Next Article in Journal
Change in Growth Mode of BGaN Layers Grown on GaN
Previous Article in Journal
The Wannier-Mott Exciton, Bound Exciton, and Optical Phonon Replicas of Single-Crystal GaSe
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multifunctional Experimental Studies of Sm-Ion-Influenced Pseudo-Cubic Morphotropic Phase Boundary Regional BiFeO3-xSrTiO3 Ceramics for High-Temperature Applications

by
Ahmad Hussain
1,*,
Nawishta Jabeen
2,*,
Aasma Tabassum
3,
Muhammad Usman Khan
4,
Laiba Basharat
1 and
Islam H. El Azab
5
1
Department of Physics, The University of Lahore, Sargodha Campus, Sargodha 40100, Pakistan
2
Department of Physics, Fatima Jinnah Women University Rawalpindi, Rawalpindi 46000, Pakistan
3
Herbert Gleiter Institute of Nanoscience, School of Material Science and Engineering, Nanjing University of Science and Technology, 200 Xiaolingwei Street, Nanjing 210094, China
4
National Key Laboratory of Tunable Laser Technology, Institute of Optoelectronics, Department of Electronics Science and Technology, Harbin Institute of Technology, Harbin 150080, China
5
Department of Food Science and Nutrition, College of Science, Taif University, Taif 21944, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(6), 540; https://doi.org/10.3390/cryst14060540
Submission received: 22 May 2024 / Revised: 4 June 2024 / Accepted: 7 June 2024 / Published: 9 June 2024
(This article belongs to the Section Polycrystalline Ceramics)

Abstract

:
In this manuscript, for the first time, the exploration of the microstructural, ferroelectric, piezoelectric, and dielectric performances are measured for Sm-ion-influenced pseudo-cubic, morphotropic phase boundary (MPB) regional 0.62BiFeO3−0.38SrTiO3:xwt%Sm2O3 (BFST:xSm) ceramics with x = 0–0.25. All the compositions maintained their pseudo-cubic MPB structural stability. The composition of BFST:0.15Sm ceramics exhibited an excellent remnant polarization (Pr) of ~52.11 μC/cm2, an enhanced d33 of 101 pC/N, and the highest relative dielectric constant (ɛr) of ~1152, which are much improved as compared to that of pure BFST ceramics. BFST:0.15Sm ceramics demonstrated a Curie temperature (TC) of 378 °C. Moreover, the composition exhibited high thermal stability for d33 72 pC/N (only a 28% decrease), even at a high temperature of 300 °C. Such outstanding outcomes make BFST:0.15Sm ceramics an ideal applicant for high-temperature piezoelectric applications.

1. Introduction

Recently, perovskite-based ceramics have gained significant importance for high-temperature piezoelectric applications due to their tremendous thermal stability and high piezoelectric coefficient (d33). They are considered ideal for offering enhanced durability in harsh and fatigued conditions for aerospace, energy sector, and automotive applications. Perovskite-based materials are renowned due to their unique structures and versatile properties; the novel crystalline structural arrangement of such materials makes them ideal to dope/add or substitute suitable elements, resulting in a wide range of amplified tunable properties [1,2,3,4]. Researchers have paid attention to analyzing the thermal stability of their piezoelectric performances in the high-temperature range and utilizing them for applications in harsh or fatigued conditions, like high-temperature piezoelectric devices, industrial processes, automotive applications, aerospace applications, and energy harvesting systems [5,6,7]. Scientists are focused on exploring and stimulating the properties of lead-free dielectric ceramics due to the rising health concerns associated with lead-based ceramics [8,9].
Among perovskite-based ceramics, BiFeO3 (BF) has special importance due to its ability to demonstrate electro-magnetic effects simultaneously. It possesses a rhombohedral structure with space group R3c. BF shows the ferroelectric to paraelectric transition near its Curie temperature (TC) of ~820 °C [10]. The reported d33 of pure ceramic BF is ~40 pC/N [11]. However, the voltaic nature of Bi-ions at high temperatures always makes it a challenging task for researchers to fabricate single-phase BF ceramics without the appearance of a secondary phase [12]. To overcome the challenge of low d33, BF is mostly doped with its A- or B-site elements, with the suitable elements keeping the valance and ionic radii in consideration for d33 improvement. Another significant effort made to utilize the best merit of BF (its high TC) is to combine one of its ends with another different phased structure perovskite to explore the MPB regions and amplify the ferroelectric, dielectric, and piezoelectric performances, like BaTiO3 [13], Bi0.5K0.5TiO3 [14], Bi0.5Na0.5TiO3 [15], SrTiO3 [16], etc. Among such counterparts of BF, SrTiO3 (ST) has rarely been explored by scientists; it possesses a cubic structure with the space group Pm3m [17]. ST is regarded among the linear dielectric perovskites, which exhibit unique properties like a moderate relative dielectric constant (ɛr) and low dielectric loss (tanδ) [18]. ST is reported to be an incipient ferroelectric material that acts paraelectric at 0K, but there is the possibility of ferroelectricity by modifying the A- or B-sites with some chemical substitution or strain [19]. From a literature survey, it is observed that very limited research has been performed on BiFeO3-SrTiO3 ceramics at their MPB vicinity. Makarovic et al. have reported a d33 of 69 pC/N for BF–0.375ST ceramics and a Pr of ~30–50 μC/cm2 [20]. Recently, Wang et al. fabricated BF–0.38ST ceramics to exploit the MPB region and reported a Pr of 51.2 μC/cm2 and a d33 of 72 pC/N [21], which is the maximum reported to date according to the best of our knowledge.
Rare-earth (RE)-ion doping/addition is found to improve the dielectric permittivity and limit the dielectric loss. RE-ions are considered to be good substitutes at the A-site cations of perovskite materials due to their suitable ionic radii. Such comparable ionic radii of RE-ions as dopants/additives can accommodate lattice sites well, as it involves little formation energy to simply inhabit the A-site [22,23,24]. Samarium (Sm) is found to be a soft metal that possesses a silvery white color, and it gradually oxidizes to Sm2O3 in open air [25]. Hence, the introduction of Sm2O3 as an additive in the pseudo-cubic MPB regional BFST host matrix will possibly reduce the defected concentrations (occurring due to the voltaic nature of Bi) and increase the density of ceramics, which will result in diffusion control at the Bi-lattice sites during high-temperature sintering and will avoid undesirable phase transitions. The addition of Sm-ions will also lead to well-organized grain boundaries, resulting in the variation of dipole moments and strong ferroelectric and piezoelectric properties.
Herein, 0.62BiFeO3–0.38SrTiO3:xwt%Sm2O3 (BFST:xSm) with x = 0–0.25 ceramic series was synthesized to investigate its microstructural, ferroelectric, piezoelectric, and dielectric performances. BFST:0.15Sm’s ceramic composition demonstrated outstanding multifunctional properties with a Pr of ~52.11 μC/cm2, a d33 of 101 pC/N, and an ɛr of ~1152, which are better than those of pure BFST ceramics. Moreover, its composition demonstrated a thermally stable piezoelectric response of 72 pC/N at 300 °C, which makes it a potential candidate for high-temperature sensors or actuators.

2. Fabrication and Characterization

The conventional solid-state reaction technique was followed for the fabrication of 0.62BiFeO3–0.38SrTiO3:xwt%Sm2O3 (BFST:xSm) with x = 0–0.25 ceramic series. High-purity (~99.9%) Sigma-Aldrich chemical Bi2O3, Fe2O3, SrO, and TiO2 powders were weighted with BFST stoichiometric ratios, mixed in ethanol, and milled for 15 h. The mixed powder was dried at 150 °C overnight and then calcinated at 850 °C for 2 h with 20 °C/min increase/decreases in temperature. Calcinated powder was ground, and Sm2O3 (99.9%, Sigma-Aldrich, St. Louis, MO, USA) was added to the ground BFST powder with the stoichiometric ratios of xwt%Sm2O3 (x = 0–0.25), milled in ethanol for 12 h, and then dried at 150 °C overnight. The dried powder was pressed into disc-shaped ceramics of diameter 11 mm under 150 MPa hydraulic pressure and sintered at 1050 °C for 3 h with a rapid increase/decrease in sintering temperature (20 °C/min), and the extra content of compositional powders was spread around the ceramic discs to avoid the undesired diffusion of Bi-ions.
Structural stability of ceramic series, phase purity, and detection of any secondary phases were characterized by X-ray Diffraction (XRD, PANalytical, Almelo, The Netherlands, 40 kV) Cu Ka 1:l = 1.54056 Å and step size: 0.02°. The Rietveld method was used in Material Studio for the simulation of the achieved XRD patterns. Morphological analysis and grain size variations were analyzed using a scanning electron microscope (FE-SEM, FEI Quanta 200, Hillsboro, OR, USA). For the ferroelectric, piezoelectric, and dielectric measurements, ceramics were polished to a thickness of 0.7 mm, and Ag electrodes of specific diameters were pasted on both faces of the ceramic samples. Polarization versus electric field (P-E) analysis was performed for the ferroelectric measurements by a (aix-ACCT TF Analyzer 2000; Germany) ferroelectric analyzer at room temperature and 1 Hz frequency. BFST:xSm ceramics were poled in a silicon oil bath, applying a 100 kV/cm DC electric field, and at 60 Hz, the piezoelectric d33 coefficient was measured by (IAAS ZJ-30, Institute of Acoustics of CAS, Beijing, China) piezo-d33 m. The thermal piezoelectric stability of the ceramic series was measured by annealing the poled ceramics in a temperature range of 50–400 °C with a step size of 50 °C for 30 min, and then d33 was measured at room temperature. The dielectric properties of the fully poled ceramic series, like dielectric permittivity versus frequency (ɛr-f) and dielectric loss versus frequency (tanδ-f), were measured for the frequency range 102–107 Hz by an HP4284A precision impedance analyzer (Hewlett-Packard, Palo Alto, CA, USA). For the measurement of the Curie temperature of the ceramic samples, dielectric permittivity versus temperature (ɛr-T) and dielectric loss versus temperature (tanδ-T) were measured by an LCR analyzer (HP4980A, Agilent, Santa Clara, CA, USA) attached to a programmable furnace and computer.

3. Results and Discussion

XRD patterns of BFST:xSm with x = 0–0.25 ceramics were measured at room temperature and are presented in Figure 1a. All the samples maintained their MPB-existent pseudo-cubic perovskite structures; no phase impurities or secondary phases were detected in the patterns. For confirmation of the dual structural rhombohedral R from BiFeO3 and cubic C from the SrTiO3 phases, an amplified XRD image from 37 to 48° was taken and is presented in Figure 1b, where the duality of structures is confirmed by the dual (111) and (200) peaks. Herein, it can be clearly described that the additive ions of xwt%Sm2O3 have incorporated at the lattice sites of the BFST host successfully without damaging the structure of the material. All the visible peaks matched well with the PDF# 71–2494 for BiFeO3 and PDF# 89–4934 for SrTiO3. For further confirmation of the structural stability, XRD patterns were further simulated by the Rietveld method by using Material Studio, keeping the rhombohedral R3c and cubic Pm3m as reference structures (Figure 2a–d). The RWP, RP, and χ were the simulation parameters, which remained less than 10%, confirming the consistency of the refined data. The refinement outcomes disclose the co-existence of dual phases (R and C) for all the fabricated compositions, demonstrating the strong existence of an MPB region, which is not disturbed by the addition of Sm-ions even at varying contents.
SEM micrographs of BFST:xSm with x = 0–0.2 ceramics are presented in Figure 3a–d. No holes or cracks were observed in all the images, confirming the compactness and high density of the ceramic compositions. The grain size dispersion and average grain size of the ceramics were calculated using ImageJ software (version 1.53v). The average grain size of the compositions varied from 4.35 μm to 5.2 μm (Insets Figure 3) for pure BFST ceramics to BFST:0.15Sm ceramics by increasing the Sm content, and then a slight decrease occurred for the BFST:0.20Sm ceramics to 4.86 μm for BFST:0.20Sm ceramics. The Sm-ions can have a strong influence on the grain size of the host BFST; it only depends on how evenly the Sm-ions are dispersed in the host matrix. A low content of Sm-ions promotes grain growth due to the pinning effect and helps to establish the uniform structure. But, as the Sm-ion content increases to a specific ratio (percolation threshold), the ions can form a cluster or form other phases, reducing the grain size at a higher sintering temperature.
Ferroelectric measurements of the ceramics were performed to investigate the ability of the materials in several applications, like sensors, actuators, capacitors, energy harvesting, and memory devices. The P-E loops of pseudo-cubic structural MPB regional BFST:xSm with x = 0–0.20 ceramics were measured at room temperature, 1 Hz frequency, and 120 kV/cm electric field. Well-saturated and open P-E loops of the ceramics were achieved and are presented in Figure 4a. All the compositions demonstrated typical ferroelectric behavior without signs of any leakage current. Maximum remnant polarization (Pr~52.11 μC/cm2) and saturated polarization (PS~58.4 μC/cm2) were observed for the BFST:0.15Sm ceramic composition (Figure 4b), which are much improved over those of pure BFST ceramic (Pr~42 μC/cm2 and PS~44.55 μC/cm2). From Figure 4b, it can be observed that for the BFST:0.20Sm ceramic composition, the polarization values (Pr~47.1 μC/cm2 and PS~55.6 μC/cm2) have reduced; this reduction in the polarization values can be associated with the percolation threshold value [26,27]. Keeping the Sm-ion content below percolation threshold values, the Sm-ions incorporate the BFST host ceramics and adjust at the vacant lattice sites successfully, causing an improvement in the ferroelectric behavior till BFST:0.15Sm ceramics. Afterwards, by performing complete diffusion at the lattice sites, Sm-ions start to accumulate at the grain boundaries, which increases the conduction of the ceramics and results in the production of a leakage current to reduce the ferroelectric behavior. From Figure 4c, the corrosive electric field (EC) as a function of x values of Sm-ions in BFST ceramics is presented, which shows the reduction in the values of EC with the variation in the Sm-ion content, confirming that fabricated ceramics are well oriented and Sm-ions helped the electric domains saturate easily at lower electric fields to make the ceramic compositions completely saturated. From the ferroelectric analysis, it can be suggested that BFST:0.15Sm ceramics have the best composition and are capable of being utilized in capacitors, energy harvesting, and memory devices.
The piezoelectric coefficient (d33) as a function of the x values of Sm-ions in the BFST host matrix is presented in Figure 5a. Here, it can be analyzed that the BFST:0.15Sm ceramic composition demonstrates a maximum d33 of 101 pC/N, which is 22% higher than the pure BFST ceramic, consistent with the ferroelectric analysis. Exceptional ferroelectric and piezoelectric performances can be associated with the dual-structure MPB regional composition of the R and C phases, which is consistent for all the compositions and facilitates domain switching and polarization rotation because of an applied electric field [28]. The reduction in the d33 value (95 pC/N) for the BFST:0.20Sm ceramics is attributed to the percolation threshold limit, where the leakage current plays an important role in reducing the piezoelectric properties [29]. For the practical utilization of BFST:0.15Sm ceramics, the thermal stability of the piezoelectric performances for all the BFST:xSm (x = 0–0.2) ceramics is measured by annealing the ceramics for 50–400 °C with a step size of 50 °C for 30 min each, followed by the measurement of d33 at room temperature. A similar kind of trend is observed for all the compositions; the d33 value of BFST:0.15Sm ceramics is reduced to 72 pC/N at 300 °C, which is just a 28% reduction from the initial value but still strong enough to operate, which confirms the piezoelectric thermal stability of the ceramics at high temperatures. The thermal stability of piezoelectric materials is no doubt an important parameter in their performance and is strongly associated with the diffusion temperature (Td). Td marks the specific value of temperature where ferroelectric properties start the transition towards paraelectric properties and the reduction of piezoelectric properties. High temperatures cause variations in the thermal vibration within the crystal lattice, resulting in the disturbance of dipoles’ alignments and causing the diminution of piezoelectric performances. The specific explanation in terms of a schematic diagram is presented in Supplementary Figure S1. This analysis confirms the ability of the ceramic composition for high-temperature piezoelectric applications. Moreover, for the utilization of such ceramics in frequency-dependent sensors, the dielectric permittivity as a function of the frequency range of 102 Hz to 107 Hz of poled BFST:xSm (x = 0–0.2) ceramics was measured (Figure 5c). The maximum dielectric constant (ɛr of ~1152) was observed for the BFST:0.15Sm ceramic composition, which is much better than all other compositions, consistent with the ferroelectric and piezoelectric properties of the BFST:0.15Sm ceramics. Dielectric anomalies were observed for all the ceramic compositions at the frequency ~2.1 × 105 Hz. Poled ceramics demonstrate deformation at an applied 0.5 V AC voltage because of the piezoelectric effect present in them. Such deformation appears in the form of propagation and resonance, causing the corresponding dielectric anomalies at a specific frequency [30]. The frequency range where such dielectric anomalies appear due to the planar or thickness piezoelectric effect is dependent on the diameter and thickness of the ceramic discs and acoustic velocity. The dielectric loss values of the ceramics define the compactness and density of the fabricated ceramics. Herein, the dielectric losses as a function of frequency plots are presented in Figure 5d, where BFST:0.15Sm ceramics exhibit the lowest dielectric loss, confirming the limited leakage current, high density, excellent ferroelectric properties, outstanding piezoelectric properties, and better dielectric outcomes.
For the measurement of the Curie temperature, the dielectric permittivity as a function of temperature plots of the BFST:xSm (x = 0–0.20) ceramics measured at 1 kHz is presented in Figure 6a, demonstrating the relaxor ferroelectric behavior. As the content of xwt%Sm2O3 increases, the peak shape of the dielectric peak gradually becomes sharp (Figure 6a), meaning that the change in Sm2O3 content regulates the dispersion characteristic of the BFST:xSm system. In general, as the additive content increases, the long-range ordered state of the ceramics is disrupted to a greater extent, which causes the system to become more relaxed (expressed by green arrow in Figure 6a) [31,32,33]. The TC values of BFST, BFST:0.1Sm, BFST:0.15Sm, and BFST:0.20Sm ceramics are 395 °C, 388 °C, 378 °C, and 367 °C, respectively. However, the dielectric permittivity (ɛr) values improve until the BFST:0.15Sm ceramic composition (Figure 6b), consistent with the dielectric permittivity versus frequency plots. The values of dielectric permittivity versus the temperature of the frequency plots for all the compositions showed the same trend at the frequency of 1 kHz and room temperature (Supplementary Figure S2). Moreover, thermally stable behavior of dielectric loss as a function of temperature is observed (Figure 6c); even at the high temperature of 350 °C, the lowest dielectric loss of BFST:0.15Sm ceramic is observed (Figure 6d). The thermally lowest dielectric loss and the high ferroelectric, good piezoelectric, and significant dielectric permittivity performances of the BFST:0.15Sm ceramics confirm the potential of the material for energy harvesting and memory-based and high-temperature piezoelectric devices.
A detailed comparison of our article’s outcomes with the literature survey is presented in Table 1. It is evident from the analysis that the BFST:0.15Sm ceramics in this study achieved the highest d33 value as compared to other ST-based ceramics. But it is still lower than that of the BiFeO3–xBaTiO3-based ceramics. Efforts are continuing towards the improvement, but in this manuscript the BFST:0.15Sm ceramics demonstrated a strong ability to be utilized in high-temperature piezoelectric applications.

4. Conclusions

A successful fabrication of the pseudo-cubic structural MPB regional BFST:xSm (x = 0–0.25) ceramic series was fabricated by the conventional solid-sate reaction method. Among different compositions, BFST:0.15Sm ceramics demonstrated the best merits, like a PS of ~58.4 μC/cm2, a Pr of ~52.11 μC/cm2, a d33 of 101 pC/N, and an ɛr of ~1152, making the composition suitable for many applications, like sensors, actuators, and memory-based devices. The BFST:0.15Sm ceramics exhibited a Curie temperature of 378 °C. Moreover, the thermally stable piezoelectric performance of the BFST:0.15Sm ceramics was observed. The composition showed a high d33 of 72 pC/N even at the high temperature of 300 °C with just a 28% decrement from the initial value, making it an ideal candidate for high-temperature piezoelectric devices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14060540/s1, Figure S1: Schematic diagram of diffusion temperature (Td) phenomenon on the properties of materials.; Figure S2: Comparison between the dielectric permittivity values at frequency 1 kHz and at room temperature (a) ɛr versus frequency plot, (b) ɛr versus temperature plot.

Author Contributions

Conceptualization, N.J. and A.H.; methodology, L.B.; software, A.T.; validation, A.H., N.J., and M.U.K.; formal analysis, L.B.; investigation, A.T.; resources, L.B.; data curation, A.T.; writing—original draft preparation, A.H. and N.J.; writing—review and editing, I.H.E.A.; visualization, M.U.K.; supervision, I.H.E.A.; project administration, N.J. and I.H.E.A.; funding acquisition, I.H.E.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to Taif University, Saudi Arabia, for supporting this work through project number (TU-DSPP-2024-105).

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Materials, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cao, J.; Ji, Y.; Shao, Z. Perovskites for protonic ceramic fuel cells: A review. Energy Environ. Sci. 2022, 15, 2200–2232. [Google Scholar] [CrossRef]
  2. Hussain, A.; Jabeen, N.; Rahman, A.U.; Qaiser, M.A.; Tariq, Z.; Abd El-Gawad, H.H. Experimental and theoretical study of Gd2O3 added pseudo-tetragonal Bi3TaTiO9-based ceramics for ferroelectric, electric and high-temperature piezoelectric applications. Ceram. Int. 2024, 50, 18177–18184. [Google Scholar] [CrossRef]
  3. Assirey, E.A.R. Perovskite synthesis, properties and their related biochemical and industrial application. Saudi Pharm. J. 2019, 27, 817–829. [Google Scholar] [CrossRef] [PubMed]
  4. Goel, P.; Sundriyal, S.; Shrivastav, V.; Mishra, S.; Dubal, D.P.; Kim, K.H.; Deep, A. Perovskite materials as su-perior and powerful platforms for energy conversion and storage applications. Nano Energy 2021, 80, 105552. [Google Scholar] [CrossRef]
  5. Chen, L.; Liu, H.; Qi, H.; Chen, J. High-electromechanical performance for high-power piezoelectric applications: Fundamental, progress, and perspective. Prog. Mater. Sci. 2022, 127, 100944. [Google Scholar] [CrossRef]
  6. Meng, Y.; Chen, G.; Huang, M. Piezoelectric materials: Properties, advancements, and design strategies for high-temperature applications. Nanomaterials 2022, 12, 1171. [Google Scholar] [CrossRef] [PubMed]
  7. Lv, X.; Wang, X.; Ma, Y.; Zhang, X.X.; Wu, J. Temperature stability of perovskite-structured lead-free piezoceramics: Evaluation methods, improvement strategies, and future perspectives. Mater. Sci. Eng. R Rep. 2024, 159, 100793. [Google Scholar] [CrossRef]
  8. Chen, K.-H.; Cheng, C.-M.; Chen, Y.-J.; Chen, M.-L. Lead-Free Piezoelectric Ceramic Micro-Pressure Thick Films. Crystals 2023, 13, 201. [Google Scholar] [CrossRef]
  9. Bell, A.J.; Deubzer, O. Lead-free piezoelectrics—The environmental and regulatory issues. MRS Bull. 2018, 43, 581–587. [Google Scholar] [CrossRef]
  10. Catalan, G.; Scott, J.F. Physics and applications of bismuth ferrite. Adv. Mater. 2009, 21, 2463–2485. [Google Scholar] [CrossRef]
  11. Lv, J.; Wu, J.; Wu, W. Enhanced Electrical Properties of Quenched (1 − x)Bi1−ySmyFeO3–xBiScO3 Lead-Free Ceramics. J. Phys. Chem. C 2015, 119, 21105–21115. [Google Scholar]
  12. Narayan, B.; Adhikari, S.; Madras, G.; Ranjan, R. Trapping a metastable ferroelectric phase by size reduction in semiconducting ferroelectric BiFeO3−PbTiO3 and its implications for photocatalytic response. Phys. Rev. Appl. 2017, 7, 024018. [Google Scholar] [CrossRef]
  13. Wang, D.; Wang, G.; Murakami, S.; Fan, Z.; Feteira, A.; Zhou, D.; Sun, S.; Zhao, Q.; Reaney, I.M. BiFeO3-BaTiO3: A new gen-eration of lead-free electroceramics. J. Adv. Dielectr. 2018, 8, 1830004. [Google Scholar] [CrossRef]
  14. Zheng, L.; Niu, Z.; Zheng, P.; Zhang, K.; Luo, C.; Zhang, J.; Wang, N.; Bai, W.; Zhang, Y. Simultaneously achieving high energy storage performance and remarkable thermal stability in Bi0.5K0.5TiO3-based ceramics. Mater. Today Energy 2022, 28, 101078. [Google Scholar] [CrossRef]
  15. Cheng, S.; Zhang, B.P.; Zhao, L.; Wang, K.K. Enhanced insulating and piezoelectric properties of BiFeO3-BaTiO3-Bi0.5Na0.5TiO3 ceramics with high Curie temperature. J. Am. Ceram. Soc. 2019, 102, 7355–7365. [Google Scholar] [CrossRef]
  16. Wang, Y.; Liu, H.; Yan, T.; Zhao, J.; Guo, S.; Sun, R.; Lu, Z.; Wang, D. Microstructure and electrical properties of Nb-doped SrTiO3-BiFeO3 based lead-free ceramics. J. Am. Ceram. Soc. 2022, 105, 2020–2028. [Google Scholar] [CrossRef]
  17. Saad, Y.; Álvarez-Serrano, I.; López, M.; Hidouri, M. Structural and dielectric characterization of new lead-free perovskites in the (SrTiO3)–(BiFeO3) system. Ceram. Int. 2016, 42, 8962–8973. [Google Scholar] [CrossRef]
  18. Wang, Z.; Cao, M.; Yao, Z.; Zhang, Q.; Song, Z.; Hu, W.; Xu, Q.; Hao, H.; Liu, H.; Yu, Z. Giant permittivity and low dielectric loss of SrTiO3 ceramics sintered in nitrogen atmosphere. J. Eur. Ceram. Soc. 2014, 34, 1755–1760. [Google Scholar] [CrossRef]
  19. Haeni, J.H.; Irvin, P.; Chang, W.; Uecker, R.; Reiche, P.; Li, Y.L.; Choudhury, S.; Tian, W.; Hawley, M.E.; Craigo, B.; et al. Room-temperature ferroe-lectricity in strained SrTiO3. Nature 2004, 430, 758–761. [Google Scholar] [CrossRef]
  20. Makarovic, M.; Bencan, A.; Walker, J.; Malic, B.; Rojac, T. Processing, piezoelectric and ferroelectric properties of (x)BiFeO3-(1 − x)SrTiO3 ceramics. J. Eur. Ceram. Soc. 2019, 39, 3693–3702. [Google Scholar]
  21. Wang, S.; Liu, H.; Wang, Y.; Qin, H.; Zhao, J.; Lu, Z.; Mao, Z.; Wang, D. Microstructure, dielectric, and piezoelectric properties of BiFeO3–SrTiO3 lead-free ceramics. J. Am. Ceram. Soc. 2024, 107, 205–213. [Google Scholar] [CrossRef]
  22. Arman, M.M. The effect of the rare earth A-site cation on the structure, morphology, physical properties, and application of perovskite AFeO3. Mater. Chem. Phys. 2023, 304, 127852. [Google Scholar] [CrossRef]
  23. Zeng, Z.; Xu, Y.; Zhang, Z.; Gao, Z.; Luo, M.; Yin, Z.; Zhang, C.; Xu, J.; Huang, B.; Luo, F.; et al. Rare-earth-containing perovskite nanomaterials: Design, synthesis, properties and applications. Chem. Soc. Rev. 2020, 49, 1109–1143. [Google Scholar] [CrossRef]
  24. Jang, S.; Lim, H.; Wi, S.; Park, J.; Jeong, Y.; Chung, J.-S.; Kang, W.; Kwark, Y.-J.; Kim, I.W.; Noh, H.; et al. A-site and B-site substitutions and the emission properties of Eu3+ ions in ABO3–type cubic perovskite: A case study of BaZrO3. Curr. Appl. Phys. 2020, 20, 1110–1117. [Google Scholar] [CrossRef]
  25. Firdaus, M. Oxidation of Rare Earth Permanent Magnets. Ph.D. Thesis, Swinburne University of Technology Melbourne, Melbourne, Australia, 2018. [Google Scholar]
  26. Jabeen, N.; Hussain, A.; Nawaz, S.; Qaiser, M.A.; Zafar, R.; Ishtiaq, S.; Ahmed, F.; Masud, U.; Sultan, T.; Muntaha, S.T. Thermally stable piezoelectric performance of MnO2 inserted pseudo-tetragonal phase existent CaBi2Nb2O9-based ceramics. Mater. Technol. 2022, 37, 2702–2710. [Google Scholar] [CrossRef]
  27. Hussain, A.; Jabeen, N.; Hassan, N.U.; Rahman, A.U.; Khan, M.U.; Naz, A.; Yousef, E.S. Influence of Mn Ions’ Insertion in Pseudo-Tetragonal Phased CaBi4Ti4O15-Based Ceramics for Highly Efficient Energy Storage Devices and High-Temperature Piezoelectric Applications. Int. J. Mol. Sci. 2022, 23, 12723. [Google Scholar] [CrossRef] [PubMed]
  28. Heron, J.T.; Bosse, J.L.; He, Q.; Gao, Y.; Trassin, M.; Ye, L.; Clarkson, J.D.; Wang, C.; Liu, J.; Salahuddin, S.; et al. Deterministic switching of ferro-magnetism at room temperature using an electric field. Nature 2014, 516, 370–373. [Google Scholar] [CrossRef] [PubMed]
  29. Abedi, M.; Fangueiro, R.; Correia, A.G. Effects of electrodes layout and filler scale on percolation threshold and piezoresistivity performances of a cementitious-based geocomposite. Nanomaterials 2022, 12, 1734. [Google Scholar] [CrossRef]
  30. Hiruma, Y.; Aoyagi, R.; Nagata, H.; Takenaka, T. Ferroelectric and piezoelectric properties of (Bi1/2K1/2) TiO3 ceramics. Jpn. J. Appl. Phys. 2005, 44, 5040. [Google Scholar] [CrossRef]
  31. Liu, H.; Chen, J. Effect of Li2CO3 addition on structural and electrical properties of 0.7BiFeO3–0.3BaTiO3 piezo-electric ceramic. J. Adv. Dielectr. 2022, 12, 2241003. [Google Scholar] [CrossRef]
  32. Qin, H.; Zhao, J.; Chen, X.; Li, H.; Wang, S.; Du, Y.; Li, P.; Zhou, H.; Wang, D. Investigation of lead-free BiFeO3–BaTiO3 piezoelectric ceramics through precise composition control. J. Adv. Dielectr. 2023, 13, 2350018. [Google Scholar] [CrossRef]
  33. Li, S.; Wang, F.; Tang, L.; Tan, D.Q.; Chen, Y. Structures, dielectric properties and AC impedance characteristics of BiFeO3–BaTiO3 high-temperature lead-free piezoceramics synthesized by the hydrothermal method. J. Adv. Dielectr. 2023, 13, 2340006. [Google Scholar] [CrossRef]
  34. Cui, Y.; Li, C.; Lei, X.; Li, F.; Sun, H.; Wang, Z.; Wang, S.; Peng, B.; Wang, D.; Yan, F.; et al. Achieving high Curie temperature and prominent piezoelectric response via site engineering for BiFeO3-BaTiO3 ceramics. J. Alloys Compd. 2023, 943, 169137. [Google Scholar] [CrossRef]
Figure 1. (a) Pseudo-cubic, MPB regional XRD analysis of BFST:xSm (x = 0–0.25) ceramic series. (b) Amplified XRD analysis of ceramics to analyze existing MPB region with pseudo-cubic structure.
Figure 1. (a) Pseudo-cubic, MPB regional XRD analysis of BFST:xSm (x = 0–0.25) ceramic series. (b) Amplified XRD analysis of ceramics to analyze existing MPB region with pseudo-cubic structure.
Crystals 14 00540 g001
Figure 2. Rietveld refinement results of measured XRD patterns: (a) BFST, (b) BFST:0.1Sm, (c) BFST:0.15Sm, and (d) BFST:0.20Sm ceramics.
Figure 2. Rietveld refinement results of measured XRD patterns: (a) BFST, (b) BFST:0.1Sm, (c) BFST:0.15Sm, and (d) BFST:0.20Sm ceramics.
Crystals 14 00540 g002
Figure 3. Morphological SEM images to analyze the grain sizes of (a) BFST, (b) BFST:0.1Sm, (c) BFST:0.15Sm, and (d) BFST:0.20Sm ceramics.
Figure 3. Morphological SEM images to analyze the grain sizes of (a) BFST, (b) BFST:0.1Sm, (c) BFST:0.15Sm, and (d) BFST:0.20Sm ceramics.
Crystals 14 00540 g003
Figure 4. (a) Ferroelectric polarization versus electric field (P-E loops) analysis of BFST:xSm ceramics with x = 0–0.25, (b) remnant polarization (Pr), and saturated polarization (PS). (c) Comparison of corrosive electric field (EC) values of BFST:xSm ceramics with x = 0–0.25 as a function of x values of Sm in BFST host matrix.
Figure 4. (a) Ferroelectric polarization versus electric field (P-E loops) analysis of BFST:xSm ceramics with x = 0–0.25, (b) remnant polarization (Pr), and saturated polarization (PS). (c) Comparison of corrosive electric field (EC) values of BFST:xSm ceramics with x = 0–0.25 as a function of x values of Sm in BFST host matrix.
Crystals 14 00540 g004
Figure 5. (a) Piezoelectric coefficient (d33) value comparison as a function of x values of Sm in BFST host matrix. (b) Thermal stability of d33 value analysis of BFST:xSm ceramics with x = 0–0.25 as a function of annealing temperature 50 °C to 400 °C. (c) Dielectric permittivity as a function of frequency plots of fully poled BFST:xSm with x = 0–0.25 ceramics. (d) Dielectric loss (tanδ) as a function of frequency plots of fully poled BFST:xSm with x = 0–0.25 ceramics.
Figure 5. (a) Piezoelectric coefficient (d33) value comparison as a function of x values of Sm in BFST host matrix. (b) Thermal stability of d33 value analysis of BFST:xSm ceramics with x = 0–0.25 as a function of annealing temperature 50 °C to 400 °C. (c) Dielectric permittivity as a function of frequency plots of fully poled BFST:xSm with x = 0–0.25 ceramics. (d) Dielectric loss (tanδ) as a function of frequency plots of fully poled BFST:xSm with x = 0–0.25 ceramics.
Crystals 14 00540 g005
Figure 6. (a) Dielectric permittivity (ɛr) as a function of temperature plots of BFST:xSm with x = 0–0.25 ceramics. (b) Comparison of Curie temperature (TC) and maximum dielectric constant (ɛr) values of BFST:xSm with x = 0–0.25 ceramics as a function of x values of Sm in BFST host matrix. (c) Dielectric loss (tanδ) as a function of temperature plots of BFST:xSm with x = 0–0.25 ceramics. (d) Comparison of dielectric loss (tanδ) values at 350 °C of BFST:xSm with x = 0–0.25 ceramics as a function of x values of Sm in BFST host matrix.
Figure 6. (a) Dielectric permittivity (ɛr) as a function of temperature plots of BFST:xSm with x = 0–0.25 ceramics. (b) Comparison of Curie temperature (TC) and maximum dielectric constant (ɛr) values of BFST:xSm with x = 0–0.25 ceramics as a function of x values of Sm in BFST host matrix. (c) Dielectric loss (tanδ) as a function of temperature plots of BFST:xSm with x = 0–0.25 ceramics. (d) Comparison of dielectric loss (tanδ) values at 350 °C of BFST:xSm with x = 0–0.25 ceramics as a function of x values of Sm in BFST host matrix.
Crystals 14 00540 g006
Table 1. Piezoelectric co-efficient comparison of the present study with the literature.
Table 1. Piezoelectric co-efficient comparison of the present study with the literature.
Sr. NoMaterialFabrication TechniquePiezoelectric Coefficient
d33 (pC/N)
Ref.
1(1 − x)Bi1−ySmyFeO3–xBiScO3Rapid thermal quenching51[16]
20.625BiFeO3–0.375SrTiO3Conventional solid-state synthesis69[20]
30.62BiFeO3–0.38SrTiO3Conventional solid-state synthesis72[21]
40.66BiFeO3–0.34BaTiO3Conventional solid-state synthesis217[32]
5K+ and Nb5+ co-doped (0.75 − x)BiFeO3–0.25BaTiO3Conventional solid-state synthesis163[34]
60.62BiFeO3–0.38SrTiO3:0.15wt%Sm2O3Conventional solid-state synthesis101This 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

Hussain, A.; Jabeen, N.; Tabassum, A.; Khan, M.U.; Basharat, L.; El Azab, I.H. Multifunctional Experimental Studies of Sm-Ion-Influenced Pseudo-Cubic Morphotropic Phase Boundary Regional BiFeO3-xSrTiO3 Ceramics for High-Temperature Applications. Crystals 2024, 14, 540. https://doi.org/10.3390/cryst14060540

AMA Style

Hussain A, Jabeen N, Tabassum A, Khan MU, Basharat L, El Azab IH. Multifunctional Experimental Studies of Sm-Ion-Influenced Pseudo-Cubic Morphotropic Phase Boundary Regional BiFeO3-xSrTiO3 Ceramics for High-Temperature Applications. Crystals. 2024; 14(6):540. https://doi.org/10.3390/cryst14060540

Chicago/Turabian Style

Hussain, Ahmad, Nawishta Jabeen, Aasma Tabassum, Muhammad Usman Khan, Laiba Basharat, and Islam H. El Azab. 2024. "Multifunctional Experimental Studies of Sm-Ion-Influenced Pseudo-Cubic Morphotropic Phase Boundary Regional BiFeO3-xSrTiO3 Ceramics for High-Temperature Applications" Crystals 14, no. 6: 540. https://doi.org/10.3390/cryst14060540

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