You are on page 1of 6

Optical/ferroelectric characterization of BaTiO3

and PbTiO3 colloidal nanoparticles


and their applications in hybrid
materials technologies

Yuriy Garbovskiy and Anatoliy Glushchenko*


Department of Physics & Biofrontiers Program, University of Colorado
Colorado Springs, Colorado Springs, Colorado 80918, USA
*Corresponding author: aglushch@uccs.edu

Received 18 March 2013; revised 22 April 2013; accepted 22 April 2013;


posted 24 April 2013 (Doc. ID 187267); published 28 May 2013

In this paper, we will explore how optical and ferroelectric properties of the stressed ferroelectric nano-
particles prepared through ball milling set a limit on the performance of optical and electro-optical de-
vices based on such materials. It was found that suspensions of BaTiO3 nanoparticles exhibit a blue shift
in the optical band gap with a decrease in particle size. The optical band gap of PbTiO3 nanoparticles is
not affected by the milling time. Polarization switching is composed of slow and fast components. A slow
component is threshold-less and is caused by the particle reorientation while a fast component has a
threshold, and its rise time is inversely proportional to the electric field. The absorption edge of these
suspensions accounts for the applications in the near UV range, while kinetics of the polarization switch-
ing governs the speed of electro-optical devices. © 2013 Optical Society of America
OCIS codes: (160.2260) Ferroelectrics; (260.7190) Ultraviolet; (300.1030) Absorption; (300.2530)
Fluorescence, laser-induced; (230.3720) Liquid-crystal devices.
http://dx.doi.org/10.1364/AO.52.000E34

1. Introduction materials: (a) chemical synthesis (liquid crystals


Applied optics of liquid crystals is an exciting and are formed as a result of chemical reactions), and
continuously expanding field. Without exaggeration, (b) nonsynthetic methods (modification of the basic
optical devices based on liquid crystals (i.e., displays, physical properties of liquid crystals by doping them
sensors, tunable retarders and filters, macro- and with nanoparticles).
microlenses, prisms, shutters, etc.) are among rou- Recent reviews indicate that hybrid materials
tine tools widely used in many research labs around based on liquid crystals and inorganic ferroelectric
the globe. Overall performance of such devices nanoparticles are very promising candidates for
strongly relies on physical and chemical properties the next generation of liquid crystal-based devices
of liquid crystals (elastic constants, viscosity, [2,3]. Ferroelectric nanoparticles can modify basic
birefringence, dielectric permittivity) [1]. Therefore, physical properties of liquid crystals (alter tempera-
design of novel liquid crystalline materials with ture of the phase transitions, increase birefringence
improved properties is a prerequisite for further and dielectric anisotropy, decrease viscosity); reduce
progress in the field. Basically, there are two undesirable processes (image sticking) by trapping
approaches used to design advanced mesomorphic ions [2,3]; and even new effects, such as linear
electro-optical response in nematics [4], and asym-
metric Freedericksz transitions [5] come up.
1559-128X/13/220E34-06$15.00/0 In most cases, in order to make liquid crystal col-
© 2013 Optical Society of America loid of nanoparticles, a suspension of ferroelectric

E34 APPLIED OPTICS / Vol. 52, No. 22 / 1 August 2013


nanoparticles (covered with surfactant to prevent ag- ratio of ferroelectric nanoparticles:oleic acid:heptane
gregation) in isotropic fluid is mixed with pure liquid by weight, respectively. Milling time was varied from
crystals (detailed procedures can be found in original 5 to 50 h. As shown in [9,10] the size of the nanopar-
papers [4], see recent reviews for more references ticles was determined by the time of milling.
[2,3]). Analysis of the published data allows us to con- UV–visible spectrometer (Ocean Optics HR4000)
clude that the performance of the final product (liquid was used to (a) measure the absorption of light in
crystal colloid), among other important factors, also the near UV region and (b) analyze spectral compo-
depends on the quality of the initial suspension of sition of the photoluminescence. A CW solid state
ferroelectric nanoparticles (average size and size dis- laser (408 nm, 20 mW) was used as excitation light
tribution, degree of aggregation, optical transparency). source.
Particle aggregation is a very serious issue, especially Average sizes of the particle aggregates were
in the case of ferroelectric nanoparticles [6] because of determined by applying method of dynamic light
the interactions between permanent dipoles. This scattering (DLS) (ZetaPALS—Brookhaven Instru-
leads to weaker net spontaneous polarizations, re- ments Corporation) as described in [11].
duced dipole moments, and smaller displacement DLS and spectral measurements were carried out
currents [6,7]. As a result, overall performance of by using standard 1 cm UV cuvettes, which were
any optical device based on ferroelectric nanoparticles filled with 10 times the diluted fresh suspensions
and liquid crystals will be reduced. It’s worthy to say of ferroelectric nanoparticles BaTiO3 and PbTiO3 .
that possible applications of ferroelectric nanopar- In the case of absorption spectroscopy, the measured
ticles in material technologies are very broad and extinction of light near the absorption edge of ferro-
are not limited just to liquid crystals [8]. In light of electric nanoparticles was associated with light ab-
this, characterization of the suspensions of ferroelec- sorption. Loss of light caused by the scattering was
tric nanoparticles in isotropic fluids became a hot topic negligibly small compared with the absorption loss.
of the research during the last few years [6,7]. Conventional experimental technique inspiringly
For any optical and electro-optical applications of described in [6,7] was applied to study switching ki-
ferroelectric nanoparticles, among other things, we netics. A sandwich-like cell was made of 25 mm ×
need to know: (1) how much light is absorbed/ 25 mm ITO glasses with an electrode of area
scattered, and (2) how fast is the operating device 20 mm × 20 mm, then connected in series with a
(switching times). This knowledge is very important reference resistor. A low frequency (0.05–0.25 Hz),
for both the production and the performance of square waveform AC electric field was applied across
hybrid materials based on liquid crystals, polymers, the cell (an AC voltage was switched between VA
and ferroelectric nanoparticles. Intensive absorption and −VA ); and current through the reference resistor
of the near UV light by hybrid materials may impose was measured by means of an oscilloscope. Ampli-
certain limitations on their use as materials for UV tude VA of the applied voltage varied from 0 to 100 V.
optics (absorbed light heats the sample, decreases
order parameter, and affects device performance). 3. Results and Discussion
Reorientation of the ferroelectric nanoparticles
under the action of the electric field may set the limit A. Light Absorption and Photoluminescence
on switching times of hybrid materials. To our knowl-
Figures 1 and 2 show near UV–visible transmission
edge, there is no such experimental data available in
spectra of the suspensions of BaTiO3 (Fig. 1) and
the literature for isotropic suspensions of stressed
ferroelectric nanoparticles prepared through ball
milling [9].
This paper will investigate the optical (near UV
absorption) and ferroelectric (switching kinetics)
properties of two sets of stressed ferroelectric
nanoparticles (BaTiO3 and PbTiO3 ) suspended in
isotropic fluid (heptane).

2. Experimental Methods and Materials


Ferroelectric nanoparticles (BaTiO3 or PbTiO3 ) were
prepared by wet grinding three micron powders of
BaTiO3 or PbTiO3 in a zirconia ball mill (two station
planetary ball mill PM200 manufactured by Retsch
GmbH, Germany) according to procedures described
in [9]. The carrier fluid was heptane, to which oleic
acid was added as a surfactant for the nanoparticles.
The addition of a surfactant was necessary to prevent Fig. 1. Optical transmission spectra of the suspensions of BaTiO3
agglomeration of the ferroelectric nanoparticles and stressed nanoparticles for different milling times (dotted curve—
to enable small sizes to be achieved by milling. Typ- 5 h; dashed curve—10 h; solid curve—25 h). Tauc’s plot is shown in
ically, the solutions were prepared using a 1∶1∶20 the inset for the same suspensions.

1 August 2013 / Vol. 52, No. 22 / APPLIED OPTICS E35


cannot account for the observed blue shift shown
in Fig. 1.
Reduction in the nanoparticle size can cause an
increase in the optical band gap of ferroelectric
through the narrowing of the valence and conduction
bands [13]. The other important factors that can af-
fect optical band gap are defect centers, mechanical
stress, and changes in the crystallinity. An existence
of the mechanical stress in BaTiO3 nanoparticles
prepared through ball milling was reported in [19].
In addition, it was shown that such ferroelectric
nanoparticles are in the tetragonal phase and are
not amorphous [19].
Surface defects such as oxygen vacancies are very
common in perovskite materials [20]. Mechanical
Fig. 2. Optical transmission spectra of the suspensions of PbTiO3 milling is very likely to generate additional oxygen
stressed nanoparticles for different milling times (dotted curve— vacancies thus increasing their total density. Elec-
5 h; dashed curve—10 h; solid curve—50 h). Tauc’s plot is shown in
trons, originating from oxygen vacancies, can occupy
the inset for the same suspensions.
the lowest states for transitions near the conduction
band edge resulting in the increase of the effective
PbTiO3 (Fig. 2) nanoparticles in heptane for different optical band gap (the Burstein–Moss effect [21,22]).
milling time (5, 10, 25, 50 h). Experimental data Green photoluminescence observed in both BaTiO3
clearly indicates very different behavior of these sus- and PbTiO3 suspensions (Fig. 3) is an indirect indica-
pensions. BaTiO3 nanoparticles exhibit a blue shift tion for the generation of the surface defects (mostly
in the absorption edge (Fig. 1). The absorption edge oxygen vacancies) in ferroelectric nanoparticles dur-
of the PbTiO3 nanoparticles does not depend on the ing their preparation (ball milling). An intensity of
milling time (Fig. 2). the photoluminescence increases when milling time
Absolute values of the optical band gap Eg can be increases for both materials while the position of
estimated by assuming direct interband transitions the photoluminescence peak (∼505 nm) does not
in these nanoparticles (BaTiO3 [12,13], PbTiO3 shift with this increase. These two features reflect the
[14]) and using the following relation: “surface” origin of the photoluminescence.
Since the wavelength of the excitation light source
α · ℏ · ω2 ∼ const · ℏ · ω − Eg  (408 nm) lies outside the absorption edge of the stud-
ied materials [∼320–350 nm for BaTiO3 (Fig. 1) and
where ℏ · ω is photon energy, and α is an absorption ∼370 nm for PbTiO3 (Fig. 2)], emitted light cannot be
coefficient related to the measured transmission T: produced by band-to-band transitions. This means
that photoluminescence states are located within
T  e−α·d the energy band gap (in the forbidden band). And
such states can be associated with surface defects
(d is the cell gap). Dependence α · ℏ · ω2 versus ℏ · ω generated during the ball milling. These defects
(Tauc’s plot) is shown in the insets in Figs. 1 and 2. are randomly distributed on the particle surface,
With an increase in the milling time that corre-
sponds to a decrease in the average particle size,
the absorption edge shifts to a shorter wavelength
(blue shift) for BaTiO3. The optical band gap of the
bulk BaTiO3 is 3.0–3.2 eV [12,13,15]. The observed
blue shift is more than 0.5 eV for 25 h milled
BaTiO3 , this milling time corresponds approximately
to 9 nm particles [9,10].
A similar blue shift for mesocrystals of BaTiO3 was
reported in [16] and explained in terms of quantum
confinement, which requires ballistic motion of elec-
trons and holes where the mean free path of the car-
riers is long compared with the cell dimensions.
However, as was pointed out repeatedly in [17,18],
there is no evidence for quantum confinement in
an oxide ferroelectric at room temperature, where Fig. 3. Room temperature photoluminescence spectra of PbTiO3
the mean free path is 0.1–1.0 nm. Since the average stressed nanoparticles for different milling times (dashed curve—
sizes of ferroelectric nanoparticles BaTiO3 in hep- 5 h; dotted curve—10 h; solid curve—25 h; dashed–dotted curve—
tane are at least 10 times greater than the mean free 50 h). Inset in the figure shows the same data for BaTiO3
path of charge carriers, quantum confinement effect suspensions (dashed curve—5 h; solid curve—10 h).

E36 APPLIED OPTICS / Vol. 52, No. 22 / 1 August 2013


and, as a result, there is no shift in the position of the polarization [6,7]. Suspension of such nanoparticles
photoluminescence peak (Fig. 3). On the other hand, in isotropic fluid can be considered as a system of
the total number of these defects increases as milling free-to-rotate electric dipoles. Under the action of
time gets longer, and causes increase in the intensity the uniform external electric field these dipoles are
of the emitted light (Fig. 3). Stressed ferroelectric reoriented, and this physical reorientation accounts
nanoparticles are very complicated objects; they for switching current [6,7]. Previous studies [6,7]
have a tetragonal structure. Thus, the approach used analyzed ferroelectric response in the relatively
to describe photoluminescence in the ferroelectric low electric fields (0–1 V∕μm). The observed switch-
nanocrystals can be used [23,24]. However, their sur- ing current was associated with physical reorienta-
face is stressed and has many defects. This allows tion of the ferroelectric nanoparticles. In this paper
partial application of the models developed for disor- it was decided to extend the amplitude of the applied
dered nanopowders [25]. A more detailed analysis electric field by the order of magnitude (0–10 V∕μm)
of the photoluminescence in fluid suspensions of with focus on the time scale of the switching.
stressed ferroelectric nanoparticles will be reported It was found that both suspensions (BaTiO3 and
in a separate paper. PbTiO3 ) exhibit the same ferroelectric response.
B. Dynamic Light Scattering
Figure 4 shows typical switching current measured
for two cases: low electric field (dashed curve) and
Papers [6,7] revealed detrimental effects of the par- relatively high electric field (solid curve). Under
ticle aggregation on the ferroelectric response of the action of low electric field (<1 V∕μm) only one
the BaTiO3 suspensions. Since information about peak in the switching current is observed. This peak
average size of the aggregate is very helpful for is caused by the reorientation of ferroelectric nano-
hybrid material design, it was decided to study this particles and is in agreement with already published
question experimentally. Table 1 shows average data [6,7]. However, after the electric field reaches
sizes, effective hydrodynamic diameters, of BaTiO3 a certain threshold value (0.8–1.0 V∕μm), a second
and PbTiO3 suspensions obtained from DLS mea- peak appears that is faster than the peak caused
surements. As seen in Table 1, suspensions of by particle reorientation (Fig. 4, solid curve). Thus,
BaTiO3 and PbTiO3 nanoparticles exhibit different in the high electric fields (1–10 V∕μm) switching cur-
degree of aggregation. BaTiO3 nanoparticles tend rent of the studied ferroelectric suspensions has two
to aggregate more likely as compared to PbTiO3 .
components: a slow component and a fast component
Comparison of the data from Table 1 with TEM data
reported in [9,10] allows us to estimate number
of BaTiO3 nanoparticles per one aggregate (there
is no TEM data available for stressed PbTiO3 nano-
particles). In the case of BaTiO3, 5 and 10 h milled
suspensions are composed mostly of single nanopar-
ticles and dimers. 25 h milled suspension of BaTiO3
is composed mostly of big aggregates. The number of
particles per one aggregate depends on the aggregate
shape and may vary from 50–100 (needle-like aggre-
gates) to ∼503 (spherical aggregates) for 25 h milled
BaTiO3 . Table 1 also indicates that aggregation
degree is almost the same for all milling times for
suspensions of PbTiO3 nanoparticles, which are
composed mostly of single nanoparticles and their
dimers.
C. Polarization Switching Kinetics
Fig. 4. Kinetics of the switching current in low (dashed curve)
Ferroelectric nanoparticles produced through ball and high (solid curve) electric fields for the 10 h milled PbTiO3
milling exhibit enhanced permanent spontaneous suspension.

Table 1. DLS Data (Effective Hydrodynamic Diameter)

BaTiO3 Milling Time


5h 10 h 25 h
Hydrodynamic Diameter
36.2  18.4 nm 19.8  6.3 nm 484.4  301.7 nm (aggregation)
PbTiO3 Milling Time
5h 10 h 25 h 50 h
Hydrodynamic Diameter
20.1  6.5 nm 16.8  5.2 nm 16.2  6.2 nm 13.9  6.2 nm

1 August 2013 / Vol. 52, No. 22 / APPLIED OPTICS E37


a conclusion is very hypothetical, and requires con-
firmation by using independent experimental tech-
niques. Nevertheless, reported findings are quite
important for electro-optical applications of stressed
ferroelectric nanoparticles since they set a limit on
the device speed.
4. Conclusions
Physical properties of stressed ferroelectric nanopar-
ticles set a limit on the performance of optical and
electro-optical devices based on these materials.
Absorption edge of the BaTiO3 and PbTiO3 nanopar-
ticles allows their operation in the near UV–visible
range. Moreover, optical band gap of BaTiO3 sus-
Fig. 5. Rise time of the slow and fast components of the switching pensions can be tuned toward shorter wavelength
current as a function of the applied electric field for BaTiO3 sus- (blue shift of the optical band gap) by increasing
pensions. Slow component: (solid square) 5 h; (solid circle) 10 h; milling time.
(solid triangle) 50 h. Fast component: (open square) 5 h; (open Mechanical milling generates surface defects
circle) 10 h; (open triangle) 50 h. Inset shows the dependence of (mostly oxygen vacancies). These defects account
the fast component on the inverse electric field. for green photoluminescence observed at room tem-
perature for both suspensions.
Suspensions of stressed ferroelectric nanoparticles
are composed mostly of single particles and their
dimers for milling times 5 and 10 h. Further increase
in milling time enhances aggregation degree for
BaTiO3 nanoparticles, and does not affect that quan-
tity for PbTiO3 nanoparticles.
Speed of many electro-optical devices based on
stressed ferroelectric nanoparticles is governed by
the polarization switching. It was found that polari-
zation switching current exhibits two components.
One component is slow and threshold-less while an-
other component is fast and has a threshold that can
be associated with a coercive field. Particle reorien-
tation under the action of electric field accounts for
slow components of the polarization switching. Rise
Fig. 6. Rise time of the slow and fast components of the switching time of the fast component is inversely proportional
current as a function of the applied electric field for PbTiO3 sus-
to the electric field and can be attributed to the
pensions. Slow component: (solid square) 5 h; (solid circle) 10 h;
(solid triangle) 50 h. Fast component: (open square) 5 h; (open
domain frontal growth.
circle) 10 h; (open triangle) 50 h. Inset shows the dependence of We thank Dr. Dean Evans for helpful discussions.
the fast component on the inverse electric field. This work was supported by the grants from the NSF
no. 1102332 “Liquid Crystal Signal Processing Devi-
(Fig. 4). The slow component is threshold-less, while ces for Microwave and Millimeter Wave Operation”
the fast component has certain threshold. A slow and by the Biofrontiers program at the University
component is associated with particle reorientation. of Colorado.
To unfold the origin of the fast component, additional References
studies were carried out.
1. V. G. Chigrinov, Liquid Crystal Devices: Physics and Applica-
Figures 5 and 6 show the dependence of the rise tions (Artech House, 1999), pp. 21–80.
time of both the fast and slow components versus 2. Y. Reznikov, “Ferroelectric colloids in liquid crystals,” in
the applied electric field. As can be seen from the Liquid Crystals Beyond Displays: Chemistry, Physics, and
insets in Figs. 5 and 6, rise time of the fast compo- Applications, Q. Li, ed. (Wiley, 2012), pp. 402–426.
3. Y. Garbovskiy and A. Glushchenko, “Liquid crystalline col-
nent is inversely proportional to the applied electric loids of nanoparticles: preparation, properties, and applica-
field E: tions,” in Solid State Physics R. E. Camley, ed. (Academic,
2010), pp. 1–74.
1 4. Y. Reznikov, O. Buchnev, O. Tereshchenko, V. Reshetnyak, A.
tfast ∼ :
E Glushchenko, and J. West, “Ferroelectric nematic suspension,”
Appl. Phys. Lett. 82, 1917–1919 (2003).
This proportionality can be attributed to the domain 5. G. Cook, V. Yu. Reshetnyak, R. F. Ziolo, S. A. Basun, P. P.
Banerjee, and D. R. Evans, “Asymmetric Freedericksz transi-
frontal growth [26,27]. In this case, the found thresh- tions from symmetric liquid crystal cells doped with harvested
old value of the electric field can be associated with ferroelectric nanoparticles,” Opt. Express 18, 17339–17345
a coercive field of the ferroelectric material. Such (2010).

E38 APPLIED OPTICS / Vol. 52, No. 22 / 1 August 2013


6. S. A. Basun, G. Cook, V. Yu. Reshetnyak, A. V. Glushchenko, and dielectric properties for mesocrystals of BaTiO3 and
and D. R. Evans, “Dipole moment and spontaneous polariza- SrBi2 Ta2 O9 ,” J. Appl. Phys. 87, 474–478 (2000).
tion of ferroelectric nanoparticles in a nonpolar fluid suspen- 17. J. F. Scott, Comment on “Quantum confinement effects on the
sion,” Phys. Rev. B 84, 024105 (2011). optical and dielectric properties for mesocrystals of BaTiO3
7. D. R. Evans, S. A. Basun, G. Cook, I. P. Pinkevych, and V. Yu. and SrBi2 Ta2 O9 ,” J. Appl. Phys. 88, 6092 (2000).
Reshetnyak, “Electric field interactions and aggregation dy- 18. J. F. Scott, “Nano-scale ferroelectric devices for memory appli-
namics of ferroelectric nanoparticles in isotropic fluid suspen- cations,” Ferroelectrics 314, 207–222 (2005).
sions,” Phys. Rev. B 84, 174111 (2011). 19. G. Cook, J. L. Barnes, S. A. Basun, D. R. Evans, R. F. Ziolo, A.
8. Y. Garbovskiy, O. Zribi, and A. Glushchenko, “Emerging appli- Ponce, V. Reshetnyak, A. Glushchenko, and P. P. Banerjee,
cations of ferroelectric nanoparticles in materials technolo- “Harvesting single ferroelectric domain stressed nanopar-
gies, biology and medicine,” Advances in Ferroelectrics, A. ticles for optical and ferroic applications,” J. Appl. Phys.
Peláiz-Barranco, ed. (InTech, 2012), pp. 475–498. 108, 064309 (2010).
9. H. Atkuri, G. Cook, D. R. Evans, C.-I. Cheon, A. Glushchenko, 20. S. C. Roya, G. L. Sharma, and M. C. Bhatnagar, “Large blue
V. Reshetnyak, Y. Reznikov, J. West, and K. Zhang, “Prepara- shift in the optical band-gap of sol–gel derived Ba0.5 Sr0.5 TiO3
tion of ferroelectric nanoparticles for their use in liquid crys- thin films,” Solid State Commun. 141, 243–247 (2007).
talline colloids,” J. Opt. A 11, 024006 (2009). 21. E. Burstein, “Anomalous optical absorption limit in InSb,”
10. G. Cook, A. V. Glushchenko, V. Reshetnyak, A. T. Griffith, Phys. Rev. 93, 632–633 (1954).
M. A. Saleh, and D. R. Evans, “Nanoparticle doped organic- 22. T. S. Moss, “The interpretation of the properties of indium
inorganic hybrid photorefractives,” Opt. Express 16, antimonide,” Proc. Phys. Soc. B 67, 775–782 (1954).
4015–4022 (2008). 23. M.-Sh. Zhanga, Z. Yina, Q. Chena, W. Zhangb, and W. Chen,
11. W. Tscharnuter, “Photon correlation spectroscopy in particle “Study of structural and photoluminescent properties in ba-
sizing,” in Encyclopedia of Analytical Chemistry, R. A. Meyers, rium titanate nanocrystals synthesized by hydrothermal
ed. (Wiley, 2012), pp. 5469–5485. process,” Solid State Commun. 119, 659–663 (2001).
12. G. A. Cox, G. G. Roberts, and R. H. Tredgold, “The optical 24. P. Barik, T. K. Kundu, and S. Ram, “Light emission from
absorption edge of barium titanate,” Br. J. Appl. Phys. 17, ferroelectric barium titanate nanocrystals,” Philos. Mag. Lett.
743–745 (1966). 89(9), 545–555 (2009).
13. J. S. Zhu, X. M. Lu, W. Jiang, W. Tian, M. Zhu, M. S. Zhang, 25. E. Orhan, J. A. Varela, A. Zenatti, M. F. C. Gurgel, F. M. Pontes,
X. B. Chen, X. Liu, and Y. N. Wang, “Optical study on the size E. R. Leite, E. Longo, P. S. Pizani, A. Beltràn, and J. Andrès,
effects in BaTiO3 thin films,” J. Appl. Phys. 81, 1392–1395 “Room-temperature photoluminescence of BaTiO3 : joint exper-
(1997). imental and theoretical study,” Phys. Rev. B 71, 085113
14. D. Bao, Xi Yao, K. Shinozaki, and N. Mizutani, “Crystalliza- (2005).
tion and optical properties of sol–gel-derived PbTiO3 thin 26. E. Fatuzzo and W. J. Merz, “Switching mechanism in trigly-
films,” J. Phys. D 36, 2141–2145 (2003). cine sulfate and other ferroelectrics,” Phys. Rev. 116, 61–68
15. M. Cardona, “Optical properties and band structure of SrTiO3 (1959).
and BaTiO3 ,” Phys. Rev. 140, A651–A655 (1965). 27. T. Volk and M. Wöhlecke, Lithium Niobate: Defects, Photore-
16. S. Kohiki, S. Takada, A. Shimizu, K. Yamada, H. Higashijima, fraction and Ferroelectric Switching, Vol. 115, Series: Springer
and M. Mitome, “Quantum confinement effects on the optical Series in Materials Science (Springer, 2009), pp. 153–213.

1 August 2013 / Vol. 52, No. 22 / APPLIED OPTICS E39

You might also like