You are on page 1of 7

Journal of Alloys and Compounds 632 (2015) 103–109

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jalcom

Low sintering temperature and high piezoelectric properties of Li-doped


(Ba,Ca)(Ti,Zr)O3 lead-free ceramics
Xiaoming Chen a,b,c, Xuezheng Ruan a, Kunyun Zhao a, Xueqing He d, Jiangtao Zeng a,⇑, Yongsheng Li c,
Liaoying Zheng a, Chul Hong Park e, Guorong Li a
a
Key Laboratory of Inorganic Functional Materials and Devices, Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, China
b
University of Chinese Academy of Sciences, Beijing 100039, China
c
School of Materials Science and Engineering, East China University of Science and Technology, Shanghai 200237, China
d
School of Materials and Metallurgy, Northeastern University, Shenyang 110004, China
e
Department of Physics Education, Pusan National University, Pusan 609735, South Korea

a r t i c l e i n f o a b s t r a c t

Article history: Li-doped Ba0.85Ca0.15Ti0.9Zr0.1O3 (BCZT) lead-free piezoelectric ceramics were prepared by the two-step
Received 4 October 2014 synthesis and the solid-state reaction method. The density and grain size of ceramics sufficiently
Received in revised form 4 January 2015 increases by Li-doped sintering aid, and their sintering temperature decreases from about 1540 °C down
Accepted 15 January 2015
to about 1400 °C. X-ray diffraction reveals that the phase structure of Li-doped BCTZ ceramics is changed
Available online 22 January 2015
with the sintering temperature, which is consistent with their phase transition observed by the temper-
ature-dependent dielectric curves. The well-poled Li-doped BCZT ceramics show a high piezoelectric con-
Keywords:
stant d33 (512 pC/N) and a planar electromechanical coupling factor kp (0.49), which have the
Ferroelectrics
Sintering
characteristics of soft Pb(Zr,Ti)O3 (PZT) piezoceramic, on the other hand, the mechanical quality factor
Phase transitions Qm is about 190, which possesses the features of hard PZT piezoceramics. The enhanced properties of
X-ray diffraction the Li-doped BCZT are explained by the combination of Li-doped effect and sintering effect on the micro-
structure and the phase transition around room temperature.
Ó 2015 Elsevier B.V. All rights reserved.

1. Introduction tric coefficient d33 of BCTZ ceramics can be achieved about 650 pC/
N when BCZT ceramics is sintered at the optimizing temperature
Pb(Zr1 xTix)O3 (PZT) based piezoelectric ceramics are widely 1540 °C. However, it needs high energy consumption to sinter at
used in transducers and actuators for more than five decades due such high temperature. Therefore, many researchers manage to
to their excellent piezoelectric properties. However, the environ- lower the sintering temperature of the BCTZ piezoelectric ceramics
mental concerns urgently require lead-free ceramics to substitute by doping with metal oxides, such as, divalent oxides (ZnO [13],
for the PZT with high content of toxic lead [1,2]. BaTiO3 (BT) is CuO [14]), trivalent oxides (Y2O3 [15]), and tetravalent oxides
one of the good candidates for lead-free ceramics due to its (CeO2 [16]). Moreover, phase transition of BCTZ ceramics occurs
promising piezoelectric properties and solid solution with other more easily when dopants have lower valence, such as, Zn2+ [13]
perovskites, such as BT–KNbO3 [3], BT–Bi(Mg1/2Ti1/2)O3 [4], and Cu2+ [14].
BT–Bi0.5Na0.5TiO3–Bi0.5K0.5TiO3 [5], BT–Bi(Zn1/2Ti1/2)O3–BiFeO3 [6], The Li+-containing additives as a sintering aid of low valence
BT–Bi0.5Na0.5TiO3–KNbO3 [7], Bi0.5Na0.5TiO3–(Ba,Ca)(Ti,Zr)O3 [8], can decrease the sintering temperature without sacrificing the
(Ba,Ca)(Ti,Sn)O3–(Ba,Ca)(Ti,Zr)O3 [9], BT–CaTiO3–BaHfO3 [10]. electric properties of BaTiO3 lead-free ceramics, where Li ions can
Among them, the BaTi0.8Zr0.2O3–(Ba0.7Ca0.3)TiO3 (BCTZ) composi- incorporate into the lattice of BaTiO3 [17–19] during the sintering
tion lead-free ceramics, which were reported by Liu and Ren process. Haussonne et al. [20] reported that a perovskite phase
[11], have been attracted great attention because of their high pie- (BaTi(1 x)LixO(3 3x)F3x) was formed by addition of 1–2 wt.% LiF to
zoelectricity arising from the polymorphic phase transition (PPT) BaTiO3 during liquid phase sintering, which contributes to the den-
at room temperature. Wang et al. [12] reported that the piezoelec- sification behavior and good dielectric properties. Binhayeeniyi
et al. [21] observed that the Li2O-doped Ba(ZrxTi1 x)O3 ceramics
can obviously reduce the densification temperature from 1250 °C
⇑ Corresponding author.
to 900 °C and keep their high electromechanical properties
E-mail addresses: zjt@mail.sic.ac.cn, chenxiaoming@student.sic.ac.cn, grli@mail.
sic.ac.cn (J. Zeng). (kp = 44%). For BCTZ ceramics, Tan et al. [22] reported that

http://dx.doi.org/10.1016/j.jallcom.2015.01.088
0925-8388/Ó 2015 Elsevier B.V. All rights reserved.
104 X. Chen et al. / Journal of Alloys and Compounds 632 (2015) 103–109

LiF-doped BCTZ ceramics sintered at 1350 °C show enhanced elec- tor Qm were calculated from the resonance–antiresonance method on the basis of
IEEE standards using an impedance analyzer (HP4294A). Ferroelectric hysteresis
trical properties (d33 = 380 pC/N, kp = 39.5%) compared with the
loops (P–E) and electromechanical strain were measured by a modified Sawyer–
pure BCTZ ceramics (d33 = 300 pC/N) sintered at the same temper- Tower circuit at room temperature (10 Hz, TF Analyzer 2000, aixACCT Systems
ature. In spite of these investigations, there are few reports on the GmbH, Aachen, Germany).
influence of a low valence Li+-containing additive on the phase
transition of BCTZ lead-free ceramics, which can influence electric 3. Results and discussion
properties.
In this work, Li-doped BCTZ lead-free piezoelectric ceramics Fig. 1 shows the SEM micrographs of the BCTZL ceramics sin-
were prepared by the two-step synthesis and the solid-state reac- tered at different temperature. All ceramics sintered above
tion method. The influence of a Li+-containing additive of low 1350°C are very dense and have clear grain boundaries. Moreover,
valence on crystal phase, microstructure and electric properties the average grain sizes increase from about 2.8 to 11.0 lm with the
of Li-doped BCTZ ceramics were systematically investigated. increase of sintering temperature from 1200 °C to 1475 °C. The
increase of grain sizes should be attributed to the part occupation
2. Experimental procedure of Li+ for the lattice of ABO3 perovskite structure, which generates
proper oxygen vacancies in BCTZL ceramics, and these defects can
Ba0.85Ca0.15Ti0.9Zr0.1O3 + 0.3 wt.% Li2CO3 (BCTZL) ceramics were prepared by the
two-step synthesis and conventional ceramic fabrication technique, in which BCTZ
source powders were calcined firstly and then mixed with Li2CO3. The purpose of
two-step synthesis is to achieve the homogeneity of the prepared BCTZ powders
and promote the formation of liquid phase from Li+-containing additives only at
the sintering process. To prepare BCTZ source powders, analytical-grade metal oxi-
des or carbonate powders of BaCO3 (99.8%), TiO2 (99.48%), CaCO3 (99.5%), and ZrO2
(99.84%) were mixed in anhydrous ethanol by ball-milling for 6–7 h, and then were
calcined at 1200 °C for 2 h. After calcination, Li2CO3 (99.31%) were added, ball-
milled for 6–7 h, and dried again. The reground powders were mixed with the poly
vinyl alcohol (PVA: 5 wt.%) solution, then pressed into pellets with a diameter of
12 mm. The pellets were sintered at 1200–1475 °C for 2 h in the air atmosphere.
The specimens coated with a silver paste to form electrodes on both sides were fired
at 710 °C, and then poled in a stirred silicone oil bath under a DC field of 4–5 kV/mm
at 30 °C for 15–30 min.
The crystalline phase of the sintered discs was identified by X-ray diffraction
using Cu Ka radiation (XRD, D8 ADVANCE, Bruker AXS, Germany) at room temper-
ature (26 °C). The grain size was determined by averaging over the total number of
grains in a scanning electron microscopy. The microstructure of the ceramics was
studied by the scanning electron microscope (SEM, KYKY-EM3200, KYKY Technol-
ogy Development Ltd., Beijing, China). The density of the sintered samples was
measured by the Archimedes method. The temperature dependence of dielectric
properties was measured by using an LCR meter (Novocontrol Alpha-A BroadBand
Dielectric Spectrometer, Germany) in the temperature range from 100 °C to
250 °C with a heating rate of 3 K/min. Piezoelectric coefficient d33 was measured
by a Burlincourt-type d33 meter (ZJ-4AN, Institute of Acoustic Academia Sinica) Fig. 2. Relative density of the BCZTL ceramics as a function of the sintering
and the planar electromechanical coupling factor kp and the mechanical quality fac- temperature.

Fig. 1. SEM images of BCTZL ceramics sintered at (a) 1250 °C, (b) 1300 °C, (c) 1350 °C, (d) 1400 °C, (e) 1450 °C, (f) 1475 °C for 2 h.
X. Chen et al. / Journal of Alloys and Compounds 632 (2015) 103–109 105

Fig. 3. (a) X-ray diffraction patterns of the BCTZL ceramics sintered at different temperatures; (b) the expanded XRD patterns of BCTZL ceramics in the range of 44–46°; (c) the
expanded XRD patterns of BCTZ ceramics sintered at different temperatures in the range of 44–47°.

accelerate the mass transport process, further promoting the grain


growth [23]. On the other hand, the ceramics can be well sintered
at a relatively low temperature range (1350–1450 °C) by doping
with Li+ (vs BCTZ, 1540 °C [12]). It may be attributed to the liquid
phase sintering mechanism. Li2CO3 has a low melting point of
723 °C [24], which can form a liquid phase to lower the sintering
temperature, accelerate the grain growth and improve the density.
Fig. 2 shows the relative density, which was calculated based on
theoretical density of BCZT (5.687 g/cm3) [12]. The relative densi-
ties of BCTZL ceramics increase with the sintering temperature. It
is higher than 90% of the theoretical ones when Tsin P 1350 °C.
Therefore, the dense microstructure of BCTZL ceramics can be
obtained at low sintering temperature, which is advantageous for
piezoelectric properties of ceramics.
Fig. 3(a) shows the XRD patterns of BCTZL ceramics sintered at
different temperatures of 1200–1450 °C. The X-ray diffraction pat-
terns in Fig. 3(a) indicate that there is no second impurity phase in
the solid solutions above 1350 °C. Fig. 3(b) is the expanded XRD
patterns in the range of 44–46° for BCTZL ceramics. Diffraction
Fig. 4. Ferroelectric properties of BCTZL ceramics as a function of sintering peaks at 45° almost keep unchanged when Tsin < 1350 °C (PDF#
temperature.
81-2200, orthorhombic (O) phase). However, tetragonal (T) phase
106 X. Chen et al. / Journal of Alloys and Compounds 632 (2015) 103–109

Fig. 5. Temperature dependence of er for BCTZL ceramics (Tsin = 1200 °C, 1250 °C, 1300 °C, 1400 °C, 1450 °C, 1475 °C).

can be identified by the separation of (0 0 2) and (2 0 0) peaks that low Ec is able to make the ceramics easily poled and large Pr
(PDF#05-0626) at around 2h of 45° when Tsin P 1350 °C, while is beneficial to the piezoelectric properties of ceramics [25,26].
the phase structure of pure BCTZ almost keep unchanged with Therefore, it is expected that the BCTZL ceramics show better pie-
the variation of sintering temperature (Fig. 3(c)). The result indi- zoelectric properties at low sintering temperature.
cates that the structure of BCTZL ceramics is changed by the grad- Fig. 5 shows the temperature dependence of er and tan d (1 kHz)
ual diffusion of Li+, and the BCTZL ceramics occur the phase for BCTZL ceramics sintered at different temperatures. It can be
transition from O phase to T phase with the sintering temperature seen that the ceramics exhibit dielectric anomalies, which are sim-
(Tsin) increasing. ilar to the pure BCTZ ceramics [11]. The peaks of the tan d curve are
Fig. 4 shows the P–E hysteresis loops for BCTZL ceramics sin- used to determine the rhombohedral (R)–O, O–T and T-cubic (C)
tered at different temperatures. With the increase of the sintering phase transition. With the increase of sintering temperature, R–O
temperature, the P–E loops become saturated. The values of rem- and O–T phase transition temperatures gradually shift to low tem-
nant polarization (Pr) increase from 3 lC/cm2 at Tsin = 1200 °C to peratures. It may be attributed to the structural evolution induced
10 lC/cm2 at Tsin = 1450 °C, and coercive field (Ec) slightly by the Li doping, or the internal stresses which was reported by
decreases from 2.4 kV/cm to 1.8 kV/cm. Previous reports depicted Armstrong and Buchanan [27] and Park and Kim [28], that
X. Chen et al. / Journal of Alloys and Compounds 632 (2015) 103–109 107

Fig. 8. Impedance spectra of BCTZL ceramics sintered at 1400 °C.

Fig. 6. Phase diagram of BCZTL ceramics as a function of sintering temperature.

Based on the results of the XRD and temperature-dependent


dielectric response, the phase diagrams of BCZTL ceramics are plot-
depresses the ferroelectric peak intensity and shifts the transition ted in Fig. 6. At room temperature, BCZTL ceramics experience
temperature to low temperatures, but the detailed reason should structural changes from O phase to T phase with Tsin increasing.
be complicated and will be further investigated. Fu et al. [29] Furthermore, all samples above Tsin = 1350 °C have O–T transition
reported that the PPT (orthorhombic–T phase transformations) of temperature, which are close to room temperature.
BCTZ ceramics occurs near room temperature, which should be Fig. 7 shows the electric properties of BCTZL ceramics as a func-
the origin of a very large piezoelectric response. In this work, the tion of the sintering temperature. The piezoelectric constant d33 of
orthorhombic (O) phase can be easily discerned in BCTZL ceramics, BCTZL ceramics reaches the maximum value of 512 pC/N at
which is consistent with the result of BCTZ ceramics [29–32]. In Tsin = 1450 °C, while the maximum planar electromechanical cou-
addition, the Curie temperature Tc firstly decreases, and then pling factor kp (the value is evaluated to be approximately 0.49 from
increases to 82 °C for BCTZL ceramics sintered at 1475 °C. It is very the resonance–antiresonance frequencies in Fig. 8) occurs
similar to the reported result of the BCTZ–LiNbO3 system [33]. at Tsin = 1400 °C, which have the characteristics of soft PZT

Fig. 7. Electric properties of BCTZL ceramics as functions of sintering temperature.


108 X. Chen et al. / Journal of Alloys and Compounds 632 (2015) 103–109

piezoceramics (d33 > 400 pC/N). Similar to the dependence of d33 on O to T symmetry. The densities and grain sizes of ceramics
Tsin, the relative permittivity reaches the maximum value, at increase, and the range of sintering temperature of BCTZ ceramics
Tsin = 1400 °C, and then decreases with the increase of Tsin. It is con- has been broadened around 1350–1450 °C by the addition of Li+
sistent well with the results of microstructure and ferroelectric due to the liquid phase sintering mechanism. The BCTZL ceramics
properties. On the other hand, the Qm reaches the maximum value not only have the characteristics of hard piezoceramics because of
(190) at Tsin = 1450 °C, and slightly decreases with the increase of the high mechanical quality factor Qm (190) caused by the defect of
Tsin, which possesses the features of relatively hard PZT piezoceram- oxygen vacancies, but also possesses the features of soft piezoce-
ics (Qm of soft PZT piezoceramics is usually about 50). The loss tan- ramics due to the large electric properties induced by the polymor-
gent tan d monotonically decreases with Tsin increasing. The phic phase transition around room temperature. The BCTZL
improved Qm and tan d may be due to the hardening effect of Li+ ceramics sintered at 1450 °C exhibit optimum properties, which
doping which is similar to the CuO-doped (K0.5Na0.5)NbO3 [34– are as follow: d33 = 512 pC/N, kp = 0.45, er = 4394, tan d = 1.2%,
36]. Acceptor dopant Li+ enters the lattice of ABO3 perovskite struc- Qm = 190, Tc = 79.6 °C, Pr = 10 lC/cm2 and Ec = 1.8 kV/cm. The
ture, which will contribute to the creation of oxygen vacancies for results imply that electric properties of BCTZ ceramics could be
ionic charge compensation. Oxygen vacancies can pin the motion enhanced by a low valence of Li+ doping at low sintering tempera-
of the ferroelectric domain walls and harden the ceramics [37,38]. ture, which keeps the structural evolution around room tempera-
Although the enhanced electric properties of BCTZL ceramics ture. It can guide the design of BCTZ-based piezoceramics.
can be related to dense ceramics [17,39], oxygen vacancies
[37,38], grain size effects [40], the phase evolution [29], it can be Acknowledgments
mainly attributed to the fact that the PPT temperature (O–T transi-
tion temperature) of BCTZL ceramics above Tsin = 1350 °C are This work is supported by the National Key Project for Basic
approximate to room temperature (as shown in Fig. 6), and then Research of China (Nos. 2012CB619406 and 2013CB632900), the
the ceramics can easily go through the phase evolution from O to National High-Technology Research and Development Program of
T phase at room temperature, which make ferroelectric domains China (No. 2013AA030801), International Science & Technology
easily switch and produce large piezoelectric response. Cooperation Program of China (No. 2013DFR50800), National Nat-
For Li+-substituted piezoelectric ceramics, it has been reported ural Science Foundation of China (Nos. 61137004, 61275181 and
that the piezoelectrics can be enhanced by the large off-centering 61311120086), the External Cooperation Program of BIC, CAS
of Li ions in A site of K0.5Li0.5NbO3 [41] and K1 xLixTaO3 ceramics (No. 121631KYSB20130003), the National Research Foundation of
[42]. For the ceramics sintering theory, the doping elements or sin- Korea (NRF) funded by the Ministry of Education (No.
tering aids are usually located in the surfaces of grains in the initial 2013R1A1A2065742).
state, and then gradually diffuse into grains which results in the
grain growth, as the sintering temperature increases, the doping
References
elements or sintering aids in grain surfaces will continuously dif-
fuses into the lattice of grain, and ceramics become more density [1] T. Takenaka, H. Nagata, J. Eur. Ceram. Soc. 25 (2005).
[43–46]. In our case, the Li+ mainly occupies the B-site of the BCTZL [2] P.K. Panda, J. Mater. Sci. 44 (2009) 5049–5062.
ceramics, and combines with the oxygen vacancies to form the [3] S. Wada, M. Nitta, N. Kumada, D. Tanaka, M. Furukawa, S. Ohno, C. Moriyoshi,
Y. Kuroiwa, Jpn. J. Appl. Phys. 47 (2008) 7678–7684.
defect dipole (LiTi–Vo) , which is supported by our first-principles [4] S. Wada, K. Yamato, P. Pulpan, N. Kumada, B.Y. Lee, T. Iijima, C. Moriyoshi, Y.
energy calculations that the formation enthalpy of defect dipole Kuroiwa, J. Ceram. Soc. Jpn. 118 (2010) 683–687.
(LiTi–Vo) is lower than that of Li+ occupation in A-site. The Li+ sub- [5] W.L. Li, W.P. Cao, D. Xu, W. Wang, W.D. Fei, J. Alloys Comp. 613 (2014) 181–
186.
stitution in B-site is consistent with our experimental results: as [6] C.Z. Xu Shan, Zhenyong Cen, Huabin Yang, Qin Zhou, Weizhou Li, Ceram. Int. 39
Tsin increases up to 1350 °C, the grain sizes of BCTZL ceramics grad- (2013) 6707–6712.
ually increase (as shown in Fig. 1), and the diffusion of Li+ in grains [7] H.M. Ni, L.H. Luo, W.P. Li, Y.J. Zhu, H.S. Luo, J. Alloys Comp. 509 (2011) 3958–
3962.
surface should become dominant than the occupation behavior of [8] Q. Gou, A. Li, B. Wu, D. Xiao, J. Zhu, J. Alloys Comp. 521 (2012) 4–7.
Li+ in lattice of BCTZL ceramics, which is consistent with XRD [9] J.G. Wu, A. Habibul, X.J. Cheng, X.P. Wang, B.Y. Zhang, Mater. Res. Bull. 48
results that the position of 45° diffraction peaks almost keep (2013) 4411–4414.
[10] D.L. Wang, Z.H. Jiang, B. Yang, S.T. Zhang, M.F. Zhang, F.F. Guo, W.W. Cao, J.
unchanged when Tsin < 1350 °C (as shown in Fig. 3(b)); as
Mater. Sci. 49 (2014) 62–69.
Tsin > 1350 °C, Li+ enters the lattice of grain and the densities of [11] W. Liu, X. Ren, Phys. Rev. Lett. 103 (2009) 257602.
BCTZL ceramics increase greatly, the occupation of Li+ with large [12] P. Wang, Y. Li, Y. Lu, J. Eur. Ceram. Soc. 31 (2011) 2005–2012.
radius (r = 0.76 Å, CN = 6) for Ti4+ with small radius (r4+ [13] J.G. Wu, D.Q. Xiao, W.J. Wu, Q. Chen, J.G. Zhu, Z.C. Yang, J. Wang, Scr. Mater. 65
Ti = 0.61 Å)
(2011) 771–774.
[47] leads to the enlargement of lattice parameters of the BCTZL [14] T. Chen, T. Zhang, G. Wang, J. Zhou, J. Zhang, Y. Liu, J. Mater. Sci. 47 (2012)
ceramics according to Bragg’s equation (2d sin h = k), which is con- 4612–4619.
sistent with the XRD result that the 45° diffraction peaks move to [15] Y. Cui, C. Yuan, X. Liu, X. Zhao, X. Shan, J. Mater. Sci.-Mater. El 24 (2013) 654–
657.
the low diffraction angle, as shown in Fig. 3(b), and the structure [16] Y. Cui, X. Liu, M. Jiang, X. Zhao, X. Shan, W. Li, C. Yuan, C. Zhou, Ceram. Int. 38
phase of BCTZL is changed from O phase to T phase. Li+ occupation (2012) 4761–4764.
in B-site (Ti4+) of the BCTZL ceramics produce piezoelectric harden- [17] X. Chen, Y. Liao, L. Mao, Y. Deng, H. Wang, Q. Chen, J. Zhu, D. Xiao, J. Zhu, G. Xu,
Phys. Stat. Sol. A-Appl. Mater. Sci. 206 (2009) 1616–1619.
ing effect, which result in the increase of Qm, as shown in Fig. 7. The [18] T. Kimura, Q. Dong, S. Yin, T. Hashimoto, A. Sasaki, T. Sato, J. Eur. Ceram. Soc. 33
mechanism of Li+ substitution in the B-site of the ABO3 oxides will (2013) 1009–1015.
be further discussed in the other paper. [19] H.Y. Xie, S. Yin, T. Takatoshi, Y. Tokano, A. Sasaki, T. Sato, Mater. Res. Bull. 45
(2010) 1345–1350.
[20] J.M. Haussonne, G. Desgardin, P.H. Bajolet, B. Raveau, J. Am. Ceram. Soc. 66
(1983) 801–807.
4. Conclusions [21] N. Binhayeeniyi, P. Sukvisut, C. Thanachayanont, S. Muensit, Mater. Lett. 64
(2010) 305–308.
[22] C.K.I. Tan, K. Yao, J. Ma, Int. J. Appl. Ceram. Technol. 10 (2013) 701–706.
The structure and electrical properties of BCTZL lead-free piezo- [23] Y.C. Lee, Y.L. Huang, J. Am. Ceram. Soc. 92 (2009) 2661–2667.
electric ceramics prepared at different sintering temperatures from [24] N. Ma, B.-P. Zhang, W.-G. Yang, J. Electroceram. 28 (2012) 275–280.
1200 °C to 1475 °C have been studied. The ceramics possess pure [25] L. Dunmin, K.W. Kwok, H.W.L. Chan, J. Phys. D: Appl. Phys. 40 (2007) 7523.
[26] Y. Zhang, R. Chu, Z. Xu, Q. Chen, Y. Liu, G. Zhang, Curr. Appl. Phys. 12 (2012)
single phase of perovskite structure as Tsin is higher than 1350 °C, 204–209.
and with the effect of Li+ doping, the phase structure evolves from [27] T.R. Armstrong, R.C. Buchanan, J. Am. Ceram. Soc. 73 (1990) 1268–1273.
X. Chen et al. / Journal of Alloys and Compounds 632 (2015) 103–109 109

[28] Y. Park, H.-G. Kim, J. Am. Ceram. Soc. 80 (1997) 106–112. [37] S.-J. Yoon, A. Joshi, K. Uchino, J. Am. Ceram. Soc. 80 (1997) 1035–1039.
[29] D.S. Fu, Y. Kamai, N. Sakamoto, N. Wakiya, H. Suzuki, M. Itoh, J. Phys.-Condens. [38] Y.-D. Hou, L.-M. Chang, M.-K. Zhu, X.-M. Song, H. Yan, J. Appl. Phys. 102 (2007).
Matter 25 (2013) 5. [39] X. Chen, Y. Liao, H. Wang, L. Mao, D. Xiao, J. Zhu, Q. Chen, J. Alloys Comp. 493
[30] F. Benabdallah, A. Simon, H. Khemakhem, C. Elissalde, M. Maglione, J. Appl. (2010) 368–371.
Phys. 109 (2011). [40] K. Okazaki, K. Nagata, J. Am. Ceram. Soc. 56 (1973) 82–86.
[31] D.S. Keeble, F. Benabdallah, P.A. Thomas, M. Maglione, J. Kreisel, Appl. Phys. [41] D.I. Bilc, D.J. Singh, Phys. Rev. Lett. 96 (2006) 147602.
Lett. 102 (2013). [42] S. Prosandeev, E. Cockayne, B. Burton, Phys. Rev. B 68 (2003) 014120.
[32] L. Zhang, M. Zhang, L. Wang, C. Zhou, Z. Zhang, Y. Yao, L. Zhang, D. Xue, X. Lou, [43] A.G. Evans, J. Am. Ceram. Soc. 65 (1982) 497–501.
X. Ren, Appl. Phys. Lett. 105 (2014). [44] M.F. Ashby, Acta Metall. 22 (1974) 275–289.
[33] T. Chen, T. Zhang, J. Zhou, J. Zhang, Y. Liu, G. Wang, Ceram. Int. 38 (2012) 3591– [45] G.C. Kuczynski, Trans. Am. Inst. Mining Met. Eng. 185 (1949) 169–178.
3594. [46] H. Xianchang, Essentials of ceramic, Shanghai Popular Science Press, China,
[34] M. Matsubara, K. Kikuta, S. Hirano, J. Appl. Phys. 97 (2005). 2005.
[35] M. Matsubara, T. Yamaguchi, W. Sakamoto, K. Kikuta, T. Yogo, S.-I. Hirano, J. [47] David R. Lide (Ed.), Handbook of Chemistry and Physics, 90th ed., CRC Press,
Am. Ceram. Soc. 88 (2005) 1190–1196. Boca Raton, Florida, 2009.
[36] T. Shrout, S. Zhang, J. Electroceram. 19 (2007) 113–126.

You might also like