Next Article in Journal
A Study on Cu Thin-Film Electroplated TiO2 Photoanodes for Applications in Natural Dye-Sensitized Solar Cells
Next Article in Special Issue
Aerosol Deposited Polycrystalline PbZr0.53Ti0.47O3 Thick Films with a Large Transverse Piezoelectric Coefficient
Previous Article in Journal
Progress in Boron Nitride-Based Materials as Catalysts for Energy Storage and Electrochemical Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dielectric Temperature Stability and Enhanced Energy-Storage Performance of Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2O6 High-Entropy Ferroelectric Ceramics

1
College of Materials Science and Engineering, Xi’an University of Science and Technology, Xi’an 710054, China
2
Jiangxi Key Laboratory of Extreme Manufacturing Technology for High-End Equipment, School of Materials Science and Engineering, Nanchang Hangkong University, Nanchang 330603, China
3
State Key Laboratory for Mechanical Behavior of Materials, Frontier Institute of Science and Technology, Future Industrial Innovation Institute of Emerging Information Storage and Smart Sensor, Xi’an Jiaotong University, Xi’an 710049, China
*
Authors to whom correspondence should be addressed.
Crystals 2025, 15(1), 26; https://doi.org/10.3390/cryst15010026
Submission received: 27 November 2024 / Revised: 21 December 2024 / Accepted: 27 December 2024 / Published: 29 December 2024

Abstract

:
In this research, we employed a high-entropy approach in tungsten-bronze-structured ferroelectric ceramics, preparing Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2 (denoted as SBN40-H) ceramics through the traditional solid-state reaction technique. By utilizing the high-entropy approach, the resulting SBN40-H ceramics demonstrated extremely fine grains, averaging 0.58 μm in size. Furthermore, these ceramics possessed a high bandgap of 3.35 eV, which, when combined with the small grain size, contributed to a remarkable breakdown strength of 570.01 kV/cm. The dielectric characteristics demonstrated typical relaxation behavior and outstanding temperature stability, with a capacitance temperature coefficient (TCC) of less than 5% within the temperature range of 111–317 °C. Additionally, the SBN40-H ceramics exhibit slim P–E loops with negligible hysteresis, which is considered to be related to the existence of weakly coupled relaxors. This results in exceptional overall energy-storage properties in the SBN40-H ceramics, exhibiting a notable recoverable energy density (Wrec) of 2.68 J/cm3 and an efficiency (η) of 93.7% at 390 kV/cm, and finally achieving a remarkable temperature stability in terms of energy-storage performance with variations in Wrec and η being less than 3.5% and 4.4% from 25 to 150 °C. It is worth noting that the high-entropy approach is highly effective in reducing grain size, increasing the breakdown field strength and enhancing the dielectric temperature stability of tungsten-bronze-structured ferroelectric ceramics.

1. Introduction

As traditional energy sources are consumed, the development of new energy sources and storage technologies has emerged as a crucial challenge [1,2,3,4]. As crucial components in energy-storage devices, ferroelectric ceramics have attracted considerable attention in the field of energy storage owing to their swift charging and discharging rates, extended operational lifespan, high power density, and impressive thermal stability [5,6,7,8,9,10]. However, their practical applications are limited by their relatively low energy-storage density and temperature stability. Consequently, the advancement of lead-free ferroelectric materials with outstanding energy-storage performance is increasingly important, driven by the trend of miniaturizing electronic devices and the growing emphasis on environmental protection [11,12,13]. The performance metrics for energy storage, including the recoverable energy density (Wrec) and energy-storage efficiency (η), can be derived using the formulas outlined in Equations (1)–(3) [14,15]:
W total = 0 P m E d P
W rec = P r P m E d P
η = W rec / W total
where Pm, Pr, and E represent the maximum polarization, remnant polarization, and applied electric field, respectively. Dielectric energy-storage materials are mainly grouped into four distinct categories: linear dielectric, normal ferroelectric, relaxor ferroelectric, and anti-ferroelectric materials [16,17]. Linear dielectrics exhibit a high η because they exhibit a direct, linear correlation between polarization and the applied electric field. However, their low dielectric constant leads to a correspondingly low recoverable energy density Wrec. The anti-ferroelectric materials own double ferroelectric polarization–electric field (P–E) loops, which result in a lower remnant polarization (Pr) and a high recoverable energy density Wrec [18,19]. Despite this advantage, anti-ferroelectric materials suffer from relatively large energy losses that dissipate as heat, posing potential risks for practical applications in energy-storage devices [20,21].
In contrast, relaxor ferroelectric materials exhibit slim P–E loops with a small Pr, contributing significantly to their high η and Wrec values [22,23,24]. These materials have been engineered to exhibit polar nanoregions through doping, which disrupts the long-range ordered structure, increasing compositional disorder and resulting in slimmer P–E loops [25,26]. Recently, high-entropy ceramics have come to prominence as an innovative and promising strategy in the design and development of ferroelectric ceramics [27,28,29]. Generally, high-entropy ceramics are composed of five or more elements, each present in equal molar ratios, disrupting the original ordered state and enhancing atomic, lattice, and compositional disorder to achieve relaxor behavior [30,31]. The high-entropy strategy has been extensively applied in perovskite ferroelectric materials to boost their energy-storage capabilities [32,33,34,35]. Notably, perovskite-structured, BNT-based, high-entropy ceramics have demonstrated an increase in Wrec of nearly 10 times, attributed to the reduced nanodomain size and an enhanced random field resulting from increased entropy [36]. Similarly, in traditional BaTiO3-based ceramics, the introduction of high entropy led to the formation of a local polymorphic distortion structure and further reduction in the size of polar nanoregions, resulting in an ultrahigh η [37]. However, research on high-entropy tungsten-bronze-structured ferroelectric materials remains relatively limited due to the complexity of their structure.
Strontium barium niobate (SBN), which possesses a tetragonal tungsten-bronze structure and exhibits exceptional ferroelectric and optical properties as a lead-free ferroelectric material, has garnered significant attention [38]. The entropy related to its atomic configuration, denoted as Sconfig, can be calculated as follows [39]:
S config =   R ( i = 1 N x i ln x i ) cation - site + ( j = 1 M x j ln x j ) anion - site
where R signifies the ideal gas constant, while N and M represent the number of atomic species occupying the cation and anion sites, respectively. The atomic proportions are indicated by x i and x j . When the Sconfig value exceeds 1.5R, it is categorized as high entropy [40,41]. Therefore, in this work, we propose the incorporation of various elements of Zr4+, Ti4+, Sn2+, and Ta5+ into the B-site of Sr0.4Ba0.6Nb2O6 to elevate the entropy, and the Sconfig value designed in this work is 2.28R. A thorough investigation into the impacts of this high-entropy design on the microstructure, dielectric properties, and energy-storage capabilities were conducted. By adopting the high-entropy approach, the strontium-barium-niobate ferroelectric ceramics achieved a refined grain structure and remarkable dielectric temperature stability. Additionally, they exhibited a wide bandgap and exceptionally high breakdown field strength. Consequently, these ceramics possess superior energy-storage density and efficiency. This research offers valuable insights for the design and development of tungsten-bronze-structured ferroelectric materials with high energy-storage performance.

2. Experimental Procedures

The high-entropy ferroelectric Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2O6 (abbreviated as SBN40-H) ceramics were synthesized using the conventional solid-state reaction method. High-purity starting materials, including BaCO3, SrCO3, Nb2O5, TiO2, ZrO2, SnO2, and Ta2O5, with a purity exceeding 99%, obtained from Alfa Aesar company located at Ward Hill (MA, USA), were weighed in accordance with the stoichiometric ratio and mixed through ball milling with ethanol for a period of 12 h. Afterwards, the calcination of the dried powders was conducted at 1200 °C for 3 h in a muffle furnace. Then, a second ball-milling process was carried out for 12 h. Following that, the powders underwent another drying process and were subsequently compressed into pellets, each with a diameter of 10 mm, using isostatic pressing at a pressure of 200 MPa for a duration of 3 min. The pressed ceramics were sintered at 1350 °C for 3 h.
The sintered ceramics underwent polishing and were then thermally etched at a temperature that was approximately 80 °C lower than the sintering temperature for a period of 10 min to determine the surface microstructure, which was detected using a Field Emission Scanning Electron Microscope (FESEM, ZEISS, Sigma 300, Oberkochen, Germany). The composition uniformity was detected by energy dispersive spectroscopy (EDS) mapping conducted using an Oxford Xplore analyzer (Oxford, UK). The phase structures of the samples at room temperature were determined using X-ray diffraction (XRD) analysis, employing a Rigaku SmartLab SE instrument (Tokyo, Japan) with Cu Kα radiation of λ = 0.15406 nm. The sintered samples were polished to a thickness of 0.7 mm and subsequently coated with a Ag electrode for the purpose of evaluating their dielectric properties. The temperature-dependent dielectric performance was evaluated using an LCR meter (Agilent, 4980 A, Santa Clara, CA, USA) within a temperature range of 0–500 °C. For ferroelectric measurements, the sintered pellets were refined to a thickness of 0.1 mm and sputtered with gold electrodes on both surfaces. The ferroelectric P–E hysteresis loops were characterized using a Precision Premier II system from Radiant Company (Redmond, WA, USA), in conjunction with a high-voltage amplifier. Additionally, the charge–discharge characteristics were tested using a capacitor charge/discharge facility (CFD-003, TG Technology, Shanghai, China) with a load resistance of 240 Ω.

3. Results and Discussion

Figure 1a illustrates the micro-morphology of the component, revealing a uniform grain distribution devoid of notable porosity. It is noteworthy that the grain shape of the high-entropy ceramics, composed of SBN40-H isometric crystals, contrasts with the columnar crystals observed in pure strontium-barium-niobate ceramics, as previously reported in our work [42,43]. Additionally, the average grain size, determined using Nano Measure software 1.2.0 is presented in the inset of Figure 1a. Interestingly, the average grain size is only 0.58 μm, markedly smaller than that of strontium-barium-niobate ceramics produced through the identical process. This outcome was expected, as the distortion in the lattice structure resulting from the multiple cations present in the TTB structure had been predicted to cause such a reduction. The smaller grain size is advantageous for achieving a high breakdown strength (BDS), as grain boundaries effectively hinder carrier transport due to the high potential barrier between them [44,45]. The significantly decreased grain size and the compact microstructure have been proven to be helpful in improving the breakdown strength. Moreover, the EDS analysis reveals elemental mapping for the Sr, Ba, Ti, Zr, Nb, Sn, Ta, and O elements through distribution visualization, which is displayed in Figure 1b. It can be obviously seen that all of the elements are evenly distributed and there is no agglomeration phenomenon. This confirms the successful incorporation and uniform distribution of doping elements within the B-site of the tungsten-bronze structure. The smaller grain size achieved through high-entropy design suggests a possible enhancement in breakdown strength, which exhibits an inverse relationship with grain size. Moreover, the Rietveld refinement of the XRD pattern of as-prepared SBN40-H high-entropy ceramics analyzed by Highscore plus software (ver 4.0) is shown in Figure 1c. The refined results exhibit excellent reliability, as evidenced by the low values of the refinement parameters (Rwp, Rp, and χ2) shown in Figure 1c. The lattice parameters have been determined to be a = b = 1.2552 nm, and c = 0.3977 nm. Additionally, two secondary phases, identified as SnO2 and SrTi0.45Zr0.55O3, are observable. These phases may result from exceeding their respective solid solubility thresholds, with the impurity concentrations estimated to be roughly 16.7% for SnO2 and 19.8% for SrTi0.45Zr0.55O3, respectively.
The temperature dependence of the dielectric constant and loss tangent (tanδ) of SBN40-H high-entropy ceramics measured across various frequencies within a temperature range of 0–500 °C are shown in Figure 2a,b, respectively. In addition, the enlarged view of dielectric loss over a temperature range of 0–300 °C is illustrated in the inset of Figure 2b. It can be seen that the temperature-dependence curves of the dielectric constant exhibit quite a wide peak, and it is observable that the temperature of this peak value (Tm) shifts to higher temperatures as the frequency increases. The temperature-dependent dielectric loss also presents a corresponding frequency dispersion phenomenon with the dielectric anomaly peak shifting towards to a higher temperature with the frequency increasing. It is worth noting that the dielectric loss is very low (<0.08), while it increases sharply at high temperatures, indicating the onset of the conduction processes. Compared to SBN ceramics, the dielectric constant of SBN40-H ceramics decreases significantly, indicating strong relaxation characteristics. The low dielectric constant of these high-entropy ceramics is attributed to their small grain size. This is due to the presence of multiple ion species, which enhances the system’s disorder and causes fluctuations in B-site ion charge, generating random electric fields (RFs) that amplify the ferroelectric relaxation behavior [22]. To assess the diffuseness degree of SBN40-H high-entropy ceramics, the modified Curie–Weiss law is employed, as illustrated below, to quantify their relaxor characteristics [43].
(1/ε − 1/εm) = (TTm)γ/C
Here, εm denotes the maximum dielectric permittivity at Tm, and C is assumed to be constant. The parameter γ, which ranges from 1 to 2, describes the diffuseness degree. Specifically, γ = 1 characterizes the Curie–Weiss behavior typical of normal ferroelectric materials, whereas γ = 2 is indicative of ideal relaxor ferroelectric materials. The fitting result, depicted in Figure S1, yields a γ value of 1.70, indicative of typical relaxor behavior. Furthermore, the relaxation behavior of ferroelectric materials constitutes a thermally activated process, often likened to a spin or dipole glass that only exhibits dynamic activity above a specific freezing temperature. In such scenarios, the Vogel–Fulcher (V–F) law expressed in Equation (6) is usually utilized to delineate this behavior [46,47,48].
f = f0exp[−Ea/kB(TmTf)]
Here, f denotes the probing frequency, f0 is related to the attempt frequency, Ea represents the activation energy of the relaxation process, kB denotes the Boltzmann’s constant, Tm is the temperature at which the dielectric permittivity reaches its maximum, and Tf is the freezing temperature. A nonlinear fitting analysis was conducted using the V–F law, with the results presented in Figure 2c. It is evident that the measured data align closely with the V–F law, and the fitted parameters f0, Ea, Tf are also depicted in Figure 2c. Notably, the activation energy Ea is approximately one order higher than that observed in classical relaxor ferroelectric behavior. This intriguing phenomenon has also been observed in perovskite systems, such as BaTiO3-BiScO3 (Ea 0.24–0.26 eV) [49], BaTiO3-Bi(Mg1/2Ti1/2)O3 (Ea 0.17–0.22 eV) [50], BaTiO3-Bi(Zn1/2Ti1/2)O3-BiScO3 (Ea 0.37–0.50 eV) [51], and BaTiO3-Bi(Zn2/3Nb1/3)O3 (Ea 0.20–0.22 eV) [52]. High activation energies suggest that achieving long-range dipole alignment under field-cooled conditions is challenging, resulting in isolated and frustrated polar clusters with minimal coupling between adjacent clusters. However, macroscopic switching of the polarization in these so-called “weakly coupled relaxors” becomes feasible at low temperatures with the application of high fields [49]. This behavior is particularly advantageous for energy-storage applications, as it is typically associated with low dielectric nonlinearity (or a high polarization saturation field) and minimal hysteresis (or low energy loss).
Moreover, to assess the stability of dielectric performance across various temperatures, the temperature coefficient of capacitance (TCC) is defined as follows:
TCC = C T   C base C base
where CT denotes the capacitance over the measured temperature range, and the base capacitance is referred to as Cbase. In the case of high-temperature capacitors, the chosen base temperature is typically 150 °C. As is illustrated in Figure 2d, the TCC value of SBN40-H high-entropy ceramics initially decreases, then increases, and finally decreases again. Notably, these ceramics exhibit excellent TCC stability, with values below 5% over a wide temperature range of 111–317 °C. Given the temperature range where tanδ is less than 0.08 (0–286 °C), the combined range of both low loss and high capacitance stability extends up to 175 °C. Overall, the SBN40-H high-entropy ceramics demonstrate impressive dielectric properties, characterized by a broad temperature range where both the temperature coefficient of capacitance (TCC) is less than 5% and tanδ is less than 0.08.
As far as is known, the breakdown strength (BDS), which exhibits a close correlation with the bandgap, is crucial for the energy-storage capabilities of ferroelectric materials. Generally, a larger bandgap (Eg) indicates that promoting electrons from the valence band to the conduction band is more demanding, leading to reduced conductivity and a higher BDS value. Consequently, Figure 3a displays the UV–vis absorption spectra of the SBN40-H high-entropy ceramics, while Figure 3b illustrates the deduced bandgap (Eg) calculated using the equation listed as follows:
( α h υ ) 2 = A ( h υ E g )
where A refers to a proportionality constant, α denotes the absorption coefficient, and h υ represents the photon energy. From Figure 3b, it is evident that the SBN40-H high-entropy ceramics possess a notably high Eg of 3.35 eV, contributing to their enhanced breakdown strength. Dielectric ceramics’ breakdown strength (BDS) is typically assessed using the two-parameter Weibull distribution, owing to the statistical characteristics associated with failure phenomena. Generally speaking, the formulation of the Weibull distribution for evaluating BDS is articulated as follows [36,53,54,55,56]:
P ( E i ) = 1 exp E i E b β
P i =   i n + 1
Here, P(Ei) denotes the probability of dielectric breakdown for each sample, with Ei representing the breakdown field strength (Eb) of each sample. The variable i serves as the serial number of the sample, and n indicates the total count of samples. The term Eb corresponds to the field strength at which the cumulative failure probability reaches 63.2%. The parameter β is associated with the distribution of breakdown strength (BDS), where a higher β value suggests a more tightly concentrated BDS. Equations (9) and (10) can be converted to the following:
ln ( ln ( 1 i / ( n + 1 ) ) ) = β ln E i β ln E b
Hence, Equations (12) and (13), derived from Equation (11), incorporate two parameters, Xi and Yi, which exhibit a linear relationship as referenced in [57,58,59].
X i = ln ( E i )
Y i = ln ( ln ( 1 i / ( n + 1 ) ) )
The graph in Figure 3c illustrates Xi and Yi values based on these equations. As is depicted in Figure 3c, when Xi is plotted as a function of Yi, the resulting linear fit yields a slope (β) of 6.22 (significantly greater than 1). This indicates a high degree of reliability for the Weibull distribution and suggests minimal variation in the quality of the samples studied. The SBN40-H high-entropy ceramics exhibit a remarkable BDS value of 570.01 kV/cm, which is advantageous for enhancing energy-storage density. This high BDS can be ascribed to the small grain size and the high bandgap.
To assess the energy-storage characteristics, Figure 4a displays the unipolar P–E hysteresis loops of the SBN40-H high-entropy ceramics measured at a frequency of 100 Hz and across a range of electric field intensities. And, the unipolar P–E hysteresis loops plotted separately and the bipolar P–E loops measured at 100 Hz under different electric fields are also shown in Figures S2 and S3 in the Supporting Information. It can be observed that all of the unipolar and bipolar P–E loops at various electric fields exhibit distinctive slim P–E loops with neglectable hysteresis, which can be attributed to the increased B-site disorder and strong relaxation induced by the high-entropy strategy. Additionally, Figure 4b illustrates the variations in Pm, Pr, and PmPr with respect to the electric field. The PmPr values are almost identical to Pm, which is technologically meaningful because large PmPr values are often favorable for high discharged energy densities. The nearly linear P–E loops with almost the same value between with PmPr and Pm are similar to those of the relaxor ceramics (1 − x)BaTiO3−xBi(Zn2/3Nb1/3)O3 (x > 0.05) reported in ref. [52]. This is considered to be related to the weakly coupled behavior of these ceramics. Additionally, using Equations (1)–(3), the calculated recoverable energy density (Wrec) and efficiency (η) were then plotted in Figure 4c. Due to the small Pr value, a polarization difference of 14.14 μC/cm2 is achieved at 390 kV/cm, resulting in exceptional energy-storage performance with a Wrec of 2.68 J/cm3 and η of 93.7%. Additionally, the stability of the energy-storage performance across varying temperatures is crucial for the practical utilization of ferroelectric ceramics. Therefore, Figure 4d–f displays the unipolar hysteresis loop, Wrec, and η of the sample as they vary across a temperature range of 25–150 °C, when measured at an electric field strength of 200 kV/cm and a frequency of 100 Hz. Over this temperature range, the hysteresis loops remain slim and Pm, Pr, and PmPr exhibit minimal temperature-dependent changes. As is displayed in Figure 4f, although Wrec and η decrease slightly within the temperature range of 25–150 °C, the energy-storage performance of SBN40-H high-entropy ceramics exhibits a good temperature stability and the variation range of Wrec and η is less than 3.5% and 4.4%, respectively.
To further evaluate the practical applicability of SBN40-H high-entropy ceramics, we examined underdamped and overdamped charge–discharge curves measured at different electric fields, which are presented in Figure 5a,c, respectively. The energy-storage property, denoted as Wd, can be calculated using Equation (14) derived from the overdamped discharge curves.
W d = R i 2 ( t ) dt V
where R represents the load resistor (240 Ω), i stands for the discharge current, and V signifies the volume of the sample. As observed, W d increases monotonically as the electric field increases, and the discharge time τ0.9 is extremely short at 13.3 ns. The W d value obtained from this method is lower compared to that derived from the quasi-static P–E loop, primarily owing to the fundamental differences between the measurement principles of the two techniques [12,60]. Distinct and well-defined underdamped discharge curves are observable. The current density (CD) and power density (PD) shown in Figure 5b can be calculated using the following specified formulas [61]:
C D = I max S
P D = E I max 2 S
where S is the electrode area. With an increase in the electric field, there is a notable rise in Imax, CD, and PD. The charge–discharge test results underscore the considerable potential of SBN40-H high-entropy ceramics in energy-storage applications.
Additionally, in order to confirm the advancement of this work, the performance comparison between the current system and some other reports is shown in Table 1. It can be seen that Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2 high-entropy ceramics developed in this study exhibit an exceptionally high η value coupled with a moderate Wrec, when compared to other lead-free systems that are especially attractive in TTB systems. Furthermore, our work has achieved a remarkably high breakdown strength along with superb dielectric stability, suggesting promising prospects for their application in energy-storage devices.

4. Conclusions

In conclusion, we introduce elements with different valences and radii into the B-site of strontium-barium-niobate ceramics to increase entropy. The Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2 high-entropy ceramics were synthesized successfully using the conventional solid-state reaction method. The results demonstrate that the high-entropy strategy leads to a significantly reduced grain size and increasing relaxation. The dielectric curves show an obvious relaxor characteristic and excellent temperature stability with the temperature coefficient of capacitance (TCC) <5% over a wide temperature range of 111–317 °C. The SBN40-H high-entropy ceramics exhibit a high breakdown strength of 570.01 kV/cm, which is ascribed to the fine grain achieved through the high-entropy design and the high bandgap of 3.35 eV. Finally, a superior overall energy-storage performance was achieved in SBN40-H high-entropy ceramics with the Wrec of 2.68 J/cm3 and η of 93.7% at 390 kV/cm, and an extraordinary temperature stability with Wrec and η varying less than 3.5% and 4.4% from 25 to 150 °C. This study once again demonstrates the important potential of a high-entropy strategy on enhancing the energy-storage performance of tungsten-bronze-structured ferroelectric ceramics.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst15010026/s1, Figure S1: Plots of ln(1/ε − 1/εm) versus ln(TTm) at 1 kHz; Figure S2: The unipolar P-E loops measured at 100 Hz under various electric fields; Figure S3: (a) The bipolar P-E loops measured at 100 Hz under various electric fields (b) the corresponding variations of Pm, Pr, and PmPr with respect to the electric field.

Author Contributions

Conceptualization, Y.Z. and R.K.; methodology, Z.L. and S.Y.; investigation, Z.L. and S.Y.; writing—original draft preparation, Y.Z.; writing—review and editing, P.M. and R.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (52302155) Postdoctoral Fellowship Program of CPSF under Grant Number GZC20241336.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Zhou, S.; Pu, Y.; Zhang, X.; Shi, Y.; Gao, Z.; Feng, Y.; Shen, G.; Wang, X.; Wang, D. High energy density, temperature stable lead-free ceramics by introducing high entropy perovskite oxide. Chem. Eng. J. 2022, 427, 131684. [Google Scholar] [CrossRef]
  2. 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]
  3. Zhang, X.; Zhang, F.; Niu, Y.; Zhang, Z.; Lei, X.; 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]
  4. Chen, X.; Zhang, H.; Cao, F.; Wang, G.; Dong, X.; Gu, Y.; He, H.; Liu, Y. Charge-discharge properties of lead zirconate stannate titanate ceramics. J. Appl. Phys. 2009, 106, 034105. [Google Scholar] [CrossRef]
  5. Wang, C.; Lou, X. High energy storage properties of 0.94Bi0.5Na0.5TiO3-0.06BaTiO3 ceramics by incorporating Sr0.8Bi0.1γ0.1Ti0.8Zr0.2O2.95. Microstructures 2023, 3, 2023023. [Google Scholar]
  6. Kang, R.; Wang, Z.; Yang, W.; Zhao, Y.; Zhang, L.; Lou, X. Enhanced energy storage performance of Bi0.5Na0.5TiO3-based ceramics via grain tuning and relaxor construction. Chem. Eng. J. 2023, 455, 140924. [Google Scholar] [CrossRef]
  7. Xu, S.; Du, H.; Zhang, Y.; Hu, Y.; Tang, Z.; Liu, Y.; Du, X. Dielectric tunability of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 porous ceramics with oriented pore structure. J. Mater. Sci. Mater. Electron. 2024, 35, 284. [Google Scholar] [CrossRef]
  8. Zhou, S.; Pu, Y.; Zhao, X.; Ouyang, T.; Ji, J.; Zhang, Q.; Zhang, C.; Sun, S.; Sun, R.; Li, J.; et al. Dielectric temperature stability and energy storage performance of NBT-based ceramics by introducing high-entropy oxide. J. Am. Ceram. Soc. 2022, 105, 4796–4804. [Google Scholar] [CrossRef]
  9. Gordon, R. Energy Storage in Ceramic Dielectrics. J. Am. Ceram. Soc. 1990, 73, 323–328. [Google Scholar]
  10. Viola, G.; Ning, H.; Reece, M.J.; Wilson, R.; Correia, T.M.; Weaver, P.; Cain, M.G.; Yan, H. Reversibility in electric field-induced transitions and energy storage properties of bismuth-based perovskite ceramics. J. Phys. D Appl. Phys. 2012, 45, 355302. [Google Scholar] [CrossRef]
  11. Huang, Y.; Shang, K.; Yang, Y.; Shi, W.; Zhang, L.; Laletin, V.; Shur, V.; Jing, R.; Jin, L. Ultrahigh energy storage capacities in high-entropy relaxor ferroelectrics. J. Mater. Chem. A 2024, 12, 18224–18233. [Google Scholar] [CrossRef]
  12. Jiang, J.; Meng, X.; Li, L.; Guo, S.; Huang, M.; Zhang, J.; Wang, J.; Hao, X.; Zhu, H.; Zhang, S.-T. Ultrahigh energy storage density in lead-free relaxor antiferroelectric ceramics via domain engineering. Energy Storage Mater. 2021, 43, 383–390. [Google Scholar] [CrossRef]
  13. Hao, X. A Review on the dielectric materials for high energy-storage application. J. Adv. Dielectr. 2013, 3, 14. [Google Scholar] [CrossRef]
  14. Ning, Y.; Pu, Y.; Zhang, X.; Chen, Z.; Wu, C.; Zhang, L.; Wang, B.; Li, X. Remarkable energy-storage density together with efficiency of above 92% in high-entropy ferroelectric ceramics. Mater. Today Phys. 2024, 43, 101418. [Google Scholar] [CrossRef]
  15. Zhou, J.; Zheng, P.; Bai, W.; Fan, Q.; Zheng, L.; Zhang, Y. Breaking the mutual constraint between polarization and voltage resistance with nanograined high-entropy ceramic. ACS Appl. Mater. Interfaces 2024, 16, 2530–2538. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, M.-H.; Qi, J.-L.; Liu, Y.-Q.; Lan, S.; Luo, Z.-X.; Pan, H.; Lin, Y.-H. High energy storage capability of perovskite relaxor ferroelectrics via hierarchical optimization. Rare Met. 2021, 41, 730–744. [Google Scholar] [CrossRef]
  17. 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] [PubMed]
  18. Chauhan, A.; Patel, S.; Vaish, R.; Bowen, C. Anti-ferroelectric ceramics for high energy density capacitors. Materials 2015, 8, 8009–8031. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, H.; Chen, X.; Cao, F.; Wang, G.; Dong, X. Charge-discharge properties of an antiferroelectric ceramics capacitor under different electric fields. J. Am. Ceram. Soc. 2010, 93, 4015–4017. [Google Scholar] [CrossRef]
  20. Li, D.; Zeng, X.; Li, Z.; Shen, Z.-Y.; Hao, H.; Luo, W.; Wang, X.; Song, F.; Wang, Z.; Li, Y. Progress and perspectives in dielectric energy storage ceramics. J. Adv. Ceram. 2021, 10, 675–703. [Google Scholar] [CrossRef]
  21. Chen, D.; Zhu, X.; Yang, X.; Yan, N.; Cui, Y.; Lei, X.; Liu, L.; Khaliq, J.; Li, C. A review on structure–property relationships in dielectric ceramics using high-entropy compositional strategies. J. Am. Ceram. Soc. 2023, 106, 6602–6616. [Google Scholar] [CrossRef]
  22. Xie, A.; Fu, J.; Zuo, R.; Jiang, X.; Li, T.; Fu, Z.; Yin, Y.; Li, X.; Zhang, S. Supercritical relaxor nanograined ferroelectrics for ultrahigh-energy-storage capacitors. Adv. Mater. 2022, 34, 2204356. [Google Scholar] [CrossRef]
  23. Yang, H.; Yan, F.; Lin, Y.; Wang, T.; Wang, F.; Wang, Y.; Guo, L.; Tai, W.; Wei, H. Lead-free BaTiO3-Bi0.5Na0.5TiO3-Na0.73Bi0.09NbO3 relaxor ferroelectric ceramics for high energy storage. J. Eur. Ceram. Soc. 2017, 37, 3303–3311. [Google Scholar] [CrossRef]
  24. Abdelmoula, N.; Chaabane, H.; Khemakhem, H.; Von der Mühll, R.; Simon, A. Relaxor or classical ferroelectric behavior in A site substituted perovskite type Ba1−x(Sm0.5Na0.5)xTiO3. Phys. Status Solidi 2006, 203, 987–996. [Google Scholar] [CrossRef]
  25. Yang, H.; Lu, Z.; Li, L.; Bao, W.; Ji, H.; Li, J.; Feteira, A.; Xu, F.; Zhang, Y.; Sun, H.; et al. Novel BaTiO3-Based, Ag/Pd-compatible lead-free relaxors with superior energy storage performance. ACS Appl. Mater. Interfaces 2020, 12, 43942–43949. [Google Scholar] [CrossRef] [PubMed]
  26. Guo, J.; Yu, H.; Ren, Y.; Qi, H.; Yang, X.; Deng, Y.; Zhang, S.-T.; Chen, J. Multi-symmetry high-entropy relaxor ferroelectric with giant capacitive energy storage. Nano Energy 2023, 112, 108458. [Google Scholar] [CrossRef]
  27. Xiong, W.; Zhang, H.; Cao, S.; Gao, F.; Svec, P.; Dusza, J.; Reece, M.J.; Yan, H. Low-loss high entropy relaxor-like ferroelectrics with A-site disorder. J. Eur. Ceram. Soc. 2021, 41, 2979–2985. [Google Scholar] [CrossRef]
  28. Qiao, W.; Xu, Z.; Yuan, W.; Xu, J.; Gao, Y.; Bai, M.; Zhu, X.; Hu, Y.; Lou, X. High-performance energy storage in BNST-based lead-free ferroelectric ceramics achieved through high-entropy engineering. Chem. Eng. J. 2023, 477, 147167. [Google Scholar] [CrossRef]
  29. Tian, Z.; Zhang, Y.; Zhang, J.; Shi, P. Facile Preparation, Microstructure and dielectric properties of La(Cr0.2Mn0.2Fe0.2Co0.2Ni0.2)O3 perovskite high-entropy ceramics. Crystals 2022, 12, 1756. [Google Scholar] [CrossRef]
  30. Ning, Y.; Pu, Y.; Zhang, Q.; Zhou, S.; Wu, C.; Zhang, L.; Shi, Y.; Sun, Z. Achieving high energy storage properties in perovskite oxide via high-entropy design. Ceram. Int. 2023, 49, 12214–12223. [Google Scholar] [CrossRef]
  31. Zhang, R.-Z.; Reece, M.J. Review of high entropy ceramics: Design, synthesis, structure and properties. J. Mater. Chem. A 2019, 7, 22148–22162. [Google Scholar] [CrossRef]
  32. Liu, Z.; Xu, S.; Li, T.; Xie, B.; Guo, K.; Lu, J. Microstructure and ferroelectric properties of high-entropy perovskite oxides with A-site disorder. Ceram. Int. 2021, 47, 33039–33046. [Google Scholar] [CrossRef]
  33. Fang, J.; Wang, T.; Li, K.; Li, Y.; Gong, W. Energy storage properties of Mn-modified (Na0.2Bi0.2Ca0.2Sr0.2Ba0.2)TiO3 high-entropy relaxor-ferroelectric ceramics. Results Phys. 2022, 38, 105617. [Google Scholar] [CrossRef]
  34. Ning, Y.; Pu, Y.; Wu, C.; Chen, Z.; Zhang, X.; Zhang, L.; Wang, B. Design strategy of high-entropy perovskite energy-storage ceramics: A review. J. Eur. Ceram. Soc. 2024, 44, 4831–4843. [Google Scholar] [CrossRef]
  35. Ding, Y.; Ren, K.; Chen, C.; Huan, L.; Gao, R.; Deng, X.; Chen, G.; Cai, W.; Fu, C.; Wang, Z.; et al. High-entropy perovskite ceramics: Advances in structure and properties. Process. Appl. Ceram. 2024, 18, 1–11. [Google Scholar] [CrossRef]
  36. Chen, L.; Yu, H.; Wu, J.; Deng, S.; Liu, H.; Zhu, L.; Qi, H.; Chen, J. Large energy capacitive high-entropy lead-free ferroelectrics. Nano-Micro Lett. 2023, 15, 65. [Google Scholar] [CrossRef]
  37. Ning, Y.; Pu, Y.; Chen, Z.; Zhang, L.; Wu, C.; Zhang, X.; Wang, B.; Zhang, J. Novel high-entropy relaxors with ultrahigh energy-storage efficiency and density. Chem. Eng. J. 2023, 476, 146673. [Google Scholar] [CrossRef]
  38. Duan, J.; Wei, K.; Du, Q.; Ma, L.; Qi, H.; Li, H. High-entropy tungsten bronze ceramics for large capacitive energy storage with near-zero Losses. Adv. Funct. Mater. 2024, 34, 2409446. [Google Scholar] [CrossRef]
  39. Sarkar, A.; Wang, Q.; Schiele, A.; Chellali, M.R.; Bhattacharya, S.S.; Wang, D.; Brezesinski, T.; Hahn, H.; Velasco, L.; Breitung, B. High-entropy oxides: Fundamental aspects and electrochemical properties. Adv. Mater. 2019, 31, 1806236. [Google Scholar] [CrossRef] [PubMed]
  40. Wright, A.J.; Luo, J. A step forward from high-entropy ceramics to compositionally complex ceramics: A new perspective. J. Mater. Sci. 2020, 55, 9812–9827. [Google Scholar] [CrossRef]
  41. Wright, A.J.; Wang, Q.; Huang, C.; Nieto, A.; Chen, R.; Luo, J. From high-entropy ceramics to compositionally-complex ceramics: A case study of fluorite oxides. J. Eur. Ceram. Soc. 2020, 40, 2120–2129. [Google Scholar] [CrossRef]
  42. Zhao, Y.; Wang, J.; Zhang, L.; Liu, S.; Zhang, D.; Wang, X. Large activation energy in aged Mn-doped Sr0.4Ba0.6Nb2O6 ferroelectric ceramics. RSC Adv. 2017, 7, 26894–26902. [Google Scholar] [CrossRef]
  43. Zhao, Y.; Li, Z.; Du, H.; Liu, X.; Liu, J.; Kang, R.; Mao, P.; Zhang, L. Enhanced energy storage properties promoted by the synergistic effects of aging effects and relaxor behavior in Ce-Mn co-doped Sr0.4Ba0.6Nb2O6 ferroelectric ceramics. Ceram. Int. 2024, 50, 38462–38470. [Google Scholar] [CrossRef]
  44. Yan, F.; Bai, H.; Ge, G.; Lin, J.; Shi, C.; Zhu, K.; Shen, B.; Zhai, J.; Zhang, S. Composition and Structure Optimized BiFeO3-SrTiO3 Lead-Free Ceramics with Ultrahigh Energy Storage Performance. Small 2022, 18, 2106515. [Google Scholar] [CrossRef]
  45. Lin, J.; Ge, G.; Zhu, K.; Bai, H.; Sa, B.; Yan, F.; Li, G.; Shi, C.; Zhai, J.; Wu, X.; et al. Simultaneously achieving high performance of energy storage and transparency via A-site non-stoichiometric defect engineering in KNN-based ceramics. Chem. Eng. J. 2022, 444, 136538. [Google Scholar] [CrossRef]
  46. Venet, M.; Guerra, J.d.L.S.; Santos, I.A.; Eiras, J.A.; Garcia, D. Diffuse phase transition and relaxor behaviour of textured Sr0.63Ba0.37Nb2O6 ceramics. J. Phys. Condens. Matter 2007, 19, 026207. [Google Scholar] [CrossRef]
  47. Sun, E.; Qi, X.; Yuan, Z.; Sang, S.; Zhang, R.; Yang, B.; Cao, W.; Zhao, L. Relaxation behavior in 0.24Pb(In1/2Nb1/2)O3-0.49Pb(Mg1/3Nb2/3)O3-0.27PbTiO3 ferroelectric single crystal. Ceram. Int. 2016, 42, 4893–4898. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Wang, J.; Zhang, L.; Shi, X.; Liu, S.; Zhang, D. Relaxor transition and properties of Mn-doped SrxBa1−xNb2O6 ferroelectric ceramics. Ceram. Int. 2016, 42, 16697–16702. [Google Scholar] [CrossRef]
  49. Ogihara, H.; Randall, C.A.; Trolier-McKinstry, S. Weakly coupled relaxor behavior of BaTiO3-BiScO3 ceramics. J. Am. Ceram. Soc. 2009, 92, 110–118. [Google Scholar] [CrossRef]
  50. Choi, D.H.; Baker, A.; Lanagan, M.; Trolier-McKinstry, S.; Randall, C.; Johnson, D. Structural and dielectric properties in (1 − x)BaTiO3-xBi(Mg1/2Ti1/2)O3 ceramics (0.1 ≤ x ≤ 0.5) and potential for high-voltage multilayer capacitors. J. Am. Ceram. Soc. 2013, 96, 2197–2202. [Google Scholar] [CrossRef]
  51. Raengthon, N.; Sebastian, T.; Cumming, D.; Reaney, I.M.; Cann, D.P.; Roedel, J. BaTiO3-Bi(Zn1/2Ti1/2)O3-BiScO3 ceramics for high-temperature capacitor applications. J. Am. Ceram. Soc. 2012, 95, 3554–3561. [Google Scholar] [CrossRef]
  52. Wu, L.; Wang, X.; Li, L. Lead-free BaTiO3-Bi(Zn2/3Nb1/3)O3 weakly coupled relaxor ferroelectric materials for energy storage. RSC Adv. 2016, 6, 14273–14282. [Google Scholar] [CrossRef]
  53. Luo, J.; Du, J.; Tang, Q.; Mao, C. Lead sodium niobate glass-ceramic dielectrics and internal electrode structure for high energy storage density capacitors. IEEE Trans. Electron Devices 2008, 12, 3549. [Google Scholar] [CrossRef]
  54. Li, M.; Fan, P.; Ma, W.; Liu, K.; Zang, J.; Samart, C.; Zhang, T.; Tan, H.; Salamon, D.; Zhang, H. Constructing layered structures to enhance the breakdown strength and energy density of Na0.5Bi0.5TiO3-based lead-free dielectric ceramics. J. Mater. Chem. C 2019, 7, 15292–15300. [Google Scholar] [CrossRef]
  55. Jin, L.; Pang, J.; Luo, W.; Huang, Y.; Jing, R.; Xu, J.; Guo, D.; Hu, Q.; Tian, Y.; Lu, X.; et al. Enhanced breakdown strength and improved ferroelectric properties in lead-containing relaxor ferroelectric ceramics with addition of glass. Mater. Res. Express 2019, 6, 116310. [Google Scholar] [CrossRef]
  56. Abdullahi Hassan, Y.; Hu, H. Current status of polymer nanocomposite dielectrics for high-temperature applications. Compos. Part A Appl. Sci. Manuf. 2020, 138, 106064. [Google Scholar] [CrossRef]
  57. Liu, G.; Li, Y.; Guo, B.; Tang, M.; Li, Q.; Dong, J.; Yu, L.; Yu, K.; Yan, Y.; Wang, D.; et al. Ultrahigh dielectric breakdown strength and excellent energy storage performance in lead-free barium titanate-based relaxor ferroelectric ceramics via a combined strategy of composition modification, viscous polymer processing, and liquid-phase sintering. Chem. Eng. J. 2020, 398, 125625. [Google Scholar] [CrossRef]
  58. Zhang, C.; Chen, Y.; Zhou, M.; Li, X.; Wang, L.; Xia, L.; Shen, Y.; Dong, X. Achieving ultrahigh dielectric breakdown strength in MgO-based ceramics by composite structure design. J. Mater. Chem. C 2019, 7, 8120–8130. [Google Scholar] [CrossRef]
  59. Yang, H.; Yan, F.; Lin, Y.; Wang, T. Novel strontium titanate-based lead-free ceramics for high-energy storage applications. ACS Sustain. Chem. Eng. 2017, 5, 10215–10222. [Google Scholar] [CrossRef]
  60. Li, C.; Liu, J.; Bai, W.; Wu, S.; Zheng, P.; Zhang, J.; Pan, Z.; Zhai, J. Superior energy storage performance in (Bi0.5Na0.5)TiO3-based lead-free relaxor ferroelectrics for dielectric capacitor application via multiscale optimization design. J. Mater. Chem. A 2022, 10, 9535–9546. [Google Scholar] [CrossRef]
  61. Zuo, C.; Yang, S.; Cao, Z.; Jie, W.; Wei, X. Combustion synthesis of high-performance high-entropy dielectric ceramics for energy storage applications. Ceram. Int. 2023, 49, 25486–25494. [Google Scholar] [CrossRef]
  62. Xu, S.; Hao, R.; Yan, Z.; Hou, S.; Peng, Z.; Wu, D.; Liang, P.; Chao, X.; Wei, L.; Yang, Z. Enhanced energy storage properties and superior thermal stability in SNN-based tungsten bronze ceramics through substitution strategy. J. Eur. Ceram. Soc. 2022, 42, 2781–2788. [Google Scholar] [CrossRef]
  63. Xu, S.; Shen, S.; Huang, C.; He, Y.; Chao, X.; Wu, D.; Liang, P.; Yang, Z.; Lu, J.; Wei, L. Enhancing energy storage performance in lead-free bismuth sodium niobate-based tungsten bronze ceramics through relaxor tuning. ACS Appl. Mater. Interfaces 2023, 15, 11642–11651. [Google Scholar] [CrossRef]
  64. Lv, J.; Song, S.; Jiao, Y.; Yang, Z.; Zhang, Y.; Fan, Y.; Song, C.; Liu, G. Enhancing the dielectric and energy storage properties of lead-free Na0.5Bi0.5TiO3–BaTiO3 ceramics by adding K0.5Na0.5NbO3 ferroelectric. Ceram. Int. 2022, 48, 22–31. [Google Scholar] [CrossRef]
  65. Pu, Y.; Yao, M.; Liu, H.; Frömling, T. Phase transition behavior, dielectric and ferroelectric properties of (1 − x)(Bi0.5Na0.5)TiO3-xBa0.85Ca0.15Ti0.9Zr0.1O3 ceramics. J. Eur. Ceram. Soc. 2016, 36, 2461–2468. [Google Scholar] [CrossRef]
  66. Ren, P.; He, J.; Yan, F.; Wang, X. Temperature stable dielectric and energy storage properties of (1 − x)(0.94Bi0.5Na0.5TiO3-0.09BiAlO3)-xSrTiO3 ceramics. J. Alloys Compd. 2019, 807, 151676. [Google Scholar] [CrossRef]
  67. Zhang, L.; Pu, Y.; Chen, M.; Wei, T.; Peng, X. Novel Na0.5Bi0.5TiO3 based, lead-free energy storage ceramics with high power and energy density and excellent high-temperature stability. Chem. Eng. J. 2020, 383, 123154. [Google Scholar] [CrossRef]
  68. Li, Q.; Li, M.; Wang, C.; Zhang, M.; Fan, H. Enhanced temperature stable dielectric properties and energy-storage density of BaSnO3-modified (Bi0.5Na0.5)0.94Ba0.06TiO3 lead-free ceramics. Ceram. Int. 2019, 45, 19822–19828. [Google Scholar] [CrossRef]
  69. Li, D.; Lin, Y.; Zhang, M.; Yang, H. Achieved ultrahigh energy storage properties and outstanding charge–discharge performances in (Na0.5Bi0.5)0.7Sr0.3TiO3-based ceramics by introducing a linear additive. Chem. Eng. J. 2020, 392, 123729. [Google Scholar] [CrossRef]
  70. Cao, L.; Yuan, Y.; Li, E.; Zhang, S. Relaxor regulation and improvement of energy storage properties of Sr2NaNb5O15-based tungsten bronze ceramics through B-site substitution. Chem. Eng. J. 2021, 421, 127846. [Google Scholar] [CrossRef]
Figure 1. (a) SEM micrographs of SBN40-H ceramics after thermal etching treatment (the inset shows the grain size distribution); (b) EDS mapping images of SBN40-H ceramics; (c) the Rietveld fitting results of experimental and calculated values of XRD patterns of SBN40-H high-entropy ceramics.
Figure 1. (a) SEM micrographs of SBN40-H ceramics after thermal etching treatment (the inset shows the grain size distribution); (b) EDS mapping images of SBN40-H ceramics; (c) the Rietveld fitting results of experimental and calculated values of XRD patterns of SBN40-H high-entropy ceramics.
Crystals 15 00026 g001
Figure 2. (a) The variation in the dielectric constant with temperature and (b) the temperature dependence of dielectric loss for SBN40-H ceramics; (c) the Vogel–Fulcher fitting for SBN40-H ceramics; (d) capacitance variation as a function of temperature at 10 kHz for SBN40-H ceramics.
Figure 2. (a) The variation in the dielectric constant with temperature and (b) the temperature dependence of dielectric loss for SBN40-H ceramics; (c) the Vogel–Fulcher fitting for SBN40-H ceramics; (d) capacitance variation as a function of temperature at 10 kHz for SBN40-H ceramics.
Crystals 15 00026 g002
Figure 3. (a) UV–vis absorbance spectra and (b) (αhυ)2 vs. plots of SBN40-H high-entropy ceramics; (c) Weibull distribution of the SBN40-H high-entropy ceramics.
Figure 3. (a) UV–vis absorbance spectra and (b) (αhυ)2 vs. plots of SBN40-H high-entropy ceramics; (c) Weibull distribution of the SBN40-H high-entropy ceramics.
Crystals 15 00026 g003
Figure 4. (a) Unipolar P–E loops of the SBN40-H high-entropy ceramics at different electric fields; (b,c) corresponding Pm, Pr, PmPr, Wrec, and η variation with electric field; (d) the temperature stability of P–E loops under 200 kV/cm for SBN40-H high-entropy ceramics; (e,f) corresponding Pm, Pr, PmPr, Wrec, and η variation with temperature.
Figure 4. (a) Unipolar P–E loops of the SBN40-H high-entropy ceramics at different electric fields; (b,c) corresponding Pm, Pr, PmPr, Wrec, and η variation with electric field; (d) the temperature stability of P–E loops under 200 kV/cm for SBN40-H high-entropy ceramics; (e,f) corresponding Pm, Pr, PmPr, Wrec, and η variation with temperature.
Crystals 15 00026 g004
Figure 5. Pulse charge–discharge curves for SBN40-H high-entropy ceramics: (a) electric-field-dependent overdamped discharge curves and (b) the corresponding Wd; (c) electric-field-dependent underdamped discharge curves and (d) the corresponding Imax, CD, and PD.
Figure 5. Pulse charge–discharge curves for SBN40-H high-entropy ceramics: (a) electric-field-dependent overdamped discharge curves and (b) the corresponding Wd; (c) electric-field-dependent underdamped discharge curves and (d) the corresponding Imax, CD, and PD.
Crystals 15 00026 g005
Table 1. Comparison of the energy-storage performance and dielectric stability of this work and others.
Table 1. Comparison of the energy-storage performance and dielectric stability of this work and others.
CompositionsWrec
(J/cm3)
H (%)BDS
(kV/cm)
Temperature Range (°C)Article
Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)22.6893.7570.1±5%, 111–317 °CThis work
Sr2Ag0.2Na0.8Nb5−xTaxO151.4482175 ± 15 % , −9.6–232.7 °C[62]
Sr2Na0.85Bi0.05Nb5−xTaxO151.680200 ± 15 % , −97.3–165.2 °C[63]
NBT–BT–KNN2.8194.4180--[64]
(1−x)BNT-xBCZT0.66583.270±15%, 81–330 °C[65]
(1−x)BNT-0.09BA-xST0.8448.390±5%, 132.8–391.8 °C[66]
0.93Na0.5Bi0.5TiO3-0.07LiTaO32.374.2200±15%, 15–600 °C[67]
(1−x)BNBT-xBS1.286.7115±15%, 45–338 °C[68]
(Na0.5Bi0.5)0.7Sr0.3TiO33.1391.14262--[69]
Sr1.88La0.12NaNb4.88−xTaxTi0.12O152.039435--[70]
Sr(Zr0.2Sn0.2Hf0.2Ti0.2Nb0.2)O33.5286.5283±5%, 52.4–362 °C[8]
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

Zhao, Y.; Li, Z.; Yang, S.; Mao, P.; Kang, R. Dielectric Temperature Stability and Enhanced Energy-Storage Performance of Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2O6 High-Entropy Ferroelectric Ceramics. Crystals 2025, 15, 26. https://doi.org/10.3390/cryst15010026

AMA Style

Zhao Y, Li Z, Yang S, Mao P, Kang R. Dielectric Temperature Stability and Enhanced Energy-Storage Performance of Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2O6 High-Entropy Ferroelectric Ceramics. Crystals. 2025; 15(1):26. https://doi.org/10.3390/cryst15010026

Chicago/Turabian Style

Zhao, Yingying, Ziao Li, Shiqiang Yang, Pu Mao, and Ruirui Kang. 2025. "Dielectric Temperature Stability and Enhanced Energy-Storage Performance of Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2O6 High-Entropy Ferroelectric Ceramics" Crystals 15, no. 1: 26. https://doi.org/10.3390/cryst15010026

APA Style

Zhao, Y., Li, Z., Yang, S., Mao, P., & Kang, R. (2025). Dielectric Temperature Stability and Enhanced Energy-Storage Performance of Sr0.4Ba0.6(Zr0.2Ti0.2Sn0.2Ta0.2Nb0.2)2O6 High-Entropy Ferroelectric Ceramics. Crystals, 15(1), 26. https://doi.org/10.3390/cryst15010026

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