Next Article in Journal
Enhancement Mode Ga2O3 Field Effect Transistor with Local Thinning Channel Layer
Previous Article in Journal
Annealing Engineering in the Growth of Perovskite Grains
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Different Ca2+ and Zr4+ Contents on Microstructure and Electrical Properties of (Ba,Ca)(Zr,Ti)O3 Lead-Free Piezoelectric Ceramics

1
School of Materials Science and Engineering, Yancheng Institute of Technology, Yancheng 224051, China
2
State Key Laboratory of Mechanics and Control of Mechanical Structures, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Crystals 2022, 12(7), 896; https://doi.org/10.3390/cryst12070896
Submission received: 29 May 2022 / Revised: 16 June 2022 / Accepted: 22 June 2022 / Published: 24 June 2022

Abstract

:
In the preparation of (Ba,Ca)(Zr,Ti)O3 lead-free piezoelectric ceramics, different Ca2+ and Zr4+ contents will greatly affect the phase structure, microstructure, and electrical properties of the ceramics. XRD shows that all samples have pure perovskite phase structure, and the (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics morphotropic phase boundary region from tetragonal phase to rhombohedral phase near 0.08 ≤ y ≤ 0.1. From the dielectric temperature curve, the phase transition temperature (TO-T) was found near room temperature at 0.12 ≤ x ≤ 0.18 for the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics. Both Ca2+ and Zr4+ increase have a significant decrease on the Curie temperature Tc. All samples were revealed as relaxers with diffusivities in the range 1.29 ≤ γ ≤ 1.82. Different from the undoped ceramics, ceramics doped with Ca and Zr ions exhibit saturated PE hysteresis loops, and their ferroelectric properties are significantly optimized. In particular, the (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 ceramic demonstrated optimal properties, namely d33 = 330 pC/N, kp = 0.41, εr = 4069, Pr = 4.8 μC/cm2, and Ec = 3.1 kV/cm, indicating that it is a viable lead-free piezoelectric contender. Variations in Ca and Zr content have a significant effect on the crystal grain sizes and densities of ceramics, which is strongly associated with their piezoelectricity.

1. Introduction

Lead-based piezoceramics are extensively utilized in various electronic devices, including actuators and sensors, owing to their high electrical activity, and their demand in the electronics sector has been steadily increasing in recent years [1,2]. However, as people are becoming increasingly aware of environmental challenges, many countries have issued restrictions on the use of substances involving lead, which limits the development of lead-based piezoelectric ceramics [3,4]. The development of environmentally friendly lead-free piezoelectric ceramics, such as (KxNa1−x)NbO3-based ceramics [5,6], (BixNa1−x)TiO3-based ceramics [7,8], BiFeO3-based ceramics [9,10], and BaTiO3 (BT)-based ceramics [11,12], is undoubtedly an important research topic for future piezoelectric applications.
Among the abovementioned ceramics, BT-based piezoelectric ceramics have been most extensively investigated. In 2009, Ren et al. [13] reported Ba(Zr0.2Ti0.8)O3x(Ba0.7Ca0.3)TiO3 (BCZT) piezoelectric ceramics with a d33 value up to 620 pC/N at optimal composition, which attracted considerable interest from the research community. Several efforts were devoted to further enhancing the d33 of BCZT ceramics, including the use of different synthesis methods [14], ion substitutions [15], and compositions [16]. Chen et al. [17] used the molten-salt synthesis method (MSS) to synthesize lead-free BCZT ceramics, indicating that MSS is a promising design for achieving improvements in the electrical properties of BCZT ceramics, as well as lowering the reaction temperature. Ji et al. [18] tailored the microstructure and properties of BCZT ceramics using the hydrothermal process, obtaining excellent dielectric properties with εm = 14,548. Bai et al. [19] studied a broad range of processing factors including composition, sintering conditions, particle size of the calcined ceramic powder, structure, and microstructure, with the objective of achieving optimum properties in BCZT ceramics. In addition to the fabrication and processing, the construction of phase boundaries also plays an essential role in enhanced electrical properties for lead-free ceramics. Regarding the origin of the large piezoelectricity of the system, Ren et al. [13] attributed the enhanced piezoelectricity of the BCZT pseudo-binary system to the approach of the morphotropic phase boundary (MPB) from the tricritical point of the rhombohedral (R), tetragonal (T), and cubic (C) phases. Recent research has discovered that the mixed phase region is an intermediate orthorhombic phase (O), resulting in excellent piezoelectric and electromechanical properties [20,21,22]. Zhou et al. [20] discovered the coexistence of the O and T phases in BCZT ceramics doped with 0.3 mol% LiTaO3 and obtained a high d33 value of 433 pC/N. Kaddoussi et al. [21] investigated the structural transitions in different x of Ca content for BCZT ceramic, revealing two types of phase-transition sequences: R–O–T–C occurring at x = 0.05 and O–T–C occurring at x = 0.2. Tian et al. [22] showed that the MPB of (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 ceramics was closely related to the occurrence of an intermediate phase (considered as O phase) at narrow regions between R and T phases. Moreover, the high piezoelectric of BCZT ceramics are the result of the microstructural instability and elastic softening by Guo et al. reported [23]. Cai et al. [24] attribute the enhancement of the piezoelectric properties of BCZT ceramics to the synergistic effect of grain size and phase boundaries. Hence, the underlying mechanisms for the phase diagram and piezoelectric performance enhancement of BCZT ceramic systems remain under-investigated.
The objective of this work is thus to understand the reasons for the effect of the Ca and Zr content variation in the (Ba,Ca)(Zr,Ti)O3 ceramics on their piezoelectric properties. The contribution of the phase structure, microstructure, dielectric, and ferroelectric properties to the piezoelectric properties of the (Ba,Ca)(Zr,Ti)O3 ceramics are explored.

2. Experimental

2.1. Sample Preparation

The (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics with x = 0, 0.06, 0.12, and 0.18, as well as (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics with y = 0, 0.08, 0.1, and 0.12, were designed for this study. The specimens were prepared via the traditional solid-state reaction method. The reagent-grade powders were used as the starting materials: BaCO3 (99%), CaCO3 (99%), ZrO2 (99%), TiO2 (99.9%). The powders were mixed, and agate balls were used to ball mill them in alcohol for 12 h. After desiccating the slurry, it was calcined for 4 h at 1180 °C in a closed crucible. Post-calcination, the powders were reground for 10 h. Polyvinyl alcohol (PVA) was used as a binder to compress the powders into disks at 200 MPa. To evaporate the PVA, all samples were sintered at 800 °C for 4 h, followed by 3 h in a covered crucible at 1400 °C. Silver electrodes were applied to the two surfaces of the polished ceramic disks before firing at 550 °C for 10 min. Polarization of the samples in silicone oil by imposition of a DC field of 2 kV/mm for 30 min at 50 °C.

2.2. Characterization

The samples were examined for the evolution of their crystalline structure via X-ray diffraction (XRD, D/Max-2500, Rigaku, Tokyo, Japan). The Archimedes method was used to identify the bulk density (ρ) of the samples (electronic density detector, ZMD-2, Shanghai Fangrui Instrument Limited Company, Shanghai, China). The microstructural characteristics of all the samples were observed using scanning electron microscopy (SEM, JMS-5610LV, Tokyo, Japan) after ultrasonic cleaning and vacuum drying of the sample surface, as well as spraying with gold. Piezoelectric constant (d33) was measured by a quasistatic piezoelectric d33 m (ZJ-3A, Institute of Acoustics, Chinese Academy of Sciences, China). The electromechanical coupling factor (kp) was determined through the resonance and antiresonance method using an LCR analyzer (HP4294A, Agilent, CA, USA) based on the IEEE standards [25]. The dielectric behavior and curie temperature (Tc) of the samples from room temperature to 160 °C were analyzed using the HP4294A impedance analyzer with an automatic temperature controller system. Measurement of polarization hysteresis loops (PE) at 10 Hz with a ferroelectric analysis tester (TF2000, aixACCT Systems GmbH, Aachen, Germany).

3. Results and Discussion

3.1. Crystal Phase and Structure

Figure 1a,b illustrates the XRD patterns of the (Ba1−xCax)(Zr0.1Ti0.9)O3 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics with different Ca and Zr contents. All samples exhibit a single perovskite structure, suggesting that Ca and Zr ions have spread into the BT lattice and formed a new solid solution. Figure 1c–f provides a magnified view of the XRD patterns around 2θ = 45°, and 56° for both samples, so that the peak positions and shapes can be easily compared. From Figure 1c, the 200 diffraction peak shifts toward a higher 2θ value with increasing Ca concentration for (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics. The same phenomenon is also observed for the 211 diffraction peaks. Conversely, increasing the Zr content causes the peak to shift toward a lower 2θ value for (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics. According to the Bragg equation 2dsinθ = , this is caused by the fact that Ca2+ has a smaller ionic radius (99 pm) than Ba2+ (135 pm), which results in a reduction in the lattice parameters. Equivalently, Zr4+ has a larger ionic radius (72 pm) than Ti4+ (60.5 pm), which leads to an increase in the lattice parameters. Figure 1e,f shows that the (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics present peaks splitting phenomenon, which corresponds to the tetragonal (T) phase structure at y = 0. With the increase in the Zr content (y ≤ 0.1), the distance between the (002) and (200) peaks and also between the (112) and (211) peaks become gradually smaller, indicating the co-existence of the tetragonal and rhombohedral (R-T) phases [22,26]. The estimated increase in the Zr content shows that the (200) and (211) peaks merge into a single peak, which corresponds to the rhombohedral (R) phase structure. Therefore, it can be assumed that a phase transformation from the T to the R with increasing Zr content at room temperature, where MPB between the R and T phases occurs in the region of 0.08 ≤ y ≤ 0.1.
Figure 2 illustrates the SEM morphology images of the (Ba1−xCax)(Zr0.1Ti0.9)O3 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics with different Ca and Zr contents. The ceramics show clear grain boundaries, a dense microstructure, and a low porosity. With the introduction of the Ca and Zr ions, the distribution of the grain sizes of the BCZT ceramics becomes more uniform compared with that of the pure BZT (x = 0) and BCT (y = 0) ceramics. In order to investigate the influences of different Ca and Zr contents on the grain size and density of ceramics, the linear intercept method and the Archimedes method, as illustrated in Figure 3, were used to estimate the average grain size and density, respectively. The addition of Ca to the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics results in the diminution of the grain size and an increase in the density, but a considerable reduction in the density is observed at x > 0.12. Furthermore, the addition of Zr increases the grain size and density of the (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics from 9.9 to 14.9 μm and from 5.61 to 5.69 g/cm3, respectively. The substitution of Zr assists in enhancing the grain growth of the materials [26]. Because the rhombohedral phase content progressively rises with increasing Zr content from Figure 1d, the change in microstructure with the introduction of Zr in this study may be explained by the phase transition from the T to the R for BCT ceramics. In addition, the greater radius of Zr4+ occupies the B-site of the ABO3 structure, which contributes to the steady rise in grain size.

3.2. Dielectric Properties

Figure 4 displays the dielectric behavior of (Ba1−xCax)(Zr0.1Ti0.9)O3 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics for various temperatures at 0.1, 1, 10, 100, and 1000 kHz. All of the samples show a distinct dielectric anomaly, which is attributed to the shift from the ferroelectric-to-paraelectric phase. This dielectric anomaly occurs at a temperature that corresponds to the Curie temperature (Tc) and reduces along with increasing Ca and Zr content. As presented in Table 1, the addition of the Ca and Zr ions reduces the Tc of the (Ba1−xCax)(Zr0.1Ti0.9)O3 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics from 87 °C to 73 °C and from 120 °C to 66 °C, respectively. The dielectric peaks shift toward a lower temperature, which may be due to the reduced internal stress in the structure [27]. The results demonstrate that the addition of Zr ions has a more pronounced impact on the Tc than that of Ca ions, which can be attributed to the reduction in the tetragonality shown in Figure 1d. As the relatively active Ti ions are replaced by the relatively passive Zr ions, the cell parameter ratio decreases, the polarization becomes extremely irregular, and the overall domain structure is disrupted, causing the Tc to shift toward a lower temperature [28].
In particular, an extra dielectric anomaly is observed in Figure 4b,c corresponding to the temperature of the O–T phase transition (TO–T) of (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics. The dielectric peak of TO–T is not observed in Figure 4a,d because TO–T is close to TC and covered by εm due to the diffuse phase transition behavior at x = 0 [22], and TO–T shifts to below room temperature at x = 0.18. The frequency dispersion is negligible for all samples, demonstrating their excellent frequency stability.
Furthermore, Figure 4 shows that the morphologies of the curves at the Curie peak exhibit a broad trend with adding Ca and Zr ions. The Curie-–Weiss law is suitable for describing the dielectric behavior of relaxer ferroelectrics, and the parameter γ can be used to estimate the degree of diffuseness by plotting ln(1/εr−1/εm) as a function of ln(TTm) at a frequency of 1 kHz. The dielectric characteristics can be expressed by the following modified Curie-Weiss relationship:
1 ε r 1 ε m = ( T T m ) γ C ( 1 γ 2 )
where εr is the dielectric constant at temperature T, εm is the maximum dielectric constant at the transition temperature (Tm), C is the Curie constant, and γ is the indicator of the degree of diffuseness, which takes a value between 1 (for a normal ferroelectric) and 2 (for a completely diffuse phase transition). As shown in Figure 5, the addition of the Ca and Zr ions increases the γ values of the samples from 1.39 to 1.80 and from 1.29 to 1.82, respectively, thus suggesting that the as-prepared (Ba1−xCax)(Zr0.1Ti0.9)O3 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics are typical relaxer ferroelectrics. The addition of the Ca and Zr ions drastically affects the degree of the dielectric relaxation of the BCZT piezoelectric ceramics.

3.3. Piezoelectric Properties

Figure 6 plots the piezoelectric coefficient d33 and electromechanical coupling coefficient kp for the (Ba1−xCax)(Zr0.1Ti0.9)O3 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics. Figure 6a shows that the d33 value incrementally increases with Ca content until a peak value of 300 pC/N at x = 0.12, and declines subsequently with further increase in x. The kp value’s fluctuating pattern with x is comparable to the d33 value, and the highest value of kp at x = 0.12 being 0.44. The enhanced piezoelectric properties are closely related to the gradual approach to room temperature of the TO–T dielectric peak observed in Figure 4a–d. The polarization state becomes unstable due to the composition-induced O–T phase transition, permitting the polarization direction to simply be turned by an external stress or electric field, leading to the strong piezoelectricity.
Regarding the (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics, the change in the piezoelectric properties shown in Figure 6b is consistent with that of the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics, with maximum d33 and kp values of 330 pC/N and 0.41 obtained at y = 0.1, respectively. The improvement of the piezoelectric properties can be attributed to the coexistence of the T and R phases at y = 0.1, as revealed by the XRD patterns shown in Figure 1d. Figure 6c,d presents the contour plots of the d33 and kp values for the (Ba1−xCax)(ZryTi1−y)O3 ceramics with 0 ≤ x ≤ 0.18 and 0 ≤ y ≤ 0.12. In particular, the (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 ceramic has optimal piezoelectric properties. Tian et al. [22] also reported the best properties at this composition, although both ceramics were prepared by the conventional solid-phase method, but differences in the pre-firing, sintering, and polarization processes can cause the piezoelectric properties to be affected [19]. In our work, the (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 ceramic exhibits high d33 and kp values, which can be attributed to the existence of the O–T and R–T MPBs at room temperature and well-formulated microstructure.

3.4. Ferroelectric Properties

Figure 7 displays polarization–electrical field (PE) hysteresis loops of the (Ba1−xCax)(Zr0.1Ti0.9)O3 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics with different Ca and Zr contents. Figure 7a displays that the BZT (x = 0) ceramic exhibits no ferroelectric hysteresis loop. However, ferroelectric properties appear after the addition of the Ca ions, and the best performance is achieved at x = 0.12, namely Pr = 6.3 μC/cm2 and Ec = 2.3 kV/cm. The room-temperature Pr and Ec values are summarized in Table 1. The enhancement of the ferroelectric properties may be explained by the fact that the Ca substitution diffuses the phase transition of BZT and causes the dielectric peak of TO–T to shift toward room temperature as the Ca content increases. As shown in Figure 7b, the PE loops of the (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics show a small ferroelectric contribution with a remnant polarization 2Pr < 10 μC/cm2 at y = 0. However, as the Zr content increases, the loops become well-saturated and Pr increases, approaching 5 μC/cm2 for all ceramics at 0.08 ≤ y ≤ 0.12. Additionally, Ec decreases gradually, reaching values of 4.6, 3.1, and 1.1 kV/cm at y = 0.08, 0.1, and 0.12, respectively, indicating an easy polarization mechanism of the material. In summary, both Ca and Zr elements have a significant effect on the ferroelectric properties of the BCZT ceramics.

4. Conclusions

In summary, the influences of Ca and Zr contents on the phase transition and electrical performances of (Ba1−xCax)(Zr0.1Ti0.9)O3 (x = 0, 0.06, 0.12, and 0.18) ceramics and (Ba0.85Ca0.15)(ZryTi1−y)O3 (y = 0, 0.08, 0.1, and 0.12) ceramics have been systematically researched. The phase of (Ba1−xCax)(Zr0.1Ti0.9)O3 transforms from orthogonal to tetragonal near room temperature at 0.12 ≤ x ≤ 0.18 via dielectric temperature curve. For (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics, MPB region from tetragonal phase to rhombohedral phase at composition 0.08 ≤ y ≤ 0.1 via XRD patterns. Moreover, the doped ceramics exhibited saturated PE hysteresis loops and significantly optimized ferroelectric properties compared with the undoped ceramics. Results indicate that optimal electrical properties of d33 = 330 pC/N, kp = 0.41, εr = 4069, Pr = 4.8 μC/cm2, and Ec = 3.1 kV/cm are achieved for the (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 ceramic. Our study demonstrates the contribution of the Ca and Zr additions to the piezoelectric properties of BCZT ceramics in terms of the phase transition, microstructure, dielectric properties, and ferroelectricity, demonstrating a potential value of these lead-free ceramics.

Author Contributions

Conceptualization, J.D.; methodology, K.Z.; formal analysis, J.D., L.Q. and C.Y.; resources, K.Z. and L.W.; writing—original draft preparation, L.Q. and C.Y.; writing—review and editing, L.Q., Y.C. and J.D.; visualization, J.D., L.Q. and Y.C.; supervision, L.W. and K.Z.; project administration, J.D. and Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (2019YFC1907103-04), Funding for school-level research projects of Yancheng Institute of Technology (No. xj201528, xjr2021016), General Program of Natural Science Foundation of the Higher Education Institutions of Jiangsu Province (No. 16KJB430030) and Postgraduate Research & Practice Innovation Program of Jiangsu Province (SJCX22_XZ001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. He, Y.X.; Yang, Q.; Luo, M.D.; Liu, R.Y. Effect of the physical parameters of longitudinally polarized PZT tubes on PZT sensors. J. Phys. D Appl. Phys. 2020, 53, 275501. [Google Scholar] [CrossRef]
  2. Sabri, M.F.M.; Ono, T.; Said, S.M.; Kawai, Y.; Esashi, M. Fabrication and Characterization of Microstacked PZT Actuator for MEMS Applications. J. Microelectromech. Syst. 2015, 24, 80–90. [Google Scholar] [CrossRef]
  3. Liu, W.F.; Cheng, L.; Li, S.T. Prospective of (BaCa)(ZrTi)O3 Lead-free Piezoelectric Ceramics. Crystals 2019, 9, 179. [Google Scholar] [CrossRef] [Green Version]
  4. Nguyen, T.N.; Thong, H.C.; Zhu, Z.X.; Nie, J.K.; Liu, Y.X.; Xu, Z.; Soon, P.S.; Gong, W.; Wang, K. Hardening effect in lead-free piezoelectric ceramics. J. Mater. Res. 2021, 36, 996–1014. [Google Scholar] [CrossRef]
  5. Li, J.F.; Wang, K.; Zhu, F.Y.; Cheng, L.Q.; Yao, F.Z. (K, Na) NbO3-Based Lead-Free Piezoceramics: Fundamental Aspects, Processing Technologies, and Remaining Challenges. J. Am. Ceram. Soc. 2013, 96, 3677–3696. [Google Scholar] [CrossRef]
  6. Shalini, K.; Giridharan, N.V. Structural, dielectric and magnetic properties of K0.5Na0.5NbO3 and K0.5 Na0.5Nb0.975Co0.025O3 lead free ceramics. Ferroelectrics 2017, 518, 52–58. [Google Scholar] [CrossRef]
  7. Dong, N.N.; Gao, X.Y.; Xia, F.Q.; Liu, H.X.; Hao, H.; Zhang, S.J. Dielectric and Piezoelectric Properties of Textured Lead-Free Na0.5Bi0.5TiO3-Based Ceramics. Crystals 2019, 9, 206. [Google Scholar] [CrossRef] [Green Version]
  8. Trelcat, J.F.; d’Astorg, S.; Courtois, C.; Champagne, P.; Rguiti, M.; Leriche, A. Influence of hydrothermal synthesis conditions on BNT-based piezoceramics. J. Eur. Ceram. Soc. 2011, 31, 1997–2004. [Google Scholar] [CrossRef]
  9. Vila-Fungueirino, J.M.; Gomez, A.; Antoja-Lleonart, J.; Gazquez, J.; Magen, C.; Noheda, B.; Carretero-Genevrier, A. Direct and converse piezoelectric responses at the nanoscale from epitaxial BiFeO3 thin films grown by polymer assisted deposition. Nanoscale 2018, 10, 20155–20161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Bai, L.Q.; Li, X.L.; Wang, S.; Liu, H.J.; Shi, L.; Luo, Q.Y.; Song, W.L.; Chen, D. Fabrication and Photoelectrochemical Performances of BiFeO3 Thin Film Photoelectrodes: Effect of Annealing Temperature. Int. J. Electrochem. Sci. 2021, 16, 210721. [Google Scholar] [CrossRef]
  11. Zhu, K.J.; Qiu, J.H.; Kajiyoshi, K.; Takai, M.; Yanagisawa, K. Effect of washing of barium titanate powders synthesized by hydrothermal method on their sinterability and piezoelectric properties. Ceram. Int. 2009, 35, 1947–1951. [Google Scholar] [CrossRef] [Green Version]
  12. Liu, Y.C.; Chang, Y.F.; Li, F.; Yang, B.; Sun, Y.; Wu, J.; Zhang, S.T.; Wang, R.X.; Cao, W.W. Exceptionally High Piezoelectric Coefficient and Low Strain Hysteresis in Grain-Oriented (Ba, Ca)(Ti, Zr)O3 through Integrating Crystallographic Texture and Domain Engineering. ACS Appl. Mater. Interfaces 2017, 9, 29863–29871. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, W.F.; Ren, X.B. Large Piezoelectric Effect in Pb-Free Ceramics. Phys. Rev. Lett. 2009, 103, 257602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Zhao, Z.H.; Li, X.L.; Dai, Y.J.; Ji, H.M.; Su, D. Highly textured Ba0.85Ca0.15Ti0.90Zr0.10O3 ceramics prepared by reactive template grain growth process. Mater. Lett. 2016, 165, 131–134. [Google Scholar] [CrossRef]
  15. Sun, M.Z.; Du, J.; Chen, C.; Fu, P.; Li, P.; Hao, J.G.; Yue, Z.X.; Li, W. Enhanced piezoelectric properties in M (M = Co or Zn)-doped Ba0.99Ca0.01Ti0.98Zr0.02O3 ceramics. Ceram. Int. 2020, 46, 17351–17360. [Google Scholar] [CrossRef]
  16. Nie, X.; Yan, S.G.; Guo, S.B.; Cao, F.; Yao, C.H.; Mao, C.L.; Dong, X.L.; Wang, G.S. Influence of Ca2+ concentration on structure and electrical properties of (Ba1-xCax)(Zr0.2Ti0.8)O3 ceramics. Mater. Res. Express 2018, 5, 036301. [Google Scholar] [CrossRef]
  17. Chen, X.F.; Peng, Z.H.; Chao, X.L.; Yang, Z.P. Structure, electrical properties and reaction mechanism of (Ba0.85Ca0.15) (Zr0.1Ti0.9)O3 ceramics synthesized by the molten salt method. Ceram. Int. 2017, 43, 11920–11928. [Google Scholar] [CrossRef]
  18. Ji, W.W.; Fang, B.J.; Lu, X.Y.; Zhang, S.; Yuan, N.Y.; Ding, J.N. Tailoring structure and performance of BCZT ceramics prepared via hydrothermal method. Phys. B 2019, 567, 65–78. [Google Scholar] [CrossRef]
  19. Bai, Y.; Matousek, A.; Tofel, P.; Bijalwan, V.; Nan, B.; Hughes, H.; Button, T.W. (Ba,Ca)(Zr,Ti)O3 lead-free piezoelectric ceramics—The critical role of processing on properties. J. Eur. Ceram. Soc. 2015, 35, 3445–3456. [Google Scholar] [CrossRef]
  20. Zhou, M.X.; Liang, R.H.; Zhou, Z.Y.; Xu, C.H.; Nie, X.; Dong, X.L. Enhanced Curie temperature and piezoelectric properties of (Ba0.85Ca0.15) (Zr0.10Ti0.90)O3 lead-free ceramics after the addition of LiTaO3. Mater. Res. Bull. 2018, 106, 213–219. [Google Scholar] [CrossRef]
  21. Kaddoussi, H.; Lahmar, A.; Gagou, Y.; Manoun, B.; Chotard, J.N.; Dellis, J.L.; Kutnjak, Z.; Khemakhem, H.; Elouadi, B.; El Marssi, M. Sequence of structural transitions and electrocaloric properties in (Ba1-xCax)(Zr0.1Ti0.9)O3 ceramics. J. Alloy. Compd. 2017, 713, 164–179. [Google Scholar] [CrossRef]
  22. Tian, Y.; Wei, L.L.; Chao, X.L.; Liu, Z.H.; Yang, Z.P. Phase Transition Behavior and Large Piezoelectricity Near the Morphotropic Phase Boundary of Lead-Free (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 Ceramics. J. Am. Ceram. Soc. 2013, 96, 496–502. [Google Scholar] [CrossRef]
  23. Guo, H.Z.; Voas, B.K.; Zhang, S.J.; Zhou, C.; Ren, X.B.; Beckman, S.P.; Tan, X.L. Polarization alignment, phase transition, and piezoelectricity development in polycrystalline 0.5Ba(Zr0.2Ti0.8)O3-0.5(Ba0.7Ca0.3)TiO3. Phys. Rev. B 2014, 90, 014103. [Google Scholar] [CrossRef] [Green Version]
  24. Cai, W.; Zhang, Q.W.; Zhou, C.; Gao, R.L.; Zhang, S.L.; Li, Z.D.; Xu, R.C.; Chen, G.; Deng, X.L.; Wang, Z.H.; et al. Synergistic effect of grain size and phase boundary on energy storage performance and electric properties of BCZT ceramics. J. Mater. Sci. Mater. Electron. 2020, 31, 9167–9175. [Google Scholar] [CrossRef]
  25. ANSI/IEEE Std 176-1987; IEEE Standard on Piezoelectricity. IEEE: Piscataway, NJ, USA, 1988.
  26. Bai, W.F.; Chen, D.Q.; Zhang, J.J.; Zhong, J.S.; Ding, M.Y.; Shen, B.; Zhai, J.W.; Ji, Z.G. Phase transition behavior and enhanced electromechanical properties in (Ba0.85Ca0.15)(ZrxTi1-x)O3 lead-free piezoceramics. Ceram. Int. 2016, 42, 3598–3608. [Google Scholar] [CrossRef]
  27. Tian, Y.S.; Li, S.Y.; Gong, Y.S.; Yu, Y.S.; Tang, Y.T.; Liu, P.; Jing, Q.S. Diversified electrical properties of (1-x)Ba0.90Ca0.10Ti0.95Zr0.05O 3-xRuO2 ceramics with defect electron complexes. Mater. Chem. Phys. 2018, 204, 163–170. [Google Scholar] [CrossRef]
  28. Wu, J.G.; Xiao, D.Q.; Wu, W.J.; Chen, Q.; Zhu, J.G.; Yang, Z.C.; Wang, J. Composition and poling condition-induced electrical behavior of (Ba0.85Ca0.15)(Ti1-xZrx)O3 lead-free piezoelectric ceramics. J. Eur. Ceram. Soc. 2012, 32, 891–898. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the (a) (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics. Magnified view of the XRD patterns of the (c,d) BCxZT ceramics and (e,f) BCZyT ceramics around 2θ = 45°, and 56°, respectively.
Figure 1. XRD patterns of the (a) (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics. Magnified view of the XRD patterns of the (c,d) BCxZT ceramics and (e,f) BCZyT ceramics around 2θ = 45°, and 56°, respectively.
Crystals 12 00896 g001aCrystals 12 00896 g001b
Figure 2. SEM images of the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics for Ca content at (a) x = 0, (b) x = 0.06, (c) x = 0.12, and (d) x = 0.18, and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics for Zr content at (e) y = 0, (f) y = 0.08, (g) y = 0.1, and (h) y = 0.12.
Figure 2. SEM images of the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics for Ca content at (a) x = 0, (b) x = 0.06, (c) x = 0.12, and (d) x = 0.18, and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics for Zr content at (e) y = 0, (f) y = 0.08, (g) y = 0.1, and (h) y = 0.12.
Crystals 12 00896 g002
Figure 3. Average grain size and density of (a) (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics.
Figure 3. Average grain size and density of (a) (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics.
Crystals 12 00896 g003
Figure 4. The temperature and frequency dependence of εr and tanδ of the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics with (a) x = 0, (b) x = 0.08, (c) x = 0.1, (d) x = 0.12 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics with (e) y = 0, (f) y = 0.08, (g) y = 0.1, (h) y = 0.12.
Figure 4. The temperature and frequency dependence of εr and tanδ of the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics with (a) x = 0, (b) x = 0.08, (c) x = 0.1, (d) x = 0.12 and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics with (e) y = 0, (f) y = 0.08, (g) y = 0.1, (h) y = 0.12.
Crystals 12 00896 g004
Figure 5. Diffusivity of dielectric behavior about ln(1/εr−1/εm) function versus ln (TTm) for (a) (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics at 1 kHz.
Figure 5. Diffusivity of dielectric behavior about ln(1/εr−1/εm) function versus ln (TTm) for (a) (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics at 1 kHz.
Crystals 12 00896 g005
Figure 6. The d33 and kp of (a) (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics at room temperature. Contour maps of (c) d33 and (d) kp of the (Ba1−xCax)(ZryTi1−y)O3 ceramics with 0 ≤ x ≤ 0.18 and 0 ≤ y ≤ 0.12.
Figure 6. The d33 and kp of (a) (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics at room temperature. Contour maps of (c) d33 and (d) kp of the (Ba1−xCax)(ZryTi1−y)O3 ceramics with 0 ≤ x ≤ 0.18 and 0 ≤ y ≤ 0.12.
Crystals 12 00896 g006
Figure 7. PE loops of (a) the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics.
Figure 7. PE loops of (a) the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (b) (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics.
Crystals 12 00896 g007
Table 1. Room-temperature physical properties of the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics.
Table 1. Room-temperature physical properties of the (Ba1−xCax)(Zr0.1Ti0.9)O3 ceramics and (Ba0.85Ca0.15)(ZryTi1−y)O3 ceramics.
Compositionx = 0x = 0.06x = 0.12x = 0.18y = 0y = 0.08y = 0.1y = 0.12
d33 (pC/N)70102300231122302330211
kp0.160.200.440.310.240.430.410.27
εr2398174426323928928294240693832
tanδ9.8%3.9%1.1%1.2%9.7%1.1%1.5%1.6%
Tc (°C)87.186.678.573.1119.39180.566.05
Pr (μC/cm2)-4.56.33.63.55.74.85.1
Ec (kV/cm)-0.92.33.13.34.63.11.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Du, J.; Qiu, L.; Yang, C.; Chen, Y.; Zhu, K.; Wang, L. Effect of Different Ca2+ and Zr4+ Contents on Microstructure and Electrical Properties of (Ba,Ca)(Zr,Ti)O3 Lead-Free Piezoelectric Ceramics. Crystals 2022, 12, 896. https://doi.org/10.3390/cryst12070896

AMA Style

Du J, Qiu L, Yang C, Chen Y, Zhu K, Wang L. Effect of Different Ca2+ and Zr4+ Contents on Microstructure and Electrical Properties of (Ba,Ca)(Zr,Ti)O3 Lead-Free Piezoelectric Ceramics. Crystals. 2022; 12(7):896. https://doi.org/10.3390/cryst12070896

Chicago/Turabian Style

Du, Jianzhou, Long Qiu, Cong Yang, Yuansheng Chen, Kongjun Zhu, and Luming Wang. 2022. "Effect of Different Ca2+ and Zr4+ Contents on Microstructure and Electrical Properties of (Ba,Ca)(Zr,Ti)O3 Lead-Free Piezoelectric Ceramics" Crystals 12, no. 7: 896. https://doi.org/10.3390/cryst12070896

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