Next Article in Journal
Influences of Initial Stresses on Formation of Shear Bands and Mechanical Properties in Binodal Decomposed Metallic Glass Composites
Previous Article in Journal
Engineering the Morphology and Properties of MoS2 Films Through Gaseous Precursor-Induced Vacancy Defect Control
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

BaTiO3–(Na0.5Bi0.5)TiO3 Ceramic Materials Prepared via Multiple Design Strategies with Improved Energy Storage

1
Guangdong Provincial Key Laboratory of Electronic Functional Materials and Devices, Huizhou University, Huizhou 516001, China
2
Shenzhen Key Laboratory of Advanced Materials, School of Materials Science and Engineering, Harbin Institute of Technology, Shenzhen 518055, China
3
School of Mechanical Engineering, College of Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
4
Department of Physics, Saveetha School of Engineering, Saveetha Institute of Medical and Technical Sciences (SIMATS), Chennai 602105, India
5
Hubei Key Laboratory of Micro-Nanoelectronic Materials and Devices, Hubei University, Wuhan 430062, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(22), 1724; https://doi.org/10.3390/nano15221724 (registering DOI)
Submission received: 23 October 2025 / Revised: 10 November 2025 / Accepted: 13 November 2025 / Published: 15 November 2025
(This article belongs to the Special Issue Perspectives on Physics of Advanced Nanomaterials and Interfaces)

Abstract

The investigation of environmentally friendly, Pb-free ceramic dielectric materials with excellent energy storage capability represents a fundamental yet challenging research direction for the development of next-generation high-power capacitors. In this study, linear dielectric Ca0.7La0.2(Mg1/3Nb2/3)O3 was added into [0.65BaTiO3–0.35(Na0.5Bi0.5)TiO3] to form a solid solution. The introduction of Ca0.7La0.2(Mg1/3Nb2/3)O3 modified the crystal structure, enhanced insulation performance and breakdown strength, and reduced hysteresis loss. These improvements collectively contributed to higher energy storage density and efficiency (η). The ceramic pellet with the optimal 10 mol% Ca0.7La0.2(Mg1/3Nb2/3)O3 demonstrated a higher retrievable energy density (~3.40 J cm−3) and efficiency (~81%) at a breakdown strength of 340 kV cm−1 compared to BaTiO3-based ferroelectric ceramics. The sample also exhibited good stability across a temperature range of 30–90 °C and a frequency range of 0.5–300 Hz. Thus, the as-prepared ceramics sample exhibited significant potential for pulsed power device applications.

Graphical Abstract

1. Introduction

The rapid rise in global energy consumption over recent decades has driven the demand for renewable, efficient, and environmentally friendly energy storage devices [1,2,3,4]. Dielectric capacitors, as essential components of modern electronic systems, can store energy electrostatically and are widely employed for energy storage due to their ability to rapidly and reliably store and release energy [5,6]. When an electric field is applied, the positive and negative charges inside the material’s atomic structure are slightly pulled apart. Energy is stored within this induced polarization. After the electric field is removed, the material releases this stored energy as it discharges. However, their relatively low energy storage density remains a significant limitation for pulsed power applications. Consequently, there is an urgent need for environmentally friendly (lead-free) energy storage materials that have both a high recoverable energy storage density (Wrec) and high energy storage efficiency (ƞ) [7,8]. Generally, the total energy storage density (W), Wrec, and ƞ are expressed as Equations (1)–(3) [9]:
W = 0 P max E d P
W rec = P r P max E d P
η = W rec W
where Pmax, Pr, and P indicate the maximum, remnant, and normal polarization, respectively; E denotes the applied electric field. Wrec can be improved by simultaneously achieving ΔP (PmaxPr) and a high E.
Energy storage ceramic materials with dielectric properties are generally classified into four categories: ferroelectric (FEc), linear dielectric, antiferroelectric (AFEc), and relaxor ferroelectric (RFc) ceramics [10,11]. RFc and AFEc ceramics are regarded as excellent energy storage materials owing to their high Pmax. However, many AFEc systems contain expensive noble metals such as silver or environmentally hazardous elements like lead [12,13,14]. RFc materials are particularly attractive because they contain polar nano-regions (PNRs). These PNRs enable rapid domain switching under an applied E, resulting in a small Pr. This, combined with their high Pmax values, allows them to simultaneously achieve excellent Wrec and η. Consequently, lead-free RFc materials are gaining significant attention for both environmental and economic reasons.
Barium titanate (BaTiO3, BT), the first discovered perovskite-type piezoelectric material, has been extensively utilized in commercial multilayer ceramic capacitors [15]. However, the low intrinsic breakdown electric field (Eb) and high coercivity field (Ec) of these devices limit a desirable W value. The fast response of polar nanoregions (PNRs) reduces Ec in the presence of electric fields, which substantially improves the efficiency (η) and energy density (W) [16,17,18,19]. Lv et al. [20] added 40 mol% of strontium bismuth titanate (Sr0.7Bi0.2TiO3) to a 0.6 [0.94 sodium bismuth titanate (Na0.5Bi0.5TiO3, NBT)–0.06 BT] system. The macroscopic domain was disrupted and Pr and Ec were reduced, which enhanced W and η. In addition, Yuan et al. [21] enhanced the threshold field and decreased Ec for introducing a long-range order in a BT–bismuth magnesium zirconate [Bi(Mg0.5Zr0.5)O3] system by forming PNRs. While various modification strategies exist, a long-term challenge remains in further improving the energy storage capacity of these ceramics [16].
Sodium bismuth titanate (Na0.5Bi0.5TiO3, NBT) is a lead-free ceramic material for energy storage. Its high Pmax (>40 μC cm−2) arises from the hybridization between the O2− 2p orbitals and the Bi3+ 6s2 lone pair electrons [22]. Incorporating NBT into BaTiO3 effectively enhances polarization and stabilizes dielectric performance. This additional O 2p-Bi 6s orbital hybridization compensates for the polarization deficiency in BT [23]. As a result, 0.65BT-0.35NBT-based ceramics have been extensively researched for their excellent W. High energy storage performance has been reported for (1–x)(0.65BT–0.35NBT)– xBi(Mg2/3Nb1/3)O3 materials with η of 90.3% and Wrec of 1.6 J cm−3 [24]. Similarly, Dai et al. [25] et al. achieved η of 90.18% and Wrec of 2.02 J cm−3 at 206 kV cm−1 by adding Sr(Sc0.5Nb0.5)O3 into 0.65BT-0.35NBT ceramics.
Herein, 0.65BT–0.35NBT was selected as a basic component, while Ca0.7La0.2(Mg1/3Nb2/3)O3 (CLMN), a linear dielectric material, was used for doping 0.65BT–0.35NBT. This is a rational, multifunctional design strategy. The effect of CLMN on the energy storage performance of 0.65BT–0.35NBT was investigated. Calcium titanate (CaTiO3) is a linear dielectric material with a high Eb and negligible Pr. Incorporating CaTiO3 into ferroelectric matrices to form solid solutions can enhance W [26,27]. The addition of the rare earth element lanthanum (La) is known to refine grains [28], while the presence of magnesium oxide (MgO) and niobium pentoxide (Nb2O5) with their wide bandgaps (Eg, 7.8 eV and 3.4 eV, respectively) improves the material’s insulation performance. These improved relaxation behavior, and reduced Pr in CLMN-doped 0.65BT–0.35NBT, optimizing its energy storage properties. Introducing 10 mol% of CLMN into the ceramic significantly enhanced its energy storage performance, with an Eb of 340 kV cm−1, η of ~81%, and a Wrec of ~3.40 J cm−3. The resulting ceramic also demonstrated good stability over a temperature range of 30–90 °C and a frequency range of 0.5–300 Hz. This approach provides an effective strategy to optimize the energy storage capability of BT-based ceramics, making them suitable for advanced power systems.

2. Experimental Procedure

(1–x)[0.65BaTiO3–0.35(Na0.5Bi0.5)TiO3]–xCa0.7La0.2(Mg1/3Nb2/3)O3 (abbreviated as xCLMN, x = 0.08, 0.10, 0.12, 0.14 and 0.16) lead-free ceramic samples were synthesized using high-purity raw materials: bismuth oxide (Bi2O3, AR 99%), sodium carbonate (Na2CO3, AR 99.8%), barium carbonate (BaCO3, AR 99%), calcium carbonate (CaCO3, AR 99%), lanthanum oxide (La2O3, metal basis 99.9%), magnesium oxide (MgO, AR 98%), niobium oxide (Nb2O5, Metals basis 99.9%) and titanium oxide (TiO2, AR 99%) using a standard solid-state method. The chemicals were procured from Alladin Scientific, Shanghai, China. The powders and ceramic pellets were prepared according to the procedure described in our previous work [29]. The powders were weighted according to the stoichiometric ratio and ball milled in ethanol for 24 h. After drying, the mixtures were calcined at 900 °C for 4 h in air. The calcined powders were milled again and pressed into disks with 12 mm in diameter and 1.5 mm in thickness using 6 wt% PVA aqueous solution as binder. The green pellets were heated to 600 °C for 6 h to burn out the binder. The pellets were sintered in a covered alumina crucible at 1050–1170 °C for 3 h to obtain high density.
The crystal structures were characterized using a Renishaw Raman spectrometer and X-ray diffraction (XRD) on SmartLab (Rigaku, Tokyo, Japan) instruments under the following conditions: (i) step size, 0.02°; (ii) 2θ range, 20–80°; and (iii) scanning speed, 2° min−1. Raman spectroscopy (Renishaw InVia Reflex) with a radiation of Ar+ laser (λ = 514.5 nm) was carried out to investigate the structural properties of ceramics at room temperature. The thermally etched surface morphology of the ceramic pellets was examined via scanning electron microscopy (SEM; VEGA3/XUM, TESCAN, Brno, Czechia), and elemental composition was analyzed using energy-dispersive X-ray spectroscopy (EDS) during SEM analysis. Grain size distribution of the samples was determined using Nano Measurer software 1.2. The ceramic samples were manually ground and polished to a thickness of 0.5 mm for electrode fabrication. Silver electrodes were applied and subsequently heated at 550 °C for 30 min. Impedance spectra and dielectric properties were measured using an impedance analyzer (Keysight E4990A, Santa Rosa, CA, USA) at a heating rate of 3 °C min−1 over a temperature range of 30–200 °C and at various frequencies. The ferroelectric properties were measured using unipolar triangular waveform at 1 Hz and under varying electric fields using a Premier II ferroelectric tester (Radiant Technologies Inc., New Mexico, NW, USA). The green pellets were polished to a appropriately 0.15 mm thickness prior to testing. Au electrodes with a 3.14 mm2 area were sputtered.

3. Results and Discussion

The phase structure of xCLMN samples was analyzed via XRD, and the results are presented in Figure 1. The XRD patterns of the samples exhibited a pure perovskite structure. The absence of secondary phases indicated that Ca, La, Mg, and Nb ions were successfully incorporated into the BT-NBT lattice, forming a solid solution. With increasing CLMN doping concentration, the symmetry of the (200) diffraction peak at the 2θ value of 45° increased (Figure 1b), suggesting a transition from the tetragonal phase (T, P4mm) to the pseudo-cubic phase (PC, Pm-3m) [30]. The presence of the Ka2 reflection, often mistaken for non-cubic distortion, is a key observation [31]. This reflection is typically difficult to resolve in tetragonal or rhombohedral symmetries. The distinct appearance of the (200) peak at low diffraction angles (<50°) indicates that non-cubic distortions were negligible within the detection limits of the X-ray apparatus. Further, with increasing CLMN content, the peaks shifted toward higher 2θ values, suggesting a reduction in the lattice constant and unit cell volume. This effect likely arises because the ionic radii of the dopants (Ca2+, Mg2+, Nb5+, and La3+) are smaller than those of the original A-site and B-site ions [23,30].
The structure and phase evolution of xCLMN samples were analyzed via Raman spectroscopy (Figure 1c). BT-based ferroelectric ceramics exhibit four main vibrational modes [30,31,32,33], which were also observed in our sample spectra: (i) A-site cations vibrations (<200 cm−1), (ii) B–O bond vibrations (200–400 cm−1), (iii) BO6 octahedra vibrations (400–650 cm−1), and (iv) A1 + E (>700 cm−1) overlapping bands. The Raman bands at ~179 and 120 cm−1 correspond to A–O vibrations, confirming the presence of Na+, Ba2+, Bi3+, La3+, and Ca2+. Furthermore, the bending vibrational mode (A1) of the Ti–O bond at ~300 cm−1 shifted to a lower wavenumber. With increasing CLMN concentration, the Raman peaks near 300 cm−1 became broader and more diffuse. Multiple cations occupy the A-site causing a mismatch in the valency, which induces a high local disorder and reduced unit cell polarity [34,35]. It can be induced that CLMN doping progressively replaces the Ti4+ ions, which reduces the material’s P [30]. The creation of local nanoregions is proved by the appearance of the Raman peaks at ~750 cm−1 [36]. Consequently, adding CLMN breaks down the long-range ferroelectric order, leading to the formation of PNRs [37]. This makes the material exhibit a high Pmax and a nearly negligible Pr during charging and discharging. The transformation of the Raman peaks from active modes to complex/diffused modes is similar to those of the XRD peaks. This indicates a phase transition from the tetragonal (ferroelectric) state to a pseudo-cubic (relaxor ferroelectric) state.
The thermally etched surface morphology of the ceramic samples was analyzed using SEM (Figure 2). The xCLMN ceramics exhibited a dense microstructure with well-defined grain boundaries and a few pores. Grain-size distributions were calculated using analytical software, and variations in grain size among the as-sintered samples were assessed (Figure 2a−1–f−1). The grain sizes of CLMN-doped ceramics decreased from 1.48 μm to 0.9 μm. This reduction is crucial for achieving a high Eb and is closely related to grain-boundary migration and the sintering temperature. Moreover, the ceramic grain size was highly sensitive to the CLMN content. EDS mapping of 0.10CLMN (Figure 2g) revealed compositional homogeneity, with all elements uniformly distributed across the observed region, showing no signs of segregation or aggregation.
The behavior of different polar components in a material under an external field with a low alternating current is indicated by the dielectric properties of the material. Based on this phenomenon, the relaxation properties and phase transition of ferroelectrics were investigated. The loss tangent (tan δ) and dielectric constant (ε) of xCLMN were measured as a function of temperature at various frequencies, and the results are shown in Figure 3. For ferroelectric materials, the sharp peaks representing dielectric components are independent of frequencies. However, an increase in the CLMN concentration gradually alters the sharp high ε-temperature peak (Tm) to a diffuse and broad peak. This phenomenon is attributed to frequency dispersion, highlighting the existence of pronounced relaxor characteristics [38]. Furthermore, Tm and the maximum dielectric constant (εm) decreased with increasing x amount. The reduction in εm is likely due to changes in the lattice constant and A-site ionic radii [39]. The presence of different cations at the A-site compresses the oxygen octahedra and reduces the mobility of B-site cations. Therefore, ε and P are lowered. The domain structure might also be involved. An increase in the x content diminishes the phase proportion of the T phase whereas that of the PC phase increases, forming a relaxor ferroelectric phase. The presence of these nanodomain structures might decrease εm [39]. Notably, Tm shifts to room temperature for samples with x ≥ 0.10.
The conductivity, dielectric behavior, and relaxation-type characteristics of xCLMN ceramic materials were investigated via impedance spectroscopy. The impedance spectra of the samples from 400 °C to 460 °C in the range of 20–106 Hz are presented in Figure 4. The samples exhibited a semicircular feature in the examined temperature range, likely due to grain boundaries [40]. The impedance data were fitted using a single parallel resistor–capacitor equivalent circuit to extract reliable resistance values for the grain boundary phases. Grain boundary conductivity displayed an evident decrease in impedance at high temperatures, indicating a thermally activated conduction behavior [40]. Plotting the composition-dependent impedance spectra at 445 °C [Figure 4f] reveals a dramatic increase in the semicircle radius as the CLMN concentration rises. This clearly indicates a corresponding increase in insulating resistance. In addition, the Arrhenius law was used to calculate the required activation energy (Ea) for electrical conductivity [Equations (4) and (5)] [17,40]:
σ = σ 0 exp E a k B T
σ = l R S
where kB denotes the Boltzmann constant, σ0 denotes a pre-exponential factor, and σ denotes electrical conductivity. S denotes the cross-sectional sample area, and R represents the real-part impedance (Z′) extrapolated intercept. T and l denote the measured temperature and sample thickness, respectively. Figure 4g shows the plot of ln ρ against 1000/T; The solid line is fit using Equation (4). The Ea of the xCLMN samples was ascertained by fitting the conductivity at high temperatures linearly (1.10–1.30 eV) [Figure 4h]. The B- and A-site values of cation transport are 14 and 4 eV, respectively, and the resulting Ea values of the samples are 1.11–1.25 eV, which is similar to Ea required for the migration of the oxygen vacancy (Ea < 2 eV) [41]. Therefore, Therefore, oxygen vacancy dictates the conductivity of the samples at high temperatures, which is in agreement with reported BT-based ceramics [19,41,42]. It is noted that when the CLMN content rises, the Ea of the 0.10CLMN ceramics peaks at 1.25 eV.
The ferroelectric performance of xCLMN ceramics (Figure 5) is exhibited by their ferroelectric hysteresis P–electric field (E) loops under the applied electric field with a maximum value. With increasing CLMN content, the P-E loop gradually shifted from one with a large Pr and high Ec to a more linear shape, indicating that the long-range ferroelectric order was gradually disrupted, leading to the formation of nanodomains or PNRs [23]. The PNRs or nanodomains are dependent on the applied electric field. Under an external electric field, they exhibit a long-range ferroelectric state, while in the absence of the field, they revert to the original ergodic state. This behavior is consistent with the dielectric measurement results shown in Figure 3. A lower Pr and moderate Pmax are obtained in 0.10CLMN ceramics, which has a peak value of Eb.
Figure 6a revealed the PE loops of xCLMN ceramic materials determined at room temperature under a critical electric field. With increasing the CLMN content, the ferroelectric hysteresis loop of the ceramic samples narrowed, accompanied by a gradual decrease in the polarization (P) value. The observed phenomenon was attributed to the presence of the PC phase, which is consistent with the findings derived from the dielectric and XRD analyses. The energy storage properties of xCLMN ceramic materials are presented in Figure 6b,c, showing the critical values Pmax, Pr, Wrec, and W of PE as a function of E. Pr demonstrated minimal variation with an increase in x, revealing the characteristic relaxor behavior. Upon removing the electric field, a transition occurred from a more coherent long-range ferroelectric ordering to PNRs. With an increase in electric field, the initially ordered short-range PNRs transition into long-range ferroelectric ordering, resulting in a continuous rise in Pmax. Thus, an increase in the external electric field led to a continuous increase in Wrec and W. The Eb values for 0.08CLMN, 0.10CLMN, 0.12CLMN, 0.14CLMN, and 0.16CLMN were determined to be 280, 340, 320, 300, and 290 kV cm−1, respectively. The Wrec value of the pure sample demonstrated an initial increase from 0.65 J cm−3 to a maximum of 3.40 J cm−3 at x = 0.10. However, further doping leads to a decline in Wrec. The 0.10CLMN composition demonstrated the best energy storage performance, achieving a high value of Wrec (3.40 J cm−3) and η (81%). Therefore, the addition of CLMN can substantially enhance the W of BT-NBT-based ceramics.
Figure 6d presents a summary of key energy storage parameters at approximately 350 kV cm−1 for several lead-free dielectric ceramic materials, including bismuth ferrite (BiFeO3, BF)–BT-based [43,44,45], BT-based [23,30,39], potassium sodium niobate [(K, Na)NbO3, KNN]-based [46,47,48], and NBT-based [49,50,51,52,53] ceramics. High ƞ and Wrec of 0.10CLMN ceramic are achieved simultaneously, indicating it as a potential component for the application in energy storage devices. The energy storage properties of 0.10CLMN are comparable with most reported studies, with wide application prospects in dielectric capacitors. Table 1 presents a comparison of the energy storage characteristics of 0.10CLMN ceramic with those of other reported lead-free materials. The energy storage properties of 0.10CLMN specimen were found to be analogous to those of other reported Pb-free ceramics.
While dielectric capacitors demonstrated high W and η, maintaining stable energy storage performance is essential for their reliable use in electrostatic capacitor applications. Moreover, the capacitors must also demonstrate consistent reliability. Since 0.10CLMN demonstrated high η and Wrec, its PE loops with an E of 160 kV cm−1 were selected for an in-depth analysis of energy storage performance under different temperature and frequency conditions (Figure 7a,c). The values of Pmax, Pr, Wrec, and ƞ of 0.10CLMN, calculated using the methods described previously [8] are presented in Figure 7b,d. Remarkably, the P–E loops of 0.10CLMN demonstrated a narrow shape. According to Figure 7b, the Pmax and Pr demonstrated a slight decrease with rising temperature, whereas the values of Wrec and ƞ ranged from 1.46–1.50 J cm−3 and 89.28–90.17%, respectively. The observed fluctuations for these parameters (ΔWrec < 2.67% and Δη < 0.82%) are small within this temperature range. Thus, the energy storage performance of 0.10CLMN might possess high temperature stability. Figure 7d shows that the fluctuation of Pmax and Pr at 0.5–300 Hz is relatively small. Thus, Wrec remains relatively stable around 1.50 J cm−3, with fluctuations not exceeding 2.63% across the frequency range of 0.5 to 300 Hz. The η of 0.10CLMN remained above 90% within this frequency range, demonstrating its potential for use across a broad spectrum of frequencies. These findings highlight the favorable energy storage performance of 0.10CLMN, characterized by its strong stability across both temperature and frequency variation.

4. Conclusions

By doping CLMN, a linear dielectric material, into BT–NBT–based ceramic materials, a substantial improvement in energy storage capability was achieved. The performed modification enhanced the insulation performance and altered the relaxation behavior, optimizing the overall energy storage properties of the ceramics. These modifications significantly enhanced the breakdown strength while reducing hysteresis loss, leading to a synergistic increase in both Wrec and η. Wrec of 3.40 J cm−3 and ƞ of 81% was achieved in the 0.10CLMN. The energy storage parameters demonstrated high thermal and frequency stability, with Wrec fluctuating by less than 2.67% and 2.63% over the 30–90 °C and 0.5–300 Hz range, respectively. Thus, 0.10CLMN is a promising material for pulsed-power applications and demonstrates a linear additive approach for developing ceramic materials with high energy storage performance.

Author Contributions

Methodology, J.D. and X.W.; Formal analysis, P.R.; Investigation, J.D., J.G. and X.W.; Data curation, J.Z.; Writing—original draft, X.Z. and P.R.; Writing—review & editing, T.W., S.V.P.V. and Q.Z.; Supervision, T.W. and Q.Z.; Project administration, T.W., X.Z. and Q.M.; Funding acquisition, T.W., Q.M. and Q.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Guangdong Basic and Applied Basic Research Foundation (Grant No. 2022A1515140002 and 2023A1515140074), Young Innovative Talents Project of General Colleges and Universities in Guangdong Province (2019GKQNCX127), Project for Independent Innovation Capability Improvement of Huizhou University (Grant No. HZU202505), Open Project Program of Hubei Key Laboratory of Micro-Nanoelectronic Materials and Devices (Grant No. K202503), and Youth Science and Technology Talent Cultivation Program of Guangdong Association for Science and Technology (Grant No. SKXRC2025215).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author(s).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, T.; Mallikarjuna, K.; Vattikuti, S.V.P.; Altaf, M.; Goud, B.S.; Koyyada, G.; Shim, J. Synergistic enhancement of electrochemical storage using g- C3N4 modified MoO2/MoO3 nanostructure electrodes via thermal decomposition. J. Energy Storage 2024, 96, 112716. [Google Scholar] [CrossRef]
  2. Chen, L.; Hu, T.F.; Shi, X.M.; Yu, H.F.; Zhang, H.; Wu, J.; Fu, Z.Q.; Qi, H.; Chen, J. Near-zero energy consumption capacitors by controlling inhomogeneous polarization configuration. Adv. Mater. 2024, 36, 10. [Google Scholar] [CrossRef]
  3. Cui, C.H.; Bai, F.; Yang, Y.N.; Hou, Z.Q.; Sun, Z.; Zhang, T. Ion-exchange-induced phase transition enables an intrinsically air sable hydrogarnet electrolyte for solid-state lithium batteries. Adv. Sci. 2024, 11, 11. [Google Scholar] [CrossRef]
  4. Duan, J.H.; Wei, K.; Du, Q.B.; Ma, L.Z.; Yu, H.F.; Qi, H.; Tan, Y.C.; Zhong, G.K.; Li, H. High-entropy superparaelectrics with locally diverse ferroic distortion for high-capacitive energy storage. Nat. Commun. 2024, 15, 8. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, B.; Liu, Y.; Jiang, R.J.; Lan, S.; Liu, S.Z.; Zhou, Z.; Dou, L.; Zhang, M.; Huang, H.; Chen, L.Q.; et al. Enhanced energy storage in antiferroelectrics via antipolar frustration. Nature 2025, 637, 1104–1110. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, M.; Lan, S.; Yang, B.B.; Pan, H.; Liu, Y.Q.; Zhang, Q.H.; Qi, J.L.; Chen, D.; Su, H.; Yi, D.; et al. Ultrahigh energy storage in high-entropy ceramic capacitors with polymorphic relaxor phase. Science 2024, 384, 185–189. [Google Scholar] [CrossRef] [PubMed]
  7. Yang, B.; Zhang, Y.; Pan, H.; Si, W.; Zhang, Q.; Shen, Z.; Yu, Y.; Lan, S.; Meng, F.; Liu, Y.; et al. High-entropy enhanced capacitive energy storage. Nat. Mater. 2022, 21, 1074–1080. [Google Scholar] [CrossRef]
  8. Chen, S.; Wang, T.; Wang, X.L.; Li, K.; Zhu, Q.F.; Gong, W.P.; Liu, G.; Wang, Q.Y.; Xie, S.X. Structural origin of enhanced storage energy performance and robust mechanical property in A-site disordered high-entropy ceramics. Rare Met. 2025, 44, 551–564. [Google Scholar] [CrossRef]
  9. Chen, L.; Deng, S.; Liu, H.; Wu, J.; Qi, H.; Chen, J. Giant energy-storage density with ultrahigh efficiency in lead-free relaxors via high-entropy design. Nat. Commun. 2022, 13, 3089. [Google Scholar] [CrossRef]
  10. Zhong, W.T.; Liu, X.Y.; Zheng, X.T.; Zheng, P.; Wang, J.Q.; Sheng, L.S.; Zheng, L.; Fan, Q.L.; Bai, W.F.; Zhang, Y. Realizing exceptional energy storage performance in tungsten bronze-based ceramics via weakly coupled relaxor and grain boundary reinforcement designs. ACS Appl. Mater. Interfaces 2025, 17, 12375–12383. [Google Scholar] [CrossRef]
  11. Zhang, L.; Pu, Y.; Chen, M.; Peng, X.; Wang, B.; Shang, J. Design strategies of perovskite energy-storage dielectrics for next-generation capacitors. J. Eur. Ceram. Soc. 2023, 43, 5713–5747. [Google Scholar] [CrossRef]
  12. Tang, T.; Liu, J.C.; Liu, D.; Han, Y.; Luan, R.D.; Wang, Q.; Liu, H.; Zhang, B.P.; Zheng, Q.; Deng, S.Q.; et al. Self-generated glass-ceramics-like structure boosts energy storage performance of AgNbO3-based MLCC. Adv. Funct. Mater. 2025, 35, 2425711. [Google Scholar] [CrossRef]
  13. Tang, T.; Liu, D.; Wang, L.; Li, J.Z.; Zhang, Z.; Zhao, L.; Zhang, B.P.; Zhu, L.F. Ultrahigh energy storage density and efficiency of antiferroelectric AgNbO3-based MLCCs via reducing the off-center cations displacement. Chem. Eng. J. 2025, 503, 158557. [Google Scholar] [CrossRef]
  14. Zhou, J.; Liu, D.K.; Chen, R.X.; Zhang, K.; Jin, R.Q.; Sun, H.C.; Feng, Y.J.; Wei, X.Y.; Xu, Z.; Xu, R. Enhanced ultra-high efficiency in high-energy-density PbHfO3-based antiferroelectric ceramics through synergistic effect design. Chem. Eng. J. 2024, 496, 154369. [Google Scholar] [CrossRef]
  15. Wu, L.W.; Cai, Z.M.; Zhu, C.Q.; Feng, P.Z.; Li, L.T.; Wang, X.H. Significantly enhanced dielectric breakdown strength of ferroelectric energy-storage ceramics via grain size uniformity control: Phase-field simulation and experimental realization. Appl. Phys. Lett. 2020, 117, 6. [Google Scholar] [CrossRef]
  16. Wang, W.; Zhang, L.; Yang, Y.; Shi, W.; Huang, Y.; Alikin, D.O.; Shur, V.Y.; Lou, Z.; Zhang, A.; Wei, X.; et al. Enhancing energy storage performance in Na0.5Bi0.5TiO3-based lead-free relaxor ferroelectric ceramics along a stepwise optimization route. J. Mater. Chem. A 2023, 11, 2641–2651. [Google Scholar] [CrossRef]
  17. Li, D.; Xu, D.; Zhao, W.; Avdeev, M.; Jing, H.; Guo, Y.; Zhou, T.; Liu, W.; Wang, D.; Zhou, D. A high-temperature performing and near-zero energy loss lead-free ceramic capacitor. Energy Environ. Sci. 2023, 16, 4511–4521. [Google Scholar] [CrossRef]
  18. Cao, W.; Lin, R.; Hou, X.; Li, L.; Li, F.; Bo, D.; Ge, B.; Song, D.; Zhang, J.; Cheng, Z. Interfacial polarization restriction for ultrahigh energy storage density in lead free ceramics. Adv. Funct. Mater. 2023, 33, 2301027. [Google Scholar] [CrossRef]
  19. Chen, L.; Li, F.; Gao, B.; Zhou, C.; Wu, J.; Deng, S.; Liu, H.; Qi, H.; Chen, J. Excellent energy storage and mechanical performance in hetero-structure BaTiO3-based relaxors. Chem. Eng. J. 2023, 452, 139222. [Google Scholar] [CrossRef]
  20. Lv, J.; Li, Q.; Li, Y.; Tang, M.; Jin, D.; Yan, Y.; Fan, B.; Jin, L.; Liu, G. Significantly improved energy storage performance of NBT-BT based ceramics through domain control and preparation optimization. Chem. Eng. J. 2021, 420, 129900. [Google Scholar] [CrossRef]
  21. Yuan, Q.B.; Li, G.; Yao, F.Z.; Cheng, S.D.; Wang, Y.F.; 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]
  22. Zhu, W.; Shen, Z.Y.; Deng, W.; Li, K.; Luo, W.Q.; Song, F.S.; Zeng, X.J.; Wang, Z.M.; Li, Y.M. A review: (Bi,Na)TiO3 (BNT)-based energy storage ceramics. J. Mater. 2024, 10, 86–123. [Google Scholar] [CrossRef]
  23. 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]
  24. Qiu, Y.; Lin, Y.; Liu, X.Y.; Yang, H.B. Bi(Mg2/3Nb1/3)O3 addition inducing high recoverable energy storage density in lead-free 0.65BaTiO3-0.35Bi0.5Na0.5TiO3 bulk ceramics. J. Alloys Compd. 2019, 797, 348–355. [Google Scholar] [CrossRef]
  25. Dai, Z.; Xie, J.; Fan, X.; Ding, X.; Liu, W.; Zhou, S.; Ren, X. Enhanced energy storage properties and stability of Sr(Sc0.5Nb0.5)O3 modified 0.65BaTiO3-0.35Bi0.5Na0.5TiO3 ceramics. Chem. Eng. J. 2020, 397, 125520. [Google Scholar] [CrossRef]
  26. Xie, A.; Fu, J.; Zuo, R.; Zhou, C.; Qiao, Z.; Li, T.; Zhang, S. NaNbO3-CaTiO3 lead-free relaxor antiferroelectric ceramics featuring giant energy density, high energy efficiency and power density. Chem. Eng. J. 2022, 429, 132534. [Google Scholar] [CrossRef]
  27. Ye, W.B.; Zhu, C.H.; Xiao, Y.M.; Bai, X.Z.; Zheng, P.; Zhang, J.J.; Bai, W.F.; Fan, Q.L.; Zheng, L.; Zhang, Y. Remarkable energy-storage performances and excellent stability in CaTiO3-doped BiFeO3-BaTiO3 relaxor ferroelectric ceramics. J. Eur. Ceram. Soc. 2023, 43, 900–908. [Google Scholar] [CrossRef]
  28. Yang, F.; Pan, Z.; Ling, Z.; Hu, D.; Ding, J.; Li, P.; Liu, J.; Zhai, J. Realizing high comprehensive energy storage performances of BNT-based ceramics for application in pulse power capacitors. J. Eur. Ceram. Soc. 2021, 41, 2548–2558. [Google Scholar] [CrossRef]
  29. Wang, T.; Li, Y.; Zhang, X.; Zhang, D.; Gong, W. Simultaneous excellent energy storage density and efficiency under applied low electric field for high entropy relaxor ferroelectric ceramics. Mater. Res. Bull. 2023, 157, 112024. [Google Scholar] [CrossRef]
  30. Chen, X.; Wang, M.; Pan, Z.; Li, H.; Zhao, J.; Tang, L.; Liu, J.; Li, P.; Xie, H.; Zhai, J. Ultrahigh energy density and efficiency of BaTiO3-based ceramics via multiple design strategies. Chem. Eng. J. 2023, 467, 143395. [Google Scholar] [CrossRef]
  31. 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]
  32. Minakshi, M.; Samayamanthry, A.; Whale, J.; Aughterson, R.; Shinde, P.A.; Ariga, K.; Kumar Shrestha, L. Phosphorous-containing activated carbon derived from natural honeydew peel powers aqueous supercapacitors. Chem. Asian J. 2024, 19, e202400622. [Google Scholar] [CrossRef]
  33. Wickramaarachchi, K.; Sundaram, M.M.; Henry, D.J.; Gao, X. Alginate biopolymer effect on the electrodeposition of manganese dioxide on electrodes for supercapacitors. ACS Appl. Energy Mater. 2021, 4, 7040–7051. [Google Scholar] [CrossRef]
  34. Xie, A.; Zuo, R.; Qiao, Z.; Fu, Z.; Hu, T.; Fei, L. NaNbO3-(Bi0.5Li0.5)TiO3 lead-free relaxor ferroelectric capacitors with superior energy-storage performances via multiple synergistic design. Adv. Energy Mater. 2021, 11, 2101378. [Google Scholar] [CrossRef]
  35. Luo, N.N.; Han, K.; Cabral, M.J.; Liao, X.Z.; Zhang, S.J.; Liao, C.Z.; Zhang, G.Z.; Chen, X.Y.; Feng, Q.; Li, J.F.; et al. Constructing phase boundary in AgNbO3 antiferroelectrics: Pathway simultaneously achieving high energy density and efficiency. Nat. Commun. 2020, 11, 10. [Google Scholar] [CrossRef]
  36. Chen, X.L.; Li, X.; Sun, J.; Sun, C.C.; Shi, J.P.; Pang, F.H.; Zhou, H.F. Achieving ultrahigh energy storage density and energy efficiency simultaneously in barium titanate based ceramics. Appl. Phys. A-Mater. Sci. Process. 2020, 126, 8. [Google Scholar] [CrossRef]
  37. Yan, F.; Huang, K.W.; Jiang, T.; Zhou, X.F.; Shi, Y.J.; Ge, G.L.; Shen, B.; Zhai, J.W. Significantly enhanced energy storage density and efficiency of BNT-based perovskite ceramics via A-site defect engineering. Energy Storage Mater. 2020, 30, 392–400. [Google Scholar] [CrossRef]
  38. Zhao, J.; Pan, Z.; Tang, L.; Shen, Y.; Chen, X.; Li, H.; Li, P.; Zhang, Y.; Liu, J.; Zhai, J. Greatly enhanced discharged energy density and efficiency of BiFeO3-Based ceramics by regulating insulation performance. Mater. Today Phys. 2022, 27, 100821. [Google Scholar] [CrossRef]
  39. Liu, Z.G.; Li, M.D.; Tang, Z.H.; Tang, X.-G. Enhanced energy storage density and efficiency in lead-free Bi(Mg1/2Hf1/2)O3-modified BaTiO3 ceramics. Chem. Eng. J. 2021, 418, 129379. [Google Scholar] [CrossRef]
  40. Guan, Z.N.; Yan, Y.; Ma, J.; Pan, T.; Li, X.; Guo, S.; Zhang, J.; Wang, J.; Wang, Y. Significantly enhanced energy storage performance of lead-free BiFeO3-based ceramics via synergic optimization strategy. ACS Appl Mater Interfaces 2022, 14, 44539–44549. [Google Scholar] [CrossRef] [PubMed]
  41. Deng, X.Z.; Zhang, L.Y.; Geng, X.Y.; Zhang, J.; Sun, L.; Wang, R.-X.; Gu, Z.-B.; Zhang, S.-T. Crystal structure, impedance, and multiferroic property of SrZrO3 and MnO2 modified 0.725BiFeO3-0.275BaTiO3 ceramics. Ceram. Int. 2017, 43, 14748–14755. [Google Scholar] [CrossRef]
  42. Weng, N.; Zhang, J.; Wang, Z.Y.; Wang, H.; Wang, L.; Wang, J.; Wang, Y.J. Moderate electric field driven ultrahigh energy density in BiFeO3-BaTiO3-based ceramics with improved relaxor behavior and breakdown strength. Chem. Eng. J. 2024, 485, 149947. [Google Scholar] [CrossRef]
  43. Zhao, J.; Li, H.; Du, Y.; Chen, X.; Qin, H.; Wang, J.; Yan, T.; Yu, S.; Hu, Y.; Wang, D. Superior energy storage performance of BiFeO3-BaTiO3-CaHfO3 lead-free ceramics. J. Mater. Chem. A 2024, 12, 5261–5268. [Google Scholar] [CrossRef]
  44. Montecillo, R.; Chien, R.R.; Chen, C.S.; Wu, P.H.; Tu, C.S.; Feng, K.C. Ultrahigh energy storage in multilayer BiFeO3-BaTiO3-NaTaO3 relaxor ferroelectric ceramics. J. Mater. Chem. A 2024, 12, 30642–30654. [Google Scholar] [CrossRef]
  45. Zhao, J.; Hu, T.; Fu, Z.; Pan, Z.; Tang, L.; Chen, X.; Li, H.; Hu, J.; Lv, L.; Zhou, Z.; et al. Delayed polarization saturation induced superior energy storage capability of BiFeO3-based ceramics via introduction of non-isovalent ions. Small 2023, 19, 2206840. [Google Scholar] [CrossRef] [PubMed]
  46. Huan, Y.; Wu, L.Z.; Xu, L.Y.; Li, P.; Wei, T. Superior energy-storage density and ultrahigh efficiency in KNN-based ferroelectric ceramics via high-entropy design. J. Mater. 2025, 11, 100862. [Google Scholar] [CrossRef]
  47. Zha, J.L.; Yang, Y.L.; Liu, J.X.; Lu, X.M.; Hu, X.L.; Yan, S.; Wu, Z.J.; Zhou, M.; Huang, F.Z.; Ying, X.N.; et al. High energy storage performance of KNN-based relaxor ferroelectrics in multiphase-coexisted superparaelectric state. J. Appl. Phys. 2024, 136, 10. [Google Scholar] [CrossRef]
  48. Sun, Z.; Zhao, S.; Wang, T.; Jing, H.; Guo, Q.; Gao, R.; Diwu, L.; Du, K.; Hu, Y.; Pu, Y. Achieving high overall energy storage performance of KNN-based transparent ceramics by ingenious multiscale designing. J. Mater. Chem. A 2024, 12, 16735–16747. [Google Scholar] [CrossRef]
  49. Mao, P.; Guo, Y.; Lu, G.; Yan, Q.; Kang, R.; Wang, T.; Xie, B.; Liu, Z.; Zhang, L. Synergistic effect of multi-phase and multi-domain structures induced high energy storage performances under low electric fields in Na0.5Bi0.5TiO3-based lead-free ceramics. Chem. Eng. J. 2023, 472, 144973. [Google Scholar] [CrossRef]
  50. Zhu, X.; Gao, Y.; Shi, P.; Kang, R.; Kang, F.; Qiao, W.; Zhao, J.; Wang, Z.; Yuan, Y.; Lou, X. Ultrahigh energy storage density in (Bi0.5Na0.5)0.65Sr0.35TiO3-based lead-free relaxor ceramics with excellent temperature stability. Nano Energy 2022, 98, 107276. [Google Scholar] [CrossRef]
  51. Liu, X.; Hou, Y.; Song, B.; Cheng, H.; Fu, Y.; Zheng, M.; Zhu, M. Lead-free multilayer ceramic capacitors with ultra-wide temperature dielectric stability based on multifaceted modification. J. Eur. Ceram. Soc. 2022, 42, 973–980. [Google Scholar] [CrossRef]
  52. Zhao, X.; Zhang, L.; Fan, Z.; Huang, Y.; Hu, Y.; Shen, M.; Wang, Z.; He, Y.; Wang, D.; Zhang, Q. Excellent high-temperature dielectric energy storage performance in bilayer nanocomposites with high-entropy ferroelectric oxide fillers. Nat. Commun. 2025, 16, 5570. [Google Scholar] [CrossRef] [PubMed]
  53. Fan, Z.; Dai, J.; Huang, Y.; Xie, H.; Jiao, Y.; Yue, W.; Huang, F.; Deng, Y.; Wang, D.; Zhang, Q.; et al. Superior energy storage capacity of polymer-based bilayer composites by introducing 2D ferroelectric micro-sheets. Nat. Commun. 2025, 16, 1180. [Google Scholar] [CrossRef] [PubMed]
  54. Dang, S.; Peng, Z.; Zhang, X.; Wang, Y.; Chai, Q.; Wu, D.; Liang, P.; Wei, L.; Chao, X.; Yang, Z. Enhanced energy storage performance of BiFeO3-BaTiO3 based ceramics under moderate electric fields via multiple synergistic design. Chem. Eng. J. 2025, 512, 162494. [Google Scholar] [CrossRef]
  55. Zhang, J.B.; Pu, Y.P.; Hao, Y.X.; Yang, Y.L.; Zhang, L.; Wang, B.; Pan, Q. Realizing excellent energy-storage performance under low electric fields in lead-free BiFeO3-BaTiO3-based ceramics with ultrahigh polarization difference. J. Energy Storage 2025, 105, 114786. [Google Scholar] [CrossRef]
  56. Kong, X.; Yang, L.T.; Meng, F.Q.; Zhang, T.; Zhang, H.J.; Lin, Y.H.; Huang, H.B.; Zhang, S.J.; Guo, J.M.; Nan, C.W. High-entropy engineered BaTiO3-based ceramic capacitors with greatly enhanced high-temperature energy storage performance. Nat. Commun. 2025, 16, 9. [Google Scholar] [CrossRef]
  57. Zheng, B.; Yuan, Q.; Lin, Y.; Li, D.; Yang, H.; Hong, Z.; Ma, Y.; Ma, Y.; Guo, J.; Wang, J. High entropy-driven large capacitive energy storage in BaTiO3-based multilayer ceramic capacitors. Adv. Energy Mater. 2025, 16, e04126. [Google Scholar] [CrossRef]
  58. Jiang, Y.; Liu, J.M.; Zhang, W.C.; Cheng, X.; Hui, K.Z.; Zhen, Y.C.; Hao, Y.N.; Bi, K.; Guo, L.M.; Zhao, P.Y.; et al. Comprehensively improved energy storage and DC-bias properties in Bi0.5Na0.5TiO3-NaNbO3 based relaxor antiferroelectric. J. Mater. 2025, 11, 8. [Google Scholar]
Figure 1. (a) Room-temperature XRD patterns of as-prepared ceramics; (b) the corresponding magnification around 2θ ∼ 45°; (c) Raman spectra of the xCLMN ceramics.
Figure 1. (a) Room-temperature XRD patterns of as-prepared ceramics; (b) the corresponding magnification around 2θ ∼ 45°; (c) Raman spectra of the xCLMN ceramics.
Nanomaterials 15 01724 g001
Figure 2. (af) SEM micrographs and grain size for ceramics sintered at optimal sintering temperature; (g) SEM-EDS images of 0.1CLMN ceramic.
Figure 2. (af) SEM micrographs and grain size for ceramics sintered at optimal sintering temperature; (g) SEM-EDS images of 0.1CLMN ceramic.
Nanomaterials 15 01724 g002
Figure 3. Temperature-dependent dielectric constant and tanδ for all ceramics at different frequency: (a) x = 0; (b) x = 0.08; (c) x = 0.10; (d) x = 0.12; (e) x = 0.14; (f) x = 0.16.
Figure 3. Temperature-dependent dielectric constant and tanδ for all ceramics at different frequency: (a) x = 0; (b) x = 0.08; (c) x = 0.10; (d) x = 0.12; (e) x = 0.14; (f) x = 0.16.
Nanomaterials 15 01724 g003
Figure 4. Impedance spectra of xCLMN ceramics measured at various temperatures: (a) x = 0.08; (b) x = 0.10; (c) x = 0.12; (d) x = 0.14; (e) x = 0.16; (f) composition-dependent impedance spectra of as-prepared ceramics at 445 °C; (g) Arrhenius-type plots of resistivity vs. 1000/T of as-prepared ceramics; (h) variation of Ea value with doping content).
Figure 4. Impedance spectra of xCLMN ceramics measured at various temperatures: (a) x = 0.08; (b) x = 0.10; (c) x = 0.12; (d) x = 0.14; (e) x = 0.16; (f) composition-dependent impedance spectra of as-prepared ceramics at 445 °C; (g) Arrhenius-type plots of resistivity vs. 1000/T of as-prepared ceramics; (h) variation of Ea value with doping content).
Nanomaterials 15 01724 g004
Figure 5. Unipolar P-E loops under different applied electric fields for xCLMN ceramics: (a) x = 0; (b) x = 0.08; (c) x = 0.10; (d) x = 0.12; (e) x = 0.14; (f) x = 0.16. (the pink line in (c): the current-electric field curve).
Figure 5. Unipolar P-E loops under different applied electric fields for xCLMN ceramics: (a) x = 0; (b) x = 0.08; (c) x = 0.10; (d) x = 0.12; (e) x = 0.14; (f) x = 0.16. (the pink line in (c): the current-electric field curve).
Nanomaterials 15 01724 g005
Figure 6. (a) Unipolar P-E loops prior to their breakdown electric fields; (b) Pmax, Pr and (c) W, Wrec change with E for xCLMN ceramics; (d) A comparison of Wrec and η in this study and reported in other lead-free bulk ceramics.
Figure 6. (a) Unipolar P-E loops prior to their breakdown electric fields; (b) Pmax, Pr and (c) W, Wrec change with E for xCLMN ceramics; (d) A comparison of Wrec and η in this study and reported in other lead-free bulk ceramics.
Nanomaterials 15 01724 g006
Figure 7. (a) Unipolar P-E curves (30–100 °C at 50 Hz at 160 kV/cm); (c) Unipolar P-E curves (50–300 Hz at room temperature at 160 kV/cm); (b,d) show the energy storage temperature, and frequency stability of the 0.10CLMN ceramic at 160 kV/cm.
Figure 7. (a) Unipolar P-E curves (30–100 °C at 50 Hz at 160 kV/cm); (c) Unipolar P-E curves (50–300 Hz at room temperature at 160 kV/cm); (b,d) show the energy storage temperature, and frequency stability of the 0.10CLMN ceramic at 160 kV/cm.
Nanomaterials 15 01724 g007
Table 1. Comparison of energy storage properties of 0.10CLMN ceramic with other ceramic capacitor systems.
Table 1. Comparison of energy storage properties of 0.10CLMN ceramic with other ceramic capacitor systems.
CompositionWrec (J cm−3)η(%)Eb (kV cm−1)Refs.
BT–NBT–SBMT7.1290720[30]
BT–NBT–CZ9.0487.2540[19]
BT–BMH3.3887240[39]
BT–NBT–BZMASZ3.7482.2273[23]
BT–NBT–BMN1.690.3182[24]
BF–BT–NNT6.181.4330[54]
BF–BT–LA5.7180.19270[55]
BCT–BMZ10.993720[56]
BNBSCTZZTN7.390.6530[57]
KNN–STZ–BZTN11.1487.1750[46]
NBT–NN8.0485630[58]
BT–NBT–CLMN3.4081340This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Deng, J.; Guo, J.; Wang, T.; Zhang, J.; Wu, X.; Zhang, X.; Vattikuti, S.V.P.; Ma, Q.; Rosaiah, P.; Zhang, Q. BaTiO3–(Na0.5Bi0.5)TiO3 Ceramic Materials Prepared via Multiple Design Strategies with Improved Energy Storage. Nanomaterials 2025, 15, 1724. https://doi.org/10.3390/nano15221724

AMA Style

Deng J, Guo J, Wang T, Zhang J, Wu X, Zhang X, Vattikuti SVP, Ma Q, Rosaiah P, Zhang Q. BaTiO3–(Na0.5Bi0.5)TiO3 Ceramic Materials Prepared via Multiple Design Strategies with Improved Energy Storage. Nanomaterials. 2025; 15(22):1724. https://doi.org/10.3390/nano15221724

Chicago/Turabian Style

Deng, Jianming, Jingjing Guo, Ting Wang, Jingxiang Zhang, Xu Wu, Xuefeng Zhang, Surya Veerendra Prabhakar Vattikuti, Qing Ma, Pitcheri Rosaiah, and Qingfeng Zhang. 2025. "BaTiO3–(Na0.5Bi0.5)TiO3 Ceramic Materials Prepared via Multiple Design Strategies with Improved Energy Storage" Nanomaterials 15, no. 22: 1724. https://doi.org/10.3390/nano15221724

APA Style

Deng, J., Guo, J., Wang, T., Zhang, J., Wu, X., Zhang, X., Vattikuti, S. V. P., Ma, Q., Rosaiah, P., & Zhang, Q. (2025). BaTiO3–(Na0.5Bi0.5)TiO3 Ceramic Materials Prepared via Multiple Design Strategies with Improved Energy Storage. Nanomaterials, 15(22), 1724. https://doi.org/10.3390/nano15221724

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