Vous êtes sur la page 1sur 8

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/215468341

SolGel Preparation of Lead Zirconate Titanate


Powders and Ceramics: Effect of Alkoxide
Stabilizers and Lead Precursors
ARTICLE in JOURNAL OF THE AMERICAN CERAMIC SOCIETY JANUARY 2000
Impact Factor: 2.61

CITATIONS

READS

29

463

4 AUTHORS, INCLUDING:
Paula Vilarinho
University of Aveiro
311 PUBLICATIONS 2,960 CITATIONS
SEE PROFILE

Available from: Paula Vilarinho


Retrieved on: 08 November 2015

journal

J. Am. Ceram. Soc., 83 [6] 1379 85 (2000)

SolGel Preparation of Lead Zirconate Titanate Powders and Ceramics:


Effect of Alkoxide Stabilizers and Lead Precursors
Aiying Wu, Paula M. Vilarinho, Isabel M. Miranda Salvado, and Joao L. Baptista*
Department of Ceramic and Glass Engineering, UIMC, University of Aveiro, 3810 Aveiro, Portugal
tetra-n-propoxide, and titanium tetra-isopropoxide.2,1115 Methoxyethanol (MOE) has been widely used as a solvent and stabilizer
for preparing PZT precursors, because of its chelating properties and
low viscosity.2,11,14 17 In more recent works, this toxic product has
been substituted by less hazardous solvents, such as propanediol,
acetic acid, acetone, and ethanolamine, also effective for preparing
PZT.3,4,12,13,18 20
For all of these PZT precursors, crystallization of the perovskite
phase occurs at temperatures ranging from 250 to 760C, depending on the zirconium content. As the zirconium content increases,
the perovskite crystallization temperature and the temperature at
which a single perovskite phase can be obtained increase. At
calcination temperatures of !600C, a homogeneous and stoichiometric PZT powder is obtained, whereas at 700C, a slight lead
deficiency occurs, adversely affecting the piezoelectric properties.7,21 To compensate for the lead loss, excess lead has been
widely used.6,15 However, excess PbO must be avoided in perovskite materials; the PbO excess either precipitates into the grain
boundary,21,22 damaging the electrical properties, or shifts the PZT
composition toward the titanium-lean side, because TiO2 has
higher solubility than ZrO2 in the PbO liquid phase.
The above-mentioned metal alkoxides used in synthesizing PZT
are very unstable, because of the high electropositive nature of the
metal atoms. Many works report that a controlled atmosphere (dry
argon or nitrogen gas) must be used to synthesize PZT.4,7,8
However, for industrial applications, ease of operation also should
be considered. Using some alkoxide stabilizers, which reduce the
reactivity of the metal alkoxides, enables synthesis in air.
For the present work, a series of perovskite PZT (52/48)
powders was prepared in air by the sol gel method, using a
non-MOE route and a stoichiometric starting composition. No
excess lead was used in the precursor. Because the choice of
precursors can affect the chemical-reaction kinetics, microstructures, and properties of the product, this paper compares the
crystallization behavior of PZT powders derived from different
precursors, stressing the influence of the alkoxide stabilizers and
the starting lead precursors on the phase assemblage of the
obtained products.
Detailed thermal analysis and X-ray diffractometry (XRD) were
conducted, and the relation between the preparation conditions and
the phase-forming process is described here. Dense ceramics were
prepared from the obtained submicrometer-sized PZT powders,
and the electrical properties of the sintered PZT ceramics were
measured and reported. The conditions required to obtain reasonably dense ceramics were experimentally set, and a preliminary
electrical characterization of these ceramics was undertaken. A
more detailed study of the nucleation and sintering process is
underway and will be reported in a later paper.

Lead zirconate titanate (PZT) (52/48) powders were prepared


by a sol gel process, using different raw materials to introduce
the lead component together with several solvents and chemical modifying agents. A study of the effect of these variables
on crystallization behavior was conducted to determine the
best conditions for preparing monophasic submicrometersized PZT perovskite powders in the morphotropic region. In
the present work, well-crystallized, submicrometer-sized
single-phase perovskite PZT powders were obtained after heat
treatment at 600C for 1 h. The dependence of this crystallization temperature on the preparation conditions was observed. The sol gel-derived submicrometer-sized PZT powders were sintered to !96% of relative density after 2 h at
950970C. The sintered ceramics exhibited a dielectric permittivity of 1000, a piezoelectric coefficient of 135 pC!N!1, a
remanent polarization of 20 "C!cm!2, and a coercive field of
10.6 kV!cm!1.
I. Introduction

(Pb(ZrxTi1x)O3 or PZT), with a


perovskite structure, is a piezoelectric ceramic widely used in
transducers, ferroelectric memories, sonars, optical filters, shutters,
actuators, and modulators.1
Various technological applications of PZT require a low sintering temperature for the ceramic. The typical sintering temperature
of PZT bulk ceramics is !1200C. In multilayer bulk stack
devices or integrated-memory thin-film capacitors, the different
layers and the substrate are cofired in one single step. Therefore,
low firing temperatures are desirable, to avoid interdiffusion
between layers and also to allow the use of inexpensive non-noblemetal-containing electrodes.
In contrast to the traditional mixed-oxide solid-state reaction
methods, wet chemical preparation allows mixing of the components at a molecular level, resulting in materials with high
compositional homogeneity and lower sintering temperatures.
Considerable work has been conducted on the chemical preparation of PZT powders and films, with different degrees of
success.29 A very recent report10 indicated that monophasic PZT
powders can be obtained at a nominal temperature of 450C from
a triethanolamine solution containing Pb2", Zr4", and Ti4" ions.
The obtained result is attributed to a high in situ temperature
caused by exothermic decomposition of the carbonaceous triethanolamine residue.
The alkoxide-based sol gel process is one of the most promising methods for synthesizing ceramic powders and, especially, films
at relatively low temperatures. The most commonly used starting
chemicals for PZT synthesis are lead acetate trihydrate, zirconium
EAD ZIRCONATE TITANATE

II.

Experimental Procedure

To study the effects of the starting chemicals, solvents, and


stabilizers on the sol gel formation of PZT (52/48) powders,
different reagents were used. Table I shows the starting chemicals,
together with solvents and stabilizers, used in the present work.
Acetic acid (HAc) and 1,2-propanediol were used to dissolve the
lead acetate. Acetylacetone (HAcAc) was used as a stabilizer for

W. A. Schulzecontributing editor

Manuscript No. 189316. Received June 1, 1999; approved November 11, 1999.
*Member, American Ceramic Society.

1379

1380

Journal of the American Ceramic SocietyWu et al.

Vol. 83, No. 6

Table I. Chemicals Used for the Preparation of PZT (52/48) Powders


Sample
designation

Chemicals to introduce
lead and solvents

Chemicals to introduce
titanium and stabilizers

Chemicals to introduce
zirconium and stabilizers

Pb(Ac)2!3H2O
1,2-Propanediol, HAc

Ti(i-Pr)4
HAcAc

Zr(n-Pr)4
HAc

Pb(Ac)2!3H2O
1,2-Propanediol, HAc

Ti(i-Pr)4
HAcAc

Zr(n-Pr)4
HAcAc

Pb(Ac)2!3H2O
1,2-Propanediol, HAc

Ti(i-Pr)4
HAcAc

Zr(n-Pr)4
HNO3

Pb(Ac)2!3H2O
1,2-Propanediol, HAc

Ti(i-Pr)4
HAcAc

Zr(n-Pr)4
1,2-Propanediol

CP1

Pb(NO3)2
Ethylene glycol

Ti(i-Pr)4
HAcAc

Zr(n-Pr)4
HNO3

CP2

Pb(NO3)2
Propanol, H2O

Ti(i-Pr)4
HAcAc

Zr(n-Pr)4
HNO3

CP3

Pb(NO3)2
H2O

Ti(i-Pr)4
HAcAc

Zr(n-Pr)4
HNO3

CP1, CP2, and CP3 are based on sample C but change the lead precursor and solvents.

the titanium alkoxide. Acetic acid, acetylacetone, nitric acid


(HNO3), and 1,2-propanediol were used as stabilizers for the
zirconium precursor compound. Ethylene glycol, 2-propanol, and
distilled water were used as solvents for the Pb(NO3)2. In each
sample, the ratio of Pb:Zr:Ti was always 1:0.52:0.48, and no
excess lead was added.
Figure 1 shows a flow chart of the synthesis process for the PZT
(52/48) powders. The obtained sols were maintained at 60C until
gelification occurred. The gels were dried at 120C for 24 h and
then crushed in an agate mortar and calcined in air between 400
and 800C for 1 h.
A Fourier transform infrared (FTIR) spectrometer with a resolution of 1 cm#1 (Model No. 7000, Mattson Technology, Inc.,
Fremont, CA) was used to study the structure of some precursors.
For FTIR analysis of the solutions, a small drop of the sol was
placed on a KBr disk, and the sample was analyzed after
evaporation. The powder samples were blended with KBr, pressed
into pellets, and analyzed.
Dried, crushed gels and calcined powders were analyzed by
XRD (Model No. Geigerflex D/Max-B, Rigaku Co., Ltd., Tokyo,

Japan). The enhanced resolution achieved through a longer counting time was used to detect the minor phases.23 Differential
thermal analysis (DTA) and thermogravimetric analysis (TGA)
(Model No. L18, Linseis International, Selb, Germany) were used
in conjunction with XRD to follow the decomposition and crystallization of the dried gels and powders. The average particle size
was measured using an instrument (Model No. ZetaSizer 4,
Malvern Instruments, Malvern, U.K.) equipped with a series 7032
Multi-8 Correlator and a scanning electron microscope (SEM;
Model No. S-4100, Hitachi, Co., Ltd., Tokyo, Japan) connected to
an image analysis system (Model No. Quantimet 500", Leica
Cambridge, Ltd., Cambridge, U.K.).
The calcined powders were isostatically pressed directly into a
disk shape at 200 MPa, without milling, and sintered at various
temperatures for 2 h in an oxygen atmosphere. To avoid lead losses
during the sintering step, a PbZrO3 powder was used to cover the
sample. The density of the ceramic samples was measured by the
Archimedes method. The dielectric properties were measured
using an impedance/gain-phase analyzer (Model No. 1260, Solartron Instruments, Farnborough, U.K.). Hysteresis curves were
obtained using a SawyerTower circuit at 60 Hz. The piezoelectric
coefficient (d33) was measured (Model No. CADT-3300 Piezo
d-meter, Berlincourt) at 60 Hz. Before the d33 value was measured,
the samples were poled with 4 kV/mm at 130C for 1 h and aged
for 24 h.
III.

Fig. 1.

Flow chart for the sol gel processing of PZT (52/48) powders.

Results and Discussion

(1) SolGel Formation


Transition-metal alkoxides are very reactive and experience
hydrolysis and condensation rapidly when water is added. Control
of the hydrolysis and condensation reaction is important in a
multicomponent system, such as PZT. Unequal hydrolysis and
condensation rates for each metal alkoxide result in phase separation during hydrolysis or thermal treatment. The preparation of
highly homogeneous gels, in which the various cations (lead,
zirconium, and titanium) are uniformly distributed at an atomic
scale through MOM$ bridges and that will have similar hydrolysis and condensation rates, requires chemical modification of the
starting alkoxides.
When titanium and zirconium alkoxides are added into a
solution to form a sol or a gel, precipitates often form. To provide
a stable and homogeneous sol, titanium and zirconium alkoxides
must be chemically modified.
(A) Stabilization of Titanium Alkoxide: HAcAc is effective
in stabilizing titanium alkoxides because the enolic form of
%-diketones contains a reactive hydroxy group that reacts readily
with metal alkoxides.24 In addition, the gelation behavior of

June 2000

SolGel Preparation of PZT Powders and Ceramics: Effect of Alkoxide Stabilizers and Lead Precursors

1381

Table II. Stabilizing Results for Zirconium Propoxide (Zr(n-Pr)4)


Stabilizers (molar ratio)

Zr(n-Pr)4
Zr(n-Pr)4
Zr(n-Pr)4
Zr(n-Pr)4

Fig. 2.

in
in
in
in

HAc (1:1)
HNO3 (1:1)
1,2-propanediol (1:11:5)
HAcAc (1:11:5)

Stirring for 30 min

Transparent solution
Transparent solution
Gelification
Turbid solution

IR spectra of Zr(n-Pr)4 modified with different stabilizers.

titanium alkoxides can be modified by HAc.24 In the present work


HAc, 1,2-propanediol, HAcAc, and 2-propanol were used as
stabilizers for titanium alkoxide. Clear and transparent sols were
obtained with all of these stabilizers, although HAcAc, suggested
to act as a chelating ligand with the formation of a stable titanium
compound,24 proved to be the most efficient.
(B) Stabilization of Zirconium Alkoxide: Although it has
been reported that HAcAc and HAc can also be used as stabilizers
for zirconium,19,24 in the present work, HNO3, as well as HAcAc,
1,2-propanediol, and HAc, were compared as stabilizers for
zirconium propoxide. The obtained results are listed in Table II,
which shows that HNO3 was the most efficient stabilizer for
zirconium, in agreement with previous results for the system
xZrO2!(100 x)SiO2.25 In the present work, a similar molar ratio,
HNO3/Zr & 1, was found sufficient to prepare homogeneous sols.
To obtain information on how the zirconium coordination
changes when different stabilizers are added, the infrared (IR)
spectra of Zr(n-Pr)4 modified with different stabilizers were
recorded; those results are shown in Fig. 2. The IR spectrum of
pure zirconium propoxide, without stabilizers, shows a band at
!1144 cm#1, attributed to the Zr(OC) stretching vibration of the
propoxy groups of zirconium propoxide.26
In the HAc-stabilized solution, the band at !1144 cm#1 is
much weaker. The HAc had partially substituted the alkoxy groups
of the zirconium alkoxide and formed acetate groups. According to
the literature,27,28 the set of two bands at !1500 cm#1 may result
from acetate ligands. The zirconium propoxide acetate was not
stable (Table II); the solution became more and more viscous and

Observations
24 h later

Transparent gel
Transparent solution
White precipitate
Yellow solution (1:1)
Precipitate (*1:1)

5 d later

Colorless gel
Transparent solution
White precipitate
Yellow gel

finally lost its fluidity, forming a clear gel after 24 h. Acetate


groups behaved, more likely, as bridging ligands (bonded to two
zirconiums) in the present work. The two IR bands at 1566 and
1450 cm#1 could result from 'COO-.28 This theory also could
explain the increase in viscosity with HAc-stabilized zirconium
solutions.
In a similar work, Yi and Sayer19 reported that the acetate
groups on zirconium propoxide acetate could probably be removed
by the alcohol created during the reaction, to form an ester. Based
on the present results, the band at !960 cm#1 can be assigned to
the '(CO) vibration of the propanol29 created in the reactions.
The formation of Zr-O-Zr links can cause gelation as time elapses.
In the spectrum of the 1,2-propanediol-stabilized zirconium
propoxide solution, the Zr(OC) stretching vibration band is still
present at 1144 cm#1, indicating that the alkoxy groups of
zirconium propoxide are not completely removed. In fact, the
spectrum is very similar to that of pure Zr(n-Pr)4. A small band at
720 cm#1 might be assigned to the Zr-O-Zr linkage formed,30 and
this result indicates that 1,2-propanediol is not effective in stabilizing this alkoxide, confirming the results reported in Table II,
although a dihydroxy alcohol (diol) has been mentioned to be
sufficiently reactive with metal alkoxides to form the corresponding glycolate or mixed alkoxide glycolate derivatives.26
In the spectrum of the HAcAc-stabilized zirconium propoxide
solution, the band at !1144 cm#1 still appears, and two bands at
!1600 and 1530 cm#1 are observed. These bands could indicate
that the alkoxy groups of zirconium propoxide Zr(OC) are only
partially removed and that a zirconium tripropoxide monoacetylacetonate chelate (bands at 1600 and 1530 cm#1) has formed.31
However, it has been observed (Table II) that precipitation occurs
after zirconium propoxide is mixed with HAcAc. The IR spectrum
of the isolated precipitates shows that the structure of zirconium
acetylacetonate is present, as indicated by comparison with an IR
spectrum obtained for pure zirconium acetylacetonate (ZrAcAc,
Merck, Darmstadt, Germany) (Fig. 3).Therefore, zirconium acetylacetonate may form in this condition:
Zr(OC3H7) 4 "xCH3COCH2COCH33
(CH3COCHCOCH3) x Zr(OC3H7) 4#x "xC3H7OH (1)
(CH3COCHCOCH3)xZr(OC3H7)4#x"(4#x)CH3COCH2COCH33
(CH3COCHCOCH3)4Zr"(4#x)C3H7OH (2)
Zirconium acetylacetonate has a poor solubility in the propanol
formed during the reaction (see band at !960 cm#1, Fig. 3), which
explains the observed precipitation.
In the spectrum obtained for the HNO3-stabilized solution, the
Zr(OC) stretching vibration band disappears, indicating that the
propoxy groups of zirconium propoxide are no longer bonded to
the zirconium cation in the same way. The NO2 vibration bands
(1697, 1311, and 660 cm#1), NO$ stretch (902 cm#1), and the
ONO$ bend band (597 cm#1) of free HNO332 are not observed.
Thus, the zirconium propoxide and HNO3 have reacted to form a
new complex ligand, which is rather stable. In fact, the solution
stabilized with HNO3 is still clear and fluid, with low viscosity
(0.009 Pa!s) after 5 d (Table II).
(2) Phase Development
(A) Effect of Stabilizers: The influence of the chemicals used
to modify the stability of the zirconium alkoxide on the phase

1382

Journal of the American Ceramic SocietyWu et al.

Vol. 83, No. 6

Fig. 3. IR spectra of the Zr(AcAc)4 powder and the precipitates isolated


from the turbid solution of Zr(n-Pr)4 with HAcAc.

development of the PZT powder also was studied. The crystallization behavior of the different gels was followed by thermal
analysis and XRD. The DTA and TGA results for the dried gels,
together with the XRD results, indicated the critical temperature
ranges for the removal of organic species and for crystallization of
the pyrochlore and perovskite phases.
In the temperature range 200 400C, all of the samples show
several exothermic peaks in the DTA curves (Fig. 4). The TGA
curves present marked weight losses in this temperature region.
The XRD results (Fig. 5) for all of the samples treated at
temperatures +400C evidence amorphous structures (except for
sample B and CP3, which will be discussed later). By combining
the DTA, TGA, and XRD results (Figs. 4 and 5), the DTA peaks
in this temperature range can be attributed to the decomposition of
most of the organic species. The different temperatures, intensities,
and shapes of the thermal peaks probably are related to the

Fig. 4.

Fig. 5. XRD patterns of dried gels (sample A) heat-treated at different


temperatures (other gels, except gel B and CP3, have similar XRD patterns
versus temperature).

different natures of the organic species and, consequently, caused


by the removal of species differently bounded in the network.
In the temperature range 400 600C, two exothermic peaks
are observed in the DTA curves of all of the samples (except for
samples B and CP3). In this region, the samples present a more
gradual weight loss. The XRD results reveal the coexistence of a
pyrochlore phase and a perovskite-type phase at 450C, for all of
the samples. When the calcination temperature is increased, the
pyrochlore phase is gradually eliminated, and the amount and
degree of crystallization of the perovskite phase increases. Thus,
the exothermic peak at !400 450C (415C for sample CP1,
426C for sample A, 440C for sample C, 447C for sample CP2,

(a) DTA and (b) TGA spectra of dried gels (heating rate of 5C/min).

June 2000

SolGel Preparation of PZT Powders and Ceramics: Effect of Alkoxide Stabilizers and Lead Precursors

Fig. 6. XRD patterns of gels derived from different stabilizers for


zirconium, calcined at 600C for 1 h.

and 434C for sample D) may be caused by crystallization of the


pyrochlore phase, and the last exothermic peak at 497570C
(497C for sample CP1, 515C for sample A, 548C for sample C,
563C for sample D, and 568C for sample CP2) may result from
perovskite phase crystallization, both probably accompanied by
the decomposition of organic residuals.
At temperatures *600C, according to DTA and TGA results,
the chemical reactions are complete. However, the complete
transformation of the pyrochlore to the perovskite phase seems to
differ for the differently prepared samples, depending on the
precursor chemicals used for each preparation, as evidenced by the
XRD results.
Figure 6 illustrates the powder phase assemblage obtained from
the gels prepared with different zirconium stabilizers after calcination at 600C for 1 h (samples A, B, C, and D). A perovskite
phase with less second phase was obtained when HNO3 was used
to stabilize the zirconium precursor.
Although HAcAc is very efficient for stabilizing titanium, it did
not work as well as HNO3 in stabilizing zirconium under the
present experimental conditions. The thermal behavior of sample
B was different from that of the other samples in the temperature
range 400 600C. The DTA spectra of sample B (Fig. 4(a)) show
only one peak in this temperature range, whereas the other sample
spectra show two peaks, suggesting different crystallization behavior for gel B. Indeed, for sample B, segregation and subsequent
multicrystallization occurred during the drying of the gel; the
phases ZrTiO4, ZrO2, TiO2, and Ti3.2Pb0.8 already were observed
in the dried gels. After calcination, a single-phase perovskite PZT
was difficult to obtain from this gel, even at temperatures as high
as 800C.
(B) Effect of Lead Precursor and Solvent Used to Dissolve
Pb(NO3)2: Two lead precursors, lead acetate and lead nitrate,
were used in the present work, and their effect on the final powder
phase assemblage was analyzed. The XRD pattern of these
samples calcined at 600C revealed a pure perovskite phase for
sample CP1 (within the detection limits of the equipment) and
small amounts of pyrochlore, together with the perovskite phase,
for sample C (Fig. 7).
In the DTA spectra of samples C and CP1 (Fig. 4(a)), the last
exothermic peak of the sample derived from Pb(Ac)2!3H2O (sample C) occurred at higher temperatures (548C) than that of the
sample derived from Pb(NO3)2 (sample CP1) (497C), which
seems to correlate with the ease of formation for the PZT
perovskite phase.
Dissolution of Pb(NO3)2 into different solvents also influenced
the phase-assemblage development. Figure 8 shows the XRD
patterns of PZT gels prepared using lead nitrate, with different
solvents, after calcination at 600C for 1 h. Samples CP2 and CP3,

1383

Fig. 7. XRD patterns of gels derived from different chemicals used to


introduce lead, calcined at 600C for 1 h.

in which 2-propanol plus water and water alone were used,


respectively, to dissolve the lead nitrate, did not form a pure
perovskite PZT, even after calcination for 1 h at 800C. In the
dried CP3 gel, in which water only was used as a solvent,
Pb(NO3)2 and PbTiO3 segregated out from the precursor gel.
However, when ethylene glycol was used as a solvent for lead
nitrate (sample CP1), the pure PZT perovskite phase was obtained
at 600C. The differences observed are probably attributable to the
polar nature of the solvents, because water and 2-propanol plus
water are much more polar than ethylene glycol. The formation of
(Zr(n-Pr)4)n (n ! 4) oligomers is favored in a nonpolar solvent,
and oligomeric species are less reactive than their monomers.33
Furthermore, the hydrolysis of a transition-metal alkoxide is an SN2
(nucleophilic substitution) reaction, and the polarity of the solvent
medium also has an effect on the reaction rate.33 The solvents used
for the lead precursor may, in this way, interfere with the
hydrolysis rate of the different components of this multicomponent
system, causing phase segregation when more polar solvents are
used.
(3) Characteristics of SolGel-Derived Ceramics
After calcination, the powders have similar morphologies, as
shown in Fig. 9. They consist of small agglomerates of fine
particles that are easily broken after dispersal in water by ultrasonic agitation. The individual particles have sizes of !30 60 nm,
and the agglomerates measure 200 300 nm (Fig. 9).

Fig. 8. XRD patterns of Pb(NO3)2-derived gels prepared with different


solvents, calcined at 600C for 1 h.

1384

Journal of the American Ceramic SocietyWu et al.

Fig. 9.

Vol. 83, No. 6

SEM photographs of PZT powders of sample CP1: (a) as-calcined powder and (b) after dispersal in water.

Table III. Dielectric and Piezoelectric Properties of PZT Ceramics Derived from Powder CP1
r

tan ,

d33 (pC!N#1)

1060
1130
1100
1080
1070
1050
1010

0.0001
0.0252
0.0124
0.0102
0.0125
0.0245
0.1518

135 (60 Hz)

20 (60 Hz)

730 (Ref. 34)

0.008 (Ref. 9)

121 (Ref. 9)

1728 (Ref. 35)

Frequency

1 Hz
10 Hz
100 Hz
1 kHz
10 kHz
100 kHz
1 MHz
Literature values

Pr (-C!cm#2)

Ps (-C!cm#2)

Ec (kV!cm#1)

38.3 (60 Hz)

10.58 (60 Hz)

814 (Ref. 35)

Sintered at 970C for 2 h.

The submicrometer-sized PZT powders derived from the CP1


gel calcined at 600C for 1 h were used to prepare PZT ceramics.
Dense ceramics were obtained (!96% of theoretical density) after
sintering for 2 h at temperatures between 950 and 970C. A
weight loss of !1.5% was observed.
The dielectric and piezoelectric properties of the samples
sintered at 970C for 2 h are shown in Table III. The dielectric
constants (r) of CP1-derived PZT ceramics in the frequency range
1 Hz1 MHz are !1000. The dielectric loss (tan ,) of these
ceramics is close to 0.01 in the frequency range 100 Hz1 kHz.
Compared with traditionally prepared PZT ceramics,34 the present
ceramic shows a higher dielectric constant, possibly related to its
dense microstructure and, because no excess lead oxide was used,
the absence of a lead-rich intergranular phase. Excess lead oxide
used to control the stoichiometry of a lead-based perovskite may
form an intergranular phase rich in lead, which damages the
dielectric permittivity of the ceramic.21,22 The present piezoelectric coefficient (d33) is slightly lower than that of conventionally
prepared PZT ceramics.34 The remanent polarization (Pr) and
coercive field (Ec), calculated from the hysteresis loop, are all in
agreement with literature values.35
IV.

Summary

The sol gel processing of PZT powders involves numerous


variables, namely the starting reagents and types of solvents and
stabilizers, which affect the hydrolysis and condensation reaction
of the metal alkoxides and, consequently, the final phase assemblage, morphology, microstructure, and properties of the material.
In the present study, we investigated the role of the starting
chemicals that introduce the lead component, the role of their
solvents, and the role of the zirconium precursor stabilizers on PZT
perovskite phase formation. The substitution of Pb(Ac)2!3H2O,

widely used in the preparation of PZT powders, by Pb(NO3)2,


together with the use of HNO3 as a stabilizer for the zirconium
alkoxide and ethylene glycol as a solvent for the lead nitrate,
allowed the formation of the pure perovskite PZT phase at lower
temperatures (!600C).
Although the pure perovskite PZT phase does not form directly
from the precursor solution, achievement of the desired phase at a
low calcination temperature seems to be related to the ability to
maintain, in such a multicomponent system, the highest possible
degree of homogeneity in solution, during the gelification process,
in order to avoid precipitation and dissociation from the atomic
mixture level attainable in solution.
The high reactivity of sol gel-derived submicrometer-sized
PZT powders enables them to lower the sintering temperature of
ceramics. The ceramics derived from these powders exhibit a high
dielectric constant, 1000 at room temperature, and piezoelectric
and ferroelectric characteristics similar to those mentioned in the
literature.
The simplicity of the process described here, together with the
low temperatures of phase formation and sintering, could eventually lead to an attractive method for the industrial fabrication of
PZT materials. A more detailed study of the nucleation and
sintering process is underway and will be reported in a later paper.
Acknowledgments
The authors thank the Portuguese FCT (Foundation for Science and Technology)
for financial support. A. W. thanks the Portuguese Program PRAXIS XXI for
Maintenance Grant No. Praxis XXI/BD/9007/96.

References
1

R. C. Buchanan, Ceramic Materials for Electronics, 2nd ed.; p.182. Marcel


Dekker, New York, 1991.

June 2000

SolGel Preparation of PZT Powders and Ceramics: Effect of Alkoxide Stabilizers and Lead Precursors

2
C. D. E. Lakeman and D. A. Payne, Processing Effects in the SolGel
Preparation of PZT-Derived Gels, Powders, and Ferroelectric Thin Layers, J. Am.
Ceram. Soc., 75 [11] 309196 (1992).
3
M. L. Calzada and S. J. Milne, Lead Zirconate Titanate Films from a Diol-Based
SolGel Method, J. Mater. Sci. Lett., 12, 122123 (1993).
4
Y. Takahashi, Y. Matsuoka, K. Yamaguchi, M. Matsuki, and K. Kobayashi, Dip
Coating of PT, PZ, and PZT Films Using an Alkoxide-Diethanolamine Method, J.
Mater. Sci., 25, 3960 64 (1990).
5
R. Zimmermann-Chopin and S. Auer, Spray Drying of SolGel Precursors for the
Manufacturing of PZT Powders, J. SolGel Sci. Technol., 3, 101107 (1994).
6
C. T. Lin, B. W. Scanlan, J. D. McNeill, J. S. Webb, L. Li, R. A. Lipeles, P. M.
Adams, and M. S. Leung, Crystallization Behavior in a Low-Temperature Acetate
Process for Perovskite PbTiO3, Pb(Zr,Ti)O3, and (Pb1x,Lax)(Zry,Ti1y)1x/4O3 bulk
Powders, J. Mater. Res., 7 [9] 2546 54 (1992).
7
B. Malic and M. Kosec, Electron Microscope Study of Alkoxide-Derived
Compositions within the PbZrO3PbTiO3 Phase Diagram, J. SolGel Sci. Technol.,
2, 443 46 (1994).
8
S. P. Faure, P. Barboux, P. Gaucher, and J. P. Ganne, Synthesis of Ferroelectric
Thin Films and Ceramics from Solution Processes, Ferroelectrics, 128, 19 24
(1992).
9
B. G. Muralidharan, A. Sengupta, G. S. Rao, and D. C. Agrawal, Powders of
Pb(ZrxTi1x)O3 by SolGel Coating of PbO, J. Mater. Sci., 30, 323137 (1995).
10
R. Das, A. Pathak, and P. Pramanik, Low-Temperature Preparation of Nanocrystalline Lead Zirconate Titanate and Lead Lanthanum Zirconate Titanate Powders
Using Triethanolamine, J. Am. Ceram. Soc., 81 [12] 3357 60 (1998).
11
J. S. Kim, D. S. Yoon, and K. No, Physical and Electrical Properties of
MnO2-Doped Pb(ZrxTi1x)O3 Ceramics, J. Mater. Sci., 29, 6599 603 (1994).
12
C. Livage, A. Safari, and L. C. Klein, Glycol-Based SolGel Process for the
Fabrication of Ferroelectric PZT Thin Films, J. SolGel Sci. Technol., 2, 605 609
(1994).
13
G. Yi, Z. Wu, and M. Sayer, Preparation of Pb(Zr,Ti)O3 Thin Films by SolGel
Processing: Electrical, Optical, and Electro-Optic Properties, J. Appl. Phys., 64 [5]
271724 (1988).
14
C. J. Kim, D. S. Yoon, J. S. Lee, C. G. Choi, W. J. Lee, and K. No, Electrical
Characteristics of (100), (111), and Randomly Aligned Lead Zirconate Titanate Thin
Films, J. Appl. Phys., 76 [11] 7478 82 (1994).
15
J. Chen, M. P. Harmer, and D. M. Smyth, Compositional Control of Ferroelectric Fatigue in Perovskite Ferroelectric Ceramics and Thin Films, J. Appl. Phys., 76
[9] 5394 98 (1994).
16
L. M. Reaney, K. Brooks, R. Klissuraka, C. Pawlaczyk, and N. Setter, Use of
Transmission Electron Microscopy for the Characterization of Rapid Thermally
Annealed, SolutionGel, Lead Zirconate Titanate Films, J. Am. Ceram. Soc., 77 [5]
1209 16 (1994).
17
V. K. Seth and W. A. Schulze, Fabrication and Characterization of Ferroelectric
PLZT 7/65/35 Ceramic Thin Films and Fibers, Ferroelectrics, 112, 283307 (1990).

1385

18
A. Wu, I. M. Miranda Salvado, P. M. Vilarinho, and J. L. Baptista, Lead
Zirconate Titanate Prepared from Different Zirconium and Titanium Precursors by
SolGel, J. Am. Ceram. Soc., 81 [10] 2640 44 (1998).
19
G. Yi and M. Sayer, An Acetic Acid/Water-Based SolGel PZT Process, I:
Modification of Zr and Ti Alkoxide with Acetic Acid, J. SolGel Sci. Technol., 6,
6574 (1996).
20
T. Fukui, C. Sakurai, and M. Okuyama, Preparation of Pb(ZrxTi1x)O3 powders
from Complex Alkoxides and Their Lower-Temperature Crystallization, J. Mater.
Res., 7 [4] 79194 (1992).
21
M. Villegas, C. Moure, J. R. Jurado, and P. Duran, Processing and Properties of
Pb(Mg1/3Nb2/3)O3PbZrO3PbTiO3 Ceramic Relaxors, J. Mater. Sci., 29, 4975 83
(1994).
22
K. Kakegawa, O. Matsunaga, T. Kato, and Y. Sasaki, Compositional Change
and Compositional Fluctuation in Pb(Zr,Ti)O3-Containing Excess PbO, J. Am.
Ceram. Soc., 78 [4] 107175 (1995).
23
J. A. T. Taylor, Microphase Detection Using XRD, Am. Ceram. Soc. Bull., 74
[6] 81 82 (1995).
24
C. Sanchez, J. Livage, M. Henry, and F. Babonneau, Chemical Modification of
Alkoxide Precursors, J. Non-Cryst. Solids, 100, 56 76 (1988).
25
I. M. Miranda Salvado, Preparation and Characterization of Materials in the
System of ZrO2SiO2, TiO2SiO2, and Al2O3SiO2. Application as Coatings; Ph.D.
Thesis. University of Aveiro, Portugal, Spain, 1990.
26
D. C. Bradley, R. C. Mehrotra and D. P. Gaur, Metal Alkoxides. Academic Press,
New York, 1978.
27
K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination
Compounds, 4th ed.; pp. 23139. Wiley, New York, 1986.
28
S. Doeuff, M. Henry, C. Sanchez, and J. Livage, Hydrolysis of Titanium
Alkoxides: Modification of the Molecular Precursor by Acetic Acid, J. Non-Cryst.
Solids, 89, 206 16 (1987).
29
C. J. Pouchert, The Aldrich Library of Infrared Spectra, 2nd ed. Aldrich
Chemical Company, Milwaukee, WI, 1975.
30
Y. Kanno and T. Suzuki, Synthesis of Fine Spherical ZrO2SiO2 Particles by
Ultrasonic Spray Pyrolysis, J. Mater. Sci. Lett., 7, 386 88 (1988).
31
U. B. Saxena, A. K. Rai, V. K. Mathur, R. C. Mehrotra, and D. Radford,
Reaction of Zirconium Isopropoxide with %-diketones and %-keto-Esters, J. Chem.
Soc. (A), 904 907 (1970).
32
W. A. Guillory and M. L. Bernstein, Infrared Spectral Frequencies for Matrix
Nitric Acid in an N2 Matrix, J. Chem. Phys., 62, 1059 (1975).
33
C. J. Brinker and G. W. Scherer, SolGel Science: The Physics and Chemistry of
SolGel Processing. Academic Press, San Diego, CA, 1990.
34
B. Jaffe, W. R. Cook Jr., and H. Jaffe, Piezoelectric Ceramics; pp. 145 48.
Academic Press, New York, 1971.
35
T. Yamamoto, Optimum Preparation Methods for Piezoelectric Ceramics and
Their Evaluation, Am. Ceram. Soc. Bull., 71 [6] 978 85 (1992).
"

Vous aimerez peut-être aussi