• Aucun résultat trouvé

Dynamics of a free boundary problem with curvature modeling electrostatic MEMS

N/A
N/A
Protected

Academic year: 2021

Partager "Dynamics of a free boundary problem with curvature modeling electrostatic MEMS"

Copied!
25
0
0

Texte intégral

(1)

HAL Id: hal-00794036

https://hal.archives-ouvertes.fr/hal-00794036

Submitted on 24 Feb 2013

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Dynamics of a free boundary problem with curvature modeling electrostatic MEMS

Joachim Escher, Philippe Laurencot, Christoph Walker

To cite this version:

Joachim Escher, Philippe Laurencot, Christoph Walker. Dynamics of a free boundary problem with curvature modeling electrostatic MEMS. Transactions of the American Mathematical Society, Amer- ican Mathematical Society, 2015, 367, pp.5693–5719. �hal-00794036�

(2)

ELECTROSTATIC MEMS

JOACHIM ESCHER, PHILIPPE LAURENC¸ OT, AND CHRISTOPH WALKER

Abstract. The dynamics of a free boundary problem for electrostatically actuated microelectrome- chanical systems (MEMS) is investigated. The model couples the electric potential to the deformation of the membrane, the deflection of the membrane being caused by application of a voltage difference across the device. More precisely, the electrostatic potential is a harmonic function in the angular domain that is partly bounded by the deformable membrane. The gradient trace of the electric potential on this free boundary part acts as a source term in the governing equation for the mem- brane deformation. The main feature of the model considered herein is that, unlike most previous research, the small deformation assumption is dropped, and curvature for the deformation of the membrane is taken into account what leads to a quasilinear parabolic equation. The free boundary problem is shown to be well-posed, locally in time for arbitrary voltage values and globally in time for small voltages values. Furthermore, existence of asymptotically stable steady-state configura- tions is proved in case of small voltage values as well as non-existence of steady-states if the applied voltage difference is large. Finally, convergence of solutions of the free boundary problem to the solutions of the well-established small aspect ratio model is shown.

1. Introduction

The focus of this paper is the analysis of a model describing the dynamics of an electrostat- ically actuated microelectromechanical system (MEMS) when the deformation of the devices are not assumed to be small. More precisely, consider an elastic plate held at potentialV and suspended above a fixed ground plate held at zero potential. The potential difference between the two plates generates a Coulomb force and causes a deformation of the membrane, thereby converting electrostatic energy into mechanical energy, a feature used in the design of several MEMS-based devices such as micropumps or microswitches [26]. An ubiquitous phenomenon observed in such devices is the so called “pull-in” instability: a threshold value of the applied voltageVabove which the elastic response of the membrane cannot balance the Coulomb force and the deformable membrane smashes into the fixed plate. Since this effect might either be use- ful or, in contrast, could damage the device, its understanding is of utmost practical importance and several mathematical models have been set up for its investigation [7, 21, 26].

In the following subsection we give a brief description of an idealized device as depicted in Figure 1, where the state of the device is characterized by the electrostatic potential in the region between the two plates and the deformation of the membrane which is not assumed to be small from the outset, cf. [5].

1.1. The Model. To derive the model for electrostatic MEMS with curvature we proceed sim- ilarly to [5, 7, 22]. We consider a rectangular thin elastic membrane that is suspended above a rigid plate. The (x, ˆˆ y, ˆz)-coordinate system is chosen such that the ground plate of dimen- sion [−L,L]×[0,l] in (x, ˆˆ y)-direction is located at ˆz = −H, while the undeflected membrane

Date: February 24, 2013.

2010Mathematics Subject Classification. 35R35, 35M33, 35Q74, 35B25, 74M05.

Key words and phrases. MEMS, free boundary problem, curvature, well-posedness, asymptotic stability, small aspect ratio limit.

1

(3)

with the same dimension [−L,L]×[0,l] in(x, ˆˆ y)-direction is located at ˆz = 0. The membrane is held fixed along the edges in ˆy-direction while the edges in ˆx-direction are free. Assuming homogeneity in ˆy-direction, the membrane may thus be considered as an elastic strip and the

ˆ

y-direction is omitted in the sequel. The mechanical deflection of the membrane is caused by a voltage difference that is applied across the device. The membrane is held at potentialV while the rigid plate is grounded. We denote the deflection of the membrane at position ˆxand time ˆtby

ˆ

u=uˆ(t, ˆˆ x)>−Hand the electrostatic potential at position(x, ˆˆ z)and time ˆtby ˆψ=ψˆ(t, ˆˆ x, ˆz). We do not indicate the time variable ˆtfor the time being. The electrostatic potential ˆψis harmonic, i.e.

ψˆ =0 in Ωˆ(uˆ) (1.1)

and satisfies the boundary conditions

ψˆ(x,ˆ −H) =0 , ψˆ(x, ˆˆ u(xˆ)) =V, xˆ∈(−L,L), (1.2) where

ˆ

Ω(uˆ):={(x, ˆˆ z);−L<xˆ<L,H<zˆ<uˆ(xˆ)}

is the region between the ground plate and the membrane. The total energy of the system constitutes of the electric potential and the elastic energy and reads

E(uˆ) =Ee(uˆ) +Es(uˆ).

Theelectrostatic energy Ee in dependence of the deflection ˆuis given by Ee(uˆ) =−ǫ20

Z L

L Z u(ˆ x)ˆ

H |∇ψˆ(x, ˆˆ z)|2d ˆzd ˆx

withǫ0being the permittivity of free space while theelastic energy Esonly retains the contribution due to stretching (in particular, bending is neglected) and is proportional to the tensionTand to the change of surface area of the membrane, i.e.

Es(uˆ) =T Z L

L

q

1+ (xˆuˆ(xˆ))21

d ˆx . Introducing the dimensionless variables

x= xˆ

L , z= zˆ

H , u= uˆ

H , ψ= ψˆ V

and denoting the aspect ratio of the device byε= H/L, we may write the total energy in these variables in the form

E(u) =TL Z 1

1

q

1+ε2(xu(x))21

dx

ǫ0V

2

Z

(u)

ε2|xψ(x,z)|2+|zψ(x,z)|2d(x,z),

(1.3)

with

Ω(u):={(x,z)∈(−1, 1)×(−1,) : −1<z<u(x)} , so that, formally, the corresponding Euler-Lagrange equations are

0=ε2x xu p1+ε2(xu)2

!

ε2λε2|xψ(x,u(x))|2+|zψ(x,u(x))|2 (1.4) forx∈(−1, 1), where we have set

λ= ǫ0V

2

TLε3 .

(4)

We now take again time into account and derive the dynamics of the dimensionless deflection u = u(t,ˆ x) by means of Newton’s second law. Letting ρ and δ denote the mass density per unit volume of the membrane and the membrane thickness, respectively, the sum over all forces equalsρδ∂2tˆu. The elastic and electrostatic forces, given by the right hand side of equation (1.4), are combined with a damping force of the form−a∂ˆtubeing linearly proportional to the velocity.

This yields

ρ δ ∂2ˆtu+a∂tˆu=ε2x xu p1+ε2(xu)2

!

ε2λε2|xψ(x,u(x))|2+|zψ(x,u(x))|2 . Finally, scaling time based on the strength of damping according to t = ˆ 2/a and setting γ:=

ρδε

a , we derive for the dimensionless deflection the evolution problem γ22tu+tu = x xu

p1+ε2(xu)2

!

λε2|xψ(x,u(x))|2+|zψ(x,u(x))|2 (1.5) for t > 0 and xI := (−1, 1). Instead of considering this hyperbolic equation, however, we assume in this paper that viscous or damping forces dominate over inertial forces, i.e. we assume that γ ≪ 1 and thus neglect the second order time derivative term in (1.5). The membrane displacementu=u(t,x)∈(−1,)then evolves according to

tux p xu 1+ε2(xu)2

!

=−λ ε2|xψ(t,x,u(x))|2+|zψ(t,x,u(x))|2 , (1.6) fort>0 andxIwith clamped boundary conditions

u(t,±1) =0 , t>0 , (1.7)

and initial condition

u(0,x) =u0(x), xI. (1.8)

In dimensionless variables, equations (1.1)-(1.2) read

ε22xψ+2zψ=0 , (x,z)∈(u(t)), t>0 , (1.9) subject to the boundary conditions (extended continuously to the lateral boundary)

ψ(t,x,z) = 1+z

1+u(t,x) , (x,z)∈(u(t)), t>0 . (1.10) In the following we shall focus our attention on (1.6)-(1.10), its situation being depicted in Fig- ure 1.

1.2. Simplified Models. Besides assuming that damping forces dominate over inertial forces and thus reducing equation (1.5) to (1.6), other simplifications of the model above have been considered as well in the literature. For instance, restricting attention to small deformations of the membrane yields a linearized stretching term 2xu in (1.6). The corresponding semilinear evolution problem with (1.6) being replaced by

tu2xu=−λ ε2|xψ(t,x,u(x))|2+|zψ(t,x,u(x))|2 , xI, t>0 , (1.11) is investigated in [6]. It is shown therein that the problem (1.7)-(1.11) is well-posed locally in time. Moreover, solutions exist globally for small voltage values λ while global existence is shown not to hold for high voltage values. It is also proven that, for small voltage values, there is an asymptotically stable steady-state solution. Finally, as the parameterεapproaches zero, the solutions are shown to converge toward the solutions of the so-called small aspect ratio model,

(5)

Figure1. Idealized electrostatic MEMS device.

see (1.13) below. Indeed, lettingε=0 (and applying a potentialVwithV2ε3as suggested in [5]), one can solve (1.9)-(1.10) explicitly for the potentialψ=ψ0, that is,

ψ0(t,x,z) = 1+z

1+u0(t,x) , (t,x,z)∈[0,∞)×I×(−1, 0), (1.12) and the displacementu=u0satisfies

tu02xu0 = −(1+λu0)2, xI, t∈(0,∞), u0(t,±1) = 0 , t∈(0,∞),

u0(0,x) = u0(x), xI.

(1.13)

In the limit ε →0, the free boundary problem is thus reduced to the singular semilinear heat equation (1.13) which has been studied thoroughly in recent years, see [7] for a survey as well as e.g. [8, 9, 13, 14, 15, 16, 18, 22, 25]. It is noteworthy to remark that the picture regarding pull-in voltage for the small aspect ratio model (1.13) is rather complete.

Let us point out that [6] is apparently the first mathematical analysis of the parabolic free boundary problem (1.7)-(1.11) while the corresponding elliptic (i.e. steady-state) free boundary problem is investigated in [20]. Moreover, we shall emphasize that the inclusion of non-small deformations is a feature of great physical relevance and, even though the results presented herein are reminiscent of the ones in [6], the quasilinear structure of (1.6) is by no means a trivial mathematical extension of (1.11).

(6)

1.3. Main Results. To state our findings on (1.6)-(1.10), we introduce the spaces

Wq,D(I):=





{uWq(I);u1) =0} for 1q, 2i , Wq(I) for 0≤< 1

q .

(1.14)

We shall prove the following result regarding local and global existence of solutions:

Theorem 1.1(Well-Posedness). Let q∈ (2,∞) >0, and consider an initial value u0Wq,D2 (I) such that u0(x)>−1for xI. Then, the following are true:

(i) For each voltage value λ > 0, there is a unique maximal solution(u,ψ) to(1.6)-(1.10)on the maximal interval of existence[0,Tmε)in the sense that

uC1 [0,Tmε),Lq(I)C [0,Tmε),Wq,D2 (I) satisfies(1.6)-(1.8)together with

u(t,x)>−1 , (t,x)∈[0,TmεI,

andψ(t)∈W22 Ω(u(t))solves(1.9)-(1.10)onΩ(u(t))for each t∈[0,Tmε).

(ii) If for eachτ>0there isκ(τ)∈(0, 1)such that u(t)≥ −1+κ(τ)andku(t)kWq2(I)κ(τ)1 for t∈[0,Tmε)∩[0,τ], then the solution exists globally, that is, Tmε =.

(iii) If u0(x) ≤ 0 for xI, then u(t,x) ≤0 for(t,x) ∈ [0,TmεI. If u0 = u0(x)is even with respect to xI, then, for all t∈[0,Tmε), u=u(t,x)andψ=ψ(t,x,z)are even with respect to xI as well.

(iv) Givenκ ∈(0, 1), there areλ(κ)>0and r(κ)>0such that Tmε = with u(t,x)≥ −1+κ for (t,x) ∈ [0,∞)×I provided that λ ∈ (0,λ(κ)) and ku0kWq2(I)r(κ). In that case, u enjoys the following additional regularity properties:

uBUCρ([0,∞),Wq,D2ρ(I))∩L([0,∞),Wq,D2 (I)) for someρ>0small.

Note that part (iv) of Theorem 1.1 provides uniform estimates onu in theWq2(I)-norm and ensures thatunever touches down on -1, not even in infinite time. In contrast to the semilinear case considered in [6], the global existence result for the quasilinear equation (1.6) requires ini- tially a small deformation, see also Remark 3.3 below. The proof of Theorem 1.1 is the content of Section 3. It is based on interpreting (1.6) as a abstract quasilinear Cauchy problem which allows us to employ the powerful theory of evolution operators developed in [4]. Let us emphasize at this point that the regularity properties of the right-hand side of equation (1.6) established in [6] are not sufficient to handle the quasilinear character of the curvature operator and we conse- quently have to derive Lipschitz properties of the right-hand side of (1.6) in weaker topologies than in [6]. This is the purpose of Section 2.

Regarding existence and asymptotic stability of steady-state solutions to (1.6)-(1.10) we have a similar result as in [20, Thm. 1] and [6, Thm.1.3].

Theorem 1.2(Asymptotically Stable Steady-State Solutions). Let q∈(2,∞)andε>0.

(i) Letκ∈(0, 1). There areδ=δ(κ)>0and an analytic function [λ7→Uλ]:[0,δ)→Wq,D2 (I)

such that(Uλλ)is the unique steady-state to(1.6)-(1.10)satisfyingkUλkWq,D2 (I)1/κwith Uλ≥ −1+κandΨλW22((Uλ))whenλ∈(0,δ). The steady-state possesses the additional

(7)

regularity

UλC2+α [−1, 1],

ΨλW22 Ω(Uλ)C (Uλ)C2+α (Uλ)∪Γ(Uλ), (1.15) whereα∈[0, 1)is arbitrary andΓ(Uλ)denotes the boundary ofΩ(Uλ)without corners. More- over, Uλis negative, convex, and even with U0=0andΨλ =Ψλ(x,z)is even with respect to xI.

(ii) Let λ ∈ (0,δ). There are ω0,m,R > 0 such that for each initial value u0Wq,D2 (I) with ku0UλkWq,D2 <m,(1.6)-(1.10)has a unique global solution(u,ψ)with

uC1 [0,∞),Lq(I)C [0,∞),Wq,D2 (I), ψ(t)∈W22 Ω(u(t)), t0 , and

u(t,x)>−1 , (t,x)∈[0,∞)×I. Moreover,

ku(t)−UλkWq,D2 (I)+ktu(t)kLq(I)Reω0tku0UλkWq,D2 (I), t0 . (1.16) Part (ii) of Theorem 1.2 shows local exponential stability of the steady-states derived in part (i).

We also point out that the potentialψconverges exponentially toΨλin theW22-norm ast, see Remark 4.1 for a precise statement. The proof of Theorem 1.2 is given in Section 4 and relies on the Implicit Function Theorem for part (i) and the Principle of Linearized Stability for part (ii).

Clearly, Theorem 1.2 is just a local result with respect toλvalues. However, we next show that there is an upper threshold forλabove which no steady-state solution exists. This is expected on physical grounds and is related to the “pull-in” instability already mentioned in the introduction.

Theorem 1.3 (Non-Existence of Steady-State Solutions). Let q ∈ (2,∞) and ε > 0. There is λ¯(ε) > 0 such that, ifλλ¯(ε), then there is no steady-state solution(u,ψ)to(1.6)-(1.10)such that uWq,D2 (I)W22((u)), and u(x)>−1for xI. In addition,λ¯(ε)→2asε0.

Similar results have already been obtained for related models, including the small aspect ratio model [5, 7] and for the stationary free boundary problem corresponding to (1.7)-(1.11), see [20].

The proof of Theorem 1.3 relies on a lower bound onzψ(x,u(x))established in the latter paper and is given in Section 5.

The final issue we address is the connection between the free boundary problem (1.6)-(1.10) and the small aspect ratio limit (1.13). More precisely, we show the following convergence result:

Theorem 1.4 (Small Aspect Ratio Limit). Let λ > 0, q ∈ (2,∞), and let u0Wq,D2 (I) with

1 < u0(x) ≤ 0 for xI. Forε > 0 we denote the unique solution to(1.6)-(1.10) on the maximal interval of existence[0,Tmε)by (uε,ψε). There areτ >0,ε0 >0, andκ0∈ (0, 1)depending only on q and u0such that Tmετwith uε(t)≥ −1+κ0andkuε(t)kWq2(I)κ01for all(t,ε)∈[0,τ]×(0,ε0). Moreover, asε0,

uε −→u0 in C1θ [0,τ],Wq(I), 0<θ<1 , and

ψε(t)1(uε(t))−→ψ0(t)1(u0(t)) in L2 I×(−1, 0), t∈[0,τ], (1.17) where

u0C1 [0,τ],Lq(I)C [0,τ],Wq,D2 (I)

(8)

is the unique solution to the small aspect ratio equation (1.13)and ψ0 is the potential given in(1.12).

Furthermore, there is Λ(u0) > 0 such that the results above hold true for each τ > 0 provided that λ∈(0,Λ(u0)).

The proof is given in Section 6. A similar result has been established in [20, Thm. 2] for the stationary problem and in [6, Thm.1.4] for the semilinear parabolic version (1.11). As in the latter paper, the crucial step is to derive theε-independent lower boundτ >0 onTmε, which is not guaranteed by the analysis leading to Theorem 1.1. The proof of Theorem 1.4 uses several properties of (1.9)-(1.10) with respect to theε-dependence shown in [6].

2. On theEllipticEquation(1.9)-(1.10)

We shall first derive properties of solutions to the elliptic equation (1.9)-(1.10) in dependence of a given (free) boundary. To do so, we transform the free boundary problem (1.9)-(1.10) to the fixed rectangleΩ:=I×(0, 1). More precisely, letq∈(2,∞)be fixed and consider an arbitrary functionvWq,D2 (I)taking values in(−1,). We then define a diffeomorphismTv:Ω(v)→ by setting

Tv(x,z):=

x, 1+z 1+v(x)

, (x,z)∈(v), (2.1) withΩ(v):={(x,z)∈ I×(−1,); 1<z<v(x)}. Clearly, its inverse is

Tv1(x,η) = x,(1+v(x))η1, (x,η)∈, (2.2) and the Laplace operator from (1.9) is transformed to thev-dependent differential operator

Lvw :=ε22xw2η xv(x)

1+v(x) xηw+1+ε2η2(xv(x))2 (1+v(x))2

2ηw

+ε2η

"

2

xv(x) 1+v(x)

2

2xv(x) 1+v(x)

#

ηw.

An alternative formulation of the boundary value problem (1.9)-(1.10) is then Lu(t)φ

(t,x,η) =0 , (x,η)∈, t>0 , (2.3) φ(t,x,η) =η, (x,η)∈, t>0 , (2.4) forφ=ψTu(t)1. With this notation, the quasilinear evolution equation (1.6) forubecomes

tux xu p1+ε2(xu)2

!

=−λ

1+ε2(xu)2 (1+u)2

|ηφ, 1)|2, xI , t>0 , (2.5) where we have usedxφ(t,x, 1) =0 forxIandt>0 due toφ(t,x, 1) =1 by (2.4). The inves- tigation of the dynamics of (2.5) involves the properties of its nonlinear right hand side as well as the properties of the quasilinear curvature term. We shall see that these two features of (2.5) are somewhat opposite as the treatment of the former requires a functional analytic setting in Wq2(I)to handle the second order terms ofLu(t)in (2.3), while a slightly weaker setting has to be chosen to guarantee H ¨older continuity ofuwith respect to time which is required in quasilinear evolution equations (see Remark 3.3 for further details). To account for these features of (2.5) we have to refine the Lipschitz property of the right-hand side of (2.5) derived in [6] as stated in (2.8) below.

Defining forκ ∈(0, 1)the open subset Sq(κ):=

uWq,D2 (I);kukWq,D2 (I)<1/κ and1+κ<u(x)forxI

(9)

ofWq,D2 (I)(defined in (1.14)) with closure Sq(κ) =

uWq,D2 (I);kukW2

q,D(I)1/κ and1+κu(x)forxI

,

the crucial properties of the nonlinear right-hand side of (2.5) are collected in the following proposition:

Proposition 2.1. Let q∈(2,∞) ∈ (0, 1), andε>0. For each v∈Sq(κ)there is a unique solution φvW22()to

Lvφv(x,η) =0 , (x,η)∈, (2.6) φv(x,η) =η, (x,η)∈. (2.7) Ifv is defined by˜ v˜(x) := v(−x)for xI, then φv˜(x,η) =φv(−x,η)for(x,η) ∈ . Moreover, for 2σ∈[0, 1/2), the mapping

gε:Sq(κ)−→W2,D (I), v7−→ 1+ε2(xv)2

(1+v)2 |ηφv, 1)|2

is analytic and bounded with gε(0) =1. Finally, ifξ ∈ [0, 1/2)and ν ∈ [0,(1−)/2), then there exists a constant c1(κ,ε)>0such that

kgε(v)−gε(w)kW2,Dν (I)c1(κ,ε)kvwkW2ξ

q,D(I), v,wSq(κ). (2.8) According to [6, Prop. 2.1] we actually only have to prove (2.8). Notice that this global Lipschitz property is in the weaker topology ofWq,D2ξ(I) instead ofWq,D2 (I) and improves [6, Prop. 2.1] where it was established for ξ = 0. The property (2.8) will be a consequence of a sequence of lemmas. For the remainder of this section we fixε>0,κ∈(0, 1), andq∈(2,∞).

In the following, ifα>1/2 we letW2,Dα ()denote the subspace of elements inW2α()whose boundary trace is zero, and if 0≤α<1/2 we setWα

2,D():=W2α(). We equipW2,D1 ()with the norm

kΦkW2,D1 ():=kxΦk2L2()+kηΦk2L2()

1/2

, and introduce the notation

W2,Dθ():= (W2,Dθ ()), 0≤θ1 .

Lemma 2.2. For each vSq(κ) and FW2,D1() there is a unique solutionΦ ∈ W2,D1 ()to the boundary value problem

−LvΦ =F inΩ, (2.9)

Φ =0 on∂Ω, (2.10)

and there is a constant c2(κ,ε)>0such that

kΦkW2,D1 ()c2(κ,ε)kFkW1

2,D(). (2.11)

Furthermore, if FL2(), thenΦ∈W2,D2 ()and

kΦkW2,D2 ()c2(κ,ε)kFkL2(). (2.12) Proof. According to [12, Def. 1.3.2.3, Eq. (1,3,2,3)], we may write anyFW2,D1()in the form F= f0+xf1+ηf2 with (f0,f1,f2) ∈ L2()3. Consequently, [10, Thm. 8.3] ensures that the

(10)

boundary value problem (2.9)-(2.10) has a unique solutionΦ∈W2,D1 (). Furthermore, takingΦ as a test function in the weak formulation of (2.9)-(2.10) gives

hF,Φi= Z

ε2(xΦ)22ε2η1xv

+v∂xΦηΦ−ε2η ∂x

xv 1+v

ΦηΦ−ε21xv

+v ΦxΦ

d(x,η) +

Z

"

1+ε2η2(xv)2

(1+v)2 (ηΦ)2+2ε2η xv

1+v 2

ΦηΦ

#

d(x,η)

+ Z

ε2η

"

x

xv 1+v

xv

1+v 2#

ΦηΦd(x,η)

= Z

ε2(xΦ)22ε2η1xv

+v∂xΦηΦ+ 1+ε2η2(xv)2

(1+v)2 (ηΦ)2

d(x,η)

Z

ε2 xv 1+v

xΦ−η1xv +v∂ηΦ

Φd(x,η) and thus

Z

"

ε2

xΦ −η1+xvv∂ηΦ 2

+ ηΦ

1+v 2#

d(x,η)

ε2

xv 1+v

L

(I)

xΦ−η1+xvv∂ηΦ L

2(Ω) kΦkL2(Ω) + kFkW1

2,D(Ω)kΦkW2,D1 ().

(2.13)

Note then that by definition ofSq(κ)and Sobolev’s embedding, there isc3>0 such that 1+v(x)≥κ, xI, kvkC1([1,1])cκ3 (2.14) for allvSq(κ). Also, ifζ= (ζ1,ζ2)∈R2, Young’s inequality ensures that, for(x,η)∈,

ε2ζ212ε2

ζ1η xv(x) 1+v(x)ζ2

2

+2ε2η2

xv(x) 1+v(x)

2

ζ22

21+2ε2kxvk2

ε2

ζ1η xv(x) 1+v(x)ζ2

2

+1 2

ζ2

1+v(x) 2!

. Therefore, introducing

ν(κ,ε):= 1 2 min

( ε2κ2

κ2+2ε2c23, κ2 (κ+c3)2

) , we infer from (2.14) that

ν(κ,ε) ζ21+ζ22

ε2

ζ1η xv(x) 1+v(x)ζ2

2

+ ζ2

1+v(x) 2

. (2.15)

Consequently, (2.13), (2.14), and (2.15) give ε2

xΦ − η1+xvv ηΦ

2 L2()

+

ηΦ 1+v

2 L2()

ε2κc32

xΦ − η1+xvv∂ηΦ L

2()kΦkL2()

+kFkW1 2,D()

pν(κ,ε) ε

2

xΦ− η1xv +v∂ηΦ

2 L2()

+

ηΦ 1+v

2 L2()

!1/2

,

(11)

whence, using again (2.15),

kΦkW2,D1 () ≤ p ε ν(κ,ε)

c3

κ2kΦkL2()+kFkW1 2,D()

ν(κ,ε) . (2.16)

We now proceed as in [10, Lem. 9.17] and argue by contradiction to show (2.11) (see also [6, Lem. 6.2] for a similar argument in a slightly different functional setting). The last statement of

Lemma 2.2 is proved in [6, Lem. 6.2].

Now, introducing

fv(x,η):=Lvη=ε2η

"

2

xv(x) 1+v(x)

2

2xv(x) 1+v(x)

#

, (x,η)∈, (2.17) forvSq(κ)given, we readily deduce that fvL2()with

kfvkL2()c4(κ,ε). (2.18) Consequently, Lemma 2.2 provides a unique solutionΦvW2,D2 ()to

− LvΦv(x,η) = fv, (x,η)∈, Φv(x,η) =0 , (x,η)∈. Clearly, defining

φv(x,η):=Φv(x,η) +η, (x,η)∈¯ , (2.19) gives then the unique solutionφvW22()to (2.6)-(2.7). To prove a Lipschitz dependence ofφv

onvSq(κ), we introduce a bounded linear operator

A(v)∈ L W2,D1 (),W2,D1()∩ L W2,D2 (),L2() by setting

A(v)Φ:=−LvΦ, ΦW2,D1 ().

Note thatA(v)is invertible according to Lemma 2.2 and thatΦv = A(v)1fv. For the inverse A(v)1we have:

Lemma 2.3. Givenθ∈[0, 1]\ {1/2}, there is a constant c5(κ,ε)>0such that kA(v)1kL(Wθ1

2,D(),W2,Dθ+1())c5(κ,ε), vSq(κ).

Proof. Due to Lemma 2.2,A(v)1belongs for eachvSq(κ)to bothL(W2,D1(),W2,D1 ())and L(L2(),W2,D2 ())with

kA(v)1kL(W1

2,D(),W2,D1 ())+ kA(v)1kL(L2(),W2,D2 ())2c2(κ,ε). Hence, using complex interpolation, we derive forθ∈[0, 1],

kA(v)1kL [W1

2,D(),L2()]θ,[W2,D1 (),W2,D2 ()]θ2c2(κ,ε),

and it then remains to characterize the interpolation spaces forθ ∈ [0, 1]\ {1/2}. For this we first invoke [28, Thm. 1.1.11] to obtain that

[W2,D1(),L2()]θ = [(W2,D1 ()),(L2())]θ = [W2,D1 (),L2()]θ

with equivalent norms and hence, using [12, Cor. 1.4.4.5] and [19, Prop. 2.11 & Prop. 3.3] to characterize the last interpolation space,

[W2,D1(),L2()]θ= (W2,D1θ())=W2,Dθ1().

(12)

Finally, since the embedding

[W2,D1 (),W2,D2 ()]θ֒→W2,Dθ+1()

is obviously continuous by [28, Thm. 4.3.2/2] and definition of interpolation, the assertion fol-

lows.

Next, we show that A(v) is Lipschitz continuous with respect to v in a suitable topology.

More precisely:

Lemma 2.4. Givenξ∈[0,(q1)/q)andα∈(ξ, 1), there exists c6(κ,ε)>0such that kA(v)− A(w)kL(W2,D2 (),W2,Dα())c6(κ,ε)kvwkW2ξ

q (I), v,wSq(κ).

Proof. Considerv,wSq(κ)andΦ∈W2,D2 (). ThenA(v)ΦL2()֒→W2,Dα()and so, for anyϕW2,Dα (),

Z

A(v)− A(w)Φ ϕd(x,η) =2ε2 Z

η xv

1+v1xw +w

xηΦϕd(x,η)

Z

1+ε2η2(xv)2

(1+v)21+ε2η2(xw)2 (1+w)2

2ηΦϕd(x,η)

2 Z

η

"

xv 1+v

2

xw

1+w 2#

ηΦϕd(x,η) +ε2

Z

η 2xv

1+v

2xw 1+w

ηΦϕd(x,η)

= :J1+J2+J3+J4.

Sinceξ∈[0,(q1)/q)it follows from the continuous embeddingWq2ξ(I)֒→W1(I)that

|J1| ≤2ε2

xv

1+v1+xww L

(I)kxηΦkL2()kϕkL2()

ε2c(κ)kvwkW2ξ

q (I)kΦkW2,D2 ()kϕkW2,Dα (). Similarly, we have

|J2|+|J3| ≤(1+ε2)c(κ) kvwkW2ξ

q (I) kΦkW2,D2 ()kϕkW2,Dα ().

Finally, we infer from the embedding W21() ֒→ L2q/(q2)() and the fact that (Wqξ(I)) = Wqξ(I)sinceξ∈[0,(q1)/q), see, e.g., [3, Eq. (5.14)], that

|J4| ≤ε2 Z

2x(vw)η ∂ηΦϕ 1+v d(x,η)

+ε2k2xwkLq(I)

1

1+v1 1 +w

L

(I)kηΦkL2q/(q2)()kϕkL2()

ε2k2x(vw)kWξ q (I)

1 1+v

Z 1

0 η ∂ηΦ(·,η)ϕ,η)dη Wξ

q(I)

+ε2c(κ)kvwkW2ξ

q (I)kηΦkW1

2,D()kϕkW2,Dα ().

Now, choosingβ∈[ξ,α), pointwise multiplication is continuous as mappings Wq2(IW2β(I)֒→Wqξ(I), C2(¯W21(W2α()֒→W2β()

Références

Documents relatifs

The occurrence of a finite time singularity is shown for a free boundary problem mod- eling microelectromechanical systems (MEMS) when the applied voltage exceeds some value.. The

A variational approach is employed to find stationary solutions to a free boundary problem mod- eling an idealized electrostatically actuated MEMS device made of an elastic plate

Local and global well-posedness results are obtained for the corresponding hyperbolic and parabolic evolution problems as well as a criterion for global existence excluding

Idealized modern MEMS devices often consist of two components: a rigid ground plate and a thin and deformable elastic membrane that is held fixed along its boundary above the

Lieberman, Boundary regularity for solutions of degenerate elliptic equations, Nonlinear Anal. Magenes, Problemi ai limiti

Thus, Theorem 1.1 may be seen as a rigidity result on an fully nonlinear fluid jet model which determines the shape of the fluid (i.e., the domain ) and the stream lines of the

WOLANSKI, Uniform estimates and limits for a two phase parabolic singular perturbation problem, Indiana Univ. WOLANSKI, Pointwise and viscosity

angle between free and fixed boundary (fig. Using blow-up techniques we prove that the free boundary is tan-.. Chang- ing the porous medium afterwards one can construct