Processing math: 100%
Research article Special Issues

Linear stability analysis of overdetermined problems with non-constant data

  • Received: 17 March 2022 Revised: 16 July 2022 Accepted: 19 July 2022 Published: 09 August 2022
  • We study an overdetermined problem that arises as the Euler-Lagrange equation of a weighted variational problem in elasticity. Based on a detailed linear analysis by spherical harmonics, we prove the existence and local uniqueness as well as an optimal stability estimate for the shape of a domain allowing the solvability of the overdetermined problem. Our linear analysis reveals that the solution structure is strongly related to the choice of parameters in the problem. In particular, the global uniqueness holds for the pair of the parameters lying in a triangular region.

    Citation: Michiaki Onodera. Linear stability analysis of overdetermined problems with non-constant data[J]. Mathematics in Engineering, 2023, 5(3): 1-18. doi: 10.3934/mine.2023048

    Related Papers:

    [1] Rolando Magnanini, Giorgio Poggesi . Interpolating estimates with applications to some quantitative symmetry results. Mathematics in Engineering, 2023, 5(1): 1-21. doi: 10.3934/mine.2023002
    [2] Luigi Provenzano, Alessandro Savo . Isoparametric foliations and the Pompeiu property. Mathematics in Engineering, 2023, 5(2): 1-27. doi: 10.3934/mine.2023031
    [3] Filippo Gazzola, Gianmarco Sperone . Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040
    [4] Bruno Bianchini, Giulio Colombo, Marco Magliaro, Luciano Mari, Patrizia Pucci, Marco Rigoli . Recent rigidity results for graphs with prescribed mean curvature. Mathematics in Engineering, 2021, 3(5): 1-48. doi: 10.3934/mine.2021039
    [5] Chao Xia, Jiabin Yin . Two overdetermined problems for anisotropic $ p $-Laplacian. Mathematics in Engineering, 2022, 4(2): 1-18. doi: 10.3934/mine.2022015
    [6] Antonio Greco, Francesco Pisanu . Improvements on overdetermined problems associated to the $ p $-Laplacian. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022017
    [7] Yu Chen, Jin Cheng, Giuseppe Floridia, Youichiro Wada, Masahiro Yamamoto . Conditional stability for an inverse source problem and an application to the estimation of air dose rate of radioactive substances by drone data. Mathematics in Engineering, 2020, 2(1): 26-33. doi: 10.3934/mine.2020002
    [8] Giovanni S. Alberti, Yves Capdeboscq, Yannick Privat . On the randomised stability constant for inverse problems. Mathematics in Engineering, 2020, 2(2): 264-286. doi: 10.3934/mine.2020013
    [9] Connor Mooney, Arghya Rakshit . Singular structures in solutions to the Monge-Ampère equation with point masses. Mathematics in Engineering, 2023, 5(5): 1-11. doi: 10.3934/mine.2023083
    [10] YanYan Li . Symmetry of hypersurfaces and the Hopf Lemma. Mathematics in Engineering, 2023, 5(5): 1-9. doi: 10.3934/mine.2023084
  • We study an overdetermined problem that arises as the Euler-Lagrange equation of a weighted variational problem in elasticity. Based on a detailed linear analysis by spherical harmonics, we prove the existence and local uniqueness as well as an optimal stability estimate for the shape of a domain allowing the solvability of the overdetermined problem. Our linear analysis reveals that the solution structure is strongly related to the choice of parameters in the problem. In particular, the global uniqueness holds for the pair of the parameters lying in a triangular region.



    The main purpose of the present paper is to derive a sharp quantitative stability estimate for the rigidity of the spherical configuration of Ω under non-radial perturbations of the boundary data g, where a bounded domain ΩRn with n2 admits a solution u to the overdetermined problem

    {Δu=fin Ω,u=0on Ω,uν=gon Ω. (1.1)

    Here, ν is the unit outer normal vector to Ω, and f,g are prescribed functions of the form

    f(x)=(n+α)|x|α,g(x)=g0(ξ)|x|β,ξ=x|x|, (1.2)

    where α,βR, g0 is a function defined on the unit sphere S and the coefficient n+α is only for the normalization. In order to clarify the sense in which (1.1) is to be understood, we shall additionally assume α>n, so that

    u(x)=1|x|α+2α+2 (1.3)

    is a solution to (1.1) in the sense of distributions when Ω is the unit ball B, g0=1 and βR. Equation (1.1) arises as the Euler-Lagrange equation of a variational problem for a weighted torsional rigidity (see Section 2). The particular case α=β=0 has been extensively studied in the literature and is sometimes referred to as Serrin's overdetermined problem.

    In the case where f,g are positive constants (i.e., α=β=0 and g0>0 a constant), it is well-known that Ω must be a ball if (1.1) has a solution uC2(¯Ω). This rigidity result was proved in a seminal paper [34] by Serrin, with an innovative argument called the method of moving planes based on Alexandrov's reflection principle and a refined boundary point lemma for corners, which in fact applies to nonlinear equations. Weinberger [37] provided an alternative proof based on the observation that, if u satisfies (1.1), the Cauchy-Schwarz deficit

    d(u):=|D2u|2(Δu)2n0

    becomes identically zero and thus u is a quadratic function as (1.3) (see [22,23,31] for refined arguments). Another interesting proof was introduced by Brandolini, Nitsch, Salani and Trombetti [7] using an integral quantity related to Newton's inequalities involving elementary symmetric functions of the eigenvalues of the Hessian matrix D2u.

    There have been numerous studies on the stability of the spherical configuration of Ω when f=n (i.e., α=0) and g is slightly perturbed from a constant. Here we recall a few relevant results (without mentioning technical assumptions) from a methodological point of view, but not in chronological order. In order to describe the results in a unified manner, let Ωρ be the bounded star-shaped domain enclosed by

    Ωρ:={(1+ρ(ξ))ξξS} (1.4)

    for ρC2+γ(S) with 1<ρ(ξ)< and 0<γ<1, and let uρ denote a unique solution to the Dirichlet problem consisting in the first two equations in (1.1) for Ω=Ωρ. Aftalion, Busca and Reichel [3] initiated the stability analysis of (1.1) by developing a quantitative version of the method of moving planes and proved that, up to translation,

    ρL(S)C|loguρν+1C1(Ωρ)|1/n (1.5)

    holds if the quantity νuρ+1C1(Ωρ) is sufficiently small. This inequality shows that the deviation ρ of domain Ωρ from the unit ball Ω0=B can be controlled by that of the Neumann data g from the constant c=1. This method was further developed by Ciraolo, Magnanini and Vespri [13] using a quantitative Harnack's inequality, and the logarithmic estimate (1.5) was sharpened to a power-type estimate. In fact, these two results apply to general nonlinear equations. For the particular case f=n, in a series of papers [22,23,24,25], Magnanini and Poggesi improved (1.5) by establishing an integral identity that relates d(u) to the deviation νu+1 and estimating both sides of the identity. The resulting estimate is

    ρL(S)Cuρν+1τnL2(Ωρ),

    where τ2=1, τ3=1ε for any ε>0, and τn=4/(n+1) for n4 (see [25] for a sharper estimate in the case n=3). In particular, this estimate is optimal for n=2, and almost optimal for n=3, as one can confirm by choosing Ωρ as ellipsoids that linear estimates (i.e., τn=1) are sharpest. Optimal linear estimates for any spatial dimensions n2 have been established either when the norm of the left hand side is weakened or when the norm of the right hand side is strengthened. Indeed, Feldman [16] proved

    ρL1(S)Cuρν+1L2(Ωρ)

    by refining an argument of Brandolini, Nitsch, Salani and Trombetti [8], in which a power-type stability estimate was obtained for the first time by exploiting their own proof of the symmetry in [7]. Another optimal estimate obtained by Gilsbach and the present author [18] states that

    ρC2+γ(S)Cuρν+1C2+γ(Ωρ). (1.6)

    This estimate is a consequence of the detailed linear analysis of (1.1) for α=β=0 based on a new implicit function theorem for triplets of Banach spaces, and it also provides the existence and local uniqueness of Ωρ for small perturbations of g from c=1.

    For non-constant f,g, the overdetermined problem (1.1) or other variants were treated in [1,2,4,5,9,19,20,28,29,30,35]. Bianchini, Henrot and Salani [6] studied the existence, uniqueness and geometric properties of Ω for (1.1) with α=0, β>0 and β1 by a variational method and the maximum principle. In particular, for α=0 and β>1, they proved the stability estimate

    ρL(S)Cuρν+|x|βL(Ωρ), (1.7)

    or equivalently

    ρL(S)Cg01L(S).

    The restriction β>1 hinges on the availability of the comparison principle for domains Ω with different values of g, and indeed (1.7) was proved by a comparison of Ωρ with radial domains. Note that a priori estimates for α=β=0 such as (1.6) cannot yield the corresponding estimates for general α,β by a direct use of the triangle inequality.

    Our purpose in this paper is to prove an optimal quantitative stability estimate of the radial configuration of Ω for non-radial perturbations of g in any spatial dimensions n2 by linearization approach as in [18]. Since this approach relies only on the non-degeneracy of the linearized problem, we can treat general α,β unless

    αβ+1N{0}.

    In fact, our linear analysis suggests that a symmetry-breaking bifurcation should occur at these exceptional values of α,β. We refer to [10,11,12,14,15,21,26,27,33,36] for the linearization approach to bifurcation phenomena in overdetermined problems. We also emphasize that our approach yields the existence and local uniqueness of Ω for given perturbations g0 even if β0 for which the variational method in general fails (see [6]).

    To state our main result, for kN and 0<γ<1, we set hk+γ(¯Ω), called the little Hölder space, to be a closed proper subspace of Ck+γ(¯Ω) defined as the closure of C(¯Ω) in Ck+γ(¯Ω), and similarly we define hk+γ(Γ) for a hypersurface Γ. The little Hölder space hk+γ(S) is suitable for our linearization approach, since the set of spherical harmonics spans a dense subspace of hk+γ(S).

    Theorem 1.1. Let α>n, βR and 0<γ<1 satisfy

    αβ+1N{0}.

    Then, there are δ,ε>0 such that, for any g0h2+γ(S) with g01h2+γ(S)<δ, there exists a unique ρh3+γ(S) with ρh3+γ(S)<ε such that (1.1) is solvable in Ω=Ωρ with f,g defined by (1.2). Moreover, ρ=ρ(g0) satisfies the following:

    (i) If g01 in h2+γ(S), then ρ0 in h3+γ(S).

    (ii) There is a constant C>0 such that

    ρh2+γ(S)Cg01h2+γ(S) (1.8)

    holds for any g0 with g01h2+γ(S)<δ.

    In the case where αβ+1<0, the uniqueness in fact holds among all bounded domains Ω having C1-boundary with 0Ω.

    Remark 1.2. Theorem 1.1 contains, as a special case, the radial symmetry of Ω for g0=1. In fact, the global rigidity/symmetry of Ω for αβ+1<0 has its counterpart for the endpoint case αβ+1=0 (see Proposition 2.2). Moreover, we prove a global uniqueness result in this special case αβ+1=0, where the solvability of (1.1) is invariant under rescaling of Ω (see Proposition 5.2). $

    Remark 1.3. The stability estimate (1.8) still holds for hk+γ(S) with arbitrary k2 in both sides. This can be easily verified, as all the succeeding arguments equally proceed with hk1+γ,hk+γ,hk+1+γ instead of h1+γ,h2+γ,h3+γ.

    The structure of the present paper is as follows. In Section 2, we shall discuss the radial symmetry of Ω in the case g0=1 with various different techniques. In Section 3, we introduce a functional analytic formulation of (1.1) and derive the linearized problem. A detailed linear analysis is carried out by spherical harmonics. In Section 4, we derive the stability estimate (1.8) as well as the existence and local uniqueness of Ω by an implicit function theorem in [18]. Lastly, in Section 5, we study the global uniqueness of Ω when αβ+10.

    This section concerns the radial symmetry of Ω when it admits a solution u to (1.1) for radial data f,g. Although some of the symmetry results presented in this section are well-known or easily deduced from existing methods, we briefly discuss them so as to compare the well-known arguments with ours.

    Let us begin with a variational structure of (1.1). Indeed, (1.1) is derived as the Euler-Lagrange equation of the minimization problem of the generalized torsion functional

    J(Ω)=infuH10(Ω){0}Ω|u|2dx(Ωufdx)2

    among all sets Ω of equal weighted volume

    V(Ω)=Ωg2dx.

    In the case where g0 is a positive constant and

    n+22<α0β,

    we can show that J(Ω) is minimized when Ω is a ball centered at the origin by a rearrangement inequality as in Pólya [32]. Indeed, Sobolev's inequality implies that J(Ω) is attained by a nonnegative function uΩH10(Ω) for α>(n+2)/2 and thus

    J(Ω)=Ω|uΩ|2dx(ΩuΩfdx)2.

    If we denote by Ω the ball centered at the origin having the same volume as Ω, and by uΩ the symmetric decreasing rearrangement of uΩ, we see that

    J(Ω)Ω|uΩ|2dx(ΩuΩfdx)2J(Ω),V(Ω)V(Ω)

    Thus, choosing a larger ball BΩ with V(B)=V(Ω), we have

    J(B)J(Ω)J(Ω),

    with equality only if Ω is a ball. We emphasize that this symmetry result only holds for the minimizer Ω, but not for every critical point Ω that admits a solution u to (1.1).

    The method of moving planes can be used to deduce the radial symmetry of any bounded domain Ω having C2-boundary in which (1.1) has a solution u with

    n<α0β. (2.1)

    Indeed, the method is based on the comparison between the solution u and its reflection ˜u(x):=u(x) in the hyperplane x1=λ in a maximal cap

    Ωλ:={x=(x1,x2,,xn)Ωx1>λ},

    where λ0 is chosen to be the smallest number so that the reflected caps

    Ωμ:={x=(2μx1,x2,,xn)RnxΩμ}(μλ)

    are all contained in Ω. Since f is non-increasing and g is non-decreasing in the radial direction if (2.1) holds, the difference u˜u is subharmonic in Ωλ, and Hopf's boundary lemma or its refined version by Serrin [34] derives a contradiction at a boundary point xΩλ unless λ=0 or u is symmetric with respect to the hyperplane x1=λ. The same argument with x1=λ moved from the opposite side, i.e., from λ= toward λ=0, deduces the symmetry of Ω with respect to x1. Hence, choosing the moving plane in every direction, we can obtain the radial symmetry of Ω.

    The aforementioned arguments by Weinberger [37] and Brandolini, Nitsch, Salani and Trombetti [7] using integral quantities and algebraic inequalities apparently work only for α=β=0.

    Our approach here is based on the existence of a spherical foliation of Rn{0} consisting of the boundaries of parametrized solutions Ω(t) (see [29,30], where a similar argument was used for a different overdetermined problem). This argument only relies on the structure of spherical solutions and is irrelevant to the monotonicity of f,g; and thus our result applies even to the case α>0 or β<0. Moreover, the result holds under a minimal regularity assumption on Ω. However, we point out that the result does not fully cover the case (2.1).

    Proposition 2.1. Let α>n, βR satisfy

    αβ+1<0,

    and let Ω be a bounded domain having C1-boundary and 0Ω. If (1.1) has a solution uC1(¯Ω{0})C2(Ω{0}) for f,g defined by (1.2) with g0=1, then Ω must be the unit ball B.

    Proof. For 0<t<, let us consider the parametrized overdetermined problem

    {Δu=(n+α)|x|αin Ω(t),u=0on Ω(t),uν=tαβ+1|x|βon Ω(t). (2.2)

    It is easy to check that Ω(t)=Bt, the ball centered at the origin with radius t, has a solution u=ut to (2.2) given by

    ut:=tα+2|x|α+2α+2.

    Now let us suppose that there is a bounded domain Ω admitting a solution uC2(¯Ω{0}) to (1.1) with g0=1 and 0Ω. We choose the largest number t>0 and the smallest number t>0 such that

    BtΩBt. (2.3)

    We will show by contradiction that t1 and t1; and thus Ω=B. Let us suppose t<1 and take a point x0ΩBt. Then, w:=uut satisfies

    {Δw=0in Bt,w=u0on Bt,w(x0)=0=minx¯Btw(x). (2.4)

    Hence we arrive at a contradiction as

    0wν(x0)=uν(x0)utν(x0)=(t)β+(t)α+1>0.

    Similarly, t>1 leads to a contradiction by considering w:=uut in Ω.

    We also obtain a symmetry result in the endpoint case αβ+1=0. In this particular case, Bt for arbitrary radius t>0 allows the solvability of (1.1) for g0=1. Indeed, ut solves (2.2) in Ω(t)=Bt with tαβ+1=1 for any t>0.

    Proposition 2.2. Let α>n, βR satisfy

    αβ+1=0,

    and let Ω be a bounded domain having C1-boundary and 0Ω. If (1.1) has a solution uC1(¯Ω{0})C2(Ω{0}) for f,g defined by (1.2) with g0=1, then Ω must be a ball centered at the origin.

    Proof. The proof proceeds similarly as before, except that in the inclusion (2.3) we will only prove that t=t. If this is not true, then BtΩ and w:=uut satisfies (2.4) and w>0 in Bt by the strong maximum principle. Hence by Hopf's lemma (used in Bt) we arrive at a contradiction as

    0>wν(x0)=uν(x0)utν(x0)=(t)β+(t)α+1=0.

    Thus Bt=Ω as desired.

    In the proofs above, the crucial step for the radial symmetry of Ω is the construction of a foliated family of domains Ω(t). This technique will be used in Section 5 to prove the uniqueness of Ω with a different foliation.

    In this section, we will first derive a functional analytic formulation of the problem of finding Ω for a prescribed g such that (1.1) is solvable, as presented in [18]. For this purpose, throughout this section we fix α>n, βR and 0<γ<1, and set

    Uk+γδ={ρhk+γ(S)ρhk+γ(S)<δ}

    for δ>0 and kN. If δ>0 is sufficiently small, for ρU2+γδ, we may define the domain Ωρ by (1.4) and a diffeomorphism θρh2+α(¯B,¯Ωρ) by

    θρ(x)={x+η(|x|1)ρ(x|x|)x|x|(x0),0(x=0),

    where ηC(R) is a cut-off function satisfying 0η1, |η|4, η(r)=1 for |r|1/4, and η(r)=0 for |r|3/4. This induces the pullback and pushforward isomorphisms θρIsom(hk+γ(¯Ωρ),hk+γ(¯B)), θρIsom(hk+γ(¯B),hk+γ(¯Ωρ)) for 0k2, as well as the corresponding boundary isomorphisms, defined by

    θρu=uθρ,θρu=uθ1ρ.

    For each ρU2+γδ, the Dirichlet problem consisting in the first two equations in (1.1) has a unique solution uρh2+γ(¯Ωρ) by the Schauder theory. Consequently, we can define the mapping FC(U2+γδ×h1+γ(S),h1+γ(S)) by

    F(ρ,g0)=θρ[uρνρ+g0|x|β], (3.1)

    where νρh1+γ(Ωρ,Rn) is the unit outer normal vector field on Ωρ. Thus, for a given g0h1+γ(S), our problem reduces to finding a solution ρU2+γδ to

    F(ρ,g0)=0.

    Indeed, for such a ρ, uρ additionally satisfies the Neumann boundary condition in (1.1). In terms of these notations, the spherical solution Ω=B for g0=1 corresponds to F(0,1)=0.

    In order to construct a solution ρ for g01 by an implicit function theorem, we will differentiate (3.1) with respect to ρ. At this point, we encounter a regularity issue, that is, we need to impose the higher regularity assumption ρh3+γ(S) for the differentiability of F as stated in the following lemma. Note that uρh3+γ(¯Ωρ) under this assumption. Here, we shall use the notation

    Nρ(x)=|x|1ρ(x|x|)(x0), (3.2)

    by which νρ and the normal and tangential components μρ,τρh2+γ(Ωρ,Rn) of the vector field θρν0 on Ωρ are represented by

    νρ=Nρ|Nρ|,μρ=νρ|Nρ|,τρ=x|x|νρ|Nρ|, (3.3)

    where we have used θρν0(x)=ν0(ξ)=ξ=x/|x| for x=ξ+ρ(ξ)ξΩρ.

    Lemma 3.1. For sufficiently small δ>0, we have

    FC1(U3+γδ×h1+γ(S),h1+γ(S))C(U2+γδ×h1+γ(S),h1+γ(S)),

    and the following hold:

    (i) The Fréchet derivative of F with respect to ρ is given by, for ˜ρh3+γ(S),

    ρF(ρ,g0)[˜ρ]=θρ[HΩρp+pνρfθρ˜ρ|Nρ|+2uρτρνρθρ˜ρ]+β(1+ρ)β1g0˜ρ (3.4)

    where ph2+γ(¯Ωρ) is the unique solution to

    {Δp=0in Ωρ,p=uρνρθρ˜ρ|Nρ|on Ωρ, (3.5)

    and HΩρh1+γ(Ωρ) is the mean curvature of Ωρ normalized in such a way that HB=n1.

    (ii) ρF(ρ,g0) has a continuous extension in L(h2+γ(S),h1+γ(S)) and

    ρFCω(U3+γδ×h1+γ(S),L(h2+γ(S),h1+γ(S))).

    Proof. We give a sketch of proof and refer to [18,Lemma 2.1] for details. Let us first derive the formula (3.4) in the simplest case where ρ=0. To this end, for ˜ρh3+α(S) and small εR, we substitute ρ=ε˜ρ and the formal expansion uρ=u0+εp+o(ε) into

    {Δuρ=fin Ωρ,uρ=0on Ωρ

    and use the corresponding equation for u0 to obtain

    f(x)=Δuρ(x)=f(x)εΔp(x)+o(ε)for xΩ0=B,0=uρ(x+ε˜ρν0)=εu0ν0(x)˜ρ+εp(x)+o(ε)for xΩ0=S.

    Thus, letting ε0, we see that p satisfies (3.5) for ρ=0. Moreover, for xS,

    F(ε˜ρ,g0)(x)=uρνρ(x+ε˜ρν0)+g0(x)(1+ε˜ρ)β=uρν0(x+ε˜ρν0)+εuρτ(x+ε˜ρν0)+g0(x)(1+εβ˜ρ)+o(ε)=F(0,g0)(x)+ε2u0ν02(x)˜ρ+εpν0(x)+εu0τ(x)+εβg0(x)˜ρ+o(ε)=F(0,g0)(x)εf(x)˜ρ+εHΩ0p(x)+εpν0(x)+εβg0(x)˜ρ+o(ε),

    where we used the fact that νρ and 2u0/ν20 can be represented by a tangent vector τ to S and the Laplace-Beltrami operator ΔS on S as

    νρ=ν0+ετ+o(ε),2u0ν02(x)=Δu0(x)ΔSu0(x)HSu0ν0(x)=f(x)HSu0ν0(x).

    This shows (i) for ρ=0 by letting ε0. For general ρ0, we use the same argument as above with the reference domain Ωρ instead of Ω0=B. In particular, every occurrence of ˜ρ must now be replaced by θρ˜ρ/|Nρ| and the extra term ε(θρ˜ρ)τρνρuρ appears in the expansion of F(ρ+ε˜ρν0,g), since for xS

    x+(ρ(x)+ε˜ρ(x))ν0=θρ(x)+ε˜ρ(x){μρ(θρ(x))+τρ(θρ(x))}.

    For (ii), we observe that the formula (3.4) with (3.5) still makes sense for ρU3+γδ and ˜ρh2+γ(S); and the extension in (ii) is thus defined. Finally, the analyticity of ρF follows from that of U3+γδρθρHΩρh1+γ(S), θρuρh3+γ(¯B) and θρph2+γ(¯B).

    Remark 3.2. The required higher regularity ρU3+γδ is adequate in view of the formula in (i), since, if ρhk+γ(S), then at most HΩρhk2+γ(Ωρ), uρhk+γ(¯Ωρ) and phk1+γ(¯Ωρ).

    As stated in the next lemma, the representation formula (3.4) of the Fréchet derivative of F in Lemma 3.1 yields a characterization of the invertibility of the extended operator ρF(ρ,g0) in terms of the elliptic boundary value problem

    {Δp=0in Ωρ,(HΩρfg+1ggνρ)p+pνρ=φon Ωρ. (3.6)

    Note that F(ρ,g0)=0 implies that g=νρuρh2+γ(Ωρ) and g>0 on Ωρ, where the latter follows from the maximum principle.

    Lemma 3.3. Suppose that ρU3+γδ and g0h2+γ(S) satisfy F(ρ,g0)=0. Then, the inverse

    ρF(ρ,g0)1L(h1+γ(S),h2+γ(S))

    exists if and only if (3.6) has a unique solution ph2+γ(¯Ωρ) for φh1+γ(Ωρ). Furthermore, the inverse is then given by

    ρF(ρ,g0)1[θρφ]=θρ[p|Nρ|g]. (3.7)

    Proof. By assumption, νρuρ=g on Ωρ. Moreover, in view of (1.2) and (3.3),

    θρ2uρτρνρ=θρgτρ=β(1+ρ)β1g0+θρ[1|Nρ|gνρ]on S.

    Hence, the boundary condition in (3.5) becomes

    θρ˜ρ=p|Nρ|gon Ωρ, (3.8)

    and the remaining condition in (3.5) and (3.4) are

    {Δp=0in Ωρ,(HΩρfg+1ggνρ)p+pνρ=θρρF(ρ,g0)[˜ρ]on Ωρ.

    Since θρ,θρ are isomorphisms and (3.8) yields a one-to-one correspondence between ˜ρ and p, the invertibility of ρF(ρ,g0)L(h2+γ(S),h1+γ(S)) is equivalent to the unique existence of a solution ph2+γ(¯Ωρ) to (3.6) for any given boundary data φh1+γ(Ωρ). The formula (3.7) follows from the above equations.

    A well-known sufficient condition (see Gilbarg and Trudinger [17,Theorem 6.31]) for the unique solvability of (3.6) is

    HΩρfg+1ggνρ>0on Ωρ.

    In particular, for ρ=0 and g0=1, this positivity condition is nothing but

    αβ+1<0.

    In fact, we can classify all the values of α,β for which (3.6) is uniquely solvable by virtue of spherical harmonics. To this end, we recall some basic properties of spherical harmonics. Let us denote by Hl the vector space of all homogeneous harmonic polynomials of degree lN{0} on Rn. The dimension d(n)l of Hl is given by d(n)0=1, d(n)1=n and

    d(n)l=(l+n1l)+(l+n3l2)(l2).

    If we regard Hl as a subspace of L2(S) and choose an orthonormal basis

    {hl,1,hl,2,,hl,d(n)l}Hl,

    then it is known that

    B=l=0{hl,1,hl,2,,hl,d(n)l}

    forms a complete orthonormal system of L2(S). In particular, uhk+γ(S) can be expressed by its Fourier series in L2(S) as

    u=l=0d(n)lm=1ˆul,mhl,m.

    Moreover, if uC(S), the coefficients ˆul,m are rapidly decaying so that the series on the right hand side converges in the norm in Ck+γ(S). Thus, the linear subspace spanned by B is dense in hk+γ(S). We also note that hl,m satisfies

    hl,mν=xhl,m=lhl,mon S.

    In other words, the Dirichlet-to-Neumann operator NL(hk+γ(S),hk1+γ(S)), for k2, defined by

    Nφ=vν,

    where vhk+γ(¯B) is the unique solution to the Dirichlet problem

    {Δv=0in B,v=φon S,

    satisfies

    Nhl,m=lhl,m.

    Lemma 3.4. Let ρ=0, g0=1, α>n and βR. The boundary value problem (3.6) has a unique solution ph2+γ(¯B) for any φh1+γ(S) if and only if

    αβ+1N{0}. (3.9)

    Proof. The boundary condition in (3.6) can be written as

    (1α+β)p+pν=φ.

    Hence, (3.6) is uniquely solvable if and only if

    (1α+β)I+NL(h2+γ(S),h1+γ(S)) (3.10)

    is invertible. As remarked earlier, the latter holds if 1α+β>0; and in particular (I+N)1L(h1+γ(S),h2+γ(S)) exists. Thus, by the Fredholm theory (see e.g., [17, Theorem 5.3]) applied to

    ((1α+β)I+N)(I+N)1=I+(2α+β)(I+N)1,

    where (I+N)1 is compact as a mapping from h1+γ(S) to itself, the range of (3.10) is closed and the invertibility follows from the surjectivity of (3.10). Now, since

    (1α+β)hl,m+Nhl,m=(l1α+β)hl,m,

    the condition (3.9) implies that the range of (3.10) contains the linear span of B and hence its closure h1+γ(S). On the other hand, if there is an lN{0} such that αβ+1=l, then obviously (3.10) is not injective from the above computation.

    Our goal in this section is to derive the existence, uniqueness and regularity of a solution ρ to the nonlinear equation F(ρ,g0)=0, based on the linear analysis in the previous section. Lemmas 3.3 and 3.4 show that the linearized operator ρF(0,1) has the inverse

    ρF(0,1)1L(h1+γ(S),h2+γ(S)) (4.1)

    as long as (3.9) holds. This indicates that F(ρ,g0)=0 can be locally solved and the solution map g0ρ is differentiable.

    However, a classical perturbation method generally fails for our nonlinear equation defined by (3.1). Indeed, in contrast to F(,g0)C1(U3+γδ,h1+γ(S)), the inverse (4.1) only recovers a partial regularity that is not sufficient for a successive approximation of the form

    ρj+1=ρjρF(0,1)1F(ρj,g0),ρ0=0

    to converge, since ρjh3+γ(S) only results in ρj+1h2+γ(S). This regularity deficit called the loss of derivatives can be circumvented by the following implicit function theorem introduced by Gilsbach and the author [18,Theorem 4.2]. It requires some additional regularity assumptions on F at the single point (ρ,g0)=(0,1), but provides the existence, local uniqueness and differentiability of g0ρ.

    Proposition 4.1. Let X2X1X0, Z2Z1Z0 and Y be Banach spaces with inclusions being continuous embeddings, and let Dj be a neighborhood of a point (x0,y0)X2×Y in Xj×Y with D2D1D0 for j=0,1,2. Suppose that

    FC1(D2,Z1)C(D1,Z1)C1(D1,Z0)C(D0,Z0) (4.2)

    satisfies the following conditions:

    (a) F(x0,y0)=0;

    (b) xF(x,y)L(X2,Z1)L(X1,Z0) has an extension with

    xFC(D2,L(X1,Z1))C(D1,L(X0,X0));

    (c) F is differentiable at (x0,y0) as a mapping from Dj to Zj for j=1,2;

    (d) xF(x0,y0)1L(Zj,Xj) exists for j=0,1,2.

    Then, there are a neighborhood U1 of x0X1, a neighborhood V of y0Y and a mapping υ:VU1 satisfying

    (i) F(υ(y),y)=0 for all yV;

    (ii) υ(y0)=x0 and υ(y)x0 in X1 as yy0 in Y;

    (iii) F(x,y)=0, xU1 and yV imply that x=υ(y);

    (iv) υC1(V,X0) and

    υ(y)=xF(υ(y),y)1yF(υ(y),y)L(Y,X0).

    Remark 4.2. The continuity in (ii) is not explicitly stated in [18,Theorem 4.2]. But it is clear from the proofs of [18,Theorems 4.1 and 4.2]: choose an arbitrarily small neighborhood U1U1 of x0 and then take a small VV accordingly and use the contraction mapping principle to get υ(y)U1 for yV. $

    This proposition enables us to handle nonlinear problems having a particular type of loss of derivatives specified in the conditions (b) and (d). The assumption (c) can be regarded as no loss of derivatives occurring at the single point (x0,y0).

    In order to apply Proposition 4.1 to our problem with F defined by (3.1), we set, for j=0,1,2,

    Xj=hj+2+γ(S),Zj=hj+1+γ(S),Y=h2+γ(S),Dj=Uj+2+γδ×Y.

    As in Lemma 3.1, we have

    FC1(U4+γδ×h2+γ(S),h2+γ(S))C(U3+γδ×h2+γ(S),h2+γ(S))C1(U3+γδ×h1+γ(S),h1+γ(S))C(U2+γδ×h1+γ(S),h1+γ(S)).

    This implies that F meets the regularity assumption (4.2). Similarly, (a), (b) are easily confirmed with (x0,y0)=(0,1). Now we use Lemma 3.4 and its variant in higher regular spaces to conclude that the non-degeneracy condition (d) is satisfied if (3.9) holds. For the remaining condition (c), we recall that the loss of derivatives is caused by the regularity of the mean curvature HΩρ (see Remark 3.2) and by several other terms in (3.4). However, at (ρ,g0)=(0,1) where Ω0=S, non-smooth terms vanish and we have

    ρF(0,1)[˜ρ]=(n1)p+pν(n+α)˜ρ+β˜ρ

    and p is as smooth as ˜ρ by (3.5). Thus one can check that FC(Dj,Zj), with the image space having the stronger topology, is still differentiable at (ρ,g0)=(0,1) for j=1,2. As a conclusion, we obtain open sets U1h3+γ(S) and Vh2+γ(S) with (0,1)U1×V and a solution map ρ:VU1 such that

    (i) F(ρ(g0),g0)=0 for all g0h2+γ(S);

    (ii) ρ(1)=0 and ρ(g0)0 in h3+γ(S) as g01 in h2+γ(S);

    (iii) F(˜ρ,g0)=0, ˜ρU1 and g0V imply that ˜ρ=ρ(g0).

    Moreover, ρC1(V,h2+γ(S)) and hence the linear stability estimate

    ρ(g0)h2+γ(S)Cg01h2+γ(S)

    holds for any g0V in a small neighborhood of 1h2+γ(S). This concludes the proof of Theorem 1.1, except the last assertion on the global uniqueness of Ω.

    Our remaining task is to prove that Ωρ constructed in the previous section is the only possible bounded domain having a solution u to (1.1) in the case where

    αβ+1<0.

    The technique we employ is based on the construction of a foliation of Rn{0} by the boundaries of particular solutions Ω(t) to (1.1), as in the proof of Propositions 2.1 and 2.2. However, for non-constant g0, the spherical foliation is no longer suitable and we need to construct a non-spherical one. We rely on an argument used by Bianchini, Henrot and Salani [6,Theorem 3.4], where the uniqueness is proved for α=0 by constructing a non-spherical foliation as the boundaries of the rescaled family

    tΩ:={txRnxΩ}.

    The following proposition is a generalization of [6,Theorem 3.4] to the case of arbitrary α>n, which completes the proof of Theorem 1.1.

    Proposition 5.1. Let α>n, βR and g0C(S) satisfy

    αβ+1<0,g0(ξ)>0,

    and let Ω and ˜Ω be bounded domains having C1-boundaries and 0Ω˜Ω. Suppose that (1.1) in Ω and ˜Ω respectively admit solutions

    uC1(¯Ω{0})C2(Ω{0}),˜uC1(¯˜Ω{0})C2(˜Ω{0})

    with f,g defined by (1.2). Then Ω=˜Ω.

    Proof. The proof is similar to that of Proposition 2.1 with Ω(t)=tΩ for 0<t<. It is easy to see that the parametrized overdetermined problem

    {Δu=(n+α)|x|αin Ω(t),u=0on Ω(t),uν=tαβ+1g0(ξ)|x|βon Ω(t).

    has a solution

    ut(x):=tα+2u(xt)(xΩ(t)).

    As in the proof of Proposition 2.1, we can choose the largest number t>0 and the smallest number t>0 such that

    Ω(t)˜ΩΩ(t),

    and prove that t1 and t1 and hence ˜Ω=Ω(1)=Ω by contradiction. Indeed, if t<1, we take a point x0˜ΩΩ(t) and observe that w:=˜uut satisfies

    {Δw=0in Ω(t),w=˜u0on Ω(t),w(x0)=0=minx¯Ω(t)w(x). (5.1)

    Hence we arrive at a contradiction as

    0wν(x0)=˜uν(x0)utν(x0)={1+(t)αβ+1}g0(x0|x0|)|x0|β>0.

    Similarly, t>1 leads to a contradiction by considering w:=˜uut in ˜Ω.

    In the endpoint case αβ+1=0, the existence of Ω for g01 is not guaranteed due to the degeneracy of ρF(0,1). However, we can prove the uniqueness of Ω up to dilation as in Proposition 2.2.

    We say that Ω satisfies the interior sphere condition if for any point xΩ there is a ball BΩ such that xB. In particular, this condition is fulfilled if Ω is of class C2. The interior sphere condition allows us to use Hopf's lemma.

    Proposition 5.2. Let α>n, βR and g0C(S) satisfy

    αβ+1=0,g0(ξ)>0,

    and let Ω and ˜Ω be bounded domains having C1-boundaries and 0Ω˜Ω, and moreover suppose that Ω or ˜Ω satisfies the interior sphere condition. If (1.1) in Ω and ˜Ω respectively admit solutions

    uC1(¯Ω{0})C2(Ω{0}),˜uC1(¯˜Ω{0})C2(˜Ω{0})

    with f,g defined by (1.2), then ˜Ω=tΩ for some t>0.

    Proof. We may suppose that Ω satisfies the interior sphere condition. The same argument as in the proof of Proposition 5.1 yields Ω(t)˜Ω with a common boundary point x0˜ΩΩ(t). If Ω(t)˜Ω, then w:=˜uut satisfies (5.1) and w>0 in Ω(t) by the strong maximum principle. Hence Hopf's lemma derives a contradiction as

    0>wν(x0)=˜uν(x0)utν(x0)=0.

    Thus Ω(t)=˜Ω as desired.

    The author was supported in part by the Grant-in-Aid for Scientific Research (C) 20K03673, Japan Society for the Promotion of Science.

    The author declares no conflict of interest.



    [1] A. Acker, Interior free boundary problems for the Laplace equation, Arch. Rational Mech. Anal., 75 (1981), 157–168. https://doi.org/10.1007/BF00250477 doi: 10.1007/BF00250477
    [2] A. Acker, Uniqueness and monotonicity of solutions for the interior Bernoulli free boundary problem in the convex, n-dimensional case, Nonlinear Anal. Theor., 13 (1989), 1409–1425. https://doi.org/10.1016/0362-546X(89)90102-8 doi: 10.1016/0362-546X(89)90102-8
    [3] A. Aftalion, J. Busca, W. Reichel, Approximate radial symmetry for overdetermined boundary value problems, Adv. Differential Equations, 4 (1999), 907–932.
    [4] V. Agostiniani, R. Magnanini, Symmetries in an overdetermined problem for the Green's function, Discrete Cont. Dyn. Syst. S, 4 (2011), 791–800. https://doi.org/10.3934/dcdss.2011.4.791 doi: 10.3934/dcdss.2011.4.791
    [5] C. Bianchini, A Bernoulli problem with non-constant gradient boundary constraint, Appl. Anal., 91 (2012), 517–527. https://doi.org/10.1080/00036811.2010.549479 doi: 10.1080/00036811.2010.549479
    [6] C. Bianchini, A. Henrot, P. Salani, An overdetermined problem with non-constant boundary condition, Interfaces Free Bound., 16 (2014), 215–241. https://doi.org/10.4171/IFB/318 doi: 10.4171/IFB/318
    [7] B. Brandolini, C. Nitsch, P. Salani, C. Trombetti, Serrin-type overdetermined problems: allternative proof, Arch. Rational Mech. Anal., 190 (2008), 267–280. https://doi.org/10.1007/s00205-008-0119-3 doi: 10.1007/s00205-008-0119-3
    [8] B. Brandolini, C. Nitsch, P. Salani, C. Trombetti, On the stability of the Serrin problem, J. Differ. Equations, 245 (2008), 1566–1583. https://doi.org/10.1016/j.jde.2008.06.010 doi: 10.1016/j.jde.2008.06.010
    [9] L. A. Caffarelli, H. W. Alt, Existence and regularity for a minimum problem with free boundary, J. Reine Angew. Math., 325 (1981), 105–144. https://doi.org/10.1515/crll.1981.325.105 doi: 10.1515/crll.1981.325.105
    [10] L. Cavallina, T. Yachimura, On a two-phase Serrin-type problem and its numerical computation, ESAIM: COCV, 26 (2020), 65. https://doi.org/10.1051/cocv/2019048 doi: 10.1051/cocv/2019048
    [11] L. Cavallina, Local analysis of a two phase free boundary problem concerning mean curvature, Indiana Univ. Math. J., in press.
    [12] L. Cavallina, T. Yachimura, Symmetry breaking solutions for a two-phase overdetermined problem of Serrin-type, In: Current trends in analysis, its applications and computation (Proceedings of the 12th ISAAC Congress, Aveiro, Portugal, 2019), Birkhäuser, 2022, in press.
    [13] G. Ciraolo, R. Magnanini, V. Vespri, Hölder stability for Serrin's overdetermined problem, Annali di Matematica, 195 (2016), 1333–1345. https://doi.org/10.1007/s10231-015-0518-7 doi: 10.1007/s10231-015-0518-7
    [14] M. Del Pino, F. Pacard, J. Wei, Serrin's overdetermined problem and constant mean curvature surfaces, Duke Math. J., 164 (2015), 2643–2722. https://doi.org/10.1215/00127094-3146710 doi: 10.1215/00127094-3146710
    [15] M. M. Fall, I. A. Minlend, T. Weth, Unbounded periodic solutions to Serrin's overdetermined boundary value problem, Arch. Rational Mech. Anal., 223 (2017), 737–759. https://doi.org/10.1007/s00205-016-1044-5 doi: 10.1007/s00205-016-1044-5
    [16] W. M. Feldman, Stability of Serrin's problem and dynamic stability of a model for contact angle motion, SIAM J. Math. Anal., 50 (2018), 3303–3326. https://doi.org/10.1137/17M1143009 doi: 10.1137/17M1143009
    [17] D. Gilbarg, N. S. Trudinger, Elliptic partial differential equations of second order, Berlin, Heidelberg: Springer, 2001. https://doi.org/10.1007/978-3-642-61798-0
    [18] A. Gilsbach, M. Onodera, Linear stability estimates for Serrin's problem via a modified implicit function theorem, Calc. Var., 60 (2021), 241. https://doi.org/10.1007/s00526-021-02107-1 doi: 10.1007/s00526-021-02107-1
    [19] A. Henrot, M. Onodera, Hyperbolic solutions to Bernoulli's free boundary problem, Arch. Rational Mech. Anal., 240 (2021), 761–784. https://doi.org/10.1007/s00205-021-01620-z doi: 10.1007/s00205-021-01620-z
    [20] A. Henrot, H. Shahgholian, The one phase free boundary problem for the p-Laplacian with non-constant Bernoulli boundary condition, Trans. Amer. Math. Soc., 354 (2002), 2399–2416. https://doi.org/10.1090/S0002-9947-02-02892-1 doi: 10.1090/S0002-9947-02-02892-1
    [21] N. Kamburov, L. Sciaraffia, Nontrivial solutions to Serrin's problem in annular domains, Ann. Inst. H. Poincaré C Anal. Non Linéaire, 38 (2021), 1–22. https://doi.org/10.1016/j.anihpc.2020.05.001 doi: 10.1016/j.anihpc.2020.05.001
    [22] R. Magnanini, G. Poggesi, On the stability for Alexandrov's soap bubble theorem, J. Anal. Math., 139 (2019), 179–205. https://doi.org/10.1007/s11854-019-0058-y doi: 10.1007/s11854-019-0058-y
    [23] R. Magnanini, G. Poggesi, Serrin's problem and Alexandrov's soap bubble theorem: enhanced stability via integral identities, Indiana Univ. Math. J., 69 (2020), 1181–1205. https://doi.org/10.1512/iumj.2020.69.7925 doi: 10.1512/iumj.2020.69.7925
    [24] R. Magnanini, G. Poggesi, Nearly optimal stability for Serrin's problem and the Soap Bubble theorem, Calc. Var., 59 (2020), 35. https://doi.org/10.1007/s00526-019-1689-7 doi: 10.1007/s00526-019-1689-7
    [25] R. Magnanini, G. Poggesi, Interpolating estimates with applications to some quantitative symmetry results, Mathematics in Engineering, 5 (2023), 1–21. https://doi.org/10.3934/mine.2023002 doi: 10.3934/mine.2023002
    [26] F. Morabito, P. Sicbaldi, Delaunay type domains for an overdetermined elliptic problem in Sn×R and Hn×R, ESAIM: COCV, 22 (2016), 1–28. https://doi.org/10.1051/cocv/2014064 doi: 10.1051/cocv/2014064
    [27] F. Morabito, Symmetry breaking bifurcations for two overdetermined boundary value problems with non-constant Neumann condition on exterior domains in R3, Commun. Part. Diff. Eq., 46 (2021), 1137–1161. https://doi.org/10.1080/03605302.2020.1871363 doi: 10.1080/03605302.2020.1871363
    [28] M. Onodera, Geometric flows for quadrature identities, Math. Ann., 361 (2015), 77–106. https://doi.org/10.1007/s00208-014-1062-2 doi: 10.1007/s00208-014-1062-2
    [29] M. Onodera, On the symmetry in a heterogeneous overdetermined problem, Bull. Lond. Math. Soc., 47 (2015), 95–100. https://doi.org/10.1112/blms/bdu098 doi: 10.1112/blms/bdu098
    [30] M. Onodera, Dynamical approach to an overdetermined problem in potential theory, J. Math. Pure. Appl., 106 (2016), 768–796. https://doi.org/10.1016/j.matpur.2016.03.011 doi: 10.1016/j.matpur.2016.03.011
    [31] L. E. Payne, P. W. Schaefer, Duality theorems in some overdetermined boundary value problems, Math. Method. Appl. Sci., 11 (1989), 805–819. https://doi.org/10.1002/mma.1670110606 doi: 10.1002/mma.1670110606
    [32] G. Pólya, Torsional rigidity, principal frequency, electrostatic capacity and symmetrization, Quart. Appl. Math., 6 (1948), 267–277. https://doi.org/10.1090/QAM/26817 doi: 10.1090/QAM/26817
    [33] F. Schlenk, P. Sicbaldi, Bifurcating extremal domains for the first eigenvalue of the Laplacian, Adv. Math., 229 (2012), 602–632. https://doi.org/10.1016/j.aim.2011.10.001 doi: 10.1016/j.aim.2011.10.001
    [34] J. Serrin, A symmetry problem in potential theory, Arch. Rational Mech. Anal., 43 (1971), 304–318. https://doi.org/10.1007/BF00250468 doi: 10.1007/BF00250468
    [35] H. Shahgholian, Diversifications of Serrin's and related symmetry problems, Complex Var. Elliptic Equ., 57 (2012), 653–665. https://doi.org/10.1080/17476933.2010.504848 doi: 10.1080/17476933.2010.504848
    [36] P. Sicbaldi, New extremal domains for the first eigenvalue of the Laplacian in flat tori, Calc. Var., 37 (2010), 329–344. https://doi.org/10.1007/s00526-009-0264-z doi: 10.1007/s00526-009-0264-z
    [37] H. F. Weinberger, Remark on the preceding paper of Serrin, Arch. Rational Mech. Anal., 43 (1971), 319–320. https://doi.org/10.1007/BF00250469 doi: 10.1007/BF00250469
  • This article has been cited by:

    1. Filomena Pacella, Giorgio Poggesi, Alberto Roncoroni, Optimal quantitative stability for a Serrin-type problem in convex cones, 2024, 307, 0025-5874, 10.1007/s00209-024-03555-z
    2. Giorgio Poggesi, Remarks about the mean value property and some weighted Poincaré-type inequalities, 2024, 203, 0373-3114, 1443, 10.1007/s10231-023-01408-w
    3. Tynysbek Kalmenov, Nurbek Kakharman, An overdetermined problem for elliptic equations, 2024, 9, 2473-6988, 20627, 10.3934/math.20241002
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2174) PDF downloads(250) Cited by(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog