Japanese Journal of Applied Physics Vol. 46, No. 9A, 2007, pp. 5674–5685

Review Paper

#2007 The Japan Society of Applied Physics

Recent Advances in Electromechanical Imaging on the Nanometer Scale: Polarization Dynamics in Ferroelectrics, Biopolymers, and Liquid Imaging Sergei V. K ALININ1;2 , Stephen J ESSE1 , Brian J. RODRIGUEZ1;2 , Katyayani SEAL2 , Arthur P. BADDORF2 , Tong ZHAO3 , Y. H. C HU3 , Ramamoorthy R AMESH3 , Eugene A. ELISEEV4 , Anna N. M OROZOVSKA4 , B. M IRMAN5 , and Edgar KARAPETIAN5 1

Materials Sciences and Technology Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, U.S.A. Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, TN 37831, U.S.A. 3 Department of Materials Science and Engineering, University of California, Berkeley, CA 94720, U.S.A. 4 National Academy of Science of Ukraine, Kiev 03028, Ukraine 5 Department of Mathematics and Computer Science, Suffolk University, Boston, MA 02114, U.S.A. 2

(Received February 13, 2007; accepted April 22, 2007; published online September 7, 2007)

Coupling between electrical and mechanical phenomena is ubiquitous in nature, with examples ranging from piezoelectricity in polar perovskites and chemical bonds to complex pathways of electromechanical transformations underpinning the functionality of electromotor proteins, cells, and tissues. Piezoresponse force microscopy (PFM) had originally emerged as a technique to study electromechanical phenomena in ferroelectric perovskites on the nanoscale. In recent years, the applicability of PFM for studying a broad range of non-ferroelectric polar materials has been demonstrated, necessitating further development of the technique, including theory of the image formation mechanism as well as probe and controller development. Here, we review the basic principles of PFM and summarize some of the recent advances, including switching spectroscopy, mapping of polarization dynamics in ferroelectric and multiferroic nanostructures, imaging of biopolymers in calcified and connective tissues and PFM in liquid environments. [DOI: 10.1143/JJAP.46.5674] KEYWORDS: piezoresponse force microscopy, ferroelectric, multiferroic, atomic force microscopy, switching spectroscopy, electromechanics, polymers, biological imaging, liquid imaging

1.

Introduction

Coupling between electrical and mechanical phenomena is ubiquitous in nature. Examples of this coupling range from piezoelectricity in polar materials such as quartz, III– V nitrides and biopolymers, to a rich gamut of electromechanical phenomena linked with polarization dynamics in inorganic ferroelectrics and ferroelectric polymers. An extremely broad and fascinating spectrum of complex electromechanical behaviors emerges in biological systems, with examples ranging from simple piezoelectricity in calcified and connective tissues, to proton-driven conformational changes in cellular membranes, electromotor proteins and ion channels, and, on larger length scales, electromechanical responses on cellular and tissue levels.1) In piezoelectric and ferroelectric materials, tremendous progress in understanding electromechanical properties and their relationship with polarization dynamics and soft-mode behavior was achieved in the last 50 years.2) This was made possible by the (a) availability of single crystal samples that allow macroscopic property measurements and (b) emergence of X-ray and neutron scattering techniques that enabled probing structure and lattice dynamics on the unitcell level. The key enabling component in this case is the spatial uniformity of the material that allows comparison and interpretation of the macroscopic properties and scattering data. This is no longer the case for biological systems, in which understanding electromechanical functionality requires measurements on cellular, sub-cellular, and ultimately molecular level. Remarkably, similar challenges emerge for nanoscale ferroelectrics. It has been long known that symmetry breaking at ferroelectric surfaces and interfaces can both affect the surface phenomena and the stability of ferro

Corresponding author. E-mail address: [email protected]

electric phases. The examples of polarization-mediated surface phenomena are multiple and range from well-known domain-dependent chemical reactivity in acid dissolution3) and metal photodeposition reactions4) to chemical adsorption.5) At the same time, surface and size effects strongly affect the stability of a ferroelectric phase.6,7) The examples include symmetry breaking between equivalent polarization orientations in thin films (imprint), size-induced transitions between ferroelectric and paraelectric phases, and recently predicted novel vortex polarization states.8) Furthermore, several important classes of ferroelectric materials including relaxor ferroelectrics and materials close to morphotropic phase boundary exhibit complex nanoscale phase separation behavior even in the bulk form.9) Probing electromechanical coupling in biological, macromolecular, and inorganic systems opens a pathway for studies of underlying materials functionality from polarization dynamics in ferroelectrics to endlessly fascinating sets of electrochemical and proton-driven transformations in biomolecular systems. However, for both biological and inorganic materials this task generally requires electromechanical measurements on the nanometer scale of a single ferroelectric domain or cell, polar nanodomain or voltage channel, and ultimately single molecule or unit cell. Electromechanical imaging and spectroscopy on the nanometer scale has been enabled by the emergence of piezoresponse force microscopy (PFM) slightly over a decade ago.10) For a decade, PFM has been developed and studied exclusively in the context of inorganic ferroelectric materials for non-volatile memory and data storage applications.11–14) In these materials, strong electromechanical coupling (50 {1000 pm/V), good mechanical properties (Young’s moduli  100 GPa), and relative insensitivity to atmospheric conditions allowed broad applicability of PFM using large indentation forces, large voltage amplitudes, and frequencies well below resonance. This direct version of

5674

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

Vtip = Vdc

Vtip = Vdc + Vac sin ω t Tip Sample

Actuator

(a)

(b)

Fc = Fc (h, Vtip) PFM

Fnc = C z' ( z)

AFAM

(Vtip

UFM

Force

HUEFM

Apply Bias

Apply Force

− V surf )

2

STM

EFM KPFM

Measure Current

Contact Bias

d = d 0 + d1 sin ω t

PFM KPFM

AFM

Measure Displacement

Non-contact Distance

(b)

(d)

Fig. 1. (a) Experimental set-up for PFM. In PFM, detected is the first harmonic of the mechanical response of the AFM tip induced by the periodic bias applied to the tip. (b) Experimental set-up for atomic force acoustic microscopy (AFAM). In AFAM, measured is the first harmonic of the mechanical response of the AFM tip induced by the mechanical oscillations applied to the sample. (c) Voltage modulation SPMs can be described using a force–distance bias surface. In the small signal limit, the signal in techniques such as PFM, AFAM, EFM, and KPFM is directly related to the derivative in the bias or distance direction. (d) Schematic relating PFM to other scanning probe techniques.

PFM can be readily incorporated on most commercial SPM systems (with proper cross-talk compensation) and is relatively straightforward both in terms of instrumentation and qualitative data interpretation. Recently, PFM has been demonstrated as a powerful tool for imaging materials systems such as III–V nitrides,15) ferroelectric polymers,16) and piezoelectric biopolymers.17–19) In these cases, the response amplitudes are significantly smaller (1{ 5 pm/V). This necessitates the use of resonance enhancement to amplify weak electromechanical signals, and also results in a significant electrostatic contribution to the signal due to capacitive tip–surface interactions. The latter scales reciprocally with the cantilever spring constant (non-local part) and the spring constant of the tip–surface junction (local part) and can be minimized by using stiff (10 – 40 N/m) cantilevers. This strategy, however, is inapplicable for soft biological and polymer materials (Young moduli  1{10 GPa), necessitating the search for alternative probes and driving and detection schemes.20) Finally, high resolution imaging of clean ferroelectric surfaces requires imaging under ultra high vacuum conditions, while electromechanical studies of biological systems requires imaging in conductive liquid environments under controlled electrochemical conditions. Here, we briefly discuss image the formation mechanism in PFM and present several recent advances in PFM imaging

of polarization dynamics in ferroelectric materials, imaging of biological systems, and imaging in liquid environments, and formulate some of the pathways for the development of the field. 2.

Piezoresponse Force Microscopy

2.1 Principles of piezoresponse force microscopy Piezoresponse force microscopy is based on the detection of the bias-induced piezoelectric surface deformation. The tip is brought into contact with the surface, and the piezoelectric response of the surface is detected as the first harmonic component, A1! , of the tip deflection, A ¼ A0 þ A1! cosð!t þ ’Þ, induced by the application of the periodic bias, Vtip ¼ Vdc þ Vac cosð!tÞ, to the tip [Fig. 1(a)]. Here, the deflection amplitude, A1! , is assumed to be calibrated and given in the units of length. When applied to the pyroelectric or ferroelectric materials, the phase of the electromechanical response of the surface, ’, yields information on the polarization direction below the tip. Application of the bias to the tip results in the surface displacement, w, with both normal and in-plane components, w ¼ ðw1 ; w2 ; w3 Þ. The normal displacement of the tip apex in contact with the surface is generally equal to the surface displacement, since the effective spring constant of the tip– surface junction is typically 2 – 3 orders of magnitude higher than the cantilever spring constant (for imaging below first

5675

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

(a)

100 nm (b)

(c)

(d)

200 nm

(e)

(f)

Fig. 2. (a) Topography, (b) PFM amplitude, and (c) PFM phase of a BFO–CFO film. (d) Topography, (e) PFM amplitude, and (f) PFM phase of 1  1 mm2 area following the acquisition of the SS-PFM data shown in Fig. 3.

resonance). The lateral piezoresponse component in the direction normal to the cantilever axis (lateral transversal displacement), determined as torque of the cantilever, can be significantly smaller than that of the surface, e.g., due to the onset of sliding friction. The dynamic behavior of the cantilever and frequency dependence of these signal contributions are discussed in detail elsewhere.21,22) With proper calibration of detector sensitivity and imaging the sample surface at two orthogonal in-plane orientations, the full electromechanical response vector can be obtained. 2.2 Contact mechanics of PFM The foundation of any scanning probe microscopy (SPM) method is the fundamental image formation mechanism, which interrelates local surface and material properties, describes the dominant tip–surface interactions and is required for the (a) analysis of data in terms of relevant material properties, (b) separation of the required interaction from other interaction (cross-talk) and instrumental artifacts, and (c) formulation of strategies for instrumental platform, probe, controller, and software development for increasing the precision, resolution and sensitivity. In PFM, the underlying fundamental image formation mechanism is the voltage dependent mechanics of the tip– surface junction, which can be conveniently represented as a force-bias-distance surface (for purely normal force), Fc ¼ Fc ðh; Vtip Þ, where h is indentation depth. The image formation mechanism of various SPM methods on piezoelectric materials can then be represented as derivatives in the voltage or distance directions.23) In the small signal approximation the PFM signal is given by ð@h=@Vtip ÞF = const and the atomic force acoustic microscopy (AFAM) [Fig. 1(b)] signal is related to ð@h=@FÞV = const [Fig. 2(c)]. The determination of the force-bias-distance surface is equivalent to the voltage dependent contact mechanics of the

piezoelectric material, representing an extension of Hertzian (or Maugis–Dugdale) contact mechanics to piezoelectric materials. Given the dearth of analytically solvable contact mechanics models even for purely elastic materials, generally undefined geometry of an SPM probe, and poorly defined surface conditions, this necessitates the search for both exact and approximate solutions, as described below. 2.2.1 Exact solutions The rigorous solution of piezoelectric indentation, e.g., the shape of the force-distance-bias surface, is available only for the case of transversally isotropic materials.24–26) Kalinin et al.27) and Karapetian et al.28) derived a rigorous description of contact mechanics in terms of stiffness relations between applied force, P, and concentrated charge, Q, with indenter displacement, w0 , indenter potential, 0 , indenter geometry and material properties. The solutions were obtained for flat, spherical, and conical indenter geometries, and have the following phenomenological structure: 2 ½hnþ1 C1 þ ðn þ 1Þhn 0 C3   2 Q ¼ ½hnþ1 C3 þ ðn þ 1Þhn 0 C4 ;  P¼

ð1Þ ð2Þ

where h is total indenter displacement and  is a geometric factor ( ¼ a for flat indenters,  ¼ ð2=3ÞR1=2 for spherical indenters, and  ¼ ð1=Þ tan  for conical indenters) and n ¼ 0 for flat, n ¼ 1=2 for the spherical, and n ¼ 1 for the conical indenters, respectively. These stiffness relations provide an extension of the corresponding results of Hertzian mechanics and continuum electrostatics to the transversely isotropic piezoelectric medium. All indentation stiffnesses are complex functions of electroelastic constants of the material, Ci ¼ Ci ðcij ; eij ; "ij Þ, where cij are elastic stiffnesses, eij are piezo-

5676

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

electric constants, and "ij are dielectric constants. Detailed analysis of the stiffness relations for the spherical indentation and the effect of materials constants on values of coupling coefficients is given elsewhere.27,28) It has been shown that for most materials C3 =C1  d33 and C4  pffiffiffiffiffiffiffiffiffiffiffiffi "11 "33 , validating earlier approximations in PFM. Equations (1) and (2) yield a number of important conclusions on the information that can be obtained from SPM or nanoindentation experiments on transversally isotropic piezoelectric materials (e.g., cþ , c domains in tetragonal perovskites). For all simple tip geometries, materials properties are described by three parameters, indentation elastic stiffness, C1 , indentation piezocoefficient, C3 , and indentation dielectric constant, C4 . Thus, the maximum information on electroelastic properties for a transversally isotropic material that can be obtained from an SPM experiment is given by these three quantities and mapping of the Ci distributions provides a comprehensive image of the surface electroelastic properties. Experimentally, the PFM signal is deff ¼ @h=@Vtip ¼ C3 h=ðC1 h þ n 0 C3 Þ. The AFAM signal is related to the stiffness of the tip–surface junction, k1 ¼ @P=@h ¼ 2ðn þ 1Þhn ðC1 h þ n 0 C3 Þ=ðhÞ. For small tip biases, deff ¼ C3 =C1 and k1 ¼ 2ðn þ 1Þhn C1 =. Due to the smallness of the corresponding capacitance, indentation dielectric constant, C4 , can not be directly determined in the SPM experiment; however, it might be accessible on larger length scales, e.g., using the nanoindentation approach. From this analysis, the PFM signal does not depend on the tip–surface contact area and effective tip geometry, thus allowing for quantitative imaging. On the contrary, the AFAM signal is strongly dependent on the tip geometry, necessitating complex calibration procedures. Another implication is that PFM is only weakly dependent on surface topography, whereas AFAM will exhibit strong topographic cross-talk. 2.2.2 Approximate solutions The necessity for calculating the PFM signal for materials of general symmetry, as well as calculation of the response at microstructural elements such as domain walls, cylindrical domains, and topographically inhomogeneous ferroelectrics such as nanoparticles have stimulated theoretical attempts to derive approximate solutions for PFM contact mechanics. A general approach for the calculation of the electromechanical response is based on the decoupling approximation. In this case, (a) the electric field in the material is calculated using a rigid electrostatic model (no piezoelectric coupling, dijk ¼ eijk ¼ 0), (b) the strain or stress field is calculated using constitutive relations for a piezoelectric material, Xij ¼ Ek ekij , and (c) the displacement field is evaluated using the appropriate Green’s function for an isotropic or anisotropic solid. A simplified one-dimensional (1D) version of the decoupled model was originally suggested by Ganpule et al.29) to account for the effect of 90 domain walls on PFM imaging. A similar 1D approach was adapted by Agronin et al.30) to yield closed-form solutions for the PFM signal. The three dimensional (3D) version of this approach was developed by Felten et al.31) using the analytical form for the corresponding Green’s function. Independently, Scrymgeour and Gopalan32) have

used the finite element method to model PFM signals across domain walls. In the decoupled approximation, the PFM signal, i.e., the surface displacement ui ðx; yÞ at location x induced by the tip at position y ¼ ðy1 ; y2 Þ is given by Z1 Z1 Z1 d1 d2 d3 ui ðx; yÞ ¼ 1

1

0

@Gij ðx1  1 ; x2  2 ; 3 Þ  @k  El ðÞckjmn dlnm ðy1 þ 1 ; y2 þ 2 ; 3 Þ:

ð3Þ

Here coordinate x ¼ ðx1 ; x2 ; zÞ is linked to the indentor apex, coordinates y ¼ ðy1 ; y2 Þ denote indentor position in the sample coordinate system y. Coefficients dlmn and ckjmn are position dependent components of the piezoelectric strain constant and elastic stiffness tensors, respectively. Ek ðxÞ is the electric field strength distribution produced by the probe. The Green’s function for a semi-infinite medium G3j ðx  Þ links the eigenstrains cjlmn dmnk Ek to the displacement field. This approach is rigorous for materials with small piezoelectric coefficients. The use of the decoupling approximation reduces an extremely complex coupled contact mechanics problem to solutions of much simpler electrostatic and mechanical Green’s function problems, and numerical integration of the result. We note that the dielectric and, particularly, the elastic properties described by positively defined second- and fourthrank tensors (invariant with respect to 180 rotation) are necessarily more isotropic than the piezoelectric properties described by third-rank tensors (anti-symmetric with respect to 180 rotation). Hence, elastic and dielectric properties of a material can often be approximated as isotropic. Recently, Kalinin et al.,33) Eliseev et al.,34) and Morozovska et al.35) have applied the decoupled theory to systematically describe the image formation mechanism in PFM. In particular, the electromechanical response of a fully anisotropic material was found as ui ðxÞ ¼ VQ Wijlk ðxÞdkjl , where the tensor Wijlk ðxÞ describes materials properties and VQ is the electrostatic potential in the point of contact. In the limit of electric and elastic isotropy Wijlk ðxÞ depends only on the Poisson ratio of the material and the analytical expressions for its components were derived. This analysis opens the pathway for orientational mapping, in which three Euler angles, ð; ’; Þ, determining molecular or crystallographic orientation can be found from three components of the electromechanical response vector. Image formation described by eq. (3) belongs to the class of so-called linear imaging mechanisms. The phenomenological resolution theory in PFM has recently been developed by Kalinin et al.36) The corresponding analytical theory has been developed by Morozovska et al.35) In particular, linearity of the PFM image formation mechanism allows to . establish the unambiguous definitions for spatial resolution (minimal feature size required for quantitative response measurements) and information limits (minimal detectable feature size) and its dependence on tip geometry and materials properties, suggesting strategies for high-resolution imaging . develop the pathways for calibration of tip geometry for quantitative data interpretation, i.e., determine the set of image charges fQm ; dm gN representing the tip.37)

5677

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

. interpret the imaging and spectroscopy data in terms of intrinsic domain wall widths and the size of the nascent domain below the tip; and . deconvolute the ideal image from experimental data and establish applicability limits and errors associated with such deconvolution processes. 2.3 PFM vs other SPM The fundamental factors underpinning any SPM method are (a) the tensorial nature of the signal, (b) the signal dependence on contact radius (contact) or tip–surface separation (non-contact), and (c) the signal dependence on cantilever spring constant. These factors determine the strategies for instrumentation and technique development and the potential for quantitative measurements. The relationship between externally controlled parameters (bias and force) and measured responses (current and displacement) in a generalized SPM experiment is shown in Fig. 1(d). In current-based techniques such as scanning tunneling microscopy (STM) and conductive atomic force microscopy (AFM), the excitation signal (applied bias), and detected signal (current magnitude) are scalar quantities. In the classical (Maxwell) regime the conductance scales with the contact area as   a1 .38) In most applications of conductive AFM, the variations in local conductivities between dissimilar materials (e.g., differently doped regions of a semiconductor) dominate relatively small changes in contact geometry, the condition ensured by operation at large indentation forces on planar substrates. This, in turn, allows for calibration and quantitative measurements using currentbased techniques. In force-based techniques, both the excitation signal (force) and the response signal (displacement) are vectors. Hence, the AFM signal is a rank two tensor. In the continuum mechanics limit, the contact stiffnesses are proportional to the contact radius, aij  a1 . When the contact area is a single molecule, as in protein unfolding spectroscopy, or a single atom as in atomic resolution imaging, aij  a0 . Hence, quantitative force measurements are generally limited to the cases when the contact geometry is well characterized, as in nanoindentation techniques, or is weakly dependent on the probe, as in molecular unfolding spectroscopy. Finally, in PFM the excitation signal is scalar, whereas the electromechanical response of the surface is a vector. Hence, the PFM response is a vector. From simple dimensionality considerations, the electromechanical response does not depend on the contact area, di  a0 , as confirmed by the rigorous theory for special classes of material symmetry.27,28) This weak dependence of the PFM signal on the contact area suggests that electromechanical measurements are (a) intrinsically quantitative and do not require extensive probe calibration and (b) the signal is relatively insensitive to topographic cross-talk. On the other hand, resonance enhancement cannot be directly employed in PFM, since the resonance frequency is determined by mechanical, rather than electromechanical properties of the material.20) Finally, quadratic bias dependence of electrostatic forces, as compared to linear piezoelectric interactions, suggests that the

electromechanical signal cannot be unambiguously distinguished from the electrostatic signal and can only be minimized through the use of stiff cantilevers, shielded probes, imaging at independently determined nulling potentials (bias control) or antiresonances for electrostatic contribution (frequency control). 3.

Polarization Dynamics in Ferroelectrics

Multiple recent applications of PFM include high-resolution imaging of ferroelectric and piezoelectric materials,39,40) tip-induced polarization switching for data storage14,41) and ferroelectric lithography,42) and local hysteresis loop measurements.43,44) In the latter, the local electromechanical response is measured as a function of a dc bias applied to the probe. The resulting local electromechanical hysteresis loop provides information on the size of the biasinduced domain formed below the tip, and hence the thermodynamics and kinetics of local tip-induced phase transitions between opposite polarization states. However, point spectroscopy is insufficient to study spatial variability of the switching behavior and the role of local defects, topographic inhomogeneities, etc. on switching. To overcome this limitation, we have recently developed switching spectroscopy PFM (SS-PFM).45) In this technique, the tip approaches the surface vertically with the deflection signal used as a feedback, until the deflection set-point is achieved thus indicating contact between the tip and the surface. When the set-point is achieved, a hysteresis loop of the local piezoresponse vs the dc bias is acquired. The tip is then moved to the next location so that an M  M point mesh with spacing, l, between points is scanned. The hysteresis curves are collected at each point and stored in a 3D data array for subsequent examination and analysis. The acquisition time for a 128  128  256 point data array is typically 2 – 4 h. In order to extract switching parameters from the large number of hysteresis loops acquired during a single scan, automated routines based on statistical analysis and parametric curve fitting were developed.46) Phenomenological parameters describing the switching process such as positive and negative coercive bias, imprint voltage, and saturation response can be extracted from the data sets and plotted as two-dimensional (2D) maps. Alternatively, hysteresis loops from selected point(s) can be extracted and analyzed to establish the presence of fine structure indicative of domaindefect interactions. Application of SS-PFM for ferroelectric nanostructure is illustrated on an example of multiferroic BiFeO3 –CoFe2 O4 (BFO–CFO) self assembled nanostructures. Topography and PFM amplitude and phase images of a 0:5  0:5 mm2 region are shown in Figs. 2(a)–2(c), respectively. From the PFM amplitude and phase images [Figs. 2(b) and 2(c)], the ferromagnetic nanopillars are not piezoelectric, while the ferroelectric BFO thin film exhibits clear electromechanical contrast, as anticipated. A domain wall can be seen in Fig. 2(b), separating two domains in Fig. 2(c). The effective resolution estimated from domain wall width is 30 nm, while the electromechnical width of the BFO/ CFO boundary is 10 nm. These values provide upper and lower estimates for effective tip size in these measurements.

5678

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

(c)

(b)

(a)

1

(d)

piezoresponse, a.u.

piezoresponse, a.u.

1

0

-1

-2

-3 -20

-10

0

10

(e)

20

Vdc, V

0

-1

-2

-3 -20

-10

0

10

20

Vdc, V

Fig. 3. 2D SS-PFM images of (a) the initial PFM response, (b) imprint, and (c) the work of switching for the area shown in Fig. 1. (d, e) Representative loops from the BFO and BFO–CFO interface regions, respectively.

To address the local switching behavior in the BFO regions confined between CFO nanopillars, SS-PFM is employed to measure local hysteresis loops on the equidistant N  N mesh. Shown in Fig. 3 are 2D SS-PFM maps of initial PFM response, imprint, work of switching, and representative loops from two regions. The loops in Fig. 3(d) are from the BFO region indicated by an oval in Fig. 3(a), while the loops in Fig. 3(e) are from the boundary between the BFO film and a CFO nanopillar as indicated by the oval in Fig. 3(c). The loops in Fig. 3(d) are not well saturated at 20 V bias and show a significant asymmetry along the piezoresponse axis. The loops measured near the BFO–CFO interface have a lower piezoresponse at the maximum negative bias as indicated by comparing the lower branches of the loops in Figs. 3(d) and 3(e). The remarkable feature of the hysteresis loops shown in Figs. 3(c) and 3(e) is the gradual decay of the effective loop area from the ferroelectric BFO phase to the non-ferroelectric CFO phase. Notably, repeated local hysteresis measurements of the film have changed the domain structure, indicative of the switching process. Shown in Figs. 2(d)–2(f) are topography and PFM amplitude and phase images of a 1  1 mm2 region after the SS-PFM scan. A reduction in piezoresponse amplitude and apparent backswitched domains in the piezoresponse phase can be seen as indicated by the ovals in Figs. 2(e) and 2(f), respectively. However, there is no visible imprint in these regions sandwiched between nanopillars from the 2D SSPFM imprint map [Fig. 3(b)]. The evolution of the ferroelectric properties as dependent on the distance to the BFO–CFO interface is shown in Fig. 4. Shown are 2D SS-PFM maps of negative remanent piezoresponse [Fig. 3(a)] and switchable piezoresponse [Fig. 4(b)], with corresponding line profiles of all of the switching parameters to demonstrate the spatial variations in

switching properties near BFO–CFO interfaces. In Fig. 4(c) line profiles of PFM, imprint, remanent piezoresponses, and switchable piezoresponse are plotted, while in Fig. 4(d), line profiles of positive and negative coercive biases and the work of switching are plotted. There appears to be a gradual change in switching properties (as opposed to an abrupt change at the interface) near the interface, suggesting that the heterostructure interface has significant influence on the switching properties as detected by PFM over a length scale much larger than the width interface. In particular, while, e.g., the coercive biases appear to reach equilibrium after 3 or 4 pixels (96 –128 nm), the line profiles for work and switchable piezoresponse continue to change over longer length scales. This behavior can be understood from the mechanism of hysteresis loop formation as analyzed recently by Morozovska et al.47) In particular, it was shown that the coercive bias, i.e., domain size for which response of the domain formed below the tip is exactly inverse of that of the surrounding matrix, corresponds to a domain radius of r ¼ ð0:3 { 0:7ÞRtip , where Rtip is the characteristic tip size. At the same time, the saturation of electromechanical response to a value within 10% of the bulk occurs for a domain size of r ¼ ð5 {15ÞRtip . This extremely slow saturation of the hysteresis loop tails stem from the Rtip =r behavior of the corresponding Green’s function. Thus, the observed evolution of the hysteresis loops can be attributed primarily to the confinement effect of the nascent ferroelectric domain by the BFO–CFO interface. This conclusion is further corroborated by the fact that other parameters describing switching, including nucleation and coercive biases, do not demonstrate any anomalies in the vicinity (30 { 50 nm) of the interface. Thus, within the resolution of the SS-PFM method used here, switching mechanism in the BFO component is not

5679

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

1

0 Responses, a.u.

-1 -2 -3

0 -1 PFM Imprint Rplus Rminus Psp

-2 -3

(a)

(c)

-4

18 16 14 12 10 8 6 Position

4

50

2

(b)

Responses, a.u.

30 20 10 0

-10 18 16 14 12 10 8 6 (d) Position

0

0

Vminus Vplus work

40

1

2

4

2

0

Fig. 4. 2D SS-PFM images of (a) the negative remanent piezoresponse and (b) the switchable piezoresponse. (c, d) Line profiles of the switching properties across the section of the SS-PFM dataset indicated by a dashed line in (a).

affected by the presence of BFO–CFO interfaces, evidencing the high quality of the multiferroic structure. This example illustrates the applicability of SS-PFM to study spatial variability of switching properties in ferroelectric and multiferroic heterostructures, necessary to understanding the coupling mechanisms between dissimilar phases, i.e., the role of short-range interface effects and longrange strain48) and electric field interactions, and the knowledge of the mechanism of switching in ferroelectric components under spatial constraints imposed by the epitaxial film, including the role of interfaces and strain field on the stability of a given polarization orientation and switching behavior. 4.

PFM Studies of Biological Systems

One of the most important manifestations of electromechanical behavior is piezoelectricity, which stems from the crystal structure of most biopolymers including cellulose, collagen, keratin, etc. Piezoelectric behavior has been observed in a variety of biological systems including bones,49–52) teeth,53) wood,54,55) and seashells.56) It is postulated that the piezoelectric coupling, via mechanical stress that generates the electric potential, controls the mechanisms of tissue development.57,58) Understanding the relationship between physiologically generated electric fields and mechanical properties on the molecular, cellular and tissue levels has become the main motivation of studying piezoelectricity in biological systems. However, the complex hierarchical structure of biological materials combined with strong orientation dependence of the piezoelectric effect precluded quantitative studies.

Recently, variation in piezoelectric activity between piezoelectric proteins and non-piezoelectric hydroxiapatite was suggested as a functional basis for high-resolution imaging of calcified and connective tissues. Shown in Fig. 5 is the mechanical and electromechanical imaging in a human tooth. A strong response signal of the dentine region is consistent with a high density of piezoelectrically active collagen.53) Several isolated regions with a high piezoresponse signal observed in the enamel region are indicative of a low fraction of protein fibers. To get further insight into the structure of the collagen fibril, we have combined measurements of the vertical (outof-plane) electromechanical response with lateral (in-plane) response measurements [Fig. 6(a)]. Neither topographic [Fig. 6(b)] nor elastic [Fig. 6(c)] images of an enamel surface show any significant contrast difference between the grains. In comparison, both vertical PFM (VPFM) and lateral PFM (LPFM) images show a very strong electromechanical response that we attribute to a protein (presumably collagen) fibril embedded within a non-piezoelectric matrix [Figs. 6(d) and 6(e)]. The spatial resolution of PFM, determined as a half-width of the boundary between differently piezoelectric regions, is about 5 nm. Note that the resolution achieved is an order of magnitude better than 50 – 100 nm typical for single crystals and is comparable to the best results achieved to date for thin films of ferroelectric perovskites. Comparison of the VPFM and LPFM images shows a different pattern of piezoelectric domains, suggesting a complicated structure of fibril, consisting of several protein molecules. To visualize electromechanical response data, we

5680

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

(a)

1 µm

(b)

S. V. KALININ et al.

1 µm

(c)

(d)

(e)

1 µm

(f)

Fig. 5. Topographic (a, c, e) and piezoelectric (b, d, f) images of enamel (a, b), dentin (c, d), and central (e, f) regions of dry tooth. Partially reprinted with permission from S. V. Kalinin et al. [Ultramicroscopy 106 (2006) 334].

employ vector representation for PFM. The VPFM and LPFM images are normalized with respect to the maximum and minimum values of the signal amplitude so that the intensity changes between 1 and 1, i.e., vpr; lpr 2 ð1; 1Þ. Using commercial software,59) the 2D vector data ðvpr; lprÞ is converted to the amplitude/angle pair, A2D ¼ Absðvpr þ I lprÞ, 2D ¼ Argðvpr þ I lprÞ. The data are plotted so that the color corresponds to the orientation, while color intensity corresponds to the magnitude, as shown in the color wheel diagram. The color encoded vector response map (Vector PFM), shown in Fig. 6(f), clearly delineates the helical structure, visualizing the electromechanically active protein fibril conformation in real space. These examples illustrate the broad potential of PFM for electromechanical imaging of biological tissues. Since the introduction of bioelectromechanical imaging in 2005, several systems including teeth, bones, canine femur, and butterfly wigs have been studied.17–19,60,61) Clearly, there is tremendous potential for fundamental studies of these materials. However, the studies of electromechanical activity in cells and biomolecules require PFM imaging in liquid environment, as discussed below. 5.

PFM in Liquid Environment

To date, PFM and closely related Kelvin probe force microscopy have been implemented primarily in ambient or ultra-high vacuum (UHV) environments. Both in ambient and in UHV, tip–surface interaction is controlled by short range van der Waals forces, resulting in a characteristic jump to contact behavior for static cantilevers. In addition, in ambient the large negative curvatures and strong electric field in the tip–surface junction lead to the formation of liquid necks due to capillary and electrocapillary effects, further increasing tip–surface forces. Hence, the minimal

tip–surface force in ambient and UHV is limited, resulting in a lower achievable limit on the contact stiffness of tip– surface junction. On approaching the surface, the effective spring constant of the systems changes abruptly from the value corresponding to free cantilever (0:1{10 N/m) to a much higher value (1000 N/m) corresponding to a cantilever in contact with the surface. Hence, the contact (PFM) and non-contact (KPFM) regimes are well-separated in UHV and ambient, thus allowing these techniques to be unambiguously distinguished. The distinguishing feature of scanning probe microscopy in liquids is the complexity of tip–surface interactions.62) In particular, depending on relative polarizabilities of tip and surface materials and liquid, van der Waals can be both attractive and repulsive. At small distances, short-range hydrophobic interactions, and structural component of disjoining pressure become significant, often resulting in the formation of multiple minima on the force–distance curve. These interactions control the effective spring constant of the system. Particularly relevant for PFM and KPFM imaging are electrostatic interactions in liquid. Similar to van der Waals interactions, capacitive forces in liquid can be both attractive and repulsive, depending on the relative magnitude of the static dielectric constant of liquid and surface. In addition, electrostatic interactions are strongly mediated by the presence of mobile ions and decay exponentially at distances above the Debye length in solution. Thus, in the interpretation of electromechanical interactions in solution, the capacitive interactions, coupling in the double layer and electromechanical response of the piezoelectric material must be taken into account. While the understanding of electromechanical imaging of biological and biomolecular systems in liquid environments is clearly a fruitful area of future research, below we

5681

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

Displacement

z

Topography

Elasticity

(b)

(c)

Vertical PFM

Lateral PFM

(d)

(e)

VPFM LPFM

x

y Bias

(a) 2D Vector Map



ψ Map

θ Map

-100°

(f)



-100°

(h)

(g)

Fig. 6. (a) Electromechanical response of the surface to the tip bias is a vector measure, components of which are related to the local orientation of protein molecules. In PFM, detection of vertical (VPFM) and torsional (LPFM) components of the cantilever response allows two vector components, in-plane and out-of-plane, to be simultaneously measured. (b) Surface topography (vertical scale 20 nm) and (c) elasticity map of 400  400 nm2 region on an enamel surface. Vertical (d) and lateral (e) PFM images of the same region. (f) Vector PFM map of local electromechanical response. Color indicates the orientation of the electromechanical response vector, while the intensity provides the magnitude (color wheel diagram). (g, h) Semiquantitative map of local molecular orientation. Reprinted with permission from S. Kalinin et al. [Ultramicroscopy 106 (2006) 334].

summarize two examples of imaging and manipulation of model ferroelectric systems. 5.1 Imaging in liquid environment High-resolution liquid PFM imaging of a polycrystalline PZT sample is shown in Fig. 7. Resolution of the order of 3 nm in the PFM signal (0.2 nm for phase signal), approaching the theoretical domain wall width, as compared to a resolution of 30 nm (2 nm for phase) in ambient, is observed for several domain walls.63) Tip–surface interactions in liquid PFM were studied using force–distance curves, amplitude frequency and amplitude bias spectroscopy. Force distance curves have shown an almost complete absence of capillary hysteresis. At low frequencies, the PFM response in liquid is reduced compared to ambient by more than an order of magnitude and the first resonant peak is shifted. For high order resonances the Q factor increases, achieving Qi ¼ 13:6 for i ¼ 3, comparable to Q ¼ 20 in ambient. This behavior agrees with known models for the frequency dependence of viscous damping and added mass effects. In particular, the maximal signal at the second resonance in liquid is within a factor of two of the first

ambient resonance, potentially providing comparable sensitivity for resonant-enhanced ambient and liquid PFM. The observed high resolution is attributed to the absence of long-range electrostatic forces due to screening by mobile ions in liquid and the absence of capillary interactions, resulting in localization of the ac field in the tip–surface junction. The tip and the surface are separated by a thin (1– 3 nm) liquid layer due to hydrophobic tip–surface interactions, resulting in the disappearance of contrast in solutions with high (>102 M) concentrations. The contrast decreases for high salt concentrations, and disappears when tip–surface separation becomes larger than the corresponding Debye length (Fig. 8). At low frequencies, the signal is reduced by an order of magnitude compared to the ambient. However, the viscous damping and added mass effects on cantilever dynamics are reduced at high frequencies corresponding to resonances, allowing sensitivity comparable to ambient conditions to be achieved. 5.2 Switching in liquid environment The conductive properties of liquids on the switching behavior was investigated by using alcohols with variable

5682

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

175 nm

10 nm

10 V

+90°

0

-90°

0 2 µm

(b)

(c)

mixed signal phase

60 40

Ambient 0.0

20 0 -20 -40 -60

30 nm

-0.5

0.5

-80

0

20

40

60

80

100

120

140

0.0

Liquid

(e)

160

100 80 60 40 20 0 -20

-0.5

-40 -60

3 nm

-80

-1.0

-100

(d)

mixed signal phase

80

phase, degrees

0.5

100

phase, degrees

piezoresponse, a.u.

1.0

piezoresponse, a.u.

(a)

-100 0

2

4

6

8

10

12

14

16

distance, nm

distance, nm

Ambient/Vacuum

Fig. 7. (a) Topography, (b) PFM amplitude, and (c) PFM phase images of polycrystalline PZT in water. Mixed PFM and PFM phase line profiles across a domain wall in air and in water are shown in (d) and (e), respectively. Reprinted with permission from B. Rodriguez et al. [Phys. Rev. Lett. 96 (2006) 237602]. Copyright 2006 by the American Physical Society.

Non-contact Tip

Contact N=1

Liquid

Surface (a)

(c)

(b)

N=2

N=3

λd δ

(d)

(e)

Fig. 8. Schematic illustrating PFM contrast enhancement in liquid. In ambient (a, b) electrostatic forces are present between the tip and the sample. Electromechanical coupling [double arrow in (b, d)] is present only for contact mode, when tip and surface are touching. In liquid, however, the electrostatic forces are minimized (c). (e) Oscillation of cantilever at high eigenmodes reduces hydrodynamic damping and added mass effects on cantilever dynamics. Reprinted with permission from B. Rodriguez et al. [Phys. Rev. Lett. 96 (2006) 237602]. Copyright 2006 by the American Physical Society.

alkyl chain lengths to control the concentration and diffusion length of the mobile ions and thereby vary the spatial extent of the potential distribution in the solution, and hence spatial distribution of electric field in the film. In an ambient or non-polar liquid environment, the biased tip establishes a highly-localized electric field that decays rapidly with distance from the tip–surface junction. This field results in highly localized polarization switching above a certain threshold value, generating well-defined

2D domains that grow in area with time and pulse length, Figs. 9(a) and 9(b). In this case, domain dimensionality is controlled by the field structure and is unaffected by largescale disorder. In liquid, the PFM contrast is strongly mediated by the presence of mobile ions that result in an increase of the effective area affected by the tip field. For solvents with intermediate polarities, the electric field is localized to the tip surface junction, but the characteristic length scale is significantly larger than the tip size and is

5683

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

(e)

(c)

(a)

(b)

S. V. KALININ et al.

400 nm

2 µm

(d)

(f)

2 µm

Fig. 9. Schematics (a, c, e) and PFM phase images (b, d, f) of switching in (a, b) local, (c, d) fractal, and (e, f) non-local cases. Depending on the characteristic spatial extent of electric field vs length scale of disorder, the transition between switching modes is observed. Adapted with permission from B. J. Rodriguez et al. [Phys. Rev. Lett. 98 (2007) 247603].

mediated by pulse length via the mobile ion diffusion length. The switching in this case results in the formation of irregular fractal domains with scale-dependent dimensionality determined by the relationship between the domain size and characteristic length scale of the frozen disorder, Figs. 9(c) and 9(d). In conductive solvents, the solution is uniformly biased, resulting in a homogeneous electric field across the film thickness (similar to ferroelectric capacitors), as shown in Figs. 9(e) and 9(f). In this non-local case, switching is dominated by the frozen disorder in the film, since the driving force for switching is independent of the position of the tip. Variation of the pulse length and magnitude in this case allows the regions where the switching is most likely to happen, i.e., the nucleation centers, to be visualized. Notably, high resolution imaging is possible even in polar solvents because of the much higher excitation frequencies, minimization of the diffusion paths, and high localization of the strain that transmit predominantly through the mechanical (rather than electrical) contact area. This allows imaging of domains independently of whether the switching is nonlocal or local. 6.

Conclusions

In a decade and a half since its invention, PFM has evolved from a relatively obscure technique for domain imaging to a broadly accepted tool for nanoscale characterization, domain patterning, and spectroscopy of ferroelectric material. PFM is applicable for imaging of other piezoelectric functional materials such as III–V nitrides, and electromechanically active proteins in calcified and connective tissues. The pronounced difference in signal formation mechanism between PFM and force- and currentbased SPMs suggests that there is a tremendous potential for further development, including probe development (e.g., including electrostatic shielding64,65) to minimize electrostatic force and allow reliable imaging in conductive liquid environments), instrumentation (avoiding the instrumental

cross-talks and also enabling frequency-tracking methods based on non-sinusoidal excitation signals66)), and theoretical effort. This progress will bring the understanding of electromechanical coupling at the nanometer level, establish the role of surface defects on polarization switching (Landauer paradox), probe nanoscale polarization dynamics in phaseordered materials and unusual polarization states. In biosystems, PFM potentially opens pathways for studies of electrophysiology on cellular and molecular levels, signal propagation in neurons, etc. Ultimately, at the molecular level, PFM can allow reactions and energy transformation pathways to be understood and become the enabling component of molecular electromechanical machines. Acknowledgements Research supported in part (SVK, SJ, BJR) by the Division of Materials Science and Engineering, Basic Energy Sciences, U.S. Department of Energy at Oak Ridge National Laboratory (ORNL), which is managed by UTBattelle, LLC. The work by ABP and KS is supported by the Center for Nanophase Materials Sciences, which is sponsored at Oak Ridge National Laboratory by the Division of Scientific User Facilities, U.S. Department of Energy. The work of BM and EK are supported by ORNL SEED funding. Authors acknowledge multiple discussions with E. W. Plummer (UT/ORNL), A. Gruverman (NCSU), and multiple colleagues at ORNL, NCSU, and UT. S. H. Kim (Inostek) is acknowledged for BFO films used for data in Fig. 9.

5684

1) R. Plonsey and R. C. Barr: Bioelectricity: A Quantitative Approach (Kluwer, Dordrecht, 2000). 2) M. E. Lines and A. M. Glass: Principles and Applications of Ferroelectric and Related Materials (Clarendon Press, Oxford, 1977). 3) F. Jona and G. Shirane: Ferroelectric Crystals (Dover Publications, New York, 1993). 4) J. L. Giocondi and G. S. Rohrer: Chem. Mater. 13 (2001) 241. 5) S. V. Kalinin, C. Y. Johnson, and D. A. Bonnell: J. Appl. Phys. 91

Jpn. J. Appl. Phys., Vol. 46, No. 9A (2007)

S. V. KALININ et al.

(2002) 3816. 6) M. Dawber, K. M. Rabe, and J. F. Scott: Rev. Mod. Phys. 77 (2005) 1083. 7) N. Sai, A. M. Kolpak, and A. M. Rappe: Phys. Rev. B 72 (2005) 020101. 8) I. Naumov, L. Bellaiche, and H. X. Fu: Nature 432 (2004) 737. 9) A. A. Bokov and Z. G. Ye: J. Mater. Sci. 41 (2006) 31. 10) A. Gruverman, O. Auciello, and H. Tokumoto: Appl. Phys. Lett. 69 (1996) 3191. 11) A. Gruverman, O. Auciello, and H. Tokumoto: Annu. Rev. Mater. Sci. 28 (1998) 101. 12) L. M. Eng, S. Grafstrom, Ch. Loppacher, F. Schlaphof, S. Trogisch, A. Roelofs, and R. Waser: Advance in Solid State Physics (Springer, Berlin, Heidelberg, 2001) Vol. 41, p. 287. 13) Nanoscale Characterization of Ferroelectric Materials, ed. M. Alexe and A. Gruverman (Springer, Heidelberg, 2004). 14) Nanoscale Phenomena in Ferroelectric Thin Films, ed. S. Hong (Kluwer, Dordrecht, 2004). 15) B. J. Rodriguez, A. Gruverman, A. I. Kingon, R. J. Nemanich, and O. Ambacher: Appl. Phys. Lett. 80 (2002) 4166. 16) B. J. Rodriguez, S. Jesse, S. V. Kalinin, J. Kim, S. Ducharme, and V. M. Fridkin: cond-mat/0611539. 17) C. Halperin, S. Mutchnik, A. Agronin, M. Molotskii, P. Urenski, M. Salai, and G. Rosenman: Nano Lett. 4 (2004) 1253. 18) S. V. Kalinin, B. J. Rodriguez, S. Jesse, T. Thundat, and A. Gruverman: Appl. Phys. Lett. 87 (2005) 053901. 19) S. V. Kalinin, B. J. Rodriguez, S. Jesse, T. Thundat, V. Grichko, A. P. Baddorf, and A. Gruverman: Ultramicroscopy 106 (2006) 334. 20) S. Jesse, B. Mirman, and S. V. Kalinin: Appl. Phys. Lett. 89 (2006) 022906. 21) S. Jesse, A. P. Baddorf, and S. V. Kalinin: Nanotechnology 17 (2006) 1615. 22) S. V. Kalinin, B. J. Rodriguez, S. Jesse, J. Shin, A. P. Baddorf, P. Gupta, H. Jain, D. B. Williams, and A. Gruverman: Microsc. Microanal. 12 (2006) 206. 23) S. V. Kalinin, A. Rar, and S. Jesse: IEEE Trans. Ultrason. Ferroelectr. Freq. Control 53 (2006) 2226. 24) A. E. Giannakopoulos and S. Suresh: Acta Mater. 47 (1999) 2153. 25) W. Chen and H. Ding: Acta Mech. Solida Sin. 12 (1999) 114. 26) E. Karapetian, I. Sevostianov, and M. Kachanov: Philos. Mag. B 80 (2000) 331. 27) S. V. Kalinin, E. Karapetian, and M. Kachanov: Phys. Rev. B 70 (2004) 184101. 28) E. Karapetian, M. Kachanov, and S. V. Kalinin: Philos. Mag. 85 (2005) 1017. 29) C. S. Ganpule, V. Nagarajan, H. Li, A. S. Ogale, D. E. Steinhauer, S. Aggarwal, E. Williams, R. Ramesh, and P. DeWolf: Appl. Phys. Lett. 77 (2000) 292. 30) A. Agronin, M. Molotskii, Y. Rosenwaks, E. Strassburg, A. Boag, S. Mutchnik, and G. Rosenman: J. Appl. Phys. 97 (2005) 084312. 31) F. Felten, G. A. Schneider, J. Mun˜oz Saldan˜a, and S. V. Kalinin: J. Appl. Phys. 96 (2004) 563. 32) D. A. Scrymgeour and V. Gopalan: Phys. Rev. B 72 (2005) 024103. 33) S. V. Kalinin, E. A. Eliseev, and A. N. Morozovska: App. Phys. Lett. 88 (2006) 232904. 34) E. A. Eliseev, S. V. Kalinin, S. Jesse, S. L. Bravina, and A. N.

Morozovska: J. Appl. Phys. 102 (2007) 014109. 35) A. N. Morozovska, S. L. Bravina, E. A. Eliseev, and S. V. Kalinin: cond-mat/0608289. 36) S. V. Kalinin, S. Jesse, B. J. Rodriguez, J. Shin, A. P. Baddorf, H. N. Lee, A. Borisevich, and S. J. Pennycook: Nanotechnology 17 (2006) 3400. 37) S. Jesse, S. V. Kalinin, B. J. Rodriguez, E. A. Eliseev, and A. N. Morozovska: submitted. 38) N. N. Lebedev, I. P. Skal’skaya, and Ya. S. Uflyand: Problems in Mathematical Physics (Pergamon Press, New York, 1966). 39) B. J. Rodriguez, A. Gruverman, A. I. Kingon, R. J. Nemanich, and O. Ambacher: Appl. Phys. Lett. 80 (2002) 4166. 40) S. V. Kalinin, B. J. Rodriguez, S. Jesse, T. Thundat, and A. Gruverman: Appl. Phys. Lett. 87 (2005) 053901. 41) H. Shin, S. Hong, J. Moon, and J. U. Jeon: Ultramicroscopy 91 (2002) 103. 42) S. V. Kalinin, D. A. Bonnell, T. Alvarez, X. Lei, Z. Hu, and J. H. Ferris: Adv. Mater. 16 (2004) 795. 43) T. Hidaka, T. Maruyama, M. Saitoh, N. Mikoshiba, M. Shimizu, T. Shiosaki, L. A. Wills, R. Hiskes, S. A. Dicarolis, and J. Amano: Appl. Phys. Lett. 68 (1996) 2358. 44) H. Y. Guo, J. B. Xu, I. H. Wilson, Z. Xie, E. Z. Luo, S. B. Hong, and H. Yan: Appl. Phys. Lett. 81 (2002) 715. 45) S. Jesse, A. P. Baddorf, and S. V. Kalinin: Appl. Phys. Lett. 88 (2006) 062908. 46) S. Jesse, H. N. Lee, and S. V. Kalinin: Rev. Sci. Instrum. 77 (2006) 073702. 47) A. N. Morozovska, E. A. Eliseev, and S. V. Kalinin: Appl. Phys. Lett. 89 (2006) 192901. 48) J. Van Suchtelen: Philips Res. Rep. 27 (1972) 28. 49) E. Fukada and I. Yasuda: J. Phys. Soc. Jpn. 12 (1957) 1158. 50) L. Yasuda: J. Jpn. Orthop. Surg. Soc. 28 (1957) 267. 51) S. B. Lang: Nature 212 (1966) 704. 52) J. C. Anderson and C. Eriksson: Nature 227 (1970) 491. 53) A. A. Marino and B. D. Gross: Arch. Oral Biol. 34 (1989) 507. 54) E. Fukada: J. Phys. Soc. Jpn. 10 (1955) 149. 55) V. A. Bazhenov: Piezoelectric Properties of Wood (Consultants Bureau, New York, 1961). 56) E. Fukada: Biorheology 32 (1995) 593. 57) C. A. L. Bassett: Calcif. Tissue Res. 1 (1968) 252. 58) A. A. Marino and R. O. Becker: Nature 228 (1970) 473. 59) Mathematika 5.0, Wolfram Research. 60) J. Shin, B. J. Rodriguez, A. P. Baddorf, T. Thundat, E. Karapetian, M. Kachanov, A. Gruverman, and S. V. Kalinin: J. Vac. Sci. Technol. B 23 (2005) 2102. 61) A. Gruverman, D. Wu, B. J. Rodriguez, S. V. Kalinin, and S. Habelitz: Biochem. Biophys. Res. Commun. 352 (2007) 142. 62) H. J. Butt, B. Capella, and M. Kappl: Surf. Sci. Rep. 59 (2005) 1. 63) B. J. Rodriguez, S. Jesse, A. P. Baddorf, and S. V. Kalinin: Phys. Rev. Lett. 96 (2006) 237602. 64) K. Stephenson and T. Smith: in Scanning Probe Microscopy: Electrical and Electromechanical Phenomena on the Nanoscale, ed. S. V. Kalinin and A. Gruverman (Springer, New York, 2006). 65) B. T. Rosner and D. W. van der Weide: Rev. Sci. Instrum. 73 (2002) 2505. 66) S. Jesse and S. V. Kalinin: patent pending.

5685

Recent Advances in Electromechanical Imaging on the ...

Sep 7, 2007 - further development of the technique, including theory of the image formation mechanism as well as probe .... Application of the bias to the tip results in the surface .... between applied force, P, and concentrated charge, Q, with.

753KB Sizes 2 Downloads 252 Views

Recommend Documents

Recent advances on organocatalysed asymmetric Mannich ... - Arkivoc
of asymmetric Mannich reactions covering from 2007 to now under different organocatalytic ... introduced the one-pot catalytic asymmetric synthesis of pseudo-C2 -symmetric ... reported the one-pot organocatalytic reactions between α-.

Recent advances on organocatalysed asymmetric Mannich ... - Arkivoc
O. L-proline. DMSO up to 99% ee up to 90% yield up to >20:1 dr. Scheme 3. List. 19 introduced the one-pot catalytic asymmetric synthesis of pseudo-C2 - ...

Recent Advances in Dependency Parsing
Jun 1, 2010 - auto-parsed data (W. Chen et al. 09) ... Extract subtrees from the auto-parsed data ... Directly use linguistic prior knowledge as a training signal.

Recent Advances in Surfactant EOR.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Recent Advances in Surfactant EOR.pdf. Recent Advances in Surfactant EOR.pdf. Open. Extract. Open with. Sign

Recent advances in ketene chemistry - Arkivoc
Flash vacuum pyrolysis of N-(2-pyridyl)acetamides 78 generates ketenes 79 by elimination, and these are ...... Sun, L.; Liang, Z.; Ye, S. Acta Chim. Sinica 2014 ...

Recent advances in ketene chemistry - Arkivoc
method and used to interpret the excited state of the molecule.16 Electron scattering ... Ketene formation from ground-state oxygen atom reaction with ethylene.

Recent Advances in Nutritional Sciences
men. Subjects consumed a meal containing 15% of energy from protein, 55% from carbohydrate and 30% from fat, in the form of corn oil (CO) and animal fat or MCT oil (56% octanoate, 40% decanoate) in random order. Energy expendi- ture measurements were

Recent advances in ipso-nitration reactions - Arkivoc
©ARKAT USA, Inc. The Free ..... indicated, ipso-nitration with acetic anhydride/nitric acid ensures a good yield of p-nitrocalix[n]arenes; ...... Iyer, L. M. Formation and reactions of adducts from ipso nitration of nitroarenes, University of Victor

Recent advances in ipso-nitration reactions - Arkivoc
presence of a protonable site at the γ-position of the phenolic oxygen atom. Due to .... In those reactions, aryl-/heteroaryl-/polyaryl carboxylic acids with electron ...

Recent advances in asymmetric Strecker reactions - Arkivoc
of ketimines bearing electron-withdrawing and electron-donating substituents at different positions ...... Yan, H. L. Oh, J. S.; Lee, J. -W.; Song, C. E. Nat. Commun ...

Bioluminescence Recent Advances in Oceanic Measurements and ...
Bioluminescence Recent Advances in Oceanic Measurements and Laboratory Applications.pdf. Bioluminescence Recent Advances in Oceanic Measurements ...

Recent advances in asymmetric Strecker reactions - Arkivoc
Scheme 3. In 2011, Brigaud and coworkers32 reported a concise synthesis of enantiopure (S)- and (R)- α-Tfm-aspartic acid (Scheme 4) and α-Tfm-serine ...

Recent advances in asymmetric Strecker reactions - Arkivoc
Lu, W-Y.; Chen, P-R.; Lin, G-Q. Tetrahedron, 2008, 64, 7822. http://dx.doi.org/10.1016/j.tet.2008.05.113. 25. García, R. J. L.; Martín-Castro, A. M.; Tato, F.; Alonso, ...

Simultaneous elastic and electromechanical imaging ...
Both imaging and quantitative interpretation of SPM data on complex ... Stanford Research Instruments, and Model 7280, Signal Re- covery as ... Typical values were 1. =99 kHz .... nal in the center and enhanced PFM amplitude at the circum-.

Simultaneous elastic and electromechanical imaging by scanning ...
Received 3 March 2005; accepted 15 August 2005; published 20 September 2005. An approach for combined imaging of elastic and electromechanical ...

pdf-1425\advances-in-information-retrieval-recent-research-from-the ...
... the apps below to open or edit this item. pdf-1425\advances-in-information-retrieval-recent-resear ... formation-retrieval-the-information-retrieval-series.pdf.

recent advances in the ibm gale mandarin transcription ...
2 The author was at IBM T. J. Watson Research Center during this work. ABSTRACT. This paper ... The IBM GALE Mandarin broadcast speech transcription system operates in multiple .... We call this the dynamic topic LM. As shown in Fig.

Nanoscale Electromechanical and Mechanical Imaging ...
beyond visualization of the scale structure and actually allows evaluation of ... ning the sample, a three-dimensional map of the surface topography can be .... Also, AFAM data shows significantly higher level of the detail than the topographic ...

High Resolution Electromechanical Imaging of ...
Jun 16, 2006 - and Model 7280, Signal Recovery) and an external signal generation and data acquisition system. A custom- modified commercial liquid cell ...

cme on applied radiological physics and recent advances - IRIA Delhi
Nov 27, 2015 - Please mail your registration form along with cheque to –. Dr. Atin Kumar. Organising Secretary. Room No. 52, Department of Radiology,.

cme on applied radiological physics and recent advances - IRIA Delhi
Nov 27, 2015 - A K Singh beyond. 12.30-01.00 PM. Dual energy CT: Principle and ... Planning of radiology department. Dr. Vijay Kumar. 03.40-04.00 PM. TEA.

Recent Advances in Micro/Nanoscale Heat Transfer ...
Applications in Clean Energy Technologies. December 21-22, 2013. Indian Institute of Technology Ropar. Rupnagar, Punjab – 140001, INDIA. Please note: ...

Recent Advances in Batch Mode Reinforcement Learning - Orbi (ULg)
Nov 3, 2011 - Illustration with p=3, T=4 .... of the Workshop on Active Learning and Experimental Design 2010 (in conjunction with AISTATS 2010), 2-.