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

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/274013286

Low-Temperature Sintering of Bi0.5(Na,K)0.5TiO3 for Multilayer Ceramic


Actuators

Article  in  Journal of the American Ceramic Society · March 2015


DOI: 10.1111/jace.13564

CITATIONS READS

14 894

11 authors, including:

Chang Won Ahn Hee Sung Kim


University of Ulsan 151 PUBLICATIONS   3,753 CITATIONS   
215 PUBLICATIONS   4,487 CITATIONS   
SEE PROFILE
SEE PROFILE

Sung Sik Won Hae Jin Seog


Brown University University of Ulsan
28 PUBLICATIONS   684 CITATIONS    18 PUBLICATIONS   179 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Enhancing piezoelectric properties of RF sputtered KNN based thin film by structural curing View project

BNT-ST-BF ternary View project

All content following this page was uploaded by Chang Won Ahn on 08 June 2018.

The user has requested enhancement of the downloaded file.


J. Am. Ceram. Soc., 98 [6] 1877–1883 (2015)
DOI: 10.1111/jace.13564
© 2015 The American Ceramic Society

Journal
Low-Temperature Sintering of Bi0.5(Na,K)0.5TiO3 for Multilayer
Ceramic Actuators
Chang Won Ahn,‡ Hee Sung Kim,‡ Won Seok Woo,‡ Sung Sik Won,‡ Hae Jin Seog,‡ Song A Chae,‡
Bong Chan Park,‡ Ki Bong Jang,§ Yun Po Ok,§ Hyon Ho Chong,§ and Ill Won Kim‡,†

Department of Physics and EHSRC, University of Ulsan, Ulsan 680-749, Korea
§
R&D Center, Samjeon Co. Ltd, Ulsan 689-934, Korea

We investigated the influence of CuO amount (0.5–3.0 mol%), the active layers and internal electrodes. Therefore, it is
sintering temperature (900°C–1000°C), and sintering time important to find combinations of materials that do not have
(2–6 h) on the low-temperature sintering behavior of CuO- these problems.
added Bi0.5(Na0.78K0.22)0.5TiO3 (BNKT22) ceramics. Normal- Ag–Pd internal electrodes are widely used in MLCAs.
ized strain (Smax/Emax), piezoelectric coefficient (d33), and rem- However, Bi ions usually react with Pd elements at high
anent polarization (Pr) of 1.0 mol% CuO-added BNKT22 cofiring temperatures above 1100°C.16,17 Bi0.5Na0.5TiO3-
ceramics sintered at 950°C for 4 h was 280 pm/V, 180 pC/N, based ceramics should be sintered at temperatures higher
and 28 lC/cm2, respectively. These values are similar to those than 1150°C to obtain complete densification. Therefore, if a
of pure BNKT22 ceramics sintered at 1150°C. In addition, we MLCA is produced at a high sintering temperature using
investigated the performance of multilayer ceramic actuators BNT-based ceramics as the active layer material and Ag–Pd
made from CuO-added BNKT22 in acoustic sound speaker as the internal electrode, the actuating performance of the
devices. A prototype sound speaker device showed similar out- MLCA may suffer.
put sound pressure levels as a Pb(Zr,Ti)O3-based device in the Nguyen et al. tried to fabricate MLCAs using BNT-based
frequency range 0.66–20 kHz. This result highlights the feasi- ceramics with Ag–Pd (7:3) internal electrodes.18 However,
bility of using low-cost multilayer ceramic devices made of the Ag–Pd electrodes were not perfectly active at cofiring
lead-free BNKT-based piezoelectric materials in sound speaker temperatures of 1100°C–1140°C, and the electrical properties
devices. of the MLCAs deteriorated compared to those of bulk
ceramics. Low-temperature sintering of ceramics is therefore
important to prevent a decrease in performance of MLCAs
I. Introduction and to produce them at low cost. High sintering tempera-
tures can decrease the stability of piezoelectric properties
I N the past decade, lead-free piezoelectric ceramics such as
(K,Na)NbO3 (KNN) and Bi0.5Na0.5TiO3 (BNT)-based
materials have been widely studied as replacements for lead-
because of the formation of interfacial microdefects and
internal electrode loss. Moreover, the ratio of expensive Pd
based materials because of environmental concerns about in Ag–Pd internal electrodes has to be increased. Therefore,
Pb-containing materials such as lead zirconate titanate, the sintering temperature of BNT-based ceramics needs to be
which contains more than 60 wt% Pb.1–3 decreased for practical application of MLCAs.
Recently, many studies on lead-free piezoelectric ceramics In this study, we investigated the low-temperature sinter-
have been published.3–5 Among various lead-free piezoelec- ing behavior of CuO-added Bi0.5(Na0.78K0.22)0.5TiO3
tric materials, BNT-based ceramics are of particular interest (BNKT22) ceramics. In particular, we evaluated the influence
because of their high electric-field-induced strain6–8 and of the amount of CuO (0.5–3.0 mol%), sintering temperature
blocking force,9,10 which play important roles in piezoelectric (900°C–1000°C), and sintering time (2–6 h). In addition, we
actuator applications. evaluated the performance of acoustic sound speaker devices
Even though many studies have reported the physical containing MLCAs made from CuO-added BNKT22.
properties of BNT-based lead-free ceramics, only a few
reports are available about their practical applications.11–13 II. Experimental Procedures
Multilayer ceramic devices such as multilayer ceramics actua-
tors (MLCAs) and multilayer ceramics capacitors are widely
used in industry.14,15 In this study, we developed a MLCA
using BNT-based materials. When developing a MLCA, the (1) Low-Temperature Sintering of
difference in thermal expansion coefficients and reaction Bi0.5(Na0.78K0.22)0.5TiO3 Ceramics
between the internal electrode and active material should be Bi0.5(Na0.78K0.22)0.5TiO3 (BNKT22) powder was prepared
considered because the MLCA is cofired with the internal using a conventional solid-state reaction method. Bi2O3,
electrode. Differences in thermal expansion coefficients or Na2CO3, TiO2, and K2CO3 (99.9% High Purity Chemicals,
reactions between the internal electrode and active material Saitama, Japan) were the starting raw materials. Before
can cause significant problems in the MLCA, such as internal weighing, the raw powders were dried in an oven at 80°C for
cracks in the active ceramics layers or microdefects between 24 h to remove any moisture. Starting materials were
weighed according to the stoichiometric ratio and then ball-
milled for 24 h in ethanol with zirconia balls. Upon drying
at 80°C, the powders were ground and calcined at 850°C for
J. Jones—contributing editor
2 h in a covered alumina crucible. Powders were ground and
ball-milled again using the same conditions as described for
the initial milling step to prepare fine and uniform particles.
Manuscript No. 35978. Received November 26, 2014; approved February 18, 2015. Then, an excess amount of CuO at 0.5, 1.0, 2.0, or 3.0 mol%

Author to whom correspondence should be addressed. e-mail: kimiw@mail.ulsan.ac.kr was introduced to the calcined pure BNKT22 powder, and
1877
1878 Journal of the American Ceramic Society—Ahn et al. Vol. 98, No. 6

the mixed powders were ball-milled again for 24 h, then trical properties. Silver paste was applied through screen
dried at 100°C for 24 h. The obtained powders were mixed printing of both surfaces of the disks as electrodes. After
with an aqueous polyvinyl alcohol solution, and made into applying the silver paste, disks were fired at 700°C for
green disks with diameters of 10 mm at a pressure of 30 min. Ferroelectric hysteresis loops were measured in sili-
100 MPa. After binder burnout at 500°C, the disks were sin- con oil with the aid of a Sawyer–Tower circuit to apply an
tered at sintering temperatures ranging from 900°C to electric field with a sinusoidal waveform. Electric-field-
1000°C and sintering times ranging from 2 to 6 h in closed induced strain was measured with a linear variable differen-
alumina crucibles. tial transducer (Mitutoyo MCH-331 & M401, Kawasaki,
Purity and formation of the desired phase in sintered Japan). The piezoelectric charge coefficient d33 was mea-
samples were monitored by powder X-ray diffraction sured using a piezo d33 meter (ZJ-3A; Institute of Acous-
(XRD; X’ Pert-PRO, PANalytical, Almelo, the Nether- tics, Chinese Academy of Sciences, Beijing, China).
lands) using CuKa1 radiation. For microstructural analysis,
the surface of the as-sintered samples was removed by lap-
ping. Lapped samples were then mirror-polished and ther- (2) Fabrication of Multilayer Ceramic Actuators
mally etched at 900°C–1000°C for 10 min depending on Slurries for tape-casting were prepared by mixing the cal-
their composition. Finally, Field-emission scanning electron cined powders, solvent, binder, and plasticizer. As inner elec-
microscopy (FE-SEM; S-4200, Hitachi, Tokyo, Japan) was trodes, silver-palladium paste (SJAP-12-510X; Sung Jee
used to examine the microstructure of the polished and Tech. Co., Suwon, Korea) was screen-printed on the sheet.
thermally etched samples. Sintered disk-type samples were Ag and Pd contents were 90% and 10%, respectively. CuO
polished down to a thickness of 1 mm to assess their elec- 1.0 mol%-added BNKT22 ceramic green sheets were pre-
pared by tape-casting a slurry containing CuO 1.0 mol%-
added BNKT22 powder, ethyl alcohol (94% purity, ethanol;
SK Chemical, Sungnam, Korea), methyl ethyl ketone
(99.5% purity, MEK; Daejung Chemical & Metals Co., Ltd)
as a solvent, polyvinyl butyral (BM-SZ; Sekisui Chemical
Co., Ltd, Tokyo, Japan) as a binder, dibutyl phthalate
(94%; Daejung Chemical & Metals Co., Ltd) as a plasti-
cizer, and BYK-112 (BYK Chemical, Wesel, Germany) as a
dispersant.
The slurry was tape-casted using a doctor blade to a thick-
ness of about 40 lm and then dried at room temperature.
Ag–Pd paste was screen-printed on the ceramic sheet as the
inner electrode. Ceramic sheets with thick film electrodes
were stacked and then pressed at 50°C under 50 MPa pres-
sure for 5 s to enhance adhesion between layers. Each sheet
was laminated and cut to form a green compact. The final
Fig. 1. X-ray diffraction patterns of BNKT22 ceramics sintered at dimensions of the green compact were about
1150°C and BNKT22-CuO 1 mol% ceramics sintered at 900°C, 25.0 mm 9 20.0 mm 9 0.25 mm. Six layers of sheets were
950°C, and 1000°C for 4 h. laminated and internal electrodes were connected to the

(a) (b)

(c) (d)

Fig. 2. FE-SEM images of (a) pure BNKT22 ceramics sintered at 1150°C and BNKT22:CuO 1.0 mol% ceramics sintered at (b) 1000°C, (c)
950°C, and (d) 900°C for 4 h. All sample surfaces were thermally etched at each sintering temperature for 10 min after polishing.
June 2015 Bi0.5(Na,K)0.5TiO3 MLCA 1879

(a) (b) (c)

(d) (e) (f)

Fig. 3. EDX mapping results of BNKT22:CuO 1.0 mol% ceramics sintered at 950°C for 4 h: (a) Surface morphology (b) Cu-mapping, (c) Bi-
mapping, (d) Na-mapping, (e) K-mapping, (f) Ti-mapping.

(a) (b)

(c) (d)

Fig. 4. (a) Piezoelectric coefficient (d33) and normalized strain (Smax/Emax), (b) bipolar S–E loops, (c) unipolar S–E loops, and (d) P–E loops of
BNKT22:CuO 1.0 mol% ceramics samples as a function of sintering temperature. Pure BNKT22 sintered at 1150°C was used as the reference
sample.
1880 Journal of the American Ceramic Society—Ahn et al. Vol. 98, No. 6

external electrode to synthesize piezoelectric speakers. A III. Results and Discussion


schematic diagram of the piezoelectric speaker is presented in
fig. 10(b). Before sintering, the binder was burnt out in air X-ray diffraction patterns of BNKT22 and BNKT22:CuO
by heating the samples at 500°C for 10 h. Cofiring was car- 1 mol% ceramics in the 2h range 20°–80° are shown in
ried out in a covered alumina crucible at a temperature of Fig. 1. All samples exhibited a single-phase perovskite struc-
950°C with a soaking time of 4 h in air. ture. The XRD profiles suggest that no crystalline second
Sound pressure level (SPL) of the piezoelectric speakers phase is induced with the addition of CuO within the resolu-
was measured using an omnidirectional microphone (UMIK- tion limit of the apparatus used.
1; miniDSP Ltd., Hong Kong, China) and multichannel FE-SEM images of pure BNKT22 ceramics sintered at
dynamic signal analyzer (Electroacoustic toolbox; Faber 1150°C are shown in Fig. 2(a). Pure BNKT22 ceramics had
Acoustical, LLC, UT) with a 30 Vp-p input signal. The dis- a dense surface morphology with round grains without any
tance between the speaker and microphone was 10 cm. Sam- pores. BNKT22:CuO 1.0 mol% ceramics sintered at 1000°C
ples were prepared to conform to the KS CIEC 60268-5 and 950°C also had a dense surface morphology, but the
standard. number of pores was much greater in the specimen sintered

(a) (b)

(c) (d)

Fig. 5. Piezoelectric coefficient (d33) and normalized strain (Smax/Emax) of CuO-added BNKT22 ceramics as a function of CuO content at
sintering temperatures of (a) 925°C, (b) 950°C, (c) 975°C, and (d) 1000°C. All specimens were sintered for 4 h.

(a) (b)

Fig. 6. Piezoelectric coefficient (d33) and normalized strain (Smax/Emax) of BNKT22:CuO 1.0 mol% ceramics as a function of sintering time at a
sintering temperature of (a) 925°C and (b) 950°C.
June 2015 Bi0.5(Na,K)0.5TiO3 MLCA 1881

S–E loops and P–E loops of BNKT22:CuO 1.0 mol%


ceramics samples in the sintering temperature range 950°C–
1000°C showed similar behaviors to pure BNKT22 sintered
at 1150°C, as shown in Figs. 4(b)–(d). However, these prop-
erties deteriorated drastically at the lower sintering tempera-
ture of 900°C due to the low density of the specimen, as
shown in Fig. 2.
Changes in the d33 and Smax/Emax values of CuO-added
BNKT22 ceramics as a function of CuO content at various
sintering temperature are presented in Fig. 5. Values of d33
and Smax/Emax were stable in specimens within the experi-
mental CuO content range sintered at 925°C and 950°C. The
d33 and Smax/Emax values were about 175 pC/N and 270 pm/
V, respectively. However, the specimens sintered at 975°C
and 1000°C showed a drastic decline in d33 and the Smax/
Emax values at a CuO content above 2 mol%. These
decreases in d33 and Smax/Emax values at relatively high sin-
tering temperatures and high CuO content were likely caused
by the hardening effect of acceptor doping due to replace-
Fig. 7. Temperature-dependent dielectric constant and the ment of Ti4+ B-site ions with Cu2+.19
corresponding loss factor of BNKT22 and BNKT22:CuO 1.0 mol% Figure 6 shows d33 and Smax/Emax of BNKT22:CuO
ceramics. 1.0 mol% ceramics as a function of sintering time at a sinter-
ing temperature of 925°C and 950°C. While samples sintered
at 925°C had low d33 values after the relatively short sinter-
ing time of less than 4 h, as shown in Fig. 6(a), samples sin-
tered at 950°C had stable d33 and Smax/Emax values over
sintering times ranging from 2 to 6 h, as shown in Fig. 6(b).
In conclusion, the appropriate sintering temperature for
CuO-added BNKT22 ceramics is 950°C based on the results
shown in Figs. 5 and 6. And, the 1.0 mol% CuO-added sam-
ples show relatively reliable piezoelectric properties over wide
sintering temperature and time rang compare with other
CuO contents samples. Therefore, the appropriate fabrication
condition of MLCA specimens was decided as 1.0 mol%
CuO contents and sintering temperature of 950°C.
Temperature-dependent dielectric constant and the corre-
sponding loss factor for the poled materials were measured
from room temperature to 500°C. Profiles of the selected
compositions are presented in Fig. 7. Pure BNKT22 ceramics
exhibited two dielectric anomaly peaks at around 120°C and
Fig. 8. Temperature stability of the piezoelectric coefficient (d33) of 250°C within the measuring temperature range. The first
poled BNKT22 and BNKT22:CuO 1.0 mol% ceramics. inflection, which matched with the peak of dielectric loss, is
often designed as the depolarization temperature (Td) where
the ferroelectricity of the given system significantly decreases,
at 900°C than those sintered at the higher sintering tempera- whereas the latter inflection point, called Tm, is where the
tures. dielectric constant reaches its maximum. BNKT22:CuO
Successful incorporation of CuO into the BNKT22 ceram- 1.0 mol% ceramics exhibited dielectric anomaly peak related
ics was demonstrated by energy-dispersive X-ray spectros- with Td at 150°C which is 30°C higher than that of pure
copy (EDX). EDX mapping spectra of each element in BNKT22 ceramics. Furthermore, the peak of dielectric loss
BNKT22:CuO 1.0 mol% ceramics are shown in Fig. 3. of BNKT22:CuO 1.0 mol% ceramics at 150°C–220°C had a
Homogeneous distribution of Cu was observed [Fig. 3(b)]. It blunt shape. To more clearly observe the depolarization
has been reported that added CuO not only forms a liquid behavior, we measured the temperature stability of the piezo-
phase, but also diffuses into the lattice by replacing B-site electric coefficient using poled samples after thermal shock.
ions.19 This leads to homogeneous distribution of Cu ions. The temperature stability of piezoelectric properties (d33)
In general, the formation of a liquid phase of CuO leads to was measured in the temperature range 40°C–250°C with a
low-temperature sintering behavior and the Cu2+ ions that step of 10°C after thermal shock at each temperature for
replace Ti4+ B-site ions in BNT-based perovskite structures 10 min, as shown in Fig. 8. BNKT22 and BNKT22:CuO
act as acceptors.19 1.0 mol% ceramics had similar d33 values of 180 pC/N at
Piezoelectric coefficient (d33) and normalized strains room temperature. In the case of BNKT22 ceramics, the d33
(Smax/Emax) of pure BNKT22 ceramics sintered at 1150°C, value decreased suddenly after thermal shock at 130°C for
and BNKT22:CuO 1.0 mol.% ceramics sintered at various 10 min. However, the d33 value of BNKT22:CuO 1.0 mol%
temperatures ranging from 900°C to 1000°C with 25°C steps ceramics decreased slightly to 140 pC/N near 150°C. This d33
are shown in Fig. 4(a). Piezoelectric coefficient (d33) and value was maintained to 200°C. BNKT22:CuO 1.0 mol%
normalized strain (Smax/Emax) of pure BNKT22 sintered at ceramics had a relatively good temperature stability, due to
1150°C were 180 and 275 pm/V, respectively. These proper- the hardening effect of doping of Cu2+ ions at Ti4+ sites
ties were maintained with slight deterioration at the lower and induced ferroelectric ordering from existing Cu00Ti  V O
sintering temperature of 925°C in 1.0 mol% CuO-added defect dipoles.19
samples. The d33 and Smax/Emax of BNKT22:CuO 1.0 mol% The high-temperature stability is important for practical
sintered at 950°C were 175 pC/N and 270 pm/V, respec- application of MLCAs, because it provides a wide tempera-
tively. These values represent a degradation rate of about ture range for stable performance. Therefore, the addition of
3% compared with those of the pure BNKT22 reference CuO to BNKT-based ceramics not only decreased the sinter-
samples. ing temperature, but also increased the temperature stability
1882 Journal of the American Ceramic Society—Ahn et al. Vol. 98, No. 6

(a) (b)

Fig. 9. (a) Cross-sectional FE-SEM micrograph and (b) magnified micrograph images of the fracture surface of a multilayer ceramic actuator
(MLCA) fabricated using BNKT22:CuO 1.0 mol% powder and Ag–Pd (9:1) internal electrodes. Specimens were cofired at 950°C for 4 h in air.

(a) (b)

(c)

Fig. 10. (a) Sound pressure level (SPL) versus frequency curve of piezoelectric speakers prepared using (1) Pb-based MLCA (0.12 mm
thickness) in a 24 mm 9 30 mm frame, (2) BNKT22:CuO 1.0 mol% MLCA (0.2 mm thickness) in a 24 mm 9 30 mm frame, and (3) BNKT22:
CuO 1.0 mol% MLCA (0.2 mm thickness) in a 60 mm 9 70 mm frame with a PET diaphragm (250 lm thickness). (b) Schematic diagram of
the piezoelectric MLCA consisting of six active layers. (c) Schematic diagram of the piezoelectric speaker consisting of the piezoelectric MLCA
and PET diaphragm.

of the piezoelectric properties of the ceramics. These data 24 mm 9 30 mm metal frame. A lead-based piezoelectric
suggested that BNKT22:CuO 1.0 mol% is suitable for use in speaker was prepared for comparison. The SPL of lead-free
lead-free piezoelectric devices. To investigate the practical piezoelectric speakers at frequencies ranging from 100 and
applications of this material, we fabricated a prototype 20 000 Hz was measured, and the results are shown in
MLCA device. Fig. 10(a) along with those obtained using a lead-based pie-
To examine the thermal stability of the BNKT22:CuO zoelectric speaker that was prepared using a lead-based mate-
1.0 mol% and Ag–Pd (9:1) layers, cross-sectional microstruc- rial with a high d33 value of ~500 pC/N. The SPL of the
tures of MLCA specimens were observed by FE-SEM. speaker was calculated using the following equation:20
Cross-sectional microstructures of the fracture surface of a
MLCA specimen cofired at 950°C for 4 h in air are shown  
I
in Fig. 9. Gray areas in Fig. 9(a) correspond to BNKT22: SPL ¼ 10 log 10
CuO 1.0 mol% layers and bright lines between them indicate Iref
electrode layers. MLCA specimen shows dense BNKT22:
CuO 1.0 mol% layers with good thickness uniformity and a where I is the intensity of the sound and Iref is the intensity
thickness of about 35 lm. When producing MLCAs using of the reference.
general BNT-based materials, even if using an Ag–Pd (7: 3) The fundamental frequency (f0) of the lead-free piezoelec-
internal electrode, serious defects such as vertical cracks in tric speaker prepared with BNKT22:CuO 1.0 mol% ceramics
active layers and local disconnections in internal electrode was 810 Hz and the SPL value at f0 was about 67 dB. This
can occur because of the requirement for cofiring at a tem- speaker had a similar average SPL value to that of the lead-
perature of 1150°C, which is near the melting point of the based piezoelectric speaker in the frequency range from 2 to
electrode.18 However, in this study, no serious defects were 20 kHz. However, the f0 (810 Hz) of the BNKT22-based
evident. To more clearly observe the interface between speaker was higher than that of the lead-based speaker
BNKT22:CuO 1.0 mol% and internal electrode layers, (300 Hz). This difference in f0 was mainly due to the different
higher magnification micrographs were taken from the same thicknesses of the MLCA. Therefore, the weak SPL at low
specimens as shown in Fig. 9(b). The Ag–Pd (9:1) layer had frequencies of the BNKT22-based speaker can be modified
good thickness uniformity with a thickness of about 2 lm by changing the sample size and structure of the speaker. In
regardless of cofiring because of the low firing temperature this work, the BNKT22-based speaker was modified using a
of 950°C. PET diaphragm to improve SPL characteristics at low fre-
We next fabricated a piezoelectric speaker using a bender- quencies, as shown in Fig. 10(c), and the properties of this
type piezoelectric MLCA with six active layers of BNKT22: speaker are displayed in Fig. 10(a). The f0 of the BNKT22-
CuO 1.0 mol% and an Ag–Pd (9:1) internal electrode in a based speaker modified with a PET diaphragm was 660 Hz
June 2015 Bi0.5(Na,K)0.5TiO3 MLCA 1883
2
and the SPL value at f0 was about 76 dB. Moreover, the W. Jo, R. Dittmer, M. Acosta, J. Zang, C. Groh, E. Sapper, K. Wang, and J.
average SPL of this device was comparable with that of the R€odel, “Giant Electric-Field-Induced Strains in Lead-Free Ceramics for Actua-
tor Applications - Status and Perspective,” J. Electroceram., 29, 71–93 (2012).
lead-based speaker in the frequency range between 660 Hz 3
T. Takenaka and H. Nagata, “Current Status and Prospects of Lead-Free
and 20 kHz, which is part of the audible frequency range. Piezoelectric Ceramics,” J. Eur. Ceram. Soc., 25, 2693–700 (2005).
4
Therefore, the BNKT22:CuO 1.0 mol% piezoelectric MLCA J. R€
odel, W. Jo, K. T. P. Seifert, E. M. Anton, T. Granzow, and D. Dam-
speaker is a good candidate for lead-free audible sound janovic, “Perspective on the Development of Lead-Free Piezoceramics,” J.
Am. Ceram. Soc., 92, 1153–77 (2009).
devices. However, the SPL of the BNKT22:CuO 1.0 mol% 5
P. K. Panda, “Review: Environmental Friendly Lead-Free Piezoelectric
piezoelectric MLCA speaker was relatively low at frequencies Materials,” J. Mater. Sci., 44, 5049–62 (2009).
6
below 660 Hz. This low SPL value can be increased by A. Ullah, A. Ullah, I. W. Kim, D. S. Lee, S. J. Jeong, and C. W. Ahn,
reducing the thickness of the MLCA. “Large Electromechanical Response in Lead-Free La-Doped BNKT–BST Pie-
zoelectric Ceramics,” J. Am. Ceram. Soc., 97, 2471–8 (2014).
7
A. Ullah, R. A. Malik, A. Ullah, D. S. Lee, S. J. Jeong, J. S. Lee, I. W.
Kim, and C. W. Ahn, “Electric-Field-Induced Phase Transition and Large
IV. Conclusions Strain in Lead-Free Nb-Doped BNKT-BST Ceramics.,” J. Eur. Ceram. Soc.,
Low-temperature sintering properties of lead-free BNKT- 34, 29–35 (2014).
8
A. Ullah, C. W. Ahn, A. Ullah, and I. W. Kim, “Large Strain Under a
based piezoelectric ceramics with CuO as a sintering additive Low Electric Field in Lead-Free Bismuth-Based Piezoelectrics,” App. Phys.
were investigated. The optimal conditions for low-tempera- Lett., 103, 022906, 4pp (2013).
9
ture sintering were 1.0 mol% CuO-additive, a sintering tem- R. Dittmer, E. Aulbach, W. Jo, K. G. Webber, and J. R€ odel, “Large
perature of 950°C, and a sintering time of 4 h. Samples Blocking Force in Bi1/2Na1/2TiO3-Based Lead-Free Piezoceramics,” Scripta
Mater., 67, 100–3 (2012).
prepared under these conditions had similar piezoelectric 10
R. Dittmer, K. G. Webber, E. Aulbach, W. Jo, X. Tan, and J. R€ odel,
properties to those of pure BNKT22 ceramics sintered at “Electric-Field-Induced Polarization and Strain in 0.94(Bi1/2Na1/2)TiO3–
1150°C. Moreover, the addition of CuO to BNKT-based 0.06BaTiO3 Under Uniaxial Stress,” Acta Mater., 61, 1350–8 (2013).
11
ceramics not only decreased the sintering temperature, but S. H. Choy, X. P. Jiang, K. W. Kwok, and H. L. W. Chan, “Piezoelectric
and Dielectric Characteristics of Lead-Free BNKLBT Ceramic Thick Film
also increased the temperature stability of the piezoelectric and Multilayered Piezoelectric Actuators,” Ceram. Inter., 36, 2345–50 (2010).
properties due to induced ferroelectric ordering from existing 12
H. Nagata, Y. Hiruma, and T. Takenaka, “Electric-Field-Induced Strain
Cu00Ti  V
O defect dipoles.
for (Bi1/2Na1/2)TiO3-Based Lead-Free Multilayer Actuator,” J. Ceram. Soc.
A piezoelectric MLCA was fabricated using active layers Jpn., 118, 726–30 (2010).
13
W. Krauss, D. Sch€ utz, M. Naderer, D. Orosel, and K. Reichmann, “BNT-
of BNKT22:CuO 1.0 mol% and a Ag–Pd (9:1) internal Based Multilayer Device with Large and Temperature Independent Strain Made
electrode, and then the performance of a loudspeaker by a Water-Based Preparation Process,” J. Eur. Ceram. Soc., 31, 1857–60 (2011).
14
device with this MLCA was investigated. The average SPL K. Uchino and S. Takahashi, “Multilayer Ceramic Actuators,” Curr. Opin.
of this device was comparable to that of the lead-based Solid State Mater. Sci., 1, 698–705 (1996).
15
J. Pritchard, C. R. Bowen, and F. Lowrie, “Multilayer Actuators:
speaker in the frequency range between 660 Hz and 20 Review,” Brit. Ceram. Trans., 100, 1–9 (2001).
kHz. Therefore, BNKT22:CuO 1.0 mol% piezoelectric 16
S. Wang and W. Huebner, “Interaction of Ag/Pd Metallization with Lead
MLCA speakers are good candidates for lead-free audible and Bismuth Oxide-Based Fluxes in Multilayer Ceramic Capacitors,” J. Am.
sound devices. Ceram. Soc., 75, 2339–52 (1992).
17
S. Kuo, W. Tuan, Y. Lao, C. Wen, H. Chen, and H. Lee, “Investigation
into the Interactions Between Bi2O3-Doped ZnO and AgPd Electrode,” J. Eur.
Ceram. Soc., 28, 2557–63 (2008).
Acknowledgments 18
V. Nguyen, H. Han, H. Lee, J. Yoon, K. Ahn, and J. Lee, “Preparation
This work was supported by the development program of the local science and Properties of Bi-Based Lead-Free Ceramic Multilayer Actuators,” J.
park funded by the ULSAN Metropolitan City and the Ministry of Science, Ceram. Process. Res., 13, s282–5 (2012).
19
ICT and Future Planning (MSIP). W. Jo, J.-B. Ollagnier, J.-L. Park, E.-M. Anton, O.-J. Kwon, C. Park, H.-
H. Seo, J.-S. Lee, E. Erdem, R.-A. Eichel, and J. R€ odel, “CuO as a Sintering
Additive for (Bi1/2Na1/2)TiO3–BaTiO3–(K0.5Na0.5)NbO3 Lead-Free Piezoce-
ramics,” J. Eur. Ceram. Soc., 31, 2107–17 (2011).
References 20
T. Yanagisawa and N. Koike, “Cancellation of Both Phase Mismatch and
1
T. R. Shrout and S. T. Zhang, “Lead-Free Piezoelectric Ceramics: Alterna- Position Errors with Rotating Microphones in Sound Intensity Measurement,”
tives for PZT?” J. Electroceram., 19, 111–24 (2007). J. Sound Vib., 113, 117–26 (1987). h

View publication stats

You might also like