Simulation of atomic force microscopy operation via three-dimensional finite element modelling

Similar documents
FEM-SIMULATIONS OF VIBRATIONS AND RESONANCES OF STIFF AFM CANTILEVERS

Introductory guide to measuring the mechanical properties of nanoobjects/particles

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

Theoretical Manual Theoretical background to the Strand7 finite element analysis system

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

Features of static and dynamic friction profiles in one and two dimensions on polymer and atomically flat surfaces using atomic force microscopy

Design and Analysis of Various Microcantilever Shapes for MEMS Based Sensing

Software Verification

D : SOLID MECHANICS. Q. 1 Q. 9 carry one mark each. Q.1 Find the force (in kn) in the member BH of the truss shown.

SUPPLEMENTARY INFORMATION

Supplementary Figures

Numerical Modeling of Interface Between Soil and Pile to Account for Loss of Contact during Seismic Excitation

An Increase in Elastic Buckling Strength of Plate Girder by the Influence of Transverse Stiffeners

UNLOADING OF AN ELASTIC-PLASTIC LOADED SPHERICAL CONTACT

Mechanics of Irregular Honeycomb Structures

RHK Technology Brief

ME 2570 MECHANICS OF MATERIALS

9 MECHANICAL PROPERTIES OF SOLIDS

Imaging Methods: Scanning Force Microscopy (SFM / AFM)

Mechanical Engineering Ph.D. Preliminary Qualifying Examination Solid Mechanics February 25, 2002

Finite element prediction of the ultimate axial load capacity of V-section band clamps

five Mechanics of Materials 1 ARCHITECTURAL STRUCTURES: FORM, BEHAVIOR, AND DESIGN DR. ANNE NICHOLS SUMMER 2017 lecture

NIS: what can it be used for?

A HIGHER-ORDER BEAM THEORY FOR COMPOSITE BOX BEAMS

Supplementary Information: Nanoscale heterogeneity promotes energy dissipation in bone

FLEXIBILITY METHOD FOR INDETERMINATE FRAMES

Outline. Tensile-Test Specimen and Machine. Stress-Strain Curve. Review of Mechanical Properties. Mechanical Behaviour

FINITE ELEMENT ANALYSIS OF TAPERED COMPOSITE PLATE GIRDER WITH A NON-LINEAR VARYING WEB DEPTH

Effect of AFM Cantilever Geometry on the DPL Nanomachining process

EFFECT OF STRAIN HARDENING ON ELASTIC-PLASTIC CONTACT BEHAVIOUR OF A SPHERE AGAINST A RIGID FLAT A FINITE ELEMENT STUDY

Module-4. Mechanical Properties of Metals

A General Equation for Fitting Contact Area and Friction vs Load Measurements

Arbitrary Normal and Tangential Loading Sequences for Circular Hertzian Contact

TECHNIQUE TO EVALUATE NANOMECHANICS OF CANTILEVER AT NANOSCALE BY USING AFM

Bending Load & Calibration Module

New concept of a 3D-probing system for micro-components

SUPPLEMENTARY NOTES Supplementary Note 1: Fabrication of Scanning Thermal Microscopy Probes

3.032 Problem Set 2 Solutions Fall 2007 Due: Start of Lecture,

force measurements using friction force microscopy (FFM). As discussed previously, the

Table of Contents. Preface...xvii. Part 1. Level

Tolerance Ring Improvement for Reducing Metal Scratch

Effect of Strain Hardening on Unloading of a Deformable Sphere Loaded against a Rigid Flat A Finite Element Study

STRAIN GAUGES YEDITEPE UNIVERSITY DEPARTMENT OF MECHANICAL ENGINEERING

Special edition paper

Iraq Ref. & Air. Cond. Dept/ Technical College / Kirkuk

Intensity (a.u.) Intensity (a.u.) Raman Shift (cm -1 ) Oxygen plasma. 6 cm. 9 cm. 1mm. Single-layer graphene sheet. 10mm. 14 cm

142. Determination of reduced mass and stiffness of flexural vibrating cantilever beam

PERIYAR CENTENARY POLYTECHNIC COLLEGE PERIYAR NAGAR - VALLAM THANJAVUR. DEPARTMENT OF MECHANICAL ENGINEERING QUESTION BANK

KINGS COLLEGE OF ENGINEERING DEPARTMENT OF MECHANICAL ENGINEERING QUESTION BANK. Subject code/name: ME2254/STRENGTH OF MATERIALS Year/Sem:II / IV

Figure 1 Lifting Lug Geometry with Weld

Effect of Angular movement of Lifting Arm on Natural Frequency of Container Lifting Mechanism using Finite Element Modal Analysis

Measurement and modelling of contact stiffness

INTERNATIONAL JOURNAL OF APPLIED ENGINEERING RESEARCH, DINDIGUL Volume 2, No 1, 2011

Experiment Two (2) Torsional testing of Circular Shafts

TE 75R RESEARCH RUBBER FRICTION TEST MACHINE

D : SOLID MECHANICS. Q. 1 Q. 9 carry one mark each.

Chapter 2 Correlation Force Spectroscopy

2012 MECHANICS OF SOLIDS

2d-Laser Cantilever Anemometer

INFLUENCE OF FLANGE STIFFNESS ON DUCTILITY BEHAVIOUR OF PLATE GIRDER

Contact Modeling of Rough Surfaces. Robert L. Jackson Mechanical Engineering Department Auburn University

Computational Simulation of Dynamic Response of Vehicle Tatra T815 and the Ground

DEPARTMENT OF MECHANICAL ENIGINEERING, UNIVERSITY OF ENGINEERING & TECHNOLOGY LAHORE (KSK CAMPUS).

A SCIENTIFIC APPROACH TO A STICKY PROBLEM

Finite Element Simulation of Bar-Plate Friction Welded Joints Steel Product Subjected to Impact Loading

NUMERICAL ANALYSIS OF A PILE SUBJECTED TO LATERAL LOADS

Finite Element Modelling with Plastic Hinges

Mechanics of Materials Primer

Measuring the spring constant of atomic force microscope cantilevers: thermal fluctuations and other methods

Finite Element Models for European Testing: Side Impact Barrier to WG13 Pedestrian Impactors to WG17

FEA A Guide to Good Practice. What to expect when you re expecting FEA A guide to good practice

COURSE TITLE : THEORY OF STRUCTURES -I COURSE CODE : 3013 COURSE CATEGORY : B PERIODS/WEEK : 6 PERIODS/SEMESTER: 90 CREDITS : 6

Mechanics of Materials II. Chapter III. A review of the fundamental formulation of stress, strain, and deflection

1 Force Sensing. Lecture Notes. 1.1 Load Cell. 1.2 Stress and Strain

PES Institute of Technology

Nonlinear Finite Element Modeling of Nano- Indentation Group Members: Shuaifang Zhang, Kangning Su. ME 563: Nonlinear Finite Element Analysis.

Finite Element analysis of Laterally Loaded Piles on Sloping Ground

On Nonlinear Buckling and Collapse Analysis using Riks Method

Supplementary Information to the article entitled: Internally Architectured Materials with Directionally Asymmetric Friction

ME Final Exam. PROBLEM NO. 4 Part A (2 points max.) M (x) y. z (neutral axis) beam cross-sec+on. 20 kip ft. 0.2 ft. 10 ft. 0.1 ft.

Mechanical Properties of Materials

Dynamic Analysis of a Reinforced Concrete Structure Using Plasticity and Interface Damage Models

Finite Element Analysis of Piezoelectric Cantilever

DYNAMIC RESPONSE OF THIN-WALLED GIRDERS SUBJECTED TO COMBINED LOAD

Structural Dynamics Lecture Eleven: Dynamic Response of MDOF Systems: (Chapter 11) By: H. Ahmadian

Lecture 12: Biomaterials Characterization in Aqueous Environments

VORONOI APPLIED ELEMENT METHOD FOR STRUCTURAL ANALYSIS: THEORY AND APPLICATION FOR LINEAR AND NON-LINEAR MATERIALS

AERO 214. Lab II. Measurement of elastic moduli using bending of beams and torsion of bars

1 332 Laboratories 1. 2 Computational Exercises 1 FEA of a Cantilever Beam... 1 Experimental Laboratory: Tensile Testing of Materials...

Nanomechanics Measurements and Standards at NIST

Elastoplastic Deformation in a Wedge-Shaped Plate Caused By a Subducting Seamount

1 Introduction. Abstract

Week 10 - Lecture Nonlinear Structural Analysis. ME Introduction to CAD/CAE Tools

Influence of residual stresses in the structural behavior of. tubular columns and arches. Nuno Rocha Cima Gomes

EDEM DISCRETIZATION (Phase II) Normal Direction Structure Idealization Tangential Direction Pore spring Contact spring SPRING TYPES Inner edge Inner d

Lateral force calibration in atomic force microscopy: A new lateral force calibration method and general guidelines for optimization

QUESTION BANK DEPARTMENT: CIVIL SEMESTER: III SUBJECT CODE: CE2201 SUBJECT NAME: MECHANICS OF SOLIDS UNIT 1- STRESS AND STRAIN PART A

Experimental Modal Analysis of a Flat Plate Subjected To Vibration

LINEAR AND NONLINEAR BUCKLING ANALYSIS OF STIFFENED CYLINDRICAL SUBMARINE HULL

STRUCTURAL SURFACES & FLOOR GRILLAGES

Transcription:

IOP PUBLISHING Nanotechnology 20 (2009) 065702 (14pp) NANOTECHNOLOGY doi:10.1088/0957-4484/20/6/065702 Simulation of atomic force microscopy operation via three-dimensional finite element modelling JLChoi 1 and D T Gethin 2 1 College of Engineering, Korea University, Seoul 136-713, Republic of Korea 2 School of Engineering, Swansea University, Singleton Park, Swansea SA2 8PP, UK E-mail: jlchoi@korea.ac.kr and d.t.gethin@swansea.ac.uk Received 23 August 2008, in final form 20 November 2008 Published 15 January 2009 Online at stacks.iop.org/nano/20/065702 Abstract Numerical modelling of atomic force microscopy cantilever designs and experiments is presented with the aim of exploring friction mechanisms at the microscale. As a starting point for this work, comparisons between finite element (FE) models and previously reported mathematical models for stiffness calibration of cantilevers (beam and V-shaped) are presented and discrepancies highlighted. A colloid probe (comprising a plain cantilever on which a particle is adhered) model was developed, and its normal and shear interaction were investigated, exploring the response of the probe accounting for inevitable imperfections in its manufacture. The material properties of the cantilever had significant impact on both the normal response and the lateral response. The sensitivity of the mechanical response in both directions was explored and it was found to be higher in terms of normal rather than lateral sensitivity. In lateral measurements, generic response stages were identified, comprising a first stage of twisting, followed by lateral bending, and then slipping. This was present in the two cantilever types explored (beam and V-shaped). Additionally, a model was designed to explore the dynamic sensitivity by comparing the simulation of a hysteresis loop with a previously reported experiment, and the results show good agreement in the response pattern. The ability to simulate the scan over an inclined surface representing the flank of an asperity was also demonstrated. (Some figures in this article are in colour only in the electronic version) 1. Introduction The atomic force microscope (AFM) represents a technology that is used for measurement at the nanoscale. It is already used extensively for investigating surface topography and material surface development, principally normal adhesion, that may be categorized as a normal force (or stress). Its ability to image the topography of the surface as well as to measure force at the nanoscale has led to unique insights into surface studies. The AFM can be operated in a number of modes dependent on the application [1]; it can also be configured as a colloid probe [2]. There is considerable interest in measuring lateral forces at a surface, categorized in this case as a shear force. This has particular relevance for friction measurement where sliding takes place, or shear adhesion when a detachment force (or stress) needs to be determined. Calibration of the AFM is a crucial step in obtaining reliable data in force measurements. Procedures for performing normal force calibrations are becoming established for perfectly formed probes, such as those obtained by silicone nitride fabrication, and this will be discussed below. In the case of colloid probes, there is likely to be some geometric imperfection, for example a nonspherical colloid particle, or the particle position being offset from the cantilever centreline, see figure 1. To date, lateral force microscopy (LFM) may only be used as a comparator, rather than a method to quantify an absolute level of lateral force. This is due to the challenges that are faced in performing a calibration in this measurement mode. A particular difficulty in measuring lateral force is that, at this resolution, it will be influenced by surface topography, for example, where the tip dimensions and roughness dimensions are of the same order [3], and that separating pure force effects 0957-4484/09/065702+14$30.00 1 2009 IOP Publishing Ltd Printed in the UK

Figure 1. SEM images taken for colloid probe tip with (a) overuse of adhesive, (b) wrong position, (c) imperfect particle shape, (d) clustered particles, (e) and (f) large particle. from topography effects is difficult. In this respect, simulation of the experiment may provide some insight as these issues can be separated numerically and their influences quantified. The aim of this study is to explore how simulation may be used to estimate how geometric imperfections influence the probe response and further to explore how it may be used in conjunction with LFM measurement to estimate either friction or shear adhesion forces. In reviewing the literature, it is found that the factors that affect the probe response are not fully understood; simplifying assumptions are often made, and considerable discrepancies still exist [4]. The calibration of an AFM involves stiffness calibration as well as detector calibration accounting for optical path influence [5]. At present, many AFM experiments are performed with a set stiffness value provided by the manufacturer of the cantilever, and this is combined with the sensitivity of the detector. Lateral stiffness may be defined in a number of ways. These include definitions based on a plain lateral movement, torsion, or more realistically a combination of lateral deflection and torsion. Lateral deflection will influence the optical path for the detector, but this is likely to be secondary in comparison with torsion. However, since the accuracy in the normal mode may be improved through refinement of the optical analysis [5], it is anticipated that a detailed understanding of AFM response to lateral load will lead to improved resolution and ultimately to standard calibration procedures. Numerical modelling using finite element modelling (FEM) offers the potential to achieve insight/understanding of the response of AFM metrology, and two examples are set out in [6, 7]. The work in [6] represents a dynamic modelling approach in which the effect of contact on the resonant response of a plain simple beam cantilever is explored. This work focused on derivation of formulations reflecting the exact excitation mechanisms for different modes. The work in [7] is concerned with a V-shaped cantilever under steady loading conditions in which the impact of geometry on the response is explored through FEM. The work excludes a tip contact model and the main focus is to demonstrate the impact of geometric nonlinearity as the cantilever is subjected to large deformations. To date, the lateral mode has received less attention, and so the purpose of this paper is to explore the lateral mode response for simple beam and V-shaped cantilever designs to quantify the influence of the parameters on the probe response. The current study extends the types exemplified in [6, 7] through incorporation of a mechanical contact model coupled with exploration of lateral operation over surfaces for which topographical influences are investigated. Comparison will be drawn with the studies [6, 7] where appropriate. Modelling the probe response forms a potential solution to establish a lateral stiffness as the mechanical component of lateral sensitivity. The key focus is to build a model to simulate operations of the AFM incorporating the colloid probe system to explore parameters that affect it and investigate how it may influence the measurements and outcome. This will aim to establish a tool and quantify parameter influences that are currently understood in a qualitative manner. In order to calibrate the stiffness, definitions of normal and lateral stiffness are required. For normal stiffness, this is straightforward as it is just the ratio of force to deflection in the normal direction. There are a number of ways of determining this [4], and this has now led to standard procedures that depend to some extent on the final application. However, for most precise measurements each probe needs to be calibrated in situ. 2. Model characteristics A cantilever is a key component in the AFM, and there are two designs in common use [1]. Figure 2 displays these two types, configured as a colloid probe. The boundary conditions 2

Figure 2. Geometries of (a) rectangular beam, (b) V-shaped cantilever (dimensions in μm). Figure 3. Geometries of cantilever beam model calibration: (a) normal, (b) lateral mode (μm). reflect fixing at the probe mounting point, and the target surface is brought up to the tip to displace it by a prescribed amount and interaction is captured through a contact model. In the investigation of lateral response this is followed by lateral movement of the target surface under the probe to capture part of the raster scan. This emulates its application in a measurement cycle. The force evolution is then obtained directly from the reaction components that are induced at the fixing points. The FEM in this work used a commercial software, ELFEN [8]. This system has been developed to include advanced capability to handle elasto-plasticity and contact. The elasto-plastic model is based on a linear elastic model coupled with a plasticity that includes hardening. Where it happens, yielding is assumed to occur based on a von Mises criterion. In contact, simple interaction laws may be incorporated (e.g. stick slip and Coulomb friction) and contact areas are identified automatically as they occur. In FEM that includes contact, the solution strategy often uses an explicit time marching scheme in which small time steps are used to ensure analysis stability, and the increments adapt to ensure that stability criteria are satisfied automatically. These analyses are generally very time consuming. Implicit schemes are unconditionally stable, and in this case, the software ELFEN facilitates an implicit calculation that includes simple contact. Overall this is computationally more efficient, and it was therefore used in this work. To explore the differences that may arise from discretization, simulation for the simple beam was undertaken, with dimensions from figure 3(a). A structured mesh comprising near cubic elements yielded the best result. The unstructured mesh produced a slightly higher stiffness value compared with that of the structured counterpart. This is mainly due to the number of elements through the thickness of the beam, and ideally, at least two should be used. For the unstructured mesh model, having two layers through the thickness was computationally prohibitive and led to simulation failure due to the discretization. However, while noting this difference, the unstructured mesh system was used later since it allows discretization of a colloid probe, including the particle attached to the tip. For colloid probe operation, example material properties of the tip and the cantilever are itemized in table 1 [9 11]. These have been chosen as the initial aim was to explore the frictional response between a particle and a rigid surface. This aim was later modified to focus on calibration and the impact of geometric imperfection. In this work, the influence of the gold reflecting layer on the back of the cantilever was excluded from the analysis; however, the cantilever can be modelled as a bimetallic structure if this is required. This will present difficulties in terms of finite element meshing either requiring the application of very fine elements, or the derivation of special elements to approximate the bimetallic properties. Thus a single material has been assigned in this study. The force against displacement or force against slope data are the key results for calibration. For the normal mode, force against displacement data give the stiffness parameter, which is critical for the mechanical contribution to the calibration. A similar analysis may be carried out for the lateral mode, but a variety of lateral stiffness definitions are reported in the literature [12]. Force against slope at the cantilever tip is 3

Part Table 1. Material properties of parts in friction modelling. Material Young s modulus (N μm 2 ) Poisson s ratio Cantilever Silicon nitride 0.28 0.2 Tip (colloid probe) Steel 316L 0.2 0.3 Copper 0.115 0.35 Target surface Mild steel 0.2 0.3 important for both modes since the AFM obtains the deflection of the cantilever by reflecting a laser beam off the back surface and the slope is the principal factor in determining the beam path. This is especially the case for the lateral mode, where a number of studies [13] highlight this as the method of estimating the mechanical component for calibration. The slope is readily derived with sufficient accuracy through simple post-processing of displacement information in the region where the laser beam is reflected. 3. Calibration of stiffness 3.1. Cantilever beam model calibration The simple cantilever design allows an analytical solution based on simple bending and torsion models for both normal and lateral modes. This solution is subject to the assumptions embodied into the simple bending and torsion theories and so may be considered as first order. However, such firstorder models provide a datum and allow comparison with a finite element model as an initial benchmark for calibration. Information from Liu [14] is chosen for this purpose and the beam is modelled with the geometry summarized in figure 3. For small deflections, the normal and lateral stiffnesses can be calculated using the equations k n = 3EI L 3 (1) k l = GK (2) h 2 L where k n and k l are the spring constants in the normal and lateral directions, E is the Young s modulus (0.2 Nμm 2 ); ν is Poisson s ratio (0.3); I and K are the second bending and torsion moment of area of the cross section; ( (K wt3 16 16 3 3.36 t ) ), for t w ; w L is the beam length; w is the beam width; h is the height E 2(1+ν) of the tip; G is the shear modulus ( );andt is the beam thickness. Equation (1) is based on a simple deflection model and equation (2) on a pure torsion type equation that excludes any lateral deflection effects. To quantify both bending and twisting of the cantilever independently, load components were applied in the form of normal and lateral displacements at the end of the tip. The lateral stiffness was obtained by the distance the tip end travelled and the reaction force in the lateral direction at the cantilever fixing point. The lateral stiffness therefore includes the combined influence of lateral bending and torsion mechanisms. From calculations in which a small displacement of 10 μm was applied, the normal stiffness was found using equation (1) to be k n = 0.312 10 6 N μm 1, and for a structured finite element mesh comprising near cubical elements k n = 0.353 10 6 N μm 1. In addition, the lateral stiffness from equation (2) was k l = 0.161 10 3 and 0.106 10 3 N μm 1 from the simulation. The differences are 13% and 34%, respectively; the latter demonstrates clearly that the simple analytical model may not be acceptable. The smaller discrepancy between the simulation and normal stiffness based on a simple beam bending model may be attributed to a more accurate representation of the physics. This is a point made in Clifford [4] for analysis of normal stiffness calibration, in that the FE model does not embody the assumptions that are required for the simple bending model; thus it may be described as ab initio, and is the most accurate. It is interesting to note that the AFM beam cantilever is significantly more sensitive in the normal mode (about 300 times for the case study set out above), confirming its ability to resolve normal forces more accurately. 3.2. V-shaped cantilever model calibration An approximate analytical calibration for the V-shaped cantilever may be obtained from equations by Neumeister [12] and Noy [15]. These approximate the cantilever as a series of sections for which the stiffness is obtained through superposition. The normal stiffness constant was derived by Neumeister [12] for a V configuration as the sum of three parts: the deflection of a clamped triangular plate ( 1 ), the end deflection of the two beams ( 2 ), and rotation (θ 2, i.e. deflection angle in the normal direction), shown in figure 4(a). This has led to the equation ( ) w N = 1 + 2 + θ 2 sin α d = N 1 (3) k n where N is the force applied at the end of the cantilever to induce normal deflection of the device. E is the Young s modulus; and ν is the Poisson s ratio. Using this equation, the normal stiffness constant for the cantilever dimensions shown in figure 4(a) is k n = 1.453 10 6 N μm 1 (N = 0.0126 10 3 N, E = 0.2 Nμm 2,ν = 0.3). Noy[15] also reported an approximation for k n as a sum of the normal stiffness constants for two rectangular beams, equation (4). From the dimensions in figure 4(b), the normal stiffness constant is found to be k n = 0.574 10 6 N μm 1 (E = 0.2 Nμm 2,ν = 0.3). The differences are significant, thus highlighting the need for a more precise evaluation in which approximating simplification is reduced. k n = Ew 2t2 3 2L 3. (4) 2 For the lateral stiffness constant, Neumeister [12] reported an equation based on the representation shown in figure 4(a). 4

Figure 4. Geometries and dimensions using the (a) Neumeister [12]and(b)Noy[15] models. Figure 5. Deflection of the V-shaped cantilever with forces applied in the normal and lateral directions. The slope analysis was divided into two: the twist of the triangular plate and that of the beams. From the total rotation of these two parts, a torsional stiffness constant was calculated, following which the lateral stiffness constant was derived. Et 3 k l = 3(1 + ν)h 2 ( 1 tan α log w d sin α + L cos α w 3sin2α ) 1. 8 (5) From this equation, the lateral stiffness constant was found to be k l = 0.143 10 3 N μm 1. It is shown to be dependent on the cantilever design, but is particularly sensitive to the height of the tip (h). Instead of a direct derivation of the lateral stiffness constant, an equation relating both normal and lateral stiffness constants is reported in Noy [15]. This was done based on a hypothesis that both forces originate from the breaking of intermolecular interactions, and it could avoid the difficulties in derivation of the lateral stiffness. Thus, given the calculated ratio of normal to lateral stiffness and a good calibration of normal stiffness (including the optical path calibration), it was hypothesized that the lateral stiffness could be calculated. 2 k l = [6cos 2 θ + 3(1 + ν)sin 2 θ] ( L2 h 2 ) 2 k n. (6) By calculation from the geometry in figure 4(b), the lateral stiffness was found to be k l = 0.133 10 3 N μm 1. For comparison, an elastic model of the V-shaped cantilever was built with the dimensions shown in figure 4 with a cube shaped tip to facilitate lateral load application without any geometric complication. A very fine structured mesh was used and the result is presented below in figure 5, shownasa deflected cantilever. From this model, the normal stiffness constant was found to be k n = 1.031 10 6 N μm 1, and the lateral stiffness was found to be k l = 0.123 10 3 N μm 1. These values lie between the results from the analytical solutions. The stiffness derived from the finite element model achieved best agreement with the approximations proposed by Neumeister [12]. Since the FE model is more free from simplification, it can be used to establish the mechanical component of normal and lateral stiffness for the AFM. 4. Colloid probe model A colloid probe offers the opportunity to explore the interaction between two chosen materials. The manufacture of a colloid probe is difficult and will inevitably result in geometric imperfections, as shown in the examples in figure 1. However, at this stage emphasis is placed on exploring parameters that will influence the colloid probe response in terms of its size and particle fixing rather than exploring geometric imperfections. Work will also focus on the response as it is drawn over a sloping and then undulating surface as an exploratory step to characterize the response over real surfaces. A colloid probe model will be constructed to explore the overall cantilever response to loading when assuming the geometries shown in figure 2. First, a three-dimensional model 5

Figure 6. Geometry of the adhesive. of the simple beam with a spherical particle attached to the tip was considered, as shown in figure 2(a). The contact model was defined between the tip and the target surface, and the particle attached to the end was assumed to behave elastically. The target surface plate was then subjected to an applied displacement in the normal direction. The surfaces at the mounting end of the beam cantilever were fixed, at which reaction forces were then captured. The particle was attached to the end of the cantilever using a glass glue adhesive to form a colloid probe; see figure 1. The geometric representation is captured in figure 6, which shows the finite element model at the probe tip. This figure also shows a range of neck radii that were investigated within the sensitivity case studies that will be described below. Although the material properties of the glue are different, due to difficulties in defining the geometry and mesh generation, it was treated as having the same material properties as the particle included in the tip definition. Loading was applied through the contact model, i.e. the target surface was constrained to move, to come into contact in the normal direction. The interactions in the normal direction are presented below for steel and copper particles fixed at the cantilever tip. As is seen from figure 7, the force deflection and slope are related linearly, confirming the elastic behaviour of the system. In addition, because the deflection of the cantilever is small, the contact patch does not move sufficiently to affect the normal direction mechanical stiffness of the system. Both graphs show very small differences between the two colloid probes even though the particles that form the probe have a significant difference in elastic modulus. For the particle material properties used, this confirms the dominance of the material properties of the cantilever in determining the normal stiffness coefficient. However, this result may change in the event that very soft particles, such as biological materials, are attached to the cantilever. A similar investigation was performed for the V-shaped cantilever (figure 2(b)), and a high normal stiffness was observed, reflecting its more rigid construction. Further models were developed in which the particle at the probe tip was assumed to deform plastically according to a yield model Figure 7. Graph of normal direction reaction force against displacement for both 316L and OFHC Cu probes with elastic material properties for beam model. Table 2. Plastic properties of ductile materials. Material Yield stress (N μm 2 ) Hardening modulus (N μm 2 ) 316L 0.24 10 3 0.44 10 3 OFHC Cu 0.05 10 3 0.44 10 3 basedonavonmisescriterioninwhichshearisassumedtobe the dominant yielding mode. Although many ductile materials have nonlinear hardening characteristics, the von Mises model used has a linear approximation. The slope of the hardening curve was defined using the values in table 2. Again the dominance of cantilever material properties is highlighted, with plasticity having a very small effect on stiffness. This is an expected result as the loads and deflections remain very small, even though there may be local yielding at the tip point in the contact and some areas of the glued joint, asshowninfigure9. The latter seems to be due to geometrical description in terms of the variation of neck radii, since the tip and glue were defined as having the same material properties. The influence of plasticity on stiffness can be observed in figure 8(b), where a very soft tip (the modulus is 100 times smaller) was used. Even under this circumstance, the beam properties dominate, confirming that the probe may be used for characterizing soft materials such as tissue. Geometric factors that may affect the normal deflection were also investigated (i.e. fixity geometry and size of colloid probe). However, since the material properties of the adhesive were not reflected in these models, but investigated purely in terms of geometry, both factors will mainly show the effect of tip height on the deflection in the normal direction. The case study was modelled with elastic material properties since the effect of plasticity was negligible for the material used, as reported above. Although only a very small difference was observed for the reaction force against displacement (figure 11), some increase in force was seen against deflection slope, especially in the case of the larger colloid fixity. The deflection slope increased for both cases, because the larger the tip height, the greater the deflection angle in the normal direction with the 6

Figure 8. Graphs of normal direction reaction force against displacement for (a) 316L and OFHC Cu probes with plastic material properties and (b) a very soft tip (100 times softer) with elastic and plastic material properties for the V-shaped model. Figure 9. Plastic strain zones for OFHC Cu (V-shaped model). Figure 10. Geometries of (a) original, (b) larger colloid fixity (CF), (c) larger probe size (PS). same applied displacement of the target surface, especially when the material properties of the cantilever are most influential (figure 12). This is as expected, and demonstrates the additional stiffening effect of the fixed region, whereas for the smaller glued region there is more freedom for the cantilever end to deflect. However, in general, the response was cantilever driven compared to the effects of particle size and fixing. A case study of a cantilever engaging an inclined surface will give an insight to understanding the linking of normal and lateral sensitivity and their undesirable mutual dependence. An inclined target surface is introduced to replace the horizontal target plate; see figure 13. By bringing this vertically into contact with the probe, it will induce both normal deflection and lateral torsion. This model can also represent the sloping flank of a rough surface. The slope and friction coefficient at the contact can be varied. The slope was set at 0 and 45, and a friction coefficient of 0.0 and 0.1 was also explored. The results from these simulations are shown in figure 14. It can be seen that the normal reaction force is nearly independent of the surface inclination. This is due to no 7

Figure 11. Graphs of the normal direction reaction force against displacement for (a) larger colloid fixity, (b) larger probe size with elastic material properties for the V-shaped model (316L probe). Figure 12. Graphs of normal direction reaction force against deflection slope for (a) larger colloid fixity, (b) larger probe size with elastic material properties for the V-shaped model (316L probe). Figure 13. Schematic diagrams for the inclined surface models. The results may be contrasted against the threedimensional FEM of V-shaped cantilevers, as reported in [7], where the bending behaviour of the tip displacements was investigated and their linearity was found to occur below 100 and 10 nm for the normal and lateral directions, respectively. However, our current study showed almost constant stiffness values for both normal and lateral directions due to the influence of the cantilever characteristics and different cantilever/tip geometries. Also, since no contact model at the tip was defined in [7], the slipping stage was not taken into account, whereas this is one of the key factors for LFM measurements. slipping occurring at the contact; thus the probe is constrained to move through nearly the same distance as when the surface is not inclined. The slight reduction in displacement occurs because the probe now twists slightly, and so the normal deflection of the beam is reduced. The inclination leads to significant changes in lateral force development, as shown in figure 14(b), due to resolution at the plate contact. As a further check, zero friction was applied at the tip contact, and this led to identical normal and lateral loads as the engagement proceeded. This is the anticipated result, confirming the correctness of the model. 5. Shear interactions For both colloid probe models (i.e. beam and V-shaped), the shear interaction was explored for the geometries shown in figure 2. Models were developed to simulate a complete friction experiment through engagement of the probe, followed by its lateral movement against the target surface. The contact between the probe and the target surface was defined to have a friction coefficient of 0.1 as a datum. Additionally, since the material selection of the tip had only a very small effect on the results, a 316L particle having an elastic material response 8

Figure 14. Graphs of (a) normal direction reaction force against displacement, (b) lateral direction reaction force against test time for a 316L probe on an inclined target surface with elastic material properties for a V-shaped cantilever. Figure 15. Graphs of reaction force in the lateral direction, lateral displacement at the tip end, and lateral deflection slope against test time for the beam model (left to right respectively). was chosen. With the lateral direction motion applied while holding the normal engagement constant, a complex response is observed for the probe. The response can be generalized as the first action is the occurrence of twisting, reflected in a couple type reaction at the probe fixing point. This is followed by bending in the lateral direction that leads to the generation of a lateral load, and eventually it settles to the slipping stage, at which point no further displacement (torsion and bending) occurs. By exploring the relationship between the lateral force against displacement and the lateral force against slope, the interactions of bending and twisting can be explored. 5.1. Cantilever beam model Figure 15 shows results in terms of lateral force, displacement, and slope, the latter as a measure of twisting over the duration of the sliding part of the cycle. The probe and target surface were engaged by an applied displacement of the target surface to give an interaction force of 0.000 52 N. The lateral displacement at the tip contact point incorporates both twisting and bending of the cantilever. In order to differentiate between these, the displacement at the cantilever end has also been plotted. It can be observed that, for the beam type cantilever, although twisting had a greater influence in shear interaction, 9

Figure 16. Graphs of reaction force in the lateral direction, lateral displacement at the tip end, and lateral deflection slope against test time for the V-shaped model (left to right respectively). bending was also significant, accounting for about 45% of the deflection. This confirms findings from [6], which reports that lateral bending can only be neglected when the lateral interaction is relatively small compared to the lateral stiffness. Bending will be expected to show much less effect for the V- shaped cantilever model shown in the section below. The slope characteristics in figure 15 show how twisting dominates the initial cantilever response for a beam configuration. 5.2. Cantilever V-shaped model In this instance the tip engagement period is 0 0.5 time units and lateral loading starts to be generated following its application at 0.5 units, seen in figure 16. This pattern reflects the results for the simple beam arrangement. The negative displacement of the tip during the normal loading stage is an unexpected result. This was attributed to an offsetting of the contact at the tip due to imprecise representation of the spherical surface. Such a twisting mechanism could also be a problem in an experiment, where the colloid probe has geometric imperfections. It can be seen clearly that the lateral bending has much less effect on the V-shaped cantilever than for the beam model. This shows that twist is the dominant response of this cantilever type. One of the main factors identified that affect the response is the height of the tip (or the diameter of the sphere). For this purpose, a model with a larger colloid fixity was considered, having the geometry shown in figure 10(b); the results of the simulation are shown in figure 17. Observing figure 17, the difference that appears due to the introduction of larger colloid fixity can be seen. The final lateral force level is identical due to the equivalent mechanical engagement and contact friction coefficient choice. Also, the finite element mesh captures the tip geometry more accurately, so the contact between the tip and the target surface is on the cantilever centreline and therefore load induced twist does not occur. This highlights the importance of the discretization level and also the importance of correct positioning of the particle at the cantilever tip. The graphs of lateral displacement of the probe and the slope at the cantilever end show larger values for the greater tip height an expected result. 6. Lateral scanning over sloping surfaces Practical surfaces are not smooth, and depending on the roughness wavelength and particle size there can be strong interaction when experiments are conducted in lateral scanning mode and the tip (or colloid probe) interlocks with the surface topography. In order to capture rough surface topography, the idea of encountering a slope during shearing was explored by introducing a simple bump, see figure 18, torepresentthe flank of an asperity or a groove in the surface. The geometry for the cantilever was the same as for the previous V-shaped cantilever probe. The model was designed to apply a displacement on the target surface so that the probe will encounter a bump with slope of 0.5. The friction coefficient at the surface was again assigned a value of 0.1. As can be observed from the graphs in figure 19, an additional step increase in all of the parameters can be seen after around 3 test time units, as the probe tip encounters the sloped section as it traverses the surface. Even with a slope 10

Figure 17. Graphs of reaction force in the lateral direction, lateral displacement at the tip end, and lateral deflection slope against test time for the original and larger colloid fixity model (V-shaped) (left to right respectively). Figure 18. Schematic diagrams for the bump model. Figure 19. Graphs of (a) reaction force in the lateral direction and (b) lateral displacement on the tip end for the V-shaped bump model. 11

Figure 20. Graphs of (a) lateral direction reaction force against displacement (tip end), (b) lateral reaction force against deflection slope for the V-shaped bump model. Figure 21. Geometries of a V-shaped cantilever for the emulation model in Ogletree [13] with plastic material properties. of 0.5, the increase in lateral force and deflection slope due to further normal deflection was significant. The end process of shearing did not include the probe reaching the top of the bump; however, both the normal and lateral forces exhibit similar increases during this sliding stage (figure 20). The simulation of lateral force evolution over an undulating surface may be compared with experimental data. Ogletree [16] reported experimental data for LFM on inclined surfaces to produce a friction loop as an attempt at calibrating the response dynamically. The experiment involved a V-shaped cantilever with pyramid shape tip scanning up and down plain inclined surfaces having precise geometry. Emulation of this experimental work was performed with the geometries shown in figure 21 that represent a Veeco TR Series probe [17]. The cantilever shown in figure 21 was built into a model to simulate sliding over a faceted surface to represent a friction loop. In general, the graphs shown in figure 22 agree with the experimental results, and thus the friction loop was produced. By observing the lateral force evolution for each graph, it is clearly seen that for the positive inclined surface the graph shifted up and for the negative inclined surface it shifted down, as expected from the experimental result. However, when moving up and down the inclined surface, the results were not symmetrical as for the case of the plain surface, which follows the pattern shown from the experiment. 7. Conclusion An exploration of AFM response in both the normal and lateral directions has been carried out. The normal response was much more sensitive than the lateral response, and hence more accurate results could be obtained. The analytical approximation of the normal and lateral stiffnesses for both a simple beam [14] and a V-shaped model [12, 15] was compared with the results from the FE model and showed the best agreement with the mathematical modelling of Neumeister [12] (for both normal and lateral response equations). The differences in stiffness between analytical and FE simulation remains significant at 13% in the normal and 34% in the lateral direction for a plain beam and 29% in the normal and 14% in the lateral direction for a V-shaped cantilever. For the lateral loading characteristics, both the beam and the V-shaped cantilever were affected by twisting, caused by torsion, and bending in the lateral direction. However, the bending had more significant influence for the beam model (around 45% of the deflection), whereas twisting dominated for the V-shaped model. FE models that include a colloid probe tip revealed that the material properties of the cantilever 12

Figure 22. Graphs of lateral displacement of target surface against lateral reaction force for plain (flat), inclined (+14 ), and inclined ( 12.5 ) cases, respectively, for (a) the FE model, (b) from Ogletree [16]. were most significantly influential for both normal and lateral responses, the colloid tip had little impact, even for the softer materials that were considered. The tip height, including the area of fixity and size of the probe tip, was also found to affect the deflection angle of the cantilever. Additionally, the importance of discretization, probe position, and shape were highlighted by twisting of the cantilever when a pure normal load was applied. The relationship between normal and lateral sensitivities was explored by introducing an inclined (slope = 1) surface for the V-shaped model, where division between them was observed; this confirmed that the sensitivity in the normal direction was higher than that in the lateral direction. The overall process of lateral response for a V- shaped probe was explored with the inclusion of a bump, and three generic stages were identified (i.e. first twisting, second bending, and finally slipping). The dynamic sensitivity was also explored by emulating a previously reported experiment, and the results showed good agreement in the response pattern. References [1] Veeco Website 2006 http://www.veeco.com [2] Bowen W R et al 2001 Atomic force microscope studies of stainless steel: surface morphology and colloidal particle adhesion J. Mater. Sci. 36 623 9 [3] Strijbos S 1976 Friction between a powder compact and a metal wall Sci. Ceram. 8 415 27 [4] Clifford C A and Seah M P 2005 The determination of atomic force microscope cantilever spring constants via dimensional methods for nanomechanical analysis Nanotechnology 16 1666 80 [5] Beaulieu L Y et al 2007 A complete analysis of the laser beam deflection systems used in cantilever-based systems Ultramicroscopy 107 422 30 [6] Song Y and Bhushan B 2006 Simulation of dynamic modes of atomic force microscopy using a 3D finite element model Ultramicroscopy 106 847 73 [7] Muller M et al 2006 Finite element analysis of V-shaped cantilevers for atomic force microscopy under normal and lateral force loads Surf. Interface Anal. 38 1090 5 [8] ELFEN User Manual 3.0 2001 Rockfield Software Limited [9] Ultra Hard Materials Website 2006 http://www. ultrahardmaterials.co.uk [10] Khan A and Philip J 2004 Young s modulus of silicon nitride used in scanning force microscope cantilevers J. Appl. Phys. 95 1667 72 [11] Properties and selection: nonferrous alloys and pure metals ASM Metals Handbook 1986 vol 2 (Metals Park, Ohio: ASM International) [12] Neumeister J M and Ducker W A 1994 Lateral, normal and longitudinal spring constants of atomic force microscopy cantilevers Rev. Sci. Instrum. 65 2527 31 [13] Green C P et al 2004 Normal and torsional spring constants of atomic force microscope cantilevers Rev. Sci. Instrum. 75 1988 96 [14] Liu Y et al 1994 Lateral force microscopy study on the shear properties of self-assembled monolayers of dialkylammonium surfactant on mica Langmuir 10 2241 5 13

[15] Noy A et al 1995 Chemical force microscopy: exploiting chemically-modified tips to quantify adhesion, friction, and functional group distributions in molecular assemblies J. Am. Chem. Soc. 117 7943 51 [16] Ogletree D F et al 1996 Calibration of frictional forces inatomic force microscopy Rev. Sci. Instrum. 67 3298 306 [17] Veeco Probes Website 2007 https://www.veecoprobes.com 14