• Aucun résultat trouvé

A parabolic free boundary problem modeling electrostatic MEMS

N/A
N/A
Protected

Academic year: 2021

Partager "A parabolic free boundary problem modeling electrostatic MEMS"

Copied!
24
0
0

Texte intégral

(1)

HAL Id: hal-00757167

https://hal.archives-ouvertes.fr/hal-00757167

Submitted on 26 Nov 2012

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.

A parabolic free boundary problem modeling electrostatic MEMS

Joachim Escher, Philippe Laurencot, Christoph Walker

To cite this version:

Joachim Escher, Philippe Laurencot, Christoph Walker. A parabolic free boundary problem modeling electrostatic MEMS. Archive for Rational Mechanics and Analysis, Springer Verlag, 2014, 211, pp.389- 417. �10.1007/s00205-013-0656-2�. �hal-00757167�

(2)

JOACHIM ESCHER, PHILIPPE LAURENC¸ OT, AND CHRISTOPH WALKER

Abstract. The evolution problem for a membrane based model of an electrostatically actuated microelectromechanical system (MEMS) is studied. The model describes the dynamics of the mem- brane displacement and the electric potential. The latter is a harmonic function in an angular do- main, the deformable membrane being a part of the boundary. The former solves a heat equation with a right hand side that depends on the square of the trace of the gradient of the electric po- tential on the membrane. The resulting free boundary problem is shown to be well-posed locally in time. Furthermore, solutions corresponding to small voltage values exist globally in time 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, the small aspect ratio limit is rigorously justified.

1. Introduction

An idealized electrostatically actuated microelectromechanical system (MEMS) consists of a rigid ground plate above which a thin and deformable elastic membrane is suspended that is held fixed along its boundary, see Figure 1. Applying a voltage difference between the two components induces displacements of the membrane and thus transforms electrostatic energy into mechanical energy, a feature that has applications in the design of transistors, switches, or micro-pumps, for instance. There is, however, an upper limit for the applied voltage potential beyond which the electrostatic force cannot be balanced by the elastic response of the membrane which then touches down on the rigid plate. This phenomenon is usually referred to as “pull- in” instability. Estimating this threshold value is an important issue in applications as it may be a desirable feature of the device in some situations (e.g. switches, micropumps) or possibly damage the device in others. Mathematical models have been set up for that purpose, and we refer the reader e.g. to [26, 27, 28] and the references therein for a more detailed account of the physical background and the modeling aspects of such devices.

Denoting the displacement of the membrane and the electrostatic potential in the device byu andψ, respectively, we consider here the idealized situation where the applied voltage and the permittivity of the membrane are constant (normalized to one), and there is no variation in the horizontal direction orthogonal to thex-direction of bothψand u. Under appropriate scalings, the rigid ground plate is at z = −1, and the undeflected membrane at z = 0 is fixed at the boundaryx =−1 andx =1 of the intervalI:= (−1, 1), see Figure 1. Lettingεdenote the aspect ratio of the device before scaling, i.e. the ratio of the undeformed gap size to the device length, the membrane displacementu=u(t,x)∈(−1,)evolves according to

tu2xu=−λ ε2|xψ(t,x,u)|2+|zψ(t,x,u)|2 , xI, t>0 , (1.1)

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

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

1

(3)

Figure1. Idealized electrostatic MEMS device.

with clamped boundary conditions

u(t,±1) =0 , t>0 , (1.2)

and initial condition

u(0,x) =u0(x), xI. (1.3)

The dimensionless electrostatic potentialψ=ψ(t,x,z)satisfies Laplace’s equation

ε22xψ+2zψ=0 , (x,z)∈(u(t)), t>0 , (1.4) in the region

Ω(u(t)):={(x,z)∈ I×(−1,) : −1<z<u(t,x)}

between the rigid ground plate atz=−1 and the deflected membrane. The boundary conditions forψare then

ψ(t,x,z) = 1+z

1+u(t,x) , (x,z)∈(u(t)), t>0 . (1.5) Equation (1.1) corresponds to the situation in which viscous forces dominate over inertial forces in the system, e.g. see [6, 27]. Also, deformations due to bending are neglected in (1.1).

Of particular importance in the model is the parameter λ > 0 which characterizes the rela- tive strengths of electrostatic and mechanical forces and is proportional to the applied voltage.

According to the above discussion, the pull-in instability is expected to take place for λ large enough.

The analysis of (1.1)-(1.5) turns out to be rather complex since (1.4) is a free boundary prob- lem: indeed, the domain between the rigid ground plate and the elastic membrane changes with time. Due to this, equations (1.1) and (1.4) are strongly coupled. However, a common assump- tion made in mathematical analysis hitherto is a vanishing aspect ratioεthat reduces the free boundary problem to a heat equation with a right hand side involving a singularity when the

(4)

membrane touches down on the ground plate. More precisely, settingε=0 allows one to solve (1.4)-(1.5) explicitly for the potentialψ=ψ0, that is,

ψ0(t,x,z) = 1+z

1+u0(t,x) , (t,x,z)∈[0,∞)×I×(−1, 0), (1.6) where the displacementu=u0now satisfies the so-called small aspect ratio model

tu02xu0 = − λ

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

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

(1.7)

Several mathematical results have been obtained for (1.7), including a characterization of the critical value ofλwhich corresponds to the value beyond which no steady-state exists as well as a possible space dependence of the permittivity of the membrane, see, e.g., [5, 8, 22, 27] for the stationary problem and [5, 7, 9, 16, 13, 14, 18, 27] for the evolution problem. Inertial effects are taken into account in [15, 19].

To the best of our knowledge, the first analytical research without assumption of a small aspect ratio and thus dedicated to the original free boundary problem (1.1)-(1.5) is [21], where the existence of steady-states has been established for small voltage valuesλand a non-existence result for steady-states is obtained for large values ofλ.

Here, we address the evolution problem. A rough summary of our results reads as follows:

We prove the local well-posedness of (1.1)-(1.5) for all voltage values and show that the solu- tions exist globally in time provided the voltage value is sufficiently small. In contrast to the stationary case [21] it turns out that aW2(I)-setting is no longer suitable for the u-component of (1.1)-(1.5). This is due to the fact that the heat semigroup does not enjoy suitable properties in L(I). Instead, we are therefore lead to work in the framework ofWq2(I)-spaces for q< ∞, which generates additional difficulties as now2xu may become unbounded. For small voltage values we further prove that there is a locally asymptotically stable steady-state. For high voltage values we prove that global existence of solutions does not hold. In addition, we analyze the behavior of the solutions as the small aspect ratioε→0, showing convergence towards (1.7) as expected from a formal analysis.

To state precisely our results we introduce forq∈[2,∞)andκ∈(0, 1)the set Sq(κ):=

uWq,D2 (I);kukW2q,D(I)<1/κ and1+κ<u(x)forxI

,

where Wq,D(I) := {uWq(I);u1) = 0} for 2α ∈ (1/q, 2] and Wq,D(I) := Wq(I) for 0≤<1/q. The local existence result now reads:

Theorem 1.1 (Local Well-Posedness). Let q ∈ (2,∞), ε > 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(1.1)-(1.5)on the maximal interval of existence[0,Tmε)in the sense that

uC1 [0,Tmε),Lq(I)C [0,Tmε),Wq,D2 (I) satisfies(1.1)-(1.3)together with

u(t,x)>−1 , (t,x)∈[0,TmεI,

andψ(t)∈Wq2 Ω(u(t))solves(1.4)-(1.5)onΩ(u(t))for each t∈[0,Tmε).

(5)

(ii) If for eachτ>0there isκ(τ)∈ (0, 1)such that u(t)∈Sq(κ(τ))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.

The proof of Theorem 1.1 is performed as follows. We first transform the Laplace equation (1.4) to a fixed rectangle which results in an elliptic boundary value problem with non-constant coefficients depending on u and its derivatives up to order 2. Solving this elliptic equation (for a given u) allows us to interpret the full free boundary problem as a nonlocal semilinear heat equation foru (see (2.5)). We then employ a fixed point argument to solve this evolution problem. Since the nonlinearity in the u-equation depends on the trace of the gradient of the potential, two ingredients are essential: precise estimates based on the regularizing effects of the heat semigroup and elaborated investigations of the properties of the solution to the transformed elliptic problem. The proof is given in Section 2.

We now address global existence issues. From a physical viewpoint a “pull-in” instability occurs for high voltage values. Accordingly, for large values ofλsolutions cease to exist globally while solutions corresponding to smallλvalues exist globally in time. More precisely, we have:

Theorem 1.2 (Global Existence). Let q ∈ (2,∞), ε > 0, λ > 0, and let u0Wq,D2 (I) satisfy

1 < u0(x) ≤ 0 for xI. Let (u,ψ)be the corresponding solution to (1.1)-(1.5) on the maximal interval of existence[0,Tmε).

(i) Givenκ∈(0, 1)there existsλ :=λ(κ,ε)>0andκ0:=κ0(κ,ε)>0such that Tmε =and u(t)∈Sq(κ0)for t0provided that u0Sq(κ)andλ∈(0,λ).

(ii) There isλ(ε)>0depending only onεsuch that Tmε <∞providedλ>λ(ε).

Note that part (i) of Theorem 1.2 provides uniform estimates on u in the Wq2(I)-norm and ensures that u never touches down on -1, not even in infinite time. Its proof is contained in Section 2 and it is a consequence of the above mentioned fixed point argument. The second part of Theorem 1.2 is proven in Section 3 by constructing a suitable strict Lyapunov functional.

Let us mention that similar results as stated in Theorem 1.2 are known to hold for the small aspect ratio model (1.7), see [7, 9]. However, the nonlocal features of (1.1)-(1.5) prevents one from using similar techniques and we thus have to develop an alternative approach. Also, there is a qualitative difference of the interpretation of the finiteness ofTmε in Theorem 1.2(ii). Indeed, according to Theorem 1.1, Tmε < ∞ implies that the Wq2(I)-norm of u blows up or u touches down on −1 in finite time. This is in clear contrast to the small aspect ratio model (1.7) for which touchdown is the only mechanism for a finite time singularity. The difference stems from the fact that in (1.7) the nonlinearity is of zero order while for the free boundary problem (1.1)-(1.5) the nonlocal nonlinearity is rather of order “3/2” in theLq-sense (see Proposition 2.1).

Nevertheless, we strongly believe that finite time touchdown occurs in the present model as well whenTmε is finite.

We next turn to stability of steady-states. This is a delicate issue since it is expected in analogy to what is known for the small aspect ratio model [5, 27] that there are two steady states for small λvalues. In [21] it was shown that there is at least one steady-state to (1.1)-(1.5) for small values ofλ(and none for largeλ). We shall refine this result here and prove that, providedλis small, this steady-state is unique with a first component in the set Sq(κ) and locally asymptotically stable.

Theorem 1.3(Asymptotic Stability). Let q∈(2,∞)>0, andκ∈(0, 1).

(6)

(i) There are δ = δ(κ) > 0 and an analytic function [λ 7→ Uλ] : [0,δ) → Wq,D2 (I) such that (Uλλ) is for each λ ∈ (0,δ) the unique steady-state to (1.1)-(1.5) with UλSq(κ) and ΨλW22((Uλ)). Moreover, Uλis negative, convex, and even forλ∈(0,δ)and U0=0.

(ii) Let λ ∈ (0,δ). There are ω0,r,R > 0 such that for each initial value u0Wq,D2 (I) with ku0UλkW2

q,D

<r, the solution(u,ψ)to(1.1)-(1.5)exists globally in time and

ku(t)−UλkWq,D2 (I)+ktu(t)kLq(I)Reω0tku0UλkWq,D2 (I), t0 . (1.8) The first part of Theorem 1.3 is a consequence of the Implicit Function Theorem while the sec- ond part follows from the Principle of Linearized Stability, and the proofs are given in Section 4.

We shall point out that Theorem 1.3 provides uniqueness of steady-states with first components in Sq(κ)for fixedλsmall. A result in this spirit is also shown in [8, Thm.5.6]. But, as pointed out before, for the small aspect ratio model (1.7) it is known that below the critical threshold there are exactly two steady-states. If this would turn out to be true for the free boundary problem as well, that is, if there would be another smooth branch of steady-states emanating from λ = 0, say,Vλ 6=Uλ, then the fact thatSq(κ1)⊂Sq(κ2)for 0<κ2 <κ1<1 would imply thatδ(κ0 asκ ց0 in Theorem 1.3. Obviously,Vλ ∈/Sq(κ)forλ<δ(κ)and thus, asλց0, the minimum ofVλhas to approach−1 or theWq2-norm ofVλhas to blow up1.

We also note thatψconverges exponentially toΨλin theW22-norm ast, see Corollary 4.1 for a precise statement. Finally, both components of the steady-state enjoy more regularity than stated, see [21, Cor.10].

More insight in the connection between the free boundary model and its small aspect ratio limit is offered in the next theorem. Indeed, we show that the solution(u,ψ) = (uε,ψε)to (1.1)- (1.5) provided by Theorem 1.1 converges to the solution(u0,ψ0)of the small aspect ratio model (1.6), (1.7) asε→0. This gives a rigorous justification of the formal derivation.

Theorem 1.4(Small Aspect Ratio Limit). Let λ > 0, q ∈ (2,∞) ∈ (0, 1), and let u0Sq(κ) with u0(x) ≤0 for xI. Forε > 0let(uε,ψε)be the unique solution to(1.1)-(1.5)on the maximal interval of existence[0,Tmε). There areτ> 0,ε0 >0, andκ0∈ (0, 1)depending only on q andκsuch that Tmετand uε(t)∈Sq(κ0)for all(t,ε)∈[0,τ]×(0,ε0). Moreover, the small aspect ratio equation (1.7)has a unique solution

u0C1 [0,τ],Lq(I)C [0,τ],Wq,D2 (I) satisfying u0(t)∈Sq(κ0)for all t∈[0,τ]and such that the convergences

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], (1.9) hold as ε0, whereψ0is the potential given in(1.6). Furthermore, there is Λ(κ) > 0 such that the results above hold true for eachτ>0provided thatλ∈(0,Λ(κ)).

A similar result has been established for the stationary problem in [21, Theorem 2] and the proof of Theorem 1.4 is performed along the same lines provided one ensures anε-independent lower boundτ>0 onTmε. In addition, in [21] we took advantage of the fact that aW2(I)-bound is available for solutions to the stationary problem. We refine the arguments here by showing that aWq2(I)-bound is sufficient forq>2.

1For the case of the small aspect ratio model,Vλapproaches theV-shaped functionx7→ |x| −1 asλց0.

(7)

2. Local andGlobalWell-Posedness: Proof ofTheorem1.1andTheorem1.2(i) The starting point for the proof of Theorem 1.1 is to transform the free boundary problem (1.4)-(1.5) 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 diffeomorphism Tv:=(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 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. The boundary value problem (1.4)-(1.5) is then obviously equivalent to

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 evolution equation (1.1) forubecomes

tu2xu=−λ

1+ε2(xu)2 (1+u)2

|ηφ, 1)|2, xI, t>0 , (2.5) after noticing that we havexφ(t,x, 1) =0 forxIand t>0 due to φ(t,x, 1) =1 by (2.4). To set the stage for the proof of Theorem 1.1 we first observe:

Proposition 2.1. Letκ∈(0, 1)andε>0. For each v∈Sq(κ)there is a unique solutionφvW22() to

Lvφv(x,η) =0 , (x,η)∈, (2.6) φv(x,η) =η, (x,η)∈. (2.7) In addition, defining v by˜ v˜(x) := v(−x)for xI, we haveφv˜(x,η) = φv(−x,η)for(x,η) ∈ . Moreover, for2σ∈[0, 1/2), the mapping

gε:Sq(κ)−→W2,D (I), v7−→ 1+ε2(xv)2

(1+v)2 |ηφv, 1)|2 is analytic, globally Lipschitz continuous, and bounded with gε(0) =1.

The proof of Proposition 2.1 shares some common steps with that of [21, Lem. 5 & 6], but requires further developments, in particular establishing the Lipschitz continuity of gε which was not needed in [21]. We first derive suitable properties of the operatorLvforvin the closure

Sq(κ) =

uWq,D2 (I);kukW2

q,D(I)1/κ and1+κu(x)forxI

ofSq(κ), which we gather in the next lemma.

(8)

Lemma 2.2. Letκ ∈(0, 1)andε>0. For each v∈ Sq(κ)and FL2(), there is a unique solution Φ∈W2,D2 ()to the boundary value problem

−LvΦ =F inΩ, (2.8)

Φ =0 on∂Ω. (2.9)

Moreover, there is a constant c1(κ,ε)>0depending only on q,κ, andεsuch that

kΦkW22()c1(κ,ε)kFkL2(). (2.10) Proof. First note that the definition ofSq(κ) and Sobolev’s embedding theorem guarantee the existence of some constantc0>0 depending only onqsuch that, forvSq(κ),

1+v(x)≥κ, xI, and kvkC1([1,1])cκ0 . (2.11) It follows from the proof of [21, Lem. 5] that, due to (2.11), the operator −Lv is elliptic with ellipticity constant ν(κ,ε) > 0 being independent of vSq(κ). Moreover, writing −Lv in divergence form,

−Lvw=−x a11(v)xw+a12(v)ηw

η a21(v)xw+a22(v)ηw +b1(v)xw+b2(v)ηw,

with

a11(v):=ε2, a22(v):= 1+ε2η2|xv(x)|2 (1+v(x))2 , a12(v):=−ε2η xv(x)

1+v(x), a21(v):=a12(v), b1(v):=ε2 xv(x)

1+v(x), b2(v):=−ε2η

xv(x) 1+v(x)

2

, we see from (2.11) and the definition ofSq(κ)that

2

i,j=1

kaij(v)kWq1()+

2

i=1

kbi(v)kL(Ω)c2(κ,ε) (2.12) for allvSq(κ). Moreover, the embedding ofWq1(I)inC([−1, 1])ensures thataij(v)belongs to C()for 1≤i,j2 andvSq(κ). It then follows from [10, Thm. 8.3] that, givenvSq(κ)and FL2(), the boundary value problem (2.8)–(2.9) has a unique weak solutionΦ ∈ W2,D1 (). Furthermore, the regularity ofΦand (2.12) ensure thatG:=Fb1(v)xΦ−b2(v)ηΦbelongs toL2(), and we are in a position to apply [20, Chapt. 3, Thm. 9.1] to conclude thatΦis actually the unique solution inW2,D2 ()to the boundary value problem

−L0vΦ=G in Ω, Φ=0 on Ω, whereL0vdenotes the principal part of the operatorLv, that is,

−L0vw :=−x a11(v)xw+a12(v)ηw

η a21(v)xw+a22(v)ηw .

In addition, it follows from [20, Chapt. 3, Thm. 10.1] that there is a constantc3(κ,ε)>0 depend- ing only onq,κ, andεsuch that

kΦkW22()c3(κ,ε) kΦkL2()+kGkL2()

.

(9)

Combining the previous inequality with (2.12) and the inequality kxΦk2L2()+kηΦk2L2()=−

Z

Φ

2xΦ+2ηΦ

d(x,η)≤δ2kΦk2W2

2()+ 1

2 kΦk2L2()

which is valid for allδ>0, we are led to

kΦkW22()c3(κ,ε) hkΦkL2()+kFkL2()+c2(κ,ε) kxΦkL2()+kηΦkL2()

i

c3(κ,ε)

kΦkL2()+kFkL2()+2c2(κ,ε)

δkΦkW22()+ 1

2δ kΦkL2()

whence, after choosingδsufficiently small,

kΦkW22()c4(κ,ε) kΦkL2()+kFkL2()

. (2.13)

We finally prove (2.10) and argue as in the proof of [10, Lemma 9.17]. Assume for contradiction that (2.10) is not true. Then, for eachn ≥1, there arevnSq(κ),ΦenW2,D2 (),Φen 6≡ 0, and FenL2()such that

−LvnΦen=Fen in Ω and eΦnW2

2()n eFn

L

2() . (2.14) SettingΦn :=Φen/ eΦnL

2()andFn :=Fen/ eΦnL

2(), we realize that (2.13) and (2.14) imply

−LvnΦn = Fn in Ω, ΦnW2,D2 (), (2.15)

kΦnkL2() = 1 , (2.16)

and

nkFnkL2()≤ kΦnkW22()c4(κ,ε) kΦnkL2()+kFnkL2()

=c4(κ,ε) 1+kFnkL2()

. Consequently, we have forn2c4(κ,ε),

nkFnkL2(Ω)2c4(κ,ε) and kΦnkW22()

1+ 2 n

c4(κ,ε). (2.17) SinceW22() and Wq2(I)are compactly embedded in W21() and C1([−1, 1]), respectively, we infer from (2.17) and the boundedness of Sq(κ) in Wq2(I) that there are (Φ,v) ∈ W2,D2 (Wq,D2 (I)and a subsequence of(Φn,vn)n (not relabeled) such that

(Φn,vn)⇀(Φ,v) in W2,D2 (Wq,D2 (I), (2.18) (Φn,vn)−→(Φ,v) in W2,D1 (C1([−1, 1]). (2.19) It readily follows from (2.18) and (2.19) thatvSq(κ)and

aij(vn),bi(vn)−→ aij(v),bi(v) in C([−1, 1]) (2.20) for all 1 ≤i,j2. SinceFn −→0 in L2()by (2.17), the convergences (2.18), (2.19), and (2.20) allow us to pass to the limit as n in the weak formulation of (2.15) and conclude that Φ ∈ W2,D2 ()is a weak solution to −LvΦ = 0 inΩ. Using again [10, Thm. 8.3], this implies Φ≡0 contradictingkΦkL2()=1 as follows from (2.16) and (2.19).

Proof of Proposition 2.1. ForvSq(κ)and(x,η)∈, we set fv(x,η):=Lvη=ε2η

"

2

xv(x) 1+v(x)

2

2xv(x) 1+v(x)

# .

(10)

Sinceq>2, the function fvbelongs toL2()with

kfvkL2()c5(κ,ε), (2.21) and Lemma 2.2 ensures that there is a unique solutionΦvW2,D2 ()to

−LvΦv = fv inΩ, (2.22)

Φv =0 onΩ, (2.23)

satisfying

kΦvkW22()c1(κ,ε)kfvkL2(). (2.24) Lettingφv(x,η) =Φv(x,η) +ηfor(x,η)∈ , the functionφv obviously solves (2.6)-(2.7), and from (2.21) and (2.24) we obtain

kφvkW22()c6(κ,ε). (2.25) In addition, if vSq(κ) and ˜v denotes the function defined by ˜v(x) := v(−x) for xI, we obviously have ˜vSq(κ)and the properties ofLv˜ensure that(x,η)7−→φv(−x,η)solves (2.22)- (2.23) with ˜v instead of v. The uniqueness of the solution to (2.22)-(2.23) then readily implies thatφv˜(x,η) =φv(−x,η)for(x,η)∈and thus that

gε(v˜)(x) =gε(v)(−x), xI. (2.26) Next, givenvSq(κ), we define a bounded linear operatorA(v)∈ L W2,D2 (),L2()by

A(v)Φ:=−LvΦ, ΦW2,D2 ().

Lemma 2.2 guarantees that A(v)is invertible with inverseA(v)1 ∈ L L2(),W2,D2 ())satis- fying

A(v)1

L L2(),W2,D2 ())c1(κ,ε). (2.27) We then note that

kA(v1)− A(v2)kL(W2,D2 (),L2())c7(κ,ε)kv1v2kWq2(I), v1,v2Sq(κ), (2.28) which follows from the definition ofLvand the continuity of pointwise multiplication

Wq1(IWq1(I)֒→Wq1(I)֒→L(I)

except for the terms involving2xvi,i=1, 2, where continuity of pointwise multiplication Lq(W21()֒→L2()

is used. Now, forv1,v2Sq(κ), we infer from (2.27) and (2.28) that kA(v1)1− A(v2)1kL(L2(),W2,D2 ())

A(v1)1

L(L2(),W2,D2 ()) kA(v2)− A(v1)kL(W2,D2 (),L2())

A(v2)1

L(L2(),W2,D2 ())

c1(κ,ε)2c7(κ,ε)kv1v2kWq2(I),

which, combined with (2.21), the observation that 0∈Sq(κ)and

kfv1fv2kL2()c8(κ,ε)kv1v2kWq2(I), v1,v2Sq(κ),

(11)

ensures that

kφv1φv2kW22()=kΦv1Φv2kW22()=kA(v1)1fv1− A(v2)1fv2kW2,D2 ()

c1(κ,ε)2c7(κ,ε)kv1v2kWq2(I)kfv1kL2()

+ c8(κ,ε)kv1v2kW2q(I)kA(v2)1kL(L2(),W2,D2 ())

c9(κ,ε)kv1v2kWq2(I) (2.29) forv1,v2Sq(κ). We may then invoke [24, Chapt. 2, Thm. 5.4] and the continuity of pointwise multiplication

W21/2(IW21/2(I)֒→W21(I) for 2σ1<1/2 according to [1, Thm. 4.1] to conclude that the mapping

Sq(κ)→W21(I), v7→ηφv, 1)2 (2.30) is globally Lipschitz continuous. Thanks to the continuity of the embedding ofWq2(I)inW1(I), the mapping

Sq(κ)→Wq1(I), v7→ 1+ε2(xv)2

(1+v)2 (2.31)

is globally Lipschitz continuous with a Lipschitz constant depending only on κ and ε, and the Lipschitz continuity of gε stated in Proposition 2.1 follows at once from (2.30), (2.31), and continuity of pointwise multiplication

Wq1(IW21(I)֒→W2(I) =W2,D(I),

where 2σ < 1 < 1/2, see again [1, Thm. 4.1]. Finally, to prove that gε is analytic, we note that Sq(κ) is open in Wq,D2 (I) and that the mappings A : Sq(κ) → L(W2,D2 (),L2()) and [v 7→ fv] : Sq(κ) → L2() are analytic. The analyticity of the inversion map ℓ 7→ ℓ1 for bounded operators implies that also the mapping[v7→φv]:Sq(κ)→W22()is analytic, and the assertion follows as above from the results on pointwise multiplication . Let p ∈ (1,∞). We define Ap ∈ L(W2p,D(I),Lp(I))by Apv := −2xv for vW2p,D(I). Since ArAp forrp, we suppress the subscript in the following and write A := Ap. Note then that (2.5) subject to the boundary condition (1.2) and the initial condition (1.3) may be recast as an abstract parameter-dependent Cauchy problem

˙

u+Au=−λgε(u), t>0 , u(0) =u0, (2.32) where we recall that the function gε was defined in Proposition 2.1. To prove Theorem 1.1 it then suffices to focus on (2.32). For that purpose, let {etA;t0}denote the heat semigroup onLp(I)corresponding to −A. In order to state suitable regularizing properties we recall that we have setWp,D (I) =Wp(I) for 2β∈ (0, 1/p)and Wp,D (I) = {uWp (I);u1) =0}for 2β∈(1/p, 2]. Then we have:

Lemma 2.3. Let1< pr<∞. There existsω>0such that the following hold.

(i) If0≤αβ1with2α, 2β6=1/p, then ketAkL(W

p,D(I),Wp,D(I))Meωttαβ, t>0 , for some number M1depending on p,α, andβ.

(12)

(ii) If0≤αβ1with6=1/p and2β6=1/r, then ketAkL(W

p,D(I),Wr,D(I))Meωttαβ12(1p1r), t>0 , for some number M1depending on p, r,α, andβ.

Proof. (i) Since−A∈ L(W2p,D(I),Lp(I))is the generator of the analytic semigroup{etA;t0} onLp(I)with a negative spectral bound, it follows from [3, Chapt. V, Thm.2.1.3] that there are ω>0 andM≥1 such that

ketAkL(Eα,Eβ)Meωttαβ, t>0 ,

where, for θ ∈ [0, 1], Eθ := (Lp(I),Wp,D2 (I))θ with (·,·)θ chosen as real interpolation functor (·,·)θ,pif 2θ6=1 and as complex interpolation functor[·,·]1/2if 2θ=1. SinceEθ=Wp,D(I)with equivalent norms for 2θ∈[0, 2]\ {1/p}by [11, 29], assertion (i) follows.

(ii) From Sobolev’s embedding we haveWp,D2 (I)֒→Wr,D (I)for 2θ=2−(1/p−1/r)6=1/r, whence

ketAkL(W

p,D(I),Wr,D(I))cke2tAkL(W

r,D(I),Wr,D(I))ke2tAkL(Wp,D(I),Wp,D2 (I)), t>0 ,

and so assertion (ii) follows from (i).

We are now in a position to prove Theorem 1.1 and Theorem 1.2(i).

Proof of Theorem 1.1 and Theorem 1.2(i). Let λ > 0, q ∈ (2,∞), ε > 0, and consider u0Wq,D2 (I)withu0(x)>−1 forxI. Clearly, there isκ∈(0, 1/2)such that

u0Sq(2κ). (2.33)

We now fix 121q << 1

2 with 2σ6=1/qand putκ0:=κ/M, whereM≥1 is such that ketAkL(Wq,D2 (I))+tσ+1+12(121q)ketAkL(W2,D(I),Wq,D2 (I))Meωt, t0 (2.34) withω>0 according to Lemma 2.3. By Proposition 2.1 there isc10(κ,ε)>0 such that

kgε(v1)−gε(v2)kW2,D(I)c10(κ,ε)kv1v2kW2q,D(I), v1,v2Sq(κ0). (2.35) Since 0∈Sq(κ0)andgε(0) =1, we deduce from (2.35) that

kgε(v)kW

2,D(I)1+c10(κ,ε)

κ0 =c11(κ,ε), vSq(κ0). (2.36) Now, forτ>0, defineVτ :=C([0,τ],Sq(κ0))and

F(v)(t):=etAu0λ Z t

0 e(ts)Agε v(s)ds

for 0≤tτandv∈ Vτ. Considerv1,v2∈ Vτ andt∈[0,τ]. Then, introducing I(τ):=

Z τ

0 eωssσ112(121q)ds≤ I():= Z

0 eωs sσ112(121q)ds, (2.37)

Références

Documents relatifs

PONCE, G., Global existence of small solutions to a class of nonlinear evolution equations, NonL Anal.. B., Existence and non-existence of global solutions for a semilinear

Global regularity for solutions of the minimal surface equation with continuous boundary values.. Annales

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 large positive deflection is then noticed on one side (up state) whereas on the other side a smaller negative deflection of the membrane is obtained (down state). Finally,

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

Finally, convergence of solutions of the free boundary problem to the solutions of the well-established small aspect ratio model is shown..

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