Sei sulla pagina 1di 21

Accepted Manuscript

Title: Enhanced Curie temperature and piezoelectric


properties of (Ba0.85 Ca0.15 )(Zr0.10 Ti0.90 )O3 lead-free ceramics
after the addition of LiTaO3

Authors: Mingxing Zhou, Ruihong Liang, Zhiyong Zhou,


Chenhong Xu, Xin Nie, Xianlin Dong

PII: S0025-5408(17)34209-5
DOI: https://doi.org/10.1016/j.materresbull.2018.05.036
Reference: MRB 10033

To appear in: MRB

Received date: 9-11-2017


Revised date: 9-2-2018
Accepted date: 29-5-2018

Please cite this article as: Zhou M, Liang R, Zhou Z, Xu C, Nie X, Dong X, Enhanced
Curie temperature and piezoelectric properties of (Ba0.85 Ca0.15 )(Zr0.10 Ti0.90 )O3 lead-
free ceramics after the addition of LiTaO3 , Materials Research Bulletin (2018),
https://doi.org/10.1016/j.materresbull.2018.05.036

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Enhanced Curie temperature and piezoelectric properties of
(Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 lead-free ceramics after the addition of LiTaO3

Mingxing Zhou, a,b Ruihong Liang, a,* Zhiyong Zhou, a Chenhong Xu, a,b Xin Nie,
a,b and Xianlin Donga,**

a
Key Laboratory of Inorganic Functional Materials and Devices, Shanghai Institute of

T
Ceramics, Chinese Academy of Sciences, 1295 Dingxi Road, Shanghai 200050,

IP
People’s Republic of China

R
b
University of Chinese Academy of Sciences, Beijing 100049, People’s Republic of

SC
China

*Corresponding

**Corresponding U
author: E-mail address: liangruihong@mail.sic.ac.cn
author: E-mail address: xldong@mail.sic.ac.cn
N
A
Graphical Abstract
M
ED
E PT
CC
A

1
Research Highlights

 The Curie temperature and piezoelectric properties of BCZT ceramics can be


obviously enhanced by the addition of LiTaO3.
 The Curie temperature of (Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 ceramics improved from
93 ℃ to 102 ℃.
 (Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 ceramic with x = 0.3 mol. % LiTaO3 demonstrates

T
optimal piezoelectric properties.

IP
 The giant electrostrictive coefficients (Q33 ≈ 0.040-0.050 m4/C2) were also
obtained in LT-doped BCZT ceramics.

R
SC
Abstract

U
Polycrystalline (Ba0.85Ca0.15)(Zr0.10Ti0.90)O3–xLiTaO3 ceramics were designed
N
and synthesized using a solid-state reaction to improve both the Curie temperature and
A
the electrical properties of (Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 lead-free piezoelectric
ceramics. The effect that the addition of LiTaO3 has on the structure and electrical
M

properties of BCZT ceramics was studied systemically. The results show that the
ED

Curie temperature and the piezoelectric properties of BCZT can be substantially


enhanced by adding LiTaO3. The Curie temperature increased from 93 °C to 102 °C
after adding LT of 0.3 mol. %. X-ray diffraction and Raman spectral analysis of the
PT

ceramics reveal the coexistence of orthorhombic (O) and tetragonal (T) phases. This
coexistence of O and T phases at room temperature helps improve the ferroelectric
E

and piezoelectric properties. The ceramic with x = 0.3 mol. % exhibits the highest
CC


values for d33 = 433 pC/N, 𝑑33 = 858 pm/V (at 1 kV/mm) and kp = 44.6 %.
Furthermore, giant electrostrictive coefficients (Q33 ≈ 0.040-0.050 m4/C2) were also
A

obtained for these LT-doped BCZT ceramics. These properties confirm that
BCZT-xLT is a promising candidate system for lead-free piezoelectric ceramic
materials.

Keywords: Piezoelectric ceramics; Electrical properties; BCZT; Curie temperature;


2
Coexisting orthorhombic and tetragonal phase

1. Introduction

Piezoelectric ceramics have been widely-used in actuators, sensors, transducers


and many other applications for several decades due to their outstanding
electromechanical properties [1]. For a long time, Pb(Zr,Ti)O3 (PZT) ceramics have

T
played a dominant role in this field because of its excellent piezoelectric properties.

IP
However, the strong toxicity of lead oxides has the potential to create serious health

R
and environmental problems [2, 3]. Therefore, there is a need for good lead-free

SC
piezoelectric ceramics to replace PZT [4]. Recently, Ren et al. has discovered a large
piezoelectric effect (with d33 = 500–620 pC/N) in modified BaTiO3

U
(Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 (BCZT) [4]. It is a great breakthrough for lead-free
N
piezoelectric research and BCZT is expected to replace the widely used PZT family in
A
some applications. However, there are still drawbacks with BCZT. For example, the
Curie temperature (Tc) is only 85 ℃, which limits its practical application [5, 6]. For
M

these reasons, it would be important to improve the Tc, while maintaining the strong
ED

piezoelectric properties.
Previously, Randall et al. studied the crystallographic relation between the end
member ABO3 perovskite tolerance factor and the Curie temperature of PbTiO3-based
PT

boundary systems with a morphotropic phase. They found that the Curie temperature
increases in these PbTiO3-based systems with decreasing tolerance factor. The
E

improvement of the Curie temperature is attributed to the incorporation of the end


CC

member composition ABO3 with a small perovskite tolerance factor into PbTiO3. For
example, a solid solution of BiScO3 (t = 0.907) with PbTiO3 shows MPBs with Tc’s
A

significantly greater than PZT (Tc = 386 ℃) [7]. Hence, it is reasonable to infer that
the Tc of BCZT can be enhanced by decreasing the tolerance. It is known that LiTaO3
thanks to its ferroelectric phase (FE) possesses a high Curie temperature (665 ℃),
large electromechanical coupling coefficient (k33 = 0.60, k31 = 0.32), and strong
ferroelectric properties (Pmax ~ 50 µC/cm2) [1]. The cationic radius of Li+ (rLi+ =1.18
3
Å) is clearly smaller than that of A-site Ba2+ (rBa2+ =1.61 Å), and the radius of Ta5+
(rTa5+ =0.64 Å) is larger than that of B-site Ti4+ (rTi4+ = 0.605 Å) [8, 9]. Therefore, the
introduction of LiTaO3 into the BCZT ceramics can decrease its tolerance factor.
Additionally, almost all studies of LiTaO3 modified (K0.5Na0.5)NbO3 systems show
improved piezoelectric properties and Curie temperatures [10, 11]. Therefore, it is can
be anticipated that introducing LiTaO3 into BCZT can improve both its Curie
temperature and piezoelectric properties.

T
On the basis of the above considerations, a new lead-free piezoelectric ceramic,

IP
(Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 modified with LiTaO3, was synthesized through solid-state

R
reaction in this work (abbreviated as BCZT-xLT, x = 0.0, 0.1, 0.2, 0.3, 0.4 mol.%).

SC
The effect of LiTaO3 doping on the phase structure, microstructure, and the electrical
properties (including dielectric, ferroelectric, and piezoelectric) were studied
systemically.
U
N
A
2. Results and discussion
2.1. Phase and microstructure
M

Fig. 1(a) displays the X-ray patterns of the BCZT-xLT ceramics for the 2θ range
ED

of 10°–90°. All synthesized BCZT-xLT ceramics exhibit a pure perovskite structure


with no signs of any secondary phases, which indicates that LiTaO3 diffused into the
(Ba0.85Ca0.15)(Zr0.10Ti0.90)O3 lattice to form a complete solid solution in the studied
PT

composition. Fig. 1(b) shows the enlarged 2θ range for 44°–46°. When the content of
LiTaO3 increases, the peaks near 45° clearly shift toward higher diffraction angles,
E

which should be attributed to the shrinkage of lattice parameters of the BCZT-xLT


CC

ceramics according to Bragg’s equation (2d sinθ = λ). The change of lattice
parameters can be explained by the effect of LiTaO3 diffusing into the crystal structure.
A

Considering the cationic radius and valence of Li+ and Ta5+ [8, 9], Li+ should enter the
12-fold coordinated A-site (rLi+ =1.18 Å, rBa2+ =1.61 Å, rCa2+ =1.34 Å, Coordination
Number = 12) and Ta5+ can occupy the 6-fold coordinated B-site (rTa5+ =0.64 Å, rZr4+
= 0.72 Å, rTi4+ = 0.605 Å, Coordination Number = 6). Therefore, the shrinkage of
lattice parameters of the BCZT-xLT ceramics possibly occurs because the radius
4
effects caused by the occupation of smaller Li+ at the A sites are more dominant than
the ones caused by larger Ta5+ at the B sites (Fig. S1).
As shown in Fig. 1(b), it is clear that orthorhombic and tetragonal phases coexist
in the BCZT-xLT. This confirms that the polymorphic phase boundary nature of
these compositions, which is further confirmed by the slight broadening and splitting
trend of the diffraction reflection as shown in Fig. 1(b). Similar XRD results were
found in previous reports [12, 13]. In order to clarify the variation of phase structure

T
and lattice parameters XRD powder patterns were collected at room temperature and

IP
refined using the GSAS software [14] with the graphical interface EXPGUI [15]. The

R
background correction was performed using the Chebyshev polynomial of the first

SC
kind, and the diffraction peak profiles were fitted using the pseudo-Voigt/FCJ peak
asymmetry function. Rietveld refinement of XRD data was performed using a

U
combination of tetragonal and orthorhombic symmetries with the space groups P4mm
N
(JCPDS: 05-0626) and Amm2 (JCPDS: 81-2200) [16], respectively. The X-ray
A
powder patterns were fitted to the calculated ones using a full-profile analysis
program to minimize the profile discrepancy factor Rp, which converged well for all
M

compositions – see Fig. 2(a)-(e). The corresponding lattice parameters, phase content,
ED

and fit quality are shown in Table S1. As the LT mole content increases, the lattice
parameters decrease. This is consistent with the diffraction peak position shifting
toward a higher diffraction angle in Fig. 1(b).
PT

Raman spectroscopy is a powerful technique for the study of ferroelectric


materials because of the close relationship between ferroelectricity and lattice
E

dynamics. It is an effective method to study the structure effect thanks to its


CC

sensitivity to local symmetry [17]. To further confirm the phase structure of


BCZT-xLT ceramics, room temperature Raman spectra of the samples were also
A

conducted for the range of 90-1000 cm-1 - see Fig. 2(f). Nine characteristic vibration
modes were traced and numbered from 1 to 9. The phonon modes 2, 4, 5, 7, and 8 of
all samples were assigned to the T phase, with detailed phonon assignments, including
the detection of an interference dip at 170 cm−1 (mode 2), a broad peak at 242 cm−1
(A1(TO), mode 4), a sharp peak at 293 cm−1 (E(TO + LO) + B1, mode 5), a broad peak
5
at 523 cm−1 (E(TO)+A1(TO), mode 7), and 724 cm−1 (E(LO)+(A1(LO), mode 8) [18].
The clear spectral dip at 170 cm−1 (mode 2) is known to result from the interference of
three A1(TO) an-harmonic coupling modes [18]. Mode 5 at 293 cm−1 is assigned to be
characteristic of the T phase [19]. A clear mode was labeled as mode 3–206 cm−1 and
reported to be another sub-band assigned to the A1(LO) mode [20, 21]. This band is
reported to be more prominent in the rhombohedral (R) and O phases, and it
disappears in the T phase [20-22]. Thus, the appearance of mode 3 A1(LO) and a shift

T
of mode 4 A1(TO) couples with the decoupling of three A1(TO) modes, which

IP
suggests the presence of an O phase. The 470 cm−1 (E(LO)+A1(LO)+E(TO), mode 6)

R
peak loses its intensity around the orthorhombic/tetragonal (O/T) transition [23].

SC
Alternatively, the relative intensity of mode 2 at 170 cm−1 becomes higher, which can
be induced by phonon damping due to the coexistence of the orthorhombic phase [24].

U
This indicates that all samples coexist with O and T phases at room temperature since
N
its coexistence in BCZT is well known [13, 25]. This behavior of the O phase can be
A
observed by monitoring the relative intensity of Raman modes 3 and 4 (I206/I242). The
values of I206/I242 decrease from 1.002 to 0.863 for x = 0.0 and x = 0.4 ceramics, which
M

indicates a decreasing fraction ratio for the O phase. These results reveal the
ED

coexistence of orthorhombic and tetragonal phases in all samples, which is consistent


with the results of the Rietveld refinement shown in Fig. 2(a)-(e).
Fig. 3(a)-(e) shows the surface morphology of BCZT-xLT with different LiTaO3
PT

contents. All samples have surfaces with very few pores and exhibit high density with
sufficient grain growth. In order to easily identify the average grain size of BCZT-xLT,
E

grain size distributions were measured using analytical software (Nano Measurer).
CC

The statistics results on the grain size distributions are shown in the inset in Fig.
3(a)-(e). The effect of LT content on the variation in sintered density and average
A

grain size of BCZT-xLT ceramics is shown in Fig. 3(f). After the addition of LiTaO3,
the density of the samples slightly increased. The average grain size increased from
5.40 µm to 7.30 µm, while similar result were observed in NaNbO3-LiTaO3 [26] and
BiFeO3-BaTiO3-LiTaO3 ceramics [27].
2.2.Dielectric properties
6
The temperature and frequency variations of the dielectric constant of BCZT-xLT
are shown in Fig. 4(a)-(e). To highlight the phase transitions, we plotted the curve for
the differential dielectric constant (∂𝜀𝑟 / ∂T) as a function of temperature in the insets
of Fig. 4(a)-(e). The BCZT-xLT at x = 0.0, 0.1, 0.2, 0.3, and 0.4 mol.% exhibit two
obvious polymorphic phase transitions that correspond to the orthorhombic–tetragonal
(TO–T) and tetragonal–cubic transitions (Tc), respectively. In Fig. 4(f), we show a
simplified phase diagram based our dielectric measurements. The Tc increased first

T
and then decreased with increasing LT content. A maximum Tc value (102 ℃) can be

IP
obtained for x = 0.3 mol. %. The improvement of Tc can be attributed to the addition of

R
LiTaO3 with a high Curie temperature (665 ℃). In addition, the lattice symmetry

SC
increases with the increase of tetragonal phase content, Therefore, the required energy
to induce a phase transition from ferroelectric to paraelectric increased, which results

U
in an increased Tc. However, the Tc of BCZT-xLT shifts to lower temperatures when x
N
changes from 0.3 to 0.4. This Tc decrease may be caused by the appearance of greater
A
ion disorder and inner stress. Ion disorder affects long-range interactions, which
results in lower transition temperatures. Similar results were reported for
M

(Ba0.85Ca0.15)(Zr0.1Ti0.9)O3-(Ca0.28Ba0.72)Nb2O6 ceramics [28], and


ED

(Ba0.85Ca0.15)(Zr0.1Ti0.9)O3-LiNbO3 ceramics [29, 30].


2.3.Ferroelectric properties
Fig. 5(a) shows the P-E hysteresis loops for BCZT-xLT with different LT
PT

contents measured at room temperature and 1 Hz. All ceramics exhibit a


well-saturated P-E loop in an electric field of 4 kV/mm. P-E loops become slimmer
E

with increasing LT content, which can be attributed to the increase in grain size of
CC

BCZT-xLT - see Fig. 3(f). The basic variation in grain size affects the domain size of
the ceramics. A larger grain size is associated with larger domains and fewer domain
A

walls. Weak ferroelectric and piezoelectric properties are the results of a large number
of grain boundaries and domain walls. These make domain reorientation more
difficult and severely constrain domain wall motion [31]. Larger grains have fewer
domain walls and thus a small electric field (Ec) is sufficient to reorient the domains,
and the loops become narrower with increasing LT content. The composition depends
7
on the remnant polarization Pr and the coercive field Ec - see Fig. 5(b). It can be seen
that Pr of BCZT-xLT increases with increasing LT content up to x = 0.3, then it
decreases. Ceramics with 0.3 mol. % LT content show a maximum Pr of 9.16 µC/cm2
and a low Ec of 0.32 kV/mm. The low Ec and large Pr indicate easy reversal of
polarization under direct current bias, which improves the piezoelectric properties of
BCZT-xLT ceramics. The decrease in Pr for x = 0.4 mol. % can be because the
difference in valence and ionic radius affects the long-range ferroelectric order in

T
large doping amounts, which reduces the ferroelectric properties [6].

IP
The S-E curves after bipolar cycling at 1 kV/mm are shown in Fig. 5(c). An

R
electric field-induced strain-loop with a typical V-shaped curve was observed for all

SC
compositions. This type of bipolar strain shape is different from the one observed in
soft KNN-based [32] and BNT-based lead-free ceramics [33]. The negative strain in

U
all compositions is close to zero, which can be attributed to the relaxor behavior by
N
the addition of LiTaO3. For a driving field of 1 kV/mm, the electric-field induced
A
strain increased from 0.0659 % for undoped BCZT to 0.0858 % for the 0.3 mol. % LT
modified BCZT sample. The corresponding dS/dE values for these compositions were
M

659 and 858 pm/V, respectively. The variations of dS/dE as a function of LT content
ED

for BCZT-xLT are shown in Fig. 5(d). The increase in useable strain (Smax) may be
ascribed to an intrinsic piezoelectric strain effect. This effect is caused by the
chemical inhomogeneity at unit cell level and the addition of a small amount of LT.
PT

Furthermore, the increase of electro-strain also originates from the decreasing TO–T
close to room temperature, which causes the easy rotation of non-180° domains [34].
E

However, upon further increase in LT content, this decrease in Smax was attributed to
CC

the presence of a short-range relaxor state for 0.4 mol. % LT content, which is evident
from the observed decrease in Pr and Ec. A similar behavior was also observed in
A

BNT-based and BFO-based ceramics [27, 35]. A small asymmetry appears in the S-E
loops. The absence of an “imprint” effect may be attributed to the presence of crystal
defects and a space-charge field in the ferroelectric [36]. Fig. 5(e) shows the S−E
curves for bipolar cycling at 4 kV/mm. A higher electric field resulted in a smaller
hysteresis, similar to the one reported for Fe3+-doped BCZT [37]. The electrostrictive
8
coefficients in this work are between 0.04 and 0.05 m4/C2, which indicates that
BCZT-xLT also could be a good candidate for electrostrictive applications.
2.4.Piezoelectric properties
The composition dependence of d33 and kp is given in Fig. 6(a). It can be seen
that the piezoelectric properties of LiTaO3 doped BCZT is superior to undoped BCZT.
Both d33 and kp exhibit similar trends within the LT content range. The d33 and kp
increase before decreasing with increasing LT content, which is similar to the effect of

T
LT content on dS/dE. The largest value of d33 is 433 pC/N, and kp reaches 44.6 % for x

IP
= 0.3 mol. %. This is consistent with the maximum low-field piezoelectric response of

R
the BZT-BCT system, which occurs slightly above the O-T phase boundary [13, 38].

SC
The improvement of the piezoelectric properties is related to the orthorhombic and
tetragonal phases that can coexist at room temperature. The vanishing of polarization

U
anisotropy in the orthorhombic-tetragonal phase boundary composition causes a low
N
energy barrier for polarization rotation, which is the most important factor responsible
A
for enhanced piezoelectric properties [4]. The enhanced piezoelectric properties also
can be partly attributed to the dense microstructures. Furthermore, the lattice
M

distortion induced by LT doping may be beneficial to ferroelectric domain


ED

reorientation during the poling process [39]. However, the value of d33 and kp
decreased slightly for the 0.4 mol. % LT sample, which may be due to the increased
relaxor behavior in this sample [6].
PT

Fig. 6(b) summarizes the Curie temperature and piezoelectric properties of


previous studies of conventionally (solid-state) synthesized BCZT samples with
E

different additions, allowing immediate comparison with the results of this study. All
CC

data were taken from the references [5, 6, 12, 29, 40-47]. Fig. 6(b) shows that LiTaO3
doped BCZT not only improves the Curie temperature, but it also maintains its
A

relative strong piezoelectric properties. Therefore, LiTaO3 modified BCZT is a


promising candidate to replace lead-based piezoelectric materials for many
applications.
3. Conclusions

9
BCZT-xLT ceramics were synthesized using a conventional solid-state reaction.
The effect of added LiTaO3 on the phase structure, microstructure, dielectric,
ferroelectric, and piezoelectric properties of BCZT-xLT ceramics were investigated
systematically. The XRD patterns and Raman spectra analysis reveal the coexistence
of orthorhombic and tetragonal phases in BCZT-xLT. The Curie temperature increased
from 93 °C (pure BCZT) to 102 °C (x = 0.3 mol. %). The multiphase containing both
O and T phases at room temperature improves the ferroelectric and piezoelectric

T
properties of BCZT-xLT. Both ferroelectric and piezoelectric properties increase with

IP
increasing LT content up to 0.3 mol. %. However, any further increase reduces the

R
piezoelectric properties. The highest direct and converse piezoelectric constants as

SC

well as the planar coupling factor d33= 433 pC/N, Smax = 0.0858 % (at 1 kV/mm), 𝑑33

= 858 pm/V (at 1 kV/mm) Smax = 0.21 % (at 4 kV/mm), 𝑑33 = 518 pm/V (at 4

U
kV/mm) and kp = 44.6 % were obtained for the sample with 0.3 mol. % LT. A giant
N
electrostrictive coefficient (Q33 ≈ 0.040-0.050 m4/C2) was also obtained in LT-doped
A
BCZT. These results will help find future practical applications of BCZT-xLT.
M

Acknowledgements
ED

This work was supported by the National Natural Science Foundation of China
(Grant No. 11574334) and the National Key Basic Research Program of China (973
PT

Program, Grant No. 2015CB057502).


E
CC
A

10
References

[1] Y. Saito, H. Takao, T. Tani, T. Nonoyama, K. Takatori, T. Homma, T. Nagaya, M.


Nakamura, Lead-free piezoceramics, Nature. 432 (2004) 84-87.
[2] F.Z. Yao, K. Wang, W. Jo, K.G. Webber, T.P. Comyn, J.X. Ding, B. Xu, L.Q.
Cheng, M.P. Zheng, Y.D. Hou, Diffused phase transition boosts thermal stability of
high-performance lead-free piezoelectrics, Adv. Funct. Mater. 26 (2016) 1217-1224.

T
[3] X. Wang, J. Wu, D. Xiao, J. Zhu, X. Cheng, T. Zheng, B. Zhang, X. Lou, X. Wang,

IP
Giant piezoelectricity in potassium-sodium niobate lead-free ceramics, J. Am. Chem.

R
Soc. 136 (2014) 2905-2910.

SC
[4] W. Liu, X. Ren, Large piezoelectric effect in Pb-free ceramics, Phys. Rev. Lett.
103 (2009) 257602.

U
[5] X. Liu, Z. Chen, B. Fang, J. Ding, X. Zhao, H. Xu, H. Luo, Enhancing
N
piezoelectric properties of BCZT ceramics by Sr and Sn co-doping, J. Alloy. Compd.
A
640 (2015) 128-133.
[6] R. Hayati, M.A. Bahrevar, T. Ebadzadeh, V. Rojas, N. Novak, J. Koruza, Effects of
M

Bi2O3 additive on sintering process and dielectric, ferroelectric, and piezoelectric


ED

properties of (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 lead-free piezoceramics, J. Eur. Ceram. Soc.


36 (2016) 3391-3400.
[7] R.E. Eitel, C.A. Randall, T.R. Shrout, P.W. Rehrig, W. Hackenberger, S.-E. Park,
PT

New high temperature morphotropic phase boundary piezoelectrics based on


Bi(Me)O3–PbTiO3 ceramics, Jpn. J. Appl. Phys. 40 (2001) 5999.
E

[8] R. Ubic, Revised method for the prediction of lattice constants in cubic and
CC

pseudocubic perovskites, J. Am. Ceram. Soc. 90 (2007) 3326-3330.


[9] R. Ubic, G. Subodh, The prediction of lattice constants in orthorhombic
A

perovskites, J. Alloy. Compd. 488 (2009) 374-379.


[10] Y.P. Guo, K. Kakimoto, H. Ohsato, (Na0.5K0.5)NbO3-LiTaO3 lead-free
piezoelectric ceramics, Mater. Lett. 59 (2005) 241-244.
[11] L. Li, Y.Q. Gong, L.J. Gong, H. Dong, X.F. Yi, X.J. Zheng, Low-temperature
hydro/solvothermal synthesis of Ta-modified K0.5Na0.5NbO3 powders and
11
piezoelectric properties of corresponding ceramics, Mater. Design. 33 (2012) 362-366.
[12] W. Li, Z. Xu, R. Chu, P. Fu, P. An, Effect of Ho doping on piezoelectric
properties of BCZT ceramics, Ceram. Int. 38 (2012) 4353-4355.
[13] N. Chaiyo, D.P. Cann, N. Vittayakorn, Lead-free (Ba,Ca)(Ti,Zr)O3 ceramics
within the polymorphic phase region exhibiting large, fatigue-free piezoelectric
strains, Mater. Design. (2017).
[14] A.C. Larson, R.B. Von Dreele, Gsas, General structure analysis system. LANSCE,

T
MS-H805, Los Alamos, New Mexico (1994).

IP
[15] B.H. Toby, EXPGUI, A graphical user interface for GSAS, J. Appl. Crystallogr.

R
34 (2001) 210-213.

SC
[16] I. Coondoo, N. Panwar, R. Vidyasagar, A.L. Kholkin, Defect chemistry and
relaxation processes: effect of an amphoteric substituent in lead-free BCZT ceramics,
Phys. Chem. Chem. Phys. 18 (2016) 31184-31201.
U
N
[17] M. DiDomenico Jr, S. Wemple, S. Porto, R. Bauman, Raman spectrum of
A
single-domain BaTiO3, Phys. Rev. 174 (1968) 522.
[18] P. Fleury, P. Lazay, Acoustic-soft-optic mode interactions in ferroelectric BaTiO3,
M

Phys. Rev. Lett. 26 (1971) 1331.


ED

[19] C. Han, J. Wu, C. Pu, S. Qiao, B. Wu, J. Zhu, D. Xiao, High piezoelectric
coefficient of Pr2O3-doped Ba0.85Ca0.15Ti0.90Zr0.10O3 ceramics, Ceram. Int. 38 (2012)
6359-6363.
PT

[20] C. Xiao, Z. Chi, W. Zhang, F. Li, S. Feng, C. Jin, X. Wang, X. Deng, L. Li, The
phase transitions and ferroelectric behavior of dense nanocrystalline BaTiO3 ceramics
E

fabricated by pressure assisted sintering, J. Phys. Chem. Solids. 68 (2007) 311-314.


CC

[21] G. Singh, V. Sathe, V. Tiwari, Investigation of orthorhombic-to-tetragonal


structural phase transition in (Ba1−xCax)(Zr0.05Ti0.95)O3 ferroelectric ceramics using
A

micro-Raman scattering, J. Appl. Phys. 115 (2014) 044103.


[22] S. Liu, L. Zhang, J. Wang, X. Shi, Y. Zhao, D. Zhang, Rapid stability of
ferroelectric polarization in the Ca, Ce hybrid doped BaTiO3 ceramics, Sci. Rep. 6
(2016) 38354.

12
[23] P. Dobal, R. Katiyar, Studies on ferroelectric perovskites and Bi‐layered
compounds using micro‐Raman spectroscopy, J. Raman. Spectrosc. 33 (2002)
405-423.
[24] I.B. Ouni, D. Chapron, H. Aroui, M.D. Fontana, Ca doping in BaTiO3 crystal:
Effect on the Raman spectra and vibrational modes, J. Appl. Phys. 121 (2017) 114102.
[25] D.S. Keeble, F. Benabdallah, P.A. Thomas, M. Maglione, J. Kreisel, Revised
structural phase diagram of (Ba0.7Ca0.3TiO3)-(BaZr0.2Ti0.8O3), Appl. Phys. Lett. 102

T
(2013) 092903.

IP
[26] H. Sun, Q. Zheng, Y. Wan, Q. Li, Y. Chen, X. Wu, K.W. Kwok, H.W.L.W. Chan,

R
D. Lin, Microstructure, electrical properties, and electric field‐induced phase

SC
transitions in NaNbO3-LiTaO3 lead‐free ceramics, Phys. Status. Solidi. A. 211 (2014)
869-876.

U
[27] F. Akram, A. Hussain, R.A. Malik, T.K. Song, W.-J. Kim, M.-H. Kim, Synthesis
N
and electromechanical properties of LiTaO3-modified BiFeO3–BaTiO3 piezoceramics,
A
Ceram. Int. (2017).
[28] X. Chao, J. Wang, P. Liang, T. Zhang, L. Wei, Z. Yang, Phase transition and
M

improved electrical performance of Ba0.85Ca0.15Zr0.1Ti0.9O3-Ca0.28Ba0.72Nb2O6


ED

ceramics with high Curie temperature, Mater. Design. 89 (2016) 465-469.


[29] T. Chen, T. Zhang, J. Zhou, J. Zhang, Y. Liu, G. Wang, Microstructure and
electrical properties of (Ba0.85Ca0.15)1−xLix(Ti0.90Zr0.10)1−xNbxO3 ceramics with a low
PT

dielectric loss and a low sintering temperature, Ceram. Int. 38 (2012) 3591-3594.
[30] S. Ye, C. Wu, J. Fuh, L. Lu, Structure, Dielectric and piezoelectric properties of
E

(Ba0.85Ca0.15)(Zr0.1Ti0.9)O3–LiNbO3 lead-free ceramics, J. Nanoeng. Nanomanuf. 4


CC

(2014) 85-92.
[31] J. Hao, W. Bai, W. Li, J. Zhai, Correlation between the microstructure and
A

electrical properties in high-performance (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 lead‐free


piezoelectric ceramics, J. Am. Ceram. Soc. 95 (2012) 1998-2006.
[32] Y. Zhao, Z. Xu, R. Chu, J. Hao, J. Du, G. Li, Improved piezoelectricity and high
strain response of (1−x)(0.948K0.5Na0.5NbO3−0.052LiSbO3)−xBi2O3 ceramics, J.
Mater. Sci.: Mater. Electron. 28 (2017) 1211-1216.
13
[33] P. Fu, Z.J. Xu, H.M. Zhang, R.Q. Chu, W. Li, M.J. Zhao, Structure and electrical
properties of Er2O3 doped 0.82Bi0.5Na0.5TiO3-0.18Bi0.5K0.5TiO3 lead-free piezoelectric
ceramics, Mater. Design. 40 (2012) 373-377.
[34] L.E. Cross, Materials for adaptive structural acoustic controls, Pennsylvania State
Univ Report, (1994).
[35] Z. Arif, H. Ali, M. Rizwan Ahmed, M. Adnan, N. Sahn, K. Myong-Ho, Dielectric
and electromechanical properties of LiNbO3-modified (BiNa)TiO3–(BaCa)TiO3

T
lead-free piezoceramics, J. Phys. D:. Appl. Phys. 49 (2016) 175301.

IP
[36] X. Wen, C. Feng, L. Chen, S. Huang, Dielectric tunability and imprint effect in

R
Pb(Mg1/3 Nb2/3)O3-PbTiO3 ceramics, Ceram. Int. 33 (2007) 815-819.

SC
[37] L. Jin, R. Huo, R. Guo, F. Li, D. Wang, Y. Tian, Q. Hu, X. Wei, Z. He, Y. Yan,
Diffuse phase transitions and giant electrostrictive coefficients in lead-free Fe3+-doped

U
0.5Ba(Zr0. 2Ti0. 8)O3-0.5(Ba0.7Ca0.3)TiO3 ferroelectric ceramics, ACS. Appl. Mater.
N
Inter. 8 (2016) 31109-31119.
A
[38] M. Acosta, N. Novak, W. Jo, J. Rödel, Relationship between electromechanical
properties and phase diagram in the Ba (Zr0.2Ti0.8)O3–x(Ba0.7Ca0.3)TiO3 lead-free
M

piezoceramic, Acta. Mater. 80 (2014) 48-55.


ED

[39] W. Li, Z. Xu, R. Chu, P. Fu, G. Zang, Temperature Stability in Dy‐Doped (Ba0.
99Ca0. 01)(Ti0. 98Zr0. 02) O3 Lead‐Free Ceramics with High Piezoelectric
Coefficient, J. Am. Ceram. Soc. 94 (2011) 3181-3183.
PT

[40] Y. Cui, X. Liu, M. Jiang, X. Zhao, X. Shan, W. Li, C. Yuan, C. Zhou, Lead-free
(Ba0.85Ca0.15)(Ti0.9Zr0.1)O3–CeO2 ceramics with high piezoelectric coefficient obtained
E

by low-temperature sintering, Ceram. Int. 38 (2012) 4761-4764.


CC

[41] C. Han, J.G. Wu, C.H. Pu, S. Qiao, B. Wu, J.G. Zhu, D.Q. Xiao, High
piezoelectric coefficient of Pr2O3-doped Ba0.85Ca0.15Ti0.90Zr0.10O3 ceramics, Ceram. Int.
A

38 (2012) 6359-6363.
[42] X. Xia, X. Jiang, C. Chen, X. Jiang, N. Tu, Y. Chen, Effects of Cr2O3 doping on
the microstructure and electrical properties of (Ba,Ca)(Zr,Ti)O3 lead-free ceramics,
Front. Mater. Sci. 10 (2016) 203-210.
[43] X.Y. Huang, C.-H. Gao, Z.W. Zhu, L. Pan, Z.G. Chen, Influence of Co2O3 doped
14
amount on the properties of (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 lead-free piezoelectric
ceramics, Piezoelectricity, Acoustic Waves and Device Applications (SPAWDA), 2012
Symposium on, IEEE, 2012, pp. 9-12.
[44] X. Chen, Y. Li, J. Zeng, L. Zheng, C.H. Park, G. Li, Phase transition and large
electrostrain in lead-free Li‐doped (Ba,Ca)(Ti,Zr)O3 Ceramics, J. Am. Ceram. Soc.
(2016).
[45] P. Parjansri, U. Intatha, S. Eitssayeam, Dielectric, ferroelectric and piezoelectric

T
properties of Nb5+ doped BCZT ceramics, Mater. Res. Bull. 65 (2015) 61-67.

IP
[46] Q. Lin, M. Jiang, D. Lin, Q. Zheng, X. Wu, X. Fan, Effects of La-doping on

R
microstructure, dielectric and piezoelectric properties of Ba0.85Ca0.15Ti0.90Zr0.10O3

SC
lead-free ceramics, J. Mater. Sci.: Mater. Electron. 24 (2013) 734-739.
[47] Z. Zhao, X. Li, H. Ji, Y. Dai, T. Li, Microstructure and electrical properties in

U
Zn-doped Ba0.85Ca0.15Ti0.90Zr0.10O3 piezoelectric ceramics, J. Alloy. Compd. 637 (2015)
N
291-296.
A
M
ED
E PT
CC
A

15
Figures and Captions

Fig. 1. (a) The XRD patterns of BCZT-xLT at x = 0.0, 0.1, 0.2, 0.3, 0.4. (b)
Magnification of the peaks at 2θ = 44°–46°.

T
R IP
SC
U
N
A
Fig. 2. XRD profiles for the Rietveld refinement results for BCZT-xLT with different
M

amounts of LT: (a) x = 0.0 mol.%, (b) x = 0.1 mol.%, (c) x = 0.2 mol.%, (d) x = 0.3
mol.%, (e) x = 0.4 mol.%, (f) Room-temperature Raman spectra of BCZT-xLT with
ED

different LT contents.
E PT
CC
A

16
T
R IP
SC
U
N
A
M
ED
PT

Fig. 3. SEM images of the surface of the BCZT-xLT: (a) x = 0.0 mol.%, (b) x = 0.1
E

mol.%, (c) x = 0.2 mol.%, (d) x = 0.3 mol.%, (e) x = 0.4 mol.%, (f) The density,
CC

relative density and average grain-size of BCZT-xLT, sintered at 1430 ℃ for 3h.
A

17
T
R IP
SC
Fig. 4. Temperature dependence of the dielectric constant of BCZT-xLT ceramics: (a)
x = 0.0 mol.%, (b) x = 0.1 mol.%, (c) x = 0.2 mol.%, (d) x = 0.3 mol.%, (e) x = 0.4

U
mol.%, the insets show the derivative of the dielectric constant as a function of
N
temperature for each composition, (f) phase diagram of BCZT-xLT based on the
A
present dielectric measurements.
M
ED
E PT
CC
A

18
T
IP
R
SC
U
N
A
M
ED
PT

Fig. 5. Ferroelectric properties of BCZT-xLT.


E
CC
A

19
T
R IP
SC
U
N
A
M
ED
PT

Fig. 6. (a) Piezoelectric properties of the BCZT-xLT. (b) d33 vs Tc of BCZT-based


materials, derived from other publications [5, 6, 12, 29, 40-47].
E
CC
A

20

Potrebbero piacerti anche