Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

J Mater Sci: Mater Electron (2014) 25:2540–2545

DOI 10.1007/s10854-014-1907-1

Temperature-dependent electrical properties


of 0.5Pb(Ni1/3Nb2/3)O3–(0.5 2 x)PbTiO3–xPbZrO3 piezoceramics
near the morphotropic phase boundary
Jia-Jun Zhou • Ke Wang • Jing-Feng Li •
Xiao-Wen Zhang • Hong Liu • Jing-Zhong Fang

Received: 25 January 2014 / Accepted: 27 March 2014 / Published online: 3 April 2014
Ó Springer Science+Business Media New York 2014

Abstract 0.5Pb(Ni1/3Nb2/3)O3–(0.5 - x)PbTiO3–xPbZrO3 1 Introduction


(x = 0.10–0.20) polycrystalline ceramics were fabricated by
solid-state reaction method. The phase structures, micro- Pb(Zr, Ti)O3 (PZT)-based piezoceramics have been widely
structure and temperature dependence of the electrical used in sensors, transducers, actuators, and other important
properties were investigated in detail. Both X-ray diffraction electric devices for decades due to their excellent electrical
and Raman spectroscopy results indicated that the morpho- properties [1]. The morphotropic phase boundary (MPB)
tropic phase boundary (MPB) separating tetragonal from plays a key role in the composition–structure–property
rhombohedral phases located around the region with relationship of the ceramics that the piezoelectric proper-
x = 0.14–0.16. Well-saturated ferroelectric P–E loops were ties can be significantly enhanced when the composition
obtained for all compositions and the compositions near locates in the vicinity of MPB. It is commonly known that
MPB possessed the relative large remnant polarization the free energy profile demonstrates a flat behavior near
Pr * 24 lC/cm2 and low coercive field Ec * 5 kV/cm. MPB, which induces the flexible phase transition between
High performance of electric-induced-strain property with two different phases, i.e. tetragonal and rhombohedral.

d33 (Smax/Emax) up to 1,140 pm/V could be achieved in the Thus, the polarization directions will increase as a result of
composition with x = 0.15. When the content of PbZrO3 the combined directions of the two phases, leading to the
was 14 mol%, the ceramics showed slightly lowered pie- improvement of electrical properties [2, 3]. Two kinds of
 MPB known as MPB-I and MPB-II exist in the composi-
zoelectric coefficient d33 of 898 pm/V. But excellent tem-
perature stability of piezoelectric property was exhibited that tions of PT–PZ and relaxor–PT, respectively [4]. Because
 of their better tailored electrical properties than binary PZT
the d33 changed\10 % when the temperature increased from
room temperature to 120 °C. The piezoelectric property of system, relaxor–PT–PZ ternary systems (e.g. Pb(Ni1/3Nb2/3)
this solid solution makes it promising for application in O3–PT–PZ [5–9], Pb(Mg1/3Nb2/3)O3–PT–PZ [10, 11] and
multilayer piezoactuators. Pb(Zn1/3Nb2/3)O3–PT–PZ [12, 13] ) have been investigated
extensively, whose MPB follows a linear region between
MPB-I and MPB-II.
Pb(Ni1/3Nb2/3)O3 (PNN) is a relaxor ferroelectric with a
broad maximum of dielectric permittivity around -120 °C
and a pseudocubic crystal structure at room temperature
[14]. Considerable efforts have been made to obtain high
J.-J. Zhou (&)  H. Liu  J.-Z. Fang
electrical performance of PNN–PT–PZ system [5–9]. A
Institute of Optics and Electronics, Chinese Academy of
Sciences, Chengdu 610209, China peak value of piezoelectric constant d33 and electrome-
e-mail: [email protected] chanical coupling factor k33 were obtained in the vicinity of
0.5PNN–0.35PT–0.15PZ composition by Luff et al. [15]
K. Wang  J.-F. Li  X.-W. Zhang
and Kondo et al. [16], respectively. Vittayakorn et al. [17]
State Key Laboratory of New Ceramics and Fine Processing,
School of Materials Science and Engineering, Tsinghua investigated the PNN–Pb(Zr0.5T0.5)O3 solid solution
University, Beijing 100084, China and found that the MPB existed in the vicinity of

123
J Mater Sci: Mater Electron (2014) 25:2540–2545 2541

0.2PNN–0.8PZT. Another composition near the MPB of


0.3PNN–0.39PT–0.31PZ with excellent piezoelectric
properties was found by Cao et al. [18]. These studies
demonstrate the promising applications of this type of
ferroelectric ceramic in multilayer piezoactuators. For
applications in piezoactuators, the electric-induced-strain

property (converse piezoelectric constant d33 ) is the pri-
mary parameter. However, limited information about the
electric-induced strain behavior of this system has been
reported in the literature [19–21]. In addition, the temper-
ature stability of piezoelectric property is another important
parameter for piezoactuators ceramics which needs to be
thoroughly investigated. Obtaining desired strains in an
Fig. 1 XRD patterns of 0.5PNN–(0.5 - x)PT–xPZ piezoceramics
operation window is substantially important for multilayer with 0.1 B x B 0.2, a 2h from 20° to 60°, b enlarged area for (002)
piezoactuators since the working temperature varies with peaks
the environment and self-heating phenomenon.
This work aims to investigate the temperature depen-
dence of electric properties of 0.5Pb(Ni1/3Nb2/3)O3– employed to investigate the phase structures of 0.5PNN–
(0.5 - x)PbTiO3–xPbZrO3 perovskite solid solution and (0.5 - x)PT–xPZ system in this work. The Raman spectra
obtain high-strain piezoceramics with good thermal sta- was measured by the LabRAM HR800 (France) with
bility. The ceramics were fabricated by solid-state reaction 633 nm radiation. The cross-sectional microstructures of
method and systematic experiments were carried out to the samples were observed by field emission scanning
investigate the phase structure, microstructure and tem- electron microscopy (FESEM, S-7001F, JEOL, Japan). The
perature-dependent piezoelectric properties. ferroelectric properties and temperature-dependent elec-
tric-field-induced strains were measured by TF ANA-
LYZER 1000 (aixACCT Systems GmbH, Germany)
2 Experimental procedure equipped with a laser interferometer. The converse piezo-

electric constant d33 was calculated by the value of electric-
0.5Pb(Ni1/3Nb2/3)O3–(0.5 - x)PbTiO3–xPbZrO3 piezoce- field-induced strain (Smax/Emax) under 1 kV/mm. These
ramics (abbreviated as 0.5PNN–(0.5 - x)PT–xPZ, measurements involved the application of a triangular
x = 0.10, 0.12, 0.14, 0.15, 0.16, 0.18, 0.20) were prepared voltage waveform with a frequency of 1 Hz. Evaluation of
by modified columbite method. The Relaxor B-site oxide the temperature-dependent dielectric permittivity was car-
NiNb2O6 was synthesized by mixing Ni(CH3COO)24H2O ried out in a temperature-regulated chamber, which was
with Nb2O5 and calcining at 1,000 °C for 6 h. PbO connected with Agilent 4194 (Hewlett-Packard, Palo Alto,
(1.8 mol% excess), TiO2 and ZrO2 powders of high purity CA) at 500 Hz, 1, 10 and 100 kHz.
(99.0 wt%) were mixed with NiNb2O6 according to the
stoichiometric composition, and milled for 24 h in ethanol
using ZrO2 balls. Then, the slurry was dried and calcined at 3 Results and discussions
850 °C for 3 h. The synthesized powder were ball milled
again for 24 h also in ethanol and dried, after that the Figure 1 shows the representative XRD patterns of
powders were pressed into small disks of 10 mm in 0.5PNN–(0.5 - x)PT–xPZ piezoceramics in the 2h range
diameter and 1.2 mm in thickness. Finally, these green from 20° to 60°. As can be seen, well crystallized perov-
pellets were sintered around 1,250 °C for 2 h in a lead skite structures were formed for all compositions. Although
atmosphere by using PbZrO3 powders. After being painted the columbite method has been adopted to synthesize the
with sliver electrodes, the samples were poled at 75 °C for ceramics, the trace of Pb2Nb2O7 was still detected in the
10–20 min under an electric field of 2 kV/mm in silicone sample because this pyrochlore phase easily forms in ter-
oil. nary relaxor ferroelectric ceramics [8, 19]. The (002)/(200)
The crystal structure of the sintered samples was peaks of the samples vary with the content of PbZrO3, as
determined by X-ray diffraction (XRD) characterization enlarged in Fig. 1b, which indicates a phase transition from
with monochromatic Cu Ka radiation (Rigaku, D/Max tetragonal to rhombohedral [18]. The dominant tetragonal
2500, Tokyo, Japan). Raman spectroscopy has been and rhombohedral phases coexist in the sample with low
suggested as a powerful tool to study the structural changes content of PZ (x = 0.12), identified by the split of (002)
in the PZT-based piezoceramics [17, 22], and was peaks around 2h = 45°. However, with the increasing

123
2542 J Mater Sci: Mater Electron (2014) 25:2540–2545

Fig. 2 a Representative Raman spectra of 0.5PNN–(0.5 - x)PT–xPZ


piezoceramics and the inset denotes the fitted Raman spectrum
(650–900 cm-1), b Raman shifts and the relative intensity ratios of
I(p1)/I(p2) for A1g mode with the variations of x

content of PZ, the split of (002) peaks decreases, and there Fig. 3 Representative SEM images of the fracture surfaces of
is almost only one peak observed for 0.5PNN–0.3PT– 0.5PNN–(0.5 - x)PT–xPZ piezoceramics, a x = 0.10, b x = 0.18
0.2PZ sample, indicating the existence of a dominant
rhombohedral structure. The suggested MPB consisting as the content of PZ increases from 0.10 to 0.20, as does
rhombohedral and tetragonal phases is in the vicinity of the relative intensity ratio I(p1)/I(p2), which indicates a
x = 0.15, according to their XRD patterns. phase transition from tetragonal to rhombohedral phase.
Figure 2a shows the representative Raman spectra of Between x = 0.14–0.16, a flat behavior of I(p1)/I(p2) as well
0.5PNN–(0.5 - x)PT–xPZ system and the fitted spectrum as the shift of p1 and p2 versus compositions is observed,
between 650 and 900 cm-1 of each sample (inset). The indicating the existence of a intermediate state and the
analyzed results are presented in Fig. 2b. From the whole location of MPB. The phase transition derived from the
spectra in the range from 100 to 1,000 cm-1, it appears that analysis of the Raman spectra agrees well with the XRD
there is little difference among the all compositions. But a results.
continuous change with the variations of compositions Figure 3 shows the representative SEM images of the
could still be found as careful investigations were taken. fracture surfaces of 0.5PNN–(0.5 - x)PT–xPZ piezoce-
The peaks in the Raman spectra of the PZT-based samples ramics. Dense microstructures have been obtained for all
between 650 and 900 cm-1 originated from the stretching samples since few pores are left and the grains are well-
vibration mode of B–O, which is known as A1g mode. The grown. Non-uniform distribution of grain size is found in
A1g mode comprises two peaks centered at around the ceramics, in which the grain size varies from 2 to 6 lm.
730 cm-1 (denoted as p1) and 810 cm-1 (denoted as p2), The PZ content seems to have little influence on the grain
which represent the existence of tetragonal and rhombo- size of ceramics, which is consistent with the results of a
hedral symmetry, respectively [23]. Thus, the phase previous report [16].
structures of the investigated compositions could be iden- Figure 4 shows the temperature dependence of relative
tified by the variations of A1g mode. As shown in Fig. 2b, dielectric permittivity (er) of the 0.5PNN–(0.5 - x)PT–
the locations of p1 and p2 for A1g mode decrease obviously xPZ system measured at 500 Hz, 1, 10 and 100 kHz. Two

123
J Mater Sci: Mater Electron (2014) 25:2540–2545 2543


Fig. 5 Room-temperature electrical properties of d33 , kp and Qm of
0.5PNN–(0.5 - x)PT–xPZ piezoceramics

Fig. 4 Temperature dependence of relative permittivity (er) for


0.5PNN–(0.5 - x)PT–xPZ piezoceramics
ref [19]. The Qm shows a minimum value in the vicinity of
MPB, whereas it decreases from 106 to 50 when the con-
peaks can be observed for the samples with x = 0.14–0.20 tent of PZ increases from 0.10 to 0.14.
at each frequency. The peak located at lower temperature The representative strain–electric-field (S–E) curves and
 
represents the ferroelectric phase transition from rhombo- temperature dependence of d33 =d33RT of the 0.5PNN–
hedral phase to tetragonal phase (TR–T) [17], while the (0.5 - x)PT–xPZ system are shown in Fig. 6. All the
other peak at higher temperature is the ferroelectric–para- compositions show typical linear S–E curves and the hys-
electric transition point (Tc). The Tm shifts to higher tem- teresis at room temperature is a result of domain wall
perature as the frequency increases, demonstrating the movement. The sample with x = 0.15 shows the largest
pronounced relaxor ferroelectric behavior. For x = 0.12, strain (up to 0.11 %) than any other composition under
the Tm increases by 2 °C when the frequency increases E = 1 kV/mm, which is attributed to the MPB effect. The
from 500 Hz to 100 kHz, whereas the DTm is 4 °C for the temperature dependence of the piezoelectric property plays
0.5PNN–0.3PT–0.2PZ composition under the same condi- a very important role in the actual applications, especially
tion. This result indicates a strengthened relaxor behavior for piezoactuator ceramics. The temperature dependence of
 
with increased PZ amount. The Tm (500 Hz) of this system d33 =d33RT for each sample varies with the PZ content, as
shifts from 166 to 132 °C as the content of PT decreases shown in Fig. 6b. The 0.5PNN–0.35PT–0.15PT sample
from 0.40 to 0.30, which is quite reasonable since the PT demonstrates superior electric-induced strain behavior at
has a higher Curie temperature (Tc) than that of PZ room temperature; however, its piezoelectric property is
(230 °C) or PNN (-120 °C). The composition with quite thermally unstable and decreases drastically when the
x = 0.15 demonstrates the highest er up to 5,215 (1 kHz) temperature increases, which is undesirable for piezoactu-
among the all compositions at room temperature, indicat- ator applications since the self-heating is inevitable [24,

ing the optimized dielectric property around MPB. 25]. For samples with 14 and 18 mol% PZ, the d33
Figure 5 displays the room-temperature electrical increases initially and then decreases with the increasing

properties of the converse piezoelectric constant d33 , the temperature, showing a peak value at 60 and 90 °C
planar electromechanical coupling factor kp and the respectively, corresponding to the TR–T. The 0.5PNN–
mechanical quality factor Qm for 0.5PNN–(0.5 - x)PT– 0.36PT–0.14PT composition demonstrates a slightly lower

xPZ piezoceramics with 0.10 B x B 0.20. The electric- piezoelectric constant (d33 = 898 pm/V) than the sample

induced strain method was used to measure d33 under the with x = 0.15, but has greater thermally-stable piezoelec-
 
electric field of 1 kV/mm. Both the d33 and kp show peak tric properties. The change in d33 of the sample is less than
values of 1,140 pm/V and 0.56 for the composition with ±10 % at 120 °C, indicating the suitability for applications
x = 0.15 due to the MPB effect, which is consistent with in piezoactuators.
the analysis of XRD patterns. As mentioned earlier, the Figure 7 shows the representative P–E loops of the
polarization would become much easier at MPB region 0.5PNN–(0.5 - x)PT–xPZ piezoceramics measured at a
because of the increased polarization directions, leading to frequency of 1 Hz. Well-saturated hysteresis loops are
the improvement of piezoelectric properties. The electric- obtained for all the compositions under an electric field of
induced strain property of this composition in the present 3 kV/mm. The coercive field (Ec) of this system decreases
work is superior to that of a similar composition reported in markedly in the vicinity of MPB, in accordance with the

123
2544 J Mater Sci: Mater Electron (2014) 25:2540–2545

decrease of Qm. The Ec decreases from 9.85 to 4.96 kV/cm


as the PZ content increases from 0.10 to 0.14, whereas the
Pr slightly increases from 22.64 to 24.92 lC/cm2. The
compositions near MPB seem to have the same Pr and
small differences in Ec. The extremely low internal bias
fields (Ei) of all compositions show the ‘‘soft’’ character-
istic of this type of piezoceramics.

4 Conclusions

0.5PNN–(0.5 - x)PT–xPZ piezoceramics with 0.10 B


x B 0.20 were successfully fabricated by columbite pre-
cursor method. The phase structures and temperature-
dependent electrical properties were investigated in detail.
The increasing content of PZ induced a phase transition from
tetragonal to rhombohedral and the MPB was located around
x = 0.14–0.16, identified by the XRD patterns and Raman
spectra. The sample with x = 0.15 showed the peak electric

properties of d33 * 1,140 pm/V, kp * 0.56, er * 5,215
(1 kHz) at room temperature. Furthermore, the optimum
piezoelectric properties with improved temperature stability
could be obtained in the 0.5PNN–0.36PT–0.14PZ ceramics,

whose d33 (*898 pm/V) changed \10 % when the tem-
perature increased from room temperature to 120 °C. This
system demonstrates the promising applications in high-
temperature multilayer piezoactuators.

Acknowledgments This work was supported by West Light


Foundation of the Chinese Academy of Sciences and State Key
Fig. 6 The representative room-temperature S–E curves (a) and Laboratory of New Ceramics and Fine Processing Tsinghua Univer-
 
temperature-dependent d33 =d33RT (b) of 0.5PNN–(0.5 - x)PT–xPZ sity (KF201309).
piezoceramics

References

1. B. Jaffe, W.R. Cook, H. Jaffe, Piezoelectric Ceramics (Academic


Press, London, 1971)
2. D. Damjanovic, Appl. Phys. Lett. 97, 062906 (2010)
3. C.M. Lonkar, D.K. Kharat, H.H. Kumar, S. Prasad, K. Bala-
subramanian, J. Mater. Sci. Mater. Electron. 24, 411 (2013)
4. Y. Yamashita, Jpn. J. Appl. Phys. 33, 5328 (1994)
5. E.A. Buyanova, P.L. Strelets, I.A. Serova, V.A. Isupov, Bull.
Acad. Sci. USSR Phys. Ser. 29, 1877 (1965)
6. M.S. Yoon, H.M. Jang, J. Appl. Phys. 77, 3991 (1994)
7. G. Robert, M.D. Maeder, D. Damjanovic, N. Setter, J. Am.
Ceram. Soc. 84, 2869 (2001)
8. J. Luo, J. Qiu, K. Zhu, J. Du, X. Pang, H. Ji, Ferroelectrics 425,
90 (2011)
9. L. Zhang, Q. Sun, W. Ma, Y. Zhang, H. Liu, J. Mater. Sci. Mater.
Electron. 23, 688 (2012)
10. H. Ouchi, K. Nagano, S. Hayakawa, J. Am. Ceram. Soc. 48, 630
(1965)
11. Z. Xia, Q. Li, S. Zhang, Solid State Commun. 145, 34 (2008)
12. H. Fan, H.E. Kim, J. Appl. Phys. 91, 317 (2002)
13. G.F. Fan, M.B. Shi, W.Z. Lu, Y.Q. Wang, F. Liang, J. Eur.
Fig. 7 Representative P–E loops of 0.5PNN–(0.5 - x)PT–xPZ Ceram. Soc. 34, 23 (2014)
piezoceramics 14. E.F. Alberta, A.S. Bhalla, Mater. Lett. 54, 47 (2002)

123
J Mater Sci: Mater Electron (2014) 25:2540–2545 2545

15. D. Luff, R. Lane, K.R. Brown, H.J. Marshallsay, Trans. J. Br. 21. M.J. Hoffmann, H. Kungl, Curr. Opin. Solid State Mater. Sci. 8,
Ceram. Soc. 73, 251 (1974) 51 (2004)
16. M. Kondo, M. Hida, M. Tsukada, K. Kurihara, N. Kamehara, Jpn. 22. A. Slodczyk, P. Daniel, A. Kania, Phys. Rev. B 77, 4114 (2008)
J. Appl. Phys. 36, 6043 (1997) 23. N. Luo, Q. Li, Z. Xia, X. Chu, J. Am. Ceram. Soc. 95, 3246
17. N. Vittayakorn, G. Rujijanagul, X. Tan, M.A. Marquardt, D.P. (2012)
Cann, J. Appl. Phys. 96, 5103 (2004) 24. E. Sapper, A. Gassmann, L. Gjødvad, W. Jo, T. Granzow, J.
18. R. Cao, G. Li, J. Zeng, S. Zhao, L. Zheng, Q. Yin, J. Am. Ceram. Rödel, J. Eur. Ceram. Soc. 34, 653 (2014)
Soc. 93, 737 (2010) 25. J. Zheng, S. Takahashi, S. Yoshikawa, K. Uchino, J.W.C. De
19. S. Mahajan, C. Prakash, O.P. Thakur, J. Alloys Compd. 471, 507 Vries, J. Am. Ceram. Soc. 79, 3193 (1996)
(2009)
20. M. Kondo, M. Tsukada, K. Kurihara, Jpn. J. Appl. Phys. 38, 5539
(1999)

123

You might also like