Next Article in Journal
Enhancing High-Alloy Steel Cutting with Abrasive Water Injection Jet (AWIJ) Technology: An Approach Using the Response Surface Methodology (RSM)
Previous Article in Journal
Suitability Study of Optical Coordinate Measuring Machine for Quality Assessment and Wear Phenomena Identification of Blade Edge and Surface of Planer Technical Knives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Energy Storage Performance of 0.85Bi0.5Na0.5TiO3-0.15LaFeO3 Lead-Free Ferroelectric Ceramics via Buried Sintering

1
School of Materials Science and Physics, China University of Mining and Technology, Xuzhou 221116, China
2
Key Laboratory of Inorganic Functional Materials and Devices, Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, China
3
Key Laboratory of Polar Materials and Devices, Ministry of Education, East China Normal University, Shanghai 200241, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(16), 4019; https://doi.org/10.3390/ma17164019
Submission received: 8 July 2024 / Revised: 26 July 2024 / Accepted: 7 August 2024 / Published: 13 August 2024
(This article belongs to the Special Issue Ferroelectric, Magnetic, and Multiferroic Materials and Applications)

Abstract

:
Bismuth sodium titanate (Bi0.5Na0.5TiO3, BNT) ceramics are expected to replace traditional lead-based materials because of their excellent ferroelectric and piezoelectric characteristics, and they are widely used in the industrial, military, and medical fields. However, BNT ceramics have a low breakdown field strength, which leads to unsatisfactory energy storage performance. In this work, 0.85Bi0.5Na0.5TiO3-0.15LaFeO3 ceramics are prepared by the traditional high-temperature solid-phase reaction method, and their energy storage performance is greatly enhanced by improving the process of buried sintering. The results show that the buried sintering method can inhibit the formation of oxygen vacancy, reduce the volatilization of Bi2O3, and greatly improve the breakdown field strength of the ceramics so that the energy storage performance can be significantly enhanced. The breakdown field strength increases from 210 kV/cm to 310 kV/cm, and the energy storage density increases from 1.759 J/cm3 to 4.923 J/cm3. In addition, the energy storage density and energy storage efficiency of these ceramics have good frequency stability and temperature stability. In this study, the excellent energy storage performance of the ceramics prepared by the buried sintering method provides an effective idea for the design of lead-free ferroelectric ceramics with high energy storage performance and greatly expands its application field.

Graphical Abstract

1. Introduction

Ferroelectric materials are widely used in the detection, conversion, processing, and storage of various kinds of information because of their excellent ferroelectric and piezoelectric properties [1,2,3,4,5]. However, traditional ferroelectric ceramics contain the Pb element, which will cause serious harm to the human body and the environment in the production, use, and waste of ceramics [6,7]. Among many lead-free ferroelectric materials, because Bi2+ and Pb2+ have a similar single-pair electron 6S2 structure, Bi0.5Na0.5TiO3 (BNT) ceramic has excellent ferroelectric properties [8]. BNT-based ceramics are considered to be one of the most likely to replace lead-free ferroelectric ceramics [9,10]. However, the low breakdown electric field strength and poor energy storage performance of pure BNT ceramic limit its application in electrical fields [11]. In order to improve the electrical properties of BNT ceramics, the modification of BNT-based ferroelectric ceramics has become the main research direction in the ferroelectric field. Some of these new BNT-based ferroelectric materials are already well known. For example, Jiang et al. [12] constructed 0.8Bi0.5Na0.5TiO3-0.2Ba0.3Sr0.4TiO3 and mixed 0.1NaNbO3 on this basis to obtain a good solid solution structure. A high energy storage density of 2.26 J/cm3 at 180 kV/cm was obtained, and the relaxation characteristics and dielectric temperature stability of the material were enhanced. Guo et al. [13] engineered and synthesized (1−x)(0.94Bi0.5Na0.5TiO3-0.06BaTiO3)-xBiMg2/3Nb1/3O3 solid solution to achieve the co-existence of tetragonal- and rhombohedral-phase PNRs in the perovskite structure, and an ultra-high energy density of 6.3 J/cm3 and an energy efficiency of 79.6% were obtained. Zhang et al. [14] found that the La element can improve the energy storage density and efficiency of 0.93 (Bi0.5Na0.5)TiO3-0.07Ba (Ti0.945Zr0.055)O3 ceramics. Gong et al. [15] found that Bi0.9La0.1FeO3 ceramics have a low leakage current density and high saturation polarization. At present, a large number of research works have made certain progress, such as controlling the sintering temperature [16,17], using a N2 atmosphere for sintering [18], adding the sintering aid CuO [19,20], and other methods. Improving the preparation process, such as the sintering process, is also expected to improve the performance of ceramics. Buried sintering refers to when the ceramic structure is buried under something (powder, etc.) for sintering [21,22]. For example, Fujii et al. [23] found that sintering BaTiO3-Bi(Mg1/2Ti1/2)O3-BiFeO3 ceramic samples in a bismuth-rich atmosphere, that is, sintering in a closed calcination powder crucible with the same composition, inhibited the volatilization of Bi2O3, thus enhancing the electrical properties of the ceramics. Li et al. [24] prepared Li3+xMg2NbO6 ceramics with excellent microwave dielectric properties using the buried powder sintering method. The purpose of the buried sintering method is to prevent the material loss, volatilization, and combustion of the ceramic at a high temperature during sintering, resulting in the loss of bonding and disintegration of the billet and the inability to maintain its shape. Secondly, buried sintering can have the effect of isolating the air, which can prevent the reaction of the ceramic powder and the components in the air at high temperatures.
Therefore, in this study, 0.85Bi0.5Na0.5TiO3-0.15LaFeO3 (BNT-LFO) binary solid solution is constructed to study the relationship between its microstructure and electrical properties. More importantly, compared with ordinary sintering, the micro-structure of ceramics is optimized by using buried sintering, which greatly improves the energy storage properties of BNT-LFO ceramics. Besides doping modification, it provides a new design idea for improving the energy storage performance of BNT-based ceramics.

2. Materials and Methods

BNT-LFO ceramics are prepared by the high-temperature solid reaction method. The experimental raw materials are Bi2O3 (99%), Na2CO3 (99.9%), TiO2 (99%), La2O3 (99.99%), and Fe2O3 (99.9%). The experimental steps are as follows: According to the stoichiometric ratio, ball milling is carried out for 8 h. After drying, the pre-burning temperature is 650 °C, heat preservation is carried out for 4 h, and the heating rate is 3 °C/min. Next, 5%PVA is used for granulation, and then tablet pressing is carried out (applied pressure: 0.6 Pa, ceramic area: 12.56 mm2, and thickness: 0.3 mm). The heat is kept at 650 °C for 3 h to discharge the glue (PVA), and then it is raised to 1100 °C for 2 h for sintering, and the heating rate is 3 °C/min, and then it is naturally cooled to room temperature. The powder used for buried sintering is made of pre-fired powder. The crystal structures of the ceramic samples are analyzed by X-ray diffractometry (XRD, D8 ADVANCE) with Cu Kα1 radiation (λ = 1.5406 Å). The surface morphology of the samples is characterized by scanning electron microscopy (SEM, JEOL JEM-7800F, Akishima, Japan). The dielectric test system (Wayne Kerr 6500B, Shenzhen, China) is used to test the dielectric properties of ceramics at room temperature. The ferroelectric and energy storage properties of ceramics are tested by a ferroelectric analysis facility (TF-3000, aixACCT, Aachen, Germany) at room temperature, and variable temperature energy storage is carried out at 20–100 °C.

3. Results and Discussion

The XRD patterns of BNT-LFO ceramics under ordinary sintering and buried sintering are shown in Figure 1a. All ceramics have a pure perovskite structure without introducing the second phase, indicating that LFO has completely entered into the BNT lattice, forming a good solid solution. The main diffraction peak (110) is amplified, as shown in Figure 1b, and it is found that the diffraction peak position of the buried sintered ceramic is shifted to a lower angle compared with that of the ordinary sintered ceramic. According to the Bragg equation, the spacing between the crystal faces increases. Buried sintering can reduce the evaporation of Bi2O3 and the generation of oxygen vacancy, and the smaller the vacancy, the larger the crystal plane spacing. Therefore, the distance between the ceramic crystal faces after buried sintering is larger, and the position of the main diffraction peak is slightly shifted to the left side. In addition, the bottom of the XRD diffraction peak of the buried sintered ceramic shows a wider peak, which is caused by the reduction in the grain size of the ceramic after buried sintering.
The surface microscopic morphologies of the ordinary sintered and buried sintered BNT-LFO ceramics are shown in Figure 2. The insets are the grain distribution data obtained by using the Nano Measurer (2009) software and origin (2018) software. The mean grain size of the buried sintered ceramic is 1.28 μm, which is significantly lower than that of the ordinary sintered ceramic (1.50 μm), showing a reduction of 14.7%. In addition, the grain size distribution of the ceramics prepared by the buried sintering method is more concentrated, the grain size is more uniform, and the number of abnormally coarse grains is significantly reduced, which is expected to improve the electrical properties of BNT-LFO ferroelectric ceramics.
The dielectric constant (εr) and dielectric loss (tanδ) curves of the ordinary sintered and buried sintered BNT-BFO ceramics are shown in Figure 3. The measurement frequencies are 1 kHz, 10 kHz, 100 kHz, and 1 MHz. Two dielectric anomaly peaks can be found, where Ts occurs at lower temperatures (~100 °C), which is believed to be characteristic of relaxed ferroelectrics and is associated with thermal relaxation in the PNRs. Tm occurs at higher temperatures (~380 °C), which corresponds to the transition from the rhombohedral phase to the tetragonal phase [25,26,27]. The εr of the buried sintered ceramic decreased, mainly because this method inhibits the volatilization of Bi2O3, which causes charge fluctuation. After the ceramic undergoes buried sintering, the tanδ is reduced, which can effectively reduce the energy loss, providing the basis for obtaining excellent electrical properties.
At room temperature, the bipolar P-E loops of the ceramics under different electric fields are shown in Figure 4a,b. The bipolar P-E loop of the ceramics under a 150 KV/cm electric field is summarized in Figure 4c. The ferroelectric property of the buried sintered ceramic is obviously better than that of the ordinary sintered one. The key parameters, maximum polarization (Pmax), remanent polarization (Pr), and ΔP (Pmax − Pr) are summarized, as shown in Figure 4d. By comparison, it is found that the Pmax value increased from 12.422 µC/cm2 to 13.793 µC/cm2, and the Pr value decreased from 0.880 µC/cm2 to 0.708 µC/cm2 for the buried sintered BNT-LFO ceramics. ΔP increased from 11.542 µC/cm2 to 13.086 µC/cm2. The increase in ΔP is expected to improve the energy storage performance of the ceramics.
In order to explore the influence of the buried sintering method on the energy storage performance of BNT-LFO ceramics, we tested the unipolar P-E loops of ordinary sintered and buried sintered ceramics, and the test frequency was 10 Hz. The test results are shown in Figure 5a,b, respectively. The breakdown electric field strength of the buried sintered ceramic is up to 310 kV/cm, which is 1.48 times as large as that of the ordinary sintered ceramic (210 kV/cm). The unipolar P-E loops of the ordinary sintered and buried sintered ceramics under their maximum electric field is shown in Figure 5c. Pmax increases from 18.741 µC/cm2 (ordinary sintering) to 36.968 µC/cm2 (buried sintering), which is an increase of 97.3%. According to the unipolar P-E loops, the discharge energy storage density (Wrec) and energy storage efficiency (η) are calculated. The total energy storage density (W), Wrec, and η of the ferroelectric materials are calculated as follows [28]:
W = 0 P m a x E d P
W rec   P r P m a x E d P
η = (Wrec/W) × 100%
The Wrec and η values of the ordinary sintered and buried sintered ceramics are shown in Figure 5d. After buried sintering, the Wrec of the BNT-LFO ceramics can be increased from 1.759 J/cm3 to 4.923 J/cm3, which is a huge increase of 179.9%. The η decreased slightly from 81.8% to 77.4%, which is a decrease of about 5.7%. The microstructure of the buried sintered BNT-LFO ceramic is more uniform and compact, and the average grain size decreases, so it shows a larger breakdown electric field strength than ordinary ceramic. The Wrec and η values of the ceramics are improved. This result shows that BNT-LFO ceramics can be changed by buried sintering to obtain better energy storage performance, which greatly expands its application in the field of electricity.
The Wrec of buried sintered BNT-LFO ceramic is compared with other previously published lead-free BNT-based ceramics, as shown in Figure 6 [29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52]. The energy storage ability is compared between this work and other previously published lead-free BNT-based ceramics, and the results are shown in Table 1. The breakdown electric field strength of ceramic can reach 310 kV/cm, and the Wrec is as high as 4.923 J/cm3. The results show that the buried sintering progress is a good strategy for improving the energy storage performance of lead-free ferroelectric ceramics.
Excellent temperature stability and frequency stability are the prerequisites for the material to ensure the stable operation of the device in practical applications. Therefore, we test the energy storage temperature stability and frequency stability of the buried sintered BNT-LFO ceramic. Figure 7a shows the unipolar P-E loops of the ceramic at different temperatures. The change curves of Wrec and η under the electric field of 150 kV/cm with the increase in temperature are shown in Figure 7b. In the temperature range of 30 °C to 100 °C, the fluctuation in Wrec is less than 2.7%, and the fluctuation in η is less than 13.0%. Figure 7c shows the unipolar P-E loops of the ceramic at different frequencies. The change curves of Wrec and η under the electric field of 150 kV/cm with the increase in frequency are shown in Figure 7d. The fluctuations in Wrec and η are less than 3.5% and 9.5%, respectively. The results show that the ceramics have excellent energy storage temperature stability and frequency stability.

4. Conclusions

In summary, BNT-LFO ceramics are prepared by the traditional high-temperature solid state reaction method, and the ceramics are made into porcelain by ordinary sintering and buried sintering. The microstructure and dielectric, ferroelectric, and energy storage properties of the ceramics are studied. The results show that buried sintered ceramic can inhibit the volatilization of Bi2O3, reduce the generation of oxygen vacancy, make the microstructure more uniform, and reduce the main grain size. The breakdown electric field strength increases from 210 kV/cm to 310 kV/cm. Wrec increases from 1.759 J/cm3 to 4.923 J/cm3, which is an increase of 179.9%. It shows good stability in the energy storage temperature and frequency. In this study, the advantages of buried sintered ceramic from the microscopic level to electrical properties are deeply discussed, which provides a new method for further improving the energy storage performance of BNT-based lead-free ferroelectric ceramics and lays a theoretical foundation for the development of BNT-based lead-free ferroelectric ceramics with high storage performance.

Author Contributions

Data curation, Y.Z., J.Y. and Z.F.; Formal analysis, Y.Z., Y.J., J.Y., Z.F., S.S., X.Z., H.W., S.Y. and M.Z.; Funding acquisition, M.Z.; Methodology, Y.Z., Y.J. and J.Y.; Writing—original draft, Y.Z.; Writing—review and editing, Y.Z. and M.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (Grant No. 12004423), the Opening Project of the Key Laboratory of Inorganic Functional Materials and Devices, Chinese Academy of Sciences (Grant No. KLIFMD202303), the Open Research Fund of the Key Laboratory of Polar Materials and Devices, Ministry of Education (Grant No. CLPM-2024-01), the State Key Laboratory of Advanced Technology for Materials Synthesis and Processing (Wuhan University of Technology) (Grant No. 2024-KF-21), the Materials Science and Engineering Discipline Guidance Fund of China University of Mining and Technology (Grant No. CUMTMS202201) and College Students’ Innovative Entrepreneurial Training Plan Program (China University of Mining and Technology) (Grant No. X2024202).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Zhu, Z.Y.S.; Persson, A.E.O.; Wernersson, L.E. Reconfigurable signal modulation in a ferroelectric tunnel field-effect transistor. Nat. Commun. 2023, 14, 2530. [Google Scholar] [CrossRef] [PubMed]
  2. He, T.F.; Cao, Z.Z.; Li, G.R.; Jia, Y.M.; Peng, B.L. High efficiently harvesting visible light and vibration energy in (1−x)AgNbO3xLiTaO3 solid solution around antiferroelectric-ferroelectric phase boundary for dye degradation. J. Adv. Ceram. 2022, 11, 1641–1653. [Google Scholar] [CrossRef]
  3. Xue, F.; He, X.; Ma, Y.C.; Zheng, D.X.; Zhang, C.H.; Li, L.J.; He, J.H.; Yu, B.; Zhang, X.X. Unraveling the origin of ferroelectric resistance switching through the interfacial engineering of layered ferroelectric-metal junctions. Nat. Commun. 2021, 12, 7291. [Google Scholar] [CrossRef] [PubMed]
  4. Jin, T.Y.; Mao, J.Y.; Gao, J.; Han, C.; Wee, A.T.S.; Loh, K.P.; Chen, W. Ferroelectrics-integrated two-dimensional devices toward next-generation electronics. ACS Nano 2022, 16, 13595–13611. [Google Scholar] [CrossRef] [PubMed]
  5. Niu, X.; Jian, X.D.; Gong, W.P.; Zhang, G.Z.; Jiang, S.L.; Yu, K.; Zhao, X.B.; Yao, Y.B.; Tao, T.; Liang, B.; et al. Field-driven merging of polarizations and enhanced electrocaloric effect in BaTiO3-based lead-free ceramics. J. Adv. Ceram. 2022, 11, 1777–1788. [Google Scholar] [CrossRef]
  6. Zhong, X.C.; Lin, Z.C.; Chen, C.; Wang, R.X.; Zhong, S.P.; Xu, Z.F. The role of ZnFe2O4 in the electrochemical performance of Pb-ceramic composite anode in sulfuric acid solution. Hydrometallurgy 2021, 201, 105587. [Google Scholar] [CrossRef]
  7. Chen, Y.; Li, L.F.; Zhou, Z.; Wang, Y.Y.; Chen, Q.; Wang, Q.Y. La2O3-modified BiYbO3-Pb(Zr,Ti)O3 termary piezoelectric ceramics with enhanced electrical properties and thermal depolarization temperature. J. Adv. Ceram. 2023, 12, 1593–1611. [Google Scholar] [CrossRef]
  8. Kounga, A.B.; Zhang, S.T.; Jo, W.; Granzow, T.; Rödel, J. Morphotropic phase boundary in (1 − x)Bi0.5Na0.5TiO3−xK0.5Na0.5NbO3 lead-free piezoceramics. Appl. Phys. Lett. 2008, 92, 222902. [Google Scholar] [CrossRef]
  9. Zidani, J.; Alaoui, I.H.; Zannen, M.; Birks, E.; Chchiyai, Z.; Majdoub, M.; Manoun, B.; Marssi, M.E.; Lahmar, A. On the lanthanide effect on functional properties of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramic. Materials 2024, 17, 1783. [Google Scholar] [CrossRef]
  10. Uddin, S.; Ahmad, A.; Nasir, M.F.; Zaman, A.; Algahtani, A.; Tirth, V.; Zheng, G.P. Effect of BiFeO3 on the ferroelectric and energy storage properties of (Bi1/2Na1/2)0.94Ba0.06TiO3 based compositions. Inorg. Chem. Commun. 2024, 159, 111746. [Google Scholar] [CrossRef]
  11. Pan, Y.; Dai, Z.H.; Liu, C.X.; Zhao, X.; Yasui, S.; Cong, Y.; Gu, S.T. High energy storage properties of Nd(Mg2/3Nb1/3)O3 modified Bi0.5Na0.5TiO3 lead-free ceramics. J. Mater. Sci. 2024, 59, 3284–3296. [Google Scholar] [CrossRef]
  12. Jiang, Z.H.; Yang, Z.Y.; Yuan, Y.; Tang, B.; Zhang, S.R. High energy storage properties and dielectric temperature stability of (1 − x)(0.8Bi0.5Na0.5TiO3-0.2Ba0.3Sr0.4TiO3)-xNaNbO3 lead-free ceramics. J. Alloys Compd. 2021, 851, 156821. [Google Scholar] [CrossRef]
  13. Guo, B.; Yan, Y.; Tang, M.Y.; Wang, Z.Y.; Li, Y.; Zhang, L.Y.; Zhang, H.B.; Jin, L.; Liu, G. Energy storage performance of Na0.5Bi0.5TiO3 based lead-free ferroelectric ceramics prepared via non-uniform phase structure modification and rolling process. Chem. Eng. J. 2021, 420, 130475. [Google Scholar] [CrossRef]
  14. Zhang, X.R.; Xiao, Y.A.; Du, B.N.; Li, Y.M.; Wu, Y.D.; Sheng, L.Y.; Tan, W.C. Improved non-piezoelectric electric properties based on La modulated ferroelectric-ergodic relaxor transition in (Bi0.5Na0.5)TiO3-Ba(Ti, Zr)O3 ceramics. Materials 2021, 14, 6666. [Google Scholar] [CrossRef] [PubMed]
  15. Gong, Y.F.; Wu, P.; Liu, W.F.; Wang, S.Y.; Liu, G.Y.; Rao, G.H. Switchable ferroelectric diode effect and piezoelectric properties of Bi0.9La0.1FeO3 Ceramics. Chin. Phys. Lett. 2012, 29, 047701. [Google Scholar] [CrossRef]
  16. Kwon, Y.H.; Lee, G.H.; Koh, J.H. Effects of sintering temperature on the piezoelectric properties of (Bi,Na)TiO3-based composites for energy harvesting applications. Ceram. Int. 2015, 41, S792–S797. [Google Scholar] [CrossRef]
  17. Qiu, Y.Z.; Yu, Z.D. Effect of sintering temperature on structure and electrical properties of ZnO-added (Bi0.5Na0.5)0.94Ba0.06TiO3 lead-free ceramics. J. Mater. Sci.-Mater. Electron. 2023, 34, 88. [Google Scholar] [CrossRef]
  18. Leng, S.L.; Jia, F.H.; Zhong, Z.K.; Yang, Q.F.; Li, G.R.; Zheng, L.Y. Fabrication of High Tc BaTiO3-(Bi0.5Na0.5)TiO3 Lead-free positive temperature coefficient of resistivity ceramics. J. Inorg. Mater. 2015, 30, 576–580. [Google Scholar]
  19. Ahn, C.W.; Kim, H.S.; Woo, W.S.; Won, S.S.; Seog, H.J.; Chae, S.A.; Park, B.C.; Jang, K.B.; Ok, Y.P.; Chong, H.H.; et al. Low-temperature sintering of Bi0.5(Na,K)0.5TiO3 for multilayer ceramic actuators. J. Am. Ceram. Soc. 2015, 98, 1877–1883. [Google Scholar] [CrossRef]
  20. Tian, H.Y.; Kwok, K.W.; Chan, H.L.W.; Buckley, C.E. The effects of CuO-doping on dielectric and piezoelectric properties of Bi0.5Na0.5TiO3-Ba(Zr,Ti)O3 lead-free ceramics. J. Mater. Sci. 2007, 42, 9750–9755. [Google Scholar] [CrossRef]
  21. Kong, D.K.; Guo, A.F.; Hu, Y.B.; Zhou, X.Y.; Wu, H.L.; Li, X.J.; Qu, P.; Wang, S.Q.; Guo, S. Alumina-based ceramic cores prepared by vat photopolymerization and buried combustion method. Mater. Today Commun. 2023, 37, 107434. [Google Scholar] [CrossRef]
  22. Chen, R.Y.; Xie, K.S.; Zhu, H.P.; He, Q.; Li, S.S.; Wen, H.M. Improving strength and microstructure of SiC reticulated porous ceramic through in-situ generation of SiC whiskers within hollow voids. Ceram. Int. 2023, 49, 40414–40420. [Google Scholar] [CrossRef]
  23. Fujii, I.; Mitsui, R.; Nakashima, K.; Kumada, N.; Yabuta, H.; Shimada, M.; Watanabe, T.; Miura, K.; Wada, S. Effect of sintering condition and V-doping on the piezoelectric properties of BaTiO3-Bi(Mg1/2Ti1/2)O3-BiFeO3 ceramics. J. Cearm. Soc. Jpn. 2013, 121, 589–592. [Google Scholar] [CrossRef]
  24. Li, H.; Liu, Y.S.; Liu, Y.S.; Zeng, Q.F.; Hu, K.H.; Lu, Z.G.; Liang, J.J. Effect of burying sintering on the properties of ceramic cores via 3D printing. J. Manuf. Process. 2020, 57, 380–388. [Google Scholar] [CrossRef]
  25. Long, C.; Su, Z.; Song, H.; Xu, A.; Liu, L.; Li, Y.; Zheng, K.; Ren, W.; Wu, H.; Ding, X. Excellent energy storage properties with ultrahigh Wrec in lead-free relaxor ferroelectrics of ternary Bi0.5Na0.5TiO3-SrTiO3-Bi0.5Li0.5TiO3 via multiple synergistic optimization. Energy Storage Mater. 2024, 65, 103055. [Google Scholar] [CrossRef]
  26. Fan, J.T.; He, G.; Cao, Z.Z.; Cao, Y.F.; Long, Z.; Hu, Z.G. Ultrahigh energy-storage density of a lead-free 0.85Bi0.5Na0.5TiO3-0.15Ca(Nb0.5Al0.5)O3 ceramic under low electric fields. Inorg. Chem. Front. 2023, 10, 1561–1573. [Google Scholar] [CrossRef]
  27. Che, Z.Y.; Ma, L.; Luo, G.G.; Xu, C.; Cen, Z.Y.; Feng, Q.; Chen, X.Y.; Ren, K.L.; Luo, N.N. Phase structure and defect engineering in (Bi0.5Na0.5)TiO3-based relaxor antiferroelectrics toward excellent energy storage performance. Nano Energy 2022, 100, 107484. [Google Scholar] [CrossRef]
  28. Guan, P.F.; Zhang, Y.X.; Yang, J.; Zheng, M. Effect of Sm3+ doping on ferroelectric, energy storage and photoluminescence properties of BaTiO3 ceramics. Ceram. Int. 2023, 49, 11796–11802. [Google Scholar] [CrossRef]
  29. Shi, X.H.; Li, K.; Shen, Z.Y.; Liu, J.Q.; Chen, C.Q.; Zeng, X.J.; Zhang, B.; Song, F.S.; Luo, W.Q.; Wang, Z.M.; et al. BS0.5BNT-based relaxor ferroelectric ceramic/glass-ceramic composites for energy storage. J. Adv. Ceram. 2023, 12, 695–710. [Google Scholar] [CrossRef]
  30. Li, Z.P.; Li, D.X.; Shen, Z.Y.; Zeng, X.J.; Song, F.S.; Luo, W.Q.; Wang, X.C.; Wang, Z.M.; Li, Y.M. Remarkably enhanced dielectric stability and energy storage properties in BNT-BST relaxor ceramics by A-site defect engineering for pulsed power applications. J. Adv. Ceram. 2022, 11, 283–294. [Google Scholar] [CrossRef]
  31. Zhang, X.; Zhang, F.; Niu, Y.W.; Zhang, Z.Q.; Lei, X.Q.; Wang, Z.J. Excellent energy storage performance of perovskite high-entropy Oxide-modified (Bi0.5Na0.5)TiO3-based ceramics. ACS Appl. Electron. Mater. 2024, 6, 4698–4708. [Google Scholar] [CrossRef]
  32. Lin, Y.Z.; Wan, R.F.; Zheng, P.; Li, Z.H.; Wang, Y.K.; Fan, Q.L.; Zheng, L.; Zhang, Y.; Bai, W.F. Achieving remarkable energy storage performances under low electric field in Bi0.5N0.5TiO3-SrTiO3-based relaxor ferroelectric ceramics via a heterostructure doping strategy. ACS Appl. Electron. Mater. 2023, 5, 4576–4586. [Google Scholar] [CrossRef]
  33. Cai, Z.M.; Yang, H.; Zhu, C.Q.; Li, S.H.; Luo, B.C.; Li, A.Y.; Li, X.H.; Tian, Z.B.; Feng, P.Z. Local heterogeneous polarization enhanced superior low-field energy storage performance in lead-free relaxor ferroelectric ceramics. ACS Sustain. Chem. Eng. 2023, 11, 13729–13735. [Google Scholar] [CrossRef]
  34. He, J.; Liu, X.; Zhao, Y.; Du, H.; Zhang, T.; Shi, J. Dielectric stability and energy-storage performance of BNT-based relaxor ferroelectrics through Nb5+ and its excess modification. ACS Appl. Electron. Mater. 2022, 4, 735–743. [Google Scholar] [CrossRef]
  35. Chu, B.K.; Hao, J.G.; Li, P.; Li, Y.C.; Li, W.; Zheng, L.M.; Zeng, H.R. High-energy storage properties over a broad temperature range in La-modified BNT-based lead-free ceramics. ACS Appl. Mater. Interfaces 2022, 14, 19683–19696. [Google Scholar] [CrossRef] [PubMed]
  36. Yang, J.; Zhu, X.L.; Wang, H.T.; Zhang, Y.X.; Guan, P.F.; Yan, S.G.; Zheng, M. Achieving outstanding energy storage behaviors via combinatorial optimization design in BNT-based relaxor ferroelectric ceramics under medium-low electric fields. J. Mater. Chem. C 2024, 12, 6479–6486. [Google Scholar] [CrossRef]
  37. Liu, T.Y.; Yan, B.; Ma, J.X.; He, Q.; An, L.N.; Chen, K.P. Enhanced energy storage properties in BNT-based ceramics with a morphotropic phase boundary modified by Sr(Mg1/3Nb2/3)O3. J. Mater. Chem. C 2023, 11, 15294–15302. [Google Scholar] [CrossRef]
  38. Li, X.H.; Zhu, C.Q.; Li, S.H.; Li, A.Y.; Liang, L.Q.; Cai, Z.M.; Feng, P.Z. Enhancing energy storage density of BNT-ST-based ceramics by a stepwise optimization strategy on the breakdown strength. J. Eur. Ceram. Soc. 2024, 44, 6422–6429. [Google Scholar] [CrossRef]
  39. Sun, M.Z.; Wang, X.M.; Li, P.; Du, J.; Fu, P.; Hao, J.G.; Li, W.; Zhai, J.W. Realizing ultrahigh breakdown strength and ultrafast discharge speed in novel barium titanate-based ceramics through multicomponent compounding strategy. J. Eur. Ceram. Soc. 2023, 43, 974–985. [Google Scholar] [CrossRef]
  40. Chen, Y.; Huang, Y.; Zuo, Y.D.; Wang, H.S.; Liu, K.; Fan, B.Y.; Zhang, Q.F.; Zhang, G.Z.; Jiang, S.L.; Shen, M. Enhanced energy storage property achieved in Na0.5Bi0.5TiO3-based ferroelectric ceramics via composition design and grain size tuning. J. Eur. Ceram. Soc. 2022, 42, 6985–6996. [Google Scholar] [CrossRef]
  41. Shi, W.J.; Zhang, L.Y.; Jing, R.Y.; Hu, Q.Y.; Zeng, X.Y.; Alikin, D.O.; Shur, V.Y.; Wei, X.Y.; Gao, J.H.; Liu, G.; et al. Relaxor antiferroelectric-like characteristic boosting enhanced energy storage performance in eco-friendly (Bi0.5Na0.5)TiO3-based ceramics. J. Eur. Ceram. Soc. 2022, 42, 4528–4538. [Google Scholar] [CrossRef]
  42. Singh, A.; Kharangarh, P.; Gupta, V. Enhanced energy storage efficiency with superior thermal stability under low electric field and large electric field driven strain in environment- friendly Bi0.5Na0.5TiO3 based ferroelectric modified with LiNbO3. J. Alloys Compd. 2023, 945, 169181. [Google Scholar] [CrossRef]
  43. Yang, J.; Guan, P.F.; Zhang, Y.X.; Zhu, X.L.; Wang, H.T.; Yang, C.; Zheng, M. High energy storage density achieved in BNT-based ferroelectric translucent ceramics under low electric fields. J. Am. Ceram. Soc. 2024, 107, 6294–6306. [Google Scholar] [CrossRef]
  44. Shi, X.H.; Li, Z.P.; Shen, Z.Y.; Song, F.S.; Luo, W.Q.; Zeng, X.J.; Wang, Z.M.; Li, Y.M. Ba2+/Sr2+ regulation in A-site vacancy-engineered B0.015+1.5xS0.245-1.5x0.03BNT relaxor ceramics for energy storage. J. Am. Ceram. Soc. 2024, 107, 2325–2336. [Google Scholar] [CrossRef]
  45. Zhu, W.; Guo, H.H.; Shen, Z.Y.; Song, F.S.; Luo, W.Q.; Wang, Z.M.; Li, Y.M. Boosting dielectric temperature stability in BNBST-based energy storage ceramics by Nb2O5 modification. J. Am. Ceram. Soc. 2023, 106, 3633–3642. [Google Scholar] [CrossRef]
  46. Luo, W.X.; Wu, M.X.; Han, Y.F.; Zhou, X.; Liu, L.J.; He, Q.W.; Ren, P.R.; Yang, H.M.; Yang, H.; Wang, Q.; et al. Enhanced optical transmittance and energy-storage performance in NaNbO3-modified Bi0.5Na0.5TiO3 ceramics. J. Am. Ceram. Soc. 2023, 106, 4723–4731. [Google Scholar] [CrossRef]
  47. Jiang, Z.H.; Yuan, Y.; Yang, H.C.; Li, E.Z.; Zhang, S.R. Excellent thermal stability and energy storage properties of lead-free Bi0.5Na0.5TiO3-based ceramic. J. Am. Ceram. Soc. 2022, 105, 4027–4038. [Google Scholar] [CrossRef]
  48. Lian, H.L.; Liang, X.J.; Shi, M.; Liu, L.N.; Chen, X.M. Improved dielectric temperature stability and energy storage properties of BNT-BKT-based lead-free ceramics. Ceram. Int. 2024, 50, 5021–5031. [Google Scholar] [CrossRef]
  49. Wang, D.; Chu, B.K.; Li, P.; Han, W.F.; Kong, Y.X.; Fu, P.; Li, Y.C.; Hao, J.G.; Li, W. Improving the energy storage performance of (Bi0.5Na0.5)TiO3-BaTiO3 based ceramics via (Sr0.7Bi0.2)TiO3 modification. Ceram. Int. 2023, 49, 37486–37493. [Google Scholar] [CrossRef]
  50. Zheng, M.; Guan, P.F.; Yang, J.; Zhang, Y.X. Microstructure and composition driven ferroelectric properties of Er3+ doped lead-free multifunctional 0.94Bi0.5Na0.5TiO3-0.06BaTiO3 ceramics. Ceram. Int. 2023, 49, 30481–30489. [Google Scholar] [CrossRef]
  51. Yang, Y.; Jing, R.; Wang, J.; Lu, X.; Du, H.; Jin, L. Large electrostrain and high energy-storage properties of (Sr1/3Nb2/3)4+-substituted (Bi0.51Na0.5)TiO3-0.07BaTiO3 lead-free ceramics. Ceram. Int. 2022, 48, 23975–23982. [Google Scholar] [CrossRef]
  52. Shang, K.L.; Shi, W.J.; Yang, Y.L.; Zhang, L.Y.; Hu, Q.Y.; Wei, X.Y.; Jin, L. Medium electric field-induced ultrahigh polarization response and boosted energy-storage characteristics in BNT-based relaxor ferroelectric polycrystalline ceramics. Ceram. Int. 2022, 48, 37223–37231. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of the BNT-LFO ceramics under ordinary sintering and buried sintering in the 2θ ranges of (a) 20–80° and (b) 30–36°.
Figure 1. The XRD patterns of the BNT-LFO ceramics under ordinary sintering and buried sintering in the 2θ ranges of (a) 20–80° and (b) 30–36°.
Materials 17 04019 g001
Figure 2. SEM pictures of (a) ordinary sintered BNT-LFO ceramic and (b) buried sintered BNT-LFO ceramic; the insets are the distributions of the grain size.
Figure 2. SEM pictures of (a) ordinary sintered BNT-LFO ceramic and (b) buried sintered BNT-LFO ceramic; the insets are the distributions of the grain size.
Materials 17 04019 g002
Figure 3. Dielectric temperature spectra of (a) ordinary sintered BNT-LFO ceramic and (b) buried sintered BNT-LFO ceramic.
Figure 3. Dielectric temperature spectra of (a) ordinary sintered BNT-LFO ceramic and (b) buried sintered BNT-LFO ceramic.
Materials 17 04019 g003
Figure 4. The bipolar P-E loops of (a) ordinary sintered and (b) buried sintered BNT-LFO ceramics. (c) The bipolar P-E loops of ordinary sintered and buried sintered BNT-LFO ceramics at 150 kV/cm. (d) The variation in the Pmax and Pr values of ordinary sintered and buried sintered BNT-LFO ceramics at 150 kV/cm. The curves of different colors represent the bipolar P-E loops under different electric fields. Due to the different materials a and b, the bipolar P-E loops are different.
Figure 4. The bipolar P-E loops of (a) ordinary sintered and (b) buried sintered BNT-LFO ceramics. (c) The bipolar P-E loops of ordinary sintered and buried sintered BNT-LFO ceramics at 150 kV/cm. (d) The variation in the Pmax and Pr values of ordinary sintered and buried sintered BNT-LFO ceramics at 150 kV/cm. The curves of different colors represent the bipolar P-E loops under different electric fields. Due to the different materials a and b, the bipolar P-E loops are different.
Materials 17 04019 g004
Figure 5. The unipolar P-E loops of (a) ordinary sintered and (b) buried sintered BNT-LFO ceramics. (c) The unipolar P-E loops of ordinary sintered and buried sintered BNT-LFO ceramics at their maximum electric fields. (d) The variation in the Wrec and η values of ordinary sintered and buried sintered BNT-LFO ceramics.
Figure 5. The unipolar P-E loops of (a) ordinary sintered and (b) buried sintered BNT-LFO ceramics. (c) The unipolar P-E loops of ordinary sintered and buried sintered BNT-LFO ceramics at their maximum electric fields. (d) The variation in the Wrec and η values of ordinary sintered and buried sintered BNT-LFO ceramics.
Materials 17 04019 g005
Figure 6. A comparison of the Wrec in this work with that of other previously published lead-free BNT-based ceramics [29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52].
Figure 6. A comparison of the Wrec in this work with that of other previously published lead-free BNT-based ceramics [29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52].
Materials 17 04019 g006
Figure 7. (a) The unipolar P-E loops of buried sintered BNT-LFO ceramic under different temperatures. (b) The variation in Wrec and η of buried sintered BNT-LFO ceramic under different temperatures. (c) The unipolar P-E loops of buried sintered BNT-LFO ceramic under different frequencies. (d) The variation in Wrec and η of buried sintered BNT-LFO ceramic under different frequencies.
Figure 7. (a) The unipolar P-E loops of buried sintered BNT-LFO ceramic under different temperatures. (b) The variation in Wrec and η of buried sintered BNT-LFO ceramic under different temperatures. (c) The unipolar P-E loops of buried sintered BNT-LFO ceramic under different frequencies. (d) The variation in Wrec and η of buried sintered BNT-LFO ceramic under different frequencies.
Materials 17 04019 g007
Table 1. A comparison of the energy storage ability in this work and other previously published lead-free BNT-based ceramics.
Table 1. A comparison of the energy storage ability in this work and other previously published lead-free BNT-based ceramics.
CompositionE (kV/cm)Wrec (J/cm3)η (%)Ref.
(Ba0.3Sr0.7)0.5(Bi0.5Na0.5)0.5TiO3 + 14 wt% GS1702.165.2[29]
Ba0.105Na0.325Sr0.185Bi0.385TiO31101.8~73[30]
0.45(Bi0.5Na0.5)TiO3-0.55(Sr0.7Bi0.2)TiO3-0.3(Bi0.2Na0.2Ba0.2Sr0.2Ca0.2)(Ti0.9Nb0.1)O34106.0485[31]
Bi0.5N0.5TiO3-SrTiO3-0.04Sr2NaNb5O153405.2293.87[32]
(65%(0.92Bi0.5Na0.5TiO3-0.08Bi(Mg0.3Zr0.6)O3)-35%(0.6BaTiO3-0.4NaNbO3)2804.1195.6[33]
0.54Bi0.5Na0.5TiO3-0.06BaTiO3-0.4Bi0.2Sr0.7Ti0.96875Nb0.125O31702.0794.5[34]
[(Bi0.5Na0.5)0.94Ba0.06]0.82La0.12TiO34405.9377.6[35]
0.7Bi0.5Na0.5TiO3-0.2BaZr0.3Ti0.7O3-0.1NaNbO32203.5393.5[36]
0.85(0.8525BNT–0.10995BKT–0.03755BT)0.15Sr(Mg1/3Nb2/3)O33103.5386.3[37]
0.6Bi0.5Na0.5TiO3-0.4SrTiO3(@Si-TSS)2703.1486.21[38]
0.9[0.88Ba0.6Ca0.4TiO3-0.12Bi(Mg2/3(Nb0.85Ta0.15)1/3)O3]-0.1Bi0.5Na0.5TiO34303.5990.86[39]
0.7Na0.5Bi0.5TiO3-0.3NaNbO3/7 wt%CaZr0.5Ti0.5O34104.9393.3[40]
0.92Bi0.5Na0.5TiO3–0.08LiNbO31803.6280.8[41]
0.65Bi0.5Na0.4K0.1TiO3-0.35[2/3SrTiO3-1/3Bi(Mg2/3Nb1/3)O3]2904.4386[42]
0.9Bi0.5Na0.5TiO3−0.1BaZr0.3Ti0.7O3:0.6mol%Er3+1902.9551.3[43]
Ba0.087Sr0.176Bi0.385Na0.325TiO31302.3364.5[44]
(Bi0.5Na0.5)0.65(Ba0.3Sr0.7)0.35(Ti0.98Ce0.02)O3+2wt%Nb2O5901.4484.1[45]
0.5Bi0.5Na0.5TiO3-0.5NaNbO32865.1479.65[46]
0.75Bi0.5Na0.5TiO3-0.25CaTiO33102.7491[47]
0.7[0.85(0.84Bi0.5Na0.5TiO3-0.16Bi0.5K0.5TiO3)-0.15BiMg2/3Nb1/3O3]-0.3Sr0.7La0.2TiO33004.0385.2[48]
0.53(Bi0.5Na0.5)TiO3-0.07BaTiO3-0.4(Sr0.7Bi0.2)TiO32603.2690.3[49]
0.94Bi0.5Na0.5TiO3-0.06BaTiO3:1mol%Er3+800.429~48[50]
[0.93(Bi0.51Na0.5)0.07Ba)]Ti0.9925(Sr1/3Nb2/3)0.0075O31001.36~61[51]
0.85(0.75Bi0.5Na0.4K0.1TiO3-0.25SrTiO3)-0.15Bi(Mg0.5Ti0.5)O32704.8284.9[52]
0.85Bi0.5Na0.5TiO3-0.15LaFeO3 (buried sintering)3104.92377.4This work
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

Zhang, Y.; Jia, Y.; Yang, J.; Feng, Z.; Sun, S.; Zhu, X.; Wang, H.; Yan, S.; Zheng, M. Enhancing Energy Storage Performance of 0.85Bi0.5Na0.5TiO3-0.15LaFeO3 Lead-Free Ferroelectric Ceramics via Buried Sintering. Materials 2024, 17, 4019. https://doi.org/10.3390/ma17164019

AMA Style

Zhang Y, Jia Y, Yang J, Feng Z, Sun S, Zhu X, Wang H, Yan S, Zheng M. Enhancing Energy Storage Performance of 0.85Bi0.5Na0.5TiO3-0.15LaFeO3 Lead-Free Ferroelectric Ceramics via Buried Sintering. Materials. 2024; 17(16):4019. https://doi.org/10.3390/ma17164019

Chicago/Turabian Style

Zhang, Yixiao, Yuchen Jia, Jian Yang, Zixuan Feng, Shuohan Sun, Xiaolong Zhu, Haotian Wang, Shiguang Yan, and Ming Zheng. 2024. "Enhancing Energy Storage Performance of 0.85Bi0.5Na0.5TiO3-0.15LaFeO3 Lead-Free Ferroelectric Ceramics via Buried Sintering" Materials 17, no. 16: 4019. https://doi.org/10.3390/ma17164019

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