Вы находитесь на странице: 1из 25

18

Atomic-Scale Friction
Studies Using Scanning

Force Microscopy

18.1 Introduction
18.2 The Scanning Force Microscope as a Tool for
Nanotribology
18.3 The Mechanics of a Nanometer-Sized Contact
18.4 Amontons Laws at the Nanometer Scale
18.5 The Inuence of the Surface Structure on Friction

Frictional Contrast Caused by Local Changes in the Chemical
Composition Friction of Surfaces Possessing Identical
Chemical Composition Direction Dependence of Sliding
Friction

18.6 Atomic Mechanism of Friction
18.7 The Velocity Dependence of Friction
18.8 Summary

18.1 Introduction

Although frictional forces are familiar to everyone from daily life experiences, little progress has been
made in nding an exact physical description of friction since the following phenomenological friction
laws were established by Amontons (1699) and Coulomb (1783):
1. Friction is

proportional

to the load (i.e., to the applied normal force).
2. Friction is

independent

of the (apparent) contact area.
3. Sliding friction is

independent

of the sliding velocity.
These laws hold surprisingly well in the case of

dry friction

or

Coulomb friction,

if there is no lubrication
between the two interacting bodies. Nevertheless, they could not be derived from rst principles up to
now. Moreover, the understanding of the fundamental mechanisms of friction on the atomic scale is
poor since most macroscopic and microscopic frictional effects are normally dominated by the inuence
of wear, plastic deformation, lubrication, surface roughness, and surface asperities. Macroscopic friction
experiments (see Figure 18.1a), especially if performed under ambient conditions are therefore difcult
to analyze in terms of a universal theory. Moreover, the complexity of the phenomena prevented the
observation of pure wearless friction for many years, thus disabling all attempts to understand the origins
of friction on the molecular and atomic scale.
In the last decades, however, the eld of

nanotribology

was established by introducing new experi-
mental tools which opened the nanometer and the atomic scale to tribologists. One of the basic ideas of

Udo D. Schwarz

University of Hamburg

Hendrik Hlscher

University of Hamburg
2001 by CRC Press LLC
nanotribology is that for a better understanding of friction in macroscopic systems, the frictional behavior
of a single asperity contact should be investigated rst. Macroscopic friction could then possibly be
explained with the help of statistics, i.e., by adding the interactions of a large number of individual
contacts, which together form the macroscopic roughness of the interface between the two bodies which
are in relative motion (see Figure 18.1b for illustration). The rst big step toward such experiments at
the atomic level was taken when the surface force apparatus was equipped with special stages enabling
the measurement of lateral forces between two molecularly smooth surfaces sliding against each other
(Israelachvili and Tabor, 1973; Briscoe and Evans, 1982). The biggest step concerning the reduction of
the asperity size was accomplished with the development of the

friction force microscope

(FFM) (Mate
et al., 1987). The FFM is derived from the

scanning force microscope

(SFM) (Binnig et al., 1986) and
enables the local measurement of lateral forces by moving a sharp tip, representing (approximately) a

point contact,

over a sample surface.
In this chapter, we intend to describe how the FFM can contribute to the understanding of the origin
and nature of the fundamental laws of friction by the investigation of atomic-scale frictional effects. As
shown in the following sections, friction at the nanometer scale manifests itself in a signicantly different
manner than the Coulomb friction at the macroscopic scale. The friction laws (1) and (2) are found to
be

not valid

for point contact friction, since the frictional forces on the nanometer scale are proportional
to the actual area of contact, which is generally not proportional to the load (Section 18.4). On the other
hand, friction law (3) could be conrmed for moderate sliding velocities also on this scale (Section 18.7).
The analysis of the frictional behavior of such approximate point contacts in terms of a simple mechanical
model exhibits a two-dimensional nature of friction on the atomic scale, which is determined by a stick-
slip type movement of the foremost tip atoms (Section 18.6). Molecular-scale effects involving wear or
lubrication, however, will be discussed in the next chapter.

18.2 The Scanning Force Microscope as a Tool

for Nanotribology

The principle of scanning force microscopy is rather simple and resembles that of a record player. A force
microscope (Figure 18.2) detects forces acting between a sample surface and a sharp tip which is mounted
on a soft leaf spring (the

cantilever

). A feedback system which controls the vertical

z

-position of the tip
on the sample surface keeps the deection of the cantilever (and thus the force between tip and sample)
constant. Moving the tip relative to the sample in the

x


y

-plane of the surface by means of piezoelectric
drives, the actual

z

-position of the tip is recorded as a function of the lateral

x


y

-position with (ideally)

FIGURE 18.1

(a) The well-known experimental setup to conrm the frictional laws (1) and (2) of dry friction on
the macroscopic scale. (b) The apparent contact area observed on the macroscopic scale consists of many individual
small asperities. A nanometer-sized single asperity contact (point contact ) can be realized in a scanning force
microscope.
2001 by CRC Press LLC
sub-ngstrm precision. The obtained three-dimensional data represent a map of equal forces; the
necessary conversion of the cantilever deection into the normal force is basically performed by applying
Hooks law. The data can be analyzed and visualized through computer processing. A more detailed
description of the method and the generally used instrumentation is given, e.g., by Meyer and Heinzel-
mann (1992), Wiesendanger (1994), Bhushan and Ruan (1994), or Schwarz (1997).
The experimental setup for the deection detection measurement, which is most frequently used in
SFM studies on friction, is the beam deection scheme sketched in gure 18.2a (Marti et al., 1990; Meyer
and Amer, 1990). With this detection scheme, not only the deection, but also the torsion of the cantilever
can be measured (see the caption of Figure 18.2a for further details). Force microscopes, which can record
bending and torsion of the cantilever simultaneously, are often referred to as

friction force microscopes

(FFMs) or

lateral force microscopes

(LFMs).
What are the specic advantages of friction force microscopy in comparison with other methods?As
we have seen in the introduction, macroscopic friction experiments are often difcult to analyze due to
the complexity of the experimental system. For the unambiguous analysis of physical phenomena, the
experiments performed to clarify their nature are preferably arranged as simply as possible. Friction force
microscopy is an especially suitable tool for nanotribological investigations because it meets this require-
ment quite well:
1. Measurements within the

purely elastic

regime prevent plastic deformation or wear of tip or sample.
Consequently, all experiments presented in this chapter were performed at loading forces which
were sufciently low such that no effects due to plastic deformation could be observed.
2. The contact area of the sliding tip with the sample surface is very small (typically some nm

2

), thus
enabling the

localized detection of frictional forces.

3. The

apparent

and the

actual

contact area are identical. Therefore, problems occurring from surface
roughness and surface asperities can be neglected if the experiments are performed on atomically
smooth surfaces.
4. If measured under controlled conditions (e.g., in ultra-high vacuum or in an argon atmosphere),
an inuence of adsorbates on the measured friction can be reduced or completely avoided.

FIGURE 18.2

(a) Principle of the scanning force microscope. Bending and torsion of the cantilever are measured
simultaneously by measuring the lateral and vertical deection of a laser beam while the sample is scanned in the

x


y

-plane. The laser beam deection is determined using a four-quadrant photo diode: (A+B)(C+D) is a measure
for the bending and (A+C)(B+D) a measure for the torsion of the cantilever, if A, B, C, and D are proportional to
the intensity of the incident light of the corresponding quadrant. (b) The torsion of the cantilever (middle) is solely
due to lateral forces acting in the

x

-direction, whereas both forces acting normal to the surface (

F

z

) as well as acting
in plane in the

y

-direction (

F

y

) cause a bending of the cantilever (bottom).
2001 by CRC Press LLC
Finally, two more points should be noted: (1) During scanning, the torsion of the cantilever is
exclusively induced by lateral forces acting in the

x

-direction (

F

x

), whereas for corrugated samples, the
bending of the cantilever is mainly caused by forces acting perpendicular to the sample surface in the

z

-direction (

F

z

). Nevertheless, a bending of the cantilever induced by lateral forces acting in the

y

-direction
(

F

y

) cannot be excluded (see Figure 18.2b, bottom). For atomically at surfaces, Fujisawa et al. (1993a,b;
1995) pointed out that the variation of

F

z

is so small that the measured signal of an SFM equipped with
the beam deection method is mostly determined by lateral forces. Therefore, on the atomic scale, the
lateral forces both in scan direction and perpendicular to the scan direction are measured. (2) The correct
calibration of the bending and the torsion of the cantilever in terms of lateral forces is much more difcult
than the conversion of the cantilever deection into the normal force, which can simply be performed
by applying Hooks law. Many parameters such as the cantilever dimensions, the elastic moduli of the
cantilever material, the tip length, the position on the cantilever backside where the laser beam is reected
and the sensitivity of the four-quadrant photo diode have to be known. Different procedures for the
quantitative analysis of lateral force microscopy experiments are described, e.g., by Marti et al. (1993),
Bhushan and Ruan (1994), Lthi et al. (1995), Putman et al. (1995), Ogletree et al. (1996), and Schwarz
et al. (1996a).

18.3 The Mechanics of a Nanometer-Sized Contact

Since it is the attempt of nanotribology to reconstruct the macroscopic frictional behavior from the
frictional properties of an individual nanometer-sized contact, it is clear that a knowledge of the mechan-
ics of such small single asperity contacts is essential. For example, the exact contact area of tip and sample
as a function of the applied loading force has to be known in order to compare the data obtained by
scanning force microscopy with theoretical models, as we will see in Section 18.4. Scanning force micros-
copy, however, does not allow an independent measurement of the contact area. In order to circumvent
this restriction, it is a common approach in SFM to use tips with a geometrically well-dened apex and
to calculate the effective contact area (and other important parameters such as deformation or stiffness
of the contacts) on the basis of continuum elasticity theory. A detailed description of

contact mechanics

is given by Johnson (1994).
Here, we will restrict ourselves to the mathematically simplest and most frequently used case of a spherical
tip which is in contact with a at surface (the so-called

Hertzian contact,

Figure 18.3), in spite of the fact
that there also exist extensions to other geometries (see, e.g., Carpick et al., 1996). Without adhesion, the
contact area of such a Hertzian contact can be determined using the Hertzian theory (Hertz, 1881), which
was already developed in 1881. It predicts a contact area-load dependence of the following form:
(18.1)

FIGURE 18.3

(a) Macroscopic tipsample contact;

F

l

denotes the externally applied loading force. The contact
area-load dependence is difcult to describe due to the irregular shape of both tip and sample surface. (b) Geometry
of a Hertzian contact (see text).

R

= radius of the tip apex;

a

= radius of the contact area.
A
RF
K
l

j
(
,
\
,
(
2 3
2001 by CRC Press LLC
Here,

R

is the radius of the tip,

F

l

the loading force, and
(18.2)
the effective elastic modulus (with

E

1,2

and


1,2

as Youngs moduli and Poissons ratios of sphere and
at, respectively). Other useful quantities are, e.g., the contact radius

a

of the interface region
(18.3)
the vertical displacement


of the two bodies in contact
(18.4)
and the pressure or stress distribution

p

(

r

) within the contact region
(18.5)
Since the early 1970s, however, theories have been developed which include attractive forces (Johnson
et al., 1971; Derjaguin et al., 1975; Muller et al., 1983; Maugis, 1987; Fogden and White, 1990; Maugis,
1992; Maugis and Gauthier-Manuel, 1994). The general mathematical formulation of the contact area-
load dependence of a Hertzian contact including adhesion cannot be solved analytically (Fogden and
White, 1990; Maugis, 1992). Nevertheless, depending on how the attractive forces act, approximations
can be derived.
In order to decide which of the above-mentioned approximations applies under which conditions,
Tabor (1977) introduced the nondimensional parameter


(18.6)
(


: surface energy of sphere and at [ for equal surfaces] and

z

0

: equilibrium distance in contact), which
represents basically a measure for the magnitude of the elastic deformation

outside

the effective contact
area compared with the range of the surface forces. When


is large (


> 5 [ Johnson, 1997] ), the
JohnsonKendallRoberts (JKR) theory (Johnson et al., 1971) applies, which has been conrmed in many
experiments (see, e.g., Israelachvili et al. [ 1980] and Homola et al. [ 1990] ). If


< 0.1, however, contacts
should be appropriately described by the analysis of Derjaguin, Muller, and Toporov (DMT) (Derjaguin
et al., 1975; Muller et al., 1983). Nevertheless, since the small value of


indicates that there is only
negligible elastic deformation outside the effective contact area, the inuence of attractive forces can in
further simplication be considered by introducing an effective normal force

F

n

=

F

1

+

F

0

, where

F

0

is
the sum of all acting attractive forces. Replacing

F

1

with

F

n

in Equation 18.1 leads to
(18.7)
K
E E

j
(
,
\
,
(

4
3
1 1
1
2
1
2
2
2
1

a
RF
K
l
3


a
R
F
Ka
l
2
p r
Ka
R
r
a
F
a
r
a
l
( )

j
(
,
\
,
(

j
(
,
\
,
(
3
2
1
3
2
1
2
2
2

j
(
,
\
,
(
9
4
2
2
0
3
1 3
R
K z

A
R F F
K
RF
K
l
n

+
( )
j
(
,
\
,
(

j
(
,
\
,
(
0
2 3
2 3
2001 by CRC Press LLC
This model will be referred to as the

Hertz-plus-offset model

in the following to distinguish between the
original and more precise DMT formalism and the present further simplication. Detailed mathematical
derivations of Equation 18.7 for the case of capillary forces were published by Fogden and White (1990)
and Maugis and Gauthier-Manuel (1994).
By introducing typical values for contacts as they occur in SFMs into Equation 18.6 (

R

= 30 nm,


=
25 mJ/m

2

[ typical for van der Waals surfaces] ,

K

= 50 GPa, and

z

0

= 3 ), we nd a value for


of


0.09. Therefore, Equation 18.7 should represent the correct contact area-load dependence for typical
contacts found in SFMs. A more detailed discussion of the contact mechanics of a nanometer-sized
Hertzian contact is given by Schwarz et al. (1996a, 1997a).
However, Equation 18.7 only applies for tips with

exactly spherical

apex. Using SFM tips as supplied
by the manufacturer, such a shape is only accidentally realized, and thus all kinds of power laws

A





F

n

with

n

ranging from

n



0.4 to

n



1.2 have been found (Schwarz et al., 1997a). Consequently, tips of
exactly dened spherical tip apex and a tip radius which is known with nanometer accuracy are mandatory
for quantitatively reproducible friction force measurements.
Mainly two different methods to prepare such well-dened tips have been introduced in the past. The
rst method is to heat a very sharp tip in high vacuum (Binh and Uzan, 1987; Binh and Garcia, 1992).
Surface diffusion will then induce migration of atoms from regions of higher curvature to regions of
lower curvature. Another method was realized by covering doped single-crystalline silicon tips (apex
radii 5 to 15 nm without coating) with a layer of amorphous carbon in a transmission electron microscope
(Schwarz et al., 1997b). Molecules from the residual gas are ionized in the electron beam and accelerated
to the tip end. There, the molecules spread out evenly due to their charge, forming a well-dened spherical
tip end. With this method, tips with radii from 7 nm up to 112 nm could be successfully produced. Three
examples of tips prepared using the second method described above are shown in Figure 18.4.

18.4 Amontons Laws at the Nanometer Scale

Amontons macroscopic friction laws (1) and (2) can be condensed in the well-known equation
(18.8)
where

F

f

denotes the observed friction force.


represents a value which is constant for a given material
combination in a wide range of applied normal forces

F

n

and which is usually referred to as the

friction
coefcient.

Nevertheless, for the description of the frictional behavior of materials at the nanometer scale,
it is practical to introduce the mean friction per unit area (the so-called

shear stress S

)

FIGURE 18.4

Three examples of tip specially prepared in a transition electron microscope with spherical tip apexes
of (a) 21


5 nm, (b) 35


5 nm, and (c) 112


5 nm radius.
F F
f n

2001 by CRC Press LLC
(18.9)
Combining Equations 18.7 and 18.9, we nd
(18.10)
with


S

/

K

2/3

.
There have already been many reports in the literature when the frictional force has been measured
as a function of the normal force using an FFM (Mate et al., 1987; Mate, 1993; Hu et al., 1995; Bhushan
and Kulkarni, 1995; Carpick et al., 1996; Meyer et al., 1996; Lantz et al., 1997; Schwarz et al., 1997a;
Enachescu et al., 1998). In most of these reports, the frictional force showed a strongly nonlinear depen-
dence on the normal force. Figure 18.5 features results published by Schwarz et al. (1997a), which were
acquired in argon atmosphere on amorphous carbon with tips showing exactly spherical tip apexes (tip
radii of 17


5 nm [ a] and 58


10 nm [ b] , respectively). The ts according to Equation 18.10 show
excellent agreement between the experimental results and theory. The dashed lines illustrate the deviation
of the present friction law from the macroscopic linear law Equation 18.8.
It is interesting to note that not only can the relationship

F

f





F

2/3
n

be conrmed, but also the

F

f



R

2/3

-dependence. Comparing Figure 18.5a with 18.5b, much higher friction is observed in (b); the fric-
tional force at

F

n

= 10 nN, for example, is about three times larger (


15 nN) than the frictional force of
only 5.5 nN observed in (a). However, the calculated numerical value of

= 0.17


0.07 GPa

1/3

for the
second measurement is in good agreement with the value of

= 0.17


0.09 GPa

1/3

obtained in the rst
measurement.
Summarizing the main results obtained from the different research groups on a large variety of
materials, it was found that the friction as a function of the normal load showed good agreement of
experimental data and theoretical ts using Equation 18.9 combined with contact mechanical models
appropriate for the specic tip-sample contact, if

S

is set constant. This nding leads to the following
remarkable consequences:

FIGURE 18.5

The frictional force

F

f

as a function of the normal force

F

n

, measured on amorphous carbon with
geometrically well-dened spherical tips in argon atmosphere (Schwarz et al., 1997a). The data presented in (a) were
obtained using a tip with a radius of

R

= 17


5 nm; the data in (b) using a tip with

R

= 58


10 nm. Both curves
are in excellent agreement with ts according to Equation 18.10 (solid lines); the dashed lines illustrate the deviation
from the linear macroscopic model (Equation 18.8). The offsets of the solid lines from the zero point of the normal
force are caused by the experimental uncertainty in the determination of the zero point by means of force-vs.-distance
curves.
S F A
f
F S
R
K
F R F
f n n

j
(
,
\
,
(

2 3
2 3 2 3 2 3

2001 by CRC Press LLC


The observed frictional force F
f
is in opposition to law (2) proportional to the crack area A (see
Equation 18.9). In the special case of a Hertzian contact, this leads to a F
f
(F
n
)-dependence of F
f
~
F
2/3
n
.
The shear stress S depends only on the materials used and on environmental conditions such as
temperature and humidity, but not on the mean contact pressure p = F
n
/A.*
The independence of the shear stress S from the mean contact pressure p = F
n
/A is in agrant
contradiction to Amontons law (1).
The continuum elasticity theory also applies at the nanometer scale.
Since the continuum elasticity theory also applies at the nanometer scale, Equation 18.7 can be used
to calculate the actual contact area of an SFM tip and the sample surface. Typical normal forces F
n
occurring in an SFM during measurement are between 1 nN and 100 nN. This results in contact areas
of a few nm
2
, with the effect that only some tens or some hundreds of atoms are in direct contact. The
deformations of tip and sample are then fully elastical, i.e., fully reversible. Therefore, with an SFM, it
is indeed possible to study wearless friction of a quasipoint contact.
Another outcome from the above considerations is the nding that it is obviously useless to determine
friction coefcients (see Equation 18.8) from point contact measurements, since these values will
heavily depend on the specic geometry of the tipsample contact. For a classication of the microscopic
frictional properties of materials, Schwarz et al. (1997a) have therefore proposed the factor

= S/K
2/3
(see Equation 18.10), which combines the frictional and elastic properties of tip and sample and which
can be regarded as an effective friction coefcient for point contact-like single-asperity friction in the case
of a nanometer-sized Hertzian contact. The denition of such a coefcient is advantageous for two main
reasons: (1) If materials with identical intrinsic frictional properties (i.e., the materials show the same
friction per unit area) but different Young moduli are examined in measurements performed with
identical tips and normal forces, higher friction will be found on the softer material due to the larger
contact area. Materials which have the same

, however, will always show the same friction during an


experiment. (2) If the same sample is investigated with tips featuring different apex radii, lower friction
will be found in the experiment carried out with the sharper tip. Nevertheless, consideration of the
geometry by the calculation of

will lead to identical values of

, as we have seen in Figure 18.5.


How can the above results now be correlated with the macroscopic laws of friction?As we have seen
in the introduction, it is the general attempt of nanotribology to conclude from the frictional properties
of an individual nanocontact to the behavior of macroscopic bodies by the statistical adding of the
interactions of a large number of individual contacts with represent the macroscopical roughness of the
contact interface. This means for the given situation that the contradiction between Amontons laws (1)
and (2) and their corresponding counterparts on the nanometer scale could be entirely eliminated if the
effective contact area of such a macroscopic contact increased proportionally to the externally applied
loading force. Then, the observed frictional force would also increase linearly with the load, according
to the nanoscopic friction law found above. Such a linear F
f
(A
eff
)-dependence for macroscopically at,
but microscopically rough surfaces has indeed been demonstrated by Greenwood (1967, 1992). Interest-
ingly, the apparent contact area of the two bodies does not matter in this case.
18.5 The Inuence of the Surface Structure on Friction
In the preceding sections, we dealt with the mechanical and tribological properties of nanometer-sized
contacts. Questions concerning the fundamental mechanisms of friction at the atomic scale, however,
*The same result has been found by Carpick et al. (1997) and Enachescu et al. (1998) with different approaches,
which both circumvent the assumption of a specic contact mechanical model. Additionally, experiments performed
with the surface force apparatus and consequently much larger contact areas A also show the same behavior
(Israelachvili and Tabor, 1973; Homola et al., 1990).
2001 by CRC Press LLC
are still unanswered. Since we focus in this chapter on the analysis of wearless friction on atomically at
surfaces without defects, it is reasonable to expect that the origin for the different friction observed on
the different materials can be found in the local arrangement of the atoms at the tipsample contact area.
If this expectation holds, it should be possible to distinguish areas of different chemical composition
by lateral force microscopy. This ability is sometimes called chemical imaging; examples for such
investigations will be discussed in Section 18.5.1. Another interesting consequence of the above expec-
tation is that the friction should depend on the direction with which the tip proles the sample surface.
This frictional anisotropy is especially easy to verify experimentally if a material exhibiting an asymmetric
surface potential is used as a sample, as presented in Section 18.5.3. Finally, we will see in Section 18.5.2
that even small conformational (i.e., purely geometrical) changes within a surface unit cell give rise to
different friction in lateral force maps.
18.5.1 Frictional Contrast Caused by Local Changes
in the Chemical Composition
The ability of the FFM to provide a contrast between surfaces of different chemical composition is not
only important for tribologists, but also of considerable general interest for users from the whole eld
of scanning force microscopy, where contrast mechanisms are needed which provide information in
addition to the pure topography. Since the lateral forces acting on the FFM tip are not independent of
the topography,* atomically or molecularly at terraces give the most unambiguous results. Consequently,
chemical contrast has rst been demonstrated for LangmuirBlodgett lms, which can easily be prepared
and investigated under atmospheric conditions and exhibit large, molecularly at surfaces. See, for
example, Overney et al. (1992), or Meyer et al. (1992).
However, in this chapter we will restrict the discussion to the atomic-scale frictional effects at crystalline
surfaces. In order to check the frictional behavior for such systems, Lthi et al. (1995) partially covered
the surface of an NaCl(001) single crystal with 1 to 6 monolayers of AgBr (Figure 18.6a), which has the
same crystalline structure as NaCl. In the corresponding FFM investigations (Figure 18.6b), they found
a strong frictional contrast between the AgBr islands and the NaCl substrate due to the different chemical
composition of the two materials.
FIGURE 18.6 (a) Topography and (b) lateral force map of an AgBr thin lm deposited on NaCl(001) at room
temperature (image size: 1.4 m 1.4 m). The AgBr islands are 1 to 6 monolayers high and partially cover the
NaCl substrate. In the lateral force map, the AgBr islands are revealed as areas with about ten times higher friction
than the corresponding friction observed on the NaCl surface. (From Meyer, E., Lthi, R., Howald, L., and
Gntherodt, H.-J., 1995. Friction force microscopy, in Forces in Scanning Probe Methods, Gntherodt, H.-J., Ansel-
metti, D., and E. Meyer (Eds.), Kluwer Academic Publishers, Dordrecht, The Netherlands, 285. With permission.)
*Scanning up hill causes more torsion of the cantilever than scanning down hill ; see, for example, Grafstrm
et al. (1993), Fujisawa et al. (1993b), Ruan and Bhushan (1994a), or Aim et al. (1995).
2001 by CRC Press LLC
According to the same principle, basically all areas of different chemical composition can be distin-
guished in force microscopical investigations. Fompeyrine et al. (1998), for example, could differentiate
areas terminated by SrO from areas terminated by TiO
2
on an SrTiO
3
single crystal. This ability of the
SFM is sometimes called chemical imaging. It should be noted, however, that chemical imaging does
not mean chemical identication. The reason for this has been discussed in Section 18.4. If the contact
mechanics of the tipsample contact is not exactly known, the results obtained are difcult to interpret
without additional information. In order to circumvent this restriction at least partially, chemically
modied tips have been used to sensitize the tip to certain functional groups which have been patterned
on suitable sample surfaces. See, for example, Frisbie et al. (1994) or Sasaki et al. (1998). A theoretical
study on the bases of chemical force microscopy on organic monolayers has been published recently
by Fujihira and Ohzono (1999).
18.5.2 Friction of Surfaces Possessing Identical Chemical Composition
After what has been discussed above, it seems promising to analyze the inuence of differences in the
structure of surfaces possessing identical chemical composition on the friction observed by a point probe.
Probably the rst observation on this issue was published by Ruan and Bhushan (1994b). They observed
graphite(0001) areas with unusually high friction, which could be identied as areas exhibiting graphite
planes of different orientations (other than [ 0001] ) as well as amorphous carbon. Comparative quanti-
tative studies for different carbon compounds have been performed by Mate (1993) and Schwarz et al.
(1997a) on diamond, graphite, C
60
thin lms, and amorphous carbon. High frictional forces have been
found especially on the C
60
thin lms, whereas friction nearly vanishes on graphite. These examples
demonstrate that even on materials that consist of the same chemical species, large variations in the
observed friction might occur.
As a consequence, it follows that it is not the specic atomic species which form a crystal, but factors
such as the crystalline structure, the charge distribution, and/or the binding conditions that might decide
the frictional properties of a material. The chemical differentiation postulated in Section 18.5.1 is
therefore based more on structural and electronical differences between the materials than on the fact
that they represent chemically different species.
This issue will be further illustrated with the example of the ferroelectric material guanidinium
aluminum sulfate hexahydrate (GASH) in the following. Since the Curie point of GASH is above its
decomposition temperature, the crystals always show the ferroelectric phase. The surfaces of positively
and negatively charged domains not only have the same chemical composition, the surface atoms are
also bound in the exact same manner on both domains. The difference is that the surface undergoes a
conformational (i.e., geometrical) change of the surface structure if the polarity of a certain surface area
is ipped. Certain aluminum ions which are octahedrally coordinated by water molecules stick on the
positive domain more out from the surface (by 0.5 ) than they do no the negative domain, resulting
in a different surface corrugation. Bluhm et al. (1998) showed in a combination of voltage-dependent
friction force microscopy experiments and electrostatic force microscopy experiments that the domain
contrast observed in FFM images is not caused by the electrostatic interaction between a charged tip and
the electric eld of the sample, but is explained by the geometrical differences exhibited by the surfaces
of domains with opposite polarity (Figure 18.7). This example elegantly demonstrates that even small
changes in the arrangement of the atoms located at the sample surface inuence the friction occurring
during the sliding of nanocontacts.
18.5.3 Direction Dependence of Sliding Friction
In the above sections, we have seen that friction is very sensitive to the structure of the surface. In everyday
life, it is often observed that friction depends on the sliding direction. This phenomenon can be observed
rubbing by ones hand over certain pieces of cloth. Another example is that of a saw; there, how much
2001 by CRC Press LLC
one has to pull or push depends on the direction. It is now instructive to investigate how this mechanism
also applies on the atomic scale.
The basic idea of this issue is illustrated in Figure 18.8 by contrasting the behavior of a tip moving
(a) in a symmetric and (b) in an asymmetric surface potential. The asymmetric surface potential is
represented by a sawtooth-like structure. Intuitively, one might assume that a sharp tip which proles
such an asymmetric potential would experience a larger lateral force at the steep sides of the sawtooth than
FIGURE 18.7 (a) Topographical image of GASH(0001) (image size: 30 m 30 m). The surface is featureless
and without any steps. (b) Friction force map of the same surface spot as in (a) (slight zoom-out; image size: 40 m
40 m). A well-expressed contrast is visible. (c) Image acquired with electrostatic force microscopy at the same
surface area as in (a), showing the structure and absolute sign of the ferroelectric domains. By comparison of (b)
and (c), it can be concluded that the contrast in the friction image represents the domain structure of GASH.
FIGURE 18.8 Principle of FFM contrast formation. (a) On materials that have no directional dependence of friction
(i.e., the friction coefcient is the same in the forward and backward scan directions), the contrast between surface
areas exhibiting different is reversed when scanning forward compared to scanning backward since the FFM signal
is a measure for the torsion of the cantilever. At the surface steps, peaks occur in the FFM signal due to some torque
of the tip. (b) FFM tip probing the frictional force of a surface with asymmetric surface potential, illustrated by a
sawtooth-like structure. The surface structure on the lower terrace is rotated by 180 along the surface normal
compared to the structure on the upper terrace. The friction coefcient of one single terrace is, e.g.,

in the forward
scan direction, whereas it changes to

for the backward scan direction due to effects as described in the text; i.e.,
the friction coefcient is direction-dependent. The contrast recorded in the FFM signal is the same for both directions.
The direction dependence, however, has no inuence on the torque of the FFM tip at the surface steps.
2001 by CRC Press LLC
on the less inclined sides.* Therefore, lower friction is expected for the upper terrace in comparison with
the lower terrace for the forward scan direction. Additionally, the frictional contrast between the upper
and the lower terrace should vanish if the tip is moved along the grooves of the sawtooth, i.e., perpen-
dicular to the sketched gure.
For the rst time, effects due to the atomic-scale asymmetry of a surface potential were reported by
Overney et al. (1994) for a lipid bilayer on oxidized Si(100). A similar domain contrast was also found
by Gourdon et al. (1997) and Liley et al. (1998) with a thiolipid monolayer on mica. As a possible
explanation, the observed contrast was considered to be caused by different molecular tilts within the
individual domains. On crystalline surfaces, however, such effects can be studied in a more controlled
manner. Here, we will use the ferroelectric material triglycine sulfate (TGS) as an example; similar results
have recently been published by Shindo et al. (1999) for alkaline earth sulfate crystals.
Figure 18.9 displays a perspective view on two neighboring terraces separated by a b/2 step within a
negative domain. The glycine molecules form a sawtooth-like pattern perpendicular to the c-axis. The
most remarkable feature of the surface structure in this context is that the arrangement of the molecules
on the upper terrace is rotated by 180 compared to the structure on the lower terrace. This sawtooth-
like surface structure is very similar to the case considered above, and an asymmetry in friction is therefore
expected.
Measurements by Bluhm et al. (1995) fully conrm the expected behavior. Figure 18.10 shows the
relative frictional contrast between neighbored terraces (i.e., the difference of the friction on the upper
and on the lower terrace) in arbitrary units as a function of the sample rotation angle; 0 is orthogonal
to the a-axis. Due to a 105 angle between the a and the c-axis of the TGS crystal, the sample orientation
is for 165 and 345 parallel to the grooves of the sawtooth structure. For these angles, the frictional
contrast vanishes as predicted. A detailed discussion of the frictional anisotropy on TGS(010), including
a simple theoretical model, can be found by Schwarz et al. (1996b).
18.6 Atomic Mechanism of Friction
Although the experiments described above were performed with sliding distances in the micrometer
range, it is clear that the origins for the studied effects can only be found on the atomic scale. In
Section 18.2, we have seen that it is possible to measure frictional forces with a scanning force microscope
down to the atomic scale. The analysis of such measurements gives interesting insight into the basic
mechanism of friction.
FIGURE 18.9 Perspective view of the (010) surface of a negative domain exhibiting a surface step of half of the
unit cell height.
*Such a correlation between surface slope and lateral force can indeed not only be observed on the macroscopic
scale, but has also been demonstrated at length scales of some 10 or 100 nanometers (Fujisawa et al., 1993b; Grafstrm
et al., 1993; Bhushan et al., 1994; Ruan and Bhushan, 1994a), making our assumption even more substantial.
2001 by CRC Press LLC
Many issues observed on the atomic scale can be illustrated by using a simple but instructive mechanical
model based on the so-called Tomlinson (Tomlinson, 1929) or independent oscillator (IO) model (Prandtl,
1928; McClelland, 1989; Helman et al., 1994). A schematic view of this simple spring model is shown in
Figure 18.11. A point-like tip is coupled elastically to the main body M with a spring possessing a spring
constant c
x
in the x-direction, and it interacts with the sample surface via a periodic potential V
int
(x
t
),
where x
t
represents the actual position of the tip. During sliding, the body M is moved with the sliding
velocity
M
in the x-direction. All energy dissipation independent of the actual dissipation channel
(phonons or electronic excitations) is considered by a simple velocity-dependent damping term (
x
.
x
t
).
Within this model, the point-like tip represents the average of the real tipsample contact or single
asperity contact, where up to hundreds of atoms might be involved (see Section 18.4). In principle, the
real tipsample contact could also be treated as a system consisting of many individual atomically sharp
tips interacting through springs, leading to more complex models of friction. However, we will restrict
ourselves to the simple model introduced above, which has proven to describe successfully the tip motion
for many materials (Table 18.1).
The resulting equation of motion for the tip in the interaction potential is given by
(18.11)
FIGURE 18.10 Relative frictional contrast for neighbored terraces on the negative domain of TGS(010) as a function
of the sample rotation angle.
FIGURE 18.11 (a) A simple model for a tip sliding on an atomically at surface based on the Tomlinson model.
A point-like tip is coupled elastically to the body M by a spring with spring constant c
x
in the x-direction; x
t
represents
its position within an external potential V(x
t
) with periodicity a. If x
t
= x
M
, the spring is in its equilibrium position.
For sliding, the body M is moved with the velocity
M
in the x-direction. (b) A schematic view of the tip movement
in a sinusoidal interaction potential. If condition (18.13) holds, the tip shows the typical stick-slip-type movement,
i.e., it jumps from one potential minimum to another. (c) If the tip moves with stick-slips over the sample surface,
the lateral force F
x
manifests as a sawtooth-like function.
m x c x x
V x
x
x
x x x

int
t M t
t
t
t

( )

( )


2001 by CRC Press LLC
where m
x
is the effective mass of the system, x
M
=
M
t the equilibrium position of the spring, and
x
the
damping constant. The solution of this differential equation is the path of the tip x
t
(t). The lateral force
F
x
to move the tip in the x-direction can be calculated from F
x
= c
x
(x
M
x
t
), whereas the friction force F
f
is dened as the averaged lateral force (F
x
).
If the sliding velocities are very low, analytical solutions of Equation 18.11 can be derived. In this case,
the tip will always be in its stable equilibrium position, and Equation 18.11 can be solved for x
t
= 0 and
.
x
t
= 0:
(18.12)
From this equation, a path x
t
(x
M
) can be determined.* For a stiff spring, there is only one solution for
all x
M
, resulting in a continuous tip movement and vanishing friction, since the averaged lateral force
(F
x
) is zero (Figure 18.12). However, the tip movement changes dramatically if the condition
(18.13)
is fullled. Now the tip moves discontinuously in a stick-slip-type motion over the sample surface and
jumps from one potential minimum to another. This specic movement of the tip results in a sawtooth-
like function for the lateral force F
x
(see Figures 18.11c and 18.12). Since the averaged lateral force (F
x
)
is nonzero in this case, the friction force F
f
is necessary to move the body M in the x-direction.
To simulate experimental data obtained with a scanning force microscope, it is important to remember
that an SFM can measure the lateral forces along and perpendicular to the scan direction (Section 18.2).
Therefore, it is necessary to extend the one-dimensional model introduced above to two dimensions in
a manner that considers explicitly the tip movement perpendicular to the scan direction (Gyalog et al.,
1995). In the further analysis, we will show that this two-dimensional movement of the tip is signicant
and thus has a considerable effect on SFM images.
The extension of the model to two dimensions leads to the following coupled second-order differential
equation
(18.14)
TABLE 18.1 Some Examples of Materials where the Tomlinson or
Independent Oscillator Model Has Been Successfully Applied to Simulate
Experimental Data
Sample Reference
Graphite(0001) Sasaki et al., Phys. Rev. B, 54, 2138 (1996)
Toussaint et al., Surf. Interface Anal., 25, 620 (1997)
Hlscher et al., Phys. Rev. B, 57, 2477 (1998)
KBr(001) Lthi et al., J. Vac. Sci. Technol. B, 14, 1280 (1996)
MoS
2
(001) Hlscher et al., Surf. Sci., 375, 395 (1997)
-MoTe
2
(001) Hlscher et al., Phys. Rev. B, 59, 1661 (1999)
NaF(001) Hlscher et al., Europhys. Lett., 39, 19 (1996)
Organic Monolayers Ohzono et al., Jpn. J. Appl. Phys., 37, 6535 (1999)
*An alternative way to determine the path of the tip is to calculate the minimum of its energy (McClelland, 1989).
c x x
V x
x
x M t
t
t

( )

( )

int
c
V
x
x
<

,
,
]
]
]
]
2
2
int
min
t
m x c t x
V x y
x
x
x x x

int
t M t
t t
t
t

( )

( )


2001 by CRC Press LLC
(18.14)
which describes the movement of the tip in the tipsample potential of an idealized SFM (Hlscher et al.,
1997). The simulation of experimental data is performed by moving the support M with constant velocity

M
in the forward direction (i.e., from left to right) as well as the backward direction (i.e., from right
to left) continuously along a certain line in the x-direction while y
M
is held constant. Calculating the
position of the tip (x
t
, y
t
) from Equation 18.14, the lateral forces F
x
= c
x
(x
M
x
t
) and F
y
= c
y
(y
M
y
t
) can
be determined as a function of the support position (x
M
, y
M
). Then, after increasing y
M
, a new scan line
is computed parallel to the previous line.
The application of the model to a realistic sample system (and thus the movement of a sharp single-
asperity tip in general) is illustrated in the following with the example of the MoS
2
(001) surface, where
the simulated results are compared with experimental data of Fujisawa et al. (1994, 1995). For the
calculations, it is necessary to choose a suitable tipsample interaction potential V
int
(x
t
, y
t
), which has to
be used to solve Equation 18.14. The exact tipsample interaction potential is difcult to determine, since
the precise structure of the tip is usually unknown. However, it can be shown that all basic features of
SFM measurements are reproduced by the simulation as long as V
int
reects the translational symmetry
FIGURE 18.12 The tip movement in the Tomlinson model, illustrated with the example of the simple sinusoidal
potential displayed in (a). If the tip moves in such a potential, its path can be determined from the solution of
Equation 18.12. The graphical solution of this equation is shown in (b) for two different cases. For a stiff spring,
there is only one point of intersection between c
x
(x
M
x
t
) (dotted lines in [ b] ) and (solid line). Consequently,
the obtained path x
t
(x
M
) and the lateral force F
x
displayed by dotted lines in (c) and (d) are continuous, with
the effect that the lateral force averages to zero and no friction occurs. If condition (18.13) holds, however, there is
more than one point of intersection for certain support positions x
M
(dashed lines in [ b] ). This leads to instabilities,
which manifest as jumps of the tip, as shown in (c) and (d). The occurrence of more than one intersection in (b)
means that more than one value x
t
is possible for the same support position x
M
(dashed line in [ c] ). Thus, x
t
jumps
for increasing x
M
instantaneously from one stable value to the next (solid line with arrows), leading to a discontinuous
lateral force which looks like a sawtooth (solid line with arrows in [ d] ). The averaged lateral force F
f
is now nonzero.
V
sin
x
t

x
t
---------------
m y c y y
V x y
y
y
y y M y

int
t t
t t
t
t

( )

( )


2001 by CRC Press LLC
of the sample surface. A comparison with the surface structure of the MoS
2
(001) sample surface displayed
in Figure 18.13 shows that this requirement is met by
(18.15)
where a
x
= 3.16 and a
y
= 5.48 ; the actual value of V
0
depends on the loading force.
With this potential, numerical solutions of Equation 18.14 were computed. The resulting images and
line plots presented here were obtained with the following set of parameters: V
0
= 1 eV, c
x
= c
y
= 10 N/m,
m
x
= m
y
= 10
8
kg,
M
= 400 /s,
x
=
y
= 10
3
Ns/m ~ 2 (critical damping is assumed). These
parameters have typical magnitude, but qualitatively similar results are obtained with other choices.
Figure 18.14 shows a comparison of lateral force images calculated using the two-dimensional IO-
model with experimental data acquired in the [ 100] - and [ 120] -directions. The good agreement dem-
onstrates that the presented method has the capability to simulate experimental lateral force maps.
Moreover, it suggests that the model is able to describe the basic mechanism of atomic-scale friction,
although it is based on comparably simple assumptions.
A better insight into the scan process allows Figure 18.15a, where both lateral forces F
x
and F
y
are
plotted as a function of the support position x
M
for the [ 100] -direction and scan positions y
M
= 0.0, 0.2,
1.0, 1.3, 1.4, 1.7, 2.5, and 2.7 (from left to right) according to the denition of y
M
given in Figure 18.13.
The lateral force in scan direction F
x
(x
M
) exhibits a sawtooth-like shape, whereas the force perpendicular
to the scan direction, F
y
(x
M
), looks like a step function. All features of the calculated forces are in good
agreement with experimental data, which is shown for comparison in Figure 18.15b.
The origin of this behavior is analyzed in Figure 18.16. Figure 18.16a shows the paths of the tip in the
potential of V
MoS
2
(x
t
, y
t
) for different values of y
M
. For y
M
= 0 , no movement of the tip perpendicular
to the scan direction can be observed, as expected from Figure 18.15. However, with increasing y
M
-values,
the path of the tip becomes zigzag. A smaller area of the potential is shown in Figure 18.16b, where the
path of the tip for y
M
= 1.4 and 0.7 is plotted time-resolved by points of equal time distance with
t = 62.5 s. It can be seen that the tip only moves slowly as long as it is in the areas which are framed
by white lines. These areas are obtained by calculating the surface areas where the tip has a stable position
for very slow scan velocities (
M
0) (Gyalog et al., 1995). If the tip leaves these areas of stability, it
becomes unstable and jumps into the next area of stability. This explains the stick-slip-type behavior
of the lateral force F
x
as well as the step-function-like shape of the force F
y
.
FIGURE 18.13 A schematic view of the MoS
2
surface structure. The spheres represent the positions of the sulfate
surface atoms. The dashed line marks the position y
M
= 0 used in Figure 18.15.
V x y V
a
x
a
y
x y
MoS t t t t
2
, cos cos
( )

j
(
,
\
,
(

j
(
,
\
,
( 0
2 2
c
x
m
x

2001 by CRC Press LLC


One consequence of this behavior is that the maximum of F
x
might appear at different positions than
the maximum values of F
y
(see Hlscher et al., 1998), as it was rst observed experimentally by Ruan
and Bhushan (1994a). Another consequence is that a large part of the surface is skipped by the tip and,
therefore, no information about these skipped surface areas can be extracted from the acquired images.
This latter behavior has been experimentally veried for mica by Kawakatsu and Saito (1996); a visual-
ization of this issue based on the MoS
2
data presented in Figures 18.14 and 18.15 is shown in Figure 18.16c.
Here, the position probability density of the tip during the scan is plotted, i.e., only the parts of the
surface that show a high density of black points have contact with the tip for a signicant time; the white
areas are skipped by the tip.
FIGURE 18.14 A comparison between the simulated (top) and measured (bottom) lateral force images for the scan
directions indicated in Figure 18.13. The movement of the support was in positive x-direction (forward ); the scan
size was 25 25 . (Data from Fujisawa, S., Kishi, E., Sugawara, Y., and Morita, S. (1995), Atomic-scale friction
observed with a two-dimensional frictional-force microscope, Phys. Rev. B, 51, 7849-7857.)
FIGURE 18.15 A comparison between the calculated (a) and measured (b) lateral forces F
x
and F
y
for a scan in the
[ 100] -direction and different support positions y
M
as dened in Figure 18.13. (Data from Fujisawa, S., Kishi, E.,
Sugawara, Y., and Morita, S. (1995), Atomic-scale friction observed with a two-dimensional frictional-force micro-
scope, Phys. Rev. B, 51, 7849-7857.)
a)
b)
4
2
0
-4
-2
25
x
M
[A
o
]
F
y

[
n
N
]
F
x

[
n
N
]
LFM (fx)
4
.
8
x
1
0
-
7
N

(
8
.
8
A o

)
1
.
1
x
1
0
-
1
0

N

(
1
.
5
A o

)
AFM (fY )
25 A
o
2001 by CRC Press LLC
At the end of this section, the question arises of how far the situation described above corresponds
with the tribological behavior of a nanocontact which is not connected to a cantilever. By the comparison
of experiment and simulation, however, it can be demonstrated that elastic deformations between the
atoms at or close to the tipsample contact area are responsible for the stick-slip movement of the tip;
astonishingly the cantilever has only a minor inuence. Additionally, in molecular dynamics simulations,
atomic stick-slips as found in the above analysis are found to take place as a direct consequence of the
interplay between the surface forces and interatomic interactions within tip and sample (Landman et al.,
1989; Harrison et al., 1992, 1993a; Schimizu et al., 1998). Even the zigzagging of the tip atoms in the
interaction potential of the surface has been reproduced (Harrison et al., 1993a,b; Shluger et al., 1999).
Thus, it can be concluded that the scanning force microscope indeed measures the real frictional
behavior of a nanocontact.
18.7 The Velocity Dependence of Friction
So far, we have discussed different phenomenological aspects as well as theoretical concepts of friction
on the nanometer scale, but we still have to nd an explanation for Coulombs law (3) on the independence
of sliding friction from the sliding velocity
M
. As already mentioned in the introduction, this law has
also been conrmed on the nanometer scale within a wide velocity range for crystalline surfaces (Hu
et al., 1995; Bouhacina et al., 1997; Zwrner et al., 1998).* Figure 18.17 shows two examples of experi-
mental data sets for amorphous carbon and diamond. It is found that the friction force F
t
is nearly
constant for the velocities that can be reached with an SFM.
To get a basic idea of the mechanism which causes the velocity independence of the frictional forces,
we again apply the simple mechanical model introduced in Section 18.6. For this purpose, we are
interested in the solutions of Equation 18.11 for moderate scan velocities. The qualitative behavior within
the one-dimensional case is shown for a sinusoidal potential V
int
(x
t
) = V
0
sin(2x
t
). All data presented
here are obtained with the parameters V
0
= 1.0 eV, a = 3 , c
x
= 10 N/m (condition [ 13] holds for these
FIGURE 18.16 (a) The interaction potential V(x
t
, y
t
) (image size: 25 25 ) and the calculated paths of the tip
(x
t
(t), y
t
(t)) in this potential for the support positions y
M
= 0.0, 0.2, 0.7, 1.0, and 1.4 (from top to bottom; for the
denition of y
M
see Figure 18.13). If y
M
0, the tip performs a zigzag movement. (b) A detail of (a) (image size:
10 10 ). The contours of the areas of stability (see text) are marked by white lines. In order to show that the
tip stays in these areas most of time, the path of the tip is drawn for y
M
= 1.4 (top) and 0.7 (bottom) by points
separated by equal time intervals t = 62.5 s. (c) Visualization of the tip position during the scanning of the sample.
All calculated paths (54 lines per period) are plotted as in (b). Consequently, the density of points can be interpreted
as the position probability density of the tip on the surface during the scan. The areas of stability are framed by
black lines. Only parts of the stable areas the dark areas are in contact with the tip for a signicant time.
*There is, however, experimental evidence that friction has a logarithmic or even more complicated dependence
on the sliding velocity for noncrystalline overlayers (e.g., polymeric, self-assembled, or LangmuirBlodgett-type
layers), probably due to a certain viscosity of the layers due to a liquid-like structure (Liu et al., 1994; Koinkar and
Bhushan, 1996; Bouhacina et al., 1997). Additionally, it should be mentioned that there have been reports where
even on crystalline surfaces, a slight logarithmic dependence of the friction on the sliding velocity was observed by
SFM (Liu et al., 1994; Koinkar and Bhushan, 1996).
2001 by CRC Press LLC
parameters, i.e., stick-slips will occur), m
x
= 10
10
kg (with this effective mass, the system [ tip and sample
in contact] has a typical resonance frequency of 1/(2) 50 kHz, which is typical for SFM
experiments), and
x
= 2 (critical damping). With this set of parameters, we calculate the numerical
solutions of Equation 18.11 and determine the lateral forces F
x
as well as the frictional force F
f
for different
sliding velocities (
M
= 10 nm/s 100 m/s).
In Figure 18.18, the calculated lateral force F
x
is displayed for different sliding velocities
M
. It is obvious
that the obtained results change only little within the wide range of sliding velocities
M
= 10 nm/s
10 m/s. Only for high sliding velocities, where the solution of Equation 18.11 is dominated by the
velocity-dependent damping term (see F
x
for 100 m/s in Figure 18.18), the lateral force increases
signicantly. Consequently, since the friction force is the average of the lateral force, F
f
is approximately
independent of the sliding velocity in a wide range.
What is now the reason for the relative independence of the lateral forces on the sliding velocities at
moderate values of
M
?This issue is illustrated in Figure 18.11b. We have seen above that the tip usually
moves in a stick-slip-type motion over the sample surface. With this type of motion, the tip stays in the
minima of the interaction potential most of the time where it slides very slowly or sticks. Therefore,
almost no energy is dissipated in this stick-state, since we assumed that the damping is proportional
to the sliding velocity. In the slip-state, however, the tip jumps from one to another minimum. During
this jump, the tip reaches very high peak velocities and dissipates signicant amounts of energy due to
the velocity-dependent damping mechanism.
It is now important to note that the maximum velocity reached during the slip depends for low values
of
M
in rst approximation only on the spring constant c
x
, the height and the shape of the potential,
the chosen damping constant, and the effective mass of the system, but not on the sliding velocity
M
.
As a consequence, the total amount of energy dissipated during sliding does not change signicantly as
long as
M
is much smaller than the slip velocity of the tip. From Figure 18.18, we obtain a slip velocity
of about 60 m/s for sliding speeds up to 10 m/s.* For larger sliding velocities, the mechanism of energy
dissipation through the stick-slip effects breaks down, as can be seen for
M
= 100 m/s.
However, despite its success, the model introduced above is still unsatisfactory from an atomistic view,
since it is fully phenomenological in the way that it does not give any insight into the dissipation
mechanism involved. It is clear that the energy dissipation process will be much more complicated in a
real physical system than that considered here; more complete theories can be found in, for example,
Sokoloff (1990, 1993) or Colchero et al. (1996).
FIGURE 18.17 The friction forces F
f
as a function of the sliding velocity
M
measured with a silicon tip on
(a) amorphous carbon, and (b) diamond. F
f
is independent of the sliding velocity in good approximation.
*The numerical value of the slip velocity obtained within this simple mechanical model is determined by the one-
dimensional interaction potential and the chosen numerical values of the parameters, which partially might vary in
a very large range without changing the obtained results qualitatively. Consequently, the slip velocity of 60 m/s is
not a prediction for an experimental measurement and could actually be much higher without contradiction of the
general mechanism discussed here.
c
x
m
x

c
x
m
x

2001 by CRC Press LLC


18.8 Summary
In this chapter, we have reviewed some of our present understanding of atomic-scale friction gained from
scanning force microscopy experiments. Our analysis, however, has always been restricted to fundamental
aspects of wearless Coulomb friction (i.e., dry friction); effects due to boundary lubrication and wetting
of surfaces were not considered.
It has been shown that the main advantage of the SFM in this context is its ability to enable detailed
studies of the friction of point contacts (which have been found to exhibit, in reality, contact areas of
some nm
2
, depending on the applied load, the elastic properties of tip and sample, and the tip radius)
with high spatial resolution as a function of the contact geometry, the applied load, the effective contact
area, the sliding velocity, and the surface structure of the proled sample. Remarkably, it has been found
that some phenomena manifest similarly on the nanometer scale compared to their macroscopic behavior,
whereas others show a very different behavior.
Most strikingly, friction on the nanometer scale is proportional to the effective contact area, contrary
to the macroscopic case. Consequently, the concept of the friction coefcient, which is independent of
the actual contact geometry, fails on this scale. In the corresponding experiments, contact pressures of
up to 1 GPa are easily reached. Such pressures are already above the yield stress of many materials and
therefore seem very high at rst sight, but might match the realistic pressures of individual nanocontacts
within a macroscopical contact far better than lower contact pressures. However, despite these high
FIGURE 18.18 The lateral force F
x
and the tip velocity
t
for different sliding velocities
M
= 10 nm/s 100 m/s.
2001 by CRC Press LLC
pressures and the small dimensions of the individual contact with only some tens or some hundreds of
atoms in intimate contact, contact mechanics, which has been derived from continuum elasticity theory,
has proved to be still valid.
Another interesting point is that the shear stress is independent of the mean contact pressure at
least in the case of nanometer-sized Hertzian or approximately Hertzian contacts. Again, very high shear
stresses of up to 1 GPa are found, orders of magnitude more than typically obtained from other friction
experiments. These high values can possibly be accounted for as representing approximately the theoretical
shear stress of the contact, which is about
1
/30 of the shear modulus of the contact (Cottrell, 1953) and
therefore has about the correct order of magnitude.
Much more difcult to answer is the question of which factors determine whether a given contact
shows high or low friction. In many experiments, it has been demonstrated that friction is very sensitive
to even small changes in the surface structure and might additionally show a strong dependence on the
sliding direction relative to the crystal lattice. This feature has been used frequently to distinguish between
chemically different areas of samples or to obtain a contrast between individual domains that differ due
to structural effects, but has not yet resulted in signicant progress in our understanding of why certain
surfaces have higher friction than others.
Nevertheless, studying such issues, the atomic-scale movement of the individual atoms which are in
intimate contact has been claried to a certain degree. By comparison with theoretical models, it can be
shown that the tip atoms move in a stick-slip-type motion over the sample surface, jumping from
one minima of the tipsample interaction potential to the next. Since they try to avoid passing over the
position of a sample atom which represents a maximum of the interaction potential, their paths often
resemble a zigzag curve. Due to this specic behavior, it follows that the atomically resolved images of
contact force microscopy represent solely the periodicity of the potential minima of the interaction
potential, which does not generally match the atomic structure of a sample for nontrivial unit cells.
Another intention of this chapter was to bridge the gap between the macroscopic and the micro-
scopic/nanoscopic scale. As we have seen in the introduction, it is one of the basic aims of nanotribology
to explain the macroscopic frictional behavior of materials and contacts with fundamental processes
occurring on the atomic scale. And it seems indeed possible to derive Amontons and Coulombs laws
of friction from nanoscopic effects. Greenwood (1967, 1992) showed that the effective contact area
between macroscopically at bodies, which nevertheless exhibit a statistical microscopical roughness,
increases linearly with the load, which eliminates any contradiction between the corresponding macro-
scopic and the nanoscopic friction laws. Moreover, the relative independence of friction on the sliding
velocity on both the macroscopic and the nanoscopic scale might be explained with the stick-slip-type
motion of the atoms in the tipsample interaction potential. As long as the relative velocity of the two
bodies is much lower than the intrinsic slip velocity of the atoms, the dissipated energy will be
independent of variations in the sliding velocity.
To conclude, nanotribology is a young eld which has developed rapidly during the last one or two
decades. Considerable progress has been made during this time in our understanding of fundamental
frictional processes on the nanometer scale, as we hope to have demonstrated above, and rst attempts
to connect the atomic-scale effects with macroscopic phenomena seem to be successful. However, in spite
of this progress, there are still many more open than answered questions. For example, it is still impossible
with our present knowledge to predict the frictional properties of a specic contact from rst principles.
Nevertheless, the optimization of the frictional properties of surfaces for a given purpose will be increas-
ingly important in the future for the further miniaturization of motors and other moving parts in, for
example, hard discs or micro electromechanical systems. A further unsolved question is how energy is
dissipated during wearless friction. It seems that the excitation of phonons represents the main channel
for energy dissipation, even though electronic processes might be equally important in the case of
conducting materials. Thus, the exact control of the dissipation channels might be a prerequisite for the
tailoring of surfaces with given frictional properties and simultaneously offers plenty of room for
original and innovative nanotribological research.
2001 by CRC Press LLC
References
Aim, J.P., Elkaakour, Z., Gauthier, S., Michel, D., Bouhacina, T., and Curly, J. (1995), Role of the force
of friction on curved surfaces in scanning force microscopy, Surf. Sci., 329, 149-156.
Bhushan, B., Koinkar, V.N., and Ruan, J. (1994), Microtribology of magnetic media, Proc. Inst. Mech.
Part J: J. Eng. Tribol., 208, 17-29.
Bhushan, B. and Ruan, J. (1994), Atomic-scale friction measurements using friction force microscopy:
Part I General principles and new measurement techniques, ASME J. Tribol., 116, 378-388.
Bhushan, B. and Kulkarni, A.V. (1995), Effect of normal load on microscale friction measurements, Thin
Solid Films, 278, 49-56.
Binh, V.T. and Garcia, N. (1992), On the electron and metallic ion emission from nanotips fabricated by
eld-surface-melting technique: experiments on W and Au tips, Ultramicroscopy, 42-44, 80-90.
Binh, V.T. and Uzan, R. (1987), Tip-shape evolution: capillary induced matter transport by surface
diffusion, Surf. Sci., 179, 540-560.
Binnig, G., Quate, C.F., and Gerber, Ch. (1986), Atomic force microscope, Phys. Rev. Lett., 56, 930-933.
Bluhm, H., Schwarz, U.D., Meyer, K.-P., and Wiesendanger, R. (1995), Anisotropy of sliding friction on
the triglycine sulfate (010) surface, Appl. Phys. A, 61, 525-533.
Bluhm, H., Schwarz, U.D., and Wiesendanger, R. (1998), Origin of the ferroelectric domain contrast
observed in lateral force microscopy, Phys. Rev. B, 57, 161-169.
Bouhacina, T., Aim, J.P., Gauthier, S., Michel, D., and Heroguez, V. (1997), Tribological behavior of a
polymer grafted on silanized silica probed with a nanotip, Phys. Rev. B, 56, 7694-7703.
Briscoe, B.J. and Evans, D.C.B. (1982), The shear properties of LangmuirBlodgett layers, Proc. R. Soc.
Lond. A, 380, 389-407.
Carpick, R.W., Agrait, N., Ogletree, D.F., and Salmeron, M. (1996), Measurement of the interfacial shear
(friction) with an ultrahigh vacuum atomic force microscope, J. Vac. Sci. Technol. B, 14, 1289-1295.
Carpick, R.W., Ogletree, D.F., and Salmeron, M. (1997), Lateral stiffness: a new nanomechanical mea-
surement for the determination of shear strengths with friction force microscopy, Appl. Phys. Lett.,
70, 1548-1550.
Colchero, J., Bar, A.M., and Marti, O. (1996), Energy dissipation in scanning force microscopy friction
on an atomic scale, Tribol. Lett., 2, 327-343.
Cottrell, A.H. (1953), Dislocations and Plastic Flow in Crystals, Oxford University Press, Oxford, U.K.
Derjaguin, B.V., Muller, V.M., and Toporov, Y.P. (1975), Effect of contact deformations on the adhesion
of particles, J. Colloid Interface Sci., 53, 314-326.
Enachescu, M., van den Oetelaar, R.J.A., Carpick, R.W., Ogletree, D.F., Flipse, C.F.J., and Salmeron, M.
(1998), Atomic force microscopy study of an ideally hard contact: the diamond(111)/tungsten
carbide interface, Phys. Rev. Lett., 81, 1877-1880.
Fogden, A. and White, L.R. (1990), Contact elasticity in the presence of capillary condensation, J. Colloid
Interface Sci., 138, 414-430.
Fompeyrine, J., Berger, R., Lang, H.-P., Perret, J., Mchler, E., Gerber, Ch., and Locquet, J.-P. (1998),
Local determination of the stacking sequence of layered materials, Appl. Phys. Lett., 72, 1697-1699.
Frisbie, C.D., Rozsnyai, L.F., Noy, A., Wrighton, M.S., and Lieber, C.M. (1994), Functional group imaging
by chemical force microscopy, Science, 265, 2071-2074.
Fujihara, M. and Ohzono, T. (1999), Basis of chemical force microscopy by friction: energetics and
dynamics of wearless friction between organic monolayers in terms of chemical and physical
properties of molecules, Jpn. J. Appl. Phys., 38, 3918-3931.
Fujisawa, S., Sugawara, Y., Ito, S., Mishima, S., Okada, T., and Morita, S. (1993a), The two-dimensional
stick-slip phenomenon with atomic resolution, Nanotechnology, 4, 138-142.
Fujisawa, S., Sugawara, Y., and Morita, S. (1993b), Origins of the forces measured by atomic force/lateral
force microscope, Microbeam Analysis, 2, 311-316.
Fujisawa, S., Kishi, E., Sugawara, Y., and Morita, S. (1994), Fluctuation in the two-dimensional stick-slip
phenomenon observed with two-dimensional frictional microscope, Jpn. J. Appl. Phys., 33, 3752-3755.
2001 by CRC Press LLC
Fujisawa, S., Kishi, E., Sugawara, Y., and Morita, S. (1995), Atomic-scale friction observed with a two-
dimensional frictional-force microscope, Phys. Rev. B, 51, 7849-7857.
Gourdon, D., Burnham, N.A., Kulik, A., Dupas, E., Oulevey, F., Gremaud, G., Stamou, D., Liley, M.,
Dienes, Z., Vogel, H., and Duschl, C. (1997), The dependence of friction anisotropies on the
molecular organisation of LB lms as observed by AFM, Tribol. Lett., 3, 317-324.
Grafstrm, S., Neitzert, M., Hagen, T., Ackermann, J., Neumann, R., Probst, O., and Wrtge, M. (1993),
The role of topography and friction for the image contrast in lateral force microscopy, Nanotech-
nology, 4, 143-151.
Greenwood, J.A. (January 1967), The area of contact between rough surfaces and ats, J. Lub. Technol., 81-91.
Greenwood, J.A. (1992), Contact of rough surfaces, in Fundamentals of Friction: Macroscopic and Micro-
scopic Processes, Singer, I. L. and Pollock, H.M. (Eds.), Kluwer Academic Publishers, Dordrecht,
The Netherlands, 37.
Gyalog, T., Bammerlin, M., Lthi, R., Meyer, E., and Thomas, H. (1995), Mechanism of atomic friction,
Europhys. Lett., 31, 269-274.
Harrison, J.A., Colton, R.J., White, C.T., and Brenner, D.W. (1993a), Effect of atomic-scale surface
roughness on friction: a molecular dynamics study of diamond surfaces, Wear, 168, 127-133.
Harrison, J.A., White, C.T., Colton, R.J., and Brenner, D.W. (1993b), Effects of chemically-bound, exible
hydrocarbon species on the frictional properties of diamond surfaces, J. Chem. Phys., 97, 6573-6576.
Helman, J.S., Baltensberger, W., and Holyt, J.A. (1994), Simple model for dry friction, Phys. Rev. B, 49,
3831-3838.
Hertz, H. (1881), ber die Berhrung fester elastischer Krper, J. Reine Angew. Math., 92, 156-171.
Hlscher, H., Schwarz, U.D., and Wiesendanger, R. (1997), Modelling of the scan process in lateral force
microscopy, Surf. Sci., 375, 395-402.
Hlscher, H., Schwarz, U.D., Zwrner, O., and Wiesendanger, R. (1998), Consequences of the stick-slip
movement for the scanning force microscopy imaging of graphite, Phys. Rev. B, 57, 2477-2481.
Homola, A.M., Israelachvili, J.N., McGuiggan, P.M., and Gee, M.L. (1990), Fundamental experimental
studies in tribology: the transition from interfacial friction of undamaged molecularly smooth
surfaces to normal friction with wear, Wear, 136, 65-83.
Hu, J., Xiao, X.-D., Ogletree, D.F., and Salmeron, M. (1995), Atomic scale friction and wear of mica, Surf.
Sci., 327, 358-370.
Israelachvili, J.N. and Tabor, D. (1973), The shear properties of molecular lms, Wear, 24, 386-390.
Israelachvili, J.N., Perez, E., and Tandon, R.K. (1980), On the adhesion force between deformable surfaces,
J. Colloid Interface Sci., 78, 260-261.
Johnson, K.L. (1994), Contact Mechanics, Cambridge University Press, Cambridge, U.K.
Johnson, K.L. (1997), Adhesion and friction between a smooth elastic spherical asperity and a plane
surface, Proc. R. Soc. Lond. A, 453, 163-179.
Johnson, K.L., Kendall, K., and Roberts, A.D. (1971), Surface energy and the contact of elastic solids,
Proc. R. Soc. London A, 324, 301-313.
Kawakatsu, H. and Saito, T. (1996), Scanning force microscopy with two optical levels for detection of
deformations of the cantilever, J. Vac. Sci. Technol. B, 14, 872-876.
Koinkar, V.N. and Bhushan, B. (1996), Microtribological studies of unlubricated and lubricated surfaces
using atomic force/friction force microscopy, J. Vac. Sci. Technol. A, 14, 2378-2391.
Landman, U., Luedtke, W.D., and Nitzan, A. (1989), Dynamics of tipsubstrate interactions in atomic
force microscopy, Surf. Sci., 210, L177-L184.
Lantz, M.A., OShea, S.J., Welland, M.E., and Johnson, K.L. (1997), Atomic-force-microscope study of
contact area and friction on NbSe
2
, Phys. Rev. B, 55, 10776-10785.
Liley, M., Gourdon, D., Stamou, D., Meseth, U., Fischer, T.M., Lautz, C., Stahlberg, H., Vogel, H.,
Burnham, N.A., and Duschl, C. (1998), Friction anisotropy and asymmetry of a compliant mono-
layer induced by a small molecular tilt, Science, 280, 273-275.
Liu, Y., Wu, T., and Evans, D.F. (1994), Lateral force microscopy study on the shear properties of self-
assembled monolayers of dialkylammonium surfactant on mica, Langmuir, 10, 2241-2245.
2001 by CRC Press LLC
Lthi, R., Meyer, E. Haefke, H., Howald, L., Gutmannsbauer, W., Guggisberg, M., Bammerlin, M., and
Gntherodt, H.-J. (1995), Nanotribology: an UHVSFM study on thin lms of C
60
and AgBr, Surf.
Sci., 338, 247-260.
Marti, O., Colchero, J., and Mlynek, J. (1990), Combined scanning force and friction microscopy of mica,
Nanotechnology, 1, 141-144.
Marti, O., Colchero, J., and Mlynek, J. (1993), Friction and forces on the atomic scale, in Nanosources
and Manipulation of Atoms Under High Fields and Temperatures: Applications, Binh V.T. et al. (Eds.)
Kluwer Academic Publishers, Dordrecht, The Netherlands, 253.
Mate, C.M., McClelland, G.M., Erlandsson, R., and Chiang, S. (1987), Atomic-scale friction of a tungsten
tip on a graphite surface, Phys. Rev. Lett., 59, 1942-1945.
Mate, C.M. (1993), Nanotribiology studies of carbon surfaces by force microscopy, Wear, 168, 17-20.
Maugis, D. (1987), Adherence of elastomers: fracture mechanics aspects, J. Adhesion Sci. Technol., 1, 105-134.
Maugis, D. (1992), Adhesion of spheres: the JKRDMT transition using a Dugdale model, J. Colloid
Interface Sci., 73, 294-269.
Maugis, D. and Gauthier-Manuel, B. (1994), JKRDMT transition in the presence of a liquid meniscus,
J. Adhesion Sci. Technol., 8, 1311-1322.
McClelland, G.M. (1989), Friction at weakly interacting interfaces, in Adhesion and Friction, Grunze, M.
and Kreuzer, H.J. (Eds.), Springer Series in Surface Science, Vol. 17, Springer Verlag, Heidelberg, 1.
Meyer, E., Overney, R., Brodbeck, D., Howald, L., Lthi, R., Frommer, J., and Gntherodt, H.-J. (1992),
Friction and wear of LangmuirBlodgett lms observed by friction force microscopy, Phys. Rev.
Lett., 69, 1777-1780.
Meyer, E. and Heinzelmann, H. (1992), Scanning force microscopy, in Scanning Tunneling Microscopy II,
Wiesendanger, R. and Gntherodt, H.-J. (Eds.), Springer Series in Surface Science, Vol. 28,
Springer-Verlag, Heidelberg, 99.
Meyer, E., Lthi, R., Howald, L., Bammerlin, M., Guggisberg, M., and Gntherodt, H.-J. (1996), Site-
specic friction force microscopy, J. Vac. Sci. Technol. B, 14, 1285-1288.
Meyer, G. and Amer, N.M. (1990), Simultaneous measurement of lateral and normal forces with optical-
beam-deection atomic force microscope, Appl. Phys. Lett., 57, 2089-2091.
Muller, V.M., Yuschenko, V.S., and Derjaguin, B.V. (1983), General theoretical consideration of the
inuence of surface forces on contact deformations and the reciprocal adhesion of elastic spherical
particles, J. Colloid Interface Sci., 92, 92-101.
Ogletree, D.F, Carpick, R.W., and Salmeron, M. (1996), Calibration of frictional forces in atomic force
microscopy, Rev. Sci. Instrum., 67, 3298-3306.
Overney, R.M., Meyer, E., Frommer, J., Brodbeck, D., Lthi, R., Howald, L. Gntherodt, H.-J., Fujihira,
M., Takano, H., and Gotoh, Y. (1992), Friction measurements on phase-separated thin lms with
a modied atomic force microscopy, Nature, 359, 133-135.
Overney, R.M., Takano, H., Fujihira, M., Paulus, W., and Ringsdorf, H. (1994), Anisotropy in friction
and molecular stick-slip motion, Phys. Rev. Lett, 72, 3546-3549.
Prandtl, L. (1928), Ein Gedankenmodell zur kinetischen Energie der festen Krper, Z. Angew. Math.
Mech., 8, 85-106.
Putman, C., Igarashi, M., and Kaneko, R. (1995), Quantitative determination of friction coefcients by
friction force microscopy, Jpn. J. Appl. Phys., 34, L264-L267.
Ruan, J. and Bhushan, B. (1994a), Atomic-scale and microscale friction studies of graphite and diamond
using friction force microscopy, J. Appl. Phys., 76, 5022-5035.
Ruan, J. and Bhushan, B. (1994b), Frictional behavior of highly oriented pyrolytic graphite, J. Appl. Phys.,
76, 8117-8120.
Sasaki, K., Koike, Y., Azehara, H., Hokari, H., and Fujihara, M., (1998), Lateral force microscope and
phase imaging of patterned thiol self-assembling monolayer using chemically modied tips, Appl.
Phys. A, 66, S1275-S1277.
Schluger, A.L., Livshits, A.I., Foster, A.S., and Catlow, C.R.A. (1999), Models of image contrast in scanning
force microscopy on insulators, J. Phys: Condens, Matter, 11, R295-R322.
2001 by CRC Press LLC
Schwarz, U.D. (1997), Scanning force microscopy, in Handbook of Microscopy II, Amelinckx, S., Van Dyck,
D., Van Landuyt, J.F., and Van Tendeloo, G., (Eds.), VCH Verlagsgesellschaft, Weinheim, Germany,
827.
Schwarz, U.D., Kster, P., and Wiesendanger, R. (1996a), Quantitative analysis of lateral force microscopy
experiments, Rev. Sci. Instrum., 67, 2560-2567.
Schwarz, U.D., Bluhm, H., Hlscher, H., Allers, W., and Wiesendanger, R. (1996b), Friction in the low-
load regime: studies on the pressure and direction dependence of frictional forces by means of
friction force microscopy, in Physics of Sliding Friction, Persson, B.N.J. and Tosatti, E. (Eds.), Kluwer
Academic Publishers, Dordrecht, The Netherlands, 369.
Schwarz, U.D., Zwrner, O., Kster, P., and Wiesendanger, R. (1997a), Quantitative analysis of the frictional
properties of solid materials at low loads. I. Carbon compounds, Phys. Rev. B, 56, 6987-6996.
Schwarz, U.D., Zwrner, O., Kster, P., and Wiesendanger, R. (1997b), Preparation of probe tips with
well-dened spherical apexes for quantitative scanning force spectroscopy, J. Vac. Sci. Technol. B,
15, 1527-1530.
Shimizu, J., Eda, H., Yoritsune, M., and Ohmura, E. (1998), Molecular dynamics simulation of friction
on the atomic scale, Nanotechnology, 9, 118-123.
Shindo, H., Shitagami, T., and Kondo, S.-I. (1999), Evidence of the contribution of molecular orientations
on the surface force friction of alkaline earth sulfate crystals, Phys. Chem. Chem. Phys., 1, 1597-1600.
Sokoloff, J.B. (1990), Theory of energy dissipation in sliding crystal surfaces, Phys. Rev. B, 42, 760-765.
Sokoloff, J.B. (1993), Fundamental mechanisms for energy dissipation at small solid sliding surfaces,
Wear, 167, 59-68.
Tabor, D. (1977), Surface forces and surface interactions, J. Colloid Interface Sci., 58, 2-13.
Tomlinson, G.A. (1929), A molecular theory of friction, Philos. Mag. S., 7, 7, 905-939.
Wiesendanger, R. (1994), Scanning Probe Microscopy and Spectroscopy: Methods and Applications, Cam-
bridge University Press, Cambridge, U.K.
Zwrner, O., Hlscher, H., Schwarz, U.D., and Wiesendanger, R. (1998), The velocity dependence of
frictional forces in point-contact friction, Appl. Phys. A, 66, S263-S267.
2001 by CRC Press LLC

Вам также может понравиться