714
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Improvement of electrical and energy harvesting properties of new lead-free BST modified 0.995BNKT–0.005LN ceramics

, , , , & ORCID Icon
Pages 88-97 | Received 06 Oct 2022, Accepted 05 Dec 2022, Published online: 04 Jan 2023

ABSTRACT

New lead-free (1-y)[0.995Bi0.5(Na0.80K0.20)0.5TiO3-0.005LiNbO3]-y[(Ba0.7Sr0.3)TiO3] or (1-y)[0.995BNKT-0.005LN]-y[BST] (where y = 0, 0.01, 0.02, 0.03, 0.04 and 0.05 mol fraction) ceramics were fabricated by a conventional sintering method with two calcination steps. All ceramics sintered at 1125°C for 2 hours showed an optimum relative density of ~98% and a linear shrinkage of ~18%. X-ray diffraction patterns indicated that all ceramics possessed a pure perovskite phase with coexisting rhombohedral and tetragonal structures. Electrical and energy harvesting properties of 0.995BNKT–0.005LN ceramics were improved with 0.03 mol fraction BST addition (d33 = 287 pC/N, d*33  = 440 pm/V, g33 = 13.40 × 10−3 V/mN and FoM = 3.85 pm2/N). From the observed results, the 0.97[0.995BNKT-0.005LN]-0.03[BST] ceramic is a promising candidate for lead-free piezoelelectric sensor-actuator and energy harvester.

1. Introduction

Piezoelectric materials are well-known smart materials that can convert mechanical energy to electrical energy and vice versa. With these characteristics, the piezoelectric materials have been used in a variety of technological applications such as sensor, actuator, transducer, filter, resonator, micro-electromechanical system (MEMS), electromechanical device, high energy storage capacitor, and so on [Citation1–7]. Pb(Zr,Ti)O3 or PZT is a commercial piezoelectric material that has been selected as a material for piezoelectric applications due to its high piezoelectric performance. However, the toxicity of lead compounds released to the environment during a production process has been a concern [Citation8–11]. This problem has led to the development of new lead-free piezoelectric materials with useable piezoelectric properties.

Perovskite-type lead-free piezoelectric materials such as BaTiO3 (BT), (Bi0.5Na0.5)TiO3 (BNT), K0.5Na0.5NbO3 (KNN)-based compounds have been observed to possess high piezoelectric coefficients. However, these kinds of piezoelectric materials have always suffered from some drawbacks, i.e. a low Curie temperature (Tc) and a difficulty in poling treatment [Citation12]. Therefore, many efforts have been paid to solve these problems [Citation13]. Among lead-free piezoelectric materials, a bismuth-based Bi0.5Na0.5TiO3 (BNT) material is known to be a potential candidate for applications due to its high Tc (~ 320°C) and large remnant polarization (Pr ~ 32 µC/cm2). However, it has limited use due to high electrical conductivity, a large coercive field (Ec), and a lower normalized strain coefficient [Citation14–16]. Previous research revealed that dielectric, ferroelectric, and piezoelectric properties of BNT-based materials were enhanced by the addition of Bi0.5K0.5TiO3 (BKT). This was due to a formation of a morphotropic phase boundary (MPB) at BKT content of 16 to 20% [Citation13,Citation16–18]. Moreover, Zhou et al. have synthesized a complex (1-x-y)Bi0.5Na0.5TiO3-xBi0.5K0.5TiO3-yLiNbO3 ternary system to improve electrical properties of BNKT-based materials. They found that the optimum piezoelectric constant (d33) and electromechanical coupling factor (kp) of 195 pC/N and 0.336, respectively, were observed at x = 0.18 and y = 0.01 [Citation19]. It was also found in our previous study that electrical properties of BNKT ceramics were significantly modified by the addition of a small amount of LN (0.005 mol fraction [Citation20].

(Ba0.7Sr0.3)TiO3 (BST) is an insulating material with a large relative dielectric permittivity and small dielectric loss near ambient temperature, which is acceptable for a wide range of electrical applications [Citation21–23]. Previous research has shown improved electrical properties of BNKT-based materials by BST modification. For example, room temperature dielectric constant, εr, and d33, bipolar field-induced strain, dynamic piezoelectric coefficient, d*33, of BNKT ceramics were improved by BST addition [Citation24,Citation25].

Based on previous studies mentioned above, in this study, a new lead-free (1-y)[0.995Bi0.5(Na0.80K0.20)0.5TiO3-0.005LiNbO3]-y[(Ba0.7Sr0.3)TiO3] or (1-y)[0.995BNKT-0.005LN]-y[BST] (where y = 0, 0.01, 0.02, 0.03, 0.04 and 0.05 mol fraction) ternary system ceramics were fabricated by a conventional mixed oxide technique using two calcination steps. The effects of BST modification on phase, microstructure, dielectric, ferroelectric, and piezoelectric properties, and strain behavior of 0.995BNKT–0.005LN ceramics were investigated and discussed in detail.

2. Experimental procedures

Lead-free piezoelectric ceramics with the composition of (1-y)[0.995 Bi0.5(Na0.80K0.20)0.5TiO3-0.005LiNbO3]-y(Ba0.7Sr0.3)TiO3 or (1-y)[0.995BNKT-0.005LN]-y[BST] (where y = 0, 0.01, 0.02, 0.03, 0.04 and 0.05 mol fraction) were fabricated by a conventional mixed oxide technique using two calcination steps. Bi2O3 (purity 98%, Fluka), Na2CO3 (purity 99.5%, Carlo Erba), K2CO3 (purity 99%, Sigma-Aldrich), TiO2 (98.5%, Sigma-Aldrich), Li2CO3 (purity 99.0%, Sigma-Aldrich), and Nb2O3 (purity 99.0%, Sigma-Aldrich) for the synthesis of 0.995BNKT–0.005LN. BaCO3 (purity 98.5%, Fluka), SrCO3 (purity 99.9%, Sigma-Aldrich) and TiO2 (purity 98.5%, Sigma-Aldrich) were chosen as the reactant powder materials for (Ba0.7Sr0.3)TiO3. Prior to the reaction process, in order to eliminate any moisture, all carbonate powders, i.e. Na2CO3, K2CO3, BaCO3, Li2CO3, SrCO3, were dried in an oven at 120°C for 24 hours. All the reactant powders were weighed according to the stoichiometric formula, wet mixed by ball milling methods for 24 hours in 99% ethanol solution and dried in an oven at 120°C for 24 hours. The dried powder was separately calcined at 900°C for 2 hours with 0.995BNKT–0.005LN and 1100°C for 2 hours for BST with a heating/cooling rate of 5°C/min. The calcined 0.995BNKT–0.005LN and BST powders were mixed following the formula of (1-y)[0.995BNKT-0.005LN]-y[BST] (where y = 0, 0.01, 0.02, 0.03, 0.04 and 0.05 mol fraction) by the ball milling method again for 24 hours and dried in an oven at 120°C for 24 hours. To obtain the (1-y)[0.995BNKT-0.005LN]-yBST ceramics, the polyvinyl alcohol (PVA) binder with a concentration of 4 wt% was added to the mixed powders, which were pressed under 150 MPa pressure into the metal disks having 10 mm in diameter and sintered in an alumina crucible under normal atmosphere at 1125°C for 2 hours at a heating-cooling rate of 5°C/min.

The theoretical densities of all ceramics were calculated based on the theoretical densities of BNKT (5.84 g/cm3) [Citation26] and BT (6.01 g/cm3) [Citation27]. The linear shrinkage percentage of all ceramics was calculated using the ratio of the sample diameter before and after the sintering process. The phase formation was studied by X-ray diffractometer (XRD-Phillip, X-pert). Raman spectra were analyzed using Raman spectroscopy (Raman Jobin Yvon Horiba, T64000). The microstructure of all ceramics was examined using scanning electron microscopy (SEM, JEOL JSM-6335 F). The grain size distribution was measured using an ImageJ program. The average grain size of all ceramics was estimated using a linear intercept method following the ASTM E112-88 procedure. All ceramics were polished to a thickness of 1 mm and the silver paste was coated on both circular surfaces of the sintered sample and heated at 400°C for 15 minutes to formed electrodes before electrical properties measurement. The dielectric constant (εr) and dielectric loss (tanδ) as a function of temperature (50–350°C) were measured using a high precision LCR meter (LCR 821, GW INSTEK) under frequencies of 1, 10 and 100 kHz. The ferroelectric properties at room temperature of all ceramics were analyzed using a Radiant Precision ferroelectric tester. The electric field of 30, 40, and 50 kV/cm under a frequency of 1 Hz was applied to all studied ceramics. The remanent polarization (Pr), maximum polarization (Pmax), coercive field (Ec), and maximum coercive field (Emax) values were determined from the hysteresis loop curve. The strain-electric field (S-E) behavior was obtained by the optical displacement sensor (Fotonic Sensor model MTI-2100) at room temperatures under frequency of 0.1 Hz. The maximum strain (Smax) and the negative strain (Sneg) values were obtained from the bipolar curve at a maximum electric field (Emax) of 50 kV/cm. The normalized strain coefficient (d*33) was also calculated using the following equation: d*33 = Smax/Emax. Before piezoelectric measurement, all ceramics were poled in a silicone oil bath at 55°C under an applied DC electric field of 4 kV for 15 minutes. The piezoelectric coefficient (d33) of all ceramics was obtained by d33 meter (KCF technologies, S5865). The piezoelectric voltage constant (g33) was calculated using the equation, g33 = d33/εr ε0, where g33 is the piezoelectric voltage constant, εr the dielectric constant of the piezoelectric material, εo the dielectric constant in vacuum, and d33 the piezoelectric charge constant [Citation28,Citation29]. The off-resonance figure of merit (FoM) values for energy harvesting was calculated from the multiplication between piezoelectric charge constant (d33) and the piezoelectric voltage constant (g33) values according to the equation, FoM = d33g33 [Citation30].

3. Results and discussion

Relative density values of all (1-y)[0.995BNKT-0.005LN]-yBST ceramics were found to be about 98% while those of linear shrinkage were about ~18%. These results illustrated that the BST-added ceramics having near theoretical densities were obtained with the fabrication technique used in this study. It can be seen from the results given in that the relative density and linear shrinkage of samples gradually increased with increasing BST content.

Table 1. Physical and microstructure properties of the (1-y)[0.995BNKT-0.005LN]-yBST ceramics.

The X-ray diffraction patterns of all (1-y)[0.995BNKT-0.005LN]-yBST samples with the 2θ = 20–60 degrees are shown in ). All XRD patterns are similar to the previous reports of the 0.995BNKT–0.005LN-based ceramics [Citation20,Citation31] where a pure perovskite structure with no secondary phases was obtained. This suggested that the BST diffused into the 0.995BNKT–0.005LN to form the final compound. The A-site position of Li+ ion had no confirmed reported ionic radius value but the expectation value should be between rLi+ (0.92 Å, CN = 8 [Citation32] and rNa+ (1.39 Å, CN = 12 [Citation32])). Ba2+ (rBa2+ = 1.61 Å [Citation32]) ions likely diffused into Bi3+ (rBi3+ = 1.40 Å [Citation33]), Na+ (rNa+ = 1.39 Å [Citation32]), K+ (rK+ = 1.64 Å [Citation32]) and Sr2+ (rSr2+ = 1.44 Å [Citation32]) while Nb5+ (rNb5+ = 0.64 Å [Citation32]) could enter into B-site (CN = 6) of Ti4+ position (rTi4+ = 0.61 Å [Citation32]). From the narrow angular ranges of 2θ = 39°- 41° (See )) and 2θ = 45°- 48° (See )), the y = 0 sample represented mixed rhombohedral and tetragonal structure with the overlapping of (111)R/(111)T/(11ˉ1)R peaks. The tetragonal phase seen near 2θ ~ 46 degree from the splitting of (002)T/(200)T peaks was similar to the results previously reported [Citation16,Citation24,Citation25,Citation34]. With increasing BST content, we observed a phase change from mixed rhombohedral and tetragonal to a tetragonal dominant phase. The tetragonality (c/a) value confirmed the appearance of phase transformation (see ). The c/a values increased from 1.0118 (y = 0 sample) to 1.0136 (y = 0.05 sample). Moreover, the increase of BST concentration caused the diffraction peaks to slightly shift toward a lower angle, particularly those associated with the ab-plane of tetragonal phase. This was attributed to the substitutional effects between Bi3+, Na+, K+, Li+, Ba2+, and Sr2+ at position of A-site and Ti4+ and Nb5+ at position of B-site. This change induced a structural distortion i.e. enlargement of unit cell size as listed in . A similar peak shift was observed in previous reports of BNKT-based compounds [Citation24,Citation25,Citation31,Citation34].

Figure 1. X-ray diffraction patterns of the (1-y)[0.995BNKT-0.005LN]-y[BST] ceramics where (a)  = 20–60°, (b)  = 39–41° and (c)  = 45–48°.

Figure 1. X-ray diffraction patterns of the (1-y)[0.995BNKT-0.005LN]-y[BST] ceramics where (a) 2θ = 20–60°, (b) 2θ = 39–41° and (c) 2θ = 45–48°.

Raman spectroscopy is a technique used for investigating the vibrational modes in oxide materials, and probing the stretching and bending modes as well as the local structure in perovskite materials [Citation35] and the large length-scale average structures [Citation36]. Furthermore, this technique is used to understand the phase transformation of the ceramics. shows the Raman spectra measured at room temperature in a wavenumber range of 100–1200 cm−1 for all samples. The observed Raman spectra were similar to the results previously reported in references [Citation16,Citation24,Citation37]. The Raman modes below ~ 150 cm−1 were related to the A-site vibrations. The Raman modes between 200 and 400 cm−1 indicated the Ti-O vibration and those of 500–650 cm−1 were due to the octahedral Ti-O6 vibrations [Citation38–41]. The high-frequency region above 700 cm−1 specified the A1 (longitudinal optical) and E (longitudinal optical) overlapping bands [Citation42,Citation43]. In this study, the Raman peak around 100–200 cm−1 wavenumbers could be associated with the vibration of A-site cations such as Bi (Bi-O bond), Na (Na-O bond), and K (K-O bond) in the ABO3 perovskite structure. The Raman peak centered at ~ 272 cm−1 was attributed to the vibrations of Ti-O. The peak became broader and started to split into two bands with slight shift when the BST concentration increased. This was due to the co-occupancy of different cationic size at the position of B-site, suggesting a unit cell distortion. The two overlapping bands associated with the vibration of the TiO6 octahedra in the range of 450–650 cm−1 showed a separation into two distinct bands with the increase of BST concentration (y = 0.03–0.05 sample), indicating a particular phonon behavior in the structure evolution [Citation42,Citation43] and the phase change from mixed rhombohedral and tetragonal structures to be more tetragonal phase, in agreement with in the XRD result. The peak at the wavenumber above 700 cm−1 suggested the overlapping of longitudinal optical of A1 and E bands. This result also agreed with that observed by Hao et al [Citation42] and Malik et al [Citation44].

Figure 2. Raman spectra of the (1-y)[0.995BNKT-0.005LN]-y[BST] ceramics at room temperature.

Figure 2. Raman spectra of the (1-y)[0.995BNKT-0.005LN]-y[BST] ceramics at room temperature.

depicts the SEM images and histograms of the grain size distribution for the as-sintered surfaces of all (1-y)[0.995BNKT-0.005LN]-yBST samples. The SEM data confirmed that all samples were dense with nearly no pores. Moreover, the grain boundaries of all samples indicated the cubic-like morphology. For y = 0 sample, the average grain size was approximately 0.63 µm. The grain size was found to decrease with increasing BST concentration with the minimum value of 0.40 µm observed for the y = 0.05 sample. The average grain size values of all ceramics are given in . As observed from the density, shrinkage, tetragonality, and unit cell volume results, the properties suddenly changed when 0.01 mol fraction of BST was added. But the slight change in the properties was found with a BST content above 0.01 mol fraction. Due to the difference in diffusibility of solvent and solute ions in this system as well as the possibility of compositional variation near the grain boundaries, the mass transfer was then inhibited during the sintering process. Consequently, the grain growth process of the ceramics was restricted [Citation16,Citation45,Citation46].

Figure 3. SEM micrograph and histograms of the grain size distribution of the (1-y)[0.995BNKT- 0.005LN]-y[BST] ceramics.

Figure 3. SEM micrograph and histograms of the grain size distribution of the (1-y)[0.995BNKT- 0.005LN]-y[BST] ceramics.

shows temperature dependence of dielectric constant (εr) and dielectric loss (tanδ) measured under frequencies of 1, 10, and 100 kHz at temperature range 25–350°C. For all samples, εr values measured at lower frequency were greater as compared to those at higher frequency. This was due to a restriction of large-scale polarization such as space charge polarization which could not follow higher applied frequencies [Citation47,Citation48]. The maximum dielectric constant at high temperature (εm) of the unmodified 0.995BNKT-0.005LN ceramic was found to be 5570 with tan δ of 0.0187. This result corresponded to the previous studies reported on the BNKT-based system [Citation16,Citation34]. For the BST-added samples, the εm initially increased with BST content and reached a maximum value at y = 0.03 (εm = 5764, tan δ = 0.0155) and then dropped after y = 0.04. Furthermore, the dielectric characteristic of all samples exhibited two dielectric anomaly peaks. The first anomaly peak was observed at a lower temperature which is a depolarization temperature (Td). Td is determined to be a temperature at which a phase of a ceramic transforms from rhombohedral to tetragonal [Citation24,Citation48,Citation49]. The second anomaly peak observed at higher temperature is a maximum dielectric constant temperature (Tm) where the dielectric constant reaches a maximum value [Citation50,Citation51]. All samples exhibited a slight broadening of both Td and Tm peaks which could be attributed to compositional fluctuation of A-site and/or B-site lattices due to the occupation of different cations [Citation52]. The 0.995BNKT–0.005LN ceramic studied in this work possessed Td and Tm of 147°C and 255°C, respectively. The observed values were similar to those reported in previous studies of other BNKT-based materials [Citation16,Citation47]. The Td of the ceramic decreased to 97°C for y = 0.01 composition. The Td was nearly unchanged with further increasing y, as shown in . The decrease in Td could be resulted from the ferroelectric order destabilization and may be associated with the possible presence of nonpolar phase in BST-doped sample [Citation53,Citation54]. This behavior was in good agreement with the previous work for Zr-doped Bi0.5(Na0.78K0.22)0.5TiO3 system [Citation55]. The Curie point or Tm value was found to gradually increase with increasing BST content. This was believed to be due to the decreased grain size which resulted in the increased surface area of space charge layer and subsequently, both the space charge field and the locked-in ferroelectric polarization increased. This phenomenon was also observed in the study of the PLZT system [Citation56].

Figure 4. (a) Temperature dependence of the dielectric constant (εr) and dielectric loss (tan δ) under various of frequencies from 1–100 kHz and (b) Plots of Tm and Td as a function of BST concentration of the (1-y)[0.995BNKT-0.005LN]-y[BST] ceramics (inset show: Plots of εm and tan δ as a function of BST content).

Figure 4. (a) Temperature dependence of the dielectric constant (εr) and dielectric loss (tan δ) under various of frequencies from 1–100 kHz and (b) Plots of Tm and Td as a function of BST concentration of the (1-y)[0.995BNKT-0.005LN]-y[BST] ceramics (inset show: Plots of εm and tan δ as a function of BST content).

The ferroelectric properties and polarization-electric field (P-E) curves at room temperature and at a frequency of 1 Hz under electric fields of 30, 40, and 50 kV/cm for all samples are shown in . Plots of remanent polarization (Pr) and coercive field (Ec) values as a function of electric field are also shown. The Pr and Ec values increased with increasing electric field. All P-E curves were well-saturated with typical P-E hysteresis loops at the electric field of 50 kV/cm. The BTS addition significantly influenced the ferroelectric hysteresis loop shape and the polarization response of 0.995BNKT–0.005LN-based material. Hysteresis loop shape became more squared when y > 0.01 which indicated a strong ferroelectric interaction [Citation57]. An unmodified-0.995BNKT-0.005LN ceramic had Pr = 23.10 μC/cm2 and Ec = 18.28 kV/cm. The Pr tended to increase with increasing BST content and reached a maximum value of 32.92 μC/cm2 at y = 0.03. However, with further increasing BST content (y = 0.04–0.05), the Pr value decreased. Likewise, the Ec value reached a maximum value of 26.85 kV/cm at y = 0.04 and the value slightly dropped to 25.98 kV/cm at y = 0.05 sample. The increasing tendency of Pr could be explained by the increased density of material due to BST addition. In case of the increased Ec value, this suggested that the material became harder, which was consistent with the assumption that oxygen vacancies were formed, and the so-called complex defects were produced by these acceptor dopants [Citation57]. This finding was also in agreement with previous reports [Citation24,Citation54,Citation58].

Figure 5. Polarization-electric field (P-E) hysteresis loop, and plots of Pr and Ec values as a function of electric field of the (1-y)[0.995BNKT- 0.005LN]-y[BST] ceramics, measured at RT under various electric fields of 30, 40, and 50 kV/cm and a frequency of 1 Hz.

Figure 5. Polarization-electric field (P-E) hysteresis loop, and plots of Pr and Ec values as a function of electric field of the (1-y)[0.995BNKT- 0.005LN]-y[BST] ceramics, measured at RT under various electric fields of 30, 40, and 50 kV/cm and a frequency of 1 Hz.

In order to study the bipolar electric field-induced strain curve of all (1-y)[0.995BNKT-0.005LN]-yBST samples, strain-electric field (S-E) curves measured at room temperature were evaluated. The S-E curves measured at a frequency of 0.1 Hz with a maximum electric field of 50 kV/cm of the ceramics are shown in . The typical S-E curve of this work, which was in agreement with BNKT-based ceramics of the butterfly-like curve [Citation16,Citation24,Citation42,Citation44], indicated the typical piezoelectric response [Citation59]. The maximum strain (Smax), the negative strain (Sneg) and the normalized strain coefficient (d*33 = Smax/Emax) are also summarized in . The variation of Smax with increasing BST content showed a similar trend to that of d*33 values. The undoped sample had the Smax of 0.15% and calculated d*33 of 278 pm/V, indicating that the ferroelectric domain switching was involved at Ec where the largest Sneg of −0.09% was observed. The Smax and d*33 tended to increase with the higher concentration of BST as shown in ). The maximum value was found in the y = 0.03 sample (Smax = 0.22% and d*33 = 440 pm/V). The Smax and d*33 values were then dropped with further increasing BST content. The decreased Smax and d*33 after y = 0.03 sample might be due to the dominating tetragonal phase and the increasing trend in tetragonality (c/a) which clearly indicated the increase of lattice anisotropy of modified-0.995BNKT-0.005BST samples (i.e. rhombohedral phase decreased while tetragonal phase increased) [Citation60,Citation61]. The increasing trend in lattice anisotropy which showed slight distortion of XRD patterns resulted in the enlargement of lattice constant and lattice energy which induced a phase transformation in order to stabilize the structure [Citation62]. This suggested that ferroelectric order was over a nonpolar region. In addition, the decreasing trend of the nonpolar phase at the higher BST concentration likely suppressed the domain switching ability [Citation63].

Table 2. Electrical properties of the (1-y)[0.995BNKT-0.005LN]-yBST ceramics.

Figure 6. Bipolar strain electric-field S-E loop of the (1-y)[0.995BNKT-0.005LN]-yBST ceramics.

Figure 6. Bipolar strain electric-field S-E loop of the (1-y)[0.995BNKT-0.005LN]-yBST ceramics.

Figure 7. (a) Plot of the Smax and d33* as a function of BST content and (b) Plot of the g33 and FOM values as a function of BST content of the (1-y)[0.995BNKT-0.005LN]-y[BST] ceramics.

Figure 7. (a) Plot of the Smax and d33* as a function of BST content and (b) Plot of the g33 and FOM values as a function of BST content of the (1-y)[0.995BNKT-0.005LN]-y[BST] ceramics.

The piezoelectric coefficient (d33) values of all (1-y)[0.995BNKT-0.005LN]-yBST samples are summarized in . The 0.995BNKT-0.005LN sample had a d33 value of 152 pC/N. The d33 value of the modified 0.995BNKT-0.005LN samples increased with increasing BST content, and the composition of y = 0.03 showed the highest value of 287 pC/N. The energy-harvesting utilizing piezoelectric materials has recently attracted extensive attention due to the strong demand of self-powered electronics [Citation64]. The important factor for evaluating piezoelectric energy harvester is the piezoelectric voltage coefficient (g33) value [Citation65]. The result from the calculation of the g33 value are also listed in . The g33 value in this work was found to be improved by the BST doping. The undoped-0.995BNKT-0.005LN sample had the g33 value of 8.30 × 10−3 Vm/N and increased with increasing BST content. The maximum value of 13.40 × 10−3 Vm/N was found for the y = 0.03 sample. The d33 value decreased with further increasing of the modifier content greater than y = 0.03. The off-resonance figure of merit (FoM) of all (1-y)[0.995BNKT-0.005LN]-yBST samples are listed in . The energy harvesting properties were also improved by the BST doping. The undoped-0.995BNKT-0.005LN sample had the FoM value of 1.26 pm2/N. The FoM value increased with the increasing BST content and showed the maximum value of 3.85 pm2/N for the y = 0.03 sample and then dropped with a further increase of the modifier content. It can be seen that the trend of the FoM value was similar to that of the trend of g33 value [Citation66] as shown in ). The comparison of the FoM values with the previously studied lead-free piezoelectric ceramics are shown in . Based on the obtained data, the piezoelectric characteristics and energy harvesting behavior improvement was achieved in this system. The addition of BST into the 0.995BNKT-0.005LN system had potential to be one of the promising lead-free compounds for further use in piezoelectric devices.

Figure 8. Comparison of FoM values of (1-y)[0.995BNKT-0.005LN]-yBST ceramics with other lead-free ceramics.

Figure 8. Comparison of FoM values of (1-y)[0.995BNKT-0.005LN]-yBST ceramics with other lead-free ceramics.

4. Conclusions

The new lead-free (1-y)[0.995BNKT-0.005LN]-yBST system (where y = 0, 0.01, 0.02, 0.03, 0.04 and 0.05 mol fraction) was successfully fabricated by a solid-state mixed oxide method using two calcination steps. X-ray diffraction results showed that all ceramic samples exhibited a single-phase perovskite structure with no impurity phases detected. The decrease of grain size was observed when the BST dopant increased. For y = 0.03 sample, it showed the maximum dielectric properties (εm = 5764, tan δ = 0.0155, Td = 100°C, Tm = 262°C), and ferroelectric parameters (Pr = 32.92 µV/cm2, Ec = 24.82 kV/cm). The highest piezoelectric coefficient (d33 = 287 pC/N), piezoelectric voltage constant (g33 = 13.40 × 10−3 Vm/N), and the off-resonance figure of merit for energy harvesting (FoM = 3.85 pm2/N) were also obtained for this sample. This material system has the potential to be one of the promising new lead-free piezoelectric candidates for further use in sensor-actuator and energy harvesting applications [Citation67,Citation68]

Acknowledgments

This research was financially supported by Department of Physics and Materials Science, Faculty of Science, Chiang Mai University, Thailand Science Research and Innovation (TSRI), the National Research Council of Thailand (NRCT): NRCT5-RSA63004-15. Partial supports from the Center of Excellence in Materials Science and Technology, Faculty of Science, Chiang Mai University are also acknowledged. MP would like to thank Rubber Product and Innovation Development Research Unit (SCIRU63002).

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by the Department of Physics and Materials Science, Faculty of Science, Chiang Mai University, Thailand Science Research and Innovation (TSRI), the National Research Council of Thailand (NRCT): <#award-id;temp>NRCT5-RSA63004-15. The Center of Excellence in Materials Science a [NRCT5-RSA63004-15]; Rubber Product and Innovation Development Research Unit (SCIRU63002) [SCIRU63002];Department of Physics and Materials Science, Faculty of Science, Chiang Mai University, Thailand Science Research and Innovation (TSRI), the National Research Council of Thailand (NRCT): NRCT5-RSA63004-15. The Center of Excellence in Materials Science a [NRCT5-RSA63004-15]; P. Jaiban would like to thank partial funding by National Science, Research and Innovation Fund (NSRF), and King Mongkut’s University of Technology North Bangkok with Contract no. KMUTNB-FF-65-43.

References

  • Kulkarni H, Zohaib K, Khusru A, et al. Application of piezoelectric technology in automotive systems. Mater Today: Proc. 2018;5:21299–21304.
  • Panda S, Hajra S, Mistewicz K, et al. Piezoelectric energy harvesting systems for biomedical applications. Nano Energy. 2022;100:107514.
  • Liu J, Zhang X, Wang J, et al. Global structure search for new 2D PtSSe allotropes and their potential for thermoelectirc and piezoelectric applications. Chem Phys Lett. 2022;805:139913.
  • Kumar A, Varghese A, Sharma A, et al. Recent development and futuristic applications of MEMS based piezoelectric microphones. Sens Actuator A Phys. 2022;347:113887.
  • Bijalwan V, Tofel P. Synthesis of high density sub-10 µm (Ba0.85Ca0.15) (Zr0.1Ti0.9)O3-xCeO2 lead-free ceramics using a twostep sintering technique. J Am Ceram Soc. 2019;7(4):536–543.
  • Zeb A, Hall DA, Aslam Z, et al. Structure-property relationships in the lead-free piezoceramic system K0.5Bi0.5TiO3-BiMg0.5Ti0.5O3. Acta Mater. 2019;168:100–108.
  • Patel S, Chauhan A, Vaish R, et al. Enhanced energy storage performance of glass added 0.715Bi0.5Na0.5TiO3-0.065BaTiO3-0.22SrTiO3 ferroelectric ceramics. J Asian Ceram Soc. 2015;3:383–389.
  • Wang X, Huan Y, Ji S, et al. Ultra-high piezoelectric performance by rational tuning of heterovalent-ion doping in lead-free piezoelectric ceramics. Nano Energy. 2022;101:107580.
  • Kim KB, Lee J. Review: a lead-free piezoelectric fiber generator with a high energy conversion constant. Energies. 2022;15(18):6787.
  • Palneedi H, Peddogari M, Hwang GT, et al. Hight performance dielectric ceramic films for energy storange capacitors: progress and outlook. Adv Mater. 2018;28:1803665.
  • Khan HU, Ullah A, Ullah M, et al. Structural, electrical and electromechanical properties of [(Bi0.5Na0.5)0.935Ba0.065)]1-yLay (Ti1-zTaz)O3 lead-free piezoelectric ceramics. Physica B. 2022;584:412102.
  • Takenaka T, Nagata H. Current status and prospects of lead-free piezoelectric ceramics. J Eur Ceram Soc. 2005;25(12):2693–2700.
  • Tran VDN, Vu LH, Van VL, et al. Structure evolution and electrical properties of lead-free Bi0.5Na0.41K0.09TiO3 piezoceramics by isovalent La doping. J Mater Sci: Mater Electron. 2021;32:4363–4371.
  • Wang C, Li Q, Zhang W, et al. Large electric field-induced strain in the novel BNKTAN-BNBLTZ lead-free ceramics. J Mater Sci Technol. 2020;45:15–22.
  • Ren X, Yin H, Tang Y, et al. The large electro-strain in BNKT-BST-100xTa lead-free ceramics. Ceram Int. 2020;46(2):1876–1882.
  • Jaita P, Jarupoom P. Enhanced electric field-induced strain andelectrostrictive response of lead-free BaTiO3-modified Bi0.5(Na0.80K0.20)0.5TiO3 piezoelectric ceramics. J Asian Ceram Soc. 2021;9(3):975–987.
  • Wang X, Gao H, Hao X, et al. Enhanced piezoelectric, electrocaloric and energy storage properties at high temperature in lead-free Bi0.5(Na1-xKx)0.5TiO3 ceramics. Ceram Int. 2019;45(4):4274–4282.
  • Sasaki A, Chiba T, Mamiya Y, et al. Dielectric and piezoelectric properties of (Bi, Na)TiO3-(Bi K)TiO3 systems, Jpn. J Appl Phys. 1999;38:5564–5567.
  • Zhou C, Liu X, Li W, et al. Microstructure and electrical properties of Bi0.5Na0.5TiO3-Bi0.5K0.5TiO3-LiNbO3 lead-free piezoelectric ceramics. J Phys Chem Solids. 2009;70(3–4):541–545.
  • Wannasut P, Jaita P, Watcharapasorn A, et al. X-ray photoelectron spectroscopy analysis and electrical properties of Bi0.5(Na0.80K0.20)0.5TiO3-LiNbO3 lead-free piezoelectrics. Ferroelectrics. 2015;489(1):118–128.
  • Ezhilvalavan S, Tseng T. Progress in the developments of (Ba,Sr)TiO3 (BST) thin films for Gigabit era DRAMs. Rev Mater Chem. Phys. 2000;65(3):227–248
  • Su B, Button TW. The processing and properties of barium strontium titanate thick films for use in frequency agile microwave circuit application. J Eur Ceramic Soc. 2001;21(15):2641–2645.
  • Guo H, Gao W, Yoo J. Barium strontium titanate (Ba0.7Sr0.3TiO3) ferroelectric films produced by electrophoretic deposition. Curr Appl Phys. 2004;4(2–4):385–388.
  • Jaita P, Watcharapasorn A, Kumar N, et al. Lead-free (Ba0.70Sr0.30)TiO3-modified Bi0.5(Na0.80K0.20)0.5TiO3 ceramics with large electric field-induced strains. J Am Ceram Soc. 2016;99(5):1615–1624.
  • Jaita P, Watcharapasorn A, Jiansirisomboon S. Investigation of a new lead-free Bi0.5(Na0.40K0.10)TiO3-(Ba0.7Sr0.3)TiO3 piezoelectric ceramic. Nanoscale Res Lett. 2012;7(1):24.
  • Zhang YR, Li JF, Zhang BP. Enhancing electrical properties in NBT-KBT lead-free piezoelectric ceramics by optimizing sintering temperature. J Am Ceram Soc. 2008;91(8):2716–2719.
  • Swanson HE, Fuyat RK. Standard X-ray diffraction powder patterns. Natl Bur Stand Circ (US). 1954;539:44–45.
  • Ahn CW, Kim HS, Woo WS, et al. Low-temperature sintering of Bi0.5(Na,K)0.5TiO3 for multilayer ceramic actuators. J Am Ceram Soc. 2015;98(6):1877–1883.
  • Liu B, Li P, Shen B, et al. Simultaneously enhanced piezoelectric response and piezoelectric voltage coefficient in textured KNN‐based ceramics. J Am Ceram Soc. 2018;101(1):265–273.
  • Shin DJ, Kang WS, Koh JH, et al. Comparative study between the pillar‐ and bulk‐type multilayer structures for piezoelectric energy harvesters. Phys Status Solidi A. 2014;211(8):1812–1817.
  • Wannasut P, Watcharapasorn A, Jaita P, et al. Microstructure and electrical properties of (1-x)Bi0.5(Na0.80K0.20)0.5TiO3-xLiNbO3 lead-free piezoelectric ceramics. 2015 Joint IEEE International Symposium on the Applications of Ferroelectric (ISAF), International Symposium on Integrated Functionalities (ISIF), and Piezoelectric Force Microscopy Workshop (PFM). 2015;167–170.
  • Shannon RD. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976;32(5):751–767.
  • Karthik T, Durga Rao T, Srinivas A, et al. A-Site cation disorder and size variance effects on the physical properties of multiferroic Bi0.9RE0.1FeO3 Ceramics (RE = Gd3+, Tb3+, Dy3+). Accessed on 2 October 2022. https://arxiv.org/abs/1206.5606
  • Jaita P, Sanjoom R, Lertcumfu N, et al. Improvement of electric field-induced strain and energy storage density properties in lead-free BNKT-based ceramics modified by BFT doping. RSC Adv. 2019;9(21):11922.
  • Chauhan V, Ghosh SK, Hussain A, et al. Rout influence of niobium substitution on structural and opto-electrical properties of BNKT based piezoelectric ceramics. J Alloys Compd. 2016;674:413–424.
  • Eerd BWV, Damjanovic D, Klein N. Structural complexity of (Na 0.5 Bi 0.5) TiO 3 -BaTiO 3 as revealed by Raman spectroscopy. Phys Rev B. 2010;82(10):104112.
  • Kreisel J, Glazer AM, Jones G, et al. An x-ray diffraction and Raman spectroscopy investigation of A-site substituted perovskite compounds: the (Na1-xKx)0.5Bi0.5TiO3 (0 ≤ x ≤ 1) solid solution. J Phys: Condens Matter. 2000;12:3267–3280.
  • Wei Y, Zhang N, Jin C, et al. Bi0.5K0.5TiO3</sub>–CaTiO3 ceramics: appearance of the pseudocubic structure and ferroelectric-relaxor transition characters. J Am Ceram Soc. 2019;102(6):3598–3608.
  • Suchanicz J, Kania A, Czaja P, et al. Structural, thermal dielectric and ferroelectric properties of K0.5Bi0.5TiO ceramics. J Eur Ceram Soc. 2018;38(2):567–574.
  • Rout D, Moon KS, Kang SJL, et al. Dielectric and Raman scattering studies of phase transitions in the (100-x)Na0.5Bi0.5TiO3-xSrTiO3 system. J Appl Phys. 2010;108(8):084102.
  • Jaita P, Jarupoom P, Sweatman DS, et al. Energy harvesting, electrical, and magneticproperties of potassium bismuth titanate-based lead-free ceramics. J Asian Ceram Soc. 2021;9(3):947–963.
  • Hao J, Shen B, Zhai J, et al. Switching of morphotropic phase boundary and large strain response in lead-free ternary (Bi0.5Na0.5)TiO3-(K0.5Bi0.5)TiO3-(K0.5Na0.5)NbO3 system. J Appl Phys. 2013;113(11):114106.
  • Aksel E, Forrester JS, Kowalski B, et al. Structure and properties of Fe-modified Na0.5Bi0.5TiO3 at ambient and elevated temperature. Phys Rev B. 2012;85:024121.
  • Malik RA, Hussaina A, Zaman, et al. Structure–property relationship in lead-free A- and B-site co-doped Bi0.5(Na0.84K0.16)0.5TiO3–SrTiO3 incipient piezoceramics. RSC Adv. 2015;5(117):96953–96964.
  • Shukla A, Choudhary RNP. Study of electrical properties of La3+/Mn4+-modified PbTiO3 nanoceramics. J Mater Sci. 2012;47(13):5074–5085.
  • Li Q, Wang C, Yadav AK, et al. Large electrostrictive effect and energy storage density in MnCO3 modified Na0.325Bi0.395Sr0.245□0.035TiO3 lead-free ceramics. Ceram Int. 2020;46(3):3374–3381.
  • Hussain A, Zaman A, Iqbal Y, et al. Dielectric, ferroelectric and field induced strain properties of Nb-modified Pb-free 0.99Bi0.5(Na0.82K0.18)0.5TiO3-0.01LiSbO3 ceramics. J Alloys Compd. 2013;574:320–324.
  • Chiang YM, Birnie DP, Kingery WD. 1997. physical ceramics: principles for ceramics science and engineering. New York: John Wiley & Sons, Inc.
  • Jaita P, Watcharapasorn A, Jiansirisomboon S. Effect of composition on electrical properties of lead-free Bi0.5(Na0.80K0.20)0.5TiO3-(Ba0.98Nd0.02)TiO3 piezoelectric ceramics. J Appl Phys. 2013;114(2):027005.
  • Yang Z, Hou Y, Liu B, et al. Structure and electrical properties of Nd2O3-doped 0.82Bi0.5Na0.5TiO3-0.18Bi0.5K0.5TiO3 ceramics. Ceram Int. 2009;35:1423–1427.
  • Maqbool A, Hussain A, Malik RA, et al. Evolution of phase structure and giant strain at low driving fields in Bi-based lead-free incipient piezoelectrics. Mater Sci Eng B. 2015;199:105–112.
  • Malik RA, Hussain A, Maqbool A, et al. Temperature-Insensitive High Strain in Lead-Free Bi0.5(Na0.84K0.16)0.5TiO3−0.04SrTiO3 Ceramics for Actuator Applications. J Am Ceram Soc. 2015;98(12):3842–3848.
  • Fan G, Lu W, Wang X, et al. Phase transition behaviour and electromechanical properties of (Na1/2Bi1/2)TiO3-KNbO3 lead-free piezoelectric ceramics. J Phys D Appl Phys. 2008;41(3):035403–035406.
  • Dai YJ, Zhang XW, Chen KP, et al. An approach to improve the piezoelectricproperty of (Bi0.5Na0.5)TiO3-(Bi0.5K0.5)TiO3-BaTiO3 lead-free ceramics. Int J Appl. Ceram Technol. 2011;8(2):423–429.
  • Hussain A, Ahn CW, Lee S, et al. Large electric-field-induced strain in Zr-modified lead-free Bi0.5(Na0.78K0.22)0.5TiO3 piezoelectric ceramics. Sens Actuator A Phys. 2010;158(1):84–89.
  • Okazaki K, Nagata K. Effects of grain size and porosity on electrical andoptical properties of PLZT Ceramics. J Am Ceram Soc. 1973;56(2):82–86.
  • Das S, Ghara S, Mahadevan P, et al. Designing a lower band gap bulk ferroelectric material with a sizable polarization at room temperature. ACS Energy Lett. 2018;3(5):1176–1182.
  • Vendrell X, García JE, Bril X, et al. Improving the functional properties of (K0.5Na0.5)NbO3 piezoceramics by acceptor doping. J Eur Ceram Soc. 2015;35(1):125–130.
  • Haertling GH. Ferroelectric ceramics: history and technology. J Am Ceram Soc. 1999;82(4):797–818.
  • Rase DE, Roy R. Phase equilibria in the system BaO-TiO2. J Am Ceram Soc. 1955;38(3):102–113.
  • Li Y, Chen W, Xu Q, et al. Piezoelectric and dielectric properties of CeO2-doped Bi0.5Na0.44K0.06TiO3 lead-free ceramics. Ceram Int. 2007;33(1):95–99.
  • Lee WC, Huang CY, Tsao LK, et al. Crystal structure, dielectric and ferroelectric properties of (Bi0.5Na0.5)TiO3-(Ba, Sr)TiO3 lead-free piezoelectric ceramics. J Alloy Compd. 2010;492(1–2):307–312.
  • Wang B, Luo L, Ni F, et al. Piezoelectric and ferroelectric properties of (Bi1-xNa0.8K0.2Lax)0.5TiO3 lead-free ceramics. J Alloy Compd. 2012;526:79–84.
  • Liu Y, Chang Y, Sun E, et al. Significantly enhanced energy harvesting performance and superior fatigue resistant behavior in [001]c-textured BaTiO3 based lead-free piezoceramics. ACS Appl Mater Interfaces. 2018;10(37):31488–31497.
  • Manotham S, Butnoi P, Jaita P, et al. Role of ZnO nanoparticle doping on depolarization temperature, piezoelectric and energy harvesting properties of lead-free Bi0.5(Na0.84K0.16)0.5TiO3 ceramics. Mater Res Bull. 2020;128:110859.
  • Shin DJ, Kim J, Koh JH, et al. Piezoelectric properties of(1-x)BZT-xBCT system for energy harvesting applications. J Euro Ceram Soc. 2018;38(13):4395–4403.
  • Jaita P, Manotham S, Rujijanagul G. Influence of Al2O3 nanoparticle doping on depolarization temperature, and electrical and energy harvesting properties of lead-free 0.94(Bi0.5 Na0.5)TiO3-0.06BaTiO3ceramics. RSC Adv. 2020;10(53):32078–32087.
  • Coondoo I, Panwar N, Amorın H, et al. Enhanced piezoelectric properties of praseodymium-modified leadfree (Ba0.85Ca0.15)(Ti0.90Zr0.10)O3 ceramics. J Am Ceram Soc. 2015;98(10):3127–3135.