Chaos in Atomic Force Microscopy

Similar documents
STM and AFM Tutorial. Katie Mitchell January 20, 2010

1 Introduction. 1.1 Historical Perspective

Lecture 4 Scanning Probe Microscopy (SPM)

NONLINEAR TIME SERIES ANALYSIS

7/3/2014. Introduction to Atomic Force Microscope. Introduction to Scanning Force Microscope. Invention of Atomic Force Microscope (AFM)

Scanning Near-Field Optical Microscopy for Measuring Materials Properties at the Nanoscale

Scanning Probe Microscopy

It has long been a goal to achieve higher spatial resolution in optical imaging and

Signal to Noise Instrumental Excel Assignment

Nano-Spectroscopy. Solutions AFM-Raman, TERS, NSOM Chemical imaging at the nanoscale

Fluid structure interaction of a vibrating circular plate in a bounded fluid volume: simulation and experiment

1 The water molecule and hydrogen bonds in water

EXPERIMENTAL INVESTIGATION OF A TWO-DEGREE-OF-FREEDOM VIBRO-IMPACT SYSTEM

Calibration of AFM with virtual standards; robust, versatile and accurate. Richard Koops VSL Dutch Metrology Institute Delft

A METHOD OF PRECISE CALIBRATION FOR PIEZOELECTRICAL ACTUATORS

ATOMIC FORCE MICROSCOPY

BIFURCATION PHENOMENA IN THE 1:1 RESONANT HORN FOR THE FORCED VAN DER POL - DUFFING EQUATION

SEMICONDUCTOR lasers with optical feedback have

Ferromagnetic resonance imaging of Co films using magnetic resonance force microscopy

INTRODUCTION TO SCANNING TUNNELING MICROSCOPY

QUANTITATIVE INFRARED SPECTROSCOPY. Willard et. al. Instrumental Methods of Analysis, 7th edition, Wadsworth Publishing Co., Belmont, CA 1988, Ch 11.

RANDOM VIBRATION AN OVERVIEW by Barry Controls, Hopkinton, MA

High flexibility of DNA on short length scales probed by atomic force microscopy

NEAR FIELD OPTICAL MICROSCOPY AND SPECTROSCOPY WITH STM AND AFM PROBES

No Evidence for a new phase of dense hydrogen above 325 GPa

Experiment 5. Strain Gage Measurements

III. MEMS Projection Helvetica 20 Displays

Tecniche a scansione di sonda per nanoscopia e nanomanipolazione 2: AFM e derivati

The Sonometer The Resonant String and Timbre Change after plucking

Piezoelectric Scanners

Atomic Force Microscope

A. Ricci, E. Giuri. Materials and Microsystems Laboratory

Force Spectroscopy with the Atomic Force Microscope

INTERFERENCE OF SOUND WAVES

ENS 07 Paris, France, 3-4 December 2007

Scanning Near Field Optical Microscopy: Principle, Instrumentation and Applications

Measuring Line Edge Roughness: Fluctuations in Uncertainty

13C NMR Spectroscopy

Surface Analysis with STM and AFM

Experiment 5. Lasers and laser mode structure

Raman Spectroscopy Basics

Figure Test set-up

Lab 1: The Digital Oscilloscope

State Newton's second law of motion for a particle, defining carefully each term used.

Non-Contact Vibration Measurement of Micro-Structures

Ultrasonic Wave Propagation Review

Breeding and predictability in coupled Lorenz models. E. Kalnay, M. Peña, S.-C. Yang and M. Cai

Part IV. Conclusions

Agilent AN 1316 Optimizing Spectrum Analyzer Amplitude Accuracy

Experiment #5: Qualitative Absorption Spectroscopy

High Speed Atomic Force Microscopy of Biomolecules by Image Tracking

Modern Construction Materials Prof. Ravindra Gettu Department of Civil Engineering Indian Institute of Technology, Madras

5. Scanning Near-Field Optical Microscopy 5.1. Resolution of conventional optical microscopy

AIR RESONANCE IN A PLASTIC BOTTLE Darrell Megli, Emeritus Professor of Physics, University of Evansville, Evansville, IN dm37@evansville.

Raman spectroscopy Lecture

Laser Based Micro and Nanoscale Manufacturing and Materials Processing

Chapter 3 Non-parametric Models for Magneto-Rheological Dampers

Microscopie à champs proche: et application

Raman Scattering Theory David W. Hahn Department of Mechanical and Aerospace Engineering University of Florida

Keysight Technologies How to Choose your MAC Lever. Technical Overview

Atomic Force Microscopy Observation and Characterization of a CD Stamper, Lycopodium Spores, and Step-Height Standard Diffraction Grating

The Design and Characteristic Study of a 3-dimensional Piezoelectric Nano-positioner

Linear Parameter Measurement (LPM)

Optimizing IP3 and ACPR Measurements

APPLICATION NOTE ULTRASONIC CERAMIC TRANSDUCERS

Indiana's Academic Standards 2010 ICP Indiana's Academic Standards 2016 ICP. map) that describe the relationship acceleration, velocity and distance.

Comparison of flow regime transitions with interfacial wave transitions

Chapter 7. Lyapunov Exponents. 7.1 Maps

Nano Optics: Overview of Research Activities. Sergey I. Bozhevolnyi SENSE, University of Southern Denmark, Odense, DENMARK

FUNDAMENTALS OF MODERN SPECTRAL ANALYSIS. Matthew T. Hunter, Ph.D.

1) The time for one cycle of a periodic process is called the A) wavelength. B) period. C) frequency. D) amplitude.

A PC-BASED TIME INTERVAL COUNTER WITH 200 PS RESOLUTION

GIANT FREQUENCY SHIFT OF INTRAMOLECULAR VIBRATION BAND IN THE RAMAN SPECTRA OF WATER ON THE SILVER SURFACE. M.E. Kompan

The Fundamentals of Infrared Spectroscopy. Joe Van Gompel, PhD

Preface Light Microscopy X-ray Diffraction Methods

DualBeam Solutions for Electrical Nanoprobing

Lateral Resolution of EDX Analysis with Low Acceleration Voltage SEM

Determination of source parameters from seismic spectra

Active noise control in practice: transformer station

Mesh Moving Techniques for Fluid-Structure Interactions With Large Displacements

MEMS mirror for low cost laser scanners. Ulrich Hofmann

Transducer Mounting and Test Setup Configurations. Rick Bono The Modal Shop

Instability, dispersion management, and pattern formation in the superfluid flow of a BEC in a cylindrical waveguide

5. Measurement of a magnetic field

Numerical and experimental exploration of phase control of chaos

Atomic Force Microscopy ISC-CNR

COMPUTATIONAL TOOLS FOR SIGNAL PROCESSING APPLICATIONS FROM INDUSTRY AND BIOINFORMATICS 02/2006

Application Note Noise Frequently Asked Questions

Physical Science Study Guide Unit 7 Wave properties and behaviors, electromagnetic spectrum, Doppler Effect

BNAM 2008 BERGEN BYBANE NOISE REDUCTION BY TRACK DESIGN. Reykjavik, august Arild Brekke, Brekke & Strand akustikk as, Norway

The Volumetric Erosion of Electrical Contacts

The excitation in Raman spectroscopy is usually. Practical Group Theory and Raman Spectroscopy, Part II: Application of Polarization

(Nano)materials characterization

Alternative Linear Motion Systems. Iron Core Linear Motors

Objectives. Experimentally determine the yield strength, tensile strength, and modules of elasticity and ductility of given materials.

E190Q Lecture 5 Autonomous Robot Navigation

International Year of Light 2015 Tech-Talks BREGENZ: Mehmet Arik Well-Being in Office Applications Light Measurement & Quality Parameters

State Newton's second law of motion for a particle, defining carefully each term used.

Calibration and Use of a Strain-Gage-Instrumented Beam: Density Determination and Weight-Flow-Rate Measurement

Transcription:

Purdue University Purdue e-pubs Birck and NCN Publications Birck Nanotechnology Center January 26 Chaos in Atomic Force Microscopy Shuiqing Hu Birck Nanotechnology Center and School of Mechanical Engineering, Purdue University, shuiqing@purdue.edu Arvind Raman Birck Nanotechnology Center and School of Mechanical Engineering, Purdue University, raman@purdue.edu Follow this and additional works at: http://docs.lib.purdue.edu/nanopub Hu, Shuiqing and Raman, Arvind, "Chaos in Atomic Force Microscopy" (26). Birck and NCN Publications. Paper 39. http://docs.lib.purdue.edu/nanopub/39 This document has been made available through Purdue e-pubs, a service of the Purdue University Libraries. Please contact epubs@purdue.edu for additional information.

PRL 96, 3617 (26) 27 JANUARY 26 Chaos in Atomic Force Microscopy Shuiqing Hu and Arvind Raman Birck Nanotechnology Center and School of Mechanical Engineering, Purdue University, West Lafayette, Indiana 4797, USA (Received 3 June 25; published 27 January 26) Chaotic oscillations of microcantilever tips in dynamic atomic force microscopy (AFM) are reported and characterized. Systematic experiments performed using a variety of microcantilevers under a wide range of operating conditions indicate that softer AFM microcantilevers bifurcate from periodic to chaotic oscillations near the transition from the noncontact to the tapping regimes. Careful Lyapunov exponent and noise titration calculations of the tip oscillation data confirm their chaotic nature. AFM images taken by scanning the chaotically oscillating tips over the show small, but significant metrology errors at the nanoscale due to this deterministic uncertainty. DOI: 1.113/PhysRevLett.96.3617 Deterministic chaos underpins the dynamics of many nonlinear systems that display a very high degree of sensitivity to initial conditions. In what follows we observe and characterize chaos in the oscillations of nanoscale tips in modulated atomic force microscopy (AM- AFM). Specifically, we find that chaotic tip dynamics often occur upon the transition from the noncontact to the tapping regimes indicating the presence of complex grazing dynamics. Chaotic dynamics are also observed under conditions of hard tapping. Furthermore, we show that the presence of chaos can lead to errors while imaging s using dynamic AFM, thereby introducing an element of deterministic uncertainty in nanometrology. In dynamic AFM [1], a micromechanical oscillator, a silicon microcantilever with an attached nanoscale tip, is driven near the oscillator resonance frequency in the close vicinity of the surface. During each oscillation cycle the nanoscale tip can experience a variety of short and long range forces between the tip and the such as [2] electrostatic, van der Waals, capillary, specific chemical, Pauli repulsion, and nanoscale contact and elastic forces. These interaction forces are highly nonlinear, and consequently it is possible that the oscillating tip exhibits chaotic spectra that contain a wide range of constituent frequencies. Because tip oscillations at frequencies other than the driving frequency are not controlled in AM- AFM [3,4], chaotic tip oscillation spectra can lead to fundamental and important errors in both the metrology and the force spectroscopy of the scanned. Prior to the present work, both Burnham et al. [5] and Salapaka et al. [6] have reported complicated microcantilever dynamics in experiments where vibrating s were made to impact a stationary microcantilever. The present work deals more directly with nanoscale imaging and spectroscopy in dynamic AFM where the microcantilever is driven near its fundamental resonance. The experimental setup consists of a Picoplus TM AFM system, together with an external Signal Recovery TM lockin amplifier, and a data acquisition system based on National Instruments 5911 boards. To minimize the effects of water meniscus, the experiments are performed under dry nitrogen environment on the hydrophobic surface of PACS numbers: 68.35.Ja, 5.45.Tp, 7.79.Lh, 68.37.Ps freshly cleaved highly ordered pyrolytic graphite (HOPG) supplied by SPI Supplies TM. For comparison, the same sets of experiments are also performed under ambient conditions. The experiments are performed with three different sets of microcantilevers, each set consisting of approximately ten microcantilevers with properties that are shown in Table I. This represents a wide range of microcantilever bending stiffnesses and fundamental resonance frequencies (! ). The goal of the experimental procedure is to demonstrate the existence of chaos and its effects on nanometrology within the context of AM-AFM. Accordingly, for each microcantilever, the tip at the driving frequency (setpoint ) is maintained constant via the lock-in amplifier and the AFM control system. The driving of the dither piezo is gradually increased in increments of.49 nm by increasing the voltage supplied to the dither piezo; this effectively decreases the setpoint ratio. At each driving voltage, both the tip oscillation and the piezo excitation signal are recorded. These experiments are performed for two setpoint s [7], namely, 8 and 14 nm. Moreover, these experiments are performed for cases when the driving frequency :98!,!, and 1:2!. Finally, each experiment was repeated 6 times to verify the repeatability of the results. The observed phenomena are robust over the six repeats of each experiment, and also consistent between the approximately ten microcantilevers tested within each set. Put together, this represents a substantial body of experimental data [8] representing dynamic AFM under a wide range of operating conditions and using a wide variety of microcantilevers. A summary of all the experimental results may be found in Table I; however, in this Letter we focus our attention on experiments performed under dry nitrogen using the set of Nanosensor TM Force Modulation silicon microcantilevers (! 66:3 khz) when! and with a setpoint of 8 nm. The phenomena observed in this set of microcantilevers under these operating conditions span the phenomena observed for all other sets of microcantilevers. In what follows we analyze the time series tip oscillation 31-97=6=96(3)=3617(4)$23. 3617-1 26 The American Physical Society

PRL 96, 3617 (26) 27 JANUARY 26 TABLE I. Properties and observed transitions of oscillation states for 3 sets of microcantilevers (the stiffness, resonance frequency, and quality factor are ascertained from experimental data and are averaged over the 1 microcantilevers within each set) as the driving is gradually increased. The onset of chaos is determined by calculating the noise limit and Lyapunov exponents of the tip oscillation data. The symbol ^! indicates a simultaneous transition from the noncontact to the tapping regime. Cantilever set Force modulation Ultrasharp NSC12-C TESP Fundamental resonance 63:3 2:1 13:2 1:8 31:1 5:9 frequency (khz) Stiffness (N=m) 1:5 :1 2:9 :2 148 5 Quality factor 142 8 13 11 148 5 Transitions :98!, Period-1 ^! period-2! period-4 Period-1 ^! chaotic Period-1 ^! period-1 under!, Period-1 ^! chaotic! period-2! chaotic Period-1 ^! chaotic Period-1 ^! period-1 nitrogen 8 nm 1:2!, Period-1 ^! chaotic Period-1 ^! chaotic Period-1 ^! period-1 Transitions in air 14 nm 8 nm 14 nm :98!, Period-1 ^! period-2! period-4 Period-1 ^! period-1 Period-1 ^! period-1!, Period-1 ^! chaotic! period-1! period-2 Period-1 ^! period-1 Period-1 ^! period-1 1:2!, Period-1 ^! chaotic Period-1 ^! period-1 Period-1 ^! period-1 :98!, Period-1 ^! period-2! period-4 Period-1 ^! chaotic Period-1 ^! period-1!, Period-1 ^! chaotic! period-1! chaotic Period-1 ^! chaotic Period-1 ^! period-1 1:2!, Period-1 ^! chaotic Period-1 ^! chaotic Period-1 ^! period-1 :98!, Period-1 ^! period-2! period-4 Period-1 ^! period-1 Period-1 ^! period-1!, Period-1 ^! period-2! period-1! chaotic Period-1 ^! period-1 Period-1 ^! period-1 1:2!, Period-1 ^! chaotic Period-1 ^! period-1 Period-1 ^! period-1 data acquired for this set of microcantilevers under the above conditions at each driving voltage. In order to detect rigorously the presence of chaos in the tip oscillation time series data at each driving, we use two different time series analysis methods calculation of the largest Lyapunov exponent, and calculation of the noise limit of titration of chaos. The Lyapunov exponents i, i 1;... ; N are the averaged rates of divergence (or convergence) of two neighboring trajectories in phase space. The number of Lyapunov exponents N equals the dimension of the phase space, and a positive Lyapunov exponent is indicative of chaos. We use the lyp_k routine of TISEAN 2.1 a nonlinear time series analysis package [9] to estimate the largest Lyapunov exponent of the tip oscillation data at each driving. Because the Lyapunov exponent may be an erroneous indicator of chaos in the presence of noise [1], we also use titration of chaos with added noise [11], a relatively new method that can robustly detect chaos with short noisy experimental data. Specifically this method detects the nonlinearity in a time series by comparing short time prediction results of nonlinear and linear Volterra- Wiener-Korenberg series. White noise of increasing standard deviation is added to the chaotic time series until its nonlinearity cannot be detected by a particular numerical indicator. We denote this value of the noise limit (NL) as the noise ceiling of the signals corresponding to the amount of chaos. NL > indicates chaos; NL indicates that the data series either is not chaotic or the chaotic component is already neutralized by the background noise in the original time series. The results for the Lyapunov exponents and the NL (calculated using a previously published algorithm [12]) are shown in Fig. 1. Clearly both the titration of the chaos method and the Lyapunov exponent method yield the same results and confirm that chaotic oscillations set in at the transition from the noncontact to the tapping regime of oscillation (at 37.2% setpoint ratio) and then again under hard tapping conditions. The presence of the noncontact to tapping transition is confirmed by measuring the phase of the oscillation [13]. Following the transition from noncontact to tapping, the tip continues to oscillate chaotically for driving s between 1.23 and 2.7 nm, then suddenly transitions to a period-2 oscillation and becomes chaotic again when the driving is larger than 4.66 nm. We now explore the consequences of imaging the freshly cleaved atomically flat HOPG surface in different regimes [14]. For the driving.49 nm the tip is in a period-1 [15] oscillation state, in the noncontact regime. The tip vibration power spectral density shows superharmonics, but the tip motion is still period-1 as shown in the Noncontact Lyapunov Exponent.15.1.5 Tapping 3 Lyapunov Exponent Noise Limit 2 4 6 8 1 12 14 16 Drive Amplitude (nm) Period1 Chaotic Period2 Chaotic FIG. 1 (color online). Largest Lyapunov exponent and noise limit versus dither piezo driving for one of the Force Modulation (! 66:3 khz) microcantilevers in a dry nitrogen environment over freshly cleaved HOPG. 2 1 Noise Limit (%) 3617-2

PRL 96, 3617 (26) 27 JANUARY 26 Velocity (nm/s) 5-5 -1 1 5-5 -1 1 5 Amplitude=8nm Amplitude=8nm Amplitude=8nm -5-1 1 Displacement (nm) 2 4 6 phase portrait plot in Fig. 2. Figure 3 shows the topology map of the freshly cleaved HOPG surface and the map of the error signal (the difference between the setpoint and the actual of the tip at the driving frequency) acquired under these conditions. Imaging the in this regime produces accurate topology images of freshly cleaved HOPG, and the error is nearly zero everywhere except at the edges of the HOPG layers. Figure 2 shows tip deflection phase portrait and power spectral density at the drive of 1.97 nm when the tip is in its first chaotic oscillation state at the Power Spectral Density (db/hz) 6 4 2 6 4 2 6 4 2 2 Side Bands 1 1 1 1 2B 2 6 2 6 2 2 4 6 Side Bands 6 7 2 4 6 Frequency (khz) FIG. 2 (color online). Tip deflection phase portrait (left) and power spectral density (right) versus dither piezo drive of.49 nm (period-1, noncontact regime), 1.97 nm (chaotic, tapping regime), and 14.76 nm (chaotic, tapping regime). The vertical line in the phase portrait represents the position of the surface. denotes the driving frequency, and 2B denotes the second bending mode frequency of the microcantilever. In the phase portrait, the tip velocity is computed by differentiating the experimental tip vibration signal using moving averages. FIG. 3 (color online). Influence of chaotic tip oscillations on the nanometrology of a freshly cleaved HOPG surface. Topology map (top) and error map (bottom) with driving s of.49 nm (period-1, noncontact regime), 1.97 nm (chaotic, tapping regime), and 14.76 nm (chaotic, tapping regime). The scan area is 2 2 m 2. transition from the noncontact to the tapping regime. The spectrum now contains superharmonics, subharmonics, and side bands, and the noise floor increases as seen in the power spectrum. A significant sixth harmonic of the driving frequency also appears in the tapping regime near the second bending mode resonance frequency [denoted by 2B in Fig. 2]. This is attributed to a nonlinear coupling between two different modes of the continuous microcantilever beam the fundamental mode and the second bending mode whose resonance frequency is close to 6 times that of the fundamental. This condition is also referred to as a 1:6 internal resonance [16]. Figure 3 shows the topology map and the error map at a driving of 1.97 nm. Clearly the error map is noisy even on the atomically flat part of the HOPG due to the chaotic nature of the tip oscillation. Finally, when the driving is at 14.76 nm and the tip is in the second chaotic oscillation state in the hard tapping regime, the phase portrait indicates chaotic dynamics and the spectrum becomes broadband as shown in Fig. 2. Both the topology and the error map in Fig. 3 are extremely noisy. We conclude that the presence of chaotic tip oscillations deteriorates noticeably the ability of AM-AFM to perform high precision nanometrology. The noise titration and Lyapunov exponent calculations were performed for each experiment for every cantilever under different operating conditions. In all cases both calculations yield the same results. The results of observed dynamics are summarized in Table I. Except for very stiff and high frequency microcantilevers [such as the set of tapping mode etched silicon probe (TESP) microcantilevers,! 31 khz], all other microcantilevers tested undergo chaotic oscillations upon transition from the noncontact to the tapping regimes for certain operating parameters. These experimental results provide compelling evidence for the underlying mechanism of chaos in this nanoscale dynamical system. Specifically the observed sequence of bifurcations in Table I is predicted by the global dynamics of the two-dimensional Nordmark map, which is often used to model single degree-of-freedom nonsmooth systems. For instance, two bifurcations of grazing trajectories are commonly observed [17,18] in the Nordmark map. One is from a stable period-1 orbit to a chaotic attractor at the gazing incidence; another is when a stable period-1 orbit connects to a stable period-m maximal orbit through an unstable period-m orbit via a saddle node bifurcation. These two bifurcations are very similar to the bifurcations period-1! chaotic and period-1! period-2! period-1 observed in our experiments at the transition from the noncontact to the tapping regime (as shown in Table I). In spite of their similarities, several aspects of the experimental results cannot be fully captured by the predictions of the Nordmark map. For instance, the experimental results clearly indicate a 1:6 internal resonance [16] between the first and second bending modes. We anticipate 3617-3

PRL 96, 3617 (26) 27 JANUARY 26 therefore that the full range of our experimental results is likely to be adequately captured in a higher dimensional equivalent of the two-dimensional Nordmark map. Finally, we note that another mechanism of chaos, namely, homoclinic chaos has been investigated in theoretical AFM dynamics models [19,2]. However, this requires the cantilever to be placed very close to the surface (<5 nm). We have not investigated in this regime because this is a relatively rare situation in AM-AFM where the cantilever is kept at least 1 nm away from the, typically 25 15 nm. It is interesting to speculate on the fundamental physical mechanisms in our experiments that could lead to effectively nonsmooth interaction forces. While the formation and breaking of nanoscale liquid necks between the tip and the could lead to such effects, the experimental conditions minimize this effect. Instead, when the tip comes very close to the, it is known from molecular dynamics simulations [21] that instabilities occur between the atoms at the very end of the tip and the atoms at the surface. These instabilities occur at time scales that are extremely small compared to the time scale of tip oscillations and could lead to effectively nonsmooth tip interaction forces. Several key conclusions can be drawn from the above results. First, the frequent onset of chaotic dynamics at the transition from the noncontact to the tapping regimes indicates the presence of complex grazing bifurcations such as those predicted in the Nordmark map. This is an important observation because the transition from the noncontact to the tapping regime often occurs under realistic operating conditions used for AFM imaging. Second, stiff cantilevers seem immune to chaotic dynamics under the operating conditions tested in this work; however, they also generate larger tip- interaction forces. Finally, in the chaotic spectra of the microcantilevers (such as for the 66.3 khz microcantilever discussed at length) more than 8% of the vibration energy is still concentrated at the driving frequency. The observed chaos is weak in this sense because it contains a strong periodic component. In conclusion, we have observed and characterized thoroughly and repeatably the onset of weak chaos in microcantilevers with nanoscale tips tapping on HOPG under a variety of operating conditions. We have also demonstrated that the presence of chaos deteriorates the image quality and introduces deterministic uncertainty in nanometrology. These metrology errors are expected to be small for most current applications; however, as nanometrology requirements become increasingly stringent, the presence of chaos is likely to pose some fundamental deterministic limits on measurement resolution. Financial support of this research by the National Science Foundation (NSF) under Contracts No. 116414- CMS and No. 4966-CMS is gratefully acknowledged. We also thank Professor R. Reifenberger (Condensed Matter Physics, Purdue University) for helpful discussions. [1] R. García and R. Pérez, Surf. Sci. Rep. 47, 197 (22). [2] J. N. Israelachvili, Intermolecular and Surface Forces (Academic Press, San Diego, 1992), 2nd ed. [3] M. Stark et al., Proc. Natl. Acad. Sci. U.S.A. 99, 8473 (22). [4] R. W. Stark, G. Schitter, M. Stark, R. Guckenberger, and A. Stemmer, Phys. Rev. B 69, 85412 (24). [5] N. A. Burnham, A. J. Kulik, G. Gremaud, and G. A. D. Briggs, Phys. Rev. Lett. 74, 592 (1995). [6] S. Salapaka, M. Dahleh, and I. Mezić, Nonlinear Dyn. 24, 333 (21). [7] We use the Z piezotube controller to maintain the setpoint in the presence of thermal drifts. However, during the data acquisition this controller is turned off. The Z piezotube controller therefore effectively resets the gap to the approximate setpoint preceding each increment in driving. Note that since only the setpoint is maintained constant, other frequency components are uncontrolled, and the gap between the cantilever and will still vary a little at different driving s. [8] The lock-in amplifier triggers the 5911 boards to initiate the data acquisition. The lock-in amplifier and AFM controller are fully controlled by VISUAL BASIC script programs. In this manner, the entire experiment sequence with a specific microcantilever is performed automatically with one program. [9] H. Kantz and T. Schreiber, Nonlinear Time Series Analysis (Cambridge University Press, Cambridge, 1997). [1] L. Arnold, P. Imkeller, and N. S. Namachchivaya, J. Sound Vib. 269, 13 (24). [11] M. Barahona and C. Poon, Nature (London) 381, 215 (1996). [12] C. Poon and M. Barahona, Proc. Natl. Acad. Sci. U.S.A. 98, 717 (21). [13] R. García and A. San Paulo, Phys. Rev. B 61, R13 381 (2). [14] For a fair comparison, the imaging feedback control parameters for these three cases are identical, so that any differences in imaging quality are entirely due to the tip oscillation state. Also optimizing control parameters does not further improve the imaging quality. [15] A period-n motion is a periodic motion with a fundamental period that is 1=n of the excitation signal. [16] A. H. Nayfeh, Nonlinear Interactions: Analytical, Computational, and Experimental Methods (Wiley-Interscience, New York, 2). [17] W. Chin, E. Ott, H. E. Nusse, and C. Grebogi, Phys. Rev. E 5, 4427 (1994). [18] F. Casas, W. Chin, C. Grebogi, and E. Ott, Phys. Rev. E 53, 134 (1996). [19] M. Basso, L. Giarré, M. Dahleh, and I. Mezić, Trans. ASME: J. Dyn. Syst. Meas. Control 122, 24 (2). [2] A. Ashhab, M. Salapaka, M. Dahleh, and I. Mezić, Nonlinear Dynamics 2, 197 (1999). [21] U. Landman and W. D. Luedtke, J. Vac. Sci. Technol. B 9, 414 (1991). 3617-4