Next Article in Journal
Recurrent Supramolecular Patterns in a Series of Salts of Heterocyclic Polyamines and Heterocyclic Dicarboxylic Acids: Synthesis, Single-Crystal X-ray Structure, Hirshfeld Surface Analysis, Energy Framework, and Quantum Chemical Calculations
Previous Article in Journal
First Examples of Metal-Organic Frameworks with Pore-Encapsulated [Co(CO)4] Anions: Facile Synthesis, Crystal Structures and Stability Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High Energy Storage Performance in Pb1−xLax(Hf0.45Sn0.55)0.995O3 Antiferroelectric Ceramics

by
Erping Wang
1,
Liqin Yue
1,
Yuanhong Chu
1,
Caixia Sun
1,
Jinyu Zhao
1,
Siyu Zhang
1,
Jiale Liu
1,
Yangyang Zhang
1,2,* and
Ling Zhang
3,*
1
Faculty of Engineering, Huanghe Science and Technology College, Zijingshan South Road, Zhengzhou 450006, China
2
Henan Key Laboratory of Nanocomposites and Applications, Huanghe Science and Technology College, Zijingshan South Road, Zhengzhou 450006, China
3
College of Mechanical and Electrical Engineering, Shihezi University, Shihezi 832061, China
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(8), 732; https://doi.org/10.3390/cryst14080732
Submission received: 12 July 2024 / Revised: 4 August 2024 / Accepted: 15 August 2024 / Published: 17 August 2024
(This article belongs to the Section Polycrystalline Ceramics)

Abstract

:
Energy storage efficiency (η) and large recoverable energy density (Wre) are necessary for antiferroelectric materials in order to develop antiferroelectric-based dielectric capacitors with exceptional energy storage capacity. In the present paper, the effect of doping La3+ on the energy storage capacity of Pb1−xLax(Hf0.45Sn0.55)0.995O3 antiferroelectric ceramics was studied. Adjusting the content of La and changing the phase structure of PLHS from antiferroelectric to relaxor ferroelectric gradually, which narrowed its hysteresis loop, yielded a high energy storage efficiency of 81.9% and the maximum breakdown field strength of 200 kV/cm when x = 2 mol%. In addition, the recoverable energy density and energy storage efficiency both showed excellent temperature stability and frequency stability in the temperature range of 10–110 °C and the frequency range of 10–100 Hz, suggesting that Pb0.98La0.02(Hf0.45Sn0.55)0.995O3 are favorable materials candidates for the preparation of pulsed-power capacitors that can be used in a wide range of conditions.

1. Introduction

As the preferred material for dielectric capacitors, antiferroelectric materials are widely used in automotive electronics, communications, aerospace, and other fields because of their reversible antiferro–ferro phase transition, exhibiting the typical traits of double P-E hysteresis loops, shown in Figure 1d, and achieving high energy density (Wst) easily [1,2,3].
The antiferroelectric materials that many researchers have focused on include PZT [4,5,6]: PbHfO3 [7], AgNbO3 [8,9], and NaNbO3 [10] systems.
As lead-free antiferroelectric materials, AgNbO3 and NaNbO3 have attracted much attention due to being environmentally friendly. However, they all exhibit complex temperature-induced phase transition behavior, generally low energy storage efficiency, low breakdown field strength, and harsh sintering process requirements, all of these factors limit their use in the energy storage industry [11,12,13,14,15,16,17]. Therefore, researchers have turned their attention back to lead-containing materials.
Because of the narrow antiferroelectric phase region of pure PZT, slight composition fluctuation can easily lead to the deviation of the FE-AFE phase boundary; thus, researchers have improved the energy storage performance through A/B-site-doping-formed PLZT, PLZS, and other systems. However, when EA-F exceeds the breakdown strength (Eb) it is challenging to acquire a high recoverable energy density (Wre).
PbHfO3(PHO)-based ceramics are similar to PZT-based ceramics in phase structure and electrical properties. Studies have shown that their valence band and conduction band electron distribution are similar, while the band gap is quite different (PZO is about 3.6 eV, whereas PHO is about 2.6 eV) [18]. The multi-phase transitions of PHO are complex, and the mesophase is also controversial. The doping modification of PHO (such as A-site doping, B-site doping, or A/B-site co-doping) can efficiently and effectively adjust the electric field strength for the phase transition of AFE-FE or optimize the breakdown strength, thereby improving the energy storage performance [19,20,21,22,23].
The phase stability of the ABO3 perovskite structure can be calculated using Formula (1) [22,24], in which, RA, RB, and RO represent the ionic radius of A-site, B-site, and oxygen cations, respectively.
t = R A + R O / 2 R B + R O
High t is conducive to the formation of the ferroelectric phase, while low t is crucial for the antiferroelectric phase. If the antiferroelectric phase is too stable to be transformed into the ferroelectric phase before breakdown, the P-E loop will appear as a linear dielectric, resulting in a low energy density [25]. As a result, it is necessary to select the appropriate radius of ions for doping at different sites for optimum electrical performance. At the same time, acquiring excellent recoverable energy density (Wre) and energy efficiency (η) and improving the stability of its antiferroelectric phase are the main research hotspots regarding PHO ceramics.
In multiple studies, La3+ with a smaller ionic radius was introduced into the A-site of Pb(Hf,Sn)O3 (PHS) to reduce its tolerance factor and strengthen the stabilization of the antiferroelectric stage, thus enhancing the energetic storing density [26,27,28,29,30,31,32,33,34,35]. In addition, the introduction of higher-valence cations into A site of Pb(Hf,Sn)O3 can produce Pb vacancies, which helps to establish chemical pressure in the lattice, resulting in the tilt and compression of the oxygen octahedron of Pb(Hf,Sn)O3 and structural heterogeneity, thereby enhancing the stabilization of the antiferroelectric stage [29,30,34].
At the same time, by adjusting the doping of La3+, the crystal plane spacing of the sample decreases as the La doping content increases, so the lattice size and lattice constant also decrease accordingly. The basic relationship between ceramic grain size and breakdown field strength (Eb) is expressed as follows:
E b 1 G
where G is the mean grain dimension. From Equation (2), it can be inferred that the lower the ceramic grain size, the larger the breakdown field strength. Meanwhile, the energy storage efficiency (η), recoverable energy density (Wre), and energy density (Wst) of the permittivity of the dielectric materials are directly related to the breakdown electric field.
W st = 0 P max E dP upon   charging
W re = P max P r E dP upon   discharging
η = W re W st × 100 %
Therefore, to obtain ceramic materials with high Wst and η, many researchers are also committed to reducing the porosity of ceramics, making grains uniform, or performing surface modification or the doping of ceramic particles.
In the current work, we examined how La3+ doping affects the energy storage characteristics of ceramics Pb1−xLax(Hf0.45Sn0.55)0.995O3. By utilizing the solid-phase sintering process, the optimal grain size of the PLHS ceramic material was determined, leading to an increase in its energy storage efficiency (η) and energy density (Wst). The recoverable energy density and the energy storage efficiency both showed excellent temperature stability and frequency stability in a wide temperature range and frequency scale.

2. Experimental Procedures

The formula Pb1−xLax(Hf0.45Sn0.55)0.995O3 (x = 0, 0.01, 0.02, 0.03) was followed in the preparation of the PLHS ceramics with a conventional solid-state reaction process. The raw materials were La2O3, PbO, SnO2, and HfO2 powders with a chemical purity of not less than 99.0%. The mixed oxides were milled with ethanol for 12 h and then dried at 120 °C. The dried material was pre-sintered with 100~150 °C/h, heated to a temperature of 800~1200 °C, and kept for 2~4 h, and then naturally reduced to room temperature. After that, the second milling and drying procedure was carried out with the same conditions as those of the first, and then the samples were mixed with 95% alcohol and 6% PVA and sieved (60~80 mesh), ensuring their good fluidity. The obtained material was pressed into cylindrical pellets of 10 mm in diameter and 2 mm in height at 6 MPa, degreased at 600 °C for 4 h, and subsequently cooled to ambient temperature naturally. The material was then subjected to sintering for 2.5 h at 1150–1250 °C, after which it was polished, and the entire piece was silver-plated on both sides and then sintered for 30 min at 550 °C. Finally, the samples were polarized in silicone oil for dielectric testing.
X-ray diffraction (D8-Advanced, Bruker AXS Inc., Madison, Germany) was applied to measure the crystal structure at ambient temperature. SEM (JSM 6510LV, Jeol, Tokyo, Japan) was utilized for observing surface morphology. The variation in dielectric performance with temperature was measured with a TH2827C LCR instrument (Changzhou Tonghui, Changzhou, China) at 1 kHz frequency, and p-e hysteresis loops were determined with a high-voltage amplifier (Trek 610E; Trek, Lockport, NY, USA) and a ferroelectric test system (PolyK Technologies, North Philipsburg, PA, USA) in silicone oil at 10 Hz. Energy storage efficiency (η) along with energy density (Wst) were detected through the p-e hysteresis loops.

3. Results and Discussion

Figure 2a presents the XRD images of PLHS ceramic samples, which were subjected to sintering for 2.5 h at 1150–1250 °C and polished with various La amounts at ambient conditions. Figure 2a displays that there are no impurities or secondary phases in any of the ceramics; they are all pure perovskite phase structures. To better show the phase structure of the ceramics, we performed a narrow-spectrum XRD scan of the sample in the range of 43.5°–45.0° in 2θ, as shown in Figure 2b. The ceramics have a tetragonal phase structure, and there are two obvious splittings at the (200) and (002) lattices in Figure 2b. At the same time, it can be seen that as the La amount increases from x = 0 to x = 3 mol%, the (200) and (002) diffraction peaks gradually move to a high angle (right direction). According to Bragg’s formula, the crystal plane spacing of samples reduces with an increase in the La doping amount, and therefore the lattice size and lattice constant decrease accordingly. The analysis shows that the ionic radius of La3+ (1.36 Å) is lower than that of Pb2+ (1.49 Å) (coordination number is 6), and lattice distortion occurs when La3+ replaces Pb2+, so the ceramic grain size reduces in microns. According to Formula (2), the smaller the ceramic grain, the greater the strength of the breakdown field, which is conducive to obtaining excellent capacity storing performance.
Figure 3 shows the SEM pictures of PLHS ceramics, which were subjected to sintering for 2.5 h at 1150–1250 °C and polished with various La contents. From the graph, we can observe that the growth of the La3+ level has a strong influence on grain size. Grain size clearly increases as the La content increases to x = 1 mol%. This may be because the ionic size of La3+ is smaller than Pb2+, and once La3+ replaces Pb2+, the internal diffusion rate is accelerated, so the growth of grains is accelerated. When the content of La increases to x = 2 mol%, the ceramics’ grain size starts to reduce significantly. Because the introduction of extra La3+ accelerates the formation of negatively charged Pb2+ vacancies, it is easy to separate at grain boundaries. These generated cavities induce the attraction of directly electrically loaded La3+ ions, which in turn generate vacancy ion couples in the vicinity of the grain boundaries via Coulombic function. These vacancy ion pairs hinder the movement of grain boundaries and suppress grain growth, causing a reduction in grain size. Moreover, once the content of La is less than 2 mol%, the surface morphology of all ceramics is relatively dense, and when the content is further increased to 3 mol%, pores begin to appear in the ceramic. It is possible that the lattice distortion of the ceramics will increase when the doping amount of La increases, resulting in more defects.
The amount of La not only influences the microstructures of PLHS but also has a significant impact on their energetic behavior. Consequently, to achieve the best energy storage performance, the appropriate amount of La doping should be determined. Figure 4 shows the P-E hysteresis loops and Pmax, Pr, and Pmax-Pr, as well as the energy density (Wst) and the recoverable energy density (Wre), of all ceramic samples that had been tested under ambient conditions and 10 Hz with 200 kV/cm of maximum electric field strength. As illustrated in Figure 4a, when x = 0, the hysteresis loop corresponds to the antiferroelectric form. However, as the content of La increases, the hysteresis loops change from antiferroelectric to relaxed ferroelectric gradually. This may be due to the addition of La3+ ions, which enhances the local random electric field and random stress field, breaks the long-range order of antiferroelectrics, and transforms antiferroelectrics into relaxor ferroelectrics [36,37], which shortens the hysteresis loop of the antiferroelectric phase and enhances its stability. As the electric field strength involved in the antiferroelectric-to-relaxed-ferroelectric phase transition (EA-RF) increases, and the electric field strength (E) reduces, all P-E hysteresis loops become thinner, showing that PLHS ceramics present excellent energetic storing properties. When x = 2 mol%, the p-e loop is the narrowest, which shows that the residual polarization is the smallest, which can minimize the heat generated inside the material and enhance the reliability of the device. As can be seen from Figure 4b, Pmax and Pr decreases from 16.9 μC/cm2 to 8.3 μC/cm2 and from 1.1 μC/cm2 to 0.9 μC/cm2, respectively, as the content of La increases from x = 0 to x = 3 mol%. Since La3+ ions occupy the crystallographic position of Pb2+, the composition is disordered, and the disordered ions diffuse and exchange positions, a process that tends to be ordered. This process will form a strong random field to break the long-range ordered antiferroelectric state and form a relaxed ferroelectric state, thereby reducing the residual polarization (Pr), enhancing the breakdown field strength (Eb), and improving the energy storage performance of the material. As illustrated in Figure 4c, the maximum η of 81.9% and Wre of 0.77 J/cm3 are observed in Pb0.98La0.02(Hf0.45Sn0.55)0.995O3.
Based on previous reports, it is almost impossible to obtain both a large Wre and high η in AFE ceramics. For example, Wang et al. [35] obtained the highest restorable storing energy intensity of 16.4 J/cm3 in (Pb0.99Nb0.02) (Zr, Sn, Ti)0.98O3 (PNZST) antiferroelectric thin films, but the efficiency was only 57%. The efficiency of other studies was mostly between 60 and 75% and rarely more than 80% [38,39,40,41,42]. This study also revealed a remarkably high energy storage efficiency (η) of 81.9% and recoverable energy density (Wre) of 0.77 J/cm3, making it generally better than alternative Pb-based AFE ceramics with respect to energy storage characteristics. Ultra-high efficiencies were obtained at moderate energy storage densities. High energy storage efficiency means limited internal dissipated energy, thus avoiding the influence of internal heat accumulation on material stability and enhancing the service life of ceramic capacitors. Figure 5 shows the recently reported performance parameters of antiferroelectric ceramics.
In practical applications, performance stabilization is also a key criterion for evaluating the usage of energy storage materials. Therefore, at an electrical amount of 150 kV/cm, the energy storage characteristic of the x = 2 mol% element varies with frequency and temperature, as presented in Figure 6 and Figure 7. In Figure 6a, the ceramics show a consistent narrow hysteresis loop at different frequencies, and the residual polarization is generally small. As shown in Figure 6b, Pmax and Pr decrease slowly from 9.74 μC/cm2 and 1.26 μC/cm2 to 9.37 μC/cm2 and 0.52 μC/cm2 in the range of 0.5–10 Hz, respectively, and then stabilize in the range of 10–100 Hz. Meanwhile, Wst decreases in the range of 0.5–10 Hz and then stabilizes at about 0.71 J/cm3 in the range of 10–100 Hz. Wre increases from 0.56 J/cm3 to 0.62 J/cm3 in the range of 0.5–10 Hz, so η also increases from 71.3% to 84.5% and then stabilizes to about 88.1% in the range of 10–100 Hz. At the same time, it can be observed from Figure 7 that in the temperature scale of 10–110 °C, the shape of the hysteresis loop almost does not change, with Pmax and Pr increasing from 8.91 μC/cm2 and 0.56 μC/cm2 to 9.42 μC/cm2 and 0.95 μC/cm2, respectively, so Pmax-Pr tends to be stable throughout the temperature range. Therefore, in the 10–100 Hz range and the 10–110 °C temperature interval, the ceramics have good frequency and temperature stability and have broad application prospects. The above findings show that when x = 2 mol%, the ratio of Pb and La at point A contributes to PLHS obtaining suitable tolerance in terms of the t factor according to Formula (1), which improves the ceramic’s AFE phase stability.

4. Conclusions

In the present study, Pb1−xLax(Hf0.45Sn0.55)0.995O3 ceramics with different La contents were successfully gained through a conventional solid-phase sintering procedure, and the relationships among microstructures, crystal structures, and energy storage performance of ceramics and the La amount were investigated systematically. With a higher La concentration, the antiferroelectric state gradually transforms into the relaxor ferroelectric state with narrower hysteresis loops, which leads to an increase in EA and Eb and a decrease in ΔE. The efficiency (η) increased from 71.4% at x = 0 to 81.9% at x = 2 mol% with a Wre of 0.77 J/cm3. However, with an increase in the La content to x = 3 mol%, pores began to appear in the ceramics; thus, the efficiency (η) decreased to 75.6%. The energy storage characteristics exceed most of antiferroelectric ceramics. When x = 2 mol%, Pb1−xLax(Hf0.45Sn0.55)0.995O3 ceramic exhibits excellent temperature and frequency stability at 10–100 Hz and 10–110 °C and has broad application prospects. The Pb0.98La0.02(Hf0.45Sn0.55)0.995O3 ceramic is a prospective material for use in improved pulsed power capacitors, according to all of these data.

Author Contributions

Conceptualization, L.Y. and Y.Z.; data curation, Y.C. and J.Z.; funding acquisition, E.W.; investigation, L.Y. and J.L.; methodology, E.W., L.Y. and L.Z.; project administration, E.W., Y.Z. and L.Z.; resources, C.S., J.Z. and S.Z.; software, Y.C., C.S. and J.L.; writing—original draft preparation, E.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The Project of Henan Province Science and Technology (232102221003, 232102210183) and The Postgraduate Education Reform and Quality Improvement Project of Henan Province (YJS2023JD67).

Data Availability Statement

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

Acknowledgments

The authors are grateful to Jinping Zhang for her help with the preparation of figures in this paper.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Liu, Z.; Lu, T.; Ye, J.; Wang, G.; Dong, X.; Withers, R.; Liu, Y. Antiferroelectrics for energy storage applications: A review. Adv. Mater. Technol. 2018, 3, 1800111. [Google Scholar] [CrossRef]
  2. Wang, G.; Lu, Z.; Li, Y.; Li, L.; Ji, H.; Feteira, A.; Zhou, D.; Wang, D.; Zhang, S.; Reaney, I.M. Electroceramics for high-energy density capacitors: Current status and future perspectives. Chem. Rev. 2021, 121, 6124–6172. [Google Scholar] [CrossRef]
  3. Wang, H.; Liu, Y.; Yang, T.; Zhang, S.J. Ultrahigh energy-storage density in antiferroelectric ceramics with field-induced multiphase transitions. Adv. Funct. Mater. 2019, 29, 1807321. [Google Scholar] [CrossRef]
  4. Shirane, G.; Sawaguchi, E.; Takagi, Y. Dielectric properties of lead zirconate. Phys. Rev. J. Arch. 1951, 84, 476. [Google Scholar] [CrossRef]
  5. Shirane, G.; Sawaguchi, E.; Takeda, A. On the phase transition in lead zirconate. Phys. Rev. J. Arch. 1950, 80, 485. [Google Scholar] [CrossRef]
  6. Viehland, D. Transmission electron microscopy study of high-Zr-content lead zirconate titanate. Phys. Rev. B 1995, 52, 778–791. [Google Scholar] [CrossRef]
  7. Shirane, G.; Pepinsky, R. Phase Transitions in Antiferroelectric PbHfO3. Phys. Rev. 1953, 91, 812–815. [Google Scholar] [CrossRef]
  8. Kania, A.; Roleder, K.; Kugel, G.; Fontana, M.D. Raman scattering, central peak and phase transitions in AgNbO3. J. Phys. C Solid State Phys. 1986, 19, 9. [Google Scholar] [CrossRef]
  9. Yashima, M.; Matsuyama, S.; Sano, R.; Itoh, M.; Tsuda, K.; Fu, D. Structure of Ferroelectric Silver Niobate AgNbO3. Chem. Mater. 2011, 23, 1643–1645. [Google Scholar] [CrossRef]
  10. Mishra, S.K.; Choudhury, N.; Chaplot, S.L.; Krishna, P.S.R.; Mittal, R. Competing antiferroelectric and ferroelectric interactions inNaNbO3: Neutron diffraction and theoretical studies. Phys. Rev. B 2007, 76, 024110. [Google Scholar] [CrossRef]
  11. Saito, T.; Adachi, H.; Wada, T.; Adachi, H. Pulsed-laser deposition of ferroelectric NaNbO3 thin films. Jpn. J. Appl. Phys. 2005, 44, 6969–6972. [Google Scholar] [CrossRef]
  12. Koruza, J.; Groszewicz, P.; Breitzke, H.; Buntkowsky, G.; Rojac, T.; Malič, B. Grain-size-induced ferroelectricity in NaNbO3. Acta Mater. 2017, 126, 77–85. [Google Scholar] [CrossRef]
  13. Shuvaeva, V.A.; Antipin, M.Y.; Lindeman, R.S.V.; Fesenko, O.E.; Smotrakov, V.G.; Struchkov, Y.T. Crystal structure of the electric-fieldinduced ferroelectric phase of NaNbO3. Ferroelectrics 2011, 141, 307–311. [Google Scholar] [CrossRef]
  14. Beppu, K.; Shimasaki, T.; Fujii, I.; Imai, T.; Adachi, H.; Wada, T. Energy storage properties of antiferroelectric 0.92 NaNbO3-0.08 SrZrO3 film on (001) SrTiO3 substrate. Phys. Lett. A 2020, 384, 126690. [Google Scholar] [CrossRef]
  15. Luo, B.; Dong, H.; Wang, D.; Jin, K. Large recoverable energy density with excellent thermal stability in Mn-modified NaNbO3-CaZrO3 lead-free thin films. J. Am. Ceram. Soc. 2018, 101, 3460–3467. [Google Scholar] [CrossRef]
  16. Tian, Y.; Jin, L.; Zhang, H.; Xu, Z.; Wei, X.; Politova, E.D.; Stefanovich, S.Y.; Tarakina, N.V.; Abrahams, I.; Yan, H. High energy density in silver niobate ceramics. J. Mater. Chem. A 2016, 4, 17279–17287. [Google Scholar] [CrossRef]
  17. Wang, J.; Wan, X.; Rao, Y.; Zhao, L.; Zhu, K. Hydrothermal synthesized AgNbO3 powders: Leading to greatly improved electric breakdown strength in ceramics. J. Eur. Ceram. Soc. 2020, 40, 5589–5596. [Google Scholar] [CrossRef]
  18. Rashid, M.; Mahmood, Q.; Babar, F.; Ramay, S.M.; Mahmood, A. Study of mechanical, electronic and optical properties of PbZrO3 and PbHfO3; DFT approach. Mater. Res. Express 2019, 6, 066311. [Google Scholar] [CrossRef]
  19. Xu, R.; Zhu, Q.; Xu, Z.; Feng, Y.; Wei, X. High energy and power density achieved in Pb0.94La0.04HfO3 antiferroelectric ceramics with multiple phase transition. Appl. Phys. Lett. 2022, 120, 052904. [Google Scholar] [CrossRef]
  20. Hu, J.; Li, W.; Tang, X.; Shen, Z.; Wang, K.; Zhang, Y.; Zhang, S.; Jiang, Y.; Guo, X. Enhancement of energy storage density and efficiency of PbHfO3 doped with La antiferroelectric thin films. ACS Appl. Energy Mater. 2022, 6, 120–126. [Google Scholar] [CrossRef]
  21. Ma, C.-H.; Liao, Y.-K.; Zheng, Y.; Zhuang, S.; Lu, S.-C.; Shao, P.-W.; Chen, J.-W.; Lai, Y.-H.; Yu, P.; Hu, J.-M.; et al. Synthesis of a new ferroelectric relaxor based on a combination of antiferroelectric and paraelectric systems. ACS Appl. Mater. Interfaces 2022, 14, 22278–22286. [Google Scholar] [CrossRef]
  22. Ge, P.-Z.; Tang, X.-G.; Meng, K.; Huang, X.-X.; Li, S.-F.; Liu, Q.-X.; Jiang, Y.-P. Energy storage density and charge–discharge properties of PbHf1−xSnxO3 antiferroelectric ceramics. Chem. Eng. J. 2022, 429, 132540. [Google Scholar] [CrossRef]
  23. Ge, P.-Z.; Tang, X.-G.; Liu, Q.-X.; Jiang, Y.-P.; Guo, X.-B.J. Superior Energy and Power Density Realized in Pb (Hf1−xTix)O3 System at Low Electric Field. Energy Mater. Adv. 2023, 4, 0025. [Google Scholar] [CrossRef]
  24. Ge, P.-Z.; Tang, X.-G.; Meng, K.; Huang, X.-X.; Liu, Q.-X.; Jiang, Y.-P.; Gong, W.-P.; Wang, T. Ultrahigh energy storage density and superior discharge power density in a novel antiferroelectric lead hafnate. Mater. Today Phys. 2022, 24, 100681. [Google Scholar] [CrossRef]
  25. Yang, L.; Kong, X.; Li, F.; Hao, H.; Cheng, Z.; Liu, H.; Li, J.-F.; Zhang, S. Perovskite lead-free dielectrics for energy storage applications. Prog. Mater. Sci. 2019, 102, 72–108. [Google Scholar] [CrossRef]
  26. Liu, Z.; Lu, T.; Yan, S.; Liao, Q.; Dong, X.; Wang, G.; Liu, Y. Large electrocaloric and pyroelectric energy harvesting effect over a broad temperature range via modulating the relaxor behavior in non-relaxor ferroelectrics. J. Mater. Chem. A 2021, 9, 22015–22024. [Google Scholar] [CrossRef]
  27. Almousa, N.; Issa, S.A.; Salem, M.M.; Darwish, M.A.; Serag, E.N.; Nazrin, S.N.; Zakaly, H.M. Tailoring perovskite ceramics for improved structure, vibrational behaviors and radiation protection: The role of lanthanum in PbTiO3. Opt. Mater. 2024, 152, 115543. [Google Scholar] [CrossRef]
  28. Wang, S.B.; Zhao, P.F.; Jian, X.D.; Yao, Y.B.; Tao, T.; Liang, B.; Lu, S.G. Large energy storage density and electrocaloric strength of Pb0. 97La0. 02 (Zr0.46−xSn0.54Tix)O3 antiferroelectric thick film ceramics. Scr. Mater. 2022, 210, 114426. [Google Scholar] [CrossRef]
  29. Ren, P.; Ren, D.; Sun, L.; Yan, F.; Yang, S.; Zhao, G. Grain size tailoring and enhanced energy storage properties of two-step sintered Nd3+-doped AgNbO3. J. Eur. Ceram. Soc. 2020, 40, 4495–4502. [Google Scholar] [CrossRef]
  30. Luo, N.; Han, K.; Zhuo, F.; Xu, C.; Zhang, G.; Liu, L.; Chen, X.; Hu, C.; Zhou, H.; Wei, Y. Aliovalent A-site engineered AgNbO3 lead-free antiferroelectric ceramics toward superior energy storage density. J. Mater. Chem. A 2019, 7, 14118–14128. [Google Scholar] [CrossRef]
  31. Mao, S.; Luo, N.; Han, K.; Feng, Q.; Chen, X.; Peng, B.; Liu, L.; Hu, C.; Zhou, H.; Toyohisa, F.; et al. Effect of Lu doping on the structure, electrical properties and energy storage performance of AgNbO3 antiferroelectric ceramics. J. Mater. Sci. Mater. Electron. 2020, 31, 7731–7741. [Google Scholar] [CrossRef]
  32. Gao, J.; Zhang, Y.; Zhao, L.; Lee, K.-Y.; Liu, Q.; Studer, A.; Hinterstein, M.; Zhang, S.; Li, J.-F. Enhanced antiferroelectric phase stability in La-doped AgNbO3: Perspectives from the microstructure to energy storage properties. J. Mater. Chem. A 2019, 7, 2225–2232. [Google Scholar] [CrossRef]
  33. Gao, J.; Liu, Q.; Dong, J.; Wang, X.; Zhang, S.; Li, J.-F. Local structure heterogeneity in Sm-doped AgNbO3 for improved energy-storage performance. ACS Appl. Mater. Interfaces 2020, 12, 6097–6104. [Google Scholar] [CrossRef] [PubMed]
  34. Li, S.; Nie, H.; Wang, G.; Xu, C.; Liu, N.; Zhou, M.; Cao, F.; Dong, X. Significantly enhanced energy storage performance of rare-earth-modified silver niobate lead-free antiferroelectric ceramics via local chemical pressure tailoring. J. Mater. Chem. C 2019, 7, 1551–1560. [Google Scholar] [CrossRef]
  35. Wang, X.; Hao, X.; Zhang, Q.; An, S.; Chou, X. Energy-storage performance and pyroelectric energy harvesting effect of PNZST antiferroelectric thin films. J. Mater. Sci. Mater. Electron. 2017, 28, 1438–1448. [Google Scholar] [CrossRef]
  36. Bokov, A.A.; Ye, Z.G. Recent progress in relaxor ferroelectrics with perovskite structure. J. Mater. Sci. 2006, 41, 31–52. [Google Scholar] [CrossRef]
  37. Nguyen, M.D.; Houwman, E.P.; Dekkers, M.; Nguyen, C.T.; Vu, H.N.; Rijnders, G. Research update: Enhanced energy storage density and energy efficiency of epitaxial Pb0.9La0.1(Zr0.52Ti0.48)O3 relaxor-ferroelectric thin-films deposited on silicon by pulsed laser deposition. APL Mater. 2016, 4, 080701. [Google Scholar] [CrossRef]
  38. Liu, Z.; Chen, X.; Peng, W.; Xu, C.; Dong, X.; Cao, F.; Wang, G. Temperature-dependent stability of energy storage properties of Pb0.97La0.02(Zr0.58Sn0.335Ti0.085)O3 antiferroelectric ceramics for pulse power capacitors. Appl. Phys. Lett. 2015, 106, 262901. [Google Scholar] [CrossRef]
  39. Xu, R.; Xu, Z.; Feng, Y.; He, H.; Tian, J.; Huang, D. Temperature dependence of energy storage in Pb0.90La0.04Ba0.04[(Zr0.7Sn0.3)0.88Ti0.12]O3 antiferroelectric ceramics. J. Am. Ceram. Soc. 2016, 99, 2984–2988. [Google Scholar] [CrossRef]
  40. Zhang, H.; Chen, X.; Cao, F.; Wang, G.; Dong, X.; Hu, Z.; Du, T. Charge–discharge properties of an antiferroelectric ceramics capacitor under different electric fields. J. Am. Ceram. Soc. 2010, 93, 4015–4017. [Google Scholar] [CrossRef]
  41. Ciuchi, I.V.; Mitoseriu, L.; Galassi, C. Antiferroelectric to ferroelectric crossover and energy storage properties of (Pb1−xLax)(Zr0.90Ti0.10)1−x/4O3(0.02 ≤ x ≤ 0.04) ceramics. J. Am. Ceram. Soc. 2016, 99, 2382–2387. [Google Scholar] [CrossRef]
  42. Xu, R.; Xu, Z.; Feng, Y.; Wei, X.; Tian, J.; Huang, D. Polarization of antiferroelectric ceramics for pulse capacitors under transient electric field. J. Appl. Phys. 2016, 119, 224103. [Google Scholar] [CrossRef]
Figure 1. Hysteresis loops of energy storage ceramic materials: (a) linear dielectric; (b) ferroelectric; (c) relaxor ferroelectric; (d) antiferroelectric.
Figure 1. Hysteresis loops of energy storage ceramic materials: (a) linear dielectric; (b) ferroelectric; (c) relaxor ferroelectric; (d) antiferroelectric.
Crystals 14 00732 g001
Figure 2. (a) XRD results of PLHS ceramics with various La amounts; (b) XRD narrow-spectrum scanning of 2θ in 43.5°–45.0°.
Figure 2. (a) XRD results of PLHS ceramics with various La amounts; (b) XRD narrow-spectrum scanning of 2θ in 43.5°–45.0°.
Crystals 14 00732 g002
Figure 3. SEM results of PLHS ceramics with various La amounts at 2000 times: (a) x = 0; (b) x = 1 mol%; (c) x = 2 mol%; (d) x = 3 mol%.
Figure 3. SEM results of PLHS ceramics with various La amounts at 2000 times: (a) x = 0; (b) x = 1 mol%; (c) x = 2 mol%; (d) x = 3 mol%.
Crystals 14 00732 g003
Figure 4. (a) PLHS P-E hysteresis loop under a large electric field of 200 kV/cm; (b) Pmax, Pr, and Pmax-Pr; (c) Wst, Wre, η.
Figure 4. (a) PLHS P-E hysteresis loop under a large electric field of 200 kV/cm; (b) Pmax, Pr, and Pmax-Pr; (c) Wst, Wre, η.
Crystals 14 00732 g004
Figure 5. The recently reported performance parameters of antiferroelectric ceramics.
Figure 5. The recently reported performance parameters of antiferroelectric ceramics.
Crystals 14 00732 g005
Figure 6. The x = 2 mol% component at the electric field strength of 150 kV/cm: (a) the hysteresis loops at a frequency of 0.5–100 Hz; (b) the changes in Pmax, Pr, and Pmax-Pr; (c) the changes in Wst, Wre, and η.
Figure 6. The x = 2 mol% component at the electric field strength of 150 kV/cm: (a) the hysteresis loops at a frequency of 0.5–100 Hz; (b) the changes in Pmax, Pr, and Pmax-Pr; (c) the changes in Wst, Wre, and η.
Crystals 14 00732 g006
Figure 7. The x = 2 mol% component at 10–110 °C: (a) the hysteresis loops at a frequency of 0.5–100 Hz; (b) the changes in Pmax, Pr, and Pmax-Pr; (c) the changes in Wst, Wre, and η.
Figure 7. The x = 2 mol% component at 10–110 °C: (a) the hysteresis loops at a frequency of 0.5–100 Hz; (b) the changes in Pmax, Pr, and Pmax-Pr; (c) the changes in Wst, Wre, and η.
Crystals 14 00732 g007
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

Wang, E.; Yue, L.; Chu, Y.; Sun, C.; Zhao, J.; Zhang, S.; Liu, J.; Zhang, Y.; Zhang, L. High Energy Storage Performance in Pb1−xLax(Hf0.45Sn0.55)0.995O3 Antiferroelectric Ceramics. Crystals 2024, 14, 732. https://doi.org/10.3390/cryst14080732

AMA Style

Wang E, Yue L, Chu Y, Sun C, Zhao J, Zhang S, Liu J, Zhang Y, Zhang L. High Energy Storage Performance in Pb1−xLax(Hf0.45Sn0.55)0.995O3 Antiferroelectric Ceramics. Crystals. 2024; 14(8):732. https://doi.org/10.3390/cryst14080732

Chicago/Turabian Style

Wang, Erping, Liqin Yue, Yuanhong Chu, Caixia Sun, Jinyu Zhao, Siyu Zhang, Jiale Liu, Yangyang Zhang, and Ling Zhang. 2024. "High Energy Storage Performance in Pb1−xLax(Hf0.45Sn0.55)0.995O3 Antiferroelectric Ceramics" Crystals 14, no. 8: 732. https://doi.org/10.3390/cryst14080732

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