New Antiferroelectric Perovskite System with Ultrahigh Energy

Jan 25, 2019 - Department of Chemistry and 4D LABS, Simon Fraser University , Burnaby , British Columbia V5A 1S6 , Canada. § Department of Applied ...
2 downloads 0 Views 13MB Size
Article pubs.acs.org/cm

Cite This: Chem. Mater. XXXX, XXX, XXX−XXX

New Antiferroelectric Perovskite System with Ultrahigh EnergyStorage Performance at Low Electric Field Pan Gao,†,‡ Zenghui Liu,† Nan Zhang,† Hua Wu,*,§,‡ Alexei A. Bokov,‡ Wei Ren,*,† and Zuo-Guang Ye*,‡,†

Chem. Mater. Downloaded from pubs.acs.org by EASTERN KENTUCKY UNIV on 01/25/19. For personal use only.



Electronic Materials Research Laboratory, Key Laboratory of the Ministry of Education & International Center for Dielectric Research, School of Electronic and Information Engineering, Xi’an Jiaotong University, Xi’an 710049, P. R. China ‡ Department of Chemistry and 4D LABS, Simon Fraser University, Burnaby, British Columbia V5A 1S6, Canada § Department of Applied Physics, Donghua University, Ren Min Road 2999, Songjiang District, Shanghai 201620, P. R. China ABSTRACT: The development of antiferroelectric (AFE) materials with high recoverable energy-storage density (Wrec) and energy-storage efficiency (η) is of great importance for meeting the requirements of miniaturization and integration for advanced pulse power capacitors. However, the drawbacks of traditional AFE materials, namely, high critical field (Ecr) and low Wrec, make them unsuitable to be utilized in practical applications. To increase Wrec and η, here we report an effective approach using the transient liquid-phase sintering and the softening of antiferroelectric order to decrease the porosity, enhance the dielectric breakdown strength (DBS), and increase the maximum electric-field-induced polarization (Pmax) of the AFE ceramics. On the basis of this concept, a novel solid solution of (1 − x)PbHfO3−xPb(Mg1/2W1/2)O3 [(1 − x)PHf−xPMW] was designed and prepared in the form of ceramics by the solid-state reaction method. Their crystal structures, phase transitions, dielectric properties, and energy-storage properties were investigated systemically. X-ray diffraction analysis indicates the formation of solid solution with a partial order on the B site at room temperature in a broad composition range. Dielectric measurements reveal that the AFE to ferroelectric (FE) phase-transition temperature shifts toward room temperature with the increasing Pb(Mg1/2W1/2)O3 (PMW) content. The optimal energy-storage performance is found for the 0.90PHf−0.10PMW ceramic with the highest Wrec of 3.7 J/cm3 (at a relatively low electric field of 155 kV/cm) and a favorable η of 72.5% among all of the studied compositions, which is much superior to that of the so far reported perovskite ceramics under the similar electric fields. This is the first reported PHf-based solid solution with ultrahigh energy-storage density. The enhanced energy-storage performance can be attributed to the improved DBS and enhanced Pmax (45 μC/cm2) due to the incorporation of PMW that leads to dense microstructure and softens the antiferroelectricity. The results show that the (1 − x)PHf−xPMW ceramics form a new family of promising AFE candidates with significantly enhanced DBS, Wrec, and η. This work also demonstrates the design methodology for developing not only the PbHfO3-based but also other new AFE−AFE solid solution material for high-energy-storage applications.

1. INTRODUCTION

In general, there are four kinds of dielectric materials for energy-storage applications: (a) linear dielectrics, (b) ferroelectrics (FEs), (c) relaxor ferroelectrics (RFEs), and (d) antiferroelectrics (AFEs). Linear dielectrics usually show a high dielectric breakdown strength (DBS) and low energy loss, but the small polarization limits their high-energy-storage applications. Ferroelectrics possess a large saturated polarization (Ps) and a moderate DBS, but their energy density and efficiency show low values due to their large remanent polarization (Pr). Therefore, relaxor ferroelectrics and antiferroelectrics are the main candidates that can be practically used as energy-storage materials because of their high

In recent years, high-performance energy-storage devices have received tremendous attention due to the increasing needs to protect environment and develop sustainable economy.1−3 Nowadays, dielectric ceramic capacitors, such as BaTiO3, Bi0.5Na0.5TiO3, and (Pb,La)(Zr,Ti)O3-based solid solutions, are well known for high energy density storage applications because of their high power density, ultrafast charge/discharge capability (95%) with a dense microstructure in the ceramics. The optimized value was achieved for x = 0.10 with the highest relative density of 96% among all of the compositions. By applying the linear interception method to data of these SEM images, the grain size was estimated to vary from about 1 μm in PHf to 3.3 μm in 0.83PHf−0.17PMW. The enhanced grain size possibly results from a transient liquid phase28 formed in the presence of some intermediate phases with low melting points (m.p.) such as Pb2WO5 (mp = 935 °C) in addition to the reagents such as PbO (m.p. = 888 °C) in the PHf−PMW F

DOI: 10.1021/acs.chemmater.8b04470 Chem. Mater. XXXX, XXX, XXX−XXX

Article

Chemistry of Materials

Figure 6. (a) SEM images of the (1 − x)PbHfO3−xPb(Mg1/2W1/2)O3 ceramics with x = 0.00−0.17. (b) Experimental procedure of ceramic quenching and the results of subsequent SEM imaging and energy-dispersive X-ray spectroscopy (EDS) analysis, which proves the presence of a transient liquid phase of Pb2WO5-type during sintering. (c) Schematic diagram of the ceramic densification during the sintering process in the presence of the transient liquid phase.

the coexistence of the AFE orthorhombic and FE orthorhombic phases in this composition, as found by XRD and dielectric results. The P−E loop for x = 0.17 shows a single hysteresis loop, indicating the presence of a typical FE state at room temperature with polar domain switching under an electric field. The decrease of antiferroelectricity and the appearance of the ferroelectric phase in the PHf−PMW solid solution formed between two antiferroelectric end members can be attributed to the dipole frustration, which is related to the lattice mismatch between PHf and PMW. The antiferroelectricity in PbHfO3 mainly arises from the antiparallel displacement of Pb2+ ions along the pseudocubic [110]pc direction (see Figure 3b), whereas the antiferroelectricity in Pb(Mg1/2W1/2)O3 mainly results from the antiparallel displacement of Pb2+ ions along both the [100]pc and [010]pc directions (see Figure 3c). These two competing types of ionic displacement create dipole frustration and possibly incommensurate modulation, which disrupt the antiferroelectric order in PHf and thereby induce the long-range ferroelectric order with the increase of PMW concentration, since the ferroelectric and antiferroelectric phases have relatively close state energy.22 More detailed studies are underway to further clarify the mechanisms underlying the induced ferroelectricity in this solid solution. Therefore, the 0.90PHf−0.10PMW ceramic exhibits the most interesting AFE behavior, which may lead to high performance for energy storage. Figure 8 shows the ferroelectric/antiferroelectric characters and the energy-storage parameters (Wst, Wrec, Wloss, and energy-storage efficiency η)

and decreased sintering temperature. The existence of the liquid phase was also observed in other solid solution systems that contain Pb- and W-based compounds, such as the PbZrO 3 −Pb(Mn 1/2 W 1/2 )O 3 22 and Pb(Mg 1/2 W 1/2 )O 3 − PbTiO329 systems. Such a transient liquid phase sintering and the compositiondriven dense microstructure mechanism is found to contribute to the high DBS and high energy-storage density, which will be shown and discussed in the following section. 4.4. Energy-Storage Performance. The polarization− electric field (P−E) relations of the (1 − x)PHf−xPMW ceramics measured at room temperature and 10 Hz are shown in Figure 7. It can be seen that the PMW concentration has an important influence on the features of P−E loops. A mainly linear behavior of P−E relation is found for the ceramics of 0.00 ≤ x ≤ 0.06, which is not favorable for energy storage because of the insignificant electric-field-induced polarization. In fact, the electric field required to induce the AFE to FE phase transition at room temperature is stronger than the breakdown strength of those ceramics. In contrast, welldeveloped double hysteresis loops are displayed for the compositions x = 0.10−0.13 at room temperature, indicating an electric-field-induced ferroelectric state. All of the aforementioned P−E features for 0.00 ≤ x ≤ 0.13 clearly show the characteristic AFE behavior at room temperature. A “triple” hysteresis loop is displayed for x = 0.15, indicating a mixed AFE and FE behavior with the FE phase becoming more stable, leading to a small remanent polarization Pr when the electric field returns to zero. This result could be explained by G

DOI: 10.1021/acs.chemmater.8b04470 Chem. Mater. XXXX, XXX, XXX−XXX

Article

Chemistry of Materials

probably related to the enhanced polarizability of the lattice and the reduced critical field. The detailed structural origin of this effect is still under investigation. It can be seen from Figure 8c,d that the composition dependences of the energy-storage parameters, namely, the stored energy density Wst, the recoverable energy density Wrec, and the energy-loss density Wloss, of the ceramics show a similar trend, i.e., they increase first and then decrease with further increase of the PMW content. Wst and Wrec reach the maximum values of 5.1 and 3.7 J/cm3, respectively, for x = 0.10, whereas Wloss reaches the maximum value of 2.1 J/cm3 at x = 0.13. The highest Wst and Wrec values of the 0.90PHf−0.10PMW ceramics are believed to result from a synergetic combination of the high Eb (155 kV/ cm), large induced Pmax (45 μC/cm2), and large polarization difference (Pmax − Pr). In practical applications, in addition to have high Wrec values, a high-energy-storage efficiency η is also desirable. In this work, η is calculated on the basis of the following formula13 η=

Wrec Wrec = × 100% W st Wrec + Wloss

(4)

Figure 8d shows that η decreases from 87.7 to 21.2% with the increase of the PMW content. Considering both the values of Wrec and η for the compositions varying from x = 0 to 0.17, it can be seen that a large Wrec (3.7 J/cm3) and a favorable η (72.5%) are simultaneously achieved for the 0.90PHf− 0.10PMW ceramics at 155 kV/cm, indicating that these PHfbased AFE ceramics are promising candidates for advanced pulsed power energy-storage applications. The dielectric breakdown strength is essential to AFE materials for high-energy-storage applications. Figure 9 shows the Weibull distribution of the dielectric breakdown strength for the (1 − x)PHf−xPMW ceramics with 0.10 ≤ x ≤ 0.17 at room temperature. This Weibull distribution is usually used for the DBS analysis and can be described by30

Figure 7. P−E hysteresis loops of the (1 − x)PbHfO3−xPb(Mg1/2W1/2)O3 ceramics displayed below their critical breakdown strength.

of the (1 − x)PHf−xPMW ceramics calculated from their P−E loops. It can be seen from Figure 8a that the critical field EF is significantly reduced with the increase of the PMW concentration, confirming an enhanced stabilization of the metastable ferroelectric phase as suggested by the dielectric measurements. In addition, the breakdown field Eb is also enhanced remarkably with the substitution of PMW for PHf and the maximum value of 155 kV/cm is achieved for x = 0.10. As shown in the SEM results, this ceramic also exhibits a homogeneous microstructure with the highest relative density. Thus, the control of the microstructure is proved to be an effective way to improve the dielectric breakdown strength. For the ceramics of compositions with x ≤ 0.06, the double P−E loops cannot be displayed due to breakdown, indicating that the AFE-to-FE switching field EF is higher than the breakdown strength. Double hysteresis loops are achieved for x = 0.10− 0.13, because the critical field is reduced below the breakdown field due to the substitution of an sufficient amount of PMW for PHf (Figure 8a), which allows the FE phase to be induced from the AFE state. Figure 8b shows that both the induced polarization Pind at 80 kV/cm and remanent polarization Pr are enhanced when the PMW content is increased up to x = 0.17. Nevertheless, the maximum polarization Pmax shows a nonmonotonic trend with composition. It increases sharply to a high value of 45 μC/cm2 at x = 0.10, followed by a slight decrease owing to the decreased Eb, and a subsequent gradual increase due to the presence of the FE phase as the PMW content further increases. The large achievable polarization is

Xi = ln(Ei)

(5)

Yi = ln(ln(1/(1 − Pi)))

(6)

and Pi = i/(n + 1)

(7)

where Xi and Yi are two parameters in the Weibull distribution function, Ei is the specific breakdown electric field of each specimen in the experiments, Pi is the probability for dielectric breakdown, n is the number of specimens of each sample, and i is the serial number of a specimen. The samples are arranged in ascending order of DBS values so that E1 ≤ E2 ≤ E3 ... ≤ Ei ... ≤ En . A total of 10 samples for each composition were measured to obtain the average breakdown strength of the ceramics. According to the Weibull distribution, there is a linear relationship between Xi and Yi; therefore, the average DBS values could be extracted from the points where the fitting lines intersect with the horizontal line at Yi = 0, and the range of DBS is related to the slope of the linear fitting of the experimental data. As shown in Figure 9, the average DBS values are found to be 151, 144, 96, and 77 kV/cm for x = 0.10, 0.13, 0.15, and 0.17, respectively. Therefore, the ceramics with the intermediate PMW concentration of x = 0.10 exhibit not only a high energy-storage density and a uniform microstructure but also a high breakdown strength. H

DOI: 10.1021/acs.chemmater.8b04470 Chem. Mater. XXXX, XXX, XXX−XXX

Article

Chemistry of Materials

Figure 8. (a) Variations of the breakdown field Eb and critical field EF of the (1 − x)PbHfO3−xPb(Mg1/2W1/2)O3 ceramics with PMW concentration; the green and brown areas define the AFE state and the electric-field-induced FE phase in terms of the composition dependence of the critical field EF. (b) Composition dependences of the maximum polarization (Pmax), the remnant polarization (Pr) before breakdown (before Eb), and the electric-field-induced polarization (Pind) at 80 kV/cm. (c) Calculated energy-storage density Wst and recoverable energy-storage density Wrec as a function of composition. (d) Calculated energy-loss density Wloss and energy-storage efficiency η as a function of composition. The top colored and shaded rows in (b)−(d) indicate the AFE and FE phase ranges and their boundary, and the composition ranges in which the critical field is higher and lower than the breakdown electric field, respectively, at room temperature.

at an electric field of 151 kV/cm or lower. In contrast, the higher Wrec values reported in refs 55−57 were achieved under a much higher electric field (DBS = 240−300 kV/cm), which represents a significant drawback. Figure 10b compares the PHf−PMW ceramics with the above-mentioned three ceramics in terms of Wrec, DBS, and Pmax as well as the preparation methods. It should be noted that the maximum polarization Pmax of the PHf−PMW ceramic achieved under the critical breakdown electric field is higher than any of the PBLYZST, 0.9KNN−0.1BMN, and Ta−AgNbO3 ceramics prepared by the conventional sintering method. (Only the PBLYZST ceramics prepared by the spark plasma sintering method55 exhibit a higher Pmax value.) These results place the PHf− PMW ceramics among the best materials for energy-storage applications at room temperature with a high Wrec and a large Pmax at a relatively low electric field. This optimum performance also confirms the validity of the strategy proposed in this work to improve Wrec and Pmax by chemical modification and microstructural engineering.

To put the energy-storage performance of the 0.90PbHfO3− 0.10PMW ceramic in perspective, we compare it with the performance of other lead-free and lead-based bulk ceramics. Figure 10a compiles the key parameters of the best-known ceramic material systems for energy storage, including the present system, in terms of recoverable energy density (Wrec) and electric field. It can be seen that the 0.90PbHfO3− 0.10PMW ceramics have a relatively high Wrec of 3.7 J/cm3 when compared with most lead-free and lead-based bulk ceramics so far reported in the literature.31−54 This high performance mainly arises from the large Pmax (45 μC/cm2) and high DBS (151 kV/cm). In fact, the Wrec of the 0.90PbHfO3−0.10PMW ceramics is superior to that of PZbased AFE−AFE solid solution ceramics, such as PbZrO3− Pb(Mg 1/2 W 1/2 )O 3 , 20 PbZrO 3 −Pb(Mn 1/2 W 1/2 )O 3 , 21 and PbZrO3−Pb(Ni1/3Nb2/3)O3.26 Although this Wrec value is slightly lower than that of the PBLYZST,55 0.9KNN− 0.1BMN,56 and Ta−AgNbO357 ceramics prepared by the conventional sintering method, it represents the highest value I

DOI: 10.1021/acs.chemmater.8b04470 Chem. Mater. XXXX, XXX, XXX−XXX

Article

Chemistry of Materials

enhance the field-induced maximum polarization (Pmax) so as to increase the recoverable energy-storage density Wrec and the energy-storage efficiency η. On the basis of this strategy, ceramics of the (1 − x)PbHfO3−xPb(Mg1/2W1/2)O3 solid solution between two AFE end members were prepared with various PMW concentrations via the solid-state reaction method, and their crystal structure, phase transitions, microstructure, and energy-storage performance were investigated. It is found that a partial order on the B site is developed with 10 mol % substitution of PMW. With the increase of the PMW content, the structure changes from an AFE orthorhombic phase to an FE orthorhombic phase at room temperature, and the AFE−FE phase-transition temperature shifts to room temperature. More interestingly, the energy-storage performance has been greatly improved with the partial substitution of PMW for PHf. The optimal energy-storage properties are achieved in the 0.90PHf−0.10PMW ceramics with a high relative density (96%), a high maximum polarization Pmax (45 μC/cm2), a large recoverable energy density Wrec (3.7 J/cm3), and a high energy-storage efficiency η (72.5% under 155 kV/ cm). Compared with the properties of the other AFE materials so far reported, the PHf−PMW ceramics demonstrate the following advantages: (1) they achieve the highest Wrec value at a practically attainable low electric field of around 150 kV/cm and (2) they show the largest Pmax among all of the ceramics prepared by the conventional sintering process. Such excellent energy-storage properties are thought to arise from the composition-driven dense microstructure following the transient liquid-phase-assisted sintering on the one hand and the softening of the antiferroelectric ordering due to the induced dipole frustration on the other hand. The concept and strategy of materials’ design and synthesis demonstrated in this work could have a profound impact on the development of new series of antiferroelectric ceramics of AFE−AFE solid solutions with improved energy-storage performance. A large energy-storage density Wrec in AFE materials requires a high critical field and a large induced polarization, but many current materials suffer from a low dielectric breakdown strength, which prevents them from achieving the field-induced AFE to FE phase transition. The formation of the PHf−PMW solid solution between two AFE end members of different antiparallel dipole arrangements has allowed us (i) to appropriately lower the critical field and at the same time enhance the induced polarization by tuning the interplay between the competing AFE/FE orderings so that the AFE order is slightly decreased whereas the FE strength is increased and (ii) to enhance the DBS by obtaining highquality ceramics with a dense and homogenous microstructure via liquid-phase sintering. These combined chemical, structural, and dipolar modifications have resulted in enhanced energy-storage performance in the PHf−PMW ceramics with a high Wrec and a large Pmax at a relatively low electric field, making them a promising family of new AFE materials for practical energy-storage applications. It is reasonable to believe that by improving the densification of the ceramics using such techniques as hot-pressing and/or plasma sintering and by fine adjusting the chemical compositions and polar structures, it would be possible to optimize the microstructure and properties in the ceramics of the PHf−PMW and related systems so as to achieve further enhanced energy-storage performance.

Figure 9. Weibull distribution for the (1 − x)PbHfO3−xPb(Mg1/2W1/2)O3 ceramics, which is used to calculate the average dielectric breakdown strength (DBS).

Figure 10. (a) Comparison of the recoverable energy density Wrec under the highest applied field of the 0.90PHf−0.10PMW ceramics with that of the other systems reported in the literature.31−57 (b) Comparison of the key energy-storage properties (Wrec, DBS, and Pmax) of the PBLYZST,55 0.9KNN−0.1BMN,56 Ta−AgNbO3,57 and 0.90PHf−0.10PMW ceramics prepared by the conventional sintering method; the properties of the PBLYZST ceramics prepared by the spark plasma sintering method are also illustrated by the shaded areas.

5. CONCLUSIONS To design novel antiferroelectric materials for energy-storage applications, we have proposed and demonstrated a combined chemical modification and microstructural engineering strategy to improve the dielectric breakdown strength (DBS) and J

DOI: 10.1021/acs.chemmater.8b04470 Chem. Mater. XXXX, XXX, XXX−XXX

Article

Chemistry of Materials



(11) Zhang, J.; Ji, L.; Jia, X.; Wang, J.; Zhai, J. W.; Zhou, Y. Crystal Structure and Effective Dielectric Response of Ba0.5Sr0.5TiO3-MgO Composites Synthesized in Situ Process. J. Am. Ceram. Soc. 2015, 98, 97−103. (12) Xu, B. M.; Moses, P.; Pai, N. G.; Cross, L. E. Charge release of lanthanum-doped lead zirconate titanate stannate antiferroelectric thin films. Appl. Phys. Lett. 1998, 72, 593−595. (13) Hao, X. H.; Zhai, J. W.; Kong, L. B.; Xu, Z. K. A comprehensive review on the progress of lead zirconate-based antiferroelectric materials. Prog. Mater. Sci. 2014, 63, 1−57. (14) Yao, Z. H.; Song, Z.; Hao, H.; Yu, Z. Y.; Cao, M. H.; Zhang, S. J.; Lanagan, M. T.; Liu, H. X. Homogeneous/InhomogeneousStructured Dielectrics and their Energy-Storage Performances. Adv. Mater. 2017, 29, No. 1601727. (15) Zhang, H.; Chen, X.; Cao, F.; Wang, G.; Dong, X.; Hu, Z.; Du, T. Charge-Discharge Properties of an Antiferroelectric Ceramics Capacitor Under Different Electric Fields. J. Am. Ceram. Soc. 2010, 93, 4015−4017. (16) Zhao, Y.; Hao, X. H.; Zhang, Q. Energy-Storage Properties and Electrocaloric Effect of Pb(1‑3x/2)LaxZr0.85Ti0.15O3 Antiferroelectric Thick Films. ACS Appl. Mater. Interfaces 2014, 6, 11633−11639. (17) Fujishita, H.; Ogawaguchi, A.; Katano, S. Analysis of Structures and Order Parameters in Antiferroelectric PbHfO3 Using Neutron Diffraction. J. Phys. Soc. Jpn. 2008, 77, No. 064601. (18) Kupriyanov, M. F.; Petrovich, E. V.; Dutova, E. V.; Kabirov, Y. V. Sequence of phase transitions in PbHfO3. Crystallogr. Rep. 2012, 57, 205−207. (19) Baldinozzi, G.; Sciau, P.; Pinot, M.; Grebille, D. Crystal structure of the antiferroelectric perovskite Pb2MgWO6. Acta Crystallogr., Sect. B: Struct. Sci. 1995, B51, 668−673. (20) Vittayakorn, N.; Charoonsuk, P.; Kasiansin, P.; Wirunchit, S.; Boonchom, B. Dielectric properties and phase transition behaviors in (1-x)PbZrO3-xPb(Mg1/2W1/2)O3 ceramics. J. Appl. Phys. 2009, 106, No. 064104. (21) Ren, Z. H.; Zhang, N.; Su, L.-W.; Wu, H.; Ye, Z.-G. Softening of antiferroelectricity in PbZrO3-Pb(Mn1/2W1/2)O3 complex perovskite solid solution. J. Appl. Phys. 2014, 116, No. 024103. (22) Viehland, D.; Dai, X. H.; Li, J. F.; Xu, Z. Effects of quenched disorder on La-modified lead zirconate titanate: Long- and shortrange ordered structurally incommensurate phases, and glassy polar clusters. J. Appl. Phys. 1998, 84, 458. (23) Swartz, S. L.; Shrout, T. R. Fabrication of perovskite lead magnesium niobate. Mater. Res. Bull. 1982, 17, 1245−1250. (24) Akbas, M. A.; Davies, P. K. Cation Ordering Transformations in the Ba(Zn1/3Nb2/3)O3-La(Zn2/3Nb1/3)O3 System. J. Am. Ceram. Soc. 1998, 81, 1061−1064. (25) Sternberg, A.; Shebanovs, L.; Birks, E.; Yamashita, Y.; Tyunina, M.; Zauls, V. Effects of structure ordering, structure defects and external conditions on properties of complex ferroelectric perovskites. Ferroelectrics 1998, 217, 307−317. (26) Wirunchit, S.; Laoratanakul, P.; Vittayakorn, N. Physical properties and phase transitions in perovskite Pb[Zr1‑x(Ni1/3Nb2/3)x]O3 (0.0 ≤ x ≤ 0.5) ceramics. J. Phys. D: Appl. Phys. 2008, 41, No. 125406. (27) Banlue, W.; Vittayakorn, N. Effect of Pb(Zn1/3Nb2/3)O3 Additions on Phase Structure, Ferroelectric and Dielectric Properties of PbZrO3 Ceramics. Ferroelectrics 2009, 382, 122−126. (28) Kingery, W. D.; Narasimhan, M. D. Densification during Sintering in the Presence of a Liquid Phase. II. Experimental. J. Appl. Phys. 1959, 30, 307−310. (29) Lu, C. H. Perovskite phase formation and microstructural evolution of lead magnesium tungstate-lead titanate ceramics. J. Mater. Sci. 1996, 31, 699−705. (30) Huang, J. J.; Zhang, Y.; Ma, T.; Li, H. T.; Zhang, L. W. Correlation between dielectric breakdown strength and interface polarization in barium strontium titanate glass ceramics. Appl. Phys. Lett. 2010, 96, No. 042902. (31) Xu, Q.; Liu, H.; Song, Z.; Huang, X.; Ullah, A.; Zhang, L.; Xie, J.; Hao, H.; Cao, M.; Yao, Z. A new energy-storage ceramics system

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (H.W.). *E-mail: [email protected] (W.R.). *E-mail: [email protected] (Z.-G.Y.). ORCID

Zuo-Guang Ye: 0000-0003-2378-7304 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The SEM and EDS measurements were carried out at International Center for Dielectric Research (ICDR), Xi’an Jiaotong University. The authors would like to thank Dr. Yijun Zhang for his help in the experiments. This work was supported by the U.S. Office of Naval Research (Grant Nos. N00014-12-1-1045 and N00014-16-1-3106), the Natural Sciences and Engineering Research Council of Canada (NSERC, Grant No. 203773), the National Natural Science Foundation of China (Grant Nos. 51332003, 51202184, and 61604123), the International Science and Technology Cooperation Program of China (Grant No. 2011DFA51880), and the “111 Project” of China (Grant No. B14040). P.G. would like to acknowledge the China Scholarship Council (CSC) for supporting his studies at SFU, and Z.L. would like to thank the China Postdoctoral Science Foundation (Grant No. 2018M643632) for support.



REFERENCES

(1) Liu, C.; Li, F.; Ma, L.-P.; Cheng, H.-M. Advanced materials for energy storage. Adv. Mater. 2010, 22, E28−E62. (2) Hadjipaschalis, I.; Poullikkas, A.; Efthimiou, V. Overview of current and future energy storage technologies for electric power applications. Renewable Sustainable Energy Rev. 2009, 13, 1513−1522. (3) Chen, H.; Cong, T. N.; Yang, W.; Tan, C.; Li, Y.; Ding, Y. Progress in electrical energy storage system: A critical review. Prog. Nat. Sci. 2009, 19, 291−312. (4) Cao, Y.; Irwin, P. C.; Younsi, K. The future of nanodielectrics in the electrical power industry. IEEE Trans. Dielectr. Electr. Insul. 2004, 11, 797−807. (5) Park, M. H.; Kim, H. J.; Kim, Y. J.; Moon, T.; Kim, K. D.; Hwang, C. S. Thin HfxZr1‑xO2 Films: A New Lead-Free System for Electrostatic Supercapacitors with Large Energy Storage Density and Robust Thermal Stability. Adv. Energy Mater. 2014, 4, No. 1400610. (6) Shi, Z. C.; Fan, R.-H.; Zhang, Z.-D.; Qian, L.; Gao, M.; Zhang, M.; Zheng, L.-T.; Zhang, X.-H.; Yin, L.-W. Random Composites of Nickel Networks Supported by Porous Alumina Toward Double Negative Materials. Adv. Mater. 2012, 24, 2349−2352. (7) 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. (8) Correia, T. M.; McMillen, M.; Rokosz, M. K.; Weaver, P. M.; Gregg, J. M.; Viola, G.; Cain, M. G. A Lead-Free and High-Energy Density Ceramic for Energy Storage Applications. J. Am. Ceram. Soc. 2013, 96, 2699−2702. (9) Lim, J. B.; Zhang, S.; Kim, N.; Shrout, T. R. High-Temperature Dielectrics in the BiScO3-BaTiO3-(K1/2Bi1/2)TiO3 Ternary System. J. Am. Ceram. Soc. 2009, 92, 679−682. (10) 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 2017, 5, 4710−4718. K

DOI: 10.1021/acs.chemmater.8b04470 Chem. Mater. XXXX, XXX, XXX−XXX

Article

Chemistry of Materials based on Bi0.5Na0.5TiO3 ternary solid solution. J. Mater. Sci.: Mater. Electron. 2016, 27, 322−329. (32) Wei, M.; Zhang, J.; Wu, K.; Chen, H.; Yang, C. Effect of BiNbO4 doping on the dielectric properties of BaTiO3 ceramics. Int. J. Appl. Ceram. Technol. 2018, 15, 197−202. (33) Chauhan, A.; Patel, S.; Vaish, R.; Bowen, C. R. AntiFerroelectric Ceramics for High Energy Density Capacitors. Materials 2015, 8, 8009−8031. (34) Song, Z.; Liu, H. X.; Zhang, S. J.; Wang, Z. J.; Shi, Y. T.; Hao, H.; Cao, M. H.; Yao, Z. H.; Yu, Z. Y. Effect of grain size on the energy storage properties of (Ba0.4Sr0.6)TiO3 paraelectric ceramics. J. Eur. Ceram. Soc. 2014, 34, 1209−1217. (35) Qu, B.; Du, H.; Yang, Z. Lead-free relaxor ferroelectric ceramics with high optical transparency and energy storage ability. J. Mater. Chem. C 2016, 4, 1795−1803. (36) Wang, T.; Jin, L.; Li, C.; Hu, Q.; Wei, X. Relaxor Ferroelectric BaTiO3-Bi(Mg2/3Nb1/3)O3 Ceramics for Energy Storage Application. J. Am. Ceram. Soc. 2015, 98, 559−566. (37) Chen, P.; Chu, B. Improvement of dielectric and energy storage properties in Bi(Mg1/2Ti1/2)O3-modified (Na1/2Bi1/2)0.92Ba0.08TiO3 ceramics. J. Eur. Ceram. Soc. 2016, 36, 81−88. (38) Wu, L.; Wang, X.; Li, L. T. Lead-free BaTiO3-Bi(Zn2/3Nb1/3)O3 weakly coupled relaxor ferroelectric materials for energy storage. RSC Adv. 2016, 6, 14273−14282. (39) Hu, Q.; Jin, L.; Wang, T.; Li, C.; Xing, Z.; Wei, X. Dielectric and temperature stable energy storage properties of 0.88BaTiO30.12Bi(Mg1/2Ti1/2)O3 bulk ceramics. J. Alloys Compd. 2015, 640, 416−420. (40) 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. (41) Yin, J.; Zhang, Y.; Lv, X.; Wu, J. Ultrahigh energy-storage potential under low electric field in bismuth sodium titanate-based perovskite ferroelectrics. J. Mater. Chem. A 2018, 6, 9823−9832. (42) Liu, G.; Fan, H.; Dong, G.; Shi, J.; Chang, Q. Enhanced energy storage and dielectric properties of Bi0.487Na0.427K0.06Ba0.026TiO3xCeO2 anti-ferroelectric ceramics. J. Alloys Compd. 2016, 664, 632− 638. (43) Yang, Z.; Du, H.; Qu, S.; Hou, Y.; Ma, H.; Wang, J. F.; Wang, J.; Wei, X.; Xu, Z. Significantly enhanced recoverable energy storage density in potassium-sodium niobate-based lead free ceramics. J. Mater. Chem. A 2016, 4, 13778−13785. (44) Zhang, L.; Jiang, S.; Zeng, Y.; Han, K.; Wang, Q.; Zhang, G.; et al. Y doping and grain size co-effects on the electrical energy storage performance of (Pb0.87Ba0.1La0.02)(Zr0.65Sn0.3Ti0.05)O3 antiferroelectric ceramics. Ceram. Int. 2014, 40, 5455−5460. (45) 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. (46) Zhang, L.; Jiang, S.; Fan, B.; Zhang, G. Enhanced energy storage performance in (Pb0.858Ba0.1La0.02Y0.008)(Zr0.65Sn0.3Ti0.05)O3(Pb0.97La0.02)(Zr0.9Sn0.05Ti0.05)O3 antiferroelectric composite ceramics by Spark Plasma Sintering. J. Alloys Compd. 2015, 622, 162−165. (47) Shen, Z.; Wang, X.; Luo, B.; Li, L. BaTiO3-BiYbO3 perovskite materials for energy storage applications. J. Mater. Chem. A 2015, 3, 18146−18153. (48) Gao, F.; Dong, X.; Mao, C.; Liu, W.; Zhang, H.; Yang, L.; Cao, F.; Wang, G. Energy-Storage Properties of 0.89Bi0.5Na0.5TiO30.06BaTiO3-0.05K0.5Na0.5NbO3 Lead-Free Anti-ferroelectric Ceramics. J. Am. Ceram. Soc. 2011, 94, 4382−4386. (49) Gao, F.; Dong, X.; Mao, C.; Cao, F.; Wang, G. c/a RatioDependent Energy-Storage Density in (0.9-x)Bi0.5Na0.5TiO3-xBaTiO30.1K0.5Na0.5NbO3 Ceramics. J. Am. Ceram. Soc. 2011, 94, 4162−4164. (50) 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.

(51) Ye, J.; Liu, Y.; Lu, Y.; Ding, J.; Ma, C.; Qian, H.; Yu, Z. Enhanced energy-storage properties of SrTiO3 doped (Bi1/2Na1/2)TiO3−(Bi1/2K1/2)TiO3 lead-free antiferroelectric ceramics. J. Mater. Sci.: Mater. Electron. 2014, 25, 4632−4637. (52) Ding, J.; Liu, Y.; Lu, Y.; Qian, H.; Gao, H.; Chen, H.; Ma, C. Enhanced energy-storage properties of 0.89Bi 0.5 Na 0.5 TiO 3 0.06BaTiO3-0.05K0.5Na0.5NbO3 lead-free anti-ferroelectric ceramics by two-step sintering method. Mater. Lett. 2014, 114, 107−110. (53) Li, Q.; Wang, J.; Liu, Z.; Dong, G.; Fan, H. Enhanced energystorage properties of BaZrO 3 -modified 0.80Bi 0.5 Na 0.5 TiO 3 0.20Bi0.5K0.5TiO3 lead-free ferroelectric ceramics. J. Mater. Sci. 2016, 51, 1153−1160. (54) Zhao, Y.; Xu, J.; Zhou, C.; Yuan, C.; Li, Q.; Chen, G.; Wang, H.; Yang, L. High energy storage properties and dielectric behavior of (Bi0.5Na0.5)0.94Ba0.06Ti1−x(Al0.5Nb0.5)xO3 lead-free ferroelectric ceramics. Ceram. Int. 2016, 42, 2221−2226. (55) Yi, J. Q.; Zhang, L.; Xie, B.; Jiang, S. L. The influence of temperature induced phase transition on the energy storage density of anti-ferroelectric ceramics. J. Appl. Phys. 2015, 118, No. 124107. (56) 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 2017, 5, 554− 563. (57) Zhao, L.; Liu, Q.; Gao, J.; Zhang, S. J.; Li, J.-F. Lead-Free Antiferroelectric Silver Niobate Tantalate with High Energy Storage Performance. Adv. Mater. 2017, 29, No. 1701824.

L

DOI: 10.1021/acs.chemmater.8b04470 Chem. Mater. XXXX, XXX, XXX−XXX