Next Article in Journal
Stent-Graft Fabrics Incorporating a Specific Corona Ready to Fenestrate
Next Article in Special Issue
Ferroelectric Phase Transition in Barium Titanate Revisited with Ab Initio Molecular Dynamics
Previous Article in Journal
Effects of Homogenization Heat Treatment on the Fe Micro-Segregation in Ti-1023 Titanium Alloy
Previous Article in Special Issue
Modulating the Performance of the SAW Strain Sensor Based on Dual-Port Resonator Using FEM Simulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Energy Storage Performance and Efficiency in Bi0.5(Na0.8K0.2)0.5TiO3-Bi0.2Sr0.7TiO3 Relaxor Ferroelectric Ceramics via Domain Engineering

1
Department of Materials Science and Engineering, Pukyong National University, 45 Yongso-ro, Nam-Gu, Busan 48513, Republic of Korea
2
School of Materials Science and Engineering, Kyungpook National University, 80 Daehak-ro, Buk-Gu, Daegu 41566, Republic of Korea
3
Department of Biofibers and Biomaterials Science, Kyungpook National University, Daegu 41566, Republic of Korea
*
Author to whom correspondence should be addressed.
Materials 2023, 16(14), 4912; https://doi.org/10.3390/ma16144912
Submission received: 5 June 2023 / Revised: 1 July 2023 / Accepted: 7 July 2023 / Published: 9 July 2023

Abstract

:
Dielectric materials are highly desired for pulsed power capacitors due to their ultra-fast charge-discharge rate and excellent fatigue behavior. Nevertheless, the low energy storage density caused by the low breakdown strength has been the main challenge for practical applications. Herein, we report the electric energy storage properties of (1 − x) Bi0.5(Na0.8K0.2)0.5TiO3-xBi0.2Sr0.7TiO3 (BNKT-BST; x = 0.15–0.50) relaxor ferroelectric ceramics that are enhanced via a domain engineering method. A rhombohedral-tetragonal phase, the formation of highly dynamic PNRs, and a dense microstructure are confirmed from XRD, Raman vibrational spectra, and microscopic investigations. The relative dielectric permittivity (2664 at 1 kHz) and loss factor (0.058) were gradually improved with BST (x = 0.45). The incorporation of BST into BNKT can disturb the long-range ferroelectric order, lowering the dielectric maximum temperature Tm and inducing the formation of highly dynamic polar nano-regions. In addition, the Tm shifts toward a high temperature with frequency and a diffuse phase transition, indicating relaxor ferroelectric characteristics of BNKT-BST ceramics, which is confirmed by the modified Curie-Weiss law. The rhombohedral-tetragonal phase, fine grain size, and lowered Tm with relaxor properties synergistically contribute to a high Pmax and low Pr, improving the breakdown strength with BST and resulting in a high recoverable energy density Wrec of 0.81 J/cm3 and a high energy efficiency η of 86.95% at 90 kV/cm for x = 0.45.

1. Introduction

Materials with high energy and power have received extensive attention for high-power applications, such as microwaves, electromagnetic devices, pulsed power devices, hybrid electric vehicles, high-frequency inverters, and other energy storage devices [1,2,3]. In particular, dielectric ceramics are the most promising materials for energy storage applications due to their super-fast charge-discharge rate and excellent temperature stability compared to electrochemical energy storage devices (batteries and electrochemical capacitors) and dielectric polymers [4,5,6]. However, the dielectric capacitor’s energy storage density and efficiency are much lower than those of polymers/batteries due to their low dielectric breakdown strength (DBS), which restricts their practical application in energy storage devices.
The recoverable energy density (Wrec) of a dielectric capacitor is governed by the applied electric field (E) and induced polarization (P), expressed by the following equation, usually estimated from the P-E loop, and it is schematically shown in Figure 1a by the shaded area with cyan color [7,8,9,10].
W r e c = P r P m a x E   d P
where Pmax and Pr are the maximum polarization and remnant polarization, respectively (Figure 1a). Energy efficiency (η) can be estimated by the following equation [8,9,10].
η = W r e c W r e c + W l o s s
where Wloss is the hysteresis loss. According to Equation (1), the energy storage properties can be significantly enhanced by increasing the difference between Pr and PmaxP). The breakdown electric field (EBD) is also an essential factor for energy storage; i.e., a higher DBS is responsible for a large energy storage density.
Linear dielectrics (LDEs), ferroelectrics (FEs), relaxor ferroelectrics (RFEs), and anti-ferroelectrics (AFEs) have been widely explored for electrostatic energy storage applications [11]. LDEs (Al2O3 and SrTiO3) typically show low relative dielectric permittivity and loss factor, high EBD, and free hysteresis loop with low polarization, resulting in poor Wrec and high η [11,12,13]. The AFEs (PbZrO3, PbHfO3, AgNbO3, and NaNbO3) display elevated polarization and substantial hysteresis because of phase transition between AFE and FE phases induced by the external field, leading to extremely high Wrec and low η [14,15,16,17,18]. The RFEs (Bi0.5Na0.5TiO3 (BNT), BaTiO3 (BT), and BiFeO3 (BFO)) display moderate Pmax and small Pr (Figure 1b), arising from widespread PNRs, which generally exhibit highly dynamic short-range FE orders. However, these PNRs gradually transform into long-range FE orders with an increasing field, resulting in large Pmax. After eliminating the electric field, the induced FE orders will easily revert to PNRs, leading to small Pr [19,20,21,22]. Hence, the RFEs usually show high Wrec and η. Therefore, lead-based RFE materials have been widely investigated for energy storage applications [23,24]. However, lead is hazardous to the environment and human health due to its toxicity, which has motivated the development of alternative lead-free materials. In recent years, lead-free perovskite-structured (ABO3) RFEs, such as BT [25,26,27], BNT [6,28,29,30,31], BFO [32,33,34], and other lead-free perovskite RFEs, such as Bi4Ti3O12 [35], Sr1.25Bi2.75Nb1.25Ti1.75O12 [36], and Sr0.6Ba0.4Nb2O6 [37] based materials with boosted energy storage performance, have been reported for applications in energy storage devices.
Perovskite-structured BNT-based ceramics exhibit a strong ferroelectric response, since Bi3+ has a lone pair of electrons (6s2), which strongly hybridizes with the oxygen 2p orbital [38]. Furthermore, the formation of highly dynamic polar nano-regions (PNRs) are facilitated by local random fields induced by valency differences and compositional inhomogeneity [39,40]. Moreover, the relaxor behavior of the material can be improved by adding another phase with a similar perovskite to form a solid solution or modifying the base compound with a suitable dopant, which enables slim P-E loops [41]. Sayyed et al. [42] investigated the local structural deformation and dielectric anomalies near the morphotropic phase boundary (MPB) of (1 − x) Na0.5Bi0.5TiO3-xSrTiO3 ceramics. The ferroelectric response of (1 − x) Na0.5Bi0.5TiO3-SrTiO3-xAgNbO3 ceramics is similar to the antiferroelectric response and improved energy storage performance [43]. Shi et al. [44] reported that Zr- and Sm-doped 0.74Na0.5Bi0.5TiO3-0.26SrTiO3 ceramics significantly enhanced energy storage performance and the DBS. Bi0.2Sr0.7TiO3 (BST) exhibits strong polarization and a wide-phase transition temperature with diffused dielectric maxima. It was incorporated into BNT ceramics, suppressing the field-generated ferroelectric phase and achieving a large Pmax and small Pr [45,46]. Recently, Li et al. reported a synergistic approach to enhance the energy storage response in BNT-based RFEs by introducing PNRs and lowering the transition temperature by stabilizing the AFE responses at low temperatures [9].
In this work, we investigate a domain engineering process to improve the energy storage performance by modifying Bi0.5(Na0.8K0.2)0.5TiO3 (BNKT) RFEs with BST, since Bi0.5(Na1−xKx)0.5TiO3 exhibits a stronger ferroelectric response with relaxor behavior at the MPB at x = 0.16–0.2 [47,48] than pure BNT. It is revealed that the addition of BST can disturb the long-range ferroelectric order and transform the ferroelectric microdomains of BNKT into highly dynamic PNRs. This results in a macroscopic ferroelectric to relaxor ferroelectric transition, as schematically illustrated in Figure 1b. The favorable relaxor ferroelectric state formed by the domain engineering method simultaneously produces a large Pmax and reduced Pr, which facilitates the enhancement in DBS with the BST, resulting in high energy density and high efficiency of the BNKT-BST RFEs.

2. Materials and Methods

(1 − x) Bi0.5(Na0.8K0.2)0.5TiO3–xBi0.2Sr0.7TiO3 (BNKT-BST; x = 0.15, 0.30, 0.40, 0.45, and 0.50) RFE ceramics were fabricated via a conventional solid-state reaction method. To prepare BNKT and BST, high-purity raw materials of Bi2O3 (Sigma-Aldrich, St. Louis, MI, USA, 99.9%), Na2CO3 (Sigma-Aldrich, St. Louis, MI, USA, 99.5%), K2CO3 (Sigma-Aldrich, St. Louis, MI, USA, 99%), TiO2, (Sigma-Aldrich, St. Louis, MI, USA, 99%), and SrCO3 (Sigma-Aldrich, St. Louis, MI, USA, 98%) were weighed according to the nominal stoichiometric compositions and then ball-milled using a planetary ball mill for 24 h with ZrO2 balls in ethanol. After the slurries were dried at 120 °C, the mixture of BNKT and BST powders was calcined at 800 °C and 950 °C for 2 h and 3 h, respectively, to form a pure phase of Bi0.5(Na0.8K0.2)0.5TiO3 and Bi0.2Sr0.7TiO3. Both BNKT and BST calcined powders were mixed and ball-milled for 12h to prepare a BNKT-BST composition. Further, these powders were granulated with 5 wt.% polyvinyl alcohol (Sigma-Aldrich, 99%, St. Louis, MI, USA,) and uniaxially pressed into disks, at a pressure of 10 MPa, of 10 mm diameter and ~0.5 mm thickness, followed by sintering at 1100 °C for 3 h. To perform electrical measurements, a silver paste (ELCOAT, Electroconductives) was coated on both sides of the sintered disks.
The phase formation of the BNKT-BST ceramic samples was examined using an X-ray diffractometer (Rigaku, Tokyo, Japan, TTRAX III 18 kW) with monochromatic Cu-Kα radiation (λ = 1.5406 Å). Raman spectra were recorded using a Raman spectrometer (JOBIN YVON, Oberursel, Germany, LABRAM HR800) with a laser wavelength of 532.06 nm. Surface morphology was investigated using a field emission scanning electron microscope (FESEM) (JEOL, Tokyo, Japan, JSM-7610F). Temperature- and frequency-dependent dielectric properties were measured from room temperature (RT) to 450 °C and 1 kHz–1 MHz using an impedance analyzer (Hewlett Packard, Palo Alto, CA, USA, 4294A). P-E, I-E loops, and fatigue behavior were measured using a ferroelectric tester (AixACCT Systems GmbH, Aachen, Germany, TF Analyzer 2000).

3. Results and Discussion

3.1. Phase Evolution and Microstructure

Figure 2a shows the X-ray diffraction (XRD) patterns of BNKT-BST ceramics (x = 0.15–0.50) in the 2θ range of 20–70°. All of the samples revealed a rhombohedral and tetragonal crystal structure, indicating the diffusion of BST into BNKT and the formation of BNKT-BST as a homogeneous solid solution. At RT, the BNKT system exhibits a rhombohedral and tetragonal crystal structure near MPB at x = 0.16–0.2 [47,48]. The formation of MPB in BNKT-BST ceramics is confirmed by the splitting of the (021)/(111) and (122)/(211) peaks at 2θ around 40° and 58°, respectively, which is shown in Figure 2b. Similar splitting and formation of MPB were observed in Bi0.5(Na1–xKx)0.5TiO3]-BiAlO3 [49], Bi0.5Na0.5TiO3-Bi0.5K0.5TiO3-Bi0.5Li0.5TiO3 [50], and (Bi0.5Na0.5)TiO3-(Bi0.5K0.5)TiO3-BaTiO3 [51] ceramics. In addition, both the (021) and (122) peaks shifted slightly toward lower angles with increasing BST into BNKT, demonstrating enhanced lattice parameters (Figure 2b). The enhancement in lattice parameters can be attributed to the ionic radius of Sr2+ (1.44 Å), which is larger than that of Bi3+ (1.36 Å), Na+ (1.39 Å), and K+ (1.38 Å), respectively, at the A-site [52,53,54].
Figure 2c shows the Raman spectra of BNKT-BST along with spectral de-convolution in the Raman shift of 50–1000 cm−1. The Raman spectra of all compositions are similar to the previous reports of BNKT-based ceramics [51,55]. The Raman active bands are divided into four Raman vibration modes, as shown at the top of Figure 2c. (i) The modes below 200 cm−1 are related to the vibration of the A-site (Bi-O, Na-O, K-O, and Sr-O); (ii) the modes between 200 and 440 cm−1 correspond to the vibrations of B-O (Ti-O); (iii) the modes between 440 and 700 cm−1 correspond to the vibrations of BO6 (TiO6)-octahedra; and (iv) the modes above 700 cm−1 are related to the A1 and E (longitudinal optical) overlapping modes [55]. The modes appearing at 124–172 cm−1 and 768 cm−1 are shifted to the higher wavenumbers of 128–189 cm−1 and 779 cm−1 with BST, associated with the A-site and A1 + E vibrations caused by A-site disorder. Such a disorder is induced by the incorporation of BST (Bi3+ and Sr2+) into the BNKT (Bi3+, Na+, and K+) system [56]. In addition, a noticeable change at 250 and 320 cm−1 shifted toward a lower wavenumber of 233 and 305 cm−1 with BST, which is caused by an increase in the B-site disorder in the BNKT-BST system [53]. Moreover, these modes are slightly broadened, confirming the disturbance of the long-range ferroelectric order and the formation of highly dynamic PNRs, improving the relaxor characteristics of BNKT-BST [44]. This result is consistent with the XRD and electrical properties presented in Section 3.2 and Section 3.3.
Figure 3 shows FESEM images of the BNKT-BST ceramics. All of the ceramics display rectangular-shaped grains, which are homogeneously distributed. Figure 3d clearly shows that the x = 0.45 composition exhibits a highly dense microstructure and is more compact with smaller grains, as compared to the other samples of BNKT-BST (x < 0.45 and x = 0.50). To prove that all of the samples are homogeneous and highly dense, the density of sintered BNKT-BST ceramic samples was calculated using the Archimedes principle. It increased with BST from 5.58 g/cm3 to 5.73 g/cm3 for x = 0.15 to 0.45 and further decreased (5.54 g/cm3) for x = 0.50. The calculated relative density of BNKT-BST ranged from 94.96% to 98.17% of the theoretical density [57], confirming that these samples are homogeneous and highly dense. Further, the average grain size of the BNKT-BST ceramics was estimated using Image-J software (Wayne Rasband and contributors, National Institutes of Health, USA, ImageJ 1.53t) via the linear intercept method and found to be 1.37 µm for the x = 0.15 composition; it gradually enhanced to 1.6 µm with the incorporation of BST. Grain size enhancement is caused by the generation of oxygen vacancies by Sr2+ entering the perovskite of BNKT and being substituted at the A-site of Bi3+, Na+, and K+ [58]. Previous reports have investigated that the fine grain size with a homogeneous and dense microstructure can withstand higher electric fields, leading to high DBS, and improve energy storage performance [59,60].

3.2. Dielectric Properties and Relaxor Behavior

Figure 4a displays the frequency variation in the relative dielectric permittivity (εr) (solid line) and loss factor (tan δ) (dot line) of BNKT-BST ceramic capacitors, measured at RT in the range of 1 kHz to 1 MHz. The sample x = 0.15 displayed a higher εr of 1481 and tan δ of 0.231 at 1 kHz than pure BNKT (εr of 1273 and tan δ of 0.047 at 1 kHz), as reported in our previous report [48]. These εr values gradually enhanced to 2664, and the tan δ values reduced to 0.058 for the x = 0.45 sample (Figure 4b). The enhancement in the dielectric properties is attributed to the incorporation of BST into BNKT and the dense microstructure.
Figure 4c,d displays the temperature dependence of εr and tan δ of BNKT-BST for the lower and higher compositions (x = 0.15 and 0.45), measured at various frequencies (0.1 kHz to 1 MHz). It was observed that the dielectric maximum temperature (Tm) shifted toward a lower temperature 53 °C with the incorporation of BST for x = 0.45 (Figure 4d), as compared to x = 0.15 (Tm = 345 °C) (Figure 4c) and pure BNKT (300 °C) [48], and this is similar to the 0.74 Na0.5Bi0.5TiO3-0.26 SrTiO3 ceramics reported by Shi et al. [44]. The incorporation of BST into BNKT can disturb the long-range ferroelectric order, resulting in a lowered Tm. This lower Tm leads to the formation of highly dynamic PNRs due to the mismatch of the ionic radius at the A-site of BNKT-BST. In addition, the Tm shifted toward higher temperatures, and dielectric peaks diffused with an increase in frequency. This frequency dispersion with a diffuse phase transition reveals typical relaxor ferroelectric characteristics [61,62]. The degree of the relaxor characteristics was determined using the modified Curie-Weiss law via the following equation [52,63].
1 ε r 1 ε r m = T T m γ C
where ε r m is the maximum relative dielectric permittivity at the maximum temperature Tm, T is the temperature, γ is the degree of relaxation, and C is the Curie constant. Generally, the γ value is 1 for normal ferroelectrics and between 1 and 2 for relaxor ferroelectrics [63]. The insets of Figure 4c,d show the log-log plots of (1/ ε r     1/ ε r m ) vs. (TTm) of BNKT-BST for x = 0.15 and 0.45, measured at 1 MHz. The value of γ slightly increased from 1.80 to 1.83, proving that there is an increase in relaxor behavior with BST from x = 0.15 to 0.45, leading to an increase in the energy storage performance. This is consistent with Raman’s results and previous reports [4,5,6].

3.3. FE-RFE Transformation, Domain Evolution, and Energy Storage Performance

Figure 5 displays the RT bipolar P-E hysteresis loops and current (I)-electric field (E) curves of BNKT-BST ceramic capacitors measured at various electric fields and 10 Hz. The BNKT-BST (x = 0.15) sample exhibits a typical ferroelectric (FE) characteristic, displaying high remnant polarization Pr of 19.89 µC/cm2, high maximum polarization Pmax of 31.46 µC/cm2, and a high coercive field Ec of 16.66 kV/cm. These values, listed in Table 1, gradually decreased, whereas the Emax or EBD increased from 57.42 kV/cm to 90 kV/cm with the incorporation of BST (x = 0.45), which is favorable for high energy storage density (Figure 5f). It is evident that the two peaks in the I-E curves (x ≥ 0.30) and slim P-E loops are attributed to the formation of highly dynamic PNRs, which can commonly be seen in RFEs [64]. In general, the P-E loops present in normal FEs are due to the macroscopic domain wall motion, while in RFEs, highly dynamic PNRs exist instead of macrodomains, resulting in slim P-E loops [19].
Further, the Wrec was calculated via Equation (1) from P-E loops, which are shown in Figure 6 (cyan shaded area). The Wloss is calculated by the enclosed area of the P-E loops in the first quadrant (magenta shaded area), and η is calculated by Equation (2); they are listed in Table 1. The Wrec values gradually increased, the Wloss values decreased with the substitution of BST, and the composition x = 0.45 displays a high energy density of 0.81 J/cm3 at an EBD of 90 kV/cm and high energy efficiency of 86.95% (Figure 6f). The improvement in the energy storage performance is achieved via the domain engineering method by modifying BNKT with BST. It can be understood that the substitution of BST can transform the ferroelectric microdomains of BNKT into highly dynamic PNRs, resulting in a macroscopic FE to RFE transition. This domain evolution and transformation of FE to RFE transition in the present samples is schematically shown in Figure 1b. The highly dynamic PNRs induced large Pmax and low Pr, which improved the DBS with the incorporation of BST, resulting in high energy storage density and high energy efficiency of the BNKT-BST RFEs [65]. The obtained Wrec and η of 0.55 BNKT-0.45 BST are comparable/superior to other lead-free RFEs and are promising for energy storage capacitors [64,65,66,67,68,69].
Electrical fatigue endurance is an important property necessary for energy storage applications. Therefore, the fatigue behavior of BNKT-BST ceramic capacitors was measured up to 106 electric cycles at a frequency of 10 Hz under an electric field of 90 kV/cm. Figure 7 shows the unipolar P-E loops of BNK-BST (x = 0.45) and corresponding Wrec (square line) and η (circle line) values measured after various electric cycles (black, red and blue colour P-E loops measured at 100 and 103 and 106, respectively, as shown in inset of Figure 7). It is observed that the slender P-E loops are without significant change, revealing an excellent fatigue-free response and negligible variations in Wrec and η.

4. Conclusions

The domain-engineered relaxor ferroelectric BNKT-BST lead-free ceramics were fabricated by a solid-state reaction method and demonstrated structural, microstructural, dielectric, and ferroelectric properties in detail. XRD, Raman spectra, and FESEM studies revealed the formation of a rhombohedral-tetragonal phase, highly dynamic PNRs, and dense microstructure. The dielectric properties were improved with BST, and a high εr of 2664 and low tan δ of 0.058 at 1 kHz were obtained for the x = 0.45 composition. The incorporation of BST into BNKT can disturb the long-range ferroelectric order, causing lowered Tm and the formation of highly dynamic PNRs. In addition, the Tm shifts toward a high temperature with frequency and diffuse phase transition, indicating relaxor ferroelectric characteristics of BNKT-BST ceramics, and is confirmed via the modified Curie-Weiss law. The rhombohedral-tetragonal phase, fine grain size, and lowered Tm with relaxor properties simultaneously contribute to a high Pmax and low Pr. This improves the DBS and gives rise to giant energy storage density and high energy efficiency of the BNKT-BST RFEs, making this material a good candidate for pulse-driving energy storage applications.

Author Contributions

Conceptualization, measurement, data curation, investigation, visualization, writing—original draft preparation, review, and editing, S.P.; data curation, formal analysis, H.C.; data curation, Y.L.; writing—review and editing, K.-I.P.; writing—review and editing, K.C.; validation, project administration, writing—review and editing, supervision, G.-T.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by an NRF grant from the Korean government (MSIT) (no. 2022R1A2C4001497).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shen, Y.; Zhang, X.; Li, M.; Lin, Y.; Nan, C.-W. Polymer Nanocomposite Dielectrics for Electrical Energy Storage. Natl. Sci. Rev. 2017, 4, 23–25. [Google Scholar] [CrossRef] [Green Version]
  2. Yao, Z.; Song, Z.; Hao, H.; Yu, Z.; Cao, M.; Zhang, S.; Lanagan, M.T.; Liu, H. Homogeneous/Inhomogeneous-Structured Dielectrics and Their Energy-Storage Performances. Adv. Mater. 2017, 29, 1601727. [Google Scholar] [CrossRef]
  3. Palneedi, H.; Peddigari, M.; Hwang, G.-T.; Jeong, D.-Y.; Ryu, J. High-Performance Dielectric Ceramic Films for Energy Storage Capacitors: Progress and Outlook. Adv. Funct. Mater. 2018, 28, 1803665. [Google Scholar] [CrossRef]
  4. Li, Q.; Chen, L.; Gadinski, M.R.; Zhang, S.; Zhang, G.; Li, H.U.; Iagodkine, E.; Haque, A.; Chen, L.-Q.; Jackson, T.N.; et al. Flexible High-Temperature Dielectric Materials from Polymer Nanocomposites. Nature 2015, 523, 576–579. [Google Scholar] [CrossRef]
  5. Kumar, N.; Ionin, A.; Ansell, T.; Kwon, S.; Hackenberger, W.; Cann, D. Multilayer Ceramic Capacitors Based on Relaxor BaTiO3-Bi(Zn1/2Ti1/2)O3 for Temperature Stable and High Energy Density Capacitor Applications. Appl. Phys. Lett. 2015, 106, 252901. [Google Scholar] [CrossRef]
  6. Li, J.; Li, F.; Xu, Z.; Zhang, S. Multilayer Lead-Free Ceramic Capacitors with Ultrahigh Energy Density and Efficiency. Adv. Mater. 2018, 30, 1802155. [Google Scholar] [CrossRef]
  7. Chen, Q.; Shen, Y.; Zhang, S.; Zhang, Q.M. Polymer-Based Dielectrics with High Energy Storage Density. Annu. Rev. Mater. Res. 2015, 45, 433–458. [Google Scholar] [CrossRef]
  8. Xie, A.; Zuo, R.; Qiao, Z.; Fu, Z.; Hu, T.; Fei, L. NaNbO3-(Bi0.5 Li0.5 )TiO3 Lead-Free Relaxor Ferroelectric Capacitors with Superior Energy-Storage Performances via Multiple Synergistic Design. Adv. Energy Mater. 2021, 11, 2101378. [Google Scholar] [CrossRef]
  9. Li, T.; Jiang, X.; Li, J.; Xie, A.; Fu, J.; Zuo, R. Ultrahigh Energy-Storage Performances in Lead-Free Na0.5 Bi0.5 TiO3 -Based Relaxor Antiferroelectric Ceramics through a Synergistic Design Strategy. ACS Appl. Mater. Interfaces 2022, 14, 22263–22269. [Google Scholar] [CrossRef]
  10. Li, B.; Yan, Z.; Zhou, X.; Qi, H.; Koval, V.; Luo, X.; Luo, H.; Yan, H.; Zhang, D. Achieving Ultrahigh Energy Storage Density of La and Ta Codoped AgNbO3 Ceramics by Optimizing the Field-Induced Phase Transitions. ACS Appl. Mater. Interfaces 2023, 15, 4246–4256. [Google Scholar] [CrossRef]
  11. Yang, Z.; Du, H.; Jin, L.; Poelman, D. High-Performance Lead-Free Bulk Ceramics for Electrical Energy Storage Applications: Design Strategies and Challenges. J. Mater. Chem A Mater. 2021, 9, 18026–18085. [Google Scholar] [CrossRef]
  12. Okhay, O.; Vilarinho, P.M.; Tkach, A. Structure, Microstructure, and Dielectric Response of Polycrystalline Sr1-XZnxTiO3 Thin Films. Coatings 2023, 13, 165. [Google Scholar] [CrossRef]
  13. Pu, Y.; Wang, W.; Guo, X.; Shi, R.; Yang, M.; Li, J. Enhancing the Energy Storage Properties of Ca0.5 Sr0.5 TiO3Based Lead-Free Linear Dielectric Ceramics with Excellent Stability through Regulating Grain Boundary Defects. J. Mater. Chem. C Mater. 2019, 7, 14384–14393. [Google Scholar] [CrossRef]
  14. Chauhan, V.; Wang, B.-X.; Ye, Z.-G. Structure, Antiferroelectricity and Energy-Storage Performance of Lead Hafnate in a Wide Temperature Range. Materials 2023, 16, 4144. [Google Scholar] [CrossRef]
  15. Zhu, L.-F.; Deng, S.; Zhao, L.; Li, G.; Wang, Q.; Li, L.; Yan, Y.; Qi, H.; Zhang, B.-P.; Chen, J.; et al. Heterovalent-Doping-Enabled Atom-Displacement Fluctuation Leads to Ultrahigh Energy-Storage Density in AgNbO3-Based Multilayer Capacitors. Nat. Commun. 2023, 14, 1166. [Google Scholar] [CrossRef] [PubMed]
  16. Moradi, P.; Taheri-Nassaj, E.; Yourdkhani, A.; Mykhailovych, V.; Diaconu, A.; Rotaru, A. Enhanced Energy Storage Performance in Reaction-Sintered AgNbO3 Antiferroelectric Ceramics. Dalton Trans. 2023, 52, 4462–4474. [Google Scholar] [CrossRef] [PubMed]
  17. An, K.; Li, G.; Fan, T.; Huang, F.; Wang, W.; Wang, J. Enhanced Energy Storage Performance of AgNbO3:XCeO2 by Synergistic Strategies of Tolerance Factor and Density Regulations. Coatings 2023, 13, 534. [Google Scholar] [CrossRef]
  18. Zhang, M.-H.; Ding, H.; Egert, S.; Zhao, C.; Villa, L.; Fulanović, L.; Groszewicz, P.B.; Buntkowsky, G.; Kleebe, H.-J.; Albe, K.; et al. Tailoring High-Energy Storage NaNbO3-Based Materials from Antiferroelectric to Relaxor States. Nat. Commun. 2023, 14, 1525. [Google Scholar] [CrossRef] [PubMed]
  19. 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]
  20. Gui, D.-Y.; Ma, X.-Y.; Yuan, H.-D.; Wang, C.-H. Mn- and Yb-Doped BaTiO3-(Na0.5Bi0.5)TiO3 Ferroelectric Relaxor with Low Dielectric Loss. Materials 2023, 16, 2229. [Google Scholar] [CrossRef]
  21. Shi, H.; Li, K.; Li, F.; Ma, J.; Tu, Y.; Long, M.; Lu, Y.; Gong, W.; Wang, C.; Shan, L. Enhanced Piezoelectricity and Thermal Stability of Electrostrain Performance in BiFeO3-Based Lead-Free Ceramics. Nanomaterials 2023, 13, 942. [Google Scholar] [CrossRef]
  22. Ji, H.; Wang, D.; Bao, W.; Lu, Z.; Wang, G.; Yang, H.; Mostaed, A.; Li, L.; Feteira, A.; Sun, S.; et al. Ultrahigh Energy Density in Short-Range Tilted NBT-Based Lead-Free Multilayer Ceramic Capacitors by Nanodomain Percolation. Energy Storage Mater. 2021, 38, 113–120. [Google Scholar] [CrossRef]
  23. Zhang, G.; Zhu, D.; Zhang, X.; Zhang, L.; Yi, J.; Xie, B.; Zeng, Y.; Li, Q.; Wang, Q.; Jiang, S. High-Energy Storage Performance of (Pb0.87 Ba0.1 La0.02 )(Zr0.68 Sn0.24 Ti0.08 )O3 Antiferroelectric Ceramics Fabricated by the Hot-Press Sintering Method. J. Am. Ceram. Soc. 2015, 98, 1175–1181. [Google Scholar] [CrossRef]
  24. Liu, Z.; Chen, X.; Peng, W.; Xu, C.; Dong, X.; Cao, F.; Wang, G. Temperature-Dependent Stability of Energy Storage Properties of Pb0.97 La0.02 (Zr0.58 Sn0.335 Ti0.085 )O3 Antiferroelectric Ceramics for Pulse Power Capacitors. Appl. Phys. Lett 2015, 106, 262901. [Google Scholar] [CrossRef]
  25. Yuan, Q.; Li, G.; Yao, F.-Z.; Cheng, S.-D.; Wang, Y.; Ma, R.; Mi, S.-B.; Gu, M.; Wang, K.; Li, J.-F.; et al. Simultaneously Achieved Temperature-Insensitive High Energy Density and Efficiency in Domain Engineered BaTiO3-Bi(Mg0.5Zr0.5)O3 Lead-Free Relaxor Ferroelectrics. Nano Energy 2018, 52, 203–210. [Google Scholar] [CrossRef]
  26. Zhao, P.; Wang, H.; Wu, L.; Chen, L.; Cai, Z.; Li, L.; Wang, X. High-Performance Relaxor Ferroelectric Materials for Energy Storage Applications. Adv. Energy Mater. 2019, 9, 1803048. [Google Scholar] [CrossRef]
  27. Asbani, B.; Gagou, Y.; Ben Moumen, S.; Dellis, J.-L.; Lahmar, A.; Amjoud, M.; Mezzane, D.; El Marssi, M.; Rozic, B.; Kutnjak, Z. Large Electrocaloric Responsivity and Energy Storage Response in the Lead-Free Ba(GexTi1−x)O3 Ceramics. Materials 2022, 15, 5227. [Google Scholar] [CrossRef]
  28. Yin, J.; Zhang, Y.; Lv, X.; Wu, J. Ultrahigh Energy-Storage Potential under Low Electric Field in Bismuth Sodium Titanate-Based Perovskite Ferroelectrics. J. Mater. Chem. A Mater. 2018, 6, 9823–9832. [Google Scholar] [CrossRef]
  29. Qi, H.; Zuo, R. Linear-like Lead-Free Relaxor Antiferroelectric (Bi0.5 Na0.5 )TiO3-NaNbO3 with Giant Energy-Storage Density/Efficiency and Super Stability against Temperature and Frequency. J. Mater. Chem. A Mater. 2019, 7, 3971–3978. [Google Scholar] [CrossRef]
  30. Lu, Y.; Zhang, H.; Yang, H.; Fan, P.; Samart, C.; Takesue, N.; Tan, H. SPS-Prepared High-Entropy (Bi0.2Na0.2Sr0.2Ba0.2Ca0.2)TiO3 Lead-Free Relaxor-Ferroelectric Ceramics with High Energy Storage Density. Crystals 2023, 13, 445. [Google Scholar] [CrossRef]
  31. Jiang, Y.; Niu, X.; Liang, W.; Jian, X.; Shi, H.; Li, F.; Zhang, Y.; Wang, T.; Gong, W.; Zhao, X.; et al. Enhanced Energy Storage Performance in Na0.5Bi0.5TiO3-Based Relaxor Ferroelectric Ceramics via Compositional Tailoring. Materials 2022, 15, 5881. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, G.; Li, J.; Zhang, X.; Fan, Z.; Yang, F.; Feteira, A.; Zhou, D.; Sinclair, D.C.; Ma, T.; Tan, X.; et al. Ultrahigh Energy Storage Density Lead-Free Multilayers by Controlled Electrical Homogeneity. Energy Env. Sci. 2019, 12, 582–588. [Google Scholar] [CrossRef]
  33. Pan, H.; Li, F.; Liu, Y.; Zhang, Q.; Wang, M.; Lan, S.; Zheng, Y.; Ma, J.; Gu, L.; Shen, Y.; et al. Ultrahigh-Energy Density Lead-Free Dielectric Films via Polymorphic Nanodomain Design. Science 1979 2019, 365, 578–582. [Google Scholar] [CrossRef]
  34. Yang, Z.; Wang, B.; Li, Y.; Hall, D.A. Enhancement of Nonlinear Dielectric Properties in BiFeO3-BaTiO3 Ceramics by Nb-Doping. Materials 2022, 15, 2872. [Google Scholar] [CrossRef]
  35. Wang, D.; Liu, J.; Zeng, M.; Zhang, C.; Li, H.; Yu, H.; Yuan, Y.; Zhang, S. Stabilizing Temperature-capacitance Dependence of (Sr, Pb, Bi)TiO3-Bi4Ti3O12 Solutions for Energy Storage. J. Am. Ceram. Soc. 2019, 102, 4029–4037. [Google Scholar] [CrossRef]
  36. Wendari, T.P.; Zulhadjri. Emriadi Observation of Relaxor Ferroelectric Behavior and Energy Storage Performances in Sr1.25Bi2.75Nb1.25Ti1.75O12 Aurivillius Ceramic Synthesized by Molten Salt Method. J. Solid State Chem. 2023, 325, 124150. [Google Scholar] [CrossRef]
  37. Luo, C.; Zheng, X.; Zheng, P.; Du, J.; Niu, Z.; Zhang, K.; Bai, W.; Fan, Q.; Zheng, L.; Zhang, Y. Realizing Excellent Energy Storage Performances in Tetragonal Tungsten Bronze Ceramics via a B-Site Engineering Strategy. J. Alloys Compd. 2023, 933, 167809. [Google Scholar] [CrossRef]
  38. Li, F.; Hou, X.; Wang, J.; Zeng, H.; Shen, B.; Zhai, J. Structure-Design Strategy of 0–3 Type (Bi0.32Sr0.42Na0.20)TiO3/MgO Composite to Boost Energy Storage Density, Efficiency and Charge-Discharge Performance. J. Eur. Ceram. Soc. 2019, 39, 2889–2898. [Google Scholar] [CrossRef]
  39. Liu, N.; Liang, R.; Zhou, Z.; Dong, X. Designing Lead-Free Bismuth Ferrite-Based Ceramics Learning from Relaxor Ferroelectric Behavior for Simultaneous High Energy Density and Efficiency under Low Electric Field. J. Mater. Chem. C Mater. 2018, 6, 10211–10217. [Google Scholar] [CrossRef]
  40. Wu, Y.; Fan, Y.; Liu, N.; Peng, P.; Zhou, M.; Yan, S.; Cao, F.; Dong, X.; Wang, G. Enhanced Energy Storage Properties in Sodium Bismuth Titanate-Based Ceramics for Dielectric Capacitor Applications. J. Mater. Chem. C Mater. 2019, 7, 6222–6230. [Google Scholar] [CrossRef]
  41. Hu, Q.; Tian, Y.; Zhu, Q.; Bian, J.; Jin, L.; Du, H.; Alikin, D.O.; Shur, V.Y.; Feng, Y.; Xu, Z.; et al. Achieve Ultrahigh Energy Storage Performance in BaTiO3-Bi(Mg1/2Ti1/2)O3 Relaxor Ferroelectric Ceramics via Nano-Scale Polarization Mismatch and Reconstruction. Nano Energy 2020, 67, 104264. [Google Scholar] [CrossRef]
  42. Sayyed, S.; Acharya, S.A.; Kautkar, P.; Sathe, V. Structural and Dielectric Anomalies near the MPB Region of Na0.5 Bi0.5 TiO3-SrTiO3 Solid Solution. RSC Adv. 2015, 5, 50644–50654. [Google Scholar] [CrossRef]
  43. Ma, W.; Zhu, Y.; Marwat, M.A.; Fan, P.; Xie, B.; Salamon, D.; Ye, Z.-G.; Zhang, H. Enhanced Energy-Storage Performance with Excellent Stability under Low Electric Fields in BNT-ST Relaxor Ferroelectric Ceramics. J. Mater. Chem. C Mater. 2019, 7, 281–288. [Google Scholar] [CrossRef]
  44. Shi, L.N.; Wang, Y.G.; Ren, Z.H.; Jain, A.; Jiang, S.S.; Chen, F.G. Significant Improvement in Electrical Characteristics and Energy Storage Performance of NBT-Based Ceramics. Ceram. Int. 2022, 48, 26973–26983. [Google Scholar] [CrossRef]
  45. Kong, X.; Yang, L.; Cheng, Z.; Zhang, S. Bi-modified SrTiO3-based Ceramics for High-temperature Energy Storage Applications. J. Am. Ceram. Soc. 2020, 103, 1722–1731. [Google Scholar] [CrossRef]
  46. Wei, T.; Liu, K.; Fan, P.; Lu, D.; Ye, B.; Zhou, C.; Yang, H.; Tan, H.; Salamon, D.; Nan, B.; et al. Novel NaNbO3-Sr0.7Bi0.2TiO3 Lead-Free Dielectric Ceramics with Excellent Energy Storage Properties. Ceram. Int. 2021, 47, 3713–3719. [Google Scholar] [CrossRef]
  47. Pham, K.-N.; Hussain, A.; Ahn, C.W.; Kim, W.; Jeong, S.J.; Lee, J.-S. Giant Strain in Nb-Doped Bi0.5(Na0.82K0.18)0.5TiO3 Lead-Free Electromechanical Ceramics. Mater. Lett. 2010, 64, 2219–2222. [Google Scholar] [CrossRef]
  48. Pattipaka, S.; James, A.R.; Dobbidi, P. Dielectric, Piezoelectric and Variable Range Hopping Conductivity Studies of Bi0.5(Na, K)0.5TiO3 Ceramics. J. Electron. Mater. 2018, 47, 3876–3890. [Google Scholar] [CrossRef]
  49. Ullah, A.; Ahn, C.W.; Hussain, A.; Lee, S.Y.; Kim, J.S.; Kim, I.W. Effect of Potassium Concentration on the Structure and Electrical Properties of Lead-Free Bi0.5 (Na,K)0.5 TiO3-BiAlO3 Piezoelectric Ceramics. J. Alloys Compd. 2011, 509, 3148–3154. [Google Scholar] [CrossRef]
  50. Sumang, R.; Cann, D.P.; Kumar, N.; Bongkarn, T. Large Strain in Lead-Free Piezoelectric (1 − x − y)Bi0.5Na0.5TiO3-XBi0.5K0.5TiO3-YBi0.5Li0.5TiO3 System near MPB Prepared via the Combustion Technique. Ceram. Int. 2015, 41, S127–S135. [Google Scholar] [CrossRef]
  51. Fernandez-Benavides, D.; Gutierrez-Perez, A.; Benitez-Castro, A.; Ayala-Ayala, M.; Moreno-Murguia, B.; Muñoz-Saldaña, J. Comparative Study of Ferroelectric and Piezoelectric Properties of BNT-BKT-BT Ceramics near the Phase Transition Zone. Materials 2018, 11, 361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Gupta, S.K.; McQuade, R.; Gibbons, B.; Mardilovich, P.; Cann, D.P. Electric Field-Induced Strain in Sr(Hf0.5Zr0.5)O3-Modified Bi0.5(Na0.8 K0.2)0.5 TiO3 Piezoelectric Ceramics. J. Appl. Phys. 2020, 127, 074104. [Google Scholar] [CrossRef]
  53. Pattipaka, S.; James, A.R.; Dobbidi, P. Enhanced Dielectric and Piezoelectric Properties of BNT-KNNG Piezoelectric Ceramics. J. Alloys Compd. 2018, 765, 1195–1208. [Google Scholar] [CrossRef]
  54. Tong, X.-Y.; Song, M.-W.; Zhou, J.-J.; Wang, K.; Guan, C.-L.; Liu, H.; Fang, J.-Z. Enhanced Energy Storage Properties in Nb-Modified Bi0.5Na0.5TiO3-SrTiO3 Lead-Free Electroceramics. J. Mater. Sci. Mater. Electron. 2019, 30, 5780–5790. [Google Scholar] [CrossRef]
  55. Jaita, P.; Saenkam, K.; Rujijanagul, G. Improvements in Piezoelectric and Energy Harvesting Properties with a Slight Change in Depolarization Temperature in Modified BNKT Ceramics by a Simple Technique. RSC Adv. 2023, 13, 3743–3758. [Google Scholar] [CrossRef]
  56. Chu, B.; Hao, J.; Li, P.; Li, Y.; Li, W.; Zheng, L.; Zeng, H. 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]
  57. Jaita, P.; Watcharapasorn, A.; Cann, D.P.; Jiansirisomboon, S. Dielectric, Ferroelectric and Electric Field-Induced Strain Behavior of Ba(Ti0.90Sn0.10)O3-Modified Bi0.5(Na0.80K0.20)0.5TiO3 Lead-Free Piezoelectrics. J. Alloys Compd. 2014, 596, 98–106. [Google Scholar] [CrossRef]
  58. Zhang, H.; Zhou, J.; Shen, J.; Yang, X.; Wu, C.; Han, K.; Zhao, Z.; Chen, W. Enhanced Piezoelectric Property and Promoted Depolarization Temperature in Fe Doped Bi1/2(Na0.8K0.2)1/2TiO3 Lead-Free Ceramics. Ceram. Int. 2017, 43, 16395–16402. [Google Scholar] [CrossRef]
  59. Huang, Y.; Zhao, C.; Wu, B.; Wu, J. Multifunctional BaTiO3 -Based Relaxor Ferroelectrics toward Excellent Energy Storage Performance and Electrostrictive Strain Benefiting from Crossover Region. ACS Appl. Mater. Interfaces 2020, 12, 23885–23895. [Google Scholar] [CrossRef]
  60. Sun, Z.; Wang, Z.; Tian, Y.; Wang, G.; Wang, W.; Yang, M.; Wang, X.; Zhang, F.; Pu, Y. Progress, Outlook, and Challenges in Lead-Free Energy-Storage Ferroelectrics. Adv. Electron. Mater 2020, 6, 1900698. [Google Scholar] [CrossRef]
  61. Shvartsman, V.V.; Lupascu, D.C. Lead-Free Relaxor Ferroelectrics. J. Am. Ceram. Soc. 2012, 95, 1–26. [Google Scholar] [CrossRef]
  62. Liu, G.; Wang, Y.; Han, G.; Gao, J.; Yu, L.; Tang, M.; Li, Y.; Hu, J.; Jin, L.; Yan, Y. Enhanced Electrical Properties and Energy Storage Performances of NBT-ST Pb-Free Ceramics through Glass Modification. J. Alloys Compd. 2020, 836, 154961. [Google Scholar] [CrossRef]
  63. Uchino, K.; Nomura, S.; Cross, L.E.; Jang, S.J.; Newnham, R.E. Electrostrictive Effect in Lead Magnesium Niobate Single Crystals. J. Appl. Phys. 1980, 51, 1142–1145. [Google Scholar] [CrossRef]
  64. Li, Q.; Wang, C.; Yadav, A.K.; Fan, H. Large Electrostrictive Effect and Energy Storage Density in MnCO3 Modified Na0.325Bi0.395Sr0.2450.035TiO3 Lead-Free Ceramics. Ceram. Int. 2020, 46, 3374–3381. [Google Scholar] [CrossRef]
  65. Wu, J.; Mahajan, A.; Riekehr, L.; Zhang, H.; Yang, B.; Meng, N.; Zhang, Z.; Yan, H. Perovskite Srx(Bi1-xNa0.97-xLi0.03)0.5TiO3 Ceramics with Polar Nano Regions for High Power Energy Storage. Nano Energy 2018, 50, 723–732. [Google Scholar] [CrossRef]
  66. Yan, Y.; Zeng, X.; Deng, T.; Chen, F.; Yang, L.; He, Z.; Jin, L.; Liu, G. Rheological, Mechanical, and Electrical Properties of Sr0.7 Bi0.2 TiO3 Modified BaTiO3 Ceramic and Its Composites. J. Am. Ceram. Soc. 2023, 106, 3052–3065. [Google Scholar] [CrossRef]
  67. Nayak, S.; Venkateshwarlu, S.; Budisuharto, A.S.; Jørgensen, M.R.V.; Borkiewicz, O.; Beyer, K.A.; Pramanick, A. Effect of A-site Substitutions on Energy Storage Properties of BaTiO3-BiScO3 Weakly Coupled Relaxor Ferroelectrics. J. Am. Ceram. Soc. 2019, 102, 5919–5933. [Google Scholar] [CrossRef]
  68. Cao, W.P.; Li, W.L.; Dai, X.F.; Zhang, T.D.; Sheng, J.; Hou, Y.F.; Fei, W.D. Large Electrocaloric Response and High Energy-Storage Properties over a Broad Temperature Range in Lead-Free NBT-ST Ceramics. J. Eur. Ceram. Soc. 2016, 36, 593–600. [Google Scholar] [CrossRef]
  69. Jabeen, N.; Hussain, A.; Qaiser, M.A.; Ali, J.; Rehman, A.; Sfina, N.; Ali, G.A.; Tirth, V. Enhanced Energy Storage Performance by Relaxor Highly Entropic (Ba0.2Na0.2K0.2La0.2Bi0.2)TiO3 and (Ba0.2Na0.2K0.2Mg0.2Bi0.2)TiO3 Ferroelectric Ceramics. Appl. Sci. 2022, 12, 12933. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of (a) recoverable energy density and hysteresis loss from P-E hysteresis loop of a dielectric material. (b) Domain evolution and formation of FE to RFE transition with the substitution of BST into BNKT, resulting in enhanced Wrec and η.
Figure 1. Schematic diagram of (a) recoverable energy density and hysteresis loss from P-E hysteresis loop of a dielectric material. (b) Domain evolution and formation of FE to RFE transition with the substitution of BST into BNKT, resulting in enhanced Wrec and η.
Materials 16 04912 g001
Figure 2. (a) XRD patterns of the (1 − x)BNKT-xBST ceramics for x = 0.15–0.50, where (b) 2θ = 39–59°. (c) Raman spectra of BNKT-BST ceramics along with spectral deconvolution.
Figure 2. (a) XRD patterns of the (1 − x)BNKT-xBST ceramics for x = 0.15–0.50, where (b) 2θ = 39–59°. (c) Raman spectra of BNKT-BST ceramics along with spectral deconvolution.
Materials 16 04912 g002
Figure 3. FESEM images of (1 − x) BNKT-xBST ceramics for (a) x = 0.15, (b) x = 0.30, (c) x = 0.40, (d) x = 0.45, and (e) x = 0.50. The inset of (ae) shows the average grain size versus counts (grain size distribution histogram). (f) The variation in grain size with composition (x).
Figure 3. FESEM images of (1 − x) BNKT-xBST ceramics for (a) x = 0.15, (b) x = 0.30, (c) x = 0.40, (d) x = 0.45, and (e) x = 0.50. The inset of (ae) shows the average grain size versus counts (grain size distribution histogram). (f) The variation in grain size with composition (x).
Materials 16 04912 g003
Figure 4. (a) Frequency variation of relative dielectric permittivity and loss factor of BNKT-BST ceramics for x = 0.15–0.50. (b) Composition vs. relative dielectric permittivity and loss factor. (c,d) Temperature variation of relative dielectric permittivity and loss factor of BNKT-BST for x = 0.15 and 0.45 (The left and right sides of the arrows with circles enclosed by curves indicate relative dielectric permittivity and loss factor, respectively). The inset of (c,d) shows the l o g T T m versus l o g 1 ε r 1 ε r m of BNKT-BST for x = 0.15 and 0.45, respectively, at 1 MHz.
Figure 4. (a) Frequency variation of relative dielectric permittivity and loss factor of BNKT-BST ceramics for x = 0.15–0.50. (b) Composition vs. relative dielectric permittivity and loss factor. (c,d) Temperature variation of relative dielectric permittivity and loss factor of BNKT-BST for x = 0.15 and 0.45 (The left and right sides of the arrows with circles enclosed by curves indicate relative dielectric permittivity and loss factor, respectively). The inset of (c,d) shows the l o g T T m versus l o g 1 ε r 1 ε r m of BNKT-BST for x = 0.15 and 0.45, respectively, at 1 MHz.
Materials 16 04912 g004
Figure 5. RT P-E and I-E curves of BNKT-BST ceramics for (a) x = 0.15, (b) x = 0.30, (c) x = 0.40, (d) x = 0.45, and (e) x = 0.50. (f) Composition (x) versus polarization and electric field.
Figure 5. RT P-E and I-E curves of BNKT-BST ceramics for (a) x = 0.15, (b) x = 0.30, (c) x = 0.40, (d) x = 0.45, and (e) x = 0.50. (f) Composition (x) versus polarization and electric field.
Materials 16 04912 g005
Figure 6. P-E loops of BNKT-BST ceramics measured at EBD and 10 Hz for (a) x = 0.15, (b) x = 0.30, (c) x = 0.40, (d) x = 0.45, and (e) x = 0.50. (f) Composition (x) versus Wrec, Wloss, and η.
Figure 6. P-E loops of BNKT-BST ceramics measured at EBD and 10 Hz for (a) x = 0.15, (b) x = 0.30, (c) x = 0.40, (d) x = 0.45, and (e) x = 0.50. (f) Composition (x) versus Wrec, Wloss, and η.
Materials 16 04912 g006
Figure 7. Fatigue behavior of BNKT-BST for x = 0.45 composition measured up to 106 electric cycles.
Figure 7. Fatigue behavior of BNKT-BST for x = 0.45 composition measured up to 106 electric cycles.
Materials 16 04912 g007
Table 1. List of all values of ferroelectric properties (Pr, Pmax, Ec, and EBD) and energy storage performance (Wrec and η) of BNKT-BST ceramics for x = 0.15–0.50.
Table 1. List of all values of ferroelectric properties (Pr, Pmax, Ec, and EBD) and energy storage performance (Wrec and η) of BNKT-BST ceramics for x = 0.15–0.50.
CompositionPr (µC/cm2)Pmax (µC/cm2)Ec (kV/cm)EBD (kV/cm)Wrec (J/cm3)η (%)
x = 0.1519.8931.4616.6657.420.2018.67
x = 0.303.9532.185.5968.180.5752.02
x = 0.400.9223.912.9874.570.7385.30
x = 0.450.7822.51.58900.8186.95
x = 0.500.9617.792.9471.540.5685.23
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

Pattipaka, S.; Choi, H.; Lim, Y.; Park, K.-I.; Chung, K.; Hwang, G.-T. Enhanced Energy Storage Performance and Efficiency in Bi0.5(Na0.8K0.2)0.5TiO3-Bi0.2Sr0.7TiO3 Relaxor Ferroelectric Ceramics via Domain Engineering. Materials 2023, 16, 4912. https://doi.org/10.3390/ma16144912

AMA Style

Pattipaka S, Choi H, Lim Y, Park K-I, Chung K, Hwang G-T. Enhanced Energy Storage Performance and Efficiency in Bi0.5(Na0.8K0.2)0.5TiO3-Bi0.2Sr0.7TiO3 Relaxor Ferroelectric Ceramics via Domain Engineering. Materials. 2023; 16(14):4912. https://doi.org/10.3390/ma16144912

Chicago/Turabian Style

Pattipaka, Srinivas, Hyunsu Choi, Yeseul Lim, Kwi-Il Park, Kyeongwoon Chung, and Geon-Tae Hwang. 2023. "Enhanced Energy Storage Performance and Efficiency in Bi0.5(Na0.8K0.2)0.5TiO3-Bi0.2Sr0.7TiO3 Relaxor Ferroelectric Ceramics via Domain Engineering" Materials 16, no. 14: 4912. https://doi.org/10.3390/ma16144912

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