) and tanδ (f, T) as a
magnetic phases are also coexisting with the long range ordered spins function of the log f at various fixed temperatures between 180 and 400
(AFM phase) at low temperatures in the specimen [15,23]. K. Fig. 6(A) shows that ε′ r (f, T) decreases monotonically with increasing
frequency at 180 K. With increasing T, a broad hump appeared from
lower frequency side for T = 240 K, which typically implies the relax
3.3. Temperature dependent dielectric spectroscopic studies ation of the thermally activated defect charges (space charge/hopping
polarization) originating in the grain (G)/grain-boundaries (GB) of the
The complex dielectric constant [ε*(f, T)] of SYFM (58–42) polycrystalline material [32]. This broad hump apparently shifted to
measured within the T interval of 90–400 K with fixed frequencies and higher frequencies with increasing T also. Interestingly, with further
are displayed in Fig. 5(A)–(E). Fig. 5(A) shows the T variation of real increasing temperature for T > 260 K another low frequency relaxation
part (ε′ r (f, T)) of complex dielectric constant, exhibiting two anomalous appeared, manifesting itself as a nearly f independent plateau at the
set of steps in two different thermal regimes above 150 K. The first ε′ r (f, lower f domain (<1 kHz). The f independent ‘plateau’ extends over de
T) step [indicated by red arrow in Fig. 5(A)]) manifest itself with low cades of frequency with increasing T, indicating strengthening of the
steepness. The second ε′ r (f, T) anomaly begins from the end of the first relaxation process [25]. The tanδ vs log f plots also displayed humps as
anomaly and is manifested as huge dielectric steps (indicated by black illustrated in Fig. 6(B) that coincide with both the weak relaxation hump
arrow in Fig. 5 (A)) having a higher slope than the first one. The second and strong relaxation plateau in ε′ r (f, T) plots (Fig. 6(A)).
anomalous ε′ r (f, T) step spanned in magnitudes between ~ 220 and 950 In real dielectric materials, the dielectric relaxation cannot be
at measured frequency of 1 kHz. Since both the anomalies occurred modelled with Debye equation that consist single relaxation time of the
around the vicinity of the magnetic transition temperatures (viz. TSR1, polarization charges. The polarization of the trapped mobile charges,
TSR2 and TN) it indicates that a substantial magnetoelectric effect [24] ions as defects at grain and grain boundaries causes distribution of
may be present in the material similar to that observed in YFe0⋅6Mn0⋅4O3 relaxation times. Such non-debye relaxation processes led to the for
[5]. On the other hand as both the dielectric anomalies are found to be mation of depressed semi circles in the argand plane of complex
frequency dependent i.e. the dielectric steps decrease in magnitude with dielectric functions [33]. According to the Cole–Cole model the dielec
increasing frequency of the measurement, it indicates that it should have tric functions ε′ r (f, T) and ε′′ r (f, T) can be separately modelled as [33]:
contribution from some low frequency relaxation process [25]. /
The tanδ (f, T) vs T plot, as displayed in Fig. 5(B), also shows two set ′
εr − εr∞ 1 + (ωτ)(1− α) sin απ 2
of peaks, that are coincident with the dielectric steps in ε′ r (f, T) [Fig. 5 = / (2)
εrs − εr∞ 1 + 2(ωτ)(1− α) sin απ 2 + (ωτ)2(1− α)
(A)]. The low and high T sets of peaks are designated as anomaly-I and
anomaly-II respectively, and are found to be frequency dependent also. /
In Fig. 5(B), the increased values of tanδ in the anomaly-II suggest an
α)
ε′′r (ωτ)(1− cos απ 2
enhanced dc conductivity in the medium [26–28]. Inset of Fig. 5(A) and = / (3)
εrs − εr∞ 1 + 2(ωτ)(1− α)
sin απ 2 + (ωτ)2(1− α)
(B) displays the T variation of dε′ r (f, T)/dT and dtanδ (f, T)/dT plots
respectively. They suggest relaxor-like behaviour of the inflection point Here εrs and εr∞ are respectively the static and high frequency limits
of the ε′ r (f, T) and tanδ (f, T) plots. Fig. 5(C) and (D) displays the peak of dielectric constant, τ is the most probable relaxation time and α is the
temperature (Tp) [obtained from dε′ r/dT vs T plots in Fig. 5(A)] de broadening parameter that assumes the values 0 = α ≤ 1. The modelling
pendency of the relaxation time (τ), (calculated from the measured of the frequency explicit plots of ε′ r (f, T) with Eq. (2) are shown by solid
frequencies) for the anomalies I and II, respectively. In order to seek the lines in Fig. 6(A). However, within the entire T range, single set of pa
dipolar cluster glass dynamics associated with these anomalies the plots rameters in Eq. (2) cannot reproduce the experimental data for T ≥ 230
of τ were fitted with Vogel-Fulcher’s (VF) law τ= K and two similar set of right hand terms of Eq. (2) were used to fit ε′ r (f,
( ) T) vs log f plots satisfactorily (shown as the red solid lines). Solid lines in
/
τ∗ exp Ea k (T − T ) , where Ea is the activation of energy of the blue in Fig. 6(A) represents the fit with single set of parameters of Eq. (2)
B p 0 to reproduce ε′ r (f, T) data. Fig. 6(C) and (D) illustrates the thermal
relaxation process, kB is the Boltzmann’s constant, τ* is the character variation of ln (τ1) and ln (τ2) respectively and Fig. 6 (E) shows T de
istic relaxation time, and T0 is the freezing temperature for dipolar pendency of α1 and α2. Here 1 and 2 represents the high and low fre
dynamics. quency relaxations respectively [Fig. 6(A) and (B)]. Fig. 6(C) and (D),
From the VF fit [Fig. 5(C) and (D)], thermal activation energy for demonstrated no changes in the slope either at SR or at TN of the spec
anomaly I are found to be (Ea)I ≈ 154 meV along with the characteristic
imen. Each ln (τ) vs 1000/T plots had been treated by Arrhenius law τ =
time (τ*)I ≈ 3.75 × 10− 10 s and temperature (T0)I = 59.78 (1.2) K, while ( )
the fitting yielded the parameters as (Ea)II ≈ 236 meV, (τ*)II ≈ 2.215 × τ0 exp Ea/k T in order to extract the activation energy Ea for the re
b
10− 10 s and (T0) = 79.73 (3.28) K for anomaly II. This reveals the vit laxations and are shown by the solid line in the figures. The activation
reous nature of the dipolar correlations with medium-range length scale energies are calculated as Ea1 = 316.19 (3.8) meV and Ea2 = 295.31 (5)
in the material. The low values of the activation energies for the relaxor meV [34]. The calculated Ea values indicate the p-type small polaron
dynamics of the material are of the same order as that obtained in hopping relaxation in the medium [35]. Fig. 6(E) illustrating the T
SmFeO3 nanoparticles [29], in multiglass FeTiO5 [26] and also for the variation of α1 and α2, shows critical behaviour at TN, TSR1 and TSR2. A
solid solution of the BaTi1− xZrxO3 relaxor system [30]. Moreover, the dip is observed also at 220 K in the T variation of α1 as indicated by
activation energy (~100 meV) is also observed in other relaxor ferro arrow in Fig. 6(E). The origin of such dip in α1-T plot at 220 K has been
electrics [28,31,30]. Fig. 5(E) shows the thermal variation of the ε′ r (f, T) ascribed to the onset of electron phonon interaction causing charge
at 0.5 kHz measured in warming and cooling cycle. A wide thermal localization as observed from Raman spectroscopy and resistivity mea
irreversibility between the cooling and warming curves is apparent at surements in the later section. As there are no anomalous changes in T
TSR2 also validates the first order nature of the SR2 transition. On the variation of τ, but there are apparent anomalies in the thermal variation
other hand, the reversibility between the curves at TN also establishes its
7
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 6. (A) and (B)shows the frequency dependence of the real part ε′ r (f, T) and tanδ between 100 Hz and 2MHz at various selected T between 180 and 380 K.
Joining lines are guide to the eye. Panel (C) and (D) shows the logarithm of the relaxation time (ln τ1, ln τ2) vs. 1000/T. solid lines are Arrhenius fit described in the
text. (E) Shows the T variation of the parameters αi (i = 1, 2) demonstrating critical behaviors around TSR1, TSR2 and TN.
of the width of the distribution of the relaxation times, it is suggestive response of the material, the combined impedance and modulus spec
that a weak coupling may exist between the magnetic and dielectric troscopic analysis had been employed [25,35] (See supplementary ma
properties in this studied specimen. terial). Quiet astonishingly, it can be observed from the combined
Since the time constant of the relaxations are comparable it is diffi frequency explicit plots of Z′′ and M′′ [Fig. S3, supplementary material]
cult to designate them as G or GB contributions. Moreover, at high that for T < 180 K only the G effect exists in the measured frequency
temperatures electrode-material interface polarization effects (EP) may window. For 190 K ≤ T ≤ 270 K, both G and GB effects contributed in the
co-contribute with the other microstructures also. In order to clearly ac response.
distinguish the effect of the G, GB and EP effects in the ac electrical For 270 K < T ≤ 360 K, all the microstructural regions contributed to
8
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
the electrical response. Hence the weak relaxation “broad hump” Rgb and Re between 90 and 400 K have been plotted as ln (Ri/T) vs.
observed in ε′ r (f, T) plot for T > 200 K [Fig. 6(A)] consists of both the G 1000/T (i = g, gb and e) as displayed in Fig. 7(A)–(C). Fig. 7(A) shows
and GB relaxation, while the frequency independent “plateau” for T > the entire plot of ln Rg vs 1000/T is being sub-divided into different T
260 K arises from the EP effects. These observed features of co- regimes designated as I, II and III in the figure, having different slopes.
contribution of G and GB in a single dielectric relaxation is exactly First apparent changes in the slope in the plot occurred for T > 140 K (T
similar to that observed in CFO-PZT thick films [25]. In order to further region I) and the second one for T > 220 K (II). It is to be noted that the
extract the individual G, GB and electrode-material interface resistances slope change in T region II is also coincident with the dip in α1-T vari
and capacitances, equivalent circuit modelling of the Nyquist plots had ation [Fig. 6(E)].
been employed, the details of which are given in supplementary section Fig. 7(B) and (C) displays the thermal variation of resistances as ln
[Fig. S2]. Rgb and ln Re vs 1000/T respectively. Fig. 7(B) displays the a similar
The thermal variation of individual microstructural resistances Rg, changes in Rgb with increasing T, leading to two distinct thermal regimes
I (190 K ≤ T ≤ 270 K) and II (270 K < T ≤ 380 K) also having different
activation energies. The linear regions in Fig. 7(A), (B) and 7(C) have
been fitted with Arrhenius law, Ri/T =Ri0 exp (Eact/kBT), where
subscript i stands for g, gb and e in the expression and Ri0 is a pre-
exponential factor [36]. The excellent fitting of the logarithmic resis
tance plots with Arrhenius law are represented as the solid lines in Fig. 7
(A)–(C) also. The activation energies Eact. For conduction, obtained from
linear fitting [Fig. 7(F)] are (79.71 ± 4.65) meV, (292.52 ± 19.35) meV,
and (317.35 ± 10.32) meV for regions I, II and III respectively. Similarly
values of Eact. Calculated from ln (Rgb/T) vs. 1000/T plot, are (336.23 ±
9.96) meV and (340.32 ± 8.79) meV in the regions I and II respectively,
while the calculated activation energy from the EP charges are obtained
as (416.32 ± 15.33) meV. The obtained values of the Eact. For conduc
tion are similar to that obtained from dielectric relaxations in the ma
terial [Fig. 6(c) and (d)] that suggests the same conducting charge
species to be involved in dielectric relaxation in SYFM (58–42). From the
above ac electrical analysis, it is clear that bulk and grain boundary
conduction are dominant at low temperatures (T < 280 K) for exhibiting
dielectric relaxation under the ac field. However at the high tempera
tures, the EP effects dominates the ac response of the material especially
for T ≥ 280 K at lower frequencies, although G and GB polarization
charges continues to contribute to the dielectric relaxations at mid and
high frequencies respectively.
The linear variation of Rg, Rgb and Re plots against 1000/T [Fig. 7
(A)–(C)] suggests the small polaron hopping (SPH) conduction of the
charge carriers based on the strong electron-phonon coupling at these
temperatures. Hence the plots are fitted with the Mott’s SPH model [37]:
( )
/ W
(4)
′
ρi = kT R ν e2 c (1 − c) exp(2α R)exp
0 kT
Here ρi stands for the resistivity of the concerned microstructure, T is
the temperature, k is the Boltzmann constant, ν0 is the optical phonon
frequency, R is the average intersite separation, e is the electrical charge,
c is the fraction of the transition metal ion concentration in the lower
valence state namely, in our case is [Fe 2+]/[Fe 2+ + Fe 3+] or [Mn 3+]/
[Mn 3+ + Mn 4+]. The fitting proceeded with α′ and ν0 as parameters and
there values are listed in Table 2. The possible intersite separation
within the unit cell can be nearest Fe/Mn–Fe/Mn distances 3.78 Å or
5.67 Å which corresponds to the ion centres along b axis and in the ab
plane along a direction of the unit cell. The results given in Table 2 are
agreeable with earlier reports on similar system [38,39]. The
non-vanishing value of α′ indicates a non-adiabatic small polaron hop
ping conduction mechanism in the microstructures. However, a hopping
process is of nonadiabatic small polaronic in nature, requires several
restrictions on the electron transfer integral between the neighbouring
hopping sites. These restrictions then serve as the criteria of judgement
whether the hopping conduction mechanism is adiabatic or nonadia
batic [39,40].
An alternative conduction mechanism [41] as demonstrated in
semiconducting manganites [39] is the variable range hopping (VRH)
conduction. According to the VRH model, the resistivity can be
expressed as [41]:
Fig. 7. (A)–(C) shows the T variation of individual microstructural resistances
ρ( T ) = ρ0 exp [(T0 / T)κ ] (5)
Rg, Rgb and Re, respectively along with SPH and VRH model fits (solid lines) as
described in the text.
9
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Table 2
1/4
The parameters deduced from the SPH modelling and VRH modelling of the logarithm of the ln (Ri/T) versus 1000/T and ln (Ri) versus T− plots respectively.
Micro-structures T regions Parameters
SPH VRH
where, T0 is a characteristic temperature, ρ0 is an exponential factor and Table 2 for T < 220 K, it can be concluded that dominant dc conductivity
κ assumes fixed values of 1/4 or 1/2, according to the Mott regime of is due to both the intergranular and intragranular conduction. The
uncorrelated hopping carriers or for a system of carriers with a gap due change in the activation energy at 220 K with a much lower one above
to correlations according to the Efros–Shklovskii mechanism [41,42] this temperature point indicates the change in conductivity mechanism
respectively. The plots of ln Ri (T) vs. 1/T− 1/4 for G, GB and EP as dis from p-type to n-type hopping conduction [34]. We have also modelled
played in Fig. 7(A)–(C) (right –bottom axis) reveals excellent agreement the ρ(T) data with VRH model [Eq. (5)].
with Eq. (5) [red lines]. The values of the characteristic temperatures T0 As displayed in Fig. 8(C), ln ρ(T) vs 1/T− 1/4 plot showed linearity
associated with each slopes are also listed in Table 2. The tabulated (shown by the solid line in the figure) only for T < 220 K while VRH
values shows T0 of the order ~108 K for each microstructural regions mechanism failed to be obeyed for temperatures above it. Thus the VRH
and is agreeable with earlier literatures [39,41]. Density of states near of the charge carriers satisfactorily explains the ρ(T) behaviour for T ≤
the fermi level N (EF) in each T regimes of G, GB and EP effects have been 220 K, while the non-adiabatic SPH mechanism explains the resistivity
calculated using the relation T0 = 16 α3/kB N (EF) and are also listed in variation for T > 220 K. It is because since the thermal energy of the
Table 2. The values of N (EF) are also found comparable to self-doped charge carriers is not enough to allow electrons to hop to their nearest
LaMnO3 and similar systems [39,43]. Thus both the SPH and VRH neighbours, the electrons favourably hop farther to find a smaller po
models can satisfactorily describe the individual T variations of the tential difference [43]. Such changes from VRH to SPH mechanism has
microstructural resistances in the present polycrystalline system. How been found in Fe doped LaMnO3 [45], heterovalent doped NdFeO3 [46],
ever due to the co-contribution of the G, GB/EP, the overall resistance in in mixed valent manganites [47], that can arise from electron localiza
the sample may exhibit a complex behaviour. tion due to electron-phonon (e-p) coupling. Moreover, the grain
boundaries may also act as potential barriers and contribute to the
localization of carriers. The e-p coupling in mixed manganite systems
3.4. Resistivity and magneto resistance measurements can occur via three different kinds of lattice distortions:
We have studied the spin-electronic correlation through measure (a) Jahn-Teller (JT) distortion of the Mn3+ ions octahedra which
ment of dc resistivity ρ (H, T) in the material as a function of temperature raises the energy of its outermost eg electrons.
and external magnetic field. Fig. 8(A)–(C) shows the results of the zero- (b) Breathing-type distortion due to the presence of formally two
field ρ (T) vs T measurement of the specimen. The thermal variation of different valence states, Mn3+ and Mn4+ [48].
the resistivity of the specimen measured at zero magnetic field in the (c) Distortion from the A-site cation size mismatch, which is valid for
cooling and warming cycles between 150-400 K is displayed in Fig. 8(A). our present system also.
The system exhibited a typical behaviour of an insulator/semiconductor
within the temperature interval of 150–400 K, with a steep rise on The lattice distortion arising from factors given in (a) –(c) can lead to
cooling below 170 K, similar to other insulating or semiconducting strong e-p coupling in the specimen. The overall ρ(T) behaviour of the
manganite systems [20,44]. Below 150 K, ρ (T) values becomes larger material can explained by considering the carriers to be ‘small polarons’
and so it couldn’t be detected by the instrument. The inset of Fig. 8(A) as had been established for mixed valence manganite systems [47].
highlights the T region in ρ (T) with the magnetic transitions viz. SR1, Corresponding to electron localization due to e-p coupling, the car
SR2 and WFM transition occurs. No sharp anomaly is vividly observed riers are localized as small polarons with a scale of about Fe/Mn–O bond
across TN and TSR1, but a hump can be observed in the cooling curve length ~2 Å [47]. At low temperatures and within the magnetically
across TSR2. This implies that FOMT weakly influences electronic de ordered state, the electrons as small polarons are self-trapped in a deep
grees of freedom of the carriers. The sharp decrease in ρ(T) for T > 170 K potential well. The thermal energy of the carriers being insufficient, they
is due to the increase in charge carrier concentration/mobility from cannot hop out from their site, while it is more likely to be activated into
thermally activated carriers, at G and GB as concluded in section III(C), an intermediate state first, which is still a localized state but with higher
which enhances the overall ρ(T) of the specimen. energy. Thus the thermal energy becomes enough for a polaron to hop to
In order to verify the conduction mechanism through thermally an energetically equivalent site under the influence of the magnetic
activated charge carriers, the ln ρ(T) of the material has been plotted localization due to spin disorder on the interatomic scale (~1 nm) in the
against 1000/T as displayed in Fig. 8(B). The observed plot is found material. However the charge carriers can either hop to a NN site or to
linear with a sharp change in the slope around T~ 220 K. Each linear T further site. This explains the satisfactory agreement of the ρ(T)
domains of Fig. 8(B) viz. Regime I and Regime II, have been fitted with behaviour from both the SPH as well as VRH models in the thermal
SPH model [Eq. (4)], using the respective slopes for calculation of the regime II [Fig. 8(B) and (C) respectively].
activation energies. The values extracted for α and ν0 are 0.1921 Å− 1 and With further increasing T until 220 K, the enhancement of the charge
1.47337 × 1013 Hz for Regime I with Eact = 0.026 eV, while for Regime II carrier concentration occurs from EP charges and also strength of elec
with Eact = 0.291 eV, the extracted parameter values are 1.33684 Å− 1 tron phonon interaction gets altered. This causes further electron
and 8.34461 × 1013 Hz. localization and the carriers gain sufficient thermal energy to hop to its
Comparing the results of the activation energies with that listed in nearest neighbouring sites without undergoing into the intermediate
10
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 8. (A) The T variation of resistivity ρ(T) of Sm0.5Y0⋅5Fe0⋅58Mn0⋅42O3 between 150 and 390 K in the cooling and warming cycles. Inset of the figure shows the
magnified region around the magnetic transitions at high T. (B) Log (ρ(T)/T) vs. 1000/T plot and the SPH modelling with change in the activation energy at T ~220
K. Inset shows the deviation of the SPH model near 220 K in Regime II. (C) The ln ρ vs. T− 1/4 plot showing linear portion for T < 220 K. (D) The resistivity and
magnetoresistance vs. field cycle at 300 K. (E) The resistivity vs. field cycle at 260 K. Inset shows the MR% vs field variation at 260 K.
state. Hence ρ(T) obeys the SPH model of charge carriers in regime I T to 6 T [Fig. 8(D)]. Astonishingly in Fig. 8(D), in the metastable region,
[Fig. 8(B)]. The extracted T0 of the VRH model [49,50] in T regime II ρ (T)/MR increases linearly for applied magnetic field variation from 0
[Fig. 8(B)] gives the value of density of states near Fermi level N (EF) = → 6 T. Since from the ρ (T) measurements it is revealed that although
(0.98 ± 0.014) × 1020 eV− 1cm− 3 that is agreeable with earlier litera both the AFM and WFM phases are highly resistive, yet the resistivity of
tures [47,48]. the AFM state is higher than that of the WFM state. Within the meta
The magnetoresistances (MR) at two different temperatures 300 K stable region, both the AFM and WFM states coexists with minimum free
(in the metastable region) and at 260 K in the AFM state as displayed in energies and both have the same free energy [51–53]. Their free energy
Fig. 8(D) and (E). Both at 300 K and also at 260 K, the sample exhibited potential wells are separated by an energy barrier whose height repre
small values of MR of nearly 0.01% at 6 T on initial field increase from 0 sents the energy required for the formation of stable nuclei of the AFM
11
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
phase inside the WFM phase [53,54]. On decreasing temperature, the irreversibility during the positive half cycle. This can be due to the
free energy of the AFM state becomes lower than that of the WFM as presence of some fraction of arrested WFM phases in the stable AFM
shown in Fig. 9(A) for T* < T < TSR2. Upon crossing T*, if the kinetic matrix. The application of the magnetic field of 6 T is also not capable to
arrest temperature Tg < T* of the supercooled WFM phase, WFM phases transform the arrested WFM phases into the low T higher resistive AFM
transform completely into the stable AFM phase [52]. The trans phases, while the application of the same H in the negative cycle
formation of the supercooled state can also take place even in the completely transform it. This is because due to some quenched dis
metastable region, upon the application of a sufficient magnetic field orderness, kinetic arrest of the FOMT lead to the some fraction of un
called the critical field Hcri. [52–55]. For several materials including transformed WFM phases in the AFM ground state. At T = 260 K (<T*)
Heusler alloys [51–53], doped manganites [56,57], the application of the arrested phase being lower in concentration than the AFM phase,
magnetic field in the FOMT region enhances the difference between the application of H transform almost entire WFM phase during the positive
free energy of the low T and high T states and therefore will further half cycle. Thus in negative field cycle, ρ(T) and MR remains nearly
reduce the free energy barrier as shown in Fig. 9(B). So the material constant [Fig. 8(E)]. Hence the above results of zero-field resistivity and
undergoes a field induced transformation from supercooled high tem isothermal MR measurements confirms a very weak scattering of the
perature phase into the low T state above Hc. This field induced mag conduction electrons by the spins and also arrested WFM phases can also
netic transformation within FOMT is widely manifested as irreversibility be present in the AFM state in the specimen.
and reversibility between the forward and backward curves for H < Hc
and H ≥ Hc respectively. 3.5. Magnetodielectric, pyroelectric and Raman spectroscopic
In our specimen, application of H in the forward and backward di measurements
rection in the positive half cycle exhibits a wide irreversibility. This
implies that the initial increase of the magnetic field up to 6 T is still not In order to visualize directly the degree of spin-charge coupling in
sufficient to convert all of the supercooled WFM phases into stable AFM the specimen, the measurement of the dielectric constant and loss under
phase. Upon reducing the field, the transformed AFM phase cannot go variable magnetic field and at fixed temperatures had been conducted at
back into the metastable phase. As such the resistivity remains higher four excitation frequencies viz. 1 kHz, 10 kHz, 100 kHz and 1 MHz. The
when the field is reduced to zero. It is to be noted from Fig. 8(D) that temperature of the measurement are selected within 5–300 K. The
ρ(T)/MR increases also during the backward paths i. e ± 6 T→ 0 T that symbols with black, red and blue colour respectively represents the
suggests the phase transformation of the supercooled WFM state during datas recorded for 0 → +5 T, -5 T → +5 T and -5 T → 0 T as shown in
decreasing the field, but the irreversibility between the forward and the Fig. 10(C) and (D), 10(G) and 10(H). As shown in Fig. 10(A) large
backward process is lesser during the negative cycle than the positive changes can be observed in the ε′ r (H, T) between 1 kHz and 1 MHz
one which implies that application of H > 6 T can transform the meta because of the dielectric relaxation at the concerned temperature. As
stable WFM phases. After a complete cycle, MR increased up to ~0.02% displayed in Fig. 10(C), the compound exhibited a robust magneto
at 300 K [Fig. 8(D)]. dielectricity at 300 K, with MD % [ = {(ε′ r (H) - ε′ r (0))/ε′ r (0)} × 100] of
Fig. 8(E) displays ρ(T) and MR variation with H at 260 K (in the ~0.8% with initial H increase of 0 T→ +5 T measured at f ≥ 10 kHz.
stable AFM state). Like at 300 K both ρ(T) and MR shows huge With further successive field branches, MD% increases linearly for f ≥
10 kHz in contrast to that found for 1 kHz.
The Magneto loss (ML) defined as ML% [ = {(tanδ (H) - tanδ (0))/
tanδ (0)} × 100] also displays a linear increase in the first increasing
branch while, it decreases in the successive field branches measured at 1
kHz. At 300 K the Maxwell-Wagner (M − W) relaxation is also present
along with the FOMT. Hence the observed magnetocapacitive response
may comprise contributions from both MR in conjunction with the M −
W effects [58,59] (extrinsic) and also from q dependent spin-pair cor
relation function <MqM-q> (intrinsic) [60]. It is to be noted respectively
from Fig. 10(A) and (C) that both MD % vs. H and ML % vs. H for f > 1
kHz mimics the MR vs H at 300 K which suggest that magnetic phase
coexistence and/metastabilty greatly influences the MD behaviour of
the material. For MD and ML at 1 kHz, the dependency on H is different
from that at mid and higher frequencies, which implies that charges
contributing to EP relaxation also affect the magnetodielectric behav
iour of the system especially at lower frequency.
In the Maxwell–Wagner relaxation model, the real (ε′ ) and imaginary
(ε") parts of the dielectric permittivity are given as [59,61]:
1 τ i + τ b − τ + ω2 τ i τ b τ
(6a)
′
ε (ω) =
C0 (Ri + Rb ) 1 + ω2 τ2
1 1 − ω2 τi τb + ω2 (τi + τb )τ
ε′′ (ω) = (6b)
Fig. 9. Schematic diagram of the free energies of the AFM and WFM phases
ωC0 (Ri + Rb ) 1 + ω2 τ2
across the first order phase transition. F denotes the free energy. The filled Here suffixes i and b refers to the interfacial-like (GB and EP) and
circles indicate the state in which the system is existing. (A) Shows the variation bulk-like layers, respectively, R = resistance, C = capacitance, = ac
of the magnetic phases with temperature across the first order phase transition
frequency, τi = CiRi, τb = CbRb, = (τi Rb +τb Ri ) /(Ri +Rb ) , C0 = ε0A/t, A =
region. Above the superheating limit T**, the WFM is the stable phase. While
area, and t = thickness of the capacitor. Clearly, according to Eq. (6),
during cooling, the WFM phase becomes metastable in the FOPT region within
the limit of supercooling T*. For T < T*, the supercooled WFM phases transform change in resistance of one layer invoke changes in the dielectric con
completely into the stable AFM phases if the kinetic arrest temperature Tg < T*. stant and dielectric loss in the system measured at a particular fre
(B) Shows the magnetic field induced transitions in the metastable region. With quency. Hence the combination of MR and Maxwell-Wagner (MW) effect
the enhancement of the field for H > Hc the supercooled WFM phase can be can lead to magnetocapacitive effects. This effect of combined MR and
transformed completely into the AFM phase. MW is explicit to the phenomenon of true magnetoelectric effect in the
12
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 10. Panel (A) and (B) shows the ε′ r and tanδ vs. H field variation at 300 K respectively at fixed frequencies. Panel (C), (D) shows the MD% and ML% vs H for a
complete field cycle at 300 K. Panel (E) and (F) shows the εʹr and tanδ vs. in a field cycle at 260 K at fixed frequencies. Panel (G) and (H) shows the MD % and ML %
vs. in the field cycle at fixed frequencies at 260 K. Panel (I) shows the MD% vs. H plot at 260 K at 1 MHz. For Panels (C), (D), (G) and (H) the symbols (#), (.), (9) and
(δ) respectively stands for the datas measured at 1 MHz, 100 kHz, 10 kHz and 1 kHz respectively.
material. On the other hand, the intrinsic MD (that comes from the g(q) is the wave vector dependent coupling strength in the medium, and
bound charges) can arise from the q dependency of the magnetodi <Mq M-q> is the thermal average of the instantaneous spin-spin corre
electric coupling term in the free energy (F) [60]: lation, which obeys the sum rule:
∑
P2 ∑ < Mq M− q >= Ng2 μ2B S(S + 1) (8)
F= − PE + P2 g(q) < Mq M− q > (T) (7)
2ε0 q
q
13
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
E relation < Mq M− q > (T) as well as coupling strength g(q) are altered and
P= ∑ ≡ εE (9)
1
ε0 +2 g(q) < Mq M− q > (T) so the magnetodielectricity changes its sign accordingly. Fig. 10(H)
q
displays the ML% vs H at 260 K in the same fixed frequencies. It can be
observed that ML% measured at 1 kHz, 10 kHz and 100 kHz mimics the
Where dielectric constant
respective MD% similar and MR% at 260 K in a field cycle. Again since
∑
ε=
ε0
, I(T) = g(q) < Mq M− > (T) (10) the MR at 260 K is also ~10 times less than the MD in the field cycle, the
q
1 + 2ε0 I(T) q observed MD and ML cannot be driven by the resistive components
(along with MW relaxation) itself and hence must have a majority
Hence the dielectric constant at a particular temperature greatly contribution from spin-pair correlation in the magnetic phases.
depends on the q dependent magnetic correlations. The results of the MD measurements at 5 K are displayed in Fig. 11
The ML% vs H at 300 K as dispayed in Fig. 10(D), exhibited following (A)–(C). Fig. 11(A) shows a nearly frequency independent values of ε′ r
marked changes in the polarity with the field cycle as the measuring f is (T, H) due to absence of any relaxation at low temperatures. ε′ r (T, H) vs
varied from 1 kHz to 1 MHz: H however exhibits a linear increase in the forward and backward field
variation at positive and negative field cycles. The tanδ (T, H) assuming
(i) The ML% at 1 kHz increases nonlinearly towards positive values very small values at 5 K, exhibits field independency at all measured
with the initial H increase during the successive H variations. frequencies, as displayed in Fig. 11(B). The MD % vs H at 5 K is increased
(ii) The increasing trend at 10 kHz and 100 kHz, ML% during the first linearly with increasing f in the entire field cycle as shown in Fig. 11(C)
half cycle i. e 0→+5 T→0 T, is changed into a decreasing one where it attains a value of ~0.33% at the end of the cycle. As the ma
during the second half field cycle i. e 0→-5 T→0 T. terial is highly resistive (i.e. no relaxation effects present) for T < 100 K
(iii) Lastly, the ML% vs H plot at 1 MHz becomes entirely negative for showing negligible dielectric losses at 5 K, it can be concluded that the
the entire H cycle. In order to explain the features from (i)-(iii) obtained MD % at 5 K arises entirely from the intrinsic (true) spin-charge
above, the conclusions from the dielectric and impedance spec coupling in the bulk of the material [58,59].
troscopy [sec.(C)] should be recalled. The analysis of IS unveiled An important feature to notice is the peculiar linear increase of MD%
that all the three microstructural regions co-contributed to the ac with H cycle at 5 K that is also observable at 260 K and 300 K. Such open
electrical response [Fig. 10 (D)] at 300 K with the EP effects loops in the MD% vs H plots have been previously attributed to the
dominating the low frequency response (<5 kHz) of the material. spurious MD signals from the charges accumulated at the grains and
With the enhancement of f > 40 kHz, the polarization from grain boundaries i. e in presence of M − W relaxation and MR effects
electrode-material contact as well as at the grain boundaries both [62,63]. Although 300 K is the temperature point where the MW effects
relaxes while the intrinsic polarization effects from G remained are present, this increasing trend with H of MD% still occur at 5 K where
and govern the ac response at higher frequencies (>40 kHz). neither the MW relaxation nor the MR is present that can confer about
Indeed Catalan et al. [58,59] have showed that intrinsic magne erroneous MD signals [Fig. 11(C)]. However, at these temperatures the
tocapacitance should be measurable at frequencies higher than appearance of the coexisting magnetic phases at 300 K (AFM/WFM) and
the conductivity cutoff (RC time constant). For MR% at 300 K, an 5 K (SG/WFM) is a common feature [15]. As stated earlier in this section,
order less than the MD% and ML%, it can be concluded that the such mixed coexisting magnetic phases renders the spin-pair correlation
observed MC and ML for f ≥ 100 kHz are consequences of the true function to vary over finite regions in q-space, this also results in varying
ME coupling [24,35,60] in SYFM (58–42). The intrinsic ME coupling constant g (q) associated with different magnetic phases. Thus
within FOMT should arise from the coupling of the dielectric phase coexistence of WFM/AFM at 300 K or SG/WFIM at 5 K also
constant and spin-pair correlation in the AFM and WFM magnetic changes g (q) over finite region in q-space. Hence we suggest that the q
phases. At 300 K, since the field induced transition of supercooled dependent spin-spin correlation directly affects the dielectric state of the
WFM → AFM state occurs, < Mq M− q > (T) also undergoes spatial codoped system and plays a significant role in the open loop of MD% vs.
and temporal variation. In such spatial variation of the magnetic H response against field cycling like MR% vs H [Fig. 8(D) and (E)].
interactions, ME coupling in the material depends largely on the The Intrinsic MD effect phenomenologically described by the simple
strength of the competing AFM-FM interactions. The resultant < Ginzburg-Landau theory for phase transition and attributed to the ME
Mq M− q > (T) may causes the MD% to increase with H in the coupling term γP 2M2 in the thermodynamic potential (Φ) given as [64]:
material.
′
β β
(11)
′
φ = φ0 + αP2 + α M 2 + P4 + M 4 − PE − MH + γP2 M 2
Similar to that observed at 300 K, the ε r (T, H) at 260 K changes
′
2 2
vastly because of the presence of the relaxation, as displayed in Fig. 10
(E) and (F). In the entire field cycle, MD% at 260 K is found to be Where α, β, α′ , β′ , and γ are the constants and functions of temperature.
opposite in polarity to that at 300 K, especially for f >1 MHz As dis The influence of the magnetic order on the magnetic field driven mag
played in Fig. 10(G). Astonishingly, the MD% measured at 1, 10 and 100 netodielectric can be followed from the linear variation of the MD% vs
kHz remained negative during the entire field cycle, while for the 1 MHz M2 curve [65–67]. The square of the magnetization M2 against the MD
[Fig. 10(I)], the MD% during H variation from 0 → 5 T becomes response at different temperatures for the specimen showed linear
increasingly negative attaining a saturated value of ~ − 0.05% for 1.25 behaviour in the high field regimes i. e for H > 1.5 T at 300 K and H >
T < μ0H < 3.5 T. Then after, it increases with further H enhancement 2.75 T at 5 K as depicted in Fig. 11(D) and (E). This proves that the linear
turning to positive values for μ0H > 4.75 T. With field variation +5 T→ coupling term γP2 M2 term of the Ginzburg-Landau theory [Eq. (11)] is
0 T, the MD % increases and remains positive, attaining a value of significant for SYFM (58–42) similar to spinels MCr2O4 (M = Mn, Ni, Co)
0.135% at the end of the positive cycle. In the second half cycle, the MD [66,67].
% loop becomes mirror image of the positive one, resulting in a closed We have also investigated the occurrence of ferroelectricity by
loop in the MD% vs. H plot. Thus in the present material both the measuring the pyroelectric current. Initially to seek the genuine ferro
temperature and field induced sign reversal of the magnetodielectricity electric transition in the material, the bias electric field (BE) method had
occurs in the vicinity of the FOMT. The sign (negative or positive) on the been employed as recently described by N. Terada et al. [68]. After
magnetodielectric effect is determined by the product of spin-pair cor wards the pyroelectric current measurement had been conducted within
relation of neighbouring spins and the coupling constant [35]. the T range of 50–175 K. As displayed in Fig. 12(A), the current IDC
Since the magnetic states in the material suffer rigorous alteration measured in the bias electric field method under applied field of 3
when the temperature/magnetic field is varied, both the spin-pair cor kV/cm with a T sweep rate of 8 K/min reveals a slight dip around 108 K
14
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 11. Panel (A) and (B) shows the ε′ r and tanδ vs H plot respectively for the field cycle at 5 K. Panel (C) shows the MD % at various frequencies for the field cycle at
5 K. Panel (D) and (E) shows the M2 vs. MD % plots at 300 K and 5 K respectively at 100 kHz. The linear fit is represented as solid lines in the plots.
as indicated in the inset of Fig. 12(A). The appearance of the dip in IDC(T) cm2 under ± 5 kV/cm poling field respectively below TFE. Reversal of P
[Fig. 12(A)] confirms the occurrence of true ferroelectric state in SYFM due to a change in sign of E signifies ferroelectric behaviour of SYFM
(58/42) [96] with TFE~ 108 K as the ferroelectric transition (58–42). The value of the Ps obtained in the material is comparable to
temperature. several improper ferroelectrics [66–68] and hence confirms the
The pyroelectric current (IP) measurements under ±5 kV/cm be involvement of long-range ordering of electric polarization.
tween 50 and 175 K, revealed identical ± IP peaks at T~ 108 K as dis To seek the intrinsic origin of the magnetodielectricity, ME coupling
played in Fig. 12(B). Fig. 12(C) displays the time integrated IP that gives and ferroelectricity, Raman spectroscopic measurements as function of
intrinsic saturation electric polarization (Ps) values of about ±0.06 μC/ temperature had been conducted inside the T interval of 83–503 K. The
15
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 12. (A) Shows the bias electric field (BE) measurement showing the true ferroelectric transition at TFE = 108 K with a temperature sweep of 8 K/min. Inset
shows the zoomed portion around TFE. (B) Shows the pyroelectric current measurements under ±5 kV/cm electric field. with T ramp of 5 K/min between 50 and 175
K. (C) Shows the T variation of the ferroelectric polarization (P) of the material between 50 and 175 K, under poling fields of ±5 kV/cm. With T ramp of 5 K/min.
dielectric constant of a material usually depends on the long wavelength Phonon mode frequencies line widths which is suggested in recent lit
longitudinal and transverse optic phonon frequencies through the eratures also [71]. As the hence reveal the presence of significant
Lydian-Sachs relation. Hence, affecting the phonon mode at the mag spin-phonon coupling i. e phonon modulation of the spin-exchange in
netic transition through spin-phonon coupling [69], alters the dielectric tegral in the system [71–73]. Similar convincing anomalies in the
constant of the material, thereby giving rise to phonon mediated ME phonon mode frequencies are also evident around TFE/Tcomp, indicates
effects that is intrinsic in nature. Indeed, for isostructural SeCuO3 and that lattice modulation at the compensation temperature is stabilizing the
TeCuO3 it has been grounded both theoretically and experimentally that ferroelectric ground state in SYFM (58–42). This appearance of electric
coupling of the spin fluctuations affecting the optical phonons can give polarization in conjunction with the magnetization reversal convinc
rise to magnetodielectric effects [60]. Unpolarised Raman spectra of ingly suggest that SYFM (58–42) to be a Type II multiferroic [13,21].
SYFM (58–42) are displayed in Fig. 13(A) at selected temperatures. The Fig. 13(D) and (E) displays the T variation of the linewidths (FWHM)
spectra at 86 K, as shown in Fig. 13(B) illustrates the peak synthesis of of the B2g (7) and B1g (5) modes respectively in the measured T range.
the high intensity broad peak centered around 649.45 cm− 1. Its Lor Same convincing anomalous T variation of the linewidths of Raman
entzian peak deconvolution yielded two synthetic peaks corresponding modes are also evident at the magnetic and ferroelectric transitions. As
to M − O (M = Fe/Mn) octahedral stretching modes of B2g (7) and B1g shown in Fig. 13(D) there is an apparent temperature independency of
(5) symmetries centered around 632.34 ± 3.35 cm− 1 and 657.37 ± the linewidth of B2g (7) mode far above TN, that can be attributed to the
2.13 cm− 1 respectively at 86 K. Most importantly, these modes couple competition between a decrease in the linewidths due to the absence of
with the atomic spins in several Mn doped systems and hence plays the magnon–phonon interaction above TN and an increase in the line widths
key role in displaying spin-phonon coupling in these systems [33,34]. As with increased site disordering with increasing temperature [73]. Unlike
evidenced from Fig. 13(C), the T variation of Raman shift (RS) of both the B2g (7) mode the FWHM of the B1g (5) mode exhibited a monotonous
B2g (7) and B1g (5) modes suffer significant softening at TN (blue dashed increase above 373 K due to anharmonicity as dictated by Eq. (12b).
line in Fig. 13(C)). The cubic anharmonic variation of the RS and line These anomalous features in RS and FWHM occurring at the magnetic
widths (LW) with temperature are described as [70]: and ferroelectric transitions confirms that SPC gives rise to the observed
( ) MD and ferroelectricity in the SYFM (58–42) [5]. Previous reports in cho
2
ω (T) = ω(0) − A 1 + ℏω0 /2KB T (12a) et al. [74] showed thermally stimulated depolarization currents (TSDC)
e − 1
is the actual origin of the pyroelectric current for T ~ 110 K in
( ) YFe0⋅8Mn0⋅2O3 single crystals, thereby generalizing that no long range
2
Γ (T) = Γ(0) + B 1 + (12b) polar order of electric dipoles occur in YFe0⋅6Mn0⋅4O3 [5]. However, in
eℏω0 /2KB T − 1
our present study, the substitution of Sm in high percentage at Y site
where ω(0) and Γ(0) are the intrinsic frequency and linewidth of the induces a true ferroelectric state which is also in the vicinity of the
optical mode due to the defect respectively; A and B are coefficients for compensation point through SPC. Hence we suggest the necessity to
cubic anharmonic processes. It is clear from Fig. 13(C) that Eq. (12a) can re-investigation several doped and undoped RFeO3 systems such as
satisfactorily describe T variation of the RS of the B2g (7) and B1g (5) SmFeO3 [29], regarding occurrence of intrinsic electric polarization in
modes [solid (red) lines] for T ≥ 373 K only. For T < 373 K, significant conjunction with the NM state.
phonon softening occurs in both the Raman modes. With decreasing T, a It is worth notable that there exist certain critical anomalies in the T
dip appeared at TN causing the change in slope in T variation of the variation of linewidths and the phonon frequencies of both the modes at
phonon frequency from usual anharmonicity. With further decreasing T 220 K [indicated as dashed yellow line in Fig. 13(C)–(E)] and around
similar features appeared at TSR1 and TSR2 as well. 153 K [indicated by arrow in Fig. 13(D) and (E)]. Since the temperatures
This implies that q dependent spin-spin correlation arising from the of these anomalies coincides with the slope changes in thermal variation
phase coexistence anomalously affects the optical phonon modes. The of the overall resistivity and grain resistance as shown in Figs. 11 (B) and
competing SPC strengths from AFM/WFM correlations within the ma 10 (G) respectively, it confirms the earlier suggestion of electron local
terial at different temperatures can give rise to anomalous changes in ization caused by electron-phonon interactions at 220 K.
16
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Fig. 13. Panel (A) shows the unpolarised Raman spectra of SYFM (58–42) at several fixed temperatures. Panel (B) shows the B2g (7) and B1g (5) Raman modes at 86
K. Panel (C) shows the T variation of the Raman shift of the B2g (7) and B1g (5) Raman modes between 83 and 506 K. The solid lines are phonon anharmonicity
variation with T and broken lines are guide to the eye. Panel (D) and (E) shows the T variation of the FWHM of the B2g (7) and B1g (5) Raman modes between 83 and
506 K.
4. Conclusion for T below 70 K and existing with the long range ordered magnetic
phase in the material. Robust magnetodielectric effects can be observed
In conclusion, the polycrystalline samples of Sm0.5Y0.5 at RT as illustrated from the magnetic field dependent dielectric con
Fe0⋅58Mn0⋅42O3 below TN exhibits an incomplete second order spin stant that scales linearly to the squared magnetization in the high field
reorientation transition at TSR1 that is immediately completed by a first regime for H > 1.5 T (at 300 K) as described by the Ginzburg-Landau
order spin reorientation at TSR2, leading to completion of the spin theory. One important property exhibited is the occurrence of true
reorientation in to a nearly collinear antiferromagnetic state. The deli ferroelectric polarization whch coulldont be found neither in YFeO3 or
cate interplay between the Sm3+- Fe3+/Mn3+ anisotropic exchange in its Mn doped compositions. Significant spin phonon coupling is
interaction and anisotropy nature of the Mn3+ ions causes the two observed at TN, TSR1 and across FOMT, involving magnetoelectric
consecutive spin reorientation transitions below TN in the present sys coupling and ferroelectricity above liquid N2 temperatures. All-over the
tem. Astonishingly a re-entrant spinglass like state have been observed intercorrelation of the physical properties of the material
17
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
Y0.5Sm0.5Fe0⋅58Mn0⋅42O3 reveals a delicate interplay of the spin, charge [4] I. Fita, A. Wisniewski, R. Puzniak, V. Markovich, G. Gorodetsky, Phys. Rev. B 93
(2016), 184432.
and lattice degrees of freedom, that suggest the material to be a potential
[5] P. Mandal, V.S. Bhadram, Y. Sundarayya, C. Narayana, A. Sundaresan, C.N. Rao,
candidate for multifunctional applications. Phys. Rev. Lett. 107 (2011), 137202.
[6] Y.K. Jeong, J.-H. Lee, S.-J. Ahn, H.M. Jang, Solid State Commun. 152 (2012) 1112.
[7] J. Lohr, F. Pomiro, V. Pomjakushin, J.A. Alonso, R.E. Carbonio, R.D. Sánchez, Phys.
Author contribution Rev. B 98 (2018), 134405.
[8] Dong Gun oh, Jong Hyuk Kim, Hyun Jun Shin, Young jai choi & nara lee, Sci. Rep.
I. S. Raut- Manuscript writing (Draft & Final), work methodology, 10 (2020), 11825.
[9] F. Pomiro, R.D. Sánchez, G. Cuello, A. Maignan, C. Martin, R.E. Carbonio, Phys.
Data Acquisition and analysis. S. Chakravarty- Experimental support,
Rev. B 94 (2016), 134402.
Data Acquisition and analysis H.S Mohanty- - Manuscript writing (Draft [10] J. Wang, J. Liu, J. Sheng, W. Luo, F. Ye, Z. Zhao, X. Sun, S.A. Danilkin, G. Deng,
& Final), Data analysis. S. Mahapatra- Manuscript writing (Draft & Final, W. Bao, Phys. Rev. B 93 (R) (2016), 140403.
work methodology, Experimental support. V., S. Bhardwaj-, Experi [11] J.-H. Lee, Y.K. Jeong, J.H. Park, M.A. Oak, H.M. Jang, J.Y. Son, J.F. Scott, Phys.
Rev. Lett. 107 (2011), 117201.
mental support, Manuscript writing (Draft) A.M. Awasthi- Experimental [12] Juan P. Bolletta, Fernando Pomiro, R.D. Sánchez, V. Pomjakushin, G. Aurelio,
resources, Funding, Manuscript writing (Draft) bimankar92@gmail. A. Maignan, Christine Martin, Raúl E. Carbonio, Phys. Rev. B 98 (2018), 134417.
com- Manuscript writing (Draft & Final), Data Acquisition kpatyal@ [13] K. Dey, A. Indra, S. Mukherjee, S. Majumdar, J. Strempfer, O. Fabelo, E. Mossou,
T. Chatterji, S. Giri, Phys. Rev. B 100 (2019), 214432.
gmail.com, - work methodology, Experimental support mohitsem [14] P. Mandal, C.R. Serrao, E. Suard, V. Caignaert, B. Raveau, A. Sundaresan, C.N.
[email protected] methodology, Experimental support, Manuscript R. Rao, J. Solid State Chem. 197 (2013) 408.
writing (Draft). [email protected] methodology, Data [15] S. Raut, S. Chakravarty, H.S. Mohanty, S. Mahapatra, S. Bhardwaj, A.M. Awasthi,
B. Kar, K. Singh, M. Chandra, A. Lakhani, V. Ganesan, M. Mishra Patidar, R.
Acquisition and analysis [email protected] Data Acquisi K. Sharma, Velaga Srihari, H.K. Poswal, S. Mukherjee, S. Giri, S. Panigrahi,
tion and analysis. Xii. [email protected] Data Acquisition and J. Magn. Magn Mater. 563 (2022), 169950.
analysis, experimental resources. [email protected] Data [16] R.A. Young, The Rietveld Method, Oxford University Press, Oxford, 1993.
[17] Subhajit Raut, P. D Babu, R.K. Sharma, R. Pattanayak, S. Panigrahi, J. Appl. Phys.
Acquisition and analysis, experimental resources. [email protected]. 123 (2018), 174101.
in- Data Acquisition and analysis. [email protected] Data [18] Weiyao Zhao, Shixun Cao, Ruoxiang Huang, Yiming Cao, Kai Xu, Baojuan Kang,
Acquisition and analysis. [email protected] Data Acquisition and Jincang Zhang, Wei Ren, Phys. Rev. B 91 (2015), 104425.
[19] A.H. Cooke, D.M. Martin, M.R. Wells, J. Phys. C Solid State Phys. 7 (1974)
analysis, experimental resources, Manuscript writing (Draft & Final xvii.
3133–3144.
[email protected] Funding, Manuscript writing (Draft & Final), [20] S. Harikrishnan, C M Naveen Kumar, H.L. Bhat, Suja Elizabeth, U.K. Rößler,
work methodology. K. Dörr, S. Rößler, S. Wirth, J. Phys. Condens. Matter 20 (2008), 275234.
[21] K. Dey, S. Majumdar, S. Giri, Phys. Rev. B 90 (2014), 184424.
[22] S. Chattopadhyay, S. Giri, S. Majumdar, V. Ganesan, D. Venkateshwarlu, J. Appl.
Declaration of competing interest Phys. 115 (2014) 17E104.
[23] Subhajit Raut, Somnath Mahapatra, Simanchalo Panigrahi, J. Magn. Magn Mater.
529 (2021), 167887.
The authors declare the following financial interests/personal re [24] R. Tackett, G. Lawes, Phys. Rev. B 76 (2007), 024409.
lationships which may be considered as potential competing interests: [25] W. Chen, W. Zhu, O.K. Tan, X.F. Chen, J. Appl. Phys. 108 (2010), 034101.
[26] Shivani Sharma, Tathamay Basu, Aga Shahee, 1 K. Singh, N.P. Lalla 1, E.
Simanchalo Panigrahi reports financial support, administrative support,
V. Sampathkumaran, Phys. Rev. B 90 (2014), 144426.
article publishing charges, equipment, drugs, or supplies, statistical [27] K Devi Chandrasekhar, A.K. Das, A. Venimadhav, J. Phys. Condens. Matter 24
analysis, travel, and writing assistance were provided by University (2012), 376003.
[28] Md G. Masud, Arijit Ghosh, J. Sannigrahi, B.K. Chaudhuri, J. Phys. Condens. Matter
Grants Commission Department of Atomic Energy Consortium for Sci
24 (2012), 295902.
entific Research. Simanchalo Panigrahi reports a relationship with [29] Smita Chaturvedi, Priyank Shyam, Amey Apte, Jitender Kumar,
University Grants Commission Department of Atomic Energy Con Arpan Bhattacharyya, A.M. Awasthi, Sulabha Kulkarni, Phys. Rev. B 93 (2016),
sortium for Scientific Research that includes: funding grants and travel 174117.
[30] V.V. Shvartsman, J. Zhai, W. Kleemann, Ferroelectrics 379 (2009) 77–85.
reimbursement. [31] D.C. Arnold, F.D. Morison, J. Mater. Chem. 19 (2009) 6485–6488.
[32] Chi Kao Kwan, DIELECTRIC PHENOMENA IN SOLIDS, Elsevier Academic Press,
Data availability London, 2004.
[33] K.S. Cole, R.H. Cole, J. Chem. Phys. 9 (1941) 341.
[34] S. Kumari, N. Ortega, A. Kumar, S.P. Pavunny, J.W. Hubbard, C. Rinaldi,
Data will be made available on request. G. Srinivasan, J.F. Scott, R.S. Katiyar, J. Appl. Phys. 117 (2015), 114102.
[35] Shalini Kumari, Dhiren K. Pradhan, Proloy T. Das, Nora Ortega, Kallol Pradhan,
Ashok Kumar, J.F. Scott, Ram S. Katiyar, J. Appl. Phys. 122 (2017), 144102.
Acknowledgement [36] M. Idrees, M. Nadeem, M.M. Hassan, J. Phys. D Appl. Phys. 43 (2010), 155401.
[37] N.F. Mott, J. Non-Cryst. Solids 1 (1969) 1.
S.R. acknowledges Dr. Srihari Velaga for support in the high tem [38] R.P. Maiti, S. Dutta, M.K. Mitra, D. Chakravorty, J. Phys. D Appl. Phys. 46 (2013),
415303.
perature synchrotron measurements. We acknowledges UGC-DAE CSR [39] S. Pal, A. Banerjee, E. Rozenberg, B.K. Chaudhuri J, Appl. Phys. 89 (2001) 4955.
Indore for measurement support. S.R. and S.P. also acknowledges Mr. [40] N. Nakamura, E. Eguchi, Solid State Chem 145 (1999) 58.
Balram Thakur for helping in SQUID-VSM measurements in UGC-DAE [41] N.F. Mott, E.A. Davis, Conduction in Non-crystalline Materials, Clarendon, Oxford,
1971.
CSR Kalapakkam Node. The whole work was supported by the Minis [42] B.I. Shklovskii, A.L. Efros, Electronic Properties of Doped Semiconductors,
try of Human Resource and Development (MHRD), India, under the Springer, Berlin, 1984.
indigenous fellowship scheme. This work is also partially funded by [43] M. Viret, L. Ranno, J.M.D. Coey, Phys. Rev. B 55 (1997) 8067.
[44] J.M. De Teresa, M.R. Ibarra, J. Garc′ ıa, J. Blasco, C. Ritter, P.A. Algarabel,
project approved under collaborative research scheme (UGC DAE- CRS)
C. Marquina, A. del Moral, Phys. Rev. Lett. 76 (1996) 3392.
vide project no.: CSR–IC–255/2017–18/1336 dated 31-03- 2018. [45] W. Khan, A.H. Naqvi, M. Gupta, S. Husain, R. Kumar, J. Chem. Phys. 135 (2011),
054501.
[46] I. Ahmad, M.J. Akhtar, R.T.A. Khan, M.M. Hasan, J. Appl. Phys. 114 (2013),
Appendix A. Supplementary data 034103.
[47] Young Sun, Xiaojun Xu, Yuheng Zhang, J. Phys. Condens. Matter 12 (2000)
Supplementary data to this article can be found online at https://doi. 10475–10480.
[48] A.J. Millis Phil, Trans. R. Soc. Lond. A 356 (1998) 1473.
org/10.1016/j.physb.2022.414593.
[49] V.H. Crespi, Li Lu, Y.X. Jia, K. Khazeni, A. Zetll, M.L. Cohen, Phys. Rev. B 53
(1996), 14303.
References [50] M. Jaime, M.B. Salamon, K. Pettit, M. Rubinstein, R.E. Trece, J.S. Horwitz, D.
B. Chirisoy, Appl. Phys. Lett. 68 (1996) 1576.
[51] Sindhunil Barman Roy, J. Phys. Condens. Matter 25 (2013), 183201.
[1] T. Yamaguchi, K. Tsushima, Phys. Rev. B 8 (1973) 5187.
[52] A. Lakhani, A. Banerjee, P. Chaddah, X. Chen, R.V. Ramanujan, J. Phys. Condens.
[2] T. Yamaguchi, J. Phys. Chem. Solid. 35 (1974) 479.
Matter 24 (2012), 386004.
[3] S. Cao, H. Zhao, B. Kang, J. Zhang, W. Ren, Sci. Rep. 4 (2014) 5960.
18
S. Raut et al. Physica B: Condensed Matter 651 (2023) 414593
[53] S.B. Roy, G.K. Perkins, M.K. Chattopadhyay, A.K. Nigam, K.J.S. Sokhey, [65] T. Kimura, S. Kawamoto, I. Yamada, M. Azuma, M. Takano, Y. Tokura, Phys. Rev. B
P. Chaddah, A.D. Caplin, L.F. Cohen, Phys. Rev. Lett. 92 (2004), 147203. 67 (R) (2003), 180401.
[54] M. Manekar, C. Mukherjee, S.B. Roy, EPL 80 (2007), 17004. [66] T.D. Sparks, M.C. Kemei, P.T. Barton, R. Seshadri, E.-D. Mun, V.S. Zapf, Phys. Rev.
[55] Y. Imry, M. mortis, Phys. Rev. B 19 (1979) 3580–3585. B 89 (2014), 024405.
[56] A. Banerjee, A.K. Pramanik, Kranti Kumar, P. Chaddah, J. Phys. Condens. Matter [67] N. Mufti, A.A. Nugroho, G.R. Blake, T.T.M. Palstra, J. Phys. Condens. Matter 22
18 (2006) L605–L611. (2010), 075902.
[57] A. Banerjee, Kranti Kumar, P. Chaddah, J. Phys. Condens. Matter 21 (2009), [68] N. Terada, Y.S. Glazkova, A.A. Belik, Phys. Rev. B 93 (2016), 155127.
026002. [69] B. Poojitha, A. Rathore, A. Kumar, S. Saha, Phys. Rev. B 102 (2020), 134436.
[58] G. Catalan, Appl. Phys. Lett. 88 (2006), 102902. [70] J.S. Bae, I.S. Yang, Phys. Rev. 73 (2006), 052301.
[59] G. Catalan, D. O’Neill, R.M. Bowman, J.M. Gregg, Appl. Phys. Lett. 77 (2000) [71] P. Prakash, V. Sathe, C.L. Prajapat, A.K. Nigam, P.S.R. Krishna, A. Das, A. Pal,
3078. S. Ghosh, A.G. Joshi, Shiv Kumar, S. Patil, P.K. Gupta, P. Singh, V.K. Gangwar,
[60] G. Lawes, A.P. Ramirez, C.M. Varma, M.A. Subramanian, Phys. Rev. Lett. 91 P. Prakash, R.K. Singh, E.F. Schwier, M. Sawada, K. Shimada, A.K. Ghosh, A. Das,
(2003), 257208. S. Chatterjee, J. Phys. Condens. Matter 31 (2019), 275802. J. Phys.: Condens.
[61] A. von Hippel, Dielectrics and Waves, Artech House, London, 1995. Matter 32, 095801 (2020).
[62] B. Ramachandran, A. Dixit, R. Naik, G. Lawes, M.S. Ramachandra Rao, Appl. Phys. [72] J. Laverdière, S. Jandl, A.A. Mukhin, V. Yu Ivanov, V.G. Ivanov, M.N. Iliev, Phys.
Lett. 100 (2012), 252902. Rev. B 73 (2006), 214301.
[63] Md F. Abdullah, P. Pal, S. Lal, S.R. Mohapatra, K. Chandrakanta, S.D. Kaushik, C. [73] Somdutta Mukherjee, Ashish Garg, Rajeev Gupta, J. Phys. Condens. Matter 23
S. Yadav, A.K. Singh, Mater. Res. Express 5 (2018), 116101. (2011), 445403.
[64] J.K. Dey, A. Chatterjee, S. Majumdar, A.-C. Dippel, O. Gutowski, M. [74] Kwanghee Cho, Soomin Hur, Soonyong Park, Appl. Phys. Lett. 110 (2017), 162905.
v. Zimmermann, S. Giri, Phys. Rev. B 99 (2019), 144412.
19