Ceramics International xx (xxxx) xxxx–xxxx Contents lists available at ScienceDirect Ceramics International journal homepage: www.elsevier.com/locate/ceramint B2O3 doping in 0.94(Bi0.5Na0.5)TiO3-0.06BaTiO3 lead-free piezoelectric ceramics ⁎ Metin Ozgul , Abdullah Kucuk Afyon Kocatepe University, Department of Materials Science and Engineering, Afyonkarahisar, 03200 Turkey A R T I C L E I N F O A BS T RAC T Keywords: C. Dielectric properties D. BaTiO3 and titanates Piezoelectric properties Perovskites The effects of B2O3 doping on the dielectric and piezoelectric properties of 0.94(Bi0.5 Na0.5) TiO3-0.06BaTiO3 (BNT-6BT) ceramics were investigated. Pre-reacted BNT-6BT powders synthesized by solid state reaction method were doped with B3+ ions in 0, 0.1, 0.2, 0.3, 0.4, 0.5, 1, and 2 mol%. All samples were sintered at 1150 ° C for 12 h in air. The effects of boron doping on structure and electrical properties were characterized as a function of dopant concentration. The results indicate a substantial increase in piezoelectric coefficient (d33), dielectric constant (εr ), and dielectric loss (tan δ ) approximately 19%, 42%, and 95%, respectively. The positive effects of doping were more significant in BNT-6BT samples with 0.3% and more B2O3 addition. The highest piezoelectric coefficient was obtained as d33=173 pC/N in 1% doped BNT-6BT. In terms of dielectric properties, both εr and tan δ values increased with B3+% reaching 1075 and 0.0421, respectively in 2% doped samples. Even though the measured density values remain relatively unchanged, the increase of all these coefficients together implies a donor dopant role of B3+ ions. 1. Introduction Piezoelectric ceramics are very important for many electronic applications and many researchers have shown great interest in the synthesis and electrical properties of these materials. For more than a half century lead-based piezoelectric ceramics, such as Pb(Zr1−xTix)O3 (PZT), have been the most commonly studied and used piezoelectric materials in sensors, actuators, and transducers due to their excellent piezoelectric properties [1–3]. The outstanding properties of PZT ceramics are mostly obtained in the compositions near the morphotropic phase boundary (MPB) where Zr/Ti ratio is ~53/47. Due to coexistence of both rhombohedral and tetragonal phases around MPB, a high degree of alignment of ferroelectric dipoles takes place resulting in striking enhancement in the piezoelectric properties [4]. However, the strong toxicity and high volatility of lead oxide during synthesis and processing at elevated temperatures cause severe environmental and ecological problems. Therefore, synthesis and use of PZT ceramics were started to be regulated in many countries [5–9]. For this reason piezoelectric materials research started to focus on finding alternatives to lead-based systems in recent years. Significant improvements have been made in the development of lead-free piezoelectric ceramics with good properties which are comparable with PZT. The search for alternative piezoelectric ceramics has been concentrated on alkali niobates (KNaNbO3; KNN), modified barium titanates (BaTiO3; BT), and bismuth titanates (BiNaTiO3; BNT) and other systems with MPBs ⁎ [4–9]. Among these materials, a solid solution of 0.94(Bi0.5Na0.5)TiO30.06BaTiO3 (BNT-6BT) with a MPB has been considered as a potential alternative for piezoelectric applications [10]. Similar with PZT, MPB separates rhombohedral and tetragonal phases in BNT-6BT. BNT is a perovskite ferroelectric material with a large remanent polarization (Pr=38 μC/cm2). However, due to its relatively large coercive field (Ec=73 kV/cm) pure BNT piezoelectric ceramics are difficult to pole effectively and therefore provide low piezoelectric properties (d33=83 pC/N), which is significantly lower than that of the Pb (Zr0.52Ti0.48)O3 ceramic (d33=223 pC/N) [1,10,11]. Much higher piezoelectric properties (d33=155 pC/N) were reported for BNT-6BT in contrast to pure BNT [9–13]. The coercive field values reported to be much lower for BNT–BT compositions near the MPB in comparison to pure BNT, providing substantially improved piezoelectric properties [11–13]. On the other hand, in order to meet the strict requirements for specific applications, the electrical properties need to be improved, for example, by doping of various oxides [14–24]. Modifications of BNT-BT include lithium [14], cerium [15], lanthanum [16], zirconium [17], niobium [18], tantalum [19], antimony [20], hafnium [21], cobalt [22], silver [23] and manganese [24]. All enhance some combination of piezoelectric and dielectric properties by improving densification and/ or controlling microstructure and chemistry. In various lead-free perovskite systems B2O3 was used as sintering aid and/or dopant to improve properties [25–32]. Studies in BaTiO3 based compositions have reported that both dielectric and piezoelectric coefficients increase Correspondence to: Afyon Kocatepe University, Department of Materials Science and Engineering, ANS Kampusu, 03200 Afyonkarahisar, Turkey. E-mail address: metinozgul@aku.edu.tr (M. Ozgul). http://dx.doi.org/10.1016/j.ceramint.2016.09.073 Received 6 August 2016; Received in revised form 8 September 2016; Accepted 10 September 2016 Available online xxxx 0272-8842/ © 2016 Elsevier Ltd and Techna Group S.r.l. All rights reserved. Please cite this article as: Ozgul, M., Ceramics International (2016), http://dx.doi.org/10.1016/j.ceramint.2016.09.073 Ceramics International xx (xxxx) xxxx–xxxx M. Ozgul, A. Kucuk 3. Results and discussion with B2O3 addition mainly due to the lowered sintering temperature and formation of boron interstitials acting as donors [25,26]. Recently the influence of B2O3 addition in K0.5Na0.5NbO3 (KNN) ceramics has also been investigated and improved dielectric and piezoelectric properties reported [29,30,32]. However dielectric and piezoelectric properties of B2O3 added Bi0.5Na0.5TiO3 (BNT) ceramics have not been studied previously. In this work, 0.94(Bi0.5Na0.5)-0.06BaTiO3 (BNT6BT) was selected as a representative of MPB compositions in the BNT–BT system to study. BNT-6BT ceramics were doped with varied amounts of B2O3 and the influence of B3+ on the structure and electrical properties was examined. Fig. 1(a) shows the XRD diffraction pattern of ceramics as a function of boron (B) concentration (mol%) in BNT-6BT+xB ceramics, measured at room temperature and in the 2θ range of 20–70°. All samples possess a pure perovskite structure and no secondary phase was detected, implying that both Ba2+ and B3+ have diffused into BNT lattice to form a homogeneous solid solution. Fig. 1(b) shows the expanded XRD patterns of ceramics in the 2θ range of 39–47°. It is known that the rhombohedral symmetry of BNT at room temperature is characterized by a (003)/(021) peak splitting of between 38° and 42° and a single peak of (202) between 45° and 48°. For tetragonal symmetry, it is characterized by the profile of the (111) XRD line with a single peak in the range of 38.5–41° and splitting of the (002) and (200) peaks in the range of 44.5–47.5° [34]. As observed in Fig. 1(b), for samples of all compositions (x=0–2 mol% B) both (111) and (200) peaks were found to be split, demonstrating that these compositions were located in the MPB region, with the coexistence of rhombohedral and tetragonal phases. It is also worth to note that a substantial shift to higher 2θ angles was observed when boron (B3+) doping is 0.4 mol% and over. This may be due to a reduced lattice constant as new vacancies form in the perovskite crystal as a function of doping. There may be at least two different sources of vacancies in BNT-BT system as shown in the following defect chemistry Eqs. (1) and (2) based on Kröger-Vink notation: 2. Experimental Bi2O3 (% 99.9, Sigma-Aldrich), Na2CO3 (% 99.9 Merck), BaCO3 (% 99, Sigma-Aldrich), TiO2 (% 99.8, Sigma-Aldrich), and B2O3 (%99.999, Alfa Aesar) were used as raw materials to prepare 0.94(Bi0.5Na0.5)TiO30.06BaTiO3+(x mol%)B2O3 (x=0, 0.1, 0.2, 0.3, 0.4, 0.5, 1.0, 2.0) ceramics by the solid-state reaction method. The choice of the B3+ dopant concentrations in this study was made in light of earlier research for similar perovskite materials reported in the literature [25–32]. The amount of B2O3 was systematically increased from 0.1 to 2 mol% step by step to determine the optimum dopant concentration based on primary piezoelectric coefficient (d33) measurements for every composition. First, BNT-6BT powders were synthesized by using high purity Bi2O3, Na2CO3, BaCO3, and TiO2 based on the stoichiometric formula. The carbonates were dried at 150 °C for 4 h prior to use. The weighed batch was ball-milled in ethanol with 3 mm zirconia balls in a polyethylene bottle for 24 h, then dried, and calcined at 900 °C for 4 h. The phase formation of the calcined mixture was checked by x-ray diffraction technique (XRD). At the second step, pre-reacted BNT-6BT powders were mixed thoroughly with different amount of B2O3 by remilling to prepare compositions doped up to 2 mol%. After a second milling, powders of all the compositions were further mixed with 5 wt% polyvinyl alcohol (PVA) as a binder and then uniaxially pressed into discs with a diameter of 12 mm at 100 MPa pressure. Pre-shaped specimens were further densified by a cold isostatic press (CIP) at 150 MPa. After burning out PVA at 600 °C for 3 h with 1 °C/min heating rate, the green samples were placed on a Pt foil in an alumina crucible and sintered in air at 1150 °C for 12 h. Prolonged soaking time was chosen to ensure the full crystallization of B2O3 containing ceramics. Bulk density (ρb) of the sintered samples was measured by the Archimedes method (ASTM C373)[33]. As one of the easy methods for determining bulk density it involves measuring the dry weight (D) of the sample, boiling it in distilled water for 5 h, cooling it in water for 24 h, and measuring the suspended weight in water (S) and wet weight in air (W). The bulk density is then calculated as follows: ρb=D/V=D/ (W−S) where V represents the volume. Phase formation and the crystal structure of all the samples with different boron content were analyzed using an x-ray diffractometer (Bruker D8 Advance) with CuKα radiation (λ=1.5406 Å). In order to determine the influence of doping on phase structure of the ceramics, fine scanning was also recorded with a step interval of 0.02° in the range of 38–48°. The microstructure of the thermally etched (900 °C for 12 h) surfaces of sintered ceramics was analyzed by using a scanning electron microscope (LEO 1430VP). For electrical characterization, the parallel surfaces of samples were polished using 1200-grit SiC paper before applying a sputtered gold electrode. The thickness of samples used in this study ranged from 0.824 to 1.193 mm. The ceramics were poled for piezoelectric measurement in silicone oil bath at room temperature for 20 min and DC electric field strength of 30 kV/cm by using a high voltage supply amplifier/controller (Trek Model 610 E, Lockport, NY). The room temperature dielectric properties (εr and tan δ) were determined using an LCR-meter (Instek LCR-816) at a frequency of 1 kHz. The piezoelectric coefficients d33 of the samples were measured using a piezoelectric d33-meter (APC YE2730A d33 METER). Fig. 1. XRD patterns of undoped and doped BNT-6BT as a function of x=B3+ content (mol%) in the 2θ range of (a) 20–70° and (b) 39–47°. (Note the shift in peaks with increasing B3+% in (b)). 2 Ceramics International xx (xxxx) xxxx–xxxx M. Ozgul, A. Kucuk BaO B2 O3 Na2O • → BaNa + V/Na + O Ox BNT − BT → 2B••• + 6V/Na + 3O Ox i interstitial position in the BNT-6BT perovskite lattice [26]. As the amount of B3+ mol% increases, higher concentration of positive charges will be added to the system. These positively charged boron interstitials, B••• i , may suppress the negatively charged V′Na defects even formed in undoped BNT-6BT due to volatilization of Na+ alkali ions. However since each B••• can neutralize three of V′Na defects, excessive i boron doping may also cause the formation of sodium vacancies in the system. This is consistent with the shifting of (111) and (200) peaks towards higher 2θ angles when the boron dopant concentration is 0.4 mol% and over. Also in the case of boron doping, since B••• may i neutralize V′Na defects in BNT-6BT lattice, the need for the formation of oxygen vacancies, V•• O , will be reduced. Suppression of oxygen vacancies, V•• O , is known to influence both sintering and properties of piezoelectric ceramics [36–38]. (1) (2) The ionic radii of Ba2+, Na+, Bi3+, Ti4+, and B3+ are 0.160, 0.139, 0.128, 0.061, and 0.020 nm, respectively, so Ba2+ ions likely to substitute Na+ alkali ions with similar size creating negatively charged cation vacancies, V′Na [35]. Such excess negative charges will have to be balanced by positively charged defects in the system. One strong possibility is the formation of oxygen vacancies,V•• O , which would consequently increase total vacancy concentration in the crystal lattice. Another source of vacancies in the BNT-BT crystals in this study can be the boron impurity ion as a dopant. Due to its extremely small ionic radius of approximately 0.020 nm, B3+ ion is expected to occupy Fig. 2. SEM images of undoped and doped BNT-6BT ceramics as a function of B3+ content (mol%): (a) 0, (b) 0.1, (c) 0.2, (d) 0.3, (e) 0.4, (f) 0.5, (g) 1.0, (h) 2.0. (Scale bar is for 1 µm). 3 Ceramics International xx (xxxx) xxxx–xxxx M. Ozgul, A. Kucuk Fig. 2 shows the SEM micrographs of boron doped BNT-6BT ceramics as a function of dopant concentration. As can be seen from the figures, undoped and boron doped samples exhibit a dual microstructure containing large size grains (up to 6 µm) with smaller grains (as small as 1 µm). BNT-6BT samples doped with B in 0.3, 0.4, 1.0, and 2.0 mol% exhibits much more uniform microstructures. It can be concluded, with the exception of 0.5 mol%, that the average grain size of the ceramics is reduced when the content of boron is 0.3 mol% and over. Table 1 shows the density values and electrical properties as a function of dopant concentration in BNT-6BT ceramics sintered at 1150 °C for 12 h. To compare with the measured bulk densities, theoretical densities are calculated as a function of dopant percent by using reported theoretical densities of BNT-6BT and B2O3 given as 5.99 and 2.46 g/cm3, respectively [39,40]. Relative density values were also calculated from the ratio of measured bulk density to calculated theoretical density. As presented in Table 1 and shown in Fig.3, the density increases from 5.57 g/cm3 for the undoped samples to 5.67 g/ cm3 for the 0.1 mol% B doped samples. The equally highest density was obtained in 1 and 2 mol% doped samples as 5.71 g/cm3 (~96% relative density). Electrical properties (d33 and εr ) were also observed to be higher for these relatively high density samples. On the other hand, in contrast to the lower density value of 5.51 g/cm3, 0.5 mol% doped samples also exhibit improved electrical properties. B2O3 is usually reported to act as a glass-forming agent in barium titanate based ceramics. With its relatively low melting temperature, ~450 °C, B2O3 can form a liquid phase and promote densification during sintering Fig. 4. Piezoelectric coefficient (d33) values of undoped and doped BNT-6BT as a function of B3+ content (mol%). [25,26,30]. However, the purpose of B2O3 addition into BNT-6BT ceramics in the current study is not merely obtaining better densification. Besides improving the densification behavior, B3+ ions can also play a different role as a dopant. In terms of improving electrical properties both high density and donor doping can contribute. Fig. 4 indicates the room temperature piezoelectric coefficient (d33) values as a function of dopant concentration in BNT-6BT ceramics. If the dopant concentration is less than 0.3 mol%, a decrease in d33 values in contrast to undoped samples was observed. Measured d33 value of 145 pC/N for undoped samples dropped down to 136 pC/N for 0.2 mol % B doped samples. When samples doped with 0.3 mol% B, a substantial increase measured in d33 values reaching 156 pC/N. And then, with further increasing B addition, the d33 remarkably increased, giving a maximum value of 173 pC/N at 1 mol% B. Thereafter, d33 values started to decrease when B amount increased up to 2 mol%. It is clearly observed that d33 coefficient values were higher for all the compositions doped with 0.3 mol% and more B in contrast to undoped BNT-6BT. The effect of B doping can be explained in two ways; first one is its role as a sintering aid by providing liquid phase and the second one is its role as a dopant modifying crystal chemistry. The highest d33 value of 173 pC/N was obtained in the composition (1 mol% B) having the highest density of 5.71 g/cm3. For the 0.3 mol% B doped ceramics, the density and the d33 values were measured as 5.67 g/cm3 and 156 pC/N, respectively. However in contrast to their lower density values than 0.3 mol% B doped composition, higher d33 values of 161 and 163 pC/N were observed in 0.4 and 0.5 mol% B doped samples, respectively. This may be explained by the modified crystal chemistry of the samples as higher amount of B interstitials, B••• , formed in i heavily doped ceramics. These excessive positively charged ions will act as donors for the BNT-6BT system and improve the domain wall contribution to piezoelectric properties [36–38]. Since the higher amount of positively charged boron interstitials, B••• i , would form in the lattice structure, the need for oxygen vacancies, V•• O , will be diminished in 1 and 2 mol% B-doped BNT-6BT ceramics providing higher piezoelectric coefficients. This result also weakens the possibility of B3+ ions to substitute with Ti4+ ions,BT/i , which would create more oxygen vacancies, V•• O , instead of suppression. Fig. 5. shows the dielectric constant (εr ) and dielectric loss (tan δ) values as a function of dopant concentration in BNT-6BT ceramics, measured at room temperature and a frequency of 1 kHz. Both dielectric constant and loss values show an increase trend with the increase of B-dopant concentrations. Highest εr and tan δ values were recorded for 2 mol% B doped BNT-6BT ceramics. The significant increase in both coefficients is consistent with the well known effect of donor dopants in perovskite systems [1]. Table 1 Density values, piezoelectric, and dielectric property coefficients of undoped and doped BNT-6BT as a function of B3+ content (mol%). B3+ (mol %) ρth (g/ cm3) ρb (g/ cm3) ρr (=ρb/ρth) d33 (pC/ N) Ɛr (1 kHz) tan δ (1 kHz) 0.0 0.1 0.2 0.3 0.4 0.5 1.0 2.0 5.99 5.99 5.99 5.99 5.99 5.98 5.98 5.97 5.57 5.67 5.60 5.67 5.59 5.51 5.71 5.71 0.93 0.95 0.94 0.95 0.93 0.92 0.96 0.96 145 140 136 156 161 163 173 171 756 762 707 838 874 808 946 1075 0.0216 0.0264 0.0231 0.0272 0.0298 0.0269 0.0326 0.0421 ρth; Theoretical density, ρb; Measured bulk density, ρr; Relative density. Fig. 3. Bulk density (ρb) values of undoped and doped BNT-6BT as a function of B3+ content (mol%). 4 Ceramics International xx (xxxx) xxxx–xxxx M. Ozgul, A. Kucuk [11] C. Xu, D. Lin, K.W. Kwok, Structure, electrical properties and depolarization temperature of (Bi0.5Na0.5)TiO3-BaTiO3 lead-free piezoelectric ceramics, Solid State Sci. 10 (2008) 934–940. [12] Y. Guo, Y. Liu, R.L. Withers, F. Brink, H. Chen, Large electric field-induced strain and antiferroelectric behavior in (1−x)(Na0.5Bi0.5)TiO3–xBaTiO3 ceramics, Chem. Mater. 23 (2011) 219–228. [13] B.J. Chu, D.R. Chen, G.R. Li, Q.R. Yin, Electrical properties of Na1/2Bi1/2TiO3– BaTiO3 ceramics, J. Eur. Ceram. Soc. 22 (2002) 2115–2121. [14] D. Lin, D. Xiao, J. Zhu, P. Yu, Piezoelectric and ferroelectric properties of lead-free [Bi1−y(Na1−x−yLix)]0.5BayTiO3 ceramics, J. Eur. Ceram. Soc. 26 (2006) 3247–3251. [15] J. Shi, W. Yang, Piezoelectric and dielectric properties of CeO2-doped (Bi0.5Na0.5)0.94Ba0.06TiO3 lead-free ceramics, J. Alloy. Compd. 472 (2009) 267–270. [16] P. Jarupoom, K. Pengpat, N. Pisitpipathsin, S. Eitssayeam, U. Intatha, G. Rujijanagul, T. Tunkasiri, Development of electrical properties in lead- free bismuth sodium lanthanum titanate–barium titanate ceramic near the morphotropic phase boundary, Curr. Appl. Phys. 8 (2008) 253–257. [17] Y.Q. Yao, T.Y. Tseng, C.C. Chou, H.H.D. Chen, Phase transition and piezoelectric property of (Bi0.5Na0.5)0.94Ba0.06ZryTi1−yO3(y=0–0.04) ceramics, J. Appl. Phys. 102 (9) (2007) 094102. [18] E.A. Zereffa, A.V.P. Rao, Effect of B site substitutions of Nb on microstructure, phase transition temperatures and electrical properties of (Bi0.5Na0.5)(0.94)Ba0.06TiO3 ceramics, Ceram. Silik. 58 (1) (2014) 28–32. [19] R. Zuo, C. Ye, X. Fang, J. Li, Tantalum doped 0.94Bi0.5Na0.5TiO3–0.06BaTiO3 piezoelectric ceramics, J. Eur. Ceram. Soc. 28 (2008) 871–877. [20] Z.B. Yu, V.D. Krstic, B.K. Mukherjee, Microstructure and properties of lead-free (Bi0.5Na0.5)TiO3, J. Mater. Sci. 42 (2007) 3544–3551. [21] T. Chen, H. Wang, T. Zhang, G. Wang, J. Zhou, J. Zhang, Y. Liu, Effect of HfO2 content on the microstructure and piezoelectric properties of (Bi0.5Na0.5)0.94Ba0.06TiO3 lead-free ceramics, Ceram. Int. 40 (2014) 2959–2963. [22] Q. Xu, X.L. Chen, W. Chen, S.T. Chen, B. Kim, J. Lee, Synthesis, ferroelectric and piezoelectric properties of some (Na0.5Bi0.5)TiO3 system compositions, Mater. Lett. 59 (2005) 2437–2441. [23] L. Wu, D.Q. Xiao, D.M. Lin, J.G. Zhu, P. Yu, Synthesis and properties of [Bi0.5(Na1 −xAgx)0.5]1−yBayTiO3 piezoelectric ceramics, Jpn. J. Appl. Phys. 44 (2005) 8515–8518. [24] G.F. Fan, W.Z. Lu, X.H. Wang, F. Liang, Effects of manganese additive on piezoelectric properties of (Bi1/2Na1/2)TiO3-BaTiO3 ferroelectric ceramics, J. Mater. Sci. 42 (2) (2007) 472–476. [25] S.M. Rhim, S. Hong, H. Bak, O.K. Kim, Effects of B2O3 addition on the dielectric and ferroelectric properties of Ba0.7Sr0.3TiO3 ceramics, J. Am. Ceram. Soc. 83 (2000) 1145–1148. [26] J.Q. Qi, W.P. Chen, Y. Wang, H.L.W. Chan, L.T. Li, Dielectric properties of barium titanate ceramics doped by B2O3 vapor, J. Appl. Phys. 96 (2004) 6937–6939. [27] N. Tawichai, G. Rujijanagul, Influence of sintering temperature on dielectric and piezoelectric properties of B2O3 doped lead-free Ba(Ti0.9Sn0.1)O3 ceramics, Ferroelectrics 385 (2009) 128–134. [28] P. Jarupoom, K. Pengpat, G. Rujijanagul, Enhanced piezoelectric properties and lowered sintering temperature of Ba(Zr0.07Ti0.93)O3 by B2O3 addition, Curr. Appl. Phys. 10 (2010) 557–560. [29] R.X. Huang, Y.Z. Zhao, Y.J. Zhao, R.Z. Liu, H.P. Zhou, Low-temperature sintering of ZnO and B2O3 co-doped (K0.5Na0.5)NbO3 lead-free piezoelectric ceramics, Mater. Sci. Forum 687 (2011) 315–320. [30] P. Bharathi, K.B.R. Varma, Effect of the addition of B2O3 on the density, microstructure, dielectric, piezoelectric and ferroelectric properties of K0.5Na0.5NbO3 ceramics, J. Electron. Mater. 43 (2) (2014) 493–505. [31] S.H. Shin, J.D. Han, J. Yoo, Piezoelectric and dielectric properties of B2O3-added (Ba0.85Ca0.15) (Ti0.915Zr0.085)O3 ceramics sintered at low temperature, Mater. Lett. 154 (2015) 120–123. [32] K.P. Chen, F.L. Zhang, J.Q. Zhou, X.W. Zhang, C.W. Li, L.N. An, Effect of borax addition on sintering and electrical properties of (K0.5Na0.5)NbO3 lead-free piezoceramics, Ceram. Int. 41 (2015) 10232–10236. [33] M. Bengisu, Engineering Ceramics, Springer, Berlin, 2001. [34] Y. Xua, X. Liuc, G. Wanga, X. Liua, Y. Fengb, Antiferroelectricity in tantalum doped (Bi0.5Na0.5)0.94Ba0.06TiO3 lead-free ceramics, Ceram. Int. 42 (2016) 4313–4322. [35] Y. Qu, D. Shan, J. Song, Effect of A-site substitution on crystal component and dielectric properties in Bi0.5Na0.5TiO3 ceramics, Mater. Sci. Eng. B 121 (2005) 148–151. [36] W.L. Warren, K. Vanheusden, D. Dimos, G.E. Pike, B.A. Tuttle, Oxygen vacancy motion in perovskite oxides, J. Am. Ceram. Soc. 79 (1996) 536–538. [37] J. Kim, H. Lee, Lattice distortion and electrical properties of x(Na0.5K0.5)NbO3–(1 −x)BaTiO3 dielectrics, J. Mater. Sci. – Mater. Electron. 27 (2016) 2315–2320. [38] S. Swain, S.K. Kar, P. Kumar, Dielectric, optical, piezoelectric and ferroelectric studies of NBT–BT ceramics near MPB, Ceram. Int. 41 (2015) 10710–10717. [39] J.-F. Trelcat, C. Courtois, M. Rguiti, A. Leriche, P.H. Duvigneaud, T. Segato, Morphotropic phase boundary in the BNT-BT-BKT system, Ceram. Int. 38 (2012) 2823–2827. [40] A. Andrews, M. Herrmann, T.C. Shabalala, I. Sigalas, Liquid phase assisted hot pressing of boron suboxide materials, J. Eur. Ceram. Soc. 28 (2008) 1613–1621. Fig. 5. Dielectric constant (εr ) and dielectric loss (tan δ) values of undoped and doped BNT-6BT as a function of B3+ content (mol%). 4. Conclusions Phase structure, density, microstructure, dielectric and piezoelectric properties of undoped and up to 2 mol% B2O3 doped 0.94 (Bi0.5Na0.5)TiO3-0.06BaTiO3 (BNT-6BT) ceramics sintered at 1150 °C for 12 h have been investigated as a candidate for lead-free piezoelectric ceramics. The incorporation of B3+ ions into the perovskite structure resulted in an increase in piezoelectric coefficient (d33) from 145 to 173 pC/N. Dielectric constant (εr ) and loss (tan δ) values also showed an enormous jump from 756 to 1075 and 0.0216–0.0421, respectively with doping. The effect of doping on density values was about 2.5% improvement for 1 and 2 mol% B2O3 containing ceramics. The results indicate that, besides its well known role as a sintering additive, B2O3 is a very effective dopant in improving the piezoelectric properties of BNT-6BT by modifying the crystal defect chemistry. As the density values are slightly affected, a remarkable increase of all the dielectric and piezoelectric parameters measured in this study was attributed to the donor dopant role of B3+ interstitial ions forming B••• i . Acknowledgments This work was supported by the Department of Scientific Research Projects (BAPK) of Afyon Kocatepe University, Turkey (Project funding no. 15.FEN.BIL.09). References [1] B. Jaffe, W.R. Cook, H. Jaffe, Piezoelectric Ceramics, Academic Press, New York, 1971. [2] G.H. Haertling, Ferroelectric ceramics: history and technology, J. Am. Ceram. Soc. 82 (1999) 797–818. [3] K. Uchino, Ferroelectric Devices, CRC Press, New York, 2009. [4] P.K. Panda, B. Sahoo, PZT to lead free piezo ceramics: a review, Ferroelectrics 474 (2015) 128–143. [5] Y. Saito, H. Takao, T. Tani, T. Nonoyama, K. Takatori, T. Homma, T. Nagaya, M. Nakamura, Lead-free piezoceramics, Nature 432 (2004) 84–87. [6] L. Yi, K. Moon, C.P. Wong, Electronics without lead, Science 308 (2005) 1419–1420. [7] T.R. Shrout, S.J. Zhang, Lead-free piezoelectric ceramics: alternatives for PZT?, J. Electroceram. 19 (2007) 111–124. [8] P.K. Panda, Review: environmental friendly lead-free piezoelectric materials, J. Mater. Sci. 44 (2009) 5049–5062. [9] J. Rödel, W. Jo, K.T.P. Seifert, E.M. Anton, T. Granzow, D. Damjanovic, Perspective on the development of lead-free piezoceramics, J. Am. Ceram. Soc. 92 (2009) 1153–1177. [10] T. Takenaka, K. Maruyama, K. Sakata, (Bi1/2Na1/2)TiO3–BaTiO3 system for leadfree piezoelectric ceramics, Jpn. J. Appl. Phys. 30 (1991) 2236–2239. 5