• No results found

Tribology Using Homogenization

N/A
N/A
Protected

Academic year: 2021

Share "Tribology Using Homogenization"

Copied!
51
0
0

Loading.... (view fulltext now)

Full text

(1)

2009:027

M A S T E R ' S T H E S I S

Tribology Using Homogenization

Betuel Canhanga Afonso Tsandzana

Luleå University of Technology Master Thesis, Continuation Courses Mathematics

Department of Mathematics

(2)
(3)

Tribology Using Homogenization

Betuel Canhanga Afonso Tsandzana and

Department of Mathematics Lule˚a University of Technology

SE-971 87 Lule˚a, Sweden

(4)

Supervisors:

Peter Wall, Ove Edlund and Lars-Erik Persson.

Lule˚a University of Technology, Sweden.

(5)

”As far the laws of mathematics refer to reality, they are not certain, and as far they are certain, they do not refer

to reality.”

Albert Einstein.

(6)
(7)

Table of Contents

1. INTRODUCTION . . . 5

2. DERIVATION OF REYNOLDS EQUATIONS . . . 7

2.1. THE EQUATION OF CONTINUITY OF THE FLUID 7 2.2. THE EQUATION OF MOTION . . . 8

2.3. CONSTITUTIVE RELATIONS . . . 10

2.4. NAVIER-STOKES EQUATIONS . . . 12

2.5. REYNOLDS EQUATIONS . . . 14

3. HOMOGENIZATION OF REYNOLDS EQUATIONS BY MULTIPLE SCALE EXPANSION . . . 19

3.1. ROUGH STATIONARY AND SMOOTH MOVING SURFACES . . . 21

3.2. BOTH ROUGH MOVING SURFACES . . . 25

4. NUMERICAL RESULTS . . . 31

5. THE FRICTION FORCES - F . . . . 39

5.1. CONVERGENCE OF ∇pεAND CONVERGENCE OF F . . . 40

5.2. CONCLUDING REMARKS . . . 42

(8)
(9)

1. INTRODUCTION

Many machine components for example, bearings, pistons rings, gearboxes, breaks consists of parts that to operate must rub against each other causing moving parts which can increase the production costs. To minimize costs of moving parts and to make the machines more reliable it is necessary to lubricate the machine elements.

Lubrication is the action of viscous fluids to diminish friction and wear between solid surfaces. It is fundamental to the operation of all engineer machines and biological process’s. A fluid film can separate two surfaces in relative motion pressed together under and external load, when a fluid film act on this way is called lubricant. Because of the lubricant resistance to motion, the hydrodynamic pressure is built from a lubricant, and this pressure helps to prevent contact between two solid surfaces.

The field of science which deals with practice and technology of lu- brication is named Tribology. In the Hydrodynamic lubrication, the flow of fluid through machine elements may be governed by an mathematical model. In 1886 Osborne Reynolds published a theory of lubrication [25]

that is considered the main guide of Tribology.

Consider a liquid flowing through a thin film region separated by two closely spaced moving surfaces, assuming that the fluid pressure is not vary- ing across the film thickness and fluid inertia effects are ignored; from the momentum transport and continuity equation (conservation of mass), the analysis of the fluid movement lead to a elliptic partial differential equation named Reynolds equation, that describe the hydrodynamic pressure used in lubrication theory, this equation is motivated in this master thesis.

Homogenization is a branch within mathematics that involves the study of partial differential equations PDE’s (Reynolds equations for example) with rapidly oscillating coefficients. The main purpose of homogenization of PDE’s, is to approximate PDE’s that have rapidly varying coefficients with equivalent ”homogenized” PDE’s that more easily lend themselves to numerical treatment in a computer.

In this thesis we use the homogenization theory in tribology by homog- enizing the Reynolds type equations. Recent researches in Reynolds type equations and Homogenization theory can be found in [2], [3], [4], [5], [6], [7], [8], [9], [10], [11], [12], [14], [16], [17], [18], [19], [21], [22], [23].

(10)
(11)

2. DERIVATION OF REYNOLDS EQUATIONS

Let Ω be a domain in R3which consists of a fluid in motion. In the following discussion we assume that there are no source or sinks in the domain, i.e.

no fluid is created or disappears.

2.1. THE EQUATION OF CONTINUITY OF THE FLUID

Consider a fix domain D ⊂ Ω of the fluid bounded by the surface S. The fundamental law of conservation of mass tells that, the rate of change of mass in D equals the flow of mass across S into D. Let ρ = ρ(x1, x2, x3, t) and v = v(x1, x2, x3, t) denote the density and velocity of the fluid in the point x = (x1, x2, x3) at time t. The mass of D at time t is then:

m = Z

D

ρ dx, and since the mass changes at rate

m0= Z

D

∂ρ

∂t dx. (2.1)

By considering the flow of mass across the boundary S, we can define the rate of change of mass in D in a different way. Let n be the unit normal at position x on S pointing out of D. The flow of mass across S into D is equal to the rate of change of mass in D and will be:

m0= − Z

S

ρvini ds, i = 1, 2, 3; (2.2) from divergence theorem, see [1], applied in (2.2)

m0 = − Z

D

∂xi(ρvi) dx, (2.3)

comparing (2.1) and (2.3) gives that Z

D

·∂ρ

∂t +

∂xi(ρvi)

¸

dx = 0,

(12)

and since D was an arbitrary fixed subset of Ω, the equation above implies

that ∂ρ

∂t +

∂xi(ρvi) = 0 (2.4)

must hold in Ω. The equation (2.4) is called the equation of continuity of the fluid, which can be expanded and expressed as

∂ρ

∂t + vi ∂ρ

∂xi = 0 when ∂vi

∂xi = 0, i.e. when the density of the fluid is constant (incomprehen- sible fluid).

2.2. THE EQUATION OF MOTION

Considering a fix domain D of a fluid bounded by the surface S with the outward directed unit normal n. The motion of the fluid obeys Newton’s second law, that is, the rate of increase of momentum of any part of the fluid is equal to the sum of

1. total body force acting on D, 2. total force acting on S,

3. rate of momentum crossing S into D.

Let the body forces be described by the force density f = f (x, t) (force per unit mass). The total body force acting on D at time t is then given by

Z

D

ρf dx. (2.5)

The total force acting on S at time t is Z

S

σn ds, (2.6)

where σn is the stress vector in the point x = (x1, x2, x3) at time t. The rate at which momentum is flowing across S into D at time t is

Z

S

viρvjnj ds. (2.7)

(13)

Summing up the Newton’s second law equations (2.5), (2.6) and (2.7) im- plies that

Z

D

∂t(ρvi) dx = Z

D

ρfi dx + Z

S

σni ds − Z

S

viρvjnj ds. (2.8) We will now use this equation to derive the equation of motion, which is the main purpose of this section. The main step in this process is to convert the surface integrals in (2.8) into volume integrals by applying the divergence theorem. The second surface integral in (2.8) can be converted into volume integral directly by using divergence theorem, more precisely

Z

S

viρvjnj ds = Z

D

∂xj(viρvj) dx = Z

D

µ ρvj∂vi

∂xj + vi

∂xj(ρvj)

dx.

(2.9) Using (2.9) in (2.8) gives

Z

D

µ

∂t(ρvi) − ρfi+ ρvj∂vi

∂xj + vi

∂xj(ρvj)

dx =

Z

S

σin ds. (2.10) Making use of the equation of continuity (2.4), (2.10) can be reduced to

Z

D

µ ρ∂vi

∂t − ρfi+ ρvj∂vi

∂xj

dx =

Z

S

σin ds. (2.11) Remember that our aim is to convert the surface integrals into volume integral, to do this with the surface integral in (2.11), we need first to use the fact that the stress vector on any surface with normal n can be expressed in terms of the stress vectors corresponding to the surfaces with normal vectors e1 = (1, 0, 0), e2 = (0, 1, 0) and e3 = (0, 0, 1); see more details in [20] pages 510-514. Using this fact

σin= σ1in1+ σ2in2+ σ3in3 = σjinj, where i = 1, 2, 3. (2.12) Substituting (2.12) in the surface integral in (2.11) we will have

Z

D

µ ρ∂vi

∂t − ρfi+ ρvj∂vi

∂xj

dx =

Z

S

σjinj ds, and applying divergence theorem we produce the equation

Z

D

µ ρ∂vi

∂t − ρfi+ ρvj∂vi

∂xj

dx =

Z

D

∂σji

∂xj dx, (2.13)

(14)

and since (2.13) holds for any D, it follows that ρ∂vi

∂t + ρvj∂vi

∂xj = ∂σji

∂xj + ρfi, (2.14)

which is called equation of motion of the fluid holds in Ω.

Up to this point we have four equations; equation of continuity of the fluid (2.4) and three equations of motion of the fluid (2.14) which relates the unknowns functions ρ, the three components vi and the nine components of σji, i.e. we have thirteen unknowns and four equations. It is obvious that these equations are too few compared to the number of unknowns, but we can reduce the number of unknowns from thirteen to ten by considering the fact that the σjimatrix is symmetric, the symmetry of matrix σjiis proved in [20], pages 515-519; and we proceed more simplifications by introducing constitutive relations which express σji in terms of vi and the thermody- namic pressure p; and assuming an equation of the state ρ = ρ(p, T ), where T is the temperature. Let us have this in mind in the proceeding discussion.

2.3. CONSTITUTIVE RELATIONS

Consider an ideal fluid, that is homogeneous, isotropic and not viscous. If p denotes the thermodynamical pressure, then

σni = σjinj = −pni for any n. Therefore σij = 0 if i 6= j and −p = 1

3σkk. Thus

σij = 1

3σkkδij = −pδij, here δij is a identity matrix.

We will now generalize this, taking in account the viscous effects. Lets assume that the shear stress is of the form

σij = −pδij+ sij, (2.15) where p is the thermodynamic pressure and sij the components in the viscous stress tensor. The viscous stress tensor represent the contribution to σ caused by viscous effect. The motion of the fluid may be divided into

(15)

local translation, local rigid rotation and local deformation; where the local deformation is described by strain tensor

eij = 1 2

µ∂vi

∂xj +∂vj

∂xi

.

Lets assume now that it is only the local deformation which contributes to s. We also assume that the constitutive relation is linear and isotropic, i.e, the fluid is Newtonian and satisfy the equation

sij = λδijekk+ 2µeij, (2.16) where λ and µ are Lame Viscous coefficient. In general we have that s also depends on the temperature T and the density ρ, that is, λ = λ(T, ρ) and µ = µ(T, ρ).

Using (2.16) in (2.15) we have

σij = −pδij+ λδijekk+ 2µeij. (2.17) Then it follows that

σii= −pδii+ λδiieii+ 2µeii= −3p + 3 µ

λ + 2 3µ

eii.

If we define K = λ +2

3µ, then (2.17) can be expressed as σij = −(p − Kekkij + 2µ

µ eij 1

3δijekk

, (2.18)

the second term in the right hand side has zero trace, and we get that

1

3σii= p − Kekk.

The mechanical pressure P (mean normal stress) is by definition equal to

1

3σii, thus, we have the relation P = p − Kekk between the thermody- namical pressure p and mechanical pressure P. The relation (2.18) with p replaced with respect of P is

σij = −P δij + 2µ µ

eij1 3δijekk

. (2.19)

(16)

2.4. NAVIER-STOKES EQUATIONS

Inserting constitutive relation (2.19) into (2.14) gives that ρ∂vi

∂t + ρvj∂vi

∂xj (2.20)

= ρfi ∂p

∂xi +

∂xj µ

µ∂vi

∂xj + µ∂vj

∂xi

¶ +

∂xi µµ

K − 2 3µ

∂vk

∂xk

. Now, we introduce the equations of state ρ = ρ(p, T ) and derive one equa- tion which correspond to the conservation of energy. We will restrict our discussion to the situation where ρ = ρ(p), K = K(p) and µ = µ(p) are known. With this restrictions in mind the fluid motion is described by equations bellow, equation of continuity of the fluid and the Navier-Stokes equations respectively

∂ρ

∂t +

∂xi(ρvi) = 0, ρ∂vi

∂t + ρvj∂vi

∂xj

= ρfi ∂p

∂xi +

∂xj µ

µ)∂vi

∂xj + µ∂vj

∂xi

¶ +

∂xi µµ

K − 2 3µ

∂vk

∂xk

.

If the fluid is incompressible, ∂vi

∂xi = 0, then the Navier-Stokes equations are reduced to

ρ∂vi

∂t + ρvj∂vi

∂xj = ρfi ∂p

∂xi +

∂xj µ

µ∂vi

∂xj + µ∂vj

∂xi

, (2.21)

in the other hand if µ is constant the equation (2.21) is reduced to ρ∂vi

∂t + ρvj∂vi

∂xj = ρfi ∂p

∂xi + µ∂2vi

∂x2j , (2.22) and if µ = K = 0 (Lame viscous coefficients equals zero) then (2.22) can be simplified, and

ρ∂vi

∂t + ρvj∂vi

∂xj = ρfi ∂p

∂xi (2.23)

which are called Euler equations.

(17)

Now if we consider a viscous fluid with ρ and µ, constants, then, from (2.22) the governing Navier-Stokes equations are:

∂vi

∂t + vj∂vi

∂xj = fi 1 ρ

∂p

∂xi +µ ρ

2vi

∂x2j . The ratio µ

ρ is called kinematic viscosity. Considering two different dimen- sionless forms of this equation:

1. xi = xi

l0, vi= vi

v0, t = tv0

l0 , p = p

ρv02, fi = l0fi v20 ; 2. xi = xi

l0, vi= vi

v0, t =

l20ρ, p= l0p

µv0, fi = l02ρfi µv0 , we produce respectively the dimensionless Navier-Stokes equations:

∂vi

∂t + vj∂vi

∂xj = fi−∂p

∂xi + 1 R

2vi

∂x2j , (2.24)

∂vi

∂t + Rvj∂vi

∂xj = fi ∂p

∂xi +2vi

∂x2j , (2.25) where R = ρv0l0

µ is the Reynolds number. From (2.24) and (2.25) it seems reasonable that for large or small values of the Reynolds number R, the Navier-Stokes equation can be approximated by neglecting the viscous or inertia term. Indeed, for fluid problems where R is large we obtain the governing Euler equations

∂vi

∂t + vj∂vi

∂xj = fi−∂p

∂xi, for situation where R is small we get Stokes equations

∂vi

∂t = fi ∂p

∂xi +2vi

∂x2j .

(18)

2.5. REYNOLDS EQUATIONS

The derivation of Reynolds equations is based on certain appropriate as- sumptions, which reduces the Navir-Stokes equations and the law of mass conservation. Note that this assumptions can be motivated by writing the Navier-Stokes equations in dimensionless form and analyzing the order of magnitude of different terms. For simplicity we assume that one of the two surfaces correspond to the domain Ω in the plane x3 = 0. The distance between the two surfaces is denoted by h = h(x1, x2). The upper and lower surfaces are moving with velocity (Uu, Vu) and (Ul, Vl) respectively. Recall the Navier-Stokes equation (2.20), it is reasonable to assume that the iner- tia effects with respect to time at a given point can be neglected, i.e. the left hand said in the equation (2.20). Moreover, also assume that the body force fi can be neglected and that the only significance velocity gradients

are ∂v1

∂x3 and ∂v2

∂x3, then equation (2.20) is reduced to:

∂p

∂x1 =

∂x3 µ

µ∂v1

∂x3

, (2.26)

∂p

∂x2 =

∂x3 µ

µ∂v2

∂x3

, (2.27)

and ∂p

∂x3 = 0;

assuming that the pressure is constant in the x3 direction, i.e. the pressure depends only in x1 and x2, p = p(x1, x2). It is possible to express the velocity in terms of the pressure. Indeed, integrating (2.26) and (2.27) with respect to x3 twice gives

v1(x) = ∂p

∂x1 x23

+A(x1, x2)

µ x3+ B(x1, x2); (2.28) v2(x) = ∂p

∂x2 x23

+A(x1, x2)

µ x3+ B(x1, x2).

To find A and B we use the no slip boundary conditions both at the upper and low surfaces, that is

(v1, v2) = (Uu, Vu) when x3= h,

(19)

(v1, v2) = (Ul, Vl) when x3= 0, and we obtain that

∂p

∂x1 h2 +A

µh + B = Uu; B = Ul. From this it is clear that

A

µ = Uu− Ul h ∂p

∂x1 h 2µ(p), and inserting into (2.28) gives that

v1(x) = ∂p

∂x1 x23

+Uu− Ul

h x3 ∂p

∂x1 h

2µ(p)x3+ Ul (2.29)

= 1

∂p

∂x1(x23− hx3) + Uu− Ul

h x3+ Ul. Similarly we get that

v2(x) = 1

∂p

∂x2(x23− hx3) +Vu− Vl

h x3+ Vl. (2.30) Since p = p(x1, x2), we know that ρ and µ are constants in the x3- direction (across the fluid film).

Let w be a domain in the lower surface Ω. We denote by D the three dimensional domain obtained by extending Ω between the two surfaces and by S the part of the boundary of D which does not include the surface.

The fundamental law of conservation of mass tells that the rate of change of mass in D is equal to the flow of mass across S into D. The mass m of D at time t is then

m = Z

D

ρdx = Z

w

hρ dx1dx2,

and since the mass changes at rate m0 =

Z

D

∂ρ

∂tdx = Z

w

∂t(hρ) dx1dx2, (2.31) An another expression for m0 can be found by considering the flow of mass across the boundary S (there is no flow through the surface). Let n be the

(20)

unit normal at x on S pointing out of D. The flow of mass across S into D is then

m0 = − Z

S

(ρv1n1+ ρv2n2) ds,

if we use the notation ds = dx3dτ where dτ is the measure on ∂w, and use that ρ does not depend on x3 we get that

m0 = − Z

∂w

ρ Zh

0

v1 dx3 n1+ ρ Zh

0

v2 dx3 n2

 dτ.

Using divergence theorem, see in [1] page 947, we get that

m0= − Z

w

 ∂∂x1

ρ Zh

0

v1dx3

 +

∂x2

ρ Zh

0

v2dx3

 dx1 dx2. (2.32)

comparing (2.31) and (2.32)

∂t(hρ) = −

∂x1

ρ Zh

0

v1dx3

 −

∂x2

ρ Zh

0

v2 dx3

 . (2.33)

Solving (2.33), using (2.29) gives Zh

0

v1dx3 = Zh

0

µ 1

∂p

∂x1(x23− hx3) + Uu− Ul

h x3+ Ul

dx3

= − h3 12µ

∂p

∂x1 + (Uu+ Ul)h 2. Similarly using (2.30) gives

Zh

0

v2dx3= Zh

0

µ 1

∂p

∂x2(x23− hx3) +Vu− Vl

h x3+ Vl

dx3

= − h3 12µ

∂p

∂x2 + (Vu+ Vl)h 2.

(21)

Inserting into (2.33) gives

∂t(hρ) = ∇ · µρh3

12µ∇p

−Uu+ Ul 2

∂x1(ρh) −Vu+ Vl 2

∂x2(ρh).

This is the Reynolds equation which is frequently used to model the pressure distribution in a thin fluid film between two surfaces in relative motion. In many situations the relative motion only take place in one direction, lets say the x1-direction, which the further reduction

∂t(hρ) = ∇ · µρh3

12µ∇p

−Uu+ Ul 2

∂x1(ρh). (2.34) For an incompressible fluid (constant density) the equation (2.34) is re- duced to

∂th = ∇ · µ h3

12µ∇p

−Uu+ Ul 2

∂x1h. (2.35)

Assuming that the film thickness is constant with respect to time we will have

∇ · µ h3

12µ∇p

= Uu+ Ul 2

∂h

∂x1. (2.36)

(22)
(23)

3. HOMOGENIZATION OF REYNOLDS EQUATIONS BY MULTIPLE SCALE

EXPANSION

As follows from the introduction, the Reynolds type equation are used to analyze the influence of the surface geometry and roughness on the hydrodynamic performance of different machine elements when lubricant is flowing into it. The two surfaces through which the lubricant flows between may have different characteristics, but in this study we will only consider the following cases:

1. one surface is smooth and moving, the other is rough and stationary, 2. both surfaces are rough and moving;

that describe the real situation of most of machine elements.

In the case (1), since the film thickness within the machine elements does not change when time changes, i.e. it is constant at some position x, the equation modeling the hydrodynamic performance of machine elements when the lubricant is flowing through it will be time independent.

In Figure 3.1, the surface s1, is moving with velocity Ul from the left to the right, the rough surface, s2 is stationary, h is the film thickness, that will be a function independent of time. In the case (2), the film thickness within the machine elements h, is a function depending on time and space because of the movements of rough surfaces. In Figure 3.2, s1 is moving with velocity Ul, and s2 with velocity Uu, both from the left to the right. In both cases, h will rapidly change and this changes will cause rapidly oscil- lations of hydrodynamic pressure within the machine elements, therefore a mathematical model describing this phenomena will be composed by differ- ential equations with rapidly oscillating coefficients that can not be easily solved even by numerical methods. To avoid this problem it is reasonable to introduce some well accuracy averaging methods that will transform the problem into another that can be solved (at least numerically) and that the solutions will fit the real problem as much as possible, there is where

(24)

s1 s2

Uu=0

Ul h(x)

rough stationary surface global film thickness smooth moving surface

Figure 3.1. smooth moving and rough stationary surfaces

s1 s2

h(x,t)

Uu

Ul

rough stationary surface global film thickness rough moving surface

Figure 3.2. both smooth moving surfaces

(25)

homogenization appears with strong importance. In this chapter we will introduce the homogenization procedure describing the situation (1) and (2).

3.1. ROUGH STATIONARY AND SMOOTH MOVING SURFACES The Reynolds type equation will be of the form

∇ ·

·ρ(p(x))h3(x) ∇p(x)

¸

= u

∂x1 [ρ(p(x))h(x)] , (3.1) with u = Ul+ Uu. Because of the roughness on s2 the film thickness h will depend on the roughness wavelength ε, where ε is a positive sequence that converges to zero and will also affect the oscillation of hydrodynamic pressure, if we assume that the fluid is incompressible, i.e. ρ is a constant, then the equation (3.1) can be expressed as

∇ ·£

h3ε(x)∇pε(x)¤

= λ

∂x1[hε(x)] , λ = 6µu, (3.2) The film thickness will be result of a periodic function h1 representing the effect of the surface roughness and the global film thickness h0 representing the geometry of the surfaces,

hε(x) = h0(x) + h1(y) , y = x ε,

y is the local variable, this means that for small values of ε the function hε is rapidly oscillating.

The first step to introduce the homogenization procedure of the equation (3.2) is to assume multiple scale expansion, see in [13] and [24] of the solutions in the following forms:

pε(x) = p0(x, y) + εp1(x, y) + ε2p2(x, y) + · · · (3.3) where pi(x, y) is periodic in y for every x ∈ Ω; Ω ⊂ R2 is the domain in which the fluid is flowing and (i = 0, 1, 2, · · · ). Assume that the cells of periodicity is Y = (0, 1) × (0, 1). This means that y is a local variable, describing the behavior of the solution on the unit cell scale and the global behavior is described by the variable x. The homogenized Reynolds equa- tions describes the limiting results when the wavelength of the modelled surface roughness tends to zero.

(26)

Applying the Chain rule on the smooth function fε(x) = f (x, y)

the partial derivatives with respect to xj will be:

∂fε

∂xj(x) = µ∂f

∂xj + ε−1∂f

∂yj

(x, y) , j = 1, 2; (3.4) which means that

xfε= ∇xf + ε−1yf (3.5) Using (3.3), (3.4) and (3.5) in (3.2) we get

(∇x+ ε−1y) · [h3(∇x+ ε−1y)(p0+ εp1+ ε2p2+ · · · )]

= λ µ

∂x1 + ε−1

∂y1

h, expanding this, we will have

λ µ

∂x1 + ε−1

∂y1

h (3.6)

= {ε−2y· (h3y) + ε−1[∇y· (h3x) + ∇x· (h3y)] + ∇x· (h3x)}(pε(x)), defining

A0= ∇y· (h3(x)∇y),

A1 = ∇y· (h3(x)∇x) + ∇x· (h3(x)∇y), A2 = ∇x· (h3(x)∇x);

(3.6) can be written as:

λ µ ∂h

∂x1 + ε−1∂h

∂y1

= (ε−2A0+ ε−1A1+ A2)(p0+ εp1+ ε2p2+ · · · ). (3.7) Equating the three lowest power of ε we will have the following system of equations:

A0p0= 0, (3.8)

A0p1+ A1p0 = λ∂h

∂y1, (3.9)

A0p2+ A1p1+ A2p0 = λ∂h

∂x1. (3.10)

In order to solve equations (3.8), (3.9) and (3.10) we need the following lemma:

(27)

Lemma 3.1.1. Let F (y) ∈ L2(Ω) be Y -periodic. Consider the boundary value problem

A0ϕ(y) = F (y)

where ϕ(y) is y -periodic, then, the following holds : 1. There exist a solution ϕ(y) if and only if |Y |1 R

Y F dy = 0

2. If there exist a solution it is unique up to an additive constant, it means that, if ϕ0(y) is another solutions of the boundary value pro- blem, then ϕ(y) = ϕ0(y) + C.

The proof of this lemma can be found in [24]. Now we note that (3.8) has the trivial solution p0 = 0, since the variable x is just a parameter in (3.8). Lemma (3.1.1) says that, if p0(x, y) is another solution of (3.8) then p0(x, y) is a constant with respect to variable y, i.e.

p0(x, y) = p0(x),

where p0(x) is sufficiently differentiable. From equation (3.9) we see that A0p1= λ∂h

∂y1 − A1p0, i.e.

y· (h3yp1) = λ∂h

∂y1 − ∇x· (h3yp0) − ∇y· (h3xp0), since ∇yp0 = 0, we have

y· (h3yp1) = λ∂h

∂y1 − ∇y· µ

h3∂p0

∂x1e1+ h3∂p0

∂x2e2

, (3.11)

where e1, e2are the canonical basis of R2. Since the right hand side consists of three linear terms, we expect that p1(x, y) should be a linear function of three terms. By linearity we will let

p1(x, y) = ∂p0

∂x1v1(x, y) + ∂p0

∂x2v2(x, y) + v3(x, y), (3.12) and using (3.12) in (3.11) we will have:

y ·

½ h3

·

y µ∂p0

∂x1v1(x, y) + ∂p0

∂x2v2(x, y) + v3(x, y)

¶¸¾

(3.13)

(28)

= λ∂h

∂y1 − ∇y· µ

h3∂p0

∂x1e1+ h3∂p0

∂x2e2

, that will give us the following cell problems:

y· (h3yv3− λhe1) = 0, (3.14)

y· (h3yv1+ h3e1) = 0, (3.15)

y· (h3yv2+ h3e2) = 0, (3.16) where vi = vi(x, y) are their solutions.

Equation (3.10), averaged over the period Y, gives:

Z

Y

µ

A0p2+ A1p1+ A2p0− λ∂h

∂x1

dy = 0. (3.17) Because of the periodicity of p2, R

Y A0p2dy = 0; using (3.12), expanding (3.17) and the fact that h3xp1 is periodic, R

Y y· (h3xp1)dy = 0, the above integral can be written as:

Z

Y

·

x· µ

h3y µ∂p0

∂x1v1+ ∂p0

∂x2v2

¶¶

+ ∇x· µ

h3∂p0

∂x1e1+ h3∂p0

∂x2e2

¶¸

dy

= Z

Y

· λ∂h

∂x1 − ∇x· (h3yv3)

¸ dy.

Collecting identical terms we get:

Z

Y

x·

·∂p0

∂x1(h3yv1+ h3e1) +∂p0

∂x2(h3yv2+ h3e2)

¸ dy

= Z

Y

· λ∂h

∂x1 − ∇x· (h3yv3)

¸ dy, or

x·

·∂p0

∂x1 Z

Y

(h3yv1+ h3e1)dy + ∂p0

∂x2 Z

Y

(h3yv2+ h3e2)dy

¸

(3.18)

= ∇x· Z

Y

λh

0

à h3 ∂v∂y31 h3 ∂v∂y32

!#

dy, and expanding (3.18) we produce a matrix

(29)

A(x) =

 R

Y

³

h3 ∂v∂y11 + h3

´

dy R

Y h3 ∂v∂y22dy R

Y h3 ∂v∂y1

2dy R

Y

³ h3 ∂v∂y2

2 + h3

´ dy

 ,

and the vector

b(x) =

" R

yλh − h3 ∂v∂y31dy

R

yh3 ∂v∂y32dy

#

; and finally we construct the equation

x· [A(x)∇xp0] = ∇x· b(x). (3.19) Summarizing, we are looking for ∇xp0. For that, we need to solve (3.19), but before, we need to have A(x) and b(x); therefore, we solve first the cell problems, ”equations (3.14), (3.15) and (3.16)” that will give v1, v2and v3, then we apply this solutions to find A(x) and b(x). This steps summarize, the homogenization algorithm.

3.2. BOTH ROUGH MOVING SURFACES

If one, or both moving surfaces are rough, the governing Reynolds type equation will then involve time, which means that the hydrodynamic pres- sure between the two surfaces will be changing according to position and time. Here we also use the formal method of multiple scale expansion.

Let

h(x, y, t, τ ) = h0(x, t) + h1(y − τ Ul) + h2(y − τ Uu), where x = (x1, x2), y = (y1, y2), y = x

ε, assuming that ui is constant and τ = t

ε. By using the auxiliary function h define the film thickness hε by hε(x, t) = h(x, y, t, τ ) and the governing Reynolds equation will have the form

γ∂hε

∂t + λ∂hε

∂x1 − ∇ ·¡

h3ε∇pε¢

= 0, (3.20)

γ = 12η and λ = 6ηu, u = Ul+ Uu. Lets perturb a litle our system by assuming the multiple scale expansion of the solution pε

pε(x, y, t, τ ) = p0(x, y, t, τ ) + εp1(x, y, t, τ ) + ε2p2(x, y, t, τ ) + · · · , (3.21)

(30)

pi = pi(x, y, t, τ ). The elements in this equation are defined on the same way as in the preview section and with the same assumptions. The Chain rule then implies that, for

fε= f (x, y, t, τ ) the partial derivative with respect to xj becomes:

∂fε

∂xj = µ∂f

∂xj + ε−1∂f

∂yj

(x, y, t, τ ) , (3.22)

and ∂fε

∂t = µ∂f

∂t + ε−1∂f

∂τ

(x, y, t, τ ) . (3.23) Using (3.21), (3.22) and (3.23) into (3.20) we get:

γ µ

∂t+ ε−1

∂τ

h + λ

µ

∂x1 + ε−1

∂y1

h

¡

x+ ε−1y¢

·£ h3¡

x+ ε−1y¢ ¡

p0+ εp1+ ε2p2+ · · ·¢¤

= 0, that can be expanded,

γ µ

∂t+ ε−1

∂τ

h + λ

µ

∂x1 + ε−1

∂y1

h (3.24)

x·¡ h3x¢

+ ε−1x·¡ h3y¢

+ ε−1y·¡ h3x¢

+ ε−2y·¡

h3y¢¤

(pε), and simplifying by defining

A0 = ∇y·¡ h3y¢

, A1 = ∇y·¡

h3x¢

+ ∇x·¡ h3y¢

, A2= ∇x·¡

h3x¢

; and applying them into (3.24) we will have:

γ µ

∂t+ ε−1

∂τ

h + λ

µ

∂x1 + ε−1

∂y1

h

A2+ ε−1A1+ ε−2A0¢ ¡

p0+ εp1+ ε2p2+ · · ·¢

= 0.

References

Related documents

This leads us to conclude that neither one of the vortex glass and vortex molasses models is in fact a good model for how the vortices are moving in the sample, and by extension for

In 2008 they created an API called Social Network Architecture (SONAR) [12] that could be installed on various internal applications at IBM that held social information. The API

There are over 500 different quasicrystalline phases that have been found today. These can be separated into different elemental families or in different

Linköping Studies in Science and Technology, Dissertation

of Science and Technology Norrköping 2011 Lars Her logsson Nor rköping 2011 Elec tr olyt e-G ated O rganic T hin-F ilm T ransist ors.. Linköping Studies in Science and

The origin of compressive stress and the dynamics of the early growth

After talking to my respondents about loyalty as a concept, it is clear that loyalty in an organisation is strongly affected by what Holmberg refers to as

Katardjiev, "Micromachined thin film plate acoustic resonators utilizing the lowest order symmetric lamb wave mode," Ultrasonics, Ferroelectrics and Frequency Control, IEEE