You are on page 1of 12

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

net/publication/26841493

Adhesive contact of rough surfaces: Comparison between numerical


calculations and analytical theories

Article  in  The European Physical Journal E · September 2009


DOI: 10.1140/epje/i2009-10508-5 · Source: PubMed

CITATIONS READS

75 322

3 authors, including:

Giuseppe Carbone Michele Scaraggi


Politecnico di Bari Università del Salento
196 PUBLICATIONS   3,879 CITATIONS    62 PUBLICATIONS   1,046 CITATIONS   

SEE PROFILE SEE PROFILE

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

Collective Intelligence in Human Groups View project

BioContact - Marie Skłodowska-Curie Individual Fellowship View project

All content following this page was uploaded by Giuseppe Carbone on 26 May 2014.

The user has requested enhancement of the downloaded file.


EPJ E
Soft Matter and
Biological Physics EPJ .org
your physics journal

Eur. Phys. J. E 30, 65–74 (2009) DOI: 10.1140/epje/i2009-10508-5

Adhesive contact of rough surfaces: Compari-


son between numerical calculations and analyt-
ical theories

G. Carbone, M. Scaraggi and U. Tartaglino


Eur. Phys. J. E 30, 65–74 (2009)
DOI 10.1140/epje/i2009-10508-5
THE EUROPEAN
PHYSICAL JOURNAL E
Regular Article

Adhesive contact of rough surfaces: Comparison between


numerical calculations and analytical theories

G. Carbone1,a , M. Scaraggi1 , and U. Tartaglino2


1
DIMeG - Politecnico di Bari, v.le Japigia 182, 70126 Bari, Italy
2
IFF Forschungszentrum Juelich, 52425 Juelich, Germany

Received 18 May 2009 and Received in final form 15 July 2009


c EDP Sciences / Società Italiana di Fisica / Springer-Verlag 2009
Published online: 26 September 2009 – 

Abstract. The authors have employed a numerical procedure to analyse the adhesive contact between a soft
elastic layer and a rough rigid substrate. The solution to the problem, which belongs to the class of the free
boundary problems, is obtained by calculating Green’s function which links the pressure distribution to the
normal displacements at the interface. The problem is then formulated in the form of a Fredholm integral
equation of the first kind with a logarithmic kernel. The boundaries of the contact area are calculated
by requiring the energy of the system to be stationary. This methodology has been employed to study
the adhesive contact between an elastic semi-infinite solid and a randomly rough rigid profile with a self-
affine fractal geometry. We show that, even in the presence of adhesion, the true contact area still linearly
depends on the applied load. The numerical results are then critically compared with the predictions of
an extended version of Persson’s contact mechanics theory, which is able to handle anisotropic surfaces,
as 1D interfaces. It is shown that, for any given load, Persson’s theory underestimates the contact area
by about 50% in comparison with our numerical calculations. We find that this discrepancy is larger than
for 2D rough surfaces in the case of adhesionless contact. We argue that this increased difference might
be explained, at least partially, by considering that Persson’s theory is a mean-field theory in spirit, so
it should work better for 2D rough surfaces rather than for 1D rough surfaces. We also observe that the
predicted value of separation is in agreement with our numerical results as well as the exponents of the
power spectral density of the contact pressure distribution and of the elastic displacement of the solid.
Therefore, we conclude that Persson’s theory captures almost exactly the main qualitative behaviour of
the rough contact phenomena.

PACS. 46.55.+d Tribology and mechanical contacts – 68.35.Np Adhesion – 46.50.+a Fracture mechanics,
fatigue and cracks – 81.40.Pq Friction, lubrication, and wear

1 Introduction one of the authors (G.C.) has shown that GW-type theo-
ries predict linearity only for vanishingly small contact ar-
Numerical studies [1–4] have shown that, in case of non- eas and load, and that when the load is increased the theo-
adhesive contacts, when an elastic body is brought into retical predictions rapidly deviate from the asymptotic lin-
contact with a rough surface the true contact area in- earity. This behaviour has been shown not to be followed
creases proportionally to the applied load. To predict such by Persson’s theory, which predicts linearity between con-
behaviour, two main approaches have been developed: tact area and load up to values of about 15-20% of the
i) multiasperity contact theories [5–9] where the contact nominal contact area. This is in agreement with some ex-
between the surfaces is modelled as an ensemble of ran- perimental and numerical results. Numerical calculations
domly distributed Hertzian contacts between the asperi- by Campañá et al. [13] have shown that the Hertzian-type
ties, and ii) Persson’s theory of contact mechanics [10, 11] regime, which is the basis on which GW and similar the-
where the probability distribution of the contact pressure ories have been developed, occurs only at relatively small
is shown to be governed by a diffusive process as the mag- loads, thus indicating the inadequacy of GW-type theories
nification at which we observe the interface is increased. at higher loads. As already observed, the original version
The scientific community is debating about which theory of Persson’s theory describes the interfacial contact pres-
gives the most accurate results. In a previous paper [12] sure through a parabolic partial differential equation. The
diffusive term of such equation is calculated under the ap-
a
e-mail: carbone@poliba.it proximation that the Power Spectral Density (PSD) of
66 The European Physical Journal E

the elastically deformed surfaces is equal to the PSD of


the underlying rough surfaces [10, 14]. This assumes that
the diffusive term that one would obtain in the case of
full-contact conditions is exactly the same as in the case
of partial contact conditions. The stored elastic energy is,
instead, calculated assuming that the PSD of the deformed
surface is the product of the PSD of the underlying rough
surface multiplied by the fraction of contact area at the
given resolution. This, in particular, can be shown to be
coherently derived by the theory itself, see ref. [14]. Of
course in full contact conditions Persson’s theory is exact,
but in case of partial contact it has to be verified. For this Fig. 1. An elastic layer of thickness d in adhesive contact with
reason, there is not yet clear evidence about the correct- a rough periodic substrate of wavelength λ.
ness of the proportionality factor between the contact area
and load predicted by Persson’s theory. Indeed, there are
numerical investigations [3, 15] of non-adhesive contacts very important to test Persson’s theoretical prediction for
between rough surfaces, which show that Persson’s theory anisotropic surfaces.
underestimates the contact area, although its main qual- In the last years scientists have been developing ad hoc
itative prediction seems to be in agreement with the nu- numerical methods to treat the problem of contact me-
merical calculations, as proved in ref. [13]. In ref. [13] the chanics between randomly rough surfaces. Here we would
Green Function Molecular Dynamics (GFMD) numerical like to recall the methodology proposed by Robbins and
calculations have been employed to show that the PSD of co-workers [3], who developed a Coarse-Graining FEM
the deformed surface has the same power law exponent as (CGFEM) approach, and the methodology conceived by
predicted by Persson’s theory [14]. Campañà and Müser [17], who developed a Green’s Func-
In this paper the authors make an attempt to give tion Molecular Dynamics (GFMD) approach to deal with
additional contributions in this direction. We extend the such a problem.
analysis to include adhesive interactions at the interface of Here we employ a different methodology to deal with
the contacting bodies, which become more and more im- adhesive contact. The methodology, already presented by
portant as the length scale of observation is decreased and one of us in ref. [18], is based on a pure continuum mechan-
may dominate the contact behaviour of micro- and nano- ics approach and belongs to the class of Boundary Element
mechanical and biomechanical systems. We focus on the Methods (BEM), since it also makes use of Green’s func-
adhesive contact between a semi-infinite half-space and a tion to solve the problem. This allows us to reduce the
randomly rough surface with roughness in only one di- problem to Fredholm integral equation of the first kind
rection, and compare our numerical predictions with the with a logarithmic kernel. We stress that the position of
results of Persson’s theory. We observe that the origi- the edges of each contact patch is not known a priori and
nal version of Persson’s theory [10, 11] was conceived to must be determined by requiring that the total energy of
deal with isotropic surfaces, but in our case the surface the system is stationary, i.e. we are dealing with a free-
is strongly anisotropic, being rough only in one direction. boundary problem.
Therefore, in order to compare theoretical and numerically The numerical procedure has been designed in such a
calculated data, we have employed an extended version of way to never lose resolution even when the single contact
Persson’s theory [16], which is able to handle anisotropic spot is below the smallest length scale of observation. We
surfaces as in the case of 1D rough surfaces. There are also show that the numerical complexity of the problem
two main reasons for studying a 1D rough surface. First can be strongly reduced since the thermodynamic state
of all one should consider that surface roughness is char- of the system only depends on the size of each contact
acterised by a large number of length scales, which can area and on the pressure distribution in the contact area.
cover 3-4 decades and even more. Therefore, in order to Thus, in our boundary element approach, only the contact
get physically meaningful results, one needs to include all patches need to be discretised (we use an ad hoc adaptive
the spectral components of the surface roughness in the grid) and the solution to the Fredholm equation, which is
analysis. However, increasing the number of length scales obtained by means of matrix inversion, has to be deter-
rapidly increases the number of points where the numer- mined only for a limited number of points, that is only
ical solution has to be sought, and, in turn, the com- those belonging to the true contact area.
putation time. This problem is strongly reduced in case
of 1D roughness so that one can include in the analysis
more than 3 decades of length scales. The second reason 2 The numerical model
is that rough surfaces, encountered in many practical ap-
plications, are often strongly anisotropic mainly as a result We consider a periodic contact as shown in fig. 1 where
of machining and surface treatments (for example, unidi- an elastic layer of thickness d is interposed between a flat
rectional polished surfaces present wear tracks along the rigid plate (upper surface) and a periodically rough rigid
polishing direction, although the resulting roughness is not substrate with a wavelength λ (bottom surface). We as-
strictly 1D). Thus, from a practical point of view, it is also sume that the rough surface has roughness in only one
G. Carbone et al.: Adhesive contact of rough surfaces: Comparison between numerical calculations . . . 67

where the fundamental wave vector is q0 = 2π/λ. We have


also defined in eq. (2) the quantity hmax = max[h(x)],
which is the maximum height of the substrate rough-
ness. Once the pressure distribution is known, the elastic
displacements at the interface can be easily determined
through the equations

u(x) − um = − G(x − s)σ(s)ds; x ∈ D − Ω, (5)

u(x) − um = h(x) − hmax + ∆; x ∈ Ω,

where D = [−λ/2, λ/2]. Of course for a infinitely thick


Fig. 2. The definition of substrate displacement utot , elastic layer (d → +∞ as in our case) the average displacement
layer average displacement um , and substrate penetration ∆. um is also infinitely large except when the ν = 0.5, but the
difference u(x)−um is always finite [18, 19], and can be in-
terpreted as the additional elastic displacement of the solid
direction and is smooth in the orthogonal direction. Un- due to the presence of roughness at the interfaces. In or-
der these conditions, the problem at hand is a periodic der to close the system of equations, we need an additional
plane problem: the stress, displacement and strain fields condition to determine the yet unknown contact domain
only depend on the x and y coordinates shown in fig. 2 Ω. To this end (see also ref. [18]), we first observe that for
and are periodic functions of period λ. Figure 2 shows, any penetration ∆, we can calculate the pressure distribu-
in particular, the total displacement utot of the substrate, tion at the interfaces through eq. (2), and the interfacial
the average displacement um of boundary of the deformed elastic displacement through eq. (5), as functions of the
layer and the penetration ∆ of the rigid substrate into the unknown coordinates ai and bi of the i-th contact area. To
elastic slab. These three quantities are shown to satisfy calculate the exact values of the quantities ai and bi , given
the following relation: isothermal conditions, we need to find the stationary point
of the free interfacial energy Utot (a1 , b1 , . . . , aL , bL , ∆) of
utot = ∆ + um . (1) the system for a fixed value of the penetration ∆; this is
the same as requiring that
We will focus on the pressure distribution σ(x) and the
displacement u(x) of the elastic solid at the interface. In 
∂Utot
 
∂Utot

refs. [18] and [19] G.C. has shown that the unknown pres- = 0, = 0. (6)
sure distribution in the contact area Ω can be determined ∂ai ∆,bj ∂bi ∆,aj
by solving the following Fredholm integral equation of the
first kind with a logarithmic kernel as: The interfacial energy (see ref. [18]) is

Utot = Uel + Uad , (7)
− G(x − s)σ(s)ds = [h(x) − hmax ] + ∆; x ∈ Ω, (2)

where we have defined the interfacial elastic energy Uel as
where Ω = ∪L i=1 [ai , bi ] is the unknown contact domain to the amount of elastic energy stored in the solid as a conse-
be determined as shown below. The quantities ai and bi quence of the elastic deformations caused by the substrate
are the unknown coordinates of the i-th contact patch, asperities, i.e.
with ai < bi and i = 1, 2, . . . , L, where L is the unknown
number of contacts. In eq. (2), assuming that the elastic Uel (a1 , b1 , . . . , aL , bL , ∆) =
layer is infinitely thick (i.e. d → +∞), the kernel is L 
1  bi
σ(x) [h(x) − hmax + ∆] dx (8)
2(1 − ν 2 )
   
 kx  2 i=1 ai
G(x) = log 2 sin (3)
πE 2 
and the adhesion energy is
and represents the Green’s function of the semi-infinite
elastic body under a periodic loading, i.e. it represents L 
 bi
the displacement u(x) − um caused by the application of a

Uad (a1 , b1 , . . . , aL , bL ) = −γ 1 + [h′ (x)]2 dx,
Dirac comb with peaks δ(x−nλ) separated by a distance λ. i=1 ai
Here E and ν are the Young modulus and the Poisson ratio (9)
of the elastic layer. In eq. (2) the quantity h(x) represents where γ is the Dupré energy of adhesion per unit area.
the heights of the rough profile measured from its mean Equations (2), (5) and (6) constitute a set of closed equa-
plane. Since we are considering a periodic problem, h(x) tions which allows, for any given penetration ∆, to de-
can be written as Fourier series termine the coordinates ai and bi of each contact patch,
+∞ the pressure distribution at the interface, and all other
thermodynamical quantities. For the numerical implemen-

h(x) = hm cos(mq0 x + φm ), (4)
m=1
tation the reader is referred to ref. [18]. Here we describe
68 The European Physical Journal E

some numerical techniques which are peculiar to the prob- penetration starting from 0 (non-contact) up to desired
lem we discuss in this paper. Let us assume that we know value, while optimising the solution at every intermediate
the solution of eqs. (6) for a given penetration ∆: the step. The solution of the contact problem in the presence

N
knowledge of the contact region Ω = i=1 [ai , bi ] is suf- of adhesion is not unique, that is, for the same penetration
ficient to fully characterise the system. Equation (2) de- more configurations are possible depending on the loading
termines the stress σ(x); we solve it iteratively, through history. Nonetheless the contact pattern occurring with
the Gauss-Seidel algorithm. Special care has been taken to increasing penetration is uniquely identified and it is the
guarantee stability and accuracy of the solution by choos- most suitable solution to represent the non-adhesive con-
ing a suitable sampling grid: the domain Ω is discretised in tact in the limit of vanishingly small adhesive bonds. As
a non-uniform, adaptive way, so to ensure that more points a final remark, we observe that this algorithm cannot be
are employed close to the edges of the contacts, where the used to solve the problem without adhesion: in this case
stress distribution, because of adhesion, presents a square the solution is still a stationary point satisfying eq. (6), but
root singularity. As long as the stress is known, the de- the profile of the elastic layer in the non-contact regions
formed profile of the elastic layer follows from eq. (5). The is always tangent to the substrate near the crack tips ai
interfacial energy is then given by eqs. (7), (8) and (9). and bi . An infinitesimal motion of any of the crack points
In summary, at a low level in our numerical implemen- decreasing the contact region would cause an intersection
tation of the algorithm there is the Solver, a software code between elastic layer and substrate. In other words, the
that, given the penetration and the contact regions, calcu- solution always lies on the boundary of the subset of R2N
lates everything else. On top of it we built another piece of identified by the physical constraint of non-intersection.
software in charge of adjusting the position of the contact The solution would not be a minimum without such a
boundaries a1 , b1 , . . . , aN , bN , so to minimise the interfa- constraint.
cial energy. We employed a conjugate gradient method in
the version given by Polak and Ribiére [20]. Unfortunately
the problem is more complex than a minimisation in a 3 Persson’s theory for anisotropic surfaces
2N -dimensional space. Starting from an arbitrary config-
The aim of this paper is to compare the numerical results
uration, some of the contact boundaries can acquire the
with analytical predictions of one of the most promising
same value, meaning that either a contact is detaching or
and also strongly debated theories of contact mechanics,
two contacts are coalescing together into a single bigger
that is the recent theory by Persson [10, 11]. However, the
one. If this happens, the minimisation has to be restarted
surface we are considering is strongly anisotropic, in fact
in a different number of dimensions.
Furthermore, a constraint must be accounted for: in it is rough in only one direction and smooth in the or-
principle we can determine the configuration correspond- thogonal direction. The original theory proposed by Pers-
ing to any penetration and contacts, but we must impose son was, instead, conceived and developed for perfectly
that in the non-contact regions the elastic layer never in- isotropic surfaces. Therefore in order to carry out the
tersects the substrate. For instance, if we consider a phys- analysis, we need to extend the theory to the case of
ical configuration minimising the interfacial energy, and anisotropic surfaces. This extension has been obtained in
then we increase the penetration pushing the substrate ref. [16]. Here we briefly summarise the main equations.
against the elastic layer, the starting point for the new con- Persson’s theory removes the assumption, which is im-
jugate gradient minimisation may show an intersection be- plicit in the multiasperity contact theories, that the area
tween the elastic layer and some peaks of the substrate in of real contact is small compared to the nominal contact
the non-contact regions. This indicates that a new contact area. On the contrary, Persson focuses on the probabil-
has to be added before starting the minimisation. A spe- ity distribution P (σ, ζ) of normal stresses at the interface,
cific procedure inside our software is in charge of detecting which depends on the magnification at which the contact
all the intersections between elastic layer and substrate, so interface is observed. To calculate the governing equation
to enforce the physical constraints of the problem. Given of P (σ, ζ), Persson moves from the limiting case of full
the penetration ∆, the search of the contact domain Ω contact conditions between a rigid rough surface and an
that minimizes the interfacial energy is challenging: not initially flat elastic half-space [10]. In such conditions the
only are the 2N variables a1 , b1 , . . . aN , bN unknown, but PSD of the deformed elastic surface is equal to C2D (q)
where C2D (q) = (2π)−2 d2 xh(0)h(x)e−iq·x is the PSD

also the number N of contact regions is unknown! The
solution to the problem resorts to conjugate gradient min- of the rigid rough substrate (the quantity h(x) is the rough
imisation alternated to searches for intersections between substrate height distribution, x = (x, y) is the in-plane po-
the elastic layer and the block. The minimisation proce- sition vector and the symbol   stands for the ensemble
dure stops when the conjugate gradient ends successfully, average). Considering that for a perfectly elastic mate-
i.e. it is not interrupted by a coalescence of detachment of rial the elastic modulus is frequency independent, it can
contacts, and the successive search for intersections con- be easily shown (see ref. [16]) that even for the general
firms that there are no intersections in the non-contact case of anisotropic surfaces Persson’s theory states that
regions. Although we took special care to guarantee that the stress probability distribution P (σ, ζ) must satisfy the
the minimisation procedure would converge for moder- following relation:
ately large variations of penetration, we observed that the ∂P (σ, ζ) ∂ 2 P (σ, ζ)
most reliable approach involves many small increments of = f (ζ) , (10)
∂ζ ∂σ 2
G. Carbone et al.: Adhesive contact of rough surfaces: Comparison between numerical calculations . . . 69

where the magnification ζ = q/q0 , and σ is the interfacial follows from his theory itself [14]. Analogously the adhe-
stress in the apparent contact area at the magnification sion energy Uad (ζ) is calculated as
ζ. The diffusivity function f (ζ) = qL G′ (q)σ02 , where σ0 is

the average normal stress in the contact area and G(q) is

1/2 −x
dx 1 + ξ 2 x

calculated in full contact conditions as Uad (ζ) = −γA (ζmax ) e , (17)
0
 2
1 E 1 where
G(q) = [∇h(x)]2 q , (11)
2 σ02

8 1−ν
ξ2 = d2 qq 2 C2D (q)
|q|>ζq0
where 
[∇h(x)]2 q = d2 q ′ q ′2 C(q′ ) (12) and ζmax = q1 /q0 , q0 = 2π/λ, and q1 = 2π/λ1 , where λ1
|q′ |<q is the shortest length scale of the rough surfaces. Equa-
tions (10), (13), (14), (15), (16) and (17) can be solved to
is the mean square value of the slope when the surface
calculate the stress probability distribution P (σ, ζ) and
is observed at the magnification ζ = q/q0 , that is when
hence the apparent contact area A(ζ)/A0 as a function of
all harmonic components of the spectrum with wave vec-
the magnification ζ [11, 22]
tor above q are filtered out. Then, Persson assumes that
eqs. (10) and (11) hold true also in partial contact (this is ∞
A(ζ)

of course an approximation since the PSD of the deformed = P (σ, ζ)dσ. (18)
surface in partial contact conditions cannot be the same as A0 −σa (ζ)
that of the rigid rough substrate) and to account for par-
tial contact the following initial and boundary conditions The separation s = hmax − ∆ can be calculated observing
are enforced: that the change of the total interfacial energy Utot = Uel +
Uad must be equal to the work done by the applied load
P (σ, 1) = δ(σ − σ0 ),
P (−σa , ζ) = 0, (13) dUtot = σ0 A0 d∆ = −σ0 A0 ds, (19)
P (∞, ζ) = 0. which gives
Here σa (ζ) is the tensile stress needed to cause detachment +∞
1 dUtot ′

over a strip of length 2π/q, recalling the Griffith criterion s= dσ0 . (20)
in plain strain [21], we get σ0 A0 σ0′ dσ0′

2 ∗
1/2 In case of non-adhesive contact i.e. when γ = 0, Persson
σa (ζ) = E γeff (ζ)q , (14) has shown that
π2
1 
−(σ−σ0 )2 /4G −(σ+σ0 )2 /4G

where E ∗ = E/(1 − ν 2 ). In eq. (14) the quantity γeff (ζ) is P (σ, ζ) = e − e (21)
2(πG)1/2
the apparent energy of adhesion at the interface defined
as [11, 22] and eq. (18) simply becomes
−γeff (ζ)A(ζ) = Uad (ζ) + Uel (ζ), (15) A(ζ)

1

= erf √ . (22)
where A(ζ) is the apparent contact area at the magnifica- A0 2 G(ζ)
tion ζ. The calculation of the elastic energy Uel (ζ) must
take into account that because of partial contact condi- The above formulation also holds true for anisotropic sur-
tions the elastic energy stored at the interface is less than faces, in particular if the substrate has roughness only
what would be stored in case of full-contact conditions. along the x-direction we get [∇h(x)]2 q = (∂h/∂x)2 q .
This is necessarily true because only where contact occurs In such a case we have h(x) = h(x) and one can easily
the elastic solid conforms with the underlying substrate, show that
whereas outside of the contact regions the elastic surface C2D (q) = C(qx )δ(qy ), (23)
is much less deformed. Persson accounts for this fact by
where C(q) = (2π)−1 dxh(0)h(x)e−iqx is the PSD of

assuming that, when calculating the interfacial elastic en-
ergy Uel , the PSD of the deformed (initially flat) elastic the x-profile of the surface. Using eq. (23) one simply ob-
surface is equal to C2D (q)A(q)/A0 , where A(q)/A0 is the tains  q
fraction of apparent contact area at the length scale 2π/q, 2
dqx qx2 C(qx ).
 
(∂h/∂x) q = (24)
i.e. −q
1 ∗

Uel (ζ) = E d2 qqC2D (q)A(q). (16) Equation (18) gives the apparent contact area at the res-
4 |q|>ζq0
olution λ(q) = 2π/q as a function of the applied load σ0 .
At the beginning this result was conjectured by Pers- However, we are interested in calculating the real contact
son [10], who very recently demonstrated that it directly area, which can be obtained by replacing ζ with ζmax .
70 The European Physical Journal E

4 Rough profile generation 100

In order to carry out the numerical simulations and com- 50


pare the results with the theoretical predictions, we need
to numerically generate a rough profile. We have opted for

h [µm]
0
a fractal self-affine rough profile. For any self-affine fractal
profile h(x) the statistical properties are invariant under rough profile
−50 s = 86 µm
the transformation
s = 38 µm
x → tx; h → tH h, (25) −100
s = 9.6 µm

in such a case it can be shown that the PSD of the profile 0.0 0.2 0.4 0.6 0.8 1.0
is  −(2H+1) x [cm]
|q|
C(q) = C0 , (26) Fig. 3. (Colour on-line) The deformed shape of the elastic
q0
body at three different separations, s = 86 µm (blue), s =
where H is the Hurst exponent of the randomly rough pro- 38 µm (red), and s = 9.6 µm (green). The rough rigid substrate
file, and is related to the fractal dimension Df = 2 − H. profile is shown in black.
In order to carry out the numerical calculations, we have
utilised a periodic profile with Fourier components up
to the value q1 = N q0 (in this case ζ1 = N ). How- Therefore, if one knows the r.m.s. roughness of the sur-
ever we need to determine the amplitudes hm and the faces hrms = h(x)2 , one can calculate h21  and there-
phases φm of the harmonic terms (see eq. (4)). It can fore all the other quantities h2m . However to completely
be shown that, in order to satisfy the translational in- characterise the rough profile, we still need the probabil-
variance of the profile’s statistical properties (which im- ity distribution of the amplitudes hm . There are several
plies that the autocorrelation function satisfies the relation choices, however the simplest assumption, as suggested by
h(x′ )h(x′ +x) = h(0)h(x)), it is enough to assume that Persson et al. in ref. [23], is that the probability density
the random phases φm are uniformly distributed on the function of hm is just a Dirac’s delta function centered at
√ 1/2
interval [−π, π[. In such a case the autocorrelation of the [4Cm /δ(0)]1/2 ≈ 2 2πCm /L1/2 , i.e.
profile takes the form  

N  2  2π 1/2
hm p(hm ) = δ hm − 2 C , (31)
L m

′ ′
h(x )h(x + x) = cos(mq0 x). (27)
m=1
2
where we have used δ(q = 0) ≈ L/(2π). In can be
Now we need to calculate the quantities h2m . To obtain shown [23] that this choice also guarantees that the ran-
this, let us calculate the PSD of the periodic profile of dom profile h(x) has a Gaussian random distribution.
eq. (4). By using the definition of the PSD we get
N   2
h
dx m cos(mq0 x)e−iqx

−1
C(q) = (2π) 5 Results
m=1
2
N We assume that the elastic block is a soft perfectly elastic
 1  2 
hm δ(q − mq0 ) + h2m δ(q + mq0 ) , material with elastic modulus E = 1 MPa and Poisson’s
  
= (28)
m=1
4 ratio ν = 0.5, i.e. we assume that the material is incom-
pressible. We assume also that the change of surface en-
from which it follows that ergy upon contact between the two surfaces (i.e. the Dupré
 2
h energy of adhesion) is γ = 0.03 J/m2 . Calculations have
C(−mq0 ) = C(mq0 ) = Cm = m δ(0). been carried out for 11 different realizations of a rough
4 self-affine fractal 1D profile. The profile has a fractal di-
Using eq. (26) and observing that C0 = h21 δ(0)/4, one mension Df = 1.3 (i.e. the Hurst coefficient is H = 0.7),
then obtains with root-mean-square roughness h2 1/2 = 50 µm. The
 2   2  −(2H+1) self-affine profiles have spectral components in the range
hm = h 1 m . (29) q0 < q < q1 . We have used λ = 2π/q0 = 0.01 m and
q1 = 103 q0 . The numerical calculations have been carried
Hence, the quantity h2m  can be determined once h21 
out for different values of the separations s = hmax − ∆,
and the Hurst exponent of the surface are known. Now
N which is defined as the distance between the mean plane
observe from eq. (27) that h(x)2  = m=1 h2m /2, and of the deformed surface and the mean plane of the rough
using eq. (29) one obtains surfaces. In fig. 3 we show three different shapes of the
 2 N deformed profiles at three different values of the separa-
h  −(2H+1) tion: s = 86 µm (blue), s = 38 µm (red), and s = 9.6 µm
h(x) = 1
2
 
m . (30)
2 m=1 (green). The black line instead represents the rigid rough
G. Carbone et al.: Adhesive contact of rough surfaces: Comparison between numerical calculations . . . 71

−12 −7
log10 [Cu (q)/(1m3)]

log10 [Cσ (q)/E* 2 ]


−14
−8
−16
profile PSD −9
−18 σ 0 /E*= 0.005 full contact
σ 0 /E*= 0.020 σ 0 /E* =0.005
− 10 σ 0 /E* =0.020
−20 Persson's prediction
Persson's prediction
−22 − 11
3.0 3.5 4.0 4.5 5.0 5.5 3.0 3.5 4.0 4.5 5.0 5.5
log10 [q/(1m−1)] log10 [q/(1m−1)]
Fig. 4. (Colour on-line) The PSD Cu (q) of the rigid substrate Fig. 5. (Colour on-line) The quantity Cσ (q)/E ∗2 (in meters)
profile (black solid line) compared to that of the deformed for full contact conditions (black solid line) and for two differ-
shape of the elastic body for two different dimensionless loads ent dimensionless loads σ0 /E ∗ = 0.005 (blue), and σ0 /E ∗ =
σ0 /E ∗ = 0.005 (blue), and σ0 /E ∗ = 0.020 (red). Solid lines 0.020 (red). Solid lines refer to numerical calculations whereas
refer to numerical calculations whereas dashed lines refer to dashed lines refer to Persson’s theory. The agreement between
Persson’s theory. The agreement between Persson’s theory and Persson’s theory and numerical calculated predictions is quali-
numerical calculated predictions is qualitatively very good. In- tatively very good, at least in the mid range of q-vectors where
deed the slope of the PSD predicted by Persson (in the mid the influence of adhesion is negligible: both calculations predict
range of q-vectors where the influence of adhesion is negligi- Cσ (q) ≈ q −H .
ble) is almost the same as the numerically calculated one (see
text): in both cases the PSD of the elastically deformed solid
is Cu (q) ≈ q −(2+H) .
is the PSD of the rough substrate. Now, in case of self-
affine fractal surfaces, using eq. (26) and observing that if
substrate profile. A deeper analysis of the figure shows one neglects adhesion (this is correct in the mid range of
that, not depending on the separation s, full contact oc- wave vectors q) eqs. (11) and (22) give A(q)/A0 ≈ q H−1 ,
curs between the elastic block and the short wavelength one obtains from Persson’s theory that Cu (q) ≈ q −(2+H) .
corrugation of the rough rigid profile. This is in agree- Since, in our case, H = 0.7, we have Cu (q) ≈ q −2.7 in
ment with some theoretical arguments [11, 22] which pre- perfect agreement with our numerical calculations. The
dict that this situation should occur when the Hurst expo- same conclusion can be found if one observes fig. 5 which
nent of the rough profile is larger than 0.5, as in our case. shows in a log-log diagram the power spectral density
This is also confirmed in fig. 4 which shows the PSD of the Cσ (q) = (2π)−1 dxσ(0)σ(x)e−iqx of the stress distri-
rough surface and that of the deformed elastic surface as bution at the interface in units of E ∗2 . Of course this is
a function of the wave vector q (in a log-log diagram) for not unexpected since the Cσ (q) and Cu (q) are related to
two values of the applied stress σ0 /E ∗ = 0.005 (blue), and each other through Cσ (q) = 41 E ∗2 q 2 Cu (q) [14, 16]. Using
σ0 /E ∗ = 0.020 (red). Indeed, we observe that for large that Cu (q) ≈ q −(2+H) , and assuming that the adhesion in-
q-vectors the PSD Cu (q) = (2π)−1 dxu(0)u(x)e−iqx

teraction is not important (which occurs in the mid range
of the numerically calculated deformed profile u(x) be- of wave vectors q), one obtains, Cσ (q) ≈ q −H . This is in-
comes almost perfectly parallel to the PSD C(q) of the deed confirmed in fig. 5 which shows that Persson’s predic-
rigid substrate: this means that the spectral content of tion (dashed lines) and numerical calculations (solid lines)
the deformed body profile at short wavelength is just the run parallel to each other in low-mid range of q-vectors.
same as that of the rough rigid profile, and therefore that However, as q is increased, the numerical calculated PSD
full contact occurs between the elastic body and the sub- Cσ (q) rapidly changes its slope. This, in turn, becomes al-
strate at short wavelengths. Figure 4 shows also, as ex- most equal to that of the PSD of interfacial normal stress
pected, that, as the load is increased, the quantity Cu (q) distribution that would be obtained in full contact condi-
continuously approaches the PSD C(q) of the rigid rough tions (black solid line in fig. 5), and confirms what we have
profile and must become equal to C(q) at very high loads, already observed above, i.e. that because of adhesion the
that is when full contact occurs. In fig. 4 we also compare short-wavelength roughness of the underlying rigid profile
the numerically calculated PSD of the deformed surface is in full contact with the elastic body. Figure 6 shows
(solid line) with Persson’s theoretical predictions (dashed the true contact area A(ζmax )/A0 vs. the dimensionless
line). We first observe that there is a non-negligible shift load σ0 /E ∗ calculated through Persson’s theory for adhe-
between Persson’s results and our numerically calculated sive contact and the one computed by our numerical code.
ones. However the two curves run almost perfectly parallel Figure 6 confirms the linearity between contact area and
especially in the mid range of wave vectors where the best load, however it also shows a significant disagreement be-
fit of our numerical results gives Cu (q) ≈ q −2.73 . The value tween Persson’s theory and our numerical calculations. In
−2.73 is very close to the value −2.7 predicted by Persson. particular, Persson’s theory predicts a contact area which
Persson’s theory relies on the assumption that the PSD of is about 50% less than that calculated with our numerical
the deformed surfaces Cu (q) = C(q)A(q)/A0 , where C(q) code. This is in agreement with some molecular dynamics
72 The European Physical Journal E

0.30 2.0
numerical calculations
Persson's theory with adhesion numerical calculations
0.25
Persson's theory without adhesion
1.5 Persson's prediction
0.20
AC /A0

s/h rms
0.15
1.0
0.10
0.05 0.5
0.00
0.000 0.005 0.010 0.015 0.020 0.025
0.0
σ 0 /E* −2.6 −2.4 −2.2 −2.0 −1.8
log10 [σ0 /E*]
Fig. 6. (Colour on-line) The true contact area A(ζmax ) in units
of the nominal contact area A0 as a function of the dimen- Fig. 7. (Colour on-line) The separation s in units of roughness
sionless applied load σ0 /E ∗ . Numerical predictions are in red, hrms as a function of the dimensionless applied load σ0 /E ∗
whereas Persson’s theoretical calculations are in black. Observe in a log-linear scale. The agreement between numerical data
that both numerical calculations and Persson’s theory predict (black curve) and the Persson’s theoretical predictions (red
linearity between contact area and load. However Persson’s curve) is almost perfect. The numerical calculations deviate
theory predicts a coefficient of proportionality which is about from theoretical predictions only at small loads. The reason
half of the numerically calculated one. The dashed line repre- is that our system is a finite system, so a finite value of the
sents Persson’s calculations in the absence of adhesion interac- separation s necessarily exists at which the contact area goes
tions. Notice that because of the low amount of adhesion the to zero and therefore the applied load too. Persson’s theory
dashed curve does not differ significantly from the solid line. instead has been developed for infinite systems. In this case
indeed the rough surface has arbitrarily many and arbitrarily
high asperities, which always allows the contact between the
two solids to occur for arbitrarily large surface separations.
simulations [15] where Persson himself has found a differ-
ence between the numerically calculated contact area and
the theoretical value of about 30% for a two-dimensional 0.7
rough surface. The dashed curve in fig. 6 represents Pers- 0.6 σ 0 /E* = 0.004
son’s theoretical predictions when adhesion is not included
P (σ/ E*,ζ max )

0.5 σ 0 /E* = 0.011


in calculations. As expected, adhesive interactions lead to
an increase of the contact area. However, being in our case 0.4 Double Gaussian Fit

the amount of adhesion energy relatively small, this effect 0.3 Persson's prediction
is only marginal. We observe that the large discrepancy
0.2
between Persson’s predictions and our numerical calcu-
lations may be explained, at least partially, by the fact 0.1
that Persson’s theory is a mean field theory in spirit, so 0.0
it should work better for 2D rough surfaces rather than −0.2 0.0 0.2 0.4 0.6
for 1D rough surfaces. Figure 7 shows the dimensionless σ /E*
separation s/hrms as a function of the dimensionless load Fig. 8. The probability function P (σ, ζmax ) of interfacial
σ0 /E ∗ . Numerical predictions (black line) are compared pressure distribution σ(x). Points are numerical predictions
to Persson’s ones. The agreement is very good except at whereas dashed lines are Persson’s results. The trend is qual-
lower applied loads. Indeed at small loads the numerically itatively the same, although it is quantitatively different. The
calculated separation drops off faster than predicted by reason for such a difference is that numerical calculations and
Persson’s theory. The same effect has also been observed Persson’s theory do not predict the same value of the contact
in molecular dynamics calculations [15]. The explanation area for any given applied load (see fig. 6). Notice that the tail
for this behaviour is that numerical calculations have been of numerically calculated probability distribution at negative
carried out for a finite system, whereas Persson’s theory loads is an effect of the adhesion interaction which has been
is for an infinite system. Indeed an infinite system has introduced only through the surface energy, i.e. with the inclu-
many (arbitrarily) high asperities, which always allows sion of an interaction force with an infinitesimally short range.
the contact between the two solids to occur for arbitrarily We also present a best fit based on double Gaussian probability
large surface separations. But a finite system has asper- distribution (see text).
ities with the height below some finite length hmax , and
for u > hmax no contact occurs between the solids and
σ0 → 0. Figure 8 shows the probability density function be easily understood if one recalls eq. (18) which states
P (σ, ζmax ) of interfacial normal stress distribution σ(x) that the integral of P (σ, ζmax ) must be proportional to
at the highest magnification. We observe that Persson’s the true contact area A(ζmax ) and considers that Pers-
predictions and numerical data agree only at a qualitative son’s theory predicts a contact area smaller by a factor
level, but strongly deviate from a quantitative point of ≈ 1/2, if compared to numerical calculations. We also
view. The reason for this quantitative disagreement can show with solid lines the best fit obtained with the double
G. Carbone et al.: Adhesive contact of rough surfaces: Comparison between numerical calculations . . . 73

0 been shown that, for any given load, the value of true
contact area predicted by Persson’s theory significantly
log10 [P(σ /E*,ζmax)]

−2
differs from the numerically calculated ones, the first be-
−4
ing smaller by a factor ≈ 1/2. This may also depend on
σ 0 /E*=0.011 the fact the Persson’s theory is of the mean-field type and,
−6 σ 0 /E*=0.004 therefore, should work well in higher dimensions than in
Double Gaussian Fit 1D. However, the predicted value of separation matches
−8 Persson's prediction almost perfectly the numerical data. We have also com-
pared the power spectral density of the deformed elastic
− 10 surface and of the stress distribution at the interface as ob-
− 0.2 0.0 0.2 0.4 0.6 tained by numerical calculations and Persson’s theory. We
σ /E* observe that both theory and numerical calculations pre-
Fig. 9. The logarithm of the probability function P (σ, ζmax ). dict the PSDs to follow a power law with almost the same
Points are numerical predictions whereas dashed lines are Pers- exponents. This extends to the case of adhesive contact
son’s results. We observe that the tail of the probability distri- what has been found previously for adhesionless contact
bution at large values of σ follows exactly a Gaussian distribu- by other authors. However, we also observe, in agreement
tion, whereas the tail obtained for negative values of σ is not with Persson’s theory of adhesive contact, that at high
Gaussian. magnification the exponent of the power law changes in
such a way to suggest that the elastic solid lies in full con-
tact condition with the fine microstructure of the rough
Gaussian distribution
surfaces. We conclude that Persson’s theory is able to cap-
P (σ, ζmax ) = ture the main physics behind contact mechanics of rough
1   surfaces independently of whether adhesion is present or
−(σ−σ0 +σa )2 /4G −(σ+σ0 +σa )2 /4G
e − e , (32) not. However, we also observe that from an engineering
2(πG)1/2
point of view a better estimation of the contact area would
where we have relaxed the quantities σa and G. Figure 8 be very useful in practical applications as in case of tires,
and even more fig. 9 show that, at least when the amount mixed lubricated interfaces, and seals, where the kinetic
of adhesion energy is small (as in our case), eq. (32) is a friction or the amount of leakage should be accurately pre-
good approximation of the numerically calculated stress dicted for design purposes. Therefore an improvement of
probability function P (σ, ζmax ). Indeed fig. 9 shows that, the theory, which allows to better estimate the real con-
at high values of σ, the tail of the stress probability den- tact area, would be strongly appreciated by the engineer-
sity function is Gaussian. Notice that the tail of numeri- ing community.
cally calculated stress probability distribution at negative
loads is an effect of the adhesion interaction which has This work, as part of the European Science Foundation EU-
been introduced only through the surface energy, namely ROCORES Programme FANAS, was supported from the EC
by means of an interaction force with an infinitesimally Sixth Framework Programme, under contract N. ERAS-CT-
short range. This even allows for negative values of σ at 2003-980409.
the interface. We also observe that the value of σa calcu-
lated by fitting the interfacial stress probability density
function differs from that calculated in the spirit of Pers-
son’s theory through eq. (14). However, it is possible to
References
show that a better estimation of the tensile stress σa can
1. C. Yang, U. Tartaglino, B.N.J. Persson, Eur. Phys. J. E
be obtained by assuming in eq. (14) that the width of the
19, 47 (2006).
detached region is a factor 1/2 smaller than that originally
2. M. Borri-Brunetto, B. Chiaia, M. Ciavarella, Comput.
assumed in Persson’s theory [11]. In such a case the esti- Methods Appl. Mech. Eng. 190, 6053 (2001).
mated tensile stress σa would increase by roughly a factor 3. S. Hyun, L. Pei, J.-F. Molinari, M.O. Robbins, Phys. Rev.
1.4, thus making σa closer to the value we have found in E 70, 026117 (2004).
fig. 8. 4. C. Campañá, Phys. Rev. E 78, 026110 (2008).
5. J.A. Greenwood, J.B.P. Williamson, Proc. R. Soc. London,
Ser. A 295, 300 (1966).
6 Conclusions 6. A.W. Bush, R.D. Gibson, T.R. Thomas, Wear 35, 87
(1975).
The authors have carried out detailed numerical calcula- 7. T.R. Thomas, Rough Surfaces, Chapt. 8 (Longman Group
tions to determine the contact area, the stress distribution, Limited, New York, 1982).
the penetration and elastic deformation of an infinitely 8. J.A. Greenwood, Wear 261, 191 (2006).
thick layer in adhesive contact with a rough strongly 9. G. Carbone, J. Mech. Phys. Solids 57, 1093 (2009).
anisotropic rigid surface. The numerical predictions have 10. B.N.J. Persson, J. Chem. Phys. 115, 3840 (2001).
been compared in detail with those of an extended ver- 11. B.N.J. Persson, Eur. Phys. J. E 8, 385 (2002).
sion of Persson’s theory, which is able to deal with ad- 12. G. Carbone, F. Bottiglione, J. Mech. Phys. Solids 56, 2555
hesive contact between anisotropic rough surfaces. It has (2008).
74 The European Physical Journal E

13. C. Campañà, M.H. Müser, M.O. Robbins, J. Phys.: Con- 19. G. Carbone, L. Mangialardi, J. Mech. Phys. Solids 52,
dens. Matter 20, 354013 (2008). 1267 (2004).
14. B.N.J. Persson, J. Phys.: Condens. Matter 20, 312001 20. E. Polak, G. Ribière, Rev. Fr. Informat. Rech. Opér. 16,
(2008). 35 (1969).
15. C. Yang, B.N.J. Persson, Phys. Rev. Lett. 100, 024303 21. A.A. Griffith, Phil. Trans. R. Soc. A 221, 163 (1920).
(2008). 22. G. Carbone, L. Mangialardi, B.N.J. Persson, Phys. Rev. B
16. G. Carbone, B. Lorenz, B.N.J. Persson, A. Wohlers, Eur. 70, 125407 (2004).
Phys. J. E 29, 275 (2009). 23. B.N.J. Persson, O. Albohr, U. Tartaglino, A.I. Volokitin,
17. C. Campañá, M.H. Müser, Phys. Rev. B 74, 075420 (2006). E. Tosatti, J. Phys.: Condens. Matter 17, R1 (2005).
18. G. Carbone, L. Mangialardi, J. Mech. Phys. Solids 56, 684
(2008).

View publication stats

You might also like