Next Article in Journal
Zinc Oxide Nanostructure Deposition into Sub-5 nm Vertical Mesopores in Silica Hard Templates by Atomic Layer Deposition
Next Article in Special Issue
Piezoelectric Actuators in Smart Engineering Structures Using Robust Control
Previous Article in Journal
An Innovative Design to Enhance Osteoinductive Efficacy and Biomechanical Behavior of a Titanium Dental Implant
Previous Article in Special Issue
A Polymorph of Dipeptide Halide Glycyl-L-Alanine Hydroiodide Monohydrate: Crystal Structure, Optical Second Harmonic Generation, Piezoelectricity and Pyroelectricity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ceramic-Based Dielectric Materials for Energy Storage Capacitor Applications

by
Srinivas Pattipaka
1,
Yeseul Lim
1,
Yong Hoon Son
1,
Young Min Bae
1,
Mahesh Peddigari
2 and
Geon-Tae Hwang
1,*
1
Department of Materials Science and Engineering, Pukyong National University, 45, Yongso-ro, Nam-Gu, Busan 48513, Republic of Korea
2
Department of Physics, Indian Institute of Technology Hyderabad, Kandi 502284, Telangana, India
*
Author to whom correspondence should be addressed.
Materials 2024, 17(10), 2277; https://doi.org/10.3390/ma17102277
Submission received: 8 April 2024 / Revised: 1 May 2024 / Accepted: 8 May 2024 / Published: 11 May 2024

Abstract

:
Materials offering high energy density are currently desired to meet the increasing demand for energy storage applications, such as pulsed power devices, electric vehicles, high-frequency inverters, and so on. Particularly, ceramic-based dielectric materials have received significant attention for energy storage capacitor applications due to their outstanding properties of high power density, fast charge–discharge capabilities, and excellent temperature stability relative to batteries, electrochemical capacitors, and dielectric polymers. In this paper, we present fundamental concepts for energy storage in dielectrics, key parameters, and influence factors to enhance the energy storage performance, and we also summarize the recent progress of dielectrics, such as bulk ceramics (linear dielectrics, ferroelectrics, relaxor ferroelectrics, and anti-ferroelectrics), ceramic films, and multilayer ceramic capacitors. In addition, various strategies, such as chemical modification, grain refinement/microstructure, defect engineering, phase, local structure, domain evolution, layer thickness, stability, and electrical homogeneity, are focused on the structure–property relationship on the multiscale, which has been thoroughly addressed. Moreover, this review addresses the challenges and opportunities for future dielectric materials in energy storage capacitor applications. Overall, this review provides readers with a deeper understanding of the chemical composition, physical properties, and energy storage performance in this field of energy storage ceramic materials.

1. Introduction

Energy storage devices such as batteries, electrochemical capacitors, and dielectric capacitors play an important role in sustainable renewable technologies for energy conversion and storage applications [1,2,3]. Particularly, dielectric capacitors have a high power density (~107 W/kg) and ultra-fast charge–discharge rates (~milliseconds) when compared to electrochemical capacitors and batteries (Figure 1b) [2,3,4,5,6,7,8,9,10,11,12,13]. These advantages of dielectric capacitors make them promising for applications in power electronics and pulsed power systems, as shown in Figure 1a. For instance, more than three trillion multilayer ceramic capacitors (MLCCs) are manufactured annually and are used in cell phones or electric vehicles [6,7,8,9,14,15]. However, dielectric capacitors have a lower energy storage density of 10−2 to 10−1 Wh/kg than electrochemical capacitors and batteries, which limits their practical applications. Therefore, high-performance dielectric materials in terms of high energy storage density, high energy efficiency, fast charge–discharge capabilities, better thermal or frequency stability, fatigue resistance, lifetime reliability, equivalent series resistance, and low manufacturing costs are needed for power electronics and pulse power applications.
The storage performance depends on the charge accumulation in dielectric materials, which are a key component of capacitors. Dielectric materials, including organic (polyvinylidene fluoride (PVDF), biaxially oriented polypropylene (BOPP), polyimide (PI), etc.), and inorganic (ceramics, glass, and glass-based ceramics) materials, have been widely investigated to improve the energy storage performance [9,16,17,18,19,20]. In recent years, significant improvements to dielectric materials have been made, although each material still has limitations. The polymers offer a high breakdown strength (BDS), low relative dielectric permittivity, and weak thermal stability, making dielectric materials for energy storage a long-term goal. Meanwhile, ceramic-based dielectric materials are popular research topics due to their application in energy storage, adaptability to various environments, fundamentality, and other factors. Therefore, the topic of dielectrics will be discussed further in this review. These materials are classified into linear dielectrics (LDs), ferroelectrics (FEs), antiferroelectrics (AFEs), and relaxor ferroelectrics (RFEs) [17,20]. They are considered viable candidates for energy storage due to their differing properties in BDS and polarization, which primarily influence energy storage performance.
This review paper presents fundamental concepts of energy storage in dielectric capacitors, including an introduction to dielectrics and key parameters to enhance energy storage responses. We also summarize recent progress in dielectrics, such as bulk ceramics, ceramic films, and multilayer ceramic capacitors, including the phase, local structure, microstructure, domain evolution, layer thickness, stability, and electrical homogeneity; fabrication methods, dopants/composites, and various strategies for enhancing energy storage properties in dielectric capacitors are also briefly discussed.

2. Fundamental Concepts for Energy Storage in a Dielectric Capacitor

2.1. Dielectric Capacitor

A parallel plate capacitor is composed of two parallel conducting plates that are separated by a ceramic layer, as schematically shown in Figure 2. When a dielectric capacitor is placed in an external electric field, the electric dipoles will be displaced and oriented due to polarization (Figure 2b). The capacitance of a dielectric capacitor (C) is the ability to store electric charge and is given by the following equation:
C = Q V
where Q is the charge and V is the voltage applied to the capacitor.
According to Gauss’s law,
V = Q d ε 0 ε r A
The capacitance of a parallel plate capacitor can be calculated in terms of the sample area and thickness via comparing Equations (1) and (2).
C = ε 0 ε r A d
where ε0 is the permittivity of free space, εr is relative dielectric permittivity, A is the area of metal plates, and d is the thickness of the ceramic sample (Figure 2a).

2.2. Evaluation of Energy Storage Performance

The energy storage density (W) of a linear dielectric material is determined with the following equation [21]:
W = 1 2 ε 0 ε r E 2
where ε0 is the permittivity of free space, εr is dielectric permittivity, and E is the applied electric field. In contrast, the nonlinear dielectric materials (FEs, AFEs, and RFEs) exhibit energy loss. Therefore, the total energy storage density (Wtot), recoverable energy density (Wrec), and energy storage efficiency (η) of these materials are calculated from the hysteresis loops as follows [22,23,24]:
W t o t = 0 P m E d P   ( C h a r g i n g )
W r e c = P r P m E d P   ( D i s c h a r g i n g )
η = W r e c W r e c + W l o s s × 100 %
where E is the applied electric field, P is polarization, Pmax is maximum polarization, Pr is remnant polarization, and Wloss is energy loss, as schematically shown in Figure 3.

2.3. Key Parameters for Energy Storage Performance

2.3.1. Energy Storage Density and Efficiency

Wrec and η are the most important parameters for evaluating the energy storage performance of dielectric materials, which are related to dielectric permittivity and polarization. A high Wrec of dielectric materials means that more energy can be stored in a given volume, promoting miniaturization and lightweight and low-cost materials being utilized in consumer power electronics and pulse power systems. It can be concluded from Equations (4)–(7) and Figure 3 that a higher εr, Pmax, and BDS lead to higher energy density, whereas low dielectric or hysteresis losses and low Pr improve energy storage efficiency in dielectric materials. Moreover, the material should have low electronic or ionic conductivity to resist higher electric fields.

2.3.2. Polarization Difference

The energy storage density and efficiency of a ceramic capacitor’s are mostly related to the shape of the P-E loop due to the area under the curve providing the Wrec (Figure 3). Therefore, the energy storage performance depends on the value of ΔPP = Pmax Pr), and the Wrec increases with ΔP [25,26]. However, some of the stored energy in dielectrics will be dissipated during the depolarization/discharge process, which will be equal to the area of the P-E loop (i.e., Wloss can be seen in Figure 3) [27,28]. Such energy loss causes heat generation, consequently deteriorating the capacitor’s thermal stability and lifespan. The heat generation is attributed to the dielectric loss (tanδ) and temperature rise (ΔT), as provided using the following equations:
w h = π ε 0 ε r E 2 t a n δ
T = f V e h A W h
where f is the driving frequency, Ve is the effective volume to an applied electric field, h is the heat transfer coefficient, and A is the total surface area of a sample [27,28]. Therefore, a low tanδ and a large ΔP (i.e., low Pr and high Pmax) are critical parameters for achieving high energy storage performance in ceramic capacitors.

2.3.3. Dielectric Breakdown Strength

The energy storage response of ceramic capacitors is also influenced by the Eb, as the Wrec is proportional to the E, as can be seen in Equation (6) [29]. The BDS is defined as the maximum electric field over which the electrical resistance of a dielectric significantly decreases. The Eb of these capacitors strongly depends on intrinsic (bandgap, grain size, phase, defect dipoles, material thickness, microstructure, and porosity) and extrinsic (working/environmental conditions and electrode configuration) properties [5,30,31,32,33]. Therefore, a dense microstructure is a critical factor for the fabrication of a high-quality ceramic capacitor to achieve greater capacitance under high electric fields. However, dielectric breakdown is caused by pores, cracks, interfaces, and compositional inhomogeneity [31,33,34]. The porosity in dielectrics affects the dielectric breakdown strength and can cause overheating and thermal stress, resulting in breakdowns at higher electric fields [31,33].

2.3.4. Discharge Time

The discharge time is another critical parameter for energy storage. The discharging speed of a ceramic capacitor is calculated in terms of the discharge time, represented by τ0.90. It is defined as the time required for a capacitor to discharge 90% of its stored energy. The discharge time is 0.15 µs at an infinite time, and it depends on the dielectric permittivity and thickness of the material, load resistance, and applied voltage [6]. The discharge time should be very short for pulsed power energy storage capacitor applications.

2.3.5. Reliability

Dielectric capacitors are interconnected with their embedded system and operating conditions, influenced by various factors such as temperature, frequency, and voltage fluctuations. Therefore, better reliability, often called high electric fatigue endurance, protects the physical integrity of pulsed power systems, particularly dielectric capacitors, during energy storage under harsh circumstances. The evaluation of the resistance to stimuli can be conducted through observing the distinct features of P-E loops under particular investigation conditions.

2.4. Categories of Dielectric Materials

Based on polarization versus the electric field response, dielectric materials are categorized into linear dielectrics (LDs) and nonlinear dielectrics (NLDs), such as ferroelectrics (FEs), anti-ferroelectrics (AFEs), and relaxor ferroelectrics (RFEs). Figure 4a–d show a schematic of the electric field-dependent polarization response and corresponding ferroelectric domain structures with dipole orientation for the LDs and NLDs. LDs display an almost linear polarization response via the application of an electric field, owing to the lack of permanent dipoles (Figure 4a). Ferroelectric materials display superior polarization responses even in the absence of an external electric field due to the presence of a net dipole moment. Therefore, they have a strong nonlinear relation with the applied electric field (Figure 4b). In AFE materials, the adjacent dipoles are oriented in antiparallel directions, resulting in zero net polarization. They show double hysteresis loops at higher electric fields caused by AFE to FE phase transitions (Figure 4c). In RFEs, the existence of polar nanoregions (PNRs)/nanodomains greatly reduces cooperative coupling between ferroelectric domains, which limits spontaneous polarization and, consequently, slim P-E loops (Figure 4d).

2.5. Energy-Storage Mechanism of the Materials

Ferroelectric materials are a fascinating class of dielectrics with unique properties, making them promising in the field of energy storage, conversion, and harvesting applications due to their electrical, mechanical, and thermal properties being intrinsically interrelated. All ferroelectrics are piezoelectric and pyroelectric materials, which make ferroelectrics extremely useful in multiple applications. The coupling of ferroelectric polarization to temperature, stress, and electric field enables various energy storage and conversion approaches that rely on diverse stimuli. The polarization is used in three ways, namely capacitive-energy storage (i.e., energy is stored in the form of polarization), piezoelectric-energy harvesting (i.e., vibration-induced stress on a piezoelectric material is converted into charge via a change in polarization), and pyroelectric-energy conversion (i.e., thermodynamic cycles can be utilized to convert temperature fluctuations into current) [35].
Based on the energy storage mechanism and the charge–discharge process, there is a substantial variation in the power density and energy density in dielectric capacitors, electrochemical capacitors, and batteries (see Figure 1b). Batteries offer higher energy density, but lower power density because of the slow movement of charges, which are used for long-term, stable energy supplies and applications with a maximum of 5 V [2,3,12,36]. Electrochemical capacitors have moderate power density and energy storage density with a slow charge–discharge rate and a low operating voltage (<3 V) [36]. Dielectric capacitors have high power density but limited energy storage density, with a more rapid energy transfer than electrochemical capacitors and batteries; this is because they store energy via dielectric polarization in response to the external electrical fields rather than chemical reactions [3,12,13,35]. Therefore, dielectric capacitors have received great interest due to their low price and high operating voltages (kV/MV range) for longer durations, making them ideal for a wide range of applications, including consumer electronics and advanced pulsed power devices.

3. Dielectric Materials for Energy Storage

3.1. Bulk Ceramics

3.1.1. Linear Dielectrics

LDs exhibit low energy loss, low relative dielectric permittivity, and a high breakdown electric field, and are promising for energy storage device applications under certain working conditions. Various LDs, such as Al2O3 [37], TiO2 [38], SrTiO3 (ST) [39,40,41], and CaTiO3 (CT) [41,42,43,44,45], have been reported to improve their energy storage performances. Pure ST ceramics exhibited a relative dielectric permittivity of 300, a breakdown electric field of 1600 kV/mm, and a dielectric loss of 0.01 at RT, and are utilized for integrated circuit applications [39,42,46]. Chemical modifications have been adopted to enhance the energy storage properties in ST ceramic capacitors. Notably, 2 mol% of Ca doping in the ST system was improved energy density of 1.95 J/cm3 and an efficiency of 72.3% at a breakdown field of 333 kV/cm, which is nearly three times higher than pure ST [41]. These improved energy storage properties in titanium-based ceramics are attributed to the insulation attenuation property caused by electronic hopping from the valence band to the conduction band. The substitution of Zr ions at the Ti site of Sr0.98Ca0.02TiO3 boosted the energy storage density to 2.77 J/cm3 and yielded an efficiency of 77.7% by reducing the dielectric loss and leakage current density, which is attributable to the higher chemical durability [47]. Mg-doped ST ceramics showed an enhanced Wrec of 1.86 J/cm3 and η of 72.3% at a BDS of 362 kV/cm by lowering the dielectric loss to 0.001 with a moderate dielectric constant of 280 [45]. Interestingly, a binary composite of CaZrO3-0.05SrTiO3 exhibited a high Wrec of 5 J/cm3 at 1000 kV/cm, caused by a low dielectric loss of 0.001 and dielectric constant of 35 [48]. It is well known that the BDS is directly proportional to the bandgap energy, and a higher bandgap energy enables a higher BDS [43,48]. Shay et al. [43] reported a binary composition of 0.8CaTiO3-0.2CaHfO3 (with 0.5 mol% of Mn doping) by modulating their bandgap energies, and showed a high Wrec of 9 J/cm3 at 1200 kV/cm (9.6 J/cm3 at 1300 kV/cm). In a similar vein, BaZrO3-CaTiO3 and SrZrO3-CaTiO3 binary compositions have shown improved energy storage performance [43,48].

3.1.2. Ferroelectrics

In comparison with LDs, FE materials show strong nonlinear behavior with high polarization, high dielectric permittivity, high energy loss, and a low BDS. Various rare earth elements and dopants (such as Sr, Ca, Nd, Mn, and Zr) were substituted at A/B-sites of the BT system to enhance the BDS and energy storage responses. Sr-doped BT (Ba1−xSrxTiO3, BST) ceramics were investigated, showing a high dielectric constant of 650, a low dielectric loss of 7.6 × 10−4 @ 1kHz, a low Wrec of 0.23 J/cm3, and the Curie temperature being lowered far below RT [49]. Choi et al. [50] reported a defect dipole engineering method to enhance the energy storage performance by co-doping Nd and Mn in Ba0.7Sr0.3TiO3 ceramics. Figure 5 presents a schematic illustration of a defect dipole concept between acceptor ions and oxygen vacancies in Ba0.7Sr0.3TiO3 ceramics. These defect dipoles with a uniform and small-grained microstructure enable a high difference between Pmax and PrP~10.39 µC/cm2) and capture electrons, improving the BDS to 110.6 kV/cm with co-doping of Nd and Mn; this in turn leads to improvements in the Wrec to 0.41 J/cm3 and a high η of 84.6% in Ba0.7Sr0.3TiO3 ceramics. Interestingly, Dong et al. [33] reported 1.6 wt% ZnO doped in Ba0.3Sr0.7TiO3 ceramics with an enhanced Wrec of 3.9 J/cm3 at 40 kV/mm. Taking a theoretical approach, Wang et al. [51] reported first-principles calculations and molecular dynamic simulations to study the effects of the chemical composition, phase under temperature, and electric fields on the ferroelectric and energy storage properties of ABO3 perovskite FEs. These simulation results revealed a Wrec of 2.8 J/cm3 and a η of 95% at Eb of 350 kV/cm in Ba0.6Sr0.4TiO3 ceramics, and, furthermore, a Wrec of 30 J/cm3 and a η of 92% obtained at an Eb of 2750 kV/cm in the same composition of Ba0.6Sr0.4TiO3. However, practically, a BDS on the order of a thousand kV/cm is not achievable in most FEs because of numerous defects, an internal mechanical field, internal stress, and the influence of crystallographic lattice constants, phase transition, and grain size. Song et al. [52] reported the effect of grain sizes from 0.5 µm to 5.6 µm in Ba0.6Sr0.4TiO3 ceramics to investigate the energy storage performance, and the samples with a grain size of 0.5 µm showed a high Wrec of 1.28 J/cm3 at an Eb of 243 kV/cm.

3.1.3. Anti-Ferroelectrics

Antiferroelectric materials differ from typical ferroelectrics in their distinctive crystal structure, with adjacent diploes aligned in opposite orientations. To generate a strong ferroelectric state, diploes are subjected to a high electric field in order to realign their polarization orientation. This results in the formation of double hysteresis loops which consist of a linear polarization response in the AFE state and a ferroelectric hysteresis loop in the FE state. The huge reversible polarization would increase the energy storage density. However, thermal runaway and high energy dissipation due to hysteresis remain major challenges in building high energy density AFEs. To improve energy storage properties, enhancing the linear polarization response area and decreasing hysteresis loss by changing the phase transition parameters is recommended.
PbZrO3 (PZ) AFE materials have been widely investigated due to their diverse phase transition features [53]. Chemical substitution affects reform polarization properties by altering the switching electric field between the AFE and FE phases. As per the phase diagram of La2O3-PbZrO3-PbTiO3 [54], Peixin et al. [55] reported the energy storage properties with the substitution of Ti4+ with Zr4+ at the B-site of in (Pb1−yLay)(ZrxTi1−x)O3 (PLZT) ceramics. The substitution of Zr4+ at Ti4+ can decrease the tolerance factor and improve the AFE properties. The P-E loops of PLZT AFEs become very slim with the substitution of the Zr concentration, and a high Wrec of 3.38 J/cm3 and a high η of 86.5% were achieved with the optimized composition of x = 0.9 and y = 0.07 (Figure 6). Similarly, the substitution of La3+ at Pb2+ (the A-site) of (Pb1−1.5xLax)(Zr0.5Sn0.43Ti0.07)O3 improved the AFE phase stability and provided slim P-E loops, resulting in the highest Wrec of 4.2 J/cm3 and a high η of 78% for the x = 0.03 composition [56]. On the basis of the phase diagram of PbZrO3-PbTiO3-PbSnO3 [57], Wang et al. [58] reported field-induced multiphase transitions (AFE-FE and FE-FE) at weak and high electric fields in (Pb0.98La0.02)(Zr0.55Sn0.45)0.995O3 AFE ceramics, yielding superior energy storage properties of a Wrec of 10.4 J/cm3 and a η of 87% at 400 kV/cm. Moreover, Liu et al. reported the substitution of Sr2+ in (Pb0.98-xLa0.02Srx)(Zr0.9Sn0.1)0.995O3 AFE ceramics to improve the BDS and the switching of electric fields between the AFE and FE phase, resulting in an ultrahigh Wrec of 11.18 J/cm3 and a high η of 82.2% [59].
In spite of the excellent features of AFE lead-based ceramics, various AFE lead-free ceramics have garnered attention due to environmental concerns. Zhao et al. [26] reported lead-free AFE AgNbO3 (AN) ceramics with a Ta substitution to improve their energy storage properties. Figure 7 presents the P-E loops of pure, Ta-doped AN ceramics and energy storage properties of Ag(Nb1−xTax)O3 ceramics as a function of the Ta concentration (x = 0 to 20 mol%). A high Wrec of 4.2 J/cm3 (260% higher than that of pure AN) and a η of 69% were achieved in Ag(Nb1−xTax)O3 ceramics for x = 0.15. The substitution of Ta into the Nb site improves antiferroelectricity due to the lower polarizability of B-site cations, and also reduces grain size and enhances density, resulting in a high BDS of 240 kV/cm (Figure 7c). Researchers recently investigated the underlying mechanism between AFE properties and the energy barrier (EB), where increased and decreased EB for the AFE-FE phase transition via the doping of Sm3+, Ca2+, and the co-doping of Sm3+/Ta5+ at the A- and A/B-sites of AN-based ceramics, which exhibited high Wrecs of 5.2, 4.87, and 3.55 J/cm3, respectively [60,61,62]. Luo et al. [63] reported a high Wrec of 6.3 J/cm3 and a high η of 90%, realized by the M2-M3 phase boundary, the stabilized AFE phase, the presence of relaxor properties, and slim double P-E loops. In a similar way, Li et al. [5] reported 0.55(Bi0.5Na0.5)TiO3-0.45(Bi0.2Sr0.7)TiO3 relaxor-antiferroelectric ceramics with a Wrec of 2.5 J/cm3 for bulk ceramics and 9.5 J/cm3 for multilayer ceramic capacitors, respectively. In addition, Qi et al. [64,65] fabricated 0.78(Bi0.5Na0.5)TiO3-0.22NaNbO3 and 0.76NaNbO3–0.24(Bi0.5Na0.5)TiO3 relaxor-antiferroelectric ceramics with giant energy storage properties as follows: a Wrec of 7.02 and 12.2 J/cm3 and a η of 85% and 69%, respectively. Instead of the chemical substitution/composition method, Wang et al. [66] utilized a hydrothermal method to enhance the energy storage performance of AN ceramics and form a fine-grain size of 3 µm, which resulted in a high BDS of 250 kV/cm.

3.1.4. Relaxor Ferroelectrics

Relaxor ferroelectric materials, a significant subclass of ferroelectric materials, have drawn the attention of researchers because of their intriguing and little-known physics since Smolenskii’s first discovery of the relaxor properties in a BaTiO3 (BT)-based system [67]. The RFEs are thought to be the most promising energy storage materials for applications in electrostatic energy storage because of their distinct and slim P-E loops, in contrast with regular ferroelectrics, and are beneficial for energy storage. It has been established that the vast differences between RFEs and FEs are closely related to the dynamics of their domain structure. The nanodomains/PNRs, which range in size from several nm to µm and are more responsive to external electric fields, are predicted to facilitate a moderate P and slight Pr in RFEs, and these features are expected to contribute to a high Wrec and η [68]. In this regard, various lead-based and lead-free perovskite RFEs, namely (Pb(Zn1/3Nb2/3)O3-PbTiO3 (PZN-PT) [69,70], Pb(Mg1/3Nb2/3) O3-PbTiO3 (PMN-PT) [70], (Pb, La)(Zr, Ti)O3 (PLZT) [71] and BT [72,73,74], (Na, K)NbO3 (KNN) [75,76], and (Bi, Na)TiO3 (BNT) [75,76], have been explored for energy storage applications, respectively.
In lead-based RFEs, the PLZT has received strong attention for energy storage applications because of their phase structure (paraelectric phase, rhombohedral FEs, tetragonal FEs, orthorhombic AFEs, and RFEs) through chemical composition design. It is observed that relaxor properties showing slim P-E loops can be obtained via the formation of a pseudocubic structure with a c/a ratio approaching one when exceeding 7 mol% of La3+ ions [77]. Thick/thin films have been fabricated to improve the BDS of the PLZT system. Hao et al. fabricated PLZT bulk ceramics with a thickness of 1 mm using a sol–gel synthesis process and an enhanced Wrec of 28.7 J/cm3 and a η of 60% with a La:Zr:Ti ratio of 9:65:35 [78]. Furthermore, a Mn-doped PLZT thick film with the same ratio and same thickness showed a high Wrec of 30.8 J/cm3 and a η of 68.4% at an electric field of 1185 kV/cm [79,80]. To date, the energy storage properties of PLZT with other lead-based RFEs and various chemical compositions have been reported, such as PZN-PT, PMN-PT, and Pb(Sn,Ti)O3 (PST), exhibiting Wrec values ranging from 1 to 50 J/cm3 for energy storage device applications [81,82,83,84,85]. However, the utilization of lead-based dielectrics has a strong impact on human health and the environment due to their toxicity. Thus, researchers have been developing lead-free RFEs for energy storage applications.
Over the past 20 years, since dielectric constant/polarization is independent of the applied electric field, temperature, and frequency, lead-free BT-based and weakly coupled RFEs have been explored in efforts to achieve high energy density and high efficiency based on the domain tailoring concept [4,23,86,87,88,89,90,91,92,93,94]. Ogihara et al. [86] reported a high Wrec of 6.1 J/cm3 at 73 kV/mm in BT-BiScO3 thick films that were sustained until 300 °C. Yuan et al. [93] reported a domain evaluation using chemical composition and improvements in the energy storage of BT-based ceramics. Furthermore, lead-free BNT-based and strongly coupled RFEs with a high polarization response via minimizing hysteresis loss and leakage currents have been reported. Qiao et al. [95] demonstrated a high Wrec of 4.14 J/cm3 in a Sr and La-co-doped BNT system. The enhanced Wrec is attributed to the small grains and delays in polarization produced by La doping, whereas remnant polarization is decreased following Sr doping. Zhai et al. [96,97] utilized an A-site defect engineering method (nonstoichiometric ratio of Bi and Na) to reduce the electric conductivity and enhance the grain size, which resulted in a high Wrec (5.63 J/cm3 and 3.72 J/cm3) and a high η (94% and 90.7%) in binary and ternary systems, such as 0.75Bi0.58Na0.42TiO3-0.25SrTiO3 and BNT-Bi0.1Sr0.85TiO3-KNbO3. Wu et al. [98] reported the incorporation of Sr0.85Bi0.1□0.05TiO3 (SBT) and NaNbO3 (NN) into a BNT system via a compositional design. The substitution of Sr2+ ions and A-site vacancies constructed RFEs on the basis of the order–disorder theory, enabling a high Wrec of 3.08 J/cm3 and a high η of 81.4%. Liu et al. [99] presented an intrinsic defect and polarization mechanism in A-site-deficient 0.66(Bi0.5Na0.5)TiO3-0.06BaTiO3-0.28(BixSr1–3x/2)TiO3 (BNT-BT-BST) relaxors, favoring polarization behavior, which resulted in a Wrec of 1.61 J/cm3 and a η of 90.5%. Hwang et al. [100] demonstrated the electric energy storage density and energy efficiency of (1 − x)Bi0.5(Na0.8K0.2)0.5TiO3-xBi0.2Sr0.7TiO3 (BNKT-BST; x = 0.15–0.50) RFEs via a domain engineering method. The substitution of BST composition into the BNKT system can disturb the long-range ferroelectric order, reducing the dielectric maximum temperature Tm, which leads to the formation of dynamic PNRs (Figure 8a). Additionally, the Tm was shifted to a higher temperature with increasing frequency, signifying RFE behavior in BNKT-BST ceramics, which is supported by the modified Curie Weiss law (Figure 8b). The relaxor properties contribute to a higher Pmax and a lower Pr, enhancing the BDS with the incorporation of BST, and leading to a high Wrec of 0.81 J/cm3 and high η of 86.95% at an electric field of 90 kV/cm for a x = 0.45 composition (Figure 8c,d). Ma et al. [101] utilized a morphotropic phase boundary (MPB) 0.76Bi0.5Na0.5TiO3-0.24SrTiO3 (BNT-ST) RFE with the incorporation of AFE AN to a lower Pr and retained the same Pmax in order to achieve a Wrec of 2.03 J/cm3. Furthermore, lead-free KNaNbO3-based RFEs have been explored to enhance their energy storage properties. Yang et al. [102] reported composition-driven grain size to a sub-micrometer scale (~100–200 nm) to enhance the breakdown strength of (K0.5Na0.5)NbO3−xSrTiO3 (KNN-ST) RFEs, and showed a high Wrec of 4.03 J/cm3 at 400 kV/cm. Similarly, KNN has been modified with BiFeO3, Sr(Sc0.5Nb0.5)O3, and Bi(Mg2/3Nb1/3)O3 ceramics, and high Wrec values of 2 J/cm3, 2.60 J/cm3, and 4.08 J/cm3 were achieved [34,65,103,104]. Xie et al. [105] reported an ultra-high Wrec of 8.73 J/cm3 and a high η of 80.1% in 0.68 NaNbO3-0.32Bi0.5Li0.5TiO3 ceramics, achieved via exploiting the stable orthorhombic FE phase instead of the AFE orthorhombic phase (Figure 9a,b). In addition, they introduced the AFE relaxor concept to discuss the energy storage performance of 0.78NN-0.24BNT systems. They reported that the local AFE was transformed/reversed into the FE phase at an electric field of 400 kV/cm, inducing a large Pmax (50 µC/cm2) and a low Pr of 5 µC/cm2, which together provided an enhanced ultra-high Wrec of 12.2 J/cm3 and a high η of 69% at an electric field of 680 kV/cm, as shown in Figure 9c,d [65]. The energy storage properties of ceramic-based dielectric materials are listed in Table 1.

3.2. Ceramic Films

In Section 3.1.4, we presented lead-free RFE materials, which are good candidates for energy storage device applications, owing to their ultra-high energy storage density, excellent BDS, and eco-friendliness. However, the miniaturization of electronic devices is necessary for real-world applications, such as hybrid electric vehicles, defense artillery, and smart and wearable electronics [145,146,147]. Therefore, thin/thick film capacitors (e.g., RFEs) have received significant attention in developing high-performance ceramic capacitors for energy storage as compared to bulk ceramic capacitors (LDs, FEs, and AFEs) [1,148,149,150]. Interestingly, these film capacitors have a higher BDS due to less defects, which results in a high energy density. In addition, thin/thick film capacitors are promising for miniaturized electronic devices due to their uniform and highly dense microstructure. The thickness of ceramic capacitors plays an important role in determining the BDS. The thickness/volume ratio of a film capacitor determines its energy storage capacity. Moreover, ceramic capacitor devices with a higher BDS are safe for operation at high voltages and have a smaller likelihood of device failure [6,151].
RFE film-based dielectric capacitors that adopt various strategies for energy storage have been investigated [152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169]. Zhang et al. [170] improved the energy storage performance via a small amount of Mn doping (1 mol.%) in 0.70BNT-0.3ST RFE thin films. Mn2+ ions induce an intrinsic restoring force and enable the reversible domain switching and slim P-E loops (ΔP~56 µC/cm2), resulting in a high Wrec of 27 J/cm3. The same amount of Mn in 0.6ST-0.4BNT thin films yielded a high Wrec of 33.58 J/cm3 at a BDS of 3134 kV/cm, owing to reduced oxygen vacancies [171]. Interestingly, BNT-BT has shown excellent dielectric properties at the MPB between the coexistence of a rhombohedral FE phase and a tetragonal AFE phase for x = 0.06. Peng et al. [172] reported an ultra-high Wrec of 154 J/cm3 via the co-doping of La and Zr in 0.94BNT-0.06BT RFE thin films. The La dopant plays a critical role in enhancing the relaxor properties, whereas the Zr dopant was utilized to control the transition temperature. Pan et al. [173] reported an energy density of 70 J/cm3 in 0.55BiFeO3–0.45SrTiO3 (BF-ST) films via a domain engineering method. The substitution of ST into BF can transform the micrometer-scale FE domains into highly dynamic PNRs, resulting in a high energy storage density in the BF-ST films. In addition, they demonstrated that the coexistence of rhombohedral and tetragonal nanodomain structures in a cubic paraelectric matrix creates a flattened domain-switching pathway in BF-BT-ST films, which minimizes hysteresis loss and delivers an energy density of 112 J/cm3 [152]. Pan and co-workers carried out phase-field simulations in order to choose the proper combination of BF and BT with Sm doping to achieve high energy storage. These simulations were helpful in designing super-paraelectric RFEs with unique and smaller size nanodomains in a Sm-doped BF-BT system, which generated an ultra-high Wrec of 152 J/cm3 and a high η of 90% [174]. The energy storage properties of the ceramic films are summarized in Table 2.

3.3. Multilayer Ceramic Capacitors

MLCCs have received extensive attention in the field of energy storage capacitor applications due to their ultra-high energy density, efficiency, and fast charge–discharge rates [175,176,177,178,179]. In recent years, the energy storage performance was improved in RFE Bi0.5Na0.5TiO3 and AFE AgNbO3-based lead-free ceramics, attaining energy densities of 2.7 J/cm3 and 4.2 J/cm3, respectively [26,177,178,180,181,182,183,184,185,186]. However, high energy dissipation and poor stability are attributed to the AFE to FE phase transition, which are the main drawbacks of AFEs limiting their practical applications. In this regard, Li et al. [5] demonstrated 0.55(Bi0.5Na0.5)TiO3 (BNT)-0.45(Bi0.2Sr0.7)TiO3 (BST) MLCCs and improved their energy density and efficiency by combining RFE and AFE features. The RFE exhibits highly dynamic polar nano-regions and disrupts the long-range ferroelectric order, which results in a hysteresis-free P-E loop. The RFE BST displaying a diffused phase transition was utilized with BNT to obtain RFE features, and is expected to reduce polarization and the high ΔP.
MLCCs have been fabricated using the tape-casting technique, which has two main advantages as follows: (i) The MLCC layers offer low porosity and a fine grain size, leading to a high Eb. (ii) A higher Eb is expected in the MLCC compared to conventional ceramic capacitors because the Eb increases with the decreasing layer thickness. The fabrication process of the MLCCs entails various stages, such as ball milling, slurry formation, tape casting, screen printing, stacking/lamination, dicing, sintering, and termination dipping. Figure 10 presents a schematic illustration of the MLCC fabrication process [187,188]. The ceramic powders were ball milled, slurry dried, and calcined. This calcined powder was re-milled with a dispersant (ethyl methyl ketone), binder (poly(propylene carbonate)), and plasticizer (butyl benzyl phthalate). Furthermore, a slurry was used to prepare thick films using a tape-casting process. The films were stacked layer by layer with inner printed Pt electrodes and then sintered at the desired temperatures to obtain the MLCCs. Lastly, the sintered samples were polished to terminate the opposite ends of the MLCC, and silver paste was coated to form the outer electrodes for electrical characterizations.
Figure 11a shows the unipolar polarization and current versus electric field curve for a 0.55(Bi0.5Na0.5)TiO3 (BNT)-0.45(Bi0.2Sr0.7)TiO3 (0.55BNT-0.45SBT) ceramic sample. It exhibited a high energy storage density of 2.5 J/cm3 and a high efficiency of 95% at a high breakdown field of 20 MV/m. The temperature dependence of the relative dielectric permittivity and the loss factor of the 0.55BNT-0.45SBT sample are shown in Figure 11b. The dielectric maximum temperature (Tm) shifted towards higher temperatures, and the dielectric peaks diffused with increasing frequency, revealing the formation of high-dynamic polar nanoregions (PNRs); such materials are called relaxor ferroelectrics [189,190]. The degree of the diffuseness (γ) of the 0.55BNT-0.45SBT sample is found to be 1.85, indicating strong relaxor behavior (Figure 11c). Due to the formation of PNRs, the 0.55BNT-0.45SBT ceramic sample exhibits a high relative dielectric permittivity, high energy density, and high energy efficiency. To achieve ultrahigh energy density, 0.55BNT-0.45SBT MLCCs were fabricated using the tape-casting method. They consist of 10 dielectric layers with a total thickness of 200 µm and an inner electrode area of 6.25 mm2 (each layer has a thickness of 20 µm), as shown in Figure 11d. The surface morphology of the MLCC is shown in Figure 11e. The breakdown electric field was increased to 72 MV/m due to the advantages of the MLCCs fabricated using the tape-casting method, which offers low porosity and a fine grain size when compared to their counterpart bulk ceramics [191]. In general, the breakdown strength of the ceramics increases as the layer thickness decreases, as observed in many ceramics [192,193]. Figure 11f shows the energy density and efficiency as a function of the electric field for 0.55NBT-0.45SBT MLCCs. The energy storage density of these MLCCs exhibited a high Wrec of 9.5 J/cm3 and a η of 92% at 72 MV/m. These results indicate that combining the antiferroelectric and relaxor properties of MLCCs is a promising approach for improving the energy storage responses in order to meet the requirements of advanced energy storage devices. In recent years, various strategies, including controlled phase [177], chemical homogeneity [178], grain orientation [194], combining antiferroelectric and relaxor properties [5], heterovalent doping [195], and two-step sintering [196], etc., were adopted to enhance the energy storage performance of MLCCs, as summarized in Table 3. In general, as the layer thickness decreases, the BDS of solid dielectrics increases [197]. Thin films show a higher BDS when compared to bulk ceramics and MLCs due to the minimal thickness and less defects, but they have limitations in energy storage density and efficiency. MLCCs have a lower BDS than thin films, but they have other advantages, such as a compact size, a balance between the BDS and energy storage, and good temperature stability, which play an important role in practical applications, especially in pulsed power systems.

4. Challenges and Future Prospects

With the discovery of new materials and strategies, the energy storage density of bulk ceramics, thin films, and MLCCs has been greatly improved to 12, 159, and 52 J/cm3, respectively, as summarized in Table 1, Table 2 and Table 3. Even with the tremendous advancements, there are still certain challenges in real-world applications. Dielectric ceramics with a high energy storage density of more than 8 J/cm3 with a high efficiency of over 90% are still scarce and cannot meet the demands of miniature advanced electronic and electric power systems. To achieve a high energy storage density in dielectrics, researchers mostly focused on the enhancement of ΔP and Eb. Extensively utilized strategies for enhancing Eb are reducing the grain size with homogeneous microstructures, stimulating electrical homogeneity, raising resistance, enhancing thermal conductivity, and lowering dielectric losses. These strategies can be implemented by employing advanced sintering procedures, adding sintering aids, employing two-step sintering, adjusting the heating/cooling rate/holding time, and making composite materials. However, effective strategies for further improving the Eb remain limited. To obtain a high ΔP, the most popular method is to choose a host material with strong ferroelectricity and then decrease its Pr via composition doping. On the other hand, select a host material with a modest Pr and then add a secondary compound to enhance its Pmax. However, it remains challenging to achieve both a high Pmax and a low Pr in these solid solutions. The domain engineering method allows for the fabrication of dielectrics with a low Pr and a moderate ΔP via producing PNRs/nanodomains. However, the Pmax value remains low, restricting the raise of the Wrec. Recently introduced local region design techniques, such as designing local regions with polarization-field response behavior or constructing local regions with polymorphic PNRs via phase structure regulation, will be an excellent choice for developing dielectrics with a high Pmax and a low Pr.
Developing dielectric materials with a high Wrec and η remains the path of future research. In addition, the trade-off between the Wrec and η and the contradiction between the εr and the Eb must be resolved. New materials, new manufacturing techniques, and new design strategies must be discovered in order to achieve these goals. Further research is needed to understand the underlying mechanisms, such as sample sintering processes, dielectric breakdown strength, and dielectric polarization responses in local regions, ultimately developing a profound understanding of the material–structure–property relationship of dielectric materials for energy storage. In addition to developing a single material, more attention should be paid to composite materials, for instance, ceramic/ceramic composites, ceramic/glass composites, ceramic/polymer composites, and ceramic/glass/polymer composites, because it is challenging to develop a single material with a high Pmax, a low Pr, a high Eb, low dielectric loss, and excellent thermal stability/fatigue. Dielectric capacitors with an easy preparation technique, a simple chemical composition, and a low sintering temperature are still in great demand for practical applications. To fabricate new materials, advanced synthesis techniques (two-step sintering and pressure-assisted sintering), comprehensive characterizations (aberration-corrected scanning transmission electron microscope and piezoelectric force microscopy), various control strategies (nanodomain and grain size engineering), and theoretical calculations (machine learning and phase-field simulations) should be employed.
Ceramic-based films show an enormous performance when compared to bulk ceramics in terms of the energy storage density and dielectric breakdown strength. The energy storage properties of ceramic films have been enhanced via various methods, including solid solution formation, layered films with particular configurations (such as sandwich structures, positive/negative gradient compositions), the interface design of films/electrodes, the lattice/strain engineering of films/substrates, and more. Among them, similar to bulk ceramics, the fundamental solution is to deeply understand the inherent nature of whether AFEs/RFEs. Developing films for energy storage is challenging due to their restricted thickness and low absolute energy content. Developing various stratification and flexible scroll technologies is a viable solution for increasing the volume without losing their characteristics. Technological simplicity has the ability to accelerate manufacturing processes and boost automation, thus leading to cost savings and innovation.
MLCCs play an important role in dielectric energy storage. The macroproperties of MLCCs are mostly determined by the thickness of the dielectric layer in addition to their composition. Developing layer thinning techniques is crucial for increasing the energy density per volume. Furthermore, the expensive cost of metal electrodes, such as Au, Pt, and Ag, hinders the commercialization of MLCCs. Low-cost electrodes must be compatible with dielectrics, taking into account the sintering temperature, metal melting temperature, and interface reaction. Therefore, economical electrodes and appropriate cofire techniques should be developed. Since different metals are typically doped to internal and terminal electrodes in most cases, the method for joint connections between these electrodes should be a crucial consideration.

5. Conclusions

Dielectric materials with high power density and ultra-fast discharge rates are becoming increasingly significant in advanced electronic devices and pulsed power systems. Currently, dielectric energy-storage materials are limited in their applications due to their low energy density. Therefore, dielectric materials with excellent energy storage performance are needed. In this review paper, we discuss the fundamental concepts for energy storage in dielectric capacitors, including principles, key parameters, and influence factors for enhancing the energy storage properties. In addition, we summarize the recent progress of dielectrics, such as bulk ceramics/composites, ceramic films, and multilayer ceramic capacitors, followed by the best strategies, such as chemical modification, grain refinement, and defect engineering, for achieving a higher energy density/BDS and higher energy efficiency in dielectric materials for applications in pulsed power systems. Moreover, we present challenges and opportunities for future energy storage dielectric materials.

Author Contributions

Conceptualization, writing—original draft preparation, review, and editing, S.P.; visualization, Y.L., Y.H.S. and Y.M.B.; Review, and editing, M.P.; Validation, project administration, supervision, writing—review and editing, G.-T.H. All authors have read and agreed to the published version of the manuscript.

Funding

The work at PKNU was supported by the National Research Foundation of Korea (NRF-2022R1A2C4001497) grant funded by the Ministry of Science and ICT (MSIT) and he Pukyong National University Industry-university Cooperation Research Fund in 2023(202312480001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yao, Z.; Song, Z.; Hao, H.; Yu, Z.; Cao, M.; Zhang, S.; Lanagan, M.T.; Liu, H. Homogeneous/Inhomogeneous-Structured Dielectrics and Their Energy-Storage Performances. Adv. Mater. 2017, 29, 1601727. [Google Scholar] [CrossRef]
  2. Kang, B.; Ceder, G. Battery Materials for Ultrafast Charging and Discharging. Nature 2009, 458, 190–193. [Google Scholar] [CrossRef]
  3. Luo, X.; Wang, J.; Dooner, M.; Clarke, J. Overview of Current Development in Electrical Energy Storage Technologies and the Application Potential in Power System Operation. Appl. Energy 2015, 137, 511–536. [Google Scholar] [CrossRef]
  4. Kumar, N.; Ionin, A.; Ansell, T.; Kwon, S.; Hackenberger, W.; Cann, D. Multilayer Ceramic Capacitors Based on Relaxor BaTiO3-Bi(Zn1/2Ti1/2)O3 for Temperature Stable and High Energy Density Capacitor Applications. Appl. Phys. Lett. 2015, 106, 252901. [Google Scholar] [CrossRef]
  5. Li, J.; Li, F.; Xu, Z.; Zhang, S. Multilayer Lead-Free Ceramic Capacitors with Ultrahigh Energy Density and Efficiency. Adv. Mater. 2018, 30, 1802155. [Google Scholar] [CrossRef] [PubMed]
  6. Palneedi, H.; Peddigari, M.; Hwang, G.-T.; Jeong, D.-Y.; Ryu, J. High-Performance Dielectric Ceramic Films for Energy Storage Capacitors: Progress and Outlook. Adv. Funct. Mater. 2018, 28, 1803665. [Google Scholar] [CrossRef]
  7. Dong, J.; Hu, R.; Xu, X.; Chen, J.; Niu, Y.; Wang, F.; Hao, J.; Wu, K.; Wang, Q.; Wang, H. A Facile In Situ Surface-Functionalization Approach to Scalable Laminated High-Temperature Polymer Dielectrics with Ultrahigh Capacitive Performance. Adv. Funct. Mater. 2021, 31, 2102644. [Google Scholar] [CrossRef]
  8. Yang, H.; Tian, J.; Lin, Y.; Ma, J. Realizing Ultra-High Energy Storage Density of Lead-Free 0.76Bi0.5Na0.5TiO3-0.24SrTiO3-Bi(Ni2/3Nb1/3)O3 Ceramics under Low Electric Fields. Chem. Eng. J. 2021, 418, 129337. [Google Scholar] [CrossRef]
  9. Li, Q.; Yao, F.-Z.; Liu, Y.; Zhang, G.; Wang, H.; Wang, Q. High-Temperature Dielectric Materials for Electrical Energy Storage. Annu. Rev. Mater. Res. 2018, 48, 219–243. [Google Scholar] [CrossRef]
  10. Li, Q.; Chen, L.; Gadinski, M.R.; Zhang, S.; Zhang, G.; Li, H.U.; Iagodkine, E.; Haque, A.; Chen, L.-Q.; Jackson, T.N.; et al. Flexible High-Temperature Dielectric Materials from Polymer Nanocomposites. Nature 2015, 523, 576–579. [Google Scholar] [CrossRef]
  11. Choi, H.; Ryu, J.; Yoon, W.-H.; Hwang, G.-T. Recent Progress in Dielectric-Based Ultrafast Charging/Discharging Devices. J. Korean Inst. Electr. Electron. Mater. Eng. 2022, 35, 322–332. [Google Scholar]
  12. Tang, H.; Lin, Y.; Sodano, H.A. Synthesis of High Aspect Ratio BaTiO3 Nanowires for High Energy Density Nanocomposite Capacitors. Adv. Energy Mater. 2013, 3, 451–456. [Google Scholar] [CrossRef]
  13. Pandya, S.; Wilbur, J.; Kim, J.; Gao, R.; Dasgupta, A.; Dames, C.; Martin, L.W. Pyroelectric Energy Conversion with Large Energy and Power Density in Relaxor Ferroelectric Thin Films. Nat. Mater. 2018, 17, 432–438. [Google Scholar] [CrossRef] [PubMed]
  14. Lin, Y.; Li, D.; Zhang, M.; Zhan, S.; Yang, Y.; Yang, H.; Yuan, Q. Excellent Energy-Storage Properties Achieved in BaTiO3-Based Lead-Free Relaxor Ferroelectric Ceramics via Domain Engineering on the Nanoscale. ACS Appl. Mater. Interfaces 2019, 11, 36824–36830. [Google Scholar] [CrossRef] [PubMed]
  15. Kishi, H.; Mizuno, Y.; Chazono, H. Base-Metal Electrode-Multilayer Ceramic Capacitors: Past, Present and Future Perspectives. JPN J. Appl. Phys. 2003, 42, 1. [Google Scholar] [CrossRef]
  16. Wang, Y.; Cui, J.; Wang, L.; Yuan, Q.; Niu, Y.; Chen, J.; Wang, Q.; Wang, H. Compositional Tailoring Effect on Electric Field Distribution for Significantly Enhanced Breakdown Strength and Restrained Conductive Loss in Sandwich-Structured Ceramic/Polymer Nanocomposites. J. Mater. Chem. A Mater. 2017, 5, 4710–4718. [Google Scholar] [CrossRef]
  17. Yao, F.-Z.; Yuan, Q.; Wang, Q.; Wang, H. Multiscale Structural Engineering of Dielectric Ceramics for Energy Storage Applications: From Bulk to Thin Films. Nanoscale 2020, 12, 17165–17184. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Yang, H.; Dang, Z.; Zhan, S.; Sun, C.; Hu, G.; Lin, Y.; Yuan, Q. Multilayer Structured Poly(Vinylidene Fluoride)-Based Composite Film with Ultrahigh Breakdown Strength and Discharged Energy Density. ACS Appl. Mater. Interfaces 2020, 12, 22137–22145. [Google Scholar] [CrossRef]
  19. Wang, Y.; Wang, L.; Yuan, Q.; Chen, J.; Niu, Y.; Xu, X.; Cheng, Y.; Yao, B.; Wang, Q.; Wang, H. Ultrahigh Energy Density and Greatly Enhanced Discharged Efficiency of Sandwich-Structured Polymer Nanocomposites with Optimized Spatial Organization. Nano Energy 2018, 44, 364–370. [Google Scholar] [CrossRef]
  20. Zhou, M.; Liang, R.; Zhou, Z.; Yan, S.; Dong, X. Novel Sodium Niobate-Based Lead-Free Ceramics as New Environment-Friendly Energy Storage Materials with High Energy Density, High Power Density, and Excellent Stability. ACS Sustain. Chem. Eng. 2018, 6, 12755–12765. [Google Scholar] [CrossRef]
  21. Chu, B.; Zhou, X.; Ren, K.; Neese, B.; Lin, M.; Wang, Q.; Bauer, F.; Zhang, Q.M. A Dielectric Polymer with High Electric Energy Density and Fast Discharge Speed. Science 2006, 313, 334–336. [Google Scholar] [CrossRef]
  22. Zhu, L.-F.; Zhao, L.; Yan, Y.; Leng, H.; Li, X.; Cheng, L.-Q.; Xiong, X.; Priya, S. Composition and Strain Engineered AgNbO3—Based Multilayer Capacitors for Ultra-High Energy Storage Capacity. J. Mater. Chem. A Mater. 2021, 9, 9655–9664. [Google Scholar] [CrossRef]
  23. Hu, Q.; Tian, Y.; Zhu, Q.; Bian, J.; Jin, L.; Du, H.; Alikin, D.O.; Shur, V.Y.; Feng, Y.; Xu, Z.; et al. Achieve Ultrahigh Energy Storage Performance in BaTiO3–Bi(Mg1/2Ti1/2)O3 Relaxor Ferroelectric Ceramics via Nano-Scale Polarization Mismatch and Reconstruction. Nano Energy 2020, 67, 104264. [Google Scholar] [CrossRef]
  24. Cao, W.; Lin, R.; Chen, P.; Li, F.; Ge, B.; Song, D.; Zhang, J.; Cheng, Z.; Wang, C. 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, 54051–54062. [Google Scholar] [CrossRef]
  25. Zhang, T.; Li, W.; Zhao, Y.; Yu, Y.; Fei, W. High Energy Storage Performance of Opposite Double-Heterojunction Ferroelectricity–Insulators. Adv. Funct. Mater. 2018, 28, 1706211. [Google Scholar] [CrossRef]
  26. Zhao, L.; Liu, Q.; Gao, J.; Zhang, S.; Li, J. Lead-Free Antiferroelectric Silver Niobate Tantalate with High Energy Storage Performance. Adv. Mater. 2017, 29, 1701824. [Google Scholar] [CrossRef] [PubMed]
  27. Uchino, K.; Zheng, J.H.; Chen, Y.H.; Du, X.H.; Ryu, J.; Gao, Y.; Ural, S.; Priya, S.; Hirose, S. Loss Mechanisms and High Power Piezoelectrics. J. Mater. Sci. 2006, 41, 217–228. [Google Scholar] [CrossRef]
  28. Zheng, J.; Takahashi, S.; Yoshikawa, S.; Uchino, K.; de Vries, J.W.C. Heat Generation in Multilayer Piezoelectric Actuators. J. Am. Ceram. Soc. 1996, 79, 3193–3198. [Google Scholar] [CrossRef]
  29. Liu, F.; Li, Q.; Cui, J.; Li, Z.; Yang, G.; Liu, Y.; Dong, L.; Xiong, C.; Wang, H.; Wang, Q. High-Energy-Density Dielectric Polymer Nanocomposites with Trilayered Architecture. Adv. Funct. Mater. 2017, 27, 1606292. [Google Scholar] [CrossRef]
  30. Song, Z.; Zhang, S.; Liu, H.; Hao, H.; Cao, M.; Li, Q.; Wang, Q.; Yao, Z.; Wang, Z.; Lanagan, M.T. Improved Energy Storage Properties Accompanied by Enhanced Interface Polarization in Annealed Microwave-Sintered BST. J. Am. Ceram. Soc. 2015, 98, 3212–3222. [Google Scholar] [CrossRef]
  31. Liu, B.; Wang, X.; Zhang, R.; Li, L. Grain Size Effect and Microstructure Influence on the Energy Storage Properties of Fine-grained BaTiO3-based Ceramics. J. Am. Ceram. Soc. 2017, 100, 3599–3607. [Google Scholar] [CrossRef]
  32. Moulson, A.J.; Herbert, J.M. Electroceramics: Materials, Properties, Applications; Wiley: Hoboken, NJ, USA, 2003; ISBN 9780471497479. [Google Scholar]
  33. Dong, G.; Ma, S.; Du, J.; Cui, J. Dielectric Properties and Energy Storage Density in ZnO-Doped Ba0.3Sr0.7TiO3 Ceramics. Ceram. Int. 2009, 35, 2069–2075. [Google Scholar] [CrossRef]
  34. Yang, Z.; Gao, F.; Du, H.; Jin, L.; Yan, L.; Hu, Q.; Yu, Y.; Qu, S.; Wei, X.; Xu, Z.; et al. Grain Size Engineered Lead-Free Ceramics with Both Large Energy Storage Density and Ultrahigh Mechanical Properties. Nano Energy 2019, 58, 768–777. [Google Scholar] [CrossRef]
  35. Wu, H.; Zhuo, F.; Qiao, H.; Kodumudi Venkataraman, L.; Zheng, M.; Wang, S.; Huang, H.; Li, B.; Mao, X.; Zhang, Q. Polymer-/Ceramic-based Dielectric Composites for Energy Storage and Conversion. Energy Environ. Mater. 2022, 5, 486–514. [Google Scholar] [CrossRef]
  36. Kusko, A.; Dedad, J. Stored Energy—Short-Term and Long-Term Energy Storage Methods. IEEE Ind. Appl. Mag. 2007, 13, 66–72. [Google Scholar] [CrossRef]
  37. Li, W.-T.; McKenzie, D.R.; McFall, W.D.; Zhang, Q.-C.; Wiszniewski, W. Breakdown Mechanism of Al2O3 Based Metal-to-Metal Antifuses. Solid. State Electron. 2000, 44, 1557–1562. [Google Scholar] [CrossRef]
  38. Wang, C.; Zhang, N.; Li, Q.; Yu, Y.; Zhang, J.; Li, Y.; Wang, H. Dielectric Relaxations in Rutile TiO2. J. Am. Ceram. Soc. 2015, 98, 148–153. [Google Scholar] [CrossRef]
  39. Hanzig, J.; Zschornak, M.; Nentwich, M.; Hanzig, F.; Gemming, S.; Leisegang, T.; Meyer, D.C. Strontium Titanate: An All-in-One Rechargeable Energy Storage Material. J. Power Sources 2014, 267, 700–705. [Google Scholar] [CrossRef]
  40. Sarantopoulos, A.; Ferreiro-Vila, E.; Pardo, V.; Magén, C.; Aguirre, M.H.; Rivadulla, F. Electronic Degeneracy and Intrinsic Magnetic Properties of Epitaxial Nb:SrTiO3 Thin Films Controlled by Defects. Phys. Rev. Lett. 2015, 115, 166801. [Google Scholar] [CrossRef]
  41. Zhang, G.-F.; Liu, H.; Yao, Z.; Cao, M.; Hao, H. Effects of Ca Doping on the Energy Storage Properties of (Sr, Ca)TiO3 Paraelectric Ceramics. J. Mater. Sci. Mater. Electron. 2015, 26, 2726–2732. [Google Scholar] [CrossRef]
  42. Liu, S.; Zhai, J. Improving the Dielectric Constant and Energy Density of Poly(Vinylidene Fluoride) Composites Induced by Surface-Modified SrTiO3 Nanofibers by Polyvinylpyrrolidone. J. Mater. Chem. A Mater. 2015, 3, 1511–1517. [Google Scholar] [CrossRef]
  43. Shay, D.P.; Podraza, N.J.; Donnelly, N.J.; Randall, C.A. High Energy Density, High Temperature Capacitors Utilizing Mn-Doped 0.8CaTiO3–0.2CaHfO3 Ceramics. J. Am. Ceram. Soc. 2012, 95, 1348–1355. [Google Scholar] [CrossRef]
  44. Wang, Z.; Cao, M.; Yao, Z.; Li, G.; Song, Z.; Hu, W.; Hao, H.; Liu, H.; Yu, Z. Effects of Sr/Ti Ratio on the Microstructure and Energy Storage Properties of Nonstoichiometric SrTiO3 Ceramics. Ceram. Int. 2014, 40, 929–933. [Google Scholar] [CrossRef]
  45. Yao, Z.; Luo, Q.; Zhang, G.; Hao, H.; Cao, M.; Liu, H. Improved Energy-Storage Performance and Breakdown Enhancement Mechanism of Mg-Doped SrTiO3 Bulk Ceramics for High Energy Density Capacitor Applications. J. Mater. Sci. Mater. Electron. 2017, 28, 11491–11499. [Google Scholar] [CrossRef]
  46. Vorotilov, K.A.; Yanovskaya, M.I.; Solovjeva, L.I.; Valeev, A.S.; Petrovsky, V.I.; Vasiljev, V.A.; Obvinzeva, I.E. Ferroelectric Capacitors for Integrated Circuits. Microelectron. Eng. 1995, 29, 41–44. [Google Scholar] [CrossRef]
  47. Cheng, B.L.; Wang, C.; Wang, S.Y.; Lu, H.B.; Zhou, Y.L.; Chen, Z.H.; Yang, G.Z. Dielectric Properties of (Ba0.8Sr0.2)(ZrxTi1−x)O3 Thin Films Grown by Pulsed-Laser Deposition. J. Eur. Ceram. Soc. 2005, 25, 2295–2298. [Google Scholar] [CrossRef]
  48. Ogihara, H.; Randall, C.A.; Trolier-McKinstry, S. High Temperature and High Energy Density Dielectric Materials. J. Am. Ceram. Soc. 2009, 92, 106–113. [Google Scholar]
  49. Wang, Y.; Shen, Z.-Y.; Li, Y.-M.; Wang, Z.-M.; Luo, W.-Q.; Hong, Y. Optimization of Energy Storage Density and Efficiency in BaxSr1−xTiO3 (x ≤ 0.4) Paraelectric Ceramics. Ceram. Int. 2015, 41, 8252–8256. [Google Scholar] [CrossRef]
  50. Choi, H.; Pattipaka, S.; Son, Y.H.; Bae, Y.M.; Park, J.H.; Jeong, C.K.; Lee, H.E.; Kim, S.-D.; Ryu, J.; Hwang, G.-T. Improved Energy Storage Density and Efficiency of Nd and Mn Co-Doped Ba0.7Sr0.3TiO3 Ceramic Capacitors via Defect Dipole Engineering. Materials 2023, 16, 6753. [Google Scholar] [CrossRef]
  51. Luo, B.; Wang, X.; Tian, E.; Song, H.; Qu, H.; Cai, Z.; Li, B.; Li, L. Mechanism of Ferroelectric Properties of (BaCa)(ZrTi)O3 from First-Principles Calculations. Ceram. Int. 2018, 44, 9684–9688. [Google Scholar] [CrossRef]
  52. Song, Z.; Liu, H.; Zhang, S.; Wang, Z.; Shi, Y.; Hao, H.; Cao, M.; Yao, Z.; Yu, Z. Effect of Grain Size on the Energy Storage Properties of (Ba0.4Sr0.6)TiO3 Paraelectric Ceramics. J. Eur. Ceram. Soc. 2014, 34, 1209–1217. [Google Scholar] [CrossRef]
  53. Tagantsev, A.K.; Vaideeswaran, K.; Vakhrushev, S.B.; Filimonov, A.V.; Burkovsky, R.G.; Shaganov, A.; Andronikova, D.; Rudskoy, A.I.; Baron, A.Q.R.; Uchiyama, H.; et al. The Origin of Antiferroelectricity in PbZrO3. Nat. Commun. 2013, 4, 2229. [Google Scholar] [CrossRef] [PubMed]
  54. Hao, X.; Zhai, J.; Kong, L.B.; Xu, Z. A Comprehensive Review on the Progress of Lead Zirconate-Based Antiferroelectric Materials. Prog. Mater. Sci. 2014, 63, 1–57. [Google Scholar] [CrossRef]
  55. Qiao, P.; Zhang, Y.; Chen, X.; Zhou, M.; Yan, S.; Dong, X.; Wang, G. Enhanced Energy Storage Properties and Stability in (Pb0.895La0.07)(ZrxTi1−x)O3 Antiferroelectric Ceramics. Ceram. Int. 2019, 45, 15898–15905. [Google Scholar] [CrossRef]
  56. Dan, Y.; Zou, K.; Chen, G.; Yu, Y.; Zhang, Y.; Zhang, Q.; Lu, Y.; Zhang, Q.; He, Y. Superior Energy-Storage Properties in (Pb,La)(Zr,Sn,Ti)O3 Antiferroelectric Ceramics with Appropriate La Content. Ceram. Int. 2019, 45, 11375–11381. [Google Scholar] [CrossRef]
  57. Pan, W.; Zhang, Q.; Bhalla, A.; Cross, L.E. Field-Forced Antiferroelectric-to-Ferroelectric Switching in Modified Lead Zirconate Titanate Stannate Ceramics. J. Am. Ceram. Soc. 1989, 72, 571–578. [Google Scholar] [CrossRef]
  58. Wang, H.; Liu, Y.; Yang, T.; Zhang, S. Ultrahigh Energy-Storage Density in Antiferroelectric Ceramics with Field-Induced Multiphase Transitions. Adv. Funct. Mater. 2019, 29, 1807321. [Google Scholar] [CrossRef]
  59. Liu, X.; Li, Y.; Hao, X. Ultra-High Energy-Storage Density and Fast Discharge Speed of (Pb0.98−x La0.02Srx)(Zr0.9Sn0.1)0.995O3 Antiferroelectric Ceramics Prepared via the Tape-Casting Method. J. Mater. Chem. A Mater. 2019, 7, 11858–11866. [Google Scholar] [CrossRef]
  60. Luo, N.; Han, K.; Zhuo, F.; Xu, C.; Zhang, G.; Liu, L.; Chen, X.; Hu, C.; Zhou, H.; Wei, Y. Aliovalent A-Site Engineered AgNbO3 Lead-Free Antiferroelectric Ceramics toward Superior Energy Storage Density. J. Mater. Chem. A Mater. 2019, 7, 14118–14128. [Google Scholar] [CrossRef]
  61. Han, K.; Luo, N.; Mao, S.; Zhuo, F.; Liu, L.; Peng, B.; Chen, X.; Hu, C.; Zhou, H.; Wei, Y. Ultrahigh Energy-Storage Density in A-/B-Site Co-Doped AgNbO3 Lead-Free Antiferroelectric Ceramics: Insight into the Origin of Antiferroelectricity. J. Mater. Chem. A Mater. 2019, 7, 26293–26301. [Google Scholar] [CrossRef]
  62. Luo, N.; Han, K.; Zhuo, F.; Liu, L.; Chen, X.; Peng, B.; Wang, X.; Feng, Q.; Wei, Y. Design for High Energy Storage Density and Temperature-Insensitive Lead-Free Antiferroelectric Ceramics. J. Mater. Chem. C Mater. 2019, 7, 4999–5008. [Google Scholar] [CrossRef]
  63. Luo, N.; Han, K.; Cabral, M.J.; Liao, X.; Zhang, S.; Liao, C.; Zhang, G.; Chen, X.; Feng, Q.; Li, J.-F.; et al. Constructing Phase Boundary in AgNbO3 Antiferroelectrics: Pathway Simultaneously Achieving High Energy Density and Efficiency. Nat. Commun. 2020, 11, 4824. [Google Scholar] [CrossRef] [PubMed]
  64. Qi, H.; Zuo, R. Linear-like Lead-Free Relaxor Antiferroelectric (Bi0.5Na0.5)TiO3–NaNbO3 with Giant Energy-Storage Density/Efficiency and Super Stability against Temperature and Frequency. J. Mater. Chem. A Mater. 2019, 7, 3971–3978. [Google Scholar] [CrossRef]
  65. Qi, H.; Zuo, R.; Xie, A.; Tian, A.; Fu, J.; Zhang, Y.; Zhang, S. Ultrahigh Energy-Storage Density in NaNbO3-Based Lead-Free Relaxor Antiferroelectric Ceramics with Nanoscale Domains. Adv. Funct. Mater. 2019, 29, 1903877. [Google Scholar] [CrossRef]
  66. Wang, J.; Wan, X.; Rao, Y.; Zhao, L.; Zhu, K. Hydrothermal Synthesized AgNbO3 Powders: Leading to Greatly Improved Electric Breakdown Strength in Ceramics. J. Eur. Ceram. Soc. 2020, 40, 5589–5596. [Google Scholar] [CrossRef]
  67. Smolenskii, G.A.; Isupov, V.A.; Agranovskaya, A.I.; Popov, S.N. Ferroelectrics with Diffuse Phase Transitions. Sov. Phys.-Solid. State 1961, 2, 2584–2594. [Google Scholar] [CrossRef]
  68. Shvartsman, V.V.; Lupascu, D.C. Lead-Free Relaxor Ferroelectrics. J. Am. Ceram. Soc. 2012, 95, 1–26. [Google Scholar] [CrossRef]
  69. Perumal, R.N.; Athikesavan, V. Investigations on Electrical and Energy Storage Behaviour of PZN-PT, PMN-PT, PZN–PMN-PT Piezoelectric Solid Solutions. J. Mater. Sci. Mater. Electron. 2019, 30, 902–913. [Google Scholar] [CrossRef]
  70. Zhang, T.F.; Tang, X.G.; Liu, Q.X.; Jiang, Y.P.; Huang, X.X.; Zhou, Q.F. Energy-Storage Properties and High-Temperature Dielectric Relaxation Behaviors of Relaxor Ferroelectric Pb(Mg1/3Nb2/3)O3–PbTiO3 Ceramics. J. Phys. D Appl. Phys. 2016, 49, 095302. [Google Scholar] [CrossRef]
  71. Cross, L.E. Relaxorferroelectrics: An Overview. Ferroelectrics 1994, 151, 305–320. [Google Scholar] [CrossRef]
  72. Liu, G.; Li, Y.; Guo, B.; Tang, M.; Li, Q.; Dong, J.; Yu, L.; Yu, K.; Yan, Y.; Wang, D.; et al. Ultrahigh Dielectric Breakdown Strength and Excellent Energy Storage Performance in Lead-Free Barium Titanate-Based Relaxor Ferroelectric Ceramics via a Combined Strategy of Composition Modification, Viscous Polymer Processing, and Liquid-Phase Sintering. Chem. Eng. J. 2020, 398, 125625. [Google Scholar] [CrossRef]
  73. Triamnak, N.; Yimnirun, R.; Pokorny, J.; Cann, D.P. Relaxor Characteristics of the Phase Transformation in (1 − x)BaTiO3–xBi(Zn1/2Ti1/2)O3 Perovskite Ceramics. J. Am. Ceram. Soc. 2013, 96, 3176–3182. [Google Scholar] [CrossRef]
  74. Yang, H.; Lu, Z.; Li, L.; Bao, W.; Ji, H.; Li, J.; Feteira, A.; Xu, F.; Zhang, Y.; Sun, H.; et al. Novel BaTiO3-Based, Ag/Pd-Compatible Lead-Free Relaxors with Superior Energy Storage Performance. ACS Appl. Mater. Interfaces 2020, 12, 43942–43949. [Google Scholar] [CrossRef] [PubMed]
  75. Ibn-Mohammed, T.; Koh, S.C.L.; Reaney, I.M.; Sinclair, D.C.; Mustapha, K.B.; Acquaye, A.; Wang, D. Are Lead-Free Piezoelectrics More Environmentally Friendly? MRS Commun. 2017, 7, 1–7. [Google Scholar] [CrossRef]
  76. Ibn-Mohammed, T.; Koh, S.C.L.; Reaney, I.M.; Acquaye, A.; Wang, D.; Taylor, S.; Genovese, A. Integrated Hybrid Life Cycle Assessment and Supply Chain Environmental Profile Evaluations of Lead-Based (Lead Zirconate Titanate) versus Lead-Free (Potassium Sodium Niobate) Piezoelectric Ceramics. Energy Environ. Sci. 2016, 9, 3495–3520. [Google Scholar] [CrossRef]
  77. Khodorov, A.; Pereira, M.; Gomes, M.J.M. Structure and Dielectric Properties of Sol–Gel 9/65/35 PLZT Thin Films. J. Eur. Ceram. Soc. 2005, 25, 2285–2288. [Google Scholar] [CrossRef]
  78. Hao, X.; Wang, Y.; Yang, J.; An, S.; Xu, J. High Energy-Storage Performance in Pb0.91La0.09(Ti0.65Zr0.35)O3 Relaxor Ferroelectric Thin Films. J. Appl. Phys. 2012, 112, 114111. [Google Scholar] [CrossRef]
  79. Zhang, L.; Hao, X.; Yang, J.; An, S.; Song, B. Large Enhancement of Energy-Storage Properties of Compositional Graded (Pb1−xLax)(Zr0.65Ti0.35)O3 Relaxor Ferroelectric Thick Films. Appl. Phys. Lett. 2013, 103, 113902. [Google Scholar] [CrossRef]
  80. Liu, Y.; Hao, X.; An, S. Significant Enhancement of Energy-Storage Performance of (Pb0.91La0.09)(Zr0.65Ti0.35)O3 Relaxor Ferroelectric Thin Films by Mn Doping. J. Appl. Phys. 2013, 114, 174102. [Google Scholar] [CrossRef]
  81. Zhao, Q.L.; Cao, M.S.; Yuan, J.; Lu, R.; Wang, D.W.; Zhang, D.Q. Thickness Effect on Electrical Properties of Pb(Zr0.52Ti0.48)O3 Thick Films Embedded with ZnO Nanowhiskers Prepared by a Hybrid Sol–Gel Route. Mater. Lett. 2010, 64, 632–635. [Google Scholar] [CrossRef]
  82. Yang, L.; Kong, X.; Li, F.; Hao, H.; Cheng, Z.; Liu, H.; Li, J.-F.; Zhang, S. Perovskite Lead-Free Dielectrics for Energy Storage Applications. Prog. Mater. Sci. 2019, 102, 72–108. [Google Scholar] [CrossRef]
  83. Xie, Z.; Peng, B.; Zhang, J.; Zhang, X.; Yue, Z.; Li, L. Highly (100)-Oriented Bi(Ni1/2Hf1/2)O3-PbTiO3 Relaxor-Ferroelectric Films for Integrated Piezoelectric Energy Harvesting and Storage System. J. Am. Ceram. Soc. 2015, 98, 2968–2971. [Google Scholar] [CrossRef]
  84. Xie, Z.; Yue, Z.; Peng, B.; Zhang, J.; Zhao, C.; Zhang, X.; Ruehl, G.; Li, L. Large Enhancement of the Recoverable Energy Storage Density and Piezoelectric Response in Relaxor-Ferroelectric Capacitors by Utilizing the Seeding Layers Engineering. Appl. Phys. Lett. 2015, 106, 202901. [Google Scholar] [CrossRef]
  85. Wang, X.; Zhang, L.; Hao, X.; An, S. High Energy-Storage Performance of 0.9Pb(Mg1/3Nb2/3)O3-0.1PbTiO3 Relaxor Ferroelectric Thin Films Prepared by RF Magnetron Sputtering. Mater. Res. Bull. 2015, 65, 73–79. [Google Scholar] [CrossRef]
  86. Ogihara, H.; Randall, C.A.; Trolier-McKinstry, S. High-Energy Density Capacitors Utilizing 0.7BaTiO3–0.3 BiScO3 Ceramics. J. Am. Ceram. Soc. 2009, 92, 1719–1724. [Google Scholar] [CrossRef]
  87. Shen, Z.; Wang, X.; Luo, B.; Li, L. BaTiO3–BiYbO3 Perovskite Materials for Energy Storage Applications. J. Mater. Chem. A Mater. 2015, 3, 18146–18153. [Google Scholar] [CrossRef]
  88. Li, W.-B.; Zhou, D.; Pang, L.-X.; Xu, R.; Guo, H.-H. Novel Barium Titanate Based Capacitors with High Energy Density and Fast Discharge Performance. J. Mater. Chem. A Mater. 2017, 5, 19607–19612. [Google Scholar] [CrossRef]
  89. Zhang, L.; Pang, L.-X.; Li, W.-B.; Zhou, D. Extreme High Energy Storage Efficiency in Perovskite Structured (1 − x)(Ba0.8Sr0.2)TiO3-xBi(Zn2/3Nb1/3)O3 (0.04 ≤ x ≤ 0.16) Ceramics. J. Eur. Ceram. Soc. 2020, 40, 3343–3347. [Google Scholar] [CrossRef]
  90. Yuan, Q.; Yao, F.; Wang, Y.; Ma, R.; Wang, H. Relaxor Ferroelectric 0.9BaTiO3–0.1Bi(Zn0.5Zr0.5)O3 Ceramic Capacitors with High Energy Density and Temperature Stable Energy Storage Properties. J. Mater. Chem. C Mater. 2017, 5, 9552–9558. [Google Scholar] [CrossRef]
  91. Zhou, M.; Liang, R.; Zhou, Z.; Dong, X. Novel BaTiO3-Based Lead-Free Ceramic Capacitors Featuring High Energy Storage Density, High Power Density, and Excellent Stability. J. Mater. Chem. C Mater. 2018, 6, 8528–8537. [Google Scholar] [CrossRef]
  92. Li, F.; Zhou, M.; Zhai, J.; Shen, B.; Zeng, H. Novel Barium Titanate Based Ferroelectric Relaxor Ceramics with Superior Charge-Discharge Performance. J. Eur. Ceram. Soc. 2018, 38, 4646–4652. [Google Scholar] [CrossRef]
  93. Yuan, Q.; Li, G.; Yao, F.-Z.; Cheng, S.-D.; Wang, Y.; Ma, R.; Mi, S.-B.; Gu, M.; Wang, K.; Li, J.-F.; et al. Simultaneously Achieved Temperature-Insensitive High Energy Density and Efficiency in Domain Engineered BaTiO3-Bi(Mg0.5Zr0.5)O3 Lead-Free Relaxor Ferroelectrics. Nano Energy 2018, 52, 203–210. [Google Scholar] [CrossRef]
  94. Liu, N.; Liang, R.; Zhou, Z.; Dong, X. Designing Lead-Free Bismuth Ferrite-Based Ceramics Learning from Relaxor Ferroelectric Behavior for Simultaneous High Energy Density and Efficiency under Low Electric Field. J. Mater. Chem. C Mater. 2018, 6, 10211–10217. [Google Scholar] [CrossRef]
  95. Qiao, X.; Zhang, F.; Wu, D.; Chen, B.; Zhao, X.; Peng, Z.; Ren, X.; Liang, P.; Chao, X.; Yang, Z. Superior Comprehensive Energy Storage Properties in Bi0.5Na0.5TiO3-Based Relaxor Ferroelectric Ceramics. Chem. Eng. J. 2020, 388, 124158. [Google Scholar] [CrossRef]
  96. Yan, F.; Huang, K.; Jiang, T.; Zhou, X.; Shi, Y.; Ge, G.; Shen, B.; Zhai, J. Significantly Enhanced Energy Storage Density and Efficiency of BNT-Based Perovskite Ceramics via A-Site Defect Engineering. Energy Storage Mater. 2020, 30, 392–400. [Google Scholar] [CrossRef]
  97. Zhang, X.; Hu, D.; Pan, Z.; Lv, X.; He, Z.; Yang, F.; Li, P.; Liu, J.; Zhai, J. Enhancement of Recoverable Energy Density and Efficiency of Lead-Free Relaxor-Ferroelectric BNT-Based Ceramics. Chem. Eng. J. 2021, 406, 126818. [Google Scholar] [CrossRef]
  98. Wu, Y.; Fan, Y.; Liu, N.; Peng, P.; Zhou, M.; Yan, S.; Cao, F.; Dong, X.; Wang, G. Enhanced Energy Storage Properties in Sodium Bismuth Titanate-Based Ceramics for Dielectric Capacitor Applications. J. Mater. Chem. C Mater. 2019, 7, 6222–6230. [Google Scholar] [CrossRef]
  99. Liu, X.; Rao, R.; Shi, J.; He, J.; Zhao, Y.; Liu, J.; Du, H. Effect of Oxygen Vacancy and A-Site-Deficiency on the Dielectric Performance of BNT-BT-BST Relaxors. J. Alloys Compd. 2021, 875, 159999. [Google Scholar] [CrossRef]
  100. Pattipaka, S.; Choi, H.; Lim, Y.; Park, K.-I.; Chung, K.; Hwang, G.-T. Enhanced Energy Storage Performance and Efficiency in Bi0.5(Na0.8K0.2)0.5TiO3-Bi0.2Sr0.7TiO3 Relaxor Ferroelectric Ceramics via Domain Engineering. Materials 2023, 16, 4912. [Google Scholar] [CrossRef]
  101. Ma, W.; Zhu, Y.; Marwat, M.A.; Fan, P.; Xie, B.; Salamon, D.; Ye, Z.-G.; Zhang, H. Enhanced Energy-Storage Performance with Excellent Stability under Low Electric Fields in BNT–ST Relaxor Ferroelectric Ceramics. J. Mater. Chem. C Mater. 2019, 7, 281–288. [Google Scholar] [CrossRef]
  102. Yang, Z.; Du, H.; Qu, S.; Hou, Y.; Ma, H.; Wang, J.; Wang, J.; Wei, X.; Xu, Z. Significantly Enhanced Recoverable Energy Storage Density in Potassium–Sodium Niobate-Based Lead Free Ceramics. J. Mater. Chem. A Mater. 2016, 4, 13778–13785. [Google Scholar] [CrossRef]
  103. Shao, T.; Du, H.; Ma, H.; Qu, S.; Wang, J.; Wang, J.; Wei, X.; Xu, Z. Potassium–Sodium Niobate Based Lead-Free Ceramics: Novel Electrical Energy Storage Materials. J. Mater. Chem. A Mater. 2017, 5, 554–563. [Google Scholar] [CrossRef]
  104. Qu, B.; Du, H.; Yang, Z.; Liu, Q.; Liu, T. Enhanced Dielectric Breakdown Strength and Energy Storage Density in Lead-Free Relaxor Ferroelectric Ceramics Prepared Using Transition Liquid Phase Sintering. RSC Adv. 2016, 6, 34381–34389. [Google Scholar] [CrossRef]
  105. Xie, A.; Zuo, R.; Qiao, Z.; Fu, Z.; Hu, T.; Fei, L. NaNbO3-(Bi0.5Li0.5 )TiO3 Lead-Free Relaxor Ferroelectric Capacitors with Superior Energy-Storage Performances via Multiple Synergistic Design. Adv. Energy Mater. 2021, 11, 2101378. [Google Scholar] [CrossRef]
  106. Wang, W.; Pu, Y.; Guo, X.; Shi, R.; Yang, M.; Li, J. Enhanced Energy Storage and Fast Charge-Discharge Capability in Ca0.5Sr0.5TiO3-Based Linear Dielectric Ceramic. J. Alloys Compd. 2020, 817, 152695. [Google Scholar] [CrossRef]
  107. Wang, W.; Pu, Y.; Guo, X.; Shi, R.; Yang, M.; Li, J. Combining High Energy Efficiency and Fast Charge-Discharge Capability in Calcium Strontium Titanate-Based Linear Dielectric Ceramic for Energy-Storage. Ceram. Int. 2020, 46, 11484–11491. [Google Scholar] [CrossRef]
  108. Wang, W.; Pu, Y.; Guo, X.; Shi, R.; Shi, Y.; Yang, M.; Li, J.; Peng, X.; Li, Y. Enhanced Energy Storage Density and High Efficiency of Lead-Free Ca1−xSrxTi1−yZryO3 Linear Dielectric Ceramics. J. Eur. Ceram. Soc. 2019, 39, 5236–5242. [Google Scholar] [CrossRef]
  109. Wang, W.; Pu, Y.; Guo, X.; Ouyang, T.; Shi, Y.; Yang, M.; Li, J.; Shi, R.; Liu, G. Enhanced Energy Storage Properties of Lead-Free (Ca0.5Sr0.5)1−1.5xLaxTiO3 Linear Dielectric Ceramics within a Wide Temperature Range. Ceram. Int. 2019, 45, 14684–14690. [Google Scholar] [CrossRef]
  110. Li, F.; Si, R.; Li, T.; Wang, C.; Zhai, J. High Energy Storage Performance and Fast Discharging Speed in Dense 0.7Bi0.5K0.5TiO3-0.3SrTiO3 Ceramics via a Novel Rolling Technology. Ceram. Int. 2020, 46, 6995–6998. [Google Scholar] [CrossRef]
  111. Ouyang, T.; Pu, Y.; Ji, J.; Zhou, S.; Li, R. Ultrahigh Energy Storage Capacity with Superfast Discharge Rate Achieved in Mg-Modified Ca0.5Sr0.5TiO3-Based Lead-Free Linear Ceramics for Dielectric Capacitor Applications. Ceram. Int. 2021, 47, 20447–20455. [Google Scholar] [CrossRef]
  112. Pu, Y.; Wang, W.; Guo, X.; Shi, R.; Yang, M.; Li, J. Enhancing the Energy Storage Properties of Ca0.5Sr0.5TiO3-Based Lead-Free Linear Dielectric Ceramics with Excellent Stability through Regulating Grain Boundary Defects. J. Mater. Chem. C Mater. 2019, 7, 14384–14393. [Google Scholar] [CrossRef]
  113. Ding, Y.; Li, P.; He, J.; Que, W.; Bai, W.; Zheng, P.; Zhang, J.; Zhai, J. Simultaneously Achieving High Energy-Storage Efficiency and Density in Bi-Modified SrTiO3-Based Relaxor Ferroelectrics by Ion Selective Engineering. Compos. B Eng. 2022, 230, 109493. [Google Scholar] [CrossRef]
  114. Yu, Z.; Zeng, J.; Zheng, L.; Rousseau, A.; Li, G.; Kassiba, A. Microstructure Effects on the Energy Storage Density in BiFeO3-Based Ferroelectric Ceramics. Ceram. Int. 2021, 47, 12735–12741. [Google Scholar] [CrossRef]
  115. Balmuchu, S.P.; Malapati, V.; Chilaka, R.R.; Dobbidi, P. Effective Strategy for Achieving Superior Energy Storage Performance in Lead-Free BaTiO3–Bi(Zn2/3(Nb0.85Ta0.15)1/3)O3 Ferroelectric Ceramics. Ceram. Int. 2024, 50, 13782–13793. [Google Scholar] [CrossRef]
  116. Liu, Z.-G.; Li, M.-D.; Tang, Z.-H.; Tang, X.-G. Enhanced Energy Storage Density and Efficiency in Lead-Free Bi(Mg1/2Hf1/2)O3-Modified BaTiO3 Ceramics. Chem. Eng. J. 2021, 418, 129379. [Google Scholar] [CrossRef]
  117. Huang, K.; Ge, G.; Yan, F.; Shen, B.; Zhai, J. Ultralow Electrical Hysteresis along with High Energy-Storage Density in Lead-Based Antiferroelectric Ceramics. Adv. Electron. Mater. 2020, 6, 1901366. [Google Scholar] [CrossRef]
  118. Huang, J.; Qi, H.; Gao, Y.; Xie, A.; Zhang, Y.; Li, Y.; Wang, S.; Zuo, R. Expanded Linear Polarization Response and Excellent Energy-Storage Properties in (Bi0.5Na0.5)TiO3-KNbO3 Relaxor Antiferroelectrics with Medium Permittivity. Chem. Eng. J. 2020, 398, 125639. [Google Scholar] [CrossRef]
  119. Li, S.; Hu, T.; Nie, H.; Fu, Z.; Xu, C.; Xu, F.; Wang, G.; Dong, X. Giant Energy Density and High Efficiency Achieved in Silver Niobate-Based Lead-Free Antiferroelectric Ceramic Capacitors via Domain Engineering. Energy Storage Mater. 2021, 34, 417–426. [Google Scholar] [CrossRef]
  120. Lu, Z.; Bao, W.; Wang, G.; Sun, S.-K.; Li, L.; Li, J.; Yang, H.; Ji, H.; Feteira, A.; Li, D.; et al. Mechanism of Enhanced Energy Storage Density in AgNbO3-Based Lead-Free Antiferroelectrics. Nano Energy 2021, 79, 105423. [Google Scholar] [CrossRef]
  121. Lai, D.; Yao, Z.; You, W.; Gao, B.; Guo, Q.; Lu, P.; Ullah, A.; Hao, H.; Cao, M.; Liu, H. Modulating the Energy Storage Performance of NaNbO3-Based Lead-Free Ceramics for Pulsed Power Capacitors. Ceram. Int. 2020, 46, 13511–13516. [Google Scholar] [CrossRef]
  122. Dong, X.; Li, X.; Chen, X.; Chen, H.; Sun, C.; Shi, J.; Pang, F.; Zhou, H. High Energy Storage and Ultrafast Discharge in NaNbO3-Based Lead-Free Dielectric Capacitors via a Relaxor Strategy. Ceram. Int. 2021, 47, 3079–3088. [Google Scholar] [CrossRef]
  123. Chen, J.; Qi, H.; Zuo, R. Realizing Stable Relaxor Antiferroelectric and Superior Energy Storage Properties in (Na1−x/2Lax/2)(Nb1−xTix)O3 Lead-Free Ceramics through A/B-Site Complex Substitution. ACS Appl. Mater. Interfaces 2020, 12, 32871–32879. [Google Scholar] [CrossRef] [PubMed]
  124. Chen, H.; Shi, J.; Chen, X.; Sun, C.; Pang, F.; Dong, X.; Zhang, H.; Zhou, H. Excellent Energy Storage Properties and Stability of NaNbO3–Bi(Mg0.5Ta0.5)O3 Ceramics by Introducing (Bi0.5Na0.5)0.7Sr0.3TiO3. J. Mater. Chem. A Mater. 2021, 9, 4789–4799. [Google Scholar] [CrossRef]
  125. Dong, X.; Chen, H.; Wei, M.; Wu, K.; Zhang, J. Structure, Dielectric and Energy Storage Properties of BaTiO3 Ceramics Doped with YNbO4. J. Alloys Compd. 2018, 744, 721–727. [Google Scholar] [CrossRef]
  126. Akram, F.; Kim, J.; Khan, S.A.; Zeb, A.; Yeo, H.G.; Sung, Y.S.; Song, T.K.; Kim, M.-H.; Lee, S. Less Temperature-Dependent High Dielectric and Energy-Storage Properties of Eco-Friendly BiFeO3–BaTiO3-Based Ceramics. J. Alloys Compd. 2020, 818, 152878. [Google Scholar] [CrossRef]
  127. Meng, D.; Feng, Q.; Luo, N.; Yuan, C.; Zhou, C.; Wei, Y.; Fujita, T.; You, H.; Chen, G. Effect of Sr(Zn1/3Nb2/3)O3 Modification on the Energy Storage Performance of BaTiO3 Ceramics. Ceram. Int. 2021, 47, 12450–12458. [Google Scholar] [CrossRef]
  128. Zhou, M.; Liang, R.; Zhou, Z.; Dong, X. Combining High Energy Efficiency and Fast Charge-Discharge Capability in Novel BaTiO3-Based Relaxor Ferroelectric Ceramic for Energy-Storage. Ceram. Int. 2019, 45, 3582–3590. [Google Scholar] [CrossRef]
  129. Zeng, X.; Li, Y.; Dong, J.; Li, J.; Yang, Z.; Song, C.; Liu, G.; Yan, Y. The Polarization Contribution and Effect Mechanism of Ce-Doped 0.65BaTiO3-0.35Sr0.7Bi0.2TiO3 Pb-Free Ferroelectric Ceramics for Dielectric Energy Storage. Ceram. Int. 2021, 47, 32015–32024. [Google Scholar] [CrossRef]
  130. Li, Y.; Liu, Y.; Tang, M.; Lv, J.; Chen, F.; Li, Q.; Yan, Y.; Wu, F.; Jin, L.; Liu, G. Energy Storage Performance of BaTiO3-Based Relaxor Ferroelectric Ceramics Prepared through a Two-Step Process. Chem. Eng. J. 2021, 419, 129673. [Google Scholar] [CrossRef]
  131. Zhu, C.; Cai, Z.; Luo, B.; Guo, L.; Li, L.; Wang, X. High Temperature Lead-Free BNT-Based Ceramics with Stable Energy Storage and Dielectric Properties. J. Mater. Chem. A Mater. 2020, 8, 683–692. [Google Scholar] [CrossRef]
  132. Yang, L.; Kong, X.; Cheng, Z.; Zhang, S. Ultra-High Energy Storage Performance with Mitigated Polarization Saturation in Lead-Free Relaxors. J. Mater. Chem. A Mater. 2019, 7, 8573–8580. [Google Scholar] [CrossRef]
  133. Luo, C.; Feng, Q.; Luo, N.; Yuan, C.; Zhou, C.; Wei, Y.; Fujita, T.; Xu, J.; Chen, G. Effect of Ca2+/Hf4+ Modification at A/B Sites on Energy-Storage Density of Bi0.47Na0.47Ba0.06TiO3 Ceramics. Chem. Eng. J. 2021, 420, 129861. [Google Scholar] [CrossRef]
  134. Yang, F.; Bao, S.; Zhai, Y.; Zhang, Y.; Su, Z.; Liu, J.; Zhai, J.; Pan, Z. Enhanced Energy-Storage Performance and Thermal Stability in Bi0.5Na0.5TiO3-Based Ceramics through Defect Engineering and Composition Design. Mater. Today Chem. 2021, 22, 100583. [Google Scholar] [CrossRef]
  135. Li, D.; Zhou, D.; Liu, W.; Wang, P.-J.; Guo, Y.; Yao, X.-G.; Lin, H.-X. Enhanced Energy Storage Properties Achieved in Na0.5Bi0.5TiO3-Based Ceramics via Composition Design and Domain Engineering. Chem. Eng. J. 2021, 419, 129601. [Google Scholar] [CrossRef]
  136. Yan, F.; Zhou, X.; He, X.; Bai, H.; Wu, S.; Shen, B.; Zhai, J. Superior Energy Storage Properties and Excellent Stability Achieved in Environment-Friendly Ferroelectrics via Composition Design Strategy. Nano Energy 2020, 75, 105012. [Google Scholar] [CrossRef]
  137. Yan, F.; Bai, H.; Zhou, X.; Ge, G.; Li, G.; Shen, B.; Zhai, J. Realizing Superior Energy Storage Properties in Lead-Free Ceramics via a Macro-Structure Design Strategy. J. Mater. Chem. A Mater. 2020, 8, 11656–11664. [Google Scholar] [CrossRef]
  138. Manan, A.; Ullah, A.; Khan, M.A.; Ahmad, A.S.; Iqbal, Y.; Qazi, I.; Rehman, M.U.; Ullah, A.; Liu, H. Preparation, Characterization, and Improvement in the Energy Storage Properties of Bi(Li0.5Ta0.5)O3 Modified Na0.5K0.5NbO3 Ceramic System. Mater. Res. Bull. 2022, 145, 111521. [Google Scholar] [CrossRef]
  139. Ren, X.; Jin, L.; Peng, Z.; Chen, B.; Qiao, X.; Wu, D.; Li, G.; Du, H.; Yang, Z.; Chao, X. Regulation of Energy Density and Efficiency in Transparent Ceramics by Grain Refinement. Chem. Eng. J. 2020, 390, 124566. [Google Scholar] [CrossRef]
  140. Huan, Y.; Wei, T.; Wang, X.; Liu, X.; Zhao, P.; Wang, X. Achieving Ultrahigh Energy Storage Efficiency in Local-Composition Gradient-Structured Ferroelectric Ceramics. Chem. Eng. J. 2021, 425, 129506. [Google Scholar] [CrossRef]
  141. Zhang, Y.; Zuo, R. Excellent Energy-Storage Performances in La2O3 Doped (Na,K)NbO3-Based Lead-Free Relaxor Ferroelectrics. J. Eur. Ceram. Soc. 2020, 40, 5466–5474. [Google Scholar] [CrossRef]
  142. Li, D.; Zhou, D.; Wang, D.; Zhao, W.; Guo, Y.; Shi, Z. Improved Energy Storage Properties Achieved in (K,Na)NbO3-Based Relaxor Ferroelectric Ceramics via a Combinatorial Optimization Strategy. Adv. Funct. Mater. 2022, 32, 2111776. [Google Scholar] [CrossRef]
  143. Zhang, M.; Yang, H.; Yu, Y.; Lin, Y. Energy Storage Performance of K0.5Na0.5NbO3-Based Ceramics Modified by Bi(Zn2/3(Nb0.85Ta0.15)1/3)O3. Chem. Eng. J. 2021, 425, 131465. [Google Scholar] [CrossRef]
  144. Zhang, M.; Yang, H.; Lin, Y.; Yuan, Q.; Du, H. Significant Increase in Comprehensive Energy Storage Performance of Potassium Sodium Niobate-Based Ceramics via Synergistic Optimization Strategy. Energy Storage Mater. 2022, 45, 861–868. [Google Scholar] [CrossRef]
  145. Feng, J.; Ye, S.; Wang, A.; Lu, X.; Tong, Y.; Li, G. Flexible Cellulose Paper-based Asymmetrical Thin Film Supercapacitors with High-Performance for Electrochemical Energy Storage. Adv. Funct. Mater. 2014, 24, 7093–7101. [Google Scholar] [CrossRef]
  146. Zhao, J.; Xu, S.; Tschulik, K.; Compton, R.G.; Wei, M.; O’Hare, D.; Evans, D.G.; Duan, X. Molecular-Scale Hybridization of Clay Monolayers and Conducting Polymer for Thin-Film Supercapacitors. Adv. Funct. Mater. 2015, 25, 2745–2753. [Google Scholar] [CrossRef]
  147. Zhang, W.; Ye, Q.; Fu, D.; Xiong, R. Optoelectronic Duple Bistable Switches: A Bulk Molecular Single Crystal and Unidirectional Ultraflexible Thin Film Based on Imidazolium Fluorochromate. Adv. Funct. Mater. 2017, 27, 1603945. [Google Scholar] [CrossRef]
  148. Li, F.; Hou, X.; Wang, J.; Zeng, H.; Shen, B.; Zhai, J. Structure-Design Strategy of 0–3 Type (Bi0.32Sr0.42Na0.20)TiO3/MgO Composite to Boost Energy Storage Density, Efficiency and Charge-Discharge Performance. J. Eur. Ceram. Soc. 2019, 39, 2889–2898. [Google Scholar] [CrossRef]
  149. Jayakrishnan, A.R.; Silva, J.P.B.; Kamakshi, K.; Annapureddy, V.; Mercioniu, I.F.; Sekhar, K.C. Semiconductor/Relaxor 0–3 Type Composites: A Novel Strategy for Energy Storage Capacitors. J. Sci. Adv. Mater. Devices 2021, 6, 19–26. [Google Scholar] [CrossRef]
  150. Shi, J.; Chen, X.; Li, X.; Sun, J.; Sun, C.; Pang, F.; Zhou, H. Realizing Ultrahigh Recoverable Energy Density and Superior Charge–Discharge Performance in NaNbO3-Based Lead-Free Ceramics via a Local Random Field Strategy. J. Mater. Chem. C Mater. 2020, 8, 3784–3794. [Google Scholar] [CrossRef]
  151. Sun, Z.; Wang, Z.; Tian, Y.; Wang, G.; Wang, W.; Yang, M.; Wang, X.; Zhang, F.; Pu, Y. Progress, Outlook, and Challenges in Lead-Free Energy-Storage Ferroelectrics. Adv. Electron. Mater. 2020, 6, 1900698. [Google Scholar] [CrossRef]
  152. Pan, H.; Li, F.; Liu, Y.; Zhang, Q.; Wang, M.; Lan, S.; Zheng, Y.; Ma, J.; Gu, L.; Shen, Y.; et al. Ultrahigh–Energy Density Lead-Free Dielectric Films via Polymorphic Nanodomain Design. Science 2019, 365, 578–582. [Google Scholar] [CrossRef] [PubMed]
  153. Silva, J.P.B.; Silva, J.M.B.; Oliveira, M.J.S.; Weingärtner, T.; Sekhar, K.C.; Pereira, M.; Gomes, M.J.M. High-Performance Ferroelectric–Dielectric Multilayered Thin Films for Energy Storage Capacitors. Adv. Funct. Mater. 2019, 29, 1807196. [Google Scholar] [CrossRef]
  154. Sun, N.; Li, Y.; Zhang, Q.; Hao, X. Giant Energy-Storage Density and High Efficiency Achieved in (Bi0.5Na0.5)TiO3–Bi(Ni0.5Zr0.5)O3 Thick Films with Polar Nanoregions. J. Mater. Chem. C Mater. 2018, 6, 10693–10703. [Google Scholar] [CrossRef]
  155. Yang, C.; Lv, P.; Qian, J.; Han, Y.; Ouyang, J.; Lin, X.; Huang, S.; Cheng, Z. Fatigue-Free and Bending-Endurable Flexible Mn-Doped Na0.5Bi0.5TiO3-BaTiO3-BiFeO3 Film Capacitor with an Ultrahigh Energy Storage Performance. Adv. Energy Mater. 2019, 9, 1803949. [Google Scholar] [CrossRef]
  156. Yang, C.; Qian, J.; Han, Y.; Lv, P.; Huang, S.; Cheng, X.; Cheng, Z. Design of an All-Inorganic Flexible Na0.5Bi0.5TiO3-Based Film Capacitor with Giant and Stable Energy Storage Performance. J. Mater. Chem. A Mater. 2019, 7, 22366–22376. [Google Scholar] [CrossRef]
  157. Liang, Z.; Ma, C.; Shen, L.; Lu, L.; Lu, X.; Lou, X.; Liu, M.; Jia, C.-L. Flexible Lead-Free Oxide Film Capacitors with Ultrahigh Energy Storage Performances in Extremely Wide Operating Temperature. Nano Energy 2019, 57, 519–527. [Google Scholar] [CrossRef]
  158. Liang, Z.; Liu, M.; Shen, L.; Lu, L.; Ma, C.; Lu, X.; Lou, X.; Jia, C.-L. All-Inorganic Flexible Embedded Thin-Film Capacitors for Dielectric Energy Storage with High Performance. ACS Appl. Mater. Interfaces 2019, 11, 5247–5255. [Google Scholar] [CrossRef]
  159. Song, B.; Wu, S.; Li, F.; Chen, P.; Shen, B.; Zhai, J. Excellent Energy Storage Density and Charge–Discharge Performance of a Novel Bi0.2Sr0.7TiO3–BiFeO3 Thin Film. J. Mater. Chem. C Mater. 2019, 7, 10891–10900. [Google Scholar] [CrossRef]
  160. Kursumovic, A.; Li, W.-W.; Cho, S.; Curran, P.J.; Tjhe, D.H.L.; MacManus-Driscoll, J.L. Lead-Free Relaxor Thin Films with Huge Energy Density and Low Loss for High Temperature Applications. Nano Energy 2020, 71, 104536. [Google Scholar] [CrossRef]
  161. Sun, N.; Li, Y.; Liu, X.; Hao, X. High Energy-Storage Density under Low Electric Field in Lead-Free Relaxor Ferroelectric Film Based on Synergistic Effect of Multiple Polar Structures. J. Power Sources 2020, 448, 227457. [Google Scholar] [CrossRef]
  162. Fan, Y.; Zhou, Z.; Chen, Y.; Huang, W.; Dong, X. A Novel Lead-Free and High-Performance Barium Strontium Titanate-Based Thin Film Capacitor with Ultrahigh Energy Storage Density and Giant Power Density. J. Mater. Chem. C Mater. 2020, 8, 50–57. [Google Scholar] [CrossRef]
  163. Yang, C.; Qian, J.; Lv, P.; Wu, H.; Lin, X.; Wang, K.; Ouyang, J.; Huang, S.; Cheng, X.; Cheng, Z. Flexible Lead-Free BFO-Based Dielectric Capacitor with Large Energy Density, Superior Thermal Stability, and Reliable Bending Endurance. J. Mater. 2020, 6, 200–208. [Google Scholar] [CrossRef]
  164. Qian, J.; Han, Y.; Yang, C.; Lv, P.; Zhang, X.; Feng, C.; Lin, X.; Huang, S.; Cheng, X.; Cheng, Z. Energy Storage Performance of Flexible NKBT/NKBT-ST Multilayer Film Capacitor by Interface Engineering. Nano Energy 2020, 74, 104862. [Google Scholar] [CrossRef]
  165. Chen, J.; Tang, Z.; Yang, B.; Zhao, S. Ultra-High Energy Storage Performances Regulated by Depletion Region Engineering Sensitive to the Electric Field in PNP-Type Relaxor Ferroelectric Heterostructural Films. J. Mater. Chem. A Mater. 2020, 8, 8010–8019. [Google Scholar] [CrossRef]
  166. Yang, C.; Han, Y.; Feng, C.; Lin, X.; Huang, S.; Cheng, X.; Cheng, Z. Toward Multifunctional Electronics: Flexible NBT-Based Film with a Large Electrocaloric Effect and High Energy Storage Property. ACS Appl. Mater. Interfaces 2020, 12, 6082–6089. [Google Scholar] [CrossRef]
  167. Yang, X.; Li, W.; Zhang, Y.; Qiao, Y.; Yang, Y.; Fei, W. High Energy Storage Density Achieved in Bi3+-Li+ Co-Doped SrTi0.99Mn0.01O3 Thin Film via Ionic Pair Doping-Engineering. J. Eur. Ceram. Soc. 2020, 40, 706–711. [Google Scholar] [CrossRef]
  168. Sun, Y.; Zhang, L.; Wang, H.; Guo, M.; Lou, X.; Wang, D. Composition-Driven Inverse-to-Conventional Transformation of Electrocaloric Effect and Large Energy Storage Density in Strontium Modified Ba(Zr0.1Ti0.9)O3 Thin Films. J. Mater. Chem. C Mater. 2020, 8, 1366–1373. [Google Scholar] [CrossRef]
  169. Silva, J.P.B.; Silva, J.M.B.; Sekhar, K.C.; Palneedi, H.; Istrate, M.C.; Negrea, R.F.; Ghica, C.; Chahboun, A.; Pereira, M.; Gomes, M.J.M. Energy Storage Performance of Ferroelectric ZrO2 Film Capacitors: Effect of HfO2:Al2O3 Dielectric Insert Layer. J. Mater. Chem. A Mater. 2020, 8, 14171–14177. [Google Scholar] [CrossRef]
  170. Zhang, Y.; Li, W.; Cao, W.; Feng, Y.; Qiao, Y.; Zhang, T.; Fei, W. Mn Doping to Enhance Energy Storage Performance of Lead-Free 0.7NBT-0.3ST Thin Films with Weak Oxygen Vacancies. Appl. Phys. Lett. 2017, 110, 243901. [Google Scholar] [CrossRef]
  171. Zhang, Y.; Li, W.; Qiao, Y.; Zhao, Y.; Wang, Z.; Yu, Y.; Xia, H.; Li, Z.; Fei, W. 0.6ST-0.4NBT Thin Film with Low Level Mn Doping as a Lead-Free Ferroelectric Capacitor with High Energy Storage Performance. Appl. Phys. Lett. 2018, 112, 093902. [Google Scholar] [CrossRef]
  172. Peng, B.; Zhang, Q.; Li, X.; Sun, T.; Fan, H.; Ke, S.; Ye, M.; Wang, Y.; Lu, W.; Niu, H.; et al. Giant Electric Energy Density in Epitaxial Lead-Free Thin Films with Coexistence of Ferroelectrics and Antiferroelectrics. Adv. Electron. Mater. 2015, 1, 1500052. [Google Scholar] [CrossRef]
  173. Pan, H.; Ma, J.; Ma, J.; Zhang, Q.; Liu, X.; Guan, B.; Gu, L.; Zhang, X.; Zhang, Y.-J.; Li, L.; et al. Giant Energy Density and High Efficiency Achieved in Bismuth Ferrite-Based Film Capacitors via Domain Engineering. Nat. Commun. 2018, 9, 1813. [Google Scholar] [CrossRef]
  174. Pan, H.; Lan, S.; Xu, S.; Zhang, Q.; Yao, H.; Liu, Y.; Meng, F.; Guo, E.-J.; Gu, L.; Yi, D.; et al. Ultrahigh Energy Storage in Superparaelectric Relaxor Ferroelectrics. Science 2021, 374, 100–104. [Google Scholar] [CrossRef] [PubMed]
  175. Zhao, P.; Wang, H.; Wu, L.; Chen, L.; Cai, Z.; Li, L.; Wang, X. High-Performance Relaxor Ferroelectric Materials for Energy Storage Applications. Adv. Energy Mater. 2019, 9, 1803048. [Google Scholar] [CrossRef]
  176. Sun, Z.; Ma, C.; Liu, M.; Cui, J.; Lu, L.; Lu, J.; Lou, X.; Jin, L.; Wang, H.; Jia, C. Ultrahigh Energy Storage Performance of Lead-Free Oxide Multilayer Film Capacitors via Interface Engineering. Adv. Mater. 2017, 29, 1604427. [Google Scholar] [CrossRef] [PubMed]
  177. Wang, D.; Fan, Z.; Zhou, D.; Khesro, A.; Murakami, S.; Feteira, A.; Zhao, Q.; Tan, X.; Reaney, I.M. Bismuth Ferrite-Based Lead-Free Ceramics and Multilayers with High Recoverable Energy Density. J. Mater. Chem. A Mater. 2018, 6, 4133–4144. [Google Scholar] [CrossRef]
  178. Wang, G.; Li, J.; Zhang, X.; Fan, Z.; Yang, F.; Feteira, A.; Zhou, D.; Sinclair, D.C.; Ma, T.; Tan, X.; et al. Ultrahigh Energy Storage Density Lead-Free Multilayers by Controlled Electrical Homogeneity. Energy Environ. Sci. 2019, 12, 582–588. [Google Scholar] [CrossRef]
  179. Seo, I.; Kang, H.W.; Han, S.H. Recent Progress in Dielectric Materials for MLCC Application. J. Korean Inst. Electr. Electron. Mater. Eng. 2022, 35, 103–118. [Google Scholar]
  180. Zhang, L.; Hao, X. Dielectric Properties and Energy-Storage Performances of (1 − x)(Na0.5Bi0.5)TiO3xSrTiO3 Thick Films Prepared by Screen Printing Technique. J. Alloys Compd. 2014, 586, 674–678. [Google Scholar] [CrossRef]
  181. Gao, F.; Dong, X.; Mao, C.; Liu, W.; Zhang, H.; Yang, L.; Cao, F.; Wang, G. Energy-Storage Properties 0.89Bi0.5Na0.5TiO3–0.06BaTiO3–0.05K0.5Na0.5NbO3 Lead-Free Anti-ferroelectric Ceramics. J. Am. Ceram. Soc. 2011, 94, 4382–4386. [Google Scholar] [CrossRef]
  182. Cao, W.; Li, W.; Feng, Y.; Bai, T.; Qiao, Y.; Hou, Y.; Zhang, T.; Yu, Y.; Fei, W. Defect Dipole Induced Large Recoverable Strain and High Energy-Storage Density in Lead-Free Na0.5Bi0.5TiO3-Based Systems. Appl. Phys. Lett. 2016, 108, 202902. [Google Scholar] [CrossRef]
  183. Cao, W.; Li, W.; Zhang, T.; Sheng, J.; Hou, Y.; Feng, Y.; Yu, Y.; Fei, W. High-Energy Storage Density and Efficiency of (1 − x)[0.94 NBT–0.06 BT]–xST Lead-Free Ceramics. Energy Technol. 2015, 3, 1198–1204. [Google Scholar] [CrossRef]
  184. Cao, W.P.; Li, W.L.; Dai, X.F.; Zhang, T.D.; Sheng, J.; Hou, Y.F.; Fei, W.D. Large Electrocaloric Response and High Energy-Storage Properties over a Broad Temperature Range in Lead-Free NBT-ST Ceramics. J. Eur. Ceram. Soc. 2016, 36, 593–600. [Google Scholar] [CrossRef]
  185. Mishra, A.; Majumdar, B.; Ranjan, R. A Complex Lead-Free (Na,Bi,Ba)(Ti,Fe)O3 Single Phase Perovskite Ceramic with a High Energy-Density and High Discharge-Efficiency for Solid State Capacitor Applications. J. Eur. Ceram. Soc. 2017, 37, 2379–2384. [Google Scholar] [CrossRef]
  186. Zhao, L.; Gao, J.; Liu, Q.; Zhang, S.; Li, J.-F. Silver Niobate Lead-Free Antiferroelectric Ceramics: Enhancing Energy Storage Density by B-Site Doping. ACS Appl. Mater. Interfaces 2018, 10, 819–826. [Google Scholar] [CrossRef] [PubMed]
  187. Murakami, S.; Wang, D.; Mostaed, A.; Khesro, A.; Feteira, A.; Sinclair, D.C.; Fan, Z.; Tan, X.; Reaney, I.M. High Strain (0.4%) Bi(Mg2/3Nb1/3)O3-BaTiO3-BiFeO3 Lead-free Piezoelectric Ceramics and Multilayers. J. Am. Ceram. Soc. 2018, 101, 5428–5442. [Google Scholar] [CrossRef]
  188. Cumming, D.J.; Sebastian, T.; Sterianou, I.; Rödel, J.; Reaney, I.M. Bi(Me)O3–PbTiO3 High TC Piezoelectric Multilayers. Mater. Technol. 2013, 28, 247–253. [Google Scholar] [CrossRef]
  189. Cross, L.E. Relaxor Ferroelectrics. Ferroelectrics 1987, 76, 241–267. [Google Scholar] [CrossRef]
  190. Ma, C.; Tan, X. Phase Diagram of Unpoled Lead-Free—Ceramics. Solid. State Commun. 2010, 150, 1497–1500. [Google Scholar] [CrossRef]
  191. Jabbari, M.; Bulatova, R.; Tok, A.I.Y.; Bahl, C.R.H.; Mitsoulis, E.; Hattel, J.H. Ceramic Tape Casting: A Review of Current Methods and Trends with Emphasis on Rheological Behaviour and Flow Analysis. Mater. Sci. Eng. B 2016, 212, 39–61. [Google Scholar] [CrossRef]
  192. Kurchania, R.; Milne, S.J. Characterization of Sol-Gel Pb(Zr0.53Ti0.47 )O3 Films in the Thickness Range 0.25–10 Μm. J. Mater. Res. 1999, 14, 1852–1859. [Google Scholar] [CrossRef]
  193. Kishimoto, A.; Koumoto, K.; Yanagida, H.; Nameki, M. Microstructure Dependence of Mechanical and Dielectric Strengths—I. Porosity. Eng. Fract. Mech. 1991, 40, 927–930. [Google Scholar] [CrossRef]
  194. Li, J.; Shen, Z.; Chen, X.; Yang, S.; Zhou, W.; Wang, M.; Wang, L.; Kou, Q.; Liu, Y.; Li, Q.; et al. Grain-Orientation-Engineered Multilayer Ceramic Capacitors for Energy Storage Applications. Nat. Mater. 2020, 19, 999–1005. [Google Scholar] [CrossRef] [PubMed]
  195. Zhu, L.-F.; Deng, S.; Zhao, L.; Li, G.; Wang, Q.; Li, L.; Yan, Y.; Qi, H.; Zhang, B.-P.; Chen, J.; et al. Heterovalent-Doping-Enabled Atom-Displacement Fluctuation Leads to Ultrahigh Energy-Storage Density in AgNbO3-Based Multilayer Capacitors. Nat. Commun. 2023, 14, 1166. [Google Scholar] [CrossRef]
  196. Cai, Z.; Zhu, C.; Wang, H.; Zhao, P.; Chen, L.; Li, L.; Wang, X. High-Temperature Lead-Free Multilayer Ceramic Capacitors with Ultrahigh Energy Density and Efficiency Fabricated via Two-Step Sintering. J. Mater. Chem. A Mater. 2019, 7, 14575–14582. [Google Scholar] [CrossRef]
  197. Gerson, R.; Marshall, T.C. Dielectric Breakdown of Porous Ceramics. J. Appl. Phys. 1959, 30, 1650–1653. [Google Scholar] [CrossRef]
  198. Lu, Z.; Wang, G.; Bao, W.; Li, J.; Li, L.; Mostaed, A.; Yang, H.; Ji, H.; Li, D.; Feteira, A.; et al. Superior Energy Density through Tailored Dopant Strategies in Multilayer Ceramic Capacitors. Energy Environ. Sci. 2020, 13, 2938–2948. [Google Scholar] [CrossRef]
  199. Hu, T.-Y.; Ma, C.; Dai, Y.; Fan, Q.; Liu, M.; Jia, C.-L. Enhanced Energy Storage Performance of Lead-Free Capacitors in an Ultrawide Temperature Range via Engineering Paraferroelectric and Relaxor Ferroelectric Multilayer Films. ACS Appl. Mater. Interfaces 2020, 12, 25930–25937. [Google Scholar] [CrossRef]
Figure 1. (a) Various applications of dielectric capacitors in power electronics and pulse power applications. (b) Comparison of the power density versus energy density of batteries, electrochemical capacitors, and dielectric capacitors.
Figure 1. (a) Various applications of dielectric capacitors in power electronics and pulse power applications. (b) Comparison of the power density versus energy density of batteries, electrochemical capacitors, and dielectric capacitors.
Materials 17 02277 g001
Figure 2. Schematic diagram of (a) a dielectric capacitor, and (b) a dielectric between two conductive plates, where electric dipoles are displaced and oriented by the applied electric field due to polarization.
Figure 2. Schematic diagram of (a) a dielectric capacitor, and (b) a dielectric between two conductive plates, where electric dipoles are displaced and oriented by the applied electric field due to polarization.
Materials 17 02277 g002
Figure 3. Schematic of the recoverable energy density and energy loss from the P-E hysteresis loop of a ceramic capacitor.
Figure 3. Schematic of the recoverable energy density and energy loss from the P-E hysteresis loop of a ceramic capacitor.
Materials 17 02277 g003
Figure 4. Schematic of the electric field-dependent polarization response and ferroelectric domain structures with dipole orientation for (a) LDs, (b) FEs, (c) AFEs, and (d) RFEs.
Figure 4. Schematic of the electric field-dependent polarization response and ferroelectric domain structures with dipole orientation for (a) LDs, (b) FEs, (c) AFEs, and (d) RFEs.
Materials 17 02277 g004
Figure 5. Schematic illustration of a defect dipole concept to achieve energy storage properties of Nd and Mn-co-doped Ba0.7Sr0.3TiO3 ceramics. Defect dipoles between donor/acceptor ions and oxygen vacancies capture electrons, decrease grain size, and enable a high difference between Pmax and Pr, thereby enhancing the BDS with Nd and Mn, which results in an improved Wrec and η in Ba0.7Sr0.3TiO3 ceramics. Reproduced with permission [50]. Copyright 2023, MDPI.
Figure 5. Schematic illustration of a defect dipole concept to achieve energy storage properties of Nd and Mn-co-doped Ba0.7Sr0.3TiO3 ceramics. Defect dipoles between donor/acceptor ions and oxygen vacancies capture electrons, decrease grain size, and enable a high difference between Pmax and Pr, thereby enhancing the BDS with Nd and Mn, which results in an improved Wrec and η in Ba0.7Sr0.3TiO3 ceramics. Reproduced with permission [50]. Copyright 2023, MDPI.
Materials 17 02277 g005
Figure 6. (a) P-E loops and (b) Wst, Wre, and η of the (Pb1−yLay)(ZrxTi1−x)O3 ceramics for y = 0.07 and x = 0.82 to 0.92. (c) shows a SEM image for x = 0.9. Reproduced with permission [55]. Copyright 2019, Elsevier.
Figure 6. (a) P-E loops and (b) Wst, Wre, and η of the (Pb1−yLay)(ZrxTi1−x)O3 ceramics for y = 0.07 and x = 0.82 to 0.92. (c) shows a SEM image for x = 0.9. Reproduced with permission [55]. Copyright 2019, Elsevier.
Materials 17 02277 g006
Figure 7. (a) P-E loops of Ag(Nb1−xTax)O3 ceramics for x = 0 and 0.15, (b) Wre and η, and (c) Eb and grain size of Ag(Nb1−xTax)O3 ceramics for x = 0 to 20. The inset of Figure 7c shows SEM images of Ag(Nb1−xTax)O3 ceramics for x = 0 and 0.20. Reproduced with permission [26]. Copyright 2017, Wiley-VCH.
Figure 7. (a) P-E loops of Ag(Nb1−xTax)O3 ceramics for x = 0 and 0.15, (b) Wre and η, and (c) Eb and grain size of Ag(Nb1−xTax)O3 ceramics for x = 0 to 20. The inset of Figure 7c shows SEM images of Ag(Nb1−xTax)O3 ceramics for x = 0 and 0.20. Reproduced with permission [26]. Copyright 2017, Wiley-VCH.
Materials 17 02277 g007
Figure 8. (a) Schematic of the domain structure and formation of the FE to RFE transition with the incorporation of BST into BNKT, leading to improved Wrec and η (where the red arrows indicate the dipole orientation). (b) Temperature dependence of the relative dielectric permittivity and loss factor of 0.55BNKT-0.45BST composition. The inset of Figure 8b presents the l o g T T m versus l o g 1 ε r 1 ε r m of 0.55BNKT-0.45BST at 1 MHz. (c) P-E hysteresis loop of 0.55BNKT-0.45BST ceramics. (d) Composition versus Wrec, Wloss, and η for x = 0.15–0.50. Reproduced with permission [100]. Copyright 2023 MDPI.
Figure 8. (a) Schematic of the domain structure and formation of the FE to RFE transition with the incorporation of BST into BNKT, leading to improved Wrec and η (where the red arrows indicate the dipole orientation). (b) Temperature dependence of the relative dielectric permittivity and loss factor of 0.55BNKT-0.45BST composition. The inset of Figure 8b presents the l o g T T m versus l o g 1 ε r 1 ε r m of 0.55BNKT-0.45BST at 1 MHz. (c) P-E hysteresis loop of 0.55BNKT-0.45BST ceramics. (d) Composition versus Wrec, Wloss, and η for x = 0.15–0.50. Reproduced with permission [100]. Copyright 2023 MDPI.
Materials 17 02277 g008
Figure 9. (a) P-E loops and (b) Wrec and η values of 0.68 NN-0.32BLT ceramics at various fields. (c) P−E loops along with the current density versus electric field curve. Reproduced with permission [105]. Copyright 2021 John Wiley and Sons. (d) Wrec and η values of 0.76 NN-0.24BNT ceramics at various fields and measured at 10 Hz and RT. Reproduced with permission [65]. Copyright 2019 John Wiley and Sons.
Figure 9. (a) P-E loops and (b) Wrec and η values of 0.68 NN-0.32BLT ceramics at various fields. (c) P−E loops along with the current density versus electric field curve. Reproduced with permission [105]. Copyright 2021 John Wiley and Sons. (d) Wrec and η values of 0.76 NN-0.24BNT ceramics at various fields and measured at 10 Hz and RT. Reproduced with permission [65]. Copyright 2019 John Wiley and Sons.
Materials 17 02277 g009
Figure 10. Schematic diagram of the MLCC fabrication process.
Figure 10. Schematic diagram of the MLCC fabrication process.
Materials 17 02277 g010
Figure 11. (a) Polarization (blue curve) and current (red curve) response are a function of the electric field (the inset shows a picture of the bulk sample), (b) the temperature variation of relative dielectric permittivity and loss factor, and (c) l n T T m a x versus l n 1 ε 1 ε m a x of 0.55(Bi0.5Na0.5)TiO3-0.45(Bi0.2Sr0.7)TiO3 ceramics. (d,e) A photograph and SEM image of MLCC of 0.55(Bi0.5Na0.5)TiO3-0.45(Bi0.2Sr0.7)TiO3. (f) Energy density and efficiency versus the applied electric field of 0.55(Bi0.5Na0.5)TiO3-0.45(Bi0.2Sr0.7)TiO3 MLCC. Reproduced with permission [5]. Copyright 2018, Wiley-VCH.
Figure 11. (a) Polarization (blue curve) and current (red curve) response are a function of the electric field (the inset shows a picture of the bulk sample), (b) the temperature variation of relative dielectric permittivity and loss factor, and (c) l n T T m a x versus l n 1 ε 1 ε m a x of 0.55(Bi0.5Na0.5)TiO3-0.45(Bi0.2Sr0.7)TiO3 ceramics. (d,e) A photograph and SEM image of MLCC of 0.55(Bi0.5Na0.5)TiO3-0.45(Bi0.2Sr0.7)TiO3. (f) Energy density and efficiency versus the applied electric field of 0.55(Bi0.5Na0.5)TiO3-0.45(Bi0.2Sr0.7)TiO3 MLCC. Reproduced with permission [5]. Copyright 2018, Wiley-VCH.
Materials 17 02277 g011
Table 1. Energy storage properties of ceramic-based dielectric bulk materials.
Table 1. Energy storage properties of ceramic-based dielectric bulk materials.
Bulk
Ceramics
CompositionWrec (J/cm3)η (%)Eb (kV/cm)Refs.
Ca0.5Sr0.5(Ta0.024Ti0.97)O3-2 wt%SiO2296360[106]
LDsCa0.5Sr0.5Ti0.97Sn0.03O32.0695330[107]
Ca0.5Sr0.5Ti0.9Zr0.1O32.0585390[108]
(Ca0.5Sr0.5)0.8875La0.075TiO32.0793370[109]
0.7(Bi0.5K0.5TiO3)-0.3SrTiO32.3177.7190[110]
(Ca0.5Sr0.5)0.99Mg0.01TiO32.8890460[111]
Ca0.5Sr0.5Ti0.85Zr0.15O33.3796440[112]
0.9(Sr0.7Bi0.2TiO3)-0.1Bi(Ni2/3Nb1/3)O33.7197340[113]
Ca-doped SrTiO31.9572.3333[41]
Zr doped Sr0.98Ca0.02TiO32.7777.7-[47]
Mg-doped SrTiO31.8672.3362[45]
CaZrO3-0.05SrTiO35-1000[48]
0.8CaTiO3-0.2CaHfO39-1200[43]
FEsBa0.3Sr0.7TiO30.2395.790[49]
Nd and Mn-doped Ba0.7Sr0.3TiO30.4184.6110.6[50]
1.6 wt% ZnO doped Ba0.3Sr0.7TiO33.9-40[33]
(BaCa)(ZrTi)O31.28-243[52]
BiFeO3–BaTiO3–Bi(Mg2/3Nb1/3)O31.27-110[114]
BaTiO3–Bi(Zn2/3(Nb0.85Ta0.15)1/3)O32.0678180[115]
0.9BaTiO3-0.1Bi(Mg1/2Hf1/2)O33.3887240[116]
AFEs(Pb0.91Ba0.045La0.03)(Zr0.6Sn0.4)O38.1692.1340[117]
0.84(Bi0.5Na0.5)TiO3-0.16KNbO35.288310[118]
Ag0.76La0.08NbO37.0177476[119]
Ag0.97Nd0.01Ta0.20Nb0.80O36.571370[120]
NaNbO3-Bi(Zn2/3Nb1/3)O32.490300[121]
0.85(NaNbO3)-0.15(Bi(Ni2/3Nb1/3)O3)3.3180.9440[122]
(Na0.41La0.09)(Nb0.82Ti0.18)O36.566550[123]
0.75[0.90NaNbO3-0.10Bi(Mg0.5Ta0.5)O3]0.25(Bi0.5Na0.5)0.7Sr0.3TiO3890.4800[124]
0.68NaNbO3-0.32(Bi0.5Li0.5)TiO38.7380.1-[105]
0.76NaNbO3-0.24(Bi0.5Na0.5)TiO312.269680[65]
RFEs0.93BaTiO3-0.07YNbO40.6187173[125]
0.65Bi1.05FeO3-0.35BaTiO3-(BiNa0.84K0.16)0.48Sr0.04TiO30.8160100[126]
0.93BaTiO3-0.07Sr(Zn1/3Nb2/3)O31.4583.12260[127]
0.88BaTiO3-0.12Bi(Ni2/3Nb1/3)O32.0995.9220[128]
0.02Ce-doped 0.65BaTiO3-0.35Sr0.7Bi0.2TiO32.5781.3330[129]
(Ba0.65Sr0.245Bi0.07)0.99Nd0.01TiO34.280460[130]
0.85(0.95Bi0.5Na0.5TiO3-0.05SrZrO3)-0.15NaNbO33.1479230[131]
Na0.25Bi0.25Sr0.5)(Ti0.8Sn0.2)O33.490310[132]
0.88Bi0.47Na0.47Ba0.06TiO3-0.12CaHfO34.266.7280[133]
0.75(Bi0.45La0.05Na0.5)0.94Ba0.06TiO3-0.25Sr0.8Bi0.10.1Ti0.8Zr0.2O2.953.8490.8330[134]
0.5(Na0.5Bi0.5TiO3)-0.5(Sr0.85Sm0.1TiO3)5.0290422[135]
0.8Bi0.5Na0.5TiO3-0.2SrNb0.5Al0.5O36.6496.5520[136]
0.70Bi0.5Na0.5TiO3-0.30SrNb0.5Al0.5O36.7889.7572[137]
0.85K0.5Na0.5NbO3-0.15Bi(Li0.5Ta0.5)O31.156151[138]
0.91K0.5Na0.5NbO3-0.09SrZrO32.8180370[139]
0.9(K0.5Na0.5)NbO3-0.1Bi(Zn2/3Nb1/3)O34.0197.1326[140]
[(Na0.5K0.5)0.91Li0.03](Nb0.88Sb0.06)O3-0.06Bi(Zn1/2Zr1/2)O34.8588.2480[141]
0.85K0.5Na0.5NbO3-0.15Bi(Zn2/3Ta1/3)O36.792600[142]
0.90K0.5Na0.5NbO3-0.10Bi(Zn2/3(Nb0.85Ta0.15)1/3)O37.478800[143]
0.85K0.5Na0.5NbO3-0.15Bi(Ni0.5Zr0.5)O38.0988.46870[144]
Table 2. Energy storage properties of ceramic films.
Table 2. Energy storage properties of ceramic films.
Film CompositionWrec (J/cm3)η (%)Eb (kV/cm)Refs.
BiFeO3-BaTiO3-SrTiO3112805.3 × 103[152]
0.5Ba(Zr0.2Ti0.8)O3–0.5(Ba0.7Ca0.3)TiO3 (BCZT)99.871750[153]
0.6(Bi0.5Na0.5)TiO3–0.4Bi(Ni0.5Zr0.5)O350.163.92200[154]
Mn-doped 0.97(0.93Na0.5Bi 0.5TiO3-0.07BaTiO3)-0.03BiFeO381.964.42285[155]
Mn-doped 0.55(0.94Na0.5Bi0.5TiO3-0.06BaTiO3)-0.45SrTiO376.1802813[156]
Ba(Zr0.35Ti0.65)O365.172.96.15 × 103[157]
Sn-doped In2O3/BaZr0.35Ti0.65O340.668.94.23 × 103[158]
0.9Bi0.2Sr0.7TiO3–0.1BiFeO348.547.574800[159]
Mn-doped BiFeO3–BaTiO380783.1 × 103[160]
0.5(Bi0.5Na0.5)TiO3-0.5Bi(Zn0.5Zr0.5)O340.864.11500[161]
0.88Ba0.55Sr0.45TiO3–0.12BiMg2/3Nb1/3O386735 × 103[162]
0.3Bi(Fe0.95Mn0.05)O3-0.7(Sr0.7Bi0.2)TiO361753000[163]
(Na0.8K0.2)0.5Bi0.5TiO3/0.6(Na0.8K0.2)0.5Bi0.5TiO3-0.4SrTiO373.768.12308[164]
Na0.5Bi3.25La1.25Ti4O15/BaBi3.4Pr0.6Ti4O15159.7703450[165]
Mn-doped 0.65(0.94Na0.5Bi0.5TiO3–0.06BaTiO3)–0.35SrTiO356662738[166]
Sr0.975(Bi0.5Li0.5)0.025Ti0.99Mn0.01O347.766.53307[167]
Ba(Zr0.1Ti0.9)O315.569.81500[168]
HfO2/Al2O3/ZrO254.351.35000[169]
Table 3. Energy storage properties of MLCCs.
Table 3. Energy storage properties of MLCCs.
MLCC CompositionThickness/No. of Active LayersWrec (J/cm3)η (%)Eb (kV/cm)Refs.
0.75(Bi1-xNdx)FeO3-0.25BaTiO3 (x = 15%)32 µm/96.7477540[177]
0.87BaTiO3-0.13Bi(Zn2/3(Nb0.85Ta0.15)1/3)O317 µm/108.1395750[196]
0.55(Bi0.5Na0.5)TiO3-0.45(Bi0.2Sr0.7)TiO320 µm/109.592720[5]
0.62BF–0.3BT-0.08NdZn0.5Zr0.5O316 µm/710.587700[178]
Sm0.05Ag0.85Nb0.7Ta0.3O310 µm14851450[195]
0.5BiFeO3-0.4SrTiO3-0.03Nb-0.1 BiMg2/3Nb1/3O38 µm15.875.21000[198]
<111>Na0.5Bi0.5TiO3-Sr0.7Bi0.2TiO320 µm/1021.565103 × 103[194]
Ba0.3Sr0.7TiO3/0.85BaTiO3-0.15Bi(Mg0.5Zr0.5)O3230 nm/830.6470.933000[199]
Ba0.7Ca0.3TiO3-BaZr0.2Ti0.8O3100 nm/852.472.34.5 × 103[176]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pattipaka, S.; Lim, Y.; Son, Y.H.; Bae, Y.M.; Peddigari, M.; Hwang, G.-T. Ceramic-Based Dielectric Materials for Energy Storage Capacitor Applications. Materials 2024, 17, 2277. https://doi.org/10.3390/ma17102277

AMA Style

Pattipaka S, Lim Y, Son YH, Bae YM, Peddigari M, Hwang G-T. Ceramic-Based Dielectric Materials for Energy Storage Capacitor Applications. Materials. 2024; 17(10):2277. https://doi.org/10.3390/ma17102277

Chicago/Turabian Style

Pattipaka, Srinivas, Yeseul Lim, Yong Hoon Son, Young Min Bae, Mahesh Peddigari, and Geon-Tae Hwang. 2024. "Ceramic-Based Dielectric Materials for Energy Storage Capacitor Applications" Materials 17, no. 10: 2277. https://doi.org/10.3390/ma17102277

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop