Next Article in Journal
β-Sitosterol Oleate Synthesis by Candida rugosa Lipase in a Solvent-Free Mini Reactor System: Free and Immobilized on Chitosan-Alginate Beads
Previous Article in Journal
Transition Metal-Free Synthesis of 3-Acylquinolines through Formal [4+2] Annulation of Anthranils and Enaminones
Previous Article in Special Issue
A New 2D Metal-Organic Framework for Photocatalytic Degradation of Organic Dyes in Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Lead-Free Bismuth Based Electroceramic Compositions for High-Energy Storage Density Applications in Electroceramic Capacitors

1
State Key Laboratory for Chemistry and Chemical Engineering of Medicinal Resources, School of Chemistry and Pharmacy, Guangxi Normal University, Guilin 541004, China
2
Drug Discovery Research Center, Southwest Medical University, Luzhou 646000, China
3
Laboratory for Cardiovascular Pharmacology, Department of Pharmacology, The School of Pharmacy, Southwest Medical University, Luzhou 646000, China
4
Department of Chemistry, University of Science and Technology, Bannu 28100, Pakistan
5
Advanced Materials Research Laboratory, Department of Physics, University of Science and Technology, Bannu 28100, Pakistan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2023, 13(4), 779; https://doi.org/10.3390/catal13040779
Submission received: 23 February 2023 / Revised: 27 March 2023 / Accepted: 30 March 2023 / Published: 21 April 2023
(This article belongs to the Special Issue Metal-Organic Frameworks in Electro and Photocatalysis)

Abstract

:
Lead-based electro-ceramic compositions are excellent energy storage materials used for high-energy storage density applications in dielectric ceramic capacitors. However, these materials have lead contents in their compositions, making them toxic, with a negative impact on human health and the environment. For this reason, we synthesized a lead-free bismuth-based electro-ceramic perovskite, 0.80(0.92Bi1/5Na1/5TiO3-0.08BaTiO3)-0.20(Na0.73Bi0.09NbO3−xTa2O5), abbreviated (BNT-BT-NBN1−xTx), from mixed oxides with doping of tantalum (Ta) at different concentrations, using a conventional solid-state reaction method. The effects of Ta doping on the phase evolution, microstructure development, and energy storage applications were investigated. Detailed powder X-ray diffraction analysis revealed a pure perovskite phase with Ta doping at ≤0.05. Furthermore, it was observed that excessive addition of Ta has been resulted in secondary phase generation. Scanning electron microscopy validated the development of dense microstructures with a reduced grain size for the Ta concentration of ≤0.01. Electrochemical analysis revealed a maximum polarization (Pm) of ~22 µC/cm2 and a recoverable energy density of 1.57 J/cm3 with 80% efficiency for Ta doping at 0.05 with an applied field of 175 kV/cm. These results demonstrate the development of enhanced ferroelectric characteristics in an as-synthesized electro-ceramic perovskite for high-energy storage density applications in electro-ceramic capacitors.

1. Introduction

Currently, 11% of consumable energy is provided by renewable resources, while the remaining 84% is still obtained via the burning of fossil fuels such as oil, coal, and gas. Consequently, the burning of fossil fuels for energy purposes also discharges an enormous amount of carbon dioxide, which is considered the most harmful gas among the greenhouse gases and accountable for global warming and climate change [1]. Consequently, it is essential to reduce temperature to less than 2 °C, as signed and agreed by world leaders at the Paris Agreement, with the aim of decreasing carbon dioxide emissions by 45% by 2030 and reaching net-zero emissions by 2050 [2,3]. In this perspective, advances in renewable-harvesting technologies can play a significant role in achieving the objectives of the Paris Agreement, by obtaining greater amounts of clean energy from renewable resources such as the sun, wind, and tide. Therefore, in order to reduce the burning of fossil fuels and harvest renewables for clean energy, it is of utmost importance to integrate energy-harvesting technologies with energy storage devices, particularly for the electrification of power trains, planes, ships, electric cars, and military vehicles for shipment and transportation purposes. For this purpose, harvesting renewables is crucial for lessening the dependency on fossil fuel usage, which will ultimately result in carbon dioxide settling in the upper atmosphere, a key step towards the remediation of global warming and climate change [4,5,6].
Currently, various composites and condensed solid solutions, such as polymeric and ceramic compositions, are used in high-energy storage devices for high-energy storage density applications [7]. In this regard, lead-based electro-ceramics have exhibited excellent energy storage density characteristics and applications; however, lead-based electro-ceramic materials have toxicity due to their lead contents, posing human health risks in consumer-level devices such as capacitors and actuators. Despite having excellent properties, lead-based electro-ceramics pose human health risks and environmental concerns [8]. Therefore, lead-free alternatives having the characteristics of nontoxicity, eco-friendliness, and high-energy storage density applications are sought [9]. In this perspective, bismuth-based electro-ceramic compositions have recently received significant research interest. Consequently, a tremendous amount of current research has focused on bismuth-based electro-ceramic compositions owing to their excellent, eco-friendly ferroelectric and piezoelectric properties for high-energy storage density applications in energy storage devices [10].
Currently, various energy storage devices such as fuel cells, batteries, and capacitors are applied for high-energy storage density applications [11]. Among them, dielectric ceramic capacitors have demonstrated outstanding features of a high power density, good exhaustion and fatigue resistance, fast charging/discharging capacity, and longer life consistency [12]. Dielectric capacitors are typically manufactured either from polymeric or condensed ceramic materials [13]. Polymeric-based capacitors have a high output power at room temperature; however, they cannot withstand high-temperature (~100 °C) conditions [14]. Nonetheless, ceramic-based capacitors are equally workable under high-temperature conditions [15].
The storage of electric energy per unit volume by a polarizing dielectric material with an external electric field is known as the energy storage density (Ws), and the amount of energy released per unit volume via a depolarizing dielectric by switching off an external applied electric field is recognized as the recoverable energy density (Wrec). Both energy densities, Ws and Wrec, are determined from the polarization vs. electric field (P–E) hysteresis loop using Equations (1) and (2) [16,17], as follows:
Ws = 0 P max EdP
W rec = P r P max EdP
where Pr represents remnant polarization, also known as residual polarization; Pmax stands for maximum polarization, also known as saturation polarization; and E represents the external applied electric field. The efficiency of dielectric ceramics is determined by Equation (3), as follows:
η = W rec Ws × 100
It is obvious from Equation (2) that ceramic materials with zero residual polarization are preferable for realizing a high recoverable energy density. Further, dielectric ceramic materials that are almost linear satisfy this phenomenon; however, linear dielectric ceramics have a lower permittivity (εr), due to which their recoverable energy density is limited. Although linear dielectric ceramic materials exhibit a high dielectric breakdown strength (DBS) and efficiency (η), their Pmax is still small; hence, they are not feasible for high-recoverable-energy storage density applications [18]. It is obvious from Equation (1) that Ws depends both on a high Pmax and the applied electric field E; nonetheless, Wrec depends on the maximal value of ∆P = Pmax − Pr, which is only achievable by remanence-free ceramic materials. Consequently, it is crucial to optimize both of the above parameters in order to obtain a high energy storage density and recoverable energy density in dielectric ceramic capacitors. Based on their polarization vs. electric field (P–E) hysteresis loops, dielectric ceramic materials are classified into two main categories: linear and nonlinear dielectrics [19,20]. The polarizations and electric fields are directly proportional in linear dielectric ceramics, though linear dielectrics have a high DBS; however, they have a lower dielectric permittivity (εr) and Pmax, hence performing poorly in energy storage applications [21]. In the case of nonlinear dielectric ceramic materials, the polarization and applied electric field are independent of each other [22]. For the purpose of carrying out steady and comprehensive research, nonlinear dielectric ceramics are further divided into four types: paraelectrics, ferroelectrics (FEs), relaxor ferroelectrics (RFEs), and anti-ferroelectrics (AFEs) [16,23]. At the macroscopic level, both paraelectric and ferroelectric materials are alike; however, at the microscopic level, spontaneous polarization is absent in paraelectric materials, and the net polarization becomes zero as soon as the external applied electric field is removed. Moreover, ferroelectric materials retain their polarization even after the removal of the external electric field. This is the reason why paraelectric ceramics are not feasible for energy storage applications. A large number of ferroelectric ceramics have a large Pr and small DBS, due to which their energy storage density is low [24]. The double P–E hysteresis loop permits anti-ferroelectric (AFE) materials to store a high power density, but the majority of these materials have lead contents in their compositions that can cause toxicity and pose human health risks and severe environmental impacts [25]. In the case of lead-free AFEs, the phase transition of anti-ferroelectrics to ferroelectrics is a more challenging phenomenon due to the long-term cycling operations. Therefore, only relaxor ferroelectric materials have the characteristics of highly polar nano-regions (PNRs), minimal Pr, small Ec, and high DBS, making them feasible for high-power storage density applications. As a consequence, an increasing amount of research interest has been directed towards lead-free relaxor ferroelectric ceramics recently for high-energy storage density and recoverable energy density applications [26].
Currently, the lead-free bismuth-based ceramic composition (Na0.5Bi0.5) TiO3 (NBT) has been studied comprehensively owing to its excellent ferroelectric characteristics such as its maximal polarization (Pmax~40 μC/cm2) and high Curie temperature (Tc~320 °C); however, it has high field coercivity (Ec) and poor piezoelectric properties. Hence, it has been recommended to dope it with other perovskite solid solutions for high-energy storage density and recoverable energy density applications [27,28]. To introduce relaxor behaviors in NBT, it has been modified via doping with other perovskite materials such as Bi0.5K0.5TiO3 (BKT), K0.5Na0.5NbO3 (KNN), and BaTiO3 (BT) [29,30]. In this regard, Praharaj et al. [31] prepared 0.76Na0.5Bi0.5TiO3-0.2SrTiO3-0.04BaTiO3 by doping BaTiO3 and SrTiO3 into the NBT structure, which resulted in the development of a morphotropic phase boundary (MPB). Ceramic compositions with an MPB are typically applied as matrices in high-energy storage density applications owing to their pinched (P–E) loops, high Pmax, and minimal Pr [32,33]. In this connection, Li et al. [34] fabricated a lead-free ternary ceramic composition (1 − x) (BNBT6-xCZ) by doping CaZrO3 in a binary ceramic system, which demonstrated an energy storage density of Ws~0.70 J/cm3 at a low electric field of 70 kV/cm. Likewise, Wang et al. [35] reported a ternary ceramic system of (1 − x) Bi0.47Na0.47Ba0.06TiO3-KNbO3 (BNBT-xKN, x = 0–0.08) with Ws~0.89 J/cm3 at an applied electric field of 100 KV/cm, by adding KNbO3 in a binary ceramic system of BNT-BT. In another similar study, Zhu et al. [36] synthesized a lead-free NBT-based ternary ceramic system, BNT-ST-BMH, by adding SrTiO3 and Bi(Mg0.5Hf0.5)O3, which displayed a high Wrec of ~5.59 J/cm3 with 85.3% efficiency. Likewise, Zhang et al. [37] studied the ternary ceramic system (1 − x) (0.775NBT-0.225BSN)-xBZ by adding BaZrO3, which achieved a high Ws of ~2.08 J/cm3 and 88.8% efficiency at an applied electric field of 245 kV/cm. In another study, Nilkhao et al. [38] reported the partial doping of Ta in Nb in 0.9(0.92Bi0.5Na0.5TiO3-0.08BaTiO3)-0.01(NaNb1−3xTaxO3), which achieved an energy storage density of Ws~0.37 J/cm 3 and efficiency of 75.1%. In the same way, Li. et al. [39] fabricated a binary ceramic system with doping of La, (1 − x) (0.72Bi0.5Na0.5TiO3-0.28 Bi0.2Sr0.7-0.1TiO3)-xLa, which achieved an energy storage density of 1.2 J/cm3 at an applied field of 90 kV/cm. Similarly, Khan et al. [40] fabricated an NBT-based ternary ceramic composition, (1 − z) (0.94Na0.5Bi0.5TiO3-0.06BaTiO3)-zNd0.33NbO3, which achieved an excellent recoverable energy density of 2.01 J/cm3 and efficiency of 62.5% with a z content of 0.02 at an applied field of 250 kV/cm. Likewise, Zhang et al. [41] fabricated a lead-free binary ceramic composition, (1 − x) Na0.5Bi0.5TiO3-xBaHfO3, abbreviated as (1 − x) NBT-xBH, which showed a recoverable energy density of 2.1 J/cm3 at a low applied field of 175 kV/cm. Meanwhile, using the same conventional solid-state reaction method, Wang et al. [42] synthesized a binary ceramic system by doping CaTiO3 at a level of 0.2, [0.8Bi0.5Na0.5TiO3-0.2CaTiO3, (0.8BNT-0.2CT)], which achieved an energy storage density of Ws~1.38 J/cm3 with an excellent efficiency of 91% at a low applied field.
Here, we fabricated an electro-ceramic composition, 0.80(0.92Bi0.5Na0.5TiO3-0.08BaTiO3)-0.20Na0.73Bi0.09NbO3, abbreviated as (NBT-BT-NBN), using a conventional solid-state reaction method, by doping Ta in (BNT-BT-NBN1−xTx) in a series of x = 0.00, 0.05, 0.10, 0.15, 0.20, in order to investigate the doping effects of Ta on the as-synthesized ceramic compositions’ phase evolution, microstructure development, dielectric characteristics, and energy storage applications in electro-ceramic capacitors [43].

2. Results and Discussions

2.1. Phase Examination

Figure 1 depicts the PXRD patterns of the ceramic compositions of (NBT-BT-NBN1−xTx) in a series of (x = 0.00, 0.05, 0.10, 0.15, 0.20) recorded at room temperature with 2θ in the range of 10–70°.
The X-ray pattern of each ceramic composition matched with that of PDF Card No. 070-4760, which was attributed to the single-perovskite structure of Na0.5Bi0.5TiO3 that appeared in the tetragonal phase; however, its peaks shifted towards lower values of 2θ. The main reason for the shifts in these peaks is the substitution of bigger cations of Ba2+ (1.61 Ǻ) for Bi3+ (1.40 Ǻ), and those of Na1+ (1.60 Ǻ) and Nb5+ (0.64 Ǻ) for Ti4+ (0.60 Ǻ) [44]. Figure 1 shows a single perovskite phase that developed as the major phase of Na0.5Bi0.5TiO3, represented by the peaks marked as “*”. However, there were some very low intensity secondary peaks marked as “#” that matched with those of PDF Card No. 32-0118, indicating that Bi2Ti2O7 bismuth-based pyrochlore also appeared due to the high concentration of NaBiNbO3. The formation of these secondary phases was also reported by Ni et al. [45] while investigating the ceramic composition of Bi (0.5+x) (Na0.82K0.18)0.5−3xTiO3. This secondary bismuth-based pyrochlore phase Bi2Ti2O7 might develop due to the vaporization of Bi and Na at rising temperatures. Furthermore, the intensity of Bi2Ti2O7 was increased with the increasing z content to 0.08 in the current formulation. The second reason for the appearance of secondary phases may be due to the excessive addition of Ta as a dopant, which did not dissolve in the ternary system of x ≥ 0.10. The few strong peaks of secondary phases that appeared at x = 0.20 were possibly due to the excess Ta doping as compared to x = 0.05, where no such secondary phase developments appeared in its XRD pattern.

2.2. Microstructure Examination

Figure 2 shows SEM micrographs of the thermally etched surfaces of all dense ceramic compositions of (NBT-BT-NBN1−xTx) in a series of (x = 0.00, 0.05, 0.10, 0.15, 0.20) recorded with a scanning electron microscope. It is obvious from the SEM micrographs that all sintered samples of the ceramic compositions of (NBT-BT-NBN1−xTx) had compact, dense microstructures with well-grown grains. The average grain size of each sample was determined via a line intercept method based on SEM photographs at a lower magnification. The mean grain size gradually increased from 1.4 to 1.7 µm for x = 0.00–0.20, respectively. The minimal growth in grain size for x > 0.05 shows that the Ta substitution had little influence on the sintering behavior and grain growth on the microstructures of the as-fabricated ceramic compositions [46]. Bi and Na elements become volatile at high temperatures and induce oxygen vacancies, performing a significant role in the transportation of mass during the sintering process. In the sintering phenomenon, sodium vacancies VNa are formed in the crystal lattice through the volatilization of Na elements. There are defect dipoles formed by oxygen vacancies VNa-VO-VNa that compensate the electrical neutrality, which result in small grains due to the grain boundary pinning effect [47,48]. During the coupling of NaBiNbO3 with the BNT-BT system, Nb+5 can replace Ti+4 to form NbTi, which neutralizes a fraction of the negative charge VNa, reducing the concentration of defect dipoles VNa-VO-VNa that result in continuous grain size growth [48,49]. Secondary phases of the bismuth pyrochlore Bi2Ti2O7 that developed in the microstructure were also confirmed via energy-dispersive spectroscopy (EDS), as depicted in Figure 2e.

2.3. Dielectric Properties

Figure 3 shows the dielectric response including the permittivity and dielectric loss for the as-synthesized ternary (NBT-BT-NBN1−xTx) (x = 0.00, 0.05, 0.10, 0.15, 0.20) electro-ceramic compositions. The dielectric permittivity was measured at different frequencies, namely, 1 kHz, 10 kHz, 100 kHz, and 1 MHz, in the temperature range of 20–500 °C. Two separate dielectric peaks in the curves of the dielectric response of BNT-BT ceramic samples have already been reported by Xu et al. [48]. One peak is observed at a lower temperature, denoted as Ts, and the other is observed at a high temperature, denoted as Tm. Meanwhile, after adding 20% NaNbO3 to BNT-BT, Tm was found to shift to a lower temperature below 50 °C.
All the above ceramic compositions exhibited significant variations in the dielectric permittivity and loss tangent (tan δ) with increasing temperature, which could be attributed to the changes in the structure as well as the local ferroelectric order. A decrease in the dielectric permittivity was observed with increasing x content in the range of 0.00–0.15, along with an increase in the dielectric permittivity with a further increase in the x content to 0.2. The lower content of the Ta doping has resulted in a decrease in the dielectric polarizability per unit volume of Ta [38]. The increase in the relative permittivity (εr) with a high content of the Ta dopant may be due to the formation of secondary phases, as depicted in the XRD patterns. The Ta doping shifted Tm from 45 °C to 80 °C. Here, Tm showed broad permittivity peaks and frequency dispersion. However, at an x content of 0.20, the observed εr was higher than that of x < 0.2 compositions. Moreover, there was a decreasing trend in εr for each ceramic sample with increasing frequency. This could be ceased at certain polarizations with higher frequencies. The dielectric loss also deceased with Ta doping at x ≤ 0.05, as shown in Figure 4a,b, but further increases in Ta doping had little effect on the dielectric loss. The tan δ loss was the lowest beyond x ≥ 0.10, as depicted in Figure 4c–e.

2.4. P–E Hysteresis Loops and Energy Storage

The polarization vs. electric field (P–E) hysteresis loops for the (NBT-BT-NBN1−xTx) (x = 0.00, 0.05, 0.10, 0.15, 0.20) ceramic compositions are depicted in Figure 5a. The P–E loops were measured at a frequency of 10 Hz for each ceramic composition with a 0.3 mm thickness. Large variations in the maximum polarization (Pm) and breakdown electric fields (Eb) were observed for each ceramic composition. The energy storage characteristics, such as energy storage/charged density (Ws), recoverable or discharge energy density (Wrec), and efficiency (η), of each ceramic sample were determined from its P–E hysteresis loop area. Each sample has its own breakdown voltage related to the average grain size and density of the sample. The P–E loop for each composition was measured at its breakdown voltage.
Figure 5b shows that the maximum polarization for x = 0.00 was 20 µC/cm2. Meanwhile, there was somewhat of an increase in Pmax to 22 μC/cm2 for x = 0.05 at a breakdown field of 170 kV/cm, which then decreased after further increasing the x content to 0.20. The value of Pmax-Pr was also higher for x = 0.05, and these high values of Pmax and Pmax-Pr for x = 0.05 contributed to high charge power and recoverable energy densities at a low external electric field of 170 kV/cm.

2.5. Energy Storage Density Analysis

The energy storage density (Ws) and recoverable energy density (Wrec) were calculated by integrating the P–E hysteresis loop areas of each ceramic composition. Figure 5c displays the variation in Ws and Wrec for the (NBT-BT-NBN1−xTx) (x = 0.00, 0.05, 0.10, 0.15, 0.20) ceramic compositions as a function of the x content. In this work, Ws~1.96 J/cm3, Wrec~1.57 J/cm3, and an efficiency (η) of 80% were obtained for x = 0.05. The large values of Wrec and Ws for x = 0.05 could be credited to the maximal values of Pmax and Pm-Pr as compared to other compositions with higher values of x in this research work. The variation in the energy density loss (Wloss) and efficiency (η) for the (NBT-BT-NBN1−xTx) (x = 0.00, 0.05, 0.10, 0.15, 0.20) ceramic compositions as a function of the x content is shown in Figure 5d. In this research work, the maximum efficiency of 80% was achieved by the ceramic composition x = 0.05.
A comparative analysis of the current study with previously published works on NBT-based ceramics is presented in Table 1. The analysis shows that the Wrec obtained for (NBT-BT-NBN1-xTx) (x = 0.05) in the present study is higher than that described for other NBT-based ceramic systems.

3. Experimental Procedures

3.1. Materials

The chemicals and reagents used in this research work were sodium carbonate, Na2CO3 (99.9%), barium carbonate, BaCO3 (99.5%), bismuth oxide, Bi2O3 (99%), titanium dioxide, TiO2 (99.8%), niobium oxide, Nb2O5 (99%), tantalum oxide, Ta2O5 (99.8%), ethanol, C2H5OH, (99.7%), and poly-vinyl alcohol. All these chemicals and reagents were purchased from a Sigma Aldrich manufacturer. Anhydrous ethanol and de-ionized water were used for dissolution purposes.

3.2. Fabrication of [0.80(0.92Bi1/5Na1/5TiO3-0.08BaTiO3)-0.20(Na0.73Bi0.09NbO3−xTa2O5)]

Several processes are applied to fabricate lead-free electro-ceramic compositions, such as the sol–gel, mechanochemical, and current-assisted sintering processes; however, the majority of bulk ceramics are fabricated using a conventional solid-state reaction method, typically from mixed metal oxide–carbonate combinations. In this research work, a conventional solid-state reaction method was applied to fabricate the ceramic composition (NBT-BT-NBN1−xTx) in a series of (x = 0.00, 0.05, 0.10, 0.15, 0.20). The reactant materials were selected suitably with a preference for high purity and small particle sizes in the sub-micron range. The raw materials in powder forms were first heated at 100 °C overnight to remove the trapped water moisture contents. These reagents in powder form were weighed in stoichiometric ratios and ball-milled in plastic containers with ZrO2 grinding balls, and ethanol was added as a lubricant for 24 h. The milling time is important in enhancing the dielectric properties of ceramic compositions. After finishing the ball milling, the resulted slurries were dried at 95 °C in an oven and then grinded to produce the powders. The milled powders of each batch sample were calcined at 850 °C for 2 h and again ball-milled for 12 h, after which they were dried to obtain fine powders. During the progress of the calcination process, the ball-milled powders of each batch sample were heated at a high temperature below their melting temperature, which enabled them to react for the development of the required crystalline phase. The calcination temperature varies from one material to another; therefore, significant attention was given to the calcination temperature of each batch sample. In this work, volatilization of the alkaline elements from the pellets was controlled by placing them in completely sealed alumina crucibles. A few drops of poly-vinyl alcohol (PVA) were added as a binder in the calcined samples and again grinded for 10 min, keeping each sample in a steel die with a 12 mm diameter, followed by placing it in an isotactic pressing machine and applying a uniaxial pressure of 150 MPa, in order to produce pellets of 1 mm thickness for each batch sample. All these pellets were initially heated at a temperature of 600 °C for 2 h to remove the PVA contents, and finally sintered in the temperature range of 1050–1150 °C for 2 h. All the green body pellets were subjected to thermal treatment (sintering) below their melting temperature to enhance their mechanical strength for bonding their particles together. In this process, all the material samples were heated in the high-temperature range of 1050–1150 °C below their melting points. The best sintering temperature and most suitable stay times were used for the proper densification of each batch sample. For the optimization of the sintering temperature, each sample was subjected to various sintering temperatures, and then the sintered samples with the highest density were selected for the advanced characterization.

3.3. Characterization Techniques

At maximal density, pellets of each ceramic composition were first crushed to form powders for their phase evolution analysis with the advanced powder X-ray diffraction (PXRD) technique. The data were recorded at an angle of 2ϴ (degree) in the range of 10–70° at a scan rate of 0.02°/min using an X-ray diffractometer. Dense pellets of each sample were smoothly polished and etched thermally; then, they were coated in gold to avoid the charging effect during contact with an electron beam. The surface morphology of each sample was examined using a scanning electron microscope.
Furthermore, 0.70 mm thick ceramic samples were prepared and sintered at 1150 °C. Both faces of each sample were thoroughly polished before coating them in silver and heated at 800 °C for 2 h for the evaluation of the dielectric constant and dielectric loss for each batch sample, at frequencies of 1 kHz, 10 kHz, 100 kHz, and 1 MHz, against the temperature by recording at a scan rate of 3 °C/min with an LCR meter (Agilent E498, Santa Clara, CA, USA) in the temperature range of 20–500 °C. The thickness of each dense sample was decreased to 0.25   m m ; then, the samples were coated with silver paste and heated at 800 °C for 2 h, and the polarization vs. electric field (P–E) hysteresis loops were recorded at 10 Hz for each batch sample of the fabricated ceramic compositions at the breakdown voltage using a ferroelectric analyzer (aixACCT TF-2000, UCSU, Germany).

4. Conclusions

A series (x = 0.00, 0.05, 0.10, 0.15, 0.20) of ceramic compositions of [0.80(0.92Bi0.5Na0.5TiO3-0.08BaTiO3)-0.20 Na0.73Bi0.09NbO3-xTa2O5], (NBT-BT-NBN1−xTx), were prepared using a conventional solid-state reaction method with doping of Ta. The effects of Ta doping on the as-fabricated ceramic compositions’ phase evolution, microstructure development, electrochemical properties, and energy storage density properties were investigated using various advanced characterization techniques such as advanced powder X-ray diffractometry, scanning electron microscopy, and an LCR meter. A single perovskite phase developed in all the synthesized compositions; however, excessive addition of Ta as a dopant resulted in the development of secondary phases, appearing as impurities at x ≥ 0.1. A decreasing trend was observed in the grain size for x ≤ 0.05, and a maximum polarization (Pm) of ~22 μC/cm2 and recoverable energy of 1.57 J/cm3 with an efficiency of 80% at an applied external electric field of 175 kV/cm were obtained. These results demonstrate the enhanced ferroelectric characteristics of the as-synthesized electro-ceramic perovskite compositions, which are suitable for high-energy storage density applications in electro-ceramic capacitors.

Author Contributions

A.K. and T.M.K.: conceptualization, methodology, software, validation; J.W.: resources; A.M. and H.B.: data curation, writing—draft preparation; S.Z.K.: reviewing and editing; X.W. and N.S.G.: visualization, investigation, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Drug Discovery Research Center, Southwest Medical University, Luzhou, China, under Grant No.42-00040176 awarded to TMK. This research work was partially supported by National Science Foundation of China (Grant No.21861004 and 22165003).

Data Availability Statement

Data that support the findings of this study are included within the article.

Acknowledgments

This research work was fully supported by the Drug Discovery Research Center, Southwest Medical University, Luzhou, China, under Grant No. 42-00040176, which was awarded to Taj Malook Khan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ritchie, H.; Roser, M.; Rosado, P. Energy. OurWorldInData.org. 2020. Available online: https://ourworldindata.org/co2-emissions?utm_source=tri-city%20news&utm_campaign=tricity%20news%3A%20outbound&utm_medium=referral (accessed on 22 February 2023).
  2. Vaughan, A. UN Climate Change Summit; Elsevier: Amsterdam, Netherlands, 2019. [Google Scholar]
  3. Ibn-Mohammed, T.; Randall, C.; Mustapha, K.; Guo, J.; Walker, J.; Berbano, S.; Koh, S.; Wang, D.; Sinclair, D.; Reaney, I. Decarbonising ceramic manufacturing: A techno-economic analysis of energy efficient sintering technologies in the functional materials sector. J. Eur. Ceram. Soc. 2019, 39, 5213–5235. [Google Scholar] [CrossRef]
  4. Wang, F.; Harindintwali, J.D.; Yuan, Z.; Wang, M.; Wang, F.; Li, S.; Yin, Z.; Huang, L.; Fu, Y.; Li, L. Technologies and perspectives for achieving carbon neutrality. Innovation 2021, 2, 100180. [Google Scholar] [CrossRef] [PubMed]
  5. Kalair, A.; Abas, N.; Saleem, M.S.; Kalair, A.R.; Khan, N. Role of energy storage systems in energy transition from fossil fuels to renewables. Energy Storage 2021, 3, e135. [Google Scholar] [CrossRef]
  6. García-Olivares, A.; Solé, J.; Osychenko, O. Transportation in a 100% renewable energy system. Energy Convers. Manag. 2018, 158, 266–285. [Google Scholar] [CrossRef]
  7. Wu, H.; Zhuo, F.; Qiao, H.; Venkataraman, L.K.; 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]
  8. Zaman, S.; Chen, S. A perspective on inaccurate measurements in oxygen reduction and carbon dioxide reduction reactions. J. Catal. 2023, 1, 872. [Google Scholar] [CrossRef]
  9. Wang, D.; Wang, G.; Murakami, S.; Fan, Z.; Feteira, A.; Zhou, D.; Sun, S.; Zhao, Q.; Reaney, I.M. BiFeO3-BaTiO3: A new generation of lead-free electroceramics. J. Adv. Dielect. 2018, 8, 1830004. [Google Scholar] [CrossRef]
  10. Shen, Y.; Wu, L.; Zhao, J.; Liu, J.; Tang, L.; Chen, X.; Li, H.; Su, Z.; Zhang, Y.; Zhai, J. Constructing novel binary Bi0.5Na0.5TiO3-based composite ceramics for excellent energy storage performances via defect engineering. Chem. Eng. J. 2022, 439, 135762. [Google Scholar] [CrossRef]
  11. Abbas, Q.; Mirzaeian, M.; Hunt, M.R.; Hall, P.; Raza, R. Current state and future prospects for electrochemical energy storage and conversion systems. Energies 2020, 13, 5847. [Google Scholar] [CrossRef]
  12. Wang, G.; Lu, Z.; Li, Y.; Li, L.; Ji, H.; Feteira, A.; Zhou, D.; Wang, D.; Zhang, S.; Reaney, I.M. Electroceramics for High-Energy Density Capacitors: Current Status and Future Perspectives. Chem. Rev. 2021, 121, 6124–6172. [Google Scholar] [CrossRef]
  13. Fan, B.; Liu, F.; Yang, G.; Li, H.; Zhang, G.; Jiang, S.; Wang, Q. Dielectric materials for high-temperature capacitors. Iet Nanodielectr. 2018, 1, 32–40. [Google Scholar] [CrossRef]
  14. Tan, D.Q. Review of polymer-based nanodielectric exploration and film scale-up for advanced capacitors. Adv. Funct. Mater. 2020, 30, 1808567. [Google Scholar] [CrossRef]
  15. 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]
  16. Zaman, S.; Huang, L.; Douka, A.I.; Yang, H.; You, B.; Xia, B.Y. Oxygen reduction electrocatalysts toward practical fuel cells: Progress and perspectives. Angew. Chem. Int. Ed. 2021, 60, 17832–17852. [Google Scholar] [CrossRef] [PubMed]
  17. Zhao, P.; Tang, B.; Si, F.; Yang, C.; Li, H.; Zhang, S. Novel Ca doped Sr0.7Bi0.2TiO3 lead-free relaxor ferroelectrics with high energy density and efficiency. J. Eur. Ceram. Soc. 2020, 40, 1938–1946. [Google Scholar] [CrossRef]
  18. Zhou, H.Y.; Liu, X.Q.; Zhu, X.L.; Chen, X.M. CaTiO3 linear dielectric ceramics with greatly enhanced dielectric strength and energy storage density. J. Am. Ceram. 2018, 101, 1999–2008. [Google Scholar] [CrossRef]
  19. Peddigari, M.; Palneedi, H.; Hwang, G.-T.; Ryu, J. Linear and nonlinear dielectric ceramics for high-power energy storage capacitor applications. J. Korean Ceram. 2019, 56, 1–23. [Google Scholar] [CrossRef]
  20. Curecheriu, L.P.; Mitoseriu, L.; Ianculescu, A. Nonlinear dielectric properties of Ba1− xSrxTiO3 ceramics. J. Alloy Compd. 2009, 482, 1–4. [Google Scholar] [CrossRef]
  21. 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]
  22. Zaman, S.; Wang, M.; Liu, H.; Sun, F.; Yu, Y.; Shui, J. Carbon-based catalyst supports for oxygen reduction in proton-exchange membrane fuel cells. Trends Chem. 2022, 4, 886–906. [Google Scholar] [CrossRef]
  23. Jayakrishnan, A.; Silva, J.P.B.; Kamakshi, K.; Dastan, D.; Annapureddy, V.; Pereira, M.; Sekhar, K. Are lead-free relaxor ferroelectric materials the most promising candidates for energy storage capacitors? Prog. Mater. Sci. 2022, 2, 101046. [Google Scholar] [CrossRef]
  24. Ali, H.; Zaman, S.; Majeed, I.; Kanodarwala, F.K.; Nadeem, M.A.; Stride, J.A. Porous Carbon/rGO Composite: An Ideal Support Material of Highly Efficient Palladium Electrocatalysts for the Formic Acid Oxidation Reaction. ChemElectroChem 2017, 4, 3126–3133. [Google Scholar] [CrossRef]
  25. Wei, J.; Yang, T.; Wang, H. Excellent energy storage and charge-discharge performances in PbHfO3 antiferroelectric ceramics. J. Eur. Ceram. Soc. 2019, 39, 624–630. [Google Scholar] [CrossRef]
  26. Wang, Q.; Wang, T.; Zhang, L.; Liu, Z.; Guo, K.; Lu, J.; Xie, B. High energy-storage performance of lead-free Ba0.4Sr0.6TiO3–Sr0.7Bi0.2TiO3 relaxor-ferroelectric ceramics with ultrafine grain size. Ceramics 2022, 48, 2068–2074. [Google Scholar] [CrossRef]
  27. Fan, P.; Zhang, S.-T.; Xu, J.; Zang, J.; Samart, C.; Zhang, T.; Tan, H.; Salamon, D.; Zhang, H.; Liu, G. Relaxor/antiferroelectric composites: A solution to achieve high energy storage performance in lead-free dielectric ceramics. J. Mater. Chem. C 2020, 8, 5681–5691. [Google Scholar] [CrossRef]
  28. Zaman, S.; Tian, X.; Su, Y.-Q.; Cai, W.; Yan, Y.; Qi, R. Direct integration of ultralow-platinum alloy into nanocarbon architectures for efficient oxygen reduction in fuel cells. Sci. Bull. 2021, 66, 2207–2216. [Google Scholar] [CrossRef]
  29. Gao, F.; Dong, X.; Mao, C.; Liu, W.; Zhang, H.; Yang, L.; Cao, F.; Wang, G. Energy-storage properties of 0.89 Bi0. 5Na0. 5TiO3–0.06 BaTiO3–0.05 K0. 5Na0. 5NbO3 lead-free anti-ferroelectric ceramics. J. Am. Ceram. 2011, 94, 4382–4386. [Google Scholar] [CrossRef]
  30. Chen, S.; Zhang, Z.; Jiang, W.; Zhang, S.; Zhu, J.; Wang, L. Engineering Water Molecules Activation Center on Multisite Electrocatalysts for Enhanced CO2 Methanation. J. Am. Chem. Soc. 2022, 144, 12807–12815. [Google Scholar] [CrossRef]
  31. Praharaj, S.; Rout, D. Electrical conductivity and modulus study of 0.76 Na0.5Bi0.5TiO3-0.2SrTiO3-0.04BaTiO3 ceramic: A near morphotropic phase boundary composition. J. Phys. Chem. Solids 2019, 127, 52–59. [Google Scholar] [CrossRef]
  32. Swain, S.; Kar, S.K.; Kumar, P. Dielectric, optical, piezoelectric and ferroelectric studies of NBT–BT ceramics near MPB. Ceramics 2015, 41, 10710–10717. [Google Scholar] [CrossRef]
  33. Makarovic, M.; Bencan, A.; Walker, J.; Malic, B.; Rojac, T. Processing, piezoelectric and ferroelectric properties of (x) BiFeO3-(1-x) SrTiO3 ceramics. J. Eur. Ceram. Soc. 2019, 39, 3693–3702. [Google Scholar] [CrossRef]
  34. Li, Q.; Wang, J.; Ma, Y.; Ma, L.; Dong, G.; Fan, H. Enhanced energy-storage performance and dielectric characterization of 0.94 Bi0. 5Na0. 5TiO3–0.06 BaTiO3 modified by CaZrO3. J. Alloy. Compd. 2016, 663, 701–707. [Google Scholar] [CrossRef]
  35. Wang, B.; Luo, L.; Jiang, X.; Li, W.; Chen, H. Energy-storage properties of (1−x) Bi0. 47Na0. 47Ba0. 06TiO3–xKNbO3 lead-free ceramics. J. Alloy. Compd. 2014, 585, 14–18. [Google Scholar] [CrossRef]
  36. Zhu, X.; Shi, P.; Gao, Y.; Kang, R.; Zhao, J.; Xiao, A.; Qiao, W.; Zhao, J.; Wang, Z.; Lou, X. Enhanced energy storage performance of 0.88 (0.65 Bi0. 5Na0. 5TiO3-0.35 SrTiO3)-0.12 Bi (Mg0. 5Hf0. 5) O3 lead-free relaxor ceramic by composition design strategy. Chem. Eng. J. 2022, 437, 135462. [Google Scholar] [CrossRef]
  37. Zhang, L.; Pu, Y. Influence of BaZrO3 additive on the energy-storage properties of 0.775 Na0. 5Bi0. 5TiO3-0.225 BaSnO3 relaxor ferroelectrics. J. Alloy Compd. 2019, 775, 342–347. [Google Scholar] [CrossRef]
  38. Zaman, S.; Douka, A.I.; Noureen, L.; Tian, X.; Ajmal, Z.; Wang, H. Oxygen reduction performance measurements: Discrepancies against benchmarks. Battery Energy 2023, 1, 20220060. [Google Scholar] [CrossRef]
  39. Khan, A.; Gul, N.S.; Luo, M.; Wu, J.; Khan, S.Z.; Manan, A.; Wang, X.-J.; Khan, T.M. Fabrication of a lead-free ternary ceramic system for high energy storage applications in dielectric capacitors. Front. Chem. 2022, 10, 2762. [Google Scholar] [CrossRef]
  40. Zhang, L.; Pu, Y.; Chen, M.; Wei, T.; Keipper, W.; Shi, R.; Guo, X.; Li, R.; Peng, X. High energy-storage density under low electric fields and improved optical transparency in novel sodium bismuth titanate-based lead-free ceramics. J. Eur. Ceram. Soc. 2020, 40, 71–77. [Google Scholar] [CrossRef]
  41. Wang, J.; Zhang, F.; Peng, Z.; Chao, X.; Yang, Z. Sodium bismuth titanate-based perovskite ceramics with high energy storage efficiency and discharge performance. J. Mater. 2022, 8, 1077–1085. [Google Scholar] [CrossRef]
  42. Zaman, S.; Su, Y.Q.; Dong, C.L.; Qi, R.; Huang, L.; Qin, Y. Scalable molten salt synthesis of platinum alloys planted in metal-nitrogen-graphene for efficient oxygen reduction. Angew. Chem. Int. Ed. 2022, 61, 202115835. [Google Scholar] [CrossRef]
  43. Kumari, N.; Monga, S.; Arif, M.; Sharma, N.; Sanger, A.; Singh, A.; Vilarinho, P.M.; Gupta, V.; Sreenivas, K.; Katiyar, R.S. Multifunctional behavior of acceptor-cation substitution at higher doping concentration in PZT ceramics. Ceram. Int. 2019, 45, 12716–12726. [Google Scholar] [CrossRef]
  44. Ni, F.; Luo, L.; Pan, X.; Li, W.; Zhu, J.-Q. Effects of A-site vacancy on the electrical properties in lead-free non-stoichiometric ceramics Bi0. 5+ x (Na0. 82K0. 18) 0.5− 3xTiO3 and Bi0. 5+ y (Na0. 82K0. 18) 0.5 TiO3. J. Alloy. Compd. 2012, 541, 150–156. [Google Scholar] [CrossRef]
  45. Wang, G.; Zhang, D.; Huang, X.; Rao, Y.; Yang, Y.; Gan, G.; Lai, Y.; Xu, F.; Li, J.; Liao, Y. Crystal structure and enhanced microwave dielectric properties of Ta5+ substituted Li3Mg2NbO6 ceramics. J. Am. Ceram. Soc. 2020, 103, 214–223. [Google Scholar] [CrossRef]
  46. Xu, Q.; Liu, H.; Zhang, L.; Xie, J.; Hao, H.; Cao, M.; Yao, Z.; Lanagan, M.T. Structure and electrical properties of lead-free Bi 0.5 Na 0.5 TiO3-based ceramics for energy-storage applications. RSC Adv. 2016, 6, 59280–59291. [Google Scholar] [CrossRef]
  47. Sung, Y.; Kim, J.; Cho, J.; Song, T.; Kim, M.; Chong, H.; Park, T.; Do, D.; Kim, S. Effects of Na nonstoichiometry in (Bi 0.5 Na 0.5+ x) TiO 3 ceramics. Appl. Phys. Lett. 2010, 96, 022901. [Google Scholar] [CrossRef]
  48. Xu, Q.; Li, T.; Hao, H.; Zhang, S.; Wang, Z.; Cao, M.; Yao, Z.; Liu, H. Enhanced energy storage properties of NaNbO3 modified Bi0.5Na0.5TiO3 based ceramics. J. Eur. Ceram. Soc. 2015, 35, 545–553. [Google Scholar] [CrossRef]
  49. Xu, Q.; Lanagan, M.T.; Huang, X.; Xie, J.; Zhang, L.; Hao, H.; Liu, H. Dielectric behavior and impedance spectroscopy in lead-free BNT–BT–NBN perovskite ceramics for energy storage. Ceram. Int. 2016, 42, 9728–9736. [Google Scholar] [CrossRef]
  50. Li, F.; Liu, Y.; Lyu, Y.; Qi, Y.; Yu, Z.; Lu, C. Huge Strain and Energy Storage Density of A-site La3+ Donor Doped (Bi0.5Na0.5)0.94Ba0.06TiO3 Ceramics. Ceram. Int. 2017, 43, 106–110. [Google Scholar] [CrossRef]
  51. High energy storage properties and dielectric behavior of (Bi0.5 Na0.5) 0.94 Ba0.06Ti1−x(Al0.5Nb0.5)xO3 lead-free ferroelectric ceramics. Ceram. Int. 2016, 42, 2221–2226. [CrossRef]
  52. Li, Q.; Yao, Z.; Ning, L.; Gao, S.; Hu, B.; Dong, G.; Fan, H. Enhanced energy-storage properties of (1-x)(0.7Bi0.5Na0.5TiO3-0.3Bi0.2Sr0.7TiO3)-xNaNbO3 lead-free ceramics. Ceram. Int. 2018, 44, 2782–2788. [Google Scholar] [CrossRef]
  53. Yu, Z.; Liu, Y.; Shen, M.; Qian, H.; Li, F.; Lyu, Y. Enhanced energy storage properties of BiAlO3 modified Bi0.5Na0.5TiO3–Bi0.5K0.5TiO3 lead-free antiferroelectric ceramics. Ceram. Int. 2017, 43, 7653–7659. [Google Scholar] [CrossRef]
  54. Gao, F.; Dong, X.; Mao, C.; Cao, F.; Wang, G. c/a Ratio-Dependent Energy-Storage Density in (0.9−x)Bi0.5Na0.5TiO3–xBaTiO3–0.1K0.5Na0.5NbO3 Ceramics. J. Am. Ceram. Soc. 2011, 94, 4162–4164. [Google Scholar] [CrossRef]
  55. Lu, X.; Xu, J.; Yang, L.; Zhou, C.; Zhao, Y.; Yuan, C.; Li, Q.; Chen, G.; Wang, H. Energy storage properties of (Bi0.5Na0.5)0.93Ba0.07TiO3 lead-free ceramics modified by La and Zr co-doping. J. Mater. 2016, 42, 87–93. [Google Scholar] [CrossRef]
  56. Manan, A.; Khan, S.; Ullah, A.A.; Ahmad, A.S.; Iqbal, Y.; Qazi, I.; Ullah, A.; Yao, Z.; Liu, H.; Hao, H.; et al. Improved energy storage characteristic of Yb doped 0.98(0.94Bi0.5Na0.5TiO3-0.06BaTiO3)-0.02BiAlO3 ceramics. Mater. Res. Bull. 2021, 137, 111175. [Google Scholar] [CrossRef]
Figure 1. PXRD patterns of (NBT-BT-NBN1−xTx) ceramic compositions: (a) x = 0.00; (b) x = 0.05; (c) x = 0.10; (d) x = 0.15; (e) x = 0.20.
Figure 1. PXRD patterns of (NBT-BT-NBN1−xTx) ceramic compositions: (a) x = 0.00; (b) x = 0.05; (c) x = 0.10; (d) x = 0.15; (e) x = 0.20.
Catalysts 13 00779 g001
Figure 2. Scanning electron micrographs of (NBT-BT-NBN1−xTx) ceramic compositions: (a) x = 0.00; (b) x = 0.05; (c) x = 0.10; (d) x = 0.15. (e) EDS spectrum of the bismuth pyrochlore Bi2Ti2O7 that developed as a secondary phase, indicated by arrows in the microstructures.
Figure 2. Scanning electron micrographs of (NBT-BT-NBN1−xTx) ceramic compositions: (a) x = 0.00; (b) x = 0.05; (c) x = 0.10; (d) x = 0.15. (e) EDS spectrum of the bismuth pyrochlore Bi2Ti2O7 that developed as a secondary phase, indicated by arrows in the microstructures.
Catalysts 13 00779 g002
Figure 3. Changes in εr for (NBT-BT-NBN1−xTx) ceramic compositions: (a) x = 0.00, (b) x = 0.05, (c) x = 0.10, (d) x = 0.15, and (e) x = 0.20, as a function of the temperature at different frequencies. (f) Variations in εr for (NBT-BT-NBN1−xTx) ceramic compositions (x = 0.00, 0.05, 0.10, 0.15, 0.2) vs. temperature at 1 MHz.
Figure 3. Changes in εr for (NBT-BT-NBN1−xTx) ceramic compositions: (a) x = 0.00, (b) x = 0.05, (c) x = 0.10, (d) x = 0.15, and (e) x = 0.20, as a function of the temperature at different frequencies. (f) Variations in εr for (NBT-BT-NBN1−xTx) ceramic compositions (x = 0.00, 0.05, 0.10, 0.15, 0.2) vs. temperature at 1 MHz.
Catalysts 13 00779 g003
Figure 4. Variations in tan δ for (NBT-BT-NBN1−xTx) ceramic compositions: (a) x = 0.00, (b) x = 0.05, (c) x = 0.10, (d) x = 0.15, and (e) x = 0.2, as a function of the temperature with increasing frequencies.
Figure 4. Variations in tan δ for (NBT-BT-NBN1−xTx) ceramic compositions: (a) x = 0.00, (b) x = 0.05, (c) x = 0.10, (d) x = 0.15, and (e) x = 0.2, as a function of the temperature with increasing frequencies.
Catalysts 13 00779 g004
Figure 5. (a) Variations of polarization for a series of (NBT-BT-NBN1−xTx) (x = 0.00, 0.05, 0.10, 0.15, 0.20) ceramics composition as a function of the electric field. (b) Variations in Pmax and Pmax-Pr for (NBT-BT-NBN1−xTx) with an increasing x content. (c) Ws and Wrec with an increasing x content. (d) Changes in Wloss and η for (NBT-BT-NBN1−xTx) with an increasing x content.
Figure 5. (a) Variations of polarization for a series of (NBT-BT-NBN1−xTx) (x = 0.00, 0.05, 0.10, 0.15, 0.20) ceramics composition as a function of the electric field. (b) Variations in Pmax and Pmax-Pr for (NBT-BT-NBN1−xTx) with an increasing x content. (c) Ws and Wrec with an increasing x content. (d) Changes in Wloss and η for (NBT-BT-NBN1−xTx) with an increasing x content.
Catalysts 13 00779 g005
Table 1. The comparative analysis of the current study with previously published works on NBT-based ceramics.
Table 1. The comparative analysis of the current study with previously published works on NBT-based ceramics.
MaterialsDielectric Breakdown
Field (kV/cm)
Wrec (J/cm3)Ref.
0.95[0.94BNT-0.06BT]-0.05LT1071.65[49]
0.95[BNT-0.06BT]-0.05AN720.71[50]
0.99(0.7NBT-0.3Bi0.2Sr0.7TiO3)-0.01NN881.04[51]
0.94(0.75BNT-0.25BKT)-0.06BA1061.15[52]
0.89BNT-0.06BT-0.05KNN550.60[53]
0.837BNT-0.063BT-0.1KNN520.45[54]
0.94[0.93NBT-0.07BT]-0.04LZ991.19[55]
Bi0.4806−xYb0.05Na0.4606Ba0.0588Ti0.98Al0.02O3960.70[56]
0.80(0.92Bi0.5Na0.5TiO3-0.08BaTiO3)-0.20 Na0.73Bi0.09Nb0.95Ta0.05O31751.57this work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khan, A.; Khan, T.M.; Wu, J.; Bilal, H.; Khan, S.Z.; Manan, A.; Wang, X.; Gul, N.S. Fabrication of Lead-Free Bismuth Based Electroceramic Compositions for High-Energy Storage Density Applications in Electroceramic Capacitors. Catalysts 2023, 13, 779. https://doi.org/10.3390/catal13040779

AMA Style

Khan A, Khan TM, Wu J, Bilal H, Khan SZ, Manan A, Wang X, Gul NS. Fabrication of Lead-Free Bismuth Based Electroceramic Compositions for High-Energy Storage Density Applications in Electroceramic Capacitors. Catalysts. 2023; 13(4):779. https://doi.org/10.3390/catal13040779

Chicago/Turabian Style

Khan, Azam, Taj Malook Khan, Jianbo Wu, Hazrat Bilal, Shahan Zeb Khan, Abdul Manan, Xiujian Wang, and Noor Shad Gul. 2023. "Fabrication of Lead-Free Bismuth Based Electroceramic Compositions for High-Energy Storage Density Applications in Electroceramic Capacitors" Catalysts 13, no. 4: 779. https://doi.org/10.3390/catal13040779

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