Structure and Electrical Properties of Iro - Doped 0.5ba Ca Tio - 0.5bati ZR O Ceramics Via Low-Temperature Sintering

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

J Mater Sci (2015) 50:6134–6141

DOI 10.1007/s10853-015-9170-2

Structure and electrical properties of IrO2-doped


0.5Ba0.7Ca0.3TiO3–0.5BaTi0.8Zr0.2O3 ceramics
via low-temperature sintering
Yongshang Tian1 • Yansheng Gong1 • Dawei Meng1 • Yuanjian Li1

Received: 31 March 2015 / Accepted: 9 June 2015 / Published online: 13 June 2015
Ó Springer Science+Business Media New York 2015

Abstract Lead-free ceramics 0.5Ba0.7Ca0.3TiO3– Introduction


0.5BaTi0.8Zr0.2O3–xIrO2 (x = 0–1.2 %) were prepared at a
low temperature of 1260 °C via sintering the as-synthe- Lead-based materials have been conventionally used in
sized nanoparticles, which were synthesized by a modified piezoelectric devices that convert mechanical energy to
Pechini polymeric precursor method. All samples featured electrical energy and vice versa for more than 50 years [1].
high levels of densification under the synthesis conditions; However, the use of lead-based materials has been pro-
particularly, the ceramic prepared at IrO2 content of 0.4 % hibited owing to their high toxicity [2]. Consequently, great
achieved the highest relative density (*97.4 %). Phase efforts have been devoted to the development of eco-
transition from tetragonal to orthorhombic and a poly- friendly lead-free materials as alternatives. Among the
morphic phase transition (PPT) region were observed, lead-free candidates studied, BaTiO3 (BT)-based materials
which were influenced by the IrO2 content. The dielectric are regarded as promising materials owing to their con-
properties, temperature coefficient of capacitance, ferro- trollable sinterability, excellent thermostability, and out-
electric properties, and piezoelectric properties as a func- standing electrical properties [3]. However, the application
tion of IrO2 content were thoroughly studied. Optimal of BT-based materials is limited by their low Curie tem-
electrical properties (remnant polarization, piezoelectric perature (TC) and single functional performance [4]. To
constant, and planar electromechanical coupling factors, address these issues, many researchers have attempted
i.e., Pr = 6.28 lC/cm2, d33 = 199 pC/N, and kp = 0.258) modification of BT-based materials by doping/substituting
were obtained around the PPT region. The findings of the elements, broadening their performance capabilities by
ceramic electrical properties by IrO2 doping are believed to endowing the materials with optical, electrical, acoustic,
be insightful in the development of lead-free dielectric and or/and magnetic properties [5–8].
ferroelectric materials. Current research involves the study of BT-based mate-
rials with a perovskite structure (ABO3) via doping rare
elements, and IA, IIA, and VIIB elements in A/B sites for
the multifunctional performance [9, 10]. Among the
research, 0.5Ba0.7Ca0.3TiO3–0.5BaTi0.8Zr0.2O3 (BCT–
BZT) system materials had been received significant
attention for the comparable piezoelectric properties with
the lead-based materials, since the morphotropic phase
boundary region was demonstrated in this system [6].
& Yansheng Gong Moreover, various methods in the preparation and electri-
gongysh@cug.edu.cn
cal performance research on BCT–BZT system were
& Dawei Meng developed, which promoted the application of lead-free
dwmeng@cug.edu.cn
materials for the replacement of lead-based materials [11–
1
Faculty of Material Science and Chemistry, China University 16]. As known, platinum group elements, e.g., IrO2 and
of Geosciences, Wuhan 430074, People’s Republic of China RuO2, have potential application in catalysis,

123
J Mater Sci (2015) 50:6134–6141 6135

electrochemistry, and microelectronic devices [17, 18]. detected by field-emission scanning electron microscopy
However, comprehensive studies on BT-based materials (FESEM; SUV-1080), and the average grain size of the
containing doped platinum group elements are yet to be ceramics was estimated using the linear intercept method.
reported. Such materials are expected to have outstanding The density of the ceramics was determined based on
electrical properties and valuable uses in electrode buffer Archimedes immersion principle using a precision elec-
materials of multilayer ceramic capacitors. tronic balance (ED-124S). The relative permittivity (er),
Our previous studies reported the preparation of BT- loss tangent (tan d), and capacitance values were deter-
based ceramics at a relatively low sintering temperature of mined using a precision LCR meter (TH-2819) at 10 kHz.
1260 °C via sintering the as-synthesized nanoparticles, A Radiant Precision Workstation was used to investigate
which were synthesized by a modified Pechini polymeric polarization–electric field (P–E) hysteresis loops. The
precursor method and showed homogeneous dispersion of resonance and anti-resonance frequencies of the poled
the ingredients, narrow particles size distribution, and samples were measured on a precision impedance analyzer
stable structure [19, 20]. The prepared BT-based ceramics (Agilent 4294A) to calculate the planar and thickness
with the as-synthesized nanoparticles featured high densi- vibration electromechanical coupling factors (kp and kt)
fication, few crystal defects, and enhanced electrical and mechanical quality factor (Qm). The piezoelectric
properties, which had been reported elsewhere [20, 21]. constant (d33) was measured on a quasi-static meter (ZJ-
The properties of platinum group element, iridium (Ir), and 3A) at room temperature.
its compound were also studied in our previous studies [21,
22]. In this work, IrO2-doped BCT–BZT ceramics were
prepared using the above-discussed process. The phase Results and discussion
transition, crystal structure, and electrical properties of the
ceramics as a function of IrO2 content were systematically Figure 1 shows the XRD patterns of the BCT–BZT–xIrO2
studied. ceramics doped with different IrO2 contents (x). All
detected samples featured a pure perovskite (ABO3)
structure with no secondary phases in solid solutions,
Experimental procedure implying that iridium was successfully impregnated into
the ABO3 lattice. As observed in Fig. 1b, the unique
The as-synthesized Ba0.7Ca0.3TiO3 (BCT) and BaTi0.8- splitting of the (002) and (200) diffraction peaks at *45°
Zr0.2O3 (BZT) nanoparticles, which were synthesized by a corresponding to the tetragonal (T) phase (JCPDS #
modified Pechini polymeric precursor method that has been 05-0626) appeared when x \ 0.4 % [15] and merged into
reported elsewhere [19–21], and IrO2 (Ir C 86.0 %) pow- (200) diffraction peak corresponding to the rhombohedral
der were stoichiometric mixed to obtain 0.5BCT–0.5BZT–
xIrO2 powders (abbreviated as BCT–BZT–xIrO2; x = 0,
0.2, 0.4, 0.6, 0.8, 1.0, and 1.2 %). Then, the powders were
dispersed homogeneously and uniaxially pressed into disks
of 20 mm in diameter and *1.5 mm in thickness mixed in
the presence of 2.5 wt% polyvinyl alcohol (PVA) solution
binder at 150 MPa. Following burnt out of PVA at 650 °C,
the green bodies were sintered at 1260 °C for 5 h in air
atmosphere to obtain the BCT–BZT–xIrO2 ceramics. For
the electrode preparation, the ceramics were polished and
both sides of the disks surfaces were silver-painted, then
annealed at 580 °C for 20 min for subsequent dielectric
and ferroelectric measurements. Then, the samples were
poled in a silicone oil bath under a direct-current electric
field of 18 kV/cm at 25 °C for *40 min for the piezo-
electric measurements.
X-ray diffraction (XRD; X’Pert PRO) with Cu Ka
radiation was used to investigate the crystal phase of the
ceramics; the analysis was conducted at room temperature.
The samples were fractured and subjected to a thermal
Fig. 1 a XRD patterns and b corresponding selected enlarged regions
etching process (1050 °C, 15 min, in air atmosphere). (44°–47°) of the BCT–BZT–xIrO2 ceramics prepared with varying
Then, the fractured microstructures of the ceramics were IrO2 contents (x)

123
6136 J Mater Sci (2015) 50:6134–6141

Fig. 2 a Variations in the lattice parameters of the BCT–BZT–xIrO2 ceramics prepared with different IrO2 contents (x) and b crystalline
schematic of the ceramic prepared with x = 1.0 %

(R) phase (JCPDS # 85-1796) when x [ 0.8 % [23]. Both average lattice parameters in Table 1 were calculated by
T and R phases were observed in the intermediate x range Jade 5.0 after the Rietveld refinement using Fullprof soft-
of 0.4–0.6 %. Thus, the results showed the phase transition ware. The dimension of the lattice parameters a, b, and
of the BCT–BZT–xIrO2 ceramics with increasing x. As c decreased slightly at IrO2 content below 0.4 %. More
observed in the selected enlarged region (44°–47°) in prominent decreases were observed at IrO2 content above
Fig. 1b, subtle variations were detected. The diffraction 0.8 %. Variations in the lattice parameters were attributed
peaks shifted monotonically to low 2h angles with to the incorporation Ir4? into the B sites in the ABO3
increasing x. According to the principles of crystal chem- structure that led to crystal lattice distortion and instigated
istry relating to element doping positions in ABO3 structure phase transition [16]. A schematic of the crystal structure
reported by Lee et al. [24] and Morrison et al. [25], iridium of BCT–BZT–xIrO2 ceramics (x = 1.0 %), featuring
generally occupied the B sites. Thus, the shift in the R phase with average lattice parameters (3.9901(8) Å,
position of the diffraction peaks was attributed to the 3.9902(1) Å, 3.9901(9) Å, and 89.88°) is shown in Fig. 2b,
decreased interplanar spacing owing to the substitution of which was obtained by Crystal maker software, revealing
Zr4? (ionic radius 0.087 nm) with a smaller Ir4? the internal crystal structure of the ceramic.
(0.072 nm) in the ABO3 structure [26]. Additionally, the Figure 3 shows the FESEM images of the BCT–BZT–
subtle variation in the intensity of diffraction peaks with xIrO2 ceramics doped with different IrO2 contents. All
increasing x suggested variations in the crystalline prop- samples featured a dense microstructure with few cavities
erties of the ceramics including the onset of a distorted as a result of the sintering process. The average grain size
crystal lattice by IrO2 doping. of the ceramics was *0.7 lm, which was considerably
Figure 2a shows the variations in the average lattice smaller than that reported in the literature studies [27, 28].
parameters (a, b, c, and axial angle) of the BCT–BZT– The fine grain size of the ceramics was believed to be
xIrO2 ceramics doped with varying IrO2 contents. The related to the raw material of the as-prepared nanoparticles.

Table 1 Lattice parameters, density and relative densities (DRE), average grain size (GAV) of the BCT–BZT–xIrO2 ceramics prepared with
varying IrO2 contents (x)
x (%) a (Å) b (Å) c (Å) Axial angle (°) Phase Density (g/cm3) DRE (%) GAV (lm)

0 3.9907(7) 3.9907(7) 4.0014(3) 90 T 5.483 96.41 0.713(2)


0.2 3.9905(5) 3.9906(2) 4.0005(3) 90 T 5.514 96.96 0.711(7)
0.4 – – – – T&R 5.537 97.36 0.707(3)
0.6 – – – – T&R 5.502 96.75 0.699(5)
0.8 3.9903(6) 3.9903(9) 3.9904(7) 89.93 R 5.458 95.97 0.690(4)
1.0 3.9901(8) 3.9902(1) 3.9901(9) 89.88 R 5.437 95.61 0.685(6)
1.2 3.9897(7) 3.9898(2) 3.9897(7) 89.84 R 5.413 95.18 0.694(9)

123
J Mater Sci (2015) 50:6134–6141 6137

Fig. 3 FESEM of the fractured morphology of the BCT–BZT–xIrO2 ceramics prepared with different IrO2 contents (x) of a 0 %, b 0.2 %,
c 0.4 %, d 0.6 %, e 0.8 %, f 1.0 %, and g 1.2 %

The average grain size decreased slightly with the distorted, with the formation of some fine grains at the
increasing x. This finding was attributed to the diffusion of grain boundaries, at higher x. This finding was the result of
small Ir4? ions into the matrix of the ABO3 structure. the solubility limit of iridium in the ABO3 structure of the
Additionally, the ceramic crystal lattice was slightly BCT–BZT–xIrO2 ceramics [29].

123
6138 J Mater Sci (2015) 50:6134–6141

densification of the ceramics increased with increases in x,


reaching maximum at x = 0.4 % (density: 5.537 g/cm3;
relative density: 97.36 %), and then decreased with further
increases in x. The enhanced densification was associated
with the replacement of Ti/Zr by iridium with heavier
atomic mass in the ABO3 structure. In contrast, the
decrease in the densification at higher x was due to crystal
lattice distortion [16]. The relative densities of the samples
were all above 95 %, which is indicative of the high den-
sification of the materials featuring few cavities.
Figure 5 shows the temperature dependence of relative
permittivity (er) and loss tangent (tan d) of the BCT–
BZT–xIrO2 ceramics doped with different IrO2 contents
assessed at a frequency of 10 kHz. All samples showed
Fig. 4 Variations in the density and relative density of the BCT– broadened dielectric peeks banded with a relatively low
BZT–xIrO2 ceramics prepared with varying IrO2 contents (x) permittivity under the testing conditions, which was
attributed to fine grain size, as consistent with the above
analysis [27]. The details of dielectric properties are
shown in Table 2. From the detected results, it can be
seen that tan d first decreased and then increased with
increasing x. This trend was due to the high densification
of the ceramics at x \ 0.6 % and subsequent formation of
fine grains at the grain boundaries. A weak dielectric
relaxation behavior was detected when x [ 0.8 % that
was associated with the disordered crystal structure and
polar nanoregions [26]. Furthermore, phase transition of
the dielectric peaks of the samples was not clearly
detected mainly because of the diffused dielectric char-
acteristic with fine grain size [30]. The inset in Fig. 5
shows the variations in the temperature coefficient of
capacitance (TCC; DC/C) of the BCT–BZT–xIrO2
ceramics as a function of IrO2 content. The TC values of
Fig. 5 Temperature dependence of the relative permittivity (er) and loss
tangent (tan d) of the BCT–BZT–xIrO2 ceramics measured at 10 kHz. the samples remained unchanged, indicating that TC was
The inset (a) shows the temperature coefficient of capacitance of the independent of x. The detected DC/C values (-35 % to 0)
BCT–BZT–xIrO2 ceramics prepared with different IrO2 contents (x) and low tan d satisfied the Y5Y and X7R specifications of
Electronic Industries Association standards at high tem-
The density and relative density values of the BCT– peratures ([TC) [31, 32]. Additionally, the subtle change
BZT–xIrO2 ceramics doped with different IrO2 contents are in the DC/C values was attributed to the large internal
shown in Fig. 4 and Table 1. As observed, the level of stress [33].

Table 2 The relative permittivity (er) and loss tangent (tan d), remnant polarization (Pr) and coercive field (Ec), piezoelectric constant (d33),
planar electromechanical coupling factors (kp), mechanical quality factor (Qm) of the BCT–BZT–xIrO2 ceramics with different IrO2 contents
(x) at 25 °C
x (%) er tan d (%) Pr (lC/cm2) Ec (kV/cm) d33 (pC/N) kp Qm

0 2849 6.9 6.31 3.57 138 0.213 112.34


0.2 2883 5.8 6.52 3.63 162 0.233 99.01
0.4 2954 5.2 6.28 2.99 184 0.244 78.17
0.6 3013 4.6 5.96 3.85 199 0.258 92.78
0.8 2918 8.7 5.80 5.40 179 0.221 116.89
1.0 2877 9.5 6.21 5.96 171 0.209 117.41
1.2 2797 9.9 6.11 6.08 162 0.207 126.77

123
J Mater Sci (2015) 50:6134–6141 6139

To study the degree of relaxor-like ferroelectric charac- frequency of 10 kHz. As observed, c of the samples first
teristics of the BCT–BZT–xIrO2 ceramics, a modified decreased and then increased, which indicated that dif-
Curie–Weiss law (Eq. 1) was used as described below: fuseness was receded with increasing x (\0.6 %). However,
1 1 ðT  Tm Þc the diffuseness of the ceramics increased when x [ 0.8 %.
 ¼ ; ð1Þ This finding was related to ingredient heterogeneity and
er em C
structural disorder/crystal lattice distortion [16]. The higher
where em is the maximum relative permittivity; Tm is the relaxor state (c = *1.9) of the ceramics than the literature
temperature associated with em; C is the modified Curie– reports was believed to be due to the balance between the
Weiss constant; and c is the diffuseness exponent. Param- altered long-range and short-range forces in the large grain
eter c value ranges from 1 (normal ferroelectric) to 2 (typ- boundaries volume with fine grain sizes [14, 35].
ical relaxor ferroelectric) [34] and is calculated from the The P–E hysteresis loops of the BCT–BZT–xIrO2
slope of the graph of ln(1/er - 1/em) versus ln(T - Tm). ceramics doped with different IrO2 contents are shown in
Figure 6 shows the variation of c of the BCT–BZT–xIrO2 Fig. 7. All hysteresis loops were typically saturated, sug-
ceramics doped with different IrO2 contents assessed at a gesting good ferroelectric properties. However, the area
associated with the hysteresis loops increased when
x [ 0.6 %, indicating that the electrical loss increased with
excessive IrO2 content. This finding is consistent with the tan
d results in Fig. 5. Variations in the remnant polarization (Pr)
and coercive field (Ec) in the selected enlarged region (-6.8
to 0 kV/cm) can be observed in the inset of Fig. 7 and
Table 2. As observed, Pr first enhanced and then deteriorated
with increasing x. The enhanced Pr was associated with easy
switched domain in the consistence of T and R phases (PPT
region), and the similar results were reported in the literature
[36, 37]. The increased Ec with excessive IrO2 content was
attributed to a high restoring force of domain owing to the
dipole defects in the distorted crystal structure [38].
Figure 8 shows the variations in the piezoelectric con-
stant (d33), planar and thickness vibration electromechani-
cal coupling factors (kp and kt), and mechanical quality
Fig. 6 Plots of ln(1/er - 1/em) as a function of ln(T - Tm) of the factor (Qm) of the BCT–BZT–xIrO2 ceramics doped with
BCT–BZT–xIrO2 ceramics prepared with different IrO2 contents different IrO2 contents. As observed, d33 reached a maxi-
(x) assessed at 10 kHz. The inset (a) shows selected enlarged regions mum value of 199 pC/N at x = 0.6 % in Table 2 owing to
(1.1–1.5)
the substitution of B sites (Ti: 1.54; Zr: 1.33) with higher
electronegative Ir (2.20) in the ABO3 structure of the

Fig. 8 Variations in the piezoelectric constant (d33), planar and


Fig. 7 Polarization–electric field (P–E) hysteresis loops of the BCT– thickness vibration electromechanical coupling factors (kp and kt), and
BZT–xIrO2 ceramics prepared with different IrO2 contents (x). The mechanical quality factor (Qm) of the BCT–BZT–xIrO2 ceramics with
inset (a) shows selected enlarged regions (-6.8 to 0 kV/cm) varying IrO2 contents (x)

123
6140 J Mater Sci (2015) 50:6134–6141

ceramics. Substitution resulted in sp3 hybridization of 3. Chen J, Chen XL, He F, Wang YL, Zhou HF, Fang L (2014)
covalency with more covalent bonds and improved piezo- Thermally stable BaTiO3-Bi(Mg0.75W0.25)O3 solid solutions:
sintering characteristics, phase evolution, Raman spectra, and
electric properties [39]. Moreover, the elevated d33 at dielectric properties. J Electron Mater 43:1112–1118
x = 0.6 % was due to the easy polarization in PPT regions in 4. Hennings D, Schreinemacher H (1995) Ca-acceptors in dielectric
the presence of an external electric field. However, d33 ceramics sintered in reductive atmospheres. J Eur Ceram Soc
deteriorated considerably with excessive x, which was 15:795–800
5. Zhang Q, Sun H, Wang X, Zhang Y, Li X (2014) Strong pho-
associated with the distorted crystal structure and ingredient toluminescence and piezoelectricity properties in Pr-doped
heterogeneity [40]. Likewise to d33, kp and kt featured similar Ba(Zr0.2Ti0.8)O3-(Ba0.7Ca0.3)TiO3 ceramics: influence of con-
trend. In contrast, Qm showed an opposite tend. Specifically, centration and microstructure. J Eur Ceram Soc 34:1439–1444
Qm reached a minimum value of 78.17 at x = 0.4 %, and the 6. Liu WF, Ren XB (2009) Large piezoelectric effect in Pb-free
ceramics. Phys Rev Lett 103:257602
relative lower Qm than the literature could enhance the fre- 7. Wu J, Wu W, Xiao D, Wang J, Yang Z, Peng Z, Chen Q, Zhu J
quency band of the electron component [15, 28]. The (2012) (Ba, Ca)(Ti, Zr)O3-BiFeO3 lead-free piezoelectric
observed trend around PPT regions was attributed to the ceramics. Curr Appl Phys 12:534–538
relatively low internal stress and weak pinching effects of the 8. Kaushal A, Olhero S, Singh B, Zamiri R, Saravanan V, Ferreira J
(2014) Successful aqueous processing of a lead free
domain wall in the PPT region [36, 37]. 0.5Ba(Zr0.2Ti0.8)O3-0.5(Ba0.7Ca0.3)TiO3 piezoelectric material
composition. RSC Adv 4:26993–27002
9. Shrout TR, Zhang SJ (2007) Lead-free piezoelectric ceramics:
Conclusions alternatives for PZT? J Electroceram 19:113–126
10. Zhang L, Ren X (2006) Aging behavior in single-domain Mn-
doped BaTiO3 crystals: implication for a unified microscopic
0.5Ba0.7Ca0.3TiO3–0.5BaTi0.8Zr0.2O3–xIrO2 ceramics were explanation of ferroelectric aging. Phys Rev B 73:094121
prepared via sintering of the as-synthesized nanoparticles 11. Kaushal A, Olhero S, Singh B, Fagg D, Bdikin I, Ferreira J
at a temperature of 1260 °C. The ceramic samples featured (2014) Impedance analysis of 0.5Ba(Zr0.2Ti0.8)O3-0.5(Ba0.7-
Ca0.3)TiO3 ceramics consolidated from micro-granules. Ceram
high levels of densification with a fine grain size. The XRD Int 40:10593–10600
results confirmed the occurrence of phase transition with 12. Kaushal A, Olhero S, Ferreira J (2013) Lead-free 0.5Ba(Zr0.2-
increasing IrO2 content and coexisted T and R phases (PPT Ti0.8)O3-0.5(Ba0.7Ca0.3)TiO3 powder surface treated against
region) when x * 0.6 %. Slightly enhanced dielectric hydrolysis—a key for a successful aqueous processing. J Mater
Chem C 1:4846–4853
properties and excellent TCC at high temperatures were 13. Wu W, Cheng L, Bai S, Dou W, Xu Q, Wei Z, Qin Y (2013)
obtained that were dependent on the IrO2 content under the Electrospinning lead-free 0.5Ba(Zr0.2Ti0.8)O3-0.5(Ba0.7Ca0.3)-
synthesis conditions. The diffuseness dielectric character- TiO3 nanowires and their application in energy harvesting.
istics were weakened, and the ferroelectric and piezoelec- J Mater Chem A 1:7332–7338
14. Mishra P, Kumar SP (2012) Effect of sintering temperature on
tric properties were enhanced in the PPT region, indicating dielectric, piezoelectric and ferroelectric properties of BZT-BCT
that the electrical properties were strongly influenced by 50/50 ceramics. J Alloys Compd 545:210–215
IrO2 doping. The deteriorated electrical properties with 15. Srinivas A, Krishnaiah RV, Niranjani VL, Kamat SV, Karthik T,
excessive x were attributed to the presence of dipole Asthana S (2015) Ferroelectric, piezoelectric and mechanical
properties in lead free (0.5)Ba(Zr0.2Ti0.8)O3-(0.5)(Ba0.7Ca0.3)-
defects and the distorted crystal structure. The electrical TiO3 electroceramics. Ceram Int 41:1980–1985
properties of the ceramics obtained by IrO2 doping sug- 16. Ehmke MC, Daniels J, Glaum J, Hoffman M, Blendell JE,
gested that the prepared ceramics are promising candidates Bowman KJ (2013) In situ X-ray diffraction of biased ferroelastic
for application as lead-free dielectric and ferroelectric switching in tetragonal lead-free (1 - x)Ba(Zr0.2Ti0.8)O3-
x(Ba0.7Ca0.3)TiO3 piezoelectrics. J Am Ceram Soc 96:2913–2920
materials. 17. Liu YX, Masumoto H, Goto T (2005) Structural, electrical and
optical characterization of SrIrO3 thin films prepared by laser-
Acknowledgements The work was supported by Fundamental ablation. Mater Trans 46:100–104
Research Funds for National University (CUG120118), State Key 18. Choi KJ, Baek SH, Jang HW, Belenky LJ, Lyubchenko M, Eom
Laboratory of Advanced Technology for Materials Synthesis Pro- CB (2010) Phase-transition temperatures of strained single-crys-
cessing (Wuhan University of Technology, 2012-KF-3). The authors tal SrRuO3 thin films. Adv Mater 22:759–762
express sincere thanks to Dr. Hongquan Wang for FESEM 19. Tian YS, Gong YS, Zhang ZL, Meng DW (2014) Phase evolu-
measurements. tions and electric properties of BaTiO3 ceramics by a low-tem-
perature sintering process. J Mater Sci Mater Electron 25:5467–
5474
20. Tian YS, Gong YS, Meng DW, Li YJ, Kuang BY (2015)
References Dielectric dispersion, diffuse phase transition, and electrical
properties of BCT-BZT ceramics sintered at a low-temperature.
1. Cross E (2004) Materials science: lead-free at last. Nature J Electron Mater. doi:10.1007/s11664-015-3727-3
432:24–25 21. Tian YS, Gong YS, Meng DW, Cao SQ (2015) Structure and
2. Damjanovic D, Klein N, Li J, Porokhonskyy V (2010) What can electrical properties of Ir4?-doped 0.5Ba0.9Ca0.1TiO3-
be expected from lead-free piezoelectric materials? Funct Mater 0.5BaTi0.88Zr0.12O3-0.12%La ceramics via a modified Pechini
Lett 03:5–13 method. Mater Lett 153:44–546

123
J Mater Sci (2015) 50:6134–6141 6141

22. Tian YS, Gong YS, Wu MY, Meng DW, Jin HY (2014) Study on 32. Shen ZB, Wang XH, Gong HL, Wu LW, Li LT (2014) Effect of
the densification of BaIrO3 bulks by spark plasma sintering and MnO2 on the electrical and dielectric properties of Y-doped
its films deposition. J Electroceram. doi:10.1007/s10832-014- Ba0.95Ca0.05Ti0.85Zr0.15O3 ceramics in reducing atmosphere.
9961-x Ceram Int 40:13833–13839
23. Wu J, Xiao D, Wu W, Chen Q, Zhu J, Yang Z, Wang J (2012) 33. Maiti T, Guo R, Bhalla A (2008) Structure property phase dia-
Composition and poling condition-induced electrical behavior of gram of BaZrxTi1-xO3 system. J Am Ceram Soc 91:1769–1780
(Ba0.85Ca0.15)(Ti1-xZrx)O3 lead-free piezoelectric ceramics. J Eur 34. Merz WJ (1950) The effect of hydrostatic pressure on the Curie
Ceram Soc 32:891–898 point of barium titanate single crystals. Phys Rev 77:52–54
24. Lee YC, Yeh YY, Tsai PR (2012) Effect of microwave sintering on 35. Jeong IK, Park CY, Ahn JS, Park S, Kim DJ (2010) Ferroelectric-
themicrostructure and electric properties of (Zn, Mg)TiO3-based relaxor crossover in Ba(Ti1-xZrx)O3 studied using neutron total
multilayer ceramic capacitors. J Eur Ceram Soc 32:1725–1732 scattering measurements and reverse Monte Carlo modeling.
25. Morrison FD, Sinclair DC, Skakle JMS, West AR (1998) Novel Phys Rev B 81:214119-1–214119-5
doping mechanism for very-high-permittivity barium titanate 36. Zhang JL, Zong XJ, Wu L, Gao Y, Zheng P, Shao SF (2009)
ceramics. J Am Ceram Soc 81:1957–1960 Polymorphic phase transition and excellent piezoelectric perfor-
26. Dong L, Stone DS, Lakes RS (2012) Enhanced dielectric and mance of (K0.55Na0.45)0.965Li0.035Nb0.80Ta0.20O3 lead-free
piezoelectric properties of xBaZrO3-(1 - x)BaTiO3 ceramics. ceramics. Appl Phys Lett 95:022909-1–022909-3
J Appl Phys 111:084107 37. Dai YJ, Zhang XW, Chen KP (2009) Morphotropic phase
27. Chen T, Zhang T, Wang G, Zhou J, Zhang J, Liu Y (2012) Effect boundary and electrical properties of K1-xNaxNbO3 lead-free
of CuO on the microstructure and electrical properties of ceramics. Appl Phys Lett 94:042905-1–042905-3
Ba0.85Ca0.15Ti0.90Zr0.10O3 piezoceramics. J Mater Sci 47:4612– 38. Zhou PF, Zhang BP, Zhao L, Zhao XK, Zhu LF, Cheng LQ, Li JF
4619. doi:10.1007/s10853-012-6326-1 (2013) High piezoelectricity due to multiphase coexistence in
28. Lin W, Fan L, Lin D, Zheng Q, Fan X, Sun H (2013) Phase low-temperature sintered (Ba, Ca)(Ti, Sn)O3-CuOx ceramics.
transition, ferroelectric and piezoelectric properties of Ba1-x- Appl Phys Lett 103:172904–1–172904-5
CaxTi1-yZryO3 lead-free ceramics. Curr Appl Phys 13:159–164 39. Saito Y, Takao H, Tani T, Nonoyama T, Takatori K, Homma T,
29. Zhang L, Thakur OP, Feteira A, Keith GM, Mould AG, Sinclair Nagaya T, Nakamura M (2004) Lead-free piezoceramics. Nature
DC, West AR (2007) Comment on the use of calcium as a dopant 432:84–87
in X8R BaTiO3-based ceramics. Appl Phys Lett 90:142914- 40. Coondoo I, Panwar N, Amorı́n H, Alguero M, Kholkin AL (2013)
1–142914-3 Synthesis and characterization of lead-free 0.5Ba(Zr0.2Ti0.8)O3-
30. Tang X, Wang J, Wang X, Chan H (2004) Effects of grain size on 0.5(Ba0.7Ca0.3)TiO3 ceramic. J Appl Phys 113:214107–1–
the dielectric properties and tunabilities of sol-gel derived 214107-6
Ba(Zr0.2Ti0.8)O3 ceramics. Solid State Commun 131:163–168
31. Hong JY, Lu HY (2014) Polar nanoregions and dielectric prop-
erties of BaTiO3-based Y5V multilayer ceramic capacitors. J Am
Ceram Soc 97:2256–2263

123

You might also like