Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Ceramics International xxx (xxxx) xxx

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Enhanced strain and reduced dielectric loss in a bismuth-based high-TC


ternary solid solution system of BiScO3-Pb(Sc1/2Nb1/2)O3-PbTiO3
Zenghui Liu a, *, Zhenjun Shao a, Zeng Luo b, Jun Xu a, Yunjian Cao a, Hao Li a, Hongyan Wan a, **,
Ruihua An a, Nan Zhang a, Yijun Zhang a, Gang Niu a, c, Wei Ren a, c
a
Electronic Materials Research Laboratory, Key Laboratory of the Ministry of Education & International Center for Dielectric Research, School of Electronic Science and
Engineering, Xi’an Jiaotong University, Xi’an, 710049, China
b
Air and Missile Defense College, Air Force Engineering University, Xi’an, 710051, China
c
State Key Laboratory for Manufacturing Systems Engineering & International Joint Laboratory for Micro/Nano Manufacturing and Measurement Technology, Xi’an
Jiaotong University, Xi’an, 710049, China

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

Handling Editor: Dr P. Vincenzini To design novel materials suitable for high-temperature electromechanical applications, a bismuth-based ternary
solid solution system, (1-x-y)BiScO3-yPb(Sc1/2Nb1/2)O3-xPbTiO3, with compositions situated around the mor­
Keywords: photropic phase boundary, were designed and synthesized in the form of ceramic. The crystal structure,
High Curie temperature microstructure, dielectric, ferroelectric and piezoelectric properties, and the local polar domain structure of the
Strain
ceramics were investigated in detail. The examined ceramics exhibit large grain sizes ranging from 20 μm to 45
Dielectric loss
μm, along with noteworthy characteristics, including a relatively large strain of 0.21 %, a high large-signal
Piezo-/ferroelectric materials
BiScO3-PbTiO3 piezoelectric coefficient (d33* = 368 pm/V), a low dielectric loss (tanδ = 1.7 %), and a high Curie tempera­
Morphotropic phase boundary ture (TC ~ 423 ◦ C). These enhanced properties signify the suitability of the designed ternary ceramics for future
high-temperature electromechanical applications. This work demonstrates an effective approach for designing
ferro-/piezoelectric materials characterized by substantial strain and low dielectric loss.

1. Introduction response (d33 ≤ 40 pC/N) [9–11].


Recently, special attention has been paid to bismuth-based ferro-/
High-temperature piezoelectric materials are of paramount impor­ piezoelectric perovskites primarily due to their remarkable combination
tance across various industries, including aerospace, energy generation, of high TC, substantial polarization, and exceptional piezoelectric
and automotive engineering. These materials are central to the fabri­ properties arising from their unique crystal chemistry feature [12,13].
cation of sensors and actuators capable of functioning under elevated Among this family of compounds, the BiScO3-PbTiO3 (BS-PT) system has
temperatures (>200 ◦ C), delivering precise measurements and control emerged as a key contender for potential applications at temperatures
in situations where conventional materials would prove inadequate exceeding 300 ◦ C. Notably, a high TC up to 450 ◦ C and a large piezo­
[1–5]. Traditional lead zirconate titanate (known as PZT)-based piezo­ electric coefficient d33 of 450 pC/N have been reported for the
electric ceramics with the compositions near the morphotropic phase 0.36BS-0.64PT ceramics, with an MPB composition bridging the rhom­
boundary (MPB) region exhibit competitive piezoelectric and electro­ bohedral and tetragonal phases. The performance of BS-PT is superior to
mechanical properties. However, PZT-based ceramics are mainly used at that of the traditional PZT ceramics [14–17]. However, the BS-PT-based
lower temperatures due to their Curie temperature (TC) falling below ceramics suffer from a relatively high dielectric loss (tanδ >3 %), which
350 ◦ C, which may result in depolarization at temperatures exceeding gives rise to serious heat generation, and energy dissipation more seri­
200 ◦ C [6–8]. In contrast, bismuth layer-structured ceramics, such as ously at high temperatures and under large electric fields. This con­
Bi4Ti3O12 (TC ~ 675 ◦ C) and CaBi2Nb2O9 (TC ~ 940 ◦ C), consistently strains their suitability for high-temperature applications [17–19]. In
demonstrate substantially higher TC values. Nevertheless, their practical the efforts to mitigate the dielectric loss, strategies such as doping with
applications are hindered by their relatively modest piezoelectric MnO2 and introduction of a third end member, such as Bi(Mn1/2Zr1/2)O3

* Corresponding author.
** Corresponding author.
E-mail addresses: liu.z.h@xjtu.edu.cn (Z. Liu), wanhongyan@stu.xjtu.edu.cn (H. Wan).

https://doi.org/10.1016/j.ceramint.2024.02.008
Received 14 November 2023; Received in revised form 31 January 2024; Accepted 1 February 2024
Available online 2 February 2024
0272-8842/© 2024 Elsevier Ltd and Techna Group S.r.l. All rights reserved.

Please cite this article as: Zenghui Liu et al., Ceramics International, https://doi.org/10.1016/j.ceramint.2024.02.008
Z. Liu et al. Ceramics International xxx (xxxx) xxx

or Pb(Mn1/3Nb2/3)O3, into the BS-PT system to form ternary solid so­ compositions (y = 0, x = 0.61–0.66; y = 0.025, x = 0.59–0.64; y = 0.05,
lutions have been reported. Unfortunately, the dielectric loss in these x = 0.58–0.63) was carried out using highly pure (>99.9 %) PbO, Bi2O3,
modified ceramics is still relatively high, rendering them unsuitable for Sc2O3, TiO2, and Nb2O5 powders as reagents. Before weighing, all these
high-temperature electromechanical applications. Moreover, the powders were preheated in an oven at 80 ◦ C for 12 h to remove the
piezoelectric coefficient d33 and TC experience significant deterioration moisture and other adsorbed gases such as CO2, and to ensure the
in comparison to the unmodified BS-PT ceramics [20–24]. stoichiometry of the designed compositions. A wolframite precursor
Pb(Sc1/2Nb1/2)O3 (PSN) is a complex perovskite material that ex­ method was employed to prevent the formation of the undesired pyro­
hibits either ferroelectric or relaxor behavior, contingent upon the de­ chlore phase [29]. Initially, the ScNbO4 (SN) precursor was synthesized
gree of Sc and Nb ordering on the B-site of its structure [25,26]. PSN and by calcining the mixed Sc2O3 and Nb2O5 powders at 1150 ◦ C for 2 h.
PbTiO3 (PT) are capable of forming a continuous solid solution, which Subsequently, SN was combined with the other oxides in stoichiometric
possesses an MPB at the composition near 0.57PSN-0.43PT. The PSN-PT proportions. To compensate for the volatilization of Bi2O3 and PbO at
ceramics showcase outstanding electric properties, including a high d33 high temperatures, an excess of 2 mol% was added. The mixed powders
(>500 pC/N), a substantial strain level of up to 0.6 %, elevated elec­ were ball-milled in the presence of ethanol for 8 h using zirconia balls in
tromechanical coupling factors (kp ~ 82 % and k33 ~ 57 %), a relatively sealed jars. The resultant mixture was dried and further ground for 0.5 h
low dielectric loss (tanδ <2 %), and a moderate TC (~254 ◦ C) [27,28]. using an agate mortar and pestle. Following this, the powders were
In this work, we present a novel approach, involving the incorpo­ pressed and calcined in a sealed alumina crucible at 850 ◦ C for 4 h to
ration of the PSN end member into the BS-PT system and the fabrication form the perovskite phase. After calcination, the powders underwent
of ceramics featuring large grain sizes ranging from 20 μm to 45 μm, to another round of ball milling and drying for granulation, and 10 wt%
address the above-mentioned issue. This strategy is employed to polyvinyl acetate (PVA) was added as a binder. The granulated powder
significantly augment strain, reduce the dielectric loss, and simulta­ was then uniaxially pressed into 13 mm diameter discs with a pressure of
neously maintain a high TC. The effects of the PSN modification on the 300 MPa for sintering. A sacrificial powder with the same composition
crystal structure, microstructure, and dielectric, ferroelectric and was used during the sintering process. To entirely remove the PVA
piezoelectric properties of BS-PT are investigated in detail. Ultimately, abinder, the pellets were heated to 650 ◦ C and held at this temperature
we have successfully achieved a relatively high strain of 0.21 %, an for 1 h. Subsequently, the pellets were sintered at temperatures ranging
appealing large-signal piezoelectric coefficient (d33* = 368 pm/V), a from 1000 ◦ C to 1200 ◦ C for 2 h to produce dense ceramics.
relatively low dielectric loss (tanδ = 1.7 %), and a high Curie temper­
ature (TC ~ 423 ◦ C). 2.3. Characterization

2. Experimental procedure X-ray diffraction (XRD) was performed on crushed ceramic powder
by utilizing a PANalytical X’Pert Pro MPD diffractometer with Cu Kα1
2.1. MPB and composition design radiation to determine the crystal structure. The relative density of the
as-sintered ceramics was measured by the Archimedes’ method, and
It is crucial to determine the MPB compositions in a multicomponent their microstructure was imaged using field emission scanning electron
system to achieve high electromechanical properties. The ternary MPB microscopy (SEM). The average grain size and the distributions were
of the (1-x-y)BiScO3-yPb(Sc1/2Nb1/2)O3-xPbTiO3 solid solution system, estimated from the SEM images using the linear interception method via
denoted as (1-x-y)BS-yPSN-xPT is projected based on the linear region the Nano Measure software. Raman spectroscopy measurements were
between the two binary solid solution systems of BS-PT and PSN-PT. performed at room temperature using a 532 nm excitation laser source
Within the ternary phase diagram, three distinct lines are selected, as (RTS2-501-SMSXS, Zolix, China).
illustrated in Fig. 1, described as (1-x)BiScO3-xPbTiO3, (0.975-x)BiScO3- The samples were coated with silver paste on both circular surfaces
0.025 Pb(Sc1/2Nb1/2)O3-xPbTiO3, and (0.95-x)BiScO3-0.05 Pb(Sc1/ as electrodes and then fired at 500 ◦ C for 10 min to achieve a good ohmic
2Nb1/2)O3-xPbTiO3, respectively. contact before subsequent electrical measurements. The dielectric
properties were characterized using a broadband dielectric impedance
2.2. Samples preparation spectrometer (E4981A Precision LCR, Agilent, Hewlett-Packard, USA)
equipped with a laboratory-made high-temperature furnace using a
The synthesis of (1-x-y)BS-yPSN-xPT ceramics with various EUROTHERM 3204 PID temperature controller. The polarization-
electric field (P-E) hysteresis loops and the strain of the samples were
displayed at room temperature using a ferroelectric testing system (TF
Analyzer 2000 Systems, Aix ACCT, Germany) combined with a laser
interferometric vibrometer (SPeS 120, SIOS Messtechnik GmbH, Ger­
many) under ±45 kV/cm and 10 Hz. For piezoresponse force micro­
scopy (PFM) studies, the samples underwent mirror polishing and were
subsequently annealed at 600 ◦ C for 0.5 h to remove any residual stress.
The domain structure and its dynamics were examined using PFM,
conducted with a modified commercially atomic force microscopy sys­
tem operating in piezoresponse mode (AFM, Dimension ICON, Nano­
Scope V, Bruker, Germany).

3. Results and discussion

3.1. Crystal structure

Fig. 2 illustrates the XRD patterns of the (1-x-y)BS-yPSN-xPT ce­


ramics at room temperature. All the XRD patterns manifest a pure
perovskite structure without any trace of pyrochlore phase, under­
Fig. 1. The designed MPB in the ternary phase diagram of (1-x-y)BS-yPSN-xPT scoring the formation of a stable ternary solid solution. Furthermore, no
and the compositions studied in this work. superlattice corresponding to cations ordering or oxygen octahedral tilt

2
Z. Liu et al. Ceramics International xxx (xxxx) xxx

Fig. 2. X-ray diffraction patterns of the (1-x-y)BS-yPSN-xPT ceramics (sintered at 1150 ◦ C): (a) y = 0, (b) y = 0.025, and (c) y = 0.05. Rietveld refinements of the (d)
y = 0.025, x = 0.64 and (e) y = 0.025, x = 0.62 ceramics.

was observed in these compositions. and a goodness of fit (gof) = 2.52. The proportions of the two phases
The phase symmetry gradually evolves with the PT content, as were determined as 75.2 % for P4mm and 24.8 % for Cm, respectively.
prominently evidenced by the (111)C and (200)C peaks (with the These fitting results indicate a gradual transition from the mixed phases
subscript ‘‘C’’ denoting a cubic crystal lattice setting). In the case of y = of P4mm and Cm in the MPB region to the PT-rich P4mm phase. Neutron
0, when x ≥ 0.65, a single and symmetrical (111)C peak is observed and synchrotron diffraction experiments are in progress to provide
alongside a doublet (200)C peak, signifying a typical tetragonal struc­ further confirmation of the crystal structure within this system.
ture. The (200)C peak becomes asymmetrical at x = 0.64 and transforms Raman spectra of the (1-x-y)BS-yPSN-xPT ceramics at room tem­
into a triple feature at x = 0.63. In addition, a shoulder emerges at a perature are shown in Fig. 3. For y = 0, the intensity of the Raman peak
lower angle in the (111)C peak. As the PT content diminishes, the (111)C at approximately 610 cm-1 gradually increases, while the peak at 547
and (200)C peaks continue to exhibit splitting. This phenomenon is cm-1 decreases with an augmentation in PT content. These peaks
characteristic of an MPB feature, consistent with previous studies [17, correspond to the vibration of the BO6 octahedra. Meanwhile, the peak
30]. For the y = 0.025 and 0.05 series, a similar sequence of phase at around 282 cm-1, indicative of B-O bending and/or stretching mode,
evolution is observed. shifts to a lower frequency region, suggesting a softening of the lattice
Compared with BS-PT, the MPB region progressively shifts toward vibration [32–34]. Additionally, at x = 0.64, the Raman spectra reveal
lower PT content in the PSN-modified systems. Within the MPB region, the disappearance of the peak at around 251 cm-1, replaced by new
the coexistence of a monoclinic MA phase (Cm space group) and a peaks at approximately 126 cm-1 and 335 cm-1. The sudden change in
tetragonal phase (P4mm space group) were reported in the BS-PT system Raman spectra aligns with the composition-induced phase transition
[30,31], which is believed to facilitate the polarization rotation and from MPB to tetragonal, as indicated by XRD results. The y = 0.025 and
enhance the piezoelectric and dielectric properties. Therefore, it is 0.05 series exhibit similar variations in Raman spectra (Fig. 3(b) and
reasonable to anticipate the presence of analogous phase components in (c)), providing further confirmation of the structural analysis.
the ternary ceramic systems currently under investigation.
Rietveld refinement was performed on two selected compositions, y 3.2. Microstructure
= 0.025, x = 0.64 and y = 0.025, x = 0.62, to investigate the crystal
structure of the (1-x-y)BS-yPSN-xPT ternary ceramics by using the The surface topography of the (1-x-y)BS-yPSN-xPT ceramic samples
TOPAS Academic software. As illustrated in Fig. 2(d) and (e), a tetrag­ sintered at 1150 ◦ C is presented in Fig. 4. All of the samples exhibit a
onal structure with P4mm space group was found to a good fit for the fairly dense and homogeneous microstructure without any observable
diffraction pattern of the y = 0.025, x = 0.64 ceramic. In contrast, a pores. The high quality of the samples is further evidenced by the
single-phase model of P4mm could not well fit the y = 0.025, x = 0.62 achievement of a high relative density exceeding 96 %. The presence of a
diffraction pattern. Therefore, structural models consisting of a mixture dense and uniform morphology underscores the favorable mechanical
of P4mm and Cm phases were employed for data fitting, yielding satis­ properties of the as-prepared (1-x-y)BS-yPSN-xPT ceramics, a critical
factory refinement results with Rwp = 8.8 %, Rexp = 3.5 %, Rp = 6.6 %, aspect for applications in high-temperature devices.

3
Z. Liu et al. Ceramics International xxx (xxxx) xxx

Fig. 3. Raman spectra of the (1-x-y)BS-yPSN-xPT ceramics (sintered at 1150 ◦ C): (a) y = 0, (b) y = 0.025, and (c) y = 0.05.

Fig. 4. SEM micrographs of the (1-x-y)BS-yPSN-xPT ceramics sintered at 1150 ◦ C: (a) y = 0, (b) y = 0.025, and (c) y = 0.05.

4
Z. Liu et al. Ceramics International xxx (xxxx) xxx

It is notable that the grain size in the obtained samples is relatively The modified Curie-Weiss law is often employed to describe the
large, averaging about 20 μm to 45 μm. The presence of these large dielectric property of a ferroelectric with diffuse phase transformation:
grains effectively eliminates porosity and promotes densification,
1 1 (T − Tm )γ
which, in turn, aids in reducing dielectric loss and enhancing piezo­ − = ,
ε εm C
electric properties. The large grain size may be attributed to the inclu­
sion of additional 2 % Bi2O3 and PbO in the initial materials, resulting in where C, γ, and εm are the Curie constant, diffusion coefficient, and
the formation of a transient liquid phase that facilitates grain growth maximum dielectric permittivity, respectively. In general, the γ falls
[35,36]. Moreover, the grain size demonstrates a gradual increase with between 1 and 2, where γ = 1 signifies a typical ferroelectric behavior, γ
increasing PT content at a given PSN content. Conversely, the increase in = 2 indicates an ideal relaxor behavior, and 1 < γ < 2 corresponds to a
PSN content exerts a suppressing effect on grain growth. diffuse phase transformation. The calculated diffusion coefficients range
from 1.56 to 1.66, 1.61 to 1.69, and 1.61 to 1.78 for the y = 0, 0.025, and
3.3. Dielectric properties and phase transition 0.05 series at the MPB, respectively, supporting the notion of a diffuse
phase transformation. The increased diffuseness can be attributed to the
In order to evaluate the dielectric property and investigate the phase introduction of PSN, which augments chemical heterogeneities and
transition behavior, the dielectric permittivity and loss were measured associated disorder effects [37–39].
as the temperature increased from room temperature to 500 ◦ C, fol­ It is worth noting that the dielectric loss observed in the BS-PT ce­
lowed by subsequent cooling. Fig. 5 depicts the temperature dependence ramics prepared in this study is rather small, ranging from 0.4 % to 2.1 %
of the dielectric properties of the selected (1-x-y)BS-yPSN-xPT ceramics at room temperature, when measured under 10 kHz. This represents an
at various frequencies. improvement compared to previous reports where losses of approxi­
In the case of y = 0, the dielectric curve exhibits a relatively sharp mately 3 % were observed [17–19]. The low dielectric loss at room
peak in permittivity, with the dielectric maximum temperature (Tm) temperature is consistently observed in the (1-x-y)BS-yPSN-xPT ce­
displaying a weak (or negligible) frequency dependence. This behavior ramics, and it further decreases at elevated temperatures by introducing
signifies the phase transition from the ferroelectric to paraelectric pha­ PSN, as indicated in Table 1. This reduced loss can be attributed to the
ses at the Curie temperature TC (or Tm). The TC gradually increases from notably large grain sizes, which restrict space charge migration (inter­
422 ◦ C for x = 0.61 to 457 ◦ C for x = 0.66, thus confirming a notably facial polarization contribution) within the ceramics, in addition to the
high TC feature in the vicinity of the MPB. Concurrently, the dielectric beneficial effects of PSN substitution [40–42]. At high temperatures
permittivity initially increases, reaches its peak at x = 0.64, and sub­ (>300 ◦ C), the dielectric loss increases significantly, especially
sequently decreases across the MPB region, as detailed in Table 1. measured under lower frequencies, which could be due to the increased
The incorporation of PSN into BS-PT results in the broadening and leakage/conductivity related to the thermally activated defects [23].
suppression of the dielectric peak, whereas the TC (or Tm) remains
almost unchanged with frequency, indicating a typical diffuse phase 3.4. Ferroelectric and piezoelectric properties
transformation characteristic [37,38]. This phenomenon is particularly
pronounced in the compositions with lower PT contents. In the y = Fig. 6 illustrates the polarization-electric field (P-E) loops of the (1-x-
0.025 and 0.05 series, the permittivity and TC show a similar variation y)BS-yPSN-xPT ceramics measured at room temperature under 10 Hz. In
trend with x, while the peak position shifts gradually to the lower PT the BS-PT ceramics, full saturation cannot be achieved under ±40 kV/
side compared with the y = 0 series. Additionally, the permittivity in­ cm, beyond which the samples are prone to breakdown. However, the
creases while TC slightly decreases with the same PT content. None­ PSN-modified BS-PT ceramics exhibit saturated P-E hysteresis loops at
theless, the (1-x-y)BS-yPSN-xPT ceramics consistently maintain a ±50 kV/cm, demonstrating a significantly enhanced polarization, as
relatively high TC above 386 ◦ C, rendering them promising candidates depicted in Fig. 6.
for high-temperature electromechanical applications. A noticeable trend in the (1-x-y)BS-yPSN-xPT ceramics is that the

Fig. 5. Temperature dependences of the dielectric permittivity and loss of the representative (1-x-y)BS-yPSN-xPT ceramics (sintered at 1150 ◦ C) measured at various
frequencies upon cooling.

5
Z. Liu et al. Ceramics International xxx (xxxx) xxx

Table 1
Structure and properties of the (1-x-y)BS-yPSN-xPT ceramics.
Material Structure TC (◦ C) εr @10 kHz tanδ @10 kHz (%) Strain (%) d33* (pm/V) EC (kV/cm) Pr (μC/cm2)

30 C

300 C

30 C ◦
300 C ◦

0.39BS-0.61PT MPB 422 674 1986 2.1 19.0 0.09 180 19 10


0.38BS-0.62PT MPB 428 688 2134 1.7 18.7 0.09 200 22 15
0.37BS-0.63PT MPB 438 777 2381 0.9 31.1 0.10 226 19 10
0.36BS-0.64PT MPB 444 954 3136 1.6 21.6 0.12 238 24 16
0.35BS-0.65PT Tetragonal 452 740 2288 0.8 29.1 0.04 120 16 4
0.34BS-0.66PT Tetragonal 457 654 1975 0.4 36.1 0.03 74 23 6
0.385BS-0.025PSN-0.59PT MPB 401 761 3022 1.5 9.5 0.12 235 23 21
0.375BS-0.025PSN-0.60PT MPB 404 758 2768 1.4 10.4 0.16 293 23 20
0.365BS-0.025PSN-0.61PT MPB 414 700 2818 1.5 13.0 0.18 326 24 23
0.355BS-0.025PSN-0.62PT MPB 418 825 3358 1.5 10.1 0.18 323 24 24
0.345BS-0.025PSN-0.63PT Tetragonal 423 1124 4151 0.9 16.2 0.21 368 28 25
0.335BS-0.025PSN-0.64PT Tetragonal 428 1060 3362 1.8 15.2 0.19 325 31 21
0.37BS-0.05PSN-0.58PT MPB 386 751 3684 1.7 4.6 0.13 244 22 29
0.36BS-0.05PSN-0.59PT MPB 388 747 3769 2.1 4.8 0.15 276 22 30
0.35BS-0.05PSN-0.60PT MPB 395 850 3824 1.8 9.0 0.16 299 23 31
0.34BS-0.05PSN-0.61PT MPB 401 1207 5372 2.2 9.0 0.17 303 24 28
0.33BS-0.05PSN-0.62PT Tetragonal 415 957 3519 1.8 7.2 0.17 303 26 19
0.32BS-0.05PSN-0.63PT Tetragonal 419 926 2910 0.6 10.6 0.19 344 30 29

Fig. 6. Polarization-electric field hysteresis loops of the representative (1-x-y) Fig. 7. Unipolar strain of the representative (1-x-y)BS-yPSN-xPT ceramics
BS-yPSN-xPT ceramics (sintered at 1150 ◦ C) measured at 10 Hz at room (sintered at 1150 ◦ C) measured at 10 Hz at room temperature.
temperature.
insights into the polar domain and polarization switching, which
remnant polarization Pr reaches its maximum in the MPB region, while contribute to a profound understanding of the electromechanical
the coercive field EC gradually increases with higher PT contents. The behaviour in ferroelectric materials [48–50]. In Fig. 8, the topography,
unipolar strain and the large-signal piezoelectric coefficient d33* display and the out-of-plane PFM amplitude and phase images of the repre­
a similar compositional dependence, as detailed in Fig. 7 and Table 1. In sentative (0.975-x)BS-0.025PSN-xPT ceramics in an area of 15✕15 μm2
the composition of y = 0.025 and x = 0.63, an enhanced strain of 0.21 % are illustrated.
and a d33* of 368 pm/V are attained, nearly twice as large as those The examined samples exhibit intricate domain configurations, with
observed in BS-PT within the MPB region. Consequently, the introduc­ domain sizes ranging from ~50 nm to ~5 μm. A noticeable feature is the
tion of PSN effectively enhances the ferroelectric and piezoelectric comparatively diminutive size of polar domains, a tendency that di­
properties of BS-PT. minishes with the escalation of PT content within the MPB region and
The large Pr, enhanced strain and high piezoelectric coefficient subsequently intensifies in the tetragonal region. The composition of x
observed in the samples with MPB compositions are attributed to the = 0.63, which displays the highest strain and piezoelectric coefficient,
coexistence of tetragonal and monoclinic phases. This coexistence re­ exhibits the smallest domain and a high density of domain wall. The
sults in a synergistic effect of equivalent, multiple polarization states (6 small domains display greater flexibility in response to the applied
equivalent states for tetragonal symmetry and 24 equivalent states for electric field due to their small domain wall energy [51–55].
monoclinic symmetry), facilitating polarization rotation [43–47]. Under the application of electric fields, nano-to micrometers-sized
domains began to coalesce, forming larger domains that eventually
switched to the “down” polarization state when subjected to 40 V, as
3.5. Local polar domain structure illustrated in the central 5✕15 μm2 area in the PFM phase images.
Reorientation of these domains to the “up” polarization state was
Piezoresponse force microscopy offers valuable micro-to nanoscale

6
Z. Liu et al. Ceramics International xxx (xxxx) xxx

Fig. 8. Topography, polar domain structure and its evolution under the electric fields determined by piezoresponse force microscopy in the (0.975-x)BS-0.025PSN-
xPT ceramics (sintered at 1150 ◦ C): (a) x = 0.61, (b) x = 0.62, (c) x = 0.63, and (d) x = 0.64. The red boxes in the phase images were poled under a voltage sequence
of 20 V, 40 V, -20 V, and -40 V, respectively, for each composition. (For interpretation of the references to colour in this figure legend, the reader is referred to the
Web version of this article.)

observed upon application of a negative voltage, further confirming the associated with the MPB feature, which accommodates the coexisting
ferroelectric nature of the ceramics. Therefore, the small domain sizes phases and multiple domain states that facilitate polarization rotations
and their active response contribute significantly to the notable piezo­ and the presence of nano-to micrometer-sized domains with high flexi­
electric properties observed in x = 0.63. bility. These improved properties suggest that the BS-PSN-PT ceramics
It was observed that the downward switching of the domains cannot hold a promising potential for applications in high-temperature elec­
be complete under -40 V for x = 0.62. This behaviour indicates the ex­ tromechanical devices.
istence of an internal bias field that appears to align in an upward di­
rection within the sample. The electric field favors the complete Declaration of competing interest
switching of the ‘‘up” domains, but prevents the completion of the
reverse process [56]. The domains switching would likely be complete if We declare that we have no competing financial interests or personal
poled at a higher voltage for a longer time. relationships with other people or organizations that can influence our
work in this article.
4. Conclusions
Acknowledgements
A strategy is proposed that involves the introduction of the Pb(Sc1/
2Nb1/2)O3 end member into the bismuth-based perovskite solid solution The authors would like to thank Dr. Zuo-Guang Ye from Simon Fraser
of BiScO3-PbTiO3 and the preparation of ceramics with large grain size University for helpful discussion. This work was supported by the Na­
(ranging 20 μm ~ 45 μm). This strategy aims to significantly enhance tional Natural Science Foundation of China (Grant No. 52372124) and
the strain level, to reduce the dielectric loss, and to simultaneously the Shaanxi Province Postdoctoral Science Foundation. We acknowl­
maintain a high Curie temperature of BS-PT. A ternary solid solution edge the Instrument Analysis Center of Xi’an Jiaotong University for
system of (1-x-y)BiScO3-yPb(Sc1/2Nb1/2)O3-xPbTiO3 with the composi­ assistance with SEM work.
tions around the morphotropic phase boundary, was designed and pre­
pared in the form of ceramics using the solid-state reaction method and References
sintering process. The impact of PSN modification on the crystal struc­
ture, microstructure, dielectric, ferroelectric, and piezoelectric proper­ [1] S. Trolier-McKinstry, S. Zhang, A.J. Bell, X. Tan, High-performance piezoelectric
ties of BS-PT was systematically investigated. The introduction of PSN crystals, ceramics, and films, Annu. Rev. Mater. Res. 48 (1) (2018) 191–217.
[2] S. Zhang, F. Yu, Piezoelectric materials for high temperature sensors, J. Am. Ceram.
was found to greatly enhance the strain level and piezoelectric proper­ Soc. 94 (10) (2011) 3153–3170.
ties and to slightly reduce the Curie temperature. In the composition [3] X.N. Jiang, K. Kim, S.J. Zhang, J. Johnson, G. Salazar, High-temperature
0.32BS-0.05PSN-0.63PT, a relatively large strain of 0.21 %, a high large- piezoelectric sensing, Sensors 14 (1) (2014) 144–169.
[4] Q. Guo, X. Meng, F. Li, F. Xia, P. Wang, X. Gao, J. Wu, H. Sun, H. Hao, H. Liu,
signal piezoelectric coefficient (d33* = 368 pm/V), a low dielectric loss S. Zhang, Temperature-insensitive PMN-PZ-PT ferroelectric ceramics for actuator
(tanδ = 1.7 % at room temperature, measured at 10 kHz), and a high applications, Acta Mater. 211 (2021) 116871.
Curie temperature (TC ~ 423 ◦ C) were achieved, confirming the effec­ [5] M. Fang, S. Rajput, Z. Dai, Y. Ji, Y. Hao, X. Ren, Understanding the mechanism of
thermal-stable high-performance piezoelectricity, Acta Mater. 169 (2019)
tiveness of the proposed strategy. These enhanced properties are
155–161.

7
Z. Liu et al. Ceramics International xxx (xxxx) xxx

[6] J. Li, W. Qu, J. Daniels, H. Wu, L. Liu, J. Wu, M. Wang, S. Checchia, S. Yang, H. Lei, [32] K. Datta, A. Richter, M. Gobbels, R.B. Neder, B. Mihailova, Mesoscopic-scale
R. Lv, Y. Zhang, D. Wang, X. Li, X. Ding, J. Sun, Z. Xu, Y. Chang, S. Zhang, F. Li, structure and dynamics near the morphotropic phase boundary of (1-x)PbTiO3-
Lead zirconate titanate ceramics with aligned crystallite grains, Science 380 (6640) xBiScO3, Phys. Rev. B 92 (2) (2015) 7.
(2023) 87–93. [33] L. Kong, G. Liu, S. Zhang, W. Yang, Origin of the enhanced piezoelectric thermal
[7] B. Jaffe, C. William, H. Jaffe, Piezoelectric Ceramics, Academic Press, London, stability in BiScO3-PbTiO3 single crystals, Appl. Phys. Lett. 106 (23) (2015)
1971. 232901.
[8] L. Chen, H. Liu, H. Qi, J. Chen, High-electromechanical performance for high- [34] J.A. Lima, W. Paraguassu, P.T.C. Freire, A.G. Souza Filho, C.W.A. Paschoal, J.
power piezoelectric applications: fundamental, progress, and perspective, Prog. M. Filho, A.L. Zanin, M.H. Lente, D. Garcia, J.A. Eiras, Lattice dynamics and low-
Mater. Sci. 127 (2022) 100944. temperature Raman spectroscopy studies of PMN–PT relaxors, J. Raman Spectrosc.
[9] X. Xie, Z. Zhou, B. Gao, Z. Zhou, R. Liang, X. Dong, Ion-pair engineering-induced 40 (9) (2009) 1144–1149.
high piezoelectricity in Bi4Ti3O12-based high-temperature piezoceramics, ACS [35] A. Sehirlioglu, A. Sayir, F. Dynys, Microstructure-property relationships in liquid
Appl. Mater. Interfaces 14 (12) (2022) 14321–14330. phase-sintered high-temperature bismuth scandium oxide-lead titanate
[10] C. Long, B. Wang, W. Ren, K. Zheng, H. Fan, D. Wang, L. Liu, Significantly piezoceramics, J. Am. Ceram. Soc. 91 (9) (2008) 2910–2916.
enhanced electrical properties in CaBi2Nb2O9-based high-temperature [36] M.-S. Lee, J.-W. Park, Y.H. Jeong, Grain growth anomaly and piezoelectric
piezoelectric ceramics, Appl. Phys. Lett. 117 (3) (2020) 032902. properties of liquid phase sintered high TC 0.36BiScO3-0.64PbTiO3 ceramics with
[11] H. Chen, J. Xi, Z. Tan, F. Wang, X. Li, N. Chen, H. Li, Q. Chen, J. Xing, J. Zhu, sillenite Bi12PbO19, Ceram. Int. 47 (24) (2021) 34405–34413.
Decoding intrinsic and extrinsic contributions for high piezoelectricity of CBT- [37] A.A. Bokov, Z.-G. Ye, Recent progress in relaxor ferroelectrics with perovskite
based piezoelectric ceramics, J. Mater. Chem. C 11 (35) (2023) 12048–12056. structure, J. Mater. Sci. 41 (1) (2006) 31–52.
[12] Z. Liu, H. Wu, Y. Yuan, H. Wan, Z. Luo, P. Gao, J. Zhuang, J. Zhang, N. Zhang, J. Li, [38] A.A. Bokov, Z.-G. Ye, Dielectric relaxation in relaxor ferroelectrics, J. Adv. Dielectr.
Y. Zhan, W. Ren, Z.-G. Ye, Recent progress in bismuth-based high Curie 2 (2) (2012) 1241010.
temperature piezo-/ferroelectric perovskites for electromechanical transduction [39] F. Li, S. Zhang, D. Damjanovic, L.-Q. Chen, T.R. Shrout, Local structural
applications, Curr. Opin. Solid State Mater. Sci. 26 (5) (2022) 101016. heterogeneity and electromechanical responses of ferroelectrics: learning from
[13] Z. Liu, H. Wu, J. Zhuang, G. Niu, N. Zhang, W. Ren, Z.-G. Ye, High Curie relaxor ferroelectrics, Adv. Funct. Mater. 28 (37) (2018) 1801504.
temperature bismuth-based piezo-/ferroelectric single crystals of complex [40] G. Zhang, D. Brannum, D. Dong, L. Tang, E. Allahyarov, S. Tang, K. Kodweis, J.-
perovskite structure: recent progress and perspectives, CrystEngComm 24 (2) K. Lee, L. Zhu, Interfacial polarization-induced loss mechanisms in polypropylene/
(2022) 220–230. BaTiO3 nanocomposite dielectrics, Chem. Mater. 28 (13) (2016) 4646–4660.
[14] Y. Dong, K. Zou, R. Liang, Z. Zhou, Review of BiScO3-PbTiO3 piezoelectric [41] A.K. Jonscher, Dielectric relaxation in solids, J. Phys. D Appl. Phys. 32 (14) (1999)
materials for high temperature applications: fundamental, progress, and R57.
perspective, Prog. Mater. Sci. 132 (2023) 101026. [42] Z.-G. Ye, Handbook of Advanced Dielectric, Piezoelectric and Ferroelectric
[15] H. Zhao, Y. Hou, X. Yu, M. Zheng, M. Zhu, Giant high-temperature piezoelectricity Materials: Synthesis, Properties and Applications, Woodhead Publishing,
in perovskite oxides for vibration energy harvesting, J. Mater. Chem. A 9 (4) Cambridge, U.K., 2008.
(2021) 2284–2291. [43] H. Fu, R.E. Cohen, Polarization rotation mechanism for ultrahigh
[16] E.E. Richard, A.R. Clive, R.S. Thomas, W.R. Paul, H. Wes, P. Seung-Eek, New high electromechanical response in single-crystal piezoelectrics, Nature 403 (6767)
temperature morphotropic phase boundary piezoelectrics based on Bi(Me)O3- (2000) 281–283.
PbTiO3 ceramics, Jpn. J. Appl. Phys. 40 (10R) (2001) 5999. [44] R. Guo, L. Cross, S. Park, B. Noheda, D. Cox, G. Shirane, Origin of the high
[17] E.E. Richard, A.R. Clive, R.S. Thomas, P. Seung-Eek, Preparation and piezoelectric response in PbZr1-xTixO3, Phys. Rev. Lett. 84 (23) (2000) 5423.
characterization of high temperature perovskite ferroelectrics in the solid-solution [45] D. Damjanovic, A morphotropic phase boundary system based on polarization
(1-x)BiScO3-xPbTiO3, Jpn. J. Appl. Phys. 41 (4R) (2002) 2099. rotation and polarization extension, Appl. Phys. Lett. 97 (6) (2010) 062906.
[18] X. Ren, Y. Wang, M. Tang, X. Liu, Z. Xu, Y. Yan, Achieving large piezoelectric [46] Z. Liu, A.R. Paterson, H. Wu, P. Gao, W. Ren, Z.-G. Ye, Synthesis, structure and
response and ultrahigh electrostriction in [001] textured BiScO3-PbTiO3 high- piezo-/ferroelectric properties of a novel bismuth-containing ternary complex
temperature piezoelectric ceramics, J. Am. Ceram. Soc. 106 (9) (2023) 5331–5340. perovskite solid solution, J. Mater. Chem. C 5 (16) (2017) 3916–3923.
[19] Z. Yao, H. Liu, Y. Liu, Z. Li, X. Cheng, M. Cao, H. Hao, Morphotropic phase [47] Z. Liu, H. Wu, W. Ren, Z.-G. Ye, Complex morphotropic phase transformations and
boundary in Pb(Sc1/2Nb1/2)O3-BiScO3-PbTiO3 high temperature piezoelectrics, high piezoelectric properties in new ternary perovskite single crystals, Acta Mater.
Mater. Lett. 62 (29) (2008) 4449–4451. 149 (2018) 132–141.
[20] S. Zhang, R.E. Eitel, C.A. Randall, T.R. Shrout, E.F. Alberta, Manganese-modified [48] N. Balke, I. Bdikin, S.V. Kalinin, A.L. Kholkin, Electromechanical imaging and
BiScO3-PbTiO3 piezoelectric ceramic for high-temperature shear mode sensor, spectroscopy of ferroelectric and piezoelectric materials: state of the art and
Appl. Phys. Lett. 86 (26) (2005) 262904, 262904. prospects for the future, J. Am. Ceram. Soc. 92 (8) (2009) 1629–1647.
[21] J. Chen, Y. Dong, J. Cheng, Reduced dielectric loss and strain hysteresis in (0.97-x) [49] J. Lin, G. Ge, J. Li, J. Qian, K. Zhu, Y. Wei, C. Shi, G. Li, F. Yan, W. Li, J. Zhang,
BiScO3-xPbTiO3-0.03Pb(Mn1/3Nb2/3)O3 piezoelectric ceramics, Ceram. Int. 41 (8) J. Zhai, H. Wu, Field-induced multiscale polarization configuration transitions of
(2015) 9828–9833. mesentropic lead-free piezoceramics achieving giant energy harvesting
[22] Y. Yu, J. Yang, J. Wu, X. Gao, L. Bian, X. Li, X. Xin, Z. Yu, W. Chen, S. Dong, performance, Adv. Funct. Mater. 33 (42) (2023) 2303965.
Ultralow dielectric loss of BiScO3-PbTiO3 ceramics by Bi(Mn1/2Zr1/2)O3 [50] Z. Luo, Z. Liu, D. Walker, S. Huband, P.A. Thomas, N. Zhang, W. Ren, Z.-G. Ye,
modification, J. Eur. Ceram. Soc. 40 (8) (2020) 3003–3010. Meso- to nano-scopic domain structures in high Curie-temperature piezoelectric
[23] X. Ren, M. Tang, X. Liu, Y. Wang, Z. Xu, Y. Yan, Simultaneously achieving large BiScO3–PbTiO3 single crystals of complex perovskite structure, J. Mater. Chem. C 8
piezoelectricity and low dielectric loss in high-temperature BiScO3–PbTiO3-based (21) (2020) 7234–7243.
ceramics, ACS Appl. Mater. Interfaces 15 (35) (2023) 41614–41623. [51] X. Lv, X.-x. Zhang, J. Wu, Nano-domains in lead-free piezoceramics: a review,
[24] Y. Feng, C. Yang, X. Guo, W. Sun, W. Wang, X. Lin, S. Huang, Achieving both large J. Mater. Chem. A 8 (20) (2020) 10026–10073.
piezoelectric constant and low dielectric loss in BiScO3–PbTiO3–Bi(Mn2/3Sb1/3)O3 [52] Z. Liu, C. Zhao, J.-F. Li, K. Wang, J. Wu, Large strain and temperature-insensitive
high-temperature piezoelectric ceramics, J. Adv. Dielectr. 12 (6) (2022) 2250017. piezoelectric effect in high-temperature piezoelectric ceramics, J. Mater. Chem. C 6
[25] F. Chu, I. Reaney, N. Setter, Investigation of relaxors that transform spontaneously (3) (2018) 456–463.
into ferroelectrics, Ferroelectrics 151 (1) (1994) 343–348. [53] R. Wang, H. Xu, B. Yang, Z. Luo, E. Sun, J. Zhao, L. Zheng, Y. Dong, H. Zhou,
[26] F. Chu, I. Reaney, N. Setter, Spontaneous (zero-field) relaxor-to-ferroelectric-phase Y. Ren, C. Gao, W. Cao, Phase coexistence and domain configuration in Pb(Mg1/
transition in disordered Pb(Sc1/2Nb1/2)O3, J. Appl. Phys. 77 (4) (1995) 1671–1676. 3Nb2/3)O3-0.34PbTiO3 single crystal revealed by synchrotron-based X-ray
[27] E.F. Alberta, A.S. Bhalla, High strain and low mechanical quality factor diffractive three-dimensional reciprocal space mapping and piezoresponse force
piezoelectric Pb[(Sc1/2Nb1/2)0.575Ti0.425]O3 ceramics, Mater. Lett. 35 (3) (1998) microscopy, Appl. Phys. Lett. 108 (15) (2016).
199–201. [54] Z. Liu, Z. Luo, C. Wang, H. Wu, N. Zhang, H. Liu, W. Ren, Z.-G. Ye, Micro-/
[28] Y.H. Bing, Z.-G. Ye, Synthesis and characterizations of the (1− x)Pb(Sc1/2Nb1/2) nanodomains and their switching in a high Curie-temperature ferroelectric single
O3–xPbTiO3 solid solution ceramics, J. Electroceram. 21 (2007) 761–764. crystal of Bi(Zn2/3Nb1/3)O3-PbTiO3, Ceram. Int. 44 (2018) S189–S194.
[29] S.L. Swartz, T.R. Shrout, Fabrication of perovskite lead magnesium niobate, Mater. [55] J. Fu, R. Zuo, Z. Xu, High piezoelectric activity in (Na,K)NbO3 based lead-free
Res. Bull. 17 (10) (1982) 1245–1250. piezoelectric ceramics: contribution of nanodomains, Appl. Phys. Lett. 99 (6)
[30] Z. Luo, N. Zhang, Z. Liu, J. Zhuang, J. Zhao, W. Ren, Z.-G. Ye, Complex (2011) 062901.
morphotropic domain structure and ferroelectric properties in high-TC single [56] A.R. Paterson, J. Zhao, Z. Liu, X. Wu, W. Ren, Z.-G. Ye, High-temperature solution
crystals of a ternary perovskite solid solution, J. Mater. Chem. C 6 (34) (2018) growth and characterization of (1-x)PbTiO3-xBi(Zn2/3Nb1/3)O3 piezo-/ferroelectric
9216–9223. single crystals, J. Cryst. Growth 486 (2018) 38–44.
[31] J. Chaigneau, J. Kiat, C. Malibert, C. Bogicevic, Morphotropic phase boundaries in
(BiScO3)1-x(PbTiO3)x (0.60<x<0.75) and their relation to chemical composition
and polar order, Phys. Rev. B 76 (9) (2007) 094111.

You might also like