throbber
Eur. Phys. J. E (2012) 35: 5
`DOI 10.1140/epje/i2012-12005-2
`
`Regular Article
`
`THE EUROPEAN
`PHYSICAL JOURNAL E
`
`Elastic contact mechanics: Percolation of the contact area and
`fluid squeeze-out
`
`B.N.J. Persson1,a, N. Prodanov1,2, B.A. Krick3, N. Rodriguez4, N. Mulakaluri1, W.G. Sawyer3, and P. Mangiagalli5
`
`1 IFF, FZ J¨ulich, D-52425 J¨ulich, Germany
`2 Sumy State University, 2 Rimskii-Korsakov Str., 40007 Sumy, Ukraine
`3 Department of Mechanical and Aerospace Engineering, University of Florida, Gainesville FL 32611, USA
`4 Advanced Technologies, BD Medical-Pharmaceutical Systems, 1 Becton Drive, MC 427, Franklin Lakes, NJ 07417, USA
`5 Advanced Technologies, BD-Pharmaceutical Systems, 38800 Pont de Claix, France
`
`Received 7 November 2011 and Received in final form 5 January 2012
`Published online: 26 January 2012 – c(cid:2) EDP Sciences / Societ`a Italiana di Fisica / Springer-Verlag 2012
`
`Abstract. The dynamics of fluid flow at the interface between elastic solids with rough surfaces depends
`sensitively on the area of real contact, in particular close to the percolation threshold, where an irregular
`network of narrow flow channels prevails. In this paper, numerical simulation and experimental results for
`the contact between elastic solids with isotropic and anisotropic surface roughness are compared with the
`predictions of a theory based on the Persson contact mechanics theory and the Bruggeman effective medium
`theory. The theory predictions are in good agreement with the experimental and numerical simulation
`results and the (small) deviation can be understood as a finite-size effect. The fluid squeeze-out at the
`interface between elastic solids with randomly rough surfaces is studied. We present results for such high
`contact pressures that the area of real contact percolates, giving rise to sealed-off domains with pressurized
`fluid at the interface. The theoretical predictions are compared to experimental data for a simple model
`system (a rubber block squeezed against a flat glass plate), and for prefilled syringes, where the rubber
`plunger stopper is lubricated by a high-viscosity silicon oil to ensure functionality of the delivery device.
`For the latter system we compare the breakloose (or static) friction, as a function of the time of stationary
`contact, to the theory prediction.
`
`1 Introduction
`
`The influence of surface roughness on fluid flow at the in-
`terface between solids in stationary or sliding contact is a
`topic of great importance both in nature and technology.
`Technological applications include leakage of seals, mixed
`lubrication, and removal of water from the tire-road foot-
`print. In nature, fluid removal (squeeze-out) is important
`for adhesion and grip between the tree frog or gecko adhe-
`sive toe pads and the countersurface during raining, and
`for cell adhesion.
`Almost all surfaces in nature and most surfaces of
`interest in tribology have roughness on many different
`length scales, sometimes extending from atomic distances
`(∼ 1 nm) to the macroscopic size of the system which
`could be of order ∼ 1 cm. Often the roughness is fractal-
`like so that when a small region is magnified (in general
`with different magnification in the parallel and orthogonal
`directions) it “looks the same” as the unmagnified surface.
`Most objects produced in engineering have some par-
`ticular macroscopic shape characterized by a radius of
`
`a e-mail: b.persson@fz-juelich.de
`
`curvature (which may vary over the surface of the solid)
`e.g., the radius R of a cylinder in a combustion engine.
`In this case the surface may appear perfectly smooth to
`the naked eye, but at short enough length scale, in gen-
`eral much smaller than R, the surface will exhibit strong
`irregularities (surface roughness). The surface roughness
`power spectrum C(q) of such a surface will exhibit a roll-
`off wavelength λ0 (cid:3) R (related to the roll-off wave vec-
`tor q0 = 2π/λ0) and will appear smooth (except for the
`macroscopic curvature R) on length scales much longer
`than λ0. In this case, when studying the fluid flow be-
`tween two macroscopic solids, one may homogenize the
`microscopic fluid dynamics occurring at the interface, re-
`sulting in effective fluid flow equations describing the av-
`erage fluid flow on length scales much larger than λ0, and
`which can be used to study, e.g., the lubrication of the
`cylinder in an engine. This approach of eliminating or in-
`tegrating out short length scale degrees of freedom to ob-
`tain effective equations of motion which describe the long
`distance (or slow) behavior is a very general and power-
`ful concept often used in physics, and is employed in the
`study presented below.
`
`Regeneron Exhibit 1110.001
`Regeneron v. Novartis
`IPR2021-00816
`
`

`

`Page 2 of 17
`
`Eur. Phys. J. E (2012) 35: 5
`
`In the context of fluid flow at the interface between
`closely spaced solids with surface roughness, Patir and
`Cheng [1, 2] have shown how the Navier-Stokes equations
`of fluid dynamics can be reduced to effective equations
`of motion involving locally averaged fluid pressure and
`flow velocities. In the effective equation the so-called flow
`factors occur, which are functions of the locally averaged
`interfacial separation ¯u. The authors showed how the flow
`factors can be determined by solving numerically the fluid
`flow in small rectangular units with linear size of order
`of (or larger than) the roll-off wavelength λ0 introduced
`above, and by averaging over several realizations. How-
`ever, with the present speed (and memory) limitations of
`computers fully converged solutions using this approach
`can only take into account roughness over two or at most
`three decades in length scale. In addition, Patir and Cheng
`did not include the long-range elastic deformations of the
`solid walls in the analysis. Later studies have attempted
`to include elastic deformation using asperity contact me-
`chanics models as pioneered by Greenwood-Williamson
`(GW) [3], but it is now known that this theory (and other
`asperity contact models [4]) does not accurately describe
`contact mechanics because of the neglect of the long-range
`elastic coupling between the asperity contact regions [5,6].
`In particular, the relation between the average interfacial
`separation ¯u and the squeezing pressure p, which is very
`important for the fluid flow problem, is not accurately de-
`scribed by the GW model [7–9].
`The paper by Patir and Cheng was followed by many
`other studies of how to eliminate or integrate out the sur-
`face roughness in fluid flow problems (see, e.g., the work
`by Sahlin et al. [10]). Most of these theories involve solving
`numerically the fluid flow in rectangular interfacial units
`and, just as in the Patir and Cheng approach, cannot in-
`clude roughness on more than ∼ 2 decades in length scale.
`In addition, in some of the studies the measured rough-
`ness topography must be “processed” in a non-trivial way
`in order to obey periodic boundary conditions (which is
`necessary for the Fast Fourier Transform method used in
`some of these studies).
`Tripp [11] has presented an analytical derivation of the
`flow factors for the case where the separation between the
`surfaces is so large that no direct solid-solid contact oc-
`curs. He obtained the flow factors to first order in (cid:4)h2(cid:5)/¯u2,
`where (cid:4)h2(cid:5) is the ensemble average of the square of the
`roughness amplitude and ¯u is the average surface separa-
`tion. The result of Tripp has recently been generalized to
`include elastic deformations of the solids [12, 13].
`In this paper, the study of fluid squeeze-out from the
`region between two elastic solids with randomly rough sur-
`faces is presented. We focus on such high contact pres-
`sures that after long enough contact time the area of real
`contact percolates resulting in pockets of confined, pres-
`surized, fluid at the interface. The Bruggeman effective
`medium theory and the Persson contact mechanics theory
`are employed to calculate the interfacial fluid conductivity
`tensor. For anisotropic surface roughness the Bruggeman
`effective medium theory predicts that the contact area
`percolates when A/A0 = γ/(1+γ), where γ is the Peklenik
`
`number (the ratio between the decay length of the height-
`height correlation function along the two principle direc-
`tions) and A/A0 is the relative contact area (A0 is the
`nominal or apparent contact area). The main aim of the
`present work is to verify the theory predictions through
`the comparison with the results of molecular dynamics
`(MD) simulations and experiments. MD calculations have
`been carried out both for isotropic and anisotropic surface
`roughness, while the experiments consider only the sur-
`faces with isotropic statistical properties (where γ = 1).
`The paper outline is as follows. The theoretical approach
`and its application to the fluid squeeze-out are described
`in sects. 2–4 and 5, respectively. Sections 6 and 7 present
`simulations and experiments. In sect. 8 we apply the the-
`ory to the breakloose (or static) friction for prefillable sy-
`ringes. The work is closed by the concluding sect. 9.
`
`2 Anisotropic surface roughness
`
`importance have roughness
`Many surfaces of practical
`with isotropic statistical properties, e.g., sandblasted sur-
`faces or surfaces coated with particles typically bound by a
`resin to an otherwise flat surface, e.g., sandpaper surfaces.
`However, some surfaces of engineering interest have sur-
`face roughness with anisotropic statistical properties, e.g.,
`surfaces which have been polished or grinded in one direc-
`tion. The surface anisotropy is usually characterized by a
`single number, the so-called Peklenik number γ, which is
`the ratio between the decay length ξx and ξy of the height-
`height correlation function (cid:4)h(x, y)h(0, 0)(cid:5) along the x-
`and y-directions, respectively, i.e. γ = ξx/ξy. Here it has
`been assumed that the x-axis is oriented along one of the
`principal directions of the anisotropic surface roughness.
`Let us define the 2×2 matrix (we use polar coordinates
`so that the wave vector q = q(cosφ, sinφ)) [13]
`(cid:2)
`
`D(q) =
`
`dφ C(q)qq/q2
`(cid:2)
`dφ C(q)
`
`,
`
`(1)
`
`where the surface roughness power spectrum [14]
`(cid:3)
`d2x (cid:4)h(x)h(0)(cid:5)e
`1
`−iq·x,
`C(q) =
`(2)
`(2π)2
`where (cid:4)...(cid:5) stands for ensemble average, and h(x) is the
`height profile. For roughness with isotropic statistical
`properties, C(q) will only depend on q = |q| and in this
`case D(q) will be diagonal with D11 = D22 = 1/2.
`We will assume most of the time that D(q) is indepen-
`dent of q and in this case (1) is equivalent to
`(cid:2)
`
`D =
`
`d2q C(q)qq/q2
`(cid:2)
`d2q C(q)
`
`.
`
`(3)
`
`In this case, in the coordinate system where D is diagonal
`the flow conductivity matrix (defined below) σeff will be
`diagonal too. Note that TrD = D11 + D22 = 1, and in
`the coordinate system where D is diagonal we can write
`D11 = 1/(1 + γ) and D22 = γ/(1 + γ), where γ = ξx/ξy
`
`Regeneron Exhibit 1110.002
`Regeneron v. Novartis
`IPR2021-00816
`
`

`

`Eur. Phys. J. E (2012) 35: 5
`
`Page 3 of 17
`
`is the Peklenik number. Note that D11(1/γ) = D22(γ).
`If D(q) (see (1)) depends on q we may still define (in the
`coordinate system where D(q) is diagonal) γ = −1+1/D11
`as before, but the xy-coordinate system where D(q) is
`diagonal may depend on q (in which case the rotation
`angle, ψ(q), of the x-axis relative to some fixed axis, is
`important information too, see ref. [13]). In this case γ
`will depend on q and we will refer to γ(q) as the Peklenik
`function (and ψ(q) as the Peklenik angle function). Note
`that since D(q) is a symmetric tensor and since TrD =
`1, the D-matrix has only two independent components.
`Thus, it is fully defined by the Peklenik function γ(q) and
`the Peklenik angle function ψ(q). In this paper we will
`assume that γ(q) and ψ(q) are constant.
`
`3 Fluid flow between solids with random
`surface roughness
`
`Consider two elastic solids with randomly rough surfaces.
`Even if the solids are squeezed in contact, because of the
`surface roughness there will in general be non-contact re-
`gions at the interface and, if the squeezing force is not
`too large, there will exist non-contact channels from one
`side to the other side of the nominal contact region. We
`consider now fluid flow at the interface between the solids.
`We assume that the fluid is Newtonian and that the fluid
`velocity field v(x, t) satisfies the Navier-Stokes equation
`+ v · ∇v = − 1
`∇p + ν∇2v,

`
`∂v
`∂t
`
`where ν = η/ρ is the kinetic viscosity and ρ is the mass
`density. For simplicity we will also assume an incompress-
`ible fluid so that
`∇ · v = 0.
`We assume that the non-linear term v · ∇v can be
`neglected (this corresponds to small
`inertia and small
`Reynolds number), which is usually the case in fluid flow
`between narrowly spaced solid walls. For simplicity we as-
`sume the lower solid to be rigid with a flat surface, while
`the upper solid is elastic with a rough surface, see fig. 1.
`We introduce a coordinate system xyz with the xy-plane
`in the surface of the lower solid and the z-axis pointing to-
`wards the upper solid. Consider now squeezing the solids
`together in a fluid. Let u(x, y, t) be the separation between
`the solid walls and assume that the slope |∇u| (cid:3) 1. We
`also assume that u/L (cid:3) 1, where L is the linear size of the
`nominal contact region. In this case one expects that the
`fluid velocity varies slowly with the coordinates x and y as
`compared to the variation in the orthogonal direction z.
`Assuming also a slow time dependence, the Navier-Stokes
`equation is reduced to
`
`∂2v
`
`∂z2 = ∇p.
`(4)
`Here and in what follows v = (vx, vy), x = (x, y) and ∇ =
`(∂x, ∂y) are two-dimensional vectors. Note that vz ≈ 0 and
`

`
`Fig. 1. An elastic solid (block) with a rough surface is squeezed
`(pressure p0) in a fluid against a rigid solid (substrate) with a
`flat surface.
`
`(5)
`
`(6)
`
`that p(x) is independent of z to a good approximation.
`From (4) one can obtain the fluid flow vector
`∇p.
`J = − u3(x)
`12η
`Assuming an incompressible fluid mass conservation de-
`mands that
`+ ∇ · J = 0,
`∂u(x, t)
`∂t
`where the interfacial separation u(x, t) is the volume of
`fluid per unit area. In this last equation we have allowed
`for a slow time dependence of u(x, t) as would be the case,
`e.g., during fluid squeeze-out from the interfacial region
`between two solids.
`The fluid flow at the interface between contacting
`solids with surface roughness on many length scales is
`a very complex problem, in particular at high squeezing
`pressures where a network of flow channels with rapidly
`varying width and height may prevail at the interface.
`This is illustrated in fig. 2, which shows the contact area
`(black) between two elastic solids with randomly rough
`surfaces. At high enough pressure the contact area will
`percolate, which will have a drastic influence on the in-
`terfacial fluid flow properties. Percolation corresponds to
`the moment when the narrowest channel disappears as a
`result of squeezing. It is also visible that for anisotropic
`roughness percolation occurs later in the direction of the
`roughness elongation (which is vertical in fig. 2).
`Equations (5) and (6) describe the fluid flow at the
`interface between contacting solids with rough surfaces.
`One can show that after eliminating all the surface rough-
`ness components, the fluid current (given by (5)) takes the
`form
`¯J = −σeff∇¯p,
`(7)
`where ¯p is the fluid pressure averaged over different real-
`izations of the rough surface. The flow conductivity σeff(¯u)
`is in general (for anisotropic surface roughness) a 2×2 ma-
`trix. The ensemble average of (6) gives
`+ ∇ · ¯J = 0.
`∂ ¯u(x, t)
`∂t
`Substituting (7) in (8) gives
`= ∇ ·(σ eff∇¯p) .
`∂ ¯u(x, t)
`∂t
`
`(8)
`
`(9)
`
`Regeneron Exhibit 1110.003
`Regeneron v. Novartis
`IPR2021-00816
`
`

`

`Page 4 of 17
`
`Eur. Phys. J. E (2012) 35: 5
`
` 1.4
`
` 1.2
`φp
`
` 1
`
` 0.8
`
` 0.6
`
` 0.4
`
` 0.2
`
` 0
` 0
`
`γ = 2
`
`1
`
`0.5
`
` 3
` 2
` 1
`average separation / hrms
`
` 4
`
`Fig. 2. (Color online) A snapshot of the contact before per-
`colation in the x-direction (which is horizontal) for anisotropic
`roughness with Peklenik number 5/7. Red line indicates a fluid
`flow stream line. It is visible that fluid is able to flow from the
`left to the right part of the figure (or vice-versa) due to the
`presence of a narrow channel at some region of the contact.
`Inset presents the magnification of this region.
`
`Fig. 3. (Color online) The pressure flow factor φp as a function
`of the average interfacial separation ¯u, for anisotropic surfaces
`with the Peklenik numbers γ = 1/2, 1 and 2. In all cases the
`angular average power spectrum is of the type shown in fig. 4
`with H = 0.9 and the root-mean-square roughness hrms =
`10 μm.
`
`4 Fluid flow conductivity σeff
`
`where Pc(u) is a continuous (finite) function of u. Substi-
`tuting this in (10) gives
`
`1
`σeff
`
`= A
`A0
`
`1 +γ
`γσeff
`
`As was mentioned above, the fluid flow at the interface be-
`tween contacting randomly rough surfaces requires taking
`into account the presence of the network of many inter-
`connected flow channels. In a macroscopic approach this
`can be achieved through the use of the pressure flow fac-
`tor. Here we have employed the 2D Bruggeman effective
`medium theory [15–18] to calculate (approximately) the
`pressure flow factor (see also appendix B).
`For an anisotropic system, the effective medium flow
`conductivity σeff is a 2 × 2 matrix. Let us introduce a xy
`coordinate system and choose the x-axis along a principal
`axis of the D-matrix. In this case we can consider σeff
`as a scalar which within the Bruggeman effective medium
`theory satisfies the relation:
`(cid:3)
`
`1
`σeff
`
`=
`
`du P (u)
`
`1 + γ
`γσeff + σ(u) ,
`
`(10)
`
`where P (u) is the probability distribution of interfacial
`separations, and where
`
`σ(u) = u3
`12η0
`
`.
`
`(11)
`
`Fluid flow along the y-axis is given by a similar equation
`with γ replaced with 1/γ. The probability distribution
`P (u) of interfacial separations has been derived in ref. [21].
`Here we note that P (u) has a delta function at the origin
`u = 0 with the weight determined by the area of real
`contact:
`
`P (u) = A
`A0
`
`δ(u) + Pc(u),
`
`(12)
`
`1 +γ
`γσeff + σ(u) .
`This equation is easy to solve by iteration.
`In fig. 3 the pressure flow factor φp = 12η0σeff /¯u3 as
`a function of the average interfacial separation ¯u is dis-
`played for anisotropic surfaces with the Peklenik numbers
`γ = 1/2, 1 and 2 (see also below). Note that φp = 0
`for ¯u < ¯uc, where ¯uc is the average interfacial separation
`where the area of real contact percolates in the direction
`orthogonal to the fluid flow. In the Bruggeman effective
`medium theory this occurs when the area of real contact
`equals A/A0 = γ/(γ + 1). Thus for γ = 1/2, 1 and 2 the
`contact area percolates (so that no fluid flow occurs along
`the considered direction) when A/A0 = 1/3, 1/2 and 2/3,
`respectively. This explains why φp vanishes at much larger
`(average) interfacial separation (and hence smaller contact
`area) for γ = 1/2 as compared to γ = 2.
`In obtaining the results presented below we have used
`the Persson contact mechanics theory for the contact area
`A and the probability distribution P (u) (see refs. [19–21]).
`This theory depends on the elastic energy Uel stored in the
`asperity contact regions and in this paper we use the sim-
`plest version for Uel (see ref. [5]), where the γ-parameter
`(not the Peklenik number) = 1. Comparison of the theory
`predictions with numerical simulations for small systems
`have shown that γ ≈ 0.45 gives the best agreement be-
`tween theory and the (numerical) experiments. However,
`using γ = 0.45 (or γ (cid:8)= 1 in general) results in much
`longer computational time, with relatively small numeri-
`cal changes as compared to using γ = 1.
`For large (average) surface separation ¯u eqs. (5) and
`(6) can be solved exactly to leading order in (cid:4)h2(cid:5)/¯u2
`(where (cid:4)h2(cid:5) is the mean of the square of the surface rough-
`
`(cid:3)
`
`+
`
`du Pc(u)
`
`(13)
`
`Regeneron Exhibit 1110.004
`Regeneron v. Novartis
`IPR2021-00816
`
`

`

`Eur. Phys. J. E (2012) 35: 5
`
`Page 5 of 17
`
`-18
`
`-22
`
`-26
`
`-30
`
`-34
`
`log C (m4)
`
` 3
`
` 4
`
` 6
` 5
`log q (1/m)
`
` 7
`
` 8
`
`Fig. 4. (Color online) The logarithm (with 10 as basis) of
`angular average power spectrum as a function of the logarithm
`of the wave vector. For qr < q < q1, with the roll-off wave vector
`−1 and the cut-off wave vector q1 = 108m
`−1, the
`qr = 104m
`surface is self-affine fractal with the Hurst exponent H = 0.9.
`−1 and hrms = 10 μm.
`The low wave vector cut-off q0 = 103m
`
`φp = 1 +
`
`ness amplitude h(x, y), where we have assumed (cid:4)h(cid:5) =
`0) [11, 13]
`3(cid:4)h2(cid:5)
`(1 − 3D).
`¯u2
`In appendix A we show that the Bruggeman effective
`medium theory gives the same expression for φp to leading
`order in (cid:4)h2(cid:5)/¯u2 if we identify the γ-parameter in the ef-
`fective medium theory with the Peklenik γ defined by the
`D-matrix (see sect. 2). This result shows that the parame-
`ter γ in the effective medium theory, which was introduced
`in a phenomenological way (as the ratio between the prin-
`ciple axis of an elliptic inclusion) in the effective medium
`theory (see ref. [13]), is indeed determined by the eigen-
`values of the D-matrix as discussed in sect. 2. This is a
`very important result and completes the theory for σeff
`developed in ref. [13].
`
`5 Fluid squeeze-out
`
`At high enough squeezing pressures and after long
`enough time, the interfacial separation will be smaller
`than hrms, so that the asymptotic relation (16) will no
`longer hold. In this case the relation pcont(¯u) can be cal-
`culated using the equations given in ref. [8]. Substituting
`(15) in (14) and measuring pressure in unit of p0, sepa-
`ration in unit of hrms and time in unit of the relaxation
`time
`
`4ηa2u0
`h3
`rmsp0
`
`=
`
`4ηa2
`αh2
`rmsp0
`
`,
`
`(17)
`
`τ =
`
`one obtains
`
`≈ −α
`
`−1φp(¯u)¯u3(1 − pcont),
`
`(18)
`
`d¯u
`dt
`where α = hrms/u0. In order to study the squeeze-out over
`a large time period, t0 < t < t1, it is convenient to write
`t = t0eμ (0 < μ < μ1 with μ1 = ln(t1/t0)). In this case
`(18) takes the form
`≈ −α
`
`d¯u
`dμ
`
`−1tφp(¯u)¯u3(1 − pcont).
`
`(19)
`
`This equation, together with the relation pcont(¯u), consti-
`tutes two equations for two unknowns (¯u and pcont) which
`can be easily solved by numerical integration.
`We have studied the influence of percolation on the
`fluid squeeze-out for an elastic solid with randomly rough
`surface squeezed against a rigid flat surface in a fluid with
`the viscosity η = 12 Pa s. In most of the studies the rough
`surface has the power spectrum shown in fig. 4 with the
`root-mean-square roughness hrms = 10 μm and the large
`wave vector cut-off q1 = 108 m−1. We also present some
`results for another surface with q1 = 107 m−1. The elastic
`block has rectangular shape with the width 2a = 1.84 cm
`(x-direction) and infinite length (y-direction), and the
`squeezing pressure p0 = 2 MPa. The rubber has the
`Young’s modulus E = 3 MPa and Poisson ratio ν = 0.5.
`
`Regeneron Exhibit 1110.005
`Regeneron v. Novartis
`IPR2021-00816
`
`Let us squeeze a rectangular rubber block (height d,
`width (x-direction) 2a and infinite length (y-direction))
`against a substrate in a fluid. Assume that we can ne-
`glect the macroscopic deformations of the rubber block in
`response to the (macroscopically) non-uniform fluid pres-
`sure (which requires d (cid:3) a) [22, 23]. In this case ¯u(x, t)
`will only depend on time t. For this case from (9) we get
`− ¯u3φp(¯u)
`12η
`
`d¯u
`dt
`
`∂2 ¯p
`∂x2 = 0.
`It follows from this equation above that the fluid pressure
`is parabolic
`
`(cid:5)
`
`,
`
`(cid:4)
`
`1 − x2
`a2
`
`pfluid(t)
`
`3 2
`
`¯p(x, t) =
`
`where 2a is the width of the contact region (x-direction)
`and pfluid(t) the average fluid pressure in the nominal con-
`tact region. Combining the two equations above gives
`= − ¯u3φp(¯u)
`d¯u
`4ηa2 pfluid(t).
`dt
`If p0 is the applied pressure acting on the top surface of
`the block, we have
`pfluid(t) = p0 − pcont(t),
`where pcont is the (locally, or ensemble averaged) asperity
`contact pressure. If the pressure p0 is so small that for all
`times ¯u (cid:9) hrms, then in this case φp(¯u) ≈ 1. For ¯u (cid:9) hrms
`we also have [7]
`(cid:5)
`(cid:4)
`
`(14)
`
`(15)
`
`∗
`
`pcont ≈ βE
`− ¯u
`(16)
`exp
`,
`u0
`∗ = E/(1 − ν2) (here E is the Young’s modulus
`where E
`and ν the Poisson ratio), and u0 = hrms/α. The parame-
`ters α and β depend on the fractal properties of the rough
`surface [7].
`
`

`

`Page 6 of 17
`
`Eur. Phys. J. E (2012) 35: 5
`
` 0.6
`
`A/A0
`
` 0.4
`
` 0.2
`
`0.5
`
`1
`γ = 2
`
`φp = 1
`
`γ = 2
`
`1
`
`0.5
`
` 15
`
` 20
`
` 0
`-5
`
` 0
`
` 5
`
` 10
`log t (s)
`
` 15
`
` 20
`
`smooth
`
` 5
`
` 10
`log t (s)
`
`-4
`
`-6
`
`-8
`
`log uav (m)
`
`-10
`
`-5
`
` 0
`
`Fig. 6. (Color online) The relative area of real contact A/A0
`as a function of the logarithm (with 10 as basis) of the squeeze-
`out time. Calculations are for the Peklenik numbers γ = 0.5,
`1 and 2 and the same parameters as in fig. 5.
`
`φp = 1
`
`γ = 2
`
`1
`
`0.5
`
` 2
`
`pcont (MPa)
`
` 1
`
` 0
`-5
`
` 0
`
` 5
`
` 10
`log t (s)
`
` 15
`
` 20
`
`Fig. 7. (Color online) The nominal contact pressure pcon as a
`function of the logarithm (with 10 as basis) of the squeeze-out
`time calculated for the Peklenik numbers γ = 0.5, 1 and 2.
`The pink curve is the result with the fluid pressure flow factor
`φp = 1.
`
`than when γ = 1/2. In fig. 5 we also show the average in-
`terfacial separation for perfectly smooth surfaces and for
`the case where φp = 1 (which we refer to as the average-
`separation theory [23]), where the influence of the surface
`roughness is not included in the fluid flow equation. In
`this case limiting (for large time) average separation ¯u(∞)
`is determined by the roughness alone independent of the
`fluid.
`The time dependencies of the relative area of real con-
`tact A/A0 for the same systems as studied in fig. 5 are
`presented in fig. 6. For γ = 1/2, 1 and 2 the contact area
`saturates at A/A0 = 1/3, 1/2 and 2/3 but the squeeze-out
`time is the same in all cases.
`Figure 7 plots the nominal contact pressure pcon as a
`function of the logarithm of the squeeze-out time for the
`same systems as studied in fig. 5. The applied squeezing
`pressure p0 = 2 MPa is higher than the (nominal) contact
`pressure pcon and the difference p0 − pcon is carried by the
`
`Fig. 5. (Color online) The logarithm (with 10 as basis) of the
`average surface separation as a function of the logarithm of the
`squeeze-out time. Calculations are for the Peklenik numbers
`γ = 0.5, 1 and 2. The pink curve is the result with the fluid
`pressure flow factor φp = 1. For a thin rectangular rubber
`block (width 2a = 1.84 cm) with fractal-like surface roughness
`with the root-mean-square roughness amplitude hrms = 10 μm
`−1 (see fig. 4),
`and the large wave vector cut off q1 = 108 m
`squeezed against a flat rigid surface.
`
`Figure 3 displays the pressure flow factor φp as a func-
`tion of the average interfacial separation ¯u, for anisotropic
`surfaces with the Peklenik numbers γ = 1/2, 1 and 2. Note
`that the flow factor and the viscosity η enter the equation
`for the fluid squeeze-out as φp/η. Thus, φp > 1 has the
`same effect as decreasing the viscosity and hence speeds-
`up the squeeze-out. For γ = 2 we have φp > 1 except
`when the average interfacial separation ¯u becomes so small
`that the contact area is close to the percolation threshold
`where φp will vanish. Thus, at least for low squeezing pres-
`sures, where the interfacial separation never becomes so
`small that the area of real contact percolates, the fluid
`squeeze-out is enhanced by anisotropic roughness when
`the “groves” are in the direction of fluid flow. In a similar
`way, γ < 1 is equivalent to increased viscosity, and slower
`fluid squeeze-out. However, these results are only valid for
`the line-contact geometry. For a circular or square contact
`area any roughness will speed up the squeeze-out. For an
`elliptic contact area with the groves oriented along one
`of the ellipse axis, the squeeze-out may be enhanced or
`slowed down depending on the ratio between the ellipse
`axis, and the value of the Peklenik number.
`Figure 5 shows the logarithm of the average surface
`separation ¯u as a function of the logarithm of the squeeze-
`out time t for the Peklenik numbers γ = 0.5, 1 and 2. Note
`that in all cases the squeeze-out time is the same but the
`final surface separation is the largest for γ = 0.5. The rea-
`son is that for this γ the contact area percolates (in the x-
`direction) for A/A0 = 1/3, so that already when the area
`of real contact reaches this value fluid will be confined in
`the non-contact area and the rubber cannot come closer to
`the substrate as the fluid is essentially incompressible. For
`γ = 2 the contact area percolates for A/A0 = 2/3 which
`will occur at much smaller (average) surface separation
`
`Regeneron Exhibit 1110.006
`Regeneron v. Novartis
`IPR2021-00816
`
`

`

`Eur. Phys. J. E (2012) 35: 5
`
`Page 7 of 17
`
`-4
`
`-5
`
`-6
`
`-7
`
`log uav (m)
`
`-5
`
` 0
`
` 5
`
` 10
`log t (s)
`
` 15
`
` 20
`
` 25
`
`Fig. 9. (Color online) The logarithm (with 10 as basis) of the
`average surface separation as a function of the logarithm of the
`squeeze-out time. Calculations are for the Peklenik number γ =
`1. For a thin rectangular rubber block (width 2a = 1.84 cm)
`with fractal-like surface roughness with the root-mean-square
`roughness amplitude hrms = 10 μm and the large wave vector
`−1 (see fig. 4) (lower (red) curve) and q1 =
`cut off q1 = 108 m
`−1 (upper (green) curve), squeezed against a flat rigid
`107 m
`surface.
`
`(which can be observed only at high magnification) exist
`in this case. However, for “short” time the squeeze-out is
`identical for both surfaces, as is indeed expected because
`in this region of “large” surface separation the squeeze-out
`is dominated by the long wavelength roughness, which is
`the same in both cases.
`
`6 Computer simulation of percolation
`
`After the description of the theoretical approach, let us
`present the numerical simulation technique and results.
`Classical MD simulations have been carried out to verify
`the Bruggeman effective medium theory prediction that
`the contact area percolates when A/A0 = γ/(1 + γ).
`The model consists of an elastic block with a flat bot-
`tom surface which is brought into contact with a randomly
`rough rigid substrate. The latter (see fig. 10) contains
`Nx × Ny = 512 × 512 atoms which occupy the sites of a
`square lattice in the xy-plane with the lattice constant of
`a = 2.6 ˚A. Self-affine fractal topography with Hurst expo-
`nent value of H = 0.8 and the power spectrum analogous
`to the displayed in fig. 4 have been used. The isotropic ran-
`domly rough surface profile of the substrate was obtained
`using the procedure described in ref. [14], based on the
`adding plane waves with random phases. The anisotropic
`roughness was generated by stretching the isotropic one in
`one direction (corresponding to the vertical or y-direction
`in the contact pictures below) accordingly to the specified
`value of the Peklenik number. In the present study the
`values of Peklenik number equal to 1.0, 5/6, 5/7, 5/8 and
`1/2 are used.
`When two elastic solids with rough surfaces come into
`contact, the elastic deformations perpendicular to the con-
`
` 0
`log φp
`-10
`
`-20
`
`-30
`
`-40
`
`-50
`
`-60
`
`-7
`
`-6
`log uav (m)
`
`-5
`
`-4
`
`Fig. 8. (Color online) The logarithm (with 10 as basis) of the
`pressure flow factors φp as a function of the logarithm of the
`average interfacial separation ¯u for isotropic surfaces (γ = 1).
`In all the cases the angular average power spectrum is of the
`type shown in fig. 4 with H = 0.9 andh rms = 10 μm and
`−1 (red curve) and
`with the cut-off wave vector q1 = 108 m
`−1 (green curve).
`q1 = 107 m
`
`fluid. Note that for t > 1017 s the contact pressure is con-
`stant but smaller than p0 because some of the external
`load is carried by the (pressurized) fluid confined in the
`non-contact surface regions. The calculation with φp = 1
`does not account for the percolation of the contact area,
`so no pressurized confined fluid regions occur at the in-
`terface in this approximation, and after long enough time
`the full load is carried by the area of real contact. Note
`also that with φp = 1 the squeeze-out occurs faster since
`the high resistance to fluid interfacial flow which occurs
`close to the percolation threshold (because of the narrow
`flow channels) is absent in this approximation.
`Let us now study the influence of changes in the sur-
`face roughness power spectra on the squeeze-out. Figure 8
`displays the logarithm of the pressure flow factors φp as a
`function of the logarithm of the average interfacial sepa-
`ration ¯u, for a surface with isotropic roughness (Peklenik
`number γ = 1) for two values of the cut-off wave vec-
`tor q1 = 108 m−1 (red curve) and q1 = 107 m−1 (green
`curve) in the power spectrum. Note that removing one
`decade of the shortest wavelength roughness has no influ-
`ence on the pressure flow factor for large interfacial separa-
`tion because this region of fluid squeeze-out is dominated
`by the long-wavelength large-amplitude roughness compo-
`nents. However, the shortest-wavelength roughness is very
`important for small interfacial surface separations (which
`require high squeezing pressures and long enough contact
`time). Thus, removing short wavelength roughness moves
`the surface-area percolation threshold to larger interfacial
`separation.
`The dependence of the logarithm of the average surface
`

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket