Next Article in Journal
Effect of Recycled Coarse Aggregate and Bagasse Ash on Two-Stage Concrete
Next Article in Special Issue
Conductivity Transitions of La0.7Sr0.3MnOδ and La0.6Sr0.4Co0.2Fe0.8O3−δ in Ce0.9Gd0.1O2−δ Matrix for Dual-Phase Oxygen Transport Membranes
Previous Article in Journal
Investigation of the Wear Behavior of Surface Welding AZ91 and AZ91+Gd Alloys under Variable Loading Conditions
Previous Article in Special Issue
Influence of Doping on the Transport Properties of Y1−xLnxMnO3+δ (Ln: Pr, Nd)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Electrostrictive Coefficient and Suppressive Hysteresis in Lead-Free Ba(1−x)SrxTiO3 Piezoelectric Ceramics with High Strain

1
The State Key Laboratory of Materials-Oriented Chemical Engineering, College of Materials Science and Engineering, Nanjing Tech University, Nanjing 210009, China
2
Jiangsu National Synergetic Innovation Center for Advanced Materials (SICAM), Nanjing Tech University, Nanjing 210009, China
3
Jiangsu Collaborative Innovation Center for Advanced Inorganic Function Composites, Nanjing Tech University, Nanjing 210009, China
*
Authors to whom correspondence should be addressed.
Submission received: 16 April 2021 / Revised: 8 May 2021 / Accepted: 13 May 2021 / Published: 16 May 2021

Abstract

:
Lead-free piezoelectric ceramics with both low hysteresis and superior electrostrictive coefficient features are crucial toward providing desired performance for intelligent electrical devices, especially in high-precision displacement actuators. In this work, we propose a novel scenario, which is to design the phase transition around ambient temperature to enhance electrostrictive effect and inhibit hysteresis. In other words, the dense ceramics with cubic phases (C) and tetragonal phases (T) coexisting at RT (room temperature) were designed. According to this scenario, the Ba(1−x)SrxTiO3 (abbreviated as BT-100xST) ceramics were fabricated by the conventional solid-state reaction method. The relaxor behavior, ferroelectric properties, crystal structure and microstructure of BT-100xST ceramics have been investigated in detail. As a result, the BT-100xST ceramics with x = 0.20–0.40 present relaxor behavior which was indicated by dielectric constant as a function of temperature and (polarization–electric field) PE hysteresis loops. The BT-30ST ceramics exhibit enhanced electrostrictive coefficient Q33 (>0.034 m4/C2), and the electrostrictive strain and low hysteresis achieves 0.11% and 2%, respectively. The BT-100xST ceramics are considered as a prospective option for application in displacement actuators with high sensitivity and high precision.

1. Introduction

Piezoelectric ceramics have been widely used in electronic devices, such as sensors, actuators, and transducer and ultrasonic motors [1,2,3], which are required not only high-strain but also ultra-low hysteresis. However, many researchers just devoted themselves to realizing the critical target of achieving large electric-field-induced strain. Hao et al. [4,5] fabricated high-strain ceramics by designing a phase transition from a ferroelectric to an ergodic relaxor phase in the lead-free ternary Bi0.5Na0.4K0.1TiO3-KxNa1−xNbO3 systems. In this scenario, however, a high-strain response is commonly accompanied by large hysteresis >60% with strong nonlinearity. Based on this strategy, other works also reported on Bi0.5N0.5TiO3-based ceramics, accompanying high hysteresis that is observed simultaneously [6,7]. Recently, several studies demonstrated the addition of compound or cation doping can suppress hysteresis [8,9]. For instance, relatively small hysteresis—as low as 23%—was achieved in a Bi4Ti3O12-Bi0.5(Na0.82K0.18)0.5TiO3 system, as reported by Fan et al. [10], which is attributed to A-site vacancies to destroy the long-range ferroelectric order and promote the formation of the extremely stable relaxor phase at room temperature.
Another effective method to overcome the significant challenge of high hysteresis is to design relaxor ferroelectric with electrostrictive strain. For the canonical lead-free BNT-based relaxor, the electrostrictive strain refers to “a ferroelectric relaxor + a ferroelectric” for solid solutions [11]. However, the electrostrictive coefficient is low (Q33 = 0.020–0.024 m4/C2) [12]. A large strain predominantly based on electrostrictive effect has been demonstrated in lead-based relaxor ferroelectrics with a T–C coexisting phase structure. Meanwhile, researchers found that doping enhanced the dielectric diffuseness and relaxation and inhibited hysteresis. As an important figure of merit, the Q33 in PMN-based systems, including PbMg1/3Nb2/3O3-PbTiO3 [13], Pb(In1/2Nb1/2)O3-Pb(Mg1/3Nb2/3)O3-PbTiO3 [14], Bi(Li0.5Nb0.5)O3-Pb(Mg1/3Nb2/3)O3-PbTiO3 [15], etc., ranges from 0.03 m4/C2 to 0.04 m4/C2. A high strain of 0.16–0.20% with low hysteresis < 7% was also observed in these samples. However, in today’s environmentally conscious world, lead-free electrostrictive materials are highly desirable for safety reasons because of the toxicity of lead compounds [16,17].
BaTiO3-based ferroelectric ceramics are one of the most promising lead-free replacements, owing to the relatively high electric field strain and low hysteresis compared with other systems of piezoelectric ceramics. These superior strain properties come from remarkable piezoelectric effects around the polymorphic phase boundary (PPB) or morphological phase boundary (MPB) [18]. Wu et al. reported that the addition of Sr2+ is a modifier that could effectively shift the TRO (Rhombohedral–Orthorhombic) and TOT (Orthorhombic–Tetragonal) values to construct phase boundary; as a result, 0.13%strain with 10% hysteresis was achieved [19]. Zhao et al. [20] explored a high-strain BT-based system of (Ba1−yCay)(Ti1−xHfx)O3, which had the surprisingly enormous strain of Smax = 0.21%and piezoelectricity coefficient of d33 = 540 pC/N by constructing multiphase coexistence, including the R, O and T. Nevertheless, these samples were accompanied by hysteresis >10%. It can be expected that we can possibly design BT-based piezoelectric ceramics of electrostrictive effect near the phase boundary, which exhibit high-strain ceramics and further suppress hysteresis.
Following the above research, we designed lead-free BaTiO3-based system piezoelectric ceramics by doping a cubic SrTiO3 into tetragonal BaTiO3 [21]. The two crystal structures of cubic and tetragonal were selected as a result of both having the potential to form a ferroelectric and pharaelectric phase boundary. At this time, the strain of BT-100xST piezoelectric ceramics were generated by electrostrictive effect [22]. Meanwhile, the BT-ST solid solution was confirmed because equivalent substitution can induce relaxor ferroelectric with low hysteresis [23]. In the case of BT-100xST ceramics, although microwave-tunability [24], sintering behavior [25,26], and energy storage properties [27,28] have been researched extensively during the past decades, hysteresis and electrostrictive effects are still rarely reported. In the following, we shall expound the effect of Sr2+ addition to BaTiO3 ceramics on the crystal structure, microstructure dielectric, ferroelectric, strain and electrostrictive properties. Consequently, we have achieved an extraordinary combination of low hysteresis, enhanced electrostrictive coefficient and high strain.

2. Materials and Methods

The powders of BaCO3 (99.0%), SrCO3 (99.0%), TiO2 (99.0%) were weighted as starting raw materials according to the stoichiometric ratio of Ba(1−x)SrxTiO3 (x = 0.10, 0.15, 0.20, 0.30, 0.40, 0.50 and 0.60, abbreviated as BT-100xST). The BT-100xST ceramics were fabricated by the conventional solid-state reaction method, and the preparation process mainly carried out the following chemical reactions to form a solid solution [24]. Details of the production process can refer to Refs. [29,30].
BaCO3→BaO + CO2
SrCO3→SrO + CO2
BaO + SrO + TiO2 → (BaSr)TiO3[BaTiO3-SrTiO3]
X-ray diffraction (XRD, Smart Lab 3 kW, Rigaku, Japan) analysis using Cu Kα1 radiation was used to investigate the crystal structures of BT-100xST ceramics. The surface appearance of thermally-etched samples was imaged by scanning electron microscopy (SEM, JSM-6510, JEOL, Tokyo, Japan). A ferroelectric tester instrument (Precision Premier II, Radiant Technology, San Diego, CA, USA) was used to measure electric properties at room temperature, including bipolar polarization–electric field (PE) hysteresis loops, bipolar and unipolar strain-electric field (SE) curves and current–electric field (IE) curves. Temperature-dependent dielectric permittivity and loss tangent under different frequencies (0.1 kHz, 1 kHz, 10 kHz, 100 kHz) were measured by an LCR (Lenz Capacitor Resistance) meter (4284 Agilent, Palo Alto, CA, USA) from −100 °C to 150 °C at a heating rate of 2 °C/min.

3. Results and Discussions

3.1. Microstructure

The room temperature XRD patterns of BT-100xST ceramics are illustrated in Figure 1a. It can be seen that the detectable peaks of all samples demonstrate a pristine perovskite structure without the presence of impurity peaks, implying that the Sr element has diffused into the BT lattice to form complete, solid solutions. The diffraction peaks shift steadily to higher angles with the increase in x (Figure 1b) as a result of the small radius (1.440 Å) Sr2+ substituting for the large radius (1.610 Å) Ba2+, which in turn results in the crystal lattice shrinkage. Meanwhile, the diffraction peaks near 45° at x < 0.40 exhibit split characteristics, and show a tendency to merge with the increasing quantity of x. A single diffraction peak was observed in the composition with x ≥ 0.40. To intuitively confirm the revolution of phase structure, all the peaks around 45° were depicted by peak fitting, as shown in Figure 1c. The sharp (200)T, (002)T peaks and ultra-weaker (200)C peaks were identified in x < 0.30, which are attributed to a main tetragonal phase and infinitesimal cubic phase, while an obvious (200)C peak was detected for x = 0.30, indicating the coexistence of the T and C phase. With a further increase in x, the sharp (200)C peak and negligible (002)T peak were presented, suggesting the cubic phase nature of crystal structure.
The SEM surface images of the BT-100xST ceramics with different Sr contents are shown in Figure 2. All samples excellently exhibited dense grain structure, and the microstructures become uniform with the increasing of x. The average grain size of BT-100xST ceramics was obtained based on the statistical method and the corresponding results are illustrated in the inset of the SEM surface images. It is noted that the BT-10ST ceramics exhibit a large average grain size of 17.8 μm, with the grain size gradually decreasing as a result of increased ST concentration. Eventually, the grain size decreases to 5.2 μm for BT-60ST ceramics. Above all, results imply that the Sr2+ ions can inhibit the growth of grains and lead to uniform micromorphology.

3.2. Dielectric Behaviors

The phase structure of BT-100xST ceramics was further demonstrated by the temperature dependence of permittivity and loss tangent, as is presented in Figure 3a–g. It can be seen that several dielectric peaks and dielectric loss peaks associated with phase transition are detected, and they are denoted as Tm, T1, and T2. Following the decreasing of temperature, Tm, T1, and T2 correspond to the phase transition temperature of cubic paraelectric to tetragonal ferroelectric, tetragonal ferroelectric to orthorhombic ferroelectric, and orthorhombic ferroelectric to rhombohedral ferroelectric, respectively [31]. Increasing the ST content causes Tm, T1 and T2 to shift to low temperatures. In addition, the dielectric loss does not exceed 2%at higher frequencies. Note that the value of dielectric loss measured at 0.1 kHz rises dramatically when the temperature exceeds 80 °C. This might be due to charges or defects being activated easily at high temperature and low frequency [32]. According to the profile of permittivity and dielectric loss as a function of temperature, the phase diagram is depicted as shown in Figure 2h. The phase structure of BT-100xST ceramics is tetragonal for x ≤ 0.20 and cubic for x ≥ 0.40. For x = 0.30, the BT-30ST ceramics exhibit a tetragonal phase coexisting with the cubic phase at room temperature. This result is consistent with the phase structure determined by the XRD pattern.
The further detail regarding the frequency dependence of permittivity as a function of temperature is presented in Figure 4. By increasing Sr content, the dielectric peaks begin flattening. Furthermore, the significant frequency dispersion that causes the dielectric constant to decline and the corresponding Tm peak shifts to increase in temperature at a higher frequency were observed at x = 0.20–0.40. For x being up to 0.40, the dielectric anomaly peak becomes even narrower, which implies the feebleness of the relaxor degree. In general, a broad Tm peak exists in the composition of BT-100xST ceramics due to the diffuse phase transition (DPT).
In order to analyze the diffuseness of the DPT quantitatively, a modified Curie–Weiss law is defined, followed by Equation (4) [33,34].
1 ε 1 ε r = ( T T m ) γ C   ( T > T m )
where C is the Curie-like constant. The index γ represents the character of the phase transition; γ = 1 and γ = 2 correspond to a typical ferroelectric and absolute diffuse phase transition, respectively. For different compositions, the slope of a linear relation between ln(1/εr − 1/εm) and ln(TTm) was extracted by fitting a straight-line equation, as shown in Figure 5. With an increase in x from 0.10 to 0.30, the index γ value rises from 1.19 to 1.36, and then γ decreases to 1.10, where x = 0.60. The value of γ for x = 0.30 and 0.40 are 1.36 and 1.29, respectively. These results suggest that BT-30ST and BT-40ST ceramics have more obvious diffuse phase transitions, which corresponds to broader dielectric peaks. The fluctuation of local components leads to the occurrence of DPT, which leads to microregions with different local Curie points exhibiting a Gaussian distribution near the mean Curie temperature [35]. Above all, these results demonstrate that the ceramics with x = 0.20–0.40 have canonical relaxor behavioral properties.

3.3. Ferroelectric and Piezoelectric Properties

Figure 6a presents the room temperature PE hysteresis loops of BT-100xST (x = 0.10–0.60). The well-saturated hysteresis loops of low-doping BT-based ceramics (x ≤ 0.20) are on the verge of typical ferroelectric behavior [36], and the high values of Pmax, Pr, and Ec were measured, corresponding to 26.2 μC/cm2, 14.1 μC/cm2 and 10.3 kV/cm, respectively, at x = 0.10. In the composition of x = 0.10–0.40, it is worth noting that the hysteresis loop transformed from a square profile into a slim profile with the increase in x. The slim hysteresis loop implies the prominent relaxor characteristic. The variation of maximum polarization (Pmax), remnant polarization (Pr), and coercive electric field (Ec) as a function of x is depicted in Figure 6b. The decline of Pr and Ec is more dramatic than Pmax. As a result, the following values were achieved at x = 0.40: Pr = 0.85 μC/cm2, Ec = 1.5 kV/cm and Pmax = 16 μC/cm2. To better interpret the relaxor behavior, Figure 6c presents the IE curves of BT-100xST (x = 0.10–0.60). For the ceramics with x = 0.10–0.40, it is obviously observed that sharp current peaks gradually broaden. The broad current peaks also reflect a relaxor ferroelectric behavior of the ceramics with x = 0.30–0.40 [37]. The current peak is related to domain reversal, while a broad current peak refers to the switching of nano-domains, which is different from the sharp current peak generated by the switching of macro-domains [35]. With a further increase in ST content, the ceramic shows a linear feature, indicating that the BT-60ST ceramic is a linear dielectric material. It is worth noting that the corresponding IE curves are a square loop. These results have also been reflected in other reports [38].
Figure 7 presents the room temperature S–E curves of BT-100xST (x = 0.10–0.60). The bipolar S–E curves (Figure 7a) show the butterfly-like shape at a low degree of substitution (x ≤ 0.20), which exhibits the ferroelectric feature. In contrast, the S–E curves transform from the butterfly shape to a symmetric narrow shape (along with the disappearance of the negative electro-strain) as the x exceeds 0.30, originating from the appearance of typical relaxor ferroelectrics. The unipolar S–E curves (Figure 7b) show that the strain rises upward first and then falls down with the increasing value of x. The maximum value of unipolar strain was obtained in x = 0.20. Figure 7c clearly shows the variation of maximum strain (Smax), negative strain (Sneg), and hysteresis (Hys) as a function of x. For x being up to 0.20, the distinct negative strain is attributed to the irreversible ferroelectric domain wall switching [38]. Thereafter, it keeps a downward tendency and approximates to zero, suggesting that irreversible ferroelectric domains decrease. The strain hysteresis is defined by the equation: Hys = ΔS/Smax × 100%, where ΔS is the difference of strain at a half maximum field. The highest strain (0.20%) was obtained at x = 0.20, while accompanied by a relatively high hysteresis (11.48%). The superior property of less than 2%hysteresis with 0.11%strain was obtained at x = 0.30. With the further increase to x = 0.50, the hysteresis increases drastically.

3.4. Electrostrictive Effect

Compared with the extrinsic activity (domain wall motion) and piezoelectric effect, the electrostrictive contribution shows ultra-low hysteresis and high stability with response to the extra electric field [39,40]. The BT-30ST and BT-40ST ceramics present ultra-low hysteresis properties, which implies that the strain with the low hysteresis is mainly attributed to intrinsic electrostrictive effect rather than macro-domain switching. Figure 8 shows the strain as a function of polarization at room temperature of BT-100xST ceramics. The width trait of the S–P curves dependent of x also reflects the magnitude of the hysteresis as shown in S–E curves. According to the equation S = QP2, Figure 9 shows the electrostriction coefficient Q33 obtained by the quadratic fitting of S and P derived from the corresponding S–E and P–E curves. The magnitude of the R coefficient of quadratic fitting S–P curves reflects the degree of deviation from the quadratic relation as shown in Figure 9. The value of R being distant from one means that the electrostrictive effect is reduced.
As can be seen in Figure 8 and Figure 9, the S–P curves of ceramics where x ≤ 0.20 deviates from the quadratic relation, presenting an obvious hysteresis, which means that the strain is not only based on electrostrictive effect, but also affected by the extrinsic motion of domain walls [39]. The results are in good agreement with the result of the S–E loop. In contrast, S–P curves are well-fitted, based on a quadratic relationship for BT-30ST and BT-40ST ceramics, indicating a favorable electrostriction effect (R > 0.98). Furthermore, the Q33 of BT-30ST and BT-40ST ceramics are 0.030 m4/c2 and Q33 = 0.034 m4/c2, which are higher than most BNT-based ceramics [41,42].

3.5. Overview

Table 1 summarizes the Hys, Q33 and Smax properties in the BT-100xST and the recently reported BT-based and other system ceramics. It can be seen that the Hys value is inferior to that of other BT-based and BNT-based piezoelectric ceramics, and the value of Q33 is also superior to lead-based and BNT-based piezoelectric ceramics. In addition, the BT-30ST ceramics strain values remain satisfactory. For previously published works, firstly, the giant strain is based on transition from ferroelectric to relaxor phase or ferroelectric to aftiferroelectric phase, the result always inevitably accompanied by a large hysteresis in BNT-based ceramics. Secondly, improved Q33 and low hysteresis were observed by shifting TC to around room temperature quickly or designing a relaxor with electrostrictive effect in a lead-based ceramics system; however, the lead material is an unmet demand in practical application. Thirdly, BT-based ceramics can efficaciously suppress hysteresis by chemical modification. Finally, in this work the BT-100xST material was selected and a manner of coexisting phase was designed. The result show that enhanced Q33 and suppressive hysteresis.

4. Conclusions

In summary, the macroscopic electrostrictive strain of BT-100xST ceramics was analyzed by multi-perspective, including the microstructure, crystal structure, relaxor properties, and ferroelectric properties. A highlight of this investigation is to adjust the composition to promote relaxor behavior appearance with electrostrictive effect, and construct the T–C phase boundary at room temperature. The results demonstrate that the uniform and dense microstructure, along with the coexisting structure of cubic symmetry and tetragonal symmetry, contributes to the superior properties. For the composition of x = 0.20–0.40, the frequency shift of Tm and frequency dependence of permittivity as a function of temperature illustrates that the ceramics have canonical relaxor behavior with γ = 1.23–1.36. Simultaneously, the PE loops, S–E loops and IE curves present relaxor behavior characteristics. As a result, an ultra-low hysteresis (<2%) with a high strain (>0.11%) has been successfully acquired at a composition of x = 0.30, and accompanied by a superior electrostrictive coefficient Q33 (0.034 m4/c2). Overall, the BT-30ST ceramics would have significant advantages in high-precision ceramic actuators.

Author Contributions

Conceptualization, Y.L. (Yinong Lyu); formal analysis, X.S.; investigation, M.S.; methodology, M.S. and Q.L.; resources, Q.L. and H.Q.; software, H.Q.; supervision, Y.L. (Yunfei Liu) and Y.L. (Yinong Lyu); visualization, X.S.; writing—original draft, M.S.; Writing—review and editing, Y.L. (Yunfei Liu) and Y.L. (Yinong Lyu). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available within the article.

Acknowledgments

This work was supported by the Priority Academic Program Development (PAPD) of Jiangsu Higher Education Institutions.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Lu, H.; Liu, L.; Lin, J.; Yang, W.; Zheng, W.; Weng, L.; Zhang, X. Diffuse characteristics and piezoelectric properties of Tb-doped BCZT ceramics with CaCl2 as sintering aid. Ceram. Int. 2017, 43, 16348–16355. [Google Scholar] [CrossRef]
  2. Wu, Y.; Li, J.; Bai, H.; Hong, Y.; Shi, K.; Zhou, Z.; Guo, R.; Bhalla, A.S. Origin of the dielectric abnormities and tunable dielectric properties in doped KTN single crystals. Appl. Phys. Lett. 2017, 111, 242902. [Google Scholar] [CrossRef]
  3. Liu, Q.; Zhang, Y.; Gao, J.; Zhou, Z.; Wang, H.; Wang, K.; Zhang, X.; Li, L.; Li, J. High-performance lead-free piezoelectrics with local structural heterogeneity. Energ. Environ. Sci. 2018, 11, 3531–3539. [Google Scholar] [CrossRef]
  4. Hao, J.; Shen, B.; Zhai, J.; Chen, H. Phase transitional behavior and electric field-induced large strain in alkali niobate-modified Bi0.5(Na0.80K0.20)(0.5)TiO3 lead-free piezoceramics. J. Appl. Phys. 2014, 115, 034101. [Google Scholar] [CrossRef]
  5. Hao, J.; Shen, B.; Zhai, J.; Liu, C.; Li, X.; Gao, X. Large Strain Response in 0.99(Bi0.5Na0.4K0.1)TiO3-0.01(KxNa1−x)NbO3 Lead-Free Ceramics Induced by the Change of K/Na Ratio in (KxNa1−x)NbO3. J. Am. Ceram. Soc. 2013, 96, 3133–3140. [Google Scholar] [CrossRef]
  6. He, H.; Lu, X.; Li, M.; Wang, Y.; Li, Z.; Lu, Z.; Lu, L. Thermal and compositional driven relaxor ferroelectric behaviours of lead-free Bi0.5Na0.5TiO3-SrTiO3 ceramics. J. Mater. Chem. C 2020, 8, 2411–2418. [Google Scholar] [CrossRef]
  7. Zhao, J.; Zhang, N.; Quan, Y.; Niu, G.; Ren, W.; Wang, Z.; Zheng, K.; Zhao, Y.; Ye, Z. Evolution of mesoscopic domain structure and macroscopic properties in lead-free Bi0.5Na0.5TiO3-BaTiO3 ferroelectric ceramics. J. Appl. Phys. 2021, 129, 084103. [Google Scholar] [CrossRef]
  8. Janbua, J.; Niemchareon, S.; Muanghlua, R.; Vittayakorn, N. High Strain Response of the (1−x)(0.94Bi(0.5)Na(0.5)TiO(3)-0.06BaTiO(3))-xBaSnO(3) Lead Free Piezoelectric Ceramics System. Ferroelectrics 2016, 490, 13–22. [Google Scholar] [CrossRef]
  9. Bai, W.; Chen, D.; Huang, Y.; Zheng, P.; Zhong, J.; Ding, M.; Yuan, Y.; Shen, B.; Zhai, J.; Ji, Z. Temperature-insensitive large strain response with a low hysteresis behavior in BNT-based ceramics. Ceram. Int. 2016, 42, 7669–7680. [Google Scholar] [CrossRef]
  10. Fan, P.; Zhang, Y.; Zhang, Q.; Xie, B.; Zhu, Y.; Mawat, M.A.; Ma, W.; Liu, K.; Xiao, J.; Zhang, H. Large strain with low hysteresis in Bi4Ti3O12 modified Bi1/2(Na0.82K0.18)1/2TiO3 lead-free piezoceramics. J. Eur. Ceram. Soc. 2018, 38, 4404–4413. [Google Scholar] [CrossRef]
  11. Liu, X.; Bell, A.J.; Fan, H.; Shi, J. Large Electrostrictive Strain in (Bi0.5Na0.5)TiO3-BaTiO3-(Sr0.7Bi0.2)TiO3 Solid Solutions. J. Am. Ceram. Soc. 2014, 97, 848–853. [Google Scholar]
  12. Han, H.; Jo, W.; Kang, J.; Ahn, C.; Kim, I.W.; Ahn, K.; Lee, J. Incipient Piezoelectrics and Electrostriction Behavior in Sn-Doped Bi1/2(Na0.82K0.18)(1/2)TiO3 Lead-Free Ceramics. J. Appl. Phys. 2013, 113, 15410215. [Google Scholar] [CrossRef] [Green Version]
  13. Noblanc, O.; Gaucher, P.; Calvarin, G. Structural and dielectric studies of Pb(Mg1/3Nb2/3)O3-PbTiO3 ferroelectric solid solutions around the morphotropic boundary. J. Appl. Phys. 1996, 79, 4291–4297. [Google Scholar] [CrossRef]
  14. Qi, X.; Zhao, Y.; Sun, E.; Du, J.; Li, K.; Sun, Y.; Yang, B.; Zhang, R.; Cao, W. Large electrostrictive effect and high energy storage performance of Pr3+ doped PIN-PMN-PT multifunctional ceramics in the ergodic relaxor phase. J. Eur. Ceram. Soc. 2019, 39, 4060–4069. [Google Scholar] [CrossRef]
  15. Jin, L.; Pang, J.; Luo, W.; Lan, Y.; Du, H.; Yang, S.; Li, F.; Tian, Y.; Wei, X.; Xu, Z.; et al. Phase transition behavior and high electrostrictive strains in Bi(Li0.5Nb0.5)O3-doped lead magnesium niobate-based solid solutions. J. Alloys Compd. 2019, 806, 206–214. [Google Scholar] [CrossRef]
  16. Fu, H.R.; Wang, Y.G.; Guo, H.; Jain, A.; Chen, F.G. Enhancing photostriction in KNN-based ceramics by constructing the morphotropic phase boundary and narrowing the energy band gap. Ceram. Int. 2021, 47, 10996–11002. [Google Scholar] [CrossRef]
  17. Shi, P.; Li, T.; Lou, X.; Yu, Z.; Zhu, X.; Zhou, C.; Liu, Q.; He, L.; Zhang, X.; Yang, S. Large electric-field-induced strain and energy storage properties in Bi0.5Na0.5TiO3-(0.5Ba(0.7)Ca(0.3)TiO(3)-0.5BaTi(0.8)Zr(0.2)O(3)) lead-free relaxor ferroelectric ceramics. J. Alloys Compd. 2021, 860, 206–214. [Google Scholar]
  18. Zhao, C.; Wu, B.; Wu, J. Composition-driven broad phase boundary for optimizing properties and stability in lead-free barium titanate ceramics. J. Am. Ceram. Soc. 2019, 102, 3477–3487. [Google Scholar] [CrossRef]
  19. Wu, W.; Ma, J.; Wang, N.; Shi, C.; Chen, K.; Zhu, Y.; Chen, M.; Wu, B. Electrical properties, strain stability and electrostrictive behavior in 0.5BaZr0.2Ti0.8O3-(0.5−x)Ba0.7Ca0.3TiO3-xBa0.7Sr0.3TiO3 lead-free ceramics. J. Alloys Compd. 2020, 814, 152240. [Google Scholar] [CrossRef]
  20. Zhao, C.; Wu, W.; Wang, H.; Wu, J. Site engineering and polarization characteristics in (Ba1−yCay)(Ti1−xHfx)O3 lead-free ceramics. J. Appl. Phys. 2016, 119, 0241082. [Google Scholar]
  21. Zhao, C.; Huang, Y.; Wu, J. Multifunctional barium titanate ceramics via chemical modification tuning phase structure. Infomat 2020, 2, 1163–1190. [Google Scholar] [CrossRef]
  22. Zhu, Y. Large electrostrictive properties in lead-free BaTiO3-ZnSnO3 solid solutions. Appl. Phys. A 2019, 125, 301. [Google Scholar] [CrossRef]
  23. Maiti, T.; Guo, R.; Bhalla, A.S. Structure-property phase diagram of BaZrxTi1−xO3 system. J. Am. Ceram. Soc. 2008, 91, 1769–1780. [Google Scholar] [CrossRef]
  24. Teranishi, T.; Kanemoto, R.; Hayashi, H.; Kishimoto, A. Effect of the (Ba + Sr)/Ti ratio on the microwave-tunable properties of Ba0.6Sr0.4TiO3 ceramics. J. Am. Ceram. Soc. 2017, 100, 1037–1043. [Google Scholar] [CrossRef]
  25. Ying, K.; Hsieh, T. Sintering Behavior, Microstructure, and Dielectric Properties of Nano-Ba0.7Sr0.3TiO3 Ceramics. Jpn. J. Appl. Phys. 2008, 47, 7947–7952. [Google Scholar] [CrossRef]
  26. Wu, T.; Pu, Y.; Gao, P.; Liu, D. Influence of Sr/Ba ratio on the energy storage properties and dielectric relaxation behaviors of strontium barium titanate ceramics. J. Mater. Sci. 2013, 24, 4105–4112. [Google Scholar] [CrossRef]
  27. Liu, B.; Wang, X.; Zhang, R.; Li, L. Grain size effect and microstructure influence on the energy storage properties of fine-grained BaTiO3-based ceramics. J. Am. Ceram. Soc. 2017, 100, 3599–3607. [Google Scholar] [CrossRef]
  28. Qiang, H.; Xu, Z. Effects of sintering temperature on the properties of Mn/Y codoped Ba0.67Sr0.33TiO3 ceramics for tunable application. J. Mater. Sci. 2015, 26, 9063–9066. [Google Scholar] [CrossRef]
  29. Redhu, P.; Punia, R.; Hooda, A.; Malik, B.P.; Sharma, G.; Sharma, P. Correlation between multifunctional properties of lead free Iron doped BCT perovskite ceramics. Ceram. Int. 2020, 46, 17495–17507. [Google Scholar] [CrossRef]
  30. Pavithra, C.; Madhuri, W.; Kiran, S.R. Effects of synthesis and sintering temperature in BCT-BST ceramics. Mater. Chem. Phys. 2021, 258. [Google Scholar] [CrossRef]
  31. Jin, L.; Luo, W.; Hou, L.; Tian, Y.; Hu, Q.; Wang, L.; Zhang, L.; Lu, X.; Du, H.; Wei, X.; et al. High electric field-induced strain with ultra-low hysteresis and giant electrostrictive coefficient in barium strontium titanate lead-free ferroelectrics. J. Eur. Ceram. Soc. 2019, 39, 295–304. [Google Scholar] [CrossRef]
  32. Wang, T.; Hu, J.; Yang, H.; Jin, L.; Wei, X.; Li, C.; Yan, F.; Lin, Y. Dielectric relaxation and Maxwell–Wagner interface polarization in Nb2O5 doped 0.65BiFeO(3)-0.35BaTiO(3) ceramics. J. Appl. Phys. 2017, 121, 084103. [Google Scholar] [CrossRef]
  33. Kang, F.; Zhang, L.; Huang, B.; Mao, P.; Wang, Z.; Sun, Q.; Wang, J.; Hu, D. Enhanced electromechanical properties of SrTiO3-BiFeO3-BaTiO3 ceramics via relaxor behavior and phase boundary design. J. Eur. Ceram. Soc. 2020, 40, 1198–1204. [Google Scholar] [CrossRef]
  34. Wang, J.; Zhang, X.; Zhang, J.; Li, H.; Li, Z. Dielectric and piezoelectric properties of (1−x)Ba0.7Sr0.3TiO3-xBa0.7Ca0.3TiO3 perovskites. J. Phys. Chem. Solids 2012, 73, 957–960. [Google Scholar] [CrossRef]
  35. Jin, L.; Li, F.; Zhang, S. Decoding the Fingerprint of Ferroelectric Loops: Comprehension of the Material Properties and Structures. J. Am. Ceram. Soc. 2014, 97, 1–27. [Google Scholar] [CrossRef]
  36. Prasertpalichat, S.; Khengkhatkan, S.; Siritanon, T.; Jutimoosik, J.; Kidkhunthod, P.; Bongkarn, T.; Patterson, E.A. Comparison of structural, ferroelectric, and piezoelectric properties between A-site and B-site acceptor doped 0.93Bi0.5Na0.5TiO3-0.07BaTiO3 lead-free piezoceramics. J. Eur. Ceram. Soc. 2021, 41, 4116–4128. [Google Scholar] [CrossRef]
  37. Li, T.; Liu, C.; Ke, X.; Liu, X.; He, L.; Shi, P.; Ren, X.; Wang, Y.; Lou, X. High electrostrictive strain in lead-free relaxors near the morphotropic phase boundary. Acta Mater. 2020, 182, 39–46. [Google Scholar] [CrossRef]
  38. Chen, K.; Ma, J.; Wu, J.; Wang, X.; Miao, F.; Huang, Y.; Shi, C.; Wu, W.; Wu, B. Improve piezoelectricity in BaTiO3-based ceramics with large electrostriction coefficient. J. Mater. Sci. 2020, 31, 12292–12300. [Google Scholar] [CrossRef]
  39. Li, F.; Jin, L.; Xu, Z.; Zhang, S. Electrostrictive effect in ferroelectrics: An alternative approach to improve piezoelectricity. Appl. Phys. Rev. 2014, 1, 11103. [Google Scholar] [CrossRef] [Green Version]
  40. Jang, S.J.; Uchino, K.; Nomura, S.; Cross, L.E. Electrostrictive behavior of lead maghesium niobate based ceramics dielectric. Ferroelectric 1980, 27, 31–34. [Google Scholar] [CrossRef]
  41. Li, J.; Wang, F.; Qin, X.; Xu, M.; Shi, W. Large electrostrictive strain in lead-free Bi0.5Na0.5TiO3-BaTiO3-KNbO3 ceramics. Appl. Phys. A 2010, 104, 117–122. [Google Scholar] [CrossRef]
  42. Bai, W.; Li, L.; Wang, W.; Shen, B.; Zhai, J. Phase diagram and electrostrictive effect in BNT-based ceramics. Solid State Commun. 2015, 206, 22–25. [Google Scholar] [CrossRef]
  43. Jin, L.; Luo, W.; Jing, R.; Qiao, J.; Pang, J.; Du, H.; Zhang, L.; Hu, Q.; Tian, Y.; Wei, X.; et al. High Dielectric Permittivity and Electrostrictive Strain in a Wide Temperature Range in Relaxor Ferroelectric (1−x)[Pb(Mg1/3Nb2/3)O3-PbTiO3]-xBa(Zn1/3Nb2/3)O3 Solid Solutions. Ceram. Int. 2019, 45, 5518–5524. [Google Scholar] [CrossRef]
Figure 1. The X-ray diffraction (XRD) patterns of BT-100xST (x = 0.10–0.60) ceramics in the 2θ range of (a) 10–80° and (b) 44.5–46.5°, (c) peak fitting in the 2θ range of 44.35–46.25°.
Figure 1. The X-ray diffraction (XRD) patterns of BT-100xST (x = 0.10–0.60) ceramics in the 2θ range of (a) 10–80° and (b) 44.5–46.5°, (c) peak fitting in the 2θ range of 44.35–46.25°.
Crystals 11 00555 g001aCrystals 11 00555 g001b
Figure 2. The scanning electron microscopy (SEM) images and average grain size (the inset corresponding to the SEM surface images) of BT-100xST ceramics, (a) x = 0.10, (b) x = 0.15, (c) x = 0.20, (d) x = 0.30, (e) x = 0.40, (f) x = 0.50.
Figure 2. The scanning electron microscopy (SEM) images and average grain size (the inset corresponding to the SEM surface images) of BT-100xST ceramics, (a) x = 0.10, (b) x = 0.15, (c) x = 0.20, (d) x = 0.30, (e) x = 0.40, (f) x = 0.50.
Crystals 11 00555 g002
Figure 3. Temperature dependence of permittivity and loss tangent for BT-100xST, (a) x = 0.10, (b) x = 0.15, (c) x = 0.20, (d) x = 0.30, (e) x = 0.40, (f) x = 0.50, (g) x = 0.60 and (h) the phase diagram of BT-100xST based on Tm, T1 and T2.
Figure 3. Temperature dependence of permittivity and loss tangent for BT-100xST, (a) x = 0.10, (b) x = 0.15, (c) x = 0.20, (d) x = 0.30, (e) x = 0.40, (f) x = 0.50, (g) x = 0.60 and (h) the phase diagram of BT-100xST based on Tm, T1 and T2.
Crystals 11 00555 g003aCrystals 11 00555 g003b
Figure 4. The frequency dependence of permittivity as a function of temperature for BT-100xST, (a) x = 0.10, (b) x = 0.15, (c) x = 0.20, (d) x = 0.30, (e) x = 0.40, (f) x = 0.50, (g) x = 0.60.
Figure 4. The frequency dependence of permittivity as a function of temperature for BT-100xST, (a) x = 0.10, (b) x = 0.15, (c) x = 0.20, (d) x = 0.30, (e) x = 0.40, (f) x = 0.50, (g) x = 0.60.
Crystals 11 00555 g004aCrystals 11 00555 g004b
Figure 5. Linear fit graph of diffuseness factor γ of BT-100xST ceramics, (a) x = 0.10, (b) x = 0.20, (c) x = 0.30, (d) x = 0.40, (e) x = 0.50, (f) x = 0.60.
Figure 5. Linear fit graph of diffuseness factor γ of BT-100xST ceramics, (a) x = 0.10, (b) x = 0.20, (c) x = 0.30, (d) x = 0.40, (e) x = 0.50, (f) x = 0.60.
Crystals 11 00555 g005aCrystals 11 00555 g005b
Figure 6. (a) P–E hysteresis loops of BT-100xST ceramics at room temperature, (b) the corresponding Pmax, Pr and Ec of BT-100xST ceramics as a function of Sr contents, (c) the corresponding I–E curves.
Figure 6. (a) P–E hysteresis loops of BT-100xST ceramics at room temperature, (b) the corresponding Pmax, Pr and Ec of BT-100xST ceramics as a function of Sr contents, (c) the corresponding I–E curves.
Crystals 11 00555 g006
Figure 7. (a) Unipolar strain curves, (b) bipolar strain curves of BT-100xST ceramics with different Sr contents, (c) line chart of unipolar Smax, Hys and Sneg.
Figure 7. (a) Unipolar strain curves, (b) bipolar strain curves of BT-100xST ceramics with different Sr contents, (c) line chart of unipolar Smax, Hys and Sneg.
Crystals 11 00555 g007
Figure 8. Bipolar S–P curves of BT-100xST ceramics with different Sr contents, (a) x = 0.10, (b) x = 0.15, (c) x = 0.20, (d) x = 0.30, (e) x = 0.40, (f) x = 0.50.
Figure 8. Bipolar S–P curves of BT-100xST ceramics with different Sr contents, (a) x = 0.10, (b) x = 0.15, (c) x = 0.20, (d) x = 0.30, (e) x = 0.40, (f) x = 0.50.
Crystals 11 00555 g008
Figure 9. Q33 and R as a function of BT-100xST ceramics with different Sr contents.
Figure 9. Q33 and R as a function of BT-100xST ceramics with different Sr contents.
Crystals 11 00555 g009
Table 1. Comparison of Q33, Smax, and Hys values of other BT-based, BNT-based and lead-based ceramics.
Table 1. Comparison of Q33, Smax, and Hys values of other BT-based, BNT-based and lead-based ceramics.
SystemSmax (%)Hys (%)Q33 (m4/C2)Ref.
Bi0.5(Na0.8K0.2)0.5TiO30.4460.0-[4]
0.99(Bi0.5Na0.4K0.1)TiO3-0.01(KxNa1−x)NbO30.4655.0-[5]
Bi0.5Na0.5TiO3-Bi0.5K0.5TiO3-BaZr0.05Ti0.95O30.12400.0237[9]
Bi4Ti3O12-Bi0.5(Na0.82K0.18)0.5TiO30.2923-[10]
Bi0.5(Na0.82K0.18)0.5TiO3-xSn0.45210.023[12]
Bi0.5Na0.5TiO3-BaTiO3-KNbO30.08-0.027[41]
Pb(In1/2Nb1/2)O3-Pb(Mg1/3Nb2/3)O3-PbTiO30.0950.030[14]
Bi(Li0.5Nb0.5)O3-Pb(Mg1/3Nb2/3)O3-PbTiO30.2270.018[15]
Pb(Mg1/3Nb2/3)O3-PbTiO3-Ba(Zn1/3Nb2/3)O30.17100.025[43]
BaZr0.2Ti0.8O3-Ba0.7Ca0.3TiO3-Ba0.7Sr0.3TiO30.1310-[19]
(Ba1−yCay)(Ti1−xHfx)O30.2112-[20]
(1−x)(Bi0.5Na0.5TiO3-0.11BaTiO3)-xBaZr0.2Ti0.8O30.27260.032[37]
BaTiO3-30SrTiO30.1120.034This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, M.; Sun, X.; Li, Q.; Qian, H.; Liu, Y.; Lyu, Y. Enhanced Electrostrictive Coefficient and Suppressive Hysteresis in Lead-Free Ba(1−x)SrxTiO3 Piezoelectric Ceramics with High Strain. Crystals 2021, 11, 555. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11050555

AMA Style

Song M, Sun X, Li Q, Qian H, Liu Y, Lyu Y. Enhanced Electrostrictive Coefficient and Suppressive Hysteresis in Lead-Free Ba(1−x)SrxTiO3 Piezoelectric Ceramics with High Strain. Crystals. 2021; 11(5):555. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11050555

Chicago/Turabian Style

Song, Mu, Xiaoyuan Sun, Qiong Li, Hao Qian, Yunfei Liu, and Yinong Lyu. 2021. "Enhanced Electrostrictive Coefficient and Suppressive Hysteresis in Lead-Free Ba(1−x)SrxTiO3 Piezoelectric Ceramics with High Strain" Crystals 11, no. 5: 555. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11050555

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop