IOP PUBLISHING

NANOTECHNOLOGY

Nanotechnology 18 (2007) 424020 (10pp)

doi:10.1088/0957-4484/18/42/424020

Towards local electromechanical probing of cellular and biomolecular systems in a liquid environment Sergei V Kalinin1,2,5 , Brian J Rodriguez1,2 , Stephen Jesse1 , Katyayani Seal2 , Roger Proksch3 , Sophia Hohlbauch3 , Irene Revenko3 , Gary Lee Thompson4 and Alexey A Vertegel4 1

Materials Sciences and Technology Division, Oak Ridge National Laboratory, Oak Ridge, TN 37931, USA 2 The Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, TN 37931, USA 3 Asylum Research, Santa Barbara, CA 93117, USA 4 Department of Bioengineering, Clemson University, Clemson, SC 29634, USA E-mail: [email protected]

Received 23 April 2007, in final form 30 May 2007 Published 19 September 2007 Online at stacks.iop.org/Nano/18/424020 Abstract Electromechanical coupling is ubiquitous in biological systems, with examples ranging from simple piezoelectricity in calcified and connective tissues to voltage-gated ion channels, energy storage in mitochondria, and electromechanical activity in cardiac myocytes and outer hair cell stereocilia. Piezoresponse force microscopy (PFM) originally emerged as a technique to study electromechanical phenomena in ferroelectric materials, and in recent years has been employed to study a broad range of non-ferroelectric polar materials, including piezoelectric biomaterials. At the same time, the technique has been extended from ambient to liquid imaging on model ferroelectric systems. Here, we present results on local electromechanical probing of several model cellular and biomolecular systems, including insulin and lysozyme amyloid fibrils, breast adenocarcinoma cells, and bacteriorhodopsin in a liquid environment. The specific features of PFM operation in liquid are delineated and bottlenecks on the route towards nanometre-resolution electromechanical imaging of biological systems are identified. (Some figures in this article are in colour only in the electronic version)

etc. On a more fundamental level, cellular membranes are flexoelectric, i.e. possess linear coupling between electric field and strain gradient (or local radius of curvature) [4]. Biopolymers forming calcified and connective tissues are often piezoelectric [5], i.e. exhibit linear coupling between electric field and strain. Significant effort has been directed towards understanding the origins of electromechanical activity of these types of biological systems and its relationship to functionality [6–15]. However, due to the complex spatial structure of biological systems, this task requires real-space probing of electromechanical coupling on the nanoscale.

1. Introduction Coupling between electrical and mechanical phenomena is ubiquitous in biological systems. Multiple examples include phenomena such as voltage controlled muscular contractions [1], cell electromotility [2], and electromotor proteins [3]. In many cases, these couplings are the functional bases for processes such as transduction of acoustic signals into electrical pulses in the outer hair cell stereocilia, cardiac activity, energy storage in mitochondria, 5 Author to whom any correspondence should be addressed.

0957-4484/07/424020+10$30.00

1

© 2007 IOP Publishing Ltd Printed in the UK

Nanotechnology 18 (2007) 424020

S V Kalinin et al

Piezoresponse force microscopy (PFM) has emerged in the last decade as a primary tool for characterization of static and dynamic polarization phenomena in ferroelectric materials based on local electromechanical coupling [16–23]. PFM employs a conducting tip to probe the mechanical response of a surface to an applied bias. In ferroelectric materials, electromechanical response is directly coupled to the primary polarization order parameter through electrostrictive constants. Hence, mapping of local electromechanical activity and its evolution under applied bias (piezoresponse spectroscopy) and force provides insight into local polarization dynamics [24–30]. In the last several years, PFM has been extended to a number of polar piezoelectric materials such as III–V nitrides [31], ferroelectric polymers [32], and piezoelectric biopolymers [33–35]. In these systems, electromechanical activity is strongly related to local surface termination (III–V nitrides), presence and molecular orientation of electromechanically active proteins (calcified tissues), etc. Hence, probing the electromechanical activity can provide an insight into nanoscale surface structure and functionality. One of the key challenges in PFM imaging of biological systems is imaging in a liquid environment. This is required both to maintain a native-like environment for biomolecules and cells, and to control tip–surface forces and eliminate capillary interactions, thereby reducing sample damage. Conventional electrochemical SPMs typically operate in a very low frequency range (dc to ∼kHz) and require control of the probe and counter-electrode potentials with ∼mV resolution. In contrast, PFM operates in the 1 kHz–5 MHz frequency range and typically requires large (0.3–20 Vpp ) modulation voltages to measure small (∼3–500 pm V−1 ) electromechanical coupling coefficients. Implementation of PFM in liquid potentially allows (I) operation in the media compatible with biological systems, (II) precise control of tip–surface interactions through the elimination of capillary forces and control over Van der Waals interactions, which largely eliminates jump-to-contact instability, and (III) reduction of electrostatic interactions through matching of dielectric constants between solvent and substrate. However, factors such as (1) electrochemical reactions, (2) liquid effect on cantilever dynamics through added mass and fluid damping, (3) formation of electric double layers and (4) stray currents can significantly affect the signal formation mechanism and limit the potential for PFM imaging in a liquid environment. This broad range of contributing factors precludes unambiguous theoretical determination of the likelihood of successful electromechanical imaging by PFM in liquid. Recently, we have demonstrated PFM imaging of a model ferroelectric system (lead zirconate titanate) in a conductive liquid environment [36]. It was demonstrated that the use of high modulation frequencies (1–2 MHz) largely precludes electrochemical processes even for large biases (10 Vpp ) and minimizes liquid damping and added mass effects. Screening by mobile ions reduces the spatial extent of electrostatic fields, and allows imaging at high resolutions. Here, we present initial results on electromechanical imaging of a variety of biosystems in a liquid environment and analyse likely sources of contrast in liquid PFM images. Finally, we discuss the technical challenges, theoretical considerations, and expected outcomes for PFM in liquid environments.

2. Experiment 2.1. Principles of PFM 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, A1ω , of the tip deflection, A = A0 + A1ω cos(ωt + ϕ), induced by the application of the periodic bias, Vtip = Vdc + Vac cos(ωt), to the tip. The phase of the electromechanical response of the surface, ϕ , yields information on the sign of the piezoresponse. PFM is implemented on commercial SPM systems (Veeco NS-IIIA MultiMode, NanoMan V, and Asylum Research MFP-3D) equipped with additional function generators and lock-in amplifiers (DS 345 and SRS 830, Stanford Research Instruments, and model 7280, Signal Recovery). Measurements in ambient conditions were performed using Pt and Au coated tips (NSC-12 C, Micromasch, l ≈ 130 μm, resonant frequency ∼150 kHz, spring constant k ∼ 4.5 N m−1 ). Measurements in liquids were performed using a Cr–Au-coated CSC-38 cantilever with l ≈ 300 μm, resonant frequency ∼14 kHz, and spring constant k ∼ 0.05 N m−1 . Vertical PFM (VPFM) measurements were performed at frequencies of 50 kHz–1 MHz, which minimizes the longitudinal contribution to the measured vertical signal and allows imaging close to the resonances [37]. 2.2. Model systems 2.2.1. Lysozyme fibrils. Fibril formation and purification procedures were based on those reported by Vernaglia et al [38]. A 6 mg ml−1 chicken egg-white lysozyme (lysozyme, No L6876 from Sigma) stock solution was prepared in 20 mM potassium phosphate (KH2 PO4 ), adjusted to pH 6.3 ± 0.1 with 1 M potassium hydroxide. Dilution with 4 M guanidine hydrochloride (GuHCl, No 105696 from MP Biomedicals, Inc.) in 20 mM KH2 PO4 yielded ∼1.8 mg ml−1 lysozyme in ∼2.8 M GuHCl and 20 mM KH2 PO4 . Fibrillization was induced by incubating the solution for 24 h in a glass vial on a heated stir plate at 50 ◦ C with mild agitation (∼100 rpm). The fibrils were centrifuged at 14 100 g for 10 min, the supernatant was discarded, and the pellet was washed with the same volume of HPLC-grade water (Fisher) using gentle shaking for 10 min. This step was repeated for a total of 15 washes. The fibril suspension was diluted with HPLC-grade water to 0.25 mg ml−1 , and samples for PFM imaging were prepared by incubating 25 μl of the resulting suspension on freshly cleaved mica (No 71851-05 from Electron Microscopy Sciences) at room temperature for 2.5 min. Incubation was followed by 30 washes with 100 μl HPLC-grade water, and gentle drying with compressed nitrogen. Another 20 washes with 100 μl HPLC-grade water and finally gentle drying with compressed nitrogen completed the adsorption of lysozyme onto mica. Before imaging, 100 μl of HPLC-grade water was added onto the sample for solution PFM imaging. 2.2.2. Insulin amyloid fibrils. Insulin amyloid fibrils were formed and purified following the procedures reported by Guo and Akhremitchev [39]. Bovine insulin (No I5500 from 2

Nanotechnology 18 (2007) 424020

(a)

(d)

S V Kalinin et al

10 µm

(b)

(c)

(e)

(f)

Figure 1. (a), (d) Topography, (b), (e) PFM amplitude and (c), (f) phase images of a periodically poled lithium niobate surface in (a)–(c) air and (d)–(f) distilled water, respectively. Data scales are 20 nm, 10 V, and 20 V for (a)–(c), and 50 nm, 20 V, and 10 V for (d)–(f), respectively.

Tris, 150 mM KCl buffer, pH 8. To prepare the sample, 10 μl of the bacteriorhodopsin solution was added to 10 μl of buffer onto a freshly cleaved mica disc. The solution was incubated for 10 min then rinsed with 100 μl of buffer solution. Additional buffer was added for a final volume of 200 μl. Patches of bacteriorhodopsin were imaged with a metal coated tip in tapping mode with low drive and set-point amplitudes.

Sigma-Aldrich) was reconstituted to 5 mg ml−1 in 10 mM hydrochloric acid (HCl, Fisher Chemical). This solution was incubated in a polypropylene microcentrifuge tube in a block heater at 80 ◦ C for 48 h. The fibril suspension was diluted tenfold with 10 mM HCl to 0.5 mg ml−1 protein and was centrifuged at 3000g for 1 min to remove smaller aggregates from the longer fibrils. The precipitates were collected and washed by gentle shaking for 5 min with the same volume of 10 mM HCl as had been removed. This was repeated twice for a total of three washes. The fibril suspension was diluted 1:1 with 10 mM HCl to 0.25 mg ml−1 , and samples for PFM imaging were prepared by incubating 25 μl of the resulting suspension on freshly cleaved mica (No 71851-05 from Electron Microscopy Sciences) at room temperature for 1 min. Incubation was followed by five washes with 35 μl of 10 mM HCl and gentle drying with compressed nitrogen. The residual HCl was then removed by five additional washes with 35 μl of HPLC-grade water and finally gentle drying with compressed nitrogen. Before imaging, 100 μl of HPLC-grade water was added onto the sample for solution PFM imaging.

3. Results 3.1. Liquid PFM imaging of model ferroelectric systems To illustrate the feasibility of electromechanical imaging in liquid environment, the PFM was performed in air and in DI water on a periodically poled lithium niobate (PPLN, Crystal Technology) as a model system. The electromechanical response of PPLN in the point contact geometry of a PFM experiment is well known both from calculations [40] and direct measurements [41] and is ∼12 pm V−1 . PPLN is thus a convenient model system for calibration of the system, i.e. determination of the conversion factor between the lockin amplifier output and local electromechanical response. Furthermore, the noise level in the PFM measurements can be determined from the dispersion of the signal on a single domain. If the vertical sensitivity of the system is known independently, the electrical contact quality between the tip and the surface can be estimated. Note that both types of calibration are generally applicable only for frequencies well below the first resonance of the cantilever. Otherwise, cantilever dynamics, which is strongly dependent on the contact stiffness of the tip–surface junction, will provide a sample-dependent multiplicative factor to the signal [42]. The PFM amplitude and phase images taken from the PPLN in air and distilled water are shown in figure 1. Both images demonstrate clear domain contrast unrelated to topographic features in amplitude images. The phase

2.2.3. Breast adenocarcinoma cells. Breast adenocarcinoma cells (MCF7 cell line) were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM) with glucose and L-glutamine (No D5523 from Sigma-Aldrich) supplemented with 10% FBS (No ES1055 from Biomeda), penicillin–streptomycin (No P0781 from Sigma-Aldrich), and non-essential amino acids (No M7145 from Sigma-Aldrich). The cells were plated onto pretreated Corning 60 mm polystyrene Petri dishes (No 430166 from Fisher Scientific) and were maintained in a CO2 incubator at 37 ◦ C until imaging. 2.2.4. Bacteriorhodopsin. Bacteriorhodopsin from Halobacterium salinarium (No B0184 from Sigma-Aldrich) was diluted to a final imaging concentration of 50 μg ml−1 in 10 mM 3

Nanotechnology 18 (2007) 424020

S V Kalinin et al

10 V

5V

2V

(b)

(c)

(d)

(e)

(f)

(g)

(h)

(i)

(a)

400 nm

Figure 2. PFM images of a lysozyme fibril adsorbed onto mica. The scan size is 2.97μm × 2.97 μm, with a scan rate of 1 Hz and 512 samples per line. Ac biases, 10, 5, and 2 Vac , as shown in the left, middle, and right rows, respectively, were applied with a frequency of 383 kHz, a 0◦ phase angle, and a lock-in bandwidth of 1 kHz. The first row shows topographic images, the second row piezoresponse phase images, and the third row piezoresponse amplitude images. Z scales are 50 nm, 360◦ , and 2.5 mV for the topographic, piezoresponse phase, and piezoresponse amplitude images, respectively. The height of the feature ranged from about 10 nm to a maximum of about 25 nm, consistent with the diameters of lysozyme fibrils reported by Vernaglia et al [38]. Note that the amplitude feature disappears with decrease in applied ac bias from 10 Vac (g) to 2 Vac (i) while the topographic and phase images reveal that the fibril bundle remains present.

contrast in both cases illustrates phase change by 180◦ between antiparallel domains. The noise level and effective domain wall width in the liquid image are higher, indicative of more significant spreading of the electric field anticipated in the high dielectric constant (ε = 80 in water versus ε = 1 in air). However, clear qualitative agreement between PFM images in both environments serves as a strong indicator of the electromechanical nature of the contrast.

structure is discernable in PFM amplitude images, while the phase response appears to be uniform within a fibril. Similar behaviour is observed for the insulin fibril in figure 3, shown for two driving voltages. The strong dependence of the phase contrast on driving amplitude and small phase shifts (180◦ ) between dissimilar regions suggests significant contribution of the electrostatic forces (modified by the presence of liquid) to the PFM signal. Based on the visual inspection of the topographic, PFM phase and PFM amplitude data it is obvious that (1) the apparent spatial resolution in the PFM amplitude image is higher compared to topography and (2) phase images show clear contrast on the parts of the fibril in which topographic height is at the detection limit. Below, we explain these observations by suggesting that the primary origin of the PFM signal is the variability of the local elastic properties between the substrate and the fibril. This elastic contrast is convoluted with electromechanical and electrostatic responses, precluding unambiguous determination of the latter. However, the increased resolution and detection limit clearly illustrate the usefulness of this detection method.

3.2. Liquid PFM imaging of amyloid fibrils After verification on the ferroelectric system with known electromechanical response, PFM was applied to a variety of biological systems. Figure 2 shows topographic ((a)–(c)) and PFM images of a lysozyme fibril as a function of imaging bias, Vac . PFM phase images are shown in figures 2(d)–(f) and PFM amplitude images are shown in figures 2(g)–(i) for 10, 5, and 2 V, respectively. The piezoresponse amplitude linearly depends on the applied bias, and the phase response is stronger for higher bias. The fibril does not appear to degrade with repeated scanning or application of ac bias. Some internal 4

Nanotechnology 18 (2007) 424020

S V Kalinin et al

10 V

(a)

(b)

(c)

(e)

(f)

120 nm

2V

(d)

Figure 3. PFM images of an insulin fibril adsorbed onto mica. The scan size is 603 nm × 603 nm, with a scan rate of 1 Hz and 512 samples per line. Ac biases, 10 Vac and 2 Vac , as shown in the top and bottom rows, respectively, were applied with a frequency of 437.8 kHz, a 0◦ phase angle, and a lock-in bandwidth of 1 kHz. The first column shows topographic images, the second column piezoresponse phase images, and the third column piezoresponse amplitude images. Z scales are 25 nm, 180◦ , and 2.0 mV for the topographic, piezoresponse phase and piezoresponse amplitude images, respectively. The height of the feature ranged from 5 to 10 nm, which is consistent with the range of insulin fibril heights reported by Guo and Akhremitchev [39]. Note that the amplitude feature decreases and fades with decrease in applied ac bias from 10 Vac (e) to 2 Vac (f), while the topographic and phase images reveal that the fibril bundle remains present.

(a)

16 µm

(b)

(c)

Figure 4. Live breast adenocarcinoma cells (MCF7 cell line) imaged in DMEM culture medium. (a) Topography (180 nm z -scale), (b) PFM amplitude, and (c) PFM phase images. Images were acquired in contact mode using the MFP-3D.

3.3. Imaging of cellular systems

performed using insulated or shielded probes, in which the central part of the tip is insulated from the solution except for the region near the probe apex.

The extension of liquid PFM imaging to viable cells is illustrated in figure 4, showing topography, PFM amplitude, and PFM phase of living breast adenocarcinoma cells. Only weak contrast between different cells or at cell boundaries is observed in the PFM amplitude image. However, the cells with larger height profiles appear to have a slightly different PFM phase response than the surrounding cells. In addition, cell boundaries are resolved in the PFM phase image. Similarly to the amyloid and insulin fibrils, these observations suggest the possibility of a strong elastic contribution to the PFM signal. The reason for the relatively poor quality of cell PFM images is the presence of electrolytes in the cell culture media because high conductivity of the solution results in the uniform biasing of the liquid. To avoid this problem, the imaging should be

3.4. Electromechanical imaging of bacteriorhodopsin and PFM resolution The general analysis of PFM image formation suggests that the electromechanical contribution to the signal is favoured by high effective stiffness of the system. This is due to the fact that the electromechanical contribution is independent of spring constant, and electrostatic contribution scales reciprocally with effective spring constant [42]. Furthermore, optimal imaging conditions with minimal topographic cross-talk correspond to flat surfaces. Hence, we have chosen the bacteriorhodopsin (BR) purple membrane as a relatively stiff and topographically 5

Nanotechnology 18 (2007) 424020

S V Kalinin et al

illustrates that the use of intermittent contact mode with a metal coated tip in liquid allows molecular resolution to be achieved.

4. Towards electromechanical imaging in liquids

(a)

200 nm

The results summarized in section 3 illustrate that electromechanical imaging in a liquid environment produces measurable contrast on a variety of biological systems. The presence of the contrast per se evidences measurable electromechanical and electrostatic interactions between the probe and the surface in liquid. In many (albeit not in all) cases, the contrast is visually distinct from surface topography or its first and second derivatives, suggesting that it is unrelated or only partially related to topographic cross-talk. Furthermore, contrast scales linearly with probing voltage, consistent with its (linear) electromechanical (e.g. piezoelectric or flexoelectric) origins. Based on these observations, we analyse the possible origins of observed signal and summarize the requirements for quantitative studies of electromechanical activity in liquids. Local quantitative mapping of electromechanical activity in liquid environment requires us to

(b)

Figure 5. Dual-ac mode images of bacteriorhodopsin imaged in buffer solution with a metal coated tip. Shown are (a) surface topography (6 nm z -scale) and (b) the amplitude of the electromechanical response signal.

flat model system for a biological system. Furthermore, the strongly asymmetric structure of the BR with significant intrinsic dipole moment suggests the high likelihood of intrinsic electromechanical response. Shown in figure 5 are the topographic and PFM amplitude images of the BR. In this case, imaging was performed using ac mode (mechanical tip oscillation at first resonance) to obtain topographic data. Electrical modulation was performed at the second resonance for amplitude [43]. The top surface of the membrane exhibits islands with double height, corresponding to membrane folding. The PFM amplitude image shows the signal of equal strength within the membrane surface and the islands, with characteristic depression at the boundary. This behaviour is qualitatively similar to that expected for ferroelectric or piezoelectric domains. Based on these observations, and given minimal potential for topographic or elastic cross-talk on the uniform surface, we attribute the observed contrast to the difference in electromechanical response (e.g. opposite sign of piezoelectric coefficients) between the extracellular and cytoplasmic regions of the BR. The experiments with BR also establish the potential for molecular-level spatial resolution. The operation of PFM requires reliable control of both dc and ac tip bias in the 1 kHz–1 MHz range. Operation in liquid environment brings additional limitations on the conductance of the tip coating, which should be greater than that of the liquid to stray current paths through the solution. This necessitates the use of metallic or metal coated tips, which typically have significantly higher radii of curvature compared to Si or Si3 N4 probes. Figure 6

(a)

200 nm

(1) concentrate the probing ac field, (2) concentrate and/or control local electrochemical potential through the dc field, (3) detect local surface displacement induced by ac potential (electromechanical response), (4) separate electromechanical response from electrostatic capacitive interactions, and (5) avoid or separate topographic and elastic cross-talk that can contribute to the measured signal. The theory of PFM, including electrostatic and strain field structure induced by the probe, resolution theory, local and non-local contributions to signal, surface–tip signal transfer and its frequency dispersion, are well studied in the context of ambient PFM [44]. Below, we briefly discuss principles of PFM imaging in ambient environment, and analyse some of the factors that can affect imaging in liquids. 4.1. Force interactions in liquid and air The key element of any scanning probe microscopy (SPM) method is the mechanism of tip–surface interactions that determine the strength of the force signal detected by the cantilever, the locality of the imaging, and the

(b)

60 nm

(c)

20 nm

Figure 6. Topography images of bacteriorhodopsin acquired in buffer solution with a metal coated tip using the NanoScope MultiMode at ˚ respectively. different magnifications. Z scales are 12 nm, 2 nm, and 4 A,

6

Nanotechnology 18 (2007) 424020

S V Kalinin et al

in solution gradually decays at distances of the order of the Debye length of the solution, and the electrostatic contribution is significantly minimized compared to ambient or vacuum imaging.

primary response mechanism. In the context of PFM, primary interactions include electrostatic interactions and electromechanical coupling in the material and electric double layers. The experimentally measured piezoresponse amplitude is A = Ael + Apiezo + Anl , where Ael is the electrostatic contribution, Apiezo is the electromechanical contribution and Anl is the non-local contribution due to capacitive cantilever– surface interactions [45, 46]. Quantitative PFM imaging requires Apiezo to be maximized to achieve predominantly electroelastic contrast. The origins of electromechanical contrast and its relationship to local materials properties for piezoelectric materials have been extensively studied [47, 48]. Piezoelectric coupling is specific to materials, and thus is not affected by the environment. At the same time, electrostatic force interactions are strongly mediated by the medium. In ambient environment or in vacuum, the tip interacts with the surface through long-range electrostatic forces, Felamb = Cz (z)(Vt − Vs )2 , where Vt is the tip potential, Vs is the surface potential, and Cz (z) is the tip–surface capacitance gradient. In the low frequency limit, the piezoresponse (PR) signal, i.e. the first harmonic component of tip oscillations induced by periodic bias applied to the tip, is k1 P Ramb = αa (h)d˜33 k1 + k     Csphere + Ccone C + (1) + cant (Vdc − Vs ), k1 + k 24k

4.2. Ac and dc field concentration The second key component of the PFM imaging is the concentration of the ac and dc field in the tip–surface junction and detection of local response. Notably, simple observation of high-resolution contrast is not enough to demonstrate the high localization of an ac electric field. Rather, the locality of the strain transfer through the mechanical tip–surface contact is a sufficient explanation. In the absence of additional information, the relative size of electrical (ac field region) and mechanical contact cannot be established and the spatial resolution provides only an estimate of the lower of the two. The distribution of the dc field however can be probed directly by the polarization switching experiments in the liquid environment. A transition of ferroelectric switching behaviour from localized to uniform switching depending on the choice of the solvent established that imaging is possible at conductivities far larger than allowed for localized switching [50]. Furthermore, these results illustrated the degree to which the spatial extent of a dc field can be controlled in solution. Application of local dc pulses using conventional metal coated cantilevers is possible only in less conductive non-aqueous solvents such as isopropanol or methanol.

where d˜33 is the effective electromechanical response of the material, αa (h) is the ratio of the ac tip potential to the ac surface potential (i.e. the potential drop in the tip–surface gap), k1 is the spring constant of the tip–surface junction, k is the    , Ccone , and Ccant spring constant of the cantilever, and Csphere are capacitance gradients due to the spherical and conical parts of the tip and the cantilever. Vdc is the dc potential offset of the tip bias, and h is the tip–surface separation. In ambient conditions αa (h) = 1 for h < 0 (contact), i.e., the response is independent of penetration depth, and αa (h)  1 for h > 0 (non-contact), giving rise to well known limits of PFM and Kelvin probe force microscopy. In liquid, the PFM contrast is strongly mediated by the presence of mobile ions that screen electrostatic tip–surface interactions [36]. For the sphere–plane system [49],

Fell (z) =

εε0 R 2Vt Vs exp(h/λD ) − (Vt2 + Vs2 ) . λD exp(2h/λD ) − 1

4.3. Contact mechanics The third element determining PFM response is the signal transfer from material to the tip, determined by the cantilever spring constant, k , and tip–surface spring constant, k1 . The importance of variations of effective spring constant on signal directly follows from equation (1), in which variations in local stiffness couple to electromechanical signal through the capacitive forces. This coupling will be the origin of both topographic and elastic cross-talks through variations in tip– surface contact stiffness. These effects can be analysed using a simple Hertzian model for the tip–surface contact. The relationship between the indentation depth, h , tip radius of curvature, R0 , and load, P , is [51] 2  3 P 3 − 13 h= R0 , (4) 4E ∗

(2)

Electrostatic interaction in liquid is short range and decays exponentially for h > λD . Thus, the PFM signal in liquid (for h > 0) is

P Rl = α(h)d33

k1l

where E ∗ is the effective Young’s modulus of the tip–surface system. The √ contact radius, a , is related to the indentation depth as a = h R0 . The contact stiffness is k1 = (∂h/∂ P)−1 , and from equation (42) k1 = 2a E ∗ , or  1 2 k1 = 2 E ∗ h R0 = (6 P E ∗ R0 ) 3 . (5)

k1l εε0 R 2Vs exp(h/λD ) 1 + , (3) λD exp(2h/λD ) − 1 k1l + k +k

where the first term is the piezoelectric response, k1l is the tip–surface spring constant in liquid, and the second term is the response due to the electrostatic coupling in the double layer. The contribution from the conical part of the tip and cantilever are absent for R  λD . The screening coefficient α(h) = 1 for h  λD , i.e. when the tip touches the surface, and α(h) = 0 if the electric double layers around the tip and the surface do not overlap. Thus, the electromechanical response

Shown in figure 7(a) is the distance dependence of the electrostatic and electromechanical contributions to the PFM signal calculated for R = 50 nm, Vdc = 0.1 V, E ∗ = 100 GPa, d3 = 50 pm V−1 , and k = 1 and 40 N m−1 . The distance dependence of the capacitive tip–surface forces and tip-induced surface potential was calculated neglecting the changes in the sphere area due to contact. The electrostatic 7

Nanotechnology 18 (2007) 424020

S V Kalinin et al

Figure 7. (a) Indentation depth dependence for electromechanical (blue) and electrostatic (red) contributions for a cantilever with k = 1 N m−1 (solid) and k = 40 N m−1 (dashed). (b) Indentation depth dependence as a fraction of the electromechanical contribution.

Figure 8. (a) Indentation force–distance dependence in the JKR model. (b), (c) force dependence of indentation stiffness. (d) Force dependence of contact radius. Shown are curves for σ = 0 (solid), 10−3 (dot), 10−2 (dash), 10−1 (dash–dot), and 1 J m−2 (dash–dot–dot).

model, the contact radius is

contribution decreases rapidly with penetration depth due to changes in the tip–surface stiffness constant. In comparison, shown in figure 7(b) is the relative electromechanical contribution depending on penetration depth. Note that in the Hertzian model the electromechanical signal dominates for a ˚ corresponding to a contact radius penetration depth of ∼1 A, of the order of ∼2 nm for R = 50 nm, imposing a limit on the spatial resolution of the technique.

a3 =

  R0  P + 3 σ π R + 6σ π R0 P + (3σ π R0 )2 , 0 E∗

(6)

where σ is the work of adhesion, P is the load and the indentation depth is

    a2 2 r0 3/2 , h= 1− R0 3 a

The analysis becomes more complicated if adhesive effects are taken into account. In this case, the contact mechanics are described by the Johnson–Kendall–Roberts (JKR) model, applicable for ambient conditions. In the JKR

(7)

where r03 = 6σ π R 2 E ∗ is the contact radius at zero force. Shown in figure 8(a) are force versus indentation depth curves calculated for σ = 0 (Hertzian), 10−3 , 10−2 , 10−1 , and 8

Nanotechnology 18 (2007) 424020

S V Kalinin et al

−2

and topography can drive the system through the resonance, with associated 180◦ change in phase response and spurious changes in the amplitude. From the structure of equation (1), the following contributions to the cross-talk signal in PFM can be differentiated:

1 J m . Shown in figures 8(b) and (c) are corresponding contact stiffnesses. Note that adhesive contact results in a rapid change of contact stiffness from zero to the value corresponding to contact, resulting in a well defined boundary between free and bound cantilevers. Finally, shown in figure 8(d) is the force dependence of the contact area. Even for a small work of adhesion, the contact radii at zero force are relatively large, of the order of nanometres. Corresponding contact stiffnesses are of the order of 100–1000 N m−1 , well above the typical spring constant of the cantilever. Thus, for most cantilevers (0.01–50 N m−1 ), jump-tocontact instability in air distinguishes the regions with k1  k (free air, k1 = 0) and k1  k (in contact k1 ∼ 1000 N m−1 ). These regions correspond to the purely electrostatic interaction in Kelvin probe force microscopy (KPFM), and predominantly electromechanical contrast in PFM. Conversely, in liquid the lack of capillary forces and reduced magnitude of van der Waals (vdW) forces suggests that k1 can vary continuously when approaching the surface. Hence, there is no well defined boundary between predominantly electromechanical and purely electrostatic contrasts, and hence PFM and KPFM regimes are not well defined. Notably, similar behaviour is anticipated for the non-contact AFM and KPFM in ultrahigh vacuum. In both Hertzian and JKR models, the transition to the predominantly electromechanical contrast occurs for contact areas larger than a certain critical value. From equation (1), when non-local electrostatic components are negligible, this condition can be generalized for materials with arbitrary properties as

a > a∗ =

 (Vdc − Vs ) Csphere

2α(h)d3 E ∗

,

• variations of long-range electrostatic interactions due to  + capacitive terms dependence on topography; Csphere  Ccone , and local work functions, Vs ; • variations of effective electromechanical response with topography, d˜33 ; • variation of electrical contact quality, αa (h); • variations of local contact stiffness, k1 . Topographic effects on scanning Kelvin probe microscopy, a non-contact (weakly dependent on interaction area), lift-mode technique, have revealed that forces are larger within topographic depressions, but that these relatively weak long-range effects often show up as offsets in the signal [52]. Therefore, capacitive terms are unlikely to produce sharp cross-talk contrast on the 1–100 nm length scale. The detailed studies of the voltage-dependent contact mechanics of piezoelectric materials, as well as simple dimensionality arguments, suggest that electromechanical response is only weakly dependent on surface topography. The electrical contact quality is expected to be only weakly dependent on contact area for a good contact (fringing field in capacitor), but is highly sensitive to surface contamination. However, contamination is expected to be detectable directly from the apparent topographic image and, in the absence of such information, its presence cannot be established independently. Finally, the most significant contribution to PFM crosstalk is the effect of topography on contact stiffness, resulting in strong topography-related signal in the resonance-enhanced PFM. This will result in the fine structure on the image, since the resolution is determined by the tip–surface contact area and large-scale variations in topography do not contribute to signal. The magnitude of this effect is studied in detail elsewhere, and it is shown that due to the specificity of the PFM driving mechanism, the stable imaging in the vicinity of the resonance can be achieved only using fast sweeps, sampling a segment of the amplitude–frequency curve in the vicinity of the resonance [53]. This limitation is especially pronounced for weakly piezoelectric materials such as calcified and connective tissues with small (1–5 pm V−1 ) coupling coefficients. The detailed analysis of PFM cross-talk and optimal operation regimes will be reported later.

(8)

where a ∗ is the critical contact radius corresponding to equality of the electrostatic and electromechanical contributions to the signal. For classical ferroelectric materials (100 GPa, 50 pm V−1 ), this condition becomes a > a ∗ = 5(Vdc − Vs ) A V−1 . For soft materials (10 GPa, 5 pm V−1 ), a > a ∗ = 50(Vdc − Vs ) nm V−1 . From this simple estimate, the resolution of PFM on hard materials can potentially achieve sub-nanometre resolution provided that the electrostatic contribution to the signal is minimized. For soft systems with small response coefficients (∼1–100 pm V−1 ), the signal is likely to represent the convolution of electrostatic and electromechanical contributions. At the same time, for systems with large electromechanical responses (i.e. many cells have ∼104 higher electromechanical coupling than inorganic piezo-and ferroelectrics), electromechanical response can be expected to be observable directly.

5. Summary We have studied in liquid the electromechanical properties of a variety of biological systems, including amyloid insulin fibrils, carcinoma cells, and purple membranes, as well as a reference ferroelectric surface. The model PPLN system illustrates high veracity electromechanical contrast in ambient and in liquid, clearly associated with ferroelectric domains and observable even for relatively small (∼12 pm V−1 ) electromechanical response. For all biological systems, strong amplitude and phase contrast are observed in the PFM image. These data typically illustrate a level of detail exceeding the topographic

4.4. Topography effect (cross-talk) During imaging of multiphase materials, or objects with high aspect ratios, care must be taken to minimize topographic and elastic cross-talk with the electromechanical response through electrostatic interactions and variations in contact stiffness. These effects are expected to be particularly important when imaging at frequencies close to contact resonances of the tip. In this case, even minute changes in local properties 9

Nanotechnology 18 (2007) 424020

S V Kalinin et al

resolution, and capacity for detection of signal even from small objects invisible on the topographic image. The PFM contrast can originate from a variety of sources, including local electromechanical response, variations in contact stiffness due to topography and elasticity, and electrostatic forces. In many systems, these sources of contrast cannot be distinguished unambiguously. However, studies on the purple membranes clearly demonstrate the difference in electromechanical response between the cytoplasmic and extracellular sides. This behaviour is anticipated for the membrane with built-in dipole moment. Furthermore, these results suggest that imaging with molecular resolution in biological systems is feasible with metal-coated probes. The significant bottleneck for future electromechanical studies in liquid is the need to concentrate both ac and dc electric field produced by the probe. This is required both to improve the resolution and to minimize the stray currents in liquid and electrochemical reactions. This necessitates the development of shielded AFM probes, as suggested by a number of groups [54–56].

Acknowledgments Research sponsored in part (SJ and SVK) by the Division of Materials Sciences and Engineering (theory of PFM), ORNL LDRD funding (BJR, SJ, and SVK) (PFM imaging in liquid) and ORNL SEED funding (BJR and SVK) (biological PFM), Office of Basic Energy Sciences, US Department of Energy at Oak Ridge National Laboratory, managed and operated by UT-Battelle, LLC. Research was also supported through CNMS user proposals No 2005-075 and No 2006-049, and NSF No CMS-0619739 (GLT and AAV). The authors (BJR, SVK) are also grateful for financial support from Asylum Research and for the use of their imaging facilities.

References [1] Plonsey R and Barr R C 2000 Bioelectricity: A Quantitative Approach (New York: Kluwer) [2] Raphael R M, Popel P S and Brownell W E 2000 Biophys. J. 78 2844 [3] Liberman M C, Gao J, He D Z Z, Wu X, Jia S and Zuo J 2002 Nature 419 300 [4] Petrov A G 2001 Biochim. Biophys. Acta 1561 1 [5] Fukada E 1995 Biorheology 32 593 [6] Anderson J C and Eriksson C 1970 Nature 227 491 [7] Marino A A and Becker R O 1970 Nature 228 473 [8] Smith J W 1968 Nature 219 157 [9] Reinish G B and Nowick A S 1975 Nature 253 626 [10] Marino A A and Becker R O 1975 Nature 253 627 [11] Athenstadt H 1970 Nature 228 830 [12] Anderson J C and Eriksson C 1968 Nature 218 166 [13] Morris R W and Kittleman L R 1967 Science 158 368 [14] Lang S B 1969 Science 165 287 [15] Andrew C, Bassett L and Pawluk R J 1972 Science 178 982 [16] G¨unther P and Dransfeld K 1992 Appl. Phys. Lett. 61 1137 [17] Franke K, Besold J, Haessler W and Seegebarth C 1994 Surf. Sci. Lett. 302 L283 [18] Eng L M, Grafstrom S, Loppacher C, Schlaphof F, Trogisch S, Roelofs and Waser R 2001 Adv. Solid State Phys. 41 287 [19] Gruverman A, Auciello O and Tokumoto H 1998 Annu. Rev. Mater. Sci. 28 101 [20] Gruverman A, Auciello O and Tokumoto H 1996 J. Vac. Sci. Technol. B 14 602

10

[21] Gruverman A, Auciello O and Tokumoto H 1996 Appl. Phys. Lett. 69 3191 [22] Hidaka T, Maruyama T, Saitoh M, Mikoshiba N, Shimizu M, Shiosaki T, Wills L A, Hiskes R, Dicarolis S A and Amano J 1996 Appl. Phys. Lett. 68 2358 [23] Eng L M, G¨untherodt H-J, Schneider G A, K¨opke U and Mu˜noz Salda˜na J 1999 Appl. Phys. Lett. 74 233 [24] Guo H Y, Xu J B, Wilson I H, Xie Z, Luo E Z, Hong S B and Yan H 2002 Appl. Phys. Lett. 81 715 [25] Kalinin S V, Gruverman A and Bonnell D A 2004 Appl. Phys. Lett. 85 795 [26] Woo J, Hong S, Setter N, Shin H, Jeon J-U, Pak Y E and No K 2001 J. Vac. Sci. Technol. B 19 818 [27] Rodriguez B J, Nemanich R J, Kingon A, Gruverman A, Kalinin S V, Terabe K, Liu X Y and Kitamura K 2005 Appl. Phys. Lett. 86 012906 [28] Molotskii M I and Shvebelman M M 2005 Phil. Mag. 85 1637 [29] Abplanalp M 2001 Piezoresponse scanning force microscopy of ferroelectric domains PhD Thesis Swiss Federal Institute of Technology, Zurich [30] Agronin A et al 2006 J. Appl. Phys. 99 104102 [31] Rodriguez B J, Gruverman A, Kingon A I, Nemanich R J and Ambacher O 2002 Appl. Phys. Lett. 80 4166 [32] Rodriguez B J, Jesse S, KalininS V, Kim J, Ducharme S and Fridkin V M 2007 Appl. Phys. Lett. 90 122904 [33] Halperin C, Mutchnik S, Agronin A, Molotskii M, Urenski P, Salai M and Rosenman G 2004 Nano Lett. 4 1253 [34] Kalinin S V, Rodriguez B J, Jesse S, Thundat T and Gruverman A 2005 Appl. Phys. Lett. 87 053901 [35] Kalinin S V, Rodriguez B J, Jesse S, Thundat T, Grichko V, Baddorf A P and Gruverman A 2006 Ultramicroscopy 106 334 [36] Rodriguez B J, Jesse S, Baddorf A P and Kalinin S V 2006 Phys. Rev. Lett. 96 237602 [37] Kalinin S V, Rodriguez B J, Jesse S, Shin J, Baddorf A P, Gupta P, Jain H, Williams D B and Gruverman A 2006 Microsc. Microanal. 12 206 [38] Vernaglia B A, Huang J and Clark E D 2004 Biomacromolecules 5 1362 [39] Guo S and Akhremitchev B B 2006 Biomacromolecules 7 1630 [40] Kalinin S V, Karapetian E and Kachanov M 2004 Phys. Rev. B 70 184101 [41] Gopalan V et al 2007 submitted [42] Jesse S, Baddorf A P and Kalinin S V 2006 Nanotechnology 17 1615 [43] Proksch R 2006 Appl. Phys. Lett. 89 113121 [44] Alexe M and Gruverman A 2004 Nanoscale Characterization of Ferroelectric Materials (Berlin: Springer) [45] Franke K, Huelz H and Weihnacht M 1998 Surf. Sci. 415 178 [46] Hong S, Woo J, Shin H, Jeon J U, Park Y E, Colla E L, Setter N, Kim E and No K 2001 J. Appl. Phys. 89 1377 [47] Scrymgeour D A and Gopalan V 2005 Phys. Rev. B 72 024103 [48] Hong S 2004 Nanoscale Phenomena in Ferroelectric Thin Films (Norwell, MA: Kluwer) [49] Bowen W R, Filippov A N, Sharif A O and Starov V M 1999 Adv. Colloid Interface Sci. 81 35 [50] Rodriguez B J, Jesse S, Kim S H, Baddorf A P and Kalinin S V 2007 Phys. Rev. Lett. 98 247603 [51] Timoshenko S and Goodier J N 1951 Theory of Elasticity (New York: McGraw-Hill) [52] Efimov A and Cohen S R 2000 J. Vac. Sci. Technol. A 18 1051 [53] Jesse S, Mirman B and Kalinin S V 2006 Appl. Phys. Lett. 89 022906 [54] Smith T J and Stevenson K J 2006 Scanning Probe Microscopy: Electrical and Electromechanical Phenomena on the Nanoscale vol 1, ed S V Kalinin and A Gruverman (New York: Springer) p 280 [55] Rosner B T and van der Weide D W 2002 Rev. Sci. Instrum. 73 2505 [56] Hunt T and Westervelt R 2007 unpublished

Towards local electromechanical probing of ... - Semantic Scholar

Sep 19, 2007 - (Some figures in this article are in colour only in the electronic .... from Electron Microscopy Sciences) at room temperature for ..... These data.

2MB Sizes 0 Downloads 429 Views

Recommend Documents

Towards Regional Elastography of Intracranial ... - Semantic Scholar
to the deformation field and strain maps of the reference measurements. Figure 1 Isometric view of the patient geometry. The surface is divided in three regions: ...

Towards Regional Elastography of Intracranial ... - Semantic Scholar
to the deformation field and strain maps of the reference measurements. ... region clustering technique applied to the strain maps in order to reduce the number.

A Bidirectional Transformation Approach towards ... - Semantic Scholar
to produce a Java source model for programmers to implement the system. Programmers add code and methods to the Java source model, while at the same time, designers change the name of a class on the UML ... sively studied by researchers on XML transf

Probing the Chiral Anomaly with Nonlocal ... - Semantic Scholar
Sep 2, 2014 - linearly from degeneracy points at which two energy bands meet. .... generation magnetic field Bg, a valley imbalance Δμ is created via the chiral anomaly ...... nonlocal transport, we note that an alternative approach would be ...

A Bidirectional Transformation Approach towards ... - Semantic Scholar
to produce a Java source model for programmers to implement the system. Programmers add code and methods to ... synchronized. Simply performing the transformation from UML model to Java source model again ... In: ACM SIGPLAN–SIGACT Symposium on Pri

Aeroengine Prognostics via Local Linear ... - Semantic Scholar
The application of the scheme to gas-turbine engine prognostics is ... measurements in many problems makes application of ... linear trend thus detected in data is used for linear prediction ... that motivated their development: minimizing false.

Towards a Semantic-Aware File Store - Semantic Scholar
CDN more efficient. An- other related application is to support data hoarding for mobile users. Before disconnected from the network, all frequently used data for ...

Multi-View Local Learning - Semantic Scholar
Recently, another type of methods, which is based on data graphs have aroused considerable interests in the machine learning and data mining community. .... problem, where a real valued fi, 1 ≤ i ≤ l + u, is assigned to each data point xi. For an

Towards Automatic Generation of Security-Centric ... - Semantic Scholar
Oct 16, 2015 - ically generate security-centric app descriptions, based on program analysis. We implement a prototype ... Unlike traditional desktop systems, Android provides end users with an opportunity to proactively ... perceive such differences

Self-tracking cultures: towards a sociology of ... - Semantic Scholar
can be competitively compared with other self-trackers. (for example the cycling platform Strava). Apps are ..... tracking. Computer software and hardware developers, manufacturers and retailers, software coders, ..... living in rural and remote area

Bandwidth and Local Memory Reduction of Video ... - Semantic Scholar
Aug 6, 2009 - MB, the associated search range data is first loaded to the on- chip local SRAM, where 8-bit data is stored for each pixel. The IME processing unit, which is designed with bit-truncation technique, only accesses 5 bits for each pixel to

Local search characteristics of incomplete SAT ... - Semantic Scholar
Department of Computer Science. University of ... Of course, local search is incomplete and can- ... to keep a tabu list (Mazure, Saпs, & Grйgoire 1997) or break.

On local estimations of PageRank: A mean field ... - Semantic Scholar
to rank the most important hits in the top screen of results. One ... service and information providers an effective new marketing tool. ...... Internet Mathematics,.

Breaking Out of Local Optima with Count ... - Semantic Scholar
Many statistical learning problems in NLP call for local model search methods. ... one good way to construct such steps in a parame- ter space is by forgetting .... 2We use the short-hand DBM to refer to DBM-3, which is equivalent to DBM-2 if D ...

Breaking Out of Local Optima with Count ... - Semantic Scholar
Our complete sys- tem achieves 48.6% accuracy (directed depen- dency macro-average over all 19 languages in the 2006/7 CoNLL data) — more than 5%.

Towards High-performance Pattern Matching on ... - Semantic Scholar
such as traffic classification, application identification and intrusion prevention. In this paper, we ..... OCTEON Software Developer Kit (Cavium SDK version 1.5):.

Giving Eyesight to the Blind: Towards attention ... - Semantic Scholar
We argue that next-generation AIED systems should have the ability to monitor and ... figure out a way to directly implant the knowledge into my brain – like in the .... In summary, despite earlier calls for the importance of incorporating eye gaze

Towards a 3D digital multimodal curriculum for the ... - Semantic Scholar
Apr 9, 2010 - ACEC2010: DIGITAL DIVERSITY CONFERENCE ... students in the primary and secondary years with an open-ended set of 3D .... [voice over or dialogue], audio [music and sound effects], spatial design (proximity, layout or.

VAMO: Towards a Fully Automated Malware ... - Semantic Scholar
[10] A. K. Jain, M. N. Murty, and P. J. Flynn. Data clustering: a review. ACM Comput. Surv., 31(3):264–323, 1999. [11] J. Jang, D. Brumley, and S. Venkataraman.

VAMO: Towards a Fully Automated Malware ... - Semantic Scholar
Dept. of Computer Science. University of Georgia. Athens .... 11, 15, 18]) on M to partition it in a number of malware clusters, (b) use VAMO to build a reference.

Towards a 3D digital multimodal curriculum for the ... - Semantic Scholar
Apr 9, 2010 - movies, radio, television, DVDs, texting, youtube, Web pages, facebook, ... and 57% of those who use the internet, are media creators, having.

Towards High-performance Pattern Matching on ... - Semantic Scholar
1Department of Automation, Tsinghua University, Beijing, 100084, China. ... of-art 16-MIPS-core network processing platform and evaluated with real-life data ...

Towards Veri ed Systems: The SAFEMOS Project - Semantic Scholar
a platform for the implementation of Occam programs. The HOL (Higher Order Logic) 20, 18, 52] ...... C.A.R. Hoare. Communicating Sequential Processes. Prentice Hall. International Series in Computer Science, 1985. 32. C.A.R. Hoare. Refinement algebra