Force-distance studies with piezoelectric tuning forks below 4.2K

Similar documents
Program Operacyjny Kapitał Ludzki SCANNING PROBE TECHNIQUES - INTRODUCTION

Lecture 4 Scanning Probe Microscopy (SPM)

Point mass approximation. Rigid beam mechanics. spring constant k N effective mass m e. Simple Harmonic Motion.. m e z = - k N z

Noncontact lateral-force gradient measurement on Si 111-7Ã 7 surface with small-amplitude off-resonance atomic force microscopy

Dopant Concentration Measurements by Scanning Force Microscopy

Measurement of hardness, surface potential, and charge distribution with dynamic contact mode electrostatic force microscope

Force Measurement with a Piezoelectric Cantilever in a Scanning Force Microscope

CNPEM Laboratório de Ciência de Superfícies

The velocity dependence of frictional forces in point-contact friction

Basic Laboratory. Materials Science and Engineering. Atomic Force Microscopy (AFM)

Application of electrostatic force microscopy in nanosystem diagnostics

Scanning Tunneling Microscopy

Scanning Force Microscopy II

Tunnelling between edge channels in the quantum hall regime manipulated with a scanning force microscope

Enhancing capabilities of Atomic Force Microscopy by tip motion harmonics analysis

BDS2016 Tutorials: Local Dielectric Spectroscopy by Scanning Probes

Scanning Probe Microscopy (SPM)

Scanning Tunneling Microscopy

Imaging Methods: Scanning Force Microscopy (SFM / AFM)

Scanning Probe Microscopy. Amanda MacMillan, Emmy Gebremichael, & John Shamblin Chem 243: Instrumental Analysis Dr. Robert Corn March 10, 2010

Noncontact-AFM (nc-afm)

Distance dependence of force and dissipation in non-contact atomic force microscopy on Cu(100) and Al(111)

Integrating MEMS Electro-Static Driven Micro-Probe and Laser Doppler Vibrometer for Non-Contact Vibration Mode SPM System Design

Electronic Quantum Transport in Mesoscopic Semiconductor Structures

Scanning capacitance spectroscopy of an Al x Ga 1Àx NÕGaN heterostructure field-effect transistor structure: Analysis of probe tip effects

Atomic and molecular interactions. Scanning probe microscopy.

Functional Microcantilever for a Novel Scanning Force Microscope

Construction of a dilution refrigerator cooled scanning force microscope

Scanning Force Microscopy

SUPPLEMENTARY INFORMATION

NIS: what can it be used for?

Module 26: Atomic Force Microscopy. Lecture 40: Atomic Force Microscopy 3: Additional Modes of AFM

Magnetic Resonance Force Microscopy. Christian Degen Department of Physics, ETH Zurich, Switzerland

Instrumentation and Operation

Molecular and carbon based electronic systems

Introduction to Scanning Probe Microscopy Zhe Fei

Magnetic Force Microscopy (MFM) F = µ o (m )H

General concept and defining characteristics of AFM. Dina Kudasheva Advisor: Prof. Mary K. Cowman

Forces and frequency shifts in atomic-resolution dynamic-force microscopy

Università degli Studi di Bari "Aldo Moro"

Nitride HFETs applications: Conductance DLTS

STM: Scanning Tunneling Microscope

INTRODUCTION TO SCA\ \I\G TUNNELING MICROSCOPY

Atomic Force Microscopy imaging and beyond

Conservative and dissipative tip-sample interaction forces probed with dynamic AFM

Chapter 2 Experimental Technique and Working Modes

Application of single crystalline tungsten for fabrication of high resolution STM probes with controlled structure 1

Searching atomic spin contrast on nickel oxide (001) by force microscopy

SCANNING-PROBE TECHNIQUES OR APPARATUS; APPLICATIONS OF SCANNING-PROBE TECHNIQUES, e.g. SCANNING PROBE MICROSCOPY [SPM]

High resolution STM imaging with oriented single crystalline tips

Contents. What is AFM? History Basic principles and devices Operating modes Application areas Advantages and disadvantages

IN 1986, Binnig, Quate, and Gerber invented the atomic force

A NEW APPROACH TOWARDS PROPERTY NANOMEASUREMENTS USING IN-SITU TEM

Enrico Gnecco Department of Physics. University of Basel, Switzerland

Scanning Tunneling Microscopy

Robust high-resolution imaging and quantitative force measurement with tuned-oscillator atomic force microscopy

R. Akram a Faculty of Engineering Sciences, Ghulam Ishaq Khan Institute of Engineering Sciences and Technology, Topi, Pakistan

Lecture 26 MNS 102: Techniques for Materials and Nano Sciences

Stability of an oscillating tip in Non-Contact Atomic Force Microscopy: theoretical and numerical investigations.

Advances in atomic force microscopy

SPM 150 Aarhus with KolibriSensor TM

Amplitude dependence of image quality in atomically-resolved bimodal atomic microscopy

Optimization of Force Sensitivity in Q-Controlled Amplitude-Modulation Atomic Force Microscopy

(Scanning Probe Microscopy)

Lorentz Contact Resonance for viscoelastic measurements of polymer blends

MSE640: Advances in Investigation of Intermolecular & Surface Forces

Heat transfer in ultrahigh vacuum scanning thermal microscopy

Scanning Tunneling Microscopy

SUPPLEMENTARY INFORMATION

Scanning Probe Microscopy (SPM)

Intermittent-Contact Mode Force Microscopy & Electrostatic Force Microscopy (EFM)

Magnetic dissipation force microscopy studies of magnetic materials invited

Small-Amplitude Atomic Force Microscopy**

Characterization of MEMS Devices

Introduction to Scanning Probe Microscopy

Scanning Probe Microscopy. EMSE-515 F. Ernst

Understanding the properties and behavior of groups of interacting atoms more than simple molecules

NIST ELECTROSTATIC FORCE BALANCE EXPERIMENT

NOISE ANALYSIS IN PARAMETRIC RESONANCE BASED MASS SENSING. Wenhua Zhang and Kimberly L. Turner

Lorentz Contact Resonance for viscoelastic measurements of polymer blends

Towards nano-mri in mesoscopic transport systems

Lecture 12: Biomaterials Characterization in Aqueous Environments

Studies of the Properties of Designed Nanoparticles Using Atomic Force Microscopy

Scanning Gate Microscopy (SGM) of semiconductor nanostructures

Principle of Electrostatic Force Microscopy and Applications. Thierry Mélin.

Atomic force microscopy study of polypropylene surfaces treated by UV and ozone exposure: modification of morphology and adhesion force

Advances in AFM: Seeing Atoms in Ambient Conditions

tip of a current tip and the sample. Components: 3. Coarse sample-to-tip isolation system, and

Principle of Magnetic Force Microscopy

AFM Imaging In Liquids. W. Travis Johnson PhD Agilent Technologies Nanomeasurements Division

Introduction to Nanomechanics: Magnetic resonance imaging with nanomechanics

Practical aspects of Kelvin probe force microscopy

Role of attractive forces in tapping tip force microscopy

Magnetic imaging and dissipation force microscopy of vortices on superconducting Nb films

developed piezoelectric self-excitation and selfdetection mechanism in PZT microcantilevers for dynamic scanning force microscopy in liquid

Mapping Elastic Properties of Heterogeneous Materials in Liquid with Angstrom-Scale Resolution

Theory of magnetoelastic dissipation due to domain wall width oscillation

Theoretical basis of parametric-resonance-based atomic force microscopy

State Feedback Control for Adjusting the Dynamic Behavior of a

Supporting Information

Transcription:

submitted to APPLIED SURFACE SCIENCE nc-afm 99, Pontresina Force-distance studies with piezoelectric tuning forks below 4.2K J. Rychen, T. Ihn, P. Studerus, A. Herrmann, K. Ensslin Solid State Physics Laboratory, ETH Zurich, CH-893 Zurich, Switzerland H. J. Hug, P.J.A. van Schendel, H.J. Güntherodt Institute of Physics, University of Basel, CH-456 Basel, Switzerland (Last modified 9 September 1999 9:33 am) Abstract Piezoelectric quartz tuning forks have been employed as the force sensor in a dynamic mode scanning force microscope operating at temperatures down to 1.7 K at He-gas pressures of typically 5 mbar. An electrochemically etched tungsten tip glued to one of the tuning fork prongs acts as the local force sensor. Its oscillation amplitude can be tuned between a few angstroms and tens of nanometers. Quality factors of up to 12 allow a very accurate measurement of small frequency shifts. Three calibration procedures are compared which allow the determination of the proportionality constant between frequency shift and local force gradient based on the harmonic oscillator model and on electrostatic forces. The calibrated sensor is then used for a study of the interaction between the tip and a HOPG substrate. Force gradient and dissipated power can be recorded simultaneously. It is found that during approaching the tip to the sample considerable power starts to be dissipated although the force gradient is still negative, i.e. the tip is still in the attractive regime. This observation concurs with experiments with true atomic resolution which seem to require the same tip-sample separation. PACS: Keywords: Atomic force microscopy; Frequency modulation force microscopy; Dynamic force microscopy; Tip-sample interaction; Tuning forks; Kelvin force microscopy 1. Introduction The interaction of the tip with the surface of a sample is of crucial importance for the operation of scanning force microscopes [1]. A renewed interest in this subject has arisen since atomic resolution was achieved in the dynamic mode on the surfaces of insulators measured in UHV which were previously not accessible for local investigations on the atomic scale [2]. Recently, atomic resolution was achieved in the dynamic mode on highly oriented pyrolytic graphite (HOPG), a so-called van der Waals surface, at temperatures of 1 K [3]. The low temperatures were necessary in order to reduce the thermal noise and provide stable imaging conditions necessary for atomic resolution. Piezoelectric tuning forks have been employed as force sensors in the dynamic mode by many authors [4,5]. In this paper we report force distance studies on Created using Adobe FrameMaker 1/5 Power Macintosh Release 5.5.3

HOPG substrates at temperatures down to 1.7 K using a tuning fork sensor. Three methods for calibrating this sensor are compared in section 3. In section 4 we discuss simultaneous measurements of the force gradient and of the power dissipation vs. tip-sample separation. 2. Setup 2.1 The Microscope setup We use a commercial cryo-sxm from Oxford Instruments which has been modified for working as a scanning force microscope utilizing a piezoelectric quartz tuning fork sensor for operation in the dynamic mode. The microscope is run in the He-gas atmosphere of the variable temperature insert of a cryostat at temperatures below 4.2 K and pressures of typically 5 millibar. The microscope is designed for the local investigation of semiconductor nanostructures in high magnetic fields, which requires a large scan range at low temperatures at the cost of lateral spatial resolution. Hence, atomic resolution cannot be achieved. Further details about the microscope can be found in Ref. 5. 2.2 Tuning fork and amplitude calibration The tuning fork sensor is prepared by attaching a tungsten wire (diameter 13µm, lenght 3µm) at the end of one tuning fork prong and subsequent electrochemical etching of the wire to form a sharp tip with a tip radius of typically 1 nm as determined from transmission electron microscope images. The tip is electrically isolated from the tuning fork contacts and separately connected to ground via a current preamplifier. On one hand this allows the measurement of the current, I T, induced on the tip at constant tip-sample bias by the oscillating tip-sample capacitance. This capacitance oscillation is caused by the oscillation of the tip itself. On the other hand we can directly measure the tip-sample capacitance, Cz ( ), by applying an AC-voltage between the tip and the sample and ramping the tip-sample separation z. The tuning fork is part of a phase-locked loop (PLL) which drives it always at its resonance frequency [5]. The PLL delivers the frequency shift of the tuning fork sensor and the dissipated power caused by tipsample interactions. In order to determine the mechanical and piezoelectric characteristics of the tuning fork sensor experimentally we have measured its mechanical vibration amplitude and its admittance as a function of frequency [6]. Applying a harmonic oscillator model we determined the effective mass of the oscillator m =.33mg, the mechanical stiffness k = 15.4 nn/pm and the piezoelectric coupling constant α = 4.2 µc/m for room temperature operation [6]. At cryogenic temperatures these characteristic values will in general be different. However, we can argue that the effective mass of the oscillator should experience little change since the kinetic energy stored in the fork for a given tip speed at zero tip deflection will be barely changed. Using the above value for the effective mass we are then able to determine k and α from the measurement of the admittance of the tuning fork at low temperatures alone, without the need of a simultaneous mechanical calibration of the amplitude. We will refer to this indirectly deduced low-temperature amplitude as the nominal amplitude of the tip. In our experiments below 4.5 K nominal tip oscillation amplitudes between.56 nm and 2 nm have been used. The quality factors of the oscillators ranged between 1 and 12 depending on temperature and pressure [6]. 3. Force calibration In order to translate measured frequency shifts f into a quantity describing the tip-sample interaction we have calibrated our sensor. We use the force gradient approximation as the basic assumption, i.e. f η F z = --------, (1) z where ( F z ) ( z) is the gradient of the tip-sample interaction force and η is the proportionality constant which is to be determined by the calibration procedure. This approximation is valid for small tip-oscillation amplitudes. It is the limiting case of more general treatments discussed in the literature [7]. The calibration constant η is independent of the details of the tip shape and depends only on the mechanical properties of the tuning fork. In the following we will compare Created using Adobe FrameMaker 2/5 Power Macintosh Release 5.5.3

three different methods of calibration and report the determined values of the calibration constant η. 3.1 Harmonic oscillator model If we apply the harmonic oscillator model mentioned above for the resonance of our tuning fork, we can calibrate the sensor simply by determining the spring constant k from the electrical resonance curve. We obtain the relation f ----- f ( F) ( z) = ------------------------- 4k which differs by a factor of 1/2 from the result obtained for conventional cantilevers. This factor reflects the fact that only one prong of the fork senses the interaction but both prongs are oscillating. This method gives η = f ( 4k). For the sensor used for the studies described here we measured the resonance frequency f = 32596.739 Hz and the stiffness k = 14.6 nn/pm and therefore obtained a value of η =.6 Hz/(N/m). Since the resonance frequency of our sensors can be determined with an accurcacy better than 1 mhz the main source of error in this method may be found in the determination of k. It is calculated from k = 4π 2 f 2 m, thus any error in the assumption of an unaltered m at low temperatures will cause an error in η. 3.2 Calibration with the electrostatic force The two other calibration methods make use of the attractive electrostatic force acting between the tip and the sample if a constant tip-sample voltage U is applied. It is given by the expression Fz ( ) = 1 2 ( C) ( z) ( U µ ( ch e) ) 2, where C(z) is the tip-sample capacitance and µ ch is the difference of the chemical potentials of the two materials which is usually taken to be equal to the work function difference between the two materials. Using eq. 1 we obtain for the frequency shift η 2 C f = -- -------- ( U µ ( (2) 2 z 2 ch e) ) 2 The measurement of f( U) shown in Fig. 1 was performed at a large tip-sample separation and with a nominal tip-oscillation amplitude of 2 nm. For this tip-sample separation the electrostatic force is by far the dominant interaction. The data can be fitted extremely well by a parabola and therefore allows us to determine µ ch = 64 mev. From the measurement in Fig. 1 the calibration constant η could be determined using eq. 2 if the second derivative of the capacitance were known at the same distance. For the determination of this quantity we have used two distinct techniques which will be described in the following. f (mhz) 5 1 15 Figure 1: f(u) measured at 4.5 K (left axis) and Kelvin current (right axis). 3.2.1 Direct measurement of the capacitance We have measured the capacitance Cz ( ) depicted in Fig. 2 by applying an AC tip-sample voltage of 1mV(rms) at a frequency of 95 khz. The measurement was fitted with a smooth analytical curve in order to allow a better determination of the second derivative. Inserting the result in eq. 2 we obtain η 1 Hz/(N/m). The accuracy of this method is mainly determined by the limited accuracy of the determination of the very small tip-sample capacitance which results in an uncertaintiy of the η-value of more than a factor of two for the measurement shown. 3.2.2 Kelvin method 5 5 U (V) 1 15 The second approach for determining the second derivative of the capacitance employs the Kelvin 15 1 5 5 I t (parms @ f res ) Created using Adobe FrameMaker 3/5 Power Macintosh Release 5.5.3

method in which the gradient of the tip-sample capacitance is directly measured by oscillating the tip over the sample surface at a constant bias voltage and recording the current induced on the tip by the motion. This current is given by I T = 2πif A C ------ ( U ( µ. z ch ) e) Fig. 1 shows an example of an I( U) curve measured at the resonance frequency and a nominal tip-oscillation amplitude A = 2 nm for a large tip-sample distance. A linear fit to the curve allows the determination of ( C) ( z). The second derivative can be obtained by repeating this measurement as a function of z. Inserting the result into eq. 2 leads us to the calibration η =.8 Hz/(N/m). For this calibration method the tip-oscillation amplitude has to be known and any error in this number will lead to an error in η. smallest tip-oscillation amplitude. At a given z, larger amplitudes lead to smaller frequency shifts in accordance with the theoretical results in Ref. 7. The minimum in f shifts to larger z and becomes more shallow when the amplitude is increased. For the smallest amplitudes below.4 nm the curves are almost identical. We therefore conclude that for these amplitudes the force-gradient approximation is still valid even when the tip is very close to the sample. P (pw) 1.5 1.5 (a) 1.2 1.6 2..81.56 C (af) 1 5 5 1 f (Hz) 2 4 6 8 (b) 2..56 2 4 6 df/dz (N/m) 15 2 25 4 2 2 4 z (nm) Figure 2: Tip-sample capacitance C(z) as a function of the tip-sample separation z measured at 1.7 K. 4. Force distance curves Fig. 3a shows the frequency shift f as a function of the tip-sample separation z at zero bias voltage for different tip-oscillation amplitudes between.56 nm and 2. nm. All these curves have been measured for both sweep directions and no significant hysteresis was found. At large z, all the curves exhibit the wellknown slow decrease of f with decreasing z until a minimum is reached. Beyond this minimum f increases steeply with a rate of up to 1 Hz/Å at the 41 42 43 44 Figure 3: (a) Frequency shift vs. distance measured at 1.7 K for various tip-oscillation amplitudes. (b) Dissipated power for different amplitudes. Fig. 3b shows the corresponding measurements of the dissipated power P as a function of the tip-sample separation z. It is calculated from P = U exc I TF, where U exc is the excitation voltage applied to the tuning fork and I TF is the tuning fork current measured in phase with the excitation. We use the fact that the power dissipated in the system, in particular the power dissipated during tip-sample interaction, has to be supplied by the electrical excitation of the tuning fork. At large z the dissipated power is constant. When the tip is approached to the surface the power starts to increase into the pw-range. For larger tip-oscillation amplitudes the onset of the dissipation is at larger tipsample separations. Created using Adobe FrameMaker 4/5 Power Macintosh Release 5.5.3

Even at the smallest tip-oscillation amplitudes the onset of the power dissipation occurs at a tip-sample separation at which the frequency shift is negative and the tip experiences an overall attractive force. We believe that at these tip-sample separations the foremost tip atoms start to come into intimate contact with the surface atoms and in addition to the conservative tipsample interaction forces there occur dissipative phenomena. We speculate that phonon emission into the tip and the sample, or electrical losses due to the induced Kelvin current might be responsible for this effect. Further investigations are necessary to clarify the origin of the power dissipation in the system. However, it is interesting to notice that the power increase occurs at a position in the f( z) curves where other authors have reported true atomic resolution [2,3]. At smaller tip-sample separations the tip must deform massively but the reproducibility of our f( z) curves suggest that this deformation is elastic. 5. Conclusions In this paper we have reported force-distance studies measured below 4.5 K using piezoelectric tuning fork sensors. Very high quality factors up to 12 allow the detection of frequency shifts of less than 1 mhz. Small tip-oscillation amplitudes down to.5nm have been realized. Three methods for relating frequency shifts with force gradients led to similar calibrations of the force sensor. The setup performs well in magnetic fields up to 8 T and is therefore ideally suited for the future investigation of semiconductor nanostructures. Measurements were performed in which the force gradient and the dissipated power were measured simultaneously as a function of the tip-sample separation on a HOPG surface. The data suggest that already at relatively large tip-sample separations when the tip experiences a net attractive force the foremost tip atoms touch the surface and power is dissipated. At smaller tip-sample separations the tip is significantly elastically deformed and at the same time the power dissipation increases strongly. Acknowledgements We thank K. Karrai for very fruitful discussions on the use of tuning forks as force sensors and W. Allers for very stimulating discussions about force-distance studies. Financial support by ETH Zurich is gratefully acknowledged. References [1] G. Binnig, C.F. Quate, C. Gerber, Phys. Rev. Lett. 56 (1986) 93. [2] F.J. Giessibl, Science 267 (1995) 68; P. Güthner, J. Vac. Sci. Technol. B 14 (1996) 2428; R. Lüthi, E. Meyer, M. Bammerlin, A. Baratoff, T. Lehmann, L. Howald, Ch. Gerber, H.-J. Güntherodt, Z. Phys. B 1 (1996) 165; R. Erlandsson, L. Olsson, P. Martensson, Phys. Rev. B 54 (1996) R839; S. Kitamura, M. Iwatsuki, Jpn. J. Appl. Phys. 35 (1996) L668. [3] W. Allers, A. Schwarz, U.D. Schwarz, R. Wiesendanger, Appl. Surf. Sci. 14 (1999) 247. [4] P. Günther, U.Ch. Fischer, K. Dransfeld, Appl. Phys. B 48 (1989) 89; K. Karrai, R.D. Grober, Appl. Phys. Lett. 66 (1989) 1842; H. Edwards, L. Taylor, W. Duncan, A.J. Melmed, J. Appl. Phys. 82 (1997) 98; F.J. Giessibl, Appl. Phys. Lett. 73 (1998) 3956; M. Todorovic, S. Schultz, J. Appl. Phys. 83 (1998) 6229. [5] J. Rychen, T. Ihn, P. Studerus, A. Herrmann, K. Ensslin, Rev. Sci. Instr. 7 (1999) 2765. [6] J. Rychen, T. Ihn, P. Studerus, A. Herrmann, K. Ensslin, H.J. Hug, P.J.A. van Schendel, H.J. Güntherodt, submitted to Rev. Sci. Instr. [7] F. J. Giessibl, Phys. Rev. B 56 (1997) 161; H. Hölscher, U.D. Schwarz, R. Wiesendanger, Appl. Surf. Sci 14 (1999) 344. Created using Adobe FrameMaker 5/5 Power Macintosh Release 5.5.3