Anda di halaman 1dari 6

Mechanochemical Synthesis of Lead Zirconate Titanate from

Mixed Oxides
Junmin Xue, Dongmei Wan, See-Ee Lee,
*
and John Wang
*
,
Department of Materials Science, Faculty of Science, National University of Singapore, Singapore 119260
High-density lead zirconate titanate (PZT) ceramics have
been successfully prepared by using a novel mechano-
chemical fabrication technique, which skips the phase-
forming calcination at an intermediate temperature that is
always required in the industrial processes currently in use.
The fabrication technique starts with mixing of the low-cost
industrial oxide powders, and the designed PZT perovskite
phase is formed by reacting the oxide constituents in a
mechanochemical chamber that consists of a cylindrical
alumina vial and one stainless-steel ball inside it. The solid-
state reaction among constituent oxides is activated via me-
chanical energy instead of high temperature. When mecha-
nochemically activated for 20 h, an ultrafine PZT powder
of perovskite structure with a minimized degree of particle
agglomeration is obtained. The resulting PZT powder sin-
ters to 99.0% of theoretical density at 1100C for 1 h. The
sintered PZT ceramic exhibits a dielectric constant of 1340
and a dielectric loss of 0.6% at a frequency of 1 kHz at
room temperature.
I. Introduction
L
EAD ZIRCONATE TITANATE (PZT) is technologically impor-
tant in electronics and microelectronics, because of its ex-
cellent piezoelectric, ferroelectric, and many other electrical
properties.
1
High density and uniform microstructure are
among the most desirable features for PZT and PZT-based
electroceramics in obtaining many of these most-wanted elec-
trical properties. However, it is difficult to achieve a sintered
density close to the theoretical density via a conventional ce-
ramic processing route, using mixed oxides as the starting ma-
terials.
2
This problem, together with many of the undesirable
features of sintered PZT, such as the nonstoichiometry in com-
position and wide fluctuation in composition and poor micro-
structure, are due to the loss in lead content during the calci-
nation and sintering processes, as a result of the high volatility
of PbO at elevated temperatures.
3
Therefore, it is desirable to
process PZT and PZT-based ceramic materials at as low a
temperature as possible, to alleviate, if not completely elimi-
nate, these problems.
One approach is the use of ultrafine ceramic powders as the
starting materials, which may be synthesized via several chem-
istry-based novel preparation routes, such as the oxalate route,
4
coprecipitation,
5
hydrothermal reaction,
6
and alkoxide hydro-
lysis,
7
to name a few of the many methodologies. Nanosized
PZT powders of high chemical homogeneity and purity can be
made, in principle, by adjusting the processing parameters that
are involved in each of these powder-preparation techniques.
However, many of these wet-chemistry-based processing
routes require the precursor powders be calcined at a tempera-
ture in the range of 500900C, to develop the designed PZT
phase. This requirement often leads to the occurrence of par-
ticle agglomerates/aggregates in the resulting PZT powder. The
presence of hard particle agglomerates/aggregates will reduce
the sinterability of the powder compact significantly and lead
to the formation of microstructural defects in the sintered PZT
ceramic, as a result of the differential sintering between ag-
glomerates of differing green density.
8
The second notable ap-
proach in reducing the sintering temperature of PZT ceramics
is to dope them with a sintering additive that consists of an
oxide or a mixture of oxides with a low melting point. Unfor-
tunately, many of these sintering additives are detrimental to
the electrical properties of PZT when a nonferroelectric phase
is formed at the grain boundaries and grain junctions, although
the sintering temperature may be reduced.
9
A novel mechanochemical processing route has been devel-
oped recently for the production of intermetallic and alloy com-
pounds.
10,11
It was designed to synthesize various novel al-
loys
12
and nanocrystalline powders.
13,14
It also had been used
to synthesize ferrite-based magnetic materials.
15
The intrinsic
advantage of this technique is that the solid-state reaction is
activated via mechanical energy instead of temperature. More
recently, it was applied to prepare a few ceramic powders,
including oxides and non-oxides, such as ZrO
2
,
16
PbTiO
3
,
17,18
and BaTiO
3
,
19
although, in many cases, it was not very suc-
cessful. It was shown that the chemical reactivity of starting
materials could be improved significantly after an appropriate
mechanochemical treatment, and, subsequently, the calcination
temperature for forming the designed ceramic phase was re-
duced. Senna et al.
20
used a high-energy milling technique to
produce a 0.9PMN0.1PT ceramic powder (here, PMN denotes
lead magnesium niobate and PT denotes lead titanate). Only a
minor amount of perovskite phase was formed when a mixture
of PbO, TiO
2
, Nb
2
O
5
, and Mg(OH)
2
was milled for 60 min in
a specially designed multiring-type high-energy mill. On sub-
sequent heat treatment, the formation of perovskite phase in the
milled powder occurred at a sufficiently high calcination tem-
perature via solid-state reaction, with a pyrochlore phase being
involved. A pure perovskite PMNPT powder was obtained at
a temperature as high as 850C. It is apparent that the mecha-
nochemical treatment can enhance the reactivity of constituent
oxides and hydroxide; however, further heat treatment cannot
be avoided to develop the desired electroceramic phase. Thus
far, no one has shown that the multicomponent compound PZT
can be synthesized directly via a mechanochemical reaction
from mixed oxides without further thermal treatment. It was
even suggested that this feat could not be achieved via mecha-
nochemical action as a result of the high positive enthalpy of
formation that is involved.
21
In this work, a novel mechanochemical technique for syn-
thesizing fine PZT powders is described. The process starts
with low-cost oxide powders as the starting materials, and the
perovskite phase of PZT is formed at room temperature in a
mechanochemical cell, rather than via calcination at an el-
evated temperature. The resulting PZT powders are studied for
W. A. Schulzecontributing editor
Manuscript No. 190031. Received July 14, 1998; approved November 9, 1998.
*
Member, American Ceramic Society.

Author to whom correspondence should be addressed (E-mail address:


maswangj@nus.edu.sg).
J. Am. Ceram. Soc., 82 [7] 168792 (1999)
J
ournal
1687
powder characteristics, sintering behavior, and dielectric prop-
erties. The novel mechanochemical technique offers many ad-
vantages over both the conventional ceramic (solid-state reac-
tion) and wet-chemistry-based processing routes for PZT,
including (i) using low-cost and widely available oxides as the
starting materials, (ii) simplifying the processing route by skip-
ping the calcination(s) at an intermediate temperature, and (iii)
minimizing the loss in lead content as the formation of PZT is
completed at a temperature close to room temperature in an
enclosed chamber. More importantly, the mechanochemically
synthesized PZT powders exhibit a much-higher sinterability
than those of PZT powders that have been prepared via both the
conventional solid-state reaction and most of the chemistry-
based processing routes. The likely contamination that is in-
troduced by high-energy mechanical activation has been mini-
mized by selecting a wear-resistant milling vial and stainless-
steel ball. Therefore, contamination does not pose any threat to
the success of this novel processing technique, as has been
demonstrated by the excellent dielectric properties for PZT.
II. Experimental Procedure
The PZT composition selected for this study was
Pb(Zr
0.52
Ti
0.48
)O
3
, which is near the morphotropic phase
boundary. The starting materials were commercially available
PbO (>99% purity, J. T. Baker, Philipsburg, NJ) ZrO
2
(>99%
purity, MEL Limited, Manchester, U.K.), and TiO
2
(>99% pu-
rity, Merck, Darmstadt, FRG). These oxide powders exhibited
a particle-size distribution covering a range from submicrome-
ter to tens of micrometers. To make a 30 g batch of PZT
powder, 20.56 g of PbO, 5.91 g of ZrO
2
, and 3.53 g of TiO
2
were weighed out and placed in a laboratory plastic ball mill,
together with 30 g of ethanol and 60 g of zirconia balls (5.0 mm
in diameter) as milling media. They were then milled together
for 48 h, followed by drying of the slurry at 80C using a
heating lamp. The as-dried powder mixture was ground using
an agate mortar and pestle, to eliminate large powder lumps,
and then passed through 40-mesh sieves. Six grams of the
powder mixture was then loaded into a cylindrical alumina vial
40 mm in diameter and 40 mm in length, together with a
stainless-steel ball 12.7 mm in diameter.
22
Mechanochemical
synthesis was conducted in a shaker mill (Model 8000, Spex
Industries, Edison, NJ) that was operated at 900 rpm for 5, 10,
15, and 20 h. The level of contamination (mainly iron and
aluminum) in the resulting powders was initially monitored by
checking the weight loss for both the alumina vial and the
stainless-steel ball at regular intervals. Then, inductively
coupled plasma (ICP) spectroscopy (Model IRIS/AP, Thermo
Jarrell Ash, Franklin, MA) was used to measure the impurity
levels for iron and aluminum in these powders. They were
compacted uniaxially in a hardened steel die 10 mm in diam-
eter at a pressure of 50 MPa and then isostatically pressed at
350 MPa. Sintering of the isostatically pressed powder pellets
was performed in air at temperatures in the range of 950
1150C for a fixed duration of 1 h, with both heating and
cooling rates being fixed at 5C/min.
Phase development in the mechanochemically synthesized
powder, with increasing mechanochemical treatment time, was
monitored using X-ray diffractometry (XRD) (CuK radiation;
Model PW 1729 diffractometer, Philips Research Laboratories,
Eindhoven, The Netherlands). The particle sizes were esti-
mated on the basis of the BrunauerEmmettTeller (BET) spe-
cific surface area (Model Nova 2002, Quantachrome, Boynton
Beach, FL). Transmission electron microscopy (TEM) (Model
100CX, JEOL, Tokyo, Japan) was used to study the changes in
powder morphology and crystallinity with increasing mecha-
nochemical treatment time. Scanning electron microscopy
(SEM) (Model XL30, Philips Research Laboratories) was used
to characterize PZT ceramics that were sintered at various tem-
peratures. Sintered PZT ceramics were characterized for den-
sity by using the Archimedes method in deionized water in
which a few drops of wetting agents were added. Their dielec-
tric constant and dissipation factor were measured by using an
inductancecapacitanceresistance (LCR) meter (Model HP
4284A, HewlettPackard, Tokyo, Japan) at a frequency of 1
kHz.
III. Results and Discussion
Figure 1 shows the XRD traces of the powders that have
been mechanochemically treated for 5, 10, 15, and 20 h,
together with that of the starting powder mixture. As ex-
pected, before any mechanochemical treatment, the powder
mixture exhibits sharp peaks of crystalline PbO, ZrO
2
, and
TiO
2
, which indicates that no reaction was triggered during
the mixing/milling in the conventional ball mill. For the
powder that was mechanochemically treated for 5 h, almost
all the sharp peaks of PbO, ZrO
2
, and TiO
2
have vanished
and are replaced by a few broadened peaks. The strongest
peak, at 2 29.0, corresponds to the PbO (111) peak, which
is a sharp peak before the mechanochemical treatment. The
second-largest broadened peak, at 2 31.5, is newly formed,
which is assigned to the PZT perovskite phase. The signifi-
cant broadening in XRD peaks implies that there is a rapid
reduction in the particle size of constituent oxides and
some degree of amorphization has occurred via the mechano-
chemical activation. This phenomenon is directly reflected
by a dramatic increase in the specific surface area of the pow-
der that has been treated for 5 h, in comparison with that of
the starting powder mixture, as shown in Table I. On the basis
of the specific surface values of 1.32 and 12.5 m
2
/g, equiva-
Fig. 1. XRD patterns of the PZT composition mechanochemically
treated for various times, ranging from 0 to 20 h (() PbO, () TiO
2
,
() ZrO
2
, and () PZT).
1688 Journal of the American Ceramic SocietyJ. Xue et al. Vol. 82, No. 7
lent particle sizes of 560 and 60 nm are estimated for the
starting powder and that which was mechanochemically treated
for 5 h, respectively. When the mechanochemical reaction time
was increased to 10 h, the PZT phase peaks became more
pronounced at the expense of the PbO (111) peak. This obser-
vation implies that the conversion from mixed oxides to
perovskite PZT phase occurs step by step via mechanochemical
action. The powder that has been mechanochemically treated
for 15 h consists of perovskite PZT as the predominant phase,
although the PbO (111) peak is still visible. In the powder that
has been mechanochemically treated for 20 h, PZT is the only
XRD detectable phase. As shown in Table I, there is a dramatic
decrease in the average particle size when the powder mixture
was mechanochemically treated for the initial 5 h. Prolonged
mechanochemical treatment, up to 20 h, resulted in little further
refinement in the average particle size. Therefore, the perovsk-
ite PZT phase can be considered to form via nucleation in the
nanocrystalline/amorphous matrix of mixed oxides, followed
by crystalline growth, as a result of continued mechanical
activation.
Figures 24 are TEM micrographs, with the selected-area
electron diffraction (SAED) patterns, for the powders that have
been mechanochemically treated for 0, 5, and 20 h, respec-
tively. All three types of oxide particles were observed in the
ball-milled powder: PbO and TiO
2
particles are irregular in
morphology and are larger in size than the approximately
spherical ZrO
2
particles. The polycrystalline nature of the pow-
der mixture that was not subjected to any mechanochemical
treatment is indicated by the SAED pattern that is shown in
Fig. 2. In contrast, Fig. 3 shows that the powder that has been
mechanochemically treated for 5 h is a mixture of amorphous
and nanocrystalline phases, although particle agglomerates of
submicrometer size are visible. This observation indicates that
some degree of amorphization occurred when the powder mix-
ture was mechanochemically treated for 5 h. When treated for
20 h, a PZT powder that consisted of well-dispersed nanosized
particles 2040 nm in size was obtained. Apparently, the av-
erage crystallite size is smaller than the observed particle sizes,
as shown by the SAED pattern.
Mechanochemical reaction among mixed oxides is a com-
plicated process, and no well-established theories exist to ac-
count for many interesting observations.
23
However, from the
results of the XRD and TEM studies shown in Figs. 14, two
steps are believed to occur when PZT perovskite phase is
formed from the mixed oxides: (i) a dramatic refinement in
particle size and some degree of amorphization of mixed ox-
ides at the initial stage; and (ii) solid-state reaction in the ac-
tivated matrix of mixed oxides, which leads to nucleation and
growth of PZT crystallites. With initial increases in the mecha-
nochemical treatment time, the refinement of mixed oxides
down to a nanometer scale will significantly increase the num-
ber of point and lattice defects, the accumulation of which will
result in some degree of amorphization via mechanical activa-
Table I. Particle Characteristics of PZT Powders
Mechanochemically Treated for Various Times
Sample Specific surface area (m
2
/g) Equivalent particle size (nm)
Starting materials 1.32 560
Treated for 5 h 12.5 60
Treated for 10 h 13.3 56
Treated for 15 h 14.2 51
Treated for 20 h 12.8 59
Fig. 2. TEM micrograph and corresponding SAED diffraction pat-
tern, showing the powder mixture before mechanochemical treatment.
Fig. 3. TEM micrograph and corresponding SAED pattern, showing
the PZT powder mechanochemically treated for 5 h.
Fig. 4. TEM micrograph and corresponding SAED pattern, showing
the PZT powder mechanochemically treated for 20 h.
July 1999 Mechanochemical Synthesis of PZT from Mixed Oxides 1689
tion, although it is difficult to suggest how the solid-state re-
action proceeds. Localized heating at the points of collison may
be a contributing factor;
24
even the overall temperature of the
mechanochemical reaction vial does not exceed 70C. Appar-
ently, the in-situ impact temperature is much higher than
70C, which may be sufficient to trigger a solid-state reaction
on a nanometer scale, although this is not easily monitored.
Also, the effects of high-pressure impact at the collision points
may have an important role in facilitating the reaction.
25
It was
estimated that the pressure that is generated via a mechano-
chemical activation is 3.306.18 GPa.
26
In addition, any struc-
tural defects and disorder that are induced via mechanical ac-
tivation will favor diffusion and atomic rearrangements at
considerably low temperatures.
27
All these factors may be in-
volved in the formation of the perovskite PZT phase via the
mechanochemical treatment of mixed oxides of PbO, ZrO
2
,
and TiO
2
.
Figure 5 shows the XRD trace of a PZT pellet that has been
sintered at 950C for 1 h, derived from the powder that was
mechanochemically treated for 20 h. All the sharp peaks that
are observed can be assigned to the PZT perovskite phase,
which indicates that high levels of impurity phases are absent
from the sintered PZT ceramic. This phenomenon also was
observed when the mechanochemically synthesized powder
(20 h) was sintered at 1000, 1050, 1100, and 1150C for
1 h each.
Figure 6 shows the sintered density (relative density, or the
percentage of theoretical density) as a function of sintering
temperature for PZT that has been synthesized via mechano-
chemical treatment for 20 h. The sintered density increases as
the temperature increases from 950C to 1100C, where it
maximizes. Further increasing the sintering temperature to
>1100C results in a slight decrease in the sintered density,
presumably because of the loss of PbO, together with the oc-
currence of exaggerated grain growth at a too-high tempera-
ture. The sintering temperature (1100C) at which a maximum
sintered density of 99.0% of theoretical density is achieved is
considerably lower than those (in the range of 1300C) gener-
ally required by powders that are prepared via conventional
solid-state reactions through calcination at a high tempera-
ture.
28
At the same time, the sintered density of the mechano-
chemically synthesized PZT is higher than those achieved via
conventional solid-state reaction and many chemistry-based
processing routes.
Figures 7 (a), (b), (c), (d), and (e) are the fracture surfaces of
mechanochemically synthesized PZT ceramics that have been
sintered at 950, 1000, 1050, 1100, and 1150C, respec-
tively. Considerable necking and a reasonably high degree of
densification with fairly uniform grain sizes, in the range of
12 m, can be observed for the material that has been sintered
at 950C (Fig. 7(a)). However, it is rather porous, which is in
agreement with the measured density of 93% of theoretical
density. At 1000C (Fig. 7(b)), almost full density is observed
in many localized areas, and grain sizes are in the range of 23
m, although there are many porous areas. The average grain
size (68 m) and sintered density both are significantly in-
creased when the sintering temperature is increased to 1050C.
It was estimated that 30% of the grains exhibited a trans-
granular fracture surface. The PZT that was sintered at 1100C
demonstrates a very dense fracture surface with few small
pores being present, as shown in Fig. 7(d). Its average grain
size is in the range of 810 m, with many grains being frac-
tured in a transgranular manner. The material that was sintered
at 1150C shows a further enlargement in grain size (Fig. 7(e)),
which is believed to have contributed to the slight decrease in
sintered density, as shown in Fig. 6.
The dielectric constant and dielectric loss, as a function of
sintering temperature, for the mechanochemically derived PZT
are shown in Fig. 8. As expected, the dielectric constant in-
creases as the sintering temperature increases from 950C to
1100C, where it peaks at a value of 1340. A further increase
in the sintering temperature leads to a decrease in the dielectric
constant. In contrast, the dielectric loss decreases as the tem-
perature increases from 950C to 1100C, where it reaches a
minimum value of 0.6%. Further increasing the sintering tem-
perature to 1150C results in a slight increase in the dielectric
loss. On the one hand, these dependencies of dielectric prop-
erties on sintering temperature may easily be related to the
microstructural change with increasing temperature, such as
that in sintered density and grain size. On the other hand, the
excellent dielectric properties that are exhibited by these
mechanochemically synthesized PZTs demonstrate the fact that
the likely contamination that is introduced during the high-
energy mechanical activation process does not pose a threat to
the success of this novel processing route. This observation is
supported by an impurity level of <0.2% (mainly iron and
Al
2
O
3
), as measured via ICP for the PZT powder of perovskite
Fig. 5. XRD pattern for a sintered PZT ceramic at 950C for 1 h,
derived from the mechanochemically treated PZT powder (20 h).
Fig. 6. Relative density of mechanochemically synthesized PZT ce-
ramics as a function of sintering temperature (9501150C).
1690 Journal of the American Ceramic SocietyJ. Xue et al. Vol. 82, No. 7
structure that was mechanically activated for 20 h. A similar
impurity level was calculated for the powder on the basis of
weight loss measured for the alumina vial and the stainless-
steel ball after 20 h of high-energy mechanical activation.
IV. Conclusions
A phase-pure perovskite lead zirconate titanate (PZT) pow-
der of fine particles has been successfully prepared via a novel
mechanochemical synthesis from the low-cost, widely avail-
able oxides without calcining them at an intermediate tempera-
ture. When mechanochemically treated for 5 h, the particle size
of the oxide mixture was refined to a nanosize range, and some
degree of amorphization had occurred as a result of the me-
chanical activation. At the same time, nanosized perovskite
PZT crystallites were formed in the activated oxide matrix. The
constituent oxides completely disappeared, and perovskite PZT
was the only phase that was detectable via X-ray diffractometry
when the mechanochemical treatment was extended to 20 h.
The resulting PZT powder exhibited an excellent sinterability
and can be sintered to 99% of theoretical density at 1100C
for 1 h. The sintered PZT exhibited a dielectric constant of
1340 and a dielectric loss of 0.6% when measured at 1 kHz at
room temperature.
Fig. 7. SEM micrographs showing the fracture surfaces of PZT ceramics sintered at different temperatures ((a) 950, (b) 1000, (c) 1050, (d)
1100, and (e) 1150C).
July 1999 Mechanochemical Synthesis of PZT from Mixed Oxides 1691
References
1
D. Berlincourt, Current Development in Piezoelectric Applications of Fer-
roelectrics, Ferroelectrics, 10, 11118 (1976).
2
P. Gr. Lucuta, Fl. Constantinescu, and D. Barb, Structural Dependence on
Sintering Temperature of Lead Zirconate-Titanate Solid Solutions, J. Am. Ce-
ram. Soc., 68 [10] 53337 (1985).
3
Y. Matsuo and H. Sasaki, Formation of Lead Zirconate-Lead Titanate Solid
Solutions, J. Am. Ceram. Soc., 48 [6] 28991 (1965).
4
E. R. Leite, M. Cerqueira, L. A. Perazoli, R. S. Nasar, E. Longo, and J. A.
Varela, Mechanism of Phase Formation in Pb(Zr
x
Ti
1-x
)O
3
Synthesized by a
Partial Oxalate Method, J. Am. Ceram. Soc., 79 [6] 156368 (1996).
5
J. H. Choy, Y. S. Han, and J. T. Kim, Hydroxide Coprecipitation Route to
the Piezoelectric Oxide Pb(Zr,Ti)O
3
(PZT), J. Mater. Chem., 5, 6570 (1995).
6
H. Cheng, J. Ma, B. Zhu, and Y. Cui, Reaction Mechanisms in the For-
mation of Lead Zirconate Solid Solutions under Hydrothermal Conditions, J.
Am. Ceram. Soc., 76 [3] 62529 (1993).
7
R. C. Buchanan and J. Bey, Effect of Powder Characteristics on Micro-
structure and Properties in Alkoxide-Prepared PZT Ceramics, J. Electrochem.
Soc.: Solid State Sci. Technol., 132, 167175 (1985).
8
F. F. Lang, Sinterability of Agglomerated Powders, J. Am. Ceram. Soc.,
67 [2] 8390 (1984).
9
R. P. Tandon, V. Singh, and R. Singh, Properties of Low Temperature
Sintered Neodynium Doped Lead Zirconate Titanate Ceramics, J. Mater. Sci.
Lett., 13, 81013 (1994).
10
J. S. Benjamin, Mechanical Alloying, Sci. Am., 234, 4048 (1976).
11
P. S. Gilman and J. S. Benjamin, Mechanical Alloying, Annu. Rev.
Mater. Sci., 13, 279300 (1983).
12
E. Gaffet, N. Malhouroux, and M. Abdellaoui, Far from Equilibrium
Phase Transition Induced by Solid-State Reaction in the Fe-Si System, J.
Alloys Compd., 194, 33960 (1993).
13
A. K. Giri, Nanocrystalline Materials Prepared Through Crystallization by
Ball Milling, Adv. Mater., 9, 16366 (1997).
14
J. Ding, T. Tsuzuki, and P. G. McCormick, Ultrafine Alumina Particles
Prepared by Mechanochemical/Thermal Processing, J. Am. Ceram. Soc., 79
[11] 295658 (1996).
15
C. Jovalekic, M. Zdujic, A. Radakovic, and M. Mitric, Mechanochemical
Synthesis of NiFe
2
O
4
Ferrite, Mater. Lett., 24, 36568 (1995).
16
D. Michel, F. Faudot, E. Gaffet, and L. Mazerolles, Stabilized Zirconias
Prepared by Mechanical Alloying, J. Am. Ceram. Soc., 76 [11] 288488
(1993).
17
S. Komatsubara, T. Isobe, and M. Senna, Effect of Preliminary Mechani-
cal Treatment on the Microhomogenization during Heating of Hydrous Gels as
Precursors for Lead Titanate, J. Am. Ceram. Soc., 77 [1] 27882 (1994).
18
K. Hamada and M. Senna, Mechanochemical Effects on the Properties of
Starting Mixtures for PbTiO
3
Ceramics by Using a Novel Grinding Equipment,
J. Mater. Sci., 31, 172528 (1996).
19
O. Abe and Y. Suzuki, Mechanochemically Assisted Preparation of Ba-
TiO
3
Powder, Mater. Sci. Forum, 225, 56368 (1996).
20
J.-G. Baek, T. Isobe, and M. Senna, Synthesis of Pyrochlore-Free
0.9Pb(Mg
1/3
Nb
2/3
)O
3
0.1PbTiO
3
Ceramics via a Soft Mechanochemical
Route, J. Am. Ceram. Soc., 80 [4] 97381 (1997).
21
L. L. Shaw, Z. Yang, and R. Ren, Mechanically Enhanced Reactivity of
Silicon for the Formation of Silicon Nitride Composites, J. Am. Ceram. Soc.,
81 [3] 76064 (1998).
22
S. E. Lee, Preparation and Characterizatioin of Lead Zirconate Titanate
via Novel Chemistry Routes; M.Sc. Thesis. Department of Materials Science,
National University of Singapore, 1998.
23
V. V. Boldyev, Mechanochemistry and Mechanical Activation, Mater.
Sci. Forum, 225, 51120 (1996).
24
L. Takacs, Multiple Combustion Induced by Ball Milling, Appl. Phys.
Lett., 69, 43638 (1996).
25
N. N. Thadhani, Shock-Induced and Shock-Assisted Solid-State Chemical
Reactions in Powder Mixtures, J. Appl. Phys., 76, 212938 (1994).
26
T. G. Shen, C. C. Koch, T. L. McCormick, R. J. Nemanich, J. Y. Huang,
and J. G. Huang, The Structure and Properties Characteristics of Amorphous/
Nanocrystalline Silicon Produced by Ball Milling, J. Mater. Res., 10, 13948
(1995).
27
M. L. Trudeau, R. Schulz, D. Dussault, and A. V. Neste, Structural
Changes during High-Energy Ball Milling of Iron-Based Amorphous Alloys: Is
High-Energy Ball Milling Equivalent to a Thermal Process?, Phys. Rev. Lett.,
64, 99102 (1990).
28
A. I. Kingon and J. B. Clark, Sintering of PZT Ceramics: I, Atmosphere
Control, J. Am. Ceram. Soc., 66 [4] 25360 (1983).
Fig. 8. Dielectric properties of PZT ceramics sintered at different
temperatures derived from the mechanochemically synthesized PZT
powder (20 h).
1692 Journal of the American Ceramic SocietyJ. Xue et al. Vol. 82, No. 7

Anda mungkin juga menyukai