Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Author’s Accepted Manuscript

Enhanced real-time high temperature piezoelectric


responses and ferroelectric scaling behaviors of
MgO-doped 0.7BiFeO3−0.3BaTiO3 ceramics

Zijing Li, Wen Peng, Changrong Zhou, Qingning


Li, Ling Yang, Jiwen Xu, Guohua Chen, Changlai
Yuan, Guanghui Rao
www.elsevier.com/locate/ceri

PII: S0272-8842(18)31188-X
DOI: https://doi.org/10.1016/j.ceramint.2018.05.057
Reference: CERI18232
To appear in: Ceramics International
Received date: 19 April 2018
Revised date: 7 May 2018
Accepted date: 8 May 2018
Cite this article as: Zijing Li, Wen Peng, Changrong Zhou, Qingning Li, Ling
Yang, Jiwen Xu, Guohua Chen, Changlai Yuan and Guanghui Rao, Enhanced
real-time high temperature piezoelectric responses and ferroelectric scaling
behaviors of MgO-doped 0.7BiFeO3−0.3BaTiO3 ceramics, Ceramics
International, https://doi.org/10.1016/j.ceramint.2018.05.057
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Enhanced real-time high temperature piezoelectric responses and

ferroelectric scaling behaviors of MgO-doped 0.7BiFeO3-0.3BaTiO3

ceramics

Zijing Lib, Wen Pengb, Changrong Zhoua,b*, Qingning Lia, Ling Yanga, Jiwen Xua,b,

Guohua Chena,b, Changlai Yuana,b, Guanghui Raoa,b*

a
Guangxi Key Laboratory of Information Materials, Guilin University of Electronic

Technology, Guilin, Guangxi, China, 541004, PR China.

b
School of Material Science and Engineering, Guilin University of Electronic

Technology, Guilin, Guangxi, 541004, PR China.

zcr750320@guet.edu.cn

rgh@guet.edu.cn

*
Corresponding authors. Tel.: +86 773 2291434; fax: +86 773 2290129.

Abstract

The thermal stability of the piezoelectric properties in BiFeO3 based high-temperature

piezoceramics is traditionally investigated through room temperature characterization

via thermal annealing measurements, which cannot reflect the actual performance at

elevated temperatures. In this work, the microstructures, electrical properties,

real-time high-temperature piezoelectric responses and ferroelectric scaling behaviors


1
of 0.7Bi1.05(Fe1-xMgx)O3-0.3BaTiO3 (BFMx-BT) ceramics were studied systematically.

Doping with 0.6% Mg resulted in enhanced piezoelectric properties with d33=170

pC/N and kp=0.29 along with a high-temperature stability (Td=440 oC). With a

significant jump near the Td, in situ high temperature piezoelectric responses

displayed a distinct behavior different from the ex situ high temperature

piezoelectricity. Scaling behaviors were presented by studying the field amplitude and

frequency dependence of P-E hysteresis loops areas. The behavior of the real-time

high temperature piezoelectricity may provide a new insight into the development of

new high-temperature piezoceramics.

Keywords: Lead-free piezoceramics; High-temperature piezoceramics; Depolarization;

Scaling behaviors

1. Introduction

The rapid development of piezoelectric applications in harsh environment

ranging from turbine engines to space exploration has put forward an urgent demand

for materials with excellent piezoelectric properties and high temperature stability [1,

2]. BiFeO3 (BFO) and its derivatives exhibiting multi-coupling between strain,

2
polarization and magnetization, have attracted considerable attention as a promising

candidate for high-temperature lead-free piezoelectric materials due to the high Curie

temperature and large piezoelectric response [2-5]. A number of compositional

modification reports have been carried out in recent years, and partial rare earth (RE)

ions substituting BFO on A-site displayed enhanced piezoelectric activity and high

depoling temperatures [6-9]. However, modified BFO ceramics manifest inferior

electrical properties (piezoelectric constant d33 <55pC/N) regardless of their ultra-high

Curie temperature.

It was reported that an addition of ABO3-type perovskites (e.g. BaTiO3, Ba(Zr,

Ti)O3 and Bi0.5K0.5TiO3) can greatly enhance the electrical properties of BFO

ceramics by forming solid solutions with rhombohedral–cubic (R–C) morphotropic

phase boundary(MPB)[10-16]. Moreover, appreciable improvement in properties has

been observed by further compositional modification and different preparation

methods [15-24]. Leontsev and Eitel reported that Mn-modified BF-BT ceramics

exhibited relatively large d33 (~116pC/N), enhanced DC resistivity (~7.6×1012·cm),

high TC (~619C), and high Td (~469C) [25]. By adding the super-tetragonal

compounds of (BiGaO3, BG) along with quenching method, Lee et al. achieved a

large piezoelectric charge constants of d33=420 pC/N with the Curie temperature of

454C in 0.67Bi1.05(Fe0.97Ga0.03)O3-0.33BaTiO3 ceramics[21]. Nevertheless, the origin

of the dramatically improved piezoelectricity remains unclear up to today.

It is well known that piezoelectric response of ferroelectric ceramics is largely

affected by extrinsic effects associated with contributions from the dynamics of

3
domain via domain wall motion and switching under an applied electric field [26-29].

Dynamic scaling analysis on ferroelectric hysteresis has been widely used to describe

the dynamics behaviors of domains in bulk ceramics [30-33]. The electric polarization

(P)-electric field (E) hysteresis loops, can provide characteristic parameters of the

dynamic behavior including hysteresis area <A>, remnant polarization Pr and coercive

field Ec [31-35]. Therefore, many studies on scaling law between <A> and E0, T0,

frequency (f0), were performed via P-E hysteresis measurements. Many efforts are

dedicated to rationalize the dynamic polarization response by investigating the scaling

behaviors of ferroelectrics such as Pb-based ceramics and lead-free ceramics based on

BaTiO3, (Na0.5K0.5)NbO3 and Bi0.5Na0.5TiO3[30-35]. However, the global polarization

reversal process and the mechanisms are still not fully clarified. Particularly, to our

knowledge, there are few reports on the dynamic polarization response and scaling

behaviors of BF-BT systems.

Generally, excellent high-temperature piezoelectric responses are important

figure of merits for applications in high-temperature piezoelectric sensor. However,

most of the thermal stability and depolarization behaviors for high-temperature

piezoceramics are typically given by characterizing room temperature piezoelectric

coefficient after thermal annealing [36-38]. The ex situ measurement may not provide

information about real performance at the operation temperatures [24, 38]. Thus, it is

necessary to investigate the real-time high-temperature piezoelectric properties which

are critical and highly desirable for monitoring the practical material performance and

for the design of relevant piezoelectric devices.

4
It is well known that MgO doping in ferroelectrics is an effective method to tune

the electrical properties [39]. In this work, the microstructure, ferroelectric and

piezoelectric properties of MgO-doped BF-BT ceramics were reported. Moreover, the

real high temperature piezoelectric response and depolarization characteristics were

analyzed by in situ measuring real-time piezoelectric performance. Compared to

conventional thermal depolarization by ex situ measuring method, different evolution

behaviors of real-time piezoelectric properties would provide a new insight into

developing new high temperature piezoceramics.

2. Experiment

0.7Bi1.05(Fe1-xMgx)O3-0.3BaTiO3+0.3wt%MnO2 (BFMx-BT, x=0-0.9%) ceramics

were prepared by the conventional ceramic processing techniques, where excess Bi

and MnO2 were added to compensate for the volatility of bismuth and reduce oxygen

vacancies formed during calcination and sintering, respectively. The stoichiometric

amounts of Bi2O3 (99.0%), Fe2O3 (99.0%), BaCO3 (99.9%), TiO2 (99.9%), MnO2

(99.0%), and MgO (98.0%) were weighed and zirconia ball mixed for 24h with

ethanol. Subsequently, the mixed powders were dried and calcined in alumina

crucible at 800 °C for 3h and pressed into pellets with 13 mm in diameter and 0.7 mm

in thickness at 90MPa pressure, followed by sintering at 980 °C for 3h in air. Silver

electrodes were pasted on both surfaces of the as-sintered samples and fired at 600 °C

for 20 min. Finally, the as-sintered samples were poled in a silicone oil bath under a

dc field of 5 kV/mm at 120 °C for 15min.

5
The crystalline phases of the as-sintered samples were identified by X-ray

diffraction (XRD) with Cu Kα radiation (D8-2-Advance, Bruker Inc., Karlsruhe,

Germany) at room temperature. Surface morphology and microstructure of the

as-sintered samples were observed by scanning electron microscopy (SEM,

JSM-5610LV, JEOL, Tokyo, Japan). The grain size distribution and average grain size

were analyzed and calculated by software of Nano measure. The piezoelectric

constant d33 of the poled samples was measured using a Berlincourt piezo-d33 meter

(ZJ-3A, CAS, Shanghai, China). At the same time, the planar coupling factor (kp ) and

the maximum phase angle (θmax) of the poled samples were determined by

piezoelectric resonance method with an impedance analyzer (Agilent 4294A, Agilent

Inc., Bayan, Malaysia). Relative dielectric constant (εr), dielectric loss (tanδ), and kp as

a function of temperature were measured in the range of 25~550 °C by an impedance

analyzer (Agilent 4294A, Agilent Inc., Bayan, Malaysia). Here, the temperature

dependence of kp (in situ measurements) of the samples was heated at a setting

temperature for 20 min in a muffle furnace. The room temperature strain-electric field

(S-E) and current-electric field (I-E) hysteresis loops were characterized using a

ferroelectric test system (TF Analyzer 2000HS, aixACCT, Aachen, Germany) at a

frequency of 1 Hz. Polarization-electric field (P-E) hysteresis loops were determined

at different frequencies (0.01~10 Hz) and electric fields (10~65 kV/cm) using the

ferroelectric test system.

3. Results and discussion

Figure 1(a) shows the XRD patterns of BFMx-BT ceramics in the 2θ range of

6
20-80° measured at room temperature. All the BFMx-BT samples crystallize in a

distorted rhombohedral structure. Apparently, the XRD diffraction peaks match well

with the diffraction data (PDFs of no. 86-1518 and no. 75-2121), except a minor low

intensity impurity peaks around 2θ=28° (marked by rhombi) associated with the

Bi2Fe4O9 (PDF of no. 72-1832). The enlarged views of XRD patterns in the range of

38-46° are shown in Fig. 1(b). The relative intensity of the (003) diffraction peaks

increases with x (0~0.6%), reflecting that the rhombohedral distortion becomes more

evident with increasing doping amount of MgO.

Surface SEM morphologies and grain size distribution of the BFMx-BT ceramics

sintered at 980 °C for 3h are displayed in Fig. 2. Polygonal grains with very few gas

pores can be observed in the sintered ceramics. The average grain size of BFMx-BT

ceramics increases from 2.03 μm to 4.68 μm as the MgO content increases from 0 to

0.6%, indicating that the moderate doping of MgO promotes the grain growth of the

ceramics.

Figure 3(a)-(f) show the weak-signal piezoelectric properties (d33 and kp),

maximum phase angle (θmax), the current-electrical field (I-E) loops, bipolar strain

curves, the polarization-electrical field (P-E) hysteresis loops, remnant polarization

(Pr), and coercive field (EC) of BFMx-BT ceramics. The piezoelectric coefficient d33

and the planar electromechanical coupling factor kp increase first, reach the maximum

values of 170 pC/N and 0.29 at x=0.60%, and then decrease sharply as the MgO

content further increases. The d33 and kp, which are correlated with the reorientation

degree of domains during the poling process, vary in a similar way to the variation of

7
θmax as shown in Fig. 3(b). The θmax, derived from the frequency dependence of the

impedance |Z| between the resonance frequency (fr) and the anti-resonance frequency

(fa), is an indication of poling induced texture degree of domains [24, 40, 41]. The

maximum θmax at x=0.6%, shown in Fig. 3(b), imply a large texture degree of domains.

Meanwhile, the remnant polarization (Pr, indicating degree of ferroelectric domains

texture) and grain size presented analogous variation trends with the weak-signal

piezoelectric properties. Therefore, the large small-signal piezoelectric responses for

x=0.60% sample are associated with the combination of electric field induced high

degree of polarization reorientation and grain size effect.

The electric-field induced strain hysteresis loops (S-E) of the BFMx-BT ceramics

exhibit characteristic shape of butterfly loop as displayed in Fig. 3(c). The

corresponding large signal strains Smax (peak-to-peak strain) are displayed in the inset.

As x increases, the variation of strain level follows the similar tendency with those of

small signal piezoelectric properties, reaches the maximum of 0.117% at x=0.60%.

The electric field induced strains are closely related to the domains switching by

applying an electric field [17, 24]. The bipolar strain with a maximum value at x=0.60%

results mainly from extrinsic contribution of domains switching due to the highest

peak current at x=0.60% as shown in Fig. 3(d).

Temperature dependence of relative permittivity (εr) and loss (tanδ) measured at

different frequencies are shown in Fig. 4(a-e). All samples exhibit a distinctive

dielectric anomaly at Tm due to the ferroelectric to paraelectric phase transition. The

loss curves of BFMx-BT ceramics also show a peak or hump below Tm. Note that the

8
dielectric loss anomaly for the samples with x=0.90% is very weak. For all BFMx-BT

ceramics, the rapidly increase in the loss above Tm should be due to the increased

conductivity. The dielectric anomalies exhibit classic ferroelectric relaxor behaviors: a

broad diffuse dielectric peak with shifting of the permittivity peak to a higher

temperature with increasing frequency. The local maxima εm and Tm as a function of x

are shown in Fig. 4(f). Both εm and Tm exhibit maximum values around x=0.60%.

Figure 5(a) displays the real-time change of planar electromechanical coupling

factor (kp) with increasing temperature by in situ measurement technique. The room

temperature ex situ kp for the annealed BFMx-BT ceramics is shown in Fig. 5(b). It

is obvious that the real-time high temperature responses manifest a rather different

temperature dependence compared with the ex situ kp. The ex situ kp is unstable and

shows a clear decreasing trend with the increase of the annealing temperature.

Correspondingly, a much faster reduction in kp occurs after annealing at a certain

temperature, which is often named as depolarization temperature Td. In contrast,

real-time high temperature kp seems to exhibit a three-stage evolution with the

increasing temperature, i.e. low, medium and high temperature stages (TL, TM and TH).

The in situ kp increases in the first stage (TL), then remains approximately constant in

the second stage (TM), and finally shows a significant jump with a prominent peak at

Td. Herein, the compositions of x=0.60% and 0.75% exhibit a striking increment near

Td. Furthermore, the Td as a function of x is shown in Fig. 5(c), which displays a

maximum of 440 °C for ceramic with x=0.6%.

Although the in situ Td is almost same as that of ex situ due to the poled sample

9
restores the nearly random domain configuration after heating to a sufficiently high

temperature, their kp values depict a different variation trend. For the ex situ kp, the

poling induced reorientation domain would recover partly to virgin state during

annealing process, thus giving rise to the decreased kp. The sharp decrease of the

real-time kp value at Td may be explained by the temperature-driven flattening of the

Gibbs free-energy profile and the polarization extension mechanism [41-47]. Td is

very close to the ferroelectric to paraelectric transition temperature Tm for the samples

with x=0.6%. Within the framework of the Landau-Ginzburg-Devonshire (LCD)

theory, a highly flattening free-energy profile is introduced when the ceramics are

heated to Tm, which will provide the conditions for minimal anisotropy energy

free-energy. With the flattened energy potential at Tm, the phase transition from a

polar phase to a non-polar state can have large response to very small external

excitation. Hence, the maximum real-time piezoelectric responses near the

depolarization temperature can be reconciled by low anisotropy energy accompanying

by polarization extension.

Typical P-E curves for the representative compositions of x=0.6% are illustrated

in Fig. 6(a). The electric filed dependence of measured remnant polarizations 2Pr,

maximum polarizations 2Pmax and coercive fields 2Ec are plotted in Fig. 6(b).

Relatively small increase in 2Pr, 2Pmax and 2Ec values under low electric fields (E < 2

kV/mm) were observed. As the electric field was increased to above 2 kV/mm, these

parameters show obvious increase with increasing electric field. Additionally, the

calculated hysteresis area <A> as a function of E0 is displayed in Fig. 6(c). The

10
evolution of the <A>–E can be divided into two stages with the increase of E0, as

indicated by the two perfectly linear relations of ln <A>=βlnE0+m shown in Fig. 6(d).

In the first stage, the fitting parameters are β=4.842 and m=-10.634 (R2=0.9994),

while in the second stage, β=1.365, m=2.055(R2=0.9966). The results indicate that the

hysteresis area <A> obeys different scaling law of <A>∝E0β in different electric field

range.

Figure 7(a) shows the frequency dependent polarization hysteresis loops. The

loops are unsaturated and show clearly f-dependent. The scaling relationships between

ln<A> and lnf0 can be described by the equation ln<A> = αlnf0+b0 as depicted in Fig.

7(b). It is obvious that the frequency scaling relationships exhibit two-stage dynamic

behaviors. The fitting coefficients are α=-0.030,n=9.147(R2=0.9994) and α=-0.078,

n=7.478(R2=0.9980), respectively. The two-stage dynamic behaviors may correspond

to motions of non-1800 and 1800 domain wall (DW), respectively [30-35]. The

non-1800 DW motion needs a lower frequency E0. However, when the applied electric

field is higher than coercive field, the polarization of both non-1800 and 1800 DW

motion can be completely switched under the high electric field, which results in a

linear frequency scaling relationship.

Conclusions

MgO doping promotes the grain growth while influences little the phase

structure transition in BFMx-BT ceramics. The maximum piezoelectric properties of

d33=170 pC/N and kp=0.29 were obtained at x=0.60%. The εr-T curves indicated

11
obvious relaxation behaviors with a highest ferroelectric to paraelectric transition

temperature Tm for x=0.60%. The real-time high temperature kp exhibits a three-stage

evolution with increasing temperature, giving rise to a significant jump with

prominent peak at Td. The in situ and ex situ piezoelectric responses display the same

Td value of 440 °C for x=0.6% ceramic. The scaling behaviors of <A>∝E0β and

<A>∝f0α with different α and β values can be established for the sample with x=0.6%,

and two distinct critical values are derived for both α and β in different ranges of E0

and f0.

Acknowledgments

This work was supported by the Natural Science Foundation of Guangxi

(2017GXNSFDA198024) and the National Nature Science Foundation of China

(11564007 and 11664006),

References

[1] X. Jiang, K. Kim, S. Zhang, J. Johnson, G. Salazar, High-temperature piezoelectric sensing,

Sensors 14 (2014) 144-169.

[2] J. Wu, Z. Fan, D. Xiao, J. Zhu, J. Wang, Multiferroic bismuth ferrite-based materials for

multifunctional applications: ceramic bulks, thin films and nanostructures, Prog. Mater. Sci.

84 (2016) 335-402.

[3] T. Rojac, A. Bencan, B. Malic, G. Tutuncu, J.L. Jones, J.E. Daniels, D. Damjanovic, BiFeO3

ceramics: processing, electrical, and electromechanical properties, J. Am. Ceram. Soc. 97

(2014) 1993-2011.
12
[4] Z.Y. Liao, F. Xue, W. Sun, D.S. Song, Q.Q. Zhang, J.F. Li, L.Q. Chen, J. Zhu, Reversible

phase transition induced large piezoelectric response in Sm-doped BiFeO3 with a composition

near the morphotropic phase boundary, Phys. Rev. B 95 (2017) 214101.

[5] D. Wang, A. Khesro, S. Murakami, A. Feteira, Q. Zhao, I.M. Reaney,

Temperature dependent, large electromechanical strain in Nd-doped BiFeO3-BaTiO3 lead-

free ceramics, J. Eur. Ceram. Soc. 37 (2017) 1857-1860.

[6] J. Lv, X.J. Lou, J.G. Wu, Defect dipole-induced poling characteristics and ferroelectricity of

quenched bismuth ferrite-based ceramics, J. Mater. Chem. C 4 (2016) 6140-6151.

[7] H. Tao, J. Lv, R. Zhang, R.P. Xiang, J.G. Wu, Lead-free rare earth-modified BiFeO3 ceramics:

Phase structure and electrical properties, Mater. Des. 120 (2017) 83-89.

[8] Z.Y. Liao, F. Xue, W. Sun, D.S. Song, Q.Q. Zhang, J.F. Li, L.Q. Chen, J. Zhu, Reversible

phase transition induced large piezoelectric response in Sm-doped BiFeO3 with a composition

near the morphotropic phase boundary, Phys. Rev. B 95 (2017) 214101.

[9] L.Y. Zhang, J. Zhang, Y.F. Chang, G.L. Yuan, B. Yang, S.T. Zhang, Composition-dependent

microstructures and properties of La-, Zn-, and Cr- modified 0.675BiFeO3-0.325BaTiO3

ceramics, J. Am. Ceram. Soc. 99 (2016) 2989-2994.

[10] S. Kim, G.P. Khanal, H.-W. Nam, I. Fujii, S. Ueno, C. Moriyoshi, Y. Kuroiwa, S. Wada,

Structural and electrical characteristics of potential candidate lead-free BiFeO3-BaTiO3

piezoelectric ceramics. J. Appl. Phys. 122 (2017) 164105.

[11] Q. Li, J. Wei, J. Cheng, J. Chen, High temperature dielectric, ferroelectric and piezoelectric

properties of Mn-modified BiFeO3-BaTiO3 lead-free ceramics, J. Mater. Sci. 52 (2016)

229-237.

13
[12] I. Calisir, A.A. Amirov, A.K. Kleppe, D.A. Hall, Optimisation of functional properties in

lead-free BiFeO3-BaTiO3 ceramics through La3+ substitution strategy, J. Mater. Chem. A 6

(2018) 5378-5397.

[13] Y.Q. Guo, P. Xiao, R. Wen, Y. Wan, Q.J. Zheng, D.L. Shi, K.H. Lam, M.L. Liu, D.M. Lin,

Critical roles of Mn-ions in enhancing the insulation, piezoelectricity and multiferroicity

of BiFeO3-based lead-free high temperature ceramics, J. Mater. Chem. C 3 (2015) 5811

-5824.

[14] I. Fujii, R. Iizuka, Y. Nakahira, Y. Sunada, S. Ueno, K. Nakashima, E. Magome, Chikako,

Electric field induced lattice strain in pseudocubic Bi(Mg1/2Ti1/2)O3-modified BaTiO3-BiFeO3

piezoelectric ceramics, Appl. Phys. Lett. 108 (2016) 172903.

[15] J. Chen, J. Cheng, High electric-induced strain and temperature-dependent piezoelectric

properties of 0.75BF-0.25BZT lead-free ceramics, J. Am. Ceram. Soc. 99 (2016) 536-542.

[16] S. Cheng, B.-P. Zhang, L. Zhao, K.-K. Wang, Enhanced insulating and piezoelectric

properties of 0.7BiFeO3-0.3BaTiO3 lead-free ceramics by optimizing calcination temperature:

analysis of Bi3+ volatilization and phase structures, J. Mater. Chem. C (2018), DOI:

10.1039/c8tc00329g.

[17] K. Tong, C. Zhou, Q. Li, J. Wang, L. Yang, J. Xu, G. Chen, C. Yuan, G. Rao, Enhanced

piezoelectric response and high-temperature sensitivity by site-selected doping of

BiFeO3-BaTiO3 ceramics, J. Eur. Ceram. Soc. 38 (2018) 1356-1366.

[18] S. Huang, Q. Li, L. Yang, J. Xu, C. Zhou, G. Chen, C. Yuan, G. Rao, Enhanced piezoelectric

properties by reducing leakage current in Co modified 0.7BiFeO3-0.3BaTiO3 ceramics,

Ceram. Int. (2018), https://doi.org/10.1016/j.ceramint.2018.02.095

14
[19] M.I. Morozov, M.A. Einarsrud, T. Grande, Control of conductivity and electric field induced

strain in bulk Bi0.5K0.5TiO3-BiFeO3 ceramics, Appl. Phys. Lett. 104 (2014) 122905.

[20] D. Wang, Z. Fan, D. Zhou, A. Khesro, S. Murakami, A. Feteira, Q. Zhao, X. Tan, I.M.

Reaney, Bismuth ferrite-based lead free ceramics and multilayers with high recoverable

energy density, J. Mater. Chem. A 6 (2018) 4133-4144.

[21] M.H. Lee, D.J. Kim, J.S. Park, S.W. Kim, T.K. Song, M.H. Kim, M.H. Kim, W.J. K, D. Do,

I.K. Jeong, High-performance lead-free piezoceramics with high Curie temperatures, Adv.

Mater. 27 (2015) 6976-6982.

[22] L.F. Zhu, B.P. Zhang, S. Li, L. Zhao, N. Wang, X.C. Shi, Enhanced piezoelectric properties of

Bi(Mg1/2Ti1/2)O3 modified BiFeO3-BaTiO3 ceramics near the morphotropic phase boundary, J.

Alloy. Compd. 664 (2016) 602-608.

[23] D. Wang, A. Khesro, S. Murakami, A. Feteira, Q. Zhao, I.M. Reaney, Temperature-dependent,

large electromechanical strain in Nd-doped BiFeO3-BaTiO3 lead-free ceramics, J. Eur. Ceram.

Soc. 37 (2017) 1857-1860.

[24] K. Tong, C.R. Zhou, J. Wang, Q. Li, L. Yang, J.W. Xu, W.D. Zeng, G.H. Chen, C.L. Yuan,

G.H. Rao, Enhanced piezoelectricity and high-temperature sensitivity of Zn-modified BF-BT

ceramics by in situ and ex situ measuring, Ceram. Int. 43 (2017) 3734-3740.

[25] S.O. Leontsev, R.E. Eitel, Dielectric and piezoelectric properties in Mn-modified

(1-x)BiFeO3-xBaTiO3 ceramics, J. Am. Ceram. Soc. 92 (2009) 2957-2961.

[26] L. Jin, F. Li, S.J. Zhang, Decoding the fingerprint of ferroelectric loops: comprehension of the

material properties and structures, J. Am. Ceram. Soc. 97 (2014) 1-27.

[27] A.N. Morozovska, E.A. Eliseev, O.V. Varenyk, S.V. Kalinin, Effective piezoelectric response

15
of twin walls in ferroelectrics, J. Appl. Phys. 113 (2013) 187222.

[28] M. Okayasu, T. Yamasaki, Effects of 90° domain switching on electric generation properties

of PZT ceramic, Ceram. Int. 43 (2017) 3590-3600.

[29] Z.F. Li, H. Wu, W.W Cao, Piezoelectric response of charged non-180° domain walls in

ferroelectric ceramics, J. Appl. Phys. 111 (2012) 024106.

[30] C.M. Wang, K. Lau, Q. Wang, Dynamic hysteresis and scaling behaviours of lead-free

0.94Bi0.5Na0.5TiO3-0.06BaTiO3 bulk ceramics, RSC Adv. 6 (2016) 30148-30153.

[31] Y.C. Zhang, C.J. Lu, Z.Z. Yang, W.N. Ye, L.H. Xia, Scaling behavior of dynamic hysteresis in

relaxor ferroelectric 0.67Pb(Mg1/3Nb2/3)O3-0.33PbTiO3 ceramics, J. Appl. Phys. 111 (2012)

084104.

[32] X.F. Chen, X.L. Dong, Z.Y. Zhou, J.X. Wang, F. Cao, G.S. Wang, H.L. Zhang, Dynamic

hysteresis and scaling behavior for Pb(Zr, Ti)O3 ceramics, J. Appl. Phys. 115 (2014) 124103.

[33] Y.A. Genenko, S. Zhukov, S.V. Yampolskii, J. Schütrumpf, R. Dittmer, W. Jo, H. Kungl, M.J.

Hoffmann, H.V. Seggern, Universal polarization switching behavior of disordered

ferroelectrics, Adv. Funct. Mater. 22 (2012) 2058-2066.

[34] J. Fu, R.Z. Zuo, Polarization reversal and dynamic scaling of (Na0.5K0.5)NbO3 lead-free

ferroelectric ceramics with double hysteresis-like loops, J. Appl. Phys. 112 ( 2012) 104114.

[35] J. Shi, H.Q. Fan, X. Liu, Q. Li, Ferroelectric hysteresis loop scaling and electric-field-induced

strain of Bi0.5Na0.5TiO3-BaTiO3 ceramics, Phys. Status Solidi A 211 (2014) 2388-2393.

[36] E.-M. Anton, W. Jo, D. Damjanovic, J. Rödel, Determination of depolarization temperature of

(Bi1/2Na1/2)TiO3-based lead-free piezoceramics, J. Appl. Phys. 110 (2011) 094108.

[37] P.M. Weaver, T. Stevenson, T. Quast, G. Bartl, T. Schmitz-Kempen, P. Woolliams, A.

16
Blumfield, M. Stewart, M.G. Cain, High temperature measurement and characterization of

piezoelectric properties, J. Mater. Sci.: Mater. Electron. 26 (2015) 9268-9278.

[38] C. Huang, K. Cai, Y. Wang, Y. Bai, D. Guo, Revealing the real high temperature

performanceand depolarization characteristics of piezoelectric ceramics by combined in situ

techniques, J. Mater. Chem. C 6 (2018) 1433-1444.

[39] M. Zhang, J. Zhai, B. Shen, X. Yao, MgO doping effects on dielectric properties of

Ba0.55Sr0.45TiO3 ceramics, J. Am. Ceram. Soc. 94 (2011) 3883–3888.

[40] C.-H. Hong, H.-P. Kim, B.-Y. Choi, H.-S. Han, J.S. Son, C.W. Ahn, W. Jo, Lead-free

piezoelectrics-where to move on? J. Materiomics 2 (2016) 1-24.

[41] M. Acosta, N. Novak, G.A.R. Jr, J. Rödel,Mechanisms of electromechanical response in

(1-x)Ba(Zr0.2Ti0.8)O3-x(Ba0.7Ca0.3)TiO3 ceramic, Appl. Phys. Lett. 107 (2015) 142906.

[42] X. Liu, S. Xue, F. Li, J. Ma, J. Zhai, B. Shen, F. Wang, X. Zhao, H. Yan, Giant electrostrain

accompanying structural evolution in lead-free NBT-based piezoceramics, J. Mater. Chem. C

6 (2018) 814-822.

[43] F. Li, D. Lin, Z. Chen, Z. Cheng, J. Wang, C. Li, Z. Xu, Q. Huang, X. Liao, L.-Q. Chen, T.R.

Shrout, S. Zhang, Ultrahigh piezoelectricity in ferroelectric ceramics by design, Nat. Mater.

17 (2018) 349-354.

[44] X. Lu, L. Zheng, H. Li, W. Cao, Theoretical study on phase coexistence in ferroelectric solid

solutions near the tricritical point, J. Appl. Phys. 117 (2015) 134101.

[45] Y. Feng, W.-L. Li, Y. Yu, H.-N. Jia, Y.-L. Qiao, W.-D. Fei,Enhanced piezoelectricity in

ABO3 ferroelectrics via intrinsic stress-driven flattening of the free-energy profile, Phys. Rev.

Mater. 1 (2017) 064405.

17
[46] K. Yan, S. Ren, M. Fang, X. Ren, Crucial role of octahedral untilting R3m/P4mm

morphotropic phase boundary in highly piezoelectric perovskite oxide, Acta Mater. 134

(2017) 195-202.

[47] D. Damjanovic, A morphotropic phase boundary system based on polarization rotation and

polarization extension, Appl. Phys. Lett. 97 (2010) 062906.

Fig. 1. Room temperature XRD patterns of (a) BFMx-BT ceramics, (b) enlarged part

of XRD patterns in the 2θ range of 38-46°

Fig. 2. Grain morphology and grain size distribution of BFMx-BT ceramics

Fig. 3 The variation of the weak-signal piezoelectric properties (d33 and kp) (a),

maximum phase angle (θmax) (b), bipolar strain curves (c), current-electrical field (I-E)

loops (d), polarization-electrical field (P-E) hysteresis loops (e) and remnant

polarization (Pr) and coercive field (EC) (f)

Fig. 4 Temperature-dependent dielectric constant εr and loss (tanδ)(a-e), and local

maxima εm and Tm (f)

Fig. 5 Real-time high temperature kp (a), temperature-dependent kp by ex situ

measurement (b), and Td as a function of the Mg content x (c)

Fig. 6 Electric field amplitude (E0) dependence of P-E hysteresis loops (a), 2Pr, 2Pmax,

and 2Ec of the hysteresis loops (b), hysteresis area <A> of the hysteresis loops (c), and

the plots of ln<A> versus lnE0 (d)

Fig. 7 Frequency (f0) dependence of P-E hysteresis loops (a), and the plots of ln<A>

versus lnf0 (b)

18
Fig. 1

(a) (b)
BiFeO3 PDF:86-1518

20 30 40 50 60 70 80

(200)
BiTiO3 PDF:75-2121
Intensity(a.u.)

(021)
(003)
20 30 40 50 60 70 80
 Bi2Fe4O9 PDF: 72-1832 x=0.9%
(110)

(200)
(021)

(211)

(220)
(100)

(003)

(210)
(112)

(208)
x=0.75%
x=0.9%
x=0.75% x=0.6%
x=0.6% x=0.3%
x=0.3%
 x=0% x=0%

20 30 40 50  60 70 80 38 
46
2 2

19
Fig. 2

20
Fig. 3

180 0.32 24 (b)


d33=170 pC/N
(a)
kp=0.29 18
160 0.28
d33 (pC/N)

kp
12

max ( )
o
140 d33 0.24

kp 6
120 0.20
0
0.0 0.3 0.6 0.9 0.0 0.3 0.6 0.9
x (%) x (%)
0.12 0.12 0.18
(c) 0.11
(d)
Smax(%)

0.10 0.12
0.08
Strain (%)

0.09
0.0 0.3 0.6 0.9
x(%) 0.06
J ( mA)

0.04 0.00
x=0
x=0
x=0.30%
-0.06 x=0.30%
0.00 x=0.60% x=0.60%
x=0.75% -0.12 x=0.75%
x=0.90%
x=0.90%
-60 -40 -20 0 20 40 60 -0.18
-60 -40 -20 0 20 40 60
E (kV/cm) E (kV/cm)
30 18
x=0% (e)
20 x=0.30%
x=0.60%
(f) 26
Pr ( C/cm )

10 x=0.75% 16
2
P( C/cm )

Ec (kV/cm)
x=0.90%
2

24
0
14
-10
Pr 22
-20 12 Ec
-30 20
-80 -60 -40 -20 0 20 40 60 80 0.0 0.3 0.6 0.9
E(kV/cm) x (%)

21
Fig. 4

20 20 30 30
(a) x=0 Tm (b) x=0.3% Tm
25 25
15 15
1kHz 1kHz
20 20
10kHz 10kHz
r( 10 )

r ( 10 )
10
3

3
100kHz 15

tan
10 15

tan
100kHz
10 10
5
5
5
5
0 0
0 0
100 200 300 400 500 100 200 300 400 500
o
T( C) T ( C)
o

40 40 40 40
(c) x=0.6% Tm (d) x=0.75% Tm
32 32 32 32
1kHz 1kHz
24 10kHz 24 10kHz
24 24
r ( 10 )

r ( 10 )
3

100kHz 100kHz
tan
3

tan
16 16
16 16
8 8
8 8
0 0
0 0
100 200 300 400 500 100 200 300 400 500
T ( C)
o
T ( C)
o

30 20 460
(e) x=0.9% 40
24 (f)
1kHz 15 450
10kHz 32
m ( 10 )

18 Tm
r ( 10 )

Tm( C)
3

100kHz
tan

10 440
o
12 24
6 5 430
16
0
0 420
100 200 300 400 500 0.0 0.3 0.6 0.9
T ( C)
o
x (%)

22
Fig. 5

0%
(a) in situ 0.30% III
0.4 0.60%
0.75%
0.90%
I II
kp

0.3

0.2
0 100 200 300 400 500

T ( C)
(b) ex situ 0%
0.30
0.30%
0.60%
0.75%
0.25 0.90%
kp

0.20

0.15
0 100 200  300 400 500
T ( C)
480
(c)
440
Td ( C)
o

400

360

320
0.0 0.3 0.6 0.9
x (%)

23
Fig. 6

10kV/cm
30 15kV/cm 60 (b) 60
20kV/cm (a)

2Pr, 2Pmax ( C/cm )


2
20 25kV/cm 50 50
30kV/cm
35kV/cm
10 40 40

2Ec (kV/cm)
P (C/cm )

40kV/cm
2

45kV/cm E0 increasing
0 50kV/cm 30 30
55kV/cm
-10 60kV/cm
20 2Pr 20
65kV/cm
-20 10 2Pmax 10
-30 2Ec
0 0
-60 -40 -20 0 20 40 60
10 20 30 40 50 60 70
2.5
E (kV/cm) E(kV/cm)
β
<A>∝E0
(c) 8
( d)
2.0 ln <A>=βlnE0+m

ln<A> ( ln( mCV/cm ) )


Hysteresis loop area <A>
3 6 β=1.365,
<A> ( CV/cm )

1.5
3

m=2.055
β=4.842, m=-10.634 2
R =0.9966
1.0 4 2
R =0.9994
I II
I II Data
0.5 2
fitting curve low E0
fitting curve high E0
0.0 0
10 20 30 40 50 60 70 2.0 2.5 3.0 3.5 4.0 4.5
E0 (kV/cm) lnE0 ( ln(kV/cm))

24
Fig. 7

0.01Hz 8.2
40 f increasing (a) (b)
0.02Hz
0.05Hz
<A>∝f α ln <A>=αlnf+n
0.1Hz 8.0 Data

ln<A> (ln(mJ/cm ))
20 0.2Hz
fitting curve

3
0.5Hz
7.8
P (C/cm )

1Hz
2

2Hz f increasing
0 5Hz
10Hz
7.6
-20
7.4
-40
7.2
-60 -40 -20 0 20 40 60 -5 -4 -3 -2 -1 0 1 2 3
E (kV/cm) lnf (ln(Hz))

25

You might also like