0% found this document useful (0 votes)
23 views11 pages

Real Hybrid Composition

This study investigates the energy storage performance of strontium and manganese co-doped 0.925(Bi1/2Na1/2)TiO3-0.075BaTiO3 ceramics, revealing that Sr doping enhances relaxor ferroelectric behavior while reducing remanent polarization and coercive field. The co-doping with manganese significantly improves energy storage density to 0.90 J/cm3 and thermal stability, attributed to the formation of defect-dipoles that create an internal field and constrict the hysteresis loop. The findings highlight the potential of BNT-based ceramics for efficient energy storage applications through compositional adjustments and structural evolution.

Uploaded by

anil
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
23 views11 pages

Real Hybrid Composition

This study investigates the energy storage performance of strontium and manganese co-doped 0.925(Bi1/2Na1/2)TiO3-0.075BaTiO3 ceramics, revealing that Sr doping enhances relaxor ferroelectric behavior while reducing remanent polarization and coercive field. The co-doping with manganese significantly improves energy storage density to 0.90 J/cm3 and thermal stability, attributed to the formation of defect-dipoles that create an internal field and constrict the hysteresis loop. The findings highlight the potential of BNT-based ceramics for efficient energy storage applications through compositional adjustments and structural evolution.

Uploaded by

anil
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 11

Materials Science and Engineering B 263 (2021) 114869

Contents lists available at ScienceDirect

Materials Science & Engineering B


journal homepage: www.elsevier.com/locate/mseb

Improved energy storage capacity in strontium and manganese co-doped


0.925(Bi1/2Na1/2)TiO3-0.075BaTiO3 ceramics
Pin-Yi Chen a, Jesuraj Anthoniappen b, *, Yi-Tsung Lee a, Chi-Shun Tu c, Cheng-Sao Chen d,
Kuei-Chih Feng a, Flora Mae Ruiz b
a
Department of Mechanical Engineering, Ming Chi University of Technology, New Taipei City 24301, Taiwan
b
Department of Physics, University of San Carlos, Cebu City 6000, Philippines
c
Department of Physics, Fu Jen Catholic University, New Taipei City 24205, Taiwan
d
Department of Mechanical Engineering, Hwa Hsia University of Technology, New Taipei City 23567, Taiwan

A R T I C L E I N F O A B S T R A C T

Keywords: Inducing slimmer P-E loops while maintaining high ferroelectric polarization is a major challenge in dielectric
Defect-dipoles capacitor applications. In this work, strontium (Sr) doped and strontium-manganese (Sr, Mn) co-doped 0.925
Constricted hysteresis loop (Bi1/2Na1/2)TiO3-0.075BaTiO3 (BNT7.5BT) ceramics were prepared to study energy storage performance. Sr-
Surface potential
doped BNT7.5BT samples showed enhanced relaxor ferroelectric behavior and reduced remanent polarization
Polarization maximum
Energy storage density
(Pr) and coercive field (Ec) which can be attributed to the disruption of ferroelectric long-range order by the
presence of non-polar cubic Pm3− m phase. (Sr, Mn) co-doped BNT7.5BT samples revealed improved energy
storage density of 0.90 J/cm3 and thermal stability, affirming the vital role of Mn ions. The presence of Mn′′Ti − VOoo
defect-dipoles can create a dipole moment (PD) which can result in an internal field causing a constricted
ferroelectric hysteresis loop with low Pr and Ec. Surface screening of polarization bound charges by Mn-induced
defects affects the overall Pmax contribution to energy storage density.

1. Introduction and higher electrical breakdown field.


Unlike ferroelectrics, relaxor ferroelectrics (RFEs) possess low
Energy storage materials have been receiving exquisite attention remnant polarization and slim hysteresis loop, which can provide high
from researchers and power industry for efficient energy-storage devices saturated polarization and superior energy conversion efficiency with
such as super-capacitors toward high power and energy densities [1]. It high discharge energy density and fast discharge ability [27]. Despite
has been found that dielectric capacitors have merits of high-power the remarkable progress made in investigating lead-free RFEs, the thrust
density, short charge/discharge and high cycling reliability compared for ceramic capacitor materials to achieve high recoverable energy
to chemical batteries and electrochemical capacitors [2–5]. Unfortu­ density and optimal efficiency is still enormous. For instance, Song et al.
nately, their low energy storage capacity due to limited usable surface found energy density increasing from 0.77 to 1.15 J/cm3 and energy
charges is a hinderance for applications. Therefore, significant efficiency from 60 to 82% in (Ba0.4Sr0.6)TO3 (BST) ceramics synthesized
improvement in energy density and energy efficiency of the ceramic by microwave sintering due to low oxygen vacancies at grain boundaries
capacitors has become the focus among many researchers [6–14]. The [28]. Zheng et al. found excellent energy-storage density of 1.56 J/
choice of material system used in these devices plays a critical role as the cm3 and a moderate energy-storage efficiency of 75% in 0.61BiFeO3-
scientific community is persistently venturing into green energy tech­ 0.33BaTiO3-0.06Ba(Mg1/3Nb2/3)O3 ceramics at 25–190 ◦ C. They
nologies to address serious concerns of health and environment [15–18]. attributed such enhanced properties to an E-field induced reversible
Lead-free perovskite antiferroelectrics [19,20], dielectric glass-ceramics ergodic-FE phase transition owing to similar free energies near a critical
[21,22], relaxor ferroelectric (FE) [23,24], and polymer-based ferro­ freezing temperature [29]. 2% Nb2O5-doped BiFeO3-BaTiO3 (BF-BT)
electrics [25,26] have been considered most promising materials own­ exhibited maximum discharge energy density of 0.71 cmJ 3 at 90 kV/cm
ing to their high saturation polarization, smaller remnant polarization due to reduction of leakage current [30]. Qi et al. found NaNbO3-doped

* Corresponding author at: Department of Physics, University of San Carlos, Talamban Campus, Cebu City 6000, Philippines.
E-mail address: janthoniappen@usc.edu.ph (J. Anthoniappen).

https://doi.org/10.1016/j.mseb.2020.114869
Received 1 July 2020; Received in revised form 14 September 2020; Accepted 5 October 2020
Available online 15 October 2020
0921-5107/© 2020 Elsevier B.V. All rights reserved.
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

Fig. 1. SEM micrographs of (0.925-x)BNT-0.075BT-xSrTiO3 (x = 0, 2, 10, and 20) ceramics with room-temperature XRD patterns.

Bi0.5Na0.5TiO3-BaTiO3 system exhibiting optimal energy storage density achieved an enhanced energy storage density even at elevated temper­
of 0.71 J/cm3 at 7 kV/cm due to increasing defect dipoles and ature by inducing defect-dipoles which greatly reduced Pr and Ec while
decreasing rhombohedral and tetragonal polar phases with a good maintaining high maximum polarization Pmax. The work proposed an
thermal stability at 25–150 ◦ C [31]. insight into enhancing energy-storage density by exploring structural
BNT-based ferroelectrics have been considered to be the most po­ evolution and role of defect-dipoles.
tential candidates in energy storage application as relaxor properties can
be tailored by composition variation [32]. The major drawback in these 2. Experimental procedure
systems is that though a slim P-E loop can be reached due to ergodic
relaxor phase, the Pmax decreases sharply as the Pr minimizes, which will Step 1: Polycrystalline (0.925-x)BNT-xSrTiO3-0.075BT(x = 0, 0.02,
reduce the energy storage density. Several researchers have adopted the 0.1, and 0.2) ceramics were prepared via solid state reaction method by
compositional variation method to improve the energy storage in BNT using Bi2O3 (99.99%), Na2CO3 (99.95%), BaCO3 (99%), TiO2 (99.99%),
based systems. Recently, Sr-doped (Bi0.70Na0.67Li0.03)0.5TiO3 system and SrCO3 (99.5%) as starting reactants. Stoichiometric amounts of
exhibited high energy density of 1.06 J/cm3 due to breaking of long- powders were mixed with zirconia balls in an agate mortar for more than
range FE order and presence of polar nano-regions (PNRs) affected by 24 h using ethanol as a medium. Subsequently, the dried powders were
A-site substituted Sr2+ ions, octahedral distortion (caused by Li+ ions), calcined at 900 ◦ C for 2 h and subjected to ball milling again before
field-induced PNRs rotations, and PNR-related reversible transitions sintering. The powder mixture was then dried, meshed and pressed into
[33]. Cao et al. proposed that introduction of defect dipoles via appro­ a 1.0 cm-diameter disk for sintering at 1175 ◦ C for 4 h. The as-sintered
priate doping can induce large recoverable strain with very little hys­ samples were polished and annealed at 600 ◦ C for 2 h in order to remove
teresis from which energy density can be improved in BNT-based residual stress.
systems [34]. Da Li et al. achieved outstanding charge–discharge per­ Grain morphologies were measured using a high-resolution field
formances in BNST-based bulk ceramics by introducing the linear ad­ emission scanning electron microscope (FE-SEM; HITCHI S-3400N). To
ditive of Sr(Ti0.85Zr0.15)O3. They attributed the ultrahigh energy storage quantify oxygen compositions, average oxygen atomic ratios were
performance to enhanced break down strength [35]. Y. Lin et al also measured from 10 to 20 spots for each compound using SEM equipped
achieved very high energy-storage capacity in (Na0.5Bi0.5)0.7Sr0.3TiO3 with wavelength dispersive spectrometry (WDS; Oxford Instruments
modified by Bi(Mg2/3Nb1/3)O3 ceramics by refining the grain size and Nanotechnology Tools Ltd, UK). X-ray diffraction spectra of as-sintered
introducing Bi3+’s lone pair electron 6 s2 configuration [36]. Similarly, ceramics were obtained using a Rigaku Multiplex Diffractometer with a
( )
reduced grain size and high relative density were also shown to be radiation wavelength Kα1 1.5406 Ȧ The Rietveld refinements were
responsible for enhanced energy-storage performance in BNT-based
performed using HighScore Plus software to determine structure space
system [37].
groups. Phase structure analysis was also carried out by using high-
In this work, the contribution of Mn4+ doping to the enhanced en­
resolution transmission electron microscope (TEM; JEOL JEM-2100
ergy storage capacity has been discussed. In the first step, various con­
LaB6) equipped with an energy-dispersive spectrometer. Wayne-Kerr
centrations of Sr2+ doped 0.925(Bi1/2Na1/2)TiO3-0.075BaTiO3
Analyzer PMA3260A was used to obtain dielectric permittivity at
(hereafter abbreviated as BNT7.5BT) was synthesized to study dielec­
zero-field heating and cooling. Temperature evolution of polarization
tric, ferroelectric and energy storage capacities. In the second step, Mn4+
vs. E field (P-E hysteresis loop) was performed using Sawyer-Tower
(2%) and Sr2+ (20%) co-doped BNT7.5BT was synthesized to study its
circuit at f = 4 Hz.
energy storage capacity. Our results showed that doping 20% Sr2+ ions
Step 2: (Sr, Mn) co-doped (Bi0.5Na0.5)0.725Sr0.20.075[BaTi0.98Mn0.02]
significantly reduced FE nature and enhanced relaxor behavior by dis­
O3 (20Sr2Mn) ceramics were prepared via solid-state reaction method
rupting long-range FE order of the parent BNT7.5BT by the presence of
by using MnO2 and other reactants as starting materials mentioned
non-polar cubic Pm3− m phase. Further doping with 2% Mn4+ ions

2
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

Fig. 2. Frequency and temperature-dependent dielectric spectra of (a) BNT7.5BT (b) 20Sr. (c) Room-temperature P-E loops and (d) P-E hysteresis loop parameters.

above. Same procedure was followed to obtain as-sintered 20Sr2Mn


Table 1
samples. To analyze Mn valence, synchrotron X-ray absorption spectra
P-E hysteresis parameters for various doping Sr2+ concentrations.
of Ti L-edge, Mn L-edge, and oxygen K-edge were measured at the
BL20A1 of National Synchrotron Radiation Research Center (NSRRC, %Sr Pr (μC/cm2 ) Pm (μC/cm2 ) Ec (kV/cm)
Taiwan). Frequency modulated-Kelvin probe force microscopy (FM- 0 42.42 49.64 22.69
KPFM) surface images were scanned in an ambient condition using a 2 34.04 42.27 24.09
scanning probe microscope (SPM, Brucker MultiMode 8) equipped with 10 27.89 34.63 21.95
antimony-doped Si conductive probe (force constant ~5 N/m and tip 20 23.83 33.86 13.63

radius ~12 nm) to evaluate surface potential. FM-KPFM is a tapping


mode AFM in which single pass-technique is used to give higher spatial favor the disruption of long-range FE order which is evident by a lower
resolution. It applies an AC signal to the probe at a low frequency (fm = Tm and pronounced frequency dispersion below Tm compared to undo­
5 kHz) while mechanically driving the probe at its resonance frequency ped BNT7.5BT. The chemical disorder induced by different ions produce
(f0 = 100 kHz). Samples were polished to 0.2 mm and silver paint was a competition between the long-range dipoles and the translational
used as contact between single edges of sample surface and sample puck. symmetry, which tends to enhance the random fields. Local random
Same sample was used to probe local out-of-plane (OP) phase and bonds generate the PNRs that will create typical dielectric relaxation
amplitude hysteresis loops using piezoresponse force microscopy (PFM) properties [32]. The polarization hysteresis (P-E loop) displays a well-
contact mode. Parameters used for the surface scanning were tip height
saturated typical FE behavior for x = 0 with Pm ~ 49.64 μC/cm2 , Pr
= 5 nm, tip velocity = 2.99 μm/s, tip bias = +10 V, and scan rate =
~ 42.42 μC/cm2 , and Ec ~ 22.69 kV/cm, and drops significantly with
0.498 Hz.
increasing Sr2+ concentration as shown in Fig. 2(c). Important P-E
hysteresis parameters for various Sr2+ concentrations are shown in
3. Results and discussion
Table 1. The reduction of polarization implies the degradation of
ferroelectricity due to the disrupted FE long-range order by the presence
The SEM micrographs of Sr-doped BNT7.5BT ceramics in Fig. 1
of dominant non-polar cubic Pm3− m phase revealed by the Rietveld
reveal dense grain distribution with well-defined grain boundaries,
refinement in Fig. 3(b) (20Sr as a reference composition). The phase
indicating that the addition of Sr can inhibit grain growth. The XRD
analysis for BNT7.5BT indicates 46.15% rhombohedral R3c phase and
spectra in the insets reveal a single perovskite structure without sec­
53.85% tetragonal P4bm phase as shown in Fig. 3(a).
ondary phase, suggesting good incorporation of Sr2+ ions into BNT7.5BT
The local symmetry analysis by TEM also reveals long range rhom­
lattice. Fig. 2 shows temperature dependent dielectric permittivity (ε’ ),
bohedral R3c and nanodomain tetragonal P4bm phase coexistence by
loss tangent (tanδ) for undoped BNT7.5BT and BNT7.5BT-20% Sr
presence of 1/2(ooo) and 1/2(ooe) superlattice reflections respectively
(hereafter abbreviated as 20Sr) at frequencies of 10 kHz− 1 MHz, and P-
as shown in Fig. 4(a) and (b) (o and e stand for odd and even miller
E loops for 0–20% Sr specimens. The permittivity spectra for 20Sr reveal
indices). It is consistent with the presence of nanoscale P4bm platelets
enhanced relaxor behavior as it showed increased frequency dispersion
embedded in the R3c matrix of pure Bi1/2Na1/2TiO3 in the previous
and decreased ε’ . Compared to BNT7.5BT, 20Sr displays dielectric studies [38,39]. Though the refinement result of 20Sr showed multi-
maximum temperature (Tm) shifting to lower temperature. The Sr2+ ions phase coexistence (Fig. 3b), its local symmetry analysis in Fig. 4(c)

3
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

R3c and P4mm phases increase significantly after an E-field poling while
Pm3− m and P4bm decrease.
To evaluate the capability of energy storage, the energy storage
density (Wrec) can be calculated from the P-E loop using the equation
[40,41],
∫ Pmax
Wrec = EdP (1)
Pr

where E is the applied electric field and Wrec can be obtained by inte­
grating the area between the polarization axis and the discharge curve
from the P-E loops. The energy storage efficiency (η) is calculated using
the equation [40,42],
Wrec
η(%) = × 100 (2)
Wrec + Wloss

where Wloss is the energy loss density which is obtained by integrating


the area between the charge and discharge curves. As for the energy
storage devices (ex. capacitors), the energy-storage density can be
improved either by enhancing electric breakdown strength (BDS) or by
increasing the limit of integration by enlarging difference between Pmax
and Pr [34]. It implies that poling a material around its BDS can result in
maximum energy storage capacity. According to Cao et al. [34], the
electric breakdown strength for BNT-BT based solid solutions is ~82–95
kV/cm. Fig. 5 shows the room temperature P-E hysteresis loops of 20Sr
obtained at 60 and 100 kV/cm with energy storage densities of 0.21 and
0.47 J/cm3, respectively, which are smaller than some BNT-BT based
Fig. 3. Rietveld refinements of (a) BNT7.5BT and (b) 20Sr ceramics. Quantitate
analysis of phase structure and refinement fitting parameters of are given in systems [43–47].
the insets. To improve energy storage density, 2% Mn4+ ions are incorporated
into the B-site as excessive Mn doping could result in reduced FE
properties due to valence shift and oxygen vacancies. In the following
does not reveal any superlattice reflection, demonstrating dominant
section, the presence and effects of defect dipoles in 2% Mn and 20% Sr
cubic Pm3− m phase in the unpoled state. After a prior E-field poling,
co-doped BNT7.5BT (hereafter abbreviated as 20Sr2Mn) are discussed
nanodomains begin to coalesce into lamellar domains indicated by the
in terms of phase structure, valence, surface potential, dielectric and FE
circle in Fig. 4(d). It signifies that the static nanodomains are irreversibly
properties. Energy storage parameters are also evaluated as functions of
switched by the E-field poling and consequently long-range polar order
temperature.
development could be evidenced by the appearance of 1/2 (0 0 0)
Fig. 6(a) and (b) shows grain morphologies for 20Sr and 20Sr2Mn
superlattice reflections along the [1 1 2] direction. The increasing polar
respectively. To analyze oxygen concentration at grain interior and
phase is further demonstrated by the (2 0 0) diffraction peak at 60 kV/cm
boundaries, micrographs and chemical compositions from several
poling as compared to the unpoled (2 0 0) peak in Fig. 4(e) and (f). The

Fig. 4. High-resolution TEM images of (a)-(b) BNT7.5BT (c) unpoled 20Sr (d) poled 20Sr ceramics. Insets are their corresponding SADP images. (e) and (f) Rietveld
refinements of (2 0 0) XRD patterns of unpoled and poled 20Sr, respectively.

4
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

Ti4+ ions and some Mn2+ ions may disperse at grain boundaries after
sintering [49,50].
Fig. 7 shows the normalized Mn L-edge X-ray absorption spectra
(XAS) for 20Sr2Mn ceramic with references of MnCO3 (Mn2+), Mn2O3
(Mn3+), and MnO2 (Mn4+). The Mn L-edge absorption is due to the
transitions from the Mn 2p to unoccupied Mn 3d orbitals. Absorption
peaks at ~640–645 eV and ~652–657 eV correspond to 2p3/2 →3dand
2p1/2 →3dtransitions, respectively. The tetravalent MnO2 powders
display two major absorption peaks at ~640 and ~642.5 eV, trivalent
Mn2O3 powders reveal single major peak at ~640 eV and divalent
MnCO3 powders reveal three major peaks at ~640, ~641.3, and
~643.3 eV. Based on the fitting result, a mixture of Mn2+, Mn3+, and
Mn4+ valence states were observed and confirms a valence shift from
Mn4+ to Mn3+ and Mn2+. A first-principles calculation suggests that Mn
ion can isovalent substitute for Ti as Mn4+ in the oxidizing limit,
whereas in the reducing limit, Mn can act as an acceptor-type as either
Mn3+ or Mn2+ compensated by the formation of oxygen vacancies [51].
Fig. 8 shows normalized Ti L-edge synchrotron X-ray absorption
spectra caused by transitions from Ti 2p to unoccupied Ti 3d orbitals
[52]. The coupling between Ti 3d and the O 2p orbitals in the TiO6 oc­
tahedron causes splitting Ti 3d orbitals into triple-degenerate low-en­
ergy t2g and doubly-degenerate high-energy eg orbitals [52]. The pair of
peaks (A1, B1) and (A2, B2) represent (t2g, eg) absorption bands of 20Sr
and 20Sr2Mn, respectively. The absorption energies of t2g and eg bands
corresponding to 20Sr and 20Sr2Mn show same as that of BaTiO3 ce­
ramics implying the stability of tetravalent Ti4+ state. The higher peak
intensity ratio of B2/A2 compared to the ratio of B1/A1 suggest a
decreased O 2p-Ti 3d orbital hybridization in 20Sr2Mn, revealing a
weaker Ti-O bond caused by Mn doping.
The spontaneous polarization in FE materials can induce surface
bound charges, creating depolarizing field which is detrimental to FE
stability in the material. This has to be compensated by screening of the
charges at the surface to achieve a thermodynamically stable state. The
field screening due to the presence of defects leads to asymmetric space
charge regions near the grain boundary, resulting in an effective dipole
layer at the grain surface. The interaction between the polarization and
Fig. 5. P-E hysteresis loops of 20Sr poled at (a) 60 kV/cm and (b) 100 kV/cm. screen charges gives rise to the surface electrostatic potential which
could be either sign [53]. Based on the semiconductor model of ferro­
electrics, potential peaks are due to positive and negative surface bound
surface area were obtained using wavelength dispersive spectrometry.
charges in the absence of charge carriers and defects. Whereas, in the
The average grain sizes and atomic ratios of different elements at grain
presence of defects and charge carriers, the potential assumes asymp­
interior and grain boundaries are shown in the Table 2. Apparently, less
totic value and produces potential step along the domain symmetry axes
oxygen contents exist at grain interior and grain boundaries for
[53]. Such a potential has significant impact on both ionic and electronic
20Sr2Mn sample compared to 20Sr sample, implying more oxygen va­
transport in FE ceramics.
cancies. An increase in the average grain size in 20Sr2Mn can be
To gain insight into electrostatic potential across grain boundaries,
attributed to different oxygen vacancies at the grain boundaries. It has
FM-KPFM measurements were carried out for 20Sr and 20Sr2Mn sam­
been suggested that grain boundaries have a higher energy level and the
ples. Fig. 9 shows the topography, potential scan maps, and potential
presence of higher oxygen vacancy concentration can promote the grain
line profiles obtained along aqua color lines in Fig. 9(b) and (e). Both
boundary migration, which leads to grain growth [48].
20Sr and 20Sr2Mn resulted in negative surface potential suggesting that
The Rietveld-refinement analysis of X-ray diffraction pattern of
polarization bound charges are positive and screen charges are negative
20Sr2Mn in Fig. 6(c) shows significantly more polar R3c (28.76%) and
probably pertaining to oxygen defects. It is found that grain boundaries
P4bm (33.16%) phases than 20Sr (10.50% R3c phase and 13.20% P4bm
result in lower surface potential as compared to the grain interiors,
phase) (in Fig. 3), indicating enhanced FE nature. It is noted that (1 1 1)
indicating non-uniform distribution of screen charges. Grain boundary
and (2 0 0) peaks shifts to the lower angles with Mn doping, manifesting
can act as a conduction path in ferroelectrics as electrical charge can
an enlarged lattice parameters in 20Sr2Mn. This result suggests that Mn
easily move through grain boundaries. Eventually, screen charge
ions possibly enter the B-site cation vacancies, leading to increased
draining at grain boundary will lead to form a potential pit with a lower
interplanar spacing. The shifting of the XRD peaks to lower angle in Mn
surface potential [54]. Compared to 20Sr, 20Sr2Mn exhibits lower po­
doping is also intrinsically connected to the smaller ionic radii of Mn
tential, indicating increased defect concentrations. It has been reported
(0.82 Å, 0.65 Å and 0.53 Å for Mn2+, Mn3+, and Mn4+ respectively)
that Mn2+ and Mn3+ ions substituting for Ti4+ valence ions can induce
compared to Sr2+ (~1.13 Å) ions. Larger ionic radii result in smaller (( )× ) ( )o
intrinsic di-valent MnTi′′ − Vooo and tri-valent MnTi − Vooo

lattice parameter and smaller ionic radii result in larger lattice param­
eter shifting the XRD diffraction peaks to higher and lower angles defect dipoles along with oxygen vacancies [53,55]. These defect di­
respectively. In MnO2-doped BNT7.5BT, three species of ions (Mn2+, poles play crucial roles on stabilization of spontaneous polarization by
Mn3+, and Mn4+) can exist due to instability of Mn valence during heat aligning with spontaneous polarization and thus reduce the electrostatic
treatment. Although Mn presents as tetravalent Mn4+ in the oxide MnO2, potential on the surface [55]. This is in agreement with the previous
it is likely that Mn3+ ions enter the B-site due to similar ionic radii with study [56], which showed the non-monotonic decrease of surface

5
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

Fig. 6. SEM micrographs of (a) 20Sr and (b) 20Sr2Mn ceramics. (c) Rietveld refinement of 20Sr2Mn and fitting parameters. Insets are the (1 1 1) and (2 0 0)
XRD peaks.

Table 2
Average grain size and atomic ratios of different elements at grain interiors and
grain boundaries for 20Sr and 20Sr2Mn ceramics.
Composition Average grain Average atomic ratio Average atomic ratio
size (%) at grain interiors (%) at grain
boundaries

20Sr 2.76 ± 0.08 μm O = 57.7 ± 1.5 O = 52.7 ± 2.1


Na = 7.1 ± 0.1 Na = 7.0 ± 0.4
Bi = 7.2 ± 0.3 Bi = 8.3 ± 0.3
Sr = 4.3 ± 0.2 Sr = 4.7 ± 0.2
Ba = 1.6 ± 0.1 Ba = 1.8 ± 0.1
Ti = 22.0 ± 1.0 Ti = 25.5 ± 2.0

20Sr2Mn 4.96 ± 0.12 μm O = 53.6 ± 0.6 O = 50.2 ± 0.8


Na = 7.8 ± 0.11 Na = 6.5 ± 0.3
Bi = 7.7 ± 0. Bi = 7.9 ± 0.4
Sr = 4.2 ± 0.1 Sr = 4.2 ± 0.2
Ba = 1.9 ± 0.1 Ba = 2.1 ± 0.1
Ti = 24.5 ± 0.4 Ti = 26.8 ± 0.7
Mn = 0.3 ± 0.1 Mn = 0.3 ± 0.1

potential with increasing defect concentration in high acceptor-doped


ferroelectrics. They argued that for a very low acceptor concentrations
(~0.005 mol%), the surface potential increases with doping concen­
Fig. 7. Normalized Mn L-edge (Mn 2p → unoccupied Mn 3d orbitals) X-ray
tration as it depends on mean surface dipole moment density. But in the absorption spectra (XAS) for 20ST2Mn ceramic and references.
case of high acceptor doping concentrations, the negative surface charge
density is fully screened due to sufficient vacancies to decrease surface
potential. The presence of defect-dipoles at grain boundary is further

6
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

characterized by unchanged Tm as frequency increases. In an acceptor


such as Mn doped materials, the defects (Vooo or MnTi′′ /MnTi ) influence

the polar displacements of Ti atoms in the neighbor TiO6 octahedra,


resulting partial or complete correlations which are responsible for the
formation of polar nano regions (PNRs). For ionic crystals in which Mn
substituting Ti in the B-site, there are some tendency for the negatively
charged MnTi′′ impurity center to be attracted to the vicinity of the
positively charged Vooo impurity center, forming [Vooo + MnTi′′ ] defect
associations. Since the binding energy of these defect associations was
estimated to be − 1.35 eV according to GULP code [59], they become
more stable in the Ti based ABO3 matrix. Therefore, the stability of
[ oo ]
Vo + MnTi′′ defect dipoles induce a weak correlation length of polar
displacements of Ti ions and restrict the motion of Vooo leading to
reduced dielectric loss. Low room-temperature permittivity in 20Sr2Mn
(~1500 and ~2000 for 20Sr2Mn and 20Sr respectively at 10 kHz) can
be attributed to the pinning effect by the dipolar defects which can
suppress dielectric permittivity.
Fig. 8. Normalized Ti L-edge (Ti 2p →unoccupied Ti 3d orbitals) absorption
Fig. 12 shows P-E loops for 20Sr and 20Sr2Mn ceramics at room
spectra of 20Sr, 20Sr2Mn and BaTiO3 ceramics.
temperature measured at 1 Hz. Pmax, Pr and Ec for each compound are
tabulated in the inset. Compared to 20Sr, the P-E hysteresis loop of
evidenced from local piezoresponse characteristics as shown in Fig. 10.
20Sr2Mn shows a 3-fold and 2-fold decrease in Pr and Ec respectively
Both phase and amplitude responses measured at the grain interior
while exhibiting same Pmax. The reduction of Ec is possibly due to the
(Fig. 10a and c) show perfect symmetric behavior with respect to
presence of multiple Mn valences (as evidenced in Fig. 7), which can
applied tip bias. On the other hand, phase and amplitude responses
compensate local charges resulted from oxygen vacancies. Previous
measured at the grain boundary (Fig. 10b and d) manifest asymmetric
studies of Mn doped BNT perovskites showed that a small amount of Mn
characteristics (center of abscissa shifting to positive-bias side) with
doping can effectively reduce the leakage current due to excess oxygen
respect to applied tip bias. This asymmetry is strongly correlated with
vacancies, thereby improving ferroelectric properties [60]. Compared to
the random-field disorder caused by the presence of defect clusters or
20Sr, 20Sr2Mn shows a constricted P-E loop. Several possibilities for
defect-dipoles at the grain boundary [57]. It is also shown that in relaxor
constricted P-E loop have been discussed previously. One possibility is
ferroelectrics, the presence of defect dipoles is strongly correlated with
due to electrically induced paraelectric-to-FE, AFE-to-FE or relaxor-to-
the local structural heterogeneity at the nanoscale [58]. On the atomic
FE transition [61–63]. Second possibility for the constricted hysteresis
scale, the compositional heterogeneity which is inherent with compo­
loop is the aging behavior in acceptor-doped poled ferroelectrics. During
sitional disorder, induces heterogeneous polar states where the random
the aging, the oriented defect dipoles formed by acceptor induced oxy­
fields prevent the formation of a long-range ordered ferroelectric state.
gen vacancies pin the domain rotation and cause a constricted loop [64].
Fig. 11 shows the frequency and temperature-dependent dielectric
It is known when the acceptor dopant Mn4+/Mn3+ ions substitute the B-
response of 20Sr and 20Sr2Mn ceramics. The dielectric response of 20Sr
site Ti4+/Zr4+ ions, oxygen vacancies are formed at the O2− sites to
exhibits diffused character at and below Tm and Tm shifts to higher ( )×
maintain the charge neutrality, resulting in divalent MnTi′′ − Vooo and
temperature with increasing frequency. Compared to 20Sr, 20Sr2Mn is

Fig. 9. (a)–(c) height, potential and line profile of 20Sr respectively and (d-f) height, potential, and line profile of 20Sr2Mn respectively obtained from FM-KPFM.

7
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

Fig. 10. Local OP-PFM phase and amplitude hysteresis loops from grain interior (a and c) and grain boundary (b and d) of 20Sr2Mn ceramics.

Fig. 12. P-E hysteresis loop of 20Sr and 20Sr2Mn poled at 100 kV/cm
and parameters.

( )o
trivalent MnTi − Vooo defect dipoles [53,65]. Therefore, in this case,

the constricted hysteresis loop is probably related to presence of P4bm


phase (as discussed in the Rietveld refinement in Fig. 5c) and
Mn′′Ti/Zr − VOoo defect dipoles. The defect dipoles tend to remain the
orientation during the P-E loop measurement because of the low
migration rate. This provides a restoring force to reverse the switched
polarization and induces the constricted P-E loop [66]. This is explained
by symmetry-conforming short-range ordering mechanism of point de­
fects [67], in which oxygen vacancies and Mn2+/Mn3+ can create a
defect dipole moment (PD), aligning along the spontaneous polarization
(Ps). When an E field is applied, Ps is switched to E-field direction while
Fig. 11. Frequency and temperature dependent dielectric spectra of (a) 20Sr PD cannot have a sudden change. The associated PD can act as an internal
(b) 20Sr2Mn ceramics.
field to switch the domain back to its original state after removing E
field. Consequently, a macroscopic pinched and double hysteresis loop
with reduced Pr and Ec is achieved. Similar results have been reported

8
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

Fig. 13. Temperature evolution of P-E hysteresis loops of 20Sr2Mn ceramics.

4. Conclusions
Table 3
Temperature-dependent energy storage parameters in 20Sr2Mn ceramics.
The incorporation of Sr ions into parent BNT7.5BT matrix disrupted
T (◦ C) Pr (μC/cm2 ) Pm (μC/cm2 ) Ec (kV/cm) Wrec (J/cm3 ) η(%) the FE long-range order by the presence of non-polar cubic Pm3− m
25 4.98 35.54 7.84 0.67 50.7 phase resulting in enhanced relaxor behavior with decreased ε’ and
100 6.55 37.10 5.38 0.85 59.1 gradual shift of Tm toward lower temperatures. This work shows that
125 5.78 36.78 5.38 0.90 62.7 20Sr2Mn resulted in a constricted P-E loop with 3-fold and 2-fold
decrease in Pr and Ec respectively while exhibiting same value for
Pmax. The reduction of Ec is possibly due to the absorption of charge
[68,69], in which pinched P-E loops were attributed to Mn ions induced
carriers by the presence of multiple Mn valencies. The constricted hys­
defect dipoles which slowly orient opposite to the poling direction.
teresis loop is related to the increase of relaxor ergodic FE P4bm phase
Fig. 13 shows temperature evolution of P-E loops to analyze energy
and the formation of Mn′′Ti − VOoo defect dipoles. Oxygen vacancies and
storage performance of 20Sr2Mn. The constricted hysteresis loop opens
Mn2+/Mn3+ ions can create a defect dipole moment PD, which acts as an
up and attains a flat and normal shape at higher temperature, suggesting
internal field to switch the new domain back to its original state after
the migration of defect dipoles affected by thermal energy. As temper­
removing E field causing macroscopic constricted hysteresis loop.
ature increases, the pinning effect of the dipoles is weakened. As a result,
softening of the FE microdomains sets in and the dipole moment asso­ 20Sr2Mn showed improved Wrec ( 0.90 J/cm3 ) and η( 62.67%)values at
ciated with PD begins to follow the Ps direction, resulting in normal FE 125 ◦ C making this potential material for energy storage applications
hysteresis loop and slight increase in Pmax. Large Pmax, small Pr, and over a wide operating temperatures.
moderate Ec are essential parameters for energy storage performance.
Temperature-dependent energy storage parameters are shown in 5. Data availability
Table 3. At room temperature, Wrec and η respectively are 0.67 J/cm3
and 50.74%, which are higher than 20Sr (0.47 J/cm3 and 35.44%, not The raw/processed data required to reproduce these findings cannot
shown here). This increase can be attributed to Mn doping induced P4bm be shared at this time as the data also forms part of an ongoing study
phase and defect-dipoles mediated lower Pr and Ec. As temperature in­
creases, slight reduction in Pr and Ec and increase in Pmax result in
Declaration of Competing Interest
increasing Wrec and η. A better thermally activated Wrec ( 0.90 J/cm3 )
with high η( 62.67%) at 125 ◦ C makes this ceramic superior for energy The authors declare that they have no known competing financial
storage applications over a wide operating temperature compared to Sr- interests or personal relationships that could have appeared to influence
doped (Bi0.70Na0.67Li0.03)0.5TiO3 system which exhibited 0.67 J/cm3 at the work reported in this paper.
100 ◦ C [33].
Acknowledgement

This project was supported by the Ministry of Science and

9
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

Technology of Taiwan (MOST) under Project Nos. 108-2221-E-131-017- capacitors, J. Phys. Condens. Matter 24 (2012), https://doi.org/10.1088/0953-
8984/24/44/445901.
MY2, 108-2221-E-030-011-MY3 and 109-2221-E-146-002.
[25] T.C.M. Chung, Functionalization of polypropylene with high dielectric properties:
application in electric energy storage, Green Sustainable Chem. 2 (2) (2012)
References 29–37, https://doi.org/10.4236/gsc.2012.22006.
[26] F. Guan, L. Yang, J. Wang, B. Guan, K. Han, Q. Wang, L. Zhu, Confined ferroelectric
[1] C. Liu, F. Li, L.-P. Ma, H.-M. Cheng, Advanced materials for energy storage, Adv. properties in poly(vinylidene fluoride-co-chlorotrifluoroethylene)-graft-
Mater. 22 (2010) E28–E62, https://doi.org/10.1002/adma.200903328. polystyrene graft copolymers for electric energy storage applications, Adv. Funct.
[2] Q. Li, L. Chen, M.R. Gadinski, S. Zhang, G. Zhang, H.U. Li, E. Iagodkine, A. Haque, Mater. 21 (2011) 3176–3188, https://doi.org/10.1002/adfm.201002015.
L.-Q. Chen, T.N. Jackson, Q. Wang, Flexible high-temperature dielectric materials [27] P. Zhao, H. Wang, L. Wu, L. Chen, Z. Cai, L. Li, X. Wang, High-performance relaxor
from polymer nanocomposites, Nature 523 (2015) 576–579, https://doi.org/ ferroelectric materials for energy storage applications, Adv. Energy Mater. 9
10.1038/nature14647. (2019), https://doi.org/10.1002/aenm.201803048.
[3] J.H. Pikul, H. Gang Zhang, J. Cho, P.V. Braun, W.P. King, High-power lithium ion [28] Z. Song, S. Zhang, H. Liu, H. Hao, M. Cao, Q. Li, Q. Wang, Z. Yao, Z. Wang, M.
microbatteries from interdigitated three-dimensional bicontinuous nanoporous T. Lanagan, J. Jones, Improved energy storage properties accompanied by
electrodes, Nat. Commun. 4 (2013), https://doi.org/10.1038/ncomms2747. enhanced interface polarization in annealed microwave-sintered BST, J. Am.
[4] Z.S. Wu, K. Parvez, X. Feng, K. Müllen, Graphene-based in-plane micro-super- Ceram. Soc. 98 (2015) 3212–3222, https://doi.org/10.1111/jace.13741.
capacitors with high power and energy densities, Nat. Commun. 4 (2013), https:// [29] D. Zheng, R. Zuo, D. Zhang, Y. Li, X. Tan, Novel BiFeO3-BaTiO3-Ba(Mg1/3 Nb2/3)O3
doi.org/10.1038/ncomms3487. lead-free relaxor ferroelectric ceramics for energy-storage capacitors, J. Am.
[5] Q. Li, K. Han, M.R. Gadinski, G. Zhang, Q. Wang, High energy and power density Ceram. Soc. 98 (2015) 2692–2695, https://doi.org/10.1111/jace.13737.
capacitors from solution-processed ternary ferroelectric polymer nanocomposites, [30] T. Wang, L. Jin, Y. Tian, L. Shu, Q. Hu, X. Wei, Microstructure and ferroelectric
Adv. Mater. 26 (2014) 6244–6249, https://doi.org/10.1002/adma.201402106. properties of Nb2O5-modified BiFeO3-BaTiO3 lead-free ceramics for energy storage,
[6] P. Banerjee, I. Perez, L. Henn-Lecordier, S.B. Lee, G.W. Rubloff, Nanotubular metal- Mater. Lett. 137 (2014) 79–81, https://doi.org/10.1016/j.matlet.2014.08.133.
insulator-metal capacitor arrays for energy storage, Nat. Nanotechnol. 4 (2009) [31] Q. Xu, T. Li, H. Hao, S. Zhang, Z. Wang, M. Cao, Z. Yao, H. Liu, Enhanced energy
292–296, https://doi.org/10.1038/nnano.2009.37. storage properties of NaNbO3 modified Bi0.5Na0.5TiO3 based ceramics, J. Eur.
[7] L.C. Haspert, S.B. Lee, G.W. Rubloff, Nanoengineering strategies for metal- Ceram. Soc. 35 (2015) 545–553, https://doi.org/10.1016/j.
insulator-metal electrostatic nanocapacitors, ACS Nano 6 (2012) 3528–3536, jeurceramsoc.2014.09.003.
https://doi.org/10.1021/nn300553r. [32] X. Liu, J. Shi, F. Zhu, H. Du, T. Li, X.C. Liu, H. Lu, Ultrahigh energy density and
[8] Z. Liu, Y. Zhan, G. Shi, S. Moldovan, M. Gharbi, L. Song, L. Ma, W. Gao, J. Huang, improved discharged efficiency in bismuth sodium titanate based relaxor
R. Vajtai, F. Banhart, P. Sharma, J. Lou, P.M. Ajayan, Anomalous high capacitance ferroelectrics with A-site vacancy, J. Materiomics 4 (2018) 202–207, https://doi.
in a coaxial single nanowire capacitor, 879–7pp, Nat. Commun. 3 (2012), https:// org/10.1016/j.jmat.2018.05.006.
doi.org/10.1038/ncomms1833. [33] J. Wu, A. Mahajan, L. Riekehr, H.F. Zhan, B. Yang, N. Meng, Z. Zhang, H.X. Yan,
[9] C.C.B. Bufon, J.D.C. González, D.J. Thurmer, D. Grimm, M. Bauer, O.G. Schmidt, Perovskite Srx(Bi1-xNa0.97-xLi0.03)0.5TiO3 ceramics with polar nano regions for high
Self-assembled ultra-compact energy storage elements based on nanomembranes, power energy storage, Nano Energy 50 (2018) 723–732, https://doi.org/10.1016/
Nano Lett. 10 (2010) 2506–2510, https://doi.org/10.1021/nl1010367. j.nanoen.2018.06.016.
[10] R. Sharma, C.C.B. Bufon, D. Grimm, R. Sommer, A. Wollatz, J. Schadewald, D. [34] W. Cao, W. Li, Y. Feng, T. Bai, Y. Qiao, Y. Hou, T. Zhang, Y. Yu, W. Fei, Defect
J. Thurmer, P.F. Siles, M. Bauer, O.G. Schmidt, Large-area rolled-up dipole induced large recoverable strain and high energy-storage density in lead-
nanomembrane capacitor arrays for electrostatic energy storage, Adv. Energy free Na0.5Bi0.5TiO3-based systems, Appl. Phys. Lett. 108 (2016), https://doi.org/
Mater. 4 (2014), https://doi.org/10.1002/aenm.201301631. 10.1063/1.4950974.
[11] C. Wang, M. Osada, Y. Ebina, B.-W. Li, K. Akatsuka, K. Fukuda, W. Sugimoto, [35] D. Li, Y. Lin, M. Zhang, H. Yang, Achieved ultrahigh energy storage properties and
R. Ma, T. Sasaki, All-nanosheet ultrathin capacitors assembled layer-by-layer via outstanding charge-discharge performances in (Na0.5Bi0.5)0.7Sr0.3TiO3-based
solution-based processes, ACS Nano 8 (2014) 2658–2666, https://doi.org/ ceramics by introducing a linear additive, Chem. Eng. J. 392 (2020), 123729,
10.1021/nn406367p. https://doi.org/10.1016/j.cej.2019.123729.
[12] J.I. Sohn, Y.-S. Kim, C. Nam, B.K. Cho, T.-Y. Seong, S. Lee, Fabrication of high- [36] Y. Lin, D. Li, M. Zhang, H. Yang, (Na0.5Bi0.5)0.7Sr0.3TiO3 modified by Bi(Mg2/3Nb1/
density arrays of individually isolated nanocapacitors using anodic aluminum 3)O3 ceramics with high energy-storage properties and an ultrafast discharge rate,

oxide templates and carbon nanotubes, Appl. Phys. Lett. 87 (2005), https://doi. J. Mater. Chem. C 8 (2020) 2258–2264, https://doi.org/10.1039/C9TC06218A.
org/10.1063/1.2048815. [37] H. Yang, F. Yan, Y. Lin, T. Wang, F. Wang, Y. Wang, L. Guo, W. Tai, H. Wei, Lead-
[13] S.-W. Chang, J. Oh, S.T. Boles, C.V. Thompson, Fabrication of silicon nanopillar- free BaTiO3-Bi0.5Na0.5TiO3-Na0.73Bi0.09NbO3 relaxor ferroelectric ceramics for high
based nano- capacitor arrays, Appl. Phys. Lett. 96 (2010), https://doi.org/ energy storage, J. Eur. Ceram. Soc. 37 (2017) 3303–3311, https://doi.org/
10.1063/1.3374889. 10.1016/j.jeurceramsoc.2017.03.071.
[14] H.-F. Ming, G.-W. Feng, F. Zhou, L. Song, X.-H. Li, X.-O. Hu, X.-G. Zhu, B. Wu, B.- [38] C. Ma, H. Guo, S.P. Beckman, X. Tan, Creation and destruction of morphotropic
Q. Wei, Dielectric capacitors with three-dimensional nanoscale interdigital phase boundaries through electrical poling: a case study of lead-free (Bi1/2Na1/2)
electrodes for energy storage, Sci. Adv. 23 (2015) 1–6, https://doi.org/10.1126/ TiO3-BaTiO3 piezoelectrics, Phys. Rev. Lett. 109 (2012), https://doi.org/10.1103/
sciadv.1500605. PhysRevLett.109.107602.
[15] J. Rödel, W. Jo, K.T.P. Seifert, E.-M. Anton, T. Granzow, D. Dam-janovic, [39] Q. Xu, J. Xie, Z. He, L. Zhang, M. Cao, X. Huang, M.T. Lanagan, H. Hao, Z. Yao,
Perspective on the development of lead-free piezoceramics, J. Am. Ceram. Soc. 92 H. Liu, Energy-storage properties of Bi0.5Na0.5TiO3-BaTiO3-KNbO3 ceramics
(2009) 1153–1177, https://doi.org/10.1111/j.1551-2916.2009.03061.x. fabricated by wet-chemical method, J. Eur. Ceram. Soc. 37 (2017) 99–106, https://
[16] J. Anthoniappen, C.S. Tu, P.-Y. Chen, C.-S. Chen, S.-J. Chiu, H.-Y. Lee, Yi Ting, S.- doi.org/10.1016/j.jeurceramsoc.2016.07.011.
F. Wang, C.-K. Chai, Structural phase stability and electric field induced [40] P. Jaita, R. Sanjoom, N. Lertcumfu, G. Rujijanagul, Improvement of electric field-
relaxor–ferroelectric phase transition in (1–x)(Bi0.5Na0.5)TiO3–xBaTiO3 ceramics, induced strain and energy storage density properties in lead-free BNKT-based
J. Alloys Compd. 618 (2014) 120–126, https://doi.org/10.1016/j. ceramics modified by BFT doping, RSC adv. 9 (2019) 11922–11931, https://doi.
jallcom.2014.08.100. org/10.1039/C9RA00956F.
[17] S.L. Candelaria, Y. Shao, W. Zhou, X. Li, J. Xiao, J.-G. Zhang, Y. Wang, J. Liu, J. Li, [41] Y. Zhao, J. Xu, L. Yang, C. Zhou, X. Lu, C. Yuan, Q. Li, G. Chen, H. Wang, High
G. Cao, Nanostructured carbon for energy storage and conversion, Nano Energy 1 energy storage and breakdown strength of Bi0.5(Na0.82K0.18)0.5TiO3 ceramics
(2) (2012) 195–220, https://doi.org/10.1016/j.nanoen.2011.11.006. modified by (Al0.5Nb0.5)4+ complex-ion, J. Alloys Compd. 666 (2016) 209–216,
[18] B. Dunn, H. Kamath, J.-M. Tarascon, Electrical energy storage for the grid: a https://doi.org/10.1016/j.jallcom.2016.01.103.
battery of choices, Science 334 (2011) 928–935, https://doi.org/10.1126/ [42] R.A. Malik, A. Hussain, M. Acosta, J. Daniels, H.S. Han, M.H. Kim, J.S. Lee,
science.1212741. Thermal-stability of electric field-induced strain and energy storage density in Nb-
[19] G.A. Samara, Pressure and temperature dependence of the dielectric properties and doped BNKT-ST piezoceramics, J. Eur. Ceram. Soc. 38 (2018) 2511–2519, https://
phase transition of the antiferroelectric perovskites: PbZrO3 and PbHfO3, Phys. doi.org/10.1016/j.jeurceramsoc.2018.01.010.
Rev. B 1 (1970) 3777–3786, https://doi.org/10.1103/PhysRevB.1.3777. [43] C. Cui, Y. Pu, Z. Gao, J. Wan, Y. Guo, C. Hui, Y. Wang, Y. Cui, Structure, Dielectric
[20] O.A. Zhelnova, O.E. Fesenko, Phase transitions and twinning in NaNbO3 crystals, and relaxor properties in lead-free ST-NBT ceramics for high energy storage
Ferroelectrics 75 (1987) 465–475, https://doi.org/10.1080/00150198708215068. applications, J. Alloys Compd. 711 (2017) 319–326, https://doi.org/10.1016/j.
[21] Q. Zhang, L. Wang, J. Luo, Q. Tang, J. Du, Improved energy storage density in jallcom.2017.04.023.
barium strontium titanate by Addition of BaOSiO2-B2O3 glass, J. Am. Ceram. Soc. [44] C. Cui, Y. Pu, R. Shi, High-energy storage performance in lead-free (0.8-x)SrTiO3-
98 (8) (2009) 1871–1873, https://doi.org/10.1111/j.1551-2916.2009.03109.x. 0.2Na0.5Bi0.5TiO3-xBa-TiO3 relaxor ferroelectric ceramics, J. Alloys Compd. 740
[22] J. Du, B. Jones, M. Lanagan, Preparation and characterization of dielectric glass- (2018) 1180–1187, https://doi.org/10.1016/j.jallcom.2017.04.023.
ceramics in Na2O-PbO-Nb2O3-SiO2 system, Mater. Lett. 59 (2015) 2821–2826, [45] C. Cui, Y. Pu, Effect of Sn substitution on the energy storage properties of
https://doi.org/10.1016/j.matlet.2005.02.090. 0.45SrTiO3-0.2Na0.5Bi0.5TiO3-0.35Ba-TiO3 ceramics, J. Mater. Sci. 53 (2018)
[23] D.K. Kwon and M.H. Lee, Temperature stable high energy density capacitors using 9830–9841, https://doi.org/10.1007/s10853-018-2282-8.
complex perovskite thin films, in: 2011 International Symposium on Applications [46] A. Mishra, B. Majumdar, R. Ranjan, A complex lead-free (Na, Bi, Ba)(Ti, Fe)O3
of Ferroelectrics (ISAF/PFM) and 2011 International Symposium on Piezoresponse single phase perovskite ceramic with a high energy-density and high discharge-
Force Microscopy and Nanoscale Phenomena in Polar Materials, BC (2011) pp. efficiency for solid state capacitor applications, J. Eur. Ceram. Soc. 37 (2017)
1–4. 2379–2384, https://doi.org/10.1016/j.jeurceramsoc.2017.01.036.
[24] N. Ortega, A. Kumar, J.F. Scott, D.B. Chrisey, M. Tomazawa, S. Kumari, D.G. [47] F. Li, J. Zhai, B. Shen, X. Liu, H. Zeng, Simultaneously high-energy storage density
B. Diestra, R.S. Katiyar, Relaxor-ferroelectric superlattices: high energy density and responsivity in quasi-hysteresis-free Mn-doped Bi0.5Na0.5TiO3-BaTiO3-

10
P.-Y. Chen et al. Materials Science & Engineering B 263 (2021) 114869

(Sr0.7Bi0.2Na0.1)TiO3 ergodic relaxor ceramics, Mater. Res. Lett. 6 (2018) 345–352, Relaxor ferroelectrics, Funct. Mater. 28 (2018) 1801504–1801521, https://doi.
https://doi.org/10.1080/21663831.2018.1457095. org/10.1002/adfm.201801504.
[48] S. Steiner, I.-T. Seo, P. Ren, M. Li, D.J. Keeble, T. Frömling, The effect of Fe- [59] I.S. Vorotiahin, A.N. Morozovska, Y.A. Genenko, Hierarchy of domain
acceptor doping on the electrical properties of Na1/2Bi1/2TiO3 and 0.94(Na1/2Bi1/2) reconstruction processes due to charged defect migration in acceptor doped
TiO3-0.06BaTiO3, J. Am. Ceram. Soc. 102 (2019) 5295–5304, https://doi.org/ ferroelectrics, Acta Mater. 184 (2020) 267–283, https://doi.org/10.1016/j.
10.1111/jace.16401. actamat.2019.11.048.
[49] J. Anthoniappen, C.S. Tu, P.-Y. Chen, C.-S. Chen, Y.U. Idzerda, S.-J. Chiu, Raman [60] W. Sakamotoa, N. Makinoa, B.-Y. Lee, T. Iijima, M. Moriyaa, T. Yogoa, Influence of
spectra and structural stability in B-site manganese doped volatile element composition and Mn doping on the electrical properties of lead-
(Bi0.5Na0.5)0.925Ba0.075TiO3 relaxor ferroelectric ceramics, J. Eur. Ceram. Soc. 35 free piezoelectric (Bi0.5Na0.5)TiO3 thin films, Sensors Actuators A. 200 (2013)
(2015) 3495–3506, https://doi.org/10.1016/j.jeurceramsoc.2015.05.002. 60–67, https://doi.org/10.1016/j.sna.2012.10.006.
[50] H. Moriwake, C.A.J. Fisher, A. Kuwabara, First-principles calculations of electronic [61] X. Yang, X. Liu, S. Ding, The defect and dielectric properties of Nb and Mn co-
structure and solution energies of Mn-doped BaTiO3, Jpn. J. Appl. Phys. 49 (2010), doping BaTiO3 ceramics, Ferroelectrics 533 (2018) 132–138, https://doi.org/
https://doi.org/10.1143/JJAP.49.09MC01, 09MC01-6. 10.1080/00150193.2018.1472995.
[51] E. Aksel, P. Jakes, E. Erdem, D.M. Smyth, A. Ozarowski, J. van Tol, J.L. Jones, R.- [62] T.F. Juliano, Y.G. Gogotsi, T.E. Buchheit, C.S. Watson, S.V. Kalinin, J. Shin, A.
A. Eichel, Processing of manganese-doped [Bi0.5Na0.5TiO3] ferroelectrics: P. Baddorf, Detection of indentation induced FE-to-AFE phase transformation in
reduction and oxidation reactions during calcination and sintering, J. Am. Ceram. lead zirconate titanate, J. Am. Ceram. Soc. 89 (2006) 3557–3559, https://doi.org/
Soc. 94 (2011) 1363–1367, https://doi.org/10.1111/j.1551-2916.2010.04249.x. 10.1111/j.1551-2916.2006.01255.x.
[52] J. Liu, N. Mohammad, B.B. Xiao, Q. Sun, L. Andrew, Li Ruying, J. Shamc, T.K. Guo, [63] H. He, X. Tan, Electric-field-induced transformation of incommensurate
X. Sun, Atomically precise growth of sodium titanates as anode materials for high- modulations in antiferroelectric Pb0.99Nb0.02[(Zr1-xSnx)1-y]0.98O3, Phys. Rev. B 72
rate and ultralong cycle-life sodium-ion batteries, J. Mater. Chem. A 3 (2015) (2005), https://doi.org/10.1103/PhysRevB.72.0241022.
24281–24288, https://doi.org/10.1039/C5TA08435K. [64] P.-Y. Chen, C.-S. Chen, C.-S. Tu, Y.-S. Wu, W.-S. Chang, S.-H. Chen, The giant strain
[53] Y.A. Genenko, O. Hirsch, P. Erhart, Surface potential at a ferroelectric grain due to response mechanism in textured Mn-modified 0.925(Bi0.5Na0.5)TiO3-0.075BaTiO3
asymmetric screening of depolarization fields, J. Appl. Phys. 115 (2014) relaxor ferroelectric ceramics, J. Alloys Compd. 737 (2018) 705–717, https://doi.
104102–104113, https://doi.org/10.1063/1.4867984. org/10.1016/j.jallcom.2017.12.173.
[54] P. Lunkenheimer, V. Bobnar, A.V. Pronin, A.I. Ritus, A.A. Volkov, A. Loidl, Origin [65] Z. Cao, A. Ding, Y. Zhang, P. Qiu, X. Zheng, Double-hysteresis-like loops in Mn-
of apparent colossal dielectric constants, Phys. Rev. B 66 (2002), https://doi.org/ doped (Pb, La)(Zr, Ti)O3 ceramics, Solid State Commun. 131 (2004) 57–60,
10.1103/PhysRevB.66.052105, 052105-4. https://doi.org/10.1016/j.ssc.2004.04.009.
[55] Y. Kim, S. Bühlma, J. Kim, M. Park, K. No, Y.K. Kim, S. Hong, Local surface [66] H. Wei, Y. Chen, Ferroelectric and pyroelectric properties of Mn-doped lead
potential distribution in oriented ferroelectric thin films, App. Phys. Lett. 91 (2007) zirconate titanate ceramics, Ceram. Int. 41 (2015) 6158–6163, https://doi.org/
052906–52913, https://doi.org/10.1063/1.2761502. 10.1016/j.ceramint.2014.02.113.
[56] R.A. Maier, T.A. Pomorski, P.M. Lenahan, C.A. Randall, Acceptor-oxygen vacancy [67] Z. Feng, S.W. Or, Aging-induced, defect-mediated double ferroelectric hysteresis
defect dipoles and fully coordinated defect centers in a ferroelectric perovskite loops and large recoverable electrostrains in Mn-doped orthorhombic KNbO3-
lattice: electron paramagnetic resonance analysis of Mn2+ in single crystal BaTiO3, based ceramics, J. Alloys Compd. 480 (2009) L29–L32, https://doi.org/10.1016/j.
J. Appl. Phys. 118 (2015) 164102–164109, https://doi.org/10.1063/1.4934505. jallcom.2009.02.043.
[57] P. Maksymovych, N. Balke, S. Jesse, M. Huijben, R. Ramesh, A.P. Baddorf, S. [68] H.B. Zhang, S.L. Jiang, Y.K. Zeng, B site doping effect on depinning in Pb(Mn1/
V. Kalinin, Defect-induced asymmetry of local hysteresis loops on BiFeO3 surfaces, 3Nb1/3Sb1/3)x(Zr0.825Ti0.175)1-xO3 ferroelectric ceramics, Appl. Phys. Lett. 93
J. Mater. Sci. 44 (2009) 5059–5101, https://doi.org/10.1007/s10853-009-3697-z. (2008), https://doi.org/10.1063/1.2978157.
[58] F. Li, S. Zhang, D. Damjanovic, L.Q. Chen, T.R. Shrout, Local structural [69] B.S. Li, G.R. Li, Q.R. Yin, Z.G. Zhu, A.L. Ding, W.W. Cao, Pinning and depinning
heterogeneity and electromechanical responses of ferroelectrics: learning from mechanism of defect dipoles in PMN-PZT ceramics, J. Phys. D Appl. Phys. 38
(2005) 1107–1111, https://doi.org/10.1088/0022-3727/38/8/001.

11

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy