Вы находитесь на странице: 1из 6

Journal of Physics and Chemistry of Solids 77 (2015) 6267

Contents lists available at ScienceDirect

Journal of Physics and Chemistry of Solids


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

Structural, ferroelectric, optical properties of A-site-modied


Bi0.5(Na0.78K0.22)0.5Ti0.97Zr0.03O3 lead-free piezoceramics
N.D. Quan a,b, N.V. Quyet c, L.H. Bac a, D.V. Thiet a, V.N. Hung b, D.D. Dung a,n
a

School of Engineering Physics, Ha Noi University of Science and Technology, 1 Dai Co Viet Road, Ha Noi, Viet Nam
ITIMS, Ha Noi University of Science and Technology, 1 Dai Co Viet Road, Ha Noi, Viet Nam
c
Hanautech Co., Ltd., 832, Tamnip-dong, Yuseong-gu, Daejeon, Republic of Korea
b

art ic l e i nf o

a b s t r a c t

Article history:
Received 25 July 2014
Received in revised form
10 October 2014
Accepted 12 October 2014
Available online 23 October 2014

We reported the role of A-site modication on the structural, ferroelectric, optical and electrical eldinduced strain properties of Bi0.5(Na0.78K0.22)0.5Ti0.97Zr0.03O3 lead-free piezoceramics. The Li ions with
concentration from 0 to 5 mol% were used to substitute at A-site. There was no phase transition when Li
ions was added up to 5 mol%. The electric eld-induced strain (Smax/Emax) values increased from 600 to
643 pm/V for 2 mol% Li -added which results from distortion both rhombohedral and tetragonal phase
structures. The band gap reduced from 2.88 to 2.68 eV and the saturation polarization decreased from
46.2 to 26.1 C/cm2 when Li ions concentration increased from 0 to 5 mol% respectively. We expect that
this work could be helpful for further understanding the role of A-site dopants in comparison with B-site
modication in lead-free Bi0.5(Na,K)0.5TiO3-based ceramics.
& Elsevier Ltd. All rights reserved.

Keywords:
A. Lead-free BNKT
B. Optical
C. Ferroelectricity
D. Lithium

1. Introduction
The lead-based ferroelectric ceramics such as Pb(Zr,Ti)O3 (PZT)
or Pb(Mg,Nb)O3PbTiO3 (PMNPT) are widely used in sensor or
actuator application [1]. However, environmental issues have
raised the need for non-hazardous materials with properties
comparable to that of Pb-based materials for use in device fabrication [2]. Thus, considerable effort has been devoted towards the
development of lead-free piezoelectric ceramics [3]. Among the
lead-free piezoelectric ceramics that developed, Bi0.5(Na,K)0.5TiO3
(BNKT) systems have received a great deal of attention due to their
excellent ferroelectric and piezoelectric properties, as well their
near rhombohedraltetragonal (MPB) composition [4]. Recently, it
was found that large strains in BNKT ceramics can be enhanced
when Ti4 ions at B-site are substituted with either isovalent ions
such as Hf4 [5] and Zr4 [6]; alionvalent ions including Nb5 [7]
and Ta5 [8]; or trivalent ions such as Y3 [9]. In addition, Nguyen
et al. reported that the co-substitution in both A- and B-site was
strongly enhanced the electric eld induced strain such as (Li,Ta)
and (Li,Sn) [10,11]. They suggested that the observation in good
piezoelectric properties of modication BNKT ceramics compared
with soft PZT results from phase transition from polar to non-polar
due to distortion of tolerance factor and/or promotion by oxygen
vacancy [7,1013]. However, we recently obtained the phase
n

Corresponding author.
E-mail address: dung.dangduc@hust.edu.vn (D.D. Dung).

http://dx.doi.org/10.1016/j.jpcs.2014.10.010
0022-3697/& Elsevier Ltd. All rights reserved.

transition from pseudocubic to tetragonal phase. This results from


distortion of the tolerance factor due to modied A-site with Li in
BNKT-based ceramics. Thus, the ferroelectric properties were enhanced [14]. Recently, Hussain et al. reported the enhancement in
both saturation polarization and electric-eld-induced giant strain
in Bi0.5(Na0.80K0.20)0.5TiO3 solid solution with SrZrO3 ceramics [15].
Interestingly, the crystal structures were maintained in pseudocubic perovskite structure, even in the Bi0.5(Na0.80K0.20)0.5TiO3
pure ceramics [15]. This result seems to be conicted with their
previously reported in Bi0.5(Na0.80K0.20)0.5TiO3 solid solution with
BiAlO3 where the coexistence of rhombohedra and tetragonal
phase in Bi0.5(Na0.80K0.20)0.5TiO3 pure ceramic and there phase
transited to pseudocubic in solid solution with BiAlO3 [16]. Thus,
the origin of phase transition from polar to non-polar is an importance key in original observation of high electrical eld induced giant strain in BNKT via (co-)dopant or solid solution with
other perovskite that was not well understood. Unlikely, other
reported for observation of the enhancement the dynamic piezoelectric coefcient (dn33) course of phase transition [713], Hussain
et al. obtained the enhancement dn33 value up to 614 pm/V by
3 mol% Zr4 replacement Ti4 in Bi0.5(Na0.78K0.22)0.5TiO3, as B-site
modication, where there were no phase transition observation
and that only distorted structure causing of different ionic radii
between Zr4 and Ti4 [6]. The results point out that the phase
transition is not a main reason for recently observation in rapidly
development the dynamic piezoelectric coefcient in lead-free
piezoelectric BNKT-based ceramics. Recently, the A-site modied
by Li in lead-free BNKT-based ceramics causing induced phase

N.D. Quan et al. / Journal of Physics and Chemistry of Solids 77 (2015) 6267

transition was found to be induce phase transition and result in


enhancement the dn33 values [10,11,14]. It was noted that Li were
found to be suppressed formation of the second phase and Ti3 /4
valence transition when it substituted at Na-site in BNKT [17,18].
In this work, we reported the role of Li -substitution at A-site
on ferroelectric, optical properties and electrical eld-induced
strain of Bi0.5(Na0.78K0.22)0.5Ti0.97Zr0.03O3 ceramics. The slight dn33
values were increased from 600 to 643 pm/V while saturation
polarization decreased graduation from 46.2 to 26.1 C/cm2. The
band gap reduced from 2.88 to 2.68 eV by adding 5 mol% Li. We
expect that this work could be helpful for further understanding
the role of A-site in comparison with B-site modication in leadfree Bi0.5(Na,K)0.5TiO3-based ceramics.

2. Experimental
The Bi0.5(Na(0.78 x)K0.22Lix)0.5(Ti0.95Zr0.03)O3 (BNKTZxLi) (x0.00,
0.01, 0.02, 0.03, 0.04, and 0.05) ceramics were prepared by a
conventional solid state reaction route. The raw materials were

63

powders composed of Bi2O3, K2CO3, TiO2, Li2CO3 (99.9%, Kojundo


Chemical), Na2CO3 (99.9%, Ceramic Specialty Inorganics) and ZrO2
(99.9%, Cerac Specialty Inorganics). The powders were weighted
according to their chemical formulas and then ball-milled for 24 h
in anhydrous ethanol with zirconia balls. The slurry was dried and
calcined at 850 C for 2 h. To prevent the vaporization of Bi, Na,
and K elements, the disks were embedded in powders of their
identical compositions and a powder of BNKT was used as the
bedding powder. The calcined powder was added with polyvinyl
alcohol as a binder and uniaxially pressed into circular disks with a
diameter of 12 mm at 98 MPa. The green compacts were sintered
in a covered alumina crucible at 11401160 C for 2 h in ambient
condition. Electrical measurements were carried out after screenprinting Ag paste on both sides of a disk-shaped specimen and
subsequent ring at 700 C for 30 min. The bulk densities were
measured by the Archimedes method. The surface morphology
was observed with a eld emission scanning electron microscope
(FE-SEM). The crystalline structures of the samples were characterized by X-ray diffraction (XRD). The optical properties were
studied by UVvis spectroscopy. The polarizationelectric elds
(PE) and electric eld-induced strain hysteresis loops were
measured in silicon oil using a modied SawyerTower circuit and
linear variable differential transducer system, respectively. The
temperature dependence of the dielectric properties was measured using an impedance analyzer.

3. Results and discussion

Fig. 1. The density of BNKTZxLi ceramics as a function of Li doping level.

Fig. 1 shows the density of the BNKTZ ceramics with difference


Li concentration. The undoped-BNKTZ samples have density of
5.714 g/cm3. The density of Li-dopants BNKTZ samples increases
and reaches maximum value of 5.808 g/cm3 at 3 mol% Li dopants.
The density was slightly decreased to 5.797 g/cm3 when Li concentration further added up to 5 mol%.
Fig. 2(a)(f) shows FE-SEM micrographs of the BNKTZxLi
ceramics with x 0, 0.01, 0.02, 0.03, 0.04 and 0.05, respectively. A
dense microstructure with some distinct pores is observed for the
BNKTZ ceramic, as seen in Fig. 1(a). The compact structure with

Fig. 2. FE-SEM images of the BNKTZxLi ceramics with different content: (a) x 0.00, (b) x 0.01, (c) x 0.02, (d) x 0.03, (e) x 0.04, and (f) x 0.05.

64

N.D. Quan et al. / Journal of Physics and Chemistry of Solids 77 (2015) 6267

Fig. 3. X-ray diffraction pattern of Li doped BNKTZ ceramics as a function of Li


doping level x in the 2 range of 1070.

appearances of small grains were obtained when Li added with


x 0.01 and 0.02, as shown in Fig. 1(b) and (c), respectively. The
samples show dense and homogeneous grains were obtained
when Li further added, as seen in Fig. 2(d)(f). The average grain
sizes of samples increased with increasing amount of Li2CO3.
However, the average grain size of undoped samples was approximately 830 nm with standard deviation 26 nm while 5 mol%
Li-doped samples was more homogeneous with average size of
1148 nm and standard deviation was only 16 nm. We suggested
that the effect of Li on the grain size resulted from liquid phase
sintering process because of low melting point of Li2CO3 and/or
promotion oxygen vacancies during sintering process.
Fig. 3 shows the XRD pattern of BNKTZxLi ceramics with
x 0.00, 0.01, 0.02, 0.03, 0.04 and 0.05. The all samples show the
single perovsktie structure without impurity phase, indicating that
Li ions were successfully diffused to lattice. The magnications of
Fig. 3 in the range from 39.0 to 40.5 and 44.0 to 48.0 are shown
in Fig. 4(a) and (b), respectively. The result shows that the peaks
were unsymmetrical with shoulder peak which were suggested
that overlap of multi-peaks. The each peak was carefully tted by
using the Lorentzian tting with mean rout square over 0.995, as
shown in the red dash line. The peaks were indexed as (003)/(021)
and (002)/(200) in the range from 39.0 to 40.5 and 44.0 to 48.0,
respectively, indicating that both rhombohedral and tetragonal
phases were coexisted. There was no phase transition observation
when Li was added. The studying about the morphotropic phase
boundary (MPB) indicated that both rhombohedral and tetragonal

phases co-existed in 20 mol% Bi0.5K0.5TiO3 (BKT) solid solution in


Bi0.5Na0.5TiO3 (BNT) [1921]. The coexistence of rhombohedral and
tetragonal phase were shown in 3 mol% Zr-modied BNKT which
were also in agreement with Hussain et al. [6]. However, the peaks
position were trended to shift to high angle when Li was added
with content up to 2 mol%, indicating that Li gave the local
compressor strain when it lled at Na site. The result can be
understood based on the difference between ionic radii of Li and
Na that ionic radius of Li (0.092 nm in 8-od coordination) was
smaller than that of Na (0.136 nm in 12-fold coordination coordination) [22]. The peaks position shifted to higher angle indicated that the phase moved to BNT peaks position as rhombohedral symmetry. In other words, the rhombohedra phase was
enhanced when Li ions replaced in Na ions. However, the peaks
position was shifted back to low angle when Li concentration
further added, up to 5 mol%. Because the ionic radii of Ti4
(0.0605 nm in 6-fold coordiantion) was comparable with ionic
radii of Li (0.076 nm in 6-fold coordination) [22], therefore, we
suggest that the Li ions can be also lled at the octahedral site
and consequently resulted in expansion the lattice constants. The
effect of multisite Li was well known in both lead-based and
lead-free ferroelectric materials [14,2325].
The room temperature absorption spectra of the Li substitution
in A-site of BNKTZ are shown in Fig. 5(a). All of the specimens
exhibited absorption in the visible light region. The absorption
spectra show a red shift slightly as the volume fraction of the Li
concentration increases, indicating that the Li ions addition
modied the band gap (Eg) of BNKTZ. The Eg values was associated
with the absorbance and photon energy by following equation
h~(h  Eg)n, where is the absorbance coefcient, h the Planck
constant, the frequency, Eg the optical band gap and n a constant
associated with different types of electronic transition (n 1/2, 2,
3/2 or 3 for direct allowed, indirect allowed, direct forbidden and
indirect forbidden transitions, respectively) [26]. The Eg values of
BNKTZxLi ceramics were evaluated by extrapolating the linear
portion of the curve or tail. The band gap estimated with n 1/2
for direct transition as shown in Fig. 5(b). The optical band gap is
found to decrease with increase in Li ions addition. It decreased
from 2.88 eV to 2.68 eV as Li ions concentration increased from
0 to 5 mol% Li. The effective of bonding type between hybridization AO and BO to band gap of perovskite ABO3 were reported
for various ferroelectric materials such as PbTiO3, SrTiO3 etc. or
solid solution ABO3ABO3 [2730]. Recently, we obtained the
reduction of Eg due to Li-modied BNKT-based ceramics resulting
from changing the ionic bonding between A-site and oxygen due
to Li ions added [14]. Thus, we suggested that the reduction of
band gap in Li-modied BNKTZ ceramics is strongly related to the
ionics bonding between A-site and oxygen.

Fig. 4. (color online) The magnied XRD patterns in the 2 ranges of (a) 39.040.5 and (b) 44.048.0.

N.D. Quan et al. / Journal of Physics and Chemistry of Solids 77 (2015) 6267

Fig. 6 shows the polarizationelectric eld (PE) hysteresis


loops of BNKTZxLi ceramics at room temperature. All samples
showed PE hysteresis loops at room temperature indicated that
all specimens were typical ferroelectric materials. However, the
maximum polarization (Pm) decreased with increasing of Li concentration, from 46.2 to 26.1 C/cm2, together with reduction of
the remanence polarization (Pr), from 22.6 to 8.4 C/cm2. The reduction values of Pm of BNKT-based ceramics were recently reported for phase transition from polar to non-polar due to dopant
and/or oxygen vacancy [510]. However, Otonicar et al. reported
that the highest values of Pr, in addition with Pm, for the MPB
samples can be attributed to the increased number of possible
directions of polarization because of the coexistence of more than
one anisotropic crystal structure, order of 6 for tetragonal and 8 for
rhombohedral structure [21]. Shuvaeva et al. reported that the
MPB composition enables the dipole moments to align efciently
with the eld, resulting in a higher polarization of material [31].
Our result showed that the structure was distorted when the Li
content added up to 2 mol% and therefore that can cause shrinking
the lattice constant. It indicated that the rhombohedral and tetragonal crystal structure symmetry reduced as decreeing the Pm

65

and Pr. However, the vacancy was created when Li lled to


Ti4 -site result in rapid decreasing both Pm and Pr values. Thomas
et al. reported that the BO6 oxygen octahedral play a critical role in
the manifestation of ferroelectric properties in ABO3-type perovskite [31]. The degree of the coupling between neighboring BO6
octahedral will be signicantly weakened by introducing defects
that acts to break the translational invariance of polarization, result in a decrease in the coupling of ferroelectrically active BO6
octahedral [3134]. Yuan et al. reported that both A-site and B-site
vacancies decreased the degree of coupling between dipoles in
ABO3-type perovskite result in reduced the coupling of ferroelectriclly active BO6 octahedra, moreover, the effect of A-site vacancies on the decoupling of BO6 octahedral was more signicant
than that of B-site vacancies [35]. In addition, theoretical predicted
that the B-cation displacement highest in case of Ti which were
expected for strong ferroelectric active [36]. Thus, we suggest that
the ions Li replaced in Ti4 and then hindered the displacement
B-site, even the tetragonal and rhombohedra structures were coexisted. The displacement of B-site can result in reduction the
maximum polarization.

Fig. 5. (a) UVvis absorption spectra of the BNKTS ceramics, and (b) the dependence of (h)2 on h of the BNKTZ ceramics. The inset of Fig. 5(b) shows the band gap Eg of the
BNKTZ ceramics as function of Li dopant.

Fig. 6. Room temperature PE hysteresis loops of the BNKTZxLi ceramics as a function of xmol% Li ions content.

66

N.D. Quan et al. / Journal of Physics and Chemistry of Solids 77 (2015) 6267

Fig. 7(a) shows the temperature dependence of the dielectric


constant and dielectric loss of BNKTZxLi ceramics at frequencies
of 1 kHz. The curves show two distinct anomalies for all samples,
which correspond to the depolarization temperature (Td) and
maximum dielectric constant temperature (Tm), respectively. The
curves for different samples look similar, all exhibiting two-phase
transition at Td and Tm. The two dielectric peaks can cause by the
phase transitions from ferroelectric to antiferroelectric (Td) and
anti-ferroelectric to paraelectric phase (Tm), which is consistent
with the previous reports of lead-free BNKT-based ceramics
[11,32,33]. The variations in the values of Td and Tm with different
amount of Li addition for BNKTZ ceramics are presented in Fig. 7
(b). From Fig. 7(b), it is found that both Td and Tm exhibited an
obvious dependency on amount of Li cations dopants concentration. However, the rst phase transition temperature seems
to change little in range from 403 to 411 K while the secondary
phase transition temperature increased from 557 to 615 K as Li
content increased from 0 to 5 mol%. Recently, it was proposed that
the dielectric maximum around 573 K (Tm) is related to relaxation
of tetragonal emerged from rhombohedral polar nanoregions, although the investigated system is Bi0.5Na0.5TiO3BaTiO3 and still
requires additional validation [34]. According to Zhu et al., the
reversed dependence of Td and Tm on Li amount dopants concentration can be attributed to the lattice distortion [35]. In addition Zhou et al. reported that the vacancies facilitate the movement of the ferroelectric domain and result in a decrease of depolarization temperature [36]. Therefore, we suggested that the
decrease of depolarization temperature was strongly related with
multisite occurrence of Li ions in BNKTZ ceramics.
Furthermore, unipolar eldinduced strain curves of BNKTZ-xLi
ceramics with x 0.000.05 are shown in Fig. 8(a). The unipolar
strain increased with increasing Li ions content up to x0.02 and
then decreased at a higher doping ratio. Fig. 8(b) showed the
maximum strain and dn33 values as function of Li ions content.
Pure BNKTZ ceramics exhibited a unipolar eld-induced strain of
0.42%, which reached 0.45% at x 0.02 at an applied electric eld
of 70 kV/cm, and then decreased by further doping. The BNKTZ-0Li
showed a normalized strain (Smax/Emax) of 600 pm/V, in agreement
with reported work for enhancement Smax/Emax value due to Zrdoped BNKT-based ceramics reported by Ali et al. [6]. The addition
of 2 mol.% Li increases the Smax/Emax up to 643 pm/V which were
comparable to high end values of soft PZT ceramics [10]. Further
Li ions doping reduced the Smax/Emax down to 442 pm/V, corresponding with strain of 0.31%, at x 0.05. The enhancement Smax
/Emax recently reported for both modications at A- and B-sites of
BNKT-based ceramics results from structure change [513]. Our
work indicated that the enhancement of Smax/Emax results from
distorted structure of BNKTZ and oxygen vacancy may play the
role to hind the electric eld-induced strain. Recently, the observation of enhancement in electric eld-induced strain was also

reported for Li doped BNKT-based ceramic [10,11]. The BNKT


ceramics show the Smax/Emax of 297 pm/V at MPB where both
tetragonal and rhombohedral symmetries were existed [37]. During doping/co-doping and/or solid solution with other perovskite
ABO3 type, the Smax/Emax was further enhanced and comparable
with soft-PZT (e.g. Smax/Emax  727 pm/V for (Li,Ta)-codoped BNKT,
Smax/Emax  715 pm/V for La-doped BNKT etc.). The enhancement
of Smax/Emax results from phase transition from polar (tetragonal/
rhombohedral) phase to non-polar (pseudocubic) phase [10,38].
Recently, the controlled phase transition from pesudocubic to
tetragonal and/or rhombohedral by multidopants was displayed a
higher Smax/Emax value than that of phase transition from polar to
non-polar [11,39]. However, the distorted tetragonal structure of
BNKT such as Zr4 -modied at Ti4 -site also shows high Smax
/Emax value of 614 pm/V [6]. Our result indicated that distorted
structure due to modifying at A-site resulted in enhancement of
Smax/Emax values which are higher than that of modidication only
B-site of BNKT-based ceramics. Based on our research, the observation of enhancement electrical eld induced giant strain in
BNKT-based materials originated most from: (i) distorted structure
of tetragonal/rhombohedral, (ii) phase transition from tetragonal
and/or rhombohedral to pseudocubic phase, and (iii) reversion of
phase transition from pseudocubic phase to tetragonal and/or
rhombohedral by doping or solid solution with other ABO3 perovskite phases. However, we note that the origin of reserved phase
transitions were still not well understood and it was required for
further theoretical investigations.
We proposed the way to enhance the Smax/Emax values of BNKTbased ceramics which were expected to approach comparable
piezoelectric properties of PZT-based ceramics, as shown in Fig. 9.
The pure BNKT ceramics had low Smax/Emax values  297 pm/V,
then these values were strongly enhanced due to replacing by
small amount Ti4 ions at octahedral-site via other element such
as Hf4 , Zr4 , Ta5 , Y3 ions etc. [416,37]. The results indicated
that the distored structure by modifying at B-site was a strong
factor to archive the enhancement of Smax/Emax values, but those
were smaller than the values which were obtained due to phase
transition from polar to non-polar [59]. However, the observation
reported recently indicated that the Smax/Emax values were further
increased when the non-polar phase transited to polar phase via
A-site and B-site co-modication [11,14]. Note that, in this case,
the B-site modication was xed at optimal dopant level [11,39].
In addition, the A-site substitution strongly affected to the structure such as distorted structure and/or promotion the phase
transition. Thus, controlling substitution at the A-site seems to be
as important as B-site modication in lead-free BNKT-based
ceramics to enhance the electric eld induced strain.

Fig. 7. (a) Relative dielectric constants at a frequency of 1 kHz of BNKTZxLi ceramics as a function of temperature, and (b) variations of Td and Tm with different amount of
Li cations additive for the BNKTZ ceramics at 1 kHz.

N.D. Quan et al. / Journal of Physics and Chemistry of Solids 77 (2015) 6267

67

Fig. 8. (a) Unipolar strain hysteresis loops of Li doped BNKTZ ceramics, (b) the normalized strain and Smax/Emax values as a function of Li content.

Fig. 9. The schematic diagram for the way of enhancement the electric-eld-induced strain in Bi0.5(Na,K)0.5TiO3 ceramics by modied structures.

4. Conclusions
The role of Li dopant on properties of lead-free
Bi0.5(Na0.78K0.22)0.5Ti0.97Zr0.03O3 ceramics has been reported. The
Li acted as multisite selections which occurred at both Na -site
and octahedral site. The Smax/Emax values were increased from 600
to 643 pm/V while saturation polarization decreased from 46.2 to
26.1 C/cm2. The band gap reduced from 2.88 to 2.68 eV by adding
5 mol% Li. We expect that this work could be helpful for further
understanding the role of A-site substitution in case of co-modication at both A- and B-sites due to limitation of the substitutions at B-site.

Acknowledgement
This work was nancially supported by the Ministry of Education and Training, Vietnam under Project no. B 2013.01.55.

References
[1] B. Jaffe, W.R. Cook, H. Jaffe, Piezoelectric Ceramics, Academics, London, 1971.
[2] Y. Li, K.S. Moon, C.P. Wong, Science 380 (2005) 14191420.

[3] L.E. Cross, Nature (London) 432 (2004) 2425.


[4] A. Sasaki, T. Chiba, Y. Mamiya, E. Otsuki, Jpn. Apl. Phys. 38 (1999) 55645567.
[5] A. Hussain, C.W. Ahn, A. Ullah, J.S. Lee, I.W. Kim, Jpn. J. Appl. Phys. 49 (2010)
(041504(6).
[6] A. Hussain, C.W. Ahn, J.S. Lee, A. Ullah, I.W. Kim, Sens. Actuators A 158 (2010)
8489.
[7] K.N. Pham, A. Hussain, C.W. Ahn, I.W. Kim, S.J. Jeong, J.S. Lee, Mater. Lett. 64
(2010) 22192222.
[8] N.B. Do, H.B. Lee, C.H. Yoon, J.K. Kang, J.S. Lee, I.W. Kim, Trans. Electric. Electron. Mater. 12 (2011) 6467.
[9] D.N. Binh, A. Hussain, H.D. Lee, I.W. Kim, J.S. Lee, I.W. Kim, W.P. Tai, J. Korean
Phys. Soc. 57 (2010) 892896.
[10] V.Q. Nguyen, H.S. Han, K.J. Kim, D.D. Dang, K.K. Ahn, J.S. Lee, J. Alloys Compd.
511 (2012) 237241.
[11] V.Q. Nguyen, C.H. Hong, H.Y. Lee, Y.M. Kong, J.S. Lee, K.K. Ahn, J. Korean Phys.
Soc. 61 (2012) 895898.
[12] H.S. Han, C.W. Ahn, I.W. Kim, A. Hussain, J.S. Lee, Mater. Lett. 70 (2012) 98100.
[13] I.K. Hong, H.S. Han, C.H. Yoon, H.N. Ji, W.P. Tai, J.S. Lee, J. Intell. Mater. Syst.
Struct. 0 (2012) 17.
[14] N.D. Quan, V.N. Hung, N.V. Quyet, H.V. Chung, D.D. Dung, AIP Adv. 4 (2014)
(017122 (7).
[15] A. Hussain, J.U. Rahman, A. Zaman, R.A. Malik, J.S. Kim, T.K. Song, W.J. Kim, M.
H. Kim, Mater. Chem. Phys. 143 (2014) 12821288.
[16] A. Ullah, C.W. Ahn, A. Hussain, I.W. Kim, H.I. Hwang, N.K. Cho, Solid State
Commun. 150 (2010) 11451149.
[17] D. Lin, D. Xiao, J. Zhu, P. Yu, Appl. Phys. Lett. 88 (2006) (062901(3).
[18] P.Y. Chen, C.C. Chou, T.Y. Tseng, H. Chen, Jpn. J. Appl. Phys. 49 (2010) 061506.
[19] Z. Yang, B. Liu, L. Wei, Y. Hou, Mater. Res. Bull. 43 (2008) 8189.
[20] H. Xie, L. Jin, D. Shen, X. Wang, G. Shen, J. Cryst. Growth 311 (2009)
36263630.
[21] M. Otonicar, S.D. Skapin, M. Spreitzer, D. Suvorov, J. Eur. Ceram. Soc. 30 (2010)
971979.
[22] R.D. Shannon, Acta Crystallogr. A 32 (1976) 751767.
[23] Y.D. Hou, L.M. Chang, M.K. Zhu, X.M. Song, H. Yan, J. Appl. Phys. 102 (2007)
(084507(7).).
[24] L.D. Vuong, P.D. Gio, Inter. J. Mater. Sci. Appl. 2 (2013) 8993.
[25] N. Lei, M. Zhu, P. Yang, L. Wang, L. Wang, Y. Hou, H. Yan, J. Appl. Phys. 109
(2011) (054102(6).).
[26] B. Parija, T. Badapanda, V. Senthil, S.K. Rout, S. Panigrahi, Bull. Mater. Sci. 35
(2012) 197202.
[27] S. Lee, R.D. Levi, W. Qu, S.C. Lee, C.A. Randall, J. Appl. Phys. 107 (2010) (023523
(6).
[28] S. Lee, W.H. Woodford, C.A. Randall, Appl. Phys. Lett. 92 (2008) (201909(3).
[29] I. Levin, E. Cockayne, V. Krayzman, J.C. Woicik, S. Lee, C.A. Randall, Phys. Rev. B
83 (2011) (094122(8).).
[30] J.C. Jan, H.M. Tsai, C.W. Pao, J.W. Chiou, K. Asokan, Y.H. Tang, M.H. Tsai, S.Y. Kuo,
W.H. Hsieh, Appl. Phys. Lett. 87 (2005) (012103(3).
[31] V.A. Shuvaeva, D. Zekria, A.M. Glezer, Q. Jiang, S.M. Weber, P. Bhattacharya,
Phys. Rev. B 71 (2005) (174114(8).
[32] N.W. Thomas, J. Phys. Chem. Solids 51 (1990) 14191431.
[33] T.Y. Kim, H.M. Jang, Appl. Phys. Lett. 77 (2000) (3824(3).
[34] X.H. Dai, A. Digiovanni, D. Viehland, J. Appl. Phys. 74 (1993) (3399(7).
[35] Y. Yuan, S. Zhang, X. Zhuo, J. Mater. Sci.: Mater. Electron. 20 (2009) 10901094.
[36] I. Grinberg, M.R. Suchomel, P.K. Davies, A.M. Rappe, J. Appl. Phys. 98 (2005)
(094111(10).
[37] M. Izumi, K. Yamamoto, M. Suzuki, Y. Noguchi, M. Miyayama, Appl. Phys. Lett.
93 (2008) (242903(3).
[38] T.H. Dinh, H.Y. Lee, C.H. Yoon, R.A. Malik, Y.M. Kong, J.S. Lee, J. Korean Phys. Soc.
62 (2013) 10041008.
[39] Q.N. Van, L.H. Bac, D.D. Dung, The way of enhancement of electrical eldinduced-strain in lead-free piezoelectric ceramics: the role of phase transition,
unpublished.

Вам также может понравиться