Next Article in Journal
Hybrid Metal-Organic Frameworks/Carbon Fibers Reinforcements for Additively Manufactured Composites
Next Article in Special Issue
Humidity Sensing Properties of (In+Nb) Doped HfO2 Ceramics
Previous Article in Journal
Synthesis of Microwave Functionalized, Nanostructured Polylactic Co-Glycolic Acid (nfPLGA) for Incorporation into Hydrophobic Dexamethasone to Enhance Dissolution
Previous Article in Special Issue
All-Water-Driven High-k HfO2 Gate Dielectrics and Applications in Thin Film Transistors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Piezoelectricity and Thermal Stability of Electrostrain Performance in BiFeO3-Based Lead-Free Ceramics

1
Information Materials and Intelligent Sensing Laboratory of Anhui Province, Anhui University, Hefei 230601, China
2
Key Laboratory of Structure and Functional Regulation of Hybrid Materials of Ministry of Education, Anhui University, Hefei 230601, China
3
Institutes of Physical Science and Information Technology, Anhui University, Hefei 230601, China
4
Guangdong Provincial Key Laboratory of Electronic Functional Materials and Devices, Huizhou University, Huizhou 516001, China
5
Laboratory of Dielectric Functional Materials, School of Materials Science & Engineering, Anhui University, Hefei 230601, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(5), 942; https://doi.org/10.3390/nano13050942
Submission received: 2 February 2023 / Revised: 24 February 2023 / Accepted: 3 March 2023 / Published: 5 March 2023
(This article belongs to the Special Issue Nanoelectronics: Materials, Devices and Applications)

Abstract

:
BiFeO3–based ceramics possess an advantage over large spontaneous polarization and high Curie temperature, and are thus widely explored in the field of high–temperature lead–free piezoelectrics and actuators. However, poor piezoelectricity/resistivity and thermal stability of electrostrain make them less competitive. To address this problem, (1 − x) (0.65BiFeO3–0.35BaTiO3)–xLa0.5Na0.5TiO3 (BF–BT–xLNT) systems are designed in this work. It is found that piezoelectricity is significantly improved with LNT addition, which is contributed by the phase boundary effect of rhombohedral and pseudocubic phase coexistence. The small–signal and large–signal piezoelectric coefficient (d33 and d 33 * ) peaks at x = 0.02 with 97 pC/N and 303 pm/V, respectively. The relaxor property and resistivity are enhanced as well. This is verified by Rietveld refinement, dielectric/impedance spectroscopy and piezoelectric force microscopy (PFM) technique. Interestingly, a good thermal stability of electrostrain is obtained at x = 0.04 composition with fluctuation η = 31% ( S max ' S RT S RT × 100 % ), in a wide temperature range of 25–180 °C, which is considered as a compromise of negative temperature dependent electrostrain for relaxors and the positive one for ferroelectric matrix. This work provides an implication for designing high–temperature piezoelectrics and stable electrostrain materials.

1. Introduction

As a mechanical–electricity conversion functional material, piezoelectric ceramics have wide applications in the defense, industrial, and medical fields, etc. [1,2,3]. Due to extraordinary piezoelectric properties and electromechanical coupling effects near morphotropic phase boundary (MPB), Pb(Zr, Ti)O3–based (PZT) ceramics have been a hot research topic [4,5]. However, with an increasingly serious environmental concern, lead–based materials are gradually replaced by lead–free materials. A series of lead–free materials such as (Bi0.5Na0.5)TiO3 (BNT–) [6,7,8], BiFeO3 (BF–) [9,10,11], (K0.5Na0.5)NbO3 (KNN–) based [12,13] ceramics have been widely developed. For example, Liu et al. reported a non-textured BNT–based ceramic with a large electrostrain value of ~0.7% at room temperature (RT), but it exhibited relatively large strain hysteresis and a decrease in electrostrain as temperature increased [14]; Due to polymorphic phase transition effect in KNN–based ceramics, the macroscopic performance usually exhibited a temperature sensitivity, although compositionally graded multilayer composite and layered distribution of dopants strategy largely alleviated this shortcoming [15,16]. In contrast, BiFeO3–BaTiO3–based (BF–BT) ceramics show a positive temperature dependence of electrostrain as their high depolarization temperature (Td) [17]. According to first–principles calculations, high spontaneous polarization and electrostrain are facilitated by the inverse rotation of oxygen octahedron and bismuth and oxygen hybridization as a result of the huge displacement of Bi3+ and Fe3+ ions [18,19]. However, a high leakage current and oxygen vacancies ( V o ¨ ) concentration is detrimental to electrical performance [20]. Zhao et al. effectively reduced V o ¨ by annealing ceramics in an oxygen atmosphere [21]. In addition, electrostrain of BF–BT–based ceramics at high temperature are frequently reported. For example, Zheng et al. reported that the unipolar strain increased from 0.10% at RT to 0.32% at 200 °C for Bi(Mg2/3Nb1/3)O3 modified BF–BT ceramics [22]. Similarly, an electrostrain of Sm–doped BF–BT ceramics increased from 0.28% at RT to 0.52% at 125 °C [23]. Although a large electrostrain was obtained at an elevated temperature, the temperature-sensitive electrostrain performance is unfavorable for practical applications.
Similar to the thermal fluctuation effect, structural and charge disorder in relaxor ferroelectrics usually produce a negative temperature dependence of electrostrain [24,25]. Also, an increased concentration of nanodomains in relaxors reduces the domain wall energy to favor ferroelectric domains switching, which can contribute to a significant increase in electrostrain [26,27]. Therefore, to obtain electrostrain with high–temperature stability, we combine the negative temperature stability of relaxors and positive temperature stability of BF–BT ferroelectric matrix. Also, an end-member La0.5Na0.5TiO3 is reported to not only increase the relaxor degree, but also reduce dielectric loss [28,29]. Interestingly, phase and domain structures of BF–BT ceramics are significantly modulated by adding La0.5Na0.5TiO3 in this work, and high piezoelectricity/resistivity and thermal stability of electrostrain are obtained. The underlying mechanism is comprehensively analyzed by Rietveld refinement, Raman, dielectric/impedance spectroscopy, and PFM technique.

2. Materials and Methods

We obtained (1 − x) (0.65BiFeO3–0.35BaTiO3)–xLa0.5Na0.5TiO3 (BF–BT–xLNT, x = 0~0.06) ceramics by conventional solid–state synthesis using Bi2O3, Fe2O3, TiO2, BaCO3, La2O3, and Na2CO3. Excessive 2%mol Bi2O3 was added for compensation. All raw materials were mixed with alcohol and ball-milled for 15 h and then calcined at 750 °C for 6 h twice. Before the second ball-milling step, 1% mol MnO2 was added to reduce the leakage current. Then, powders were mixed with 8 wt.% PVA binder and pressed into the discs with a diameter of 10 mm. After removing the binder, the discs were sintered at 1040 °C for 3 h. For electrical measurements, a silver paste was coated on both sides of the polished samples and fired at 560 °C for 10 min.
The crystal structure was measured by X-ray diffractometer (XRD, Rigaku Smart–lab). A scanning electron microscope (SEM, Regulus 8230; Hitachi Co., Tokyo, Japan) was used to observe the sample microstructures. Before observing the microstructure, all sample surfaces are polished smooth and hot corroded at 950 °C for 30 min. A Raman microscope was used to acquire Raman spectra (Horiba Jobin–Yvon HR800, France). Dielectric/impedance properties were acquired by Wayne Kerr 6500B impedance analyzer (Wayne Kerr Electronic Instrument Co., Shenzhen, China). Polarization hysteresis (P–E) loops and electrostrain (S–E) curves were measured using a ferroelectric measuring system (Precision LC, Radiant Technologies, Inc. Albuquerque, NM, USA) at a frequency of 1 Hz. The sample was poled for 15 min under an electric field (E = 60 kV/cm) at 120 °C, and then piezoelectric coefficients d33 were measured with a quasi-static d33 m (YE2730A, China). The domain structure was characterized by the PFM technique (Asylum Research).

3. Results and Discussion

Room temperature (RT) XRD patterns of BF–BT–xLNT ceramics (x = 0~0.06) are shown in Figure 1a–g, respectively. All samples have a pure perovskite structure without a second phase, indicating that LNT is completely dissolved into BF–BT matrix and forms a solid solution. To get an in-depth understanding of phase structure and content evolution with changing LNT, Rietveld refinement is performed and R3cH (R phase) and Pm 3 ¯ m (Pc phase) space group models are exploited [30,31]. Low fitted values of Rwp, Rp, and χ2 indicate that fitting results are reliable. Also, locally magnified (111) diffraction peaks are displayed in their insets. Obviously, a wide (111) peak for the x = 0 sample indicates the existence of the R phase. The (111) peak gradually becomes narrow and sharp with an increase in LNT content, indicating that the R phase is gradually substituted by a Pc phase, and finally evolves into a single Pc phase at x = 0.05, as displayed in Figure 1h. The refined lattice parameter is plotted in Figure 1i. Table 1 also shows the lattice parameters and R–factors obtained by Rietveld refinement for better understanding. Obviously, with LNT addition, the lattice parameter generally exhibits a downward shift, which is mainly attributed to smaller ionic radii of Na+ and La3+ (CN = 12, RNa+ = 1.39 Å, RLa3+ = 1.36 Å) than that of Ba2+ and Bi3+ (CN = 12, RBa2+ = 1.61 Å, RBi3+ = 1.45 Å) [32].
SEM images of selected compositions (x = 0, 0.02, 0.06) are displayed in Figure 2a–c. Clearly, all ceramic grains are uniformly distributed, and there are no obvious pores, indicating a dense ceramic microstructure. The apparent density gradually decreases with the addition of LNT, which is determined by a smaller molecular mass of LNT as compared to BF. The relative density in all compositions surpasses ~95%, which verifies their high compactness.
Raman spectra technique is a powerful tool to detect phase transition at short-range scales, and Raman spectra for BF–BT–xLNT ceramics are performed, as shown in Figure 3a. Generally, the wavenumber range in 50~1000 cm−1 is divided into three vibrational modes, namely A–site (50~200 cm−1), B–O bond (200~400 cm−1), and BO6 octahedral vibration (400~1000 cm−1) [33,34]. To clearly demonstrate vibrational mode changes, raw Raman spectra are fitted by the Lorentzian function, and a series of deconvoluted Raman peaks are delineated, as shown in Figure 3b. Here, two representative peaks (G and H bands) are selected to analyze the wavenumber and FWHM (full width at half maximum) evolution to detect phase transition, as shown in Figure 3c,d. Notably, two discontinuous changes in wavenumber and FWHM are observed (as highlighted by shadow), which strongly suggests that phase transition occurs [35]. Since the Raman shift is interrelated to crystal stress and polarization, the first abrupt change of G and H bands probably corresponds to distorted local stress and the polarization field [36]. With an increase of heterovalent ionic proportion (Na+ and La3+) and corresponding local random field, the second vibration change may be related to the ferroelectric-to-relaxor (FR) phase transition, which will be discussed infra in detail.
Temperature dependence of dielectric constant (εr) of BF–BT–xLNT ceramics (x = 0~0.06) at 1 kHz~500 kHz are shown in Figure 4a–g, respectively. For x = 0 compositions, it shows a relaxor-like behavior near 300 °C, which is related to a dipolar relaxation caused by V o ¨ hopping [37]. The dielectric drift at high temperatures is caused by a large conductivity due to an increased V o ¨ motion, as highlighted in Figure 4a [38]. As x increases to 0.01 and 0.02, the dielectric peak becomes sharp, and εr at Tm (the temperature for dielectric maxima) also increases (Figure 4h). This is also observed from an obvious hump at the imaginary part of the dielectric constant (ε′′) curve, as shown in the insets of Figure 4b,c. Notably, the high–temperature dielectric drift is significantly suppressed, indicating the resistivity is markedly improved with LNT addition, which is also observed from the dielectric loss (tanδ) in Figure 4i. As to x = 0.03~0.06, diffusive phase transition and frequency dispersion are clearly observed, exhibiting a relaxation property [39,40]. The dielectric curves for x = 0~0.06 samples at 100 kHz are collected in Figure 4h. The peak εr value peaks at x = 0.01 and 0.02 and then sharply decreases for further increasing LNT content, accompanied by a wide and diffuse dielectric shape. This is more clearly indicated by εr/εm versus T/Tm curves in Figure 4j. Therefore, an addition of an LNT component first improves the dielectric performance with an appropriate proportion of R and Pc phase. With an excessive LNT content, εr is significantly suppressed with a dominant relaxor Pc phase. In order to quantitatively describe relaxor properties for this system, two parameters of ΔTrelax [Tm(100 kHz) − Tm(1 kHz)] and ΔTspan (temperature span corresponds to εr/εm = 0.8 in Figure 4j) are adopted [41]. Obviously, ΔTrelax and ΔTspan exhibit an increasing trend as LNT content increases, as plotted in Figure 4k. Also, a modified Curie–Weiss law is used to denote the relaxor degree γ: 1 ε r 1 ε m = ( T T m ) γ C   ( 1 γ 2 ) , where C is Curie–Weiss constant, εm is dielectric maxima and the parameter γ is adopted to reveal the relaxor degree [42,43]. As shown in Figure 4l, γ value steadily increases from 1.73 for x = 0 to 1.91 for x = 0.06 composition. Therefore, an enhanced relaxor properties are obtained with LNT addition.
To determine the resistivity and conduction mechanism for BF–BT–xLNT ceramics, complex impedance spectra (CIS) are measured over 280–380 °C with an interval of 20 °C, as shown in Figure 5a–g. All ceramics exhibit a single semicircle complex impedance shape, which is related to a single relaxation mechanism of bulk response [44]. The data are well fitted by two parallel R–CPE equivalent circuits, as shown in the inset of Figure 5e, which represent the grain boundary and bulk (grain) contributions, respectively. As known, bulk resistance (Rb) often dominates ferroelectric ceramics, and an extrapolated intercept on Z axis corresponds to Rb in CIS. Obviously, Rb gradually decreases with increasing temperature, indicating a negative temperature coefficient behavior. The Rb value is obtained by fitting the experimental CIS, as shown in Figure 5a–g. For a better understanding, the Rb values are also listed in Table 2. Interestingly, the LNT addition greatly increases the Rb of the BF–BT matrix (Figure 5h,i), which is consistent with a depressed tanδ. In addition, bulk conductivity (σb) and activation energy (Ea) are calculated by the following formulas [45]:
σ b = h S × R b
σ b = σ 0 exp ( E a k B T )
where h is sample thickness, S is electrode area, kB is Boltzmann constant, and T is the measured temperature. Figure 5i depicts the σb as a function of inverse of temperature, and the fitted value of Ea is obtained between 0.93 and 1.24 eV, which is close to the Ea of V o ¨ (~1 eV) [46]. Therefore, V o ¨ dominates the high–temperature conductivity and leads to a large leakage current, which is mainly due to the volatilization of bismuth and the reduction of Fe3+ [47].
Polarization hysteresis (P–E) and corresponding current density (J–E) loops of BF–BT–xLNT ceramics at RT are measured at 1 Hz, as shown in Figure 6a,b. Low saturated/remanent polarization (Pm/Pr) with an incomplete domain switching is observed for x = 0 composition. Interestingly, for x = 0.01 and 0.02 samples, saturated PE loops and a sharp JE peak are observed, featuring a normal ferroelectric. A sharp JE peak usually indicates strong ferroelectricity with a fast domain switching behavior under E [26]. As to x = 0.03 and 0.04, splitting JE peaks are observed, indicating an emergence of an intermediate state with nonergodic/ergodic and ferroelectric phase coexistence. Meanwhile, Pm, Pr, and coercive field (Ec) decrease strikingly, as plotted in Figure 6c. Slant P–E loops and J–E platform with low Pr and Ec value indicates a pure relaxor phase for x = 0.05 and 0.06 samples, which also accords with pure Pc phase for both compositions. In addition, bipolar and unipolar electrostrain (SE) curves of BF–BT–xLNT ceramics are shown in Figure 6d,e, respectively. For x ≤ 0.03 samples, typical butterfly-shape bipolar SE curves are present, and they gradually evolve into a sprout–shape [decrease in negative strain (Sneg)] with increasing LNT content. In general, larger Sneg indicates more non-180° domain switching under E, and this predicts improvement of piezoelectric performance. For x = 0.04~0.06 samples, an almost zero–Sneg indicates a gradual FR phase transition. Clearly, positive strain (Spos) and Sneg peaks x = 0.02 sample, and the normalized d 33 * ( S max / E max ) is calculated as well, as plotted in Figure 6f. The d33 and d 33 * value simultaneously achieve optimal value (97 pC/N and 303 pm/V) at x = 0.02, which is contributed by phase boundary effect with proper proportion of R and Pc phase content. Excessive addition of LNT degrades piezoelectricity and ferroelectricity and thus d 33 * decreases.
To measure the temperature stability of piezoelectric performance for these compositions, d33 is collected at elevated annealing temperature, as shown in Figure 7a, which is considered a crucial benchmark for evaluating practical high–temperature performances. Apparently, x = 0.02 sample not only possesses a peak d33 value but also maintains a wide temperature span (≤240 °C). Also, frequency–dependent εr for poled x = 0.02 sample is also measured after different annealing temperatures, as shown in Figure 7b. Typical resonance and anti–resonance peaks are observed for sufficiently poled samples. As the temperature increases, the peak gradually degrades, indicating depolarization steadily occurs [48]. These peaks dampen gradually and finally vanish as the temperature increased to~240 °C. This is also indicated by the contour plot, as indicated in Figure 7c. To investigate the depolarization mechanism on the local structure of representative x = 0.02 and 0.03 ceramics, in situ Raman spectra is performed, as shown in Figure 7d–g. All Raman spectra are corrected by equation of I c ( ω ) = I m ( ω ) / [ n ( ω , T ) + 1 ] , where Im(ω) is Raman intensity and n ( ω , T ) = 1 / [ exp ( ω k T 1 ) ]   is Bose–Einstein temperature factor [49]. The evolution of wavenumber and FWHM of selected D and G modes are plotted in Figure 7e,f,h,i. The corresponding discontinuities are highlighted by the shadow. The discontinuity occurs around Td~240 °C and 140 °C for x = 0.02 and 0.03 samples, respectively, which indicates an FR phase transition and also accords with the Td [50].
Temperature dependent unipolar SE curves of BF–BT–xLNT ceramics are shown in Figure 8a–g. For x ≤ 0.02 compositions with normal ferroelectric character, they exhibit a positive temperature dependence of electrostrain, which mainly originates from thermally activated domain switching [30,51]. For x = 0.03 composition, the electrostrain shows a positive temperature dependence below Td~140 °C, and finally the electrostrain performance remains stable. For x = 0.04 sample with less concentration of ferroelectric domains (R/Pc = 0.25/0.75), the electrostrain shows better temperature stability, which is probably due to a compromise of thermally activated domain switching and agitation. As LNT content further increases, the temperature stability of electrostrain is maintained but strain value decreases. Here, the change rate of electrostrain with temperature is defined as η = S max ' S RT S RT × 100 % , where S max ' is unipolar electrostrain peak value in the test temperature range of 25–180 °C, S RT is the electrostrain at RT. Evolution of unipolar SE with temperature for x = 0~0.06 compositions are summarized in Figure 8h. Obviously, the x = 0.04 component exhibits a large electrostrain and strong temperature stability (E = 60 kV/cm). Within 25–180 °C, x = 0.04 sample presents a smallest η value of 31%, indicating an excellent temperature stability of electrostrain performance. A comparison of electrostrain performance with some representative lead–free piezoelectrics are present in Figure 8i. The electrostrain value fluctuates more than 100% for some BF–based ceramics, showing poor temperature stability [22,52,53]. Also, the electrostrain change exhibits a parabolic–like or monotonous decrease for BNT– and KNN–based ceramics, depending on phase transition temperature [54,55]. The enhanced temperature stability of electrostrain performance for x = 0.04 composition is due to the synergistic effect of negative temperature dependent electrostrain for relaxors with Pc phase and the positive one for ferroelectric matrix with R phase, which strongly suggest the special proportion of relaxor and ferroelectric phase can produce thermally stable electrostrain performance and thus is highly desirable for high–temperature of actuator applications.
It is well known that evolution in macroscopic performance is accompanied by changes in domain structure, and PFM images can convey domain structure information at a local scale. Figure 9a–c shows the PFM amplitude images for x = 0, 0.03, and 0.06 samples, respectively. The highlighted area in the amplitude image represents the piezoelectric response strength [56,57]. Consistent with the piezoelectric performance, x = 0 and 0.03 samples exhibit large bright areas, whereas x = 0.06 has only few bright areas, as indicated by the blue dashed box. Also, as shown in phase images of Figure 9d–f, broad strip–like long–range ordered macrodomains are clearly observed in the x = 0 and 0.03, and thin strip–shaped short–range ordered nanodomains are observed in the dim area of x = 0.03 and 0.06. This indicates that both large–sized macrodomains and small–sized nanodomains coexist in the ferroelectric and nonergodic phase composition (x = 0.03 as an example). For relaxor phase in x = 0.06, small–sized nanodomains occupy almost the entire area. Therefore, large–sized macrodomains with nanodomains blending have a positive effect on the piezoelectric effect, whereas the composition with increased concentration of nanodomains are beneficial for temperature–stable electrostrain performance.

4. Conclusions

BF–BT–xLNT (x = 0~0.06) ceramics are obtained via conventional solid-state synthesis. Rietveld refinement, Raman spectroscopy, and dielectric analysis show the system undergoes from R–to–Pc phase transition that is driven by LNT addition. Dielectric and impedance spectra show that the addition of LNT reduces the V o ¨ concentration and greatly increases the resistivity of BF–BT ceramics. At the same time, the piezoelectric performance is optimized at x = 0.02 composition (d33 = 97 pC/N), which is contributed by phase boundary effect and enhanced resistivity for efficiently poling. Notably, temperature stability of electrostrain is obtained for x = 0.04 composition, which is due to the synergistic effect of negative temperature dependent electrostrain for relaxor nanodomains with Pc phase and the positive one for ferroelectric bulk domains with R phase. This is certified by PFM images that nanodomains emerges from ferroelectric matrix with increasing doping content. It can be seen that the addition of LNT not only improves the resistivity and piezoelectric properties of BF–BT ceramics, but also enhances the temperature stability of electrostrain. Therefore, this work has implications for the design and application of high–performance piezoelectrics and temperature–stable electrostrain materials.

Author Contributions

Conceptualization, F.L. and C.W.; Data curation, H.S., K.L., J.M. and Y.L.; Investigation, H.S.; Methodology, F.L. and C.W.; Resources, F.L., C.W. and L.S.; Software, Y.T. and M.L.; Supervision, F.L., Y.T. and C.W.; Validation, Y.T.; Writing—original draft, H.S.; Writing—review and editing, M.L., W.G. and L.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (Nos. 12104001, 12174001, and 51872001), Open Project Program of Guangdong Provincial Key Laboratory of Electronic Functional Materials and Devices, Huizhou University (No. EFMD2022003Z).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jo, W.; Dittmer, R.; Acosta, M.; Zang, J.; Groh, C.; Sapper, E.; Wang, K.; Rödel, J. Giant electric–field–induced strains in lead–free ceramics for actuator applications–status and perspective. J. Electroceram. 2012, 29, 71–93. [Google Scholar] [CrossRef]
  2. Saito, Y.; Takao, H.; Tani, T.; Nonoyama, T.; Takatori, K.; Homma, T.; Nagaya, T.; Nakamura, M. Lead–free piezoceramics. Nature 2004, 432, 84–87. [Google Scholar] [CrossRef]
  3. Hao, J.G.; Li, W.; Zhai, J.W.; Chen, H. Progress in high–strain perovskite piezoelectric ceramics. Mater. Sci. Eng. R 2019, 135, 1–57. [Google Scholar] [CrossRef]
  4. Chen, C.; Wang, Y.; Li, Z.Y.; Liu, C.; Gong, W.; Tan, Q.; Han, B.; Yao, F.Z.; Wang, K. Evolution of electromechanical properties in Fe–doped (Pb, Sr)(Zr, Ti)O3 piezoceramics. J. Adv. Ceram. 2021, 10, 587–595. [Google Scholar] [CrossRef]
  5. Li, Z.; Thong, H.C.; Zhang, Y.F.; Xu, Z.; Zhou, Z.; Liu, Y.X.; Cheng, Y.Y.S.; Wang, S.H.; Zhao, C.L.; Chen, F.; et al. Defect engineering in lead zirconate titanate ferroelectric ceramic for enhanced electromechanical transducer efficiency. Adv. Funct. Mater. 2021, 31, 2005012. [Google Scholar] [CrossRef]
  6. Yin, J.; Liu, G.; Zhao, C.L.; Huang, Y.L.; Li, Z.; Zhang, X.M.; Wang, K.; Wu, J.G. Perovskite Na0.5Bi0.5TiO3: A potential family of peculiar lead–free electrostrictors. J. Mater. Chem. A 2019, 7, 13658. [Google Scholar] [CrossRef]
  7. Zhou, X.F.; Xue, G.L.; Luo, H.; Bowen, C.R.; Zhang, D. Phase structure and properties of sodium bismuth titanate lead–free piezoelectric ceramics. Prog. Mater. Sci. 2021, 122, 100836. [Google Scholar] [CrossRef]
  8. Sun, M.Z.; Li, P.; Du, J.; Han, W.F.; Hao, J.G.; Zhao, K.Y.; Zeng, H.R.; Li, W. Electric field–induced photoluminescence quenching in Pr–doped BNT ceramics across the MPB region. J. Mater. 2022, 8, 288–294. [Google Scholar] [CrossRef]
  9. Tang, Y.C.; Yin, Y.; Song, A.Z.; Liu, H.; Zhang, R.; Zhong, S.J.; Li, H.Z.; Zhang, B.P. Boosting the high performance of BiFeO3–BaTiO3 lead–free piezoelectric ceramics: One–step preparation and reaction mechanisms. ACS Appl. Mater. Interfaces 2022, 14, 30991–30999. [Google Scholar] [CrossRef]
  10. Peng, X.Y.; Tang, Y.C.; Zhang, B.P.; Zhu, L.F.; Xun, B.W.; Yu, J.R. High Curie temperature BiFeO3–BaTiO3 lead–free piezoelectric ceramics: Ga3+ doping and enhanced insulation properties. J. Appl. Phys. 2021, 130, 144104. [Google Scholar] [CrossRef]
  11. Ryu, G.H.; Hussain, A.; Lee, M.H.; Malik, R.A.; Song, T.K.; Kim, W.J.; Kim, M.H. Lead–free high performance Bi(Zn0.5Ti0.5)O3–modified BiFeO3–BaTiO3 piezoceramics. J. Eur. Ceram. Soc. 2018, 38, 4414–4421. [Google Scholar] [CrossRef]
  12. 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]
  13. Xu, K.; Li, J.; Lv, X.; Wu, J.G.; Zhang, X.X.; Xiao, D.Q.; Zhu, J.G. Superior piezoelectric properties in potassium–sodium niobate lead–free ceramics. Adv. Mater. 2016, 28, 8519–8523. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, X.M.; Tan, X.L. Giant strains in non–textured (Bi1/2Na1/2)TiO3–based lead–free ceramics. Adv. Mater. 2016, 28, 574–578. [Google Scholar] [CrossRef] [Green Version]
  15. Zheng, T.; Yu, Y.G.; Lei, H.B.; Li, F.; Zhang, S.J.; Zhu, J.G.; Wu, J.G. Compositionally graded KNN–based multilayer composite with excellent piezoelectric temperature stability. Adv. Mater. 2022, 34, 2109175. [Google Scholar] [CrossRef] [PubMed]
  16. Song, A.Z.; Liu, Y.X.; Feng, T.Y.; Li, H.T.; Zhang, Y.Y.; Wang, X.P.; Liu, L.S.; Zhang, B.P.; Li, J.F. Simultaneous enhancement of piezoelectricity and temperature stability in KNN–based lead–free ceramics via layered distribution of dopants. Adv. Funct. Mater. 2022, 34, 2204385. [Google Scholar] [CrossRef]
  17. Liu, X.; Zhai, J.W.; Shen, B. Novel bismuth ferrite–based lead–free incipient piezoceramics with high electromechanical response. J. Mater. Chem. C 2019, 7, 5122–5130. [Google Scholar] [CrossRef]
  18. Lebeugle, D.; Colson, D.; Forget, A.; Viret, M.; Bonville, P.; Marucco, J.F.; Fusil, S. Room–temperature coexistence of large electric polarization and magnetic order in BiFeO3 single crystals. Phys. Rev. B 2007, 76, 024116. [Google Scholar] [CrossRef] [Green Version]
  19. Neaton, J.B.; Ederer, C.; Waghmare, U.V.; Spaldin, N.A.; Rabe, K.M. First–principles study of spontaneous polarization in multiferroic BiFeO3. Phys. Rev. B 2015, 71, 014113. [Google Scholar] [CrossRef] [Green Version]
  20. Xun, B.W.; Song, A.Z.; Yu, J.R.; Yin, Y.; Li, J.F.; Zhang, B.P. Lead–free BiFeO3–BaTiO3 ceramics with high curie temperature: Fine compositional tuning across the phase boundary for high piezoelectric charge and strain coefficients. ACS Appl. Mater. Interfaces 2021, 13, 4192–4202. [Google Scholar] [CrossRef]
  21. Zhao, H.Y.; Hou, Y.D.; Yu, X.L.; Zheng, M.P.; Zhu, M.K. Construction of high Tc BiScO3–BiFeO3–PbTiO3 and its enhanced piezoelectric properties by sintering in oxygen atmosphere. J. Appl. Phys. 2018, 124, 194103. [Google Scholar] [CrossRef]
  22. Zheng, T.; Zhao, C.L.; Wu, J.G.; Wang, K.; Li, J.F. Large strain of lead–free bismuth ferrite ternary ceramics at elevated temperature. Scr. Mater. 2018, 155, 11–15. [Google Scholar] [CrossRef]
  23. Habib, M.; Akram, F.; Ahmad, P.; Kebaili, I.; Rahman, A.; Din, I.U.; Iqbal, M.J.; Zeb, A.; Khandaker, M.U.; Karoui, A.; et al. Ultrahigh piezoelectric strain in lead–free BiFeO3–BaTiO3 ceramics at elevated temperature. J. Alloys Compd. 2022, 919, 165744. [Google Scholar] [CrossRef]
  24. Li, F.; Jin, L.; Xu, Z.; Zhang, S.J. Electrostrictive effect in ferroelectrics: An alternative approach to improve piezoelectricity. Appl. Phys. Rev. 2014, 1, 011103. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, L.; Liang, R.H.; Zhou, Z.Y.; Dong, X.L. Thermally stable electrostrain in BiFeO3–BaTiO3–based high temperature lead-free piezoceramics. Appl. Phys. Lett. 2019, 115, 082902. [Google Scholar] [CrossRef]
  26. Li, F.; Wu, S.H.; Li, T.Y.; Wang, C.C.; Zhai, J.W. Normal–relaxor ferroelectric phase transition induced morphotropic phase boundary accompanied by enhanced piezoelectric and electrostrain properties in strontium modulated Bi0.5K0.5TiO3 lead–free ceramics. J. Eur. Ceram. Soc. 2020, 40, 3918–3927. [Google Scholar] [CrossRef]
  27. Shi, H.W.; Li, F.; Liu, W.; Liang, C.; Ji, X.L.; Long, M.S.; Gong, W.P.; Wang, C.C.; Shan, L. Composition dependent phase structure, dielectric and electrostrain properties in (Sr0.7Bi0.2□0.1)TiO3–PbTiO3–Bi(Mg0.5Ti0.5)O3 systems. J. Phys. D Appl. Phys. 2022, 55, 185301. [Google Scholar] [CrossRef]
  28. Komine, S.; Iguchi, E. Phase transitions and the weak relaxor ferroelectric phase in Ba7/8(La0.5Na0.5)1/8TiO3. J. Phys. Condens. Matter 2002, 14, 8445. [Google Scholar] [CrossRef]
  29. Wu, Y.; Bian, J.J.; Zhao, C.; Luo, S.J. Improvement of dielectric loss of Ba0.75Sr0.25TiO3 tunable material by La0.5Na0.5TiO3 addition. J. Mater. Sci. Mater. Electron. 2015, 26, 90–97. [Google Scholar] [CrossRef]
  30. Murakami, S.K.; Wang, D.W.; Mostaed, A.; Khesro, A.; Feteira, A.; Sinclair, D.C.; Fan, Z.M.; Tan, X.L.; Reaney, I.M. High strain 0.4%Bi(Mg2/3Nb1/3)O3–BaTiO3–BiFeO3 lead–free piezoelectric ceramics and multilayers. J. Am. Ceram. Soc. 2018, 101, 5428–5442. [Google Scholar] [CrossRef]
  31. Wang, D.W.; Fan, Z.M.; Li, W.B.; Zhou, D.; Feteira, A.; Wang, G.; Murakami, S.; Sun, S.; Zhao, Q.L.; Tan, X.L.; et al. High energy storage density and large strain in Bi(Zn2/3Nb1/3)O3–doped BiFeO3–BaTiO3 ceramics. ACS Appl. Energ. Mater. 2018, 1, 4403–4412. [Google Scholar] [CrossRef]
  32. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. 1976, 32, 751–767. [Google Scholar]
  33. Guo, J.J.; Zhu, M.K.; Li, L.; Qing, T.H.; Wang, C.; Liu, L.Y.; Zheng, M.P.; Hou, Y.D. Normal–relaxor ferroelectric modulation of A–site complex perovskite ferroelectric (K1/2Bi1/2)TiO3 by post–annealing. J. Appl. Phys. 2017, 121, 014101. [Google Scholar]
  34. Aksel, E.; Forrester, J.S.; Kowalski, B.; Deluca, M.; Damjanovic, D.; Jones, J.L. Structure and properties of Fe–modified Na0.5Bi0.5TiO3 at ambient and elevated temperature. Phys. Rev. B 2012, 85, 024121. [Google Scholar] [CrossRef] [Green Version]
  35. Li, F.; Zhai, J.W.; Shen, B.; Liu, X.; Yang, K.; Zhang, Y.; Li, P.; Liu, B.H.; Zeng, H.R. Influence of structural evolution on energy storage properties in Bi0.5Na0.5TiO3–SrTiO3–NaNbO3 lead–free ferroelectric ceramics. J. Appl. Phys. 2017, 121, 054103. [Google Scholar] [CrossRef]
  36. Wang, L.; Liang, R.H.; Zhou, Z.Y.; Dong, X.L. High electrostrain with high Curie temperature in BiFeO3–BaTiO3–based ceramics. Scr. Mater. 2019, 164, 62–65. [Google Scholar] [CrossRef]
  37. Wang, C.C.; Lei, C.M.; Wang, G.J.; Sun, X.H.; Li, T.; Huang, S.G.; Wang, H.; Li, Y.D. Oxygen–vacancy–related dielectric relaxations in SrTiO3 at high temperatures. J. Appl. Phys. 2013, 113, 094103. [Google Scholar] [CrossRef]
  38. Liu, W.; Li, F.; Long, M.S.; Zhai, J.W.; Wang, C.C.; Shan, L. Defect engineering boosts ferroelectricity in Pb0.9Ba0.1ZrO3 ceramic. J. Eur. Ceram. Soc. 2022, 42, 6295–6300. [Google Scholar] [CrossRef]
  39. Viehland, D.; Jang, S.J.; Cross, L.E.; Wuttig, M. Freezing of the polarization fluctuations in lead magnesium niobate relaxors. J. Appl. Phys. 1990, 68, 2916–2921. [Google Scholar] [CrossRef]
  40. Cross, L.E. Relaxor ferroelectrics: An overview. Ferroelectrics 1994, 151, 305–320. [Google Scholar] [CrossRef]
  41. Chen, W.; Yao, X.; Wei, X.Y. Tunability and ferroelectric relaxor properties of bismuth strontium titanate ceramics. Appl. Phys. Lett. 2007, 90, 182902. [Google Scholar] [CrossRef] [Green Version]
  42. Uchino, K.; Nomura, S. Critical exponents of the dielectric constants in diffused–phase–transition crystals. Ferroelectrics 1982, 44, 55–61. [Google Scholar] [CrossRef]
  43. Yu, Z.Z.; Wang, Z.G.; Shu, S.W.; Tian, T.F.; Huang, W.B.; Li, C.C.; Ke, S.M.; Shu, L.L. Local structural heterogeneity induced large flexoelectricity in Sm–doped PMN–PT ceramics. J. Appl. Phys. 2021, 129, 174103. [Google Scholar] [CrossRef]
  44. Li, F.; Zhou, M.X.; Zhai, J.W.; Shen, B.; Zeng, H.R. Novel barium titanate based ferroelectric relaxor ceramics with superior charge–discharge performance. J. Eur. Ceram. Soc. 2018, 14, 4646–4652. [Google Scholar] [CrossRef]
  45. Chen, P.Y.; Chou, C.C.; Tseng, T.Y.; Chen, H. Correlation of microstructures and conductivities of ferroelectric ceramics using complex impedance spectroscopy. Jpn. J. Appl. Phys. 2010, 49, 061505. [Google Scholar] [CrossRef] [Green Version]
  46. Tong, B.B.; Lin, J.Y.; Lin, C.H.; Chen, J.G.; Zou, X.L.; Cheng, J.R. Enhanced transduction coefficient and thermal stability of 0.75BiFeO3–0.25BaTiO3 ceramics for high temperature piezoelectric energy harvesters applications. Ceram. Int. 2022, 48, 16885–16891. [Google Scholar] [CrossRef]
  47. Li, Y.; Zheng, T.; Li, B.; Wu, J.G. Relaxation degree and defect dipoles-controlled resistivity and leakage current in BF–BT-based ceramics. J. Am. Ceram. Soc. 2023, 106, 2393–2406. [Google Scholar] [CrossRef]
  48. Li, G.H.; Shi, C.; Zhu, K.; Ge, G.L.; Yan, F.; Lin, J.F.; Shi, Y.J.; Shen, B.; Zhai, J.W. Achieving synergistic electromechanical and electrocaloric responses by local structural evolution in lead–free BNT–based relaxor ferroelectrics. Chem. Eng. J. 2022, 431, 133386. [Google Scholar] [CrossRef]
  49. Yoon, S.; Liu, H.L.; Schollerer, G.; Cooper, S.L.; Han, P.D.; Payne, D.A.; Cheong, S.W.; Fisk, Z. Raman and optical spectroscopic studies of small–to–large polaron crossover in the perovskite manganese oxides. Phys. Rev. B 1998, 58, 5. [Google Scholar] [CrossRef]
  50. Ji, X.L.; Li, F.; Long, M.S.; Wang, C.C.; Shan, L. Excellent temperature stability of energy storage performance by weak dipolar interaction strategy. Appl. Phys. Lett. 2022, 121, 023902. [Google Scholar] [CrossRef]
  51. Zhu, L.F.; Liu, Q.; Zhang, B.P.; Cen, Z.Y.; Wang, K.; Li, J.J.; Bai, Y.; Wang, X.H.; Li, J.F. Temperature independence of piezoelectric properties for high–performance BiFeO3–BaTiO3 lead–free piezoelectric ceramics up to 300 °C. RSC Adv. 2018, 8, 35794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Wei, J.X.; Fu, D.Y.; Cheng, J.R.; Chen, J.G. Temperature dependence of the dielectric and piezoelectric properties of xBiFeO3–(1–x)BaTiO3 ceramics near the morphotropic phase boundary. J. Mater. Sci. 2017, 52, 10726–10737. [Google Scholar] [CrossRef]
  53. Wang, D.W.; Khesro, A.; Murakami, S.; Feteira, A.; Zhao, Q.L.; Reaney, I.M. Temperature dependent, large electromechanical strain in Nd–doped BiFeO3–BaTiO3 lead–free ceramics. J. Eur. Ceram. Soc. 2017, 37, 1857–1860. [Google Scholar] [CrossRef] [Green Version]
  54. Zhang, S.T.; Kounga, A.B.; Aulbach, E.; Jo, W.; Granzow, T.; Ehrenberg, H.; Rödel, J. Lead–free piezoceramics with giant strain in the system Bi0.5Na0.5TiO3–BaTiO3–K0.5Na0.5NbO3. II. Temperature dependent properties. J. Appl. Phys. 2008, 103, 034108. [Google Scholar] [CrossRef] [Green Version]
  55. Wang, R.P.; Wang, K.; Yao, F.Z.; Li, J.F.; Schader, F.H.; Webber, K.G.; Jo, W.; Rödel, J. Temperature stability of lead–free niobate piezoceramics with engineered morphotropic phase boundary. J. Am. Ceram. Soc. 2015, 98, 2177–2182. [Google Scholar] [CrossRef]
  56. Jiang, X.J.; Dietz, C.; Liu, N.; Rojas, V.; Stark, R.W. Ferroelectric domain evolution in a Ba(Zr0.2Ti0.8)O3–0.5(Ba0.7Ca0.3)TiO3 piezoceramic studied using piezoresponse force microscopy. Appl. Phys. Lett. 2021, 118, 262902. [Google Scholar] [CrossRef]
  57. Zeng, F.F.; Zhou, C.; Zhang, C.; Jiang, L.; Zhang, J.J.; Guo, H.T.; Chen, Y.X.; Lv, W.Z.; Cai, W.; Zhang, G.Z.; et al. Effects of Hf4+ substitute on the enhanced electrostrain properties of 0.7BiFeO3–0.3BaTiO3–based lead–free piezoelectric ceramics. Ceram. Int. 2022, 48, 10539–10546. [Google Scholar] [CrossRef]
Figure 1. (ag) Rietveld refinement patterns of BF–BT–xLNT ceramics (x = 0~0.06); (h) R3cH (R) and Pm 3 ¯ m (Pc) phase fraction evolution and (i) lattice parameter of R and Pc phase as a function of LNT content.
Figure 1. (ag) Rietveld refinement patterns of BF–BT–xLNT ceramics (x = 0~0.06); (h) R3cH (R) and Pm 3 ¯ m (Pc) phase fraction evolution and (i) lattice parameter of R and Pc phase as a function of LNT content.
Nanomaterials 13 00942 g001
Figure 2. (ac) SEM images of selected BF–BT–xLNT ceramics (x = 0, 0.02, 0.06); (d) apparent and relative density as a function of LNT content.
Figure 2. (ac) SEM images of selected BF–BT–xLNT ceramics (x = 0, 0.02, 0.06); (d) apparent and relative density as a function of LNT content.
Nanomaterials 13 00942 g002
Figure 3. (a) Raman spectra of BF–BT–xLNT ceramics (x = 0~0.06); (b) experimental data, peak fitting curve, and the deconvolution of Raman spectra (selecting x = 0 composition as an example); (c,d) composition dependence of wavenumber and FWHM for selected G and H bands.
Figure 3. (a) Raman spectra of BF–BT–xLNT ceramics (x = 0~0.06); (b) experimental data, peak fitting curve, and the deconvolution of Raman spectra (selecting x = 0 composition as an example); (c,d) composition dependence of wavenumber and FWHM for selected G and H bands.
Nanomaterials 13 00942 g003
Figure 4. (ag) Temperature dependence of εr of fresh BF–BT–xLNT (x = 0~0.06) ceramics at 1 kHz ~500 kHz, the inset shows the ε′′ at 100 kHz; (h,i) collection of εr and tanδ curves; (j) summary of εr/εm versus T/Tm; (k) ΔTrelax, ΔTspan and (l) variation of γ at 100 kHz as a function of LNT content.
Figure 4. (ag) Temperature dependence of εr of fresh BF–BT–xLNT (x = 0~0.06) ceramics at 1 kHz ~500 kHz, the inset shows the ε′′ at 100 kHz; (h,i) collection of εr and tanδ curves; (j) summary of εr/εm versus T/Tm; (k) ΔTrelax, ΔTspan and (l) variation of γ at 100 kHz as a function of LNT content.
Nanomaterials 13 00942 g004
Figure 5. (ag) Complex impedance plots for BF–BT–xLNT (x = 0~0.06) ceramics in the temperature range of 280–380 °C with an interval of 20 °C, the inset of (e) exhibits the fitting electric circuit; (h) complex impedance spectra for x = 0~0.06 samples at 280 °C; (i) lnσ with inverse of temperature (1000/T), the inset shows the variation of Ea and Rb as a function of LNT content.
Figure 5. (ag) Complex impedance plots for BF–BT–xLNT (x = 0~0.06) ceramics in the temperature range of 280–380 °C with an interval of 20 °C, the inset of (e) exhibits the fitting electric circuit; (h) complex impedance spectra for x = 0~0.06 samples at 280 °C; (i) lnσ with inverse of temperature (1000/T), the inset shows the variation of Ea and Rb as a function of LNT content.
Nanomaterials 13 00942 g005
Figure 6. (a) RT PE loops for BF–BT–xLNT (x = 0~0.06) ceramics at 1 Hz, the inset shows JE loops; (b) the contour map of JE value; (c) evolution of Pm, Pr and Ec as a function of LNT content; (d) bipolar and (e) unipolar SE curves; (f) Spos, Sneg, d33 *, and d33 as a function of LNT content.
Figure 6. (a) RT PE loops for BF–BT–xLNT (x = 0~0.06) ceramics at 1 Hz, the inset shows JE loops; (b) the contour map of JE value; (c) evolution of Pm, Pr and Ec as a function of LNT content; (d) bipolar and (e) unipolar SE curves; (f) Spos, Sneg, d33 *, and d33 as a function of LNT content.
Nanomaterials 13 00942 g006
Figure 7. (a) Variation of d33 with thermal annealing at various temperatures for BF–BT–xLNT ceramics; (b,c) frequency dependent εr evolution for poled x = 0.02 sample at different annealing temperature and corresponding contour plot; (d,g) temperature dependent Raman spectra in 25–360 °C and (e,f,h,i) evolution of wavenumber and FWHM for selected D and G modes for x = 0.02 and 0.03 samples.
Figure 7. (a) Variation of d33 with thermal annealing at various temperatures for BF–BT–xLNT ceramics; (b,c) frequency dependent εr evolution for poled x = 0.02 sample at different annealing temperature and corresponding contour plot; (d,g) temperature dependent Raman spectra in 25–360 °C and (e,f,h,i) evolution of wavenumber and FWHM for selected D and G modes for x = 0.02 and 0.03 samples.
Nanomaterials 13 00942 g007
Figure 8. (ag) Temperature dependent unipolar SE curves at E = 60 kV/cm for BF–BT–xLNT ceramics; (h) evolution of unipolar SE with temperature; (i) comparison of electrostrain thermal stability between this work with some representative BF–[22,52,53], BNT–[54], and KNN–[55] based ceramics.
Figure 8. (ag) Temperature dependent unipolar SE curves at E = 60 kV/cm for BF–BT–xLNT ceramics; (h) evolution of unipolar SE with temperature; (i) comparison of electrostrain thermal stability between this work with some representative BF–[22,52,53], BNT–[54], and KNN–[55] based ceramics.
Nanomaterials 13 00942 g008
Figure 9. (ac) PFM amplitude and (df) phase images for selected x = 0, 0.03, 0.06 samples, respectively.
Figure 9. (ac) PFM amplitude and (df) phase images for selected x = 0, 0.03, 0.06 samples, respectively.
Nanomaterials 13 00942 g009
Table 1. Refined structural parameters and R–factors for BF–BT–xLNT ceramics.
Table 1. Refined structural parameters and R–factors for BF–BT–xLNT ceramics.
xSpace GroupLattice ParametersRwp (%)Rp (%)χ2
0R3cH
Pm 3 ¯ m
a = b = 5.69552 Å, c = 13.97396 Å, α = β = 90°, γ = 120°
a = b = c = 4.02774 Å, α = β = γ = 90°
5.704.521.31
0.01R3cH
Pm 3 ¯ m
a = b = 5.68390 Å, c = 13.99490 Å, α = β = 90°, γ = 120°
a = b = c = 4.02816 Å, α = β = γ = 90°
6.074.631.67
0.02R3cH
Pm 3 ¯ m
a = b = 5.69602 Å, c = 13.97944 Å, α = β = 90°, γ = 120°
a = b = c = 4.02837 Å, α = β = γ = 90°
6.374.811.56
0.03R3cH
Pm 3 ¯ m
a = b = 5.68895 Å, c = 13.96117 Å, α = β = 90°, γ = 120°
a = b = c = 4.02695 Å, α = β = γ = 90°
5.954.451.47
0.04R3cH
Pm 3 ¯ m
a = b = 5.69720 Å, c = 13.90696 Å, α = β = 90°, γ = 120°
a = b = c = 4.02521 Å, α = β = γ = 90°
6.134.591.64
0.05Pm 3 ¯ ma = b = c = 4.02458 Å, α = β = γ = 90°7.615.502.44
0.06Pm 3 ¯ ma = b = c = 4.02355 Å, α = β = γ = 90°7.445.302.21
Table 2. Resistance value (Ω) of Rb at different temperatures obtained by fitting circuit for BF–BT–xLNT ceramics.
Table 2. Resistance value (Ω) of Rb at different temperatures obtained by fitting circuit for BF–BT–xLNT ceramics.
x/T280 °C300 °C320 °C340 °C360 °C380 °C
03230172191541414961
0.0172,37431,54614,720675833911753
0.02177,48084,26334,95115,03574473804
0.0372,97337,54419,76610,81661133675
0.0461,70130,85416,371895850722934
0.0553,89927,06114,361794645212632
0.0654,68028,26314,881820846892739
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shi, H.; Li, K.; Li, F.; Ma, J.; Tu, Y.; Long, M.; Lu, Y.; Gong, W.; Wang, C.; Shan, L. Enhanced Piezoelectricity and Thermal Stability of Electrostrain Performance in BiFeO3-Based Lead-Free Ceramics. Nanomaterials 2023, 13, 942. https://doi.org/10.3390/nano13050942

AMA Style

Shi H, Li K, Li F, Ma J, Tu Y, Long M, Lu Y, Gong W, Wang C, Shan L. Enhanced Piezoelectricity and Thermal Stability of Electrostrain Performance in BiFeO3-Based Lead-Free Ceramics. Nanomaterials. 2023; 13(5):942. https://doi.org/10.3390/nano13050942

Chicago/Turabian Style

Shi, Hongwei, Kai Li, Feng Li, Jianxing Ma, Yubing Tu, Mingsheng Long, Yilin Lu, Weiping Gong, Chunchang Wang, and Lei Shan. 2023. "Enhanced Piezoelectricity and Thermal Stability of Electrostrain Performance in BiFeO3-Based Lead-Free Ceramics" Nanomaterials 13, no. 5: 942. https://doi.org/10.3390/nano13050942

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