High temperature piezoelectric accelerometer fabricated by 0.75BiFeO3–0.25BaTiO3 ceramics with operating temperature over 450 °C

Owing to the high Curie temperature and good piezoelectric thermal stability, BiFeO3–BaTiO3 ceramics show great potentials for high-temperature piezoelectric sensor applications. In this paper, a compression-mode piezoelectric sensor was fabricated by the lead-free and high-temperature 0.75BiFeO3–0.25BaTiO3–MnO2 (BFBT25–Mn) ceramic and its sensitivity was characterized from room temperature to 550 °C over a frequency range of 200–1000 Hz. The output charge of the BFBT25–Mn piezoelectric sensor is independent of the measuring frequency at different temperatures. The maximum working temperature of the BFBT25–Mn piezoelectric sensor is 450 °C, about 250, 150, and 100 °C higher than those of these piezoelectric sensors fabricated by PZT-5A, BSPT64–Mn, and BSPT66–Mn ceramics, respectively. The temperature sensitivity coefficient from room temperature to 350 °C of the BFBT25–Mn piezoelectric sensor is 30% of that for the BSPT66–Mn sensor. Furthermore, the sensitivity of the BFBT25–Mn piezoelectric sensor is stable with the dwelling time at 400 °C. These results indicate that the BFBT25–Mn ceramic is a strong competitor for high temperature sensing applications.

Piezoelectric accelerometers working above 400 °C are desired in aircraft, automotive smart brakes, and geothermal well drilling tools to measure and record shock, stress, and vibration because of their simple structures, fast response, and easy integration with other parts.
1–31. C. Jiang, Y. Long, F. P. Yu, X. F. Cheng, and X. Zhao, Appl. Phys. Lett. 120(11), 112901 (2022). https://doi.org/10.1063/5.00851442. V. Mortet, R. Peterson, K. Haenen, and M. D'Olieslaeger, Appl. Phys. Lett. 88, 133511 (2006). https://doi.org/10.1063/1.21904623. S. J. Zhang, Y. T. Fei, B. H. T. Bruce, E. Frantz, D. W. Snyder, X. N. Jiang, and T. R. Shrout, Appl. Phys. Lett. 92, 202905 (2008). https://doi.org/10.1063/1.2936276 Quartz crystals, such as Pb(Zr, Ti)O3 (PZT) ferroelectric ceramics, bismuth-layer-structured materials (BLS), and tourmaline or langasite single crystals, are usually used for piezoelectric accelerometers.4–64. Q. Wang and C. M. Wang, J. Am. Ceram. Soc. 103, 2686–2693 (2020). https://doi.org/10.1111/jace.169785. C. Jiang, C. Z. Chang, F. F. Li, L. Sun, Y. L. Li, F. P. Yu, and X. Zhao, J. Mater. Chem. C 10, 180–190 (2022). https://doi.org/10.1039/D1TC03192A6. S. J. Zhang, Y. Q. Zheng, H. K. Kong, J. Xin, E. Frantz, and T. R. Shrout, J. Appl. Phys. 105, 114107 (2009). https://doi.org/10.1063/1.3142429 Although perovskite PZT ceramics show excellent dielectric and piezoelectric properties, their Curie temperatures are less than 400 °C and their usage temperatures are below 200 °C.77. R. Eitel, T. Shrout, and C. Randall, “ Nonlinear contributions to the dielectric permittivity and converse piezoelectric coefficient in piezoelectric ceramics,” J. Appl. Phys. 99(12), 124110 (2006). https://doi.org/10.1063/1.2207738 Piezoelectric sensors fabricated by BLS ferroelectric ceramics and langasite piezoelectric crystals can work above 400 °C.8,98. X. C. Xie, Z. Y. Zhou, B. T. Gao, Z. Y. Zhou, R. H. Liang, and X. L. Dong, ACS. Appl. Mater. Interfaces 14, 12 (2022). https://doi.org/10.1021/acsami.1c194459. S. J. Zhang and F. P. Yu, J. Am. Ceram. Soc. 94(10), 3153–3170 (2011). https://doi.org/10.1111/j.1551-2916.2011.04792.x However, their sensitivities are low due to their inferior dielectric and piezoelectric properties.8,98. X. C. Xie, Z. Y. Zhou, B. T. Gao, Z. Y. Zhou, R. H. Liang, and X. L. Dong, ACS. Appl. Mater. Interfaces 14, 12 (2022). https://doi.org/10.1021/acsami.1c194459. S. J. Zhang and F. P. Yu, J. Am. Ceram. Soc. 94(10), 3153–3170 (2011). https://doi.org/10.1111/j.1551-2916.2011.04792.x The high-temperature piezoelectric 0.36BiScO3–0.64PbTiO3 (BSPT64) solid solution possesses both high Curie temperature of 450 °C and large piezoelectric coefficient of 450 pC/N.1010. J. G. Chen, G. X. Liu, J. R. Cheng, and S. X. Dong, Appl. Phys. Lett. 107(3), 032906 (2015). https://doi.org/10.1063/1.4927328 The resistivity and piezoelectric thermal stability of BS–PT ceramics can be significantly improved by a small amount of MnO2 additions.1111. S. Zhang, R. E. Eitel, C. A. Randall, and T. R. Shrout, Appl. Phys. Lett. 86(26), 262904 (2005). https://doi.org/10.1063/1.1968419 The high temperature resistivity of BSPT66-Mn ceramic is as high as 3 × 107 Ω cm (450 °C), and the shear-mode electromechanical coupling factor remains constant up to 440 °C, showing potential sensing applications at high temperature above 400 °C.1111. S. Zhang, R. E. Eitel, C. A. Randall, and T. R. Shrout, Appl. Phys. Lett. 86(26), 262904 (2005). https://doi.org/10.1063/1.1968419 However, the cost of Sc2O3 is too high, and the toxic lead element in the solid solution brings environment and human health concerns. Thus, lead-free ferroelectric ceramics with high Curie temperature, high piezoelectric activity, and good electric thermal stability are required for high temperature piezoelectric sensor applications.1212. J. G. Chen, B. B. Tong, J. Y. Lin, X. Y. Gao, J. R. Cheng, and S. J. Zhang, J. Eur. Ceram. Soc. 42, 3857–3864 (2022). https://doi.org/10.1016/j.jeurceramsoc.2022.03.052Lead-free BiFeO3 (BF) with high Curie temperature of 825 °C is one of the promising candidates for high-temperature piezoelectric applications.1313. R. Palai, R. S. Katiyar, H. Schmid, P. Tissot, S. J. Clark, J. Robertson, S. A. T. Redfern, G. Catalan, and J. F. Scott, Phys. Rev. B 77, 014110 (2008). https://doi.org/10.1103/PhysRevB.77.014110 However, the poor perovskite phase stability and high conductivity make BF ceramics difficult to be poled.1313. R. Palai, R. S. Katiyar, H. Schmid, P. Tissot, S. J. Clark, J. Robertson, S. A. T. Redfern, G. Catalan, and J. F. Scott, Phys. Rev. B 77, 014110 (2008). https://doi.org/10.1103/PhysRevB.77.014110 Construction of the morphotropic phase boundary by the incorporation of tetragonal structure BaTiO3 (BT), (Bi,K)TiO3 (BKT), or/and PbTiO3 (PT) with rhombohedral phase BF is an effective method to stabilize the perovskite structure and improve the electric properties.14–1714. S. O. Leontsev and R. E. Eitel, J. Am. Ceram. Soc. 92, 2957 (2009). https://doi.org/10.1111/j.1551-2916.2009.03313.x15. D. W. Wang, G. Wang, S. Murakami, Z. M. Fan, A. Feteira, D. Zhou, S. K. Sun, Q. L. Zhao, and I. M. Reaney, J. Adv. Dielectr. 08, 1830004 (2018). https://doi.org/10.1142/S2010135X1830004916. Z. H. Ning, Y. Jiang, J. Jian, J. Guo, J. R. Cheng, H. W. Cheng, J. G. Chen, and J. Euro, Ceram. Soc. 40, 2338–2344 (2020). https://doi.org/10.1016/j.jeurceramsoc.2020.01.05917. A. J. Bell, T. P. Comyn, and T. J. Stevenson, APL Mater. 9, 010901 (2021). https://doi.org/10.1063/5.0035416 Among these solid solutions, BF–BT, BF–PT–BT, and BF–PT–BKT systems show both the high Curie temperature of above 500 °C and good piezoelectric properties.14–1714. S. O. Leontsev and R. E. Eitel, J. Am. Ceram. Soc. 92, 2957 (2009). https://doi.org/10.1111/j.1551-2916.2009.03313.x15. D. W. Wang, G. Wang, S. Murakami, Z. M. Fan, A. Feteira, D. Zhou, S. K. Sun, Q. L. Zhao, and I. M. Reaney, J. Adv. Dielectr. 08, 1830004 (2018). https://doi.org/10.1142/S2010135X1830004916. Z. H. Ning, Y. Jiang, J. Jian, J. Guo, J. R. Cheng, H. W. Cheng, J. G. Chen, and J. Euro, Ceram. Soc. 40, 2338–2344 (2020). https://doi.org/10.1016/j.jeurceramsoc.2020.01.05917. A. J. Bell, T. P. Comyn, and T. J. Stevenson, APL Mater. 9, 010901 (2021). https://doi.org/10.1063/5.0035416 Furthermore, addition of a small amount of MnO2 can increase grain boundary resistivity and then improve the room temperature direct current (DC) resistivity of BF-based ceramics. For example, 0.6 wt. % MnO2 addition can improve the DC resistivity of BFBT25–0.6 wt. %Mn ceramic from 107 to 1012 Ω cm, which is much higher than that of PZT5A ceramic.14,1514. S. O. Leontsev and R. E. Eitel, J. Am. Ceram. Soc. 92, 2957 (2009). https://doi.org/10.1111/j.1551-2916.2009.03313.x15. D. W. Wang, G. Wang, S. Murakami, Z. M. Fan, A. Feteira, D. Zhou, S. K. Sun, Q. L. Zhao, and I. M. Reaney, J. Adv. Dielectr. 08, 1830004 (2018). https://doi.org/10.1142/S2010135X18300049 The Curie temperature and piezoelectric coefficient of BFBT25–0.6 wt. % Mn ceramic are 603 °C and 116 pC/N, respectively.14,1514. S. O. Leontsev and R. E. Eitel, J. Am. Ceram. Soc. 92, 2957 (2009). https://doi.org/10.1111/j.1551-2916.2009.03313.x15. D. W. Wang, G. Wang, S. Murakami, Z. M. Fan, A. Feteira, D. Zhou, S. K. Sun, Q. L. Zhao, and I. M. Reaney, J. Adv. Dielectr. 08, 1830004 (2018). https://doi.org/10.1142/S2010135X18300049 The piezoelectric coefficient of BFBT25–0.6 wt. % Mn ceramic is much higher than those of BLS ferroelectric ceramics.1818. J. Guo, J. G. Chen, J. R. Cheng, and Q. Tan, J. Am. Ceram. Soc. 104(11), 5547–5556 (2021). https://doi.org/10.1111/jace.17972 Furthermore, there are no phase transitions from room temperature to Curie temperature like in BF–BT ceramics, which is quite different from those of lead-free piezoelectric ceramics [(Bi, Na)TiO3–BaTiO3 (BNT–BT), (K, Na)NbO3 (KNN)], showing good piezoelectric thermal stability.19–2119. J. Rodel, K. G. Webber, R. Dittmer, W. Jo, and M. Kimura, J. Eur. Ceram. Soc. 35, 1659 (2015). https://doi.org/10.1016/j.jeurceramsoc.2014.12.01320. J. Rodel, W. Jo, K. T. P. Seifert, E. M. Anton, T. Granzow, and D. Damjanovic, J. Am. Ceram. Soc. 92, 1153 (2009). https://doi.org/10.1111/j.1551-2916.2009.03061.x21. G. Catalan and J. F. Scott, Adv. Mater. 21, 2463 (2009). https://doi.org/10.1002/adma.200802849 Our previous work found that the depolarization temperature and aging ratio of BFBT25–1mol%Mn ceramic were 500 °C and 1.2%, respectively, which were better than those of PZT and BS–PT ceramics.2222. J. Chen, J. Cheng, J. Guo, Z. Cheng, J. Wang, H. Liu, and S. Zhang, J. Am. Ceram. Soc. 103, 374–381 (2020). https://doi.org/10.1111/jace.16755 The good electric thermal stability and low aging ratio of BFBT25–1mol%Mn ceramic are owing to the high intrinsic lattice distortion of R3c space group. In this paper, a high temperature piezoelectric compression-mode accelerometer was fabricated by the BFBT25–Mn ceramic, and the high temperature performances were characterized from room temperature to 550 °C. For comparison, 0.57Pb(Sc0.5Nb0.5)O3–0.43PbTiO3–MnO2 (PSNT43–Mn), PZT-5A, BSPT64–Mn, and BSPT66–Mn ceramics were also used to fabricate the compression-mode sensors of the same structure.The BFBT25–Mn ceramic was prepared by the high temperature sintering method. The detailed preparation processes were reported in our previous literature.2222. J. Chen, J. Cheng, J. Guo, Z. Cheng, J. Wang, H. Liu, and S. Zhang, J. Am. Ceram. Soc. 103, 374–381 (2020). https://doi.org/10.1111/jace.16755 The raw materials of Bi2O3, Fe2O3, BaCO3, TiO2, and MnO2 are reagent-grade. The raw materials were ball milled for 6 h and calcined at 750 °C for 4 h. Then, the powders were pressed into pellets with a polyvinyl alcohol binder. After burning out the binder, the green pellets were sintered at 1000 °C for 3 h in the oxygen atmosphere. The phase structure and micro-structure of the BFBT25-Mn ceramic were characterized by x-ray diffraction (XRD, Rigaku-D/MAX-2000) and scanning electron microscopy (SEM, Zeiss G300). The piezoelectric ceramics were poled along the thickness direction under the electric field dependent on their coercive field for 30 min in the oil bath at 120 °C. The dielectric and piezoelectric properties of the BFBT25-Mn ceramic were characterized by the Agilent impedance analyzer (Agilent 4294A) and quasi-static piezoelectric meter (ZJ-3D, China), respectively. A piezoelectric ceramic pellet with a thickness of 1.2 mm and a diameter of 10 mm was used for sensor preparation. The frequency constant for a thickness mode of BFBT25–Mn ceramics is 1540 Hzm. The DC resistivity measured by a source meter (Keithley 2410C) of BFBT25–Mn ceramics at 450 °C is 5.6 MΩ cm. Figure 1(a) shows the schematic diagram of the piezoelectric sensor. Inconel and high purity alumina were chosen as mass and electrical insulation blocks, respectively. The weight of the mass is 28 g. A pair of springs was use to provide preload. The piezoelectric ceramic pellets were placed between the two insulation blocks. Two silver wires were fixed on the surface of blocks, and alumina tubes were sleeved on the silver wires. When a dynamic stress is applied on the ceramics, the electric potential will generate due to the direct piezoelectric effect, which is simulated by the finite element method (COMSOL Multiphysics software), as shown in Figs. 1(b) and 1(c). The piezoelectric ceramic suffers from stress normal under the excitation of mechanical vibration, and the sinusoidal output voltage that related to the harmonic mechanical vibration can be obtained. The theoretical output of a piezoelectric sensor is proportional to the product of the piezoelectric coefficient, weight of the mass, and its acceleration. The experimental setup for measuring the high temperature sensor is given in Fig. 2. The high temperature sensor was placed in a vertical tube furnace and connected with a vibration exciter by a screw. A function generator (Model DS340, Stanford Research Systems) generated the sinusoidal signal, which was amplified by a power amplifier (Type YE5872A, Sinocera Piezotronics, Inc.) to generate the desired vibration. A commercial accelerometer (CA-YD-182‐10, Sinocera Piezotronics, Inc.) was used to measure the acceleration of the vibration exciter. The output signal from the high temperature sensors was recorded by a lock-in amplifier (Model SR830, Stanford Research Systems). These high temperature piezoelectric sensors were measured from room temperature to 550 °C in the frequency range from 200 to 1000 Hz.The XRD pattern of the BFBT25–Mn ceramic is given in Fig. 3(a). The BFBT25–Mn ceramic shows a pure perovskite structure, and no other secondary phases can be detected. The clear splitting of PC at the 2θ of 39° implies that the BFBT25–Mn ceramic is crystallized into a rhombohedral phase. The fresh fracture surface of the BFBT25–Mn ceramic is given in Fig. 3(b). The ceramic is dense without obvious porosity, and the average grain size is about 5.6 μm. In addition, the grain boundary is very clean, indicating no secondary phases, which is consistent with the XRD result. Figure 3(c) presents the temperature dependence of the dielectric constant and loss at 10 kHz. The dielectric constant and loss are stable below 450 °C, and the Curie temperature is 630 °C. The planar electromechanical coupling factor kp as a function of temperature is present in Fig. 3(d). It is interesting that the kp is stable from room temperature to 540 °C, indicating the good piezoelectric thermal stability. Table I compares the Curie temperature, dielectric, and piezoelectric properties of the BFBT25-Mn ceramic with the other typical piezoelectric ceramics (PSNT43–Mn, PZT-5A, BSPT64–Mn, and BSPT66–Mn) with the Curie temperatures from 250 to 460 °C.22–2522. J. Chen, J. Cheng, J. Guo, Z. Cheng, J. Wang, H. Liu, and S. Zhang, J. Am. Ceram. Soc. 103, 374–381 (2020). https://doi.org/10.1111/jace.1675523. S. J. Zhang, R. Xia, L. Lebrun, D. Anderson, and T. R. Shrout, Mater. Lett. 59(27), 3471–3475 (2005). https://doi.org/10.1016/j.matlet.2005.06.01624. G. H. Haertling, J. Am. Ceram. Soc. 82(4), 797–818 (1999). https://doi.org/10.1111/j.1151-2916.1999.tb01840.x25. S. J. Zhang, E. F. Alberta, C. A. Randall, and T. R. Shrout, IEEE Trans. Ultrason Ferroelectr. Freq. Control 52(11), 2131–2139 (2005). https://doi.org/10.1109/TUFFC.2005.1561684 Although the dielectric and piezoelectric properties of the BFBT25–Mn ceramic are inferior to those of other piezoelectric ceramics, its Curie temperature of 630 °C is much higher. It is noted that piezoelectric voltage coefficient (i.e., g33) of the BFBT25–Mn ceramic is comparable to those of other piezoelectric ceramics.Table icon

TABLE I. Curie temperature, dielectric, and piezoelectric properties of selected typical piezoelectric ceramics for high temperature piezoelectric sensors.

MaterialsTc (oC)d33 (pC/N)kpɛ (1 kHz)Tanδ (1 kHz)g33 (10−3 V m/N)ReferenceBFBT25–Mn (this work)6301140.34700.0327.22222. J. Chen, J. Cheng, J. Guo, Z. Cheng, J. Wang, H. Liu, and S. Zhang, J. Am. Ceram. Soc. 103, 374–381 (2020). https://doi.org/10.1111/jace.16755BSPT66–Mn4562550.4812500.0123.12323. S. J. Zhang, R. Xia, L. Lebrun, D. Anderson, and T. R. Shrout, Mater. Lett. 59(27), 3471–3475 (2005). https://doi.org/10.1016/j.matlet.2005.06.016BSPT64–Mn4423600.5614500.0128.02323. S. J. Zhang, R. Xia, L. Lebrun, D. Anderson, and T. R. Shrout, Mater. Lett. 59(27), 3471–3475 (2005). https://doi.org/10.1016/j.matlet.2005.06.016PZT5A3653740.6017000.0124.92424. G. H. Haertling, J. Am. Ceram. Soc. 82(4), 797–818 (1999). https://doi.org/10.1111/j.1151-2916.1999.tb01840.xPSNT43–Mn2655000.6427500.0220.52525. S. J. Zhang, E. F. Alberta, C. A. Randall, and T. R. Shrout, IEEE Trans. Ultrason Ferroelectr. Freq. Control 52(11), 2131–2139 (2005). https://doi.org/10.1109/TUFFC.2005.1561684Figures 4(a)–4(f) illustrate the temperature dependent sensor charge of the BFBT25–Mn ceramic as a function of acceleration under different frequencies. The sensor charge increases linearly with the increase in the acceleration. It is noted that the sensor charge is not sensitive to the measuring frequency (200–1000 Hz) because it is much lower than the resonate frequency (1.3–1.4 MHz) of the piezoelectric sensor. The sensor charge of the BFBT25–Mn ceramic at 1 g acceleration and 1000 Hz is 49 pC. The sensor charge is sensitive to the temperature. When the measuring temperature increases from room temperature to 450 °C, the sensor charge at 1 g and 1000 Hz increases from 49 to 115 pC. The sensor charge is proportional to the piezoelectric charge coefficient of the piezoelectric ceramic. As well known, the piezoelectric properties of ferroelectric materials origin from the lattice distortion, domain motion, and phase transition.2626. H. Kungl, T. Fett, S. Wagner, and M. Hoffmann, J Appl. Phys. 101(4), 044101 (2007). https://doi.org/10.1063/1.2434836 With the increase in the testing temperature, the ferroelectric domain becomes easier to switch, leading to the enhanced the piezoelectric charge coefficient. Thereafter, the sensor charge of the piezoelectric sensor increases with the temperature.Figures 5(a) and 5(b) compare the temperature dependent sensitivity of BFBT25–Mn sensors with the other piezoelectric sensors at 1 g acceleration and 1000 Hz. The sensitivity of the sensor can be derived from the slope of the acceleration dependent sensor charge curves shown in Fig. 5. The sensitivity of piezoelectric sensors initially increases with the increased temperature, reaches up to a maximum value, and then drops dramatically, which is similar to these curves of the temperature dependent piezoelectric charge coefficient of the piezoelectric ceramics. The maximum working temperatures of piezoelectric sensors fabricated by the different piezoelectric ceramics are compared in Fig. 5(b). The ultimate temperature of the BFBT25–Mn piezoelectric sensor is as high as 450 °C, about 300 °C, 250, 150, and 100 °C higher than those of PSNT–Mn, PZT-5A, BSPT64–Mn, and BSPT66–Mn ceramics, respectively. The sensitivity of the BFBT25–Mn piezoelectric sensor at 450 °C is 115 pC/g. The planar electromechanical coupling factor kp of the BFBT25–Mn ceramic drops at 550 °C, about 100 °C higher than the ultimate temperature of the sensor. The high conductivity of the BFBT25–Mn ceramic may increase dramatically above 450 °C, which make the generated charge from the dynamic stress not able to hold enough time to be detected. Further investigation finds that the sensitivity of the BFBT25–Mn piezoelectric sensor is stabler with temperature than those of other piezoelectric ceramics, showing better thermal stability. For example, the output signal variation of the BFBT25–Mn piezoelectric ceramic sensor from room temperature to 350 °C is about 80%, which is only 30% that of the BSPT66–Mn ceramic sensor. The BFBT25–Mn ceramic shows the single rhombohedral phase of the R3c symmetry, with the oxygen octahedral titling along the [111] axis, whereas the other ferroelectric ceramics exhibit the coexistence of rhombohedral and tetragonal phases.22–2422. J. Chen, J. Cheng, J. Guo, Z. Cheng, J. Wang, H. Liu, and S. Zhang, J. Am. Ceram. Soc. 103, 374–381 (2020). https://doi.org/10.1111/jace.1675523. S. J. Zhang, R. Xia, L. Lebrun, D. Anderson, and T. R. Shrout, Mater. Lett. 59(27), 3471–3475 (2005). https://doi.org/10.1016/j.matlet.2005.06.01624. G. H. Haertling, J. Am. Ceram. Soc. 82(4), 797–818 (1999). https://doi.org/10.1111/j.1151-2916.1999.tb01840.x Due to the large lattice distortion of R3c symmetry, the domain structure of the BFBT25–Mn ceramic is much stable, and its piezoelectric properties are stable with the temperature. However, the ferroelectric domains in the other materials with two or more phases are sensitive to temperature, electric field, and stress, showing good piezoelectric properties but bad electrical thermal stability. Thereafter, the piezoelectric sensor fabricated by the BFBT25–Mn ceramic shows a better thermal stability. In order to estimate the survivability and reliability of the piezoelectric sensor at the high temperature, the piezoelectric sensors were held in the furnace at the different high temperatures for 5 h, and the data were collected every half hour at 1 g and 1000 Hz. Figures 5(c) and 5(d) present the sensitivities of the BFBT25–Mn and BSPT66–Mn piezoelectric sensors as a function of the dwelling time. The sensitivity of the BFBT25–Mn sensor is stable with time when the temperature is as high as 400 °C, which is 100 °C higher than that of BSPT66–Mn piezoelectric sensor.

In summary, a compression-mode accelerometer prototype fabricated by the BFBT25–Mn high-temperature and lead-free ceramic was characterized as a function of temperature from 25 to 550 °C over the frequency range of 200–1000 Hz. The maximum working temperature of the BFBT25–Mn piezoelectric sensor is as high as 450 °C. The thermal stability of the BFBT25–Mn piezoelectric sensor is better than those of PSNT43–Mn, PZT-5A BSPT64–Mn, and BSPT66–Mn piezoelectric sensors. The high working temperature, good thermal stability, and durability with time indicate that the BFBT25–Mn ceramic has great potentials for high temperature sensing applications.

This work was supported by the Open Fund for National Key Laboratory of Science and Technology on Underwater Acoustic Antagonizing (Grant No. JCKY2020207CH02), the National Natural Science Foundation of China (Grant No. 51872180), and the Guangdong Provincial Key Laboratory of Materials and Technologies for Energy Conversion (Grant No. MATEC2022KF002).

Conflict of Interest

The authors have no conflicts to disclose.

Author Contributions

Jianguo Chen: Conceptualization (equal); Funding acquisition (equal); Writing – review & editing (equal). Jingen Wu: Conceptualization (equal). Yun Lu: Data curation (equal). Yan Wang: Writing – review & editing (equal). Jinrong Cheng: Funding acquisition (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

REFERENCES

1. C. Jiang, Y. Long, F. P. Yu, X. F. Cheng, and X. Zhao, Appl. Phys. Lett. 120(11), 112901 (2022). https://doi.org/10.1063/5.0085144, Google ScholarScitation2. V. Mortet, R. Peterson, K. Haenen, and M. D'Olieslaeger, Appl. Phys. Lett. 88, 133511 (2006). https://doi.org/10.1063/1.2190462, Google ScholarScitation, ISI3. S. J. Zhang, Y. T. Fei, B. H. T. Bruce, E. Frantz, D. W. Snyder, X. N. Jiang, and T. R. Shrout, Appl. Phys. Lett. 92, 202905 (2008). https://doi.org/10.1063/1.2936276, Google ScholarScitation, ISI4. Q. Wang and C. M. Wang, J. Am. Ceram. Soc. 103, 2686–2693 (2020). https://doi.org/10.1111/jace.16978, Google ScholarCrossref5. C. Jiang, C. Z. Chang, F. F. Li, L. Sun, Y. L. Li, F. P. Yu, and X. Zhao, J. Mater. Chem. C 10, 180–190 (2022). https://doi.org/10.1039/D1TC03192A, Google ScholarCrossref6. S. J. Zhang, Y. Q. Zheng, H. K. Kong, J. Xin, E. Frantz, and T. R. Shrout, J. Appl. Phys. 105, 114107 (2009). https://doi.org/10.1063/1.3142429, Google ScholarScitation, ISI7. R. Eitel, T. Shrout, and C. Randall, “ Nonlinear contributions to the dielectric permittivity and converse piezoelectric coefficient in piezoelectric ceramics,” J. Appl. Phys. 99(12), 124110 (2006). https://doi.org/10.1063/1.2207738, Google ScholarScitation, ISI8. X. C. Xie, Z. Y. Zhou, B. T. Gao, Z. Y. Zhou, R. H. Liang, and X. L. Dong, ACS. Appl. Mater. Interfaces 14, 12 (2022). https://doi.org/10.1021/acsami.1c19445, Google ScholarCrossref9. S. J. Zhang and F. P. Yu, J. Am. Ceram. Soc. 94(10), 3153–3170 (2011). https://doi.org/10.1111/j.1551-2916.2011.04792.x, Google ScholarCrossref, ISI10. J. G. Chen, G. X. Liu, J. R. Cheng, and S. X. Dong, Appl. Phys. Lett. 107(3), 032906 (2015). https://doi.org/10.1063/1.4927328, Google ScholarScitation, ISI11. S. Zhang, R. E. Eitel, C. A. Randall, and T. R. Shrout, Appl. Phys. Lett. 86(26), 262904 (2005). https://doi.org/10.1063/1.1968419, Google ScholarScitation, ISI12. J. G. Chen, B. B. Tong, J. Y. Lin, X. Y. Gao, J. R. Cheng, and S. J. Zhang, J. Eur. Ceram. Soc. 42, 3857–3864 (2022). https://doi.org/10.1016/j.jeurceramsoc.2022.03.052, Google ScholarCrossref13. R. Palai, R. S. Katiyar, H. Schmid, P. Tissot, S. J. Clark, J. Robertson, S. A. T. Redfern, G. Catalan, and J. F. Scott, Phys. Rev. B 77, 014110 (2008). https://doi.org/10.1103/PhysRevB.77.014110, Google ScholarCrossref, ISI14. S. O. Leontsev and R. E. Eitel, J. Am. Ceram. Soc. 92, 2957 (2009). https://doi.org/10.1111/j.1551-2916.2009.03313.x, Google ScholarCrossref, ISI15. D. W. Wang, G. Wang, S. Murakami, Z. M. Fan, A. Feteira, D. Zhou, S. K. Sun, Q. L. Zhao, and I. M. Reaney, J. Adv. Dielectr. 08, 1830004 (2018). https://doi.org/10.1142/S2010135X18300049, Google ScholarCrossref16. Z. H. Ning, Y. Jiang, J. Jian, J. Guo, J. R. Cheng, H. W. Cheng, J. G. Chen, and J. Euro, Ceram. Soc. 40, 2338–2344 (2020). https://doi.org/10.1016/j.jeurceramsoc.2020.01.059, Google ScholarCrossref17. A. J. Bell, T. P. Comyn, and T. J. Stevenson, APL Mater. 9, 010901 (2021). https://doi.org/10.1063/5.0035416, Google ScholarScitation, ISI18. J. Guo, J. G. Chen, J. R. Cheng, and Q. Tan, J. Am. Ceram. Soc. 104(11), 5547–5556 (2021). https://doi.org/10.1111/jace.17972, Google ScholarCrossref19. J. Rodel, K. G. Webber, R. Dittmer, W. Jo, and M. Kimura, J. Eur. Ceram. Soc. 35, 1659 (2015). https://doi.org/10.1016/j.jeurceramsoc.2014.12.013, Google ScholarCrossref, ISI20. J. Rodel, W. Jo, K. T. P. Seifert, E. M. Anton, T. Granzow, and D. Damjanovic, J. Am. Ceram. Soc. 92, 1153 (2009). https://doi.org/10.1111/j.1551-2916.2009.03061.x, Google ScholarCrossref, ISI21. G. Catalan and J. F. Scott, Adv. Mater. 21, 2463 (2009). https://doi.org/10.1002/adma.200802849, Google ScholarCrossref, ISI22. J. Chen, J. Cheng, J. Guo, Z. Cheng, J. Wang, H. Liu, and S. Zhang, J. Am. Ceram. Soc. 103, 374–381 (2020). https://doi.org/10.1111/jace.16755, Google ScholarCrossref23. S. J. Zhang, R. Xia, L. Lebrun, D. Anderson, and T. R. Shrout, Mater. Lett. 59(27), 3471–3475 (2005). https://doi.org/10.1016/j.matlet.2005.06.016, Google ScholarCrossref24. G. H. Haertling, J. Am. Ceram. Soc. 82(4), 797–818 (1999).

留言 (0)

沒有登入
gif