Next Article in Journal
Determining the Annealing Temperature Dependency of Wetting and Mechanical Features on Fe3Si Films
Previous Article in Journal
On the Applicability of Modified Water-Based Yttria-Stabilized Zirconia Suspensions to Produce Plasma-Sprayed Columnar Coatings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimized Strain Response in (Co0.5Nb0.5)4+-Doped 76Bi0.5Na0.5TiO3-24SrTiO3 Relaxors

1
Engineering Department, Huanghe Science and Technology College, Zhengzhou 450006, China
2
School of Materials Science & Engineering, Hubei University, Wuhan 430062, China
*
Authors to whom correspondence should be addressed.
Coatings 2023, 13(8), 1331; https://doi.org/10.3390/coatings13081331
Submission received: 23 June 2023 / Revised: 24 July 2023 / Accepted: 26 July 2023 / Published: 28 July 2023

Abstract

:
High strain with low hysteresis is crucial for commercial applications in high precision actuators. However, the clear conflict between the high strain and low hysteresis in BNT-based ceramics has long been an obstacle to actual precise actuating or positioning applications. To obtain piezoceramics with high strain and low hysteresis, it is necessary to enhance the electrostrictive effect and develop an ergodic relaxor (ER) and nonergodic relaxor (NR) phase boundary under ambient conditions. In this work, (Co0.5Nb0.5)4+ doped 76Bi0.5Na0.5TiO3-24SrTiO3 (BNST24) relaxors were fabricated using the conventional solid state reaction route. X-ray diffraction patterns revealed the B-site substitution in BNST24 ceramics. By adjusting the (Co0.5Nb0.5)4+ doping in BNST24, we effectively tuned the TNR-ER and Td close to ambient temperature, which contributed to the development of the ergodic relaxor phase and enhanced the electrostrictive effect at ambient temperature. The I-P-E loops and bipolar strain curves verified the gradual evolution from NR to ER states, while the enhanced electrostrictive effect was verified by the nearly linear S-P2 curves and improved electrostrictive coefficient of the BNST24-xCN relaxors. An enhanced strain of 0.34% (d*33 = 483 pm/V) with low hysteresis of 8.9% was simultaneously achieved in the BNST24-0.02CN relaxors. The enhanced strain was mainly attributed to the proximity effect at the ER and NR phase boundary of BNST24-0.02CN, while the improved electrostrictive effect contributed to the reduced strain hysteresis. Our work demonstrates an effective strategy for balancing the paradox of high strain and low hysteresis in piezoceramics.

1. Introduction

Piezoelectric ceramics with perovskite structures have been the mainstay of actuator and transducer applications for the past 60 years due to their excellent electromechanical properties and stable thermal behavior [1,2]. Pb-based piezoelectric ceramics (e.g., Pb(Zr,Ti)O3 (PZT)) exhibit much higher strain than their lead-free counterparts, and therefore dominate the piezoelectric actuator market. Unfortunately, PZT ceramics contain more than 60 wt.% of the volatile element Pb, which can cause serious environmental problems. Moreover, WEEE/RoHS regulations have been widely applied in many countries to restrict the use of hazardous substances, including lead, in electrical and electronic equipment, driving the search for lead-free piezoelectric ceramics to replace PZT-based materials.
Among all of the reported lead-free piezoelectric ceramic families, Bi0.5Na0.5TiO3 (BNT)-based ceramics are considered as potential candidates to replace commercial PZT, especially in actuator applications, due to their superior strain performance, comparable to lead-based ferroelectrics [3,4,5,6,7,8,9,10]. However, the electric field level required to induce a large strain is too high for commercial applications in these BNT-based systems [7,11,12,13]. For example, a large strain of 0.43% in 90Bi1/2Na1/2TiO3-60BaTiO3-4K0.5Na0.5NbO3 ceramics at 11 kV/mm [11] and a large strain of 0.51% was achieved in <001> texture (Bi0.5Na0.5)TiO3-BaTiO3-NaNbO3 ceramics under a driving electric field of 12 kV/mm [12]. One material with potential to overcome this problem is the (1-x) BNT-xSrTiO3 (BNST100x) ceramic [14,15,16,17,18,19,20,21,22]. Krauss reported that a maximum strain of about 0.29% under 6 kV/mm was found at x = 0.25, which coincided with the transition from ferroelectric to anti-ferroelectric phases [14]. Duong observed a strain of 0.25% under 4 kV/mm in BNST28 and attributed this large strain to the phase evolution from nonergodic relaxor to ergodic relaxor states [15]. Furthermore, many researchers have investigated BNST100x ceramics with ST contents of around 24-26 mol%, which could form the polymorphic phase boundary (PPB). Samples near the PPB displayed large field-induced strain values, making BNST a promising material system for actuators in practical applications [16,17,18,19]. Therefore, we selected 76Bi0.5Na0.5TiO3-24SrTiO3 (BNST24) as the base composition, in an attempt to trigger large strain at a comparatively lower electrical field. However, the enormous strain in BNST-based relaxors has typically been accompanied by large strain hysteresis, which will cause displacement hysteresis and thermal consumption [8,20,21,22]. For example, the ((Bi1/2 (Na0.84 K0.16)1/2)0.96 Sr0.04) (Ti1-xNbx)O3 (x = 0.025) ceramic exhibited a giant strain (≈0.7%) but suffered from a huge hysteresis (≈69%) [8]. Although many studies have been carried out to suppress this large hysteresis, this has often resulted in a significant reduction in strain (<0.25%) [23,24,25,26]. Pan reported that (0.94-x) Bi0.5Na0.5TiO3-0.06BaTiO3-x(Sr0.7Bi0.18Er0.02)TiO3 ceramics exhibited lower hysteresis (26%), but with a small strain (0.11%) [23]. Therefore, it is an urgent task to balance high strain and low hysteresis in BNST-based relaxors to accelerate their actual application in environmentally friendly high-precision actuators.
In relaxors, electric-field-induced strain is derived from the internal electrostrictive effect, as well as electric-field-induced phase transition. Of these, the strain generated based on the electrostrictive effect is almost hysteresis free, with the hysteresis mainly caused by the phase transition. Therefore, strain hysteresis can be suppressed by enhanced electrostriction at room temperature (RT). An effective way to improve the electrostrictive effect is to tune the phase transition temperature of the relaxors towards RT [27,28], and a giant strain can be induced by the proximity effect at two different phase boundaries [29]. According to a previous study on the relaxor ferroelectric behaviors of (1-x)(Bi0.5Na0.5)TiO3-xSrTiO3 ceramics (0 ≤ x ≤ 0.3), the virgin state of the 76(Bi0.5Na0.5)TiO3-24SrTiO3 ceramic was found to be the NR state at RT when no driving electric field was applied, and the NR state would transfer to ER state upon heating to 150 °C [30]. To take advantage of the electrostrictive property to suppress hysteresis at RT, the TNR-ER of 76(Bi0.5Na0.5)TiO3-24SrTiO3 is supposed to be shifted to around RT. Reports [31,32] have shown that B-site complex ion doping was more effective in inducing the phase transition and improving the strain, so (Co0.5Nb0.5)4+ (CN) was introduced into BNT-24ST nonergodic relaxors as a dopant to induce the transition from NR state to ER state, and a giant strain could be induced by the proximity effect at the ER/NR phase boundary. The incorporation of CN is expected to tune the phase transition temperature to around RT and produce a coexistence of ER and NR phases at RT. Thus, the electrostrictive effect and strain will be enhanced, effectively breaking the restrictive relationship between strain and hysteresis. The structures, dielectric, ferroelectric and electromechanical strain performance of CN-modified BNST24 were systematically investigated to discover an optimized amount of CN doping in BNST24. An enhanced strain of 0.34% (d*33 = 483 pm/V) with low hysteresis of 8.9% were successfully obtained in the optimized CN-modified BNST24, which is an optimized improvement in the strain performance compared to the recent study [24,33,34]. Herein, we proposed a strategy to balance the paradox of high strain and low hysteresis for promoting commercial applications in high precision actuators.

2. Materials and Methods

2.1. Preparation of BNST24-x CN Ceramics

The 76(Bi1/2Na1/2)TiO3-24SrTi1-x(Co0.5Nb0.5)xO3 relaxors were synthesized via a conventional solid-state reaction method employing high-purity raw materials of Bi2O3, Na2CO3, TiO2, Nb2O5, Co2O3 and SrCO3. These raw materials were weighed according to the stoichiometric ratio of the compositions and homogenized with the ZrO2 balls in ethanol for 6 h. The dried slurries were pulverized and calcined at 850 °C for 2 h, and then the calcined mixture was ball milled again for 12 h and dried. The resulting powders were subsequently pressed into 10-mm-diameter disks under approximately 100 MPa. These disks were sintered in air at 1120 °C for 2 h. To minimize possible losses of volatile elements, these disks were embedded in a mixture of identical content. Finally, silver electrodes were coated on the surface of the ceramic at 580 °C for 30 min.

2.2. Characterization

The crystal structures of these ceramics were characterized via X-ray diffraction (XRD, D8-Advance, Bruker AXS Inc.,Madison, Germany) with CuK α radiation. The microstructure and morphology were tested using scanning electron microscopy (SEM, Quanta FEG 250, Frequency Electronics, Inc., Hillsboro, OR, USA). The permittivity and dielectric loss, which are dependent on temperature and frequency, were determined via a high-precision LCR meter (TH2827, Tonghui Electronic Co., Ltd., Jiangsu, China) from RT to 500 °C at 4 frequencies (0.1, 1, 10, and 100 kHz). The electric-field-induced polarization (P-E) and strain (S-E) hysteresis loops were recorded with a ferroelectric tester (RTI-LC П, Radiant Technologies Inc., Burbank, CA, USA) connected with a laser measurement system (MTI-2100, Hottinger Baldwin Messtechnik GmbH, Darmstadt, Germany). Piezoelectric constant was quantified by employing a quasi-static d33 meter (ZJ-3AN, Institute of Acoustics, Chinese Academy of Sciences, Beijing, China).

3. Results and Discussion

The XRD patterns in Figure 1 show that all compositions of the BNST24-xCN ceramics exhibited a perovskite structure without any observable secondary phases, indicating that the Co+3 ions and Nb+5 ions completely diffused into the crystal lattice of the BNST24 matrix. The enlarged diffraction patterns in the ranges of 39.5°–40.5° and 46°–47° are shown in Figure 1b,c, which were obtained to gain insight into the phase structure. The ( 111 ) / ( 1 1 ¯ 1 ) and ( 200 ) / ( 002 ) peak splitting, visible at both 40° and 46.5°, revealed the coexistence of rhombohedral and tetragonal phases in all of the BNST24-xCN samples. The rhombohedral phase was identified by the ( 111 ) / ( 1 1 ¯ 1 ) peaks at about 40° and the tetragonal phase was determined by the ( 200 ) / ( 002 ) peaks at about 46.5°, as reported in previous reports [21,35,36,37]. Simultaneously, the (111) and (200) peaks gradually shifted towards lower diffraction angles as the CN increased. This indicated that the Co+3 and Nb+5 ions replaced Ti4+ at the B-site of the perovskite unit cell. Due to the relatively larger ionic radii of Co3+ (0.745Å) and Nb5+ (0.64Å) compared to Ti4+ (0.605Å) at the B-site, the substituent (Co0.5Nb0.5)4+ ions generated lattice expansion, which then corresponded with the peak positions by shifting toward a lower angle [38,39,40,41,42].
The SEM micrographs in Figure 2a–e demonstrate the influence of various CN concentrations on the morphology and microstructure of the BNST24 ceramics. Compared to the pure BNST24 sample, BNST24-xCN samples exhibited more homogeneous and denser microstructures, where the grain size obtained by the linear intercept method first increased for 0 ≤ x≤ 0.02 and then presented a stable trend. This result showed that proper CN doping promoted grain growth, thereby improving the density and grain uniformity of the BNST24-xCN ceramics. However, the crystals have a certain solid solubility for dopants, and over-doping with CN causes dopants to aggregate at grain boundaries, and the pinning of grain boundaries would inhibit crystal growth [31]. To evaluate the distribution of CN doping in a clear manner, the plane sweeping mode of energy dispersive spectroscopy (EDS) was adopted in Figure 3b. The EDS result indicated that the Co element was distributed homogenously in the grains of the BNST24-0.02CN sample, which was consistent with the XRD results.
Temperature vs. relative dielectric constant (εr) and dielectric loss (tanδ) of BNST24-xCN (0 ≤ x ≤ 0.04) ceramics across various frequencies are illustrated in Figure 4a–e. These broad dielectric permittivity peaks relative to temperature and the apparent frequency dispersion of dielectric permittivity indicated that all of the BNST24-xCN samples exhibited the typical characteristics of relaxors. Two permittivity anomalies were detected for the 0 ≤ x ≤ 0.01 compositions. The first was located at Tm (the temperature where εr reached a maximum), reflecting the transition from paraelectric to ER phase [22], while the second anomaly was observed at a lower temperature TNR-ER, corresponding to the transition from the ER to the NR phase. With increasing CN to 0.04, Tm showed little difference in position, but TNR-ER varied considerably, as shown in Figure 4f. The TNR-ER value was 140 °C for BNST24, which then shifted down to 120 °C for BNST24-0.01CN. Interestingly, no distinctive TNR-ER could be detected for x ≥ 0.02, which implied that TNR-ER should be below room temperature. As TNR-ER approached room temperature, the interaction between PNRS became weakly correlated due to the increased dynamics of the PNRS, resulting in a transition from a short-range-ordered NR (weakly polar) state to an ER state with randomly oriented polar clusters (non-polar state), thereby stabilizing the ER phase at ambient temperature [29,43]. In addition, the depolarization temperature Td where the dielectric loss exhibited a peak gradually shifted from 114 °C down to 60 °C with increasing CN from 0 to 0.04, as we can see from the inset of Figure 4f. This suggested that an appropriate concentration of CN could effectively adjust the TNR-ER and Td close to ambient temperature, which could promote the strain value and electrostrictive effect of BNST-based ceramics [44].
To further explore the compositional dependence of the phase evolution, the I-P-E loops of the BNST24-xCN relaxors under ambient conditions were plotted, as shown in Figure 5a–e. Parameters such as Pr, Pmax, Pmax-Pr and Ec derived from the P-E loops were plotted as a function of CN content in Figure 5f. As shown in Figure 5a, an unsaturated FE-like hysteresis loop with a relatively low Pr value of 18 μC/cm2 was obtained for BNST. The I-E loop showed four current peaks at ±EF and ±ER, two for each loading quadrant. Under the cycling regime conditions, the peak at +EF indicated the critical field that the polar order produced under a positive electric field [45]. No peaks appeared upon electric field unloading until the electric field was reversed to − ER, suggesting that the residual polarization at E = 0, which could be confirmed by the P-E loop. The polar order produced at +EF was recovered at −ER. The polarization became zero at the field between −ER and −EF. With 0.01 CN doping, a slim and pinched P-E loop (Figure 5b) attributed to the sharp decrease in Pr and Ec (Figure 5f) was observed, suggesting the appearance of the ER phase around E = 0. The presence of a weak Pr value of 10 μC/cm2 indicated a retention of the weakly polar NR state in the non-polar ER state. The I-E loop of the BNST-0.01CN ceramic also showed four current peaks at ±EF and ±ER, but the current peaks corresponding to ±ER appeared during electric field unloading, implying that the polar order that occurred at +EF (−EF) could be recovered to a non-polar order at +ER (−ER) during unloading. With a further increase in CN to 0.04, the progressively decreased Pr and Ec indicated that the non-polar ER state became more stable. Of note, + EF (−EF) increased with a further increase in CN content, which reflected that the electric field required for the polarization mechanism became increasingly hindered. These effects suggested that the non-polar order became more stable with higher CN content at RT.
The CN doping induced a gradual transition from the NR to the ER phase, which was well verified by the bipolar strain measurements, as illustrated in Figure 6a. The deformed butterfly shape with a negative strain Sneg of 0.032% at x = 0 implied the NR phase for BNST24 at ambient temperature, which is consistent with the reported research [30]. With the introduction of CN to 0.01, the deformed butterfly-shaped strain hysteresis loop changed into a bud-like loop, accompanied by a sharp decrease in Sneg, d33 and a simultaneous increase in positive strain Spos, as presented in Figure 6b. With a further increase in CN (x ≥ 0.02), Sneg disappeared and d33 continued to decrease. Moreover, Spos reached its maximum at x = 0.02, but decreased with increasing CN content. The Sneg and d33 were closely correlated with domain back-switching [27], and the continuous decrease in Sneg and d33 indicated that the weakly polar NR order in BNST24 was disrupted by the addition of CN, with a decreasing degree of nonergodicity and an increasing degree of ergodicity. At a critical range around x = 0.2, where ER and NR states are competitive, a large strain of 0.35% was induced by the proximity effect at the ER and NR phase boundary [29] at 70 kV/cm, as shown in Figure 7a. For x ≥ 0.03, the reduction in strain could be attributed to the phase composition in these samples, which was away from the NR-ER phase boundary, and the ER phase was dominant.
To evaluate the suitability of the materials for actuator applications, the unipolar S-E loops and normalized strain d*33 (d*33 = Smax/Emax) abstracted from the S-E loops of all investigated BNST24-xCN ceramics were plotted in Figure 7a,b. With increasing CN, the unipolar strain and normalized strain d*33 initially increased and then diminished after the maximum at x = 0.02, while the corresponding maximum strain and d*33 were 0.34% and 483 pm/V at 70 kV/cm, respectively. Combining the I-P-E loops and bipolar strain curves, the observed maximum strain and d*33 indicated that large strain values could be induced by the proximity effect around the critical composition (x = 0.02) at ambient temperatures. Beyond the critical composition, either the NR phase or ER phase would dominate, thus reducing the strain response. To highlight the relationship between the strain and TNR-ER, the d*33 and TNR-ER abstracted from the dielectric measurements were plotted in the CN range of 0–0.04 (Figure 7b). As seen in Figure 7b, the doping of CN could decrease the TNR-ER from 140 °C for BNST24 to RT for x = 0.02 and promote the maximum d*33 of the BNST24-xCN series. The highest d*33 was achieved near the critical zone (x = 0.02) when its TNR-ER approached RT. However, the continued increase in CN beyond 0.02 resulted in decreased strain.
For actuator applications, large strain with low hysteresis η (ηShyst/Smax, where the ΔShyst is shown in Figure 8a) is very important for the design of high-precision actuators [31,46]. The BNST-0.02CN sample presented a low η value of 8.9% at 80 kV/cm, which was mainly attributed to the intrinsic electrostrictive effect. The electrostrictive effect in dielectric ceramics could be described by the electrostrictive coefficient Q33 (Q33 = S3/P23) [47]. To evaluate the electrostrictive effect of BNST24-xCN (0 ≤ x ≤ 0.04), the electrostrictive curves (S-P2 curves) at RT for x = 0, 0.01, 0.02, 0.03 and 0.04 were carried out, as presented in Figure 8b. Evidently, the S-P2 curves for x = 0 deviated from linear, while the curves with x ≥ 0.01 were almost linear. The electrostrictive coefficients were 0.0151 m4/C2 for x = 0 and 0.0245 m4/C2 for x = 0.02. These features indicated that the electrostrictive effect was enhanced greatly by the CN dopant, which contributed to TNR-ER and Td close to ambient temperature.

4. Conclusions

In this work, we utilized a conventional solid state reaction method to fabricate the BNST24-xCN ceramics (0 ≤ x ≤ 0.04). XRD and SEM were employed to investigate the structural properties, revealing the coexistence of rhombohedral and tetragonal phases and the B-site substitution of (Co0.5Nb0.5) 4+ in the perovskite BNST24 ceramics. The ε r -T curves indicated that an appropriate CN doping could effectively shift TNR-ER and Td downwards close to ambient temperatures, which contributed to an enhanced electrostrictive effect at ambient temperatures. The I-P-E and S-E loops confirmed the gradual transition from NR phase to ER phase with increasing CN doping. According to the S-P2 curves of BNST24-xCN (0 ≤ x ≤ 0.04), a high Q33 = 0.0245m4/C2 was obtained at RT for the 0.02 CN composition. As a result, the composition (x = 0.02) located near the NR-ER phase boundary exhibited a low hysteresis (8.9%) and an enhanced strain value of 0.34% (d*33 ~483 pm/V). The large strain value was attributed to the proximity effect at the boundary of the ER and NR phases, and the adjustment of TNR-ER close to RT was imperative for obtaining low hysteresis. Our work effectively balanced the conflict between high strain and low hysteresis and could stimulate further research on developing high-precision actuator applications.

Author Contributions

Conceptualization, H.L. and Y.Z.; Methodology, J.G.; Formal Analysis, H.L.; Investigation, H.L. and Y.Z.; Resources, M.L.; Data Curation, Y.Z.; Writing—Original Draft Preparation, H.L.; Writing—Review and Editing, Q.Z.; supervision, Y.Z. and Q.Z.; Project Administration, H.L.; All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Project of Henan Province Science and technology (Grant Nos. 212102210187, 212102210603, 232102221003), the Key Research Projects of Henan Higher Education Institutions (Grant No. 22B430020) and the Postgraduate Education Reform and Quality Improvement Project of Henan Province (YJS2023JD67, YJS2022JD50).

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. Yan, B.B.; Fan, H.Q.; Wang, C.; Zhang, M.C.; Yadav, A.K.; Zheng, X.K.; Wang, H.; Du, Z.N. Giant electro-strain and enhanced energy storage performance of (Y0.5Ta0.5) 4+ co-doped 0.94(Bi0.5Na0.5)TiO3-0.06BaTiO3 lead-free ceramics. Ceram. Int. 2020, 46, 281–288. [Google Scholar] [CrossRef]
  2. Si, Y.; Li, Y.; Li, L.; Li, H.; Zhao, Z.H.; Dai, Y.J. Giant electrostrain in textured Li+-doped 0.852BNT-0.11BKT-0.038BT ternary lead-free piezoelectric ceramics. J. Am. Ceram. Soc. 2020, 103, 1765–1772. [Google Scholar] [CrossRef]
  3. Liu, G.; Dong, J.; Zhang, L.; Yan, Y.; Jing, R.; Jin, L. Phase evolution in (1−x) (Na0.5Bi0.5)TiO3xSrTiO3 solid solutions: A study focusing on dielectric and ferroelectric characteristics. J. Mater. 2020, 6, 677–691. [Google Scholar] [CrossRef]
  4. Zhou, C.; Zhang, Y.; Wang, Y.; Li, Q.; Xu, J.; Chen, G.; Yuan, C.; Rao, G. An intermediate metastable ferroelectric state induced giant functional responses in Bi0.5Na0.5TiO3 ceramics. J. Mater. Chem. C 2019, 27, 8255–8260. [Google Scholar] [CrossRef]
  5. Lin, D.; Xiao, D.; Zhu, J.; Yu, P. Piezoelectric and ferroelectric properties of [Bi0.5(Na1−x−y KxLiy)0.5]TiO3 lead-free piezoelectric ceramics. Appl. Phys. Lett. 2006, 88, 062901. [Google Scholar] [CrossRef]
  6. Yin, J.; Liu, G.; Lv, X.; Zhang, Y.; Zhao, C.; Wu, B.; Zhang, X.; Wu, J. Superior and anti-fatigue electro-strain in Bi0.5Na0.5TiO3-based polycrystalline relaxor ferroelectrics. J. Mater. Chem. A 2019, 7, 5391–5401. [Google Scholar] [CrossRef]
  7. 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]
  8. Liu, X.; Tan, X. Giant strains in non-textured (Bi1/2Na1/2)TiO3-based lead-free ceramics. Adv. Mater. 2016, 28, 574–578. [Google Scholar] [CrossRef]
  9. Yin, J.; Zhao, C.; Zhang, Y.; Wu, J. Ultrahigh strain in site engineering-independent Bi0.5Na0.5TiO3-based relaxor-ferroelectrics. Acta Mater. 2018, 147, 70–77. [Google Scholar] [CrossRef]
  10. Pham, K.N.; Hussain, A.; Ahn, C.W.; Kim, I.W.; Jeong, S.J.; Lee, J.S. Giant strain in Nb-doped Bi0.5(Na0.82K0.18)0.5TiO lead-free electromechanical ceramics. Mater. Lett. 2010, 64, 2219–2222. [Google Scholar] [CrossRef]
  11. Zhang, S.T.; Kounga, A.B.; Aulbach, E. Giant strain in lead-free piezoceramics Bi0.5Na0.5TiO3-BaTiO3-K0.5Na0.5NbO3 system. Appl. Phys. Lett. 2007, 91, 112906. [Google Scholar] [CrossRef]
  12. Bai, W.F.; Zhao, X.Y.; Ding, Y.Q.; Wang, L.J.; Zheng, P.; Hao, J.G.; Zhai, J.W. Giant field-induced strain with low hysteresis and boosted energy storage performance under low electric field in (Bi0.5Na0.5)TiO3-based grain orientation-controlled ceramics. Adv. Electron. Mater. 2020, 6, 2000332. [Google Scholar] [CrossRef]
  13. Yin, J.; Zhao, C.; Zhang, Y.; Wu, J. Composition-induced phase transitions and enhanced electrical properties in bismuth sodium titanate ceramics. J. Am. Ceram. Soc. 2017, 100, 5601–5609. [Google Scholar] [CrossRef]
  14. Krauss, W.; Schütz, D.; Mautner, F.A.; Feteira, A.; Reichmann, K. Piezoelectric properties and phase transition temperatures of the solid solution of (1−x) (Bi0.5Na0.5)TiO3xSrTiO3. J. Eur. Ceram. Soc. 2010, 30, 1827–1832. [Google Scholar] [CrossRef]
  15. Duong, T.A.; Han, H.S.; Hong, Y.H.; Park, Y.S.; Nguyen, H.T.K.; Dinh, T.H.; Lee, J.S. Dielectric and piezoelectric properties of Bi1/2Na1/2TiO3-SrTiO3 lead-free ceramics. J. Electroceram. 2018, 41, 73–79. [Google Scholar] [CrossRef]
  16. Acosta, M.; Jo, W.; Rödel, J. Temperature and frequency-dependent properties of the 0.75Bi1/2Na1/2TiO3-0.25SrTiO3 lead-free incipient piezoceramic. J. Am. Ceram. Soc. 2014, 97, 1937–1943. [Google Scholar] [CrossRef]
  17. Acosta, M.; Schmitt, L.A.; Luna, L.M.; Scherrer, M.C.; Brilz, M.; Webber, K.G.; Deluca, M.; Kleebe, H.J.; Rödel, J.; Donner, W. Core-shell lead-free piezoelectric ceramics: Current status and advanced characterization of the Bi1/2Na1/2TiO3-SrTiO3 system. J. Am. Ceram. Soc. 2015, 98, 3405–3422. [Google Scholar] [CrossRef]
  18. Li, H.L.; Liu, Q.; Zhou, J.J.; Wang, K.; Li, J.F.; Liu, H.; Fang, J.Z. Grain size dependent electrostrain in Bi1/2Na1/2TiO3-SrTiO3 incipient piezoceramics. J. Eur. Ceram. Soc. 2016, 36, 2849–2853. [Google Scholar] [CrossRef]
  19. Kim, S.H.; Lee, S.H.; Han, H.S.; Lee, J.S. Piezoelectric characteristics of lead-free 0.74(Bi0.5Na0.5)TiO3-0.26SrTiO3 ceramics according to calcination temperature. J. Korean Inst. Electr. Electron. Mater. Eng. 2019, 32, 35–39. [Google Scholar]
  20. Koruza, J.; Rojas, V.; Molina-Luna, L.; Kunz, U.; Duerrschnabel, M.; Kleebe, H.J.; Acosta, M. Formation of the core-shell microstructure in lead-free Bi1/2Na1/2TiO3-SrTiO3 piezoceramics and its influence on the electromechanical properties. J. Eur. Ceram. Soc. 2016, 36, 1009–1016. [Google Scholar] [CrossRef]
  21. Zhu, Y.; Zhang, Y.; Xie, B.; Fan, P.; Marwat, M.A.; Ma, W.; Wang, C.; Yang, B.; Xiao, J.; Zhang, H. Large electric field-induced strain in AgNbO3-modified 0.76Bi0.5Na0.5TiO3-0.24SrTiO3 lead-free piezoceramics. Ceram. Int. 2018, 44, 7851–7857. [Google Scholar] [CrossRef]
  22. Hong, Y.H.; Han, H.S.; Jeong, G.H.; Park, Y.S.; Dinh, T.H.; Ahn, C.W.; Lee, J.S. High electromechanical strain properties by the existence of nonergodicity in LiNbO3-modified Bi1/2Na1/2TiO3–SrTiO3 relaxor ceramics. Ceram. Int. 2018, 44, 21138–21144. [Google Scholar] [CrossRef]
  23. Pan, H.; Zhang, J.; Jia, X.; Xing, H.; He, J.; Wang, J.; Wen, F. Large electrostrictive effect and high optical temperature sensing in Bi0.5Na0.5TiO3-BaTiO3-(Sr0.7Bi0.18Er0.02)TiO3 luminescent ferroelectrics. Ceram. Int. 2018, 44, 5785–5789. [Google Scholar] [CrossRef]
  24. Yu, D.; Zhou, C.; Chen, J.; Li, Q.; Li, L.; Yuan, C.; Xu, J.; Rao, G. Large electrostriction-like strain by tailoring relaxor degree in BNT-based ceramics. J. Mater. Sci. Mater. Electron. 2023, 34, 1153. [Google Scholar] [CrossRef]
  25. Yao, K.; Zhou, C.; Li, Q.; Xiao, Z.; Yuan, C.; Xu, J.; Chen, G.; Rao, G. Large electrostrictive coefficient with optimized Electro-Strain in BNT-based ceramics with ergodic state. Mater. Sci. Eng. B 2022, 283, 115828. [Google Scholar] [CrossRef]
  26. Jin, L.; Pang, J.; Jing, R.; Lan, Y.; Wang, L.; Li, F.; Hu, Q.; Du, H.; Guo, D.; Wei, X.; et al. Ultra-slim pinched polarization-electric field hysteresis loops and thermally stable electrostrains in lead-free sodium bismuth titanate based solid solutions. J. Alloys Compd. 2019, 788, 1182–1192. [Google Scholar] [CrossRef]
  27. Shi, J.; Fan, H.Q.; Liu, X.; Bell, A.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] [CrossRef]
  28. Lu, X.; Hou, L.; Jin, L.; Wang, D.; Hu, Q.; Alikin, D.O.; Turygin, A.P.; Wang, L.; Zhang, L.; Wei, X. Origin of composition-insensitive electrostrictive coefficient and continuous decrease of domain wall density in (1−x)NaNbO3xBaTiO3 lead-free ferroelectrics. J. Eur. Ceram. Soc. 2018, 38, 3127–3135. [Google Scholar] [CrossRef]
  29. Shvartsman, V.V.; Lupascu, D.C. Lead-free relaxor ferroelectrics. J. Am. Ceram. Soc. 2012, 95, 1–26. [Google Scholar] [CrossRef]
  30. He, H.Y.; Lu, X.; Li, M.C.; Wang, Y.M.; Li, Z.R.; Lu, Z.G.; Lu, L. Thermal and compositional driven relaxor ferroelectric behaviours of lead-free (Bi0.5Na0.5)TiO3-SrTiO3 ceramics. J. Mater. Chem. C 2020, 8, 2411. [Google Scholar] [CrossRef]
  31. Fan, P.Y.; Zhang, Y.Y.; Xie, B.; Zhu, Y.W.; Ma, W.G.; Wang, C.; Yang, B. Large electric-field-induced strain in B-site complex-ion (Fe0.5Nb0.5)4+-doped Bi1/2 (Na0.82K0.12)1/2TiO3 lead-free piezoceramics. Ceram. Int. 2018, 44, 3211–3217. [Google Scholar] [CrossRef]
  32. Cheng, R.F.; Xu, Z.J.; Chu, R.Q.; Hao, J.G.; Du, J.; Li, G.R. Electric field-induced ultrahigh strain and large piezoelectric effect in Bi1/2Na1/2TiO3-based lead-free piezoceramics. J. Eur. Ceram. Soc. 2016, 36, 489–496. [Google Scholar] [CrossRef]
  33. 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]
  34. Deng, A.; Wu, J. Optimized strain properties with small hysteresis in BNT-based ceramics with ergodic relaxor state. J. Eur. Ceram. Soc. 2021, 41, 5147–5154. [Google Scholar] [CrossRef]
  35. Lee, H.B.; Heo, D.J.; Malik, R.A.; Yoon, C.H.; Han, H.S.; Lee, J.S. Lead-free Bi1/2(Na0.82K0.18)1/2TiO3 ceramics exhibiting large strain with small hysteresis. Ceram. Int. 2013, 39, 5705–5708. [Google Scholar] [CrossRef]
  36. Malik, R.A.; Hussain, A.; Maqbool, A.; Zaman, A.; Ahn, C.W.; Rahman, J.U.; Song, T.K.; Kim, W.J.; Kim, M.H. Temperature-insensitive high strain in lead-free Bi0.5(Na0.84K0.16)0.5TiO3-0.04SrTiO3 ceramics for actuator applications. J. Am. Ceram. Soc. 2015, 98, 3842–3848. [Google Scholar] [CrossRef]
  37. Xi, H.; Yu, L.X.; Qian, H.; Chen, F.J.; Mao, M.M.; Liu, Y.F.; Lyu, Y.N. Large strain with low hysteresis in Sn-modified Bi0.5(Na0.75K0.25)0.5TiO3 lead-free piezoceramics. J. Mater. Sci. 2020, 55, 1388–1398. [Google Scholar] [CrossRef]
  38. Gozuacik, N.K.; Bayir, M.C.; Alkoy, E.M.; Alkoy, S. Origin of the large field-induced strain and enhanced energy storage response of rare-earth-doped lead-free 0.854BNT-0.12BKT-026BT ceramics, IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 2021, 68, 2576–2584. [Google Scholar] [CrossRef]
  39. Zhao, W.L.; Zuo, R.Z.; Li, F.; Li, L.T. Structural, dielectric, ferroelectric and strain properties in CaZrO3-modified Bi(Mg0.5Ti0.5)O3-PbTiO3 solid solutions. J. Alloys Compd. 2014, 591, 218–223. [Google Scholar] [CrossRef]
  40. Dinh, T.H.; Kang, J.K.; Lee, J.S.; Khansur, N.H.; Daniels, J.; Lee, H.Y.; Yao, F.Z.; Wang, K.; Li, J.F.; Han, H.S.; et al. Nanoscale ferroelectric/relaxor composites: Origin of large strain in lead-free Bi-based incipient piezoelectric ceramics. J. Eur. Ceram. Soc. 2016, 36, 3401–3407. [Google Scholar] [CrossRef]
  41. Chen, N.; Yao, W.; Liang, C.; Xiao, S.; Hao, J.; Xu, Z.; Chu, R. Phase structure, ferroelectric properties, and electric field-induced large strain in lead-free 0.99[(1−x)(Bi0.5Na0.5)TiO3x(Bi0.5K0.5)TiO3]-0.01Ta piezoelectric ceramics. Ceram. Int. 2016, 42, 9660–9666. [Google Scholar] [CrossRef]
  42. Tran, V.D.N.; Ullah, A.; Dinh, T.H.; Lee, J.S. Large field induced strain properties of Sr (K0.25Nb0.75)O3-modified Bi1/2(Na0.82K0.18)1/2TiO3 lead-free piezoelectric ceramics. J. Electron. Mater. 2016, 45, 2627–2631. [Google Scholar] [CrossRef]
  43. He, H.Y.; Lu, W.H.; Oh, J.A.S.; Li, Z.R.; Lu, X.; Zeng, K.Y.; Lu, L. Probing the coexistence of ferroelectric and relaxor states in Bi0.5Na0.5TiO3-based ceramics for enhanced piezoelectric performance. ACS Appl. Mater. Interfaces 2020, 12, 30548–30556. [Google Scholar] [CrossRef]
  44. Yin, J.; Wang, Y.M.; Zhang, Y.X.; Wu, B.; Wu, J.G. Thermal depolarization regulation by oxides selection in lead-free BNT/oxides piezoelectric composites. Acta Mater. 2018, 158, 269–277. [Google Scholar] [CrossRef]
  45. Viola, G.; McKinnon, R.; Koval, V.; Adomkevicius, A.; Dunn, S.; Yan, H. Lithium-induced phase transitions in lead-free Bi0.5Na0.5TiO3 based ceramics. J. Phys. Chem. C. 2014, 118, 8564–8570. [Google Scholar] [CrossRef]
  46. Zhang, H.B.; Xu, P.W.; Patterson, E.; Zang, J.D.; Jiang, S.L.; Rödel, J. Preparation and enhanced electrical properties of grain oriented (Bi1/2Na1/2)TiO3-based lead-free incipient piezoceramics. J. Eur. Ceram. Soc. 2015, 35, 2501–2512. [Google Scholar] [CrossRef]
  47. 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]
Figure 1. XRD patterns of the BNST24-x CN ceramics before poling in a 2θ range of (a) 20–80°, (b) 39.5–40.5°, (c) 46–47°.
Figure 1. XRD patterns of the BNST24-x CN ceramics before poling in a 2θ range of (a) 20–80°, (b) 39.5–40.5°, (c) 46–47°.
Coatings 13 01331 g001
Figure 2. SEM micrographs of the BNST24-xCN ceramics with varied CN concentrations: (a) x = 0, (b) x = 0.01, (c) x = 0.02, (d) x = 0.03, and (e) x = 0.04.
Figure 2. SEM micrographs of the BNST24-xCN ceramics with varied CN concentrations: (a) x = 0, (b) x = 0.01, (c) x = 0.02, (d) x = 0.03, and (e) x = 0.04.
Coatings 13 01331 g002
Figure 3. (a) SEM micrograph of the BNST24-0.02 CN ceramic; (b) the EDS mapping of Co element in BNST24-0.02 CN ceramic.
Figure 3. (a) SEM micrograph of the BNST24-0.02 CN ceramic; (b) the EDS mapping of Co element in BNST24-0.02 CN ceramic.
Coatings 13 01331 g003
Figure 4. ε r -T and tan δ -T: (ae) BNST24-xCN ceramics at different frequencies, and (f) for all compositions at 1 kHz.
Figure 4. ε r -T and tan δ -T: (ae) BNST24-xCN ceramics at different frequencies, and (f) for all compositions at 1 kHz.
Coatings 13 01331 g004
Figure 5. (ae): I-P-E loops of the BNST24-xCN relaxors and (f) Pmax, Pr, Pmax −Pr and Ec derived from the P-E loops as a function of CN content.
Figure 5. (ae): I-P-E loops of the BNST24-xCN relaxors and (f) Pmax, Pr, Pmax −Pr and Ec derived from the P-E loops as a function of CN content.
Coatings 13 01331 g005
Figure 6. (a) The bipolar strain curves of the BNST24-xCN relaxors measured at 1 Hz; (b) Spos, Sneg, d33 of the composites with varying CN contents.
Figure 6. (a) The bipolar strain curves of the BNST24-xCN relaxors measured at 1 Hz; (b) Spos, Sneg, d33 of the composites with varying CN contents.
Coatings 13 01331 g006
Figure 7. (a) Unipolar strain curves of the BNST24-xCN ceramics measured at 1 Hz; (b) variation of d*33 and TNR-ER with respect to CN content.
Figure 7. (a) Unipolar strain curves of the BNST24-xCN ceramics measured at 1 Hz; (b) variation of d*33 and TNR-ER with respect to CN content.
Coatings 13 01331 g007
Figure 8. (a) Unipolar strain curves of BNST24-0.02CN measured at different electric fields; (b) S-P2 curves of the BNST24-xCN ceramics.
Figure 8. (a) Unipolar strain curves of BNST24-0.02CN measured at different electric fields; (b) S-P2 curves of the BNST24-xCN ceramics.
Coatings 13 01331 g008
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

Li, H.; Gao, J.; Li, M.; Zhang, Q.; Zhang, Y. Optimized Strain Response in (Co0.5Nb0.5)4+-Doped 76Bi0.5Na0.5TiO3-24SrTiO3 Relaxors. Coatings 2023, 13, 1331. https://doi.org/10.3390/coatings13081331

AMA Style

Li H, Gao J, Li M, Zhang Q, Zhang Y. Optimized Strain Response in (Co0.5Nb0.5)4+-Doped 76Bi0.5Na0.5TiO3-24SrTiO3 Relaxors. Coatings. 2023; 13(8):1331. https://doi.org/10.3390/coatings13081331

Chicago/Turabian Style

Li, Hui, Jingxia Gao, Mingyang Li, Qingfeng Zhang, and Yangyang Zhang. 2023. "Optimized Strain Response in (Co0.5Nb0.5)4+-Doped 76Bi0.5Na0.5TiO3-24SrTiO3 Relaxors" Coatings 13, no. 8: 1331. https://doi.org/10.3390/coatings13081331

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