Next Article in Journal
Study of Microstructural and Mechanical Properties of Al/SiC/TiO2 Hybrid Nanocomposites Developed by Microwave Sintering
Next Article in Special Issue
Pyrrolylquinoline-BF2 and BPh2 BODIPY-Type Analogues: Synthesis, Structural Analysis and Photophysical Properties
Previous Article in Journal
Synthesis, Crystal Structure, Inhibitory Activity and Molecular Docking of Coumarins/Sulfonamides Containing Triazolyl Pyridine Moiety as Potent Selective Carbonic Anhydrase IX and XII Inhibitors
Previous Article in Special Issue
Synthesis of Novel Aqua ƞ4-NNNO/Cu(II) Complexes as Rapid and Selective Oxidative Catalysts for O-Catechol: Fluorescence, Spectral, Chromotropism and Thermal Analyses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Processing and Characterization of BCZT-Modified BiFeO3-BaTiO3 Piezoelectric Ceramics

by
Rizwan Ahmed Malik
1,* and
Hussein Alrobei
2,*
1
Department of Metallurgy and Materials Engineering, Faculty of Mechanical and Aeronautical Engineering, University of Engineering and Technology (UET), Taxila 47050, Pakistan
2
Department of Mechanical Engineering, College of Engineering, Prince Sattam Bin Abdul Aziz University, Al-Kharj 11942, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Crystals 2021, 11(9), 1077; https://doi.org/10.3390/cryst11091077
Submission received: 11 August 2021 / Revised: 29 August 2021 / Accepted: 30 August 2021 / Published: 6 September 2021
(This article belongs to the Special Issue New Trends in Crystals at Saudi Arabia)

Abstract

:
The synthesis of non-lead piezoelectric ceramics (1–z)(0.65Bi1.05Fe2O3-0.35BaTiO3)-z Ba(Ti0.8Zr0.2)O3-(Ba0.7Ca0.3)TiO3 using a solid state method and a quenching strategy was investigated. The processing conditions such as the sintering temperature and soaking time were optimized. The patterns of X-ray diffraction (XRD) displayed a pure perovskite structure with no secondary phases. The ferroelectric and piezoelectric characteristics of the samples were considerably improved as a result of the lattice strain. The findings of the experiment revealed that the quenching technique increases the piezoelectric sensor constant of 152 pC/N in optimized conditions. The enhanced piezoelectric sensor constant (d33) value at z = 0.020 was ascribed to the incorporation of multi-cationic BCZT, which modified the bond lengths at a unit cell level and gave rise to more flexibility in complex domain switching. This facilitated easier domain alignment in response to the applied field and resulted in an improvement in the electrical properties.

1. Introduction

Lead-based piezoelectric ceramics with superior piezoelectric characteristics are suitable for actuators, energy storage capacitor applications, nanogenerators for energy harvesting, nano-positioners, nanosensors, piezocatalysts, switching and sensing devices, and transducers [1,2,3,4,5,6,7,8,9,10,11,12,13]. Lead, on the other hand, has negative impacts on human health [14,15,16] and causes other environmental issues. Furthermore, these ceramics have a relatively high ferroelectric phase transition temperature (TC) (about 400 °C), making them appropriate for high-temperature operations [17]. Due to environmental concern, lead-based ceramics must therefore be replaced with lead-free piezoelectric materials having high working temperature. A number of experiments have been carried out on lead-free piezoelectric materials. Bismuth ferrite BiFeO3 (BFO) was discovered to be one of the most versatile and acceptable materials among all lead-free piezoelectric options, owing to its high Curie temperature (830 °C), rhombohedral perovskite structure at ambient temperature, and a high polarization of approximately 100 μC/cm2 [18,19]. However, the production of impurity phases makes it difficult to achieve a saturated ferroelectric hysteresis loop in pure BFO. Furthermore, the valence changes in Fe (Fe3+ to Fe2+) and volatilization of Bi3+ result in a large electrical leakage current [20,21,22,23]. As a result of these disadvantages, the practical applicability of BFO has been limited.
Many efforts have been made to make solid solutions using different ABO3 perovskite ceramics materials in order to mitigate the shortcomings of BiFeO3. The electrical characteristics of these solid solutions were improved to some extent. BiFeO3-BaTiO3 (BF-BT) is the most appealing and promising of these solid solutions owing to its high operating temperature [21] and relatively better electrical characteristics. However, the electrical characteristics of BF-BT are affected by excessive BFO leakage current [21]. A quenching process to address this issue has been reported in several papers [24,25,26,27]. Although the piezoelectric properties of BF-BT can be increased by the quenching procedure, they are still not enough in practice; therefore, their piezoelectrical features need to be improved. Recently, Liu et al. [28] reported a large piezoelectric sensor coefficient of d33 = 600 pC/N in lead-free Ba(Ti0.8Zr0.2)O3-(Ba0.7Ca0.3)TiO3 (BCZT) materials. However, the working temperature was found to be below 100 °C.
Based on the literature, the structure and corresponding electrical properties of BF-BT-based systems are vulnerable to processing conditions and modifier elements. In the present work, a new BF-based material was systematically investigated. Lead-free (1–z)(0.65Bi1.05Fe2O3-0.35BaTiO3) (BF-BT) was chosen as a base composition and the effect of Ba(Ti0.8Zr0.2)O3-(Ba0.7Ca0.3)TiO3 (BCZT) content on its properties was studied for the sake of improving the ferroelectric performance. Lead-free BCZT-modified BF-BT solid solutions were prepared by a conventional mixed-oxide method followed by air quenching and the effects of BCZT modification on the structural and electromechanical properties were studied in detail. The processing conditions such as calcination, sintering temperature, and soaking time were optimized, all areas yet to be explored for this piezoceramic system. The effect of BCZT modification on the crystal structure, ferroelectric, and piezoelectric properties were studied. The underlying mechanism of the increased ferroelectric characteristics and their stability, were also examined.

2. Experimental Procedure

A solid-state reaction with additional heat treatments was used for the synthesis of piezoelectric ceramics with a composition of (1–z)(0.65Bi1.05Fe2O3-0.35BaTiO3) and zBCZT. where z = 0.00, 0.010, 0.020, and 0.030. For the raw materials, commercially available carbonates, and metal oxides of Bi2O3, Fe2O3, ZrO2, TiO2, CaCO3 and BaCO3 with a purity greater than 99.9% (Sigma Aldrich Co., St. Louis MO) were used as source materials. The stoichiometric formula was employed to accurately quantify these components, and ethanol with zirconia balls was used for ball-milling for 24 h. Drying and calcining twice at 750 °C for 2 h resulted in phase development in the resultant slurry. Finally, the resultant slurry was ball-milled in ethanol for 4 h with zirconia balls. The calcined powder was pressed at 98 MPa, resulting in 10 mm diameter, disk-shaped ceramic specimens. The pressed discs were sintered at 950 °C, 960 °C, 970 °C, 980 °C, and 990 °C and 1020 °C with a soaking time of 2 h in covered alumina crucibles. Sintering took place at 1020 °C for 2 h, after which the pellets were immediately cooled to ambient temperature.
The crystal structure and phase purity were determined using X-ray diffraction (XRD, X’pert MPD3040, Philips, The Netherlands). Scanning electron microscopy (SEM, JP/JSM5200, Japan) was used to assess morphology. The samples were polarized for 15 min at room temperature in a silicone oil bath with a direct-current (DC) field of 5 kV/mm in order to determine the piezoelectric characteristics. A Berlincourt d33 meter (IACAS, ZJ-6B) was employed to determine the piezoelectric constant. The ferroelectric test system was utilized to detect the hysteresis loops of the polarization against the electric field (PE) in silicon oil at a frequency of 10 Hz and ceramics’ loss at different frequencies in the 25–500 °C temperature range.

3. Results and Discussion

The X-ray diffraction (XRD) patterns of the synthesized BCZT-modified BF-BT ceramics sintered at various temperatures are shown in Figure 1. It is well understood that the sintering temperature and soaking time are critical for producing high-density ceramics. To achieve optimized sintering conditions, the undoped BF-BT was sintered at different temperatures of 950 °C, 960 °C, 970 °C, 980 °C, and 990 °C with soaking times of 2 h and 4 h. The sintering temperature range was selected based on previous studies of BF-BT-based compositions [24,25]. The increased volatility of Bi at high temperatures destroyed the base composition. No perovskite structure was obtained at a sintering temperature T < 1000 °C. As a result, the optimized sintering temperature for undoped BF-BT ceramics was determined to be above 1000 °C. Figure 1 shows the effects of sintering temperature on the XRD analysis of undoped BF-BT.
Figure 2 depicts the XRD results at optimized sintering conditions. All the sintered ceramics show a solid solution that is homogeneous and has a pure perovskite structure with no unwanted secondary phases [29]. The optimized sintering temperature for undoped BF-BT ceramics was determined to be 1020 °C with a 2 h soaking period. At these conditions, a single pseudo-cubic structure was obtained, showing that the synthesized materials had a perovskite crystal structure [29,30]. This demonstrates that the thermal treatment was quite efficient in achieving a stable phase structure by suppressing undesirable phases.
Generally, the microstructure of the ceramics determines their electrical properties. Therefore, SEM micrographs were analyzed, as shown in Figure 3, to investigate the microstructure of the BCZT-modified BF-BT sintered ceramics. All ceramics were well-sintered with a close-packed structure with clear grains and grain boundaries. This shows that no melting occurred at these sintering conditions. For the sample z = 0.00, an inhomogeneous microstructure was observed with a mixture of small-sized and large-sized grains. However, densification improved in samples z = 0.010 and 0.020. Here, as can be seen from Figure 3, the sample z = 0.020 showed better densification with void free and relatively homogeneous micrographs as compared to the other sintered samples. A reduction in porosity is an important factor for enhancing ferroelectric or piezoelectric characteristics [31]. The sample z = 0.030 showed an inhomogeneous microstructure with small voids that suggests a decrease in densification in this sample. Overall, grain size decreased from 6.6 μm for sample z = 0.00 to 5.8 μm for sample z = 0.030.
The density of the ceramics is shown in Figure 4 at a sintering temperature of 1020 °C and a soaking time of 2 h. The unmodified ceramic sample had a density of 6.72 g/cm3, while the 2-mol.% modified sample showed an increased density value of 7.25 g/cm3 (~96% densification), which began to drop at high dopant concentrations.
Figure 5 shows the polarization-electric field (PE) hysteresis loops of BT-BF ceramics sintered at various temperatures with a soaking time of 2 h at an applied field of 2.5, 3 and 4 kV/cm, which are indicated by blue, black, and red, respectively. Up to a sintering temperature of 1010 °C, the absence of the ferroelectric PE loop of sintered samples demonstrated the paraelectric phase of the material. As the sintering temperature increased (1020 °C), the ferroelectric property of the material was progressively improved [29,30]. Hence, the optimal sintering conditions were found to be 1020 °C and a soaking time of 2 h for these ceramics.
Well-saturated PE hysteresis loops were observed with no pinching for all compositions sintered at optimized conditions, which confirmed normal ferroelectric behaviour [21,22,23,24,25,26,27]. BCZT incorporation successfully enhanced the ferroelectric characteristics of the base composition (z = 0.00), as indicated by an increase in remnant polarization Pr and a decrease in coercive field Ec, as illustrated in Figure 6. The Pr and Ec values change from ~25 μC/cm2 and ~28 kV/cm for pure ceramics to ~28 μC/cm2 and ~26 kV/cm for z = 0.020, respectively. These results are comparable to previous reports on BF-based systems [10,11,12,13]. Both Pr and Ec dropped as the BCZT content increased. This can be linked to inhomogeneous grain size, a drop in density and the presence of porosity inthe z = 0.030 sample. Also, it can be seen that very small grains were developed between larger grains. It is proposed that inhomogeneous grain size, the reduction in density, the presence of porosity and very small grains between larger grains resulted in the slanted PE loop for this sample.
Internal stresses are generated at the grain boundaries during poling. Domain re-orientation affects the re-orientation of spontaneous strain, which can modify the dimensions of specific grains. Intergranular stresses are higher in ceramics with very small grains present between larger grains due to higher grain boundary density. Due to higher intergranular stresses at these grain boundaries, back fields are exerted that inhibit domain reversal and lower saturation polarization. When the field is removed, the high intergranular stresses force the domains to switch back and lower the Pr. The polarization results are in good agreement with the microstructural results [32].
In contrast to BNT-based systems, all samples in the examined compositional range demonstrated conventional ferroelectric-like behaviour without visible pinching and the corresponding non-ergodic to ergodic transition [22,23,24,25,26,27]. This behaviour may be linked to the differences in domain morphologies and role of defects in the BF and BNT-based systems. Recently, in situ poling synchrotron X-ray diffraction revealed that the pseudo-cubic symmetry preserved during and after the application of electric fields and piezoelectric properties were linked to the presence of multi-symmetry polar nanoregions, which allowed for a high average distortion in the applied field direction [33]. In another study, by using in situ poling synchrotron XRD, the absence of long-range ferroelectric order and the retention of short-range polar order was proposed in BF-based ceramics [34]. However, the exact mechanism in BF-based systems is still unclear and needs further sophisticated studies.
All poled samples were aged for 24 h before the measurement of the piezoelectric sensor coefficient (d33). For each sample, three readings were taken, and the mean was calculated. The piezoelectric sensor coefficient was enhanced from ~50 pC/N for an unmodified sample to ~152 pC/N for a 2 mol.% modified sample, as shown in Figure 7, which is in good agreement with the ferroelectric properties. This value is better than the previously reported values for lead-free piezoelectric ceramics, as shown in Table 1 [35,36,37,38,39,40,41,42,43].
This increase in d33 at z = 0.020 may be ascribed to the incorporation of multi-cationic BZCT, which modified the bond lengths at a unit cell level and gave rise to more flexibility in the complex domain switching. Consequently, the enhancement in the d33 value was observed for very flexible (at a unit cell level) compositions. Despite the fact that no discernible structural change was observed within the XRD detection limit for all specimens, the electrical properties show that the addition of BCZT to the base BF-BT lattice has a significant effect. The observed broadening of the peaks, in contrast to a pure cubic structure, may indicate the presence of some non-cubic distortion or pseudo-cubic phase that is required for ferroelectricity to exist in materials. The variation in the electromechanical properties strongly suggests that the origin of the high piezoelectric property is linked to the crystal structure morphotropic phase boundary.

4. Conclusions

In this work, an air quenching approach and a solid-state reaction method were used to study the synthesis of lead-free BCZT-modified BF-BT piezoelectric ceramics. X-ray diffraction patterns revealed a pure perovskite structure without any secondary phases. An enhanced piezoelectric sensor constant of 152 pC/N was observed with improved remnant polarization Pr ~28 μC/cm2. The combination of grain size effect, densification, and hence improved polarization Pr is thought to be responsible for the enhanced piezoelectric properties in the optimized composition. This study suggests that the ferroelectric properties of the BF-BT system were significantly improved by BCZT incorporation.

Author Contributions

Conceptualization, methodology, formal analysis, investigation, and writing—original draft preparation, R.A.M.; writing—review and editing, R.A.M. and H.A.; supervision, R.A.M. and H.A.; project administration, R.A.M. and H.A. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by the Deanship of Scientific Research at Prince Sattam bin Abdulaziz University, under the research project no. 2020/01/17063.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hao, J.; Li, W.; Zhai, J.; Chen, H. Progress in high-strain perovskite piezoelectric ceramics. Mater. Sci. Eng. R Rep. 2019, 135, 1–57. [Google Scholar] [CrossRef]
  2. Zheng, T.; Wu, J.; Xiao, D.; Zhu, J. Recent development in lead-free perovskite piezoelectric bulk materials. Prog. Mater. Sci. 2018, 98, 552–624. [Google Scholar] [CrossRef]
  3. Malik, R.A.; Hussain, A.; Zaman, A.; Maqbool, A.; Rahman, J.U.; Song, T.K.; Kim, W.J.; Kim, M.H. Structure–property relationship in lead-free A-and B-site co-doped Bi0.5 (Na0.84K0.16)0.5 TiO3–SrTiO3 incipient piezoceramics. RSC Adv. 2015, 5, 96953–96964. [Google Scholar] [CrossRef]
  4. Rödel, J.; Webber, K.G.; Dittmer, R.; Jo, W.; Kimura, M.; Damjanovic, D. Transferring lead-free piezoelectric ceramics into application. J. Eur. Ceram. Soc. 2015, 35, 1659–1681. [Google Scholar] [CrossRef]
  5. Zeeshan; Panigrahi, B.K.; Ahmed, R.; Mehmood, M.U.; Park, J.C.; Kim, Y.; Chun, W. Operation of a low-temperature differential heat engine for power generation via hybrid nanogenerators. Appl. Energy 2021, 285, 116385. [Google Scholar] [CrossRef]
  6. Hajra, S.; Tripathy, A.; Panigrahi, B.K.; Choudhary, R.N.P. Development and excitation performance of lead-free electronic material: Eu and Fe doped Bi0.5Na0.5TiO3 for filter application. Mater. Res. Express 2019, 6, 076304. [Google Scholar] [CrossRef]
  7. Hajra, S.; Sahoo, S.; Mishra, T.; De, M.; Rout, P.K.; Choudhary, R.N.P. Studies of structural, dielectric and electrical characteristics of BaTiO3–BiFeO3–CaSnO3 electronic system. J. Mater. Sci. Mater. Electron. 2018, 29, 7876–7884. [Google Scholar] [CrossRef]
  8. Briscoe, J.; Dunn, S. Piezoelectric nanogenerators—A review of nano structured piezoelectric energy harvesters. Nano Energy 2015, 14, 15–29. [Google Scholar] [CrossRef]
  9. Hanani, Z.; Izanzar, I.; Amjoud, M.; Mezzane, D.; Lahcini, M.; Ursic, H.; Prah, U.; Saadoune, I.; Marssi, M.E.; Lukyanchuk, I.A.; et al. Lead-free nanocomposite piezoelectric nanogenerator film for biomechanical energy harvesting. Nano Energy 2021, 81, 105661. [Google Scholar] [CrossRef]
  10. Mistewicz, K. Recent advances in ferroelectric nanosensors: Toward sensitive detection of gas, mechanothermal signals, and radiation. J. Nanomater. 2018, 2018, 2651056. [Google Scholar] [CrossRef]
  11. Lee, H.S.; Chung, J.; Hwang, G.T.; Jeong, C.K.; Jung, Y.; Kwak, J.H.; Kang, H.; Byun, M.; Kim, W.D.; Hur, S.; et al. Flexible inorganic piezoelectric acoustic nanosensors for biomimetic artificial hair cells. Adv. Funct. Mater. 2014, 24, 6914–6921. [Google Scholar] [CrossRef]
  12. Qian, W.; Yang, W.; Zhang, Y.; Bowen, C.R.; Yang, Y. Piezoelectric materials for controlling electro-chemical processes. Nano-Micro Lett. 2020, 12, 149. [Google Scholar] [CrossRef]
  13. Mistewicz, K.; Kepinska, M.; Nowak, M.; Sasiela, A.; Zubko, M.; Stróz, D. Fast and efficient piezo/photocatalytic removal of methyl orange using SbSI nanowires. Materials 2020, 13, 4803. [Google Scholar] [CrossRef]
  14. Adnan, M.; Hussain, A.; Rahman, J.U.; Malik, R.A.; Song, T.K.; Kim, M.-H.; Kim, W.-J. Composition-dependent structural, dielectric and ferroelectric responses of lead-free Bi0.5Na0.5TiO3-SrZrO3 ceramics. J. Korean Phys. Soc. 2016, 68, 1430–1438. [Google Scholar]
  15. Rahman, J.U.; Hussain, A.; Adnan, M.; Malik, R.A.; Song, T.K.; Kim, M.-H.; Lee, S.; Kim, W.-J. Effect of donor doping on the ferroelectric and the piezoelectric properties of lead-free 0.97(Bi0.5Na0.5Ti1−xNbx)O3-0.03BaZrO3 ceramics. J. Korean Phys. Soc. 2015, 67, 1240–1245. [Google Scholar] [CrossRef]
  16. Rahman, J.U.; Hussain, A.; Adnan, M.; Malik, R.A.; Kim, M.S.; Kim, M.-H. Effect of sintering temperature on the electromechanical properties of 0.945Bi0.5Na0.5TiO3-0.055BaZrO3 ceramics. J. Korean Phys. Soc. 2015, 66, 1072–1076. [Google Scholar] [CrossRef]
  17. Cheon, C.I.; Choi, J.H.; Kim, J.S.; Zang, J.; Fromling, T.; Rodel, J.; Jo, W. Role of (Bi1/2K1/2) TiO3 in the dielectric relaxations of BiFeO3-(Bi1/2K1/2) TiO3 ceramics. J. Appl. Phys. 2016, 119, 15410. [Google Scholar] [CrossRef]
  18. Ryu, G.H.; Hussain, A.; Lee, M.H.; Malik, R.A.; Song, T.K.; Kim, W.J.; Kim, M.H. Lead-free high performance Bi (Zn0.5Ti0.5) O3-modified BiFeO3-BaTiO3 piezoceramics. J. Eur. Ceram. Soc. 2018, 38, 4414–4421. [Google Scholar] [CrossRef]
  19. Lee, M.H.; Kim, D.J.; Park, J.S.; Kim, S.W.; Song, T.K.; Kim, M.H.; Kim, W.J.; Do, D.; Jeong, I.K. High-performance lead-free piezoceramics with high curie temperatures. Adv. Mater. 2015, 27, 6976–6982. [Google Scholar] [CrossRef] [PubMed]
  20. Rojac, T.; Bencan, A.; Malic, B.; Tutuncu, G.; Jones, J.L.; Daniels, J.E.; Damjanovic, D. BiFeO3 ceramics: Processing, electrical, and electromechanical properties. J. Am. Ceram. Soc. 2014, 97, 1993–2011. [Google Scholar] [CrossRef]
  21. Khan, S.A.; Ahmed, T.; Akram, F.; Bae, J.; Choi, S.Y.; Thanh, T.T.; Kim, M.; Song, T.K.; Sung, Y.S.; Kim, M.H.; et al. Effect of sintering temperature on the electrical properties of pristine BF-35BT piezoelectric ceramics. J. Korean Ceram. Soc. 2020, 57, 290–295. [Google Scholar] [CrossRef]
  22. Lee, M.H.; Kim, D.J.; Choi, H.I.; Kim, M.-H.; Song, T.K.; Kim, W.-J.; Park, J.S.; Do, D. Low sintering temperature for lead-free BiFeO3-BaTiO3 ceramics with high piezoelectric performance. J. Am. Ceram. Soc. 2019, 102, 2666–2674. [Google Scholar]
  23. Lee, M.H.; Kim, D.J.; Choi, H.I.; Kim, M.-H.; Song, T.K.; Kim, W.-J.; Do, D. Thermal quenching effects on the ferroelectric and piezoelectric properties of BiFeO3–BaTiO3 ceramics. ACS Appl. Electron. Mater. 2019, 1, 1772–1780. [Google Scholar] [CrossRef]
  24. Akram, F.; Malik, R.A.; Khan, S.A.; Hussain, A.; Lee, S.; Lee, M.-H.; In, C.H.; Song, T.-K.; Kim, W.-J.; Sung, Y.S.; et al. Electromechanical properties of ternary BiFeO3−0.35BaTiO3–BiGaO3 piezoelectric ceramics. J. Electroceram. 2018, 41, 93–98. [Google Scholar] [CrossRef]
  25. Akram, F.; Hussain, A.; Malik, R.A.; Song, T.-K.; Kim, W.-J.; Kim, M.-H. Synthesis and electromechanical properties of LiTaO3-modified BiFeO3–BaTiO3 piezoceramics. Ceram. Int. 2017, 43, S209–S213. [Google Scholar] [CrossRef]
  26. Habib, M.; Lee, M.H.; Akram, F.; Kim, M.-H.; Kim, W.-J.; Song, T.K. Temperature-insensitive piezoelectric properties of lead-free BiFeO3–BaTiO3 ceramics with high Curie temperature. J. Alloys Compd. 2021, 851, 156788. [Google Scholar] [CrossRef]
  27. Habib, M.; Lee, M.H.; Choi, H.I.; Kim, M.-H.; Kim, W.-J.; Song, T.K. Phase evolution and origin of the high piezoelectric properties in lead-free BiFeO3–BaTiO3 ceramics. Ceram. Int. 2020, 46, 22239–22252. [Google Scholar] [CrossRef]
  28. Liu, W.; Ren, X. Large piezoelectric effect in Pb-free ceramics. Phys. Rev. Lett. 2009, 103, 257602. [Google Scholar] [CrossRef] [Green Version]
  29. Alzaid, M.; Alsalh, F.; Malik, R.A.; Maqbool, A.; Almoisheer, N.; Hadia, N.M.A.; Mohamed, W.S. LiTaO3 assisted giant strain and thermally stable energy storage response for renewable energy storage applications. Ceram. Int. 2021, 47, 15710–15721. [Google Scholar] [CrossRef]
  30. Malik, R.A.; Zaman, A.; Hussain, A.; Maqbool, A.; Song, T.K.; Kim, W.-J.; Sung, Y.S.; Kim, M.-H. Temperature invariant high dielectric properties over the range 200 °C–500 °C in BiFeO3 based ceramics. J. Eur. Ceram. Soc. 2018, 38, 2259–2263. [Google Scholar] [CrossRef]
  31. Fernandez-Benavides, D.A.; Gutierrez-Perez, A.I.; Benitez-Castro, A.M.; Ayala-Ayala, M.T.; Moreno-Murguia, B.; Muñoz-Saldaña, J. Comparative study of ferroelectric and piezoelectric properties of BNT-BKT-BT ceramics near the phase transition zone. Materials 2018, 11, 361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. McKinnon, R.A. Grain Size Effect in Lead-Free Bi0.5Na0.5TiO3-Based Materials: Exploring the Ferroelectric Behaviour. 2015. Available online: https://qmro.qmul.ac.uk/xmlui/handle/123456789/12899 (accessed on 24 August 2021).
  33. Wang, G.; Fan, Z.; Murakami, S.; Lu, Z.; Hall, D.A.; Sinclair, D.C.; Feteira, A.; Tan, X.; Jones, J.L.; Kleppe, A.K.; et al. Origin of the large electrostrain in BiFeO3-BaTiO3 based lead-free ceramics. J. Mater. Chem. A 2019, 7, 21254–21263. [Google Scholar] [CrossRef]
  34. Lu, Z.; Wang, G.; Li, L.; Huang, Y.; Feteira, A.; Bao, W.; Kleppe, A.K.; Xu, F.; Wang, D.; Reaney, I.M. In situ poling X-ray diffraction studies of lead-free BiFeO3-SrTiO3 ceramics. Mater. Today Phys. 2021, 19, 100426. [Google Scholar] [CrossRef]
  35. Lin, D.; Zheng, Q.; Li, Y.; Wan, Y.; Li, Q.; Zhou, W. Microstructure, ferroelectric, and piezoelectric properties of Bi0.5K0.5TiO3-modified BiFeO3–BaTiO3 lead-free ceramics with high Curie temperature. J. Eur. Ceram. Soc. 2013, 33, 3023–3036. [Google Scholar] [CrossRef]
  36. Khan, S.A.; Malik, R.A.; Akram, F.; Hussain, A.; Song, T.-K.; Kim, W.-J.; Kim, M.-H. Synthesis, and electrical properties of 0.65Bi1.05Fe1-xGaxO3–0.35BaTiO3 piezoceramics by air quenching process. J. Electroceram. 2018, 41, 60–66. [Google Scholar] [CrossRef]
  37. Akram, F.; Malik, R.A.; Lee, S.; Pasha, R.A.; Kim, M.H. Enhanced piezoelectric properties of (1 − x) [0.675BiFeO3- 0.325BaTiO3]-xLiTaO3 ternary system by air quenching. Korean J. Mater. Res. 2018, 28, 489–494. [Google Scholar] [CrossRef]
  38. Wang, D.; Fan, Z.; Zhou, D.; Khesro, A.; Murakami, S.; Feteira, A.; Zhao, Q.; Tan, X.; Reaney, I.M. Bismuth ferrite-based leadfree ceramics and multilayers with high recoverable energy density. J. Mater. Chem. A 2018, 6, 4133. [Google Scholar] [CrossRef] [Green Version]
  39. Luo, L.; Jiang, N.; Zou, X.; Shi, D.; Sun, T.; Zheng, Q.; Xu, C.; Lam, K.H.; Lin, D. Phase transition, piezoelectric, and multiferroic properties of La (Co0.5Mn0.5) O3-modified BiFeO3-BaTiO3 lead-free ceramics. Phys. Status Solidi A 2015, 212, 2012–2022. [Google Scholar] [CrossRef]
  40. Malik, R.A.; Hussain, A.; Song, T.K.; Kim, W.J.; Ahmed, R.; Sung, Y.S.; Kim, M.H. Enhanced electromechanical properties of (1-x) BiFeO3–BaTiO3–xLiNbO3 ceramics by quenching process. Ceram. Int. 2017, 43, S198–S203. [Google Scholar] [CrossRef]
  41. Liu, Z.; Zheng, T.; Zhao, C.; Wu, J. Composition design and electrical properties in BiFeO3–BaTiO3–Bi (Zn0.5Ti0.5) O3 leadfree ceramics. J. Mater. Sci. Mater. Electron. 2017, 28, 13076. [Google Scholar] [CrossRef]
  42. Leontsev, S.O.; Eitel, R.E. Dielectric and piezoelectric properties in Mn-modified (1 − x) BiFeO3-xBaTiO3 ceramics. J. Am. Ceram. Soc. 2009, 92, 2957. [Google Scholar] [CrossRef]
  43. Khan, S.A.; Akram, F.; Malik, R.A.; Hussain, A.; Kim, J.C.; Song, T.K.; Kim, W.-J.; Sung, Y.S.; Kim, M.-H.; Lee, S. Effects of cooling rate on the electrical properties of Pb-free BF-BT ceramics. Ferroelectrics 2019, 553, 76–82. [Google Scholar] [CrossRef]
Figure 1. X–ray diffraction patterns for BFBT–zBCZT (z = 0.00) ceramics in the 2θ ranges of 20–60° sintered at different temperatures.
Figure 1. X–ray diffraction patterns for BFBT–zBCZT (z = 0.00) ceramics in the 2θ ranges of 20–60° sintered at different temperatures.
Crystals 11 01077 g001
Figure 2. X–ray diffraction patterns for BFBT–zBCZT (z = 0.00–0.030) ceramics in the 2θ ranges of 30–60° sintered at optimized conditions.
Figure 2. X–ray diffraction patterns for BFBT–zBCZT (z = 0.00–0.030) ceramics in the 2θ ranges of 30–60° sintered at optimized conditions.
Crystals 11 01077 g002
Figure 3. SEM micrographs of BFBT–zBCZT (z = 0.00–0.030) samples.
Figure 3. SEM micrographs of BFBT–zBCZT (z = 0.00–0.030) samples.
Crystals 11 01077 g003
Figure 4. Density measurement of BFBT–zBCZT (z = 0.00–0.030) samples.
Figure 4. Density measurement of BFBT–zBCZT (z = 0.00–0.030) samples.
Crystals 11 01077 g004
Figure 5. Ferroelectric properties of BFBT–zBCZT (z = 0.00) sample sintered at 980 °C, 990 °C, 1010 °C and 1020 °C.
Figure 5. Ferroelectric properties of BFBT–zBCZT (z = 0.00) sample sintered at 980 °C, 990 °C, 1010 °C and 1020 °C.
Crystals 11 01077 g005
Figure 6. Ferroelectric properties of BFBT–zBCZT (z = 0.00–0.030) samples at optimized conditions.
Figure 6. Ferroelectric properties of BFBT–zBCZT (z = 0.00–0.030) samples at optimized conditions.
Crystals 11 01077 g006
Figure 7. Variation in room temperature piezoelectric constant (d33) of BFBT–zBCZT (z = 0.00–0.030) ceramics.
Figure 7. Variation in room temperature piezoelectric constant (d33) of BFBT–zBCZT (z = 0.00–0.030) ceramics.
Crystals 11 01077 g007
Table 1. Comparison of the piezoelectric constant of various BF-BT ceramics.
Table 1. Comparison of the piezoelectric constant of various BF-BT ceramics.
Materialsd33
(pC/N)
YearReferences
BiFeO3–BaTiO3–Bi0.5K0.5TiO31352013[35]
0.65BFGa–0.35BT1452018[36]
0.675BF-0.325BT-xLT1452018[37]
0.75B0.975Nd0.025F–0.25BT+Mn1402018[38]
0.73BF–0.25BT–0.02LCM + Mn1082015[39]
0.99(0.67BF–0.33BT)–0.01LN1462017[40]
0.60BF–0.40BT–0.02BZT502017[41]
0.75BF–0.25BT472009[42]
0.65Bi1.05Fe1−xGaxO3–0.35BaTiO31402019[43]
BF–BT–BCZT1522021Current Work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Malik, R.A.; Alrobei, H. Processing and Characterization of BCZT-Modified BiFeO3-BaTiO3 Piezoelectric Ceramics. Crystals 2021, 11, 1077. https://doi.org/10.3390/cryst11091077

AMA Style

Malik RA, Alrobei H. Processing and Characterization of BCZT-Modified BiFeO3-BaTiO3 Piezoelectric Ceramics. Crystals. 2021; 11(9):1077. https://doi.org/10.3390/cryst11091077

Chicago/Turabian Style

Malik, Rizwan Ahmed, and Hussein Alrobei. 2021. "Processing and Characterization of BCZT-Modified BiFeO3-BaTiO3 Piezoelectric Ceramics" Crystals 11, no. 9: 1077. https://doi.org/10.3390/cryst11091077

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