Sie sind auf Seite 1von 8

Available online at www.sciencedirect.

com

ScienceDirect
Journal of the European Ceramic Society 35 (2015) 533–540

Effect of Li2O addition on sintering and piezoelectric properties


of (Ba,Ca)(Ti,Sn)O3 lead-free piezoceramics
Lei Zhao a , Bo-Ping Zhang a,∗ , Peng-Fei Zhou a , Li-Feng Zhu a , Jing-Feng Li b
a School of Materials Science and Engineering, University of Science and Technology Beijing, Beijing 100083, China
b State Key Laboratory of New Ceramics and Fine Processing, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
Received 27 June 2014; received in revised form 22 August 2014; accepted 28 August 2014
Available online 16 September 2014

Abstract
This work investigated the enhanced sintering and electrical properties of (Ba0.95 Ca0.05 )(Ti0.90 Sn0.10 )O3 -xLi2 O (BCTSLix , x = 0, 1, 2, 3, 4, 5 and
6 mol%) lead-free piezoelectric ceramics benefiting from the Li2 O addition. Highly dense BCTS ceramics with good piezoelectric properties were
fabricated at reduced sintering temperatures by adding a small amount of Li2 O. It was found that the sintering densification is promoted mainly by
reaction liquid phase at the early sintering stage and the formation of oxygen vacancies should be considered as an additional factor. The coexistence
of R and T phases was confirmed by the Raman spectrum at a temperature range from−50 ◦ C to 40 ◦ C as x = 2 and at room temperature for all
other Li2 O added samples, while their specific ratio of R to T phase decreases with increasing Li2 O addition. Better properties with d33 = 457 pC/N,
kp = 25.18% and Qm = 491 were obtained when adding 4 mol% Li2 O, which are attributed dominantly to the appropriate specific ratio of R to T
phase because of the similar grain size and density for samples at 3 ≤ x ≤ 6.
© 2014 Elsevier Ltd. All rights reserved.

Keywords: Lead-free piezoceramics; Low sintering temperature; Li2 O; Raman spectra

1. Introduction at 1500 ◦ C have high d33 of 510 pC/N and 440 pC/N,
respectively. Higher d33 of 568 pC/N and 530 pC/N has
Pb(Zr1−x Tix )O3 (PZT) based ceramics have been widely been achieved in (Ba0.95 Ca0.05 )(Ti0.92 Sn0.08 )O3 6 and
used for more than 50 years as piezoelectric actuators, (Ba0.91 Ca0.09 )(Ti0.916 Sn0.084 )O3 7 systems, respectively.
sensors and transducers due to their excellent piezoelectric At present, the highest d33 in BaTiO3 –11BaSnO3 ceramic is
properties, but its use will be gradually replaced by lead-free 697 pC/N at 40 ◦ C in the vicinity of two converged triple points
piezoelectric ceramics such as BaTiO3 (BT), (K,Na)NbO3 or with coexisting multiphase.8 However, the sintering tempera-
(Bi1/2 Na1/2 )TiO3 for environment protecting concerns. BT tures of high-performance BCTS piezoelectric ceramics are all
was first discovered as a piezoelectric ceramic but is now over 1400 ◦ C,3–8 which impedes the practical applications of
mostly used as a dielectric material rather than a piezoelectric BT-based ceramics. Therefore, the densification temperature
material mainly because of its low Curie temperature (Tc) is need to be reduced for fabricating high-performance BT
and inferior piezoelectric properties (d33 = 191 pC/N) to PZT. based piezoelectric ceramics to realize large-scale applica-
Recently, high-performance BT based ceramics such as tions. Adding sintering aids is a simple, effective and mostly
(Ba,Ca)(Ti,Zr)O3 (BCTZ)1,2 and (Ba,Ca)(Ti,Sn)O3 (BCTS)3–7 used method to lower the sintering temperature. Li2 O as a
have been developed by co-doping Ca2+ /Zr4+ or Ca2+ /Sn4+ . Li sintering aid has been used to lower the sintering temperature
et al. [3,4] have reported that (Ba0.98 Ca0.02 )(Ti0.96 Sn0.04 )O3 of Ba(Ti0.95 Zr0.05 )O3 (from 1250 ◦ C to 900 ◦ C)9 and BT
and (Ba0.97 Ca0.03 )(Ti0.94 Sn0.06 )O3 ceramics both sintered (from 1300 ◦ C to 1000 ◦ C).10 Furthermore, the introduced
Li2 O can slightly increase Tc from 103 ◦ C to 107 ◦ C for
Ba(Ti0.95 Zr0.05 )O3 and from 120 ◦ C to 125 ◦ C for BT. Whether
∗ Corresponding author. Tel.: +86 10 62334195; fax: +86 10 62334195. the added Li2 O can not only lower the sintering temperature but
E-mail address: bpzhang@ustb.edu.cn (B.-P. Zhang). also enhance Tc in BCTS ceramics is still unclear. In the present

http://dx.doi.org/10.1016/j.jeurceramsoc.2014.08.042
0955-2219/© 2014 Elsevier Ltd. All rights reserved.
534 L. Zhao et al. / Journal of the European Ceramic Society 35 (2015) 533–540

Fig. 1. SEM images of thermally etched fracture surface for BCTSLix ceramics: (a) x = 0; (b) x = 1; (c) x = 2; (d) x = 3; (e) x = 4; (f) x = 5; (h) x = 6, and density (h).

study, the microstructure, crystal structure and electrical prop- where x is mol% of Li2 O) were prepared by conventional solid-
erties of Li2 O-added (Ba0.95 Ca0.05 )(Ti0.90 Sn0.10 )O3 ceramics state synthesis. The raw materials were BaCO3 , CaCO3 , SnO2
were investigated with a special emphasis on the influence of and TiO2 (>99%, all from ShanTouXilong Chemical Factory
Li2 O content (x). The low-temperature sintering mechanism of GuangDong, China). The weighed powders were mixed by using
Li2 O was also discussed. a planetary ball mill with anhydrous ethanol with 300 rpm for
4 h in a nylon jar with zirconia balls and then dried at 80 ◦ C for
2. Experimental procedure 8 h in an oven. After calcinating dried powders at 1250 ◦ C for
4 h, Li2 O (99%) in the range from 0 to 6 mol% was mixed with
A series of Li2 O-added (Ba0.95 Ca0.05 )(Ti0.90 Sn0.10 )O3 calcinated BCTS powders and then ball-milled again under the
ceramics (abbreviated as BCTSLix , x = 0, 1, 2, 3, 4, 5 and 6, same condition. Re-milled powders were dried at 80 ◦ C for 8 h in
L. Zhao et al. / Journal of the European Ceramic Society 35 (2015) 533–540 535

oven and pressed into disks of 10 mm in diameter and 1.5 mm in


thickness under 80 MPa using 2 wt.% polyvinyl alcohol (PVA)
as the binder. After burning out PVA at 650 ◦ C for 1 h in air,
samples were cooled to room temperature and then reheated to
1300 ◦ C with a rate of 5 ◦ C/min from room temperature and hold
for 4 h in air, and then cooled to room temperature with the same
rate. Silver electrodes were pasted on the top and bottom sur-
faces of the sintered samples and fired at 600 ◦ C for 30 min. A
poling was performed under an electric field of 3–4 kV/mm in
silicone oil bath for 30 min.
The phase structure was examined by X-ray diffraction
(XRD: D/max-RB, Rigaku Inc., Japan) with a Cu K␣ radiation
(λ = 1.5416 Å) filtered through a Ni foil. The apparent density
ρm was determined by Archimedes method. The relative density
ρr of BCTSLix ceramics was given by ρr = ρm /ρt × 100, where
ρt is the theoretical density which is calculated from the struc-
tural parameters determined from X-ray diffraction patterns by Fig. 2. Dynamic sintering curves of BT,11 BCTS and BCTSLi4 ceramics.
assuming they have pure T or R phase. The microstructure of
sintered samples was observed by field emission scanning elec-
tron microscopy (FE-SEM: SUPRATM55, Carl Zeiss, Nakano, 87%, which decrease to 4.95 g/cm3 and 81% at x = 1. Both appar-
Japan). Dielectric constant and dielectric loss were measured ent density and relative density increase and then decrease as x
using a LCR meter (TH2828S; Changzhou Tonghui Electronics increases from 2 to 6, and achieve peak values of 6.02 g/cm3 and
Co., Ltd. Changzhou, China) at 1 kHz. The Raman spectrum was 98.5% at x = 4. All samples at 2 ≤ x ≤ 6 have a relative density
measured by the LabRAM HR800 (Horiba JobinYvon, Paris, over 90%, showing a reduced sintering temperature by 150 ◦ C
France). The piezoelectric property was measured using a quasi- at least to obtain dense BCTS ceramics.
static piezoelectric coefficient d33 testing meter (ZJ-3A, Institute It is commonly considered that liquid phase contributes to
of Acoustics, Chinese Academy of Sciences, Beijing, China). grain growth and densification sintering in low temperature sin-
The planar electromechanical coupling coefficient kp was deter- tering since the liquid phase facilitates both the dissolution and
mined by resonance–antiresonance method using an Agilent migration of the species. However, the high melting point (above
4294A (Hewlett-Packard, Palo Alto, CA) precision impedance 1500 ◦ C) of Li2 O makes it impossible to form directly liquid
analyzer. Ferroelectric hysteresis loops were measured using a phase in the sintering process. The sintering mechanism of the
ferroelectric tester (RT6000HVA; Radiant Technologies, Inc., pristine BCTS and 4 mol% Li2 O doped counterpart were further
Albuquerque, NM). The shrinkage curves from 50 ◦ C to 1250 ◦ C investigated by the dynamic sintering curves as shown in Fig. 2.
for BCTS and BCTSLi8 ceramics were measured by a horizon- BCTS starts to shrink at 1050 ◦ C compared to that at 950 ◦ C
tal pushrod dilatometer (Netzsch DIL402PC, Selb, Germany) at for BT,11 indicating that BCTS ceramic has a higher sintering
a heating rate of 5 ◦ C/min. temperature than BT ceramics. While an obvious two-step sin-
tering regime is observed in BCTSLi4 ceramic. During the first
3. Results and discussion step at low temperatures (750–900 ◦ C) a reactive liquid-phase
sintering takes place in which an eutectic liquid may be formed
3.1. Microstructure and density by BaCO3 , Li2 O and TiO2 and leads to about 3% contraction.
Then, the liquid would be consumed during subsequent sinter-
Fig. 1 shows SEM images of thermally etched surface and ing. At x = 1, a small quantity of liquid is formed which is too
density of BCTSLix ceramics sintered at 1300 ◦ C for 4 h. The little to promote the grain growth, whereas the evaporation of
pristine BCTS ceramic in Fig. 1a has a porous microstruc- the liquid restrains grain growth by robbing heat which inhibits
ture with small grains of 2–4 ␮m, confirming that dense BCTS both grain growth and densification sintering. As a result, sam-
ceramic is difficult to obtain below 1300 ◦ C since its densifica- ple x = 1 shows the finest grains (Fig. 1b) and the lowest relative
tion temperature has so far been reported to be 1400–1540 ◦ C.3–8 density of 81% (Fig. 1h). As further increase in x, the content
The similar porous microstructure is formed in Fig. 1b with of the liquid is enough to promote the grain growth besides the
smaller grains of 1–2 ␮m after adding 1 mol% Li2 O. As fur- evaporated part. So the grain growth and densification sintering
ther increase in Li2 O content (x), a highly dense microstructure at 2 ≤ x ≤ 6 are ascribed to the reactive liquid-phase sintering.
along with grown grains is observed in Fig. 1c–g. Besides, intra- Apart from the reactive liquid-phase sintering mechanism
granular porosity remains and intergranular pores are trapped at during the first step at low temperatures, the emergence of
grain boundaries as x ≥ 3, along with an increased pore size as oxygen vacancies (VO•• ), generated in the matrix phase by
2BCTS
x ≥ 4. In accordance with the morphologies, the apparent den- the defect equation: Li2 O −→ 2Li ••
Ti/Sn + 3VO + TiO2 /SnO2 ,
sity and relative density of BCTSLix specimens show a similar may be another reason for the accelerated grain growth and den-
varying trend as shown in Fig. 1h. The apparent density for the sification sintering. The substitution of Li+ for high valence Ti4+
pristine BCTS is 5.26 g/cm3 along with the relative density of or Sn4+ leads to the creation of VO•• due to the ionic charge
536 L. Zhao et al. / Journal of the European Ceramic Society 35 (2015) 533–540

Fig. 3. X-ray diffraction patterns of BCTSLix ceramics.

compensation, which is favorable for the mass transport during the peak at about 2θ = 45◦ (abbreviated as peak-45) shifts to
sintering and greatly promotes grain growth and densification low angles as adding Li2 O as shown in Fig. 3b, indicating the
sintering.12 On the other hand, when all the liquid is consumed, enlargement of the lattice. The ionic radius of Li+ (0.74 Å) is
the mass transport through the liquid phase cannot be realized larger than the radii of Ti4+ (0.61 Å) and Sn4+ (0.71 Å) at B-sites,
anymore and the sintering is slowed down at 900–1020 ◦ C for but smaller than the radii of Ba2+ (1.34 Å) and Ca2+ (0.99 Å)
BCTSLi4 ceramic. Nevertheless, because of the presence of VO•• at A-sites. The substitution of Li+ for Ti4+ at B-sites would
in Li2 O doped BCTS matrix phase, the lattice-diffusion coeffi- enlarge the lattice.12,13 Besides, since cation with low valence
cient is higher than that of BCTS. This is reflected in the second has weaker binding force with O2− , the smaller binding force
sintering step, which starts at 1020 ◦ C for BCTSLi4 and is about with O2− of Li+ than that of Ti4+ and Sn4+ also results in the
30 ◦ C lower than that (1050 ◦ C) for BCTS. If the amount of expansion of lattice. So we speculate the enlarged lattice here
the liquid produced during the reaction is high enough, the sin- is due to the difference not only in the ionic size between Li+
tering would complete before all the liquid is consumed via and Ti4+ /Sn4+ but also in the binding force with O2− . Another
adding more Li2 O, which would result in the change from the characteristic in Fig. 3b is that all the peak-45 is asymmetric,
two-step sintering regime to one-step. Valant et al. [11] have ruling out the possibility of a single rhombohedral (R), pseudo-
found one-step sintering regime in 0.3 wt.% (about 2.3 mol%) cubic (PC) or cubic (C) phase where a symmetric singlet peak
Li2 O added BaTiO3 and 0.4 wt.% (about 2.9 mol%) Li2 O added existed. So BCTSLix ceramics may have a single tetragonal (T)
(Ba,Sr)TiO3 , in both cases the sintering temperature was signif- or orthorhombic (O) phase, or coexisting two or more phases
icantly reduced to 820 ◦ C and 880 ◦ C, respectively. At the same at room temperature. The full width at half maximum (FWHM)
time, the excess liquid and its subsequent evaporation would of peak-45 in Fig. 4a indicates an increased trend from 0.30◦ to
produce the formation of intergranular pores and their largen- 0.50◦ as x increases from 0 to 6. While the FWHM of peaks at
ing apart from the grown grains as x > 4 (Fig. 1e and f). Hence, about 2θ = 31.5◦ and 39◦ , abbreviated as peak-31.5 and peak-39,
the increased density and grain growth in Fig. 1 are attributed respectively, lies in 0.25–0.30◦ . The peak positions for (1 1 0)R ,
to both the reactive liquid-phase sintering and the generation of (1 1 0)R , (1 1 1)R , (1 1 1)R and (200)R in R phase (PDF#86–1569)
VO•• . and (1 0 1)T , (1 1 0)T , (1 1 1)T , (0 0 2)T and (2 0 0)T in T phase
(PDF#75–2264) are shown in Fig. 4b–d. The biggest difference
3.2. Phase structure between 2θ around 31.5◦ , 39◦ and 45◦ is 0.076◦ , 0.122◦ and
0.216◦ , respectively. The varied FWHM for peak-31.5, peak-39
Fig. 3 shows XRD patterns of BCTSLix ceramics compared and peak-45 reveals a phase transition which will be discussed
with the standard cards of R-phase (PDF#75–2264) and T-phase later rather than grain sizes, which is also well in accordance
(PDF#86–1569) BaTiO3 . All samples in Fig. 3a have a pure with those provided by FESEM.
perovskite structure without any trace of impurity within the The phase transition of BCTSLix ceramics was further
detectable limit of the XRD, indicating that added Li+ has diff- defined from the curves of dielectric permittivity versus tem-
used into BCTS lattices to form a solid solution. All the XRD perature (εr -T) as shown in Fig. 5. BCTSLi1 has a bad
patterns are similar to each other, suggesting the similar phase dielectric property due to its low density. All other sam-
symmetry at room temperature. Compared to the case of BCTS, ples have two phase-transition points and the higher one
L. Zhao et al. / Journal of the European Ceramic Society 35 (2015) 533–540 537

It is still difficult to determine the exact phase structure by εr -T


curve, even taking the XRD into consideration. The dielectric
loss (tanδ) is as high as 25% at room temperature as x = 0 and
is maintained at 0–2.5% as 1 ≤ x ≤ 6 in the whole measured
temperature. The variations of εr and tanδ with x at room tem-
perature in Fig. 5b show an increased εr and a decreased tanδ
as 2 ≤ x ≤ 6, which is attributed to the increased density. Fig. 5c
shows the frequency dependence of εr and tanδ for BCTSLi4 .
Compared to εr , tanδ is more sensitive to frequency.
Fig. 6a-1 and -2 shows the evolution of Raman spec-
tra with temperature for the pristine BCTS ceramic. Modes
at 170, 190, 250, 302, 490, 525 and 720 cm−1 observed at
−80 ◦ C in Fig. 6a-1 are all the feature of R phase especially
at about 170 and 190 cm−1 .15,16 Since most modes also can be
observed in O phase (including modes at 250, 302, 490, 525 and
720 cm−1 )15,16 and T phase (including modes at 250, 302, 525
and 720 cm−1 ),15–18 it is difficult to either exclude the exist-
ence of T or T and O phase(s) or determine their appearing
temperature. This Raman spectrum at −80 ◦ C is similar to the
case of Ba(Ti0.90 Sn0.10 )O3 16 at −100 ◦ C which has existing R,
O and T phases. We speculate that the pristine BCTS ceramic
Fig. 4. (a) The full width at half maximum (FWHM) of X-ray diffraction peaks in may also have coexisting R, O and T phases at −80 ◦ C. With
Fig. 3 abbreviated as peak-31.5, peak-39 and peak-45 as a function of x; (b–d) X-
increasing temperature to −40 ◦ C, mode at 490 cm−1 disappears
ray diffraction peaks in R (PDF#75–2264) and T phase (PDF#86–1569) around
a selected 2θ range of about 31.5◦ , 39◦ and 45◦ . or merges with the neighboring 525 cm−1 mode, indicating the
absence of O phase, which is also attributed to higher relative
intensity for mode at 190 cm−1 than that for mode at 250 cm−1
corresponds to paraelectric–ferroelectric (named TC ) phase above −40 ◦ C as marked by solid lines in Fig. 6a-2. No other
transition. TC of BCTS is 52 ◦ C and lower than 75 ◦ C of mode disappears in the temperature range of −40–20 ◦ C, indi-
(Ba0.91 Ca0.09 )(Ti0.916 Sn0.084 )O3 7 sintered at 1450 ◦ C, which is cating coexisting R and T phases, which is different from the
attributed to increased Sn4+ at B-sites.8 TC of BCTSLix ceram- single T-symmetry Ba(Sn0.10 Ti0.90 )O3 at 20 ◦ C.16 Meanwhile,
ics further decreases to 46–48 ◦ C as x increases from 2 to 6, the intensity of modes at 170 and 190 cm−1 gets weak and
owing to the substitution of Li+ for the ions at B-sites, similar their intensity gap narrowed obviously with raising tempera-
to the case for Li2 CO3 -doped PZN-PZT ceramics,14 but inverse ture to 20 ◦ C as marked by dotted lines in Fig. 6a-2, which is
to slightly increased TC in Li2 O-doped Ba(Ti0.95 Zr0.05 )O3 9 and attributed to the decreased specific ratio of R to T phase. As fur-
BaTiO3 .10 Apart from TC , another phase-transition maximum is ther increasing temperature, modes at 170 and 190 cm−1 vanish
observed at 8 ◦ C for BCTS and 32 ◦ C for BCTSLix (2 ≤ x ≤ 6). gradually, indicating the disappearance of R phase. At 60 ◦ C,

Fig. 5. Dielectric constant (εr ) and loss (tanδ): (a) temperature dependence at 1 kHz for BCTSLix ceramics, (b) x dependence at room temperature at 1 kHz for
BCTSLix ceramics, (c) temperature dependence at 1, 10 and 100 kHz for BCTSLi4 ceramic.
538 L. Zhao et al. / Journal of the European Ceramic Society 35 (2015) 533–540

Fig. 7. Variations of d33 , kp and Qm of BCTSLix ceramics as a function of x.

As shown in Fig. 6b-1 and 6b-2, BCTSLi4 indicates similar


Raman spectra with temperature to the case of BCTS, in which
R, O and T phases coexisted at −80 ◦ C, the O phase disappears at
about −50 ◦ C, R and T phases coexisted in −50–40 ◦ C accompa-
nying the decreased specific ratio of R to T phase with increasing
temperature and the coexisting T and C phases are observed over
60 ◦ C. So the phase-transition maximum appeared at 32 ◦ C in
εr -T curve as 2 ≤ x ≤ 6 are also corresponding to R-T transition.
Fig. 6c shows similar Raman spectra at room-temperature
for BCTSLix samples to each other because of the coexisting R
and T phases. With increasing x, both the intensities of modes
at 162 and 198 cm−1 get weak and the discrepancy of the rela-
tive intensity between modes at 190 and 250 cm−1 in Fig. 6c-2
decreases gradually. All that reveals the specific ratio of R to
T phase decreases with increasing x although the quantitative
analysis is hard to do, similarly to the case for the pristine BCTS
ceramic leaded by raising temperature. This result also matches
the variation of the FWHM of X-ray diffraction peaks around
31.5◦ , 39◦ and 45◦ as shown in Fig. 4a. Compared to the constant
FWHM of peak-31.5 and peak-39, the increased lift asymmetry
of peak-45 (Fig. 3b) with x along with a widened FWHM value
Fig. 6. Raman spectra of the investigated BCTSLix ceramics: (a, b) the evolution
with temperature of the Raman spectra for BCTS and BCTSLi4 ; (c) room-
(Fig. 4a) attributes to the decreased ratio of R to T phase since
temperature Raman spectra of BCTSLix specimens. the later has the lowest 2␪ = 45.091◦ around 45◦ .

3.3. Piezoelectric and ferroelectric properties

weak modes at 302 and 720 cm−1 , and broad modes at 250 and Fig. 7 shows piezoelectric constant d33 , planar mode
525 cm−1 which are the feature of high-temperature C phase17 eletromechanical coupling coefficient kp and mechanical qual-
are observed, showing the coexistence of T and C phases. And ity factor Qm for BCTSLix (0 ≤ x ≤ 6) ceramics as a function
T phase decreases gradually at higher temperature resulting in of x. d33 , kp and Qm of the pristine BCTS ceramic are 30 pC/N,
further weakened modes at 302 and 720 cm−1 . In εr -T curves 10% and 124, respectively. Due to the low density, the piezoelec-
(Fig. 5), a phase-transition maximum appears at 8 ◦ C for BCTS tric properties of BCTSLi1 ceramic cannot be measured. Except
which locates at R and T phases coexisted temperature range. So for BCTSLi1 , d33 of BCTSLix increases as x increases from 2
the phase-transition maximum at 8 ◦ C for BCTS corresponds to to 4 and achieves a maximum value of 457 pC/N at x = 4, and
partial R to T transition. In another words, R to T transition may decreases as x further increases to 6. kp and Qm have the same
happen at a temperature range rather than a temperature point. variation tendency with d33 , achieving the peak values 25.18%
Similar phenomenon also happens for T to C transition. and 491 at x = 4. In the modified BaTiO3 -based ceramics, both
L. Zhao et al. / Journal of the European Ceramic Society 35 (2015) 533–540 539

Fig. 8. Ferroelectric hysteresis loops and remnant polarization and coercive field of BCTSLix ceramics.

the increased grain size and density are helpful for enhancing coexisting R and T phases at the temperature range of −80 ◦ C to
piezoelectric property.1,19 Furthermore, the coexisting two or 40 ◦ C for BCTS sample, all BCTSLix ceramics have coexisting
more phases also contribute greatly high d33 .1–8 In the present R and T phases at room temperature, and the specific ratio of
study, both the grain size and density are similar to each other R to T phase decreases with increasing temperature or x. Bet-
for samples at 3 ≤ x ≤ 6, while all samples are of the coexistence ter piezoelectric properties of d33 = 457 pC/N, kp = 25.18% and
of R and T phases along with a decrease trend on the specific Qm = 491 were obtained at x = 4, which is attributed to the appro-
ratio of R to T with increasing x. So the better d33 of 457 pC/N priate specific ratio of R to T phase considering both the similar
at x = 4 is attributed to the appropriate specific ratio of R to T grain size and density for samples at 3 ≤ x ≤ 6.
phase.
Fig. 8 indicates the ferroelectric hysteresis (P–E) loops and
remnant polarization (Pr) and coercive field (Ec) as a function Acknowledgments
of x. A typical P-E loop is observed at x = 0. The ferroelec-
tric property is enhanced as x increases from 2 to 6. On the This work was supported by Specialized Research Fund
other hand, P–E loops at 2 ≤ x ≤ 6 show “pinched” shapes as for the Doctoral Program of Higher Education (grant no.
shown in the inset of Fig. 8a, which is similar to the case of 20130006110006) and National Natural Science Foundation of
Li2 O-doped BaTiO3 ceramic.10 This may be attributed to the China (grant no. 51472026 and 51332002).
substitution of Li+ for Ti4+ or Sn4+ at B-sites, the acceptor doping
Li+ can form defect dipoles with oxygen vacancy as mentioned
above, which disrupt the ferroelectric order and decrease the References
polarization states for the BCTS ceramics. With increasing x,
1. Wang P, Li YX, Lu YQ. Enhanced piezoelectric properties of
Pr increased and Ec decreased as shown in the inset Fig. 8b. (Ba0.85 Ca0.15 )(Ti0.9 Zr0.1 )O3 lead-free ceramics by optimizing calcination
and sintering temperature. J Eur Ceram Soc 2011;31(11):2005–12.
4. Conclusions 2. Liu WF, Ren XB. Large piezoelectric effect in Pb-free ceramics. Phys Rev
Lett 2009;103(25):257602.
3. Li W, Xu ZJ, Chu RQ, Fu P, Zang GZ. Large piezoelectric coeffi-
The adding effect of Li2 O content (x) on sintering and cient in (Ba1−x Cax )(Ti0.96 Sn0.04 )O3 lead-free ceramics. J Am Ceram Soc
electrical properties of (Ba0.95 Ca0.05 )(Ti0.90 Sn0.10 )O3 lead-free 2011;94(12):4131–3.
piezoceramics was investigated. Both the microstructure and 4. Li W, Xu ZJ, Chu RQ, Fu P, Zang GZ. Enhanced ferroelectric proper-
density of BCTS ceramic were improved by adding Li2 O, ties in (Ba1−x Cax )(Ti0.94 Sn0.06 )O3 lead-free ceramics. J Eur Ceram Soc
2012;32(3):517–20.
emerging a reduced dense sintering temperature by more 5. Chen ML, Xu ZJ, Chu RQ, Liu Y, Shao L, Li W, et al. Poly-
than 150 ◦ C for BCTS ceramic as x > 2 mol%. Two-step sin- morphic phase transition and enhanced piezoelectric properties
tering mechanism leaded by reactive liquid phase and V•• O in (Ba0.9 Ca0.1 )(Ti1−x Snx )O3 lead-free ceramics. Mater Lett
was observed in Li2 O-added BCTS ceramics. Similar to the 2013;97(15):86–9.
540 L. Zhao et al. / Journal of the European Ceramic Society 35 (2015) 533–540

6. Zhu LF, Zhang BP, Zhao XK, Zhao L, Zhou PF, Li JF. Enhanced piezoelectric 13. Yang WG, Zhang BP, Ma N, Zhao L. High piezoelectric properties of
properties of (Ba1−x Cax )(Ti0.92 Sn0.08 )O3 Lead-free ceramics. J Am Ceram BaTiO3 –xLiF ceramics sintered at low temperatures. J Eur Ceram Soc
Soc 2013;96(1):241–5. 2012;32(4):899–904.
7. Xue DZ, Zhou YM, Bao HX, Gao JH, Zhou C, Ren XB. Large piezoelec- 14. Hou YD, Chang LM, Zhu MK, Song XM, Yan H. Effect of Li2 CO3 addition
tric effect in Pb-free Ba(Ti,Sn)O3 –x(Ba Ca)TiO3 ceramics. Appl Phys Lett on the dielectric and piezoelectric responses in the low-temperature sintered
2011;99(12):122901. 0.5PZN–0.5PZT systems. J Appl Phys 2007;102(8):084507.
8. Yao YG, Zhou C, Lv DC, Wang D, Wu HJ, Yang YD, et al. Large piezo- 15. Miao S, Pokorny J, Pasha UM, Thakur OP, Sinclair DC, Reaney IM.
electricity and dielectric permittivity in BaTiO3 –xBaSnO3 system: the role Polar order and diffuse scatter in Ba(Ti1−x Zrx )O3 ceramics. J Appl Phys
of phase coexisting. EPL 2012;98(2):27008. 2009;106(11):114111.
9. Nawal B, Pisan S, Chanchana T, Supasarote M. Physical and electromechan- 16. Deluca M, Stoleriu L, Curecheriu LP, Horchidan N, Ianculescu AC, Galassi
ical properties of barium zirconium titanate synthesized at low-sintering C, et al. High-field dielectric properties and Raman spectroscopic investi-
temperature. Mater Lett 2010;64(3):305–8. gation of the ferroelectric-to-relaxor crossover in BaSnx Ti1−x O3 ceramics.
10. Ma N, Zhang BP, Yang WG. Low-temperature sintering of Li2 O- J Appl Phys 2012;111(8):084102.
doped BaTiO3 lead-free piezoelectric ceramics. J Electroceram 17. Farhi R, ElMarssi M, Simon A, Ravez J. A Raman and dielectric study of
2012;28(4):275–80. ferroelectric Ba(Ti1−x Zrx )O3 ceramics. Eur Phys J B 1999;9(4):599–604.
11. Valant M, Suvorov D, Pullar RC, Sarma K, Alford NM. A mech- 18. Dobal PS, Katiyar RS. Studies on ferroelectric perovskites and Bi-
anism for low-temperature sintering. J Eur Ceram Soc 2006;26(13): layered compounds using micro-Raman spectroscopy. J Raman Spectrosc
2777–83. 2002;33(6):405–23.
12. Kimura T, Dong Q, Yin S, Hashimoto T, Sasaki A, Sato T. Synthesis and 19. Wu JG, Xiao DQ, Wu WJ, Zhu JG, Wang J. Effect of dwell time during
piezoelectric properties of Li-doped BaTiO3 by a solvothermal approach. J sintering on piezoelectric properties of (Ba0.85 Ca0.15 )(Ti0.90 Zr0.10 )O3 lead-
Eur Ceram Soc 2013;33(5):1009–15. free ceramics. J Alloys Compd 2011;509(41):L359–61.

Das könnte Ihnen auch gefallen