Logo Passei Direto

04-Structure and electrical properties of __Na0 825K0 175_0 5Bi0 5_1_xTiO3 piezoceramics with A-s

User badge image
Heme Lata

in

Ferramentas de estudo

Material
Study with thousands of resources!

Text Material Preview

Full Terms & Conditions of access and use can be found at
https://www.tandfonline.com/action/journalInformation?journalCode=tace20
Journal of Asian Ceramic Societies
ISSN: (Print) (Online) Journal homepage: www.tandfonline.com/journals/tace20
Structure and electrical properties of
[(Na0.825K0.175)0.5Bi0.5]1+xTiO3 piezoceramics with A-
site nonstoichiometry
Xiaoming Chen, Yunwen Liao & Guorong Li
To cite this article: Xiaoming Chen, Yunwen Liao & Guorong Li (2020) Structure and electrical
properties of [(Na0.825K0.175)0.5Bi0.5]1+xTiO3 piezoceramics with A-site nonstoichiometry, Journal
of Asian Ceramic Societies, 8:4, 1036-1042, DOI: 10.1080/21870764.2020.1806192
To link to this article: https://doi.org/10.1080/21870764.2020.1806192
© 2020 The Author(s). Published by Informa
UK Limited, trading as Taylor & Francis
Group on behalf of The Korean Ceramic
Society and The Ceramic Society of Japan.
Published online: 18 Aug 2020.
Submit your article to this journal 
Article views: 931
View related articles 
View Crossmark data
Citing articles: 2 View citing articles 
https://www.tandfonline.com/action/journalInformation?journalCode=tace20
https://www.tandfonline.com/journals/tace20?src=pdf
https://www.tandfonline.com/action/showCitFormats?doi=10.1080/21870764.2020.1806192
https://doi.org/10.1080/21870764.2020.1806192
https://www.tandfonline.com/action/authorSubmission?journalCode=tace20&show=instructions&src=pdf
https://www.tandfonline.com/action/authorSubmission?journalCode=tace20&show=instructions&src=pdf
https://www.tandfonline.com/doi/mlt/10.1080/21870764.2020.1806192?src=pdf
https://www.tandfonline.com/doi/mlt/10.1080/21870764.2020.1806192?src=pdf
http://crossmark.crossref.org/dialog/?doi=10.1080/21870764.2020.1806192&domain=pdf&date_stamp=18 Aug 2020
http://crossmark.crossref.org/dialog/?doi=10.1080/21870764.2020.1806192&domain=pdf&date_stamp=18 Aug 2020
https://www.tandfonline.com/doi/citedby/10.1080/21870764.2020.1806192?src=pdf
https://www.tandfonline.com/doi/citedby/10.1080/21870764.2020.1806192?src=pdf
FULL LENGTH ARTICLE
Structure and electrical properties of [(Na0.825K0.175)0.5Bi0.5]1+xTiO3 
piezoceramics with A-site nonstoichiometry
Xiaoming Chena,b, Yunwen Liaoc and Guorong Lid
aKey Laboratory of Light Metal Materials Processing Technology of Guizhou Province, Guizhou Institute of Technology, Guiyang, China; 
bSchool of Materials and Energy, Guizhou Institute of Technology, Guiyang, China; cCollege of Chemistry and Chemical Engineering, China 
West Normal University, Nanchong, China; dShanghai Institute of Ceramics, Chinese Academy of Science, Shanghai, China
ABSTRACT
A type of nonstoichiometric piezoceramics, [(Na0.825K0.175)0.5Bi0.5]1+x TiO3 (BKNT-x, x = 0–0.009), 
was synthesized by the conventional sintering method. The effects of Bi3+, Na+ and K+ content 
on the structure and electrical properties of 0.825Bi0.5Na0.5TiO3-0.175Bi0.5K0.5TiO3 piezocera-
mics were investigated.The XRD patterns demonstrated that BKNT-x piezoceramics form a solid 
solution without a secondary phase. Meanwhile, SEM micrographs indicated that excess 
[(Na0.825K0.175)0.5Bi0.5]2+ suppresses grain growth and that BKNT-x ceramics become consider-
ably denser. BKNT-0.002 ceramics offer the following enhanced properties: d33 = 154 pC/N, kp 
= 33%, εr = 1087, tanδ = 0.036, and Qm = 116. This improvement of properties is most likely 
related to the dense structure and the maintenance of morphotropic phase boundaries 
through the introduction of a proper A-site cation excess. For ceramics containing volatile 
elements, stoichiometric control is an efficacious method of enhancing performance.
ARTICLE HISTORY 
Received 3 March 2020 
Accepted 2 August 2020 
KEYWORDS 
Structure; X-ray diffraction; 
piezoelectric properties; Bi0.5 
Na0.5TiO3; ceramics; oxygen 
vacancy
1. Introduction
To date, lead-free piezoceramics have received a great 
amount of attention from researchers due to their envir-
onmental friendliness. Smolenski et al [1]. reported Na0.5 
Bi0.5TiO3 (BNT) in 1960. At room temperature, rhombo-
hedral BNT ceramic shows a high Curie point and rela-
tively good electrical properties [2,3], which have made 
BNT ceramics become one of promising lead-free cera-
mics. BNT ceramics have a large coercive field Ec (73kv/ 
cm), however, that makes ceramic poling difficult [4,5]. 
A great deal of modification work has been conducted 
since the early 1990s. Other approaches have been 
proved to be helpful in the poling of the ceramics, as 
well as, in improving their electrical properties, include 
the formation of new solid solutions, such as, Bi0.5(Kx 
Na1-x-yLiy)0.5TiO3 [6], BaTiO3(BT)-BNT [3,7], BNT-Na NbO3 
[8], K0.5Bi0.5TiO3(BKT)-BNT [5], BT- BNT-BKT [9], BNT- 
BaTiO3-NaNbO3 [10], BKT-BNT-SrTiO3 [11], Bi0.5(K0.20 
Na0.70Li0.10)0.5TiO3 [12], Bi0.5(KxNa1-x-Agy)0.5 TiO3 [13], 
and BNT-BT-(Na0.5K0.5)NbO3 [14], and doping with 
metal oxides, such as, Dy2O3-doped (Na0.7K0.3)0.5Bi0.5 
TiO3 [15], Gd2O3-doped (Na0.5 Bi0.5)0.94Ba0.06TiO3 [16], 
Li2O-doped0.825BNT-0.175BKT [17], La2O3-doped0. 
92BNT-0.08BT [18]. Among these systems, BKT-BNT 
(BNKT) was observed by Sasaki et al [5]. They found 
that it has relatively strong piezoelectric performance 
owing to the existence of morphotropic phase bound-
aries (MPB, x=0.16-0.2). However, properties of BNT- 
based piezoceramics are usually affected by the 
volatilization of A-site cations for the sintering process. 
Zuo R et al [19]. revealed that the sintering behavior of 
(Na0.5Bi0.5)1+xTiO3 ceramics is sensitive to variants of 
A-site cations that contribute to enhancing its electrical 
properties(x=0.01, d33=83pC/N). Sung et al [20,21]. 
revealed that the effect of Bi nonstoichiometry on d33 
and Td of BNT is different from the results of Na non-
stoichiometry. Wu et al [22]. found that a proper Na/K 
excess shows good electrical properties in 0.85 BNT- 
0.15BKT systems (Pr=13.6 μC/cm2, d33* = 56 pm/V). 
Prasertpalichat [23] demonstrated that changes in Bi/ 
Na stoichiometry have a significant influence on the 
electrical properties of (Bi0.5Na0.5)0.94Ba0.06TiO3 ceramics. 
Chu et al [24]. also observed that changes in the A-site 
cations can effectively enhance the electrical properties 
of nonstoichiometric (Bi0.5Na0.5)0.92Ba0.08TiO3 systems 
(d33=149 pC/N, ε= 1230). Ni et al [25]. reported that 
Bi0.505(K0.18Na0.82)0.485TiO3 ceramics show relatively 
good piezoelectric properties when tuned with A-site 
vacancies (d33=151pC/N, Pr=29.5μC/cm2). In order to 
clarify nonstoichiometry effects, therefore, it is signifi-
cant to investigate the connection between the proper-
ties and the amount of A-site cations in BNT-based 
piezoceramics.
In this case, a type of lead-free [(Na0.825K0.175)0.5 
Bi0.5]1+xTiO3 piezoceramics was synthesized using 
a conventional mixed-oxide method. The influence of 
[(Na0.825K0.175)0.5Bi0.5]2+ nonstoichiometry on the 
microstructure, crystal phase and electrical properties 
CONTACT Xiaoming Chen chen-xm123@163.com Key Laboratory of Light Metal Materials Processing Technology of Guizhou Province, Guizhou 
Institute of Technology, Guiyang 550003, China
JOURNAL OF ASIAN CERAMIC SOCIETIES 
2020, VOL. 8, NO. 4, 1036–1042 
https://doi.org/10.1080/21870764.2020.1806192
© 2020 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group on behalf of The Korean Ceramic Society and The Ceramic Society of Japan. 
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/), which permits 
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
http://www.tandfonline.com
https://crossmark.crossref.org/dialog/?doi=10.1080/21870764.2020.1806192&domain=pdf&date_stamp=2020-12-05
of (Na0.825K0.175)0.5Bi0.5TiO3 (BKNT) piezoceramics was 
investigated in detail.
2. Experimental procedure
Piezoceramics [(Na0.825K0.175)0.5Bi0.5]1+xTiO3 (x = 0, 
0.002, 0.004, 0.006,0.009) were synthesized via 
a conventional solid state process. Bi2O3 (Chenguang 
Chemical Co., Ltd, Qishan, China, AR, 99.9%), K2CO3 
(Xilong Chemical Co., Ltd, Shantou, China, AR, 99.8%), 
Na2CO3 (Xilong Chemical Co., Ltd, Shantou, China, AR, 
99.9%), TiO2 (Zhongxing Electronic Materials Co., Ltd, 
Xiantao, China, AR, 99.5%) powders were used as raw 
materials with their water removed at 120 oC/24 h. 
According to the desired stoichiometry, all the pow-
ders were blended with anhydrous ethanol, and then 
ball-milled at a rate of 350 rotations/min for 7–8 h. 
When the powders become dry under 75 
oC conditions, the mixture was calcined at 850 °C/2 
h. The heating rate was 3 min/oC. Re-milled samples 
were blended with polyvinyl alcohol (5 wt%), and then 
pressed into discs measuring 10 mm in diameter and 
1.2 mm in thickness under 20 Mpa pressure. The 
obtained samples were sintered at 1180 oC/2 h at 
a heating rate of 5 min/oC. In order to form electrodes, 
a coating of silver conducting paste (PC-Ag-8080, Xizhi 
Electronic Materials Co., Ltd, China) was applied to two 
sides of polished compositions, which were fired at 600 
°C. In a silicone oil bath (80 oC), the annealed composi-
tions (9 mm in diameter, 1 mm in thickness) were 
poled with a DC electric field (4.5 kV/mm) for 45 min.
The sintered compositions had been determined by 
XRD (Cu Kα radiation, Rigaku Dmax/RB). The lattice 
parameters of the BKNT-x ceramics were refined by 
the Rietveld refinement method [26] and MAUD soft-
ware (version 2.94) [27–29]. The weighted profile resi-
dual factor (Rwp) was less than 15%, indicating good 
agreement of refinement. The density of the composi-
tions was calculated by the Archimedes method. The 
microstructures of the samples were characterized by 
SEM (JSM-6510, Japan). The average grain size was 
obtained by measuring the number of grains at the 
intersection lines. The measurement of piezoelectric 
properties was carried out with a d33 meter (ZJ-3A, 
Institute of Acoustics, China). The dielectric properties 
of the BKNT-x piezoceramics were measured with an 
impedance analyzer (Agilent 4294A,1 KHz) at 25–500 ° 
C. The kp and Qm were calculated by use of Onoe’s 
formula [30], employing the resonance and antireso-
nance frequencies method as a basis.
3. Results and discussion
Figure 1(a) displays the XRD patterns of BKNT-x cera-
mics. All the compositions possess typical diffraction 
peaks displayed by an ABO3-type structure. There is no 
impurity phase in these solid solutions. Ni et al [25]. 
observed that when x � 0.02, second phase Bi2Ti2O7 
forms in (K0.18Na0.82)0.5–3xBi0.5+xTiO3 ceramics. Babu 
et al. [31]
also demonstrated that excess K2CO3 (x ≤ 0.8 mol%) 
makes up for the volatility of K+ to maintain the phase 
purity of (Na0.8K0.2)0.5Bi0.5TiO3 piezoceramics. Seo et al 
[32]. observed, moreover, that Ba0.06(Bi0.5Na0.5)0.94+x 
TiO3 (−0.01 ≤ x ≤ 0.02) piezoceramics possess a homo-
geneous structure with phase purity. Our results are 
consistent with previous reports, which showed that 
excess Bi3+, Na+, and K+(x ≤ 0.009) may keep (Na0.825 
Figure 1. (a)XRD Patterns of various compositions in the BKNT-x ceramics; (b) and (c) The expanded XRD patterns of BKNT-x 
ceramics in the range of 39–47°.
JOURNAL OF ASIAN CERAMIC SOCIETIES 1037
K0.175)+/Bi3+ at 1:1 and maintain the phase purity of 
BKNT-x ceramics after the A-site cations of BKNT-x 
ceramics partly volatilize in the sintering process.
Sasaki et al [5]. reported that (Na1-xKx)0.5Bi0.5TiO3 
ceramics (BNKT) exhibit high performance due to the 
R-T MPB (0.16 ≤ x ≤ 0.2), which has a similar pseudo-
cubic structure [33]. To investigate the phase structure 
of BKNT-x ceramics, the magnified XRD patterns are 
given in the 2ɵ range of 39–47°, as shown in Figure 1(b, 
c). According to the R3 c space group (COD ID: 
2103295) and P4bm space group (COD ID: 2102068) 
[34,35], the structure could be designated by the fol-
lowing peaks: rhombohedral (003) and tetragonal 
(201) peaks at 40°, as well as rhombohedral(202) and 
tetragonal (002) peaks at 46°, repectively. The peak 
splitting reveals that BKNT-x ceramics are still in the 
rhombohedral(R)-tetragonal(T) phase transition range. 
In addition, the (002) peak moves toward lower angles 
with increases in x. The lattice parameters (aT, cT) and 
cell volume of the tetragonal phase of BKNT-x ceramics 
increase slightly with changes in the content of A-site 
cations, as shown in Table 1. In the sintering process, 
oxygen vacancies of BKNT ceramics are usually gener-
ated due to the volatilization of A-site cations. The 
addition of excess A-site cations (Bi3+, Na+, and K+, 
hereinafter BKN) balances the effect of BKN loss, 
which leads to a reduction in the number of oxygen 
vacancies. It is generally known that large numbers of 
oxygen vacancies can cause unit cells to shrink and 
lead to a shift of the diffraction peaks to high angles 
[36–39]. Therefore, the low-angle movement of (002) 
peaks for BKNT-x ceramics indicates a decrease in the 
number of oxygen vacancies.
Figure 2 displays surface micrographs of BKNT-x 
ceramics sintered at 1180 oC. The crystal grains of 
BKNT-x piezoceramics show a regular square shape. 
With increases in BKN, the average grain size decreases 
slightly. The mean crystallite sizes are about 1.28, 1.25, 
0.98, 0.76 μm, respectively. This can probably be 
ascribed to the collection of excess BNK at the A-site 
lattice of ABO3 perovskite structure due to volatiliza-
tion of A-site cations (Bi3+, Na+, and K+) at high sinter-
ing temperatures.This behavior can reduce oxygen 
vacancies, while the existence of the defects is condu-
cive to mass transport, accelerating the grain growth 
[40]. As a result, crystallite growth is restrained. This is 
similar to a previous report on (Bi0.5+xNa0.5)0.94Ba0.06 
TiO3 systems with A-site nonstoichiometry [41]. It can 
be observed, from Figure 3, that with the concentra-
tion of BKN, the relative densities of BKNT-x piezocera-
mics reach a maximum, and then descend, as was 
achieved by the theoretical density of (Na0.8K0.2)0.5 
Bi0.5TiO3 (5.889 g/cm3) [42]. Appropriate BKN excesses 
are able to decrease the evaporation of volatile A-site 
cations, and contribute to an improvement in the den-
sity of BKNT-x ceramics [43]. Bi2O3 with its melting 
Table 1. Lattice Parameters and lattice volume of BKNT-x 
ceramics.
x aT /Å cT /Å VT/Å3
0 5.51298 3.89562 118.39938
0.002 5.51347 3.89552 118.41739
0.004 5.51477 3.89960 118.59732
0.006 5.51489 3.89980 118.60856
0.009 5.51685 3.89855 118.65484
Figure 2. SEM surface micrographs of BKNT-x ceramics with various compositions sintered at 1180 oC for 2 hours.
1038 X. CHEN ET AL.
point of 825°C usually decreases sintering tempera-
tures by forming a liquid phase [44]. With further 
increases in BKN(x > 0.004), the content of bismuth 
oxide goes up, which leads to lower sintering tempera-
tures of BKNT-x piezoceramics. For this reason, the 
ceramics can’t be fully sintered at the same sintering 
temperature (1180 oC), resulting in decreased com-
pactness. The relative densities of BKNT-x piezocera-
mics are more than 95% of the theoretical values when 
0.002 � x � 0.004. A dense microstructure is particu-
larly advantageous to improvement of the electrical 
properties of BKNT-x ceramics.
Figure 4 displays the temperature dependance of εr 
and tanδ for unpoled BKNT-x ceramics under different 
frequencies. BKNT-x ceramics have two dielectric 
anomalies: a relaxor antiferroelectric-antiferroelectric 
phase transformation and an antiferroelectric- 
paraelectric phase transformation corresponding to 
TRE and Tm, respectively, which are consistent with 
NBT-based piezoceramics [5,45–48]. A phenomenon 
of mergence is exhibited by the different εr-T curves 
of BKNT-x piezoceramics, where the temperature is 
described as TRE. Frequency has a great effect on εr 
when T< TRE. Ma et al. reported that nanodomains with 
a short-range order stimulate the relaxor behavior of 
BNT-basedceramics and suggested a new term, 
“relaxor antiferroelectric,” to describe the phenom-
enon of frequency dispersion [46]. With further incre-
ments in temperature, the relative permittivity εr 
reaches a maximum value at the second dielectric 
anomaly, where the temperature is expressed as Tm. 
Ma et al. identified a long-range antiferroelectric(AFE) 
order presents in BNT-based ceramics at between TRE 
and Tm, and showed that the paraelectric phase is 
located at temperatures above Tm [46]. As soon as 
the temperature rises above Tm, tanδ increases greatly 
due to the considerable leakage conduction [49]. In 
addition, pure BKNT ceramics exhibit larger values for 
dielectric constants and loss tangents than BKNT-x 
ceramics at high temperatures (T > Tm) and low fre-
quencies, which can be ascribed to the fact that 
Figure 3. Relative density of BKNT-x ceramics sintered at 1180 
oC / 2 h.
Figure 4. Temperature dependence of relative permittivity and dielectric loss for BKNT-x ceramics (1180°C/2 h).
JOURNAL OF ASIAN CERAMIC SOCIETIES 1039
A-site vacancies and oxygen vacancies form at pre-
paration process and lead to the appearance of fre-
quency dispersion for pure BKNT ceramics [46].
Figure 5 shows the electrical properties of BKNT-x 
piezoceramics. When the content of BKN is 0.2 mol%, 
the maximum values of the piezoelectric constant d33 
and electromechanical coupling factor kp reach 154 
pC/N and 33%, respectively. The comprehensive per-
formance of BKNT-x piezoceramics is slightly superior 
to those of NBT-based ceramics with nonstoichiometry 
(see Table 2). This can be ascribed to the following 
reasons: a proper BKN excess improves the compact-
ability of BKNT-x piezoceramics which makes the polar-
ization of ceramics more completely under a broad 
electrical field [17]. Moreover, an excess amount of 
BKN can make up for the volatilization of A-site cations 
and maintain the R-T phase fractions of compositions 
around MPB, which improves the piezoelectricity of 
BKNT-x ceramics. Analogous to the close dependence 
of BKN content and d33, the maximum value of the 
relative dielectric constant εr (1087) occurs at x=0.002. 
Meanwhile, the dielectric loss tanδ initially decreases 
(x<0.002) and reaches a minimum value (0.0362), and 
then, finally increases with the concentration of BKN. 
Qm gradually decreases with the increases in BKN. This 
can be ascribed to the decrement of oxygen vacancies 
caused by the compensation of excess BKN for the loss 
of Bi3+, Na+ and K+.
The temperature dependence of the kp of BKNT-x 
ceramics is demonstrated in Figure 6. The kp of BKNT-x 
ceramics shows a decreasing trend with increases in 
the annealing temperature. The kp changes signifi-
cantly as the annealing temperature increases up to 
134 oC (x=0.004, kp=0.03), and then decreases as the 
annealing temperature increases further. When the 
annealing temperature is 134-140 oC, the piezoelectric 
responses of BKNT-0.004 ceramics almost vanish. In 
other words, the depolarization temperature (Td) of 
BKNT-0.004 ceramics is about 134 oC, while pure 
BKNT ceramics have a high Td of around 140 oC, as 
reported by Sapper [50]. As a result, the addition of 
excess BKN has a negative effect on Td. On the other 
hand, variations in Td can reflect changes in the oxygen 
vacancies. The depolarization temperature of ceramics 
increases at a high oxygen vacancy concentration 
owing to a strong coaction between the oxygen vacan-
cies and the domains [41]. Therefore, use of moderate 
amounts of BKN to compensate for the loss of Na+, Bi3+ 
Figure 5. Piezoelectric and dielectric properties of BKNT-x ceramics.
Table 2. Comparison of the properties of different NBT-based 
ceramics with nonstoichiometry.
systems
d33 
(pC/ 
N) kp Td(oC) References
0.94(Na0.5Bi0.52)TiO3-0.06BaTiO3 118 0.24 29 [41]
0.8(Na0.5Bi0.5)TiO3-0.2(K0.5Bi0.5)TiO3 
+ 0.8 mol% K2CO3
196 - - [31]
Bi0.51(Na0.82K0.18)0.5TiO3 145 0.31 ~125 [25]
Bi0.505(Na0.82K0.18)0.485TiO3 151 0.298 101 [25]
[(Na0.825K0.175)0.5Bi0.5]1.002TiO3 154 0.33 137 This work
Figure 6. The temperature dependency of kp for BKNT-x 
ceramics.
1040 X. CHEN ET AL.
and K+ during the sintering step contributes to 
a decrease in oxygen vacancies that causes Td to shift 
to a low temperature.
4. Conclusions
The influence of BKN on the structure, crystal phase and 
electrical properties of (Na0.825K0.175)0.5Bi0.5TiO3 piezo-
ceramics, which were synthesized by conventional sin-
tering, were systematically researched. All the BKNT-x 
piezoceramics have typical diffraction peaks with pure 
perovskite structures. Excess BKN compensates for the 
loss of the A-site cations, and leads to a decrease in 
oxygen vacancies which restrain the growth of crystal-
lites. All the compositions are well sintered with dense 
microstructure at 1180 oC/2 h when 0.002 � x � 0.004. 
BKNT-x piezoceramics with BKN compensation achieve 
high d33 (154 pC/N) and kp (33%). The electrical proper-
ties of BKNT-x ceramics have been enhanced by the 
introduction of excess BKN, which maintains the struc-
tural density and the R-T phase fractions of the samples 
at around MPB. As a result, in order to acquire the 
desirable electrical properties, it is essential to control 
the content of volatile A-site cations sufficiently.
Disclosure statement
No potential conflict of interest was reported by the authors.
Funding
This work was sponsored by the Natural Science Foundation 
of Department of Education of Guizhou Province, China [No. 
KY [2018]253]; Science and Technology Fund of Guizhou 
Province, China [[2020]1Y204]; and Doctor Start-up 
Foundation of Guizhou Institute of Technology [No. 
XJGC20190920].
References
[1] Smolenskii GA, Isupov VA, Agranovskaya AI, et al. 
Ferroelectrics with diffuse phase transitions. Soviet 
Phys Solid State. 1961;2:2584.
[2] Takenaka T, Sakata K. Dielectric, piezoelectric and 
pyroelectric properties of (BiNa)1/2TiO3- based 
ceramics. Ferroelectrics. 1989;95(1):153–156.
[3] Tadashi T, Kei-ichi M, Koichiro S. (Bi1/2Na1/2)TiO3- 
BaTiO3 system for lead-free piezoelectric ceramics. 
Jpn J Appl Phys. 1991;30(9S):2236–2239.
[4] Hajime N. and Tadashi T. Lead-free piezoelectric cera-
mics of (NaBi)0.5TiO3-1/2(Bi2O3·Sc2O3) system. Jpn 
J Appl Phys. 1997;36(1):6055–6057.
[5] Sasaki A, Chiba T, Mamiya Y, et al. Dielectric and piezo-
electric properties of (Bi0.5Na0.5)TiO3–(Bi0.5K0.5)TiO3 
systems. Jpn J Appl Phys. 1999;38(9S):5564.
[6] Lin D, Xiao D, Zhu J, et al. Synthesis and piezoelectric 
properties of lead-free piezoelectric [Bi0.5(Na1−x−yKx 
Liy)0.5]TiO3 ceramics. Mater Lett. 2004;58(5):615–618.
[7] Zeng WD, Li QN, Zhou CR, et al. A new insight into 
structural complexity in ferroelectric ceramics. J Adv 
Ceram. 2017;6(3):262–268.
[8] Takenaka T, Okuda T, Takegahara K. Lead-free piezo-
electric ceramics based on (Bi1/2Na1/2)TiO3-NaNbO3. 
Ferroelectrics. 1997;196(1–4):495–498.
[9] Hajime N, Masaki Y, Yoichi M, et al. Large piezoelectric 
constant and high curie temperature of lead-free piezo-
electric ceramic ternary system based on bismuth 
sodium titanate-bismuth potassium titanate-barium 
titanate near the morphotropic phase boundary. Jpn 
J Appl Phys. 2003;42(12R):7401–7403.
[10] Xu Q, Lanagan MT, Luo W, et al. Electrical properties 
and relaxation behavior of Bi0.5Na0.5TiO3-BaTiO3 cera-
mics modified with NaNbO3. J Eur Ceram Soc. 2016;36 
(10):2469–2477. .
[11] Yoo J, Oh D, Jeong Y, et al. Dielectric and piezoelectric 
characteristics of lead-free Bi0.5(Na0.84K0.16)0.5TiO3 cera-
mics substituted with Sr. Mater Lett. 2004;58 
(29):3831–3835.
[12] Guo K, Sharifzadeh Mirshekarloo M, Lin M, et al. 
Microstructure and piezoelectric properties of thermal 
sprayed Bi0.5(Na0.70K0.20Li0.10)0.5TiO3 ceramic coatings. 
Ceram Int. 2019;45(3):3570–3573.
[13] Liao Y, Xiao D, Lin D, et al. Synthesis and properties of 
Bi0.5(Na1−x−yKxAgy)0.5TiO3 lead-free piezoelectric cera-
mics. Ceram Int. 2007;33(8):1445–1448.
[14] Shi X, Kumar N, Hoffman M. Electric field-temperaturephase diagrams for (Bi1/2Na1/2)TiO3-BaTiO3-(K1/2 
Na1/2)NbO3 relaxor ceramics. J Mater Chem C. 2018;6 
(45):12224–12233.
[15] Muneeswaran M, Choi BC, Chang SH, et al. Effect of 
dysprosium doping on structural and vibrational prop-
erties of lead-free (Na0.7K0.3)0.5Bi0.5TiO3 ferroelectric 
ceramics. Ceram Int. 2017;43(16):13696–13701.
[16] Turki O, Slimani A, Seveyrat L, et al. Structural, dielec-
tric, ferroelectric, and electrocaloric properties of 2% 
Gd2O3 doping (Na0.5Bi0.5)0.94Ba0.06TiO3 ceramics. 
J Appl Phys. 2016;120(5):054102.
[17] Chen X, Liao Y, Mao L, et al. Microstructure and piezo-
electric properties of Li-doped Bi0.5(Na0.825K0.175) 
0.5TiO3 piezoelectric ceramics. Phys Status Solidi A. 
2009;206(7):1616–1619.
[18] Liu L, Zhu M, Hou Y, et al. Abnormal piezoelectric and 
dielectric behavior of 0.92Na0.5Bi0.5TiO3-0.08BaTiO3 
induced by La doping. J Mater Res. 2011;22 
(5):1188–1192.
[19] Zuo R, Su S, Wu Y, et al. Influence of A-site nonstoi-
chiometry on sintering, microstructure and electrical 
properties of (Bi0.5Na0.5)TiO3 ceramics. Mater Chem 
Phys. 2008;110(2):311–315.
[20] Sung YS, Kim JM, Cho JH, et al. Effects of Bi nonstoi-
chiometry in (Bi0.5+xNa)TiO3 ceramics. Appl Phys Lett. 
2011;98(1):012902.
[21] Sung YS, Kim JM, Cho JH, et al. Effects of Na nonstoi-
chiometry in (Bi0.5Na0.5+x)TiO3 ceramics. Appl Phys 
Lett. 2010;96(2):022901.
[22] Wu Y, Wang X, Zhong C, et al. Effect of Na/K excess on 
the electrical properties of Na0.5Bi0.5TiO3–K0.5Bi0.5TiO3 
thin films prepared by sol–gel processing. Thin Solid 
Films. 2011;519(15):4798–4803.
[23] Prasertpalichat S, Schmidt W, Cann DP. Effects of A-site 
nonstoichiometry on oxide ion conduction in 0.94Bi0.5 
Na0.5TiO3-0.06BaTiO3 ceramics. J Adv Dielectr. 
2016;6:2.
JOURNAL OF ASIAN CERAMIC SOCIETIES 1041
[24] Chu B-J, Chen D-R, Li G-R, et al. Electrical properties of 
Na1/2Bi1/2TiO3–BaTiO3 ceramics. J Eur Ceram Soc. 
2002;22(13):2115–2121.
[25] Ni F, Luo L, Pan X, et al. Effects of A-site vacancy on the 
electrical properties in lead-free non-stoichiometric 
ceramics Bi0.5+x(Na0.82K0.18)0.5−3xTiO3 and Bi0.5+y(Na0.82 
K0.18)0.5TiO3. J Alloys Compd. 2012;541:150–156.
[26] Rietveld HM. A profile refinement method for nuclear and 
magnetic structures. J Appl Crystallogr. 1969;2(2):65–71.
[27] Lutterotti L, Bortolotti M. Object oriented program-
ming and fast computation techniques in MAUD, 
a program for powder diffraction analysis written in 
java. IUCr Compcomm Newsletter. 2003;1:43–50.
[28] Lutterotti L, Matthies S, Wenk H. MAUD: a friendly Java 
program for material analysis using diffraction. IUCr: 
Newsletter CPD. 1999;21:14–15.
[29] Lutterotti L, Matthies S, Wenk H, MAUD (Material 
Analysis Using Diffraction): a user friendly Java pro-
gram for Rietveld texture analysis and more. pp. 1599 
In Proceeding of the Twelfth International Conference 
on Textures of Materials (ICOTOM-12). Vol. 1, Ottawa, 
Canada.
[30] Onoe M, Jumonji H. Useful formulas for piezoelectric 
ceramic resonators and their application to measurement 
of parameters. J Acoust Soc Am. 1967;41(4B):974–980.
[31] Veera Gajendra M, Babu SM, Abdul Kader M, et al. 
Enhanced piezoelectric constant and remnant polari-
sation in K-compensated sodium potassium bismuth 
titanate. Mater Lett. 2015;146:81–83.
[32] Seo I-T, Steiner S, Frömling T. The effect of A site 
non-stoichiometry on 0.94(NayBix)TiO3-0.06BaTiO3. 
J Eur Ceram Soc. 2017;37(4):1429–1436.
[33] Seifert KTP, Jo W, Rödel J. Temperature-insensitive 
large strain of (Bi1/2Na1/2)TiO3–(Bi1/2K1/2)TiO3–(K0.5 
Na0.5)NbO3 lead-free piezoceramics. J Am Ceram Soc. 
2010;93(5):1392–1396.
[34] Jones GO, Thomas PA. Investigation of the structure 
and phase transitions in the novel A-site substituted 
distorted perovskite compound Na0.5Bi0.5TiO3. Acta 
Crystallographica Section B. 2002;58(2):168–178.
[35] Jones GO, Thomas PA. The tetragonal phase of Na0.5 
Bi0.5TiO3 - a new variant of the perovskite structure. 
Acta Crystallographica Section B. 2000;56(3):426–430.
[36] Lei N, Zhu M, Yang P, et al. Effect of lattice occupation 
behavior of Li+ cations on microstructure and electrical 
properties of (Bi1/2Na1/2)TiO3-based lead-free 
piezoceramics. J Appl Phys. 2011;109(5):054102.
[37] Shannon R. Revised effective ionic radii and systematic 
studies of interatomic distances in halides and 
chalcogenides. Acta Crystallogr A. 1976;32(5):751–767.
[38] Akram F, Ahmed Malik R, Hussain A, et al. Temperature 
stable dielectric properties of lead-free BiFeO3–BaTiO3 
modified with LiTaO3 ceramics. Mater Lett. 
2018;217:16–19.
[39] Wang B, Luo L, Ni F, et al. Piezoelectric and ferroelectric 
properties of (Bi1−xNa0.8K0.2Lax)0.5TiO3 lead-free cera-
mics. J Alloys Compd. 2012;526:79–84.
[40] Lee YC, Huang YL. Effects of CuO doping on the 
microstructural and dielectric properties of Ba0. 6 
Sr0. 4TiO3 ceramics. J Am Ceram Soc. 2009;92 
(11):2661–2667.
[41] Qiao X-S, Chen X-M, Lian H-L, et al. Microstructure and 
electrical properties of nonstoichiometric 0.94(Na0.5 
Bi0.5+x)TiO3–0.06BaTiO3 lead-free ceramics. J Am 
Ceram Soc. 2016;99(1):198–205.
[42] Jones GO, Kreisel J, Thomas PA. A structural study of 
the (Na1−xKx)0.5Bi0.5TiO3 perovskite series as a function 
of substitution (x) and temperature. Powder Diffr. 
2012;17(4):301–319.
[43] Sung YS, Baik S, Lee JH, et al. Enhanced piezoelectric 
properties of (Na0.5+y+zK0.5−y)(Nb1−xTax)O3 ceramics. 
Appl Phys Lett. 2012;101(1):012902. .
[44] Zhao Y, Xu Z, Chu R, et al. Improved piezoelectricity 
and high strain response of (1 − x)(0.948K0.5Na0.5NbO3 
− 0.052LiSbO3) − xBi2O3 ceramics. J Mater Sci Mater 
Electron. 2017;28(2):1211–1216.
[45] Kazushige Y, Yuji H, Hajime N, et al. Electrical proper-
ties and depolarization temperature of (Bi1/2Na1/2)TiO3 
–(Bi1/2K1/2)TiO3 lead-free piezoelectric ceramics. Jpn 
J Appl Phys. 2006;45(5S):4493.
[46] Ma C, Tan X, Dul’kin E, et al. Domain structure- 
dielectric property relationship in lead-free (1−x)(Bi1/2 
Na1/2)TiO3-xBaTiO3 ceramics. J Appl Phys. 2010;108 
(10):104105.
[47] Akram F, Malik RA, Song TK, et al. Thermally-stable 
high dielectric properties of (1–x)(0.65Bi1.05FeO3 
–0.35BaTiO3)–xBiGaO3 piezoceramics. J Eur Ceram 
Soc. 2019;39(7):2304–2309.
[48] Akram F, Kim J, Khan SA, et al. Less temperature- 
dependent high dielectric and energy-storage proper-
ties of eco-friendly BiFeO3–BaTiO3-based ceramics. 
J Alloys Compd. 2020;818:152878.
[49] Sakata K, Takenaka T, Naitou Y. Phase-relations, dielec-
tric and piezoelectric properties of ceramics in the 
system (Bi0.5Na0.5)TiO3-PbTiO3. Ferroelectrics. 
1992;131(1–4):219–226.
[50] Sapper E, Schaab S, Jo W, et al. Influence of electric 
fields on the depolarization temperature of Mn-doped 
(1-x)Bi1/2Na1/2TiO3-xBaTiO3. J Appl Phys. 2012;111 
(1):014105.
1042 X. CHEN ET AL.
	Abstract
	1. Introduction
	2. Experimental procedure
	3. Results and discussion
	4. Conclusions
	Disclosure statement
	Funding
	References