Вы находитесь на странице: 1из 6

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

Nanoscale Dynamic Article Links < C

Cite this: Nanoscale, 2012, 4, 408


www.rsc.org/nanoscale COMMUNICATION
High resolution quantitative piezoresponse force microscopy of BiFeO3
nanofibers with dramatically enhanced sensitivity
Published on 21 November 2011. Downloaded by University of Windsor on 31/10/2014 06:56:21.

Shuhong Xie,ab Anil Gannepalli,c Qian Nataly Chen,b Yuanming Liu,b Yichun Zhou,a Roger Prokschc
and Jiangyu Li*b
Received 14th August 2011, Accepted 26th October 2011
DOI: 10.1039/c1nr11099c

Piezoresponse force microscopy (PFM) has emerged as the tool of dimensional ferroelectrics, the thickness is extremely small, typically
choice for characterizing piezoelectricity and ferroelectricity of low- in the range of nanometres, and thus only a small AC driving voltage
dimensional nanostructures, yet quantitative analysis of such low- can be applied to the specimen to avoid electric breakdown. In
dimensional ferroelectrics is extremely challenging. In this addition, the piezoelectric coefficient is known to decrease with the
communication, we report a dual frequency resonance tracking feature size of ferroelectric nanostructures, and is usually much
technique to probe nanocrystalline BiFeO3 nanofibers with smaller than that of bulk materials.1215 As a result, the intrinsic
substantially enhanced piezoresponse sensitivity, while simulta- piezoresponse amplitude A d33V0 of low-dimensional ferroelectric
neously determining its piezoelectric coefficient quantitatively and nanostructures is extremely small, often in the picometre region. This
correlating quality factor mappings with dissipative domain raises a formidable task for detecting piezoresponse by photodiodes,
which have a lower bound on detectable cantilever deflection of
switching processes. This technique can be applied to probe the
around 30 pm, limited by thermal noise and shot noise.16,17 Clearly, in
piezoelectricity and ferroelectricity of a wide range of low-dimen-
order to probe the piezoelectricity and ferroelectricity of low-
sional nanostructures or materials with extremely small piezoelec-
dimensional ferroelectric nanostructures or materials with extremely
tric effects.
small piezoelectric effect such as biopolymers,18 the piezoresponse of
the specimen has to be substantially magnified, and the PFM sensi-
Piezoresponse force microscopy (PFM) is a powerful tool for probing tivity has to be dramatically enhanced.
local piezoelectric and ferroelectric properties of ferroelectric mate- The piezoresponse and PFM sensitivity can usually be enhanced
rials at the nanoscale level,13 and has emerged as the tool of choice by driving the AC modulation voltage near the resonance of the
for characterizing piezoelectricity and ferroelectricity of low-dimen- cantileverspecimen system,1921 and thus magnify the piezoresponse
sional ferroelectric nanostructures.49 In a typical PFM measurement, amplitude A d33V0Q by orders of magnitude, where Q is the quality
an AC driving voltage Vac V0eiut is applied to the specimen through factor of the system typically in the range of 10 to 100 s. This makes it
the conductive cantilever tip, which induces a surface vibration uac possible to probe ferroelectric nanostructures using PFM, but such
d33V0ei(ut+4) of the specimen due to its piezoelectricity, where d33 is the a resonance enhancement comes with a penalty that compromises
out-of-plane piezoelectric coefficient of the material, and 4 is the quantitative analysisthe quality factor Q is usually unknown,
phase contrast reflecting its ferroelectric polarity. The deflection of varying from sample to sample and even varying from point to point
the cantilever induced by the piezoelectric vibration of the specimen, in the same sample. As a result, the piezoelectric coefficient d33 can no
the so-called piezoresponse, is then measured by a photodetector. longer be determined quantitatively for ferroelectric nanostructures,
Clearly, PFM offers nanoscale resolution limited only by the size of even with careful calibration against standard references, since the
the cantilever tip, and is an ideal tool for probing local piezoelectric quality factors would be different. In fact, it even becomes ambiguous
and ferroelectric properties of low-dimensional ferroelectric nano- to make a meaningful qualitative comparison of piezoresponses from
structures.411 In fact, no other technique is capable of probing sample to sample. This makes it extremely difficult to probe, for
piezoelectricity and ferroelectricity directly at the nanoscale level. example, the size effects in ferroelectric nanostructures using PFM,
The AC driving voltage of PFM is limited by the breakdown since reliable quantitative information can no longer be extracted.
strength Ebd of the specimen and its thickness t, V0 < Ebdt. For low- For that matter, it has been noticed that the piezoelectric coefficient
d33 correlates with the spontaneous polarization through the elec-
a
Faculty of Materials, Optoelectronics and Physics, and Key Laboratory of trostrictive coefficient,22 one of the key parameters of interest to
Low Dimensional Materials and Application Technology of Ministry of ferroelectric nanostructures, and there are no other techniques to
Education, Xiangtan University, Xiangtan, Hunan, 411105, China probe the spontaneous polarization directly with nanoscale
b
Department of Mechanical Engineering, University of Washington, resolution.
Seattle, WA, 98195-2600, USA. E-mail: jjli@uw.edu
c In this communication, we report a dual frequency resonance
Asylum Research, 6310 Hollister Ave., Santa Barbara, California, 93117,
USA tracking (DFRT) technique that allows us to magnify the piezores-
These authors contributed equally to the work. ponse sensitivity by orders of magnitude while determining its quality

408 | Nanoscale, 2012, 4, 408413 This journal is The Royal Society of Chemistry 2012
View Article Online

factor Q simultaneously, and thus enable high resolution quantitative point, such resonance enhancement is achieved at the expense of
PFM for low-dimensional ferroelectric nanostructures with high quantitative information.20
sensitivity. To understand the technique, let us examine how reso- To overcome this difficulty, let us examine the cantileversample
nance enhancement works first. The cantilever and sample can be system in Fig. 1(a) in more detail. For such a damped harmonic
viewed as a damped harmonic oscillator driven by an applied electric oscillator, we have four unknown parameters: resonance frequency
field Vac, as shown in Fig. 1(a). Due to the piezoelectricity of the u0, quality factor Q, amplitude An and phase shift fn, both to be
ferroelectric sample, a surface vibration uac is induced, which in turn determined at resonance frequency u0. Yet we only have two
drives the vibration of the cantilever that is in contact with the sample measurements, the amplitude Ad and phase shift fd measured at
surface through the tip. The deflection of the cantilever z is governed driving frequency ud, making it impossible to solve for the four
Published on 21 November 2011. Downloaded by University of Windsor on 31/10/2014 06:56:21.

by the following motion equation:23 unknown parameters from eqn (3) and (4). However, if we measure
u0 the amplitude A1,2 and phase shift f1,2 at two distinct frequencies u1,2
z z_ u20 z d33 V0 eiut4 ; (1) across resonance, as schematically shown in Fig. 1(c) using actual
Q
experimental data, then four measurements will be obtained, making
where u0 is the resonance frequency of the harmonic oscillator. The it possible to solve for u0, Q, An, and fn simultaneously from the
steady state solution for the deflection is given by: measured A1,2 and f1,2 using eqn (3) and (4). This is the essence of our
proposed approach, which would enable the quantitative PFM
z Aei(ut+4+f), (2) together with dramatically enhanced sensitivity, a combination that
would be impossible to achieve under the conventional single
with the piezoresponse amplitude A given by:
frequency resonance enhancement technique. As evident from Fig. 1
d33 V0 (c), eqn (3) and (4) fit the experimental data very well, especially for
A q; (3)
1  u =u20 u=u0 Q2
2 2 the amplitude data, where the error between the fitted resonance peak
and highest response measured is less than 5%. This validates the
and its phase shift f given by: quantitative analysis with DFRT. While eqn (3) and (4) cannot be
p h i solved analytically in general, they can be easily solved by numerical
f tan1 Qu=u0  u0 =u : (4) iterations.
2
The dual frequencies adopted here also allow us to overcome other
From eqn (3), it is clear that if a quasi-static voltage is applied, the tip difficulties associated with resonance enhancement techniques,
deflection is given by A d33V0, while if the AC modulation voltage particularly the cross-talk with topography and the loss of sensitivity
is driven at resonance frequency, the tip deflection is given by A when contact stiffness varies while scanning over sample surfaces.
d33V0Q, magnified by a factor of Q compared to the quasi-static For the single frequency resonance enhancement technique, the
response, as schematically shown in Fig. 1(b) and confirmed by our frequency is locked at the resonance in a particular location. As the
fitting of experimental data using eqn (3) and (4). This is the principle resonance shifts during scanning due to the change of contact stiff-
behind the resonance enhancement technique. Since Q is generally ness, such a lock-in will actually drive the oscillator away from the
unknown, and usually varies from sample to sample or point to resonance at the current location, resulting in loss of enhanced
sensitivity, as schematically shown in Fig. 1(b). This also leads to
cross-talk with topography, and severely compromises PFM resolu-
tion.20,21,24,25 The dual frequencies adopted, on the other hand, allow
us to track resonance when it varies over the sample surface and
minimize the cross-talk with topography, as schematically shown in
Fig. 1(c). Instead of locking in at a single frequency, we have two
frequencies u1,2 on both sides of the resonance, and use the difference
in the amplitudes, A1  A2, for the feedback control, which is
maintained to be zero in our implementation.21,26 When the resonance
shifts during scanning, the amplitude difference deviates from zero,
and the dual frequencies are adjusted through feedback control to
bring this difference back to zero. This allows us to track the reso-
nance frequency during PFM scanning and minimize cross-talk with
topography, while simultaneously enables high resolution quantita-
tive PFM with high sensitivity. Other approaches developed to track
resonance include frequency sweeps27 and band excitation28
techniques.
To demonstrate the concept, we applied the proposed quantitative
DFRT PFM to probe a single BiFeO3 nanofiber synthesized using
solgel based electrospinning.2931 BiFeO3 belongs to the so-called
Fig. 1 Schematics of dual frequency resonance tracking (DFRT) tech-
multiferroic materials,32,33 and ferroelectricity is essential for its
nique: (a) tip-nanofiber harmonic oscillation system; (b) resonance magnetoelectric coupling. The fiber has a diameter around 200 nm,
enhancement using single frequency; and (c) schematics of dual frequency and due to such a small size, PFM is the only method capable of
resonance tracking using actual experimental data of piezoresponses probing its piezoelectricity and ferroelectricity at the nanoscale level
versus AC driving frequency. directly. A 1 mm by 1 mm PFM scan was carried out using DFRT

This journal is The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 408413 | 409
View Article Online

under an AC modulation voltage of 1.76 V, from which mappings of the maximum piezoresponse amplitude, after the correction, is only
amplitude A1,2 and phase shift f1,2 are obtained under u1,2. These two 15.3 pm, and without the resonance enhancement, this would be
frequencies vary spatially due to the variation of contact stiffness of the impossible to detect! The dual frequency resonance tracking (DFRT)
sample, and are adjusted to track the resonance via feedback control. technique, on the other hand, makes it possible to not only detect such
This allows us to solve for u0, Q, fn, and An at each point and map their small piezoresponse, but also quantitatively determine its intrinsic
spatial distributions, as shown in Fig. 2, where the resonance frequency piezoelectric coefficient, which is estimated to be 8.70 pm V1.
u0 (a), quality factor Q (b), phase shift fn (c), and amplitude An/Q (d) The value of this piezoelectric coefficient is much smaller than that
mappings are imposed on top of the three-dimensional nanofiber of bulk BiFeO3 for a couple of reasons. First of all, the grains in the
topography. Indeed, it is observed in Fig. 2(a) that the resonance nanofiber are randomly oriented, and their spontaneous polarization
Published on 21 November 2011. Downloaded by University of Windsor on 31/10/2014 06:56:21.

frequencies vary over the surface of the nanofiber, ranging between prefers to be aligned in-plane to minimize the depolarization energy,34
227 kHz and 235 kHz. While this seems to be a rather small variation, it resulting in a much smaller out-of-plane piezoelectric response. In
can result in a reduction of sensitivity by one order of magnitude if addition, the local piezoelectric response of the nanofiber underneath
a conventional single frequency resonance enhancement technique is the conductive tip is severely constrained by its neighbors that are
used during scanning. For example, if we lock in at 235 kHz, while the subjected to a much weaker electric field, and thus is much smaller
actual resonance at a particular point is 227 kHz, then the piezores- than the piezoelectric response of a bulk material or thin film sub-
ponse is approximately 15d33V according to eqn (3) for a typical jected to a homogeneous electric field. Nevertheless, the small pie-
quality factor of 100 instead of 100d33V as expected at resonance. This zoresponse of the BiFeO3 nanofiber raises important questionsis it
would not only result in loss of sensitivity, but also lead to false contrast ferroelectric, and is it possible to switch its spontaneous polarization
in piezoresponse mapping. In addition to the relative small variation in using a conductive PFM tip? To answer these questions, we applied
resonance frequency, we also observe a large variation in quality factor a series of DC voltages up to 60 V in two complete cycles as shown in
Q ranging between 60 and 90, as shown in Fig. 2(b). This underscores Fig. 3(a) and (b) at two different spots marked by 1 and 2 on the
the importance of accurate determination of quality factor for the nanofiber, and probed the corresponding piezoresponse using a AC
quantitative PFM characterization, without which even a qualitative voltage of 1.76 V before and after the application of cyclic electric
comparison of piezoresponse at different points of a sample would be loading.35 Before the application of electric field, the two spots have
difficult to interpret. The PFM phase contrast is clearly observed in opposite phases, as seen in Fig. 3(c), indicating upward polarization
Fig. 2(c) and corresponds to the amplitude mapping in Fig. 2(d) well, (purple) at 1 and downward polarization (yellow) at 2. Thus, one of
indicating different polarity in different domains. No obvious cross- the spots should have its phase contrast reversed after the electric
talk with topography is observed, and it appears that the domains in loading, if the polarization is switched by the conductive tip. Inter-
the nanofiber span multiple grains, indicating a strong intergranular estingly, the phase contrasts of these two spots remain unchanged
interaction. Most importantly, the intrinsic piezoresponse of the after the electric cycle, as seen in Fig. 3(d). This again suggests that the
nanofiber can be now determined by correcting the quality factor Q, nanoscale electromechanical response of the BiFeO3 nanofiber is
making it possible to map the spatial distribution of piezoelectric severely constrained by its neighboring area, making it rather difficult
coefficient d33 quantitatively, as shown in Fig. 2(d). It is observed that to have either local piezoresponse or domain switching due to this

Fig. 2 Mapping of (a) frequency, (b) quality factor, (c) phase shift, and (d) amplitude of a BiFeO3 nanofiber superimposed on its three-dimensional
topography, derived using the DFRT technique proposed.

410 | Nanoscale, 2012, 4, 408413 This journal is The Royal Society of Chemistry 2012
View Article Online

and after domain switching is shown in Fig. 4(b) and (c). While there
is no clear correlation between mappings of phase contrast change
and absolute values of quality factor, the quality factor is clearly
reduced in areas where domains are switched, confirming increased
dissipation due to domain switching. In areas away from these
switched boxes, little change in the quality factor is observed.
The ferroelectricity of the nanofiber is also confirmed by the pie-
zoresponse phasevoltage hysteresis loop and amplitudevoltage
butterfly loop as shown in Fig. 5, obtained under a series of DC
Published on 21 November 2011. Downloaded by University of Windsor on 31/10/2014 06:56:21.

voltages similar to those shown in Fig. 3(a) and (b). While the PFM
phase is reversed when the DC voltage is reversed as shown in Fig. 5
(a), indicating a possible polarization reversal of ferroelectricity, the
PFM amplitude appears to be linear with respect to the applied
voltage and does not saturate even at large DC voltages up to 60 V.
Such a linear relationship is usually caused by strong electrostatic
interactions between the conductive tip and sample, but we believe
this is not the case here. The PFM switching measurement is carried
Fig. 3 Polarization switching in BiFeO3 nanofibers: (a) the DC bias out when the DC voltage is brought back to zero at each step, as
voltage applied to the cantilever tip as a function of time; (b) the detailed shown in Fig. 3(b), and thus the effect of electrostatic interactions is
variation of the DC bias voltageat every step, the DC bias is reduced to
minimized. Indeed, a zoom in on the amplitudevoltage loop shows
zero, and then increases to the next step; (c and d) the PFM phase image
characteristics of butterfly loop, and a relatively large coercivity of
of BiFeO3 nanofiber before (c) and after (d) the electric cycle.
3.5 V is observed in the hysteresis and butterfly loops, clearly shown
in the zoomed hysteresis and butterfly loops near switching in Fig. 5
constraint.36 The polarization reversal may still occur during the (c) and (d). This is a strong indication of ferroelectricity instead of
application of electric field, yet the switched domain is simply electrostatic interactions. The overall amplitudevoltage characteris-
unstable after the removal of electric field, and reverses back to its tics appear to be linear, which we believe is due to the difficulty in
original configuration. Nevertheless, ferroelectricity of the nanofiber polarization switching in the nanofiber as observed in Fig. 3(c) and
is confirmed by domain switching in areas adjacent to these two (d). As such, the switching has to proceed in a gradual incremental
spots, as highlighted in the marked boxes before and after the electric manner, instead of abrupt switching as in a bulk material. In nano-
cycle. Note that polarizations in boxes 3, 5, and 7 are switched from crystalline rhombohedral ferroelectrics such as BiFeO3, such an
upward to downward, while areas in boxes 4 and 6 are switched from incremental switching is caused not only by the two-step polarization
downward to upward, suggesting that different domain stabilities reversal process involving 71 and 109 domain switching, but also by
exist in different areas. These areas presumably have less constraint gradual polarization reversal in individual grains within a single
and more favorable orientation for domain switching, and their domain that spans multiple grains. As a result, the piezoresponse
switching also indicates that the electric field is not confined to the amplitude does not saturate even at large electric fields. Indeed, with
conductive tip. It is also interesting to note that box 3 is more than a large DC voltage of 60 V, a rather small piezoresponse is measured,
200 nm away from spot 1, where the electric field was applied, indi- only in the range of 120 pm, far smaller than the saturated piezor-
cating rather long range interactions among grains and domains. esponse expected if the domain is fully switched.
Domain switching is a dissipative process that increases dissipation In summary, a dual frequency resonance tracking (DFET) tech-
and decreases quality factor, which can also be investigated using nique is adopted to enable high resolution quantitative piezoresponse
DFRT PFM. This is indeed confirmed by the mappings of quality force microscopy of nanocrystalline BiFeO3 nanofibers with high
factor before and after application of electric field, as shown in Fig. 4. sensitivity. The BiFeO3 nanofiber has extremely small out-of-plane
The mapping of phase contrast change is shown in Fig. 4(a), with five piezoresponse, which would be rather difficult to probe by conven-
boxes highlighted, where the phase contrast change is close to 180 , tional techniques. The dual frequency resonance tracking (DFET)
a signature of domain switching; this is how we identify the five boxes technique, on the other hand, allows us to map the fine ferroelectric
shown in Fig. 3. The corresponding mapping of quality factor before domains in the nanocrystalline BiFeO3 nanofiber in detail, determine

Fig. 4 Correlation between mapping of phase contrast and quality factors: (a) change of phase contrast; (b) quality factor before the electric cycle; and
(c) quality factor after the electric cycle.

This journal is The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 408413 | 411
Published on 21 November 2011. Downloaded by University of Windsor on 31/10/2014 06:56:21. View Article Online

Fig. 5 PFM phasevoltage hysteresis (a) and amplitudevoltage butterfly (b) loops of BiFeO3 nanofiber; (c) and (d) show the zoomed hysteresis and
butterfly loops near the coercive voltage.

its out-of-plane piezoelectric coefficients quantitatively, and corre- 9 Y. M. Liu, D. N. Weiss and J. Y. Li, ACS Nano, 2009, 4, 8390.
lating the change of quality factor mappings with the dissipative 10 R. Nath, Y. H. Chu, N. A. Polomoff1, R. Ramesh and B. D. Huey,
Appl. Phys. Lett., 2008, 93, 072905.
domain switching process. The proposed approach can be applied to 11 Y. Ivry, D. P. Chu and C. Durkan, Appl. Phys. Lett., 2009, 94,
probe the piezoelectricity and ferroelectricity of a wide range of low- 162903.
dimensional nanostructures, as well as materials with extremely small 12 B. Jiang, J. L. Peng, L. A. Bursill and W. L. Zhong, J. Appl. Phys.,
piezoelectric responses. 2000, 87, 34623467.
13 H. Fujisawa, M. Okaniwa, H. Nonomura, M. Shimizu and H. Niu, J.
Eur. Ceram. Soc., 2004, 24, 16411645.
Acknowledgements 14 I. Ponomareva, I. I. Naumov and L. Bellaiche, Phys. Rev. B: Condens.
Matter Mater. Phys., 2005, 72, 214118.
We acknowledge the support of US National Science Foundation 15 V. M. Fridkin, Physics-Uspekhi, 2006, 49, 193202.
16 A. N. Morozovska, E. A. Eliseev, S. L. Bravina and S. V. Kalinin,
(DMR-1006194 and CMMI-1100339) and Natural Science Foun- Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 174109
dation of China (Approval nos. 10972189, 10902095, and 11090331). 174118.
S. H. Xie also acknowledges the partial support of China Scholarship 17 T. Jungk, A. Hoffmann and E. Soergel, Appl. Phys. Lett., 2006, 89,
163507.
Council, Q.N. Chen acknowledges the partial support of NASAs
18 S. V. Kalinin, B. J. Rodriguez, S. Jesse, T. Thundat and
Space Technology Research Fellowship, and Y. M. Liu acknowl- A. Gruverman, Appl. Phys. Lett., 2005, 87, 053901.
edges the partial support of a UIF Fellowship from the Center for 19 H. Okino, J. Sakamoto and T. Yamamoto, Jpn. J. Appl. Phys., 2003,
Nanotechnology, University of Washington. 42, 62096213.
20 S. Jesse, B. Mirman and S. V. Kalinin, Appl. Phys. Lett., 2006, 89,
022906.
References 21 B. J. Rodriguez, C. Callahan, S. V. Kalinin and R. Proksch,
Nanotechnology, 2007, 18, 475504475506.
1 H. F. Yu, H. R. Zeng, R. Q. Chu, G. R. Li and Q. R. Yin, J. Inorg. 22 C. Harnagea, A. Pignolet, M. Alexe and D. Hesse, Integr. Ferroelectr.,
Mater., 2005, 20, 257266. 2002, 44, 113124.
2 A. Gruverman and S. V. Kalinin, J. Mater. Sci., 2006, 41, 107116. 23 T. R. Albrecht, P. Grutter, D. Horne and D. Rugar, J. Appl. Phys.,
3 D. A. Bonnell, S. V. Kalinin, A. L. Kholkin and A. Gruverman, MRS 1991, 69, 668673.
Bull., 2009, 34, 648657. 24 C. Harnagea, M. Alexe, D. Hesse and A. Pignolet, Appl. Phys. Lett.,
4 A. Baji, Y. W. Mai, Q. Li and Y. Liu, Nanoscale, 2011, 3, 30683071. 2003, 83, 338340.
5 A. Kholkin, N. Amdursky, I. Bdikin, E. Gazit and G. Rosenman, 25 S. Jesse, A. P. Baddorf and S. V. Kalinin, Nanotechnology, 2006, 17,
ACS Nano, 2010, 4, 610614. 16151628.
6 K. Kathan-Galipeau, P. P. Wu, Y. L. Li, L. Q. Chen, A. Soukiassian, 26 R. Proksch, Appl. Phys. Lett., 2006, 89, 113121.
X. X. Xi, D. G. Schlom and D. A. Bonnell, ACS Nano, 2011, 5, 640 27 A. B. Kos and D. C. Hurley, Meas. Sci. Technol., 2008, 19, 015504.
646. 28 S. Jesse, S. V. Kalinin, R. Proksch, A. P. Baddorf1 and
7 S. H. Xie, F. Y. Ma, Y. M. Liu and J. Y. Li, Nanoscale, 2011, 3, 3152 B. J. Rodriguez, Nanotechnology, 2007, 18, 435503.
3158. 29 S. H. Xie, J. Y. Li, R. Proksch, Y. M. Liu, Y. C. Zhou, Y. Y. Liu,
8 B. J. Rodriguez, X. S. Gao, L. F. Liu, W. Lee, I. I. Naumov, Y. Ou, L. N. Lan and Y. Qiao, Appl. Phys. Lett., 2008, 93, 222904.
A. M. Bratkovsky, D. Hesse and M. Alexe, Nano Lett., 2009, 9, 30 F. Yao, L. Xu, B. Lin and G. D. Fu, Nanoscale, 2010, 2, 1348
11271131. 1357.

412 | Nanoscale, 2012, 4, 408413 This journal is The Royal Society of Chemistry 2012
View Article Online

31 Z. Y. Hou, Z. Y. Cheng, G. G. Li, W. X. Wang, C. Peng, C. X. Li, D. M. Kim, S. H. Baek, C. B. Eom and R. Ramesh, Nat. Mater.,
P. A. Ma, D. M. Yang, X. J. Kang and J. Lin, Nanoscale, 2011, 3, 2006, 5, 823829.
15681574. 34 A. Schilling, R. M. Bowman, G. Catalan, J. F. Scott and J. M. Gregg,
32 D. P. Dutta, O. D. Jayakumar, A. K. Tyagi, K. G. Girija, Nano Lett., 2007, 7, 37873791.
C. G. S. Pillai and G. Sharma, Nanoscale, 2010, 2, 1149 35 S. Jesse, A. P. Baddorf and S. V. Kalinin, Appl. Phys. Lett., 2006, 88,
1154. 062908.
33 T. Zhao, A. Scholl, F. Zavaliche, K. Lee, M. Barry, A. Doran, 36 J. Y. Li, R. C. Rogan, E. Ustundag and K. Bhattacharya, Nat.
M. P. Cruz, Y. H. Chu, C. Ederer, N. A. Spaldin, R. R. Das, Mater., 2005, 4, 776781.
Published on 21 November 2011. Downloaded by University of Windsor on 31/10/2014 06:56:21.

This journal is The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 408413 | 413

Вам также может понравиться