• Aucun résultat trouvé

Nonlinear diffusion with a bounded stationary level surface

N/A
N/A
Protected

Academic year: 2022

Partager "Nonlinear diffusion with a bounded stationary level surface"

Copied!
16
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

Nonlinear diffusion with a bounded stationary level surface

Rolando Magnanini

a

, Shigeru Sakaguchi

b,

aDipartimento di Matematica U. Dini, Università di Firenze, viale Morgagni 67/A, 50134 Firenze, Italy

bDepartment of Applied Mathematics, Graduate School of Engineering, Hiroshima University, Higashi-Hiroshima, 739-8527, Japan Received 2 February 2009; received in revised form 28 December 2009; accepted 28 December 2009

Available online 7 January 2010

Abstract

We consider nonlinear diffusion of some substance in a container (not necessarily bounded) with bounded boundary of classC2. Suppose that, initially, the container is empty and, at all times, the substance at its boundary is kept at density 1. We show that, if the container contains a properC2-subdomain on whose boundary the substance has constant density at each given time, then the boundary of the container must be a sphere. We also consider nonlinear diffusion in the wholeRN of some substance whose density is initially a characteristic function of the complement of a domain with boundedC2boundary, and obtain similar results.

These results are also extended to the heat flow in the sphereSNand the hyperbolic spaceHN.

©2010 Elsevier Masson SAS. All rights reserved.

Résumé

Nous considérons la diffusion non linéaire d’une substance dans un récipient (pas nécessairement borné) avec frontière bornée de classeC2. Supposons qu’initialement, le récipient soit vide et, à sa frontière, la densité de la substance soit gardée à tout moment égale à 1. Nous montrons que, si le récipient contient un sous-domaineC2propre à la frontière duquel la substance est gardée à tout moment à densité constante, alors la frontière du récipient doit être une sphère. Nous considérons aussi la diffusion non linéaire dans toutRNd’une substance dont la densité est initialement une fonction caractéristique du complémentaire d’un domaine ayant la frontière bornée etC2, et nous obtenons des résultats semblables. Ces résultats sont aussi généralisés au cas du flux de chaleur dans la sphèreSNet l’espace hyperboliqueHN.

©2010 Elsevier Masson SAS. All rights reserved.

MSC:primary 35K60; secondary 35B40, 35B25

Keywords:Nonlinear diffusion equation; Overdetermined problems; Stationary level surfaces

This research was partially supported by Grants-in-Aid for Scientific Research (B) (15340047 and20340031) and a Grant-in-Aid for Exploratory Research (18654027) of Japan Society for the Promotion of Science, and by a Grant of the Italian MURST.

* Corresponding author.

E-mail addresses:magnanin@math.unifi.it (R. Magnanini), sakaguch@amath.hiroshima-u.ac.jp (S. Sakaguchi).

0294-1449/$ – see front matter ©2010 Elsevier Masson SAS. All rights reserved.

doi:10.1016/j.anihpc.2009.12.001

(2)

1. Introduction 1.1. Background

In the paper [12], we considered the solutionu=u(x, t )of the following initial–boundary value problem for the heat equation:

ut=u inΩ×(0,+∞), (1.1)

u=1 on∂Ω×(0,+∞), (1.2)

u=0 onΩ× {0}, (1.3)

whereΩis a bounded domain inRN withN2, and we obtained the following symmetry result.

Theorem A. (See [12].) LetΩ be a bounded domain inRN, N2,satisfying the exterior sphere condition and suppose thatDis a domain, with boundary∂D,satisfying the interior cone condition, and such thatDΩ.

Assume that the solutionuof problem(1.1)–(1.3)is such that

u(x, t )=a(t ), (x, t )∂D×(0,+∞), (1.4)

for some functiona:(0,+∞)(0,+∞).ThenΩmust be a ball.

We recall some terminology from [12]. A surface satisfying (1.4) is said to be a stationary isothermic surface;

Ω satisfies theexterior sphere conditionif for everyy∂Ω there exists a ballBr(z)such thatBr(z)Ω= {y}, whereBr(z)denotes an open ball centered atz∈RNand with radiusr >0;Dsatisfies theinterior cone conditionif for everyx∂Dthere exists a finite right spherical coneKxwith vertexxsuch thatKxDandKx∂D= {x}.

In order to better understand the background of the present paper, we outline the proof of Theorem A improved by a result in [13]. The proof is essentially based on three ingredients.

The first one is a result of Varadhan [21] which states that, as t→0+,the function −4tlogu(x, t ) converges uniformly onΩto the functiond(x)2, where

d(x)=dist(x, ∂Ω), xΩ.

To apply this result one needs the boundary∂Ω to be also the boundary of the exteriorRN\Ω. The assumption that Ω satisfies the exterior sphere condition is sufficient for that to happen. Hence, by (1.4) there existsR >0 satisfying

d(x)=R for everyx∂D. (1.5)

The second ingredient consists of abalance lawproved in [10] and [11] (see [12] for another proof). It states that, in any domainGinRN,a solutionv=v(x, t )of the heat equation is zero at some pointx0Gfor everyt >0 if and only if

∂Br(x0)

v(x, t ) dSx=0, for everyr

0,dist(x0, ∂G)

andt >0. (1.6)

We use (1.6) in two different ways. In the former one, we chooseG=Ω andv=uxi, i=1, . . . , N,and obtain, by some manipulations, that the gradient∇uis zero at some pointx0Ω for everyt >0 if and only if

∂Br(x0)

(xx0)u(x, t ) dSx=0, for everyr

0, d(x0)

andt >0. (1.7)

This condition helps us show that both∂Dand∂Ωmust be analytic. Indeed, with the aid of the interior cone condition forD, by combining (1.7) and (1.5) with the short-time behavior ofudescribed in Varadhan [21], we can see that for every pointx0∂Dthere exists a timet0>0 satisfying∇u(x0, t0) =0; this implies that∂Dis analytic. Thus, by using the exterior sphere condition forΩagain, we can conclude that∂Ωis analytic and parallel to∂D.

In the latter way of using (1.6), we choose two distinct pointsP , Q∂Ωand letp, q∂Dbe the points such that BR(p)∂Ω= {P} and BR(q)∂Ω= {Q}.

(3)

Thence, we consider the functionv=v(x, t )defined by

v(x, t )=u(x+p, t )u(x+q, t ) for(x, t )BR(0)×(0,+∞).

Sincevsatisfies the heat equation andv(0, t )=a(t )a(t )=0 for everyt >0, it follows from (1.6) that tN4+1

BR(p)

u(x, t ) dx=tN4+1

BR(q)

u(x, t ) dx for everyt >0.

Therefore, by taking advantage of the boundary layer forufor short times, we lett→0+and by using a result in [13], we obtain that

C(N ) N1

j=1

1

Rκj(P )

12

=C(N ) N1

j=1

1

Rκj(Q)

12

, (1.8)

whereκj(x), j =1, . . . , N−1, denotes the j-th principal curvature of the surface∂Ω at the pointx∂Ω with respect to the inward unit normal vector to∂Ω, and whereC(N )is a positive constant depending only onN(see [13, Theorem 4.2]).

With (1.8) in hand, we are ready to use our third ingredient: Aleksandrov’s sphere theorem [1, p. 412]. (A special case of this theorem is the well-knownSoap-Bubble Theorem(see also [17]).) Since (1.8) implies thatN1

j=1[R1κj(x)]is constant forx∂Ω, by applying Aleksandrov’s sphere theorem, we conclude that∂Ω must be a sphere (see [12] and [13] for details).

1.2. Main results

In the present paper, we extend and improve the results described in Section 1.1 to the case of certainnonlinear diffusion equations. It is evident that the introduction of a nonlinearity immediately rules out the use of our second ingredient, e.g. the balance law.

Since this was crucial to prove the necessary regularity of∂Ω,we will have to change our assumptions on the domain Ω.Thus, we shall assume Ω to be a domain (not necessarily bounded) in RN, N 2, having bounded boundary of classC2,that is, ∂Ω consists ofm (m1)connected components S1, . . . , Sm∂Ω which are the boundaries of boundedC2-domainsG1, . . . , GminRN, respectively. Thus

∂Ω= m

j=1

Sj and Sj=∂Gj for eachj∈ {1, . . . , m}. (1.9)

It should also be noticed that the lack of a balance law precludes the proof of property (1.8) (unless we find an alternative proof) and hence Aleksandrov’s sphere theorem cannot be put in action. We shall overcome this difficulty by a new and more direct proof of symmetry only based on our first ingredient (conveniently modified in Theorem 1.1) and Serrin’smethod of moving planes(see [18,16,19]). It is worth mentioning that our proof does not need Serrin’s corner lemmabut simply uses the strong maximum principle and Hopf boundary lemma (see Theorems 1.2 and 1.3).

We now set up our framework. We consider the unique bounded solutionu=u(x, t )of the nonlinear diffusion equation

ut=φ(u) inΩ×(0,+∞), (1.10)

subject to conditions (1.2) and (1.3). Hereφ:R→Ris such that

φC2(R), φ(0)=0, and (1.11)

0< δ1φ(s)δ2 fors∈R, (1.12)

whereδ1,δ2are positive constants. By the maximum principle we know that 0< u <1 inΩ×(0,+∞).

Moreover, by applying the comparison principle tou(x, t+h)andu(x, t )forh >0, we get

ut0 and hence φ(u)0 inΩ×(0,+∞). (1.13)

(4)

LetΦ=Φ(s)be the function defined by Φ(s)=

s

1

φ(ξ )

ξ fors >0. (1.14)

Note that ifφ(s)s, thenΦ(s)=logs.

We extend Varadhan’s result to our setting by the following Theorem 1.1.Letube the solution of problem(1.10), (1.2)–(1.3).

Then, lim

t0+−4t Φ u(x, t )

=d(x)2 uniformly on every compact set inΩ.

The proof of this theorem is constructed by adapting well-known results of the theory of viscosity solutions [2,7,3, 4,9]. The techniques developed to prove Theorem 1.1 can be used to extend this result to the important case in which the homogeneous boundary condition (1.2) is replaced by the non-homogeneous one

u=f on∂Ω×(0,+∞), (1.15)

wheref=f (x)is a continuous function on∂Ω,bounded from above and away from zero by positive constants (see Theorem 3.7).

The following symmetry result corresponds to Theorem A and Theorem 3.1 in [14].

Theorem 1.2.LetDbe aC2domain inRN satisfyingDΩ. Assume that the solutionuof problem(1.10), (1.2)–

(1.3), satisfies(1.4).

Thenm=1and∂Ω must be a sphere.

WhenΩis limited to unbounded domains, we have

Theorem 1.3.LetDbe aC2unbounded domain inRN satisfyingDΩ.

Assume that, for any connected componentΓ of∂D, the solution uof problem(1.10), (1.2)–(1.3), satisfies the following condition:

u(x, t )=aΓ(t ), (x, t )Γ ×(0,+∞), (1.16)

for some functionaΓ :(0,+∞)(0,+∞).

Thenm=1and∂Ω must be a sphere.

Whenφ(s)=s andΩ is bounded, Theorem A is clearly stronger than Theorem 1.2, since in the former we can use the balance law to infer better regularity. Furthermore, the same techniques used for the proof of Theorem A also yield a more general version of it (see Theorem 2.1).

The paper is then organized as follows. In Section 2, we prove all our symmetry results: Theorems 1.2, 1.3 and 2.1.

In Section 3, with the help of the theory of viscosity solutions, we prove Theorem 1.1 and its extension, Theorem 3.7.

Section 4 is devoted to show similar results for the unique bounded solution of the Cauchy problem for nonlinear diffusion equations. In Section 5, we mention that this kind of results also hold for the heat flow in the sphereSNand the hyperbolic spaceHNwithN2.

2. Symmetry results

In this section, with the aid of Theorem 1.1, by applying the method of moving planes to problem (1.10), (1.2)–(1.3) directly, we prove Theorems 1.2 and 1.3.

(5)

Proof of Theorem 1.2. First of all, we consider the case whereΩis unbounded. In this caseΩis an exterior domain, that is, we have

GiGj= ∅ ifi =j, i, j=1, . . . , m, and Ω=RN\ m

j=1

Gj

(see (1.9) for the definitions ofSjandGj). Set G=

m

j=1

Gj. (2.1)

ThenGis a bounded open set inRNhavingmconnected componentsG1, . . . , Gm. Theorem 1.1 and the assumption (1.4) yield (1.5). Furthermore, with the aid of ourC2-smoothness assumption on∂D and∂Ω, we see that both∂Ω and∂D consist ofmconnected closed hypersurfaces and each component of∂Ω is parallel, at distanceR,to only one component of∂D.

We apply the method of moving planes to the open setG. The proof runs similarly to those of Serrin’s [18] — or Reichel’s [16] and Sirakov’s [19] for exterior domains — but with the major difference that, here, since the relevant overdetermination takes placeinsideΩ,Serrin’s corner lemma — an extension of Hopf boundary lemma to domains with corners — is not needed.

Letbe a unit vector inRN, λ∈R,and letπλbe the hyperplanex·=λ.For largeλ, πλwill be disjoint fromG; asλdecreases,πλwill intersectGand cut off fromGan open capGλ(on the same side ofλ→ +∞).

Denote byGλthe reflection ofGλ in the planeπλ.At the beginning,Gλwill be and remain inGuntil one of the following occurs:

(i) Gλbecomes internally tangent to∂Gat some pointP not onπλ; (ii) πλreaches a position in which it is orthogonal to∂Gat some pointQ.

Letλ denote the (minimal) value of λat which the planeπλ reaches one of these positions and suppose that Gis not symmetric with respect toπλ. LetΩ be the connected component ofΩ∩ {x∈RN: x· < λ}whose boundary contains the pointsP orQin the respective cases (i) or (ii). Since, as already observed,∂Ωand∂Dconsist of connected closed pairwise parallel hypersurfaces, we can find pointsP andQ in∂D such that |PP| or

|QQ|equalR, respectively, and we have thatPΩandQ∂Ωπλ.

Letxλ=x+2[λ(x·)]denote the reflection of a pointx∈RN in the planeπλ.For(x, t )Ω×(0,), consider the functionw=w(x, t )defined by

w(x, t )=u xλ, t

. Then it follows from (1.4) that

w(P, t )=u(P, t ) or ∂u

∂(Q, t )=0 for allt >0, (2.2)

where in the second equality we have used the fact that the vectoris tangential also to∂DatQ∂D.

Observe thatwandusatisfy

wt=φ(w) and ut=φ(u) inΩ×(0,+∞),

w=u on(∂Ωπλ)×(0,+∞),

w <1=u on(∂Ω\πλ)×(0,+∞), w=u=0 onΩ× {0}.

Hence, by the strong comparison principle,

w < u inΩ×(0,+∞). (2.3)

Indeed, (2.3) can be obtained by applying the strong comparison principle to the bounded solutionsW =φ(w)and U=φ(u)ofWt=ψ(W )1 W andUt=ψ1(U )U, respectively; here,ψis the inverse function ofφ.

(6)

If case (i) applies, (2.3) contradicts the first equality in (2.2), sincePΩ.If case (ii) applies, by using Hopf’s boundary point lemma, we can infer that

∂u

∂(Q, t ) <0 for allt >0,

which contradicts the second equality in (2.2).

In conclusion,Gis symmetric for any direction∈RN, and in view of the definition (2.1) ofG,m=1 andG must be a ball. Namely,Ω is the exterior of a ball and∂Ωmust be a sphere.

WhenΩis bounded, it suffices to apply the method of moving planes directly toΩ. 2

Proof of Theorem 1.3. With the aid of theC2smoothness assumption of both∂D and∂Ω, Theorem 1.1 and the assumption (1.16), together with the fact thatDis unbounded, yield that∂Ω and∂Dconsist ofmpairs of connected closed hypersurfaces being parallel to each other respectively. (WhenDis bounded,∂Dmay consist of two connected components being parallel to one component of∂Ω.) Hence, the proof runs similarly to that of Theorem 1.2, with the only difference that the components in each pair constituting ∂Ω∂D may be at different distance from one another. 2

We conclude this section with a more general version of Theorem A.

Theorem 2.1.LetΩ be a domain(not necessarily bounded)inRN,N 2,satisfying the exterior sphere condition and suppose that ∂Ω is bounded. Let D be a domain withDΩ, and let Γ be a connected component of∂D satisfying

dist(Γ, ∂Ω)=dist(∂D, ∂Ω). (2.4)

Suppose thatDsatisfies the interior cone condition onΓ. Assume that the solutionuof problem(1.1)–(1.3)satisfies (1.16).

Then∂Ω must be either a sphere or the union of two concentric spheres.

Proof. Because of the assumption (2.4), the proofs of Lemma 2.2 of [14] and Lemma 3.1 in [12] also work in this situation. Then, there exists a connected componentSof∂Ω such that bothΓ andSare analytic and these are parallel with distanceR=dist(Γ, ∂Ω);also,N1

j=1[R1κj(x)]is constant forxS. SinceSis bounded, by applying Aleksandrov’s sphere theorem [1] to this equation, we see thatSandΓ are concentric spheres.

LetE be the annulus with∂E=SΓ. With the help of the analyticity ofu, by proceeding as in the proof of Theorem 3.1 in [14], we see that for anyi =j

(xjaj)∂u(x, t )

∂xi +(xiai)∂u(x, t )

∂xj =0 inΩ×(0,+∞),

where the pointa=(a1, . . . , aN)∈RNis the center of the sphereS. Henceumust be radially symmetric with respect toa. 2

3. Short-time behavior of solutions of nonlinear diffusion equations

In this section, with the help of the theory of viscosity solutions, we prove our keystone result, Theorem 1.1. We begin with some preliminaries.

Lemma 3.1.Letw=φ(u),whereuis the solution of (1.10), (1.2)–(1.3). Forj=1,2,letwj solve the problem:

(wj)t=δjwj inΩ×(0,+∞), (3.1)

wj=φ(1) on∂Ω×(0,+∞), (3.2)

wj=0 onΩ× {0}. (3.3)

Then

w1ww2 inΩ×(0,+∞).

(7)

Proof. Sincewt=φ(u)w,from (1.12) and (1.13) we have:

δ1wwtδ2w inΩ×(0,+∞). (3.4)

Hence, by the comparison principle we get our claim. 2 Now, letΨ =Ψ (s)be the inverse function ofΦ. Then

s=Φ Ψ (s)

=

Ψ (s)

1

φ(ξ ) ξ and

Ψ (s)=φ Ψ (s)

Ψ(s), (3.5)

by differentiating ins.

As in Freidlin and Wentzell [5], for 0< ε <1, define the functionuε=uε(x, t )by uε(x, t )=u(x, εt ) for(x, t )Ω×(0,+∞).

Thenuεsatisfies uεt =εφ

uε

inΩ×(0,+∞), uε=1 on∂Ω×(0,+∞), uε=0 onΩ× {0}.

Moreover, the functionvε=vε(x, t )defined by vε(x, t )= −εΦ

uε(x, t )

for(x, t )Ω×(0,+∞) is such thatuε=Ψ (ε1vε)and, by (3.5), we have that

vtε=εφvε−∇vε2 inΩ×(0,+∞), (3.6)

vε=0 on∂Ω×(0,+∞), (3.7)

vε= +∞ onΩ× {0}, (3.8)

whereφ=φ(Ψ (ε1vε)).

Lemma 3.2.It holds that for(x, t )Ω×(0,+∞) δ1

δ2· 1

4td(x)2lim inf

ε0+ vε(x, t )lim sup

ε0+

vε(x, t )δ2 δ1· 1

4td(x)2,

where these limits asε→0+are uniform in every compact set contained inΩ×(0,+∞).

Proof. We observe that the following hold:

δ1sφ(s)δ2s fors0, (3.9)

δ1logsΦ(s)δ2logs for 0< s1, (3.10)

es/δ1Ψ (s)es/δ2 for − ∞< s0. (3.11)

Letwεj=wεj(x, t ) (j=1,2)be the functions defined by wεj(x, t )=wj(x, εt ),

where thewj’s are defined in Lemma 3.1. With the aid of (3.9) and (3.10), it follows from Lemma 3.1 that

εδ1log wε2

δ1

vεεδ2log wε1

δ2

inΩ×(0,+∞).

(8)

By the result of Varadhan [21], we see that, asε→0+,the functions−εδjlogwεjconverge to the function 4t1d(x)2 uniformly on every compact set contained inΩ×(0,+∞), since each scaled functionφ (1)1 wj(x, δj1t )solves problem (1.1)–(1.3). Our claim then follows at once. 2

The next lemma easily follows from Lemma 3.2.

Lemma 3.3.For any compact setKinΩ×(0,+∞), there exist three positive constantsε0, c1,andc2(0< c1c2) depending onKsuch that

0< c1vεc2 inK, for0< εε0.

The key point in the proof of Theorem 1.1 is to obtain the following gradient estimate which we shall prove at the end of this section.

Lemma 3.4.For any compact setKinΩ×(0,+∞), there exist two positive constantsε11ε0)andc3depending onK,such that

vεc3 inK, for0< εε1.

Then, by combining Lemmas 3.3 and 3.4 with Gilding’s result [6], we obtain the following uniform Hölder esti- mate.

Lemma 3.5.For any compact setKinΩ×(0,+∞), there exist two positive constantsε22ε1)andc4depending onK,such that

vε(x, t )vε(x, s)c4|ts|12 for any pair(x, t ), (x, s)K and for0< εε2.

Theorem 3.6.The following limit

εlim0+vε(x, t )= 1 4td(x)2

holds uniformly on every compact set inΩ×(0,+∞).

Proof. Lemmas 3.3, 3.4, and 3.5 together with Ascoli–Arzelà’s theorem and the Cantor diagonal process yield a positive vanishing sequence of numbers εn and a continuous function v=v(x, t ) in Ω×(0,+∞)such that, as n→ ∞, thevεn’s converge tovuniformly on every compact set contained inΩ×(0,+∞).Hence, by Lemma 3.2,

δ1 δ2· 1

4td(x)2v(x, t )δ2 δ1· 1

4td(x)2 for(x, t )Ω×(0,+∞). (3.12)

Define a functionV =V (x, t )onRN×(0,+∞)by V (x, t )=v(x, t ) ifxΩ,

0 ifx /Ω.

Since bothd2and its gradient vanish on∂Ω, (3.12) yields thatV is continuous onRN×(0,+∞), differentiable at any point on∂Ω×(0,+∞), and that bothV and∇V vanish on∂Ω×(0,+∞). Also, (3.12) yields that limt0+v(x, t )= +∞ifxΩ.

Therefore, by using the fact thatvε solves problem (3.6)–(3.8), with the help of Crandall, Ishii, and Lions [2], we see thatV is a viscosity solution of the following Cauchy problem:

(9)

Vt= −|∇V|2 inRN×(0,+∞),

V =0 on

RN\Ω

× {0},

V = +∞ onΩ× {0}. (3.13)

Moreover, since a uniqueness result of Strömberg [20] tells us that the Hopf–Lax formula provides the unique viscosity solution of the Cauchy problem (3.13), we must have that, for any(x, t )∈RN×(0,+∞),

V (x, t )=inf

ϕ(ξ )+|xξ|2

4t : ξ∈RN

=(dist(x,RN\Ω))2

4t ,

whereϕ=ϕ(ξ )is the lower semicontinuous initial data defined by ϕ(ξ )=

+∞ ifξΩ, 0 ifξ /Ω.

By the uniqueness ofV, the whole sequence{vε}converges asε→0+, and we get our claim. 2

Proof of Theorem 1.1. The desired result follows by simply settingt=1 and thenε=t in Theorem 3.6. 2 By a simple argument, we can extend Theorem 1.1 to the important case of non-homogeneous boundary values.

Theorem 3.7.Letf=f (x)be a continuous function on∂Ω such that

0< b1f (x)b2 for allx∂Ω, (3.14)

for some positive constantsb1andb2.Letube the solution of problem(1.10),(1.15), (1.3).

Then, lim

t0+−4t Φ u(x, t )

=d(x)2 uniformly on every compact set inΩ.

Proof. Consider the unique bounded solutionsuj=uj(x, t ) (j=1,2)of the following initial–boundary value prob- lems:

ujt =φ uj

inΩ×(0,+∞), uj=bj on∂Ω×(0,+∞), uj=0 onΩ× {0}.

Then it follows from (3.14) and the comparison principle that

u1uu2 inΩ×(0,+∞). (3.15)

With the help of Theorem 1.1, we see that, as t→0+,the function −4t Φ(uj(x, t ))converges to the function d(x)2uniformly on every compact set inΩ for eachj=1,2. Indeed, for eachj=1,2,we set

U=uj

bj, φ(s)˜ = 1

bjφ(bjs) fors∈R, and Φ(s)˜ = s

1

φ˜(ξ )

ξ fors >0.

Then it follows that

φ˜(s)=φ(bjs) fors∈R, Φ(s)˜ =Φ(bjs)Φ(bj) fors >0, (3.16) and

Ut=φ(U )˜ inΩ×(0,+∞), (3.17)

U=1 on∂Ω×(0,+∞), (3.18)

U=0 onΩ× {0}. (3.19)

(10)

Thus, applying Theorem 1.1 toUyields that, ast→0+,the function−4tΦ(U (x, t ))˜ converges to the functiond(x)2 uniformly on every compact set inΩ. Hence, with the aid of the second equality of (3.16), this means that, ast→0+, the function−4t Φ(uj(x, t ))converges to the functiond(x)2uniformly on every compact set inΩ.

On the other hand, sinceΦis increasing ins >0, we have from (3.15) that

−4t Φ u1

−4t Φ(u)−4t Φ u2

inΩ×(0,+∞),

which implies that, as t→0+, the function −4t Φ(u(x, t )) converges to the function d(x)2 uniformly on every compact set inΩ. 2

Proof of Lemma 3.4. We use Bernstein’s technique (see [3,7,4,9]). Let r,τ andT be positive numbers such that τ <2τ < T andKBr(0)× [2τ, T].TakeζC(B2r(0)×(τ, T])satisfying

0ζ1 and ζt0 inB2r(0)×(τ, T],

ζ =1 onBr(0)× [2τ, T], and suppζB2r(0)×(τ, T].

In the sequel of this proof, we will use the constantsε0, c1andc2of Lemma 3.3 relative to the compact setB2r(0)× [τ, T].

Consider the functionz=z(x, t )defined by

z=ζ2vε2λvε, (3.20)

whereλ >0 is a constant to be determined later, and 0< εε0.Suppose that(x0, t0)is a point inB2r(0)×(τ, T] satisfying

ζ (x0, t0) >0 and max

B2r(0)×[τ,T]z=z(x0, t0).

At(x0, t0)we then have

zt0, zxi=0, and z0, (3.21)

and hence

0ztεφ Ψ

ε1vε z.

The following inequality holds at(x0, t0)for some positive constantsA1andA2independent of(x0, t0)andε:

λvε2A1vε2+A2ζvε3−2ζ2vε2φΨvεεφζ22vε2. (3.22) It is a consequence of (3.21) and some lengthy calculations that, for the reader’s convenience, will be carried out in Appendix A.

Now, we want to bound the third and fourth summand on the right-hand side of (3.22). The bound for the latter summand,

εφζ22vε2εδ1ζ22vε2,

easily follows from (1.12). In order to bound the former one, we use the fact that φC2(R)and Lemma 3.3, the algebraic inequality 2aba2+b2,and the key inequality

0< Ψ

ε1vε

=Ψ (ε1vε)

φ 1

δ1

e

εδ2 1

δ1

e

c1

εδ2, (3.23)

which follows from (3.5), (1.12) and (3.11). With these three ingredients, we show that

−2ζ2vε2φΨvε 1 δ1e

c1 εδ2ζ2

A3vε4+∇2vε2 , at(x0, t0),for some positive constantA3independent of(x0, t0)andε.

Set

M= max

B2r(0)×[τ,T]ζvε, λ= M2+1 2(c2+1),

(11)

and chooseεin(0, ε0]so small to obtain that A3

δ1e

c1

εδ2 1

4(c2+1) and 1 δ1e

c1 εδ2 εδ1

for allε(0, ε].Then, with these choices of constants, from (3.22) and the aforementioned bounds on the second- order derivatives ofvε,we have that

M2+1

4(c2+1)∇vε2A1vε2+A2Mvε2 (3.24)

at(x0, t0),for anyε(0, ε].

Thus, if∇vε(x0, t0) =0,from (3.24) we get M2+1

4(c2+1)A1+A2M,

which yields the desired gradient estimate at once. If∇vε(x0, t0)=0, instead, we use the definition (3.20) ofzto infer that

M2maxz+λmaxvελmaxvε M2+1

2(c2+1)c2M2 2 +1

2,

sincez(x0, t0)= −λvε(x0, t0) <0.Therefore,M1 and this completes the proof. 2

Remark.Lions, Souganidis, and Vázquez [9] consider the pressure equation for the porous medium equation:

(vm)t=(m−1)vmvm+ |∇vm|2 form >1,

and consider the asymptotic behavior asm→1+. They get the interior gradient estimate forvmindependent ofmby a technique similar to ours. We follow the outline of their proof but we use inequality (3.23) in order to overcome the difficulty caused byφ=φ(Ψ (ε1vε))in Eq. (3.6).

4. On the Cauchy problem

LetΩ be a domain given in (1.9) and consider the unique bounded solutionu=u(x, t )of the following Cauchy problem:

ut=φ(u) inRN×(0,+∞), and u=χRN\Ω onRN× {0}, (4.1)

whereχRN\Ω denotes the characteristic function of the setRN\Ωandφsatisfies the assumptions (1.11)–(1.12). The purpose of this section is to prove the following result.

Theorem 4.1.Theorems1.1,1.2, and1.3also hold for the unique bounded solutionuof the Cauchy problem(4.1).

Let us start with two lemmas.

Lemma 4.2.There exist a smallδ >0and aC2-functionf=f (ξ )onRsatisfying φ(f )f

+1

2+2δ)f=0 and f<0 inR, and

1> f (−∞) > f (0) >0> f (+∞) >−∞. (4.2)

Proof. It suffices to show that there exists aC2-functionh=h(ξ )onRsatisfying φ(h)h

+1

2ξ h=0 and h<0 inR, and (4.3)

1> h(−∞) > h(0) >0> h(+∞) >−∞. (4.4)

Indeed, settingf (ξ )=h(ξ+2δ)for sufficiently smallδ >0 gives the desired solutionf.

(12)

The assumptions (1.11)–(1.12) guarantee existence and uniqueness,on the wholeR,of the solution(h, H )of the Cauchy problem for the system of ordinary differential equations

h= H

φ(h), H= −1 2ξ H

φ(h), and

h(0), H (0)

=(h0, H0) (4.5)

(obtained by lettingH=φ(h)hin (4.3)); hereh0>0 andH0<0 are given numbers. Also, by uniqueness we infer thatH <0 onRand henceh<0 onR.

Thus, with the help of (1.12), by integrating the second equation in (4.5), we have that H0exp

ξ22

H (ξ )H0exp

ξ21

<0 and

H0 δ1 exp

ξ22

h(ξ )H0 δ2 exp

ξ21

<0 forξ∈R; hence

h0+H0 δ1

ξ

0

exp

η22

dηh(ξ )h0+H0 δ2

ξ

0

exp

η21

dη, forξ >0,and

h0+H0 δ2

ξ

0

exp

η21

dηh(ξ )h0+H0 δ1

ξ

0

exp

η22

dη,

forξ <0.By lettingξ → +∞andξ→ −∞, respectively, we get h0+H0

δ1

π δ2h(+∞)h0+H0 δ2

π δ1,

h0H0

δ2

π δ1h(−∞)h0H0

δ1

π δ2.

Therefore, (4.4) is obtained by setting h0= δ13/2

2(δ13/2+δ23/2) and H0= − δ1δ2

π (δ3/21 +δ23/2) in the last two formulas. 2

Lemma 4.3.There exists a constantc0>0satisfying c0u <1 on∂Ω×(0,1].

Proof. First of all, by the strong maximum principle

0< u <1 inRN×(0,+∞). (4.6)

Consider the signed distance functiond=d(x)ofx∈RNto the boundary∂Ω defined by d(x)=

dist(x, ∂Ω) ifxΩ,

−dist(x, ∂Ω) ifx /Ω. (4.7)

Since∂ΩisC2and compact, there exists a numberρ >0 such thatd(x)isC2-smooth on a compact neighborhood N of the boundary∂Ω given by

N =

x∈RN: −ρd(x)ρ .

(13)

We set now w(x, t )=f

t12d(x)

for(x, t )∈RN×(0,+∞). (4.8)

Then, it follows from a straightforward computation and the properties off that wtφ(w)= −1

tf

t12d

δ+√

t φ(f )d

inN ×(0,+∞).

Notice that if√

t <δ δ

2maxN|d|, thenwtφ(w) <0. Hence, since∂N is compact, in view of Lemma 4.2, (4.8), and (4.1), we observe that there exists a smallτ >0 satisfying

wtφ(w) <0=utφ(u) inN×(0, τ],

wu on∂N×(0, τ],

wu onN× {0}.

Here, we note that u and w are regarded as continuous mappings from [0, τ] toL1(N) and, by taking into ac- count (4.2), the initial condition is satisfied by their limits ast→0+inL1(N).

Therefore it follows from the comparison principle thatwuinN ×(0, τ]. In particular, we have uf (0) (>0) on∂Ω×(0, τ].

Combining this with (4.6) completes the proof. 2

Proof of Theorem 4.1. Consider the unique bounded solutionsu±=u±(x, t )of the following initial–boundary value problems:

u±t =φ u±

inΩ×(0,+∞), u+=1 and u=c0 on∂Ω×(0,+∞),

u±=0 onΩ× {0}.

Then it follows from Lemma 4.3 and the comparison principle that

uuu+ inΩ×(0,1]. (4.9)

By applying Theorem 3.7 tou±, we have that, ast→0+,both functions−4t Φ(u±(x, t ))converge to the function d(x)2uniformly on every compact set inΩ.

On the other hand, sinceΦis increasing ins >0, we have from (4.9) that

−4t Φ u

−4t Φ(u)−4t Φ u+

inΩ×(0,1],

which implies that, as t →0+, the function −4t Φ(u(x, t )) converges to the function d(x)2 uniformly on every compact set inΩ. This means that Theorem 1.1 also holds for the Cauchy problem (4.1).

Finally, proceeding as in Section 2, with the aid of the strong comparison principle for the Cauchy problem, we can easily show that Theorems 1.2 and 1.3 also hold for the Cauchy problem (4.1). 2

5. SphereSSSN and hyperbolic spaceHHHN

The purpose of this section is to show that similar results hold also for the heat flow in the sphereSN and the hyperbolic spaceHN withN2. In order to handleSNandHNtogether, let us putM=SNorM=HN.

LetΩbe a domain inMwith boundedC2-smooth boundary∂Ω, and denote byLthe Laplace–Beltrami operator onM. Letu=u(x, t )be the unique bounded solution either of the following initial–boundary value problem for the heat flow:

ut=Lu inΩ×(0,+∞), (5.1)

u=1 on∂Ω×(0,+∞), (5.2)

u=0 onΩ× {0}, (5.3)

(14)

or of the following Cauchy problem for the heat flow:

ut=Lu inM×(0,+∞), and u=χM\Ω onM× {0}, (5.4) whereχM\Ω denotes the characteristic function of the setM\Ω.

Denote byd(x)=inf{d(x, y): y∂Ω}the geodesic distance betweenx and∂Ω, whered(x, y)is the geodesic distance between two pointsxandyinM. Then, with the aid of a result of Norris [15, Theorem 1.1, p. 82] concerning the short-time asymptotics of the heat kernel of Riemannian manifolds, we later prove

Theorem 5.1.Letube the solution either of problem(5.1)–(5.3)or of problem(5.4)inSNorHN. Then the function

−4tlogu(x, t )converges to the functiond(x)2ast→0+uniformly on every compact set inΩ. Theorem 5.1 yields the following symmetry results.

Theorem 5.2.Letube the solution either of problem(5.1)–(5.3)or of problem(5.4)inSNorHN.In the case ofSN, assume thatΩ is contained in a hemisphere inSN.

Then Theorems1.2and1.3hold forHN and Theorem1.2holds forSN in the sense that∂Ω must be a geodesic sphere inHNorSN,respectively.

Proof. By Theorem 5.1, each stationary isothermic surface of classC2inΩis then parallel to a connected component of∂Ωin the sense of the geodesic distance. The claims of our theorem can thus be proved by replacing the method of moving planes, used in Section 2 forRN, by a straightforward adaptation of the method of moving closed and totally geodesic hypersurfaces forSNorHNdeveloped by Kumaresan and Prajapat in [8]. 2

Proof of Theorem 5.1. Consider the signed distance functiond=d(x)ofx ∈Mgiven by the same definition as (4.7). Since ∂Ω isC2and compact, there exists a number ρ0>0 such thatd(x)isC2-smooth on a compact neighborhoodN of∂Ω given by

N =

x∈M: −ρ0d(x)ρ0

. Set

N=

x∈M:−ρ0d(x)0 (N).

Letu=u(x, t )be the unique bounded solution of the following Cauchy problem:

ut =Lu inM×(0,+∞), and u=χN onM× {0}, (5.5) whereχNdenotes the characteristic function of the setN.Moreover, for each 0< ρ < ρ0, we set

Nρ=

x∈M: −ρd(x)ρ (N).

For each 0< ρ < ρ0, letuρ+=uρ+(x, t )be the unique bounded solution of the following Cauchy problem:

uρt+=Luρ+ inM×(0,+∞), and uρ+=2χNρ onM× {0}, (5.6) whereχNρ denotes the characteristic function of the setNρ.

Letρ(0,min{1, ρ0})be arbitrarily small. Then, there exists a numbertρ>0 satisfying

uρ+>1 in∂Ω×(0, tρ]. (5.7)

Thus it follows from (5.7) and the comparison principle that for any(x, t )Ω×(0, tρ]

N

p(t, x, y) dy=u(x, t )u(x, t )uρ+(x, t )=2

Nρ

p(t, x, y) dy, (5.8)

wherep=p(t, x, y)denotes the heat kernel or the fundamental solution of the heat equation on the wholeM.

Références

Documents relatifs

ASSOCIATION OF CANADA LANDS SURVEYORS - BOARD OF EXAMINERS WESTERN CANADIAN BOARD OF EXAMINERS FOR LAND SURVEYORS ATLANTIC PROVINCES BOARD OF EXAMINERS FOR LAND SURVEYORS ---..

Considérons les expériences suivantes a) On jette un dé et on note le point obtenu. b) Un chasseur vise un lapin et le tire. c) On extrait d’un paquet de 52 cartes à jouer, une

We define a partition of the set of integers k in the range [1, m−1] prime to m into two or three subsets, where one subset consists of those integers k which are &lt; m/2,

Then the critical temperature is changed as well, since it’s like

Given a pencil of algebraic plane curves such that a general element is irreducible, the purpose of this paper is to give a sharp upper bound for the number of reducible curves in

Déterminer lesquelles des affirmations ci-dessous sont vraies ou fausses.. On justifiera chaque réponse avec une preuve ou

3. A general lower bounding technique. We now present a framework for obtaining lower bounds on edit distances between permutations in a simple and unified way. , n ), introduced

In this work, we construct an example of this type; more precisely, by solving a quasilinear elliptic boundary-value prob- lem we define a nonlinear operator on the space of