Next Article in Journal
The Methacrylate Adhesive to Double-Lap Shear Joints Made of High-Strength Steel—Experimental Study
Previous Article in Journal
A Reversible Protonic Ceramic Cell with Symmetrically Designed Pr2NiO4+δ-Based Electrodes: Fabrication and Electrochemical Features
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analogous Anti-Ferroelectricity in Y2O3-Coated (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 Ceramics and Their Energy-Storage Performance

1
School of Materials Science and Engineering, Jiangsu Collaborative Innovation Center of Photovolatic Science and Engineering, Jiangsu Province Cultivation Base for State Key Laboratory of Photovoltaic Science and Technology, National Experimental Demonstration Center for Materials Science and Engineering, Changzhou University, Changzhou 213164, China
2
Key Laboratory of Optoelectronic Material and Device, Department of Physics, Mathematics & Science College, Shanghai Normal University, Shanghai 200234, China
3
School of Material Science and Engineering, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
Materials 2019, 12(1), 119; https://doi.org/10.3390/ma12010119
Submission received: 3 December 2018 / Revised: 16 December 2018 / Accepted: 19 December 2018 / Published: 1 January 2019
(This article belongs to the Section Energy Materials)

Abstract

:
Antiferroelectric analogous (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 (PSLZSnT) ceramics were prepared by the solid-state sintering method by introducing a Y2O3-coating via the self-combustion method. The synthesized Y2O3-doped PSLZSnT ceramics present pseudo-cubic structure and rather uniform microstructural morphology accompanied by relatively small grain size. Excellent energy-storage performance is obtained in the Y2O3-doped PSLZSnT ceramics, in which the value of the energy-storage density presents a linearly increasing trend within the electric field measurement range. Such excellent performance is considered as relating to the rather pure perovskite structure, high relative density accompanied by relatively small grain size, and the antiferroelectric-like polarization-electric field behavior.

1. Introduction

With the development of miniature and compact electronic and electrical devices, and concerning sustainable development of society, designing for high power and energy-storage density dielectric materials becomes an urgent requirement [1,2]. Existing dielectric ceramics provide potentiality for high-energy-storage applications due to the fast charge/discharge capability, mature ceramic technique, and diversified composition selection, in which high dielectric constant, high dielectric breakdown strength, large polarization, and low hysteresis loss are desirable [2,3,4].
Antiferroelectric materials have gained continuous attention due to their double ferroelectric hysteresis loop characteristics and low dielectric loss, which tend to exhibit superior energy-storage density as compared with their ferroelectric counterparts [5,6,7]. Among which, PbZrO3 (PZ) was the first identified antiferroelectric compound, whereas the double polarization-electric field (P-E) hysteresis loop could not be driven at room-temperature in the polycrystalline ceramic state PZ since the critical electric field value (EF) induced the antiferroelectric to ferroelectric phase transition, exceeding the dielectric breakdown strength (Eb), therefore, PZ presented no promising applications in engineering devices [5,7].
It is well known that the Eb value depends on interfacial polarization and grain boundary density, which correlate with point charge defects and grain size, respectively [2,7,8,9]. Chemical doping provides an effective way to tailor physical performance, in which Sn and Y doping are reported as efficient in reducing the EF value, relating to the incommensurate modulation of the lattice cell [10,11,12]. Y doping also greatly decreases grain size of the (Pb0.87Ba0.1La0.02)(Zr0.65Sn0.3Ti0.05)O3 (PBLZST) ceramics as compared with the undoped PBLZST ones, which can increase the Eb value and energy-storage density [12]. Appropriate addition of additional PbO can compensate the evaporation of PbO during sintering, inhibit the formation of the pyrochlore phase, and stabilize the double ferroelectric hysteresis loop of antiferroelectrics [13].
Based on the above considerations, the (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 (PSLZSnT) ceramics were prepared by the traditional ceramic processing by introducing Y2O3-coating via the self-combustion method [14]. Forming a solid-solution by adding SrTiO3 and PbTiO3 stabilizes the perovskite structure, modifies the EF value, and then improves the antiferroelectric P-E hysteresis loop [2,5,7]. Introducing the Y2O3-coating via the self-combustion method was to thoroughly utilize its effect on inhibiting grain growth [12] since materials preparation processing presents essential influences on the crystal structure and materials’ performance [4,8,13,15,16]. The effects of the self-combustion technique and Y2O3 doping on the phase structure and electrical properties of the PSLZSnT ceramics were reported, and the electrical energy-storage performance was discussed.

2. Experimental Procedure

The Y2O3-doped (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 (PSLZSnT) ceramics were prepared by the solid-state sintering method by introducing Y2O3-coating via the self-combustion method. Stoichiometric oxides and carbonate of PbO, SrCO3, La2O3, ZrO2, SnO2, and TiO2 were weighed and well mixed according to the formula, in which an additional 5 mol% PbO was added to stabilize the antiferroelectric structure [13]. The mixture was calcined at 900 °C for 3 h and rather pure perovskite structure was obtained. Y2O3-coated PSLZSnT powder was synthesized via a self-combustion method, using Y(NO3)3⋅6H2O, citric acid, and C19H42BrN (CTAB, hexadecyl trimethyl ammonium bromide or cetyltrimethylammonium bromide), and the calcined PSLZSnT precursor powder as raw materials, and then self-combusted at 300 °C for 2 h as reported elsewhere [14]. The self-combusted Y2O3-coated PSLZSnT powder was cold-pressed into pellets after grinding and granulation, and sintered at 1275 °C for 2 h to prepare ceramics.
Crystal structure and microstructural morphology were characterized by XRD and SEM equipment (Rigaku D/max-2500/PC and Hitachi S-4800) (Akishima-shi, Tokyo, Japan and Chiyoda-ku, Tokyo, Japan), respectively. The fired-silver electrode was manually coated on both large surfaces of the sintered ceramics after polishing for physical performance measurements. The energy-storage properties were calculated using the P-E hysteresis loops measured by Radiant Precision Premier LC (Albuquerque, New Mexico, USA) [4,16].

3. Results and Discussion

The energy-storage density is an important performance parameter for high-energy-storage dielectric capacitive applications [2,5,7]. Figure 1 shows the room-temperature P-E hysteresis loop of the 0.25 mol% Y2O3-doped PSLZSnT ceramics, based on which the energy-storage density UC is calculated by the equation U C = E d D via integrating the effective shaded area accompanied by the energy-storage efficiency η calculated by the equation η = U C U C + U l o s s [1,2,7]. The waist shrunk P-E loop can be observed clearly although no typical antiferroelectric double P-E loop is driven, i.e., the nonzero remnant polarization Pr value, which exhibits antiferroelectric analogous characteristics and presents a great difference between the Pr and saturation polarization Ps values. Such a big difference indicates that the electric displacement D will not saturate quickly with an increase of the external electric field, which is especially favorable to excite large energy-storage density and high energy-storage efficiency [1,2,7]. In this work, large energy-storage density and high energy-storage efficiency can be excited at only 20 kV/cm, being 176.6 mJ/cm3 and 66.09%, respectively, providing a possibility for prospective applications.
For energy-storage applications, a high electric field is desirable, therefore, Figure 2 shows the P-E hysteresis loops of the 0.25 mol% Y2O3-doped PSLZSnT ceramics measured at different electric fields. With elevation of the external electric field, slim and antiferroelectric analogous hysteresis loops are maintained, whereas, the polarization and the difference between Pr and Ps increases greatly. Furthermore, the P-E hysteresis loop is far from saturation under 20 kV/cm, although a larger electric field was not undertaken due to the limitation of the equipment and the consideration of dielectric breakdown. Such phenomena are particularly beneficial for obtaining high energy-storage performance dielectrics [1,2,7]. Derived from Figure 2a, the P-E loops in one quadrant is presented in Figure 2b. Based on the magnified P-E loops, the parameters correlating with the energy-storage applications can be calculated.
Figure 3a shows energy-storage density and energy-storage efficiency of the 0.25 mol% Y2O3-doped PSLZSnT ceramics under different electric fields. The energy-storage density presents a linearly increasing trend within the electric field measurement range, affording promising high-performance energy-storage applications. It is noteworthy that the energy-storage density tends to present an exponentially increasing trend under a high electric field, as reported in our previous work, in which the exp ( E / 5.9193 ) relationship is excited above 40 kV/cm in the 0.075Pb(Yb1/2Nb1/2)O3-0.925Pb(Zr0.36Ti0.64)O3 ceramics [17] prepared by the conventional ceramic processing via the B-site oxide mixing route [18], indicating larger energy-storage density possibility. Such an increasing trend is not a rare phenomenon, such as the Pb0.94La0.04[(Zr0.7Sn0.3)0.9Ti0.1]O3 ceramics prepared by the solid-state reaction method reported by Xu et al., where apparent exponential elevation of the energy-storage density is observed above 60 kV/cm [19]. The energy-storage efficiency of the 0.25 mol% Y2O3-doped PSLZSnT ceramics decreases slightly at 8 kV/cm and increases greatly with a further increasing electric field, and the η value is larger than 60%, which can decrease energy dissipation during energy-storage cycling.
Figure 3b shows energy-storage performance of some lead-based and lead-free ferroelectrics and antiferroelectrics [17,19,20,21,22,23,24,25,26]. Considering the electric field used in this work, the energy-storage density of the 0.25 mol% Y2O3-doped PSLZSnT ceramics is far larger than these materials, and the energy-storage efficiency exceeds most of these materials. Furthermore, the waist shrunk P-E loops are often reported, such as in NBT-ST [21], BNT-BKT-BA [25], and PLZST [19], whether or not nominated as antiferroelectrics, the narrowed P-E loops are observed, and no apparent double ferroelectric loops are excited. The antiferroelectric-like behavior becomes evident with increasing temperature in NBT-ST, which can be attributed to the nominal ferroelectric to antiferroelectric phase transition [21]. The linear increase of energy-storage density with an increasing electric field under low electric field in NBT-ST [21] and BNT-BKT-BA [25], and the exponential increase of energy-storage density with increasing electric field under high electric field in PLZST [19], corroborate our anticipation that the Y2O3-doped PSLZSnT ceramics could present high-performance energy-storage properties.
The high energy-storage performance obtained in this work can be attributed to crystal and morphology factors as shown in Figure 4. A rather pure perovskite structure is obtained, i.e., without detectable impurity within the detection limitation of the XRD measurement. Overall, the diffraction reflections are rather symmetric and present a single peak besides the {200} and {310} diffraction reflections, which present slight splitting. The distortion from the cubic symmetry is apparent, and the Y2O3-doped PSLZSnT ceramics may exist in a rhombohedral, tetragonal, or orthorhombic system structure. However, due to the sensitivity limitation of the conventional XRD measurement, accurate crystal refinement cannot be obtained, therefore, the pseudo-cubic perovskite structure is used for calculating the crystal cell parameters. The maximum dielectric constant value of 558 at 207.3 °C and 1 kHz and the rather broad dielectric peaks without apparent dielectric frequency dispersion coincide with the antiferroelectric-like dielectric behavior [2,5,7]. The grains distribute rather homogenously without apparent gas pores, considering the free-surface-sample used, and the average grain size is determined as 2.09 μm, made statistically by SmileView software, smaller than normal Pb-based ferroelectrics [4].
Although the 0.25 mol% Y2O3-doped PSLZSnT ceramics present excellent energy-storage performance under a low electric field, an exceptionally large external electric field is required for practical dielectric capacitor applications [2,5,7]. Polymer materials tend to present a high Eb value, allowing large energy-storage density to be obtained, such as in the biaxially-oriented polypropylene reported by Chu et al., in which ~4 J/cm3 energy-storage density was acquired under 6000 kV/cm [1]. It could be predicted that such a value can increase to ~53 J/cm3 of the Y2O3-doped PSLZSnT ceramics even in the linearly increasing trend with an increase of the external electric field under a large Eb value. Furthermore, non-early saturation of electric displacement and low dielectric loss are additional necessary parameters for high energy-storage applications disclosed by Chu et al. [1], which have been the essential properties of the Y2O3-doped PSLZSnT ceramics.
Then, the major task for the Y2O3-doped PSLZSnT ceramics is to increase the Eb value to meet the requirements of energy-storage applications, which correlates with the ceramics’ density, grain size, and defect chemistry [2,5,6,7]. Since polymer materials tend to show large Eb values, polymer-based dielectrics, especially in inhomogenous structures, provide an efficient strategy to increase energy-storage performance, i.e., forming polymer matrix composites which can combine the advantages of a high dielectric constant of ferroelectric ceramics and large Eb values of polymers [2,5,6,7]. Among which, polyvinylidene fluoride (PVDF)-based polymers attract great research attention due to their large dielectric constant in the ferroelectric state. Furthermore, chemical functionalization of ferroelectric ceramic precursor powders with improving distribution and overcoming interfacial polarization are key problems in forming polymer-based composites [2,5,7].
Although the Y2O3-doped PSLZSnT ceramics present a high relative density of ~95% ( relative   density = bulk   density theoretical   density × 100 % , in which the bulk density of the ceramics was measured by the Archimedes’ water immersion method, and the theoretical density of the ceramics was calculated based on the crystal cell volume obtained by the XRD result combined with the molecular weight) with a relatively small mean grain size of 2.09 μm, as compared with the normally synthesized Pb-based ceramics via the solid-phase method [4], further increase of the density with decreasing grain size is required to enhance energy-storage performance. Such requirements encounter dilemmas since relative density and grain size change contrarily if just tailoring sintering temperature, i.e., normally, relative density increases with elevating sintering temperature within an appropriate sintering temperature range, and inevitably leads to the growth of grain size [4].
Oxide coating introduced via the self-combustion method used in this work offers the possibility of tailoring these parameters simultaneously. Great progress is obtained in synthesizing 0.55Pb(Ni1/3Nb2/3)O3-0.05PbHfO3-0.4PbTiO3/Ni0.875Zn0.125Fe2O4 (PNNHT/NZF) particulate composite ceramics in our previous work [14]. Densified PNNHT/NZF ceramics are prepared sintered at only 900 °C by simultaneously introducing nano-sized WO3 and CuO coatings via the self-combustion method, which guarantees a small grain size due to the low sintering temperature [14]. An additional promising technique is adding glass-like sintering aids, which can increase the relative density of ceramics sintered at low sintering temperatures, beneficial for obtaining a small grain size. Glass-based inhomogenous dielectrics provide another alternative choice, in which the barium boroaluminosilicate glass (SchottAF-45) is proved to exhibit the highest Eb value of 12 MV/cm in a bulk glass state as reported by Smith et al. [27], whose dielectric breakdown strength is maintained as 1300 kV/cm and charge-discharge efficiency is increased to 92.5%, as reported by Xue, when mixed with 40 mol% (4BaO-Na2O-5Nb2O5) to form niobate glass-ceramics and diminished interfacial polarization [28].
Defect chemistry is also a problem that should be considered, influencing interfacial polarization [2,5,7]. In this work, Y2O3 doped into PSLZSnT will produce donor point defects although the charge balance is taken into account in chemical formulation design, then Pb vacancies are engendered to compensate charge balance, i.e., Y 2 O 3 P S L Z S n T 2 Y P b + V P b + P b + 3 O O × considering cationic radii, expressed by the Kröger–Vink symbol system [29,30,31]. Such point defects are normally thermal-activated hopping-jumping charge carriers [31], tending to decrease resistivity of the sintered ceramics, especially at elevated temperatures. Resistivity of the 0.25 mol% Y2O3-doped PSLZSnT ceramics is 1.489 × 109 Ω⋅cm in this work, as shown in Figure 1, which should be increased further to obtain a large Eb value required for high energy-storage density, in which heterovalent ion doping with self-charge-compensation combinations provides a promising strategy [32,33].
Based on the above discussions, formation of polymer- or glass-based composites can increase the Eb value; introducing nano-sized sintering aids can increase relative density meanwhile maintaining a small grain size due to low sintering temperatures; and self-charge-compensation heterovalent ion doping can diminish point defects, which provide effective methods to obtain high-performance energy-storage properties.

4. Conclusions

Y2O3-doped (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 (PSLZSnT) ceramics were prepared by the conventional ceramic processing by introducing Y2O3-coating via the self-combustion method. The Y2O3-doped PSLZSnT ceramics exhibit a rather pure pseudo-cubic perovskite structure. Their dielectric-temperature curves are slightly broad without obvious dielectric frequency dispersion, accompanied by several hundreds of relative dielectric constants and a small dielectric loss tangent. The sintered ceramics present rather homogenous microstructural morphology with a relatively small mean grain size of 2.09 μm. The Y2O3-doped PSLZSnT ceramics present antiferroelectric analogous behavior with waist shrunk P-E loops, where the Pr and Ps values exhibit a great difference. Such characteristics are favorable for energy-storage properties, therefore, large energy-storage density and high energy-storage efficiency were obtained, being 176.6 mJ/cm3 and 66.09%, respectively, and excited at only 20 kV/cm, which exceeds many recently reported lead-based and lead-free ferroelectric and antiferroelectric ceramics. To improve the energy-storage properties further, forming polymer- or glass-based composites, introducing nano-sized sintering aids, and self-charge-compensation heterovalent ion doping provide effective strategies.

Author Contributions

B.F. and J.D. conceived and designed the experiments; S.Z. and D.W. performed the experiments and collected the measurement data; Y.J. and S.Z. analyzed the data and plotted the figures; B.F. and X.Z. polished the figures, wrote and polished the paper.

Funding

This work was supported by the National Natural Science Foundation of China (No. 51577015), the Top-notch Academic Programs Project of Jiangsu Higher Education Institutions and the Priority Academic Program Development of Jiangsu Higher Education Institutions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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] [PubMed]
  2. 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] [PubMed]
  3. Dou, R.; Lu, X.; Yang, L.; Wang, H.; Zhou, C.; Xu, J. Influence of A-site Sm doping on structural and electrical property of 0.93Na0.5Bi0.5TiO3-0.07BaTiO3 lead free ceramics. J. Inorg. Mater. 2018, 33, 528–534. (In Chinese) [Google Scholar]
  4. Zhu, R.; Yin, Y.; Fang, B.; Chen, Z.; Zhang, S.; Ding, J.; Zhao, X.; Luo, H. Optimizing structure and electrical properties of high-Curie temperature PMN-PHT piezoelectric ceramics via tailoring sintering process. Eur. Phys. J. Appl. Phys. 2016, 74, 30101. [Google Scholar] [CrossRef]
  5. Tan, X.; Ma, C.; Frederick, J.; Beckman, S.; Webber, K.G. The antiferroelectric↔ferroelectric phase transition in lead-containing and lead-free perovskite ceramics. J. Am. Ceram. Soc. 2011, 94, 4091–4107. [Google Scholar] [CrossRef]
  6. Zhao, L.; Liu, Q.; Gao, J.; Zhang, S.; Li, J.-F. Lead-free antiferroelectric silver niobate tantalate with high energy storage performance. Adv. Mater. 2017, 29, 1701824. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, Z.; Lu, T.; Ye, J.; Wang, G.; Dong, X.; Withers, R.; Liu, Y. Antiferroelectrics for energy storage applications: A review. Adv. Mater. Technol. 2018. [Google Scholar] [CrossRef]
  8. 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]
  9. Zhou, M.; Liang, R.; Zhou, Z.; Xu, C.; Nie, X.; Chen, X.; Dong, X. High energy storage properties of (Ni1/3Nb2/3)4+ complex-ion modified (Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 ceramics. Mater. Res. Bull. 2018, 98, 166–172. [Google Scholar] [CrossRef]
  10. Chang, Y.; Lian, J.; Wang, Y. One-dimensional regular arrays of antiphase domain boundaries in antiferroelectric tin-substituted lead zirconate titanate ceramics. Appl. Phys. A 1985, 36, 221–227. [Google Scholar] [CrossRef]
  11. Frederick, J.; Tan, X.; Jo, W. Strains and polarization during antiferroelectric-ferroelectric phase switching in Pb0.99Nb0.02[(Zr0.57Sn0.43)1-yTiy]0.98O3 Ceramics. J. Am. Ceram. Soc. 2011, 94, 1149–1155. [Google Scholar] [CrossRef]
  12. Zhang, L.; Jiang, S.; Zeng, Y.; Fu, M.; Han, K.; Li, Q.; Wang, Q.; Zhang, G. Y doping and grain size co-effects on the electrical energy storage performance of (Pb0.87Ba0.1La0.02)(Zr0.65Sn0.3Ti0.05)O3 anti-ferroelectric ceramics. Ceram. Int. 2014, 40, 5455–5460. [Google Scholar] [CrossRef]
  13. Zhang, G.; Liu, S.; Yu, Y.; Zeng, Y.; Zhang, Y.; Hu, X.; Jiang, S. Microstructure and electrical properties of (Pb0.87Ba0.1La0.02)(Zr0.68Sn0.24Ti0.08)O3 anti-ferroelectric ceramics fabricated by the hot-press sintering method. J. Eur. Ceram. Soc. 2013, 33, 113–121. [Google Scholar] [CrossRef]
  14. Zhao, X.; Wu, D.; Fang, B.; Wang, F.; Tang, Y.; Shi, W.; Ding, J. Enhancing magnetoelectric properties of 0–3 particulate composite ceramics by introducing nano-sized sintering aids via self-combustion method. Curr. Appl. Phys. 2017, 17, 1249–1253. [Google Scholar] [CrossRef]
  15. Huang, C.; Xu, J.; Fang, Z.; Ai, D.; Zhou, W.; Zhao, L.; Sun, J.; Wang, Q. Effect of preparation process on properties of PLZT (9/65/35) transparent ceramics. J. Alloy. Compd. 2017, 723, 602–610. [Google Scholar] [CrossRef]
  16. Zhu, R.; Zhang, Q.; Fang, B.; Wu, D.; Zhao, X.; Ding, J. Correlation of enhanced electrical properties and domain structure of high-TC PMN-PH-PT ceramics prepared by different methods. Ceram. Int. 2018, 44, 10099–10105. [Google Scholar] [CrossRef]
  17. Zheng, J.; Chen, Z.; Fang, B.; Ding, J.; Zhao, X.; Xu, H.; Luo, H. Synthesis and characterization of Pb(Yb1/2Nb1/2)O3-based high-Curie temperature piezoelectric ceramics. Eur. Phys. J. Appl. Phys. 2014, 66, 30101. [Google Scholar] [CrossRef]
  18. Liu, C.; Pan, S.; Chen, Z.; Fang, B.; Ding, J.; Zhao, X.; Luo, H. Phase diagram of ternary Pb(Mg1/3Nb2/3)O3-PbZrO3-PbTiO3 ferroelectric ceramics prepared via a B-site oxide mixing route. Ferroelectrics 2015, 482, 11–21. [Google Scholar] [CrossRef]
  19. Xu, R.; Li, B.; Tian, J.; Xu, Z.; Feng, Y.; Wei, X.; Huang, D.; Yang, L. Pb0.94La0.04[(Zr0.7Sn0.3)0.9Ti0.1]O3 antiferroelectric bulk ceramics for pulsed capacitors with high energy and power density. Appl. Phys. Lett. 2017, 110, 142904. [Google Scholar] [CrossRef]
  20. Srikanth, K.S.; Vaish, R. Enhanced electrocaloric, pyroelectric and energy storage performance of BaCexTi1-xO3 ceramics. J. Eur. Ceram. Soc. 2017, 37, 3927–3933. [Google Scholar] [CrossRef]
  21. Cao, W.P.; Lia, 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]
  22. Kumar, R.; Singh, S. Giant electrocaloric and energy storage performance of [(K0.5Na0.5)NbO3]1-x-[LiSbO3]x nanocrystalline ceramics. Sci. Rep. 2018, 8, 3186. [Google Scholar] [CrossRef] [PubMed]
  23. 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 2017, 5, 9552–9558. [Google Scholar] [CrossRef]
  24. 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]
  25. 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]
  26. Zhang, T.-F.; Huang, X.-X.; Tang, X.-G.; Jiang, Y.-P.; Liu, Q.-X.; Lu, B.; Lu, S.-G. Enhanced electrocaloric analysis and energy-storage performance of lanthanum modified lead titanate ceramics for potential solid-state refrigeration applications. Sci. Rep. 2018, 8, 396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Smith, N.J.; Rangarajan, B.; Lanagan, M.T.; Pantano, C.G. Alkali-free glass as a high energy density dielectric material. Mater. Lett. 2009, 63, 245–1248. [Google Scholar] [CrossRef]
  28. Xue, S.; Liu, S.; Zhang, W.; Wang, J.; Tang, L.; Shen, B.; Zhai, J. Dielectric properties and charge-discharge behaviors in niobate glass ceramics for energy-storage applications. J. Alloy. Compd. 2014, 617, 418–422. [Google Scholar] [CrossRef]
  29. Viehland, D.; Jang, S.J.; Cross, L.E.; Wuttig, M. Freezing of the polarization fluctuations in lead magnesium niobate relaxors. J. Appl. Phys. 1990, 68, 2916–2921. [Google Scholar] [CrossRef]
  30. Fang, B.; Zhu, M.; Ding, J.; Shan, Y.; Imoto, H. Improving ferroelectric and piezoelectric properties of PbFe1/4Sc1/4Nb1/2O3 ceramics by oxide doping prepared via a B-site oxide mixing route. Ceram. Int. 2013, 39, 1677–1681. [Google Scholar] [CrossRef]
  31. Ji, W.; Feng, S.; Fang, B.; Zhao, X.; Zhang, S.; Ding, J.; Luo, H. Facile preparation and performance of novel high-TC xBi(Ni1/2Ti1/2)O3-(1-x)Pb(Zr1/2Ti1/2)O3 piezoceramics. Curr. Appl. Phys. 2018, 18, 289–296. [Google Scholar] [CrossRef]
  32. Wu, D.; Fang, B.; Du, Q.; Ding, J. Preparation and properties of La and K co-doped BaTiO3 lead-free piezoelectric ceramics. Ferroelectrics 2012, 432, 81–91. [Google Scholar] [CrossRef]
  33. Maurya; Bahadur, A.; Dwivedi, A.; Choudhary, A.K.; Yadav, T.P.; Vishwakarma, P.K.; Rai, S.B. Optical properties of Er3+, Yb3+ co-doped calcium zirconate phosphor and temperature sensing efficiency: Effect of alkali ions (Li+, Na+ and K+). J. Phys. Chem. Solids 2018, 119, 228–237. [Google Scholar] [CrossRef]
Figure 1. Polarization-electric field (P-E) hysteresis loop of the 0.25 mol% Y2O3-doped PSLZSnT ceramics. Inset shows resistivity and the shaded area with oblique lines shows the energy-storage density.
Figure 1. Polarization-electric field (P-E) hysteresis loop of the 0.25 mol% Y2O3-doped PSLZSnT ceramics. Inset shows resistivity and the shaded area with oblique lines shows the energy-storage density.
Materials 12 00119 g001
Figure 2. P-E hysteresis loops of the 0.25 mol% Y2O3-doped PSLZSnT ceramics measured at different electric fields. (a) bipolar; (b) unipolar.
Figure 2. P-E hysteresis loops of the 0.25 mol% Y2O3-doped PSLZSnT ceramics measured at different electric fields. (a) bipolar; (b) unipolar.
Materials 12 00119 g002
Figure 3. (a) Variation of energy-storage density and energy-storage efficiency as a function of electric field of the 0.25 mol% Y2O3-doped PSLZSnT ceramics; (b) comparison of energy-storage performance of some recently reported ferroelectric and antiferroelectric ceramics, in which BCeT is BaCe0.15Ti0.85O3 [20]; NBT-ST is 0.7(Na0.5Bi0.5)TiO3-0.3SrTiO3 [21]; KNaN-LS is 0.97(K0.5Na0.5)NbO3-0.03LiSbO3 [22]; BT-BiZnZr is 0.9BaTiO3-0.1Bi(Zn0.5Zr0.5)O3 [23]; BiNBT-KN is 0.94Bi0.47Na0.47Ba0.06TiO3-0.06KNbO3 [24]; BNT-BKT-BA is 0.94(0.75Bi0.5Na0.5TiO3-0.25Bi0.5K0.5TiO3)-0.06BiAlO3 [25]; PYbN-PZT is 0.075Pb(Yb1/2Nb1/2)O3-0.925Pb(Zr0.36Ti0.64)O3 [17]; PLT is (Pb1-xLax)Ti1-x/4O3 (x = 0.28) [26]; and PLZST is Pb0.94La0.04[(Zr0.7Sn0.3)0.9Ti0.1]O3 [19].
Figure 3. (a) Variation of energy-storage density and energy-storage efficiency as a function of electric field of the 0.25 mol% Y2O3-doped PSLZSnT ceramics; (b) comparison of energy-storage performance of some recently reported ferroelectric and antiferroelectric ceramics, in which BCeT is BaCe0.15Ti0.85O3 [20]; NBT-ST is 0.7(Na0.5Bi0.5)TiO3-0.3SrTiO3 [21]; KNaN-LS is 0.97(K0.5Na0.5)NbO3-0.03LiSbO3 [22]; BT-BiZnZr is 0.9BaTiO3-0.1Bi(Zn0.5Zr0.5)O3 [23]; BiNBT-KN is 0.94Bi0.47Na0.47Ba0.06TiO3-0.06KNbO3 [24]; BNT-BKT-BA is 0.94(0.75Bi0.5Na0.5TiO3-0.25Bi0.5K0.5TiO3)-0.06BiAlO3 [25]; PYbN-PZT is 0.075Pb(Yb1/2Nb1/2)O3-0.925Pb(Zr0.36Ti0.64)O3 [17]; PLT is (Pb1-xLax)Ti1-x/4O3 (x = 0.28) [26]; and PLZST is Pb0.94La0.04[(Zr0.7Sn0.3)0.9Ti0.1]O3 [19].
Materials 12 00119 g003
Figure 4. XRD pattern of the 0.25 mol% Y2O3-doped PSLZSnT ceramics. Inset shows temperature dependence of dielectric constant measured at several frequencies upon the cooling process and SEM image of the free surface after thermal etching at 850 °C for 30 min.
Figure 4. XRD pattern of the 0.25 mol% Y2O3-doped PSLZSnT ceramics. Inset shows temperature dependence of dielectric constant measured at several frequencies upon the cooling process and SEM image of the free surface after thermal etching at 850 °C for 30 min.
Materials 12 00119 g004

Share and Cite

MDPI and ACS Style

Fang, B.; Jiang, Y.; Zhao, X.; Zhang, S.; Wu, D.; Ding, J. Analogous Anti-Ferroelectricity in Y2O3-Coated (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 Ceramics and Their Energy-Storage Performance. Materials 2019, 12, 119. https://doi.org/10.3390/ma12010119

AMA Style

Fang B, Jiang Y, Zhao X, Zhang S, Wu D, Ding J. Analogous Anti-Ferroelectricity in Y2O3-Coated (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 Ceramics and Their Energy-Storage Performance. Materials. 2019; 12(1):119. https://doi.org/10.3390/ma12010119

Chicago/Turabian Style

Fang, Bijun, Yan Jiang, Xiangyong Zhao, Shuai Zhang, Dun Wu, and Jianning Ding. 2019. "Analogous Anti-Ferroelectricity in Y2O3-Coated (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 Ceramics and Their Energy-Storage Performance" Materials 12, no. 1: 119. https://doi.org/10.3390/ma12010119

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