2019 Quantitative Studies of Domain Evolution in Tetragonal BS PT Ceramics in Electric Poling and Thermal Depoling Process

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

View Article Online

View Journal

Journal of
Materials Chemistry C
Materials for optical, magnetic and electronic devices
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: J. Wu, X. Gao, Y.
Yu, J. Yang, Z. Chu, A. A. Bokov, Z. Ye and S. Dong, J. Mater. Chem. C, 2019, DOI: 10.1039/C9TC00748B.

Volume 4 Number 1 7 January 2016 Pages 1–224 This is an Accepted Manuscript, which has been through the
Royal Society of Chemistry peer review process and has been
Journal of accepted for publication.
Materials Chemistry C Accepted Manuscripts are published online shortly after
Materials for optical, magnetic and electronic devices
www.rsc.org/MaterialsC

acceptance, before technical editing, formatting and proof reading.


Using this free service, authors can make their results available
to the community, in citable form, before we publish the edited
article. We will replace this Accepted Manuscript with the edited
and formatted Advance Article as soon as it is available.

You can find more information about Accepted Manuscripts in the


author guidelines.

Please note that technical editing may introduce minor changes


to the text and/or graphics, which may alter content. The journal’s
ISSN 2050-7526 standard Terms & Conditions and the ethical guidelines, outlined
PAPER~
in our author and reviewer resource centre, still apply. In no
Nguyên T. K. Thanh, Xiaodi Su et al.
Fine-tuning of gold nanorod dimensions and plasmonic properties using
the Hofmeister effects
event shall the Royal Society of Chemistry be held responsible
for any errors or omissions in this Accepted Manuscript or any
consequences arising from the use of any information it contains.

rsc.li/materials-c
Page 1 of 24 Journal of Materials Chemistry C

View Article Online


DOI: 10.1039/C9TC00748B

Journal of Materials Chemistry C Accepted Manuscript


Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

In electric poling process, 180° and 90° domain reversals in tetragonal BS-

PT ceramics are quantitively estimated by their charge contributions.


Journal of Materials Chemistry C Page 2 of 24

View Article Online

Quantitative studies of domain evolution in tetragonal DOI: 10.1039/C9TC00748B

BS-PT ceramics in electric poling and thermal depoling

Journal of Materials Chemistry C Accepted Manuscript


processes
Jingen Wu1,3, Xiangyu Gao1, Yang Yu1, Jikun Yang1, Zhaoqiang Chu1, Alexei A.
Bokov2, Zuo-Guang Ye2 * and Shuxiang Dong1 *
1Department of Materials Science and Engineering, College of Engineering, Peking

University, Beijing 100871, People’s Republic of China


2Department of Chemistry and 4D LABS, Simon Fraser University, Burnaby, British

Columbia, V5A 1S6, Canada


3Electronic Materials Research Laboratory, Key Laboratory of the Ministry of

Education & International Center for Dielectric Research, Xi’an Jiaotong University,
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

Xi’an 710049, China


*Authors to whom correspondence should be addressed. E-mails:
[email protected], [email protected]

Abstract
Bismuth scandate-lead titanate ceramic (abbreviated as BS-PT) is a promising
piezoelectric material for high temperature applications because of its excellent
piezoelectric performance as well as high Curie temperature. Piezoelectric properties
of ferroelectric materials are closely related to microscopic domain structure, however,
quantitative analysis on domain evaluation as well as its effect on materials properties
in both electric poling and thermal depoling processes is seldom revealed. In this work,
we investigate the effects of electric poling and thermal depoling on domain structure
and piezoelectric performance in tetragonal BS-PT ceramics. Based on the early work
of Uchida and Ikeda, a modified model is developed to quantitively study both 180°
and 90° domain reversals during electric poling process. To make clear correlation
between domain structure and piezoelectric properties, quantitative estimation of 90°
domain reversal in both electric poling and thermal depoling processes is analyzed by
XRD method. The present work gives a comprehensive insight into domain evolution
of tetragonal BS-PT ceramics, which is also of heuristic significance to other perovskite
piezoelectric ceramics.

Keywords: BS-PT ceramics, electric poling, thermal depoling, domain evolution


Page 3 of 24 Journal of Materials Chemistry C

View Article Online

1. Introduction DOI: 10.1039/C9TC00748B

The piezoelectricity of ferroelectric ceramics, which results from macroscopic

Journal of Materials Chemistry C Accepted Manuscript


polarization and domain reversal, must be induced by a strong external poling electric
field. As ferroelectric ceramics are polycrystalline, their piezoelectric properties are
tightly related to the microscopic domain structure. Before electric poling, the
spontaneous polarization (PS), or dipole, is spatially randomly-orientated because of
the random grain orientation so that the unpoled material is macroscopically nonpolar.
Consequently, ferroelectric ceramics reveal to be non-piezoelectric, with the absence
of electromechanical coupling properties or any other piezoelectric performance. Under
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

a poling electric field, the spontaneous polarization (i.e., dipole) in each grain is
reoriented to be paralleled to the poling field direction as much as possible, then
ferroelectric ceramics become piezoelectric. Crucially, researches on domain structure
provide fundamental explanations on physical properties of piezoelectric materials.
Hence, the domain evolution and dipole reorientation that account for the dielectric,
piezoelectric properties of piezoelectric ceramics are widely investigated in
conventional piezoelectric ceramics.[1-9]
Recently, bismuth scandate-lead titanate solid solution ((1-x)BiScO3-xPbTiO3,
abbreviated as BS-PT) has been continuously drawing intensive research interest since
it was first reported by Eitel et al.[10] Because of its excellent piezoelectric properties
(d33 = 460 pC/N) and high Curie temperature (TC = 450 ºC),[11] BS-PT is a promising
piezoelectric material for high temperature applications. It is reported that BS-PT has
been successfully used in various high-temperature piezoelectric devices. Piezoelectric
motors based on BS-PT ceramics are able to work stably at the temperature as high as
200 ºC.[12-13] A thin ring-shaped BS-PT actuator operating in shear-bending mode can
produce an output displacement of 20 μm at 200 ºC.[14] Piezoelectric energy harvesters
made of BS-PT piezoelectric ceramics operating in d33 or d31 mode have been proved
to work effectively, converting ambient vibration energy into electric power under
high-temperature circumstance.[15-16]
Efforts have also been made to explore the physical mechanism of the high
Journal of Materials Chemistry C Page 4 of 24

View Article Online


piezoelectric response and high ferroelectric-transition temperature in BS-PT
DOI:based
10.1039/C9TC00748B

piezoelectric materials. Íñiguez et al.[17] carried out first-principle study on the (1-

Journal of Materials Chemistry C Accepted Manuscript


x)BiScO3-xPbTiO3 piezoelectric solid solution and indicated that this material
displayed very large structural distortions and polarizations at the MPB (morphotropic
phase boundary). Jones et al.[18] reported that intrinsic contributions to the
electromechanical response were independent of frequency while extrinsic
contributions were shown to be dependent upon frequency in 36%BiScO3-64%PbTiO3
composition. Lalitha et al.[19] demonstrated a self-consistent set of piezoelectric and
structural characterizations that the piezoelectric response of (1-x)PbTiO3-xBiScO3
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

ceramic had a direct relationship with the lattice polarizability. Zhao et al.[20] studied
the high temperature performance of BS-xPT-PSN system, and revealed that the
mechanisms of its giant piezoelectricity at high temperatures were attributed to the
rotation and extension of spontaneous polarization and the MPB-related phase
transition. However, there are few researches focusing on the domain evolution in BS-
PT ceramics under the conditions of electric poling and thermal depoling. Such issues
as the domain evolution due to poling or depoling, the effects of domain configuration
on piezoelectric properties, the correlation between phase structure and domain
configuration, etc., remain to be solved in BS-PT ceramics, especially in temperature-
varying environment.
Previously, based on X-ray diffraction (XRD) method, we proposed the
diffraction-plane-transformation (DPT) model to quantitively estimate 90° domain
evolution during electric poling process, which well reveals the 90° domain reversal in
tetragonal BS-PT piezoelectric ceramics.[21] However, the DPT model cannot reveal
180° domain evolution because 180° domain reversal does not cause any interplanar
spacing change that can be detected by X-ray. Accordingly, the quantitative estimation
of 180° domain evolution cannot be realized by XRD method. In this work, by
introducing Gaussian probability distribution function of domain reversal, a modified
Uchida-Ikeda model is built for tetragonal BS-PT ceramics. 180° and 90° domain
reversals in tetragonal BS-PT ceramics are distinguished according to their charge
contributions. The contributions of 180° and 90° domain reversals to polarization
Page 5 of 24 Journal of Materials Chemistry C

View Article Online


charge under a poling field are quantitatively estimated by the modifiedDOI:
model.
10.1039/C9TC00748B

Meanwhile, the domain structure evolution of tetragonal BS-PT piezoelectric ceramics

Journal of Materials Chemistry C Accepted Manuscript


under both electric poling and thermal depoling is investigated using scanning electron
microscope (SEM) and high-resolution X-ray diffraction method. The percentage of 90°
domain reversal is quantitively estimated based on the XRD profiles, and
correspondingly, the relationship between 90° domain reversal and piezoelectric
properties is clearly established.

2. Experiment
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

In our experiments, 0.367BiScO3-0.633PbTiO3 is found to be a tetragonal


composition that close to the morphotropic phase boundary (MPB), and we use this
composition to study 180° and 90° domain reversals. Samples of 0.367BiScO3-
0.633PbTiO3 piezoelectric ceramic were prepared by conventional solid-state
synthesis,[22] using the following raw materials: PbO (99 % purity; Xilong Chemical
Co., Ltd., Shantou, China), TiO2 (98.5 % purity; Beijing Chemical Works, Beijing,
China), Bi2O3 (99 % purity; Sinopharm Chemical Reagent Co., Ltd., Beijing, China),
and Sc2O3 (99.99 % purity; Rare-chem Hi-tech Co., Ltd., Huizhou, China).
Stoichiometric mixtures were ball milled for 24 hours with ethanol as the milling
medium. The milled powder was subsequently dried at 80 °C and then calcined at
840 °C for 4 hours. The calcined powder was compacted into piece under an uniaxial
press of 10 MPa, using PVA as a binder. After burning off PVA, green compacts were
sintered at 1080 °C for 3 hours, embedded in the calcined powder having the same
composition. The sintered pellets were polished and electroded with a postfire silver
paste at 650 °C for 0.5 hours. To measure piezoelectricity, all ceramic samples were
poled in a silicone oil bath at 120 °C under a DC electric field for 20 minutes. The
resultant specimens were aged in air for 24 hours prior to any electrical measurements.
The piezoelectric constant (d33) of specimens was measured using a quasi-static
d33 meter (ZJ-3D; Institute of Acoustics, Beijing, China). The electromechanical
coupling factors were calculated based on the resonance method using an impedance
Journal of Materials Chemistry C Page 6 of 24

View Article Online


analyzer (4294A; Agilent Technologies).[23] A chemical etching technique wasDOI:
used to
10.1039/C9TC00748B

observe the domain structure. The BS-PT ceramic samples were polished and etched at

Journal of Materials Chemistry C Accepted Manuscript


room temperature in a mixed aqueous solution of HCl/HF acids.[24] The surface
microstructure of sintered pellets was imaged by scanning electron microscopy (SEM;
S-4800, Hitachi, Tokyo, Japan). The ferroelectric properties (including switching
current) of the samples were tested with a TF Analyzer 2000 Measurement System
(aixACCT Systems GmbH, Germany). To avoid the residual stress in near-surface
region that produced by mechanical polishing, the electrode on the ceramic sample was
removed by etching in HCl/HF acids, and then high-resolution X-ray diffraction (XRD)
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

measurements with a monochromatic Cu-Kα1 radiation (PANalytical X’Pert Powder,


Netherland) were carried out to determine the crystalline phase and domain texture of
resultant ceramic samples. All the samples were measured by slow scanning at
0.25°/min in the diffraction angle (2θ) range of 20° to 60° with a resolution of 0.003°.

3. Results and Discussion

3.1. Domain evolution during electric poling


During electric poling process, both 180° dipole reversal and 90° dipole reversal
in tetragonal BS-PT ceramics will be induced by the poling electric field, as
schematically shown in Fig. 1. 180° and 90° reversed domains are the microscopic
regions that respectively contain numerous 180° reversed dipoles and 90° reversed
dipoles. Accompanied by domain reversal, the macroscopic polarization will be
induced. Therefore, the electric displacement (𝐷3), which practically equals to the
macroscopic polarization in ferroelectric materials, can be calculated by using the
following equation:
𝐷3 = 𝜀0𝜀𝑟𝐸3 + (𝑃180 + 𝑃90), (1)

where 𝜀0 is the permittivity of vacuum, 𝜀𝑟 is the relative dielectric constant, 𝐸3


is the external poling electric field, 𝑃180 and 𝑃90 are the polarization contributions of
the reversible 180° and 90° domains, respectively. 𝑃180 and 𝑃90 can be estimated by
the following equations:
Page 7 of 24 Journal of Materials Chemistry C

View Article Online


𝑃180 = 𝜂180(𝐸3)𝑁180 𝑝180, (2)
DOI: 10.1039/C9TC00748B

𝑃90 = 𝜂90(𝐸3)𝑁90 𝑝90, (3)

Journal of Materials Chemistry C Accepted Manuscript


where 𝑁180 and 𝑁90 are respectively the total quantities of 180° and 90° dipoles
in reversible 180° and 90° domains. 𝜂180(𝐸3) and 𝜂90(𝐸3) are the percentages of the
reversed 180° and 90° dipoles, respectively, and they are functions of the applied poling
electric field 𝐸3. 𝑝180 and 𝑝90 are the average polarization contributions of 180° and
90° dipole reversals, respectively.
It is well known that dipoles in an unpoled piezoelectric ceramic are randomly
orientated, thus a virtual intact sphere that includes dipoles of all possible orientations
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

can be used to represent the initial distribution of dipoles.[25-26] Previously, Uchida and
Ikeda proposed such a sphere model for estimating the electrostriction and polarization
in perovskite-type ferroelectric ceramics under an arbitrary biasing field.[27] As for an
unpoled isotropic tetragonal BS-PT ceramic, all the possible dipoles are distributed in
either 180° domain or 90° domain, and their space orientations are also regarded as an
uniform distribution in the sphere (Fig. 1). Therefore, on the aspect of quantity, 180°
and 90° reversed dipoles have the same volume density. According to this sphere model,
only the dipoles distributing within the red spatial region (denoted as 𝑉180), as
presented in Fig. 1(a-i), can produce 180° reversal during electric poling. Similarly,
only the dipoles distributing within the blue spatial region (denoted as 𝑉90) can produce
90° reversal, see Fig. 1(b-i). Then, 𝑁180 and 𝑁90 can be estimated by 𝑉180 and 𝑉90,
and for the reason that dipoles have uniform distribution in the sphere, we have 𝑁180/
𝑁90 ≈ 𝑉180/𝑉90. From the sectional view of sphere model, it is also deduced that the

boundary angle between red region and blue region is around 45°, where domain
switching of tetragonal phase is energetically unfavorable.[28-29]
However, Uchida and Ikeda’s sphere model did not present a quantitative
estimation of 180° or 90° domain reversal under a poling electric field. Domain reversal
in ferroelectric ceramics is a sophisticated process, which can be regarded as the
competition results of poling electric field versus elastic distortion, steric hindrance,
point defects, etc. As for a specific dipole reversal, it is impossible to judge when and
how its reversal is completed because of its complicated ambient conditions. To make
Journal of Materials Chemistry C Page 8 of 24

View Article Online


matters worse, there are innumerable kinds of such dipole reversals in a ferroelectric
DOI: 10.1039/C9TC00748B

ceramic. Here, we introduce a Gauss probability distribution function to describe these

Journal of Materials Chemistry C Accepted Manuscript


innumerable dipole reversals. The quantitative estimation of 180° and 90° domain
reversals is accordingly given, which is confirmed by poling-induced charge
measurements.
Under a poling electric field, only dipoles within the red spatial region of the
sphere mode may produce 180° domain reversal (Fig. 1(a-iii)). 𝑃𝑆 is the spontaneous
polarization of each dipole, and 𝜑 is the angle of dipole orientation relative to the
vertical center line, i.e., the poling direction. Thus, 2𝑃𝑆𝑐𝑜𝑠𝜑 is the polarization variation
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

of each 180° reversed dipole. Assuming that the domain reversal (in a large quantity)
obeys statistical distribution, we can introduce an orientation state distribution function,
i.e., Gaussian probability distribution function 𝑓180(𝜑)𝐼 and 𝑓180(𝜑)𝐼𝐼 about 𝜑, to
describe the 180° domain reversal possibility during electric poling. Then the real
polarization contribution of each 180° reversed dipole should be 2𝑃𝑆𝑐𝑜𝑠𝜑𝑓180(𝜑)𝐼 , or
2𝑃𝑆𝑐𝑜𝑠𝜑𝑓180(𝜑)𝐼𝐼. Therefore, the average polarization contribution of each 180° dipole

reversal can be estimated by the following three equations:


𝜋
∫42𝑃𝑆𝑐𝑜𝑠𝜑𝑓180(𝜑)𝐼𝑑𝑉180
0
𝑝180 = 𝜋 , (4)
∫4𝑑𝑉180
0

𝑓180(𝜑)𝐼 =
1
2𝜋𝛿180 ( 𝜑2
)
exp ― 2𝛿2 , 𝜑 ∈ [0,4]
180
𝜋

and 𝑓180(𝜑)𝐼𝐼 =
1
2𝜋𝛿180
exp ―( (𝜑 ― 𝜋)2
2𝛿2180 ),𝜑 ∈ [ 3𝜋
4 ,𝜋] (5)

𝜋 3𝜋
𝑑𝑉180 = 2𝜋𝑟2𝑠𝑖𝑛𝜑𝑑𝜑𝑑𝑟, 𝜑 ∈ [0,4] ∪ [ 4 ,𝜋], (6)

where 𝑑𝑉180 is the volume element of 180° domain, and the total volume of red
𝜋

region can be calculated as 𝑉180 = 2∫04 𝑑𝑉180 = (4 ― 2√2)𝜋𝑟3/3, where r is the radius of

the hypothetical sphere, see Fig. 1(a-iii). 𝛿180 is the standard deviation about 𝜑 in 180°
dipole reversal, which can be numerically determined to be less than 0.25 by integrating
𝜋/4 𝜋
Eq.(5) (∫0 𝑓180(𝜑)𝐼𝑑𝜑 + ∫3𝜋/4𝑓180(𝜑)𝐼𝐼𝑑𝜑 = 1).

Analogously, for each 90° reversed dipole, its polarization contribution can be
calculated as Δ𝑝90 = 𝑃𝑆(𝑠𝑖𝑛𝜑 ― cos 𝜑)𝑓90(𝜑), and the average polarization contribution
Page 9 of 24 Journal of Materials Chemistry C

View Article Online


of each 90° dipole reversal can also be estimated by the following three equations, see
DOI: 10.1039/C9TC00748B

Fig. 1(b-iii):
3𝜋

Journal of Materials Chemistry C Accepted Manuscript


∫ 4 𝑃𝑆(𝑠𝑖𝑛𝜑 ― co𝑠𝜑)𝑓90(𝜑)𝑑𝑉90
𝜋
4
𝑝90 = 3𝜋 , (7)
∫ 4 𝑑𝑉90
𝜋
4

𝜋 2

𝑓90(𝜑) =
1
2𝜋𝛿90
exp ― ( (𝜑 ― 2)
2𝛿290 ) , (8)

𝜋 3𝜋
𝑑𝑉90 = 2𝜋𝑟2𝑠𝑖𝑛𝜑𝑑𝜑𝑑𝑟, 𝜑 ∈ [ 4 , 4 ], (9)

where 𝑓90(𝜑) is the Gaussian probability distribution function to describe the


Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

possibility of 90° dipole reversal, 𝜑 is the angle of dipole orientation relative to the
vertical center line, 𝑑𝑉90 is the volume element of 90° reversed domain. The total

volume of 𝑉90 can be calculated as 2 2𝜋𝑟3/3, and 𝛿90 is the standard deviation about

𝜑 in 90° dipole reversal, which also can be numerically determined to be less than 0.25

by the similar method of 180° dipole reversal.


According to equation (6) and equation (9), 180° reversed dipole and 90° reversed
dipole are distributed in different regions, i.e., 180° domain and 90° domain. As for 180°
reversed dipole and 90° reversed dipole, their orientation angles (i.e., 𝜑) are in the
𝜋 3𝜋 𝜋 3𝜋
ranges of [0,4] ∪ [ 4 ,𝜋] and [4, 4 ], respectively. Thus, both 180° domain and 90°

domain can switch across a broad range of spatial orientation angle. Based on numerical
simulation, 𝑝180/𝑝90 is found to be close to 1, implying that single dipole reversal
which either belongs to 180° domain or 90° domain averagely has the same contribution
to macroscopic polarization.
To further investigate the polarization contribution of 180° and 90° domain
reversals, the switching current of BS-PT ceramic samples was measured, on account
of switching current is directly related to domain reversal.[30] Based on equation (1), we
can obtain the poling-induced macroscopic charge using the following equations:
𝑄3 = ∮(𝜀0𝜀𝑟𝐸3 + 𝑃180 + 𝑃90)𝑑𝑆 = 𝑄C0 + 𝑄180 + 𝑄90, (10-a)
𝑄C0 = ∮(𝜀0𝜀𝑟𝐸3)𝑑𝑆 = 𝐶0𝑉, (10-b)
𝑄180 = ∮𝑃180𝑑𝑆, (10-c)
𝑄90 = ∮𝑃90𝑑𝑆, (10-d)
Journal of Materials Chemistry C Page 10 of 24

View Article Online


where 𝑄3 is the total induced charge along the poling direction of BS-PTDOI:
ceramic
10.1039/C9TC00748B

sample, S is the area of electrode, 𝐶0 is the capacitance of the initially unpoled BS-PT

Journal of Materials Chemistry C Accepted Manuscript


ceramic sample, and V is the applied voltage. From equation (10-a), it can be deduced
that the poling-induced charge includes three parts: (i) the capacitance stored charge
𝑄C0, (ii) 180° domain reversal induced charge 𝑄180 and (iii) 90° domain reversal

induced charge 𝑄90, which means that both 180° and 90° domain reversals induce
additional charge during electric poling process.

The charge stored in capacitance 𝐶0 is calculated by 𝑄C0 = 𝐶0𝑉, and the


𝑑𝑄C0 𝑑𝑉(𝑡)
corresponding charging current, which can be calculated by =𝐶 , is found to
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

𝑑𝑡 0 𝑑𝑡

be a constant because the applied voltage V increases linearly with respect to time, as
shown in Fig. 2(a). The current induced by 180° domain reversal is experimentally
found to be an exponential function of time, shown in Fig. 2(b). And the current induced
by 90° domain reversal reveals a peak distribution around coercive filed Ec, shown in
Fig. 2(c). Based on the measured switching currents and their time dependence, we can
obtain the domain reversal induced charge, i.e., 𝑄90 and 𝑄180, according to 𝑄 = ∫𝐼𝑑𝑡.
Fig. 2(a-i), 2(b-i) and 2(c-i) respectively present the measured switching currents
under low (0.25kV/mm), medium (2kV/mm) and high (5kV/mm) poling electric fields.
𝑑𝑉(𝑡)
The applied voltage is set as triangle wave so that 𝑑𝑡 can be a constant to simplify

the capacitance calculation. It is revealed that there is no domain reversal occurring


under low poling field, therefore, the measured switching current exhibits the behavior
of capacitance (𝐶0) charging. Thus, the stored charge totally comes from the charging
capacitance, i.e., 𝑄C0 has a nearly 100% contribution to the stored charge in the sample

under low poling electric field, demonstrating that no domain reversal occurs under low
poling electric field, see Fig. 2(a-ii). Under a medium poling electric field of 2 kV/mm,
both capacitance 𝐶0 and 180° domain reversal contribute to the stored charge, and the
180° domain reversal induced current can be well fitted by an exponential equation.
The charge induced by 180°domain reversal takes up ~76% of the total charge, see Fig.
2(b-ii). When the poling electric field increases to a sufficiently high strength, 90°
domain reversal is found to have a dominant contribution to the stored charge, because
Page 11 of 24 Journal of Materials Chemistry C

View Article Online


the switching current curve presents a prominent peak-shape distribution around the
DOI: 10.1039/C9TC00748B

coercive field where 90° domain reversal occurs. The charge contributions of

Journal of Materials Chemistry C Accepted Manuscript


capacitance 𝐶0, 180° domain reversal and 90° domain reversal are respectively ~7%,
~27% and ~66% at the poling electric field of 5 kV/mm, see Fig. 2(c-ii). These results
reveal that 180° and 90° domain reversals are strongly dependent on poling field. It is
also noted that 180° domain reversal exhibits a fast increase with respect to poling field,
while 90° domain reversal occurs only when the poling field is higher than a threshold
value, and becomes more prominent than 180° domain reversal under an adequately
high poling electric field.
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

Fig. 3(a) shows the switching current as a function of time under different poling
electric field. It is revealed that the switching current increases slowly under low poling
electric field, where 180° domain reversal is dominant. The increase of switching
current speeds up as the poling electric field further increases, gradually appearing a
peak value. This phenomenon shows that 90° domain reversal will not occur until the
poling electric field exceeds a threshold value known as coercive field. Thereafter, 90°
domain reversal becomes more dominant than 180° domain reversal in the contribution
to switching current. Fig. 3(b) shows the time dependence of the total induced charge
under different poling fields. Clearly, the induced charge is strongly related to the
poling electric field. As the poling field increases, the induced charge increases faster
with respect to time, and then gradually trends to saturation, implying that the domain
reversals at high electric field tend to become steady. Fig. 3(c) schematically shows the
calculation of induced charge that contributed from capacitance charging (𝑄C0), 180°
domain reversal (𝑄180) and 90°domain reversal (𝑄90) under the electric poling field of 3
kV/mm. The whole area as shown in Fig. 3(c) represents the total induced charge. To
separate the overlapped switching currents that resulted from 90° and 180° domain
reversals, an exponential fitting function is used to describe the 180° domain related
switching current. By excluding 180° domain switching current and 𝑄C0 charging

current, the 90° domain switching current can be obtained.


Then we can give a quantitative estimation of 180° and 90° domain reversals based
on the poling-induced charge. It has been aforementioned that 𝑝180/𝑝90 is close to 1,
Journal of Materials Chemistry C Page 12 of 24

View Article Online


meaning that single 180° or 90° dipole reversal almost has the same contribution to
DOI: 10.1039/C9TC00748B

macroscopic polarization during electric poling. According to the equations (2),

Journal of Materials Chemistry C Accepted Manuscript


equations (3), equation (10-c) and equation (10-d), we have:
𝑄180 𝑃180 𝜂180(𝐸3)𝑁180
𝑄90 = 𝑃90 = 𝜂90(𝐸3)𝑁90 , (11)

𝜂180(𝐸3) = 𝑄180(𝐸3)/𝑄saturation
180 , (12-a)
saturation
𝜂90(𝐸3) = 𝑄90(𝐸3)/𝑄90 , (12-b)
where 𝑄saturation
180 and 𝑄saturation
90 are respectively the saturation charge induced by
180° domain reversal and 90° domain reversal. In our experiment, 𝑄saturation
180 and
𝑄saturation
90 are acquired from the BS-PT ceramic sample poled under 5 kV/mm. 𝑄180(𝐸3)
and 𝑄90(𝐸3) are acquired from the samples that poled under electric fields below 5
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

𝑄180
kV/mm. In a sufficiently poled sample, we have 𝜂180(𝐸3) ≈ 1, 𝜂90(𝐸3) ≈ 1 and 𝑄90 =

𝑁180 𝑉180
𝑁90 ≈ 𝑉90 . It is interesting to note that 𝑄saturation
180 /𝑄saturation
90 ≈ 0.43 is in well accordance

with 𝑉180/𝑉90 ≈ 0.414, which further confirms that the quantitative estimation of
domain reversal based on the poling-induced charge is reasonable.
Thus, the contributions of 180° and 90° domain reversals to the total domain
reversals under different poling field 𝐸3, i.e., 𝑋180(𝐸3) and 𝑋90(𝐸3), can be

respectively estimated by their normalized charge contributions:


𝑋180(𝐸3) = 𝑄180(𝐸3)/(𝑄saturation
180 + 𝑄saturation
90 ), (13-a)
saturation saturation
𝑋90(𝐸3) = 𝑄90(𝐸3)/(𝑄180 + 𝑄90 ). (13-b)
The normalized charge contributions with respect to the external poling field 𝐸3

are presented in Fig. 3(d). It is found that both 180° and 90° domain reversals exhibit a
non-linear increase and gradually reach saturation with the increase of poling electric
field. However, compared with 90° domain reversal, 180° domain reversal exhibits a
faster saturation tendency with respect to poling field. 180° domain reversal reaches
saturation at 3 kV/mm, while the other saturates at 5 kV/mm, which indicates that 90°
domain is much more difficult to reverse than that of 180° domain due to its large strain
distortion. Our experimental results also show that the contributions of 180° and 90°
domain reversals respectively account for ~29% and ~71% of the total domain reversals
in sufficiently poled tetragonal BS-PT ceramic samples. Similar results were also found
in PZT bulk ceramics, in which 90° domain reversal contributes by 80% to the total
Page 13 of 24 Journal of Materials Chemistry C

View Article Online


reversed domains.[31] DOI: 10.1039/C9TC00748B

Journal of Materials Chemistry C Accepted Manuscript


3.2. Contribution of 90° domain reversal to piezoelectricity
As shown in the SEM images of Fig. 4, both electric poling and thermal depoling
have direct effects on domain configuration. From Fig. 4(a)-(c), it can be seen that
distinct domain structures gradually grow up when poling electric field increases. As
for an unpoled sample, its domain structure is disordered because of the randomly
orientated spontaneous polarization. While under a medium poling electric field of 2
kV/mm, the domain configuration is ordered to some extent. After sufficiently poled at
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

4 kV/mm, the domain structure becomes fully ordered, see Fig. 4(c). Fig. 4(d), (e) and
(f) respectively show the microtopographies of the ceramic samples that annealed for 1
hour at 200 ℃, 400 ℃ and 450 ℃. Before annealing, these ceramic samples have been
sufficiently poled. It is found that the domain structure in the sample annealed at 200 ℃
almost remains the same as that of a sufficiently poled sample, implying that thermal
depolarization rarely happens at or below 200 ℃. When the annealing temperature
increases to 400 ℃, the ordered domain structure partly disappears due to thermal
depoling effect, see Fig. 4(e). It is further found that the ordered domain structure totally
disappears when the annealing temperature increases to 450 ℃, as shown in Fig. 4(f).
This is ascribed to the fact that ferroelectric-transition temperature, i.e., Curie
temperature, can induce domain extinction because of the disappeared spontaneous
polarization.
Even though the surface topography reveals an intuitionistic view of the domain
evolution under electric poling and thermal depoling, we still cannot present a
quantitative analysis of domain evolution and its effect on piezoelectric properties.
Additionally, we are unable to make quantitative analysis on the whole domain
reversals, for the reason that among 90° and 180° domain reversals, only 90° domain
reversal is detectable by XRD measurement. Nevertheless, considering that 90° domain
reversal makes the greatest contribution (up to ~71%) to the total domain reversal
during poling process, it is reasonably deduced that 90° domain reversal makes a
Journal of Materials Chemistry C Page 14 of 24

View Article Online


dominant contribution to the piezoelectric properties. Therefore, from the view of10.1039/C9TC00748B
DOI: 90°
domain reversal, we quantitatively reveal the correlation between domain structure and

Journal of Materials Chemistry C Accepted Manuscript


piezoelectric properties.
The quantitative estimation of 90° domain reversal is based on the diffraction-
plane-transformation model.[21] According to this model, the intensity of {111}
diffraction peak remains unchanged whether tetragonal piezoelectric ceramics is poled
or depoled, therefore, it can be used as the intensity reference to calculate the 90°
domain reversal percentage. The 90° domain reversal percentage in poling process can
be estimated by the intensity variation of {200} diffraction peak as:
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

𝑅200 ― 𝑅′200
𝑁𝑝𝑜𝑙𝑒 = 𝑅200 , (14)

Where Npole is the 90° domain reversal percentage during poling, 𝑅200 is the intensity
ratio of 𝐼200 𝐼111 in the unpoled ceramic sample, and 𝑅′200 is the intensity ratio of
𝐼′200 𝐼′111 in a poled ceramic sample.

Conversely, the 90° domain reversal percentage in depoling process can be


estimated by the intensity variation of {002} diffraction peak as:
𝑅002 ― 𝑅′002
𝑁𝑑𝑒𝑝𝑜𝑙𝑒 = 𝑅002 , (15)

Where Ndepole is the 90° domain reversal percentage during depoling, 𝑅002 is the
intensity ratio of 𝐼002 𝐼111 in a sufficiently poled ceramic sample, and 𝑅′002 is the
intensity ratio of 𝐼′002 𝐼′111 in a depolarized ceramic sample.
In tetragonal ferroelectric ceramics, 90° domain can be classified into two kinds,
i.e., c-domain and a-domain, which are defined according to their orientations. c-
domain refers to the domain that parallels to the poling direction, and a-domain is the
horizontal domain which is perpendicular to the poling direction.[32-33] On account of
the relation between domain volume and diffraction intensity, both c-domain and a-
domain can be revealed by diffraction intensity. The domain volume of a-domain can
be reflected by {200} diffraction intensity, while the domain volume of c-domain can
be reflected by {002} diffraction intensity.[34] Under a poling field, a-domain will
transform into c-domain via 90° domain reversal, which is reflected by the intensity
decrease of {200} diffraction peak as well as the intensity increase of {002} diffraction
Page 15 of 24 Journal of Materials Chemistry C

View Article Online


peak. Vice versa, c-domain will transform into a-domain via 90° domain reversal in
DOI: 10.1039/C9TC00748B

depoling process, which is also reflected by the intensity variations of {002} and {200}

Journal of Materials Chemistry C Accepted Manuscript


diffraction peaks. The intensity variations of {002} and {200} diffraction peaks in
poling and depoling processes are experimentally revealed by XRD measurement. Fig.
5 and Fig. 6 respectively show the XRD profiles for ceramic samples poled and depoled
under different conditions. As can be seen in Fig. 5, the intensity of {200} diffraction
peak gradually decreases when poling electric field increases, which indicates that the
volume of a-domain decreases during poling. Similarly, in Fig. 6, the intensity of the
{002} diffraction peak decreases, demonstrating that the volume of c-domain decreases
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

during depoling. Based on the XRD profiles and equations (14)-(15), we can
quantitively estimate the 90° domain reversal percentage in both poling and depoling
processes.
Fig. 7(a), (b), (c) and (d) respectively show the variations of piezoelectric constant
d33, permittivity εr, electromechanical coupling factor kp and 90° domain reversal
percentage as a function of poling field and annealing temperature. Strong consistency
is found among these material properties. Both of them initially increase with the
increasing poling field, and reach their saturated values when poling electric field
increases to ~3 kV/mm, indicating that the ceramic sample gradually reaches a
sufficiently polarized state. The 90° domain reversal percentage Npole increases with the
increase of poling electric field, and it finally reaches ~53.8% at the field of 4 kV/mm.
Similar 90° domain reversal percentage in tetragonal PZT piezoelectric ceramic under
electric poling was also found in previously reported works.[35-36]
As equally important as the poling process, depolarization of piezoelectric
ceramics is also widely investigated.[37-39] The conventional methods that investigate
depolarization can be classified into two kinds, i.e., in situ and ex situ.[40] Recently, an
in situ techniques is developed to reveal the real-time high temperature performance
and depolarization characteristics of piezoelectric ceramics.[41] Differently, we
investigate the depolarization of BS-PT ceramics in an ex situ way. After sufficiently
poled and anterior to any characterization of depolarization, BS-PT ceramic samples
are annealed for 1 hour under different temperatures. Here, it must be noted that silver
Journal of Materials Chemistry C Page 16 of 24

View Article Online


electrode of ceramic sample has to be removed before the sample is thermally DOI:
depoled,
10.1039/C9TC00748B

because the electrode may have effects on the depolarization, such as the thermal

Journal of Materials Chemistry C Accepted Manuscript


expansion mismatch between ceramic sample and electrode. Fig. 7(a), (b), (c) and (d)
also show the variations of d33, εr, kp and Ndepole as a function of annealing temperature.
During thermal depoling, it is found that almost no depolarization occurs when
annealing temperature is below 200 ºC, convinced by the phenomena that d33, εr, kp and
Ndepole stably maintain their initial value. However, once the annealing temperature
increases to 300 ºC or above, a remarkable depolarization is triggered. Ndepole increases
with the increase of annealing temperature, and it becomes dominant (~34.9%) at the
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

ferroelectric-phase transition temperature of 450 ºC. Concurrently, d33, εr and kp also


reveal a fast decline when annealing temperature exceeds 300 ºC. From the view of 90°
domain reversal, it can be demonstrably deduced that depolarization is almost
negligible at temperatures below 200 °C, because Ndepole is below 5%. This is in well
accordance with the material properties that piezoelectric constant d33 also remains as
high as 95% of its initial value. Hence, it is concluded that the highest safety working
temperature for BS-PT piezoelectric ceramics is near the half of Curie temperature
(~450 ºC), which coincides well with our previous results.[15, 42] The quantitative study
of 90° domain reversal provide us with a more direct perspective to see that BS-PT
ceramic has a good thermal stability. The high temperature resistance of BS-PT ceramic
proves its promising prospect in the high-temperature applications for aerospace,
automotive industries and oil drilling, etc.

4. Conclusions
In the present work, domain evolution of tetragonal BS-PT piezoelectric ceramics,
as well as its effects on domain microtopography and piezoelectric properties, are
systematically investigated in electric poling and thermal depoling processes. By
introducing Gaussian probability distribution function of domain reversal, Uchida-
Ikeda model is modified to distinguish 90° domain reversal and 180° domain reversal
based on their charge contributions. According to the model, it is estimated that 180°
Page 17 of 24 Journal of Materials Chemistry C

View Article Online


and 90° domain reversals account for 29% and 71%, respectively, in a sufficiently
DOI:poled
10.1039/C9TC00748B

ceramic sample, which coincides well with the result observed by domain-reversal-

Journal of Materials Chemistry C Accepted Manuscript


induced charge.
Based on the quantitative study of 90° domain via XRD measurement, BS-PT
ceramics are proved to have a high-temperature stability and its piezoelectric constant
remains as high as 95% of its initial value after depoling at 200 °C, which is supported
by the insignificant 90° domain reversal percentage of 5%. The insight into domain
evolution of tetragonal BS-PT piezoelectric ceramics during electric poling and thermal
depoling provides us with a better understanding of the relation between
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

piezoelectricity and domain structure, which is also of heuristic significance to other


perovskite piezoelectric ceramics.

Acknowledgments
Jingen Wu appreciates the support from the Fundamental Research Funds for the
Central Universities. This work was supported by the National Natural Science
Foundation of China (Grant Nos. 51132001, 51072003), the Beijing Municipal Science
and Technology Projects (Grant Nos. Z131100003213020, Z151100003715003), the
United States Office of Naval Research (ONR Grants No. N00014-12-1-1045 and
N00014-16-1-3106) and the Natural Sciences & Engineering Research Council of
Canada (Grant No. 203773).

ORCID
Jingen Wu: 0000-0001-9289-5013
Zuo-Guang Ye: 0000-0003-2378-7304
Shuxiang Dong: 0000-0002-9617-6013
Journal of Materials Chemistry C Page 18 of 24

View Article Online

References DOI: 10.1039/C9TC00748B

[1] A. Roelofs, N. A. Pertsev, R. Waser, F. Schlaphof, L. M. Eng, C. Ganpule, V. Nagarajan


and R. Ramesh, Appl. Phys. Lett., 2002, 80, 1424-1426.

Journal of Materials Chemistry C Accepted Manuscript


[2] L. Chen, J. Ouyang, C. S. Ganpule, V. Nagarajan, R. Ramesh and A. L. Roytburd, Appl.
Phys. Lett., 2004, 84, 254-256.
[3] C. S. Ganpule, V. Nagarajan, H. Li, A. S. Ogale, D. E. Steinhauer, S. Aggarwal, E. Williams,
R. Ramesh and P. De Wolf, Appl. Phys. Lett., 2000, 77, 292-294.
[4] F. Xu, S. Trolier-McKinstry, W. Ren, B. Xu, Z. L. Xie and K. J. Hemker, J. Appl. Phys.,
2001, 89, 1336-1348.
[5] N. Setter, D. Damjanovic, L. Eng, G. Fox, S. Gevorgian, S. Hong, A. Kingon, H. Kohlstedt,
N. Y. Park, G. B. Stephenson, I. Stolitchnov, A. K. Taganstev, D. V. Taylor, T. Yamada
and S. Streiffer, J. Appl. Phys., 2006, 100, 51606.
[6] K. Wang and J. Li, Adv. Funct. Mater., 2010, 20, 1924-1929.
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

[7] Y. Qin, J. Zhang, Y. Gao, Y. Tan and C. Wang, J. Appl. Phys., 2013, 113, 204107.
[8] Y. Qin, J. Zhang, Y. Tan, W. Yao, C. Wang and S. Zhang, J. Eur. Ceram. Soc., 2014, 34,
4177-4184.
[9] G. Arlt and P. Sasko, J. Appl. Phys., 1980, 51, 4956-4960.
[10] R. E. Eitel, C. A. Randall, T. R. Shrout and S. Park, Jpn. J. Appl. Phys., 2002, 41, 2099-
2104.
[11] R. E. Eitel, S. J. Zhang, T. R. Shrout, C. A. Randall and I. Levin, J. Appl. Phys., 2004, 96,
2828-2831.
[12] X. Li, J. Chen, Z. Chen and S. Dong, Appl. Phys. Lett., 2012, 101, 72902.
[13] J. Chen, Z. Chen, X. Li and S. Dong, Appl. Phys. Lett., 2013, 102, 52902.
[14] J. Chen, X. Li, G. Liu, Z. Chen and S. Dong, Appl. Phys. Lett., 2012, 101, 12909.
[15] J. Wu, X. Chen, Z. Chu, W. Shi, Y. Yu and S. Dong, Appl. Phys. Lett., 2016, 109, 173901.
[16] J. Wu, H. Shi, T. Zhao, Y. Yu and S. Dong, Adv. Funct. Mater., 2016, 26, 7186-7194.
[17] J. Íñiguez, D. Vanderbilt and L. Bellaiche, Phys. Rev. B, 2003, 67, 224107.
[18] J. L. Jones, E. Aksel, G. Tutuncu, T. Usher, J. Chen, X. Xing and A. J. Studer, Phys. Rev.
B, 2012, 86, 024104.
[19] K. V. Lalitha, A. N. Fitch and R. Ranjan, Phys. Rev. B, 2013, 87, 064106.
[20] J. Wu, X. Gao, Y. Yu, J. Yang and S. Dong, J. Alloy Compd., 2018, 745, 669-676.
[21] T.-L. Zhao, A. A. Bokov, J. Wu, H. Wang, C.-M. Wang, Y. Yu, C.-L. Wang, K. Zeng, Z.-
G. Ye and S. Dong, Adv. Funct. Mater., 2019, 1807920, 1-10.
[22] J. Wu, Y. Yu, X. Li, X. Gao and S. Dong, J. Am. Ceram. Soc., 2015, 98, 3145-3152.
[23] S. J. Zhang, E. F. Alberta, R. E. Eitel, C. A. Randall and T. R. Shrout, IEEE T. Ultrason.
Ferr., 2005, 52, 2131-2139.
[24] Y. Qin, J. Zhang, W. Yao, C. Wang and S. Zhang, J. Am. Ceram. Soc., 2015, 98, 1027-
1033.
[25] F. X. Li and R. K. N. D. Rajapakse, J. Appl. Phys., 2007, 101, 54110.
[26] Y. Li, Y. Sun and F. Li, Ceram. Int., 2013, 39, 8605-8614.
[27] N. Uchida and T. Ikeda, Jpn. J. Appl. Phys., 1967, 6, 1079-1088.
[28] L. K. V., C. M. Fancher, J. L. Jones and R. Ranjan, Appl. Phys. Lett., 2015, 107, 52901.
[29] J. L. Jones, E. B. Slamovich and K. J. Bowman, J. Appl. Phys., 2005, 97, 34113.
Page 19 of 24 Journal of Materials Chemistry C

View Article Online


[30] Y. Saito, Jpn. J. Appl. Phys., 1997, 36, 5963-5969. DOI: 10.1039/C9TC00748B
[31] S. Li, A. S. Bhalla, R. E. Newham and L. E. Cross, J. Mater. Sci., 1994, 29, 1290-1294.
[32] Y. W. Li and F. X. Li, Mech. Mater., 2016, 93, 246-256.
[33] C. S. Ganpule, V. Nagarajan, H. Li, A. S. Ogale, D. E. Steinhauer, S. Aggarwal, E. Williams

Journal of Materials Chemistry C Accepted Manuscript


and R. Ramesh, Appl. Phys. Lett., 2000, 77, 292.
[34] J. L. Jones and M. H. E. D. J. Studer, Appl. Phys. Lett., 2006, 89, 092901.
[35] C. Bedaya, C. Muller, J. L. Baudour, V. Madigou, M. Anne and M. Roubin, Mat. Sci. Eng.
B, 2000, 75, 43-52.
[36] X. Zhang, C. Lei and K. Chen, J. Am. Ceram. Soc., 2005, 88, 335-338.
[37] H. Yan, H. Zhang and M. J. Reecea, Appl. Phys. Lett., 2005, 87, 082911.
[38] S. Chen, X. Dong, C. Mao and F. Cao, J. Am. Ceram. Soc., 2006, 89, 3270-3272.
[39] T. Y. Ansell, D. P. Cann, E. Sapper and J. Rödel, J. Am. Ceram. Soc., 2015, 98, 455-463.
[40] Y. Wang, K. Cai, T. Shao, Q. Zhao and D. Guo, J. Appl. Phys., 2015, 117, 164102.
[41] C. Huang, K. Cai, Y. Wang, Y. Bai and D. Guo, J. Mat. Chem. C, 2018, 6, 1433-1444.
Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

[42] J. Chen, G. Liu, X. Li, Z. Chen and S. Dong, IEEE T. Ultrason. Ferr., 2013, 60, 446-450.
Journal of Materials Chemistry C Page 20 of 24

View Article Online


DOI: 10.1039/C9TC00748B

Journal of Materials Chemistry C Accepted Manuscript


Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

Fig. 1 (a) schematics of 180° domain reversal in electric poling process, (b) schematics
of 90° domain reversal in electric poling process. Here, both 180° and 90° reversed
dipoles are assumed to obey the Gaussian distribution with respect to their spatial angle
𝜑. 𝑓180(𝜑) (including 𝑓180(𝜑)𝐼 and 𝑓180(𝜑)𝐼𝐼) and 𝑓90(𝜑) are the Gaussian

probability distribution functions of the 180° and 90° reversed dipoles, respectively.
Page 21 of 24 Journal of Materials Chemistry C

View Article Online


DOI: 10.1039/C9TC00748B

Journal of Materials Chemistry C Accepted Manuscript


Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

Fig. 2 Switching currents and charge contributions of capacity stored charge (𝑄𝐶0),
180° domain reversal induced charge (𝑄180) and 90° domain reversal induced charge
(𝑄90) under: (a) low (0.25 kV/mm), (b) medium (2 kV/mm), and (c) high (5 kV/mm)
poling electric fields. All the measurements were performed at room temperature
under 1 Hz.
Journal of Materials Chemistry C Page 22 of 24

View Article Online


DOI: 10.1039/C9TC00748B

Journal of Materials Chemistry C Accepted Manuscript


Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

Fig. 3 (a) time dependence of switching current obtained from BS-PT ceramic sample
under different poling fields; (b) time dependence of the total induced charge under
different poling fields; (c) time dependence of switching current induced by capacity
stored charge (𝑄𝐶0), 180° domain reversal induced charge (𝑄180) and 90° domain
reversal induced charge (𝑄90) under the poling filed of 3 kV/mm; (d) field dependence
of 180° domain reversal percentage (X180) and 90° domain reversal percentage (X90).
All the measurements were performed at room temperature under 1 Hz.

Fig. 4 SEM images of domain microtopography for BS-PT ceramic samples: (a)
unpoled; (b) poled at 2 kV/mm; (c) poled at 4 kV/mm; (d) annealed at 200 ºC for 1h;
(e) annealed at 400 ºC for 1h; (f) annealed at 450 ºC for 1h.
Page 23 of 24 Journal of Materials Chemistry C

View Article Online


DOI: 10.1039/C9TC00748B

Journal of Materials Chemistry C Accepted Manuscript


Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

Fig. 5 XRD patterns in the 2 theta range of 20° to 60° for BS-PT ceramic samples
poled under different electric fields. The enlarged XRD patterns in the 2 theta range
of 43° to 47° are also presented.

Fig. 6 XRD patterns in the 2 theta range of 20° to 60° for BS-PT ceramic samples
thermally depoled under different annealing temperatures. The enlarged XRD patterns
in the 2 theta range of 43° to 47° are also presented.
Journal of Materials Chemistry C Page 24 of 24

View Article Online


DOI: 10.1039/C9TC00748B

Journal of Materials Chemistry C Accepted Manuscript


Published on 14 March 2019. Downloaded on 3/17/2019 7:06:59 AM.

Fig. 7 The effects of electric poling and thermal depoling on the material properties of
BS-PT ceramic: (a) piezoelectric constant d33, (b) relative permittivity 𝜀𝑟, (c)
electromechanical coupling factor 𝑘𝑝, (d) Npole and Ndepole domain reversal
percentages.

You might also like