Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/262989024

Theory of adhesion: Role of surface roughness

Article  in  The Journal of Chemical Physics · September 2014


DOI: 10.1063/1.4895789

CITATIONS READS

52 583

2 authors, including:

Michele Scaraggi
Università del Salento
56 PUBLICATIONS   605 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Lubrication of smart textured surfaces View project

Lubrication of polymers View project

All content following this page was uploaded by Michele Scaraggi on 06 July 2014.

The user has requested enhancement of the downloaded file.


Theory of adhesion: role of surface roughness

B.N.J. Persson1 and M. Scaraggi1,2


1
PGI, FZ-Jülich, 52425 Jülich, Germany, EU and
2
DII, Universit del Salento, 73100 Monteroni-Lecce, Italy, EU

We discuss how surface roughness influence the adhesion between elastic solids. We introduce
a Tabor number which depends on the length scale or magnification, and which gives information
about the nature of the adhesion at different length scales. We consider two limiting cases relevant
for (a) elastically hard solids with weak adhesive interaction (DMT-limit) and (b) elastically soft
solids or strong adhesive interaction (JKR-limit). For the former cases we study the nature of
the adhesion using different adhesive force laws (F ∼ u−n , n = 1.5 − 4, where u is the wall-wall
separation). In general, adhesion may switch from DMT-like at short length scales to JKR-like
at large (macroscopic) length scale. We compare the theory predictions to the results of exact
numerical simulations and find good agreement between theory and the simulation results.

1 Introduction teraction, which we also find is important in the DMT


Surface roughness has a huge influence on the adhe- limit and when the surface roughness amplitude is small
sion and friction between macroscopic solid objects[1–6]. (see below). Medina and Dini[19] studied the adhesion
Most interaction force fields are short ranged and be- between an elastic sphere with smooth surface and a rigid
comes unimportant when the separation between solid randomly rough substrate surface. They observed strong
surfaces exceed a few atomic distances, i.e., at separa- contact hysteresis in the JKR-limit (relative smooth sur-
tions of order nm. This is trivially true for chemical faces) and very small contact hysteresis in the DMT-limit
bonds (covalent or metallic bonds) but holds also for which prevails for small roughness. Analytical theories of
the more long-ranged Van der Waals interaction. One contact mechanics have been compared to numerically
important exception is charged bodies. For uncharged exact calculations for two-dimensional (2D) randomly
solids, if the surface roughness amplitude is much larger rough surfaces in Ref. [20] and for 1D surface rough-
than the decay length of the wall-wall interaction po- ness in Ref. [21]. Experimental adhesion data for rough
tential and if the solids are elastically stiff enough, no surfaces have been compared to analytical theory predic-
macroscopic adhesion will prevail, as is the case in most tions in Ref. [22], [23] and [24].
practical cases. Only for very smooth surfaces, or elastic- Many practical or natural adhesive systems involve ef-
ally very soft solids (which can deform and make almost fects which usually are not considered in model stud-
perfect contact at the contacting interface without stor- ies of adhesion, and which we will not address in this
ing up a large elastic energy) adhesion will be observed paper. In particular, biological applications typically
for macroscopic solids[7]. involve complex structured surfaces (e.g., hierarchical
In this paper we will discuss how surface roughness fiber-and-plate structures) with anisotropic elastic prop-
influence adhesion between macroscopic solids. We con- erties, which are elastically soft on all relevant length
sider two limiting cases, which are valid for elastically scales[25–28]. Instead of directly relying on molecular
hard and weakly interaction solids (Deryagin, Muller, bonding over atomic dimension, many biological sys-
and Toporov, DMT-limit)[8] and for elastically soft tems adhere mainly via capillary bridges[29–31]. We will
or strongly interacting solids (Johnson, Kendall, and also not discuss either the adhesion between charged ob-
Roberts, JKR-limit)[9]. This problem has been studied jects, which must be treated by special methods which
before but usually using the Greenwood-Williamson[10, takes into account the long-range nature of the Coulomb
11] type of asperity models (see, e.g., [7, 12]), whereas interaction[32–35].
our treatment is based on the Persson contact mechanics In this paper we first briefly review (Sec. 2) two lim-
model. The latter model is (approximately) valid even iting models of adhesion for smooth surfaces. In Sec. 3
close to complete contact (which often prevail when ad- we show how the same limiting cases can be studied ana-
hesion is important)[13, 14]. Asperity models can only lytically for randomly rough surfaces using the Persson
be used as long as the contact area is small compared contact mechanics model. Numerical results obtained us-
to the nominal contact area, and even in this limit these ing the analytical theory are presented in Sec. 4, and
models have severe problems for surfaces with roughness compared to exact numerical results in Sec. 5. Sec. 6
on many length scales[15–17]. contains a discussion and Sec. 7 the summary and con-
Recently several numerical studies of adhesion between clusion.
randomly rough surfaces have been published. Pastewka
and Robbins[18] study the adhesion between rough sur- 2 Adhesion of ball on flat (review)
faces and present a criterion for macroscopic adhesion. Analytical studies of adhesion have been presented for
They emphasize the role of the range of the adhesive in- smooth surfaces for bodies of simple geometrical shape,
2

FN The JKR theory neglects the extend of the interaction


(a) FN (b) potential and assumes interaction between the solids only
in the contact area. The deformation field in the JKR
theory is obtained by minimizing the total energy given
by the sum of the (repulsive) elastic deformation energy
and the (attractive) binding energy Ead = ∆γA, where
elastic ball A is the contact area. In this theory the contribution to
binding energy from the non-contact region is neglected.
Since σad ≈ ∆γ/dc we can define the Tabor number:
dT 1/3
R∆γ 2

σad dT
µT = = = ,
σc Er2 d3c dc
DMT-limit JKR-limit
where
Figure 1: (a) In the DMT theory the elastic deformation field 1/3
R∆γ 2

is calculated with the adhesion included only as an additional dT = ,
load Fad acting on the sphere. Thus the contact area is de- Er2
termined by Hertz theory with the external load FN + Fad .
The adhesional load Fad is obtained by integrating the adhe- where Er = E/(1 − ν 2 ) is effective elastic modulus. The
sional stress over the ball non-contact area. (b) In the JKR DMT and JKR limits correspond to µT << 1 and µT >>
theory the adhesion force is assumed to have infinitesimal spa- 1, or, equivalently, dT << dc and dT >> dc , respectively.
tial extent, and is included only in the contact area as an in- In the JKR-limit the Tabor length dT can be considered
terfacial binding energy Ead = ∆γA. The shape of the elastic
as the height of the neck which is formed at the contact
body is obtained by minimizing the total energy −Ead + Uel ,
where Uel is the elastic deformation energy. line (see Fig. 1(b)). This neck height must be much
larger than the length dc , which characterizes the spatial
extend of the wall-wall interaction, in order for the JKR-
the most important case being the contact between limit to prevail.
spherical bodies. For a sphere in contact with a flat At vanishing external load, FN = 0, the JKR theory
surface two limiting cases are of particular importance, predicts the contact area:
usually referred to as the DMT theory[8] and the JKR 2/3
9πR2 ∆γ

theory[9], see Fig. 1. Analytical results for intermediate- AJKR = π .
range adhesion was presented by Maugis[36, 37] and the 2Er
ball-flat adhesion problem has also been studied in detail
using numerical methods[38, 39]. A particular detailed This contact area is a factor 32/3 ≈ 2.1 larger than ob-
numerical study was recently published by Müser who tained from the DMT theory. In the JKR theory the
also included negative work of adhesion (repulsive wall- force necessary to remove the ball from the flat (the pull-
wall interaction)[40]. off force) is given by
Consider an elastic ball (e.g., a rubber ball) with the 3π
radius R, Young’s elastic modulus E (and Poisson ratio Fc = ∆γR (1)
2
ν), in adhesive contact with a flat rigid substrate. Let
∆γ = γ1 + γ2 − γ12 be the work of adhesion and let dc be which is a factor of 3/4 times smaller than predicted by
the spatial extend of the wall-wall interaction potential the DMT theory. Also the pull-off process differs: in
(typically of order atomic distance). The DMT theory is the JKR theory an elastic instability occurs where the
valid when adhesive stress σad ≈ ∆γ/dc is much smaller contact area abruptly decreases, while in the DMT theory
than the stress in the contact region, which is of order the contact area decreases continuously, until the ball just
touches the substrate in a single point, at which point the
1/3
∆γE 2 pull-force is maximal.

σc ≈ . For the sphere-flat case the pull-off force in the DMT-
R
limit is independent of the range of the wall-wall inter-
In the opposite limit the JKR theory is valid. In the action potential. However, this is not the case for other
DMT theory the elastic deformation field is calculated geometries where in fact the contact mechanics depends
with the adhesion included only as an additional load Fad remarkably sensitively on the interaction range. As a res-
acting on the sphere. Thus the contact area is determined ult the interaction between rough surfaces in the DMT-
by Hertz theory with the external load F0 = FN + Fad , limit will depend on the force law as we will demonstrate
where FN is the actual load on the ball (see Fig. 1). The below for power law interaction pad ∼ u−n .
adhesion load Fad is obtained by integrating the adhesion In an exact treatment, as a function of the external
stress over the ball non-contact area. load FN , the total energy Etot = −Ead + Uel must have
3

a minimum at FN = 0. This is the case in the JKR the- the probability distribution P (σ, ζ) of stresses σ acting at
ory but in general not for the DMT theory. However, the interface when the system is studied at the magnific-
the DMT theory is only valid for very stiff solids and ation ζ. In Ref. [41] an approximate diffusion equation
in this limiting case the total energy minimum condi- of motion was derived for P (σ, ζ). To solve this equa-
tion is almost satisfied. Nevertheless, one cannot expect tion one needs boundary conditions. If we assume that,
dEtot /dFN (FN = 0) = 0 to be exactly obeyed in any (ap- when studying the system at the lowest magnification
proximate) theory which does not focus on minimizing ζ = 1 (where no surface roughness can be observed, i.e.,
the total energy. the surfaces appear perfectly smooth), the stress at the
The results above assume perfectly smooth surfaces. interface is constant and equal to pN = FN /A0 , where
The JKR (and DMT) theory results can, however, be FN is the load and A0 the nominal contact area, then
applied also to surfaces with roughness assuming that P (σ, 1) = δ(σ − pN ). In addition to this “initial condi-
the wavelength λ of the most longest (relevant) surface tion” we need two boundary conditions along the σ-axis.
roughness component is smaller than the diameter of the Since there can be no infinitely large stress at the inter-
contact region. In that case one only needs to replace face we require P (σ, ζ) → 0 as σ → ∞. For adhesive
the work of adhesion ∆γ for flat surfaces with an effective contact, which interests us here, tensile stress occurs at
work of adhesion γeff obtained for the rough surfaces. We the interface close to the boundary lines of the contact
will now describe how one may calculate γeff . regions. In this case we have the boundary condition
P (−σa , ζ) = 0, where σa > 0 is the largest (locally av-
3 Theory: basic equations eraged at magnification ζ) tensile stress possible. Hence,
We now show how surface roughness can be taken into the detachment stress σa (ζ) depends on the magnifica-
account in adhesive contact mechanics. We consider two tion and can be related to the effective interfacial energy
limiting cases similar to the JKR and DMT theories for (per unit area) γeff (ζ) using the theory of cracks[6]. The
adhesion of a ball on a flat. The theory presented below is effective interfacial binding energy
not based on the standard Greenwood-Williamson[10, 11]
picture involving contact between asperities, but on the γeff (ζ)A(ζ) = ∆γA(ζ1 )η − Uel (ζ),
Persson contact mechanics theory.
where A(ζ) denotes the (projected) contact area at the
3.1 JKR-limit magnification ζ, and A(ζ1 )η is the real contact area,
which is larger than the projected contact area A(ζ1 ),
In the JKR-limit the spatial extend of the wall-wall in-
i.e. η ≥ 1 (e.g. if the rigid solid is rough and the elastic
teraction potential is neglected so the interaction is fully
solid has a flat surface η > 1, see Ref. [43] for an ex-
characterized by the work of adhesion ∆γ.
pression for η). Uel (ζ) is the elastic energy stored at the
In order for two elastic solids with rough surfaces to
interface due to the elastic deformation of the solids on
make adhesive contact it is necessary to deform the sur-
length scale shorter than λ = L/ζ, necessary in order to
faces elastically, otherwise they would only make con-
bring the solids into adhesive contact.
tact in three points and the adhesion would vanish, at
The area of apparent contact (projected on the xy-
least if the spatial extend of the adhesion force is neg-
plane) at the magnification ζ, A(ζ), normalized by the
lected. Deforming the surfaces to increase the contact
nominal contact area A0 , can be obtained from
area A results in some interfacial bonding −∆γA (where
Z ∞
∆γ = γ1 + γ2 − γ12 is the change in the interfacial en- A(ζ)
ergy per unit area upon contact), but it costs elastic = dσ P (σ, ζ)
A0 −σa (ζ)
deformation energy Uel , which will reduce the effective
binding. That is, during the removal of the block from Finally, we note that the effective interfacial energy to
the substrate the elastic compression energy stored at be used in the JKR expression for the pull-off force (1) is
the interface is given back and helps to break the adhes- the macroscopic effective interfacial energy correspond-
ive bonds in the area of real contact. Most macroscopic ing to the magnification ζ = 1 (here we assume that the
solids do not adhere with any measurable force, which reference length L is of order the diameter of the JKR
implies that the total interfacial energy −∆γA + Uel van- contact region). Thus in the numerical results presented
ishes, or nearly vanishes, in most cases. in Sec. 4 we only study the area of contact A(ζ1 ) and
The contact mechanics theory of Persson[6, 41–47] can the macroscopic interfacial energy γeff = γeff (1), which
be used to calculate (approximately) the stress distribu- satisfies
tion at the interface, the area of real contact and the
γeff A0 = ∆γA(ζ1 )η − Uel (1)
interfacial separation between the solid walls[41, 42]. In
this theory the interface is studied at different magnifica-
tions ζ = L/λ, where L is the linear size of the system and 3.2 DMT-limit
λ the resolution. We define the wavevectors q = 2π/λ Let pN = FN /A0 be the applied pressure (which can
and q0 = 2π/L so that ζ = q/q0 . The theory focuses on be both positive and negative). In the DMT-limit one
4

0.1 In Ref. [45] we have derived an expression for P (u) using


the Persson contact mechanics theory. In the numerical
0 results presented below we have used the expression for
P (u) given by Eq. (17) in Ref. [45] and below.
The effective interfacial energy can in the DMT-limit
pa (GPa)

-0.1 α=1
be calculated using
Z ∞
-0.2 ∆γ = 0.2 J/m
2 A Uel
γeff = du φ(u)P (u) + ∆γ −
α=0 dc = 1 nm 0 + A 0 A0
-0.3
where A = Ar is the (repulsive) contact area and where
φ(u) is the interaction potential per unit surface area for
-0.4 flat surfaces separated by the distance u and given by
0 0.5 1 1.5 2 2.5 3
separation u (nm) Z ∞
φ(u) = du pa (u).
Figure 2: The adhesive pressure for n = 3 and m = 9 and u
α = 1 (red line) and α = 0 (blue line). In the calculation we
assumed ∆γ = 0.2 J/m2 and dc = 1 nm. Thus in the present case
 n−1  m−1
Bdc dc Bdc α dc
φ(u) = − .
assumes that the elastic deformation of the solids is the n − 1 u + dc m − 1 u + dc
same as in the absence of an adhesive interaction, except
that the external load FN is replaced with an effective For u = 0 an infinite hard wall occurs and we define
load. The latter contains the contribution to the normal the (repulse) contact area Ar when the surface separa-
force from the adhesive force acting in the non-contact tion u = 0. We also define the attractive contact area Aa
interfacial surface area: F0 = FN + Fad . If we divide this when the surface separation 0 < u < dc , but this defin-
equation by the nominal contact area A0 we get ition is somewhat arbitrary and another definition was
used in Ref. [18]. In the calculations below we use n = 3
p0 = pN + pad . and m = 9 and α = 0 (Sec. 4) and α = 1 (Sec. 5). The
The adhesive pressure interaction pressure for these two cases are shown in Fig.
2.
1
Z
pad = d2 x pa (u(x)) (2) The probability distribution of interfacial separations
A0 n.c. P (u) can be calculated as follows: We define u1 (ζ) to be
where pa (u) is the interaction force per unit area when the (average) height separating the surfaces which appear
two flat surfaces are separated by the distance u. In (2) to come into contact when the magnification decreases
the integral is over the non-contact (n.c.) area. In the from ζ to ζ − ∆ζ, where ∆ζ is a small (infinitesimal)
study below we assume that the is an attractive force per change in the magnification. u1 (ζ) is a monotonically
unit area between the surfaces given by (u ≥ 0, see e.g. decreasing function of ζ, and can be calculated from the
Fig. 2): average interfacial separation ū(ζ) and the contact area
 n  m  A(ζ) using (see Ref. [44])
dc dc
pa = B −α , (3)
u + dc u + dc u1 (ζ) = ū(ζ) + ū′ (ζ)A(ζ)/A′ (ζ).
where the cut-off dc is a typical bond length and α a The equation for the average interfacial separation ū(ζ)
number, which we take to be either 0 or 1 below. The is given in Ref. [44]. The (apparent) relative contact area
parameter B is determined by the work of adhesion (per A(ζ)/A0 at the magnification ζ is given by
unit surface area):  
Z ∞ A(ζ) p0
(m − 1) − (n − 1)α = erf ,
du pad (u) = Bdc = ∆γ, A0 2G(ζ)1/2
0 (m − 1)(n − 1)
so that where
2 Z ζq0
∆γ (m − 1)(n − 1)

π E
B= . G(ζ) = dqq 3 C(q),
dc (m − 1) − α(n − 1) 4 1 − ν2 q0
If P (u) denotes the distribution of interfacial separations
where C(q) is the surface roughness power spectrum. In
then we can also write (2) as
what follows we will denote this contact area as the re-
Z ∞
pulsive contact area Ar since the normal stress is repuls-
pad = du pa (u)P (u).
0+
ive within this area. We also define an attractive contact
5

area Aa as the surface area where the surface separation -30


0 < u < dc ; in this surface separation interval the wall- rms roghness = 0.6 nm
-32
wall interaction is attractive. The cut-off length dc is rms slope = 0.0035
H=0.8
quite arbitrary and in Ref. [18] another cut-off length (of -34

log10 C (m )
4
order dc ) was used to define the attractive contact area.
The probability distribution P (u) can be written as[45] -36

1
Z
1 -38
P (u) ≈ dζ [−A′ (ζ)] 1/2
A0 2
(2πhrms (ζ)) -40

-42
(u − u1 (ζ))2 (u + u1 (ζ))2
    
× exp − + exp − , -44
2h2rms (ζ) 2h2rms (ζ) 5 6 7 8 9
log10 q (1/m)
where h2rms (ζ) is the mean of the square of the surface
roughness amplitude including only roughness compon- Figure 3: The surface roughness power spectrum C(q) as a
ents with the wavevector q > q0 ζ, and given by function of the wavevector q (log10 − log10 scale), used in the
Z present calculations. The power spectrum corresponds to a
2 surface with the rms roughness amplitude 0.6 nm, the rms
hrms (ζ) = d2 q C(q). slope 0.0035 and the Hurst exponent H = 0.8.
q>q0 ζ

3.3 Scale-dependent Tabor length dT (q)


The contact between surfaces with roughness on many
length scales involves contact between asperities with 2.5 nm
many different radius of curvatures. Thus at low magni-
fications we only observe long-wavelength roughness and 0
the asperity radius of curvature may be macroscopic, e.g.,
∼ 1 mm or more. At high magnification, nanoscale
-2.5 nm
roughness will be observed involving asperities which
may have radius of curvature in the nm range. Thus
adhesion at long length scale may appear JKR-like while Figure 4: Surface topography of one realization of a surface
at short enough length scale the adhesion may appear with the surface roughness power spectrum shown in Fig. 3.
DMT-like. One can define a magnification or length-scale The difference between the lowest and highest point is about
dependent Tabor length dT (ζ) (ζ = q/q0 ), in the follow- 5 nm, i.e. about 10 times higher than the rms roughness
ing way: If we include only roughness components with 0.6 nm (see Appendix A in [6]).
wavevector q < ζq0 the mean summit asperity curvature
is[48]
-8
1 16 ζq0
Z
= dq q 5 C(q)
∆γ = 0.3 J/m
2
R2 (ζ) 3π q0
12
E=10 Pa
We define
log10 dT (m)

1/3 -9
R(ζ)[γeff (ζ)]2

dT (ζ) =
Er2
If dT (ζ) << dc the contact at the magnification ζ = q/q0 ,
will appear DMT-like while if dT (ζ) >> dc the contact -10
will appear JKR-like. In what follows we will sometimes
denote dT (ζ) with dT (q) (q = ζq0 ).
5 6 7 8 9
4 Theory: numerical results log10 q (1/m)
We now present numerical results which illustrates the
two adhesion theories presented above. The JKR-like Figure 5: The Tabor length parameter dT as a function of
theory has been studied before (see Ref. [43]) so we focus the wavevector (log10 − log10 scale). For the surface with the
mainly on the DMT-like theory. In the calculations we power spectrum given in Fig. 3 and with the elastic modulus
vary ∆γ and n, but we always use the cut-off dc = 0.4 nm E = 1012 Pa (ν = 0.5) and work of adhesion ∆γ = 0.3 J/m2 .
and α = 0 unless otherwise stated.
6

4.1 Surface roughness power spectrum C(q) and 1


(a) 0.4
Tabor length dT (q) 0.9 green: no adhesion
In Fig. 3 we show the surface roughness power spec- 0.8 blue: JKR-type adhesion 0.3
red: DMT-type adhesion
trum C(q) (PSD) as a function of the wavevector q 0.7
E = 10 Pa, ν = 0.5
12

(log10 − log10 scale), used in the present calculations. 0.6 0.2

A/A0
The power spectrum corresponds to a surface with the 0.5
rms roughness amplitude 0.6 nm, the rms slope 0.0035 0.4
∆γ = 0.4 J/m
2
and the Hurst exponent H = 0.8. Fig. 4 shows the sur- 0.3
0.1
face topography of one realization of a randomly rough 0.2
0.3
surface with the surface roughness power spectrum shown 0.1 0.2 0.1 0.0
in Fig. 3. The difference between the lowest and highest 0
point is about 5 nm, i.e. about 10 times higher than the
rms roughness 0.6 nm. 0.2
(b) 0.4
In the calculations below we use the Young’s modulus
E = 1012 Pa, Poisson ration ν = 0.5 and the work of ad-
hesion ∆γ = 0.1 − 0.4 J/m2 . Fig 5 shows the Tabor

γeff (J/m2)
length parameter dT as a function of the wavevector 0.1
(log10 − log10 scale), for ∆γ = 0.3 J/m2 . Note that the 0.3
contact mechanics is DMT-like for short length scales
(or large wavevectors) with dT < dc = 0.4 nm, while it is
JKR-like for long length scales (small wave vectors). 0.2
0
0.1
4.2 Results for different work of adhesion ∆γ ∆γ = 0.0
Fig. 6 shows the normalized (projected) area of con-
tact A/A0 and the effective interfacial energy γeff = -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2
applied pressure (GPa)
[Ead − Uel ]/A0 (where Ead is the (attractive) Van der
Waals interaction energy and Uel the (repulsive) elastic
Figure 6: The normalized (projected) area of contact A/A0
deformation energy) as a function of the nominal applied and the effective interfacial energy γeff = [Ead − Uel ]/A0
pressure pN acting on the block. In the DMT-like theory (where Ead is the (attractive) Van der Waals interaction en-
(red curve) A = Ar is the repulsive contact area while in ergy and Uel the (repulsive) elastic deformation energy) as a
the JKR-like theory (blue curves) A is the total contact function of the applied (nominal) pressure pN acting on the
area (which has both an attractive and a repulsive part). block. In the DMT-like theory (red curve) A = Ar is the re-
Results are shown for the work of adhesion ∆γ = 0.0 pulsive contact area while in the JKR-like theory (blue curves)
A is the total contact area (which has both an attractive and
(green curve), 0.1, 0.2, 0.3 and 0.4 J/m2 . The red and
a repulsive part). Results are shown for the work of adhe-
blue lines correspond to DMT-like and JKR-like approx- sion ∆γ = 0.0 (green curve), 0.1, 0.2, 0.3 and 0.4 J/m2 . The
imations, respectively. Note that the area of contact is red and blue lines correspond to DMT-like and JKR-like ap-
about a factor of 3 larger in the JKR-like approximation proximations, respectively. The elastic solid Young’s modulus
as compared to the DMT-like approximation. This is E = 1012 Pa and Poisson number ν = 0.5.
consistent with the results for adhesion of sphere on flat
(see Sec. 2) where the JKR theory predict about 2 times
larger contact area than the DMT theory. On the other and hence other properties like the friction force, may
hand Fig. 6(b) shows that the effective interfacial bind- be strongly enhanced by the adhesive interaction. In the
ing energies are similar, which is also consistent with the DMT-limit the effective interfacial binding energy is al-
results of Sec. 2. The effective work of adhesion to be ways non-zero if the wall-wall interaction does not van-
used in macroscopic adhesion applications, i.e., the pull- ish beyond some fix wall-wall separation. This is easy to
off of a ball from a flat (Sec. 2) is γeff for the applied understand since when the wall-wall separation is larger
pressure pN = 0, and in all cases in Fig. 6 γeff (pN = 0) than the highest asperity the solid walls will only inter-
is less than half of the work of adhesion ∆γ for smooth act with the long-ranged attractive wall-wall potential
surfaces. and increasing the separation to infinity will always re-
Fig. 6(b) shows that for ∆γ = 0.1 J/m2 in the JKR- quire a finite amount of work making γeff (pN = 0) always
limit the effective interfacial binding energy, and hence non-zero in the DMT-limit.
also the pull-off force, vanish. Nevertheless, Fig. 6(a) Let us now discuss the slopes (with increasing pN ) of
shows that in the JKR-limit the contact area as a func- the γeff (pN ) curves in Fig. 6(b) for pN = 0. As pointed
tion of pN increases much faster with increasing pN than out in Sec. 2, in an exact treatment, as a function of the
in the absence of adhesion (green line), i.e., even if no external load pN the total energy −A0 γeff = −Ead + Uel
adhesion manifests itself during pull-off, the contact area must have a minimum at pN = 0. However, the theories
7

(a)
applied pressure (GPa) 0.4
0.3 ∆γ = 0.3 J/m2

0.2 ∆γ = 0.0

A/A0
0.2 4
0 3
0.1

-0.2 0.2 0.1 2


0.3 n=1.5
0.4
-0.4
0
0 1 2 0.2
average separation (nm) (b)
n=1.5

Figure 7: The applied pressure as a function of the average


separation for the work of adhesion ∆γ = 0.0 (green curve), 2

γeff (J/m2)
0.1, 0.2, 0.3 and 0.4 J/m2 . The blue dashed curve is the Van
der Waals interaction force per unit area pa = B[dc /(u+dc )]3 , 3
0.1
where dc = 0.4 nm and u the distance from the hard wall. 4
The parameter B is chosen to reproduce the given work of
adhesion for flat surfaces.

described above are not exact, and are not based on a 0


treatment which minimize the total energy, but rather -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2
applied pressure (GPa)
focus on the force (or stress) (in the DMT-like model)
or on a combined energy and stress treatment (in the
Figure 8: The normalized (projected) repulsive area of contact
JKR-like model). This is the reason for why the slope of A/A0 and the effective interfacial energy γeff = [Ead −Uel ]/A0
the γeff (pN ) curves for large ∆γ is positive rather than (where Ead is the (attractive) Van der Waals interaction en-
negative. However, the slope is rather small compared ergy and Uel the (repulsive) elastic deformation energy) as
to the (absolute value of) the slope for the non-adhesive a function of the nominal pressure pN acting on the block.
interaction (green curve). In addition, the surface we use Results are shown for the work of adhesion ∆γ = 0.3 J/m2
has a Tabor length with dT (q) << dc for large q and and the interaction force index n = 1.5, 2, 3 and 4 (pa =
dT (q) >> dc for small q so strictly speaking neither the B[dc /(u + dc )]n ). From the DMT-like approximation (see
text). The elastic solid Young’s modulus E = 1012 Pa and
JKR-limit or the DMT-limit is correct or valid. For other Poisson number ν = 0.5.
surfaces which have dT (q) << dc or dT (q) >> dc for
all q, the JKR-like and DMT-like theories may be more
accurate and the slope of the γeff (pN ) curve negative. length u0 is of order the rms surface roughness amplitude.
Fig. 7 shows the applied pressure as a function of the
average separation for the work of adhesion ∆γ = 0.0 4.3 Results for different interaction potential
(green curve), 0.1, 0.2, 0.3 and 0.4 J/m2 . The blue exponent n and factor α
dashed curve is the Van der Waals interaction force per Fig. 8 shows the normalized (projected) repulsive
unit area pa = B[dc /(u + dc )]3 , where d = 0.4 nm and area of contact Ar /A0 and the effective interfacial en-
u the distance from the hard wall. The parameter B is ergy γeff = [Ead − Uel ]/A0 (where Ead is the (attractive)
chosen to reproduce the given work of adhesion for flat Van der Waals interaction energy and Uel the (repulsive)
surfaces. Note that the attractive interaction between elastic deformation energy) as a function of the nominal
the walls is already strong at distances where the flat pressure acting on the block. Results are shown for the
surfaces negligible wall-wall interaction would occur (as work of adhesion ∆γ = 0.3 J/m2 and the interaction
described by the blue dashed line). This is of course due force index n = 1.5 , 2 , 3 and 4 (pa = B[dc /(u + dc )]n ,
to adhesive interaction involving high asperities, which i.e. α = 0). The results are for the DMT-like approxim-
prevail even when the average wall-wall separation is rel- ation. The elastic solid Young’s modulus E = 1012 Pa
ative large. For the case of no adhesion (green curve) the and Poisson number ν = 0.5.
wall-wall interaction is purely repulsive as the asperities Fig. 8 shows that as the interaction becomes more
get compressed on decreasing the wall-wall separation. short ranged (n increases from 1.5 to 4) (at fixed work of
Asymptotically (large separation) this repulsive interac- adhesion ∆γ) the contact area increases while the effect-
tion is exponential pN ∼ exp(−u/u0 ) where the reference ive interfacial binding energy γeff decreases. The latter is
8

0.2 0.12

∆γ = 0.3 J/m
2

applied pressure (GPa)


0.08
applied pressure (GPa)

n=1.5
0
0.04
2
0
-0.2
3 α=0
-0.04 α=1

-0.4 4
-0.08

0 0.5 1 1.5 2 2.5 3


0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 average separation (nm)
average separation (nm)
Figure 11: The applied pressure pN as a function of the aver-
Figure 9: The applied pressure pN as a function of the average
age separation for the work of adhesion ∆γ = 0.2 J/m2 and
separation for the work of adhesion ∆γ = 0.3 J/m2 and the
the interaction index n = 3 and m = 9 and with dc = 1 nm.
interaction force index n = 1.5, 2, 3 and 4 (pa = B[dc /(u +
Blue curve is with α = 0 and red curve with α = 1. From the
dc )]n ). From the DMT-like approximation (see text).
DMT-like approximation (see text).

0.3
again reflect the fact that ∆γ is kept fixed.
0.25 All the numerical results presented above was for the
α=0 cut-off length dc = 0.4 nm and the repulsion factor α =
0.2 0. We now consider the case α = 1 with m = 9 (and
A/A0

n = 3). We also use dc = 1.0 nm. These are the same


0.15
α=1 parameters we will use when comparing the theory with
0.1
exact numerical results in Sec. 5. We consider a surface
with the rms roughness 0.5 nm, the roll-off wavevector
0.05 qr = 1.0 × 106 m−1 and the small and large wavevector
cut-off q0 = 2.5 × 105 m−1 and q1 = 3.2 × 107 m−1 .
0 Fig. 10 shows the normalized (projected) repulsive
-0.08 -0.04 0 0.04 0.08 0.12
applied pressure (GPa) contact area Ar /A0 as a function of the nominal pres-
sure pN acting on the block. Results are shown for the
Figure 10: The normalized (projected) repulsive area of con- work of adhesion ∆γ = 0.2 J/m2 . The blue curve is with
tact Ar /A0 as a function of the nominal pressure pN acting α = 0 and red curve with α = 1.
on the block. Results are shown for the work of adhesion Fig. 11 shows the applied pressure pN as a function of
∆γ = 0.2 J/m2 and the interaction index n = 3 and m = 9 the average surface separation for the work of adhesion
and with dc = 1 nm. Blue curve is with α = 0 and red curve ∆γ = 0.2 J/m2 . Again the blue curve is with α = 0 and
with α = 1. From the DMT-like approximation (see text).
red curve with α = 1.

5 Comparison of the DMT-like theory with ex-


easy to understand: in the limiting case when n → 0 the act numerical results
interaction potential has infinite extend (and infinites- In this section we use the interaction potential (3) with
imal strength in such a way that the work of adhesion n = 3, m = 9 and α = 1 with dc = 1 nm (see Fig. 2 and
∆γ = 0.3 J/m2 ) and in this case γeff must equal ∆γ. Appendix A). The power spectral density adopted in the
At the same time due to the weak (infinitesimal) force numerical calculations (see Appendix for the summary
the contact area at the load pN = 0 must vanish, which of the numerical model) is shown in Fig. 12.
explain the behavior observed in Fig. 8(a). In Fig. 13-15 we show, respectively, the normalized
Fig. 9 shows the applied pressure as a function of and projected area of repulsive contact Ar /A0 , of at-
the average separation for the work of adhesion ∆γ = tractive contact Aa /A0 and the total interaction area
0.3 J/m2 and the interaction force index n = 1.5 , 2 , 3 A/A0 = (Ar + Aa ) /A0 as a function of the applied (nom-
and 4 (pa = B[dc /(u + dc )]n ). The results are for the inal) pressure pN . Red dots are from the deterministic
DMT-like approximation. Note that when n decreases (numerical) model, whereas black solid lines are from the
the more long-range the effective attraction but at the mean field theory. We note that whilst the repulsive
same time the smaller the maximal attraction, which interaction area is slightly underestimated by the the-
9

0.5
-31
10 Numerical
0.4 Analytical

10-34 0.3
CHqL Hm4 L

AA0
0.2
10-37
0.1
-40
10
0.0
8 divisions at q1

10-43 -0.1
106 107 108 109 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2
-1
q Hm L applied pressure HGPaL

Figure 12: Solid line: Power spectral density C (q) as a func- Figure 13: Normalized (projected) area of repulsive contact
tion of q (solid black line). For an isotropic surface roughness Ar /A0 as a function of the applied pressure pN , for different
with cut-off frequency q0 = qr /4, root-mean-square roughness values of work of adhesion ∆γ = 0.1, 0.2, 0.3, 0.4 J/m2 . For
hrms = 0.6 nm, and with self-affine regime in the frequency an elastic solid with Er = 1.33 × 1012 GPa and for the surface
range qr = 106 m−1 to q1 = 103 qr . The Hurst exponent is roughness of Fig. 12.
H = 0.8. (dotted line): The PSD adopted in the numerical
calculations is truncated at q1 = 64q0 , with 8 divisions at the 1.0
smallest length scale (q1 ), resulting in a hrms = 0.52 nm and
mean square slope 0.00115. 0.8

0.6
ory, the pull-off pressures are remarkably accurately cap-
AA0

Numerical
tured at the different adopted values of work of adhesion. 0.4
Analytical
Moreover, the total interaction area (see Fig. 15), as a
function of applied pressure, seems to be only marginally 0.2
affected by the exact contact boundary conditions adop-
0.0
ted in the mean field theory, resulting in a perfect match
with the numerical predictions, as it could have been ex- -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2
pected. It is indeed well known that Persson’s contact applied pressure HGPaL
mechanics accurately predicts the distribution of inter-
facial separations[45]. Hence the total interaction area, Figure 14: Normalized (projected) area of attractive contact
which is evaluated from the distribution of interfacial sep- Aa /A0 as a function of the applied pressure pN , for different
aration, it is accurately captured too. Also note that the values of work of adhesion. For the same parameters as in
simulated contact is close to the DMT-limit, as shown in Fig. 13.
Fig. 16, where the repulsive area is reported as a function
of the nominal repulsive pressure (p0 = pN + pad ). 1.0
In Fig. 17 we show the applied pressure pN as a func-
tion of the average interfacial separation ū, for differ- 0.8
ent values of ∆γ, as determined from the theory (black
0.6
curves) and the numerical model. As expected from the
AA0

previous arguments, the agreement is remarkably good in Numerical


0.4
almost the entire range of average interfacial separations. Analytical

The power spectral density can be nowadays routinely 0.2


obtained with commonly available lab profilometers.
However, usually one has to adopt different acquisition 0.0
techniques depending on the range of roughness length
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2
scales needed to be investigated. Therefore, it would be
particularly interesting to appreciate the extent to which applied pressure HGPaL
the macroscopic adhesive characteristics, such as pull-off
pressure, depends on the effect of adding (or, inversely, Figure 15: Normalized (projected) contact area A/A0 =
(Ar + Aa ) as a function of the applied pressure pN , for dif-
not measuring) an increasing number of surface rough-
ferent values of work of adhesion. For the same parameters
ness frequency components. To do so, we gradually ex- as in Fig. 13.
tend the numerically calculated roughness spectral com-
10

0.5
10-31
q1 =64q0
0.4
q1 =128q0

0.3 10-34 q1 =256q0


q1 =512q0

CHqL Hm4 L
Ar A0

0.2 Numerical HDΓ=0.1 to 0.4L


10-37
Analytical
0.1 Numerical HDΓ=0L

10-40
0.0

-0.1 10-43
0.0 0.1 0.2 0.3 0.4 106 107 108 109
p0 HGPaL -1
q Hm L

Figure 16: Normalized (projected) repulsive area Ar /A0 = as Figure 18: Power spectral density C (q) as a function of q
a function of the nominal repulsive pressure p0 = pN +pad , for (solid black line). For an isotropic surface roughness with
different values of work of adhesion. For the same parameters cut-off q0 = qr /4 and root-mean-square roughness hrms =
as in Fig. 13. 0.6 nm, and with self-affine regime in the frequency range
qr = 106 m−1 to q1 = 103 qr (H = 0.8). The numerical adopted
0.2 PSD (red dots) is truncated at q1 = [64, 128, 256, 512]q0 ,
Numerical with 8 divisions at the smallest lengh scale (q1 ).
applied pressure HGPaL

0.1 Analytical

0.0 Uel ]/A0 [where Ead is the (attractive) Van der Waals in-
teraction energy and Uel the (repulsive) elastic deforma-
-0.1 tion energy] as a function of the nominal pressure pN act-
ing on the block. Results are shown for the work of adhe-
-0.2 sion ∆γ = 0.1, 0.2, 0.3 and 0.4 J/m2 . The red data points
are from the exact numerical simulation and the black
-0.3
0.0 0.5 1.0 1.5 2.0 2.5 3.0
lines from the DMT-like theory (Sec. 3.2) also shown in
Fig. 6. The elastic solid Young’s modulus E = 1012 Pa
average separation HnmL
and Poisson number ν = 0.5.
In Fig. 24 we show similar results for the effective
Figure 17: Nominal pressure pN as a function of the average
interfacial separation ū. For the same parameters as in Fig.
interfacial energy γeff but now for ∆γ = 0.2 J/m2 , and
13. for several large wavevector cut-off q1 = 64q0 , 128q0 and
256q0 . Note that the effective interfacial energy γeff is
rather insensitive to the large wavevector cut-off q1 . The
ponents of Fig. 12, as shown in Fig. 18, up to a system reason for this is that the repulsive elastic energy Uel is
size of 224 mesh points. In Fig. 19-21 we show, respect- dominated by the long-wavelength roughness. In both
ively, the normalized and projected area of repulsive con- figures 23 and 24 there is remarkable good agreement
tact Ar /A0 , the attractive contact Aa /A0 and the total between theory and the simulations.
interaction area A/A0 = (Ar + Aa ) /A0 as a function of
the applied nominal pressure pN , for different truncation 6 Discussion
wavevectors. Red dots are the predictions of the numer- In the discussion above we have neglected adhesion
ical model, whereas black solid lines are from the mean hysteresis. Adhesion hysteresis is particular important
field theory. The pull-off pressure is almost independ- for viscoelastic solids such as most rubber compounds.
ent of the large-wavevector content of the PSD, whereas However, even for elastic solids adhesion hysteresis may
the repulsive contact area, as expected, decreases by in- occur. Thus, not all the stored elastic energy Uel may be
cluding large-wavevector (small wavelength) roughness. used to break adhesive bonds during pull-off but some
Moreover, the large-wavevector roughness does not con- fraction of it may be radiated as elastic waves (phonons)
tribute significantly to the hrms , as is clear both theor- into the solids. This would result in an increase in the
etically and numerically from Fig. 22, where the applied effective interfacial binding energy during pull-off, and
pressure is reported as a function of the average interfa- would result in adhesion hysteresis.
cial separation. We note that adhesion hysteresis is observed already
Finally, let us compare the theory prediction with nu- for smooth surfaces in the JKR-limit (elastically soft
merical results for the effective interfacial energy γeff . In solids) but not in the DMT-limit (hard solids)[49]. Since
Fig. 23 the effective interfacial energy γeff = [Ead − for randomly rough surfaces the contact mechanics may
11

0.4 0.2
Numerical Numerical

applied pressure HGPaL


0.3 Analytical Analytical
0.1

0.2
AA0

0.0
0.1
q1
-0.1
0.0
8 divisions at q1

-0.1 -0.2
-0.10 -0.05 0.00 0.05 0.10 0.15 0.20 0.0 0.5 1.0 1.5 2.0 2.5 3.0
applied pressure HGPaL average separation HnmL

Figure 19: Normalized (projected) area of repulsive contact Figure 22: Applied pressure pN as a function of the average
Ar /A0 as a function of the applied pressure pN , and for ∆γ = interfacial separation ū. For the same parameters as in Fig.
0.2 J/m2 . For an elastic solid with Er = 1.33 × 1012 GPa and 19.
for the surface roughness of Fig. 18 (with q0 = 2.5 105 m−1 ,
qr = 4q0 , H = 0.8, resulting in C0 = 5.24 10−32 m4 ), at 0.4
different truncation frequencies q1 = 64, 128, 256, 512 q0 .
Effective work of adhesion
0.3
Analytical
1.0

Γeff HJm2 L
Numerical

0.8 0.2

0.6
q1 0.1
AA0

0.4 Numerical
Analytical
0.0
0.2 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2
applied pressure HGPaL
0.0

-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 Figure 23: The effective interfacial energy γeff = [Ead −
Uel ]/A0 [where Ead is the (attractive) Van der Waals interac-
applied pressure HGPaL tion energy and Uel the (repulsive) elastic deformation energy]
as a function of the applied pressure pN acting on the block.
Figure 20: Normalized (projected) area of attractive contact Results are shown for the work of adhesion ∆γ = 0.1, 0.2, 0.3
Aa /A0 as a function of the applied pressure pN . For the same and 0.4 J/m2 . The red data points are from the exact numer-
parameters as in Fig. 19. ical simulation and the black lines from the DMT-like theory
(Sec. 3.2) also shown in Fig. 6. For the same parameters as
in Fig. 13.
1.0

0.8
be close to the DMT-limit for short length scales (high
resolution) while close to the JKR-limit at large enough
0.6
length scales, as in Fig. 5, one expects in many cases
AA0

0.4 Numerical that the bond-breaking process involved at short length


Analytical scale is reversible (no hysteresis), while the elastic de-
0.2 formations at large enough length scales show hysteresis,
involving rapid (dissipative) processes during pull-off.
0.0 Contact mechanics for randomly rough surfaces is a
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2
hard problem to treat numerically in the JKR-limit (see
Appendix A) and most studies published are close to the
applied pressure HGPaL
DMT-limit. While this case may be relevant for many
hard materials, most adhesion experiments involves soft
Figure 21: Normalized (projected) contact area A/A0 =
(Ar + Aa ) as a function of the applied pressure pN . For the
materials like silicon rubber (PDMS). In this case the
same parameters as in Fig. 19. adhesion will be JKR-like in a large range of length scales.
We note that adhesion problems which are JKR-like for
12

0.20 energy ∆γ = γeff (q ∗ ) to be used in the JKR theory. This


statement does not hold when the surface roughness amp-
0.15
Effective work of adhesion litude is very small, such as in the present study, because
Analytical
the (average) surface separation in the non-contact area
Γeff HJm2 L

Numerical
is only of order ∼ 1 nm and at this separation the wall-
0.10 wall interaction potential is still important, in particular
for small index n. For charged bodies, due to the long-
0.05 range of the coulomb interaction, the wall-wall interac-
q1 =@64, 128, 256Dq0
tion potential is important for any wall-wall separation.

0.00
-0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2
7 Summary
We have discussed how surface roughness influence the
applied pressure HGPaL
adhesion between elastic solids. We have introduced a
Tabor number which depends on the length scale or mag-
Figure 24: The effective interfacial energy γeff = [Ead −
Uel ]/A0 [where Ead is the (attractive) Van der Waals interac-
nification, and which gives information about the nature
tion energy and Uel the (repulsive) elastic deformation energy] of the adhesion at different length scales. In most cases
as a function of the applied pressure pN acting on the block. the contact mechanics will be DMT-like at short length
Results are shown for the large wavevector cut-off q1 = 64q0 , scales and JKR-like at large length scales. We have con-
128q0 and 256q0 . For the work of adhesion ∆γ = 0.2 J/m2 . sidered two limiting cases relevant for (a) elastically hard
The red data points are from the exact numerical simulation solids with weak adhesive interaction (DMT-limit) and
and the black lines from the DMT-like theory (Sec. 3.2). For (b) elastically soft solids or strong adhesive interaction
the same parameters as in Fig. 19.
(JKR-limit). For the former cases we have studied the
nature of the adhesion using different adhesive force laws
(F ∼ u−n , n = 1.5 − 4, where u is the wall-wall separa-
large length scales and DMT-like for short length scales tion) and by comparing the mean field theory predictions
can be approximately treated using the theory presented with the results of exact numerical calculations. The the-
above: We plot the Tabor length dT (q) as a function of ory results have been compared to the results of exact nu-
logq as in Fig. 5 and divide the logq axis into a large merical simulations, and good agreement between theory
wavevector region q > q ∗ and a short wavevector region and the simulation results was obtained.
q < q ∗ where dT (q ∗ ) = dc . We use the DMT-like theory
to calculate γeff (q ∗ ) including only the roughness com- Appendix A: Numerical model
ponents with q > q ∗ . Next we apply the JKR-like theory We consider the case of two elastic solids patterned
for the q < q ∗ region with ∆γ = γeff (q ∗ ). This treatment with random or deterministic roughness. We assume
is of course only approximate since there will be a region the generic roughness to be characterized by a small
close to q = q ∗ which is neither DMT-like nor JKR-like, wavelength cut-off q0 = 2π/L0 with L0 ≪ L, where L is
but if this region (on the logq-scale) is small compared the representative size of the macroscopic contact region
to the total decades of length scales involved it may con- between the two solids. Given such a large difference of
stitute a good approximation. This picture of adhesion length scales, we can easily identify a representative ele-
is similar to the Renormalization Group (RG) proced- mentary volume (RVE) of interface of length scale LRVE ,
ure used in statistical physics where short wavelength with L0 ≪ LRVE ≪ L, over which we can average out
degrees of freedom (here the short wavelength roughness the contact mechanics occurring at smaller length scales
involved in the DMT-like contact mechanics) are integ- (say, at λ ≪ LRVE ). Note that the numerical or ana-
rated out (removed) to obtain effective equations relevant lytical homogenization of the high-frequency content of
at the macroscopic length scale (here the JKR-like con- a generic physical medium/process model is very com-
tact mechanics). When applying the RG procedure one mon in physics and engineering, since it allows to build a
often finds that processes or phenomena which appear mean field formulation of the model itself, characterized
very different at the microscopic (say atomistic) limit by effective (i.e. smoother) physical properties, vary-
result in the same macroscopic equations of motion e.g., ing over length scales of order ∼ L0 . This is e.g. the
the Navier Stokes equations of fluid flow does not really case of the rough contact mechanics, where the accur-
depend on the exact nature of the force law between ate knowledge of the relationship between the effective
the atoms or molecules except it determines or influence interfacial characteristics (average interfacial separation,
the fluid density and viscosity. Similar, for large surface effective work of adhesion, etc., to cite few), plays a fun-
roughness the force law between the surfaces, which is im- damental role in many physical processes, from friction
portant at short length scale (DMT-limit) does not really and thermal/electrical conduction, to adhesion and inter-
matter for the macroscopic (JKR-like) contact mechan- facial fluid flow. Here we briefly describe the novel effi-
ics except it determines the effective interfacial binding cient numerical approach devoted to simulate the contact
13

mechanics of realistically-rough interfaces at the REV 2.0


scale. 1.5
Deformed Hinitially smoothL surface
Rough rigid surface
In Fig. 25 we show a schematic of the contact geo-
metry. We assume the contact to occur under isothermal 1.0 wHxL
conditions, and the roughness to be characterized by a

@nmD
uHxL
0.5
small mean square slope, in order to make use of the u
well known half space theory. Moreover, the roughness 0.0
is assumed to be periodic with period L0 in both x- and
-0.5 hHxL
y-direction. The local separation between the mating
interfaces u (x) is shown in Fig. 25, and it can be imme- -1.0
diately agreed to be: 0 100 200 300 400 500
grid point
u (x) = ū + w (x) − h (x) , (A1)

where ū is the average interfacial separation, w (x) the Figure 25: Description of the gap [see Eq. (A1)] resulting
from a generic cross section of the contact interface.
surface out-of-average-plane displacement and h (x) the
surface roughness, with hw (x)i = hh (x)i = 0. By defin-
ing
Z
d2 x w (x) e−iq·x
−2
w (q) = (2π) Attractive interactions
@uHxL<z0 & pHxL<0D
Repulsive interactions
and @pHxL>0D
Z Non contact areasD
d2 x σ (x) e−iq·x ,
−2
σ (q) = (2π)

where σ (x) is the distribution of interfacial pressures,


it is (relatively) easy to show that w (x) can be related
to σ (x) through a very simple equation in the Fourier
space:

w (q) = Mzz (q) σ (q) ,

where Mzz (q) = −2/ (|q| Er ) for the elastic half space
[Mzz (q) can be equally determined for layered or vis-
coelastic materials, for which the reader is referred to
Ref. [50]]. Finally, the relation between separation u (x)
and interaction pressure σ (x) is calculated within the Figure 26: Example of contact map.
Derjaguin’s approximation[51], and it can be written in
term of a generic interaction law σ (u) = f (u). f (u) will
be repulsive for u ≤ uw and attractive otherwise, where Usually we adopt uw ≪ dc . Note however that for uw →
uw is a separation threshold describing the ideal equilib- 0 the repulsive term converges to an hard wall, whereas
rium separation. In this work we have adopted the L-J for uw = dc we return to the classical (integrated) L-J
potential to describe the attractive interaction, but one interaction law.
can equally make use of different interaction laws (e.g.
Eqs. (A1), (A2) and (A3) are discretized on a regular
the Morse potential, for chemical bonds). The attractive
square mesh of grid size δ, resulting in the following set
side of f (u), fa (u), reads:
of equations:
8 ∆γ  −9 
fa (u) = ε − ε−3 (A2) Lij = −uij + (ū + wij − hij ) (A4)
3 dc
ε (u) = (u − uw + dc ) /dc , σij = fa (uij ) + fr (uij ) (A5)
−1
σ (xij ) → ∆σ (qhk ) = Mzz (qhk ) w (qhk ) → w (xij ) ,
whereas the repulsive fr (u): (A6)
8 ∆γ  −9 
fr (u) = ε − ε−3 (A3) where Lij is the generic residual (related to the generic
3 dc
iterative solution uij ). In order to solve Eqs. (A4)-(A6),
ε (u) = u/uw . we rephrase Eq. A4 in term of the following ideal Mo-
14

lecular Dynamics process Α=0.5


0.2 ΜT =3.38
2
üij + 2ξij ωij u̇ij = ωij Lij (A7) Theory
Numerics
0.0
we solve with a velocity Verlet integration scheme. ξij

pN  p
and ωij are, respectively, the generic damping factor
-0.2
and modal frequency of the Residuals Molecular Dy-
Εa =1.0
namics system (RMD), which can be smartly used to Εa =1.4
damp the error dynamics. Therefore, at equilibrium -0.4 Εa =2.0
(üij = u̇ij = 0), Eq. (A7) returns the solution of Eqs. Εa =2.8 and 4.0

(A4)-(A6) at zero residuals. The adoption of the RMD -0.6


0.0 0.2 0.4 0.6 0.8 1.0
scheme allows for the (ideal time) search of the solution
to move in a (generic) non-physical error space to finally AA0
furnish, at equilibrium, the targeted (zero residuals) solu-
Figure 27: Dimensionless applied pressure pN /p̄ as a func-
tion. We have found that this very efficiently avoids
tion of the contact area, for a Westergaard like contact geo-
to be trapped in slow relaxation dynamics and/or non- metry. Red curve is from Johnson’s theory, whereas dots are
physical (uncorvergent) solution as otherwise obtained the corresponding numerical predictions at different detach-
with classical (usually very slow) relaxation approaches ment parameters. µT,λ = 3.4 for all numerical results.
(e.g. under-relaxation, often adopted in the literature
for smooth contact conditions, see e.g. Ref. [52]) applied
to realistically-rough interacting surfaces. The solution where nλ is the number of discretization points at
accuracy is set by requiring wavelength λ. It is interesting to observe from Eq. (A9)
that large Tabor numbers, i.e. adhesive interactions oc-

2 2 1/2
Lij /uij < εL (A8) curring in the full JKR regime, are numerically harsh to
D  n−1 2 1/2
E be modelled [due to the fine mesh description required to
n−1
unij − uij /uij < εu , satisfy Eq. (A9): e.g. for µT,λ = 10 and Aλ /dc = 102 ,
we have nλ ≈ 103 ]. However, as recently shown[40], a
where both errors are typically of order 10−4 . As anticip- JKR regime can be conveniently obtained for µT values
ated in previous sections, the nominal projected contact close to 1, reducing the computational complexity of JKR
area A/A0 is given by Ar /A0 + Aa /A0 , where Ar is the interactions.
area of repulsive interaction (defined by σ (x) > 0), and It would be useful to test Eq. (A9) by comparing the
where Aa is the area of attractive interaction (defined by Johnson’s solution (Ref. [53], adhesive sinus contact in
u (x) − uw < dc and σ (x) < 0). In Fig. 26 we show the JKR regime) with the corresponding numerical pre-
a typical contact area map for a DMT rough interac- dictions. In particular, the relation between nominal con-
tion, where Ar /A0 and Aa /A0 correspond, respectively, tact pressure and contact area reads[53]:
to black and gray domains.
pN /p̄ = sin2 φα − α tan φα ,
p
We also observe that for any discretized formulation of (A10)
the adhesive contact mechanics, a fracture tensile stress p
where φα = πA/ (2A0 ), α = 2Er ∆γ/ (λp̄2 ) and p̄ =
can be related to the mesh size characteristics of the con-
πEr Aλ /λ (for the adhesionless interaction, p̄ is the nom-
tact. To determine it, we make use of the penny-shaped
inal squeezing pressure to full contact). In Fig. (27) we
crack solution (see e.g. Ref. [6]), whose tensile stress σa
compare Eq. (A10) (red curve) with numerical results
reads (the π/4 takes into account the square shape of the
obtained with µT,λ ≈ 3.4 (JKR regime), at different val-
grid):
ues of detachment parameter εa . We stress that all the
πp solutions shown in Fig. (27) satisfy the accuracy require-
σa = π∆γEr /δ.
4 ments of Eq. (A8), however only at increasing detach-
σa has to be compared to the maximum tensile stress ment parameter (in particular for εa > 2) the solution
given by the interaction law, σt = ∆γ/dc and, in rapidly converges to the analytical one. Similar consid-
particular, a detachment parameter εa = σa /σt > 1 erations apply for the pressure-separation law, shown in
in order to guarantee the convergence of the numer- Fig. 28.
ical solution. By adopting a Tabor number definition
 2
1/3 Acknowledgments MS acknowledges FZJ for the
2
µT,λ = A λ(2π)2 E∆γ
2 d3 , where λ is smallest roughness support and the kind hospitality received during his
λ r c
wavelentgh and Aλ the corresponding amplitude, a con- visit to the PGI-1, where this work has been per-
vergent numerical solution will be achieved if formed. MS also acknowledges COST Action MP1303 for
 1/2 grant STSM-MP1303-090314-042252 and support from
32 3/2 Aλ MultiscaleConsulting. We thank M.H. Müser for the use-
nλ = λ/δ > 2 µT,λ , (A9)
π dc ful discussion and comments on the manuscript.
15

0.3 [22] A.G. Peressadko, N. Hosoda N and B.N.J. Persson, Phys.


Α=0.5 Rev. Lett. 95, 124301 (2005).
0.2 ΜT =3.38 [23] B. Lorenz, B.A. Krick, N. Mulakaluri, M. Smolyakova,
S. Dieluweit, W.G. Sawyer and B.N.J. Persson, J. Phys.:
0.1
Condens. Matter 25, 225004 (2013).
pN  p

[24] B.A. Krick, J.R. Vail, B.N.J. Persson, W.G. Sawyer, Tri-
0.0
bology Letters 45, 185 (2012).
-0.1
[25] K. Autumn, MRS Bull. 32, 473 (2007).
[26] L. Heepe and S. Gorb, Annu. Rev. Mater. Res. 44, 14.1-
-0.2 14.31 (2014).
[27] B.N.J. Persson, J. Chem. Phys. 118, 7614 (2003).
-0.3 [28] B.N.J. Persson and S. Gorb, J. Chem. Phys. 119, 11437
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 (2003).
uAΛ [29] B.N.J. Persson, C. Ganser, F. Schmied, C. Teichert, R.
Schennach, E. Gilli, U. Hirn, Journ. Phys. C: Condens.
Figure 28: Dimensionless average interfacial separation ū/Aλ Matter 25, (2013)
as a function of the dimensionless applied pressure pN /p̄, for a [30] B.N.J. Persson, A. Kovalev, S.N. Gorb, Tribology Letters
Westergaard like contact geometry. For the same parameters 50, 17 (2013).
of Fig. 27. [31] B.N.J. Persson, Journ. Phys. C: Condens. Matter 19,
376110 (2007)
[32] B.V. Derjaguin and V. Smilga, J. Appl. Phys. 38, 4609
(1967).
[33] A.D. Roberts, J. Phys. D: Appl. Phys. 10, 1801 (1977).
[34] B.N.J. Persson, M. Scaraggi, A.I. Volokitin, M.K.
[1] F.P. Bowden and D. Tabor, Friction and Lubrication of Chaudhury, EPL 103, 36003 (2013).
Solids (Wiley, New York, 1956). [35] K. Brörmann, K. Burger, A. Jagota and R. Bennewitz,
[2] K.L. Johnson, Contact Mechanics, (Cambridge Univer- J. Adhes. 88, 598 (2012).
sity Press, Cambridge, 1966). [36] D. Maugis, J. Colloid Interface Sci 150, 243 (1992).
[3] J.N. Israelachvili, Intermolecular and Surface Forces [37] E. Barthel, J. Phys. D: Appl. Phys. 41, 163001 (2008).
(Academic, London (1995)). [38] V.M. Müller, V.S. Yushenko and B.V. Derjaguin, J. Col-
[4] B.N.J. Persson, Sliding Friction: Physical Principles and loid Interface Sci. 77, 91 (1980).
Applications, 2nd edn. (Springer, Heidelberg, 2000). [39] J.A. Greenwood, Proc. Roy. Soc. London A453, 1277
[5] B.N.J. Persson, O. Albohr, U. Tartaglino, A.I. Volokitin (1961).
and E. Tosatti, J. Phys.: Condens. Matter 17, R1 (2005). [40] M.H. Müser, Beilstein J. Nanotechnol. 5, 419 (2014).
[6] B.N.J. Persson, Surface Science Reports 61, 201 (2006). [41] B.N.J. Persson, J. Chem. Phys. 115, 3840 (2001).
[7] K.N.G. Fuller and D. Tabor, Proc. Roy. Soc. London [42] B.N.J. Persson, Phys. Rev. Lett. 99, 125502 (2007)
A345, 327 (1975). [43] B.N.J. Persson, Eur. Phys. J E8, 385 (2002).
[8] B.V. Derjaguin, Kolloid Z 69 155 (1934). [44] C. Yang and B.N.J. Persson, J. Phys. Condens. Matter
[9] K.L. Johnson, K. Kendall and A.D. Roberts, Proc. R. 20, 215214 (2008).
Soc. Lond. A. 324, 301 (1971) [45] A. Almqqvist, C. Campana, N. Prodanov and B.N.J.
[10] J.A. Greenwood and J.B.P. Williamson, Proc. Roy. Soc. Persson, J. Mech. Phys. Solids 59, 2355 (2011).
Lond. A Mat 295, 300 (1966). [46] B.N.J. Persson, J. Phys.: Condens. Matter 24, 095008
[11] K.L. Johnson and J.A. Greenwood, J. Colloid Interface (2012).
Sci. 192, 326 (1997). [47] G. Carbone, B. Lorenz, B.N.J. Persson and A. Wohlers,
[12] D. Maugis, Journal of Adhesion Science and Technology Eur. Phys. J. E29, 275 (2009).
10(2), 161-175 (1996) [48] P.R. Nayak, J. Lubr. Technol. 93, 398 (1971).
[13] N. Prodanov, W.B. Dapp and M.H. Müser, Tribol. Lett. [49] In general, in order for elastic instabilities to occur in
53, 433 (2014). contact mechanics and sliding friction the material must
[14] W.B. Dapp, N. Prodanov and M.H. Müser, Systematic be soft enough. This is most easily understood in the
analysis of Persson’s contact mechanics theory for ran- context of a particle in a periodic potential U (x) and
domly rough surfaces, to be published. connected to an elastic spring k the free end of which is
[15] G. Carbone and F. Bottiglione, J. Mech. Phys. Solids 56, moving parallel to the corrugated potential (x-direction)
2555 (2008). (Tomlinson model of sliding friction). In this case energy
[16] C. Campana, M.H. Müser and M.O. Robbins, J. Phys.: will be dissipated in rapid slip events only if the spring
Condens. Matter 20, 354013 (2008). is soft enough (see, e.g., Ref. [4] for a discussion of this
[17] W.B. Dapp, A. Lücke, B.N.J. Persson and M.H. Müser, point).
PRL 108, 244301 (2012). [50] B.N.J. Persson, J. Phys.: Condens. Matter 24, 095008
[18] L. Pastewka and M.O. Robbins, PNAS 111, 3298 (2014) (2012)
[19] S. Medina and D. Dini, International Journal of Solids [51] B V. Derjaguin, Kolloid Z. 69155–64 (1934)
and Structures (2014) - in press [52] C. Jin, A. Jagota, CY. Hui, J. Phys. D: Appl. Phys.44,
[20] N. Mulakaluri and B.N.J. Persson, EPL 96 66003 (2011). 405303 (2011)
[21] G. Carbone, M. Scaraggi and U. Tartaglino, Eur. Phys. [53] K.L. Johnson, lnt. J. Solids Structures 32(3/4), 423430
J E30, 65 (2009). (1995)

View publication stats

You might also like