You are on page 1of 9

Available online at www.sciencedirect.

com

ScienceDirect
Acta Materialia 79 (2014) 84–92
www.elsevier.com/locate/actamat

Large piezoelectric properties induced by doping ionic pairs


in BaTiO3 ceramics
D. Xu a, W.L. Li a,b, L.D. Wang a, W. Wang a, W.P. Cao a, W.D. Fei a,⇑
a
School of Materials Science and Engineering, Harbin Institute of Technology, Harbin 150001, People’s Republic of China
b
National Key Laboratory of Science and Technology on Precision Heat Processing of Metals, Harbin Institute of Technology, Harbin 150001,
People’s Republic of China

Received 13 February 2014; received in revised form 15 June 2014; accepted 8 July 2014

Abstract

The high piezoelectric properties of piezoelectric ceramics are typically obtained by inducing a phase transition between two ferro-
electric phases, using either the morphotropic phase boundary (MPB) or the polymorphic phase transition (PPT). Here we demonstrate
that neither the MPB nor the PPT is necessary to achieve high piezoelectric properties. Our results show that the optimized distribution
of Li+–Al3+ pairs parallel to the [0 0 1] direction, as found in our xLiAlSiO4/BaTiO3 (xLAS/BT) lead-free piezoelectric ceramic system
prepared from ordinary raw materials by conventional solid-state reaction sintering, can generate a large piezoelectric constant (d33) of
378 pC/N when x = 7.5 mol.%. The d33 value of the 7.5 mol.% LAS/BT ceramic is more than three times that of the pure BT ceramic.
The distortion connected to the proposed Li+–Al3+ pairs locally creates unit cells of less than tetragonal symmetry, and these low-
symmetry cells can be responsible for the high piezoelectric response. The high-temperature stability testing reveals that these doped
ceramics are usable at temperatures as high as 120 °C. This piezoelectric mechanism coming from the doping ionic pairs provides a
new method to achieve large piezoelectric properties in a wide range of ABO3-type perovskite systems.
Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Perovskites; Piezoelectricity; Lead-free; Doping ionic pairs

1. Introduction have prompted the extensive investigation of lead-free


piezoelectric ceramics with excellent properties as replace-
Pb(Zr,Ti)O3 (PZT) ceramics have for many years repre- ments for toxic PZT ceramics. The excellent piezoelectric
sented one of the most commonly used piezoelectric mate- properties of PZT ceramics with MPB compositions have
rials and have played a vital role in various applications, led to interest in a variety of binary and ternary lead-free
such as actuators, sensors, capacitors, resonators and piezoelectric ceramics with similar MPB compositions,
transducers. The exceptional piezoelectric properties of including Na0.5Bi0.5TiO3–BaTiO3 [5–7], BiFeO3–BaTiO3
these materials are achieved by using a composition [8,9], Na0.5Bi0.5TiO3–K0.5Bi0.5TiO3–BaTiO3 [10–12], Na0.5
approximating Pb(Zr0.52Ti0.48)O3, which exhibits a compo- Bi0.5TiO3–K0.5Bi0.5TiO3–BiFeO [13] and Na0.5K0.5NbO3–
sition-induced phase transition between two ferroelectric LiSbO3–BiFeO3 [14]. However, the piezoelectric perfor-
phases known as a morphotropic phase boundary (MPB) mance of existing lead-free ferroelectric systems still cannot
[1–4]. Recently, health and environmental considerations compete with that of PZT, especially with regard to the
performance of high-end PZT ceramics (those with a high
piezoelectric constant (d33) of 600 pC/N and a high depo-
⇑ Corresponding author. Tel./fax: +86 451 86413908. larization temperature (Td) of 330 °C) [15–17]. Liu and
E-mail address: wdfei@hit.edu.cn (W.D. Fei).

http://dx.doi.org/10.1016/j.actamat.2014.07.023
1359-6454/Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
D. Xu et al. / Acta Materialia 79 (2014) 84–92 85

Ren [18] have reported that the Ba(Ti0.8Zr0.2)O3–x(Ba0.7- sintering method, so, unlike PZT, BT is used primarily as
Ca0.3)TiO3 (BZT–BCT) ceramic system exhibits a very a dielectric material rather than as a piezoelectric material.
large d33 value of 620 pC/N at an optimal composition. The reason why good piezoelectric properties cannot be
Their work involved designing an MPB material analogous obtained with ordinary BT ceramics remains a mystery.
to the PZT system and found that the high piezoelectricity To date, the basic approach to the design and optimiza-
of the BZT–BCT system originates from the compositional tion of new ABO3-related piezoceramics has been to induce
proximity of its MPB to the tricritical triple point (TCP) a phase transition between two ferroelectric phases, as
formed from combined cubic paraelectric (C), ferroelectric exemplified by MPB and PPT. The question now arises
rhombohedral (R) and tetragonal (T) phases. Unfortu- as to whether or not there is a new high piezoelectric mech-
nately, its relatively low Curie temperature (TC) limits the anism waiting to be found, and the answer may involve
overall usefulness of the BZT–BCT system in practical investigating piezoelectric ceramics with ABO3-type perov-
applications since the operational temperature of some pie- skite structures. Our recent work indicates that the addi-
zoelectric components is limited to half the TC value. More tion of eucryptite (LiAlSiO4, LAS) can greatly enhance
effort is therefore required to further increase the TC of the the piezoelectric properties of BT ceramics prepared from
BZT–BCT system. ordinary raw materials by conventional solid-state reaction
In addition to the MPB, another means of achieving a sintering. We have found that this high piezoelectricity is
high level of piezoelectricity is polymorphic phase transi- dependent on a preferential distribution of Li+–Al3+ pairs
tion (PPT), which is mainly found in the (K,Na)NbO3 in the ABO3-type lattice and is not related to the MPB and
(KNN) system [19–22]. As an example, a piezoelectric coef- PPT effects, thus providing a new means of designing lead-
ficient as high as 416 pC/N has been exhibited by a tex- free piezoelectric materials. Moreover, the high depolariza-
tured KNN-based ceramic with codopants of Li, Ta and tion temperature of these materials has encouraged us to
Sb [21], while the d33 values of non-textured KNN-based investigate xLAS/BT ceramic system for practical
ceramics are in the range of 200–300 pC/N. The enhanced applications.
piezoelectric properties evident with the KNN system are
obtained owing to the coexistence of orthorhombic (O) 2. Experimental procedure
and T ferroelectric phases when the PPT temperature
(TO-T) is shifted downward to near room temperature, xLAS/BT ceramics with LAS contents of 0, 4, 7.5 and
and a temperature-driven ferroelectric–ferroelectric phase 10 mol.% were prepared by conventional ceramic process-
transition such as this may provide an easy path for polar- ing techniques. Commercial BT (99.9%, Aladdin Chemistry
ization rotation. Although high piezoelectricity has been Co. Ltd, Shanghai, China) and LAS powders produced in
reported in these KNN-based ceramics, the strong temper- our laboratory [29] were used as the starting materials.
ature dependence of piezoelectric properties stemming These compounds were weighed out in the desired ratios
from associated PPTs means that they are unusable in and ball-milled for 6 h by planetary milling with zirconia
devices requiring high thermal stability. Zuo et al. [19], balls in alcohol. After ball-milling, the resulting slurries
for example, have found that the d33 values of Sb-doped were dried at 80 °C for 12 h, ground and sieved. The mixed
KNN ceramics decrease rapidly with increasing tempera- powders were subsequently pressed into pellets 10 mm in
ture. It is therefore worth noting that good piezoelectric diameter and 1 mm thick using a few drops of 5 wt.% poly-
properties do not always equate to high performance in vinyl alcohol (PVA) as a binder. After burning off the PVA,
piezoelectric ceramics, so there is an urgent need to develop the xLAS/BT pellets embedded in BT powder were sintered
a lead-free candidate that also exhibits high performance. in covered alumina crucibles, heating at a rate of
Barium titanate (BT) ceramic was one of the most 5 °C min1 to a final temperature of 1350 °C, which was
widely used ferroelectric materials prior to the discovery held for 2 h.
of high-performance PZT. Efforts have been devoted to The crystalline structure of each sample was analyzed
improving the piezoelectric properties of BT ceramic and using an Empyrean X-ray diffraction (XRD) system (PAN-
several effective approaches are explored. The addition of alytical) while in situ XRD measurements were carried out
some dopants into BT to replacing the A- and/or B-sites on a Philips X’Pert diffractometer with Cu Ka radiation,
is an effective means of enhancing the ceramic’s piezoelec- operating at 40 kV and 40 mA. The morphology, grain size
tric properties [23–25]. In addition, the use of special fine and local element occupancy were examined with a cold
powders and unusual sintering techniques such as spark field emission scanning electron microscope (Quanta
plasma sintering (416 pC/N) [26], microwave sintering 200F) equipped for energy-dispersive spectroscopy
(370 pC/N) [27] and two-step sintering (460 pC/N) [28] (EDS). Platinum electrodes were evaporated onto the cera-
have also resulted in a high d33. In such cases, the enhanced mic surfaces and annealed at 600 °C for 30 min to allow for
d33 of the BT ceramic is the result of extrinsic contributions electrical properties characterization. The temperature
to its polarizability, associated with submicron grain sizes dependence of the dielectric constant of each sample was
and nanodomain structures. A BT ceramic with high piezo- measured at 1 kHz across the temperature range of
electric properties (d33 > 300 pC/N), however, is difficult to 25–180 °C on a Novocontrol CONCEPT40 broadband
obtain using ordinary raw materials and a conventional dielectric spectrometer. Polarization hysteresis and
86 D. Xu et al. / Acta Materialia 79 (2014) 84–92

strain–electric field curves were determined at different elec- increasing LAS content. The formation of secondary
tric fields using a Radiant Technologies Precision worksta- phases may be associated with the low solubility limit of
tion. The samples were poled at 80 °C in a silicone oil bath LAS in the BT lattice. It is hard to determine the phase
under a DC field of 3 kV mm1 for 30 min. The piezoelec- type for these secondary phases due to the relatively low
tric constant (d33) was measured employing a quasi-static intensities of their diffraction peaks.
piezoelectric constant testing meter (ZJ-4AN, Institute of In order to further identify the impurity phases, surface
Acoustics, Chinese Academy of Science). scanning electron microscopy (SEM) micrographs of
xLAS/BT ceramics with different LAS contents sintered
3. Results and discussion at 1350 °C are displayed in Fig. 2. A remarkable effect of
the LAS addition on the microstructure of the ceramics is
3.1. Phase and microstructure characterization obvious from Fig. 2b, c and d. Firstly, the densification
of the ceramic samples is improved to some extent, with
Fig. 1a shows XRD patterns of xLAS/BT ceramic pow- only a small number of surface pores. Secondly, the mor-
ders with different LAS contents measured at room temper- phologies of the ceramic samples are modified by the intro-
ature. The major diffraction peaks match well with the duction of LAS. For pure BT ceramic (Fig. 2a),
ABO3-type perovskite phases. However, small additional inhomogeneous but mostly spherical grains are found; in
peaks indexed with an asterisk in the enlarged XRD pat- contrast, a certain number of rod-shaped grains are
terns are observed in those ceramics with added LAS, as observed in the grain boundaries, e.g. in the red circular
revealed in Fig. 1b, and their intensities increase with regions in the SEM images (Fig. 2b, c and d). The excep-
tional rod-shaped grains are probably related to the impu-
rity phases. Finally, the average grain size is seen to initially
increase before reaching a maximum value of about 40 lm
(in the case of the 4 mol.% LAS/BT ceramic), then
decreases slightly with further increasing LAS content.
The large grain sizes suggest that the enhanced d33 of the
xLAS/BT ceramics is not the result of extrinsic contribu-
tions to the polarizability, which are typically associated
with fine grains (1 lm) [30,31].
Fig. 3 shows the EDS analysis of the 7.5 mol.% LAS/BT
sample. It is very hard to determine the respective Ba/Ti
compositions due to their overlapping spectrum peaks.
Hence, the composition table shown in the inset of Fig. 3
lists the sum of the Ba and Ti contents. The corresponding
EDS analysis shown in Fig. 3a reveals a composition close
to an Al:Si ratio of 1.69:1.95 for a normal spherical grain,
confirming that the LAS with such a content had doped the
BT lattice. The possible secondary phase was also analyzed
by EDS, with the result shown in Fig. 3b. Compared with
the matrix phase, the interfacial rod-shaped inclusion has
higher signal intensities in Al and Si, especially in Al. This
indicates that the secondary phases with an Al:Si ratio of
2 may be Al- and Li-enriched, though the Li content can-
not be detected by EDS due to its low element number. It is
clear that the presence of Al- and Li-enriched phases is
caused by the excess of A-site ions after Al3+ and Li+ ions
replacement for A-site Ba2+ ions. The identified impurities
phases confirm a low solubility limit of LAS in the BT lat-
tice, which is consistent with the XRD analysis above.

3.2. Lattice distortion and role of defects

As stated above, the LAS has doped the BT lattice and


the precipitated phases are Al-enriched due to the low sol-
ubility limit of LAS in BT. To analyze the influence of LAS
Fig. 1. (a) XRD patterns of xLAS/BT ceramic powders measured at room doping on the BT lattice distortion, fine scan XRD patterns
temperature and (b) the enlarged views of XRD patterns in the selected 2h obtained at room temperature were performed, and these
range showing the additional diffraction peaks on LAS doping. are shown in Fig. 4a. The diffraction data confirms that
D. Xu et al. / Acta Materialia 79 (2014) 84–92 87

Fig. 2. Surface SEM micrographs of xLAS/BT ceramics: (a) x = 0 mol.%, (b) x = 4 mol.%, (c) x = 7.5 mol.% and (d) x = 10 mol.%. The red circular
regions correspond to the secondary phases. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

the xLAS/BT powders exhibit P4 mm symmetry (JCPDS tion- or internal stress-induced evolution can be ruled out
#83–1880), as evidenced by the relative intensity ratios of completely, since the XRD experiments were carried out
the 002/200 and 202/220 diffraction peaks and the existence on the powders of xLAS/BT ceramics. Moreover, the
of a single 111 peak at 2h  38.9°. The effect of the LAS observed variations in lattice parameters and I002/I200 with
content on the lattice parameters (a and c), as estimated LAS content cannot be explained solely by the random dis-
from XRD full pattern fitting by HighScore Plus software, tribution of Li+ and Al3+ ions. Assuming that Li+–Al3+
as well as the tetragonality (c/a) of the xLAS/BT ceramic ions randomly locate at A-sites, the same trend in the var-
powders, are shown in Fig. 4b. Both the lattice parameter iation of the diffraction intensity of the 002 and 200 peaks
c and the tetragonality are seen to decrease significantly would be evident, but reductions in the I002/I200 intensity
with increasing LAS content in spite of the slight reduction ratio with increasing LAS content would not occur since
in the lattice parameter a. the same effect would apply to the structure factor (F) asso-
The observed variations in lattice parameters are obvi- ciated with the 002 and 200 diffractions. It is therefore nec-
ously caused by the incorporation of LAS into the BT lat- essary to provide an alternative explanation for our
tice. Taking ionic valences into consideration, it is believed experimental results.
that Li+ and Al3+ ions, both of which have relatively small Since the average of the ionic radii of Li+ and Al3+ is
ionic radii, substitute the A-sites occupied by the Ba2+ ions less than the radius of Ba2+, the substitution of either
of the ABO3-type perovskite structure, while Si4+ ions Li+ or Al3+ for Ba2+ will induce lattice distortions in the
replace the B-sites occupied by Ti4+ [32]. In such cases, BT, reducing the XRD intensity [33]. In the case of the
the average A-site valence will be 2 + and hence the local (0 0 2), (0 2 0) and (2 0 0) planes, the effective structure fac-
charge neutrality is maintained. To thoroughly understand tors, F eff eff eff
002 , F 020 and F 200 ; adjusted for lattice distortions,
and identify the occupation behavior of Li+ and Al3+ ions, can be expressed as follows:
the relative intensity ratio of the 002/200 peaks (hereafter
8p2 sin2 h 2
referred to as I002/I200) was analyzed as a function of F eff
200 ¼ F 200 expð Dx Þ ¼ F 200 expðM 1 Þ ð1Þ
LAS content. Fig. 4c shows that the (I002/I200)x/(I002/ k2
I200)0 value (where the superscripts 0 and x denote pure 8p2 sin2 h 2
BT and LAS-doped BT ceramic powders, respectively) F eff
020 ¼ F 020 expð Dy Þ ¼ F 020 expðM 2 Þ ð2Þ
k2
tends to decrease with increasing LAS content. It should
8p2 sin2 h 2
be pointed out that the probability of preferred orienta- F eff
002 ¼ F 002 expð Dz Þ ¼ F 002 expðM 3 Þ ð3Þ
k2
88 D. Xu et al. / Acta Materialia 79 (2014) 84–92

Li+–Al3+ pairs tend to adopt a preferential alignment such


that most of them are parallel to the [0 0 1] direction.
Fig. 4d presents a diagram of the lattice distortion induced
by the preferential distribution of Li+–Al3+ pairs parallel
to the [0 0 1] direction. Here the lattice distortion induced
by doped Li+–Al3+ pairs with a small average ionic radius
occurs primarily along the [0 0 1] direction. As can be seen
in Fig. 4b, the (0 0 2) interplanar spacing (d002) is obviously
decreased because of the lattice distortion induced by the
preferential substitution of Li+–Al3+ pairs parallel to the
[0 0 1] direction. A decrease in the (2 0 0) and (0 2 0) interpla-
nar spacings (referred as d200 and d020, respectively), how-
ever, does not occur since the lattice distortion has
minimal influence on these spacings. The value of d002 thus
decreases more rapidly than the value of d200 with increas-
ing LAS content. In addition, the rate of reduction of the
(I002/I200)x/(I002/I200)0 ratio tends to slow at an LAS con-
tent above x = 7.5 mol.% (as in Fig. 4c), demonstrating
that the number of [1 0 0]- and [0 1 0]-oriented Li+–Al3+
pairs begins to increase as a result of elastic energy
limitations.
According to the chemical formula of LiAlSiO4, how-
ever, for such a replacement of Li+–Al3+ pairs for Ba2+
and Si4+ for Ti4+, if the excess Li and Al cannot be precip-
itated fully in the grain boundaries, there will be either an
excess of A-site ions or a deficiency of B-site ions, accom-
panied by the formation of oxygen vacancies to keep
charge neutrality. As a consequence, the occurrence of oxy-
Fig. 3. EDS analysis for (a) a normal spherical grain and (b) the gen vacancies and Si4+ replacement for Ti4+, which has a
interfacial rod-shaped inclusion in the 7.5 mol.% LAS/BT ceramic. larger radius, can reduce the elastic energy induced by
the lattice distortion stemming from the preferential distri-
where F 200 , F 020 and F 002 are the average structure factors bution of Li+–Al3+ pairs. In particular, the elastic energy
of the distortion-free crystal lattices, and Dx2 , Dy 2 and can be decreased more effectively when the smaller Si4+
Dz2 , which represent deviations from the ideal BT lattice ions and oxygen vacancies are distributed in the vicinity
sites, are the mean square displacements in the [1 0 0], of Li+–Al3+ pairs. Furthermore, the substitution of Ti4+
[0 1 0] and [0 0 1] directions, respectively. Since the diffrac- with the smaller Si4+ is very beneficial for enhancing the
tion intensity (I) is proportional to the square of the mod- solubility of LAS in BT.
ulus of the structure factor (|F|2), the following expression As shown in Fig. 4d, the preferential distribution of
can be obtained: Li+–Al3+ pairs along the [0 0 1] direction may be caused
by the effect of the electric field formed from Li+–Al3+
I i 1 expð2M i Þ ð4Þ
pairs upon C–T phase transition. The detailed discussions
where the subscript i corresponds to the (2 0 0), (0 2 0) and are as follows. Firstly, the paraelectric–ferroelectric phase
(0 0 2) planes. The 200 diffraction includes contributions transition of BT ceramics with ABO3-type perovskite struc-
from the (2 0 0) and (0 2 0) planes. ture is a first-order transition, which includes the nucle-
From the above equations, it can be concluded that lat- ation and growth of the tetragonal ferroelectric phase.
tice distortion is an important factor influencing the I002/ The c-axis orientation of the tetragonal ferroelectric phase
I200 intensity ratio. Based on the results of Fig. 4c, it is evi- is determined by the nucleus orientation. Secondly, the
dent that the values of both M1 and M2 are much smaller nucleation process can be adjusted by changing the electric
than M3, thus lattice distortion has little effect on the dif- field, and it is energetically favorable for the spontaneous
fraction intensity of the 200 peak but reduces the intensity polarization (PS) of ferroelectric nuclei parallel to the elec-
of the 002 peak with increasing LAS content, eventually tric field. Finally, Li+–Al3+ pairs can create an electric
producing the changes in the I002/I200 ratio shown in dipole moment PD, the local electric fields (ED) formed
Fig. 4c. Based on the above analysis, it can be concluded by the dipole at the head and tail of the pairs are nearly
that Li+ and Al3+ ions may form Li+–Al3+ pairs within parallel to PD, and the ED in the middle and at the sides
the same cell in order to minimize electrostatic energy, as around the dipole are nearly antiparallel to PD. The former
indicated in Fig. 4b and c. In the case of A-site replace- ED can lead to a ferroelectric phase nucleus with PS
ment, the relatively large M3 value suggests that these nearly parallel to PD, whereas the latter ED can induce a
D. Xu et al. / Acta Materialia 79 (2014) 84–92 89

Fig. 4. (a) Fine-scan XRD patterns in the 2h ranges of 38.3–39.3o, 43.7–46.3 o and 65.0–66.7o, (b) lattice constants (a and c) and tetragonality (c/a) as a
function of LAS content x, (c) (I(002)/I(200))x/(I(002)/I(200))0 intensity ratio and (d) schematic illustration of the preferential distribution of Li+–Al3+ pairs
parallel to the [0 0 1] direction. The error bars in (b) are a few tenths of the whole.

ferroelectric phase nucleus with PS nearly antiparallel to of the resulting xLAS/BT ceramics, which do not exhibit
PD. Therefore, the nucleation of the tetragonal phase in any obvious signs of leakage, are observed under a higher
the vicinity of Li+–Al3+ dipole causes the coexistence of electric field of 50 kV cm1. The offset behavior in the coer-
PS nearly parallel (PS//PD) and nearly antiparallel to PD cive field reflecting the level of internal bias, which gener-
(-PS//PD) in the xLAS/BT ceramics, which is in good agree- ally occurs in acceptor-doped ferroelectrics [16,34–36], is
ment with the observed symmetric hysteresis loops (see the not observed in the polarization vs. electric field curves of
next section and Fig. 5b). xLAS/BT ceramics measured at room temperature. This
can be attributed to the coexistence of PS//PD and -PS//
3.3. Electrical properties PD in the xLAS/BT ceramics, and PD is distributed ran-
domly in space. Consequently, the total net electric field
The temperature dependence of the dielectric constant from PD is zero, giving rise to the relatively symmetric
(e), as measured at 1 kHz for xLAS/BT ceramics with com- and no internal-bias P–E loops. The situation is different
positions of x = 0, 4, 7.5 and 10 mol.%, are shown in from that in the aged acceptor-doped ferroelectrics with a
Fig. 5a. The dielectric curves of these ceramics only exhibit severely constricted hysteresis loop and a notable bias field,
one dielectric anomaly, characterized by a phase transition where there is a defect dipole moment PD oriented along
from the ferroelectric phase (T) to paraelectric phase (C). the PS (along the crystallographic c-axis), as noted in the
The O–T phase transition cannot be observed in the dielec- previous reports [34–36].
tric–temperature curves within the measuring temperature. The polarization hysteresis loop of the pure BT ceramic
The dielectric constant (em) at TC first increases, reaching a exhibits typical ferroelectric behavior, with a remnant
maximum value of 6245 at x = 7.5 mol.%, then decreases polarization (2Pr) of 25 lC cm2, while the xLAS/BT
with increasing LAS content. In addition, the LAS doping ceramics show a relatively low 2Pr value of 17 lC cm2
has little impact on TC, as evident in Fig. 5a. at x = 4 and x = 7.5 mol.% (Fig. 5c). As the LAS content
Fig. 5b shows the room-temperature polarization hys- increases to 10 mol.%, the hysteresis loop becomes more
teresis loops of xLAS/BT ceramics with compositions of slanted in comparison with those obtained from the 4
x = 0, 4, 7.5 and 10 mol.%. A well-saturated ferroelectric and 7.5 mol.% samples, accompanied by a further decrease
loop is observed for pure BT ceramic under an electric field in 2Pr to approximately 13 lC cm2 at x = 10 mol.%
of 40 kV cm1. After adding LAS, good ferroelectric loops (Fig. 5c). The reduced 2Pr values can be accounted for
90 D. Xu et al. / Acta Materialia 79 (2014) 84–92

Fig. 5. (a) Temperature dependence of the dielectric constant at 1 kHz for xLAS/BT ceramics, in which the highest permittivity peak appears for the
composition of x = 7.5 mol.%, (b) hysteresis loops of xLAS/BT ceramics measured at room temperature and (c) 2Pr values with different LAS contents.

by the decreased tetragonality (c/a) (Fig. 4b) [37,38] and [16,17]. In addition, the high d33 value of the 7.5 mol.%
Li+–Al3+ pairs oriented antiparallel to c-axis because the LAS/BT is on a par with that of 300 pC/N reported for a
orientation of these dipole moments cannot be changed non-textured KNN-based ceramic [21]. This is indeed an
by an external electric field. exciting result, but the question still remains as to why
Fig. 6a presents unipolar strain (S)–electric field (E) these xLAS/BT ceramics have such high piezoelectric con-
curves obtained for poled xLAS/BT ceramics with various stants compared with the pure BT ceramic.
LAS contents at room temperature. In the case of the To clarify the origin of the high piezoelectricity in BT
xLAS/BT ceramics, a curve close to saturation at ceramics doped with Li+–Al3+ pairs, a possible piezoelec-
E P 3.5 kV mm1 and characteristic nonlinearity of unipo- tric mechanism is proposed here. The Li+ and Al3+ ions,
lar strain are observed, compared with the results obtained substituting Ba2+ sites, locate at the interstices formed
for the pure BT ceramic. The maximum converse piezoelec- from four immediately neighbouring BO2- 6 octahedras.
tric coefficient (dS/dE), calculated from the slope of the The replacements of Li+ and Al3+ with small radii for
unipolar S–E plot, achieves an optimum value of the large Ba2+ will lead to two important consequences.
540 pm V1 at x = 7.5 mol.%, comparable to the reported On the one hand, a large lattice distortion can be generated
high d33 values of the BT ceramics prepared by microwave in the vicinity of Li+–Al3+ pairs, as revealed in Fig. 4d, and
sintering [27] and two-step sintering [28]. this distortion can induce a deviation of spontaneous
For comparison purposes, the maximum converse pie- polarization direction from the c-axis near Li+–Al3+ pairs.
zoelectric coefficient and the piezoelectric constant mea- Most of all, the distortion connected to the Li+–Al3+
sured using a quasi-static d33 meter are shown together in dipoles locally creates unit cells of less than tetragonal sym-
Fig. 6b. When increasing the LAS content, the d33 initially metry, and these low-symmetry cells can be responsible for
increases, reaches its highest value of 378 pC/N at the com- the high piezoelectric response, in analogy to the local
position of x = 7.5 mol.%, then declines at x = 10 mol.%. monoclinic phase at the MPB in PZT [39]. On the other
The d33 value (378 pC/N) associated with the 7.5 mol.% hand, because their radii are smaller than that of Ba2+,
LAS/BT is very difficult to achieve for BT-based ceramics Li+ and Al3+ ions, with their relatively high mobility, can
prepared from ordinary raw materials by a conventional respond immediately to the external fields, which contrib-
sintering method, and is much higher than the values exhib- utes to the piezoelectric properties of xLAS/BT ceramics.
ited by existing lead-free piezoelectrics with TC The piezoelectric mechanism proposed in the present
P 120 °C, which primarily show d33 values of 200 pC/N work is different from any of the effects previously noted
D. Xu et al. / Acta Materialia 79 (2014) 84–92 91

the temperature dependence of d33 measured at room tem-


perature after annealing the poled samples at various tem-
peratures for 30 min. It is evident that, while the d33 value
of the pure BT ceramic begins to decrease remarkably
above 50 °C, this is not the case for the xLAS/BT ceramics.
As the temperature increases, the d33 values of these ceram-
ics remain almost unchanged and then drop dramatically
above a critical temperature (Td) of approximately
120 °C, which is very close to the Curie temperature of
these ceramics. Thus the piezoelectric temperature stability
of the BT ceramic can be greatly improved by adding a
small amount of LAS. More importantly, the large temper-
ature dependence and considerable degradation typically
exhibited by KNN-based ceramics are not observed in
our xLAS/BT ceramics, as shown in Fig. 7. Their high-
temperature stability demonstrates that xLAS/BT ceramics
may be used at temperatures as high as 120 °C, a range
which is very hard for pure BT ceramics to reach.
Recent studies on lead-free piezoelectric ceramics have
primarily involved three ABO3-type perovskite systems:
KNN-based, NBT-based and BT-based. In the case of
KNN-based systems, although non-textured KNN ceram-
ics have a relatively high d33 of 200–300 pC/N, their most
significant disadvantage is the large temperature depen-
dence of their piezoelectric properties, such that degrada-
tion occurs through thermal cycling between the two
distinct ferroelectric domain states. In the case of NBT-
Fig. 6. (a) Unipolar strain–electric field curves measured on poled xLAS/ based systems, the inherently low d33 of 200 pC/N is only
BT ceramics, and (b) piezoelectric constant (d33) and maximum converse halfway to the desired piezoelectric value, even though they
piezoelectric coefficient (dS/dE) with different LAS contents. have a moderate depolarization temperature in the range of
100–200 °C [16]. In the case of BT-based systems, much
in the PZT and BZT–BCT systems exhibiting high piezo- effort is still required to enhance the low TC so as to
electricity. The piezoelectric mechanism of the xLAS/BT improve their overall usefulness. In fact, in most cases, high
ceramics is different from that of the BZT–BCT system, values of d33 and Td appear to be mutually exclusive when
as reported by Liu and Ren [18], the decreased energy bar- dealing with lead-free piezoelectric ceramics such as KNN-,
rier is realized by a TCP in the BZT–BCT system and only NBT- and BT-based systems. From the above analysis, the
those ceramics with compositions near the MPB show large current piezoelectric properties of xLAS/BT ceramics are
piezoelectric coefficients. This mechanism is also different
from the high piezoelectricity of the PZT system associated
with the MPB effect. Our work indicates that, if a suitable
preferential distribution of ion pairs is designed (such as
occurred with Li+ and Al3+ in our case), lead-free ceramic
systems prepared from ordinary raw materials by a conven-
tional sintering method have a high probability of exhibit-
ing excellent piezoelectric properties.

3.4. Piezoelectric thermal stability

With regard to the development of piezoelectric ceram-


ics, one of the most serious challenges is to improve their
stability and reliability. To allow piezoelectric ceramics to
function properly under high-temperature conditions, min-
imal piezoelectric temperature dependence is vitally impor-
tant if these devices are to have practical applications.
Fig. 7 shows the temperature dependence of the normalized Fig. 7. Temperature dependence of the normalized d33 values of the poled
d33 values of poled xLAS/BT ceramics measured ex situ. xLAS/BT ceramics measured ex situ. Each datum point in Fig. 7 is the
The depolarization temperature (Td) was determined from average value of several tests.
92 D. Xu et al. / Acta Materialia 79 (2014) 84–92

superior to those of non-textured KNN-based ceramics. [8] Yang HB, Zhou CR, Liu XY, Zhou Q, Chen GH, Li WZ, et al. J Eur
Furthermore, the d33 value (378 pC/N) of the 7.5 mol.% Ceram Soc 2013;33:1177.
[9] Zhou CR, Feteira A, Shan X, Yang HB, Zhou Q, Cheng J, et al. Appl
LAS/BT ceramic is obviously higher than the values Phys Lett 2012;101:032901.
obtained with NBT-based ceramics with similar depolar- [10] Li YM, Chen W, Xu Q, Zhou J, Gu XY, Fang SQ. Mater Chem Phys
ization temperatures (Td  120 °C). These results indicate 2005;943:28.
that there is a real possibility that the xLAS/BT system [11] Dai YJ, Zhang SJ, Shrout TR, Zhang XW. J Am Ceram Soc
may represent a good lead-free piezoelectric ceramic 2010;93:1108.
[12] Shieh J, Wu KC, Chen CS. Acta Mater 2007;55:3081.
material. [13] Zhou CR, Liu XY, Li WZ, Yuan CL. Mater Chem Phys
2009;114:832.
4. Conclusions [14] Jiang MH, Liu XY, Chen GH. Scr Mater 2009;60:909.
[15] Liu W, Xu J, Wang YZ, Xu H, Xi XQ, Yang JL. J Am Ceram Soc
xLAS/BT lead-free piezoelectric ceramics were prepared 2013;96:1827.
[16] Shrout TR, Zhang SJ. J Electroceram 2007;19:111.
using ordinary raw materials and a conventional sintering [17] Takenaka T, Nagata H. J Eur Ceram Soc 2005;25:2693.
method. The 7.5 mol.% LAS/BT ceramic exhibits a very [18] Liu WF, Ren XB. Phys Rev Lett 2009;103:257602.
high piezoelectric constant (d33) of 378 pC/N, a converse [19] Zuo RZ, Fu J, Lv DY, Liu Y. J Am Ceram Soc 2010;93:2783.
piezoelectric coefficient (dS/dE) of 540 pm V1 and a depo- [20] Shen ZY, Li YM, Jiang L, Li RR, Wang ZM, Hong Y, et al. J Mater
larization temperature (Td) of 120 °C. The high piezoelec- Sci: Mater Electron 2011;22:1071.
[21] Saito Y, Takao H, Tani T, Nonoyama T, Tajatiri K, Homma T, et al.
tricity of these LAS-doped BT ceramics primarily Nature 2004;432:84.
originates from the unit cells of less than tetragonal sym- [22] Wang K, Li JF. Adv Funct Mater 2010;20:1924.
metry created by the local distortion connected to the pref- [23] Ang C, Yu Z, Jing Z, Guo RY, Bhalla AS, Cross LE. Appl Phys Lett
erential [0 0 1]-distributed Li+–Al3+ pairs. These properties 2002;80:3424.
indicate that this system represents a promising lead-free [24] Sharma S, Kumar P, Palei P. Ceram Int 2012;38:5597.
[25] Yang WG, Zhang BP, Ma N, Zhao L. J Eur Ceram Soc 2012;32:899.
piezoelectric candidate material worthy of further study. [26] Shen ZY, Li JF. J Ceram Soc Jpn 2010;118:940.
[27] Takahashi H, Numamoto Y, Tani J, Tsurekawa S. Jpn J Appl Phys
Acknowledgements 2008;47:8468.
[28] Moon SM, Wang XH, Cho NH. J Ceram Soc Jpn 2009;117:729.
This work was financially supported by the National [29] Wang LD, Fei WD, Hu M, Jiang LS, Yao CK. Mater Lett
2002;53:20.
Nature Science Foundation of China (Grant No. [30] Zhao Z, Buscaglia V, Viviani M, Buscaglia MT, Mitoseriu L, Testino
11272102). D.X. and W.L.L. contributed equally to this A, et al. Phys Rev B 2004;70:024107.
work. [31] Ghosh D, Sakata A, Carter J, Thomas PA, Han H, Nino JC, et al.
Adv Funct Mater 2013. http://dx.doi.org/10.1002/adfm.201301913.
References [32] Shannon RD. Acta Cryst 1976;A32:751.
[33] Wang YM. X-ray Diffraction of noncrystal and crystal with defects.
1st ed. Beijing: Science Press; 1988 [in Chinese].
[1] Noheda B, Gonzalo JA. Phys Rev B 2000;61:8687.
[34] Ren XB. Nature Mater 2004;3:91.
[2] Asada T, Koyama Y. Phys Rev B 2007;75:214111.
[35] Zhang LX, Erdem E, Ren XB, Eichel R-A. Appl Phys Lett
[3] Schhierholz R, Fuess H. Phys Rev B 2011;84:064122.
2008;93:202901.
[4] Cordero F, Trequattrini F, Craciun F, Galassi C. Phys Rev B
[36] Robels U, Arlt G. J Appl Phys 1993;73:3454.
2013;87:094108.
[37] Ullah A, Ahn CW, Hussain A, Lee SY, Kim IW. J Am Ceram Soc
[5] Babu JB. Appl Phys Lett 2007;90:102901.
2011;94:3915.
[6] Cernea M, Vasile BS, Capiani C, Ioncea A, Galassi C. J Eur Ceram
[38] Fan LL, Chen J, Li S, Kang HJ, Liu LJ, Fang L, et al. Appl Phys Lett
Soc 2012;32:133.
2013;102:022905.
[7] Cernea M, Galassi C, Vasile BS, Capiani C, Berbecaru C, Pintilie I,
[39] Guo R, Cross LE, Park SE, Noheda B, Cox DE, Shirane G. Phys Rev
et al. J Eur Ceram Soc 2012;32:2389.
Lett 2000;84:5423.

You might also like