2007 Orientation Control in Sol-Gel-Derived BiScO3-PbTiO3 Thin Films

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

J. Am. Ceram. Soc.

, 90 [10] 3248–3254 (2007)


DOI: 10.1111/j.1551-2916.2007.01888.x
r 2007 The American Ceramic Society

Journal
Orientation Control in Sol–Gel-Derived BiScO3–PbTiO3 Thin Films
Hai Wen, Xiaohui Wang,*,w and Longtu Li
State Key Laboratory of New Ceramics and Fine Processing, Department of Materials Science and Engineering,
Tsinghua University, Beijing 100084, China

BiScO3–PbTiO3 (BSPT) thin films were fabricated via a sol–gel and Wu,20 which were consistent with Chen’s results. However,
method on Pt(111)/Ti/SiO2/Si(111) substrates. The effects of most of the current researches are focused on PZT thin films.
different factors on the orientation of the sol–gel-derived BSPT Few researches have been reported on other lead-based perovs-
thin films were investigated. The results showed that a higher kite ferroelectric thin films.
lead excess concentration, longer drying time, higher pyrolysis (1x)BiScO3–xPbTiO3 (BSPT) is a newly developed high-
temperature, longer pyrolysis time, higher crystallization tem- temperature piezoelectric solid solution system exhibiting a
perature, and longer crystallization time could enhance the (100) higher Curie temperature (TcB4501C) and comparable proper-
orientation of the BSPT thin films. Based on the experimental ties at the MPB (x 5 0.64) with respect to PZT,21–23 which has
results, a mechanism for the orientation evolution in the sol–gel- been considered as a potential candidate for the high-tempera-
derived BSPT thin films was proposed. The production of the ture microdevices required in automotive, aerospace, and relat-
(100) orientation was attributed to the (100)-oriented PbO ed industries. The BSPT thin films were first fabricated by
nanocrystals forming during the pyrolysis process due to the Trolier-Mckinstry24,25 on LaAlO3 single-crystal substrates
lattice match. through pulsed laser deposition. In our previous study,26,27 the
BSPT thin films were prepared on platinized silicon substrates
via a sol–gel method. In this study, the effects of different factors
I. Introduction on the orientation of the sol–gel-derived BSPT thin films have
been investigated. The growth mechanism of the film orientation

F ERROELECTRIC thin films with a perovskite structure, such as


lead zirconate titanate (PZT), have attracted considerable
attention as promising materials for applications in nonvolatile
has been discussed as well.

ferroelectric random access memories (FRAM) and micro-


electromechanical systems (MEMS) due to their high ferroelec- II. Experimental Procedure
tric and piezoelectric activities near the morphotropic phase
boundary (MPB).1–3 A variety of techniques have been used to Lead acetate [Pb(COOCH3)2], tetrabutyl titanate [Ti(OC4H9)4],
fabricate the ferroelectric thin films, among which the sol–gel scandium acetate [Sc(COOCH3)3], and bismuth nitrate
method has received particular interest because of its advantages [Bi(NO)3] (Hongxing Chemical Ltd., Beijing, China) were cho-
of chemical homogeneity, facility of stoichiometry control, low sen as the raw materials, with glacial acetic acid, isopropyl al-
cost, and easy integrability into semiconductor devices. In highly cohol, and distilled water as the solvents. The preparation
oriented or textured ferroelectric thin films, due to the enhance- details of the BSPT precursor solutions have been described
ment of anisotropy, some improved properties are exhibited elsewhere.26 Three solutions with different lead excess were pre-
compared with those thin films without preferential orientation. pared, the lead excess of which were 0% (hereinafter named
For example, in tetragonal PZT, a higher piezoelectric d33 can be SG-a), 10% (hereinafter named SG-b), and 18% (hereinafter
obtained in (001)-oriented thin films, while a larger dielectric named SG-c), respectively. The resultant solutions, with the
constant can be obtained in (100)-oriented thin films. Thus, composition of 0.34BiScO3–0.66PbTiO3 and a concentration of
many approaches have been used to fabricate oriented ferro- 0.25M determined by inductive coupled plasma emission spect-
electric thin films via the sol–gel method, such as introducing a rometry, were aged for 10 days before deposition. The thin films
seeding layer4–6 or a buffer layer,7–10 using single-crystal sub- were deposited onto Pt(111)/Ti/SiO2/Si(111) substrates by spin
strates11–13 and applying external stress.14,15 Moreover, some coating at 5000 rpm for 30 s, dried, and pyrolyzed in an ambient
studies have also shown that the preferential orientation of Pb- atmosphere. The process was repeated five times to achieve a
based perovskite films can be obtained on a common platinized desired thickness (B200 nm). The amorphous films were crys-
silicon substrate Pt(111)/Ti/SiO2/Si. Chen and colleagues16,17 tallized using a rapid thermal processing (RTP) furnace in an
reported the temperature–time texture (TTT) evolution of sol– ambient atmosphere. In order to investigate the effects of dif-
gel-derived PZT thin films on Pt(111)/Ti/SiO2/Si, in which both ferent factors on the orientation of the BSPT thin films, several
(100) and (111) orientations have been obtained. The develop- groups of experiments were carried out. The details of these
ment of (111) orientation was considered to arise from lead-rich experiments were listed in Table I.
intermediate phases PtxPb due to a lattice match, while (100) The microstructures of the BSPT thin films were observed
orientation was correlated with the formation of a lead oxide using field-emission scanning electron microscopy (FE-SEM,
(100) layer, which crystallized at a low temperature before the LEO-1530, LEO, Oberkochen, Germany) and atomic force mi-
perovskite PZT. The microstructure evolutions of the PZT thin croscopy (AFM, SPI4000&SPA300HV, Seiko, Tokyo, Japan).
films with different orientations were reported by Reaney18,19 The phase structures were characterized by X-ray diffraction
(XRD, D/max-RB, Rigaku, Tokyo, Japan) with CuKa radia-
tion. Transmission electron microscopy (TEM, Tecnai G2 F20,
S. Trolier-McKinstry—contributing editor
FEI, Hillsboro, OR) was used to conduct high-resolution trans-
mission electron microscopy (HRTEM) observation and ener-
gy-dispersive spectroscopy (EDS) analysis. Fourier transform
Manuscript No. 22277. Received September 21, 2006; approved June 2, 2007. infrared spectroscopy (FTIR, Spectrum GX, Perkin-Elmer,
Supported by National 973 Project of China (No. 2002CB613301) and National Natural Waltham, MA) was used to identify the residual organic spe-
and Science Foundation of China (No. 50472006).
*Member, American Ceramic Society. cies in the thin films, which was recorded by means of a reflec-
w
Author to whom correspondence should be addressed. e-mail: wxh@tsinghua.edu.cn tion technique.
3248
October 2007 Orientation Control in Sol–Gel-Derived BSPT Thin Films 3249

Table I. Experimental Details of the BiScO3–PbTiO3 Thin Films Prepared Via the Sol–Gel Method
Group Precursor Drying temperature Drying time Pyrolysis temperature Pyrolysis time Crystallization temperature Crystallization time
number solution (1C) (min) (1C) (s) (1C) (s)

I SG-b 80 60 350–500 200 700 500


II SG-b 80 60 450 60–300 700 500
III SG-b 80 60 450 300 550–800 300
IV SG-b 80 60 450 300 700 120–600
V SG-a–c 80 60 450 300 700 300
VI SG-c 80 2–60 450 300 650 300

P
III. Results and Discussion Iðl00Þ
p¼P (2)
(1) Microstructures and Phase Structures IðhklÞ
Typical SEM and AFM micrographs for the microstructure of
the sol–gel-derived BSPT thin films are shown in Fig. 1. As the where I is the maximum intensity of each peak and p0 5 p for the
figures show, the film exhibits a dense and uniform microstruc- randomly oriented BSPT powder.28 A larger Lotgering factor f
ture without any cracks. The grain size of the film is o50 nm. indicates a higher (100) orientation. In this study, the p0 value
The XRD patterns, shown in the figures, reveal that all the thin used in calculating the Lotgering factor f is measured from a
films are single-phase perovskite without any pyrochlore peaks self-made powder sample by the sol–gel method reported in our
being observed. Moreover, the strongest peaks are either (100) previous study.29
peaks or (110) peaks for all the samples, indicating that the sol– Figure 2 presents the Lotgering factor as a function of pyroly-
gel-derived BSPT thin films are either (100) oriented or ran- sis temperature for the sol–gel-derived BSPT thin films dried at
domly oriented. The intensities of the (111) peaks are very low 801C for 60 min, pyrolyzed at different temperatures for 200 s,
or even disappeared, for all the samples, indicating that the (111) and crystallized at 7001C for 500 s using the precursor solution
orientation has been strongly suppressed. SG-b. As the figure shows, when the pyrolysis temperature
increases from 3501 to 4001C, the Lotgering factor f does not
change much. However, when it comes to 4501C, the Lotgering
(2) Effects of Different Factors on the Orientation factor f exhibits a large increase from 0.23 to 0.41 and then a
To evaluate the (100) orientation degree of the sol–gel-derived slight increase to 0.43 for 5001C. This indicates that a higher
BSPT thin films, the Lotgering factor f is used, described as pyrolysis temperature may lead to a higher (100) orientation,
follows: especially when the pyrolysis temperature increases from 4001 to
4501C. The Lotgering factor as a function of pyrolysis time for
the sol–gel-derived BSPT thin films dried at 801C for 60 min,
pyrolyzed at 4501C for different times and crystallized at 7001C
f ¼ ðp  p0 Þ=ð1  p0 Þ (1) for 500 s using the precursor solution SG-b is displayed in Fig. 3.
As shown in the figure, f increases with an increase in pyrolysis
time from 0.29 for 60 s to 0.66 for 300 s, except a small decrease
at 120 s, which is considered an experimental error.
Secondly, the effect of the crystallization process on the
orientation of sol–gel-derived BSPT thin films is investigated.
Figure 4 shows Lotgering factor as a function of crystallization
temperature for the sol–gel-derived BSPT thin films dried at
801C for 60 min, pyrolyzed at 4501C for 300 s, and crystallized
at different temperatures for 300 s using the precursor solution
SG-b. As the figure reveals, the Lotgering factor f increases with
increasing crystallization temperature, from 0.47 for 5501C to

Fig. 1. Typical (a) cross-sectional scanning electron microscopy and (b)


atomic force microscopy micrographs for the microstructure of a sol– Fig. 2. Lotgering factor as a function of pyrolysis temperature for sam-
gel-derived BiScO3–PbTiO3thin film. ple group I. The inset shows the X-ray diffraction patterns of the films.
3250 Journal of the American Ceramic Society—Wen et al. Vol. 90, No. 10

Fig. 5. Lotgering factor as a function of crystallization time for the


Fig. 3. Lotgering factor as a function of pyrolysis time for the sample sample group IV. The inset shows the X-ray diffraction patterns of the
group II. The inset shows the X-ray diffraction patterns of the films. films.

0.67 for 8001C, indicating that a higher (100) orientation degree a large rapid increase before the drying time is less than 40 min,
can be obtained at a higher crystallization temperature. Figure 5 but a slow increase after 40 min.
shows the Lotgering factor as a function of crystallization time
for the sol–gel-derived BSPT thin films dried at 801C for 60 min, (3) Detailed Study on the Microstructure Evolution:
pyrolyzed at 4501C for 300 s, and crystallized at 7001C for HRTEM and FTIR
different times using the precursor solution SG-b. Examination To gain a further understanding on the orientation growth in the
of the figure indicates that f increases with the crystallization sol–gel-derived BSPT thin films, we investigate the microstruc-
time, from 0.56 for 120 s to 0.68 for 600 s. This means that a ture evolution of the samples by HRTEM and FTIR in this
longer crystallization process may enhance the (100) orientation. section.
However, it is noticeable that the Lotgering factor f does not In the previous discussion, it is observed that the Lotgering
change much when the crystallization time is o300 s. factor f exhibits a large increase as the pyrolysis temperature
Thirdly, the effect of lead excess concentration on the orien- increases from 4001 to 4501C (see Fig. 2), which indicates some
tation of the BSPT thin films was investigated. Figure 6 shows changes occurring in the film at 4501C that may enhance the
Lotgering factor f as a function of lead excess concentration for (100) orientation. Figure 8 presents the HRTEM micrographs
the sol–gel-derived BSPT thin films dried at 801C for 60 min, for plane view of the amorphous thin films dried at 801C for
pyrolyzed at 4501C for 300 s, and crystallized at 7001C for 300 s, 60 min, pyrolyzed at (a) 4001C and (b) 4501C for 300 s, respec-
using the precursor solution SG-a (0%), SG-b (10%), and SG-c tively, without crystallization. As shown in Fig. 8(a), the film is
(18%). As the figure shows, f increases with increasing lead completely amorphous with some special dark dots, which in-
excess concentration, from 0.15 for no lead excess to 0.81 for dicates that these dark regions have a higher average atomic
18% lead excess, which means that lead may play an important number than the surrounding areas. However, for the sample
role in the formation of BSPT (100) orientation. pyrolyzed at 4501C, as shown in Fig. 8(b), there are some crys-
Figure 7 presents Lotgering factor as a function of drying tallized regions (marked by the arrows in the figure) appearing
time for the sol–gel-derived BSPT thin films dried at 801C for in the amorphous film. It has been reported that PbO seeds
different times, pyrolyzed at 4501C for 300 s, and crystallized at precipitate from the amorphous phase when pyrolyzed at 4501C
6501C for 300 s using the precursor solution SG-c. Examination in PZT thin films.30 And these PbO seeds have a good lattice
of the figure indicates that f increases with increasing drying match with PZT (100) orientation (a 5 0.3974 nm for the oxygen
time, from 0.3 for 2 min to 0.74 for 60 min. Moreover, f exhibits sublattice in lead oxide and c 5 0.4146 for PZT perovskite),

Fig. 4. Lotgering factor as a function of crystallization temperature for


the sample group III. The inset shows the X-ray diffraction patterns of Fig. 6. Lotgering factor as a function of lead excess concentration for
the films. the sample group V.
October 2007 Orientation Control in Sol–Gel-Derived BSPT Thin Films 3251

may form at both the interface and the bulk of the film, when
the pyrolysis temperature increases to 4501C. These PbO
nanocrystals formed at the interface are oriented, which can
lead to a (100)-oriented BSPT thin film due to lattice match,
while those PbO nanocrystals formed in the bulk of the film
are misoriented, which cannot lead to an oriented BSPT thin
film. This may explain the large increase of Lotgering factor f
at 4501C in Fig. 2. When the pyrolysis time is prolonged, we
consider that there may be more PbO nanocrystals formed
from the amorphous phase at the interface, which would lead
to a stronger BSPT (100) orientation (seeing Fig. 3).
It could also be observed that the drying time affects the ori-
entation of the sol–gel-derived BSPT thin films, as shown in
Fig. 7. To investigate the reason for this phenomenon, the cross-
sectional HRTEM micrograph of the amorphous BSPT thin
film dried at 801C for 2 min and pyrolyzed at 4501C for 300 s is
studied, as displayed in Fig. 10. Different from the sample dried
for 60 min, in this film most PbO nanocrystals that are misori-
ented appear in the bulk of the film (marked by white arrows in
Fig. 7. Lotgering factor as a function of drying time for the sample
the figure), with a few formed at the film–substrate interface. It
group VI.
is generally accepted that during the drying process, the solvents’
volatilization and aging of the film, such as condensation reac-
which may enhance the (100) orientation.31 Thus, we suppose tions, co-occur in the as-coated thin films. Thus, the drying time
that these crystallized regions in Fig. 8(b) may be PbO nano- actually defines the residual organic parts in the amorphous thin
crystals, while the dark regions in Fig. 8(a) are probably some films. Fe32 reported that a prior drying step before the pyrolysis
kind of transition status before the PbO nanocrystals come into might reduce the OH band concentration in the amorphous
existence. Because BSPT also has a lattice constant similar to PZT thin films, while the existence of the OH band could impart
PbO (c 5 0.4075 nm for BSPT perovskite), these PbO nano- excess thermodynamic phase energy to the film and then in-
crystals would enhance the (100) orientation of BSPT thin crease the driving force for crystallization. In this study, the
films due to lattice match as well. However, as shown in FTIR spectra of the amorphous BSPT thin films dried for dif-
Fig. 8(b), these PbO nanocrystals are all misoriented, which ferent times are shown in Fig. 11. As the figure shows, no OH
obviously cannot lead to (100)-oriented BSPT thin films and band could be observed in both samples, while peaks of acetate
will not explain the large increase of Lotgering factor f at groups around 1500 cm1 were found. Moreover, it is obvious
4501C. Hence, we continued to study the cross-sectional that the longer drying time may reduce the concentration of re-
HRTEM micrograph of the sample pyrolyzed at 4501C, as sidual acetate groups in the amorphous films. Combining this
shown in Fig. 9(a). In the figure, a well-oriented nanocrystal result with that in Fig. 10, we consider that the drying time may
with the scale of several 10 nm can be observed at the film– affect the nucleation place of the PbO nanocrsytals. When the
substrate interface. The EDS analysis on the nanocrystal, as drying time is short, the larger residual organic parts may
shown in Fig. 9(b), reveals that the nanocrystal mainly con- provide more sites for the crystallization of PbO nanocrystals
sists of merely two elements, Pb and O, which provides strong within the film bulk, which are misoriented, whereas a long
support to our previous supposition that the nanocrystals drying time will reduce the residual organic parts and more
may be PbO. Moreover, the EDS analysis on the uncrystal- PbO nanocrystals will form at the film–substrate interface,
lized region away from the interface, as presented in Fig. 9(c), which are oriented. Moreover, a longing drying time also re-
shows that the region consists of all the five elements: Pb, Ti, sults in a more rigid gel film, which may hinder the crystalliza-
Bi, Sc, and O. Nevertheless, due to the very small scale, it is tion of PbO within the film bulk as well. Because more oriented
very hard to operate select area electron diffraction on the PbO nanocrystals are formed, higher BSPT (100) orientations
PbO nanocrystal. These results suggest that PbO nanocrystals are produced.

Fig. 8. High-resolution transmission electron microscopy micrographs for the plan view of the amorphous thin films dried at 801C for 60 min,
pyrolyzed at (a) 4001C and (b) 4501C for 300 s without crystallization.
3252 Journal of the American Ceramic Society—Wen et al. Vol. 90, No. 10

Fig. 9. (a) High-resolution transmission electron microscopy micrograph for the cross-sectional view of the interface in the amorphous thin films dried
at 801C for 60 min and pyrolyzed at 4501C for 300 s without crystallization; (b) energy dispersive spectroscopy (EDS) analysis on the crystallized region
in (a); (c) EDS analysis on the uncrystallized region away from the interface in (a).

(4) Mechanism for the Orientation Evolution barrier for the two nucleation events becomes larger (see Eqs. (3)
It is well known that the crystallization of sol–gel-derived thin and (4)), which makes the heterogeneous nucleation more
films occurs by a nucleation-growth process. From a thermo- important.
dynamic perspective, the driving force (the energy difference be- Based on the experimental results and discussion above, we
tween the amorphous and crystalline states) plays a significant may propose a mechanism for the orientation evolution in sol–
role in determining the nucleation events that take place.33 From gel-derived BSPT thin films. Firstly, the as-coated thin film is
the classic Volmer–Weber nucleation theory,34 the energy dried at a low temperature (801–1001C). During this process,
barriers for homogeneous and heterogeneous nucleation, which solvents’ volatilization and aging of the film, such as condensa-
depend on the driving force, can be described by tion reactions, co-occur. The drying time may define the residual
organic parts in the as-dried thin film and gel film status. Then,
16pg3 the film is pyrolyzed at an intermediate temperature. When the
DGhomo ¼ (3) pyrolysis temperature is higher than 4501C, PbO nanocrystals
3ðDGv Þ2 precipitate from the amorphous film. These PbO nanocrystals
may form within the bulk film, which are misoriented, or at the
16pg3 film–substrate interface, which are (100) oriented. If the drying
DGhetero ¼ f ðyÞ (4)
3ðDGv Þ2

where g is the interfacial energy, DGV is the driving force for


crystallization, and f(y) is a function related to the contact angle
y. For a hemispherical nucleus, the f(y) can be described as

f ðyÞ ¼ ð2  3 cos y þ cos3 yÞ=4 (5)

The homogeneous nucleation occurs within the bulk film,


which generally results in random orientation due to the equi-
axed grains forming in this nucleation process. The heteroge-
neous nucleation occurs at the film–substrate interface, which
usually leads to textured thin films. The difference in barrier
heights for the two nucleation events is therefore defined by the
term of interfacial energy, the driving force for crystallization,
and the contact angle with the substrate. Because f(y) is always
o1 (see Eq. (5)), the energy barrier for the heterogeneous nu-
cleation is lower than that for the homogeneous nucleation (see
Eqs. (3) and (4)), which means that the heterogeneous nucle-
ation is generally favored. However, these two kinds of nucle-
ation events always take place during the crystallization process.
The temperature dependence of the crystallization driving force
is presented in Fig. 12, which was originally proposed by Roy.35 Fig. 10. Cross-sectional high resolution transmission electron micros-
As shown in the figure, crystallization at a higher temperature copy micrograph of amorphous BiScO3–PbTiO3 thin films dried at 801C
results in a lower driving force. And the difference in the energy for 2 min and pyrolyzed at 4501C for 300 s.
October 2007 Orientation Control in Sol–Gel-Derived BSPT Thin Films 3253

promotes the heterogeneous nucleation and results in a higher


BSPT (100) orientation. After the nucleation, the equiaxed
grains in the bulk film and the (100)-oriented grains at the in-
terface compete with each other in the grain growth. It is re-
ported that (100) of the perovskite structure has the lowest
surface energy.16,17 Thus, the (100)-oriented grains have more
advantages in growth and may consume the misoriented grains
in the bulk film. This would explain the larger Lotgering factor f
obtained at a long crystallization time (see Fig. 5).
It can be noticed that in the sol–gel-derived BSPT thin films,
the (111) orientation has been greatly suppressed. The reason for
this is not yet clear. The use of the RTP furnace might be a
possible reason. As reported by many researches,16–18,36 the de-
velopment of (111) orientation was considered to arise from
lead-rich intermediate phases PtxPb due to lattice match. These
intermediate phases were unstable and tended to decompose at a
high temperature. Because in this study the RTP furnace was
used, the temperature was increased very rapidly so that the
formation of the intermediate phases was skipped. Another pos-
sible reason could be the existence of BiScO3, which may sup-
Fig. 11. Fourier transform infrared spectroscopy of amorphous
BiScO3–PbTiO3 films dried at 801C for 2 and 60 min.
press the formation of the intermediate phases PtxPb at the
interface. This needs to be further proved in the future work.
Hence, the growth of the (111) orientation was suppressed.
time is long enough, there will be less residual organic parts and However, it is only a primary result of this study. More efforts
a more rigid gel film as more PbO nanocrystals form at the in- will be focused on more detailed HRTEM studies to provide
terface, which may act as seeds for the nucleation of BSPT (100) more evidences for the proposed mechanism in the future work.
orientation due to lattice match. Moreover, the amount of PbO
nanocrystals is affected by the excess lead concentration added
to the precursor solution and the pyrolysis time. Finally, the
IV. Conclusions
amorphous thin film is crystallized at a high temperature. Dur-
ing this process, crystallization of the perovskite phase takes BSPT thin films with a single perovskite phase, a dense micro-
place, which occurs by a nucleation-growth process. Two kinds structure, and uniform grains were fabricated using a sol–gel
of nucleation events may take place. One is homogeneous method. By investigating the effects of different factors on the
nucleation, which occurs within the bulk film. The other is orientation growth of sol–gel-derived BSPT thin films, we found
heterogeneous nucleation, which occurs at the film–substrate that higher lead excess concentration, longer drying time, higher
interface. As PbO nanocrystals have precipitated from the amor- pyrolysis temperature, longer pyrolysis time, higher crystalliza-
phous phase during the pyrolysis process, both nucleations of tion temperature, and longer crystallization time could enhance
the BSPT perovskite phase may take place on these PbO nano- the (100) orientation of the BSPT thin films. Based on these ex-
crystals because of lattice match. The heterogeneous nucleation perimental results, a mechanism for the orientation evolution in
of BSPT takes place on the oriented PbO nanocrystals at the BSPT thin films was proposed. During the pyrolysis process,
interface and results in a BSPT (100) orientation. Homogeneous when the pyrolysis temperature is higher than 4501C, PbO nano-
nucleation occurs on the misoriented PbO nanocrystals in the crystals precipitate from the amorphous film. As these nano-
film bulk and leads to random orientation. Usually, these two crystals occur at the film–substrate interface, they tend to be
nucleation events of the BSPT perovskite phase take place dur- (100) oriented, which may act as seeds for the nucleation of
ing the crystallization process. The energy barrier heights of the BSPT (100) due to lattice match. During the crystallization pro-
two perovskite phase nucleation events, described in Eqs. (3) cess, the heterogeneous nucleation events take place on these
and (4), determine which one is more favored and important. At seeds, and the BSPT (100) orientation is produced.
a higher crystallization temperature, the difference in the energy
barrier for the two nucleation events becomes larger, which
References
1
C. A. P. Dearaujo, J. D. Cuchiaro, L. D. McMillan, M. C. Scott, and
J. F. Scott, ‘‘Fatigue-Free Ferroelectric Capacitors with Platinum-Electrodes,’’
Nature, 374 [6523] 627–629 (1995).
2
J. F. Scott and C. A. P. Dearaujo, ‘‘Ferroelectric Memories,’’ Science, 246
[4936] 1400–1405 (1989).
3
Y. Wang, K. F. Wang, C. Zhu, and J. M. Liu, ‘‘Polarization Fatigue of Fer-
roelectric Pb(Zr0.1Ti0.9)O3 Thin Films: Temperature Dependence,’’ J. Appl. Phys.,
99 [4] 044109 (2006).
4
W. Gong, J. F. Li, X. C. Chu, Z. L. Gui, and L. T. Li, ‘‘Preparation and
Characterization of Sol–Gel Derived (100)-Textured Pb(Zr,Ti)O3 Thin Films: PbO
Seeding Role in the Formation of Preferential Orientation,’’ Acta Mater., 52 [9]
2787–2793 (2004).
5
W. Gong, J. F. Li, X. C. Chu, and L. T. Li, ‘‘Texture Control of Sol–Gel
Derived Pb(Mg1/3Nb2/3)O3–PbTiO3 Thin Films Using Seeding Layer,’’ J. Am.
Ceram. Soc., 87 [6] 1031–1034 (2004).
6
P. Muralt, T. Maeder, L. Sagalowicz, S. Hiboux, S. Scalese, D. Naumovic,
R. G. Agostino, N. Xanthopoulos, H. J. Mathieu, L. Patthey, and E. L. Bullock,
‘‘Texture Control of PbTiO3 and Pb(Zr,Ti)O3 Thin Films with TiO2 Seeding,’’
J. Appl. Phys., 83 [7] 3835–3841 (1998).
7
D. Akai, M. Yokawa, K. Hirabayashi, K. Matsushita, K. Sawada, and M.
Ishida, ‘‘Ferroelectric Properties of Sol–Gel Delivered Epitaxial Pb(Zrx,Ti1x)O3
Thin Films on Si Using Epitaxial g-Al2O3 Layers,’’ Appl. Phys. Lett., 86 [20]
202906 (2005).
Fig. 12. Schematic diagram of the free energies of a sol–gel-derived 8
A. A. Talin, S. M. Smith, S. Voight, J. Finder, K. Eisenbeiser, D. Penunuri, Z.
amorphous film, the ideal supercooled liquid, and the crystalline per- Yu, P. Fejes, T. Eschrich, J. Curless, D. Convey, and A. Hooper, ‘‘Epitaxial
ovskited phase. DGv is the thermodynamic driving force for crystalliza- PbZr0.52Ti0.48O3 Films on SrTiO3/(001)Si Substrates Deposited by Sol–Gel Meth-
tion. After Roy.35 od,’’ Appl. Phys. Lett., 81 [6] 1062–1064 (2002).
3254 Journal of the American Ceramic Society—Wen et al. Vol. 90, No. 10
9 22
Y. Wang, C. Ganpule, B. T. Liu, H. Li, K. Mori, B. Hill, M. Wuttig, R. E. Eitel, C. A. Randall, T. R. Shrout, and S. E. Park, ‘‘Preparation and
R. Ramesh, J. Finder, Z. Yu, R. Droopad, and K. Eisenbeiser, ‘‘Epitaxial Fer- Characterization of High Temperature Perovskite Ferroelectrics in the Solid-
roelectric Pb(Zr, Ti)O3 Thin Films on Si Using SrTiO3 Template Layers,’’ Appl. Solution (1x)BiScO3–xPbTiO3,’’ Jpn. J. Appl. Phys., 41 [4A] 2099–2104 (2002).
23
Phys. Lett., 80 [1] 97–99 (2002). R. E. Eitel, C. A. Randall, T. R. Shrout, P. W. Rehrig, W. Hackenberger, and
10
S. W. Wang, H. Wang, S. X. Shang, J. Huang, Z. Wang, and M. Wang, ‘‘PZT S. E. Park, ‘‘New High Temperature Morphotropic Phase Boundary Piezoelectrics
Thin Films Prepared by Chemical Solution Decomposition Using a Bi2Ti2O7 Buf- Based on Bi(Me)O3–PbTiO3 Ceramics,’’ Jpn. J. Appl. Phys., 40 [10] 5999–6002
fer Layer,’’ J. Crystal. Growth, 217 [4] 388–392 (2000). (2001).
11 24
W. Gong, J. F. Li, X. C. Chu, Z. L. Gui, and L. T. Li, ‘‘Single-Crystal T. Yoshimura and S. Trolier-McKinstry, ‘‘Growth and Properties of
Nb-Doped Pb(Zr,Ti)O3 Thin Films on Nb-Doped SrTiO3 Wafers with Different (001)BiScO3–PbTiO3 Epitaxial Films,’’ Appl. Phys. Lett., 81 [11] 2065–2066 (2002).
25
Orientations,’’ Appl. Phys. Lett., 85 [17] 3818–3820 (2004). J. C. Nino and S. Trolier-McKinstry, ‘‘Dielectric, Ferroelectric, and Piezo-
12
K. Nashimoto, D. K. Fork, and G. B. Anderson, ‘‘Solid-Phase Epitaxial- electric Properties of (001) BiScO3–PbTiO3 Epitaxial Films Near the Morphotro-
Growth of Sol–Gel Derived Pb(Zr,Ti)O3 Thin-Films on SrTiO3 and MgO,’’ Appl. pic Phase Boundary,’’ J. Mater. Res., 19 [2] 568–572 (2004).
26
Phys. Lett., 66 [7] 822–824 (1995). H. Wen, X. H. Wang, X. Y. Deng, and L. T. Li, ‘‘Fabrication and Properties
13
C. Chen, D. F. Ryder, and W. A. Spurgeon, ‘‘Synthesis and Microstructure of Sol–Gel Derived BiScO3–PbTiO3 Thin Films,’’ J. Am. Ceram. Soc., 89 [7] 2345–
of Highly Oriented Lead Titanate Thin Films Prepared by a Sol–Gel Method,’’ 2347 (2006).
27
J. Am. Ceram. Soc., 72 [8] 1493–1498 (1989). H. Wen, X. H. Wang, and L. T. Li, ‘‘Characterization of (100)-Oriented
14
G. L. Brennecka, W. Huebner, B. A. Tuttle, and P. G. Clem, ‘‘Use of Stress to BiScO3–PbTiO3 Thin Films Synthesized by a Modified Sol–Gel Method,’’ Appl.
Produce Highly Oriented Tetragonal Lead Zirconate Titanate (PZT 40/60) Thin Phys. Lett., 88, 222904 (2006).
28
Films and Resulting Electrical Properties,’’ J. Am. Ceram. Soc., 87 [8] 1459–1465 F. K. Lotgering, ‘‘Topotactical Reactions with Ferrimagnetic Oxides Having
(2004). Hexagonal Crystal Structures—I,’’ J. Inorg. Nucl. Chem., 9 [2] 113–123 (1959).
15 29
H. X. Qin, J. S. Zhu, Z. Q. Jin, and Y. Wang, ‘‘PZT Thin Films with Pre- W. Zhao, X. Wang, J. Hao, H. Wen, and L. Li, ‘‘Preparation and Charac-
ferred-Orientation Induced by External Stress,’’ Thin Solid Films, 379 [1–2] 72–75 terization of Nanocrystalline (1x)BiScO3–xPbTiO3 Powder,’’ J. Am. Ceram.
(2000). Soc., 89 [4] 1200–1204 (2006).
16 30
S. Y. Chen and I. W. Chen, ‘‘Temperature–Time Texture Transition of S. Kalpat and K. Uchino, ‘‘Highly Oriented Lead Zirconium Titanate Thin
Pb(Zr1xTix)O3 Thin-Films: 1. Role of Pb-Rich Intermediate Phases,’’ J. Am. Films: Growth, Control of Texture, and Its Effect on Dielectric Properties,’’
Ceram. Soc., 77 [9] 2332–2336 (1994). J. Appl. Phys., 90 [6] 2703–2710 (2001).
17 31
S. Y. Chen and I. W. Chen, ‘‘Temperature–Time Texture Transition of W. Gong, J. F. Li, X. C. Chu, and L. T. Li, ‘‘Effect of Pyrolysis Temperature
Pb(Zr1-xTix)O3 Thin-Films: 2. Heat-Treatment and Compositional Effects,’’ on Preferential Orientation and Electrical Properties of Sol–Gel Derived Lead
J. Am. Ceram. Soc., 77 [9] 2337–2344 (1994). Zirconate Titanate Films,’’ J. Eur. Ceram. Soc., 24 [10–11] 2977–2982 (2004).
18 32
I. M. Reaney, K. Brooks, R. Klissurska, C. Pawlaczyk, and N. Setter, ‘‘Use of L. Fe, G. J. Norga, D. J. Wouters, H. E. Maes, and G. Maes, ‘‘Chemical
Transmission Electron-Microscopy for the Characterization of Rapid Thermally Structure Evolution and Orientation Selection in Sol–Gel-Prepared Ferroelectric
Annealed, Solution–Gel, Lead–Zirconate–Titanate Films,’’ J. Am. Ceram. Soc., 77 Pb(Zr,Ti)O3 Thin Films,’’ J. Mater. Res., 16 [9] 2499–2504 (2001).
33
[5] 1209–1216 (1994). R. W. Schwartz, ‘‘Chemical Solution Deposition of Perovskite Thin Films,’’
19
K. G. Brooks, I. M. Reaney, R. Klissurska, Y. Huang, L. Bursill, and Chem. Mater., 9 [11] 2325–2340 (1997).
34
N. Setter, ‘‘Orientation of Rapid Thermally Annealed Lead–Zirconate–Titanate J. Ricote, R. Poyato, M. Aguero, L. Pardo, M. L. Calzada, and D. Chat-
Thin-Films on (111) Pt Substrates,’’ J. Mater. Res., 9 [10] 2540–2553 (1994). eigner, ‘‘Texture Development in Modified Lead Titanate Thin Films Obtained by
20
A. Wu, P. M. Vilarinho, I. Reaney, and I. M. M. Salvado, ‘‘Early Stages of Chemical Solution Deposition on Silicon-Based Substrates,’’ J. Am. Ceram. Soc.,
Crystallization of Sol–Gel-Derived Lead Zirconate Titanate Thin Films,’’ Chem. 86 [9] 1571–1577 (2003).
35
Mater., 15 [5] 1147–1155 (2003). R. Roy, ‘‘Gel Route to Homogeneous Glass Preparation,’’ J. Am. Ceram.
21
S. J. Zhang, R. E. Eitel, C. A. Randall, T. R. Shrout, and E. F. Alberta, Soc., 52 [6] 344 (1969).
36
‘‘Manganese-Modified BiScO3–PbTiO3 Piezoelectric Ceramic for High- I. M. Reaney, D. V. Taylor, and K. G. Brooks, ‘‘Ferroelectric PZT Thin Films
Temperature Shear Mode Sensor,’’ Appl. Phys. Lett., 86 [26] 262904 (2005). by Sol–Gel Deposition,’’ J. Sol–Gel Sci. Technol., 13 [1–3] 813–820 (1998). &

You might also like