Twin

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

View Article Online / Journal Homepage / Table of Contents for this issue

PAPER www.rsc.org/crystengcomm | CrystEngComm

(111) Twinned BaTiO3 microcrystallites†


Shubin Qin, Duo Liu,* Feifei Zheng, Zhiyuan Zuo, Hong Liu and Xiangang Xu
Received 8th December 2009, Accepted 14th May 2010
DOI: 10.1039/b925863a

We report here the controlled synthesis of (111) twinned BaTiO3 microcrystallites, which are widely
known for their (110) ferroelectric twins. This (111) twinned structure was obtained through
a composite hydroxide mediated process by using amorphous TiO2 powders. A twin assisted re-entrant
growth model is proposed to account for the microstructural aspects determined. It is considered that
the (111) twins originate from the Ti–O octahedra of the amorphous TiO2 powders. The role of oriented
aggregation is also discussed.
Published on 02 June 2010 on http://pubs.rsc.org | doi:10.1039/B925863A
Downloaded by Linkopings universitetsbibliotek on 02 March 2013

Introduction process, 0.5 mmol of TiO2 and 9 g of composite hydroxides


(NaOH : KOH ¼ 0.515 : 0.485) were added into a 25 ml Teflon-
Twinned crystals have been known by mankind for a long time lined autoclave. The autoclave was sealed and heated at 180  C
owing to their unique morphologies Besides, twinned structures for 5 h. After the autoclave was cooled down to room temper-
are also technologically important due to their unique mechan- ature, 2 ml of 0.3M BaCl2 solution was added. The sealed
ical, electrical, and optical properties. In particular, twins in autoclave was heat treated again at 180  C for a range of time
ferroelectric perovskites are especially of interests due to their from 4 h to 20 days. The synthesized powders were collected and
close relationships linked to ferroelectric domain switching, washed thoroughly with deionized water to remove the
which enables a variety of device applications.1 As a model impurities.
ferroelectric perovskite, barium titanate (BaTiO3) is best known The microstructures of the as-synthesized products were
for its room temperature tetragonal structures characterized by investigated by using a powder X-ray diffractometer (XRD,
(110) twins. Besides the ferroelectric domain twins, structural Bruker D8 Advance, CuKa, Ni filter, l ¼ 1.540598 A,  40 kV,
twins such as (111) twin have also been reported for single crystal 40 mA). Raman analysis was carried out at room temperature on
BaTiO32,3 and bulk ceramics.4 However, (111) twins of BaTiO3 a NEXUS 670 spectrometer equipped with a Nd:YVO4 laser
can only be formed at high temperatures (1300  C) and under (1064 nm) and an InGaAs detector. The morphologies were
the conditions of compositional non-stoichiometry.5,6 Moreover, determined by using a field emission scanning electron micro-
penetrated BaTiO3 structures similar to the (111) twinned CaF2 scope (FESEM, Hitachi S-4800) and an atomic force microscope
and diamond7,8 have not been reported. (AFM, D3100, Nanoscope IIIa, Digital Instruments).
Recently, we reported that BaTiO3 powders exhibited photo-
chromic and ferromagnetic properties.9 To better understand the
unusual observations, in the present work, we studied the
microstructural characteristics and the formation mechanism of Results and discussion
the BaTiO3 (111) twins in detail. The (111) twinned BaTiO3 General
microcrystallites can be formed at low temperatures (<200  C) by
using amorphous TiO2 as the starting material. The micro- Fig. 1a shows XRD patterns of the final products obtained after
crystallites exhibit multiple penetrated structures as evidenced by CHM treatment of various periods. All the diffraction peaks can
scanning electron microscopy (SEM) and atomic force micros- be indexed to BaTiO3 (P4/mmm, JCPD 79-2264) without any
copy (AFM). A growth model is proposed and the technical evidence of a hexagonal phase. BaTiO3 formed after reaction for
potentials of this unique structure are discussed. 4 h and the diffraction intensities increased as the reaction pro-
ceeded. The composition of the product is also characterized by
energy dispersive spectroscopy (EDS), which indicates that there
Experimental are only barium, titanium and oxygen in the sample (Fig. S1,
ESI†). As Raman spectroscopy is very sensitive to structural
The BaTiO3 microcrystallites were synthesized through a two- phase transitions, it is used to verify that only the tetragonal
step composite hydroxide mediated (CHM) approach by taking phase is present in the final product after reaction for 20 days. As
advantage of the low eutectic point of mixed NaOH and KOH shown in Fig. 1b, all the Raman modes can be attributed to
solids.10 Anatase TiO2 and BaCl2 were employed as the raw tetragonal BaTiO3.11,12 The Raman bands located at 265 cm1
materials. First, the anatase TiO2 powders were pretreated in and 518 cm1 are assigned to TO vibrations of A1 symmetry. The
molten hydroxides to form an amorphous phase. During this band at 715 cm1 can be attributed to the highest frequency LO
mode of A1 symmetry. The sharp peak located at 305 cm1 is
State Key Laboratory of Crystal Materials, Shandong University, 27 South associated with the B1 mode. The modes at 305 cm1 and
Shanda Road, Jinan, Shandong, 250100, P. R. China. E-mail: liuduo@sdu.
edu.cn; Fax: +86 531-88362807; Tel: +86 531-88363901 715 cm1 indicate that the samples are tetragonal BaTiO3.12 No
† Electronic supplementary information (ESI) available: Fig. S1 and S2. Raman bands can be attributed to hexagonal BaTiO3, which
See DOI: 10.1039/b925863a exhibits characteristic Raman peaks at 153 cm1 and 640 cm1.13

This journal is ª The Royal Society of Chemistry 2010 CrystEngComm, 2010, 12, 3003–3007 | 3003
View Article Online
Published on 02 June 2010 on http://pubs.rsc.org | doi:10.1039/B925863A
Downloaded by Linkopings universitetsbibliotek on 02 March 2013

Fig. 2 SEM images of (111) twinned BaTiO3 microcrystallites synthe-


sized at 180  C for 20 days.

using as-purchased anatase TiO2 results in separated single


crystal BaTiO3 nanocubes, as demonstrated by previous inves-
tigations.10,15 As a result, microstructural features of the pre-
treated TiO2 need to be clarified. Our XRD analysis reveals that
the pretreated product is amorphous. It is well known that
anatase TiO2 can be easily dissolved in alkaline solution at
elevated temperatures.16,17 In alkaline solutions, Ti–O octahedra
may combine with hydroxyl groups or other Ti–O octahedra to
form clusters, which yields short range ordered Ti–O octahedra
Fig. 1 (a) XRD patterns of BaTiO3 synthesized at 180  C for 4 h, 8 h, units. This result has been confirmed by previous studies, which
12 h, and 20 days,respectively. (b) Raman spectrum of BaTiO3 micro- showed that amorphous TiO2 contain large numbers of defects,
crystallites synthesized at 180  C for 20 days. (c) Raman spectrum of the i.e. oxygen vacancies and distorted Ti–O octahedra.18,19 Bursill
pretreated TiO2.
et al. showed that face-sharing octahedra can exist in non-stoi-
chiometric rutile (TiO2x).20 Recent studies by Penn et al. also
Fig. 2a shows a typical SEM image of BaTiO3 microcrystallites
revealed twinned structures in aged TiO2 nanoparticles.21,22 This
grown at 180  C for 20 days. The crystallites are 30–40 mm in
short range ordered structure is also confirmed by Raman
size and cubic in shape with penetrated structures. The cubic
analysis shown in Fig. 1c, where Ti–O vibrations of various
elements could overlap and form a chained structure (Fig. 2b).
forms can be easily identified. These vibrations are characterized
Small cubes could also be embedded into large ones (Fig. 2c)
by broadened bands with several peaks associated with either
where Ostwald coarsening could be observed. A statistical
rutile or anatase TiO2. Previous studies proved that the BaTiO3
analysis reveals that (1) all the exposed surfaces are (100) planes
(111) twin plane is composed of oxygen-deficient face-sharing
due to their relatively low surface energy,14 (2) that the cubes
Ti–O octahedra,23 so that the pretreated TiO2 could play an
intersect their neighbors in the [221] directions and form a re-
important role in the formation of (111) twins in BaTiO3.
entrant angle of 109.5 . As a result, it can be tentatively
concluded that the growth process involves (111) twins.
Formation mechanism
As mentioned previously, (111) twins have been reported in
The role of pretreated TiO2
BaTiO3 single crystals and ceramics.24 It was considered that the
The utilization of pretreated anatase TiO2 is found to be crucial (111) twins originate from the oxygen-deficient hexagonal
for obtaining (111) twinned BaTiO3. The synthesis route solely BaTiO3 and could be synthesized through control of oxygen

3004 | CrystEngComm, 2010, 12, 3003–3007 This journal is ª The Royal Society of Chemistry 2010
View Article Online

partial pressure at high temperatures (>1300  C)23,25 A twin


assisted model has been proposed to account for the high growth
rate at the re-entrant edges where the nucleation barrier is
considerably lower.26–28 However, we believe that the mechanism
governing the formation of (111) twinned BaTiO3 in a CHM
environment is fundamentally different and should be closely
linked to the microstructural aspects of the pretreated anatase
TiO2.
A schematic representation of the growth model is shown in
Fig. 3. Fig. 3a shows a pair of face-sharing Ti–O octahedra. A
BaTiO3 nucleus containing a (111) twin plane is formed when
Ba2+ cations were captured by the face-sharing Ti–O octahedra
(Fig. 3b). The re-entrant angle between the neighboring BaTiO3
lattices reduces the nucleation barrier (to 5% of the value on
Published on 02 June 2010 on http://pubs.rsc.org | doi:10.1039/B925863A

a flat interface)25 and accelerates the growth rate significantly. As


the twin plane proceeds, a penetrated structure is formed. The
Downloaded by Linkopings universitetsbibliotek on 02 March 2013

ideal morphology of a (111) penetration twin is shown in Fig. 3c,


where two identical cubes penetrate coordinatively. However,
local thermodynamic instabilities also enable an Ostwald coars-
ening process, which may lead to the growth of larger cubes in
compensation of the smaller ones. In addition, multiple (111)
twins promote growth rate in different directions. As the super-
saturation decreases, the morphologies of the BaTiO3 particles
could resemble the patterns shown in Fig. 3d,e. Fig. 4 SEM images of (111) twinned BaTiO3 microcrystallites show the
Fig. 4 shows re-entrant angles between two cubes, based on details of re-entrant edge (a) and layered growth traces (b).
which the crystallographic directions of the twin planes can be
determined. Fig. 4a shows an exposed corner of BaTiO3, similar
to the schematic representation shown in Fig. 3e. The re-entrant re-entrant edges. It is known that impurities can block the
edge at the (111) twin plane in Fig. 4a provides nucleation sites motion of surface steps on crystal surfaces, i.e., on KH2PO4 and
where favorable growths are expected. The (100) faces grow K2Cr2O7.29,30 As impurities such as metastable nuclei and Ti–O
through the conventional 2D growth model. Grooves are formed clusters can not be avoided during the growth of BaTiO3
at the interfaces between the two adjacent cubes due to lattice microcrystallites, their accumulations on the (100) surfaces create
mismatch between the (221) plane and the (100) plane. Growth pinning sites and hinder monolayer growth. It appears that
steps can be clearly visualized in Fig. 4b. Nuclei can be easily macrosteps may overcome the pinning centers due to the high
formed at re-entrant edges (indicated by arrows) and provide degree of supersaturation.29 Consequently, bunches or macro-
terrace sites for 2D growth. It is also found that in areas close to steps can be clearly visualized.
the re-entrant edges the surface steps are of small height with Contact mode AFM measurements were performed on the
higher density. Both the height and the lateral separations (100) surface of the as-grown BaTiO3 microcrystallites, as shown
of these surface steps become greater at areas far from the in Fig. 5a. It can be clearly verified that foreign particles block
the proceedings of the growth layers where a convex front can be
found. Fig. 5b shows a cross-section plot of some surface steps.
Typically, they are 2–3 nm in height, containing 5–8 unit
 c ¼ 4.018 A
lattices (a ¼ 3.9998 A,  for BaTiO3). Fig. 5c shows
a hole of 70 nm in depth. The possible reason for the formation
of holes lies in the fact that the surface layers can not cover the
inclusions completely. Another surface feature of the crystallites
is numerous nuclei randomly distributed on the (100) surfaces
(Fig. 5a). All the nuclei exhibit sizes greater than 16 nm with an
average value of 23.2 nm (based on a statistical analysis of more
than 100 nuclei). As nuclei smaller than a critical value will
spontaneously diminish, our results suggest that BaTiO3 nuclei
have a critical size greater than 16 nm. This finding agrees with
many previous studies which revealed that it is almost impossible
to synthesize BaTiO3 nanoparticles smaller than 20 nm.31–33
Theoretically, the (100) surface of BaTiO3 is characterized by
a high value of surface energy (1.237 eV per surface unit cell),34
Fig. 3 Schematic representation of the evolutionary process of the (111) which may help to explain the large size of the critical nuclei.
twinned BaTiO3 crystallites. (b,c) are the ideal models and (d,e) are the To further clarify the growth process of (111) twinned BaTiO3
practical models of the as-grown BaTiO3 microcrystallites. microcrystallites, the initial growth stages were studied. After the

This journal is ª The Royal Society of Chemistry 2010 CrystEngComm, 2010, 12, 3003–3007 | 3005
View Article Online

shows many re-entrant edges which facilitate 2D nucleation such


that the growth rate of some cubes is greatly enhanced. Because
re-entrant edges will finally disappear,26 Ostwald ripening would
become more significant, especially when supersaturation
decreases. As a result, small cubes dissolve and large cubes grow
preferentially (Fig. 6d, reaction for 4 days). Occasionally, we also
observe penetrated structures with exposed (111) faces, as shown
in Fig. 6b. It is clearly shown that these particles are (111)
twinned. A typical (111) twinned BaTiO3 particle is also shown in
Fig. S2a.† The HRTEM image (Fig. S2b†) obtained from small
BaTiO3 particles also supports the proposed twin assisted growth
mechanism, where a (111) twin can be clearly visualized. After
reaction for 4 days, the twinned structures will grow and form
aggregations (Fig. 6c and 6d). Similar to many other nano-
Published on 02 June 2010 on http://pubs.rsc.org | doi:10.1039/B925863A

structured materials,21,22,35 (111) twins could also be formed


through oriented aggregations of adjacent grains with crystal-
Downloaded by Linkopings universitetsbibliotek on 02 March 2013

lographic orientations close to the expected twinning relation.36

Conclusions
BaTiO3 microcrystallites were successfully synthesized through
a two-step CHM approach at low temperature. It is interesting to
note that these microcrystallites contain multiple (111) twins. A
twin assisted growth model is proposed to account for the
observations. The multiple (111) twins were formed through face
sharing of Ti–O octahedra, intentionally created through pre-
treating anatase TiO2 powders. These unique ferroelectric
nanostructures are potentially of importance, provided that the
Fig. 5 (a) Surface morphology measured by AFM of the (100) face of
(111) twin planes may serve as pinning sites for the motion of
BaTiO3 microcrystallites grown at 180  C for 20 days; (b) and (c) are the
height profiles along the lines indicated in (a). (110) ferroelectric twin walls. A thorough exploration of the
ferroelectric properties of this twinned structure is currently
under way.
solution was heated at 180  C for 1 day, two typical morphol-
ogies can be found, as shown in Fig. 6a,b. Fig. 6a shows that
these BaTiO3 particles are spherical and contain many pene- Acknowledgements
trated cubes of 60 nm similar to pure BaTiO3 nanocubes The authors thank National Science Foundation of China
synthesized by the CHM approach.10,15 It is evident that these (NSFC) (Grant No. 50702031, 60974117, 50872070), the Excel-
spherical particles contain multiple (111) twins. This structure lent Young Investigators Award Foundation of Shandong
Province (Grant No. BS2009CL021), and National Basic
Research Program of China (973 Program) (Grant No.
2009CB930503) for financial support.

References
1 D. Dragan, Rep. Prog. Phys., 1998, 61, 1267.
2 J. P. Remeika and W. M. Jackson, J. Am. Chem. Soc., 1954, 76, 940.
3 E. A. D. White, Acta Crystallogr., 1955, 8, 845.
4 V. J. Tennery and F. R. Anderson, J. Appl. Phys., 1958, 29, 755.
5 V. Krasevec, M. Drofenik and D. Kolar, J. Am. Ceram. Soc., 1990,
73, 856.
6 B. K. Lee, S. Y. Chung and S. L. Kang, J. Am. Ceram. Soc., 2000, 83,
2858.
7 I. Sunagawa, Crystals Growth, Morphology and perfection, Cambridge
University Press, New York, 2005.
8 M. A. Tamor and M. P. Everson, J. Mater. Res., 1994, 9, 1839.
9 S. Qin, D. Liu, Z. Zuo, Y. Sang, X. Zhang, F. Zheng, H. Liu and
X. G. Xu, J. Phys. Chem. Lett., 2010, 1, 238.
10 H. Liu, C. Hu and Z. L. Wang, Nano Lett., 2006, 6, 1535.
11 R. Asiaie, W. Zhu, S. A. Akbar and P. K. Dutta, Chem. Mater., 1996,
8, 226.
12 C. H. Perry and D. B. Hall, Phys. Rev. Lett., 1965, 15, 700.
Fig. 6 SEM images of BaTiO3 synthesized at 180  C for 1 day (a,b) and 13 N. G. Eror, T. M. Loehr and B. C. Cornilsen, Ferroelectrics, 1980, 28,
4 day (c,d), respectively. 321.

3006 | CrystEngComm, 2010, 12, 3003–3007 This journal is ª The Royal Society of Chemistry 2010
View Article Online

14 R. I. Eglitis, G. Borstel, E. Heifets, S. Piskunov and E. Kotomin, 26 K. A. Hu, B. V. Hiremath and R. E. Newnham, Phase Transitions,
J. Electroceram., 2006, 16, 289. 1986, 6, 153–164.
15 S. Qin, D. Liu, H. Liu and Z. Zuo, J. Phys. Chem. C, 2008, 112, 27 M. K. Kang, Y. S. Yoo and D. Y. Kim, J. Am. Ceram. Soc., 2000, 83,
17171. 3202.
16 Z. Y. Yuan, J. F. Colomer and B. L. Su, Chem. Phys. Lett., 2002, 363, 28 H. Y. Lee and J. S. Kim, J. Am. Ceram. Soc., 2002, 85, 977.
362. 29 T. A. Land, T. L. Martin, S. Potapenko, G. T. Palmore and
17 Z. Y. Yuan and B. L. Su, Colloids Surf., A, 2004, 241, 173. J. J. D. Yoreo, Nature, 1999, 399, 442.
18 V. V. Hoang, Phys. Status Solidi B, 2007, 244, 1280. 30 A. D. Derksen, W. J. P. van Enckevort and M. S. Couto, J. Phys. D:
19 J. P. Rino and N. Studart, Phys. Rev. B: Condens. Matter Mater. Appl. Phys., 1994, 27, 2580.
Phys., 1999, 59, 6643. 31 H. Xu, L. Gao and J. Guo, J. Eur. Ceram. Soc., 2002, 22, 1163.
20 L. A. Bursill, M. G. Blanchin, A. Mebarek and D. J. Smith, Radiat. 32 H. J. Chen and Y. W. Chen, Ind. Eng. Chem. Res., 2003, 42, 473.
Eff. Defects Solids, 1983, 74, 253. 33 T. Yan, Z. G. Shen, J. F. Chen, X. L. Liu, X. Tao and J. Yun, Chem.
21 R. L. Penn and J. F. Banfield, Am. Mineral., 1998, 83, 1077. Lett., 2005, 34, 1196.
22 R. L. Penn, J. Phys. Chem. B, 2004, 108, 12707. 34 J. Padilla and D. Vanderbilt, Phys. Rev. B: Condens. Matter, 1997, 56,
23 A. Recnik, J. Bruley, W. Mader, D. Kolar and M. R€ uhle, Philos. Mag. 1625.
B, 1994, 70, 1021. 35 R. L. Penn and J. F. Banfield, Science, 1998, 281, 969.
24 R. C. DeVRIES, J. Am. Ceram. Soc., 1959, 42, 547. 36 G. Kastner, R. Wagner and V. Hilarius, Philos. Mag. A, 1994, 69,
25 A. Recnik and D. Kolar, J. Am. Ceram. Soc., 1996, 79, 1015. 1051.
Published on 02 June 2010 on http://pubs.rsc.org | doi:10.1039/B925863A
Downloaded by Linkopings universitetsbibliotek on 02 March 2013

This journal is ª The Royal Society of Chemistry 2010 CrystEngComm, 2010, 12, 3003–3007 | 3007

You might also like