Está en la página 1de 5

Properties of epitaxial BaTiO3 deposited on GaAs

R. Contreras-Guerrero, J. P. Veazey, J. Levy, and R. Droopad


Citation: Applied Physics Letters 102, 012907 (2013); doi: 10.1063/1.4773988
View online: http://dx.doi.org/10.1063/1.4773988
View Table of Contents: http://scitation.aip.org/content/aip/journal/apl/102/1?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
Various evidences for the unusual polarization behaviors in epitaxially strained (111) BaTiO3
J. Appl. Phys. 115, 054105 (2014); 10.1063/1.4864218
90-degree polarization switching in BaTiO3 crystals without domain wall motion
Appl. Phys. Lett. 103, 232901 (2013); 10.1063/1.4832784
Orientation-dependent piezoelectric properties in lead-free epitaxial 0.5BaZr0.2Ti0.8O3-0.5Ba0.7Ca0.3TiO3 thin
films
Appl. Phys. Lett. 103, 122903 (2013); 10.1063/1.4821918
Magnetic and structural properties of BiFeO3 thin films grown epitaxially on SrTiO3/Si substrates
J. Appl. Phys. 113, 17D919 (2013); 10.1063/1.4796150
Growth of (111)-oriented BaTiO3Bi(Mg0.5Ti0.5)O3 epitaxial films and their crystal structure and electrical
property characterizations
J. Appl. Phys. 111, 084108 (2012); 10.1063/1.4704384

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
150.212.39.64 On: Fri, 03 Apr 2015 20:21:45

APPLIED PHYSICS LETTERS 102, 012907 (2013)

Properties of epitaxial BaTiO3 deposited on GaAs


R. Contreras-Guerrero,1 J. P. Veazey,2 J. Levy,2 and R. Droopad1,a)
1

Department of Physics, Texas State University, 601 University Drive, San Marcos, Texas 78666, USA
Department of Physics and Astronomy, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, USA

(Received 14 September 2012; accepted 17 December 2012; published online 10 January 2013)
Single crystal BaTiO3 (BTO) has been grown epitaxially on GaAs using molecular beam epitaxy
with a 2 unit cell SrTiO3 nucleation layer. The oxide film is lattice-matched to GaAs through an
in-plane rotation of 45 relative to the (100) surface leading to c-axis orientation of the BaTiO3.
X-ray diffraction confirmed the crystallinity and orientation of the oxide film with a full width half
maximum of 0.58 for a 7.5 nm thick layer. Piezoresponse force microscopy was used to
characterize the ferroelectric domains in the BaTiO3 layer, and a coercive voltage of 12 V and
C 2013 American Institute of Physics.
piezoresponse amplitude 5 pm/V was measured. V
[http://dx.doi.org/10.1063/1.4773988]
Ferroelectric layers grown directly on semiconductors
have been sought after for decades as a means of developing
non-volatile single transistor memory elements using the ferroelectric layer as the gate dielectric in a metal oxide semiconductor field effect transistor (MOSFET) device.1 More
recently, it has been proposed that the incorporation of a ferroelectric layer as part of a gate dielectric can be exploited to
reduce the subthreshold swing for ultra-low power MOSFET
devices, and an increase in the on-current of the transistor can
be achieved with the drawback of some hysteresis in the Id-Vd
characteristics.2,3 These performance enhancements are attributed to the negative capacitance effect of the ferroelectric
layer. Of course, the desire is to be able to integrate a single
crystal ferroelectric layer on a semiconductor to achieve spontaneous polarization of the ferroelectric layer that is aligned
along a single direction, preferably normal to the surface, as
opposed to a polycrystalline or amorphous layer.
Among the ferroelectric materials, BaTiO3 (BTO) is well
suited for integration with semiconductors. It is among the
most widely studied material systems for its many properties
that include piezoelectricity, ferroelectricity, pyroelectricity,
and electro-optics. BaTiO3 continues to be of interest for a
number of device applications46 and as one layer of twophase multiferroic systems.7,8 One of the challenges in integration of an oxide with semiconductors is the initial epitaxial
nucleation without oxidation of the surface. The pioneering
work by McKee et al.9 and later by Li et al.10 has allowed
for the growth of commensurate SrTiO3 (STO) on Si (100)
substrates using molecular beam epitaxy (MBE). These
approaches have demonstrated that with careful control of the
growth kinetics, it is possible to integrate oxides directly on Si
thereby opening up the possibility added functionality onto
complementary metal oxide semiconductor (CMOS) devices.
In 2003, Liang et al.11 have also demonstrated the growth of
high quality STO on GaAs (100) with a crystalline transition
across the GaAs/STO interface and a full width at half maximum (FWHM) of the STO (200) peak of 0.42 . Epitaxial
growth of BaTiO3 has also been attempted on Si(100) using
buffer layers1214 and on Ge-on-Si(001) substrates directly.15
a)

Author to whom correspondence should be addressed. Email: rdroopad@


txstate.edu. Tel: 512-245 6165.

0003-6951/2013/102(1)/012907/4/$30.00

Ferroelectric behavior was demonstrated for epitaxial BTO


grown on Si by Vaithyanathan et al.14 However, the authors
suggest that a thick buffer layer of BaSrO is needed for c-axis
oriented BaTiO3 when grown on Si. Using pulsed laser deposition (PLD), preferential c-axis oriented BaTiO3 was deposited on a MgO buffer layer that was grown on GaAs by
MBE16 and PLD.17 The use of MgO buffer layers were found
to lead to cracking of the BTO films due to the differences in
coefficients of thermal expansion. Nevertheless, ferroelectric
switching was demonstrated only when MBE was used to
grow the MgO layer that results in improved oxide/GaAs
interfaces. Huang et al.18 used PLD to epitaxially grow BTO
on GaAs using an STO buffer layer. TEM observations suggest that the BTO film growth is columnar in nature, and that
the crystal quality depends solely on the quality of the STO
buffer layer. The films were c-axis oriented and exhibited ferroelectric properties. In this Letter, we report the growth of
BaTiO3 on GaAs(100) by molecular beam epitaxy and examine its properties. BTO has a tetragonal (pseudo-cubic) crystal
and
structure with an a- and b-axis lattice constant of 3.992 A

c-axis lattice constant of 4.036 A at room temperature. Thus,


the lattice mismatch with GaAs is only 0.13% at room temperature and reduces to 0.08% at temperatures of 500  C, the typical growth temperature, when the BTO unit cell is rotated
45 on its (001) axis. Consequently, it is expected that BaTiO3
will be grown on GaAs with its c-axis aligned in the growth
direction.
The samples were grown in a multi-chamber interconnected UHV system that allows for the growth of the compound semiconductor in one chamber and transferring into the
oxide chamber with no exposure to atmosphere, thus maintaining a clean well-ordered semiconductor surface. The solid
source III-V chamber has a base pressure of <5  1010 mbar
pumped using a combination of ion and cryo pumping.
Growth rates were determined using the typical reflection
high-energy electron diffraction (RHEED) oscillation technique, and the substrate temperature was measured using an
optical pyrometer that was cross referenced using a KSA Bandit system. An n-buffer layer with thickness of order 0.5 lm
was grown on n GaAs substrates at a temperature of 580  C
and a growth rate of 0.8 lm/h. The Si doping level in the
buffer layer was in the range of 1  1017 cm3. The oxide

102, 012907-1

C 2013 American Institute of Physics


V

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
150.212.39.64 On: Fri, 03 Apr 2015 20:21:45

012907-2

Contreras-Guerrero et al.

chamber was configured with two electron beam evaporators,


seven effusion cells, and an oxygen plasma source. During the
growth using molecular oxygen, the plasma source was
switched off. The oxygen flux was measured using a chamber
ion gauge, which represents the incident oxygen flux onto the
substrate. This was determined by using a beam flux ion
gauge to correlate the incident oxygen flux with the chamber
pressure. A quartz crystal monitor was also used for flux
determination, and oxide stoichiometry and growth rate were
determined during growth using RHEED. The oxide growth
/min, and the stoichiometry
rate used was approximately 3 A
was maintained by ensuring the RHEED during growth displayed a (1  1) surface reconstruction. The oxidation states
of the various elements in the as-grown BTO layer were characterized using x-ray photoelectron spectroscopy (XPS) based
around a Scienta SES 2002 analyzer and a dual anode x-ray
source. The crystalline properties were determined using a
double crystal x-ray system and the ferroelectric properties
measured using piezoresponse force microscopy (PFM).
Once the GaAs wafer was introduced into the oxide chamber, it was heated to a temperature of 300  C. A 1=2 ML Ti was
then deposited onto the c(4  4) As-stabilized GaAs surface.
Figures 1(a) and 1(b) display the RHEED pattern along the
[010] direction showing the 4-fold surface reconstruction of
the GaAs surface at 300  C before and after the deposition of
Ti, respectively. During the Ti deposition, the reconstruction
changed from a c(4  4) to a (2  1). A 2 unit cell STO layer
was then deposit on the Ti/GaAs surface prior to the deposition
of BaTiO3 as outlined in Ref. 19. Figure 1(c) shows the
RHEED pattern along the oxide [110] azimuth after the growth
of the 2 unit cells STO layer indicating coherent nucleation of
the oxide film. Growth of the BTO layer was initiated by opening the Ba, Ti, and molecular oxygen at a substrate temperature
of 500  C and an oxygen flux level of 1  107 mbar. Figure
1(d) represents the RHEED pattern along the BTO[110] azimuth after the growth, showing streaky features with low background intensity indicative of a high quality commensurate
crystalline layer throughout the growth.

FIG. 1. RHEED images of the GaAs c(4  4) reconstruction along the [010]
azimuth (a) before and (b) after the deposition of 1=2 monolayer of Ti. The
RHEED patterns after 2 unit cells of STO and after 90 A BTO layer grown
on GaAs are shown in (c) and (d), respectively.

Appl. Phys. Lett. 102, 012907 (2013)

thick BaTiO3 on
FIG. 2. X-ray diffraction theta-2 theta spectra for a 75 A
GaAs.

To determine the crystalline structure of the BaTiO3


layer on GaAs and the relationship between the oxide film
and the semiconductor, x-ray diffraction measurements were
thick oxperformed. Figure 2 shows the h-2h scan of an 83 A
ide layer on GaAs. The thickness of the oxide layer includes
the STO nucleation layer. Initial TEM measurements have
with a
determined that the thickness of the BTO film is 75 A

8 A STO nucleation layer. The peaks present confirmed that


the oxide layer is (100) oriented. In fact, using the peak separation in the symmetric scan of the (400) peaks, the out-ofplane lattice constant for BTO was determined to be
, This value clearly demonstrates that the BTO layer
4.032 A
was grown with the c-axis oriented along the growth direction. A high resolution omega scan of the BTO (200) peak
produces a FWHM value of 0.58 confirming the extremely
oxide film. Also, results
high crystalline quality of the 75 A
of pole figure scans indicate that there is an in-plane 45
rotation of the oxide film with respect to the GaAs layer suggesting that the BTO (110) planes are parallel to the GaAs
(100) planes.
The oxidation states of the cations were determined
from photoemission experiments by transferring the BTO
film into the XPS chamber without exposure to air. Figure 3
shows typical peaks for the Ba 3d and Ti 2p for a thick

FIG. 3. XPS spectrum of Ba 3d (left) and Ti 2p (right) for a thick BaTiO3


layer grown on GaAs.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
150.212.39.64 On: Fri, 03 Apr 2015 20:21:45

012907-3

Contreras-Guerrero et al.

BaTiO3
FIG. 4. XPS spectrum of As 3d (left) and Ga 3d (right) for a 20 A
layer grown on GaAs probing the oxide/GaAs interface.

) BTO layer grown on GaAs. The Ba 3d5/2 peak has a


(>50 A
shoulder on the higher binding energy side and is similar in
shape to spectra reported for sintered and powder BTO.20,21
Figure 3 shows curve-fitting of the Ba 3d5/2 peaks with two
components, represented by I and II. While the Ba I peak
can be attributed to Ba in the perovskite oxide, there have
been conflicting arguments for the origin of the Ba II peak. It
has been attributed to surface contamination20,21 and impurity phases such as BaCO3 and Ba(OH)2.22 However, in our
experiment, the BTO was transferred into the XPS chamber
after growth through a buffer chamber with pressure less
than 1  1010 mbar, suggesting that the Ba II peak is not
due to contamination. In addition, we do not detect carbonaceous or hydrocarbon species in the XPS data. Consequently,
the origin of the Ba II peak is not clear. The Ti 2p spectrum
is indicative of Ti atoms that have been fully oxidized in the
4 oxidation state. Both the Ti 2p1/2 and the Ti 2p3/2 can be
fitted using a single peak.
To determine the chemical bonding at the oxide/semiconductor interface, XPS measurements were carried out on
thick BTO film on GaAs that was grown using an
a 20 A
identical deposition process. The thickness of the oxide layer
was chosen so that the GaAs/oxide interface could be
probed. Shown in Figure 4 are the As 3d and the Ga 3d spectra obtained from this sample. There is Ga-O bonding present
at the interface with the Ga being in a 3 oxidation state.
The high binding side of the main Ga-As peak can be fitted
with Sr 4d and O 2s levels. However there is no As-O bonding present from an analysis of the As 3d spectra. Even
though a 1=2 ML Ti prelayer was used prior to the nucleation
of the STO buffer layer, no evidence of Ti-As bonding can
be detected at the oxide/semiconductor interface from the
XPS investigations. It has been reported that the initial Ti
prelayer reacted primarily with the surface As11 but no
investigations were carried out at the interface to determine
the bonding configuration after the growth of the oxide layer.
However, based on TEM investigations of this interface,
Klie et al.23 found no evidence of the Ti. Instead, they
observed a sharp interface in which a SrO layer was in perfect registry with the As-terminated (1  1) substrate. It was
speculated that there was a re-arrangement of the atoms at
the interface during growth leading to a more energetically

Appl. Phys. Lett. 102, 012907 (2013)

favorable SrO layer next to the GaAs substrate.24 Ti was also


found to diffuse into GaAs. As a result, a more detailed study
with a combination of TEM and XPS investigations is
needed to determine the exact structure of the oxide/semiconductor interface.
PFM was used to study the ferroelectric properties of
the BTO film. It is a broadly applied technique to image and
manipulate ferroelectric domains of thin films.25 A small AC
excitation voltage Vac is applied between the sample and a
conductive tip, and the amplitude and phase of the resulting
deflection signal give, respectively, the magnitude and direction of the mechanical displacement due to the inverse piezoelectric effect. Ferroelectric domains of opposite polarization
give piezoresponses with 180 phase difference. The polarization of these ferroelectric domains is locally switched by
applying a DC bias Vdc above the coercive voltage, allowing
domain patterning when mapping Vdc onto the tip position
during scanning. The coercive voltage and characteristics of
domain formation are obtained from piezoresponse force
spectroscopy (PFS), where the piezoresponse is recorded
while fixing the tip position and sweeping Vdc.
Ferroelectric domain characterization of BTO/GaAs was
performed using vertical-PFM at room temperature in ambient conditions. The sample potential was fixed, and voltages
were applied to the tip. Values for all voltages are given here
with respect to the sample (V VsampleVtip). During PFS, in
order to improve signal-to-noise ratio and to reduce crosstalk
between surface topography and piezoresponse, the feedback
loop reads the difference in amplitude of two simultaneous
driving voltages with frequencies slightly above and below
resonance.26 Figure 5 shows PFM characterization of the
BTO/GaAs surface after ferroelectric domain patterning of a
square ring. The center square and outer ring were patterned
with Vdc 3 V, while the inner ring was patterned with
Vdc 3 V. Simultaneous height and piezoresponse scans
(Figs. 5(a)5(c)) show that the imaged pattern is due entirely
to the piezoresponse and not cross-coupling to the surface topography. PFS amplitude and phase curves (Figs. 5(d) and
5(e)) show that the polarization is repeatedly switchable. The
BTO shows a piezoresponse amplitude 5 pm/V, and a coercive voltage of Vc 12 V, with respect to the sample, which
was grounded. Similar patterning and characterization were
repeated several times on different areas of the sample with
similar results.
The coercive voltages observed in the PFM hysteresis
loops have a small offset along the Vdc axis. Similar offsets
have been observed previously in PFM characterization of ferroelectric/semiconductor junctions27 and can be attributed to a
voltage offset arising from the contribution of interface
dipoles.28 It is also possible that a small imprint effect from
epitaxial strain at the BTO/GaAs interface could contribute.29
The PFM signal was indeed dominated by the piezoresponse
of ferroelectric domains and not by coupling to surface or
interface charges,30 which is demonstrated by the fact that
domains of opposite polarity routinely had a 180 phase difference (Figs. 5(c) and 5(e)). Moreover, patterns imaged additional times after 1 h still exhibited stable ferroelectric
domains. The surrounding region beyond the outer ring was
not patterned but still shares the same phase as the inner ring,
which was patterned with Vdc 3 V. (Figs. 5(b) and 5(c)).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
150.212.39.64 On: Fri, 03 Apr 2015 20:21:45

012907-4

Contreras-Guerrero et al.

Appl. Phys. Lett. 102, 012907 (2013)

FIG. 5. (a) height, (b) PFM amplitude, and (c) PFM phase images acquired simultaneously after patterning a 3 lm  3 lm square ring by applying 63 V DC
bias (Scale bars: 1 lm and Vac 500 mV). The inner ring in (b) and (c) corresponds to where 3 V sample bias was applied, while the outer ring and center
square correspond to where 3 V was applied. (d) Amplitude and (e) phase curves acquired during PFM spectroscopy (Vac 4.0 V).

This result indicates that as-deposited BTO is pre-poled


(pointing toward the interface), consistent with X-ray diffraction data showing the c-axis aligned out-of-plane (Fig. 2).
In conclusion, high quality single crystal c-oriented
BaTiO3 was grown on n-type GaAs using molecular beam
epitaxy by co-deposition and molecular oxygen. RHEED
and x-ray diffraction confirmed the crystallinity with an out . The Ba and Ti were
of-plane lattice constant of 4.032 A
fully oxidized and the oxide/semiconductor interface showed
evidence of Ga-O bonding. The properties of the BTO layer
were determined using piezoresponse force microscopy, and
the results indicated that the oxide was ferroelectric, with a
coercive voltage of 12 V and a piezoresponse amplitude of
around 5 pm/V.
The authors would like to thank Manhoi Wong for help
with the XRD measurements. This work is supported by the
AFOSR under Grant Nos. FA9550-10-1-0133 (R.D.) and
NSF DMR-1104191 (J.L.).
1

J. Scott, Ferroelectric Memories (Springer, Berlin, 2000).


S. Salahuddin and S. Datta, Nano Lett. 8, 405 (2008).
3
A. I. Khan, C. W. Yeung, C. Hu, and S. Salahuddin, 2011 IEDM digest
2011, 255.
4
R. Waser, Nanoelectronics and Information Technology (Wiley,
2003).
5
N. Yanase, K. Abe, N. Fukushima, and T. Kawakubo, Jpn. J. Appl. Phys.,
Part 1 38, 5305 (1999).
6
Y. S. Kim, D. H. Kim, J. D. Kim, Y. J. Chang, T. W. Noh, J. H. Kong, K.
Char, Y. D. Park, S. D. Bu, J.-G. Yoon, and J.-S. Chung, Appl. Phys. Lett.
86, 102907 (2005).
7
S. Sahoo, S. Polisetty, C. Duan, S. Jaswal, E. Tsymbal, and C. Binek,
Phys. Rev. B 76, 092108 (2007).
8
M. K. Lee, T. K. Nath, C. B. Eom, M. C. Smoak, and F. Tsui, Appl. Phys.
Lett. 77, 3547 (2000).
9
R. A. McKee, F. J. Walker, and M. F. Chisholm, Phys. Rev. Lett. 81, 3014
(1998).
2

10

H. Li, X. Hu, Y. Wei, Z. Yu, X. Zhang, R. Droopad, A. A. Demkov,


J. Edwards, K. Moore, W. Ooms, J. Kulik, and P. Fejes, J. Appl. Phys. 93,
4521 (2003).
11
Y. Liang, J. Kulik, T. Eschrich, R. Droopad, Z. Yu, and P. Maniar, Appl.
Phys. Lett. 85, 1217 (2004).
12
R. A. McKee, F. J. Walker, J. R. Conner, E. D. Specht, and D. E. Zelmon.
Appl. Phys. Lett. 59, 782 (1991).
13
Z. Yu, J. Ramdani, J. A. Curless, C. D. Overgaard, J. M. Finder, R.
Droopad, K. W. Eisenbeiser, J. A. Hallmark, W. J. Ooms, and V. S.
Kaushik, J. Vac. Sci. Technol. B 18, 2139 (2000).
14
V. Vaithyanathan, J. Lettieri, W. Tian, A. Sharan, A. Vasudevarao, Y. L.
Li, A. Kochhar, H. Ma, J. Levy, P. Zschack, J. C. Woicik, L. Q. Chen, V.
Gopalan, and D. G. Schlom, J. Appl. Phys. 100, 24108 (2006).
15
C. Merckling, G. Saint-Girons, C. Botella, G. Hollinger, M. Heyns,
J. Dekoster, and M. Caymax, Appl. Phys. Lett. 98, 92901 (2011).
16
T. E. Murphy, D. Chen, and J. D. Phillips, Appl. Phys. Lett. 85, 3208 (2004).
17
D. Chen, T. E. Murphy, S. Chakrabarti, and J. D. Phillips, Appl. Phys.
Lett. 85, 5206 (2004).
18
W. Huang, Z. P. Wu, and J. H. Hao, Appl. Phys. Lett. 94, 032905 (2009).
19
R. Contreras-Guerrero, M. Edirisooriya, O. C. Noriega, and R. Droopad,
Interface properties of MBE grown epitaxial oxides on GaAs, J. Cryst.
Growth (submitted).
20
S. Kumar, V. S. Raju, and T. R. N. Kutty, Appl. Surf. Sci. 206, 250 (2003).
21
S. Mukhopadhyay and T. C. S. Chen, J. Mater. Res. 10, 1502 (1995).
22
C. C. Hung, R. E. Riman, and R. Caracciolo, Ceramic Powder Science III,
American Ceramic Society Vol. 12, edited by G. L. Messing, S. Hirano,
and H. Hausner (1990), p. 17.
23
Q. Qiao, R. F. Klie, S. Ogut, and J. C. Idrobo, Phys. Rev. B 85, 165406 (2012).
24
R. F. Klie, Y. Zhu, E. I. Altman, and Y. Liang, Appl. Phys. Lett. 87,
143106 (2005).
25
S. V. Kalinin, A. N. Morozovska, L. Q. Chen, and B. J. Rodriguez, Rep.
Prog. Phys. 73, 056502 (2010).
26
B. J. Rodriguez, C. Callahan, S. V. Kalinin, and R. Proksch, Nanotechnology 18, 475504 (2007).
27
M. P. Warusawithana, C. Cen, C. R. Sleasman, J. C. Woicik, Y. L. Li, L.
F. Kourkoutis, J. A. Klug, H. Li, P. Ryan, L. P. Wang, M. Bedzyk, D. A.
Muller, L. Q. Chen, J. Levy, and D. G. Schlom, Science 324, 367 (2009).
28
R. A. McKee, F. J. Walker, M. B. Nardelli, W. A. Shelton, and G. M.
Stocks, Science 300, 1726 (2003).
29
Y. Saya, S. Watanabe, M. Kawai, H. Yamada, and K. Matsushige, Jpn. J.
Appl. Phys., Part 1 39, 3799 (2000).
30
R. Shao, M. P. Nikiforov, and D. A. Bonnell, Appl. Phys. Lett. 89, 112904
(2006).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
150.212.39.64 On: Fri, 03 Apr 2015 20:21:45

También podría gustarte