Next Article in Journal
Effect of Convective Cooling on the Temperature in a Friction System with Functionally Graded Strip
Previous Article in Journal
Study of the Influence of the Irradiation Flux Density on the Formation of a Defect Structure in AlN in the Case of the Effect of Overlapping of the Heavy Ion Motion Trajectories in the Near-Surface Layer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Tunability Achieving at Low Permittivity and Electric Field in (Ba0.91Ca0.09)(SnxZr0.2xTi0.8)O3-2 mol% CuO-1 mol% Li2CO3 Ceramics

1
Department of Electrical Engineering and Automation, Luoyang Institute of Science and Technology, Luoyang 471023, China
2
School of Microelectronics, Tianjin University, Tianjin 300072, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(15), 5226; https://doi.org/10.3390/ma16155226
Submission received: 4 July 2023 / Revised: 18 July 2023 / Accepted: 21 July 2023 / Published: 25 July 2023

Abstract

:
Ferroelectric varactors should have high tunability at low permittivity and a working electric field to obtain better impedance matching and stable tunability. In this work, (Ba0.91Ca0.09)(SnxZr0.2−xTi0.8)O3-2 mol% CuO-1 mol% Li2CO3 (abbreviated as BCSZT100x, x = 0.05, 0.10, 0.15 and 0.20, respectively) are prepared to achieve high tunability at low permittivity and a working electric field. The tunable mechanisms are investigated based on crystal structure, micro-morphology and the permittivity-temperature spectrum. The results show that the shrink of oxygen octahedron and weaker interaction force between Sn4+ and O2− make BCSZT5 ceramic have a higher tunability value of 26.55% at low permittivity (1913) and a working electric field (7.3 kV/cm). The tunability value of BCSZT5 ceramic increases by 58%, while its permittivity decreases by 25%, compared with x = 0. Those advantages make BCSZT5 ceramic have substantial application prospects in varactors.

1. Introduction

With the rapid development of The Internet of Things, 5G networks and autonomous traffic, the frequency spectrum has been scarce; it requires the frequency, bandwidth and phase of front-end devices (such as antennas, filters and phase shifters) to be regulated [1,2,3,4]. It is a good strategy to fabricate front-end devices using varactors whose capacitor can be expediently controlled by external magnetic field, electric field or mechanical force [5,6,7,8]. The utilization of varactors not only reduces the volume and manufacturing cost of communication equipment, but also makes communication equipment have the ability of multi-band scanning and transmitting signals. Taking mobile phones as an example, it is usually required to work in multiple communication modes synchronously, such as the global positioning system, Bluetooth and wireless network [9]. It requires to integrate different antennas for corresponding application, and there is no signal interference between those antennas. The requirements increase the volume and manufacturing cost circuits. If varactors are used, the number of discrete transceivers can be significantly reduced, and multi-functions can be achieved at a lower cost.
According to the used technologies and materials, varactors can be divided into micro-electromechanical systems (MEMS) varactors, semiconductor varactors, ferrite varactors and ferroelectric varactors [10,11,12,13]. Among those varactors, ferroelectric varactors have attracted tremendous attention due to advantages of large power handling, high-tuning speed, long switching lifetime and so on [14]. The large nonlinear dielectric responses and lower dielectric loss of ferroelectric materials supports the stronger signal control capability. Usually, the nonlinear dielectric responses can be evaluated by the parameter tunability K, which reflects the relative variation of permittivity ( ε ( E ) ) in the electric field ( K = ( 1 ε ( E ) / ε ( 0 ) ) × 100 % , ε ( 0 ) : permittivity at zero electric field). It can be seen that large ε ( 0 ) is helpful to achieve high tunability. However, varactors should have low permittivity in order to achieve a better impedance matching in circuit, which goes against the high tunability [15,16]. To achieve high tunability at low permittivity, many works improve the working electric field. Zhai et al. reported a higher tunability value of 32% at low permittivity (875) in Ba0.4Sr0.6TiO3-ZnAl2O4 composite ceramics when electric field strength was 30 kV/cm [17]. Zhang et al. achieved a tunability value of 17.3% and low permittivity of 830 at 30 kV/cm in 30 wt% BaCuSi2O6-70 wt% Ba0.55Sr0.45TiO3 composite ceramics [18]. Though many works have achieved high tunability at low permittivity, the working electric field is usually more than 15 kV/cm [19,20]. The high-working electric field goes against the stability of tunability due to the influence of polar nano-regions (PNRs) at a repeating electric filed. Yang et al. reported that the growth of PNRs was related to the electric field [21]. However, the contribution of polar nano-regions (PNRs) plays an important role on tunability. When a higher repeating electric field is applied, the stability of tunability is brittle. Xu et al. reported that the tunability of BZT ceramics deteriorates ~3% percent at 20 kV/cm, only after two cycle numbers [22]. With increasing cycle numbers, the degradation rate of tunability increases. The similar tendency also has been found in (Ba0.6Sr0.4)TiO3 ceramics [23]. Thus, achieving high tunability at low permittivity and a working electric field is crucial for the application of ferroelectric varactors in antennas, filters and phase shifters. However, some attention is necessary to investigate reducing the working electric field.
In our previous work, (Ba0.91Ca0.09)(Zr0.2Ti0.8)O3-2 mol% CuO-1 mol% Li2CO3 ceramic (abbreviated as BCZT-CL) can achieve a tunability value of 16.76% at a low-working electric field (7 kV/cm) [24]. But the higher permittivity (2634) is adverse to impedance matching and impedes its applications. Tao et al. investigated the dielectric properties of Ba(Ti1−xSnx)O3 [25]. Sn doping can make off-centre ion displacements unstable. The unstable off-centre ion displacements are expected to enhance the tunability. Thus, (Ba0.91Ca0.09)(SnxZr0.2−xTi0.8)O3-2 mol% CuO-1 mol% Li2CO3 (abbreviated as BCSZT100x, x = 0.05, 0.10, 0.15 and 0.20, respectively) are prepared to achieve high tunability at low permittivity and a working electric field. When the electric field strength is 7.3 kV/cm, BCSZT5 ceramics have a higher tunability value of 26.55% and lower permittivity value of 1913. The tunability value of BCSZT5 ceramics increases by 58%, while its permittivity decreases by 27%, compared with BCZT-CL ceramics. Those advantages make BCSZT5 ceramics have significant application prospects in varactors. Rigaku Corporation, Japan

2. Experimental Procedures

BCSZT100x ceramics were obtained via solid-state sintering using the preparation process reported in our previous work [26]. All mixed raw materials (TiO2, ZrO2, BaCO3, CaCO3 and SnO2) were calcined at 900 °C for 3 h to obtain pre-synthesis BCSZT powders. Then, 2 mol% CuO and 1 mol% Li2CO3 were added into pre-synthesis BCSZT powders. BCSZT100x ceramic greens were sintered at 1100 °C for 3 h. Rigaku D/MAX-2500 X-ray diffractometer (XRD, Rigaku Corporation, Tokyo, Japan) and Helios NanoLab 460HP scanning electron microscopy (SEM, Thermo Fisher Scientific, Waltham, MA, USA) were used to investigate the crystal structure and micro-morphology of BCSZT100x ceramics, respectively. Permittivity (ε), dielectric loss (tan δ) and tunability (K) was measured at 1 kHz using a HP 4278A capacitance meter equipped with a HM27004H C-T-V conversion device.

3. Results and Discussion

Figure 1 shows the powder XRD pattern of BCSZT100x ceramics. According to the measured results shown in Figure 1a and the powder diffraction file of JCPDS No. 75-0213, BCSZT100x ceramics sintered at 1100 °C are cubic phase perovskite structures, and the heterodox peaks appearing at the 2θ diffraction angle, ranging from 25° to 30° (shown in Figure 1b), may be the diffraction peak of CaZrTi2O7 (PDF#81-1500). Figure 1c shows the diffraction pattern of the 2θ diffraction angle between 44° and 46°. Obviously, with the increase of Sn4+ content, the diffraction peaks of BCSZT100x ceramics move to a higher angle, indicating the reduction in cell volume of BCSZT100x ceramics. The ionic radius is responsible for the phenomenon. When Sn4+ replaces Zr4+ to occupy B-site, the smaller ionic radius of Sn4+ (0.69 Å), compared with that of Zr4+ (0.72 Å), shrinks the lattice, resulting in the shift of diffraction peak to higher angle [27,28].
Figure 2 shows the grain size distribution of BCSZT100x ceramics, and the corresponding SEM results are illustrated. According to Figure 2, it can be seen that the grain size of BCSZT100x ceramics can be effectively regulated by controlling the content of Sn4+. As shown in Figure 2a, the grain sizes of BCSZT5 ceramics ranges from 0.15 µm to 1.65 µm, and there are many pores. An average grain size value of 0.47 µm was obtained. When x increases to 0.10, the average grain size of ceramics increases to 1.06 µm, but the grain size distribution range is significantly larger than that shown in Figure 2a. When x further increases to 0.15, the grain size of ceramics ranges from 0.7 µm to 13.3 µm, the average grain size increases to 1.98 µm, and a small number of grain sizes exceed 10 µm. When x is 0.20, the grain size of ceramics ranges from 0.15 µm to 0.75 µm, and the internal grain size is uniformly distributed with an average grain size of 0.38 µm. Based on the above results, the following conclusions can be drawn: the average grain size of BCSZT100x ceramics can be effectively improved by doping with Sn4+ properly, but the grain size distribution will be uneven.
Figure 3 shows the variation of the temperature-dependent permittivity of BCSZT100x ceramics. As x gradually increases from 0.05 to 0.15, the Curie temperature gradually moves towards low temperature, and the maximum dielectric constant gradually increases. However, when x increases to 0.20, the Curie temperature of BCSZT20 ceramics increases and the maximum dielectric constant decreases. The above phenomenon is consistent with the conclusion obtained by Wu et al. in (Ba0.98Ca0.02)(Ti0.94Sn0.06−xZrx)O3 ceramics [29]. In addition, it can be seen from Figure 3 that the permittivity of BCSZT100x ceramics can be significantly improved by doping an appropriate amount of Sn4+. The same phenomenon has also been found in (Ba0.8Sr0.2)(Ti1−xSnx)O3 and (Ba0.95Ca0.05)(Ti0.92Sn0.08−xZrx)O3 ceramics [30,31]. The change of permittivity may be mainly affected by grain size. A large number of grain boundaries induced by smaller grain sizes in ceramics can impede the movement of domain, leading to the deterioration of permittivity [32,33]. According to the SEM results, the largest average grain sizes of BCSZT15 ceramics support the highest permittivity. Moreover, the depressed and dilated permittivity of BCSZT100x ceramics suggest the existence of diffused phase transition (DPT) in ceramics. For BaTiO3-based ceramics, the DPT depends on the distribution of both A site and B site ions [34,35,36]. The random occupancy of A site or B site ions is in favour of the formation of nonuniform Coulomb field, which facilitates the generation of polar nano-regions (PNRs). However, different PNRs possess diverse local dielectric properties, thus a compressed and dilated permittivity peak can be obtained readily. When A or B sites are occupied by two different ion types, the difference in electronegativity and ionic radii between two different ion types accelerates the formation of PNRs, resulting in the appearance of a more compressed permittivity peak. According to results reported in BaTiO3 and (Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 ceramics, smaller grain sizes also enhanced the diffused phase transition due to the grain size effect [37,38].
The DPT of ceramics can be discussed using empirical Lorenz formula [39]:
ε A ε = 1 + T T A 2 2 δ A 2
where diffusion coefficient δ A shown in Equation (1) represents the degree of DPT and ε A represent the permittivity at temperature T A . The fitting results about the temperature-dependent permittivity are shown in Figure 4a–d) and corresponding parameters are listed. Obviously, the diffusion coefficient δ A of BCSZT100x ceramics has little difference, indicating that there is no significant difference in the degree of DPT. In order to more intuitively compare the degree of DPT of different ceramics, the measured results in Figure 3 are normalized and shown in Figure 4e. The coincident normalized curves indicate that doping Sn4+ in BCZT ceramics cannot affect the DPT in our work, which is consistent with the result obtained by Equation (1). This phenomenon contradicts the result reported in Ba(Zr0.2Ti0.8)1−xSnxO3 ceramics. Fu et al. found that the degree of DPT in Ba(Zr0.2Ti0.8)1−xSnxO3 ceramics became stronger with increasing Sn4+ content [40]. In our work, BCSZT100x ceramics are sintered at a low temperature (1100 °C), resulting in the insufficient grain growth, uneven grain size distribution and the smaller average grain size. Thus, the grain size effect makes the DPT of BCSZT100x ceramic cannot been observed significantly.
Figure 5 shows the tunability and dielectric loss of BCSZT100x ceramics under different electric field strengths measured at room temperature. It can be seen that BCSZT5 ceramics have the largest tunability but the smallest for BCSZ20 ceramic under the same electric field intensity. According to Figure 3, it can be seen that the permittivity of BCSZT5 ceramics at room temperature (1913) is smaller than that of BCSZT20 ceramics (2192). Therefore, the tunability of BCSZT20 ceramics should be greater than that of BCSZT5 ceramics according to the previous study. It is obviously inconsistent with the measured results. For BaTiO3 based ceramics, tunability are affected not only by grain size, but also by PNRs and oxygen octahedron. Ren et al. studied the tunability of BaTi0.85Sn0.15O3/MgO ceramics [41]. The contribution of polar microregion to tunability is dominant at a lower electric field intensity, while the contribution of B-site ion polarization to tunability is dominant at a higher electric field intensity. According to the results shown in Figure 4, there is no significant difference in the degree of DPT among BCSZT100x ceramics. Therefore, it is inferred that the effect of PNRs on the tunability of BCSZT100x ceramics is basically the same. Meanwhile, Mahmoud et al. studied the tunable properties of (0.95 − x)(Bi0.5Na0.3K0.2)TiO3-xSrTiO3-0.05 (Ba0.8Ca0.2)TiO3 ceramics [42]. When the temperature was higher than Curie temperature, the permittivity was mainly contributed by the crystal structure. According to XRD results shown in Figure 1, the cell volume of BCSZT100x ceramics decreases and the oxygen octahedron shrinks when increasing the Sn4+ content, resulting in the limited movement of B-site ions. It suggests that B-site ions require more energy to generate greater displacement to achieve higher tunability. Therefore, the tunability of BCSZT100x ceramics under the same electric field intensity decreases with increasing Sn4+ content. Moreover, the interaction force of B-site ions also plays a key role on the tunability. It can be speculated that the interaction force between Sn4+ and O2− is weaker than that between Zr4+ and O2− since the melting point of SnO2 (1630 °C) is smaller than that of ZrO2 (2700 °C) [43,44]. The weaker interaction force between Sn4+ and O2− indicates that the movement of B-site ions in the oxygen octahedron will be more sensitive to the electric field after substituting Zr4+ with Sn4+. Therefore, the tunability of BCSZT5 ceramics (26.55%) is greater than that of BCZT-CL ceramics (16.76%), as reported in our previous work [24].
It can be observed from Figure 5 that BCSZT5 ceramics possess a higher tunability and lower tan δ simultaneously. In order to compare the tunable performance among those ceramics quantificationally, the figure of merit (FOM), which takes into consideration K and tan δ together, is used.
F O M = K tan δ
The relationship between the FOM value and electric field strength is shown in Figure 6. With increasing Sn4+ content, the FOM value of BCSZT100x ceramics gradually decreases at the same electric field intensity. When the electric field strength is 7.3 kV/cm, the BCSZT5 ceramics have the highest FOM value (54.29) due to their higher tunability (26.55%) and lower tan δ (0.00489), which is higher than that of BCZT-CL ceramics (34.98) [24]. The FOM value of BCSZT5 ceramics increases by 58%, while its permittivity decreases by 25%, compared with BCZT-CL ceramics at 7.3 kV/cm. The results show that substituting Zr4+ with Sn4+ is beneficial to improve the tunable performance. Table 1 shows the tunable performance of other ceramics [45,46,47,48,49,50,51,52]. Obviously, BCSZT5 ceramics not only have a low sintering temperature (1100 °C) and permittivity (1913), but also can achieve high tunability under a lower electric field strength. Those advantages make BCSZT5 ceramics have greater application prospects in varactors.

4. Conclusions

BCSZT100x ceramics are prepared to achieve high tunability at low permittivity and a working electric field. The crystal structure, micro-morphology and permittivity-temperature spectrum are used to study the diffused phase transition and tunable mechanisms. XRD results show that the cell of BCSZT100x ceramics shrinks with the increase of Sn4+ content because the ionic radius of Sn4+ (0.69 Å) is smaller than that of Zr4+ (0.72 Å). The decrease of cell supports to the shrink of oxygen octahedron. Meanwhile, SEM results indicate that the average grain size of BCSZT100x ceramics can be effectively improved by doping Sn4+ properly, but the grain size is still less than 2 µm. The smaller grain size obtained in BCSZT100x ceramics means that the DPT of BCSZT100x ceramics cannot be observed significantly. It indicates that the tunability of BCSZT100x ceramics are primarily affected by oxygen octahedron. The shrink of oxygen octahedron with increasing Sn4+ content suggests that the displacement of B-site ions is suppressed, resulting in the reduction in tunability with increasing Sn4+ content. However, the weaker interaction force between Sn4+ and O2−, which is speculated based on the melting point of SnO2 and ZrO2, makes the movement of B-site ions in the oxygen octahedron be more sensitive to the electric field. The influence, by shrinking the oxygen octahedron and weaker interaction force between Sn4+ and O2−, means that BCSZT5 ceramics have a higher tunability value of 26.55% at low permittivity (1913) and a working electric field (7.3 kV/cm). Those advantages make BCSZT5 ceramics have greater application prospects in varactors.

Author Contributions

Conceptualization, B.W., P.Y. and S.Y.; Formal analysis, B.W., P.Y., L.Z. and X.J.; Investigation, L.Z. and X.J.; Methodology, P.Y. and L.Z.; Project administration and supervision, B.W. and S.Y.; Writing—original draft, B.W. and P.Y.; Writing—review and editing, B.W., P.Y. and S.Y.; Funding acquisition, B.W. and S.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (Grant Nos. 52175525, 61971211), Science and Technology Project of Henan Province (Grant Nos. 222102230032).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this research are available from the corresponding authors on reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bronckers, L.A.; Roc’h, A.; Smolders, A.B. A new design method for frequency-reconfigurable antennas using multiple tuning components. IEEE Trans. Antennas Propag. 2019, 67, 7285–7295. [Google Scholar] [CrossRef] [Green Version]
  2. Schuster, C.; Schynol, L.; Wentzel, A.; Huhn, F.; Polat, E.; Schmidt, S.; Jakoby, R.; Heinrich, W.; Maune, H. Concept for con-tinuously tunable output filters for digital transmitter architectures. IEEE Access 2019, 7, 123490–123504. [Google Scholar] [CrossRef]
  3. Gu, Z.; Pandya, S.; Samanta, A.; Liu, S.; Xiao, G.; Meyers, C.; Damodaran, A.; Barak, H.; Dasgupta, A.; Saremi, S.; et al. Resonant do-main-wall-enhanced tunable microwave ferroelectrics. Nature 2018, 560, 622–627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ushikoshi, D.; Higashiura, R.; Tachi, K.; Fathnan, A.A.; Mahmood, S.; Takeshita, H.; Homma, H.; Akram, M.R.; Vellucci, S.; Lee, J.; et al. Pulse-driven self-reconfigurable meta-antennas. Nat. Commun. 2023, 14, 633. [Google Scholar] [CrossRef] [PubMed]
  5. Shirkolaei, M.M.; Aslinezhad, M. Substrate integrated waveguide filter based on the magnetized ferrite with tunable capability. Microw. Opt. Technol. Lett. 2020, 63, 1120–1125. [Google Scholar] [CrossRef]
  6. Aspe, B.; Castel, X.; Demange, V.; Deputier, S.; Bouquet, V.; Benzerga, R.; Sauleau, R.; Guilloux-Viry, M. Frequency-tunable slot-loop antenna based on knn ferroelectric interdigitated varactors. IEEE Antennas Wirel. Propag. Lett. 2021, 20, 1414–1418. [Google Scholar] [CrossRef]
  7. Obuh, I.; Doychinov, V.; Steenson, D.; Akkaraekthalin, P.; Robertson, I.; Somjit, N. Low-cost microfabrication for MEMS switches and varactors, IEEE Transactions on Components. Packag. Manuf. Technol. 2018, 8, 1702–1710. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, S.; Liao, J.; Huang, X.; Zhang, Z.; Wang, W.; Wang, X.; Shan, Y.; Li, P.; Hong, Y.; Peng, Z.; et al. Green fabrication of freestanding piezoceramic films for energy harvesting and virus detection. Nano-Micro Lett. 2023, 15, 131. [Google Scholar] [CrossRef]
  9. Ullah, S.; Hayat, S.; Umar, A.; Ali, U.; Tahir, F.; Flint, J.A. Design, fabrication and measurement of triple band frequency re-configurable antennas for portable wireless communications. AEU—Int. J. Electron. Commun. 2017, 81, 236–242. [Google Scholar] [CrossRef]
  10. Mirebrahimi, S.M.; Dousti, M.; Afrang, S. MEMS tunable filters based on DGS and waveguide structures: A literature review. Analog. Integr. Circuits Signal Process. 2021, 108, 141–164. [Google Scholar] [CrossRef]
  11. Chen, P.C.; Asbeck, P.M.; Dayeh, S.A. Freestanding high-power gan multi-fin camel diode varactors for wideband telecom tunable filters. IEEE Trans. Electron Devices 2023, 70, 963–970. [Google Scholar] [CrossRef]
  12. Bouca, P.; Figueiredo, R.; Matos, J.N.; Vilarinho, P.M.; Carvalho, N.B. Reconfigurable three functional dimension single and dual-band sdr front-ends using thin film bst-based varactors. IEEE Access 2022, 10, 4125–4136. [Google Scholar] [CrossRef]
  13. Berneti, E.K.; Ghalibafan, J. Tunable ferrite-based metamaterial structure and its application to a leaky-wave antenna. J. Magn. Magn. Mater. 2018, 456, 223–227. [Google Scholar] [CrossRef]
  14. Subramanyam, G.; Cole, M.; Sun, N.; Kalkur, T.; Sbrockey, N.; Tompa, G.; Guo, X.; Chen, C.; Alpay, S.; Rossetti, G.; et al. Challenges and opportunities for multi-functional oxide thin films for voltage tunable radio frequency/microwave components. J. Appl. Phys. 2013, 114, 191301. [Google Scholar] [CrossRef] [Green Version]
  15. Chou, X.; Wang, J.; Zhao, Z.; Geng, W.; Zhang, W.; Zhai, J. Dielectric tunability and microwave properties of Ba0.5Sr0.5TiO3-BaMg6Ti6O19 composite ceramics for tunable microwave device applications. J. Electroceram. 2011, 26, 185–190. [Google Scholar] [CrossRef]
  16. Menesklou, W.; Paul, F.; Zhou, X.; Elsenheimer, H.; Binder, J.R.; Ivers-Tiffée, E. Nonlinear ceramics for tunable microwave devices part I: Materials properties and processing. Microsyst. Technol. 2011, 17, 203–211. [Google Scholar] [CrossRef]
  17. Wang, W.; Zhang, M.; Xin, L.; Shen, S.; Zhai, J. Effect of interface behavior on dielectric properties of ferroelectric-dielectric composite ceramics. J. Alloys Compd. 2019, 809, 151712. [Google Scholar] [CrossRef]
  18. Zhang, M.; Xin, L.; Dong, Z.; Zhang, Y.; Jin, J.; Li, Z.; Xue, Y.; Ren, L.; Zhai, J. Structure and dielectric properties of Ba1-xSrxCuSi2O6-Ba0.55Sr0.45TiO3 ferroelectric-dielectric composite ceramics. J. Mater. Sci. Mater. Electron. 2021, 32, 21318–21325. [Google Scholar] [CrossRef]
  19. Xu, Z.; Fu, Z.; Ling, F.; Xu, Y. Simultaneously with lower permittivity and higher tunability achieved in Ba0.5Sr0.5TiO3-Mg2SiO4-MgO composite ceramics prepared by spark plasma sintering. J. Mater. Sci. Mater. Electron. 2022, 33, 23630–23638. [Google Scholar] [CrossRef]
  20. Ali, A.F.; Mahmoud, A.E.-R.; Gami, F.; Nassary, M.M.; Parashar, S.K.S. Enhancement of tunability properties of (Ba0.95Ca0.05)tio3 lead free ceramic by bawo4 for microwave applications. J. Electron. Mater. 2019, 48, 2940–2949. [Google Scholar] [CrossRef]
  21. Wang, R.-X.; Zhang, J.; Zheng, L.-M.; Sun, L.; Luo, Z.-L.; Zhang, S.-T.; Yang, B. Evolution of polar nano-regions under electric field around ferro-paraelectric transition temperature and its contribution to piezoelectric property in Pb(Mg1/3Nb2/3)O3-0.30PbTiO3 crystal. Ceram. Int. 2018, 44, 18084–18089. [Google Scholar] [CrossRef]
  22. Xu, Q.; Zhan, D.; Liu, H.-X.; Chen, W.; Huang, D.-P.; Zhang, F. Evolution of dielectric properties in BaZrxTi1−xO3 ceramics: Effect of polar nano-regions. Acta Mater. 2013, 61, 4481–4489. [Google Scholar] [CrossRef]
  23. Zhang, X.-F.; Xu, Q.; Liu, H.-X.; Chen, W.; Chen, M.; Kim, B.-H. Field history dependence of nonlinear dielectric properties of Ba0.6Sr0.4TiO3 ceramics under bias electric field: Polarization behavior of polar nano-regions. Phys. B Condens. Matter 2011, 406, 1571–1576. [Google Scholar] [CrossRef]
  24. Yang, P.; Peng, W.; Xu, K.; Li, L.; Yu, S. The effect of Li2CO3 sintering additive on the dielectric and tunable performance of (Ba0.91Ca0.09)(Zr0.2Ti0.8)O3-CuO ceramics. Ceram. Int. 2021, 47, 23571–23577. [Google Scholar] [CrossRef]
  25. Shi, T.; Xie, L.; Gu, L.; Zhu, J. Why Sn doping significantly enhances the dielectric properties of Ba(Ti1−xSnx)O3. Sci. Rep. 2015, 5, 8606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Yang, P.; Zhao, L.; Wang, B.; Zhang, L.; Yu, S. Enhanced dielectric and tunable performance in BCZT91/BCZT87 bilayer ce-ramics. Ceram. Int. 2023, 49, 16918–16923. [Google Scholar] [CrossRef]
  27. Zhao, B.; Zhu, Q.; Sun, X.; Li, J.-G. Co-doping Zn2+/Sn4+ in ZnGa2O4:Cr3+ for dynamic near-infrared luminescence and advanced anti-counterfeiting. Ceram. Int. 2021, 47, 17000–17007. [Google Scholar] [CrossRef]
  28. Li, J.; Peng, T.; Zhang, Y.; Zhou, C.; Zhu, A. Polyaniline modified SnO2 nanoparticles for efficient photocatalytic reduction of aqueous Cr(VI) under visible light. Sep. Purif. Technol. 2018, 201, 120–129. [Google Scholar] [CrossRef]
  29. Wu, B.; Ma, J.; Wu, W.; Chen, M. Microstructure, phase structure, and electrical properties of Zr-modified BaTiO3-based lead-free ceramics. J. Mater. Sci. Mater. Electron. 2017, 29, 1584–1591. [Google Scholar] [CrossRef]
  30. Wang, H.; Wu, J. Ba0.95Ca0.05Ti0.92Sn0.08-xZrxO3 lead-free ceramics: Microstructure and piezoelectricity. J. Mater. Sci. Mater. Electron. 2015, 26, 4119–4123. [Google Scholar] [CrossRef]
  31. Chihaoui, S.; Seveyrat, L.; Perrin, V.; Kallel, I.; Lebrun, L.; Khemakhem, H. Structural evolution and electrical characteristics of Sn-doped Ba0.8Sr0.2TiO3 ceramics. Ceram. Int. 2017, 43, 427–432. [Google Scholar] [CrossRef]
  32. Schultheiß, J.; Checchia, S.; Uršič, H.; Frömling, T.; Daniels, J.; Malič, B.; Rojac, T.; Koruza, J. Domain wall-grain boundary interactions in polycrystalline Pb(Zr0.7Ti0.3)O3 piezoceramics. J. Eur. Ceram. Soc. 2020, 40, 3965–3973. [Google Scholar] [CrossRef] [Green Version]
  33. Tan, Y.; Viola, G.; Koval, V.; Yu, C.; Mahajan, A.; Zhang, J.; Zhang, H.; Zhou, X.; Tarakina, N.; Yan, H. On the origin of grain size effects in Ba(Ti0.96Sn0.04)O3 perovskite ceramics. J. Eur. Ceram. Soc. 2019, 39, 2064–2075. [Google Scholar] [CrossRef]
  34. Liu, Q.; Zhang, Y.; Gao, J.; Zhou, Z.; Wang, H.; Wang, K.; Zhang, X.; Li, L.; Li, J.-F. High-performance lead-free piezoelectrics with local structural heterogeneity. Energy Environ. Sci. 2018, 11, 3531–3539. [Google Scholar] [CrossRef]
  35. Auromun, K.; Acharya, T.; Choudhary, R. Octahedral distortion-driven phase transition, relaxation and conduction process of Zr/Sn modified barium titanate relaxors. Ceram. Int. 2022, 48, 20858–20871. [Google Scholar] [CrossRef]
  36. Talanov, M.V.; Bush, A.A.; Sirotinkin, V.P.; Kozlov, V.I. Structural origin of strongly diffused ferroelectric phase transition in Ba(Ti, Zr)O3-based ceramics. Acta Mater. 2022, 227, 117734. [Google Scholar] [CrossRef]
  37. Tsuji, K.; Ndayishimiye, A.; Lowum, S.; Floyd, R.; Wang, K.; Wetherington, M.; Maria, J.-P.; Randall, C.A. Single step densifi-cation of high permittivity BaTiO3 ceramics at 300 °C. J. Eur. Ceram. Soc. 2020, 40, 1280–1284. [Google Scholar] [CrossRef]
  38. Coondoo, I.; Panwar, N.; Alikin, D.; Bdikin, I.; Islam, S.S.; Turygin, A.; Shur, V.Y.; Kholkin, A.L. A comparative study of structural and electrical properties in lead-free BCZT ceramics: Influence of the synthesis method. Acta Mater. 2018, 155, 331–342. [Google Scholar] [CrossRef]
  39. Liu, L.; Ren, S.; Zhang, J.; Peng, B.; Fang, L.; Wang, D. Revisiting the temperature-dependent dielectric permittivity of Ba(Ti1-xZrx)O3. J. Am. Ceram. Soc. 2018, 101, 2408–2416. [Google Scholar] [CrossRef]
  40. Fu, X.; Cai, W.; Cheng, G.; Gao, R. Effects of Sn doping on the microstructure and dielectric and ferroelectric properties of Ba(Zr0.2Ti0.8)O3 ceramics. J. Mater. Sci. Mater. Electron. 2017, 28, 8177–8185. [Google Scholar] [CrossRef]
  41. Ren, P.; Wang, Q.; Li, S.; Zhao, G. Energy storage density and tunable dielectric properties of BaTi0.85Sn0.15O3/MgO composite ceramics prepared by SPS. J. Eur. Ceram. Soc. 2017, 37, 1501–1507. [Google Scholar] [CrossRef]
  42. Mahmoud, A.E.-R.; Babeer, A.M. Intrinsic and extrinsic contributions in nonlinear dielectric response of (Bi0.5Na0.3K0.2)TiO3-(Ba0.8Ca0.2)TiO3-Based SrTiO3 ceramics driven by the rayleigh model. J. Electron. Mater. 2021, 51, 378–390. [Google Scholar] [CrossRef]
  43. Kumar, V.; Jaiswal, M.; Gupta, R.; Ram, J.; Sulania, I.; Ojha, S.; Sun, X.; Koratkar, N.; Kumar, R. Effect of low energy (keV) ion irradiation on structural, optical and morphological properties of SnO2-TiO2 nanocomposite thin films. J. Mater. Sci. Mater. Electron. 2018, 29, 13328–13336. [Google Scholar] [CrossRef]
  44. Xiaoqing, L.W. Synthesis of high temperature resistant ZrO2-SiO2 composite aerogels via supercritical fluid deposition. Indian J. Chem. Technol. 2023, 30, 198–204. [Google Scholar]
  45. Chou, X.; Zhai, J.; Yao, X. Dielectric tunable properties of low dielectric constant Ba0.5Sr0.5TiO3–Mg2TiO4 microwave composite ceramics. Appl. Phys. Lett. 2007, 91, 122908. [Google Scholar] [CrossRef]
  46. Lu, X.; Tong, Y.; Zhang, L.; Ma, B.; Cheng, Z.-Y. High dielectric tunability in composites prepared using SiO2 coated Ba0.5Sr0.5TiO3 nanoparticles. Ceram. Int. 2018, 44, 9875–9879. [Google Scholar] [CrossRef]
  47. Liang, R.; Zhou, Z.; Dong, X.; Wang, G.; Cao, F.; Hu, Z.; Jiang, K. Enhanced dielectric tunability of Ba0.55Sr0.45TiO3-ZnAl2O4 composite ceramic. Ceram. Int. 2015, 41, S551–S556. [Google Scholar] [CrossRef]
  48. Ren, P.; Fan, H.; Wang, X.; Tan, X. Modified tunable dielectric properties by addition of MgO on BaZr0.25Ti0.75O3 ceramics. Mater. Res. Bull. 2011, 46, 2308–2311. [Google Scholar] [CrossRef]
  49. Valant, M.; Suvorov, D. Low-temperature sintering of (Ba0.6Sr0.4)TiO3. J. Am. Ceram. Soc. 2004, 87, 1222–1226. [Google Scholar] [CrossRef]
  50. Liu, G.; Li, Y.; Gao, J.; Li, D.; Yu, L.; Dong, J.; Zhang, Y.; Yan, Y.; Fan, B.; Liu, X.; et al. Structure evolution, ferroelectric properties, and energy storage performance of CaSnO3 modified BaTiO3-based Pb-free ceramics. J. Alloys Compd. 2020, 826, 154160. [Google Scholar] [CrossRef]
  51. Liu, Z.; Fan, H.; Long, C. Dielectric nonlinearity and electrical properties of K0.5Na0.5NbO3–SrTiO3 relaxor ferroelectrics. J. Mater. Sci. 2014, 49, 8107–8115. [Google Scholar] [CrossRef]
  52. Liu, G.; Li, Y.; Shi, M.; Yu, L.; Chen, P.; Yu, K.; Yan, Y.; Jin, L.; Wang, D.; Gao, J. An investigation of the dielectric energy storage performance of Bi(Mg2/3Nb1/3)O3-modifed BaTiO3 Pb-free bulk ceramics with improved temperature/frequency stability. Ceram. Int. 2019, 45, 19189–19196. [Google Scholar] [CrossRef]
Figure 1. Powder XRD pattern of BCSZT100x ceramics at (a) 20~80°, (b) 25~30° and (c) 44~46°. The dark yellow lines shown in (b) are the diffraction peaks of CaZrTi2O7 (PDF#81-1500) and green lines shown in (c) is the moving direction of diffraction peaks.
Figure 1. Powder XRD pattern of BCSZT100x ceramics at (a) 20~80°, (b) 25~30° and (c) 44~46°. The dark yellow lines shown in (b) are the diffraction peaks of CaZrTi2O7 (PDF#81-1500) and green lines shown in (c) is the moving direction of diffraction peaks.
Materials 16 05226 g001
Figure 2. The grain size distribution of BCSZT100x ceramics: (a) BCSZT5, (b) BCSZT10, (c) BCSZT15 and (d) BCSZT20.
Figure 2. The grain size distribution of BCSZT100x ceramics: (a) BCSZT5, (b) BCSZT10, (c) BCSZT15 and (d) BCSZT20.
Materials 16 05226 g002
Figure 3. The temperature-dependent permittivity of BCSZT100x ceramics.
Figure 3. The temperature-dependent permittivity of BCSZT100x ceramics.
Materials 16 05226 g003
Figure 4. (ad) The fitting results using Equation (1) and (e) the normalized of permittivity near Cuire temperature.
Figure 4. (ad) The fitting results using Equation (1) and (e) the normalized of permittivity near Cuire temperature.
Materials 16 05226 g004
Figure 5. The tunability and dielectric loss of BCSZT100x ceramics.
Figure 5. The tunability and dielectric loss of BCSZT100x ceramics.
Materials 16 05226 g005
Figure 6. The relationship between FOM and electric field strength in BCSZT100x ceramics.
Figure 6. The relationship between FOM and electric field strength in BCSZT100x ceramics.
Materials 16 05226 g006
Table 1. The tunable performance of other ceramics reported in references.
Table 1. The tunable performance of other ceramics reported in references.
SamplesSintering Temperature (°C)ε(0)Tunability@Working Electric FieldRef.
Ba0.5Sr0.5TiO31400250022.7% (30 kV/cm)[45]
Ba0.5Sr0.5TiO31230~174647% (100 kV/cm)[46]
Ba0.55Sr0.45TiO3-20 wt% ZnAl2O41400236246.4% (20 kV/cm)[47]
BaZr0.25Ti0.75O3-10 wt%MgO1350182138.2% (10 kV/cm)[48]
Ba0.6Sr0.4TiO3+ 0.8 wt% Li2O900190016.4% (30 kV/cm)[49]
0.95BaTiO3-0.05 CaSnO31400~210021.15% (30 kV/cm)[50]
K0.5Na0.5NbO3-0.2SrTiO31250112617.4% (50 kV/cm)[51]
0.875BaTiO3-0.125Bi(Mg2/3Nb1/3)O31250~1050<3% (40 kV/cm)[52]
BCSZT51100191326.55% (7.3 kV/cm)This 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

Wang, B.; Zhao, L.; Jia, X.; Yang, P.; Yu, S. Enhanced Tunability Achieving at Low Permittivity and Electric Field in (Ba0.91Ca0.09)(SnxZr0.2xTi0.8)O3-2 mol% CuO-1 mol% Li2CO3 Ceramics. Materials 2023, 16, 5226. https://doi.org/10.3390/ma16155226

AMA Style

Wang B, Zhao L, Jia X, Yang P, Yu S. Enhanced Tunability Achieving at Low Permittivity and Electric Field in (Ba0.91Ca0.09)(SnxZr0.2xTi0.8)O3-2 mol% CuO-1 mol% Li2CO3 Ceramics. Materials. 2023; 16(15):5226. https://doi.org/10.3390/ma16155226

Chicago/Turabian Style

Wang, Bo, Le Zhao, Xiuhuai Jia, Pan Yang, and Shihui Yu. 2023. "Enhanced Tunability Achieving at Low Permittivity and Electric Field in (Ba0.91Ca0.09)(SnxZr0.2xTi0.8)O3-2 mol% CuO-1 mol% Li2CO3 Ceramics" Materials 16, no. 15: 5226. https://doi.org/10.3390/ma16155226

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