Superior energy storage performance with a record high breakdown strength in bulk Ba0.85Ca0.15Zr0.1Ti0.9O3-based lead-free ceramics via multiple synergistic strategies

Changhao Wang a, Jiaxi Hao a, Longxiao Duan a, Jianfan Zhang a, Wenfeng Yue b, Zhenhao Fan b, Dandan Han *a, Raz Muhammad c, Fanxu Meng a and Dawei Wang *b
aKey Laboratory for Special Functional Materials in Jilin Provincial Universities, Jilin Institute of Chemical Technology, Jilin 132022, PR China. E-mail: handd0502@jlict.edu.cn
bFunctional Materials and Acousto-Optic Instruments Institute, School of Instrumentation Science and Engineering, Harbin Institute of Technology, Harbin 150080, PR China. E-mail: wangdawei102@gmail.com
cDepartment of Physics, Abdul Wali Khan University Mardan, Garden Campus, 23200 Mardan, KP, Pakistan

Received 20th August 2024 , Accepted 16th October 2024

First published on 24th October 2024


Abstract

A high breakdown strength (Eb) together with a large maximum polarization (Pm) is essential for achieving a high recoverable energy density (Wrec) in energy storage dielectric ceramics. However, meeting the urgent need for practical applications remains a challenge due to the intrinsic properties of bulk dielectric ceramics. Herein, a composition and structure optimization strategy combined with a two-step sintering (TSS) process is proposed to design and fabricate (1−x)Ba0.85Ca0.15Zr0.1Ti0.9O3xBi(Mg1/2Sn1/2)O3 (BCZT-BMSx-TSS) lead-free ceramics. Highly dynamic locally polar nano-regions (PNRs) are formed via composition optimization, exhibiting a very high Pm and energy storage efficiency (η). Compared to the traditional one-step sintering (OSS) process, the TSS process results in a composition with finer grain size and higher density, dramatically increasing Eb. As a result, an ultrahigh energy storage performance with Wrec ∼ 10.53 J cm−3 and η ∼ 85.71% is achieved for the BCZT-BMSx-TSS (x = 0.08) ceramic which is attributed to a record high Eb ∼ 830 kV cm−1 and a large Pm ∼ 44.66 μC cm−2. Complex impedance spectroscopy revealed that the activation energies of the bulk and grain boundary counterparts significantly increased, suggesting an increase in insulation resistance and a decrease in oxygen vacancies, which is the main reason for the high Eb value. In addition, excellent thermal/frequency stability is achieved in both energy density and efficiency, along with good charge–discharge performance. These findings suggest that BCZT-based lead-free ceramics have the potential for practical use in the future.


1. Introduction

In recent years, the energy storage field has become an indispensable part of utilizing renewable sources. Dielectric ceramic capacitors are becoming increasingly important due to their excellent properties, such as ultrafast charging/discharging performance, high power density, and long-time stability, which are the key parameters for developing advanced pulsed power electronic systems.1–5 However, compared to commercial capacitors, the energy storage density of bulk dielectric ceramics is too low for practical applications. To evaluate the energy storage performance (ESP) of dielectric ceramics, the following basic relations are used:
 
image file: d4qi02126f-t1.tif(1)
 
image file: d4qi02126f-t2.tif(2)
 
image file: d4qi02126f-t3.tif(3)

Excellent recoverable energy storage density (Wrec) and efficiency (η) can be simultaneously obtained via achieving a high maximum polarization (Pm) and a low remanent polarization (Pr) as well as enhancing the breakdown strength (Eb).6,7

Typical ferroelectric (FE) ceramics, such as BaTiO3 (BT)-based dielectrics, exhibit a nonlinear dielectric response under a certain electric field.8–15 These ceramics possess a high total energy density (Wtot), but at the cost of a large Pr value. The large Pr and small Eb restrict the improvement of Wrec and η.11–13 Making a solid solution of BT with a relaxor end-member may enhance Eb and ΔPP = PmPr) because of the formation of highly dynamic polar nano-regions (PNRs), making it a potential candidate for energy storage applications.14,15 The formation of highly dynamic PNRs is an effective way to enhance the saturation polarization by improving the structural relaxation and nanodomains. Under an external alternating electric field, the polarity of PNRs can be rapidly reversed, leading to an approximately linear polarization–electric field curve (PE loop) with large saturation polarization, thereby ensuring high performance for energy storage ceramics.16–18 Significantly, the quasilinear characteristics of the PE loop make Eb a more crucial parameter in energy storage ceramics due to the approximate square relationship between Eb and Wrec.19–21 In other words, while ensuring a high Pm, a higher Wrec can be obtained by improving the Eb value. To maintain a high Pm and low Pr, one effective approach is to make the solid solution of Bi(M)O3 (M denotes trivalent cations) with BT-based ceramics.22–25 The introduction of an appropriate amount of a Bi containing “relaxor end-member” can disrupt the B–O coupling of perovskite ceramics and establish A(Bi)–O coupling. This results in the formation of highly dynamic PNRs and Pm, which contributes to achieving the excellent energy storage density.26–30 Although a relatively large Pm and a smaller Pr can be achieved using this method, the saturated polarization features are more dependent on the ceramics themselves. In other words, instead of further increasing the value of Pm, higher Wrec can be easily obtained by improving Eb.

The microstructural features and intrinsic electrical responses, such as grain size, density, band gap, and/or the distribution of conductive phases, play a key role in affecting Eb.1,16,31–33 Recent studies show that improving Eb can be effectively achieved by optimizing the processing method and doping strategies.34–37 It is also reported that the grain size (GS) of ceramics has an inverse relationship with Eb, i.e., image file: d4qi02126f-t4.tif. Dong et al. designed and fabricated BT-based ceramics with a chemical formula of (1−y)[0.9Ba1−xCaxTiO3−0.1Bi(Li1/3Zr2/3)O3]−yBi0.5Na0.5TiO3.38 Results showed that the ceramics exhibited excellent breakdown characteristics, which is attributed to the small and fine grains, leading to a high Wrec ∼ 4.23 J cm−3 and η ∼ 93.40%. Similarly, Long et al. obtained a giant Eb by doping and chemical defect strategies, achieving a relatively high Pm and low Pr, resulting in a high Wrec ∼ 5.96 J cm−3 and η ∼ 89.5%.39

The above discussion shows that improving Eb is an effective way to enhance the ESP, but the optimization of the sintering process can further enhance the energy storage capability of dielectric ceramics. In several cases, the two-step sintering (TSS) process has been proposed to obtain high-density fine-grain ceramics. Compared with the conventional one-step sintering (OSS) method, the TSS method cleverly controls the change in temperature to suppress the grain boundary migration (which leads to grain growth), while keeping the grain boundary diffusion (which is the driving force of the ceramic body) active to achieve sintering without grain growth. Based on this method, a high Eb (∼750 kV cm−1), ultrahigh Wrec (∼8.4 J cm−3), and η (∼90%) were achieved for BiFeO3–SrTiO3 ceramics.34 Yang et al. adopted a two-step cold sintering process for Sr0.7Bi0.2TiO3 (SBT) ceramics,35 achieving a sub-micrometer scale grain size along with a uniform and dense morphology, which leads to a high Eb (∼420 kV cm−1) and a superior η (∼92.84%). Similarly, Bai et al. obtained a high Eb of 580 kV cm−1, a large Wrec (∼5.98 J cm−3), and an ultrahigh η (∼98.6%) for SBT-based relaxor ferroelectric (RFE) ceramics using the TSS process.36 It is evident from the above discussion that the TSS process in BT-based is an effective strategy to enhance the ESP; therefore, exploring and developing BT-based lead-free ceramics by this process is indispensable.

Inspired by the above discussion, a composition optimization accompanied by a TSS strategy is proposed to improve the ESP of Ba0.85Ca0.15Zr0.1Ti0.9O3-based ceramics. The Bi(Mg1/2Sn1/2)O3 (BMS) “relaxor end-member” is doped into the BT matrix, with an aim of forming PNRs to ensure a high Pm and a low Pr, followed by a TSS process to further enhance Eb. Furthermore, the introduction of MgO, which possesses high insulating characteristics, is useful to achieve high Eb. This approach resulted in a record high Eb ∼ 830 kV cm−1, accompanied by a large Pm ∼ 44.66 μC cm−2. The calculated values of Wrec and η were as high as ∼10.53 J cm−3 and ∼85.71%, respectively. In addition, the designed and fabricated ceramics showed excellent temperature/frequency stability along with a fast charge–discharge performance. All the desired results demonstrated the potential for practical application of the designed ceramics.

2. Experimental

(1−x)Ba0.85Ca0.15Zr0.1Ti0.9O3xBi(Mg1/2Sn1/2)O3 (abbreviated as BCZT-BMSx-TSS, x = 0.03, 0.06, 0.08, 0.10) ceramics were fabricated through the TSS method. Analytical reagent grade (>99.0%) BaCO3, CaCO3, ZrO2, Bi2O3, MgO and SnO2 purchased from Sinopharm Chemical Reagent Co., Ltd and TiO2 (Shanghai Yuejiang Powders, anatase) were weighed according to the molar ratios of the compositions. The powders were ball milled for 16 h in anhydrous ethanol using a planetary milling machine at 230 rpm. After drying, the mixed powders were calcined at 950–1100 °C for 3 h at a heating rate of 300 °C h−1. An 8 wt% polyvinyl alcohol (PVA) was added to the calcined powders and then pressed into cylindrical discs, using a uniaxial press. The PVA was removed at 600 °C for 2 h. The discs were sintered by TSS technology to form densified ceramics. The obtained discs were initially heated to 1420 °C in air at a heating rate of 200 °C h−1 then straight down to 1320 °C and held at 1320 °C for 10 h to achieve maximum possible densification. The densified pellets were ground/polished to a thickness of about ∼0.1 mm, and Au electrodes with a diameter of 2 mm or 3 mm were sputtered on both surfaces to evaluate the energy storage performance. The thickness of ∼0.8 mm samples was sputtered Au as electrodes to measure dielectric properties and the electrode area is 5 mm × 5 mm.

X-ray diffraction patterns for the samples were collected using an X-ray diffractometer (DX-2700, Dandong Haoyuan). Lattice parameters were calculated by MS Modeling (Accelrys) using Cu Kα1 radiation (λ = 1.540562 Å). The Raman spectra of the samples were recorded at room temperature (RT), using a LabRAM XploRA Raman spectrometer (Horiba Jobin Yvon). The surface morphologies of the sintered discs were examined using a Scanning Electron Microscope (SEM, EVOMA 10, Zeiss). To examine the domain structure of the samples, a JEOL JEM-F200 Transmission Electron Microscope (TEM, Japan) was used. The dielectric properties, PE ferroelectric loops, and charging/discharging properties were measured using a Dielectric/Impedance spectrometer (Concept 41, Novocontrol, Germany), ferroelectric analyzer (TF2000E, aixACCT Inc., Germany), and charging/discharging measurement system (CFD-003, Shanghai Tongguo instruments technology, China), respectively. Unless otherwise stated, all the tests were carried out at room temperature in air.

3. Results and discussion

XRD patterns of the samples recorded at RT are shown in Fig. 1(a), which shows the formation of a single phase perovskite structure, confirming the solubility of Bi, Mg, and Sn in the Ba0.85Ca0.15Zr0.1Ti0.9O3 lattice. Meanwhile, there is a symmetric diffraction peak without significant splitting in the range of 44 and 46°, indicating that all the doped samples have a pseudo-cubic structure. With an increase in x from 0.03 to 0.10, a slight shift towards a lower 2θ value occurs, indicative of a relative expansion of the unit-cell, which is due to the difference in the ionic radii of the doping elements. Although the ionic radius of Bi3+ (∼1.38 Å) is less than that for (Ba0.85Ca0.15)2+ (∼1.569 Å) at the A-site but at the B-site, the doping cations (Mg1/2Sn1/2)3+ (∼0.71 Å) have larger ionic radii than (Zr0.1Ti0.9)4+ (∼0.62 Å) which increases the volume of the BO6 octahedron, causing lattice expansion which is more prominent.22,30,40 And the unit cell volume (V0) is 64.43 Å3, 64.56 Å3, 64.66 Å3, and 64.69 Å3 for BCZT-BMSx-TSS (x = 0.03, 0.06, 0.08, 0.10), respectively.
image file: d4qi02126f-f1.tif
Fig. 1 (a) XRD patterns, (b) Raman spectra, and (c) temperature-dependent Raman spectra for BCZT-BMSx-TSS ceramics.

To further understand the local structure, Raman spectroscopy of the samples was carried out at RT. The obtained spectra were divided into four main parts as shown in Fig. 1(b). With an increase in x, the peaks located near ∼113 cm−1 and ∼287 cm−1 become stronger, demonstrating that the A-site ion fluctuations and B–O vibration mode are intensified.23,41 This is because the local lattice distortion increases, which further confirms the occupancy of dopants at their sites. In addition, ν1 and ν2 peaks between 400 and 800 cm−1 shift toward a low wavenumber, which is related to the vibrations of BO6 octahedra, thus reflecting the enhanced pseudo-cubic symmetry.33,42,43 The emerging vibrational mode of ν3 is an indication of local short-range ordered structures and is often associated with enhanced relaxor behavior. Moreover, the broadening of Raman peaks indicates that long-range FE ordered is disrupted, which promotes the formation of PNRs and enhances the relaxor behavior.23,44

Temperature-dependent (from −180 °C to 180 °C) Raman spectra for BCZT-BMSx-TSS ceramics are shown in Fig. 1(c), which reflect the evolution of the local structure as a function of temperature. As the temperature increases, the Raman mode broadens and the typical ν1 mode tends to shift to a lower wavenumber, which can be ascribed to an increase in the local structural disorder and the weakening of B–O bonds.2,45,46 The cation disorders can enhance domain dynamics, inducing PNRs, and resulting in an improvement in ESP. More importantly, all the samples show similar Raman active modes, suggesting the formation of RFE behavior and the stability of local structural symmetry within the measured temperature range.

Dielectric constant (ε′) as a function of temperature at different frequencies (f) of 1 kHz–1 MHz for all the samples is shown in Fig. 2(a)–(d). All the doped ceramics show a similar trend for the maximum ε′, that is the image file: d4qi02126f-t5.tif (corresponding to Tm) decreases with an increase in the frequency, and Tm moves to high temperatures, displaying a typical dielectric relaxation behavior.9,39,47,48 The f dispersion is an important feature for relaxed ferroelectrics, and the degree of f dispersion can be characterized by calculating ΔTrelaxor as follows:

 
image file: d4qi02126f-t6.tif(4)


image file: d4qi02126f-f2.tif
Fig. 2 Temperature-dependence of ε′ and tan[thin space (1/6-em)]δ for BCZT-BMSx-TSS with x = (a) 0.03, (b) 0.06, (c) 0.08, and (d) 0.10, (e) Tm as a function of x, and the inset shows the ΔTrelaxor for BCZT-BMSx-TSS ceramics; (f) temperature-dependence of ε′ measured at 1 kHz for BCZT-BMSx-TSS ceramics, and the inset shows the TCC of BCZT-BMS8-TSS; (g) a plot of image file: d4qi02126f-t11.tif; (h) TEM image for the BCZT-BMS8-TSS ceramic.

As shown in Fig. 2(e), Tm initially decreases and then slightly increases with an increase in x, which may be due to the octahedral distortion caused by the introduction of BMS in the host lattice.13,18 The value of ΔTrelaxor increases from ∼29 °C to ∼64 °C, indicating that the dispersion phase transformation behavior is gradually enhanced with the increase in doping concentration, which further indicates the transition from the ferroelectric state to the relaxation state.38,49,50 The f dispersion and diffused phase transition behavior lead to the formation of PNRs, which is crucial to achieve a high η. One further key is all samples have a very low dielectric loss (tan[thin space (1/6-em)]δ ∼ 0.004) at RT, which is beneficial for improving Eb.51,52

In Fig. 2(f), the temperature-dependence of ε′ measured at 1 kHz for BCZT-BMSx-TSS ceramics is presented. With an increase in x, the dielectric curves become broader and the ε′ of the BCZT-BMSx-TSS ceramics decreased slightly. In particular, the BCZT-BMS8-TSS material shows a moderate RT permittivity image file: d4qi02126f-t7.tif and near independence with variation in temperature, indicating that the ceramic has a low variation in temperature coefficients of capacitance (TCC) and meets the requirements of X5R specification (−55–85 °C, image file: d4qi02126f-t8.tif). In addition, an optimal ε′ can effectively enhance Eb by reducing the likelihood of electromechanical breakdown, because ε′ is inversely related to Eb. The excellent dielectric performance and relaxor characteristics of the BCZT-BMS8-TSS ceramic can be ascribed to the compression of BO6 octahedra and the restriction of B-site cation activity after difference ionic doping, which is the basis for achieving high energy storage performance.

To further verify the relaxation behavior of the samples, the modified Curie–Weiss law is used to achieve the relaxation coefficient (γ), shown in Fig. 2(g).53 The γ values of all the doped ceramics are higher than 1.6, confirming the phase transition of the doped BMS samples having typical relaxation and dispersion characteristics.51,53,54 TEM analysis is employed to confirm the existence of the PNRs (Fig. 2(h)). The HR-TEM micrograph of the BCZT-BMS8-TSS ceramic shows that very small strip-like nano-domains of about 1.2–1.8 nm (marked by white ellipses) are evenly distributed in the sample. This complex domain pattern is usually composed of randomly distributed PNRs in the matrix with weak contrasts, which indicates the destruction of long-range ferroelectric domains and the formation of PNRs after doping the Bi-based relaxor end-member.55,56 PNRs have highly dynamic characteristics and a fast response speed under the external field, which is crucial for a slim PE loop and a high Eb, leading to both a high Wrec and η.

The unipolar PE loops measured at maximum breakdown strength for BCZT-BMSx-TSS ceramics are shown in Fig. 3(a). With an increase in BMS content, the PE loops gradually become slimmer, and the maximum polarization intensity (Pm) slightly decreases, which is consistent with the relaxation characteristics observed in the dielectric spectrum (Fig. 2). More importantly, the breakdown strength (Eb) is significantly increased. For x = 0.08, the Eb of the BCZT-BMS8-TSS ceramic reaches the maximum value of ∼830 kV cm−1, which is the record high value in bulk BT-based ceramics.4,11–15,42 A high Pm ∼ 44.66 μC cm−2 and negligible remanent polarization (Pr ∼ 2.92 μC cm−2) make a large polarization difference (ΔP) for BCZT-BMS8-TSS, which is conducive to improving the energy storage density.


image file: d4qi02126f-f3.tif
Fig. 3 (a) PE loops, (b) Wrec, Wtot, and η at maximum electric fields, (c) Weibull distribution of Eb for BCZT-BMSx-TSS ceramics, and (d) the comparison of Wrec and Eb of the recently reported ceramics with BCZT-BMS8-TSS ceramic.2–6,8,9,11–15,17,18,25,26,28,32–34,36,38,39,42,45–49,58,59

Fig. 3(b) illustrates that the BCZT-BMS8-TSS ceramic exhibits an ultra-high Wrec ∼ 10.53 J cm−3 and excellent η ∼ 85.71%, which is superior to mostly lead-free energy storage ceramics (Fig. 3(d)). To further confirm the reliability of Eb, Weibull distribution is employed as shown in Fig. 3(c). After linear fitting of the Weibull distribution, the value of β for all the samples is higher than 9, indicating that the data are highly reliable.18,22,57 The average Eb value for the BCZT-BMS8-TSS ceramic is ∼800.3 kV cm−1, which is higher than most of the reported ceramics (Fig. 3(d)).

In order to evaluate the effect of the TSS process and BMS doping on the microstructure of BCZT-based ceramics, the surface morphology and average grain size distribution are investigated (Fig. 4). All the specimens have a dense microstructure. Importantly, the BCZT-BMS8-TSS ceramic has better grain uniformity and smaller porosity than other samples, which helps to enhance the Eb.16,20,55 The energy dispersive spectroscopy (EDS) spectrum confirms that all elements are uniformly distributed in the x = 0.08 ceramic (Fig. 4(g–n)), indicating that the system has a high degree of chemical composition uniformity. With the further increase in doping concentration, the grain size of the BCZT-BMS10-TSS ceramic increases, which is ascribed to the low melting point of Bi2O3. Grain size has a significant impact on the electrical performance, ESP, and dielectric compatibility of the ceramics by regulating the internal stress, tetragonality, and phase transition temperature. The dense fine-grain ceramics have high grain boundary barriers, which hinder the electrons from crossing the barrier, leading to a high Eb.58


image file: d4qi02126f-f4.tif
Fig. 4 SEM images of BCZT-BMSx-TSS with x = (a) 0.03, (b) 0.06, (c) 0.08, and (d) 0.10, respectively; (e) EDS, (f) SEM and (g–n) elemental mapping for x = 0.08 ceramic.

To further investigate the effect of the sintering process on the energy storage performance, the PE loops of BCZT-BMS8 obtained using both the OSS60 and TSS methods are shown in Fig. 5(a) and (b), respectively. Compared to the BCZT-BMS8-OSS ceramic, the BCZT-BMS8-TSS ceramic has a larger Pm and Eb, indicating a higher Wrec. This can be explained on the basis of the TSS process, which contributes to the formation of smaller grains and lower porosity (the inset of Fig. 5(a) and (b)). Actually, the relative density of BCZT-BMS8-OSS and BCZT-BMS8-TSS is ∼93% and ∼95% of the theoretical density, and the average grain size is ∼3.8 μm and ∼1.72 μm, respectively. Fig. 5(c) presents the sintering process curve for BT-based specimens sintered by OSS and TSS, respectively. For the TSS process, the sample was heated at a high temperature (T1) to obtain a medium density (greater than 75%). It is then cooled and kept at a lower temperature (T2). The different kinetics between densification diffusion and grain boundary network mobility leave a kinetic window that can be utilized in the second-step sintering.36,37,61 Due to the slow kinetics process, grain growth during the second-step sintering is completely suppressed, which is conducive to obtaining dense fine-grained ceramics, improving Eb and ESP.61


image file: d4qi02126f-f5.tif
Fig. 5 PE loops of (a) BCZT-BMS8-OSS and (b) BCZT-BMS8-TSS ceramic; (c) sintering process curve for BaTiO3-based specimens by OSS and TSS.

Complex impedance spectroscopy was carried out to understand the defect chemistry and the mechanism of high Eb of the prepared BCZT-BMSx-TSS sample.38,55 As shown in Fig. 6(a–d), there is no semicircle at a low frequency (for electrode response), indicating the good ohmic contact between the electrode and ceramics. All the samples exhibit two well-resolved semicircles at 500–600 °C, which can be fitted by the two in-series R//CPE equivalent circuit model.55,62 The two semicircles can be ascribed to the grain and grain boundary response which is in accordance with the internal barrier layer capacitor model.63 Also, all the semicircles exhibit some degree of depression and distortion, meaning the presence of a non-Debye type of relaxation phenomenon.63,64 BCZT-BMS8-TSS ceramic has the largest resistivity, which can be extracted by the intersection of the x-axis with the semicircular arc at low frequency.23,62 The real impedance decreases with the increase of measuring temperature, which is reflected in the decrease of semicircle diameter. It also can be seen that the arc of the impedance has a contraction tendency as the temperature rises, which presents a typical thermally activated relaxor process. In order to explore the conduction mechanism, the activation energy (Ea) is calculated using the fitted resistance as a function of temperature. According to the Arrhenius relationship:

 
image file: d4qi02126f-t9.tif(5)
where σ0 is a constant, σ is the bulk conductivity, k is the Boltzmann constant, and T is the temperature in Kelvin. Ea is determined from the slope of ln(1/R) − 1000/T curves. The inset of Fig. 6(a–d) shows the Arrhenius plots for BCZT-BMSx-TSS ceramics. The activation energy of the grain boundary (EGB) is greater than that of grains (EG), and the grain boundary resistance is dominant at high temperatures. With an increase in BMS content, the EGB and EG increase, indicating that the insulation properties of ceramics are enhanced. The largest EGB of ∼1.70 eV is obtained for the BCZT-BMS8-TSS ceramic, which is significantly higher than that of most reported BT-BiMO3 systems dominated by the grain boundary.14,65 In addition, the BCZT-BMS8-TSS ceramic has the highest EG, which is attributed to its compact and uniform microstructure, small grain size, and a decrease in oxygen vacancy concentration.57,62 It can also be considered that the semiconducting grains in BCZT ceramics gradually transform into insulation grains with the increase of BMS doping concentration. Generally, the higher Ea denotes a higher barrier for oxygen vacancy jumping in the grain boundary, leading to a lower concentration of oxygen vacancies at the grain boundary, which is conducive to achieve a higher Eb. Thus, the high Eb for the sample with x = 0.08 is due to the dense microstructure with fine grain size and a high Ea.


image file: d4qi02126f-f6.tif
Fig. 6 Complex impedance and activation energy (inset) for BCZT-BMSx-TSS ceramics with x = (a) 0.03, (b) 0.06, (c) 0.08, and (d) 0.10, respectively; (e) UV-Vis absorbance spectra and (f) the curves of (αhv)2 as a function of hv for BCZT-BMSx-TSS ceramics.

Fig. 6(e) and (f) present UV-vis absorption spectra and the calculated band gap (Eg) for BCZT-BMSx-TSS ceramics. The Eg value can be obtained from the following equation:23,45,66

 
(αhv)2 = A(hvEg)(6)
where α, A, and hv are the absorbance coefficient, constant, and energy of the electromagnetic waves, respectively. As shown in Fig. 6(f), with the increase of BMS content, the Eg of BCZT-BMSx-TSS ceramics increases from 3.21 eV to 3.29 eV. The increase in Eg may be due to the doping of wider band gap SnO2 (∼3.5 eV) and MgO (∼7 eV),67,68 leading to the change of the band structure, which improves the intrinsic breakdown strength and reduces the energy loss at high electric fields. In addition, the comparison of Eg and Ea shows that intrinsic conduction takes place at the grain boundary because Eg ∼ 2Ea.69,70 The values further suggest that the concentration of oxygen vacancies decreases with an increase in x which improves the barrier strength and hence Eb.

To evaluate the operating stability of dielectric capacitors, the frequency stability and temperature stability of the BCZT-BMS8-TSS ceramic at a fixed electric field intensity of 250 kV cm−1 are studied. As shown in Fig. 7(a and b), as the frequency increases from 2 Hz to 500 Hz, neither the PE loops nor the polarization intensity changes significantly. The Wrec and η values calculated at different frequencies show high frequency stability, and the PNRs that can quickly and reversibly switch have high dynamic characteristics in the external electric field, which is the main reason for the high frequency stability of ESP. In addition, temperature-dependent ferroelectric tests were performed at a fixed frequency of 250 kV cm−1 and 10 Hz to evaluate the thermal stability of the ceramics in the temperature range of 30–130 °C. Fig. 7(c and d) demonstrate that the variation in Wrec is less than ±4.3%, and the η value is slightly decreased in the entire temperature range, which is due to the thermally activated conduction promoting PE hysteresis loops and the increase of leakage current at high temperatures.18,19 In general, the BCZT-BMS8-TSS ceramic has good stability and reliability and broad application prospects in the field of energy storage.


image file: d4qi02126f-f7.tif
Fig. 7 (a) Frequency and (c) temperature-dependence of PE loops of BCZT-BMS8-TSS ceramic; (b) and (d) show the values of W, Wrec, and η as a function of frequency and temperature, respectively.

In order to further evaluate the actual working ability of the BCZT-BMS8-TSS ceramic, the charge–discharge performance of the ceramic is evaluated from overdamped and underdamped discharge current curves. The underdamped discharge current curve at different electric field intensities is shown in Fig. 8(a). The relationship between current density (CD = Imax/S, S being the sample electrode area) and power density (PD = EImax/2S) and electric field intensity is shown in Fig. 8(b).6,28 The values of CD and PD increase monotonically with an increase in electric field intensity from 52 A cm−2 and 0.52 MW cm−3 at 20 kV cm−1 to 371.6 A cm−2 and 29.73 MW cm−3 at 160 kV cm−1, respectively. In addition, in the temperature range of 333–453 K (Fig. 8(c and d)), the values of CD and PD decreased slightly with an increase in temperature, and the change rate is controlled within ±4.8%.


image file: d4qi02126f-f8.tif
Fig. 8 The underdamped discharge current curves at (a) room temperature and (c) the temperature range of 333–453 K, and (b) and (d) the calculated CD and PD; the overdamped pulsed discharge current curves at (e) room temperature and (f) the temperature range of 333–453 K, and the inset of (e) shows the corresponding WD as a function of time for BCZT-BMS8-TSS ceramic.

Fig. 8(e) shows the overdamped discharge current curves of the BCZT-BMS8-TSS ceramic at different electric fields. The discharge energy density (WD) can be calculated from the formula of image file: d4qi02126f-t10.tif, where R (200 Ω) and V represent the load resistance and sample volume, respectively.42,63 The inset of Fig. 8(e) is the curve between WD and time (t). With the gradual increase in electric field intensity, the maximum current (Imax) and WD value rapidly increase. The time corresponding to the 90% saturation WD value is defined as the discharge time (t0.9), and the shorter the t0.9 of the capacitor, the more favorable the application in the pulse power supply system. According to the overdamped discharge current curve, WD is 0.62 J cm−3 at 160 kV cm−1, and the ultra-fast discharge time of t0.9 ∼ 142 ns. It should be noted that there is a significant difference between WD and Wrec when the electric field intensity is the same, that is,the Wrec measured from the PE loops is higher than that of the WD measured from the charging–discharging curves, mainly due to the difference of testing frequency. The BCZT-BMS8-TSS ceramic shows both high ESP and fast discharging speed, which is superior to the most reported lead-free ceramics in terms of comprehensive energy storage performance (Fig. 3(d)).

4. Conclusion

In this work, a record high breakdown strength (Eb) of ∼830 kV cm−1 has been achieved in BT-based ceramics via composition optimized and a two-step sintering (TSS) process. The results show that the introduction of BMS can effectively destroy the ferroelectric long order and form PNRs, which improves the dielectric behavior and polarization response process of the ceramics. Meanwhile, a finer grain size and higher density of ceramics are achieved by the TSS process. In addition, the doping of wider band gap MgO and SnO2 can further enlarge the Eg of the optimized ceramics, and the activation energies of the bulk and grain boundary counterparts significantly increased in the BCZT-BMS8-TSS ceramic. Benefiting from the above optimization, an unprecedented Wrec ∼ 10.53 J cm−3 and η ∼ 85.71% are achieved in the BCZT-BMS8-TSS ceramic, which is superior to the most reported ceramics. Moreover, all the samples exhibited excellent thermal/frequency stability and charge–discharge performance. In summary, the collaborative optimization strategy within this study is effective to enhance the comprehensive performance of energy storage ceramics, which provides a feasible way to develop other high ESP materials.

Author contributions

Changhao Wang and Dandan Han: conceptualization, methodology, investigation, data curation, writing – original draft, visualization, and funding acquisition. Jiaxi Hao and Longxiao Duan: investigation, formal analysis, and data curation. Jianfan Zhang, Wenfeng Yue and Zhenhao Fan: data curation. Raz Muhammad and Fanxu Meng: writing – review & editing. Dawei Wang: supervision and writing – review & editing.

Data availability

The data that support the findings of this study are available from the corresponding author and first author upon reasonable request.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the Projects of the Jilin Provincial Science and Technology Department (Grant No. YDZJ202201ZYTS420), the Projects of the Jilin Provincial Education Department (Grant No. JJKH20230298KJ), and the National Science Foundation for Yong Scientists China (Grant No. 62004081). The authors acknowledge the assistance of the JLICT Center of Characterization and Analysis.

References

  1. F. Yan, J. Qian, S. Wang and J. Zhai, Progress and outlook on lead-free ceramics for energy storage applications, Nano Energy, 2024, 123, 109394 CrossRef CAS.
  2. S. Wang, F. Yan, J. Qian, G. Ge, Z. Fu, Z. Pan, F. Zhang, J. Lin, K. Zeng, C. Chen, B. Shen, Z. Liu and J. Zhai, Temperature stability lock of high-performance lead-free relaxor ferroelectric ceramics, Energy Storage Mater., 2024, 66, 103155 CrossRef.
  3. A. Rahman, A. Xie, T. Li, Y. Zhang, M. Habib, X. Xinchun, L. Liu, X. Jiang and R. Zuo, An easy and effective strategy to design a composition in the vicinity of triple point for high energy density in dielectric ceramics, Chem. Eng. J., 2024, 486, 150091 CrossRef CAS.
  4. H. Liu, Z. Sun, J. Zhang, H. Luo, Y. Zhang, A. Sanson, M. Hinterstein, L. Liu, J. C. Neuefeind and J. Chen, Chemical framework to design linear-like relaxors toward capacitive energy storage, J. Am. Chem. Soc., 2024, 146(5), 3498–3507 CrossRef CAS PubMed.
  5. X. Zhou, G. Xue, Y. Su, H. Luo, Y. Zhang, D. Wang and D. Zhang, Optimized dielectric energy storage performance in ZnO-modified Bi0.5Na0.5TiO3-Sr0.7Bi0.20.1TiO3 ceramics with composite structure and element segregation, Chem. Eng. J., 2023, 458, 141449 CrossRef CAS.
  6. X. Zhu, Y. Gao, P. Shi, R. Kang, F. Kang, W. Qiao, J. Zhao, Z. Wang, Y. Yuan and X. Lou, 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 CrossRef CAS.
  7. M. Zhang, S. Lan, B. B. Yang, H. Pan, Y. Q. Liu, Q. H. Zhang, J. L. Qi, D. Chen, H. Su and D. Yi, Ultrahigh energy storage in high-entropy ceramic capacitors with polymorphic relaxor phase, Science, 2024, 384(6692), 185–189 CrossRef CAS PubMed.
  8. J. Zha, J. Liu, Y. Yang, X. Lu, X. Hu, S. Yan, Z. Wu, M. Zhou, F. Huang, X. Ying and J. Zhu, High energy storage performance of (1-x)Ba0.5Sr0.5TiO3-xK0.5Na0.5NbO3 ceramics via a combined strategy of fine grains and multiphase polar nanoregions, Chem. Eng. J., 2024, 486, 150441 CrossRef CAS.
  9. Y. Li, M. Y. Tang, Z. G. Zhang, Q. Li, J. L. Li, Z. Xu, G. Liu and F. Li, BaTiO3-based ceramics with high energy storage density, Rare Met., 2023, 42(4), 1261–1273 CrossRef CAS.
  10. Y. B. Adediji, A. M. Adeyinka, D. I. Yahya and O. V. Mbelu, A review of energy storage applications of lead-free BaTiO3-based dielectric ceramic capacitors, Energy, Ecol. Environ., 2023, 8(5), 401–419 CrossRef CAS.
  11. L. Chen, F. Li, B. Gao, C. Zhou, J. Wu, S. Deng, H. Liu, H. Qi and J. Chen, Excellent energy storage and mechanical performance in hetero-structure BaTiO3-based relaxors, Chem. Eng. J., 2023, 452, 139222 CrossRef CAS.
  12. M. Bai, W. Qiao, J. Mei, R. Kang, Y. Gao, Y. Wu, Y. Hu, Y. Li, X. Hao, J. Zhao, H. Hu and X. Lou, High-performance energy storage in BaTiO3-based oxide ceramics achieved by high-entropy engineering, J. Alloys Compd., 2024, 970, 172671 CrossRef CAS.
  13. C. Wang, W. Cao, C. Liang, H. Zhao, C. Cheng, S. Huang, Y. Yu and C. Wang, Ultrahigh energy-storage density of BaTiO3-based ceramics via the interfacial polarization strategy, ACS Appl. Mater. Interfaces, 2023, 15(36), 42774–42783 CrossRef CAS PubMed.
  14. C. Long, W. Zhou, H. Song, K. Zheng, W. Ren, H. Wu, X. Ding and L. Liu, Simultaneously realizing ultrahigh energy storage density and efficiency in BaTiO3-based dielectric ceramics by creating highly dynamic polar nanoregions and intrinsic conduction, Acta Mater., 2023, 256, 119135 CrossRef CAS.
  15. S. Yang, D. Zeng, Q. Dong, Y. Pan, P. Nong, M. Xu, X. Chen, X. Li and H. Zhou, Enhancement of energy storage performances in BaTiO3-based ceramics via introducing Bi(Mg2/3Sb1/3)O3, J. Energy Storage, 2024, 78, 110102 CrossRef.
  16. G. Wang, Z. Lu, Y. Li, L. Li, H. Ji, A. Feteira, D. Zhou, D. Wang, S. Zhang and I. M. Reaney, Electroceramics for high-energy density capacitors: current status and future perspectives, Chem. Rev., 2021, 121(10), 6124–6172 CrossRef CAS PubMed.
  17. Y. Huan, D. Gui, C. Li, T. Wei, L. Wu, X. Wang, X. Wang and Z. Cheng, Simultaneously enhanced energy storage performance and luminance resistance in (K0.5Na0.5)NbO3-based ceramics via synergistic optimization strategy, J. Adv. Ceram., 2024, 13(1), 34–43 CrossRef CAS.
  18. W. Wang, L. Zhang, C. Li, D. O. Alikin, V. Y. Shur, X. Wei, F. Gao, H. Du and L. Jin, Effective strategy to improve energy storage properties in lead-free (Ba0.8Sr0.2)TiO3-Bi(Mg0.5Zr0.5)O3 relaxor ferroelectric ceramics, Chem. Eng. J., 2022, 446, 137389 CrossRef CAS.
  19. M. Zhang, H. Yang, Y. Lin, Q. Yuan and H. Du, Significant increase in comprehensive energy storage performance of potassium sodium niobate-based ceramics via synergistic optimization strategy, Energy Storage Mater., 2022, 45, 861–868 CrossRef.
  20. L. Yang, X. Kong, F. Li, H. Hao, Z. Cheng, H. Liu, J. F. Li and S. Zhang, Perovskite lead-free dielectrics for energy storage applications, Prog. Mater. Sci., 2019, 102, 72–108 CrossRef CAS.
  21. Z. Shen, H. Liu, Y. Shen, J. Hu, L. Chen and C. Nan, Machine learning in energy storage materials, Interdiscip. Mater., 2022, 1, 175–195 Search PubMed.
  22. C. Wang, H. Li, Q. Zhang, J. Zhao, R. Muhammad, D. Han and D. Wang, Ultrahigh energy storage density in Ba0.85Ca0.15Zr0.1Ti0.9O3-based lead-free ceramics by introducing a relaxor end-member, J. Eur. Ceram. Soc., 2023, 43(15), 6844–6853 CrossRef CAS.
  23. G. Yan, L. Xu, B. Fang, S. Zhang, X. Lu, X. Zhao and J. Ding, Achieving high pulse charge–discharge energy storage properties and temperature stability of (Ba0.98−xLi0.02Lax)(Mg0.04Ti0.96)O3 lead-free ceramics via bandgap and defect engineering, Chem. Eng. J., 2022, 450, 137814 CrossRef CAS.
  24. L. Chen, S. Deng, H. Liu, J. Wu, H. Qi and J. Chen, Giant energy-storage density with ultrahigh efficiency in lead-free relaxors via high-entropy design, Nat. Commun., 2022, 13(1), 3089 CrossRef CAS PubMed.
  25. W. Huang, Y. Chen, X. Li, G. Wang, J. Xia and X. Dong, Superior energy storage performances achieved in (Ba,Sr)TiO3-based bulk ceramics through composition design and core-shell structure engineering, Chem. Eng. J., 2022, 444, 135523 CrossRef CAS.
  26. Z. Dai, J. Xie, Z. Chen, S. Zhou, J. Liu, W. Liu, Z. Xi and X. Ren, Improved energy storage density and efficiency of (1−x)Ba0.85Ca0.15Zr0.1Ti0.9O3-xBiMg2/3Nb1/3O3 lead-free ceramics, Chem. Eng. J., 2021, 410, 128341 CrossRef CAS.
  27. H. Zhao, X. Yang, D. Pang and X. Long, Enhanced energy storage efficiency by modulating field-induced strain in BaTiO3-Bi(Ni2/3Ta1/3)O3 lead-free ceramics, Ceram. Int., 2021, 47(16), 22734–22740 CrossRef CAS.
  28. C. Zuo, S. Yang, Z. Cao, H. Yu and X. Wei, Excellent energy storage and hardness performance of Sr0.7Bi0.2TiO3 ceramics fabricated by solution combustion-synthesized nanopowders, Chem. Eng. J., 2022, 442, 136330 CrossRef CAS.
  29. H. Liu, Z. Sun, J. Zhang, H. Luo, Q. Zhang, Y. Yao, S. Deng, H. Qi, J. Liu, L. C. Gallington, J. C. Neuefeind and J. Chen, Chemical design of Pb-free relaxors for giant capacitive energy storage, J. Am. Chem. Soc., 2023, 145(21), 11764–11772 CrossRef CAS PubMed.
  30. W. Qin, M. Zhao, Z. Li, D. Zhang, M. Zhang, Y. Xu, L. Jin and Y. Yan, High energy storage and thermal stability under low electric field in Bi0.5Na0.5TiO3-modified BaTiO3-Bi(Zn0.25Ta0.5)O3 ceramics, Chem. Eng. J., 2022, 443, 136505 CrossRef CAS.
  31. W. Cao, R. Lin, X. Hou, L. Li, F. Li, D. Bo, B. Ge, D. Song, J. Zhang, Z. Cheng and C. Wang, Interfacial polarization restriction for ultrahigh energy–storage density in lead–free ceramics, Adv. Funct. Mater., 2023, 33(29), 2301027 CrossRef CAS.
  32. A. M. Babeer and A. E.-r. Mahmoud, Suppression of the ferroelectric phase for (K0.5Na0.5)NbO3 ceramic by A/B-sites disorder for enhancement the energy storage properties and dielectric breakdown strength, Mater. Today Commun., 2023, 36, 106606 CrossRef CAS.
  33. M. Yin, G. J. Bai, P. Li, J. G. Hao, W. Li, W. F. Han, Y. C. Li, C. M. Wang, G. R. Li and P. Fu, Achieving ultrahigh energy storage properties with superior stability in novel (Ba(1−x)Bix)(Ti(1−x)Zn0.5xSn0.5x)O3 relaxor ferroelectric ceramics via chemical modification, Chem. Eng. J., 2023, 460, 141724 CrossRef CAS.
  34. F. Yan, H. Bai, G. Ge, J. Lin, C. Shi, K. Zhu, B. Shen, J. Zhai and S. Zhang, Composition and structure optimized BiFeO3−SrTiO3 lead–free ceramics with ultrahigh energy storage performance, Small, 2022, 18(10), 2106515 CrossRef CAS PubMed.
  35. S. Yang, C. Zuo, F. Du, L. Chen, W. Jie and X. Wei, Submicron Sr0.7Bi0.2TiO3 dielectric ceramics for energy storage via a two-step method aided by cold sintering process, Mater. Des., 2023, 225, 111447 CrossRef CAS.
  36. J. Liu, Y. Ding, C. Li, W. Bai, P. Zheng, S. Wu, J. Zhang, Z. Pan and J. Zhai, A synergistic two-step optimization design enables high capacitive energy storage in lead-free Sr0.7Bi0.2TiO3-based relaxor ferroelectric ceramics, J. Mater. Chem. A, 2023, 11(2), 609–620 RSC.
  37. X. H. Wang, X. Y. Deng, H. L. Bai, H. Zhou, W. G. Qu, L. T. Li and I. W. Chen, Two–step sintering of ceramics with constant grain–size, II: BaTiO3 and Ni-Cu-Zn ferrite, J. Am. Ceram. Soc., 2005, 89(2), 438–443 CrossRef.
  38. X. Dong, X. Li, X. Chen, J. Wu and H. Zhou, Simultaneous enhancement of polarization and breakdown strength in lead-free BaTiO3-based ceramics, Chem. Eng. J., 2021, 409, 128231 CrossRef CAS.
  39. C. Long, W. zhou, L. Liu, H. Song, H. Wu, K. Zheng, W. Ren and X. Ding, Achieving excellent energy storage performances and eminent charging-discharging capability in donor (1-x)BT-x(BZN-Nb) relaxor ferroelectric ceramics, Chem. Eng. J., 2023, 459, 141490 CrossRef CAS.
  40. D. Hu, Z. Pan, X. Tan, F. Yang, J. Ding, X. Zhang, P. Li, J. Liu, J. Zhai and H. Pan, Optimization the energy density and efficiency of BaTiO3-based ceramics for capacitor applications, Chem. Eng. J., 2021, 409, 127375 CrossRef CAS.
  41. H. Chen, X. Wang, X. Dong, Y. Pan, J. Wang, L. Deng, Q. Dong, H. Zhang, H. Zhou and X. Chen, Adjusting the energy-storage characteristics of 0.95NaNbO3-0.05Bi(Mg0.5Sn0.5)O3 ceramics by doping linear perovskite materials, ACS Appl. Mater. Interfaces, 2022, 14(22), 25609–25619 CrossRef CAS PubMed.
  42. M. Sun, X. Wang, P. Li, J. Du, P. Fu, J. Hao, W. Li and J. Zhai, Realizing ultrahigh breakdown strength and ultrafast discharge speed in novel barium titanate-based ceramics through multicomponent compounding strategy, J. Eur. Ceram. Soc., 2023, 43(3), 74–985 CrossRef.
  43. D. Han, L. Duan, C. Wang, L. Yuan, R. Muhammad, P. Ma, F. Meng, D. Wang and F. Meng, Effective regulation of breakdown strength through the synergistic effect of defect chemistry and energy band engineering in Ba0.85Ca0.15Zr0.1Ti0.9O3-based lead-free ceramics, Inorg. Chem. Front., 2024, 11(7), 2058–2070 RSC.
  44. Q. Hu, Y. Tian, Q. Zhu, J. Bian, L. Jin, H. Du, D. O. Alikin, V. Y. Shur, Y. Feng, Z. Xu and X. Wei, Achieve ultrahigh energy storage performance in BaTiO3-Bi(Mg1/2Ti1/2)O3 relaxor ferroelectric ceramics via nano-scale polarization mismatch and reconstruction, Nano Energy, 2020, 67, 104264 CrossRef CAS.
  45. W. Yang, H. Zeng, F. Yan, J. Lin, G. Ge, Y. Cao, W. Du, K. Zhao, G. Li, H. Xie and J. Zhai, Superior energy storage properties in NaNbO3-based ceramics via synergistically optimizing domain and band structures, J. Mater. Chem. A, 2022, 10(21), 11613–11624 RSC.
  46. L. Liu, B. Chu, P. Li, P. Fu, J. Du, J. Hao, W. Li and H. Zeng, Achieving high energy storage performance and ultrafast discharge speed in SrTiO3-based ceramics via a synergistic effect of chemical modification and defect chemistry, Chem. Eng. J., 2022, 429, 132548 CrossRef CAS.
  47. R. Kang, Z. Wang, W. Yang, Y. Zhao, L. Zhang and X. Lou, Enhanced energy storage performance of Bi0.5Na0.5TiO3-based ceramics via grain tuning and relaxor construction, Chem. Eng. J., 2023, 455, 140924 CrossRef CAS.
  48. H. Yuan, X. Fan, Z. Zheng, M. Zhao, L. Zhao, K. Zhu and J. Wang, Simultaneous enhancement of breakdown strength, recoverable energy storage density and efficiency in antiferroelectric AgNbO3 ceramics via multi-scale synergistic design, Chem. Eng. J., 2023, 456, 141023 CrossRef CAS.
  49. W. B. Li, D. Zhou, W. F. Liu, J. Z. Su, F. Hussain, D. W. Wang, G. Wang, Z. L. Lu and Q. P. Wang, High-temperature BaTiO3-based ternary dielectric multilayers for energy storage applications with high efficiency, Chem. Eng. J., 2021, 414, 128760 CrossRef CAS.
  50. Z. G. Liu, M. D. Li, Z. H. Tang and X. G. Tang, Enhanced energy storage density and efficiency in lead-free Bi(Mg1/2Hf1/2)O3-modified BaTiO3 ceramics, Chem. Eng. J., 2021, 418, 129379 CrossRef CAS.
  51. L. Liu, M. Knapp, H. Ehrenberg, L. Fang, H. Fan, L. A. Schmitt, H. Fuess, M. Hoelzel, H. Dammak, M. P. Thi and M. Hinterstein, Average vs. local structure and composition-property phase diagram of K0.5Na0.5NbO3-Bi1/2Na1/2TiO3 system, J. Eur. Ceram. Soc., 2017, 37(4), 1387–1399 CrossRef CAS.
  52. X. Dong, X. Li, H. Chen, C. Sun, J. Shi, F. Pang, X. Chen and H. Zhou, Effective strategy to realise excellent energy storage performances in lead-free barium titanate-based relaxor ferroelectric, Ceram. Int., 2021, 47(5), 6077–6083 CrossRef CAS.
  53. I. A. Santos and J. A. Eiras, Phenomenological description of the diffuse phase transition in ferroelectrics, J. Phys.: Condens. Matter, 2001, 13(50), 11733 CrossRef CAS.
  54. Z. Che, L. Ma, G. Luo, C. Xu, Z. Cen, Q. Feng, X. Chen, K. Ren and N. Luo, Phase structure and defect engineering in (Bi0.5Na0.5)TiO3-based relaxor antiferroelectrics toward excellent energy storage performance, Nano Energy, 2022, 100, 107484 CrossRef CAS.
  55. B. Yang, J. Zhang, X. Lou, Y. Gao, P. Shi, Y. Yang, M. Yang, J. Cui, L. Wei and S. Sun, Enhancing comprehensive energy storage properties in tungsten bronze Sr0.53Ba0.47Nb2O6-based lead-free ceramics by B-Site doping and relaxor tuning, ACS Appl. Mater. Interfaces, 2022, 14(30), 34855–34866 CrossRef CAS PubMed.
  56. W. Cao, R. Lin, P. Chen, F. Li, B. Ge, D. Song, J. Zhang, Z. Cheng and C. Wang, Phase and band structure engineering via linear additive in NBT-ST for excellent energy storage performance with superior thermal stability, ACS Appl. Mater. Interfaces, 2022, 14(48), 54051–54062 CrossRef CAS PubMed.
  57. Y. Ding, W. Que, J. He, W. Bai, P. Zheng, P. Li, J. Zhang and J. Zhai, Realizing high-performance capacitive energy storage in lead-free relaxor ferroelectrics via synergistic effect design, J. Eur. Ceram. Soc., 2022, 42(1), 129–139 CrossRef CAS.
  58. T. Gao, Q. Chen, X. Li, R. Lang, Z. Tan, Q. Li, Z. Tang, J. Xing and J. Zhu, Ultrafast charge–discharge and enhanced energy storage properties of lead–free K0.5Na0.5NbO3-based ceramics, J. Am. Ceram. Soc., 2024, 107(6), 4109–4121 CrossRef CAS.
  59. P. Zhao, L. Li and X. Wang, BaTiO3-NaNbO3 energy storage ceramics with an ultrafast charge-discharge rate and temperature-stable power density, Microstructures, 2023, 3(1), 2023002 CAS.
  60. D. Han, C. Wang, Z. Zeng, X. Wei, P. Wang, Q. Liu, D. Wang and F. Meng, Ultrahigh energy efficiency of (1-x)Ba0.85Ca0.15Zr0.1Ti0.9O3-xBi(Mg0.5Sn0.5)O3 lead-free ceramics, J. Alloys Compd., 2022, 902, 163721 CrossRef CAS.
  61. I. W. Chen and X. H. Wang, Sintering dense nanocrystalline ceramics without final-stage grain growth, Nature, 2000, 404(6774), 168–171 CrossRef CAS PubMed.
  62. X. Qiao, F. Zhang, D. Wu, B. Chen, X. Zhao, Z. Peng, X. Ren, P. Liang, X. Chao and Z. Yang, Superior comprehensive energy storage properties in Bi0.5Na0.5TiO3-based relaxor ferroelectric ceramics, Chem. Eng. J., 2020, 388, 124158 CrossRef CAS.
  63. S. Luan, S. Si, L. Zhang, P. Wang, G. Jian, J. Yang, S. Yu, R. Sun, Z. Fu and X. Cao, Fabrication of BaTiO3 nano-powders with high tetragonality via two-step assisted rotary furnace calcination for MLCC applications, Ceram. Int., 2023, 49(8), 12529–12539 CrossRef CAS.
  64. H. Singh, A. Kumar and K. L. Yadav, Structural, dielectric, magnetic, magnetodielectric and impedance spectroscopic studies of multiferroic BiFeO3-BaTiO3 ceramics, Mater. Sci. Eng., B, 2011, 176(7), 540–547 CrossRef CAS.
  65. X. Li, X. Chen, J. Sun, M. Zhou and H. Zhou, Novel lead-free ceramic capacitors with high energy density and fast discharge performance, Ceram. Int., 2020, 46(3), 3426–3432 CrossRef CAS.
  66. Q. Yuan, M. Chen, S. Zhan, Y. Li, Y. Lin and H. Yang, Ceramic-based dielectrics for electrostatic energy storage applications: fundamental aspects, recent progress, and remaining challenges, Chem. Eng. J., 2022, 446, 136315 CrossRef CAS.
  67. Y. Xu and M. A. A. Schoonen, The absolute energy positions of conduction and valence bands of selected semiconducting minerals, Am. Mineral., 2000, 85, 543–556 CrossRef CAS.
  68. W. Wang, L. Zhang, C. Li, D. O. Alikin, V. Ya. Shur, X. Wei, F. Gao, H. Du and L. Jin, Effective strategy to improve energy storage properties in lead-free (Ba0.8Sr0.2)TiO3-Bi(Mg0.5Zr0.5)O3 relaxor ferroelectric ceramics, Chem. Eng. J., 2022, 486, 150091 Search PubMed.
  69. R. Muhammad, Y. Iqbal and I. M. Reaney, BaTiO3-Bi(Mg2/3Nb1/3) O3 ceramics for high–temperature capacitor applications, J. Am. Ceram. Soc., 2016, 99(6), 2089–2095 CrossRef CAS.
  70. R. Muhammad and A. Khesro, Influence of A-site nonstoichiometry on the electrical properties of BT-BMT, J. Am. Ceram. Soc., 2017, 100(3), 1091–1097 CrossRef CAS.

This journal is © the Partner Organisations 2024
Click here to see how this site uses Cookies. View our privacy policy here.