Synergistically enhanced energy storage performance of Bi0.47Na0.47Ba0.06TiO3-based relaxor ferroelectrics via dual engineering of dynamic nanodomains and defect regulation

Jiangping Huang, Liang Deng*, Yu Zhang, Yue Pan, Xiuli Chen*, Xu Li* and Huanfu Zhou
Key Laboratory of New Processing Technology for Nonferrous Metal and Materials, Ministry of Education, Guangxi Key Laboratory of Optical and Electronic Materials and Devices, College of Materials Science and Engineering, Guilin University of Technology, Guilin, 541004, China. E-mail: 809188659@qq.com; cxlnwpu@163.com; lx100527@163.com

Received 15th June 2025 , Accepted 30th July 2025

First published on 19th August 2025


Abstract

Ceramic capacitors are essential components in advanced electronics known for their high-power density and outstanding operational stability. However, simultaneously achieving a large reversible polarization response and high breakdown strength remains a critical challenge for developing next-generation energy storage devices. Herein, we report a defect-engineered Bi0.47Na0.47Ba0.06TiO3-based relaxor ferroelectric ceramic that realizes a good balance between these competing parameters. Through strategically constructing heterogeneous relaxor phases with weakly coupling and precisely controlled defect concentrations, we achieve a remarkable recoverable energy density of 13.6 J cm−3 at 760.0 kV cm−1 with a high energy storage efficiency of 83.9%. The designed microstructure simultaneously enables strong field-induced polarization response while reducing hysteresis losses. Defect engineering enhances the resistivity and local electric field uniformity, leading to significantly improved dielectric breakdown strength. The optimized ceramic further demonstrates temperature/frequency-independent performance characteristics along with superior pulsed discharge capabilities. These findings provide a practical material design strategy for high-performance energy storage applications.


1. Introduction

Over the past few decades, with the massive consumption of fossil fuels and other non-renewable resources, as well as the ever-increasing environmental awareness, the development and use of renewable and environmentally friendly resources has become an urgent priority.1–3 Ceramic capacitors, known for their high-power density, fast charging and discharging capabilities, and excellent high-temperature stability, have attracted widespread attention and research in the field of energy storage.4 The availability of high-performance energy storage materials is key to the practical application of pulse power devices. Lead-based materials, known for their superior piezoelectric and dielectric properties, have been the most extensively studied ferroelectric materials.5,6 However, concerns about the environmental degradation and human health risks associated with lead-based materials, coupled with recent environmental legislation, have spurred significant research efforts to develop lead-free alternatives to replace lead-based materials. To meet the urgent demand for high-performance lead-free energy storage ceramics, which are essential components in advanced pulse power devices and high-power equipment, there is an urgent need to develop new types of energy storage ceramic capacitors with high energy storage density and high efficiency.7,8 Typically, the energy storage performance of ceramic capacitors can be described by the following formula:9–12
 
image file: d5ta04848f-t1.tif(1)
 
image file: d5ta04848f-t2.tif(2)
 
image file: d5ta04848f-t3.tif(3)
where Pmax is the maximum polarization, Pr is the remanent polarization, Wtotal represents the total energy storage density, Wrec denotes the recoverable energy storage density, η signifies the energy storage efficiency, and E denotes the applied electric field. Therefore, achieving a high polarization difference (ΔP) and a high breakdown electric field (Eb) are key factors in enhancing the energy storage performance.

Relaxor ferroelectrics, characterized by their slim hysteresis loops and high Pmax, have garnered significant attention in the field of energy storage materials.13 Bismuth sodium titanate (Bi0.5Na0.5TiO3, BNT), a lead-free ferroelectric ceramic with an ABO3 perovskite structure, exhibits superior ferroelectric properties and an easily manipulated phase structure.14 In particular, compared to other lead-free ceramics such as BaTiO3 (BT), AgNbO3 (AN), (K0.5Na0.5)NbO3 (KNN), NaNbO3 (NN), and SrTiO3 (ST), BNT with its large high polarization capacity positions it as one of the most promising lead-free candidate materials.15,16 However, pure BNT ceramic, because of its significant residual polarization and low breakdown strength, hinders its direct application in energy storage.17,18 Research indicates that BNT-based relaxor ferroelectric ceramics, modified by doping, exhibit high Pmax, low Pr, and high breakdown strength, making them suitable for advanced pulsed power applications. As a result, many optimization strategies have been proposed to improve the energy storage performance. For example, Long et al. conducted doping modifications on BNT, significantly reducing energy losses and enhancing polarization response through optimization of the grain size and domain structure, leading to a marked improvement in energy storage density. Ultimately, a high energy storage density of 8.63 J cm−3 was achieved, along with a high efficiency of 89.6%.19 Furthermore, the solid solution of Bi0.5Na0.5TiO3 with BaTiO3, which establishes a morphotropic phase boundary (MPB) where rhombohedral (space group R3c) and tetragonal (space group P4bm) phases coexist, can significantly enhance the polarization response and piezoelectric properties.20–22 Due to the possibility of extensive compositional tuning and rare earth doping modification, various superior BNT-BT-based ceramics can be obtained.23–25 Consequently, the (1 − x)BNT-xBT system has long been considered one of the important materials for the modification of energy storage ceramics. For instance, Cao et al. achieved an ultra-high energy storage density of 12.2 J cm−3 and a high η of 88.8% by incorporating Ca0.7Bi0.2(Sn0.5Ti0.5)O3 into the Bi0.5Na0.47Ba0.06O3 (BNBT) ceramic and optimizing the Eb through an interfacial polarization strategy.26 Research shows that doping modifications can regulate domain sizes, forming nanodomains and even refined polar nanoregions, which play a crucial role in improving the energy storage efficiency and achieving stability over a wide temperature range.19,27,28 For example, Dong et al. modulated the domain configuration by introducing Sr(Ni1/3Nb2/3)O3 as the second component in BNBT, ultimately achieving a Wrec of 7.0 J cm−3 and a η of 81.5%.29 Breakdown electric strength is a key factor in achieving high energy storage performance. In general, grain refinement, reduction of dielectric losses, reduction of volatilization of low melting point elements and optimization of the preparation process can significantly enhance the breakdown strength of ceramics.16 For example, Liu et al. obtained a high Wrec of 12.2 J cm−3 at 950 kV cm−1 through a grain boundary optimization strategy.30 Deng obtained an ultra-high Wrec of 12.39 J cm−3 and a η of 87.1% at 710 kV cm−1 via grain engineering in BNT-based relaxor ferroelectrics. It is shown that doping rare earth elements in BNT can enhance local heterogeneity, achieve domain structure modulation, and improve sintering, which in turn improves the energy storage performance.31–33 For instance, Tang et al. introduced Sm2O3 into BNBT, which significantly reduced the grain size and enhanced the relaxor characteristics, resulting in a Wrec of 4.41 J cm−3 and a η of 83.96%.34 Li et al. enhanced the energy storage performance of Bi0.5Na0.5TiO3 ceramic through SmFeO3 doping, achieving an ultrahigh Wrec of 9.05 J cm−3 by improving relaxor behavior and delaying polarization saturation.33 Although many studies have achieved breakthroughs, realizing ultra-high performance (Wrec >10 J cm−3) remains rare, and there is an urgent need to develop new high-performance lead-free energy storage materials. Furthermore, while many researchers have explored the evolution of domain structures in relaxor ferroelectric materials, a deeper investigation of the intrinsic relationship between micro-domain structures and macroscopic energy storage performance is still required.

To solve the problem of low energy storage density, we have made a breakthrough in achieving high energy storage performance in a novel (1 − x)(Bi0.5Na0.47Ba0.06O3)-xSm(Mg2/3Sb1/3)O3 (abbreviated as BNBT-xSMS) lead-free relaxor ferroelectric ceramics via the regulation of highly dynamic polar nanoregions and defects. Firstly, we selected Bi0.5Na0.47Ba0.06O3-based ceramics with high polarization as the matrix material for our study.35,36 The coexistence of the R3c (rhombohedral) and P4bm (tetragonal) phases at the morphotropic phase boundary (MPB) provides a foundation for constructing the polymorphic relaxor phase.37,38 However, the high Pr and low breakdown strength remain the primary limiting factors for improving the energy storage density in BNBT ceramics, necessitating further modification. Secondly, we introduced the rare earth-based complex Sm(Mg2/3Sb1/3)O3 into BNBT to form a solid solution and optimize the energy storage performance. On the one hand, Sm2O3 can optimize sintering, reduce the grain size and decrease the defect concentration, which improve the breakdown electric strength.33,39–41 On the other hand, the introduction of the B-site trivalent complex cation (Mg2/3Sb1/3)3+ effectively disrupts the long-range ferroelectric order of the matrix and induces more weakly polar P4bm phases, thereby constructing polar nanoregions (PNRs), which improve the domain response to the external electric field and consequently enhance the energy storage efficiency.42–44 Additionally, the high bandgap oxides of MgO (7.03 eV) and Sm2O3 (4.4 eV) further enhance the breakdown resistance of the ceramics.33 We have demonstrated through theoretical modeling and experimental characterization that the construction of highly dynamic polar nanodomains in BNBT-based ceramics provides the conditions for achieving high energy storage efficiency through rapid reversible switching under an external electric field. As expected, we achieve an ultrahigh Wrec of 13.6 J cm−3 with a high η of 84.9% in the BNBT-0.16SMS ceramic. In addition, the ceramic exhibits excellent temperature/frequency stability and superior charge–discharge performance. These findings provide new insights for the development of ceramics with high energy storage performance.

2. Experimental procedures

The relaxor ferroelectric ceramics (1 − x)Bi0.47Na0.47Ba0.06TiO3-xSm(Mg2/3Sb1/3)O3 (abbreviated as BNBT-xSMS with x = 0.08, 0.12, 0.16, 0.20) were prepared using traditional solid-state sintering reaction technology. The preparation methods and characterization studies of BNBT-xSMS bulk ceramics are available in the SI.

3. Results and discussion

All ceramic compositions were fabricated via conventional solid-state sintering. X-ray diffraction (XRD) analysis was conducted to characterize the crystal structure (Fig. 1a). As the doping content increased from 0.08 to 0.16, the ceramics exhibited pure perovskite structures. At x = 0.20, secondary phase peaks were detected, indicating that excessive SMS doping exceeded the solid solubility limit of the BNBT-based ceramics, leading to secondary phase formation. To further investigate the phase content of ceramics, Rietveld structural refinement was performed on representative compositions (x = 0.08 and x = 0.16) using the R3c (R) and P4bm (T) phases.29 As shown in Fig. 1b and c, the experimental data exhibited excellent agreement with the refined patterns, with low reliability factors (including the weighted profile residual factor (Rwp), pattern residual factor (Rp), and chi-square value (χ2) below 5%), confirming the robustness of the refinement results. At x = 0.08, the ceramic predominantly exhibited the R phase, while x = 0.16 showed a dominant T phase. This phase transition, driven by composition tuning, disrupts long-range ordered structures and induces a disordered local field, promoting the formation of polar nanoregions (PNRs).45 The resulting weakly polar T phase combined with a minor polar R phase not only maintains high polarization strength but also improves polarization response, thereby improving the energy storage efficiency through reduced hysteresis losses. To investigate the local structural evolution in BNBT-xSMS ceramics, room-temperature Raman spectroscopy was performed (Fig. 1d). The raw spectra exhibited four primary vibrational modes: A-site cation vibrations (100–200 cm−1), B–O stretching vibrations (180–420 cm−1), BO6 octahedral bending/tilting modes (420–670 cm−1), and A1 + E (longitudinal optical) symmetric modes (670–850 cm−1).46–49 After Gaussian–Lorentzian deconvolution (Fig. 1e and f), seven resolved peaks emerged. Progressive intensity attenuation and red shifts of ν4 and ν5 modes were observed, indicating B-site displacement induced by SMS doping. This structural disorder, arising from lattice distortion and phase fraction variation, facilitates the formation of nanodomains, thereby enhancing the energy storage performance.50 To elucidate the local structural evolution of BNBT-0.16SMS ceramic, transmission electron microscopy (TEM) was employed to analyze domain morphology, high-resolution lattice imaging, and selected area electron diffraction (SAED) patterns (Fig. 1). TEM observations revealed the presence of Moiré fringe-like nano-domain size (∼5 nm in size) with high dynamic activity,51,52 which facilitates rapid domain switching under electric fields, thereby minimizing energy loss (Fig. 1g). High-resolution TEM (HRTEM) imaging (Fig. 1h) revealed distinct crystalline fringes, confirming the excellent crystallinity of the sample. SAED patterns exhibited two distinct superlattice reflections: 1/2(ooe) along the (001) zone axis and 1/2(ooo) along the (110) zone axis, as shown in Fig. 1i and j, where “o” and “e” denote odd and even Miller indices, respectively.53 The appearance of these superlattice diffraction spots can be attributed to oxygen octahedral tilting and B-site cation displacement.54 Therefore, these superlattice spots provide direct evidence for the coexistence of the R-phase (trigonal symmetry) and T-phase (tetragonal symmetry), which further confirms the accuracy of the XRD refinement results.54 This local structural heterogeneity contributes to enhanced energy storage performance, as demonstrated by the increased recoverable energy density accompanied by reduced hysteresis loss.
image file: d5ta04848f-f1.tif
Fig. 1 (a) Room temperature XRD plots of BNBT-xSMS ceramics, (b and c) XRD refinement results of BNNBT-0.08SMS and BNBT-0.16SMS ceramics. (d) Room temperature Raman spectra of BNBT-xSMS ceramics and (e and f) the corresponding peak splitting fitting results. (g) TEM domain morphology of BNBT-0.16SMS ceramic. (h) HR-TEM image of BNBT-0.16SMS ceramic. (i and j) SAED patterns of BNBT-0.16SMS ceramic along the (001) and (110) directions, respectively.

The energy storage properties of BNBT-xSMS ceramics were characterized through unipolar PE hysteresis measurements at room temperature at 10 Hz and 350 kV cm−1 (Fig. S1a). With increasing SMS content, the loops progressively became slimmer, accompanied by reductions in both Pmax and Pr (Fig. S1b), signifying that SMS doping effectively disrupts long-range ferroelectric ordering while promoting the formation of PNRs, thereby enhancing relaxor characteristics.55 Although both Wtotal and Wrec decreased with increasing SMS content, a significant enhancement in η was observed. This apparent performance limitation likely results from the insufficient testing electric fields, as higher electric fields are expected to yield better results. The unipolar PE hysteresis loops of BNBT-xSMS ceramics tested at Eb is shown in Fig. 2a. Notably, higher Eb and Pmax values were achieved with increasing SMS content, clearly demonstrating composition-dependent optimization of energy storage performance. The composition with x = 0.12 exhibited superior Pmax compared to that with x = 0.08, owing to its enhanced Eb and improved polarization response. By contrast, the x = 0.20 sample showed degraded performance attributable to reduced Pmax and Eb, which originated from impurity phases. According to the calculated energy storage parameters, the Wrec values of BNBT-xSMS ceramics were 7.2, 11.4, 13.6, and 7.6 J cm−3 along with corresponding η values of 78.3%, 82.0%, 83.9%, and 86.8%, respectively, for increasing x. The primary factor governing these improvements was the enhancement in Eb. The reliability of Eb was evaluated using Weibull distribution analysis based on the following equation:56

 
Xi = ln(Ei) (4)
 
image file: d5ta04848f-t4.tif(5)
where n represents the sample count and Ei denotes individual breakdown fields. The calculated shape parameters (β) were all >20, confirming outstanding Eb stability. This reliability stems from high specimen density and defect reduction, as discussed later in the Eb analysis section. As shown in Fig. 2d, the unipolar PE loops for BNBT-0.16SMS were measured under different electric fields. When the electric field was increased from 100 kV cm−1 to 760 kV cm−1, Pmax values increased significantly from 6.2 μC cm−2 to 46.1 μC cm−2, while Pr remained nearly unchanged. Consequently, Wrec improved from 0.3 J cm−3 to 13.6 J cm−3 with consistently high η >83% (Fig. 2e), demonstrating excellent electric field stability. Fig. 2f provides a performance comparison between BNBT-0.16SMS and other state-of-the-art lead-free systems, revealing that it achieves competitive performance and showcasing significant application potential for high-performance energy storage capacitors.


image file: d5ta04848f-f2.tif
Fig. 2 (a) BNBT-xSMS ceramics tested up to the unipolar PE hysteresis loops of Eb at 10 Hz and (b) Wtotal, Wrec, and η as a function of x. (c) Weibull distribution of Eb for BNBT-xSMS ceramics. (d) Unipolar PE loops of BNBT-0.16SMS ceramic under different electric fields and (e) the corresponding Wtotal, Wrec, and η as a function of an electric field. (f) Comparison of Wrec and η between the BNBT-0.16SMS ceramic and other reported lead-free energy storage ceramics.

In contrast to conventional ferroelectric BNBT ceramics, the modified relaxor ferroelectric BNBT-xSMS system demonstrates significantly improved dielectric temperature and frequency stability. Fig. 3a–c present the temperature- and frequency-dependent relative permittivity (εr) and dielectric loss (tan[thin space (1/6-em)]δ) of BNBT-xSMS ceramics. With increasing x, the ceramics exhibit enhanced frequency insensitivity accompanied by reduced tan[thin space (1/6-em)]δ (Fig. 3a), which can be ascribed to the fast response and reversibility of PNRs.57 The gradual decrease in εr with SMS doping arises from weakened coupling between PNRs, confirming enhanced structural disorder. The temperature-dependent εr behavior is illustrated in Fig. 3b. Pure BNBT ceramics typically show two dielectric anomalies, which provide the basis for broad dielectric plateaus. The BNBT-0.08SMS composition displays characteristic dual peaks: a low-temperature shoulder peak (Ts ∼150 °C) associated with thermal evolution between R3c-phase PNRs and P4bm-phase PNRs, exhibiting pronounced frequency dispersion, and a high-temperature peak (Tm ∼280 °C) corresponding to the transition of R3c-phase PNRs into P4bm-phase PNRs.58 With further SMS doping, the Ts peak shifts to lower temperatures, while the Tm peak diminishes, thereby forming a broad dielectric plateau between the two peaks. This behavior results from enhanced A-/B-site ionic disorder and the disruption of long-range ferroelectric domains. This diffuse phase transition characteristic significantly improves the temperature stability of εr. The observed strong frequency dispersion and diffuse phase transition behavior provide clear evidence for the improved relaxor properties in this system.59 The enhancement of relaxor behavior can be attributed to the incorporation of heterogeneous ions into the lattice, which disrupts the original long-range ferroelectric ordering and facilitates lattice reorganization into nanodomain structures with smaller dimensions.60 This structural modification manifests macroscopically as classic relaxor ferroelectric behavior in the dielectric temperature spectra. Fig. 3c displays the temperature-dependent tan[thin space (1/6-em)]δ of BNBT-xSMS ceramics at 100 kHz. Increasing SMS content progressively suppresses tan[thin space (1/6-em)]δ, thereby mitigating energy accumulation and reducing thermal breakdown risks. The decrease in tan[thin space (1/6-em)]δ with temperature (up to 300 °C) correlates with the increasing activity of PNRs. However, the subsequent tan[thin space (1/6-em)]δ increase at elevated temperatures (>300 °C) likely originates from high-temperature leakage conduction. The characteristic frequency dispersion of dielectric response, induced by PNRs, substantially enhances the relaxor properties of BNBT-xSMS ceramics. Quantitative evaluation of relaxor behavior was conducted using the modified Curie–Weiss law:61

 
image file: d5ta04848f-t5.tif(6)
where C is the Curie constant and γ characterizes the degree of relaxor behavior (ranging from 1 to 2). Note that γ = 1 corresponds to normal ferroelectrics, while γ = 2 represents ideal relaxor ferroelectrics; therefore, the larger the γ is, the stronger the relaxation is.62 As shown in Fig. 3d, the fitted γ values for the studied compositions range from 1.84 to 1.98, confirming strong relaxor behavior. Additionally, the relaxor characteristics were further characterized by ΔTrelaxorTrelaxor = Ts (1 MHz) − Ts (5 kHz)).55 Fig. 3e demonstrates that BNBT-xSMS ceramics exhibit ΔTrelaxor values of 44–55 °C, indicating enhanced relaxor characteristics due to cation disorder and the coexistence of multiphase PNRs structures. The broadening and flattening of dielectric peaks caused by these collective relaxor effects improve dielectric thermal stability. To quantify this stability, we examined the temperature coefficient of capacitance (TCC) with 150 °C as the reference temperature (εr/ε150°C). As shown in Fig. 3f, the TCC variation remains remarkably stable within ±15% over a broad temperature range at 10 kHz, directly demonstrating that strong relaxation behavior effectively suppresses thermal fluctuations in εr.


image file: d5ta04848f-f3.tif
Fig. 3 (a) The εr as a function of frequency for BNBT-xSMS ceramics. (b) Temperature and frequency dependence of εr and tan[thin space (1/6-em)]δ for BNBT-xSMS ceramics. (c) Temperature dependence of tan[thin space (1/6-em)]δ for BNBT-xSMS ceramics at 10 kHz. (d) Modified Curie–Weiss formula fitting results of BNBT-xSMS ceramics at 10 kHz (the inset shows the variation of γ with SMS content). (e) ΔTrelaxor of BNBT-xSMS ceramics. (f) Temperature stability (TCC150°C ≤ ±15%) of BNBT-xSMS ceramics at 10 kHz. (g) Vogel–Fulcher fitting results of BNBT-xSMS samples and (h) the corresponding fitted Ea values with x content. (i) Evolution diagram of dielectric behavior dependence of SMS content.

In relaxor ferroelectrics, the thermal evolution of PNRs is accompanied by changes in their size and dynamics. Based on three characteristic temperatures—the freezing temperature (Tf), phase transition temperature (Tm), and burning temperature (TB)—the dielectric temperature spectra can be divided into non-ergodic relaxation (NER), ergodic relaxation (ER), and paraelectric (PE) states.63 In the paraelectric state (>TB), PNRs exhibit smaller sizes. As the temperature decreases in the ER state (TfTB), PNRs nucleate and grow, acquiring high dynamic activity that enables reversible polarization reversal under the electric field. Below Tf, the system transitions to the NER state, where PNRs freeze and evolve into long-range ordered ferroelectric domains.35 By tuning the ER state through composition control, enhanced dynamic responses of PNRs can be achieved, improving energy storage characteristics. To quantify these transitions, we employed Vogel–Fulcher (V–F) fitting for Tf and Curie–Weiss fitting for TB using the following formulas:55,64

 
image file: d5ta04848f-t6.tif(7)
 
image file: d5ta04848f-t7.tif(8)
where f0 is the attempt frequency, Ea is the activation energy for PNR dynamics, kB is the Boltzmann constant, and TC is the Curie temperature. A larger Ea indicates stronger energy barriers between dipole clusters, suppressing long-range dipole correlations and promoting slim hysteresis loops. Here, Tf and Ea were extracted from the dielectric frequency dispersion peak Ts using V–F fitting (Table S1). As shown in Fig. 3g, Tf decreases from 71.8 °C to −41.2 °C with increasing x, confirming the transition from NER to ER states under SMS doping. Concurrently, Ea increases from 0.096 eV (x = 0.08) to 0.264 eV (x = 0.20) (Fig. 3h), indicating smaller dipole cluster interactions that suppress long-range correlations and mitigate dielectric nonlinearities. Fig. S2a–d present TB fitting results, showing a reduction from 350 °C (x = 0.08) to 267 °C (x = 0.20). The composition-temperature phase diagram highlights how SMS doping systematically shifts BNBT ceramics into the ER regime (Fig. 3i). By engineering the ER state near room temperature, PNRs achieve weakly coupled relaxation behavior, enhanced large Pmax under a high electric field, high energy storage efficiency, and suppressed macroscopic nonlinear polarization responses, which are key factors for optimizing the energy storage performance. These findings establish that controlling PNR thermal evolution through chemical modification is a key strategy for developing high-performance energy storage dielectrics.

The energy storage characteristics of relaxor ferroelectric ceramics are fundamentally governed by the domain structure configuration and its dynamic response behavior. Hence, the domain reversible switching behavior of ceramics during voltage application and removal was characterized by piezoresponse force microscopy (PFM). The opposite voltage (±20 V) was applied to the tip of the scanned sample over a scanning area of 4 × 4 μm2. The incorporation of SMS doping into the BNBT system induces local structural disorder in the long-range ferroelectric order of the BNBT matrix, promoting the formation of fine nanodomains or PNRs. Fig. 4a and b present the domain morphology phase images of pure BNBT and BNBT-0.16SMS ceramics, respectively, where distinct color orientations correspond to different domain configurations. Pure BNBT exhibits maze-like domain structures characterized by large-scale (sub-micron) irregular domain distributions, reflecting its conventional long-range ferroelectric ordering. In striking contrast, the SMS-modified counterpart demonstrates a complete transition to nanoscale domain structures (verified by the TEM domain structure), manifested as featureless PFM response patterns due to PFM accuracy limitation. The formation of such nanodomains is mainly due to the local structural disorder and enhanced random fields resulting from the increased ionic disorder by doping SMS with different ionic radii. Detailed domain switching and relaxation property analysis uncovers the fundamentally different polarization behaviors between these systems. The litho-PFM test was used to write domains on the ceramics by applying different voltages, and the applied voltages are shown schematically in Fig. 4c. Pure BNBT shows complete domain reversal under ±20 V bias with negligible relaxation after 10 minutes, characteristic of strong ferroelectric coupling that leads to high remanent polarization and low-field saturated polarization behavior and consequently poor energy storage capability (Fig. 4d1–d3). In contrast, the BNBT-0.16SMS composition exhibits entirely distinct switching dynamics. Its displays weak domain-flipping behavior at ±20 V due to the weakly coupled characteristics of PNRs, as shown in Fig. 4e1. Higher voltages are required to induce the formation of large-size domains and achieve a high polarization response. After 5 and 10 minutes of relaxation, the domain-switching behavior fully recovers to its initial state (Fig. 4e2 and e3). This unique behavior stems from the weakly coupled, highly dynamic PNRs that enable dramatically reduced Pr.65 Therefore, the engineered PNR configuration in BNBT-0.16SMS yields exceptional energy storage performance metrics: ultrafast polarization response, minimal dielectric loss, outstanding energy storage efficiency, and remarkable field stability.


image file: d5ta04848f-f4.tif
Fig. 4 (a and b) Out-of-plane PFM phase pictures of pure BNBT and BNBT-0.16SMS ceramics, respectively. Out-of-plane PFM phase images and domain switching behavior under ±20 V for the pure BNBT and BNBT-0.16SMS ceramics: (c) schematic diagram of a 4 × 4 μm2 area applying ±20 V opposite voltage, (d1–d3) 0 min, 5 min and 10 min for pure BNBT ceramic, and (e1–e3) 0 min, 5 min and 10 min for BNBT-0.16SMS ceramic.

Additionally, achieving excellent energy storage performance is inseparable from attaining a high Eb. Generally, the Eb value is influenced by multiple comprehensive factors, including grain size, density, electrical insulation properties, and defect states. The surface morphology of the specimens was characterized by scanning electron microscopy (SEM), as shown in Fig. 5a, b, S3a and b. With increasing SMS content, the average grain size (Ga) of BNBT-xSMS ceramics initially decreases and then increases, reaching a minimum value of 2.14 μm for BNBT-0.16SMS ceramic. The reduction in Ga is attributed to the increased lattice strain energy caused by Sm3+, Mg2+, and Sb5+ ions with different ionic radii, which hinder grain boundary migration.66 However, excessive doping induces abnormal grain growth, likely associated with secondary phase formation. As shown in Fig. 5c, the correlation between various SMS contents and Eb, Ga, and bulk density (ρ) reveals that both the reduced Ga and increased ρ positively contribute to improving Eb. Complex impedance spectroscopy was employed to investigate the roles of grains and grain boundaries in charge transport mechanisms and to elucidate the intrinsic factors enhancing Eb. The impedance spectra of BNBT-xSMS ceramics with x = 0.08 and x = 0.16 measured at high temperatures (520–600 °C) are shown in Fig. 5d and e. Notably, the total resistance of the ceramics significantly increases with higher SMS doping, which favors improved Eb. For the BNBT-0.08SMS ceramic, two distinct response modes can be significantly observed, corresponding to the grain boundary and grain contributions, respectively. In contrast, the BNBT-0.16SMS ceramic exhibits a single response mode predominantly governed by grain boundaries. Further analysis of the impedance imaginary part (Z′′) versus electrical modulus (M′′) reveals that the disappearance of double peaks and peak coalescence confirm enhanced electrical homogeneity (Fig. 5f).26 Therefore, grain and grain boundary responses must be differentiated during impedance fitting. The grain and grain boundary resistances derived from impedance analysis are interpreted as thermally activated processes. Based on the Arrhenius relationship:67

 
image file: d5ta04848f-t8.tif(9)
where R is the fitted equivalent resistance, kB is the Boltzmann constant, Ea is the activation energy, and T is the temperature. The fitting results yield activation energies of 1.31 eV (grains) and 1.03 (grain boundaries) for x = 0.08, while x = 0.16 shows a unified Ea = 1.36, which indicates that the systems are all caused by the oxygen vacancy conducting mechanism (0.5–2 eV). Thus, optimal SMS doping not only refines the grain size and enhances electrical homogeneity but also elevates the thermal activation energy, thereby significantly improving Eb. The increase in Ea is inversely correlated with the reduction of oxygen vacancy (OV) content. X-ray photoelectron spectroscopy (XPS) was employed to analyze the OV concentration as a function of SMS doping levels. The O 1s XPS spectra of the ceramics and the fitting of three Gaussian–Lorentzian peak positions corresponding to lattice oxygen (OL), absorbed oxygen (OA) and oxygen vacancies are shown in Fig. 5h, where the OL, OA and OV ratios are summarized in Fig. 5i.35 For x = 0.08, the OV content reached 16.59%, whereas it decreased to 8.75% at x = 0.16. This variance underscores the positive impact of the varying SMS contents on inhibiting the generation of oxygen vacancies. In conclusion, the superior Eb of the ceramics can be primarily attributed to the decrease in Ga, increase in electrical insulation, increase in Ea and decrease in OV content.


image file: d5ta04848f-f5.tif
Fig. 5 (a and b) SEM surface topography for x = 0.08 and x = 0.16 (inset shows grain size distribution). (c) Variation of Eb, Ga and bulk density with different SMS contents. (d and e) Complex impedance plot for x = 0.08 and x = 0.16 measured at 520–600 °C. (f) Spectroscopic plots of Z′′ and M′′ at 520 °C. (g) Arrhenius plots of total electrical conductivity. (h) O 1s X-ray photoelectron spectra for x = 0.08 and x = 0.16 samples and (i) relationship between changes in OL, OA and OV.

For energy storage capacitor applications in advanced pulsed power devices, operational stability under varying conditions is paramount. Here, our systematic evaluation of the representative BNBT-0.16SMS ceramic demonstrates exceptional performance across wide temperature and frequency ranges under a high electric field (450 kV cm−1). As shown in Fig. 6a, the PE loops shape remains stable across the temperature range of 30–140 °C. The Wrec and η exhibit minimal variation, with values of 5.7 J cm−3 (84.5%) at 30 °C and 5.99 J cm−3 (84.0%) at 120 °C (Fig. 6b). This negligible fluctuation confirms the exceptional thermal stability of BNBT-0.16SMS ceramic. The ceramic exhibits similarly impressive frequency insensitivity from 10 to 100 Hz, maintaining stable Pmax values (ΔPmax <1.1 μC cm−2) (Fig. 6c). At 10 Hz, Wrec and η are 5.8 J cm−3 and 85.4%, respectively. At 100 Hz, these values decrease slightly to 5.6 J cm−3 and 84.0%, corresponding to marginal variations in Wrec and η of 3.0% and 1.6%, respectively (Fig. 6d). These results highlight the outstanding frequency insensitivity of the material. The excellent temperature stability originates from the structural design ensuring broad temperature stability. Fig. 6e presents the temperature-dependent Raman spectra of BNBT-0.16SMS ceramic. Within the temperature range of room temperature to 400 °C, no new Raman peaks emerge, and no significant peak shifts are observed, confirming that the temperature insensitivity of the local structure symmetry is a critical factor enabling superior thermal energy storage stability. Deconvolution analysis reveals that with increasing temperature, the Raman shift of the ν5 vibrational mode associated with the BO6 octahedra exhibits a red shift (lower wavenumber), accompanied by reduced Raman intensity and increased full width at half-maximum (FWHM). These spectral changes indicate enhanced local structural disorder, which facilitates the maintenance of stable Pr even at elevated temperatures.68 Pulse charge–discharge performance is critical for evaluating practical application potential. Fig. 6g presents the overdamped charge–discharge current curves of BNBT-0.16SMS ceramic with optimal energy storage performance. Under test electric fields ranging from 100 kV cm−1 to 720 kV cm−1, the discharge curves remain stable, with discharge currents gradually increasing. The discharged energy density (Wdis) can be calculated using the following formula:69

 
image file: d5ta04848f-t9.tif(10)
where R and V are the overload resistance (10 kΩ) and the volume of the ceramic sample, respectively. Fig. 6h summarizes the relationship between Wdis and t0.9 (time required to release 90% of stored energy) as a function of electric field. When the voltage increases from 100 kV cm−1 to 720 kV cm−1, Wdis sharply increases from 0.2 J cm−3 to 11.8 J cm−3, while t0.9 remains below 2 μs. The slight decrease in Wdis compared to Wrec under the same electric field is attributed to thermal dissipation during practical charge–discharge processes. The abundant PNRs induced by this composition enable rapid domain reorientation, achieving exceptional charge–discharge performance. Fig. 6i compares the Wdis of BNBT-0.16SMS ceramic with other advanced lead-free energy storage bulk ceramics. Most lead-free ceramics exhibit Wdis below 8.0 J cm−3, with only a few reaching 10.0 J cm−3. In contrast, the BNBT-0.16SMS ceramic achieves an ultrahigh Wdis of 11.8 J cm−3, demonstrating its significant potential for advanced pulsed discharge devices. Therefore, the excellent thermal and frequency stability of BNBT-0.16SMS ceramic, combined with its high pulse discharge performance, demonstrate its significant potential for practical applications in high-power devices.


image file: d5ta04848f-f6.tif
Fig. 6 (a) Unipolar PE loops of BNBT-0.16SMS ceramic, tested in the temperature range of 30–50 °C. (b) Corresponding variations in Wrec and η with different temperatures. (c) Unipolar PE loops of BNBT-0.16SMS ceramic, tested in the frequency range of 10–100 Hz. (d) Corresponding variations in Wrec and η with different frequencies. (e) Raman spectra of BNBT-0.16SMS ceramic in the temperature range of 25–400 °C. (f) Variations of wavenumber, intensity and FWHM for the ν5 Raman vibration modes. (g and h) Charge/discharge performance of BNBT-0.16SMS ceramic. (i) Comparison of Wdis between the BNBT-0.16SMS ceramic and other reported lead-free energy storage ceramics.

4. Conclusions

In this work, the domain configuration and defect structure were systematically modulated through chemical modification via SMS doping modification. This strategic chemical engineering approach simultaneously promotes highly dynamic polarization switching behavior and improved dielectric breakdown, thereby achieving superior energy storage performance. The optimized BNBT-0.16SMS ceramic demonstrates exceptional energy storage capability, with a Wrec of 13.6 J cm−3 and η of 84.9%. It also possessed outstanding temperature/frequency stability and remarkable Wdis = 11.8 J cm−3 under high electric fields up to 720 kV cm−1. Microstructural analysis reveals that SMS incorporation effectively reduces grain size and suppresses oxygen vacancy formation, leading to significantly improved Eb. Complementary Vogel–Fulcher modeling and PFM analysis confirm that the engineered fine-scale weakly coupled nanodomains substantially reduce the energy barrier for polarization reversal, which facilitates rapid polarization switching, resulting in the simultaneous achievement of Wrec and η at high electric fields. Our work provides a feasible reference for the development of dielectric materials to meet future high energy storage characteristics and promotes the research of other lead-free energy storage dielectric materials.

Author contributions

Jiangping Huang: writing – review & editing, writing – original draft, data curation, investigation, conceptualization, methodology, visualization. Liang Deng: investigation, conceptualization, visualization. Yu Zhang: formal analysis, visualization. Yue Pan: formal analysis, visualization. Xiuli Chen: supervision, funding acquisition. Xu Li: supervision, writing – review & editing, supervision, funding acquisition. Huanfu Zhou: supervision, funding acquisition, resources.

Conflicts of interest

The authors declare that they have no known competing financial interests.

Data availability

The data collected and produced in this investigation are available upon reasonable request from the corresponding author.

The P–E hysteresis loops of BNBT-xSMS ceramics, and corresponding variations in energy storage parameters; The fitting of TB value; SEM images of BNBT-xSMS (x = 0.12 and 0.20) ceramics; The V–F fitting parameters of BNBT-xSMS ceramics. See DOI: https://doi.org/10.1039/d5ta04848f.

Acknowledgements

This study was supported by the Guangxi Science and Technology Plan Project (GuikeAD25069100), the Guangxi Natural Science Foundation Project (2025GXNSFBA069167), the Seedling Talent Inclusive Policy Scientific Research Fund, and the Guangxi Postdoctoral Innovation Talent Support Program.

References

  1. H. Zubairi, Z. Lu, Y. Zhu, I. M. Reaney and G. Wang, Chem. Soc. Rev., 2024, 53, 10761–10790 RSC .
  2. F. Yan, J. Qian, S. Wang and J. Zhai, Nano Energy, 2024, 123, 109394 CrossRef CAS .
  3. S. Pattipaka, Y. Lim, Y. H. Son, Y. M. Bae, M. Peddigari and G. T. Hwang, Materials, 2024, 17, 2277 CrossRef CAS .
  4. Y. Zhang, L. Chen, H. Liu, S. Deng, H. Qi and J. Chen, InfoMat, 2023, 5, 12488 CrossRef .
  5. B. Guo, F. Jin, L. Li, Z.-Z. Pan, X.-W. Xu and H. Wang, Rare Met., 2023, 43, 853–878 CrossRef .
  6. X. Lv, T. Zheng, C. Zhao, J. Yin, H. Wu and J. Wu, Acc. Mater. Res., 2022, 3, 461–471 CrossRef CAS .
  7. F. P. Zhuo, H. Qiao, J. M. Zhu, S. Z. Wang, Y. Bai, X. P. Mao and H. H. Wu, Chin. Chem. Lett., 2021, 32, 2097–2107 CrossRef CAS .
  8. Z. Yang, H. Du, L. Jin and D. Poelman, J. Mater. Chem. A, 2021, 9, 18026–18085 RSC .
  9. G. Wang, Z. L. Lu, Y. Li, L. H. Li, H. F. Ji, A. Feteira, D. Zhou, D. W. Wang, S. J. Zhang and I. M. Reaney, Chem. Rev., 2021, 121, 6124–6172 CrossRef CAS .
  10. D. X. Li, X. J. Zeng, Z. P. Li, Z. Y. Shen, H. Hao, W. Q. Luo, X. C. Wang, F. S. Song, Z. M. Wang and Y. M. Li, J. Adv. Ceram., 2021, 10, 675–703 CrossRef CAS .
  11. B. Deka and K.-H. Cho, Materials, 2021, 14, 7188 CrossRef CAS .
  12. W. Liu, B. Fu, J. Zhang, X. Ma, Y. Mao, Q. Zong, Z. Zhu, H. Yuan, Y. Zhou and W. Bai, Chem. Eng. J., 2025, 512, 162477 CrossRef CAS .
  13. F. Z. Yao, Q. B. Yuan, Q. Wang and H. Wang, Nanoscale, 2020, 12, 17165–17184 RSC .
  14. D. Yang, J. Gao, L. Shu, Y.-X. Liu, J. Yu, Y. Zhang, X. Wang, B.-P. Zhang and J.-F. Li, J. Mater. Chem. A, 2020, 8, 23724–23737 RSC .
  15. V. Veerapandiyan, F. Benes, T. Gindel and M. Deluca, Materials, 2020, 13, 5742 CrossRef CAS .
  16. L. Yang, X. Kong, F. Li, H. Hao, Z. Cheng, H. Liu, J.-F. Li and S. Zhang, Prog. Mater. Sci., 2019, 102, 72–108 CrossRef CAS .
  17. R. Kang, Z. Wang, M. Wu, S. Cheng, S. Mi, Y. Hu, L. Zhang, D. Wang and X. Lou, Nano Energy, 2023, 112, 108477 CrossRef CAS .
  18. F. Yang, Z. Pan, Z. Ling, D. Hu, J. Ding, P. Li, J. Liu and J. Zhai, J. Eur. Ceram. Soc., 2021, 41, 2548–2558 CrossRef CAS .
  19. C. Long, Z. Su, H. Song, A. Xu, L. Liu, Y. Li, K. Zheng, W. Ren, H. Wu and X. Ding, Energy Storage Mater., 2024, 65, 103055 CrossRef .
  20. Z. F. Zhang, Y. Zhang, H. R. Bai, P. Li, H. H. Huang, Z. L. Li, M. K. Joshi, W. Li, J. G. Hao, J. Du and P. Fu, J. Energy Storage, 2024, 44, 15160–15166 Search PubMed .
  21. C. Long, Z. Su, A. Xu, H. Huang, L. Liu, L. Gu, W. Ren, H. Wu and X. Ding, Nano Energy, 2024, 124, 109493 CrossRef CAS .
  22. S. Cheng, K. Zhang, C. Li, B. Zhang, J. Chen, K. Sun, C. zhou, J. Zhao, Q. Lin, G. Rao and S. Shi, J. Mater., 2024, 10, 803–810 Search PubMed .
  23. R. Zhao, Y. Zhang, W. Li, X. Tang, K. Wang, J. Hu, Z. Shen, H. Fan, Y. Jiang and X. Guo, ACS Appl. Electron. Mater., 2023, 5, 6104–6113 CrossRef CAS .
  24. Q. Lin, L. Li, W. Dou, Y. Wu, Y. Xie and Z. Deng, Ceram. Int., 2023, 49, 16225–16234 CrossRef CAS .
  25. D. Wang, Z. Fan, D. Zhou, A. Khesro, S. Murakami, A. Feteira, Q. Zhao, X. Tan and I. M. Reaney, J. Mater. Chem. A, 2018, 6, 4133–4144 RSC .
  26. W. Cao, L. Li, K. Chen, X. Huang, F. Li, C. Wang, J. Zheng, X. Hou and Z. Cheng, Adv. Sci., 2024, 11, 2409113 CrossRef CAS .
  27. M. Zhang, S. Lan, B. B. Yang and H. Pan, Science, 2024, 384, 185–189 CrossRef CAS .
  28. H. Pan, F. Li, Y. Liu, Q. Zhang, M. Wang, S. Lan, Y. Zheng, J. Ma, L. Gu, Y. Shen, P. Yu, S. Zhang, L.-Q. Chen, Y.-H. Lin and C.-W. Nan, Science, 2019, 365, 578–582 CrossRef CAS .
  29. X. Dong, X. Wu, X. Lv and J. Wu, J. Mater. Chem. A, 2024, 12, 21772–21781 RSC .
  30. J. Liu, Y. Jiang, W. Zhang, X. Cheng, P. Zhao, Y. Zhen, Y. Hao, L. Guo, K. Bi and X. Wang, Nat. Commun., 2024, 15, 8651 CrossRef CAS .
  31. X. Qiao, A. Sheng, D. Wu, F. Zhang, B. Chen, P. Liang, J. Wang, X. Chao and Z. Yang, Chem. Eng. J., 2021, 408, 127368 CrossRef CAS .
  32. J. Chen, F. Si, P. Zhao, S. Zhang and B. Tang, Ceram. Int., 2021, 47, 26215–26223 CrossRef CAS .
  33. Z. Li, B. Xie, Z. Liu, K. Guo, K. Li, H. Zhang and H. Luo, J. Mater. Chem. A, 2025, 13, 9339–9346 RSC .
  34. X. Tang, Z. Hu, V. Koval, B. Yang, G. C. Smith and H. Yan, Chem. Eng. J., 2023, 473, 145363 CrossRef CAS .
  35. T. Shi, Q. Feng, J. Wu, Z. Cen, X. Chen, N. Luo, Y. Wei, X. Liu, J. Xu and C. Yuan, Chem. Eng. J., 2023, 470, 144205 CrossRef CAS .
  36. O. Turki, A. Slimani, Z. Sassi, H. Khemakhem, N. Abdelmoula and L. Lebrun, Appl. Phys. A, 2022, 128, 186 CrossRef CAS .
  37. Y. Pan, Q. Dong, J. Huang, X. Chen, X. Li and H. Zhou, Chem. Eng. J., 2024, 497, 154695 CrossRef CAS .
  38. X. Dong, T. Hu, X. Wu, J. Yin, Z. Fu and J. Wu, SusMat, 2023, 4, 116–125 CrossRef .
  39. Y. Pan, Z. Dai, C. Liu, X. Zhao, S. Yasui, Y. Cong and S. Gu, J. Mater. Sci., 2024, 59, 3284–3296 CrossRef CAS .
  40. Q. Liu, J. Liu, D. Lu, W. Zheng and C. Hu, J. Alloys Compd., 2018, 760, 31–41 CrossRef CAS .
  41. P. Shi, X. Wang, X. Lou, C. Zhou, Q. Liu, L. He, S. Yang and X. Zhang, J. Alloys Compd., 2021, 877, 160162 CrossRef CAS .
  42. Y. Lin, D. Li, M. Zhang and H. Yang, J. Mater. Chem. C, 2020, 8, 2258–2264 RSC .
  43. S. M. Wang, F. Yan, J. Qian, G. L. Ge, Z. Q. Fu, Z. B. Pan, F. Q. Zhang, J. F. Lin, K. Zeng, C. K. Chen, B. Shen, Z. F. Liu and J. W. Zhai, Energy Storage Mater., 2024, 66, 103155 CrossRef .
  44. H. Yang, H. Qi and R. Zuo, J. Eur. Ceram. Soc., 2019, 39, 2673–2679 CrossRef CAS .
  45. C. Long, Z. Su, A. Xu, F. Li, Y. Li, W. Ren, H. Wu, X. Ding and L. Liu, J. Adv. Ceram., 2025, 14, 9221063 CrossRef CAS .
  46. M. b. Abdessalem, A. Aydi and N. Abdelmoula, J. Alloys Compd., 2019, 774, 685–693 CrossRef CAS .
  47. A. Xie, L. Liu, Y. Zhang, A. Rahman and R. Zuo, J. Eur. Ceram. Soc., 2024, 44, 882–890 CrossRef CAS .
  48. M. K. Bilal, R. Bashir, S. U. Asif and X. Jiyang, J. Energy Storage, 2024, 97, 112841 CrossRef .
  49. H. Wang, S. Wu, B. Fu, J. Zhang, H. Du, Q. Zong, J. Wang, Z. Pan, W. Bai and P. Zheng, Chem. Eng. J., 2023, 471, 144446 CrossRef CAS .
  50. Z. Wang, R. Kang, L. Zhang, X. Lou, Y. Zhao, P. Mao and J. Wang, Chem. Eng. J., 2023, 474, 145506 CrossRef CAS .
  51. Q. Dong, D. Zeng, Y. Pan, P. Nong, X. Chen, X. Li and H. Zhou, Chem. Eng. J., 2024, 493, 152786 CrossRef CAS .
  52. Y. Pan, Q. Dong, J. Huang, Y. Zhang, X. Chen, X. Li, L. Deng and H. Zhou, J. Mater. Chem. A, 2025, 13, 3749–3764 RSC .
  53. Y. Fan, W. Qu, H. Qiu, S. Gao, L. Li, Z. Lin, Y. Yang, J. Yu, L. Wang, S. Luan, H. Li, L. Lei, Y. Zhang, H. Fan, H. Wu, S. Yu and H. Huang, Nat. Commun., 2025, 16, 3818 CrossRef CAS PubMed .
  54. J. H. Duan, K. Wei, Q. B. Du, L. Z. Ma, H. F. Yu, H. Qi, Y. C. Tan, G. K. Zhong and H. Li, Nat. Commun., 2024, 15, 6754 CrossRef CAS PubMed .
  55. X. Li, Y. Cheng, F. Wang, Q. Xu, Y. Chen, L. Xie, Z. Tan, J. Xing and J. Zhu, Chem. Eng. J., 2022, 431, 133441 CrossRef CAS .
  56. J. Chen, P. Zhao, K. Chen, F. Si, Z. Fang, S. Zhang and B. Tang, Chem. Eng. J., 2024, 502, 157866 CrossRef CAS .
  57. P. Chen and B. Chu, J. Eur. Ceram. Soc., 2016, 36, 81–88 CrossRef CAS .
  58. P. Li, H. Yang, Q. Yuan and Y. Lin, Mater. Today Phys., 2024, 43, 101420 CrossRef CAS .
  59. C. Long, Z. Su, H. Song, A. Xu, L. Liu, Y. Li, K. Zheng, W. Ren, H. Wua and X. Ding, Energy Storage Mater., 2023, 65, 103055 CrossRef .
  60. H. Xie, H. Du, L. Liu, Q. Kou, J. Xu, Y. Sun, R. Lv, Y. Chang and D. Wang, Chem. Eng. J., 2022, 450, 138432 CrossRef CAS .
  61. F. Yan, X. Zhou, X. He, H. Bai, S. Wu, B. Shen and J. Zhai, Nano Energy, 2020, 75, 105012 CrossRef CAS .
  62. D. Yang, X. Wu, X. Lv and J. Wu, J. Eur. Ceram. Soc., 2025, 45, 117240 CrossRef CAS .
  63. A. Deng and J. Wu, J. Eur. Ceram. Soc., 2021, 41, 5147–5154 CrossRef CAS .
  64. X. Dong, X. Li, X. Chen, J. Wu and H. Zhou, Chem. Eng. J., 2021, 409, 128231 CrossRef CAS .
  65. Y. Pan, Y. Zhang, Q. Dong, J. Huang, S. Zhao, X. Chen, X. Li and H. Zhou, Chem. Eng. J., 2025, 512, 162551 CrossRef CAS .
  66. L. Ma, Z. Che, Y. Luo, C. Xu, Z. Cen, F. Toyohisa, S. Zhang, J.-F. Li and N. Luo, Acta Mater., 2025, 289, 120950 CrossRef CAS .
  67. Q. Xu, M. T. Lanagan, W. Luo, L. Zhang, J. Xie, H. Hao, M. Cao, Z. Yao and H. Liu, J. Eur. Ceram. Soc., 2016, 36, 2469–2477 CrossRef CAS .
  68. F. Chen, M. Chen, J. Zhang, W. Liu, H. Du, Q. Zong, H. Yu, Y. Zhang, J. Hao, J. Wang and J. Zhai, ACS Nano, 2025, 19, 1809–1818 CrossRef CAS .
  69. Q. B. Liao, T. Deng, T. Lu, Z. Liu, N. Narayanan, S. Li, S. G. Yan, Y. Z. Bao, Y. Liu and G. S. Wang, Chem. Eng. J., 2024, 488, 150901 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.