6
A General Equation for Fitting Contact Area and Friction vs Load Measurements Robert W. Carpick, 1 D. Frank Ogletree, and Miquel Salmeron Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720 Received August 31, 1998, accepted November 30, 1998 The variation of contact area with load between adhesive elastic spheres depends upon the effective range of attractive surface forces. Relatively simple forms to describe limiting cases exist, but the gen- eral intermediate case requires a more complex analysis. Maugis, using a Dugdale model [D. Maugis, J. Colloid. Interf. Sci. 150, 243 (1992)], provides an analytic solution, but the resulting equations are cumbersome if one wishes to compare with experimental data such as atomic force microscope measurements. In this paper we present a simpler general equation that approximates Maugis’ solution ex- tremely closely. The general equation is amenable to conventional curve fitting software routines and provides a rapid method of deter- mining the value of the “transition parameter” which describes the range of surface forces. © 1999 Academic Press Key Words: contact mechanics; contact area; adhesion; Dugdale model; adhesive spheres; atomic force microscope; friction force microscope; surface forces apparatus. 1. INTRODUCTION An understanding of contact, adhesion, and friction between surfaces requires knowledge of the area of contact between them. Continuum models which predict the contact area for various geometries have been worked out starting with the pioneering work of Hertz (1). Experimental techniques that measure contact area for elastic single asperities, namely the surface forces appa- ratus (SFA) (2, 3) and the atomic force microscope (AFM) (4), can be described by the geometry of contacting spheres (approx- imated as paraboloids). In the case of the AFM, one sphere represents the sample, with an infinite radius of curvature, while the other sphere represents the tip, which in many cases conforms to a paraboloidal shape (5). In the absence of adhesion, the Hertz model has been shown to accurately describe the contact area between elastic spheres (6). However, at small scales the surface- to-bulk ratio becomes significant. Therefore, adhesion arising from attractive surface forces is generally not negligible and must be included in any description of contact area. The spatial range over which surface forces act depends upon the chemistry of the materials in contact, and may or may not be long range compared to the scale of elastic deformations due to these forces (7) (Fig. 1). Two limiting cases are appar- ent. When the surface forces are short range in comparison to the elastic deformations they cause (i.e. compliant materials, strong adhesion forces, large tip radii), the contact area is described by the Johnson–Kendall–Roberts (JKR) model (8). The opposite limit (i.e., stiff materials, weak adhesion forces, small tip radii) is referred to as the Derjaguin–Mu ¨ller–Toporov (DMT) regime (9) and the form of the contact area is presented in the work of Maugis 2 (10). It is convenient to utilize a nondimensional physical parameter to quantify these limits and cases in between. Often referred to as Tabor’s parameter m, this transition parameter is defined as (11) m 5 S 16Rg 2 9K 2 z 0 3 D 1/3 , [1] where z 0 is the equilibrium separation of the surfaces, R is the tip curvature radius, g is the interfacial energy per unit area (work of adhesion), and K is the combined elastic modulus of tip and sample, given by K 5 4/3 z ((1 2 n 1 2 )/ E 1 1 (1 2 n 2 2 )/ E 2 ) 21 where E 1 and E 2 are the tip and sample Young’s moduli, respectively, and n 1 and n 2 are the tip and sample Poisson ratios, respectively. The quantity m is in fact equal to the ratio of the elastic deformation just before the surfaces separate to the equilibrium separation z 0 . To approximate an actual interaction potential like that de- picted in Fig. 1, Maugis (10) considers a “Dugdale” (square well) potential to describe attractive forces between contacting spheres (Fig. 2), where a constant adhesive stress s 0 acts over a range d t . Thus, the work of adhesion is g 5 s 0 z d t . Maugis defines a parameter, l, which is similar to m, given by l 5 2s 0S R pgK 2D 1/3 . [2] By choosing s 0 to match the minimum adhesive stress of a Lennard–Jones potential (with equilibrium separation z 0 ), it follows that d t 5 0.97z 0 , and so l 5 1.1570m. Thus, l and m are roughly equivalent. For convenience we shall refer to l in this paper as the “transition parameter.” If l . 5, the JKR 1 Present address: Mailstop 1413, Sandia National Laboratories, Albuquer- que, NM 87185-1413. 2 As discussed by Greenwood (10), it is more appropriate to refer to the DMT regime as the “Bradley limit,” which is a model for completely rigid spheres. Here we refer to “DMT” in a conventional sense, as with Maugis (9). Journal of Colloid and Interface Science 211, 395– 400 (1999) Article ID jcis.1998.6027, available online at http://www.idealibrary.com on 395 0021-9797/99 $30.00 Copyright © 1999 by Academic Press All rights of reproduction in any form reserved.

Carpick_JColIntSci_1999!!!

Embed Size (px)

Citation preview

Page 1: Carpick_JColIntSci_1999!!!

sReu(castcmr

mm

sCgwarcirttmbtfb

und

q

Journal of Colloid and Interface Science211,395–400 (1999)Article ID jcis.1998.6027, available online at http://www.idealibrary.com on

A General Equation for Fitting Contact Areaand Friction vs Load MeasurementsRobert W. Carpick,1 D. Frank Ogletree, and Miquel Salmeron

Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720

Received August 31, 1998, accepted November 30, 1998

eetheiouerinntaap(4)prheh

forHet arfaisinmu

enmatiopa

e on tot rials,s a isd l (8).T ces,s( ntedi an andct

wt ea( oftn ’sm lePt acess

de-p arew ctings rad

B f aLfat

quo the

D ids s (9).

The variation of contact area with load between adhesive elasticpheres depends upon the effective range of attractive surface forces.elatively simple forms to describe limiting cases exist, but the gen-

ral intermediate case requires a more complex analysis. Maugis,sing a Dugdale model [D. Maugis, J. Colloid. Interf. Sci. 150, 2431992)], provides an analytic solution, but the resulting equations areumbersome if one wishes to compare with experimental data such astomic force microscope measurements. In this paper we present aimpler general equation that approximates Maugis’ solution ex-remely closely. The general equation is amenable to conventionalurve fitting software routines and provides a rapid method of deter-ining the value of the “transition parameter” which describes the

ange of surface forces. © 1999 Academic Press

Key Words: contact mechanics; contact area; adhesion; Dugdaleodel; adhesive spheres; atomic force microscope; friction forceicroscope; surface forces apparatus.

1. INTRODUCTION

An understanding of contact, adhesion, and friction betwurfaces requires knowledge of the area of contact betweenontinuum models which predict the contact area for vareometries have been worked out starting with the pioneork of Hertz (1). Experimental techniques that measure corea for elastic single asperities, namely the surface forcesatus (SFA) (2, 3) and the atomic force microscope (AFM)an be described by the geometry of contacting spheres (apmated as paraboloids). In the case of the AFM, one spepresents the sample, with an infinite radius of curvature, whe other sphere represents the tip, which in many cases cono a paraboloidal shape (5). In the absence of adhesion, theodel has been shown to accurately describe the contacetween elastic spheres (6). However, at small scales the su

o-bulk ratio becomes significant. Therefore, adhesion arrom attractive surface forces is generally not negligible ande included in any description of contact area.The spatial range over which surface forces act dep

pon the chemistry of the materials in contact, and may orot be long range compared to the scale of elastic deformaue to these forces (7) (Fig. 1). Two limiting cases are ap

1 Present address: Mailstop 1413, Sandia National Laboratories, Albuue, NM 87185-1413.

395

nm.sgctpa-,ox-reilemsrtzreace-gst

dsynsr-

nt. When the surface forces are short range in comparishe elastic deformations they cause (i.e. compliant matetrong adhesion forces, large tip radii), the contact areescribed by the Johnson–Kendall–Roberts (JKR) modehe opposite limit (i.e., stiff materials, weak adhesion formall tip radii) is referred to as the Derjaguin–Mu¨ller–ToporovDMT) regime (9) and the form of the contact area is presen the work of Maugis2 (10). It is convenient to utilizeondimensional physical parameter to quantify these limitsases in between. Often referred to as Tabor’s parameterm, thisransition parameter is defined as (11)

m 5 S16Rg2

9K2z03D 1/3

, [1]

herez0 is the equilibrium separation of the surfaces,R is theip curvature radius,g is the interfacial energy per unit arwork of adhesion), andK is the combined elastic modulusip and sample, given byK 5 4/3 z ((1 2 n1

2)/E1 1 (1 2

22)/E2)21 whereE1 and E2 are the tip and sample Youngoduli, respectively, andn1 and n2 are the tip and sampoisson ratios, respectively. The quantitym is in fact equal to

he ratio of the elastic deformation just before the surfeparate to the equilibrium separationz0.To approximate an actual interaction potential like that

icted in Fig. 1, Maugis (10) considers a “Dugdale” (squell) potential to describe attractive forces between contapheres (Fig. 2), where a constant adhesive stresss0 acts overangedt. Thus, the work of adhesion isg 5 s0 z dt. Maugis

efines a parameter,l, which is similar tom, given by

l 5 2s0S R

pgK2D 1/3

. [2]

y choosings0 to match the minimum adhesive stress oennard–Jones potential (with equilibrium separationz0), it

ollows thatdt 5 0.97z0, and sol 5 1.1570m. Thus,l andmre roughly equivalent. For convenience we shall refer tol in

his paper as the “transition parameter.” Ifl . 5, the JKR

er-

2 As discussed by Greenwood (10), it is more appropriate to refer tMT regime as the “Bradley limit,” which is a model for completely rigpheres. Here we refer to “DMT” in a conventional sense, as with Maugi

0021-9797/99 $30.00Copyright © 1999 by Academic Press

All rights of reproduction in any form reserved.

Page 2: Carpick_JColIntSci_1999!!!

m esb bt nsu byG ea

fa era ativ“ ullea f ths AFcf

T diusa

a qua-t

Sg

d tof bothc

MTm in tH udest ntsl aaf maH ora wec

bM es Thw

d inM nda it0 e for ag

396 CARPICK, OGLETREE, AND SALMERON

odel applies, and ifl , 0.1, the DMT model applies. Valuetween 0.1 and 5 correspond to the “transition regime”

ween JKR and DMT. A summary of different conventiosed for defining this “transition parameter” is providedreenwood (11). The Hertz model applies when there arttractive surface forces (g 5 0).

2. DEPENDENCE OF CONTACT AREA UPON LOADFOR THE JKR AND DMT REGIMES

The variation of contact area with load for various values ol,nd for the Hertz model, is plotted in Fig. 3. In the non-zdhesion cases, there is a well-defined “pull-off force” or negcritical load” at which the surfaces separate when being ppart. This applies for the case of “fixed loads” where one opheres is attached to a compliant spring (as in the case of anantilever). We shall refer to this negative critical load asLc, andor the two limiting cases it is given by

Lc(JKR) 5 23

2pgR [3a]

Lc(DMT) 5 22pgR. [3b]

FIG. 1. Interaction forces (per unit area) for the Hertz, JKR, and Dodels, compared to a realistic interaction. There is no attractive forceertz model, only hard wall repulsion at contact. The JKR model inclhort range adhesion which is essentially a delta function with strengthg andhus acts only within the contact zone. The DMT curve shown represeong-range surface force. A volume integrated force, like the van der Worce, can also lead to a DMT dependence, where the contact profile reertzian and the attractive forces act like an additional external load. Fctual interaction force, the integral of the force–distance attractiveorresponds to the work of adhesion,g.

FIG. 2. The force–distance relation for the Dugdale model usedaugis. A constant adhesive stress (force per unit area)s0 acts between th

urfaces over a rangedt. At greater separations, the attractive force is zero.ork of adhesion is thusg 5 s0dt.

e-

no

oedeM

he models also predict particular values for the contact rat zero loada0, given by

a0(JKR) 5 S6pgR2

K D 1/3

[4a]

a0(DMT) 5 S2pgR2

K D 1/3

. [4b]

The variation of contact radiusa with load L for the JKRnd DMT cases are each described by relatively simple e

ions:

a

a0(JKR)5 S1 1 Î1 2 L/Lc(JKR)

2 D 2/3

[5a]

a

a0(DMT)5 ~1 2 L/Lc(DMT)!

1/3. [5b]

ince the tip is axially symmetric, the contact areaA is simplyiven by

A 5 pa2. [6]

It is apparent that Eqs. [5a] and [5b] can be generalizeorm an equation which describes the contact radius forases,

a

a0~a!5 Sa 1 Î1 2 L/Lc(a)

1 1 aD 2/3

, [7]

hes

al’sinsanll

y

e

FIG. 3. The Hertz area–load curve, and the JKR–DMT transition, plotteaugis’ units (Eq. [9]). Area–load curves for the JKR limit, the DMT limit, an intermediate case are shown. These approach the Hertz curve in the limg3(no adhesion). Adhesion increases the contact area from the Hertz cas

iven load by an amount dependent upon the range of attractive forces.

Page 3: Carpick_JColIntSci_1999!!!

wc thf q.[ thf ne itior ngt

estm ) asf ion(

l

w

a tac

r hes cts).M .

ckos Eqs.[ tsw fff e ofl ntalm if noti thatu

eso thev t fit.E -t ral-i 1%( gime.F t fort itione Theo eds lizedtT angee t lowt osts sm

itioeg ,w

D ·0000000000000

J ·

397GENERAL CONTACT AREA EQUATION

herea 5 1 corresponds exactly to the JKR case, anda 5 0orresponds exactly to the DMT case. Included in Eq. [7] isact thatLc anda0 depend ona as well. We shall refer to E7] as the generalized transition equation. We now showor intermediate cases (0, a , 1), the generalized transitioquation corresponds very closely to solutions for the transegime (0.1, l , 5) elegantly worked out by Maugis usihe Dugdale model (10).

3. DEPENDENCE OF CONTACT AREA UPON LOADFOR THE MAUGIS–DUGDALE MODEL

Utilizing the Dugdale model is a reasonable method forating the value of the contact radius (and other quantities

unction of load. The solution is referred to as the “MD” solutMaugis–Dugdale). Two equations are needed to relatea andL,

a2

2 F Îm2 2 1 1 ~m2 2 2!cos21S 1

mDG1

4l2a

3 F Îm2 2 1 cos21S 1

mD 2 m 1 1G 5 1 [8a]

L 5 a3 2 la2F Îm2 2 1 1 m2cos21S 1

mDG , [8b]

hereL and a are simple parameterizations ofL anda,

L 5L

pgR[9a]

a 5 a z S K

pgR2D 1/3

, [9b]

nd the parameterm represents the ratio between the con

FIG. 4. The MD solution (heavy black lines) and the generalized transquation, Eq. [7] (thin white lines), compared for four different values ofl. Theeneralized transition equation is plotted for values ofa above the critical loadhile the MD solution includes the unstable lower branch of solutions.

e

at

n

i-a

t

adius a and an outer radiusc at which the gap between turfaces reachesdt (i.e., where the adhesive stress no longer aaugis’ equations properly predict the JKR and DMT limitsThe difficulty in utilizing the MD equations lies in the la

f a single expression relating onlya andL. To plot the MDolution or fit it to data, one needs to simultaneously solve8a] and [8b] by lettingm vary appropriately between limihich depend uponl. Furthermore, the relation for the pull-o

orce must be determined through iteration (12) if the valuis not known a priori (the usual case with experimeeasurements). In practice, this is rather cumbersome

mpossible to carry out with common software programstilize automated statistical fitting procedures.

4. COMPARISON OF THE MD SOLUTION AND THEGENERALIZED TRANSITION EQUATION

In Fig. 4, we plot the MD solution for four different valuf l. The generalized transition equation is overlaid, wherealue ofa in each case was optimized to provide the besach pair of plots share identical values ofa0 andLc, respec

ively. For all values ofl, the difference between the genezed transition equation and the MD solution is less thanand less than 0.1% in most cases) for the relevant load reurthermore, the generalized transition equation is exac

he JKR and DMT limits. Therefore, the generalized transquation is an excellent approximation to the MD solution.ptimization ofa was performed by minimizing the integratquare deviation between the MD solution and the generaransition equation for the load range2uLc(a)u to 12uLc(a)u.his load range was chosen because it is typical of the rmployed in AFM measurements (4), where loads are kep

o avoid wear. The low load regime also exhibits the mubstantial curvature of thea vs Lrelationship in general, thuaking fitting more reliable.

n

TABLE 1Conversion Table between a and l and the Associated Values

of Lc(l) and a0(l)

a l Lc(l) a0(l)

MT: 0 0 22 Î3 2 5 1.260· ·.074 0.1 21.951 1.336.158 0.2 21.881 1.394.256 0.4 21.816 1.442.433 0.5 21.718 1.517.609 0.8 21.634 1.598.692 1.0 21.601 1.636.817 1.5 21.556 1.700.886 2.0 21.535 1.738.922 2.5 21.523 1.760.944 3.0 21.517 1.775.958 3.5 21.513 1.785.967 4.0 21.510 1.792.979 5.0 21.506 1.800

KR: 1 ` 23/2 Î3 6 5 1.817· ·

Page 4: Carpick_JColIntSci_1999!!!

so topg

F mi( nsar n;n asr F5 ty il tialu itioe ]pI ntrd Ms

nau

T y tva tio

t rb ely:

E at-i ter-m eth-o

forc

a ust

alc

dcJ

s)aD

398 CARPICK, OGLETREE, AND SALMERON

Table 1 displays the optimized values ofa for various valuef l. An empirical formula was fit to this table of valuesrovide a conversion equation froma to l (Fig. 5), which isiven by

l 5 20.924z ln~1 2 1.02a!. [10]

ora . 1.0221 5 0.980, one can simply assume the JKR lia 5 1). The steep slope asa approaches the JKR limit mealarger uncertainty in determiningl from a value ofa in this

ange. Essentially this applies for values ofa greater tha0.95 or equivalently for values ofl greater than;3. This isot problematic since the change in the contact radiuslanges from 3 and up is reasonably small, apparent froma of Maugis (10). In other words, a substantial uncertainin this range doesnot lead to a correspondingly substan

ncertainty in the contact radius. The generalized transquation is therefore reliable for all values ofl, and Eq. [10rovides a reasonable and simple conversion froma to l.

mportantly, the generalized transition equation does not iuce any additional free parameters compared to theolution.Lc and a0 can be represented in Maugis’ nondimensio

nits given in Eqs. [9a] and [9b] as follows:

Lc~l! 5Lc

pgR[11a]

a0~l! 5 a0 z S K

pgR2D 1/3

. [11b]

hese nondimensional values are uniquely determined balue ofl and are listed in Table 1 and plotted vsl in Figs. 6nd 7, respectively. We have determined empirical equa

FIG. 5. Plot of l vs a from Table I (filled circles) and the empiriconversion equation, Eq. [10] (solid line), fit to the data.

t

ig.n

n

o-D

l

he

ns

hat fit these values reasonably well (within;1% accuracy oetter), which are also plotted in Figs. 6 and 7, respectiv

Lc~l! 5 27

41

1

4z S4.04 z l1.4 2 1

4.04 z l1.4 1 1D [12a]

a0~l! 5 1.541 0.279z S2.28 z l1.3 2 1

2.28 z l1.3 1 1D . [12b]

qs. [12a] and [12b] provide fast and simple ways of estimng Lc(l) and a0(l). These quantities can instead be de

ined exactly from the MD equations using numerical mds if preferred.

5. FITTING PROCEDURE

The above discussion leads to the following procedureurve fitting:

(1) Obtain a measurement of contact radiusa or friction Ff

s a function of load. For friction measurements, one m

FIG. 6. Plot of Lc(l) vs l determined from the MD solution (filleircles), and an empirical fit, Eq. [12a] (solid line). The values ofLc(l) for theKR and DMT limits are indicated by the dotted lines.

FIG. 7. Plot of a0(l) vsl determined from the MD solution (solid circlend an empirical fit, Eq. [12b] (solid line). The values ofa0(l) for the JKR andMT limits are indicated by the dotted lines.

Page 5: Carpick_JColIntSci_1999!!!

a are( ars ipla 6)t s ow .

vaia lizetf tfi cec ret ins asm a-t

thvf

seEvtd nen

thc thvf ] tod fg sit

term eg yc luef

tioe umd . Ig odc hisr mis tsr ithA o vi idap eare sibw

s sr tac

a stics nifi-cc lateralf e hasb com-p tionc . Ing n arem

cel-l tocm eL mer-i on-t ums re tod dius.F to thep n ac gapb tesi-m thec FMc fewnt nnota ato-m it isp t ana ngths ativeb blesG toc del,J ationf orG itione

tionw s af thers.T n andt ane erfv , andw lizedt helpi de-s ando

399GENERAL CONTACT AREA EQUATION

ssume that friction is directly proportional to the contactFf 5 t z pa2, where t is the constant interfacial shetrength) (4, 5). This is not expected to be true for multsperity contacts, or in the presence of wear (see Section

he case of friction measurements, replace all occurrenceaith =Ff in the relevant equations and discussion below(2) Use a curve fit optimization routine such as those a

ble in commercial software programs to fit the generaransition equation, Eq. [7], to the data, lettinga0, Lc, anda beree parameters whose values will then be extracted fromt. Alternately, measureLc independently from force–distanurves (4) and constrainLc to this value for the fit. Be awahat occasionally the last data point before separationliding friction measurement where the load is being decreay not correspond toLc if the tip and sample have prem

urely separated.(3) Use the above conversion equation, Eq. [10], or

alues in Table 1, to determinel from the value ofa extractedrom fitting the generalized transition equation.

(4) The interfacial energy,g, can now be determined. Uq. [12a] to solve forLc(l) now thatl is known. Plug thisalue into Eq. [11a] along with the value ofLc extracted fromhe generalized transition equation fit to solve forg (this can beone only if the tip radiusR is known). This can also be doumerically using the MD equations.(5) If measuring contact radius, the elastic modulus of

ontactK can be determined. Use Eq. [12b] to determinealue of a0(l). Plug this value and the value ofa0 extractedrom the generalized transition equation fit into Eq. [11betermineK. This requires knowingR and using the value odetermined above. This can also be done numerically u

he MD equations.(6) If measuring friction instead of contact radius, deine t by usingt 5 F0/pa0

2, whereF0 is extracted from theneralized transition equation fit, anda0 is determined bombining Eqs. [12b] and [11b]. This requires knowing vaor R andK.

6. DISCUSSION

One must be careful in utilizing the generalized transiquation for the case of friction measurements since it assirect proportionality between friction and contact areaeneral, SFA, AFM or other measurements where this man be applied require elastic, single-asperity contacts. Teadily determined in the case of SFA, where the smoothheets ensure a single contact, and optical measuremeneveal if wear is occurring (13). More care is required wFM measurements. Tip shapes must be characterized t

fy that they consist of a single, well-defined parabolorotrusion (4, 5). Loads must be kept low to avoid wspecially since atomic-scale wear may not be readily viith standard AFM imaging capabilities (4).The study of nanometer-scale single asperity contacts i

elatively new, and it is not certain that friction and con

a

e. Inf

l-d

he

aed

e

ee

ng

-

s

nes

nelis

cacan

er-l,le

tillt

rea will always be directly proportional even in the elaingle asperity regime. For example, there could be a sigant pressure dependence of the shear strengtht, or a signifi-ant change in the contact area due to the presence oforces, as opposed to purely normal forces. The latter issueen discussed in further detail by Johnson (14). Facilearison of friction data with the generalized transition equaould help to elucidate deviations due to these effectseneral, experiments where both contact area and frictioeasured are ideal for such purposes (12, 15, 16).Although the generalized transition equation is an ex

ent approximation for the MD solution, it is importantonsider that the MD solution is exact for anidealizedodel potential.A more physically realistic potential is thennard–Jones potential, but this can be solved only nu

cally. Greenwood has performed this calculation for cacting spheres (11). At this point the limits of continuolutions become apparent. It is not clear exactly wheefine the edge of the contact zone, i.e., the contact raor real materials, the contact zone edge correspondsoint where interfacial bonds are no longer formed. Iontinuum model, one can define it to be where theetween the continuum surfaces first displays an infinial increase, or possibly somewhere further out from

enter of the contact zone. For parameters in a typical Aontact, one may need to span radially outward aanometers to find a change in the gap of only 1 Å (i.e., less

han an atomic bond length). Continuum mechanics caddress such a question in an exact fashion when theistic nature of the materials must dominate. Therefore,ossible that the MD solution for the contact radius is noccurate prediction for real materials at nanometer lecales. Johnson, for example, has proposed an alternased on the MD model (12, 14) which closely resemreenwood’s result. The only way to test all of this isompare experimental measurements with the MD moohnson’s, and others. The generalized transition equacilitates this. A simple analytic equation for Johnson’sreenwood’s models analogous to the generalized transquation would be useful.In summary, we have provided a simple analytic equahich can be used to fit the contact radius (or friction) a

unction of load to scanning probe measurements and ohe difference between this generalized transition equatio

he MD solution is insignificant. A table of values andmpirical equation to determine the transition parametl

rom the generalized transition equation parametera is pro-ided. This approach requires a parabolic tip, a flat sampleearless elastic interactions. Application of the genera

ransition equation to friction or contact radius data candentify the accuracy of particular continuum models tocribe the contact radius for nanometer scale contactsthers.

Page 6: Carpick_JColIntSci_1999!!!

ffico tmeo s ts ana

.

.

.

A

111

1 .

11 D. F.,

1

400 CARPICK, OGLETREE, AND SALMERON

ACKNOWLEDGMENTS

This work was supported by the Director, Office of Energy Research, Of Basic Energy Sciences, Materials Sciences Division, of the U.S. Deparf Energy under Contract DE-AC03-76SF00098. R.W.C. acknowledgeupport of the Natural Sciences and Engineering Research Council of C

REFERENCES

1. Hertz, H.,J. Reine Angew. Math.92, 156 (1881).2. Israelachvili, J. N., and Tabor, D.,Proc. R. Soc. London A331,19 (1972)3. Tabor, D., and Winterton, R. H. S.,Proc. R. Soc. London A312, 435

(1969).4. Carpick, R. W., and Salmeron, M.,Chem. Rev.97, 1163 (1997).5. Carpick, R. W., Agraı¨t, N., Ogletree, D. F., and Salmeron, M.,J. Vac. Sci

Technol. B14, 1289 (1996).6. Johnson, K. L., “Contact Mechanics” Univ. Press, Cambridge, 1987

ent

heda.

7. Johnson, K. L., and Greenwood, J. A.,J. Colloid Interface Sci.192,326(1997).

8. Johnson, K. L., Kendall, K., and Roberts, A. D.,Proc. R. Soc. London324,301 (1971).

9. Derjaguin, B. V., Muller, V. M., and Toporov, Y. P.,J. Colloid InterfaceSci.53, 314 (1975).

0. Maugis, D.,J. Colloid Interface Sci.150,243 (1992).1. Greenwood, J. A.,Proc. R. Soc. London A453,1277 (1997).2. Lantz, M. A., O’Shea, S. J., Welland, M. E., and Johnson, K. L.,Phys.

Rev. B55, 10776 (1997).3. Israelachvili, J. N.,in “Fundamentals of Friction” (I. L. Singer and H. M

Pollock, Eds.). Kluwer, Dordrecht, 1992.4. Johnson, K. L.,Proc. R. Soc. London A453,163 (1997).5. Enachescu, M., van den Oetelaar, R. J. A., Carpick, R. W., Ogletree,

Flipse, C. F. J., and Salmeron, M.,Phys. Rev. Lett.81, 1877 (1998).6. Carpick, R. W., Ogletree, D. F., and Salmeron, M.,Appl. Phys. Lett.70,

1548 (1997).