Next Article in Journal
Durability of Resin Bonding to Dental 3Y-TZP Zirconia Using Different Adhesive Systems
Next Article in Special Issue
Implementation of Artificial Synapse Using IGZO-Based Resistive Switching Device
Previous Article in Journal
Influence of Silane Coupling Agent and Anionic Dispersant on the Dispersion Effect of Silicon Carbide Particles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Superparaelectric State in Relaxor Ferroelectric (Sr,Bi)TiO3-Bi(Mg,Ti)O3-Modified BaTiO3 Ceramics to Achieve High Energy Storage Performance

School of Materials Science and Engineering, Kumoh National Institute of Technology, Gumi 39177, Republic of Korea
*
Author to whom correspondence should be addressed.
Materials 2024, 17(2), 426; https://doi.org/10.3390/ma17020426
Submission received: 5 December 2023 / Revised: 8 January 2024 / Accepted: 12 January 2024 / Published: 15 January 2024
(This article belongs to the Special Issue Advanced Electrical Engineering Materials and Devices)

Abstract

:
Dielectric ceramic capacitors are highly regarded for their rapid charge–discharge, high power density, and cyclability in various advanced applications. However, their relatively low energy storage density has prompted intensive research aiming at developing materials with a higher energy density. To enhance energy storage properties, research has focused on modifying ferroelectric materials to induce relaxor ferroelectricity. The present study aims to induce a superparaelectric (SPE) state in relaxor ferroelectrics near room temperature by altering BaTiO3 ferroelectric ceramics using the (Sr,Bi)TiO3-Bi(Mg0.5Ti0.5)O3 system ((1−x)BT-x(SBT-BMT)). X-ray diffraction and Raman spectroscopy analysis demonstrated a shift in the crystal structure from tetragonal to cubic with an increasing x content. Notably, the compositions (except x = 0.1) satisfied the criteria for the SPE state manifestation near room temperature. The x = 0.2 specimen displayed characteristics at the boundary between the relaxor ferroelectric and SPE phases, while x ≥ 0.3 specimens exhibited increased SPE state fractions. Despite reduced maximum polarization, x ≥ 0.3 specimens showcased impressive energy storage capabilities, attributed to the enhanced SPE state, especially for x = 0.3, with impressive characteristics: a recoverable energy density (Wrec) of ~1.12 J/cm3 and efficiency (η) of ~94% at 170 kV/cm applied field. The good stability after the charge–discharge cycles reinforces the significance of the SPE phase in augmenting energy storage in relaxor ferroelectric materials, suggesting potential applications in high-energy density storage devices.

1. Introduction

Among various energy storage devices, dielectric ceramic capacitors exhibit advantages in advanced fields such as defense, healthcare, and smart mobility due to their rapid charge–discharge, high power density, and excellent cyclability [1,2,3]. However, the relatively low energy storage density of dielectric capacitors limits their practical applications, prompting extensive research aiming at developing dielectric ceramic materials with a higher energy density [4,5,6,7,8]. The energy storage characteristics of dielectrics depend on the changes in polarization, induced by an external electric field, and are commonly expressed by the following equations:
W = 0 P m a x E d P
W r e c = P r P m a x E d P
η = W r e c W × 100 %
Here, W represents the charging energy storage density, Wrec stands for the recoverable energy storage density (discharge energy storage density), η denotes the efficiency, and Pmax and Pr, respectively, indicate the maximum polarization during the charging process and the remnant polarization during the discharge process. To achieve a high energy density and high efficiency, a high Pmax, a high dielectric breakdown strength, and a small hysteresis area on the polarization-electric field (P-E) curve are required.
Research aimed at enhancing the energy storage properties of dielectric ceramics has primarily progressed towards inducing relaxor ferroelectricity through compositional modifications of ferroelectric materials [9,10]. While normal ferroelectrics possess high Pmax values, their large Pr values lead to lower Wrec and η. By doping various ions, it is possible to transform long-range ordered microscale ferroelectric domains into short-range ordered nanoscale domains, thereby inducing relaxor ferroelectricity [11,12,13]. The reduction in domain size and weakening inter-domain coupling reduce the domain switching energy barriers, minimizing hysteresis and consequently increasing Wrec and η [14,15,16,17,18].
In relaxor ferroelectrics, there are three characteristic temperatures that are sequentially observed at high temperatures: (1) the Burns temperature (TB), where nanodomains begin to form, (2) Tmax, where nanodomains grow, reaching the maximum dielectric permittivity, and (3) the freezing temperature (Tf), where nanodomains freeze, and which represents irreversible ferroelectricity [19]. Previous studies on relaxor ferroelectrics mainly evaluated the energy storage performance within the temperature range of Tf < T < Tmax [4,5]. In this case, although high Pmax values can be achieved, the presence of large hysteresis results in low Wrec and η. To simultaneously achieve high Wrec and η, attention needs to be given to the reported superparaelectric (SPE) relaxor ferroelectrics [7,8,9]. The SPE state occurs within the temperature range of Tmax < T < TB. In this temperature range, the size of nanodomains significantly decreases, and inter-domain coupling weakens drastically, resulting in extremely slim hysteresis due to significantly reduced domain switching energy barriers [16,17,18].
In this study, we aimed to induce an SPE state near room temperature by compositional modification of BaTiO3 (BT) ferroelectric ceramics. To harness robust relaxor ferroelectric properties from BT, which exhibits a high ferroelectric-to-paraelectric transition temperature (~120 °C), we opted for compositional modification using the (Sr,Bi)TiO3-Bi(Mg0.5Ti0.5)O3 system. The (Sr,Bi)TiO3 (SBT) ceramics exhibit relaxor ferroelectric behavior, characterized by Bi3+ ions partially substituting the A-site of cubic SrTiO3 ceramics. The introduction of Bi3+ ions aids in creating an off-center ion on the A-site, attributed to the 6s2 lone pair nature of Bi3+. This substitution of Bi3+ for Sr2+ at the A-site results in a charge imbalance, leading to the generation of Sr vacancies and subsequent defects within the system. These defect centers, along with off-center Bi-ions, induce dipoles and local random fields. These local random fields play a crucial role in the formation of polar nanoregions, consequently manifesting relaxor ferroelectric behavior [20,21,22,23,24,25,26]. Research on energy storage through compositional modification of SBT ceramics has also been extensively reported [26,27,28,29]. Moreover, it has been reported that incorporating Bi(Mg0.5Ti0.5)O3 (BMT) at levels exceeding 5 mol% into BT ceramics induces a pseudo-cubic state [30]. Therefore, we designed the (1−x)BT-x(SBT-BMT) system to explore compositions that satisfy the condition of Tmax < T < TB. We then investigated the structural, dielectric, ferroelectric, and energy storage properties concerning variations in x.

2. Materials and Methods

The (1−x)BaTiO3-x(0.75(Sr0.88Bi0.08)TiO3-0.25Bi(Mg0.5Ti0.5)O3) ((1−x)BT-x(SBT-BMT)) ceramics with x = 0.1, 0.2, 0.3, and 0.4 were prepared using solid-state synthesis. Raw material powders including BaCO3 (99.95%, Kojundo Korea Co., Uiwang-si, Korea), TiO2 (99.99%, Kojundo Korea Co., Uiwang-si, Korea), SrCO3 (99.9%, Kojundo Korea Co., Uiwang-si, Korea), Bi2O3 (99.9%, Kojundo Korea Co., Uiwang-si, Korea), and MgO (99.9%, Kojundo Korea Co., Uiwang-si, Korea) were utilized for the synthesis process. The raw powders, weighed according to each composition, were placed into a nylon container containing zirconia balls (YSZ, SciLab, Seoul, Korea) and anhydrous ethanol (99.9%, Samchun Pure Chemical Co., Seoul, Korea). These were subjected to a typical ball milling (~200 rotations per minute) process for 12 h to undergo primary grinding and mixing. After mixing, the slurries were dried and then calcined at 900 °C for 2 h. The calcined powders underwent another milling process using the same ball milling procedure for 24 h. Then, the dried powders were subjected to a uniaxial pressure of 100 MPa to shape them into discs (diameter: 10 mm, thickness: 1.5 mm). Subsequently, they were sintered at temperatures ranging from 1200 to 1275 °C for 2 h at a heating rate of 4 °C/min. Finally, they were cooled down to room temperature inside the furnace. Specifically, for x = 0.1, the sintering temperature was 1275 °C; for x = 0.2, it was 1250 °C; for x = 0.3, it was 1225 °C; and for x = 0.4, it was 1200 °C. The corresponding relative densities of these samples were found to be 94% (for x = 0.1), 95% (for x = 0.2), 96% (for x = 0.3), and 95% (for x = 0.4). To measure the electrical characteristics of the sintered specimens, Ag paste was applied to both sides. The specimens were then heated at a rate of 4 °C/min until reaching 700 °C, where the Ag electrodes were sintered for 1 h.
The phase and structural analysis of the (1−x)BT-x(SBT-BMT) ceramics was conducted by X-ray diffraction (XRD) (SmartLab, Rigaku, Tokyo, Japan) with Cu-Kα radiation (λ = 1.54056 Å). The XRD data were obtained in the range of 15–80° at a scanning rate of 1°/min with a 0.02° step. Raman spectroscopy (System1000, Renishaw, Wotton-under-Edge, UK) was used to analyze the vibrational properties, with a ± 0.001 cm−1 error limit. Scanning electron microscope (SEM) (JSM-6500F, JEOL, Tokyo, Japan) was utilized to see the surface morphology of the sintered samples with a 15 kV acceleration voltage. Temperature-dependent dielectric tests were conducted with impedance analyzer (IM3570, Hioki, Ueda, Japan) at frequencies of 10 kHz, 50 kHz, and 100 kHz, using a spring-loaded, Pt-based contact mechanism capable of withstanding temperatures up to 600 °C. The tests were performed with a 1 V oscillation potential and a heating rate of 3 °C/min in a rectangular muffle furnace, with a ±2 °C error value. Additionally, another temperature sensor (with a ±1 °C error) was positioned close to the sample during measurements near the sample stage. The corresponding values were recorded using automatic LabVIEW 2015 SP1 software connected to a computer via a cable.
The ferroelectric test was performed using a PK-CPE1801 ferroelectric test system (PolyK Technologies, Philipsburg, PA, USA), equipped with a PolyK sample stage made of stainless steel and aluminum. During the measurement, the sample stage was immersed in silicon oil.

3. Results and Discussion

Figure 1a depicts the XRD patterns of the sintered (1−x)BT-x(SBT-BMT) ceramics. All samples exhibited a perovskite structure without any detectable secondary phases within the XRD limits. To elucidate the peak shape, the range between 44.6 and 46° is magnified in Figure 1b. The diffraction peak near 45° split into two peaks for the x = 0.1 specimen, indicating a tetragonal phase (space group: P4mm, JCPDS No. 05-0626) [31]. The diffraction peaks merged into a single peak for the compositions with x ≥ 0.2, indicating a cubic phase (space group:   P m 3 ¯ m , JCPDS No. 31-0174) [32]. Additionally, the diffraction peaks exhibited a shift towards higher values with increasing x (Figure 1b), suggesting a lattice contraction due to the substitution of elements with smaller ionic radii by the perovskite A-site (r(Ba2+): 1.61 Å, r(Sr2+): 1.44 Å, and r(Bi3+): 1.45 Å [33,34]. To evaluate the lattice parameters as a function of x, Rietveld refinement was performed using the Fullprof software package version 7.95. The x = 0.1 composition fit well with the P4mm space group, whereas the compositions with x ≥ 0.2 fit well with the P m 3 ¯ m space group. The calculated lattice parameters are presented in Table 1. Additionally, Raman spectroscopy was conducted in the range of 100 to 1000 cm−1 to investigate the vibrational properties of the (1−x)BT-x(SBT-BMT) ceramics, as shown in Figure 1c. Raman modes were evidenced at 270, 306, 520, and 720 cm−1. BaTiO3 exhibits a tetragonal ( P 4 m m ) structure at room temperature, which is characterized by a sharp mode at 306 cm−1 along with asymmetric and broad modes at 270, 520, and 720 cm−1 in its Raman spectrum [35,36,37]. Furthermore, above 120 °C, BaTiO3 undergoes a phase transition to the cubic phase ( P m 3 ¯ m ), resulting in the weakening and broadening of the 306 cm−1 mode, along with further broadening of the 270, 520, and 720 cm−1 modes [38,39,40]. For the x = 0.1 specimen, a sharp mode was observed at 306 cm−1. However, for the compositions with x ≥ 0.2, the intensity of the 306 cm−1 mode decreased and broadened. Additionally, the modes at 270, 520, and 720 cm−1 also became broader, indicating a cubic phase. The modes at 270, 306, 520, and 720 cm−1 were identified as A1(TO), B1,E(TO + LO), A1,E(TO), and A1,E(LO), respectively, as presented in Figure 1c. These nomenclatures are described in detail in the reported results [35,41]. The XRD and Raman spectra results in Figure 1 both indicate that SBT-BMT was incorporated into the BT lattice successfully.
Figure 2a–d illustrate the variation in the relative dielectric permittivity (εr) and dielectric loss tangent (tan δ) of the (1−x)BT-x(SBT-BMT) ceramics at different fixed frequencies (10 kHz, 50 kHz, and 100 kHz) with increasing temperature. For x = 0.1, the measuring temperature ranges from 16 to 306 °C, while for x ≥ 0.2, it ranges from −68 to 306 °C. The x = 0.1 ceramic exhibited a ferroelectric (tetragonal) to paraelectric (cubic) phase transition temperature at 83 °C (Figure 2a), in contrast to the pure BaTiO3, which underwent the phase transition at 120 °C. Furthermore, as x increased, Tmax (the temperature at which the maximum relative dielectric permittivity (εm) occurs) dramatically decreased from 83 °C (for x = 0.1) to <2 °C (for x = 0.4, Figure 2d), and the relative dielectric permittivity curve flattened and widened, demonstrating relaxor characteristics [42]. The decrease in Tmax towards lower temperatures is associated with a reduction in the tetragonality factor (c/a ratio) due to the substitution of SBT-BMT into the BT lattice, as indicated by XRD and Raman spectroscopy [43,44]. Furthermore, the introduction of SBT-BMT into the BT host lattice increases the site disorder and charge fluctuation, thereby reducing the polar strength of TiO6 octahedra and resulting in a decrease in Tmax with increasing x. For the compositions with x ≥ 0.2, both εr and tan δ exhibited frequency dispersions in the temperature range from the starting temperature to near Tmax. These frequency dispersions exhibited opposite trends in εr and tan δ (as indicated by the arrows in Figure 2b), suggesting relaxor features in the materials [45,46]. The dielectric dispersion behavior of the (1−x)BT-x(SBT-BMT) ceramics was investigated using the modified Curie–Weiss law as described in Equation (4),
1 ε r 1 ε m = ( T T m a x ) γ C ( γ : d e g r e e   o f   d i f f u s e n e s s , C : c o n s t a n t )
where γ represents the degree of diffuseness, ranging between 1 (normal ferroelectric) and 2 (ideal relaxor ferroelectric) [47].
The γ values for all compositions were calculated to be above 1.7 (Figure 3a), which aligned well with the relaxor characteristics shown in Figure 2. The γ value increased to 1.8 until x = 0.2. However, for x ≥ 0.3, γ slightly decreased to 1.709. This decrease is attributed to the weakening of the ferroelectricity in the relaxor ferroelectric material [48,49]. The weakening of ferroelectricity in relaxor ferroelectric materials is observed during the transition from a normal relaxor ferroelectric state to the superparaelectric (SPE) state [6]. In relaxor ferroelectric materials in the SPE state, a higher energy storage efficiency can be achieved compared to the normal relaxor ferroelectric state due to an increase in the nonpolar phase and a decrease in the ferroelectric domain fraction [7,8,9,16,17,18].
We investigated the Tmax and TB of the (1−x)BT-x(SBT-BMT) ceramics to identify compositions representing the SPE state. TB refers to the onset temperature that satisfies the Curie–Weiss law in Equation (5), and in Figure 3b, we depict the Tmax and TB of the (1−x)BT-x(SBT-BMT) ceramics.
ε r = C T T 0 ( T 0 : C u r i e W e i s s   t e m p e r a t u r e )
Except for x = 0.1, all compositions satisfy the condition of Tmax < T < TB near room temperature, indicating that these compositions represent the SPE state. Based on the results in Figure 3b, we illustrate the composition vs. temperature phase diagram for normal relaxor ferroelectric, SPE, and paraelectric phases in Figure 3c. As x increases, the temperature range of the SPE phase expands, and it is confirmed that specimens x = 0.3 and 0.4 are clearly in the SPE state near room temperature.
Figure 4 represents the room temperature ferroelectric properties of the (1−x)BT-x(SBT-BMT) ceramics measured at a frequency of 10 Hz. At x = 0.1, the P-E and J-E curves of typical ferroelectric materials were obtained, indicating that the x = 0.1 specimen was in the state of T < Tf. However, at x = 0.2, the specimen exhibited broad peaks in the J-E curve, observed around E = 0 kV/cm, and a slim P-E curve, indicating the typical characteristics of the relaxor ferroelectric material. With an increase in x, the P-E curve became slimmer, and the peaks in the J-E curve broadened further, showing an increase in the squareness of the J-E curve. The increased squareness of the J-E curve indicated the enhancement of paraelectric characteristics [50,51]. Given the phase diagram depicted in Figure 3c, the x = 0.2 specimen was lying near the boundary between the relaxor ferroelectric phase and the SPE phase around room temperature. Therefore, its relative fraction of the SPE state could be small. However, for compositions with x ≥ 0.3, the phase boundary temperature decreased sufficiently below room temperature, suggesting a stronger enhancement of the SPE state in these compositions.
A high dielectric breakdown strength (DBS) is crucial in ensuring a high energy storage density. As seen in Figure 5a, the maximum available electric field (Emax) of the (1−x)BT-x(SBT-BMT) ceramics was significantly higher for specimens at x = 0.3 and 0.4 compared to those at x = 0.1 and 0.2. In each composition, the Emax value was deliberately set to be lower than the DBS. In the case of the x = 0.3 composition, five specimens were tested, resulting in Emax values of 168 kV/cm (first specimen), 170 kV/cm (second specimen), 172 kV/cm (third specimen), 170 kV/cm (fourth specimen), and 176 kV/cm (fifth specimen). The corresponding DBS value, determined using the Weibull distribution function, was approximately 172 kV/cm. Consequently, the Emax value of 170 kV/cm, which was lower than the DBS, was used to evaluate the energy storage properties of the x = 0.3 specimen. The DBS values for the other compositions were determined as follows: 102 kV/cm (for x = 0.1), 104 kV/cm (for x = 0.2), and 163 kV/cm (for x = 0.4). Subsequently, the Emax values of 100 kV/cm (for x = 0.1 and x = 0.2) and 160 kV/cm (for x = 0.4) were selected to assess their energy density properties in Figure 5. Despite the significantly larger grain size in the x = 0.4 specimen compared to that in the x = 0.1 specimen (Figure 5b), the notably higher Emax value exhibited by the x = 0.4 specimen indicates a substantial dependence of the dielectric breakdown strength of the (1−x)BT-x(SBT-BMT) ceramics on the fraction of the SPE state. The high Pmax and low Pr are also essential for achieving a high energy storage density. As observed in Figure 5c, Pmax shows a gradual decrease with increasing x, indicating that the emergence of the SPE state leads to a reduction in the fraction of polar nanoregions (PNRs). This reduction weakens the mutual interaction between ferroelectric domains [16,17,18,52,53]. However, the increase in the SPE phase and the decrease in PNRs had a more significant effect of sharply reducing the Pr (Figure 5c). Despite the decrease in Pmax, the remarkable increase in dielectric breakdown strength and the sharp reduction in Pr resulted in compositions with x ≥ 0.3 exhibiting excellent energy storage performance, with Wrec > 1 J/cm3 and η > 93% (Figure 5d).
Figure 6 represents the room temperature energy storage characteristics of the x = 0.3 (0.7BT-0.3(SBT-BMT)) specimen according to the electric field strength at a frequency of 10 Hz. Very slim unipolar P-E curves were observed at all electric field conditions (Figure 6a), resulting in a consistently high η of about 94% at all electric field conditions (Figure 6b). Wrec increased linearly with the increase in the electric field, showing a high value of 1.12 J/cm3 in an electric field of 170 kV/cm. Table 2 was designed to compare the energy storage properties of the 0.7BT-0.3(SBT-BMT) ceramic with those reported for other BT-based ceramics. From a practical application perspective, it is reasonable to regard both energy density and efficiency as equally crucial factors for high-energy-density capacitors. While the achieved energy density of the 0.7BT-0.3(SBT-BMT) ceramic is comparable to most reported values in Table 2, with only a few exceptions, its efficiency stands out notably by being particularly competitive.
Figure 7a shows the room temperature unipolar P-E curves of the 0.7BT-0.3(SBT-BMT) specimen at various frequencies under a 120 kV/cm field. In the frequency range from 2 to 20 Hz, there was no significant change in the P-E curves or energy density values, suggesting a good frequency stability of the energy storage performance. Furthermore, as seen in Figure 7b,c, the 0.7BT-0.3(SBT-BMT) specimen maintained a high Pmax and low Pr even after 105 cycles of charge–discharge tests at an electric field of 120 kV/cm. Consequently, Wrec showed a minimal decrease of within 3%, validating the good cyclability of the 0.7BT-0.3(SBT-BMT) specimen.

4. Conclusions

In this study, we investigated the structural, dielectric, ferroelectric, and energy storage properties concerning the variations in x in the (1−x)BT-x(SBT-BMT) system, designed to meet the conditions for a superparaelectric (SPE) state (Tmax < T < TB) near room temperature. Analyses from XRD and Raman spectroscopy revealed a shift in the crystal structure from tetragonal to cubic with increasing x content. The relative dielectric permittivity versus temperature curves flattened, and Tmax decreased as x increased, indicating enhanced relaxor behavior. Furthermore, except for the x = 0.1 composition, all compositions met the criteria of Tmax < T < TB near room temperature, signifying the manifestation of the SPE state in these compositions. A comparative examination of P-E curves across compositions revealed that the x = 0.2 specimen, at temperatures around room temperature, displayed characteristics at the boundary between the relaxor ferroelectric and SPE phases. It exhibited a relatively low SPE state fraction and retained attributes of normal relaxor ferroelectric behavior. On the other hand, specimens with x ≥ 0.3 exhibited slimmer P-E curves with a lower Tmax compared to the x = 0.2 specimen, indicating an increased fraction of the SPE state at room temperature. Despite the decline in Pmax, the x ≥ 0.3 specimens showed impressive energy storage capabilities, which can be attributed to significantly reduced Pr and elevated Emax values, both stemming from the enhanced SPE state. Particularly, the x = 0.3 (0.7BT-0.3(SBT-BMT)) ceramic exhibited outstanding energy storage characteristics, with Wrec = 1.12 J/cm3 and η of 94% at 170 kV/cm. Moreover, after 105 charge–discharge cycles at 120 kV/cm, the Wrec only displayed a minimal decrease of within 3%, confirming its good stability. This result highlights the significance of the SPE phase in enhancing energy storage capabilities within relaxor ferroelectric materials, showcasing the potential applicability of the (1−x)BT-x(SBT-BMT) system in high-energy-density storage devices.

Author Contributions

Conceptualization, I.-R.Y. and K.-H.C.; experiment, data acquisition, and visualization, J.-Y.P. and M.-S.K.; analysis and writing—original draft preparation, S.-H.C., I.-R.Y. and A.K.Y.; writing—review and editing, K.-H.C.; supervision, K.-H.C.; funding acquisition, K.-H.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Research Foundation (NRF) of Korea funded by the Ministry of Education (NRF-2019R1I1A3A01058105, NRF-2018R1A6A1A03025761).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chu, B.; Zhou, X.; Ren, K.; Neese, B.; Lin, M.; Wang, Q.; Bauer, F.; Zhang, Q.M. A Dielectric Polymer with High Electric Energy Density and Fast Discharge Speed. Science 2006, 313, 334. [Google Scholar] [CrossRef] [PubMed]
  2. 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] [PubMed]
  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. 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 2019, 365, 578–582. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, J.; Saremi, S.; Acharya, M.; Velarde, G.; Parsonnet, E.; Donahue, P.; Qualls, A.; Garcia, D.; Martin, L.W. Ultrahigh capacitive energy density in ion-bombarded relaxor ferroelectric films. Science 2020, 369, 81–84. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, D.; Tian, J.; Tian, S.; Yu, F.; Ren, K. Composition design of BNBT-ST relaxor ferroelectric ceramics in superparaelectric state with ultrahigh energy density. Ceram. Int. 2023, 49 Pt A, 27750–27757. [Google Scholar] [CrossRef]
  7. Wang, K.; Ouyang, J.; Wuttig, M.; Zhao, Y.-Y.; Cheng, H.; Zhang, Y.; Su, R.; Yan, J.; Zhong, X.; Zeng, F. Superparaelectric (Ba0.95,Sr0.05)(Zr0.2,Ti0.8)O3 Ultracapacitors. Adv. Energy Mater. 2020, 10, 2001778. [Google Scholar] [CrossRef]
  8. Pan, H.; Lan, S.; Xu, S.; Zhang, Q.; Yao, H.; Liu, Y.; Meng, F.; Guo, E.-J.; Gu, L.; Yi, D.; et al. Ultrahigh energy storage in superparaelectric relaxor ferroelectrics. Science 2021, 374, 100–104. [Google Scholar] [CrossRef]
  9. 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]
  10. Cho, S.; Yun, C.; Kim, Y.S.; Wang, H.; Jian, J.; Zhang, W.; Huang, J.; Wang, X.; Wang, H.; MacManus-Driscoll, J.L. Strongly enhanced dielectric and energy storage properties in lead-free perovskite titanate thin films by alloying. Nano Energy 2018, 45, 398–406. [Google Scholar] [CrossRef]
  11. Ogihara, H.; Randall, C.A.; Trolier-McKinstry, S. High-Energy Density Capacitors Utilizing 0.7BaTiO3–0.3BiScO3 Ceramics. J. Am. Ceram. Soc. 2009, 92, 1719–1724. [Google Scholar] [CrossRef]
  12. Pan, H.; Ma, J.; Ma, J.; Zhang, Q.; Liu, X.; Guan, B.; Gu, L.; Zhang, X.; Zhang, Y.-J.; Li, L.; et al. Giant energy density and high efficiency achieved in bismuth ferrite-based film capacitors via domain engineering. Nat. Commun. 2018, 9, 1813. [Google Scholar] [CrossRef] [PubMed]
  13. Peng, B.; Zhang, Q.; Li, X.; Sun, T.; Fan, H.; Ke, S.; Ye, M.; Wang, Y.; Lu, W.; Niu, H.; et al. Giant Electric Energy Density in Epitaxial Lead-Free Thin Films with Coexistence of Ferroelectrics and Antiferroelectrics. Adv. Electron. Mater. 2015, 1, 1500052. [Google Scholar] [CrossRef]
  14. Dong, X.; Li, X.; Chen, X.; Wu, J.; Zhou, H. Simultaneous enhancement of polarization and breakdown strength in lead-free BaTiO3-based ceramics. Chem. Eng. J. 2021, 409, 128231. [Google Scholar] [CrossRef]
  15. 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]
  16. Yang, Z.; Du, H.; Jin, L.; Hu, Q.; Wang, H.; Li, Y.; Wang, J.; Gao, F.; Qu, S. Realizing high comprehensive energy storage performance in lead-free bulk ceramics via designing an unmatched temperature range. J. Mater. Chem. A 2019, 7, 27256–27266. [Google Scholar] [CrossRef]
  17. Bell, A.J. Calculations of dielectric properties from the superparaelectric model of relaxors. J. Phys. Condens. Matter 1993, 5, 8773. [Google Scholar] [CrossRef]
  18. Li, F.; Zhang, S.; Damjanovic, D.; Chen, L.-Q.; Shrout, T.R. Ferroelectrics: Local Structural Heterogeneity and Electromechanical Responses of Ferroelectrics: Learning from Relaxor Ferroelectrics (Adv. Funct. Mater. 37/2018). Adv. Funct. Mater. 2018, 28, 1870262. [Google Scholar] [CrossRef]
  19. Bokov, A.A.; Ye, Z.-G. Dielectric Relaxation in Relaxor Ferroelectrics. J. Adv. Dielectr. 2012, 02, 1241010. [Google Scholar] [CrossRef]
  20. Zhang, G.-F.; Cao, M.; Hao, H.; Liu, H. Energy Storage Characteristics in Sr(1-1.5x)BixTiO3 Ceramics. Ferroelectrics 2013, 447, 86–94. [Google Scholar] [CrossRef]
  21. Li, L.; Fan, P.; Wang, M.; Takesue, N.; Salamon, D.; Vtyurin, A.N.; Zhang, Y.; Tan, H.; Nan, B.; Lu, Y.; et al. Review of lead-free Bi-based dielectric ceramics for energy-storage applications. J. Phys. D Appl. Phys. 2021, 54, 293001. [Google Scholar] [CrossRef]
  22. Zuo, C.; Yang, S.; Cao, Z.; Yu, H.; Wei, X. Excellent energy storage and hardness performance of Sr0.7Bi0.2TiO3 ceramics fabricated by solution combustion-synthesized nanopowders. Chem. Eng. J. 2022, 442, 136330. [Google Scholar] [CrossRef]
  23. Ang, C.; Yu, Z.; Vilarinho, P.M.; Baptista, J.L. Bi:SrTiO3: A quantum ferroelectric and a relaxor. Phys. Rev. B 1998, 57, 7403–7406. [Google Scholar] [CrossRef]
  24. Okhay, O.; Wu, A.; Vilarinho, P.M.; Tkach, A. Dielectric relaxation of Sr1–1.5xBixTiO3 sol-gel thin films. J. Appl. Phys. 2011, 109, 064103. [Google Scholar] [CrossRef]
  25. Zhu, X.; Shi, P.; Kang, R.; Li, S.; Wang, Z.; Qiao, W.; Zhang, X.; He, L.; Liu, Q.; Lou, X. Enhanced energy storage density of Sr0.7BixTiO3 lead-free relaxor ceramics via A-site defect and grain size tuning. Chem. Eng. J. 2021, 420, 129808. [Google Scholar] [CrossRef]
  26. Zhao, P.; Tang, B.; Si, F.; Yang, C.; Li, H.; Zhang, S. Novel Ca doped Sr0.7Bi0.2TiO3 lead-free relaxor ferroelectrics with high energy density and efficiency. J. Eur. Ceram. Soc. 2020, 40, 1938–1946. [Google Scholar] [CrossRef]
  27. Chao, M.; Liu, J.; Zeng, M.; Wang, D.; Yu, H.; Yuan, Y.; Zhang, S. High discharge efficiency of (Sr, Pb, Bi) TiO3 relaxor ceramics for energy-storage application. Appl. Phys. Lett. 2018, 112, 203903. [Google Scholar] [CrossRef]
  28. 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]
  29. Yu, Z.; Ang, C.; Guo, R.; Bhalla, A.S. Dielectric properties and tunability of (Sr,Bi)TiO3 with MgO additive. Mater. Lett. 2003, 57, 2927–2931. [Google Scholar] [CrossRef]
  30. Xiong, B.; Hao, H.; Zhang, S.; Liu, H.; Cao, M. Structure, Dielectric Properties and Temperature Stability of BaTiO3–Bi(Mg1/2Ti1/2)O3 Perovskite Solid Solutions. J. Am. Ceram. Soc. 2011, 94, 3412–3417. [Google Scholar] [CrossRef]
  31. Wu, X.; Zhao, H.; Han, W.; Wang, Z.; Li, F.; Li, J.; Xue, W. Facile preparation and dielectric properties of BaTiO3 with different particle sizes and morphologies. RSC Adv. 2023, 13, 11002–11009. [Google Scholar] [CrossRef]
  32. Yu, P.; Liu, W.; Gao, P.; Shao, T.; Zhao, S.; Han, Z.; Gu, X.; Zhang, J.; Wang, Y. Investigation on synthesis of tetragonal BaTiO3 nanopowders by a new wet chemical method. J. Mater. Sci. Mater. Electron. 2022, 33, 10828–10840. [Google Scholar] [CrossRef]
  33. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  34. Tkach, A.; Okhay, O. Comment on “Hole-pinned defect-dipoles induced colossal permittivity in Bi doped SrTiO3 ceramics with Sr deficiency”. J. Mater. Sci. Technol. 2021, 65, 151–153. [Google Scholar] [CrossRef]
  35. DiDomenico, M.; Wemple, S.H.; Porto, S.P.S.; Bauman, R.P. Raman Spectrum of Single-Domain BaTiO3. Phys. Rev. 1968, 174, 522–530. [Google Scholar] [CrossRef]
  36. Pasuk, I.; Neațu, F.; Neațu, Ș.; Florea, M.; Istrate, C.M.; Pintilie, I.; Pintilie, L. Structural Details of BaTiO3 Nano-Powders Deduced from the Anisotropic XRD Peak Broadening. Nanomaterials 2021, 11, 1121. [Google Scholar] [CrossRef]
  37. Yadav, A.K.; Fan, H.; Yan, B.; Wang, W.; Dong, W.; Wang, S. Structure evolutions with enhanced dielectric permittivity and ferroelectric properties of Ba(1−x)(La, Li)xTiO3 ceramics. J. Mater. Sci. Mater. Electron. 2021, 32, 23103–23115. [Google Scholar] [CrossRef]
  38. Shi, C.; Billinge, S.J.L.; Puma, E.; Bang, S.H.; Bean, N.J.H.; de Sugny, J.-C.; Gambee, R.G.; Haskell, R.C.; Hightower, A.; Monson, T.C. Barium titanate nanoparticles: Short-range lattice distortions with long-range cubic order. Phys. Rev. B 2018, 98, 085421. [Google Scholar] [CrossRef]
  39. Veerapandiyan, V.K.; Khosravi, H.S.; Canu, G.; Feteira, A.; Buscaglia, V.; Reichmann, K.; Deluca, M. B-site vacancy induced Raman scattering in BaTiO3-based ferroelectric ceramics. J. Eur. Ceram. Soc. 2020, 40, 4684–4688. [Google Scholar] [CrossRef]
  40. Pokorný, J.; Pasha, U.M.; Ben, L.; Thakur, O.P.; Sinclair, D.C.; Reaney, I.M. Use of Raman spectroscopy to determine the site occupancy of dopants in BaTiO3. J. Appl. Phys. 2011, 109, 114110. [Google Scholar] [CrossRef]
  41. Parsons, J.L.; Rimai, L. Raman spectrum of BaTiO3. Solid State Commun. 1967, 5, 423–427. [Google Scholar] [CrossRef]
  42. 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 2019, 7, 6222–6230. [Google Scholar] [CrossRef]
  43. Jiang, X.; Hao, H.; Yang, Y.; Zhou, E.; Zhang, S.; Wei, P.; Cao, M.; Yao, Z.; Liu, H. Structure and enhanced dielectric temperature stability of BaTiO3-based ceramics by Ca ion B site-doping. J. Mater. 2021, 7, 295–301. [Google Scholar] [CrossRef]
  44. Yao, G.; Wang, X.; Yang, Y.; Li, L. Effects of Bi2O3 and Yb2O3 on the Curie Temperature in BaTiO3-Based Ceramics. J. Am. Ceram. Soc. 2010, 93, 1697–1701. [Google Scholar] [CrossRef]
  45. Liu, J.; Ma, C.; Zhao, X.; Ren, K.; Zhang, R.; Shang, F.; Du, H.; Wang, Y. Structure, dielectric, and relaxor properties of BaTiO3-modified high-entropy (Bi0.2Na0.2K0.2Ba0.2Ca0.2)TiO3 ceramics for energy storage applications. J. Alloys Compd. 2023, 947, 169626. [Google Scholar] [CrossRef]
  46. 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]
  47. Zuo, J.; Yang, H.; Chen, J.; Li, C. Highly-reliable dielectric capacitors with excellent comprehensive energy-storage properties using Bi0.5Na0.5TiO3-based relaxor ferroelectric ceramics. J. Eur. Ceram. Soc. 2023, 43, 2452–2459. [Google Scholar] [CrossRef]
  48. Hu, D.; Pan, Z.; Zhang, X.; Ye, H.; He, Z.; Wang, M.; Xing, S.; Zhai, J.; Fu, Q.; Liu, J. Greatly enhanced discharge energy density and efficiency of novel relaxation ferroelectric BNT–BKT-based ceramics. J. Mater. Chem. C 2020, 8, 591–601. [Google Scholar] [CrossRef]
  49. Huang, N.; Liu, H.; Hao, H.; Yao, Z.; Cao, M.; Xie, J. Energy storage properties of MgO-doped 0.5Bi0·5Na0·5TiO3-0.5SrTiO3 ceramics. Ceram. Int. 2019, 45, 14921–14927. [Google Scholar] [CrossRef]
  50. Brown, E.; Ma, C.; Acharya, J.; Ma, B.; Wu, J.; Li, J. Controlling Dielectric and Relaxor-Ferroelectric Properties for Energy Storage by Tuning Pb0.92La0.08Zr0.52Ti0.48O3 Film Thickness. ACS Appl. Mater. Interfaces 2014, 6, 22417–22422. [Google Scholar] [CrossRef] [PubMed]
  51. 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]
  52. Petzelt, J. Dielectric Grain-Size Effect in High-Permittivity Ceramics. Ferroelectrics 2010, 400, 117–134. [Google Scholar] [CrossRef]
  53. Huamán, J.L.C.; Rivera, V.A.G.; Pinto, A.H.; Marega, E. 11-Multiferroic perovskite ceramics: Properties and applications. In Perovskite Ceramics; Huamán, J.L.C., Rivera, V.A.G., Eds.; Elsevier: Amsterdam, The Netherlands, 2023; pp. 339–381. [Google Scholar] [CrossRef]
  54. Shen, Z.; Wang, X.; Luo, B.; Li, L. BaTiO3–BiYbO3 perovskite materials for energy storage applications. J. Mater. Chem. A 2015, 3, 18146–18153. [Google Scholar] [CrossRef]
  55. Jiang, X.; Hao, H.; Zhang, S.; Lv, J.; Cao, M.; Yao, Z.; Liu, H. Enhanced energy storage and fast discharge properties of BaTiO3 based ceramics modified by Bi(Mg1/2Zr1/2)O3. J. Eur. Ceram. Soc. 2019, 39, 1103–1109. [Google Scholar] [CrossRef]
  56. 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]
  57. Liu, X.; Yang, H.; Yan, F.; Qin, Y.; Lin, Y.; Wang, T. Enhanced energy storage properties of BaTiO3-Bi0.5Na0.5TiO3 lead-free ceramics modified by SrY0.5Nb0.5O3. J. Alloys Compd. 2019, 778, 97–104. [Google Scholar] [CrossRef]
  58. Zhao, H.; Duan, X.; Yu, T.; Xu, D.; Zhao, W. Enhanced energy storage efficiency of Ba0.8Sr0.2TiO3 ceramics modified by BiTaO3. Solid State Commun. 2023, 360, 115050. [Google Scholar] [CrossRef]
  59. Borkar, S.N.; Deshpande, V.K. Effect of Mg substitution on microstructural, dielectric, ferroelectric and energy storage properties of BaTiO3. Phys. B Condens. Matter. 2022, 644, 414244. [Google Scholar] [CrossRef]
  60. Dong, X.; Chen, H.; Wei, M.; Wu, K.; Zhang, J. Structure, dielectric and energy storage properties of BaTiO3 ceramics doped with YNbO4. J. Alloys Compd. 2018, 744, 721–727. [Google Scholar] [CrossRef]
  61. Yi, X.; Ji, C.; Chen, G.; Yang, H.; Yong, H.; Fu, C.; Cai, W.; Gao, R.; Fan, T.; Wang, Z.; et al. Effects of Sintering Method and BiAlO3 Dopant on Dielectric Relaxation and Energy Storage Properties of BaTiO3–BiYbO3 Ceramics. Phys. Status Solidi A 2020, 217, 1900721. [Google Scholar] [CrossRef]
  62. Yin, M.; Zhang, Y.; Bai, H.-R.; Li, P.; Li, Y.-C.; Han, W.-F.; Hao, J.-G.; Li, W.; Wang, C.-M.; Fu, P. Preeminent energy storage properties and superior stability of (Ba(1–x)Bix)(Ti(1–x)Mg2x/3Tax/3)O3 relaxor ferroelectric ceramics via elongated rod-shaped grains and domain structural regulation. J. Mater. Sci. Technol. 2024, 184, 207–220. [Google Scholar] [CrossRef]
  63. Zhao, X.; Zhou, Z.; Liang, R.; Liu, F.; Dong, X. High-energy storage performance in lead-free (1-x)BaTiO3-xBi(Zn0.5Ti0.5)O3 relaxor ceramics for temperature stability applications. Ceram. Int. 2017, 43, 9060–9066. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of sintered (1−x)BT-x(SBT-BMT) ceramics. (b) Enlarged view of the XRD (200) peaks. (c) Raman spectra of (1−x)BT-x(SBT-BMT) ceramics.
Figure 1. (a) XRD patterns of sintered (1−x)BT-x(SBT-BMT) ceramics. (b) Enlarged view of the XRD (200) peaks. (c) Raman spectra of (1−x)BT-x(SBT-BMT) ceramics.
Materials 17 00426 g001
Figure 2. Relative dielectric permittivity (εr) and loss tangent (tan δ) versus temperature for (1−x)BT-x(SBT-BMT) ceramics, where (a) x = 0.1, (b) x = 0.2, (c) x = 0.3, and (d) x = 0.4. The y-axis of tan δ versus temperature is presented in logarithmic scale to enhance data visualization.
Figure 2. Relative dielectric permittivity (εr) and loss tangent (tan δ) versus temperature for (1−x)BT-x(SBT-BMT) ceramics, where (a) x = 0.1, (b) x = 0.2, (c) x = 0.3, and (d) x = 0.4. The y-axis of tan δ versus temperature is presented in logarithmic scale to enhance data visualization.
Materials 17 00426 g002
Figure 3. Dielectric characteristics of (1−x)BT-x(SBT-BMT) ceramics: (a) log(1/εr − 1/εm) vs. log(TTm) curves, (b) 1000/εr vs. T curves, and (c) phase diagram based on Tmax and TB.
Figure 3. Dielectric characteristics of (1−x)BT-x(SBT-BMT) ceramics: (a) log(1/εr − 1/εm) vs. log(TTm) curves, (b) 1000/εr vs. T curves, and (c) phase diagram based on Tmax and TB.
Materials 17 00426 g003
Figure 4. Polarization and current density versus electric field (P-E and J-E) curves of (1−x)BT-x(SBT-BMT) ceramics at a frequency of 10 Hz and an applied field of 100 kV/cm.
Figure 4. Polarization and current density versus electric field (P-E and J-E) curves of (1−x)BT-x(SBT-BMT) ceramics at a frequency of 10 Hz and an applied field of 100 kV/cm.
Materials 17 00426 g004
Figure 5. Energy storage characteristics of (1−x)BT-x(SBT-BMT) ceramics: (a) room temperature unipolar P-E curves at 10 Hz, (b) SEM microstructure (insert: grain size distribution), (c) Pmax and Pr, and (d) Wrec and η as a function of x measured at Emax.
Figure 5. Energy storage characteristics of (1−x)BT-x(SBT-BMT) ceramics: (a) room temperature unipolar P-E curves at 10 Hz, (b) SEM microstructure (insert: grain size distribution), (c) Pmax and Pr, and (d) Wrec and η as a function of x measured at Emax.
Materials 17 00426 g005
Figure 6. (a) Unipolar P-E curves of the 0.7BT-0.3(SBT-BMT) ceramic under different electric fields at a frequency of 10 Hz and (b) the calculated Wrec and η of the 0.7BT-0.3(SBT-BMT) ceramic as a function of applied electric field.
Figure 6. (a) Unipolar P-E curves of the 0.7BT-0.3(SBT-BMT) ceramic under different electric fields at a frequency of 10 Hz and (b) the calculated Wrec and η of the 0.7BT-0.3(SBT-BMT) ceramic as a function of applied electric field.
Materials 17 00426 g006
Figure 7. (a) Unipolar P-E curves at various frequencies (2–20 Hz) of 0.7BT-0.3(SBT-BMT) specimen under 120 kV/cm field, (b) unipolar P-E curves of 0.7BT-0.3(SBT-BMT) ceramic from 1 to 105 cycles at 10 Hz, and (c) calculated Pmax, Pr, and Wrec as a function of cycle.
Figure 7. (a) Unipolar P-E curves at various frequencies (2–20 Hz) of 0.7BT-0.3(SBT-BMT) specimen under 120 kV/cm field, (b) unipolar P-E curves of 0.7BT-0.3(SBT-BMT) ceramic from 1 to 105 cycles at 10 Hz, and (c) calculated Pmax, Pr, and Wrec as a function of cycle.
Materials 17 00426 g007
Table 1. Calculated lattice parameters and unit cell volume of (1−x)BT-x(SBT-BMT) ceramics.
Table 1. Calculated lattice parameters and unit cell volume of (1−x)BT-x(SBT-BMT) ceramics.
CompositionSpace GroupLattice ParametersVolume of Unit Cell
x = 0.1P4mma = b = 3.99555 ± 0.00003 Å
c = 4.01797 ± 0.00004 Å
c/a = 1.00561 ± 0.00001
64.145 ± 0.001 Å3
x = 0.2 P m 3 ¯ m a = b = c = 3.99556 ± 0.00001 Å
c/a = 1
63.787 ± 0.001 Å3
x = 0.3 P m 3 ¯ m a = b = c = 3.98836 ± 0.00001 Å
c/a = 1
63.443 ± 0.001 Å3
x = 0.4 P m 3 ¯ m a = b = c = 3.98101 ± 0.00001 Å
c/a = 1
63.093 ± 0.001 Å3
Table 2. Comparison of the energy storage properties of the 0.7BT-0.3(SBT-BMT) ceramic with other BaTiO3-based ceramics.
Table 2. Comparison of the energy storage properties of the 0.7BT-0.3(SBT-BMT) ceramic with other BaTiO3-based ceramics.
CompositionEnergy Storage Density, Wrec
(J/cm3)
Efficiency, η
(%)
Applied Field (kV/cm)Reference
0.91BaTiO3-0.09BiYbO30.7182.693[54]
0.85BaTiO3-0.15Bi(Mg1/2Zr1/2)O31.2595.4185[55]
0.85BaTiO3-0.15Bi(Zn2/3Nb1/3)O30.7993.5131[56]
0.92(0.65BaTiO3-0.35Bi0.5Na0.5TiO3)-0.08SrY0.5Nb0.5O31.3674.3152[57]
0.88Ba0.8Sr0.2TiO3-0.12BiTaO30.52698130[58]
BaTi0.95Mg0.05O31.0489.65350[59]
0.93BaTiO3-0.07YNbO40.61486.8173[60]
0.86BaTiO3-0.1BiYbO3-0.04BiAlO30.5997.44110[61]
(Ba0.9Bi0.1)(Ti0.9Mg0.2/3Ta0.1/3)O35.9787.4710[62]
0.86BaTiO3-0.14Bi(Zn0.5Ti0.5)O30.8194120[63]
0.7BaTiO3-0.3[0.75(Sr0.88Bi0.08)TiO3-0.25Bi(Mg0.5Ti0.5)O3]1.1294170This study
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

Yoo, I.-R.; Choi, S.-H.; Park, J.-Y.; Kim, M.-S.; Yadav, A.K.; Cho, K.-H. A Superparaelectric State in Relaxor Ferroelectric (Sr,Bi)TiO3-Bi(Mg,Ti)O3-Modified BaTiO3 Ceramics to Achieve High Energy Storage Performance. Materials 2024, 17, 426. https://doi.org/10.3390/ma17020426

AMA Style

Yoo I-R, Choi S-H, Park J-Y, Kim M-S, Yadav AK, Cho K-H. A Superparaelectric State in Relaxor Ferroelectric (Sr,Bi)TiO3-Bi(Mg,Ti)O3-Modified BaTiO3 Ceramics to Achieve High Energy Storage Performance. Materials. 2024; 17(2):426. https://doi.org/10.3390/ma17020426

Chicago/Turabian Style

Yoo, Il-Ryeol, Seong-Hui Choi, Je-Yeon Park, Min-Seok Kim, Arun Kumar Yadav, and Kyung-Hoon Cho. 2024. "A Superparaelectric State in Relaxor Ferroelectric (Sr,Bi)TiO3-Bi(Mg,Ti)O3-Modified BaTiO3 Ceramics to Achieve High Energy Storage Performance" Materials 17, no. 2: 426. https://doi.org/10.3390/ma17020426

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