Sie sind auf Seite 1von 10

Materials Research Bulletin 66 (2015) 16–25

Contents lists available at ScienceDirect

Materials Research Bulletin


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

Ba(Cu0.5W0.5)O3-induced sinterability, electrical and mechanical


properties of (Ba0.85Ca0.15Ti0.90Zr0.10)O3 ceramics sintered at low
temperature
Xiaolian Chao * , ZhongMing Wang, Ye Tian, Yanzhao Zhou, Zupei Yang *
Key Laboratory for Macromolecular Science of Shaanxi Province, School of Materials Science and Engineering, Shaanxi Normal University, Xi’an 710062,
Shaanxi, PR China

A R T I C L E I N F O A B S T R A C T

Article history: A low sintering temperature is achieved for (Ba0.85Ca0.15)(Ti0.90Zr0.10)O3 (BCZT) piezoelectric ceramics
Received 26 April 2014 using Ba(Cu0.5W0.5)O3 (BCW) as a sintering aid. The effects of the BCW content on the density, phase
Received in revised form 12 October 2014 structure, electrical and mechanical properties are investigated. The addition of BCW content
Accepted 6 February 2015
significantly improves the sinterability of the BCZT ceramics, which results in a reduction of the
Available online 9 February 2015
sintering temperature from 1450  C to 1220  C and an improvement in both the piezoelectric and
mechanical properties. The BCZT ceramics with x = 0.8 wt.% BCW sintered at 1220  C exhibits outstanding
Keywords:
piezoelectric, dielectric and mechanical properties: d33 = 541 pC/N, kp = 49%, Qm = 107, er = 5405, tan
A. Ceramics
C. X-ray diffraction
d = 0.014 and Tc = 78  C. The ceramics also exhibit higher hardness (78 MPa) and higher fracture toughness
D. Electrical properties (4.7 MPa m1/2). These results indicate that the BCZT–BCW lead-free ceramics sintered at low
D. Microstructure temperatures are promising for high temperature piezoelectric applications.
D. Piezoelectricity ã 2015 Elsevier Ltd. All rights reserved.

1. Introduction ceramics could be reduced if a liquid-phase is formed during


the sintering process. It is well known that the addition of a
Lead-free (Ba0.85Ca0.15)(Ti0.90Zr0.10)O3 (BCZT)-based piezoelec- sintering aid is an effective approach to reduce the sintering
tric ceramics have attracted considerable interest because of their temperature and improve the density of piezoelectric materials
excellent piezoelectric properties and environmental friendliness [12,13]. In the related literature, a sintering temperature of 1350–
[1,2]. The piezoelectric coefficient d33 (620 pC/N) of this system 1400  C has been achieved for BCZT-based ceramics while
sintered at above 1450  C has been reported, which is comparable maintaining good properties with the addition of CuO, however,
to that (500–600 pC/N) of Pb(Zr, Ti)O3 (PZT) based piezoelectric the sintering temperature remains relatively high [14,15]. Perov-
ceramics [3–5]. Some attempts have been made to improve the d33 skite-type Ba(Cu0.5W0.5)O3 (BCW) exhibits a strong ferroelectricity,
value of lead-free piezoelectric ceramics and the origin of this large a high Curie temperature and a low melting temperature (730  C).
piezoelectricity, such as by ion substitution, new preparation The addition of BCW has also been demonstrated to be very
techniques, the introduction of additives and the construction of effective in improving the piezoelectric properties of PZT ceramics
the coexistence of two or more phases [6–10]. Nevertheless, low- [16]. However, no systematic investigations have been performed
temperature sintering and the electrical properties of BCZT-based on the effects of BCW on the low-temperature sintering and
piezoelectric ceramics are seldom reported. electrical properties of BCZT ceramics to date.
Although this type of the ceramics displays good properties, During multilayer piezoelectric device operation, the piezo-
further investigation is needed to lower the sintering temperature electric device induced damage and piezoelectric device quality
of the ceramics to reduce compositional variation, energy are related to the mechanical properties of the host materials [17].
consumption and fabrication costs. In addition, low-temperature To date, only a few groups have investigated the mechanical
sintering can have several advantages, such as the reduction of properties of these ceramics. The mechanical properties of
energy consumption [11]. The sintering temperature of the piezoelectric materials are very important for equipment design
and practical use [18,19]. However, few studies have reported that
strongly enhanced mechanical properties can be obtained from
* Corresponding authors. Tel.: +86 29 81530718; fax: +86 29 8153 0702. BCZT ceramics with the addition of BCW content. In this study, our
E-mail addresses: chaoxl@snnu.edu.cn (X. Chao), yangzp@snnu.edu.cn (Z. Yang). objective is to reduce the sintering temperature of BCZT ceramics

http://dx.doi.org/10.1016/j.materresbull.2015.02.022
0025-5408/ ã 2015 Elsevier Ltd. All rights reserved.
X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25 17

and while maintaining a high-piezoelectric activity. BCW-doped both sides of the sintered samples were polished. The tempera-
BCZT ceramics are fabricated to systematically investigate the ture-dependent dielectric constant was obtained with an LCR
effect of the BCW content on the sintering temperature, phase meter (Agilent, E4980A) connected to a computer-controlled
structure, density, electrical and mechanical properties (hardness temperature chamber by measuring the capacitance at 1, 10, 100,
and fracture toughness). The underlying physical mechanisms are 100 kHz. For the temperature-dependent polarization–electric
addressed in this study. field (P–E) hysteresis and strain measurements, the top electrode
was connected to a high voltage amplifier (Model 610E, Trek, USA)
2. Experiment for the electrical loading. Before the piezoelectric properties
measurements, the samples were poled for 30 min at room
2.1. Sample preparation temperature with an electric field of 4 kV/mm. The piezoelectric
coefficient d33 was measured using a quasistatic piezoelectric
(Ba0.85Ca0.15)(Zr0.10Ti0.90)O3–xBa(Cu0.5W0.5)O3 ceramics (x = 0.0, meter (ZJ-3d, Institute of Acoustics Academic Sinica, Beijing,
0.4, 0.8, 1.2, 1.6, 2.0 and 3.0 wt.%) were synthesized using a China). The electromechanical coupling coefficient kp was deter-
conventional solid-state sintering process. BaCO3 (99%), CaCO3 mined using the resonance and anti-resonance technique using an
(99%), ZrO2 (99%), TiO2 (98%), CuO (99.5%), and WO3 (99%) powders impedance analyzer (HP/Agilent Model 4294A).
were used as starting raw materials. Ba(Cu0.5W0.5)O3 does not need Hardness measurements were performed at room temperature
to be synthesized. The powders were weighed according to the on an HXD-1000 tester with a diamond indenter under a 300 N
stoichiometric composition. The mixed powders were dried and load for 15 s. Fracture toughness tests were conducted at room
then calcined at 1100  C for 6 h. The calcined powders were mixed temperature using a PC-1036PC material tester at a testing rate of
with 5 wt.% polyvinyl alcohol (PVA) solution and then uniaxially 0.5 mm/min.
pressed into pellets with a diameter of 1.5 cm under 100 MPa. After
burning out PVA at 500  C, the samples were sintered at 1180, 1220, 3. Results and discussion
1260, 1300 and 1340  C for 4 h in air. Sliver paste was fired on both
faces of the samples at 850  C for 30 min to form electrodes. 3.1. Phase formation

2.2. Characterization Fig. 1 presents the XRD patterns of the BCZT ceramics with
various BCW contents sintered at 1220  C for 4 h. The positions of
The sintered ceramics were examined by X-ray powder the diffraction peaks can be used to evaluate the structural order at
diffraction analysis using a Cu Ka radiation (DMX-2550/PC, long range or the periodicity of the material. The BCZT phase is
Rigaku, Japan) to determine the crystalline phase at room confirmed by comparing the XRD patterns with the respective
temperature. For the measurement of the electrical properties, JCPDS card No. 50-0346. All the diffraction peaks can be assigned to
[(Fig._1)TD$IG]

Fig. 1. XRD patterns of the BCZT ceramics with various BCW contents sintered at 1220  C for 4 h.
18 X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25
[(Fig._2)TD$IG]

Fig. 2. The density and SEM images of the BCZT ceramics sintered at various temperatures as a function of the BCW content.

the perovskite structure in the ceramics with x = 0.0–2.0 wt.%. In SEM images of the 0.80 wt.% BCW-modified specimen sintered at
addition, when x  3.0 wt.% BCW, the TiO2 phase begins to appear, 1220  C reveals that the bodies are fully densified with fairly
confirmed by JCPDS card No. 21-1276 [20]. To clearly analyze the uniform microstructures, as illustrated in Fig. 2(b).
phase structure of all the ceramics, the enlarged XRD patterns with Fig. 3 displays the water-solid contact angle of the surface of
2u ranging from 44 to 46 and from 64 to 67 are presented in untreated samples with various BCW contents as a function of
Fig. 1(b) and (c), respectively. The compositions with x  0.8 wt.% time. The methods for the measured densification of the ceramics
exhibit a rhombohedral (R) structure [21]. The (2 0 0) and (0 0 2) and the liquid-solid contact angle in water have been described.
diffraction peaks of the BCZT ceramics gradually emerge with The liquid–solid contact angle is used as an index for the
increasing BCW content (x  1.2 wt.%), where both (2 2 0) and wettability or hydrophobicity: density of the ceramics. The water
(2 0 2) diffraction peaks almost appear, confirming the involve- enters the ceramics readily, which indicates that water entry is not
ment of a phase transition. Therefore, the coexistence of tetragonal restrained by the lack of pores of the ceramics. A drop of water was
and rhombohedral phases occurs for the BCZT ceramics with rapidly applied and remained at the surface of the ceramics. The
x = 0.8 wt.% at room temperature, demonstrating that the ceramics contact angle was marked every 1 s. As illustrated in Fig. 3, by
lie at the morphotropic phase boundary (MPB) [22,23]. Because the increasing the time, the contact angle decreases gradually from
peaks shift to higher angles, it is assumed that (Ti, Zr)4+ is 86.8 to 0 , which spend 330 s for the ceramics with x = 0.00 wt.%,
substituted by Cu2+ and W6+, suggesting a change of the unit cell indicating that the drop of water has seeped into the ceramics
volume. This change may be due to the formation of complete solid through the pores. In other words, the intake time of a drop of
solutions between the BCW content and the BCZT ceramics, which water is 330 s for the ceramics with x = 0.00 wt.%. The intake time is
leads to a distortion of the crystal lattice. The addition of the BCW 708 s at x = 0.8 wt.%, and then decreases to 672 s at x = 1.6 wt.%. The
solid solution promotes the substitution of the Ba2+/Ca2+ and/or Zr4 reduced porosity of the ceramics delays the intake time, indicating
+
/Ti4+ B site of the perovskite structure more effectively than the that the ceramics with x = 0.8 wt.% is dense. In addition, the inset of
addition of raw metal oxides, and then increases structure defects Fig. 3 shows that the contact angle left (CA left) and contact angle
and decreases the barrier among domains. The formation of the right (CA right) for the compositions (x = 0.0 wt.% 0.8 wt.% and
BCW solid solution is beneficial to the diffusion of ions and 1.6 wt.%) are measured. With increasing BCW, CA left and CA right
reduction of the BCZT sintering temperature. increase gradually, reaching maximum values of CA left (92.2 ) and
CA right (92.1 ) at x = 0.8 wt.%, and then decreasing. The variation
3.2. Densification behavior contact angle in Fig. 3 is consistent with the relative density in
Fig. 2.
Fig. 2 presents the density and SEM images of the BCZT ceramics
sintered at different temperatures as a function of the BCW
content. Fig. 2(a) clearly demonstrates that higher densities of all
[(Fig._3)TD$IG]
the ceramics are obtained in the sintering temperature range from
1220  C to 1260  C, demonstrating that the addition of BCW
sintering aids to BCZT ceramics reduces the sintering temperature
by 200  C. The maximum value of 5.58 g/cm3 is obtained for the
0.80 wt.% BCW-modified BCZT ceramics sintered at 1220  C. By
adding BaCO3, CuO and WO3 with low melting points, the liquid
formation promotes the sintering of BCZT ceramics and results in
an increase in the density of the ceramics. However, with
increasing the BCW content, the density decreases. This finding
may be due to a larger amount of this liquid phase accumulating at
the grain boundaries, thus resulting in voids and inhibiting
densification. It is apparent that adding a small amount of BCW
causes the generation of oxygen vacancies is beneficial to the mass
transport in the ceramics during sintering, which is responsible for
the enhanced grain growth behavior, sinterability and bulk density
of the suitable BCW-doped specimens sintered at low temperature.
Therefore, the addition of BCW plays an important role in the
Fig. 3. The water-solid contact angle of the surface of untreated samples with
improvement of the density of BCZT-based ceramics. The surface different BCW contents as a function of time.
X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25 19

3.3. Temperature dependence of the dielectric behavior of the ceramics cubic phase transition (TC) is observed in these ceramics, while the
coexistence of two phases cannot be clearly observed in the
Fig. 4 plots the temperature dependence of the dielectric temperature range involved except for the ceramics with x = 0.80,
behavior for the BCZT ceramics with different BCW contents 1.2 and 1.6 wt.%. The coexistence of two phases near room
measured at 0.1, 1, 10 and 100 kHz, respectively. The tetragonal- temperature is usually beneficial to the improvement in the

[(Fig._4)TD$IG]

Fig. 4. Temperature dependence of the dielectric behavior of the BCZT ceramics with varuous BCW contents measured at 0.1, 1, 10 and 100 kHz, respectively: (a) x = 0.0 wt.%,
(b) x = 0.4 wt.%, (c) x = 0.8 wt.%, (d) x = 1.2 wt.%, (e) x = 0.16 wt.%, (f) x = 2.0 wt.%, and (g) x = 3.0 wt.%.
20 X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25
[(Fig._5)TD$IG]
piezoelectric properties of the ceramics [22]. In addition, all the
ceramics possess a low dielectric loss in the present work, and the
peaks of the dielectric constant versus temperature become much
broader with increasing the BCW content.
The temperature dependence of the dielectric constant (a) and
em, tan d, Tc (b) for the BCW-doped BCZT ceramics is shown in Fig. 5.
As observed in Fig. 4(a), two obvious phase transitions corre-
sponding to the rhombohedral-tetragonal transition (Trt) at 38  C
and tetragonal-cubic transition (Tc) at 80  C are observed for the
samples with x = 0.0–1.6 wt.%. Trt and Tc shift to below room
temperature with increasing the BCW content, indicating its
tetragonal structure at room temperature, which is consistent with
the XRD results in Fig. 1. Usually, the Trt located below room
temperature is beneficial to the temperature stability of the BCZT-
based ceramics [3,5]. The results are in agreement with the results
in the literature [23]. It has been reported that a decrease in the
Curie temperature could be attributed to the valence mismatch
[16]. It is expected from the tolerance factors that small ions (r(R3
+
) < 0.87 Å) will occupy the B site, large ions (r(R3+) > 0.94 Å) will
occupy the A site, and intermediate ions will occupy both sites with
different portioning for each ion. Considering Ba2+ (r = 1.34 Å), Cu2+
(r = 0.72 Å) and W6+ (r = 0.62 Å), it is concluded that Cu2+ and W6+
occupy the B site, resulting in deformation in the ABO3 lattice and a
change of oxygen vacancies. However, when BCW content becomes
high, for example at 0.8 wt.%, the deformation cannot be ignored
and the structure of the ceramics becomes un-stable. Therefore,
with increasing the BCW content, the structure changes from the
tetragonal phase to the cubic phase at lower temperature, thus
Fig. 5. Temperature dependence of dielectric constant (a) and em, tan d,Tc (b) for the resulting in a decrease of Tc in this work. The Tc values of BCZT–
BCW-doped BCZT ceramics. xBCW ceramics with x = 0.0, 0.4, 0.8, 1.2, 1.6, 2.0 and 3.0 wt.% are 83,
81, 78, 70, 67, 60 and 44  C, respectively. Fig. 5(b) displays the
maximum dielectric constant (em), maximum dielectric constant

[(Fig._6)TD$IG]

Fig. 6. Plots of the inverse dielectric constant as a function of temperature at 10 kHz for the BCTZ–xBCW (x = 0.0–3.0 wt.%) ceramics.
X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25 21

Table 1
The Curie–Weiss temperature (To), the temperature above which the dielectric constant follows the Curie–Weiss law (Tcw), deviation (DTcm), and the Curie–Weiss constant (C)
for samples at 1 kHz and 100 kHz.

Samples (wt.%)

0.0 0.4 0.8 1.2 1.6 2.0 3.0

1 kHz 100 kHz 1 kHz 100 kHz 1 kHz 100 kHz 1 kHz 100 kHz 1 kHz 100 kHz 1 kHz 100 kHz 1 kHz 100 kHz
To ( C) 54 65 51 48 54 67 54 71 40 73 42 68 3 13
Tcw ( C) 95 98 90 92 98 92 87 90 83 79 71 81 60 69
DTcm ( C) 10 16 8 11 20 14 16 19 16 13 11 21 13 23
C(105) ( C) 4.52 4.62 4.98 4.59 9.59 3.53 8.83 2.61 5.36 2.29 3.76 1.89 4.96 3.57

(tan d) and Curie temperature (Tc) of specimens as a function of the increase of the relaxor behavior. The variation tendencies of the To,
BCW content. The maximum dielectric constant (13456 at Tcw, C and DTcm data of the specimen are similar between 1 kHz and
x = 0.00 wt.%) first decreases and then starts to increase at 100 kHz.
x = 0.4% in Fig. 4(b). em of the ceramics increases to 18197 at The deviation from the Curie–Weiss law can be defined by DTcm
x = 1.6 wt.% and then decreases with increasing BCW contents from as follows:
1.6 to 3.0 wt.%. Tan d exhibits a transverse “S” trend, and reaches a
minimum value of 0.0129 at x = 1.2 wt.%, which is consistent with DT cm ¼ T cw  T m ; (2)
the temperature dependence of the dielectric constant (er) in where Tcw denotes the temperature from which the dielectric
Fig. 5(a). As mentioned before, when (Cu0.5W0.5)4+ substitutes for constant starts to follow the Curie–Weiss law, and Tm represents
Zr4+ and Ti4+ at B sites, the lattice will be distorted less as the the temperature at which the dielectric constant reaches the
fraction of Ba2+ in the larger A sites decreases, making the phase maximum. As is shown in Fig. 6 and Table 1, Tcw gradually
transition at lower temperatures possible. Therefore, the formation decreases from 95  C to 60  C with increasing x from 0.00 to 3.0 wt.
of the coexistence of two phases near room temperature is %. DTcm is found to increase from 10  C to 20  C for x = 0.0, 0.40 and
beneficial to enrich the properties of BCZT-based ceramics. 0.80 wt.%, which is associated with the increased dielectric
It is of great interest to note that the dielectric peaks at Tc for all diffuseness and enhanced relaxor behavior when BCW is mixed
the ceramics are relatively broad, indicating the characteristic of into BCZT ceramics. After reaching the maximum of 20  C when
the diffuse phase transition from the tetragonal phase to the cubic x = 0.80 wt.%, DTcm gradually decreases to 13  C for the composi-
phase. It is well known that the dielectric constant of a normal tions with BCW ranging from 0.80 wt.% to 3.0 wt.%, respectively.
ferroelectric should follow the Curie–Weiss law when the
temperature exceeds the Curie temperature: [(Fig._8)TD$IG]
C
e¼ ðT > T m Þ; (1)
T  To
where C is the Curie–Weiss constant and To is the Curie–Weiss
temperature. The plot of the inverse dielectric constant versus
temperature at 1 kHz is presented in Fig. 6. It is observed that the
dielectric constant of BCZT ceramics slightly deviates from the
Curie–Weiss law with increasing BCW contents. To, Tcw, C and DTcm
obtained by the Curie–Weiss law for all specimens at 1 kHz and
100 kHz are listed in Table 1. To finally decreases from 54  C to
28  C with increasing x from 0.00 to 3.0 wt.%. By increasing x to
0.08 wt.%, C rapidly increases to 9.59  105  C, indicating an obvious
[(Fig._7)TD$IG]

Fig. 7. Ln[(em  e)/e] as a function of ln(T  Tm) for BCZT ceramics with different Fig. 8. (a) P–E loops, (b) Pr and Ec values as a function of BCW content of the BCZT
BCW contents at 1 kHz. ceramics sintered at 1220  C and measured at room temperature.
22 X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25

This result indicates that the diffuse phase transition behavior has the cationic disorder induced by both A-site and B-site sub-
been enhanced firstly and then reduced with the increase of the stitutions, such as (Cu0.5W0.5)4+ ions entering the six fold
BCW content. coordinated B-site (Zr, Ti)4+, and Ba2+ ions entering the A-site of
The dielectric behavior of complex ferroelectrics with a diffuse the perovskite structure to substitute for (Ba, Ca)2+. Heterovalent
phase transition can be explained by the modified Curie–Weiss law substitution induces quenched random electric fields because of
[24], the local charge imbalance and the local elastic fields in BCZT-
derived systems. The fields hinder long-range ordering and
em ðT  T m Þg
¼1þ ; produce polar nano-regions. These polar nano-regions are isolated
e 2
ð2D Þ and frustrated in BCZT ceramics with different BCW contents, and
where em is the maximum value of the dielectric constant at the are then embedded in the disordered matrix, which has also been
phase transition temperature Tm, and g (1 < g < 2) is the degree of demonstrated by Zheng and co workers [25–27]. The results are in
the diffuseness, which is a constant used to express the diffuseness accordance with the previous XRD analysis.
degree of the phase transition. When g = 1, the materials with this
type of phase transition belong to normal ferroelectrics; when 3.4. Electrical properties of the ceramics
1 < g < 2, the materials are relaxor ferroelectrics; when g = 2, the
materials are ideal relaxor ferroelectrics. To eliminate space- Fig. 8(a) illustrates the typical polarization levels versus applied
charge polarization contributions at high temperatures and lower electrical field (P–E) hysteresis loops of BCW-modified BCZT
measurement frequency, herein, the 1 kHz data are selected. It is ceramics sintered at 1220  C and measured at room temperature.
found that the dielectric permittivity of the BCZT ceramics with Fig. 8(a) demonstrates that the P–E hysteresis loops become much
different BCW contents obviously deviate from the Curie–Weiss slimmer when BCW additives are mixed into the BCZT ceramics.
law. Ln[(em  e)/e] as a function of Ln(T  Tm) for BCZT ceramics Fig. 8(b) depicts the remnant polarization (Pr) and coercive field
with different BCW contents at 1 kHz is plotted in Fig. 7. It can be (Ec) of the ceramics as a function of the BCW content. According to
observed that a linear relationship is obtained by linear fitting of Fig. 8(b), with increasing the BWC content, Pr first decreases, then
the experimental data. The diffuseness exponent g of all the increases and reaches the maximum value of 10.59 mC/cm2 when
ceramics is between 1 and 2 for the ceramics with x < 2.0 wt.%, x = 0.8 wt.% as increasing the BCW content, indicating that the
which indicates that these compositions show intermediate ceramics with x = 0.8 wt.% are the easiest to pole. Moreover, further
relaxor-like behavior between normal and ideal relaxor ferro-
electrics and the phase transition of all ceramics exhibits a diffuse
[(Fig._10)TD$IG]
character. When x = 3.0 wt.%, the g value is 2.001, which is higher
than 2, indicating that these ceramics are not ferroelectrics, which
is consistent with the results of Fig. 6. The relaxor behavior in the
BCZT ceramics with different BCW contents should be attributed to
[(Fig._9)TD$IG]

Fig. 10. P–E loops and current of 0.80 wt.% doped-BCZT ceramics sintered at 1220  C
Fig. 9. (a) d33, kp and (b) Qm, er, tan d of BCZT ceramics as a function of BCW content. as a function of temperature.
X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25 23

increasing the BCW content beyond 0.8 wt.% causes sudden drops ceramics, as observed in Fig. 10(b). An applied electric field can lead
in the Pr values, which can be attributed to BCW segregation at to a change in the surface charges of the sample. One contribution
boundaries which inhibits the movement of domain walls. In is from the ferroelectric domain switching, which can be
addition, it can be observed that Ec of the ceramics markedly represented by the sharp increase in the displacement currents.
decreases from 0.33 kV/cm to 0.13 kV/cm as increasing the BCW Ec can be determined from the current peak position (maximum
content. The ceramics with x = 0.80 wt.% possess a large Pr current), which decreases with the measuring temperature, as
(10.59 mC/cm2) and a relatively large Ec (0.23 kV/cm). Because shown in the inset of Fig. 10(b).
BCW doping is expected to increase the defect concentration either Fig. 11 illustrates the bipolar electric-field-induced strain of the
at the A- or B-site of the ABO3 lattice, a direct interaction of these 0.8 wt.% doped-BCZT ceramics as a function of the measuring
defects with the domain walls is expect [28,29]. Increasing the temperature. As observed in Fig. 11(a), the ceramics all exhibit the
BCW content affects the polarizability and reduces the Ec values typical butterfly-shaped curve of ferroelectrics. The hysteresis,
due to the doping effect and the decrement of the lattice which is caused by the domain switching and phase transition
parameters. A lower Ec value is advantageous in the fabrication induced by the electric field is very large. From Fig. 11(a), the
of high-density and low-operation-voltage sensors and actuators bipolar maximum strain of the ceramics is 0.82%, which is obtained
[30]. As observed in Fig. 8(b), the piezoelectric constant (d33), at T = 30  C, and then significantly decreases from 0.82% to 0.71%
electromechanical coupling factor (kp) and Pr change with the BWC with the increasing of the temperature. Generally speaking, a large
content increases, reaching the maximum values at the same time, electric-field-induced strain means a large piezoelectric response.
and then droping simultaneously. Therefore, we strongly confirm The results for the electric-field-induced strain are consistent with
that the enhanced d33 and kp are related to their ferroelectric the d33 values measured using the quasistatic method. Moreover,
properties. Moreover, the dense microstructure is also partly the largest d33 (544 pC/N) value is also obtained at T = 30  C. With
responsible for the enhanced piezoelectric properties observed in increasing the temperature from 40  C to 90  C, strain values of the
this work [14]. ceramics initially increase, and then begin to decrease at T = 50  C.
Fig. 9(a) gives the piezoelectric properties (d33 and kp) of the The peak values of the stain (0.79%) are obtained at 50  C for the
BCZT ceramics with different BCW contents. By increasing the BCW 0.80 wt.% doped composition near the rhombohedral side of the
content, the d33 and kp values of the ceramics initially increase, and MPB boundary. Because the structural activities for the compo-
then begin to decrease at x = 3.0 wt.%. The peak values of sitions in the biphase coexistent regions are high, the phase
d33 = 544 pC/N and kp = 0.49 can be obtained from the BCZT- transition from the rhombohedral ferroelectric phase to the
0.8 wt.% BCW samples. In addition, Fig. 9(b) gives the mechanical tetragonal phase, accompanied by the mobility and polarization
quality factor (Qm), dielectric constant (er) and dielectric constant of ions, can occur easily. This structural transition is very beneficial
(tan d) of BCZT ceramics with different BCW contents. er of the in enhancing the strain properties. It can also be seen in Fig. 11(b)
ceramics increases from 4248 to 8205 with increasing BCW that d33 decreases from 544 pC/N to 67 pC/N for the ceramics with
content. As increasing the BCW content, Qm and tan d value of the increasing the measuring temperature from 30  C to 90  C. The
ceramics initially decrease, and then begin to increase. Minimum lattice distortion and domain switching induced by electric field
values of Qm = 59 at x = 1.6 wt.% and tan d = 0.014 at x = 0.8 wt.% are contribute to the measured strain. Meanwhile, only lattice
obtained. The above discussions support the analysis of the phase [(Fig._1)TD$IG]
structure in Figs. 1, 4 and 5, confirming that Tr–t of the ceramics in
the MPB regions with x = 0.80 wt.% is closer to room temperature,
resulting in an enhancement of d33 and kp in the ceramics sintered
at 1220  C. Moreover, during sintering, the presence of the BCW
liquid phase enhances the density and reduces the sintering
temperature, which leads to a decrease of the energy loss and
improvement of the electrical properties. And the dense micro-
structure also contributes parameter to the enhancement in the
electrical properties of the BCZT–xBCW ceramics in Fig. 2. The
similar results were also obtained in the literature [31].

3.5. Temperature stabilities of the electrical properties of the ceramics

To obtain further details on the ferroelectric properties and


depolarization mechanism of the ceramics, P–E hysteresis loops
and the current of 0.80 wt.% doped-BCZT ceramics at different
temperatures were measured under an electric field of 3.2 kV/mm,
as shown in Fig. 10. All the ceramics exhibit a typical ferroelectric
P–E loop at room temperature, which is much slimmer between
30  C and 90  C. It can be seen from Fig. 10(a) that the P–E hysteresis
loops are more predominant at lower temperatures, and higher
temperatures lead to constriction in the loops. According to the
inset of Fig. 10(a), as the measured temperature increases from
30  C to 90  C, the ceramics present remnant polarization (Pr)
values of 9.81, 8.80, 8.09, 7.81, 3.51 1.45 and 0.92 mC/cm2 as well as
Ec values of 2.68, 2.19, 2.01, 1.81, 1.03, 0.85 and 0.63 kV/cm,
respectively. Ec of the ceramics decreases with the measured
temperature from 30  C to 90  C; thus, poling is facile due to the
low Ec. As the applied electric field is ramped to approximately
0.65–2.4 kV/cm, the currents exhibit a sharp increase, demon- Fig. 11. The strain and piezoelectric properties of 0.80 wt.% doped-BCZT ceramics
strating the occurrence of polarization switching in the BCZT sintered at 1220  C as a function of temperature.
24 X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25

distortion contributes to the d33 value measured by the quasistatic grained and relatively uniform microstructure is obtained at
method. Ec of the ceramics in Figs. 8 and 10 decreases with x = 0.80 wt.%, which is very beneficial for practical applications as
increasing the measuring temperature; therefore, the domain it is known that a fine-grained and homogenous microstructure can
switches more easily. The joint action of lattice distortion and lead to an increase in the density and mechanical strength for
domain switching results in the largest strain being obtained in the piezoelectric ceramics, as observed in Fig. 2(b).
0.80 wt.% doped-BCZT ceramics induced by the measuring
temperature. 4. Conclusions
It has been well known that the morphotropic phase boundary
(MPB) plays an important role in the improvement of piezoelectric BCZT lead-free piezoelectric ceramics with various BCW contents
properties in PZT based piezoelectric ceramics [32]. The piezo- were prepared using the conventional solid state sintering method.
electric properties of the ceramics near the MPB region reach The phase structure, microstructure, dielectric properties and
maximum values. As we know, the MPB of a specific material ferroelectric properties of the BCZT ceramics as a function of the
depends only on its composition, while the polymorphic phase BCW content were investigated in this study. The results demon-
boundary (PPB) is dependent not only on its composition but also strated that the addition of BCW content not only decreased the
on the temperature. In this work, the R–T phase transition makes sintering temperature but also retained the optimum electrical
the dielectric property peak at the corresponding temperature, as properties. The sintering temperature of BCZTcan be decreased from
shown in Figs. 4 and 5. This finding is in good agreement with the 1450 to 1220  C with the addition of BCW. The coexistence of
PZT-based single crystals with rhombohedral-tetragonal MPB tetragonal and rhombohedral phases (MPB) was identified for the
reported by Zhang and co workers [33,34]. Because the R–T phase BCZT ceramics at x = 0.8–1.6 wt.%, as confirmed by the XRD patterns
boundary of PMN-PT, PMN-PZT, and PIN-PMN-PT single crystals is and the temperature dependence of the dielectric behavior. The
well known to be an MPB, rather than a PPB, similar to KNN-based degree of relaxor behavior increased with increasing the BCW
piezoelectric ceramics, the BCZT ceramics studied in this work may content. Due to the existence of MPB and the relaxor behavior, the
also exhibit a morphotropic character. However, it has been optimum electrical properties of d33 = 541 pC/N, kp = 49%, Qm = 107,
reported that the piezoelectric properties of the BCZT-based er = 5405, tan d = 0.014 and Tc = 78  C were obtained in 0.8 wt.% BCW-
ceramics exhibit a strong temperature dependence [35]. Therefore, doped BCZT ceramics. The introduction of the diffusion mechanism
compared with pure BCZT ceramics, the ceramics with x = 0.8 wt.% into BCZT ceramics improved the temperature stability of the
BCW shift the rhombohedral-tetragonal phase transition to near piezoelectric properties, which was useful in other material
room temperature (40–50  C), and may possess more possible systems. In addition, the hardness (78 MPa) and fracture toughness
polarization states, which results in the improvement of the (4.7 MPa m1/2) of the ceramics were obtained because of the
piezoelectric and dielectric properties at room temperature. improved density by BCW additives. These results indicated that
However, inappropriate distortion is also introduced by replacing appropriate BCW additives can significantly improve the electrical
W6+ with Cu2+, which leads to the deterioration of the piezoelectric and mechanical properties of BCZT ceramics. The present study
properties. Furthermore, the real nature of the phase boundary in demonstrated that (Ba0.85Ca0.15Ti0.90Zr0.10)O3–0.8 wt.% Ba
the BCZT ceramics requires systematic investigations in the future. (Cu0.5W0.5)O3 ceramics with a lower sintering temperature were
promising candidates for lead-free piezoelectric applications.
3.6. The mechanical properties of the ceramics
Acknowledgements
The hardness and fracture toughness of the ceramics sintered at
1220  C for 4 h as a function of the BCW content are plotted in Fig.12.
It is apparent from Fig.12 that the hardness increases with increasing This work was supported by National Science Foundation of
the BCW contents, and falls to half of the free pore ones when BCW is China (NSFC) (Grant Nos. 51107077 and 51172136), the Fundamen-
0.80 wt.%. Therefore, a reduction of the bending strength occurs with tal Research Funds for the Central Universities (Program Nos.
increasing the BCW content from x = 0.00 wt.% to x = 0.80 wt.%. Both GK201305013, GK201301002, GK201102003, GK201101004 and
the hardness and fracture toughness peaked at 78 MPa and GK201403006), Science and Technology Program of Shaanxi
4.7 MPa m1/2, respectively, when the BCW concentration was Province and Xi'an City (Grant Nos. 2013K09-26 and CXY1342
0.80 wt.%. In addition, with increasing the BCW content, a fine- (4)), Fundamental Research Funds of Experimental Technology for
[(Fig._12)TD$IG] Shaanxi Normal University (Program No. SYJS201219).

References

[1] D. Damjanovic, A. Biancoli, L. Batooli, A. Vahabzadeh, J. Trodahl, Appl. Phys.


Lett. 100 (19) (2012) 192907.
[2] J.G. Wu, D.Q. Xiao, W.J. Wu, Q. Chen, J.G. Zhu, Z.C. Yang, J. Wang, Scripta Mater.
65 (2011) 771–774.
[3] J.H. Gao, D.Z. Xue, Y. Wang, D. Wang, L.X. Zhang, H.J. Wu, S.W. Guo, H.X. Bao, C.
Zhou, W.F. Liu, S. Hou, G. Xiao, X.B. Ren, Appl. Phys. Lett. 99 (2011) 092901.
[4] K. Chung, D. Lee, J. Yoo, Y. Jeong, H. Lee, H. Kang, Sens. Actuators A: Phys. 121
(2005) 142–147.
[5] Y. Tian, L.L. Wei, X.L. Chao, Z.H. Liu, Z.P. Yang, J. Am. Ceram. Soc. 96 (2) (2013)
496–502.
[6] J.G. Wu, D.Q. Xiao, W.J. Wu, Q. Chen, J.G. Zhu, Z.C. Yang, J. Wang, J. Eur. Ceram.
Soc. 32 (2012) 891–898.
[7] M.C. Ehmke, F.H. Schader, K.G. Webber, J. Rodel, J.E. Blendell, K.J. Bowman, Acta
Mater. 78 (2014) 37–45.
[8] C.Y. Chang, H.I. Ho, T.Y. Hsieh, C.Y. Huang, Y.C. Wu, Ceram. Int. 39 (2013)
8245–8251.
[9] J.G. Wu, A. Habibul, X.J. Cheng, X.P. Wang, B.Y. Zhang, Mater. Res. Bull. 48 (2013)
4411–4414.
[10] J.P. Praveen, K. Kumar, A.R. James, T. Karthik, S. Asthana, D. Das, Curr. Appl.
Fig. 12. Hardness and fracture toughness of BCZT ceramics with different BCW Phys. 14 (2014) 396–402.
contents.
X. Chao et al. / Materials Research Bulletin 66 (2015) 16–25 25

[11] I.Y. Kang, I.T. Seo, Y.J. Cha, J.H. Choi, S. Nahm, T.H. Sung, J.H. Paik, J. Eur. Ceram. [23] Y. Tsur, T.D. Dunbar, C.A. Randall, J. Electroceram. 7 (2001) 25–28.
Soc. 32 (2012) 2381–2387. [24] W.F. Liu, X.B. Ren, Phys. Rev. Lett. 103 (25) (2009) 257602.
[12] H.I. Hsiang, C.S. Hsi, C.C. Huang, S.L. Fu, Mater. Chem. Phys. 113 (2009) 658–663. [25] M.P. Zheng, Y.D. Hou, F.Y. Xie, J. Chen, M.K. Zhu, H. Yan, Acta Mater. 61 (2013)
[13] X.L. Chao, D.F. Ma, R. Gu, Z.P. Yang, L.R. Xiong, J. Alloys Compd. 49 (2010) 1489–1498.
1698–1702. [26] H.Q. Fan, W.Q. Jie, C.S. Tian, L.T. Zhang, H.E. Kim, Ferroelectrics 269 (2002)
[14] T. Chen, T. Zhang, C. Wang, J.F. Zhou, J.W. Zhang, Y.H. Liu, J. Mater. Sci. 47 (2012) 33–38.
4612–4619. [27] H.L. Du, W.C. Zhou, F. Luo, D.M. Zhu, S.B. Qu, Y. Li, Z.B. Pei, J. Appl. Phys. 104
[15] Y.R. Cui, X.Y. Liu, M.H. Jiang, Y.B. Hu, Q.S. Su, H. Wang, J. Mater. Sci. 23 (2012) (2008) 034104.
1342–1345. [28] R.S. Nasar, M. Cerqueira, E. Longo, J.A. Varela, A. Beltran, J. Eur. Ceram. Soc. 22
[16] X.L. Chao, Z.P. Yang, L.R. Xiong, Z. Li, J. Alloys Compd. 509 (2011) 512–517. (2002) 209–218.
[17] A.A. Kaminskii, M. Akchurin, R.V. Gainutdinov, K. Takaichi, Crystallogr. Rep. 50 [29] R. Eitel, C.A. Randall, Phys. Rev. B 75 (2007) 094106.
(2005) 869–873. [30] S. Zhang, R.E. Eitel, C.A. Randall, T.R. Shrout, E.F. Alberta, Appl. Phys. Lett. 86
[18] X.R. Hou, S.M. Zhou, T.T. Jia, H. Lin, H. Teng, J. Eur. Ceram. Soc. 31 (2011) 733– (2005) 262904.
738. [31] W.H. Li, X.Y. Liu, J.F. Ma, Y. Wu, Y.R. Cui, J. Mater. Sci.: Mater. Electron. 24 (5)
[19] B. Yu, K.M. Liang, S.R. Gu, Ceram. Int. 29 (2003) 695–698. (2013) 1551–1555.
[20] T.K. Lea, D. Flahaut, H. Martinez, T. Pigot, H.K.H. Nguyen, T.K.X. Huynh, Appl. [32] D. Damjanovic, J. Am. Ceram. Soc. 88 (2005) 2663.
Catal. B: Environ. 144 (2014) 1–11. [33] S. Zhang, S.M. Lee, D.H. Kim, H.Y. Lee, T.R. Shrout, Appl. Phys. 90 (2007) 232911.
[21] A.A. Bokov, Z.G. Ye, J. Mater. Sci. 41 (1) (2006) 31–52. [34] X.Z. Liu, S.J. Zhang, J. Luo, T.R. Shrout, W. Cao, J. Appl. Phys. 106 (2009) 074112.
[22] Q. Gou, J.G. Wu, A.Q. Li, B. Wu, D.Q. Xiao, J.G. Zhu, J. Alloys Compd. 521 (2012) [35] X.J. Cheng, J.G. Wu, X.P. Wang, B.Y. Zhang, J.G. Zhu, D.Q. Xiao, X.J. Wang, X.J. Lou,
4–7. Appl. Phys. Lett. 103 (2013) 052906.

Das könnte Ihnen auch gefallen