Effect of sliding velocity on capillary condensation and friction force in a nanoscopic contact

Similar documents
The 2 / 3 Power Law Dependence of Capillary Force on Normal Load in Nanoscopic Friction

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

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

RHK Technology Brief

The velocity dependence of frictional forces in point-contact friction

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

Size dependence of the mechanical properties of ZnO nanobelts

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

Nano-Scale Effect in Adhesive Friction of Sliding Rough Surfaces

Material Anisotropy Revealed by Phase Contrast in Intermittent Contact Atomic Force Microscopy

Contents. Preface XI Symbols and Abbreviations XIII. 1 Introduction 1

Interfacial forces and friction on the nanometer scale: A tutorial

SUPPLEMENTARY INFORMATION

! Importance of Particle Adhesion! History of Particle Adhesion! Method of measurement of Adhesion! Adhesion Induced Deformation

ADHESION OF AN AXISYMMETRIC ELASTIC BODY: RANGES OF VALIDITY OF MONOMIAL APPROXIMATIONS AND A TRANSITION MODEL

Single asperity friction in the wear regime

Capillary adhesion and friction : an approach with the AFM Circular Mode

Scanning Force Microscopy

NIS: what can it be used for?

M. Enachescu Materials Sciences Division, Lawrence Berkeley National Laboratory, University of California, Berkeley, California 94720

Atomic Force Microscopy imaging and beyond

Nanotribology. Judith A. Harrison & Ginger M. Chateauneuf. Chemistry Department United States Naval Academy Annapolis, MD

Numerical modeling of sliding contact

NANOTRIBOLOGY - The road to no WEAR! by Richie Khandelwal (MT03B023) & Sahil Sahni (MT03B024)

Enrico Gnecco Department of Physics. University of Basel, Switzerland

Lecture Note October 1, 2009 Nanostructure characterization techniques

Introductory guide to measuring the mechanical properties of nanoobjects/particles

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

Critical phenomena of water bridges in nanoasperity contacts

SUPPLEMENTARY INFORMATION

Frictional characteristics of exfoliated and epitaxial graphene

Molecular dynamics simulations of sliding friction in a dense granular material

DYNAMIC CONDENSATION OF WATER AT CRACK TIPS IN FUSED SILICA GLASS

Multiscale roughness and modeling of MEMS interfaces

Atomic Force Microscopy Study of an Ideally Hard Contact: The Diamond(111)/Tungsten Carbide Interface

Contact, Adhesion and Rupture of Elastic Solids

Figure 43. Some common mechanical systems involving contact.

Thermally Activated Friction

Forces Acting on Particle

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

Lecture 4 Scanning Probe Microscopy (SPM)

Department of Engineering Mechanics, SVL, Xi an Jiaotong University, Xi an

AFM Studies of Pristine PCBM Changes Under Light Exposure. Erin Chambers

Supporting Information. Interfacial Shear Strength of Multilayer Graphene Oxide Films

Force-distance studies with piezoelectric tuning forks below 4.2K

Supplementary Note 1: Grain size of the graphene grown on SiO 2 /Si

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

Roughness picture of friction in dry nanoscale contacts

CONTACT USING FRICTION HYSTERESIS UDC Adrian Kraft, Roman Pohrt

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

Improving the accuracy of Atomic Force Microscope based nanomechanical measurements. Bede Pittenger Bruker Nano Surfaces, Santa Barbara, CA, USA

Vibration Studying of AFM Piezoelectric Microcantilever Subjected to Tip-Nanoparticle Interaction

Structural and Mechanical Properties of Nanostructures

Ch. 10: Fundamental of contact between solids

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

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

RP 2.7. SEG/Houston 2005 Annual Meeting 1525

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

A van der Waals Force-Based Adhesion Model for Micromanipulation

Modeling, Simulation and Optimization of the Mechanical Response of Micromechanical Silicon Cantilever: Application to Piezoresistive Force Sensor

Characterization of MEMS Devices

Van der Waals Interaction between Polymer Aggregates and Substrate Surface Analyzed by Atomic Force Microscope (AFM)

arxiv: v1 [cond-mat.mes-hall] 11 Aug 2017

Specific ion effects on the interaction of. hydrophobic and hydrophilic self assembled

Nano-tribological characteristics of TiO 2 films on 3-mercaptopropyl trimethoxysilane sulfonated self-assembled monolayer

A Finite Element Study of the Residual Stress and Deformation in Hemispherical Contacts

MOLECULAR DYNAMICS SIMULATIONS OF FRICTION FORCE VERSUS LOAD

UNLOADING OF AN ELASTIC-PLASTIC LOADED SPHERICAL CONTACT

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

Unloading of an elastic plastic loaded spherical contact

Local friction of rough contact interfaces with rubbers using contact imaging approaches mm

Influence of the Meniscus Force for Contact Recording Head Dynamics Over a Randomly Undulating Disk Surface

Atomic scale friction in the function of modified Eyring activation energies

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

Outline Scanning Probe Microscope (SPM)

ACOUSTIC EMISSION CHARACTERISTICS OF SURFACE FRICTION IN BIO-MEDICAL APPLICATION

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

Quantitative imaging of nanoscale mechanical properties using hybrid nanoindentation and force modulation

Scanning Tunneling Microscopy

ADHESION AND FRICTION FORCE MEASUREMENTS IN AMBIENT AND HIGH VACUUM CONDITIONS

The effect of liquid-induced adhesion changes on the interfacial shear strength between self-assembled monolayers

Is Partial Slip Under Transverse Oscillatory Loading a Transient Event?

ScienceDirect. A Numerical Model for Wet-Adhesive Line Contact

MECH 466. Micro Electromechanical Systems. Laboratory Manual Laboratory #3: Stiction of MEMS and Strength of MEMS Materials

Practical Method to Limit Tip Sample Contact Stress and Prevent Wear in Amplitude Modulation Atomic Force Microscopy

AME COMPUTATIONAL MULTIBODY DYNAMICS. Friction and Contact-Impact

Materials Reliability Division, National Institute of Standards and Technology, Boulder, CO 80305

Normal contact and friction of rubber with model randomly rough surfaces

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

MSE640: Advances in Investigation of Intermolecular & Surface Forces

Deterministic repeated contact of rough surfaces

A Short Course on Nanoscale Mechanical Characterization: The Theory and Practice of Contact Probe Techniques

Microscopic Friction Studies on Metal Surfaces

FEM-SIMULATIONS OF VIBRATIONS AND RESONANCES OF STIFF AFM CANTILEVERS

Princeton University. Adhsion and Interfacial Failure in Drug Eluting Stents

Speed and Atmosphere Influences on Nanotribological Properties of NbSe2

arxiv: v2 [cond-mat.soft] 2 Jun 2016

Scanning Tunneling Microscopy

Where does a cohesive granular heap break?

FRICTION FORCES ON THE LUBRICATED SURFACES IN MICRO AND NANO SCALE

Transcription:

Materials Science and Engineering C 26 (2006) 751 755 www.elsevier.com/locate/msec Effect of sliding velocity on capillary condensation and friction force in a nanoscopic contact Pierre-Emmanuel Mazeran * Laboratoire en Mécanique Roberval, Université de Technologie de Compiègne, BP 60319, 60206 Compiègne Cedex, France Available online 28 October 2005 Abstract The influence of sliding velocity on the friction force in a nanoscopic contact was studied on a hydrophilic silicon nitride surface with an Atomic Force Microscope. By submitting a nanoscopic contact to small oscillating lateral displacements, we observe both decreases of the friction force and of the contact radius with an increase of the sliding velocity. We found experimentally that this decrease of the friction force and of the contact radius is correlated to a decrease of the capillary force. D 2005 Elsevier B.V. All rights reserved. Keywords: Scanning force microscope; Sliding friction; Velocity; Capillary force 1. Introduction Friction is an everyday life issue in industrial processes and natural phenomena [1,2]. It is involved in a great variety of fields like for example, granular media, earthquake dynamics or Micro ElectroMechanical Systems (MEMS). Nevertheless, the fundamental understanding of friction at the nanometer scale is actually too poor in front of the scientific and technological stakes. From a technological point of view, the development of nanotechnologies leads to the miniaturization of moving components for which the surface interactions (adhesion, friction and wear) become of primary importance. The development of Atomic Force Microscopy (AFM) has opened new possibilities for investigations of friction, adhesion and wear at nanometer scale [3]. The sliding velocity dependence of friction has been studied leading to contradictory results: Logarithmic increases or decreases of friction with the sliding velocity have been reported [4 8]. Recently Riedo et al. [8] suggested that the logarithmic dependence of friction to sliding velocity was due to two opposite phenomena. The first one is due to stick slip motion, which gives a logarithmic increase of friction with an increasing sliding velocity [4,6]. The second one is due to the kinetics of capillary condensation of water vapor around the contact area, which causes a logarithmic decrease of the friction with an increasing sliding velocity [8]. 2. Capillary condensation When two surfaces are close together in the presence of a vapor under saturation, water (or any other liquid) may condense to form a liquid bridge between the two surfaces [9]. At the equilibrium, the surface curvature of the meniscus is imposed by the Kelvin radius R K : 1 ¼ 1 þ 1 R K R Min R Max where R Min and R Max are the minimum and maximum curvature radius of the meniscus. In the case of a meniscus between two parallel solid surfaces, the curvature radius is minimum (R Min <0) along the planes perpendicular to the solid surfaces and maximum (R Max >0) along the planes parallel to the solid surfaces (Fig. 1). When the meniscus increases in size, R Max tends to infinity and R Min tends to the Kelvin radius: The maximum height of this capillary bridge is then twice the Kelvin radius. At fixed temperature T, the Kelvin radius depends only of the relative humidity H: ð1þ * Tel.: +33 344 235 228; fax: +33 344 235 229. E-mail address: mazeran@utc.fr. 0928-4931/$ - see front matter D 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.msec.2005.09.055 R K ¼ c LVV L RT logh ð2þ

752 P.-E. Mazeran / Materials Science and Engineering C 26 (2006) 751 755 Fig. 1. Surface curvature of a capillary meniscus. A minimum curvature (R Min <0) and a maximum curvature radius (R Max >0) describe the surface curvature of the meniscus. The average curvature radius of the meniscus is equal to the Kelvin radius (R K ): R K 1 =R Min 1 +R Max 1. where c LV is the surface tension, V L is the molar volume and R is the universal gas constant. The difference in pressure Dp between the liquid and its vapor is then: Dp ¼ c LV R K : Fogden and White [10] and independently Maugis [11] have investigated the influence of the capillary force on an elastic sphere/sphere contact. They found that, for rigid materials, the depression inside the meniscus was too low to influence the geometry of the contact, which remains Hertzian [12]. The capillary force F C just acts as an extra load that adds to the applied load F A as suggested by Derjaguin et al. [13]. The total applied load F T is then the sum of the applied load F A, the capillary force F C and others forces like for example the van der Waals force F vdw. The capillary force F C is proportional to the wet area A W : c F C ¼ A W Dp ¼ A LV W : ð4þ R K For a contact between two flat spheres, the ratio A W /R K and then the capillary force is independent of the relative humidity and is equal to F C ¼ 4pR c LV cosh ð5þ where R is the radius of the sphere and h is the static water contact angle of the meniscus. This equation is misleading in a nanoscopic contact, because the surface (typically in the nanometer range) is generally high compared to both the Kelvin radius and the indentation depth. Few publications clearly show that the capillary force is highly dependent on the relative humidity. Typically, for hydrophilic rough surface, the capillary force is maximum for a relative humidity of 20 40% [14,15]. Capillary condensation is a thermally activated process. It has been shown that for a constant relative humidity, the cohesion force of a granular media, by means of the maximum stability angle, is time-dependent [16,17]. The maximum height of water meniscus is twice the Kelvin radius. The Kelvin radius is in the nanometer range, for a relative humidity of 30% the Kelvin radius is equal to 1 nm. Water can only condense in nanometer-scale interstices or around the contact points. The presence of asperities on the two surfaces in contact generates nucleating sites for capillary bridges. At a given rest time only an increasing fraction of them is formed: The capillary force, which is proportional to the wet area, increases ð3þ (Fig. 2). When sliding occurs, the residence time depends of the sliding velocity and then the wet area and the capillary force decreases with the sliding velocity [8]. Nevertheless, only little evidences of the decrease of the capillary force with the sliding velocity have been pointed out [15]. The aim of this paper is to show experimental evidences of the decrease of the capillary force with an increasing sliding velocity. We first investigated the nature of the contact. We show that in our experimental conditions, the roughness of the surface is small in front of the indentation depth and that the contact is close to a single asperity contact. Then, we investigated the friction dependence to velocity. Experimental results show that the friction force and the contact radius decrease with an increasing velocity. We demonstrate that the decreases are the result of a decrease of the capillary force. At last, we will show a direct evidence of the decrease of the capillary force with an increasing velocity. 3. Experiments Experiments were performed with a commercial Atomic Force Microscope in contact mode (Dimension 3100, Nanoscope IIIa from Digital Instruments). Silicon nitride probes were used (DNP, Digital Instruments), the nominal curvature radius of the probe is 20 nm and the nominal stiffness of the cantilever is 0.58 N/m. The probe was brought in contact with the surface of a silicon nitride cantilever chip. The probe and sample surfaces have the same chemical and mechanical properties. The mechanical properties of the silicon nitride were measured by nanoindentation (Nanoindentor XP, MTS Systems). The hardness is equal to 25 GPa and the elastic modulus is equal to 140 GPa in good agreement with the literature [18]. The surface is hydrophilic, the water angle is equal to 30-. Experiments were performed under lab conditions. The relative humidity (typically 50 60%) was measured Fig. 2. Schematic of capillary water bridges formed in the contact area. The asperities of the surface act as nucleating site for capillary condensation. When the rest time increases, the wet area and then the capillary force increases.

P.-E. Mazeran / Materials Science and Engineering C 26 (2006) 751 755 753 using a commercial hygrometer. The AFM was modified for the tribological studies [19]: a lock-in amplifier generates a sinusoidal voltage at a typical frequency of a few khz in order to generate a sinusoidal displacement of the probe parallel to the surface sample. The lateral force generated by this lateral displacement of the probe was analyzed by the lock-in amplifier and the in-phase and out-of-phase part of the lateral force was recorded. The calibration of the lateral displacement and the lateral force at these high frequencies is particularly difficult leading to a high uncertainty of the data. 3.1. Nature of the contact To explain theirs experimental results, several authors [8,20] suggest that the contact between the probe and the sample was a polyasperities contact rather than a single asperity contact (Fig. 1). Nevertheless, the determination of the exact nature of the contact (single asperity or polyasperities contact, elastic or plastic contact) is particularly difficult and depends highly of the sample roughness. As a first approach, we consider that the contact in our experimental conditions follows the Hertz Theory. This theory gives the relation between the applied load, contact radius, indentation depth and contact pressure for an elastic contact between a flat sphere and a flat plane [12]. The plasticity and roughness of the surface are two obstacles to the validity of the Hertz theory. The theoretical average contact pressure was computed to be about 3 GPa. This value is small in front of the hardness of the sample (25 GPa) and we can conclude that the contact is theoretically elastic. To investigate the influence of the roughness, we propose to compare the roughness of the theoretical contact area to the theoretical indentation depth. If the theoretical indentation depth is high compared to the total roughness, we can conclude that the asperities are flattened out and that the contact remains Hertzian. On the other hand, if the theoretical indentation Fig. 4. 1 Am 2 scale topographic image of the silicon nitride sample (Z scale: 3 nm). depth is small compared to the total roughness, we can conclude that the asperities are not flattened out and that the contact is a polyasperities contact (Fig. 3). In a first time, we realized a topographic image of the silicon nitride sample (Fig. 4) using the so-called Tapping mode and a sharp silicon probe (MPP 11100 from Digital instruments). The nominal radius of the probe is 10 nm. In a second time we computed the peak-to-valley roughness for the whole image and decreasing fractions of it (quarter, 16th, 64th...). The peak-to-valley roughness of the fractions of the image is plotted in Fig. 5 as a function of their area. Using the Hertz theory, the indentation depth d = A /pr is plotted on the same graph as a function of the contact area A. The graph shows that the theoretical indentation depth is superior to the roughness of the contact area for a contact area of about 12 nm 2. From Hertz, we can estimate the load corresponding to this area at approximately 60 nn. This force is of the same order than the total applied load (typically 50 nn). Because the roughness in the contact area is a statistical event, we conclude that the Fig. 3. Schematic of the transition from a polyasperities contact to a single asperity contact. As the applied load increases, the contact pressure flattens the asperities: The contact follows the Hertz theory. Fig. 5. Comparison between the peak-to-valley roughness of the silicon nitride sample and the theoretical indentation depth as a function of the contact area. We observe that the indentation depth is high compared to the peak-to-valley roughness. The asperities are flattened out and the contact may obey the Hertz law.

754 P.-E. Mazeran / Materials Science and Engineering C 26 (2006) 751 755 Lateral force (nn) 18 15 12 9 6 Stick domain Sliding domain Theoretical in-phase 3 Theoretical out-of-phase Experimental in-phase Experimental out-of-phase 0 0 3 6 9 Displacement amplitude (nm) Fig. 6. Comparison between the experimental and the theoretical in-phase and out-of-phase part of the lateral force as the function of the displacement amplitude (velocity). At high displacement amplitude, we observe a decrease of both the experimental in-phase and out-of-phase component of the lateral force compared to the theoretical values. This result is interpreted as a decrease of the attractive force. contact is mainly Hertzian but that the presence of asperities could lead locally to a poly asperities contact. 3.2. Friction dependence of the velocity We investigated the friction dependence to the sliding velocity using the following set-up. For a constant applied load (10 nn), we generated a sinusoidal lateral displacement d of this single asperity contact at a frequency of 6135 Hz. This frequency corresponds to a resonance of the AFM piezo scanner. The amplitude of the lateral displacement increased linearly with time from zero to a few tens of nanometer and then decreases linearly to zero during a period of about 2 min. The lateral force generated by the lateral displacement was analyzed by the lock-in amplifier and the in-phase and out-ofphase part of the lateral force is plotted as a function of the amplitude of the lateral displacement (Fig. 6). The in-phase part of the lateral force is proportional to the ratio between the energy stored during one cycle and the displacement length. This energy is principally due to shear of the contact. The outof-phase part of the lateral force is proportional to the ratio of the energy dissipated to the displacement length. This energy is principally dissipated by friction. We used the Mindlin theory [21,22] as a model for our experiments. This theory describes the increase of the lateral force as a function of the relative displacement of a contact. This model is based on a Hertz contact with a linear dependence of the friction force to the total applied load. In our experimental conditions, the total applied load is the sum of the applied load, capillary force and van der Waals force. While the amplitude of the lateral displacement is low in front of the contact radius, the contact is principally sheared and sliding occurs only in the periphery of the contact: We are in the stick domain. The in-phase part of the lateral force is then far higher than the out-of-phase part of the lateral force. When the amplitude of the lateral displacement becomes high compared to the contact diameter, the energy is mainly dissipated by fiction: We are in the sliding domain. The out-of-phase part of the lateral force becomes higher than the in-phase part of the lateral force. Experimental results were fitted with the Mindlin model by adjusting two parameters: the contact radius and the friction force (Fig. 6). For lateral displacement amplitudes below approximately 5 nm, A very good agreement is observed between theoretical and experimental results. Beyond this value, the in-phase and out-of-phase contribution to the lateral force become quite lower than the theoretical values. This result is an evidence of the decreases of the elastic energy stored in the contact and the energy dissipated by friction. In the Mindlin model, the elastic energy stored in the contact as well as the energy dissipated by friction are proportional to the total applied load. It becomes then obvious that the total applied load decreases as the lateral displacement amplitude increases. There is a direct relation between the displacement amplitude and the maximum velocity V=2pfd during an experiment. We conclude that the total applied load decreases as the sliding velocity increases. The applied load was maintained constant during the experiment, then we may conclude that this is the attractive force, the sum of the capillary and the van der Waals force, that decreased during the experiment. 3.3. Adhesion force dependence on friction To confirm the previous experiments, we measured the adhesion force as a function of the sliding velocity of the probe. A sinusoidal lateral displacement of the probe was generated at a fixed frequency f with an amplitude d. The normal applied force is plotted as a function of the probe/surface distance in Fig. 7. While the total applied load (sum of the normal applied load and attractive force) is greater than zero, the probe is in contact with the surface. When the probe snaps off the surface, the absolute value of the applied load is equal to the attractive force. We define this force as the adhesive force, sum of capillary and van der Waals when the probe leaves the surface. Fig. 7. Force as a function of the probe/sample distance for an increasing sliding velocity. We observe a decrease of the snap-off force (marked by arrows) with an increase of the velocity. This decrease is the result of the decrease of the capillary force.

P.-E. Mazeran / Materials Science and Engineering C 26 (2006) 751 755 755 We observed a decrease of the adhesive force with an increase of the maximum sliding velocity. For a high value of the velocity, the force/distance curve looks like a Lennard Jones force curve. We conclude that for an enough high sliding velocity of the probe, the capillary force vanishes with the absence of capillary condensation. The adhesive force recorded is then only due to a van der Waal interaction. 4. Conclusion In conclusion, we showed that for a nanoscopic contact between hydrophilic surfaces, both the friction force and the contact radius decreases with the sliding velocity. These decreases have been explained experimentally by a decrease of the adhesive force with the sliding velocity due to the disappearance of the capillary force. References [1] E. Meyer, R.M. Overney, K. Dransfeld, T. Gyalog, Nanoscience: Friction and Rheology on the Nanometer Scale, World Scientific, Singapore, 1998. [2] B. Bhushan, Micro/Nanotribology and its Applications, Kluwer Academic Publishers, Dordrecht, 1997. [3] R.W. Carpick, M. Salmeron, Chem. Rev. 97 (1997) 1163. [4] T. Bouhacina, J.P. Aimé, S. Gauthier, D. Michel, V. Heroguez, Phys. Rev., B 56 (1997) 7694. [5] O. Zwörner, H. Holscher, U.D. Schwarz, R. Wiesendanger, Appl. Phys., A 66 (1998) S263. [6] E. Gnecco, R. Bennewitz, T. Gyalog, C. Loppacher, M. Bammerlin, E. Meyer, J. Güntherodt, Phys. Rev. Lett. 84 (2000) 1172. [7] H. Liu, S.I.U. Ahmed, M. Scherge, Thin Solid Films 381 (2001) 135. [8] E. Riedo, F. Lévy, H. Brune, Phys. Rev. Lett. 18 (2002) 185505. [9] J. Israelachvili, Intermolecular and Surface Forces, Academic Press, San Diego, 1977. [10] A. Fogden, L.R. White, J. Colloid Interface Sci. 138 (1990) 414. [11] D. Maugis, J. Colloid Interface Sci. 150 (1992) 243. [12] H. Hertz, J. Reine Angew. Math. 92 (1882) 156. [13] B.V. Derjaguin, V.M. Muller, Y.P. Toporov, J. Colloid Interface Sci. 53 (1975) 314. [14] M. He, A. Szuchmacher Blum, E. Aston, C. Buenviaje, R. Overney, J. Chem. Phys. 14 (2001) 1355. [15] L. Sirghi, Appl. Phys. Lett. 82 (2003) 3755. [16] F. Restagno, C. Ursini, H. Gayvallet, E. Charlaix, Phys. Rev., E 66 (2002) 021304. [17] L. Bocquet, E. Charlaix, S. Ciliberto, J. Crassous, Nature 396 (1998) 735. [18] J.E. Sader, I. Larson, P. Mulvaney, L.R. White, Rev. Sci. Instrum. 66 (1995) 3789. [19] P.E. Mazeran, J.L. Loubet, Tribol. Lett. 7 (1999) 199. [20] N.S. Tambe, B. Bhushan, J. Phys., D 38 (2005) 764. [21] R.D. Mindlin, Trans. ASME, Ser. E, J. Appl. Mech. 16 (1949) 259. [22] R.D. Mindlin, W.P. Mason, F. Osmer, H. Deresiewicz, Proc. 1st US National Congress of Applied Mechanics, New York, United States, 1954, p. 203.