This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\RedeclareSectionCommand

[ runin=true, ]subsection \RedeclareSectionCommand[ runin=true, ]subsubsection

Linear stability estimates for Serrin’s problem via a modified implicit function theorem

Alexandra Gilsbach Department of Mathematics, School of Science, Tokyo Institute of Technology Michiaki Onodera Department of Mathematics, School of Science, Tokyo Institute of Technology
Abstract

Abstract. We examine Serrin’s classical overdetermined problem under a perturbation of the Neumann boundary condition. The solution of the problem for a constant Neumann boundary condition exists provided that the underlying domain is a ball. The question arises whether for a perturbation of the constant there still are domains admitting solutions to the problem. Furthermore, one may ask whether a domain that admits a solution for the perturbed problem is unique up to translation and whether it is close to the ball. We develop a new implicit function theorem for a pair of Banach triplets that is applicable to nonlinear problems with loss of derivatives except at the point under consideration. Combined with a detailed analysis of the linearized operator, we prove the existence and local uniqueness of a domain admitting a solution to the perturbed overdetermined problem. Moreover, an optimal linear stability estimate for the shape of a domain is established.

Keywords: overdetermined problem & implicit function theorem & stability estimates

MSC2010: 35N25 & 35B35 & 47J07

1 Introduction

We study the shape of a bounded domain Ωn\Omega\subset\mathbb{R}^{n}, n2n\geq 2, in which a solution uu to the Dirichlet problem

(1.1a) Δu=1inΩ,u=0onΓ=Ω\displaystyle\begin{aligned} -\Delta u&=1\quad\text{in}\ \Omega,\\ u&=0\quad\text{on}\ \Gamma=\partial\Omega\end{aligned}
satisfies the overdetermined boundary condition
(1.1b) uν=fonΓ,\displaystyle-\frac{\partial u}{\partial\nu}=f\quad\text{on}\ \Gamma,

where ν\nu is the outer unit normal vector to Γ\Gamma and ff is a prescribed positive function defined on n\mathbb{R}^{n}.

The overdetermined problem (1.1) arises in a shape optimization problem called the Saint-Venant problem, in which one maximizes the torsional rigidity

P(Ω)=supuH01(Ω){0}(Ωu𝑑x)2Ω|u|2𝑑xP(\Omega)=\sup_{u\in H_{0}^{1}(\Omega)\setminus\{0\}}\frac{\left(\int_{\Omega}u\,dx\right)^{2}}{\int_{\Omega}|\nabla u|^{2}\,dx}

of a bar with cross section Ω\Omega, among all sets Ω\Omega of equal weighted volume

V(Ω)=Ωf2𝑑x.V(\Omega)=\int_{\Omega}f^{2}\,dx.

The Euler-Lagrange equation, after multiplying a normalizing constant, consists in (1.1). In the case where ff is a constant, Pólya [17] proved that the maximizer Ω\Omega of PP must be a ball with the prescribed volume VV, using the symmetric rearrangement of a function. This applies to a more general situation, where ff is radially symmetric and non-decreasing in the radial direction.

In fact, the same symmetry result also holds for all critical points, namely, if ff is a constant, then (1.1) has a solution uu if and only if Ω\Omega is a ball. In particular, for the normalized constant f=1nf=\tfrac{1}{n}, Ω\Omega is a ball of radius one and u(x)=12n(1|xc|2)u(x)=\tfrac{1}{2n}(1-|x-c|^{2}) with cc being the center of the ball. This well-known symmetry result is due to Serrin [18]. The proof introduces the method of moving planes motivated by Alexandrov’s reflection principle [2] originally used to establish the soap bubble theorem. This symmetry result can be alternatively proven by an ingenious combination of the Rellich-Pohozaev integral identity and elementary inequalities (see Weinberger [19], and Brandolini, Nitsch, Salani, and Trombetti [5]), or by a continuous version of the Steiner symmetrization (see Brock and Henrot [7]).

The objective of this paper is the stability of a domain Ω\Omega under a perturbation of the Neumann boundary condition (1.1b), which naturally arises if one considers the torsional rigidity in anisotropic media. Namely, setting Ω0:=𝔹\Omega_{0}:=\mathbb{B}, the nn-dimensional unit ball centered at the origin, and

(1.2) f(x)=1n+g(x|x|)(xn{0})f(x)=\frac{1}{n}+g\left(\frac{x}{|x|}\right)\quad(x\in\mathbb{R}^{n}\setminus\{0\})

with a prescribed function gg defined on Γ0:=𝕊\Gamma_{0}:=\mathbb{S}, where 𝕊=𝔹\mathbb{S}=\partial\mathbb{B}, we prove the existence and local uniqueness of Ω\Omega admitting a solution uu to (1.1), and establish a quantitative estimate of the deviation of Ω\Omega from Ω0\Omega_{0} in terms of the perturbation gg.

The domain deviation is measured by a function ρ=ρ(ζ)(1,)\rho=\rho(\zeta)\in(-1,\infty) which defines the star-shaped bounded domain Ωρ\Omega_{\rho} enclosed by

(1.3) Γρ:={ζ+ρ(ζ)ζζ𝕊}.\Gamma_{\rho}:=\left\{\zeta+\rho(\zeta)\zeta\mid\zeta\in\mathbb{S}\right\}.

A domain Ω\Omega admitting a solution to (1.1) will also be referred to as a solution of the problem.

In what follows, hk+α(Ω¯)h^{k+\alpha}(\overline{\Omega}) denotes the little Hölder space defined as the closure of the Schwartz space 𝒮\mathcal{S} of rapidly decreasing functions in Ck+α(Ω¯)C^{k+\alpha}(\overline{\Omega}), and similarly hk+α(Γ)h^{k+\alpha}(\Gamma) for a hypersurface Γ\Gamma (see Lunardi [13]).

In order to motivate our study, let us mention several related results concerning existence and stability of solutions to (1.1). The existence of Ω\Omega for non-constant ff is known (see Bianchini, Henrot and Salani [4]) in the case where ff is positively homogeneous, i.e.,

(1.4) f(tx)=tγf(x)(t>0,xn)f(tx)=t^{\gamma}f(x)\quad(t>0,\ x\in\mathbb{R}^{n})

for γ>0\gamma>0 with γ1\gamma\neq 1 with ff being Hölder continuous on n{0}\mathbb{R}^{n}\setminus\{0\}. This condition ensures the existence of a maximizer Ω\Omega of the Saint-Venant problem, and a solution uu to (1.1a) then satisfies νu=λf-\partial_{\nu}u=\lambda f on Γ\Gamma with a Lagrange multiplier λ>0\lambda>0. The γ\gamma-homogeneity of ff allows us to control λ\lambda by considering tΩ:={txxΩ}t\Omega:=\{tx\mid x\in\Omega\}, and indeed t=λ1/(1γ)t=\lambda^{1/(1-\gamma)} gives a desired domain. However, the 0-homogeneous case (1.2) cannot be treated by this variational approach, since the dichotomy of a maximizing sequence cannot be excluded in the concentration-compactness alternative.

Most of the existing stability results in the literature for (1.1), fitted into our context by translation and dilation, take inequalities of the form

(1.5) ρL(𝒮n1)C[uΩν+R]Xτ,\|\rho\|_{L^{\infty}(\mathcal{S}^{n-1})}\leq C\left[\frac{\partial u_{\Omega}}{\partial\nu}+R\right]_{X}^{\tau},

where uΩu_{\Omega} is a solution to (1.1a) in Ω=Ωρ\Omega=\Omega_{\rho} with C2+αC^{2+\alpha}-boundary, 0<τ10<\tau\leq 1 and []X[\,\cdot\,]_{X} denotes a norm or seminorm which measures the deviation of νuΩ-\partial_{\nu}u_{\Omega} from a constant R>0R>0. Aftalion, Busca and Reichel [1] adopted a quantitative version of the method of moving planes and obtained a logarithmic version of (1.5) with X=C1(Γ)X=C^{1}(\Gamma). The method was further developed by Ciraolo, Magnanini and Vespri [8], and they obtained (1.5) for some 0<τ<10<\tau<1 in terms of the Lipschitz seminorm of X=Lip(Γ)X=\text{Lip}(\Gamma). In fact, these results also hold for semilinear equations Δu=f(u)-\Delta u=f(u) with u>0u>0. On the other hand, Brandolini, Nitsch, Salani and Trombetti [6] made use of integral identities and proved (1.5) with X=L(Γ)X=L^{\infty}(\Gamma) for some 0<τ<10<\tau<1. Moreover, they obtained an estimate of the volume of the symmetric difference of Ω\Omega and a union of balls by a weaker norm, i.e., X=L1(Γ)X=L^{1}(\Gamma). Note that the problem (1.1) admits a domain Ω\Omega composed of a finite number of balls joined by tiny tentacles if we only control the extra boundary condition (1.1b) by the L1L^{1}-norm. Following this approach, Feldman [9] obtained the sharp estimate

|Ω𝔹|CuΩν+RL2(Γ),|\Omega\triangle\mathbb{B}|\leq C\left\|\frac{\partial u_{\Omega}}{\partial\nu}+R\right\|_{L^{2}(\Gamma)},

where |Ω𝔹||\Omega\triangle\mathbb{B}| is the volume of the symmetric difference of Ω\Omega and 𝔹\mathbb{B} and is considered as ρL1(𝕊)\|\rho\|_{L^{1}(\mathbb{S})} for star-shaped Ω=Ωρ\Omega=\Omega_{\rho}. The linear (i.e., τ=1\tau=1) stability estimate has also been expected in (1.5). Recently, Magnanini and Poggesi proved (1.5) with X=L2(Γ)X=L^{2}(\Gamma) and τ=1\tau=1 for n=2n=2, τ\tau arbitrarily close to 11 for n=3n=3, and τ=2n1\tau=\frac{2}{n-1} for n4n\geq 4 [14].

In general, for overdetermined problems, the super-subsolution method based on the maximum principle provides an existence criterion. In our setting, a bounded domain Ω\Omega is called a supersolution to (1.1) if the unique solution u=uΩu=u_{\Omega} to (1.1a) satisfies

uΩνfonΓ,-\frac{\partial u_{\Omega}}{\partial\nu}\leq f\quad\text{on}\ \Gamma,

and a subsolution is defined analogously with the opposite inequality. The existence of a solution, i.e., a bounded domain Ω\Omega in which uΩu_{\Omega} satisfies (1.1b), is guaranteed provided there are a supersolution Ωsup\Omega_{\sup} and a subsolution Ωsub\Omega_{\text{sub}} satisfying ΩsubΩsup\Omega_{\text{sub}}\subset\Omega_{\sup}. Typically, balls 𝔹r\mathbb{B}_{r} with large or small radii r>0r>0 give super- or subsolutions. Indeed, for Ω=𝔹r\Omega=\mathbb{B}_{r},

u𝔹r(x)=r2|x|22nu_{\mathbb{B}_{r}}(x)=\frac{r^{2}-|x|^{2}}{2n}

with νu𝔹r=rn-\partial_{\nu}u_{\mathbb{B}_{r}}=\frac{r}{n} on 𝔹r\partial\mathbb{B}_{r} solves (1.1), and we see that, in the γ\gamma-homogeneous setting (1.4) with γ>1\gamma>1, 𝔹r\mathbb{B}_{r} with large (resp. small) r>0r>0 is a supersolution (resp. subsolution); while for 0γ<10\leq\gamma<1, 𝔹r\mathbb{B}_{r} with large (resp. small) r>0r>0 is a subsolution (resp. supersolution). Hence these balls provide an appropriate pair of super- and subsolutions only if γ>1\gamma>1.

We therefore take another approach in this paper based on an implicit function theorem, yielding linear stability estimates with Hölder norms on both sides of the estimate, as well as the existence and local uniqueness of Ω\Omega for a given perturbation gg in (1.2). We will need to exploit detailed properties of the linearized equation

(1.6a) Δp=0inΩρ0,(H1f)p+pν=φonΓρ0,\displaystyle\begin{aligned} -\Delta p&=0&&\text{in}\ \Omega_{\rho_{0}},\\ \left(H-\frac{1}{f}\right)p+\frac{\partial p}{\partial\nu}&=-\varphi&&\text{on}\ \Gamma_{\rho_{0}},\end{aligned}
(1.6b) p=fρ~onΓρ0,\displaystyle p=f\tilde{\rho}\quad\,\text{on}\ \Gamma_{\rho_{0}},

where H=HΓρ0H=H_{\Gamma_{\rho_{0}}} is the mean curvature of Γρ0\Gamma_{\rho_{0}} normalized such that H=n1H=n-1 for Ω=𝔹\Omega=\mathbb{B}. The linearized equation (1.6) is derived by substituting a solution pair (Ωρ0+ερ~,uρ0+ερ~)(\Omega_{\rho_{0}+\varepsilon\tilde{\rho}},u_{\rho_{0}+\varepsilon\tilde{\rho}}) with formal expansions

Γρ0+ερ~\displaystyle\Gamma_{\rho_{0}+\varepsilon\tilde{\rho}} ={ζ+(ρ0(ζ)+ερ~(ζ)ν(ζ))+o(ε)ζ𝕊},\displaystyle=\{\zeta+\left(\rho_{0}(\zeta)+\varepsilon\tilde{\rho}(\zeta)\nu(\zeta)\right)+o(\varepsilon)\mid\zeta\in\mathbb{S}\},
uρ0+ερ~\displaystyle u_{\rho_{0}+\varepsilon\tilde{\rho}} =uρ0+εp+o(ε)\displaystyle=u_{\rho_{0}}+\varepsilon p+o(\varepsilon)

into (1.1) for a right hand side f+εφf+\varepsilon\varphi, and equating functions of order ε\varepsilon. Note that (1.6) is a decoupled system for pp and ρ~\tilde{\rho}, and we may consider only (1.6a) for the solvability of (1.6). Then (1.6b) with known pp yields a solution ρ~\tilde{\rho}.

Recall that the implicit function theorem states that the nonlinear equation F(ρ,g)=0F(\rho,g)=0 has for each gg close to g0g_{0} a unique solution ρ\rho near ρ0\rho_{0} with F(ρ0,g0)=0F(\rho_{0},g_{0})=0, if

  1. (i)

    the mapping F:X×YZF\colon X\times Y\to Z is C1C^{1} in a neighbourhood of (ρ0,g0)(\rho_{0},g_{0}) and if

  2. (ii)

    the partial derivative ρF(ρ0,g0)(X,Z)\partial_{\rho}F(\rho_{0},g_{0})\in\mathscr{L}(X,Z) is bijective.

Here, XX, YY and ZZ are Banach spaces with XZX\subset Z. In addition to the solution ρ(g)\rho(g) being locally unique, the mapping gρ(g)Xg\mapsto\rho(g)\in X is in C1C^{1}. In the current setting, the Neumann boundary condition (1.1b) yields such a mapping FF, and the linearized equation ρF(ρ0,g0)[ρ~]=φ\partial_{\rho}F(\rho_{0},g_{0})[\tilde{\rho}]=\varphi is reflected by (1.6). However, the linearized equation (1.6) has a regularity defect called loss of derivatives, i.e. ρF(ρ0,g0)1(Z,X)\partial_{\rho}F(\rho_{0},g_{0})^{-1}\not\in\mathscr{L}(Z,X). Since solutions ρ~\tilde{\rho} to (1.6) are less regular than ρ0\rho_{0}, and hence the typical iterative scheme in the classical implicit function theorem fails.

One method to overcome this regularity issue is the Nash-Moser theorem, a generalization of the classical implicit function theorem introduced by Nash in [16] and generalized by Moser in [15]. The introduction of a smoothing operator combined with Newton’s method for improved convergence was there shown to be a mean to overcome the regularity deficit. For the Nash-Moser theorem to work,

  1. (i)

    regularity properties are required for F:Xi×YZiF\colon X_{i}\times Y\to Z_{i}, where (Xi,Zi)i(X_{i},Z_{i})_{i} is a family of pairs of Banach spaces such that XiXi1X_{i}\subset X_{i-1}, ZiZi1Z_{i}\subset Z_{i-1}. Furthermore,

  2. (ii)

    a (right) inverse of ρF(ρ,g)\partial_{\rho}F(\rho,g) has to exist for (ρ,g)(\rho,g) in a neighbourhood of (ρ0,g0)(\rho_{0},g_{0}).

In this setting, for every gg in a neighbourhood of g0g_{0}, the existence of ρ(g)\rho(g) in X0X_{0} is then given. Note that there are various versions of the Nash-Moser theorem, also referred to as Nash-Moser-Hörmander theorem. We refer as an example to the work of Baldi and Haus [3] and the references therein.

Instead of applying the Nash-Moser theorem, we introduce a new modified version of the classical implicit function theorem, which has the constraint that a loss of derivatives may take place except at the point (ρ0,g0)(\rho_{0},g_{0}). We require for a pair of Banach triplets X2X1X0X_{2}\subset X_{1}\subset X_{0} and Z2Z1Z0Z_{2}\subset Z_{1}\subset Z_{0} that

  1. (i)

    for j=1,2j=1,2, FF is continuous in a neighbourhood of (ρ0,g0)(\rho_{0},g_{0}) from Xj1×YX_{j-1}\times Y to Zj1Z_{j-1}, and that it is in C1C^{1} in a neighbourhood of (ρ0,g0)(\rho_{0},g_{0}) from Xj×YX_{j}\times Y to Zj1Z_{j-1}, For (ρ,g)(\rho,g) in a neighbourhood of (ρ0,g0)Xj×Y(\rho_{0},g_{0})\in X_{j}\times Y, we have ρF(ρ,g)(Zj1,Xj1)\partial_{\rho}F(\rho,g)\in\mathscr{L}(Z_{j-1},X_{j-1}). Further,

  2. (ii)

    F:Xj×YZjF\colon X_{j}\times Y\to Z_{j} is Fréchet-differentiable at (ρ0,g0)(\rho_{0},g_{0}) for j=1,2j=1,2 and ρF(ρ0,g0)(Xj,Zj)\partial_{\rho}F(\rho_{0},g_{0})\in\mathscr{L}(X_{j},Z_{j}) is invertible for j=0,1,2j=0,1,2.

The first point reflects the loss of regularity, the second point reflects that it does not occur at the point under consideration. Under these assumptions, we derive a modified implicit function theorem that yields local uniqueness of a solution ρ(g)X1\rho(g)\in X_{1} for all gg in a neighbourhood of g0g_{0}, and the mapping gρ(g)X0g\mapsto\rho(g)\in X_{0} is in C1C^{1}. Note that in the setting of (1.1) and (1.6), the loss of derivatives does indeed not occur in the case of the solution of (1.1) for constant ff, as then Γ\Gamma and uu are smooth.

However, a second obstacle apart from the loss of derivatives arises. Due to the translational invariance of (1.1), the linearized equation (1.6a) for g=0g=0 and Ωρ0=𝔹\Omega_{\rho_{0}}=\mathbb{B} is not solvable for arbitrary φh2+α(𝕊)\varphi\in h^{2+\alpha}(\mathbb{S}), and for φ=0\varphi=0 it has an nn-dimensional space of solutions. This implies that the partial derivative of FF at (0,0)(0,0) is not invertible, which is necessary also for the modified implicit function theorem. We will remove this degeneracy by imposing an additional condition

(1.7) Ωxj𝑑x=0(j=1,,n),\int_{\Omega}x_{j}\,dx=0\quad(j=1,\ldots,n),

so that the barycenter of Ω\Omega is fixed to be the origin, and by decomposing the space hk+α(𝕊)h^{k+\alpha}(\mathbb{S}) into hk+α(𝕊)=XlKh^{k+\alpha}(\mathbb{S})=X_{l}\oplus K, where

(1.8) Xl:={ρhl+α(𝕊)ρ,xjL2(𝕊)=0,j=1,,n},K:=span{x1,,xn}.\displaystyle\begin{aligned} X_{l}&:=\left\{\rho\in h^{l+\alpha}(\mathbb{S})\mid\langle\rho,x_{j}\rangle_{L^{2}(\mathbb{S})}=0,\quad j=1,\ldots,n\right\},\\ K&:=\text{span}\left\{x_{1},\ldots,x_{n}\right\}.\end{aligned}

This allows for a decomposition of a domain perturbation ρhl+α(𝕊)\rho\in h^{l+\alpha}(\mathbb{S}) into ρ1Xl\rho_{1}\in X_{l}, ρ2K\rho_{2}\in K, as well as a decomposition of the function gg in (1.2) likewise, and we examine F(ρ1+ρ2,g1+g2)=0F(\rho_{1}+\rho_{2},g_{1}+g_{2})=0.

With these preparations, we present the main result of this work.

Theorem 1.1.

There exist neighbourhoods of zeros

VX2,U2h2+α(𝕊)×K and U3h3+α(𝕊)×K,V\subset X_{2},\,U_{2}\subset h^{2+\alpha}(\mathbb{S})\times K\text{ and }U_{3}\subset h^{3+\alpha}(\mathbb{S})\times K,

such that for all g1Vg_{1}\in V there are unique (ρ,g2)=(ρ(g1),g2(g1))U3(\rho,g_{2})=(\rho(g_{1}),g_{2}(g_{1}))\in U_{3} such that the following holds.

  1. (i)

    Ωρ(g1)\Omega_{\rho(g_{1})} defined by (1.3) admits a solution uh3+α(Ωρ(g1)¯)u\in h^{3+\alpha}(\overline{\Omega_{\rho(g_{1})}}) to (1.1) for (1.2) with g=g1+g2(g1)g=g_{1}+g_{2}(g_{1}), and satisfies (1.7).

  2. (ii)

    Ωρ(g1)\Omega_{\rho(g_{1})} is locally unique up to translations in the sense that there is (ρ(g1),g2(g1))U3(\rho(g_{1}),g_{2}(g_{1}))\in U_{3} for g1Vg_{1}\in V, and if Ωρ\Omega_{\rho} with ρU3\rho\in U_{3} admits a solution uu to (1.1) for (1.2) with g=g2(g1)+g1g=g_{2}(g_{1})+g_{1} and satisfies (1.7), then ρ=ρ(g1)\rho=\rho(g_{1}).

  3. (iii)

    For the mapping (ρ,g2):VU3(\rho,g_{2})\colon V\to U_{3}, we have (ρ,g2)C1(V,U2)(\rho,g_{2})\in C^{1}(V,U_{2}) and the stability estimates

    (1.9) ρ(g1)h2+α(𝕊)\displaystyle\left\|\rho(g_{1})\right\|_{h^{2+\alpha}(\mathbb{S})} Cg1h2+α(𝕊),\displaystyle\leq C\left\|g_{1}\right\|_{h^{2+\alpha}(\mathbb{S})},
    g2(g1)h2+α(𝕊)\displaystyle\left\|g_{2}(g_{1})\right\|_{h^{2+\alpha}(\mathbb{S})} Cg1h2+α(𝕊)\displaystyle\leq C\left\|g_{1}\right\|_{h^{2+\alpha}(\mathbb{S})}

    hold.

While the existence of ρ\rho is guaranteed in h3+α(𝕊)h^{3+\alpha}(\mathbb{S}), only the weaker norm, i.e. the h2+α(𝕊)h^{2+\alpha}(\mathbb{S}) norm, of ρ\rho is estimated in (1.9). This is due to the fact that the linear stability estimate requires C1C^{1}-regularity of the mapping g1ρg_{1}\mapsto\rho, and this regularity is expected only when the image space is h2+α(𝕊)h^{2+\alpha}(\mathbb{S}) due to the loss of derivatives.

Remark 1.2.

The translational invariance of (1.1) is mirrored in that theorem by using the decomposition into the translational part and its orthogonal complement. In that regard, it also becomes clear why the setting of the little Hölder spaces hl+αh^{l+\alpha} instead of the Hölder spaces Cl+αC^{l+\alpha} is necessary: The decomposition into subspaces is induced by the so-called spherical harmonics on 𝕊\mathbb{S}. They are dense in hl+α(𝕊)h^{l+\alpha}(\mathbb{S}), but not in Cl+α(𝕊)C^{l+\alpha}(\mathbb{S}). This will be further discussed in Section 3.

The paper is organised as follows. In Section 2, we introduce the perturbed problem as well as derive in detail the formulation via the linearized equation (1.6). We motivate the application of an implicit function theorem to a mapping FF that is derived from the Neumann boundary condition. This application is obstructed by the degeneracy of the derivative of FF as well as the loss of derivatives. The degeneracy of the derivative of FF stemming from the inherent symmetry of (1.1) will be addressed in Section 3. There, also the decomposition for the little Hölder spaces is motivated as well as the necessity of using the setting of the little Hölder spaces. In Section 4, we will revisit the implicit function theorem and establish a modified version fitting our setting. This is then applied to the perturbed problem in Section 5 to prove Theorem 1.1.

2 Preliminaries

We formally set up the perturbed problem defined in (1.1). We want to know whether for a perturbation gg there exists an open bounded domain Ω=Ω(g)\Omega=\Omega(g) admitting a solution uΩu_{\Omega} to (1.1) with (1.2), i.e. f=1n+gf=\frac{1}{n}+g.

We restrict the domain Ω\Omega to be in such a way that it may be modelled as a deviation of 𝔹\mathbb{B}, the domain admitting a solution to (1.1) with f=1nf=\frac{1}{n}. For this reference domain Ω0=𝔹\Omega_{0}=\mathbb{B}, with Ω0=Γ0=𝕊\partial\Omega_{0}=\Gamma_{0}=\mathbb{S}, we define the perturbed domain Ωρ\Omega_{\rho} by its hm+αh^{m+\alpha}-boundary Γρ=Ωρ\Gamma_{\rho}=\partial\Omega_{\rho} in the following way. We set for mm\in\mathbb{N}

Uγ,m:={vhm+α(𝕊)|vhm+α(𝕊)<γ},U_{\gamma,m}:=\left\{v\in h^{m+\alpha}(\mathbb{S})\,\big{|}\,\left\|v\right\|_{h^{m+\alpha}(\mathbb{S})}<\gamma\right\},

with γ1\gamma\leq 1 sufficiently small. Next, we define

θ:𝕊×(1,)θ(𝕊×(1,)),θ(ζ,r):=ζ+rν0(ζ)=ζ+rζ.\theta:\,\mathbb{S}\times(-1,\infty)\to\theta\left(\mathbb{S}\times(-1,\infty)\right),\quad\theta(\zeta,r):=\zeta+r\nu_{0}(\zeta)=\zeta+r\zeta.

In general, νρ\nu_{\rho} denotes the outer unit normal vector of Γρ\Gamma_{\rho}; for ρ=0\rho=0 we have ν0(x)=x\nu_{0}(x)=x. Then we set

Γρ={ζ+ρ(ζ)ν0(ζ)n|ζΓ0}={ζ+ρ(ζ)ζn|ζ𝕊},\Gamma_{\rho}=\left\{\zeta+\rho(\zeta)\nu_{0}(\zeta)\in\mathbb{R}^{n}\,\big{|}\,\zeta\in\Gamma_{0}\right\}=\left\{\zeta+\rho(\zeta)\zeta\in\mathbb{R}^{n}\,\big{|}\,\zeta\in\mathbb{S}\right\},

and ρUγ,m\rho\in U_{\gamma,m}. ρ\rho models the velocity of the boundary, and will be used to measure how much Γρ\Gamma_{\rho} deviates from Γ0\Gamma_{0}. Using this, we define the diffeomorphism

θρ(x):={x+φ(|x|1)ρ(x|x|)x|x|,for x0,0for x=0,\theta_{\rho}(x):=\begin{cases}x+\varphi\left(|x|-1\right)\rho\left(\frac{x}{|x|}\right)\frac{x}{|x|},\quad&\text{for }x\neq 0,\\ 0\quad&\text{for }x=0,\end{cases}

from Ω0=𝔹\Omega_{0}=\mathbb{B} to Ωρ\Omega_{\rho}, where φ:\varphi\colon\mathbb{R}\to\mathbb{R} is a smooth cut-off function with 0φ(r)10\leq\varphi(r)\leq 1, φ(r)=1\varphi(r)=1 for |r|14|r|\leq\frac{1}{4} and φ(r)=0\varphi(r)=0 for |r|34|r|\geq\frac{3}{4}, as well as |dφdr(r)|4\left|\frac{d\varphi}{dr}(r)\right|\leq 4. The diffeomorphism θρ\theta_{\rho} induces pullback and pushforward operators

θρu\displaystyle\theta_{\rho}^{*}u :=uθρ,\displaystyle:=u\circ\theta_{\rho},\quad θρ:hk+α(Γρ)hk+α(𝕊),\displaystyle\theta_{\rho}^{*}\colon h^{k+\alpha}(\Gamma_{\rho})\to h^{k+\alpha}(\mathbb{S}),
θρv\displaystyle\theta_{*}^{\rho}v :=vθρ1,\displaystyle:=v\circ\theta_{\rho}^{-1},\quad θρ:hk+α(𝕊)hk+α(Γρ),\displaystyle\theta_{*}^{\rho}\colon h^{k+\alpha}(\mathbb{S})\to h^{k+\alpha}(\Gamma_{\rho}),

with k{0}k\in\mathbb{N}\cup\{0\}.

Our problem now becomes the following:

Problem 2.1.

For gh1+α(𝕊)g\in h^{1+\alpha}(\mathbb{S}), is there a ρ=ρ(g)Uγ,2\rho=\rho(g)\in U_{\gamma,2} such that Ωρ\Omega_{\rho} as defined above admits a solution uρu_{\rho} to (2.1)?

(2.1a) Δuρ=1in Ωρ,uρ=0on Γρ,\displaystyle\begin{aligned} -\Delta u_{\rho}&=1\quad&\text{in }\Omega_{\rho},\\ u_{\rho}&=0\phantom{+g\left(\frac{x}{|x|}\right)}\quad&\text{on }\Gamma_{\rho},\end{aligned}
(2.1b) uρνρ(x)=1n+g(x|x|)on Γρ.\displaystyle\frac{\partial u_{\rho}}{\partial\nu_{\rho}}(x)=\frac{1}{n}+g\left(\frac{x}{|x|}\right)\quad\text{on }\Gamma_{\rho}.

By elliptic regularity theory, we have, for given ρUγ,2\rho\in U_{\gamma,2}, the existence and uniqueness of a solution uρh2+α(Ω¯ρ)u_{\rho}\in h^{2+\alpha}(\overline{\Omega}_{\rho}) when only considering (2.1a). Therefore, for the examination of this problem, it is sufficient to focus on the perturbation, i.e. (2.1b).

We define FC(Uγ,2×h1+α(𝕊),h1+α(𝕊))F\in C(U_{\gamma,2}\times h^{1+\alpha}(\mathbb{S}),h^{1+\alpha}(\mathbb{S})) by

(2.2) F(ρ,g):=θρ(uρνρ)+1n+g,F(\rho,g):=\theta_{\rho}^{*}\left(\frac{\partial u_{\rho}}{\partial\nu_{\rho}}\right)+\frac{1}{n}+g,

where uρu_{\rho} is the unique solution of (2.1a). Then Ωρ\Omega_{\rho} admits a solution to (2.1) for given gh1+α(𝕊)g\in h^{1+\alpha}(\mathbb{S}) if and only if F(ρ,g)=0F(\rho,g)=0.

This structure tempts to use the implicit function theorem to arrive at solutions in a neighbourhood of (0,0)(0,0). However, we shall arrive at two obstacles. The first is the derivative ρF(0,0)\partial_{\rho}F(0,0) not being bijective due to the inherent translational invariance, an observation that will be treated in Section 3. The second is the loss of derivatives, a regularity issue of the ρ\rho-derivative of FF that will be discussed in the following.

In view of this, note that for gh2+α(𝕊)g\in h^{2+\alpha}(\mathbb{S}) and for m=2,3m=2,3, we have

(2.3) FC(Uγ,m×h2+α(𝕊),hm1+α(𝕊)).F\in C(U_{\gamma,m}\times h^{2+\alpha}(\mathbb{S}),h^{m-1+\alpha}(\mathbb{S})).

2.1 Derivative of FF

We turn to the ρ\rho-differentiability of F at a point (ρ0,g)(\rho_{0},g). Due to the loss of derivatives, we need to assume ρ0Uγ,3\rho_{0}\in U_{\gamma,3}. We consider

F(ρ0+ερ~,g)F(ρ0,g)=A(ρ0,g)[ερ~]+o(ε)F(\rho_{0}+\varepsilon\tilde{\rho},g)-F(\rho_{0},g)=A(\rho_{0},g)[\varepsilon\tilde{\rho}]+o(\varepsilon)

for ρ~Uγ,3\tilde{\rho}\in U_{\gamma,3} and ε0\varepsilon\to 0. Since uρu_{\rho} in F(ρ,g)F(\rho,g) lives on Ω¯ρ\overline{\Omega}_{\rho}, which varies for ρ\rho, we consider the following approach.

Let xΩ¯0x\in\overline{\Omega}_{0}. We define the mapping u(ρ,x):=uρ(θρ(x))u(\rho,x):=u_{\rho}(\theta_{\rho}(x)) and y=θρ0(x)Ω¯ρ0y=\theta_{\rho_{0}}(x)\in\overline{\Omega}_{\rho_{0}}, as well as z=θρ0+ερ~(x)Ω¯ρ0+ερ~z=\theta_{\rho_{0}+\varepsilon\tilde{\rho}}(x)\in\overline{\Omega}_{\rho_{0}+\varepsilon\tilde{\rho}}. One may show that u(ρ,θρ(x))u(\rho,\theta_{\rho}(x)) is differentiable with respect to ρ\rho, for the procedure see e.g. [12, Sect. 5.6]. Therefore, the following calculations are well-defined.

Using the Taylor expansion, we write

uρ0+ερ~(z)\displaystyle u_{\rho_{0}+\varepsilon\tilde{\rho}}(z) =u(ρ0+ερ~,θρ0+ερ~(x))\displaystyle=u(\rho_{0}+\varepsilon\tilde{\rho},\theta_{\rho_{0}+\varepsilon\tilde{\rho}}(x))
=u(ρ0,y)+ρu(ρ0,y)ρ~(θρ01(y)|θρ01(y)|)=:p(y)ε\displaystyle=u(\rho_{0},y)+\underbrace{\partial_{\rho}u(\rho_{0},y)\tilde{\rho}\left(\frac{\theta_{\rho_{0}}^{-1}(y)}{|\theta_{\rho_{0}}^{-1}(y)|}\right)}_{=:p(y)}\varepsilon
+yu(ρ0,y)Vρ0(y)ν0(θρ01(y)|θρ01(y)|)ρ~(θρ01(y)|θρ01(y)|)ε+o(ε)\displaystyle\quad+\partial_{y}u(\rho_{0},y)V_{\rho_{0}}(y)\nu_{0}\left(\frac{\theta_{\rho_{0}}^{-1}(y)}{|\theta_{\rho_{0}}^{-1}(y)|}\right)\tilde{\rho}\left(\frac{\theta_{\rho_{0}}^{-1}(y)}{|\theta_{\rho_{0}}^{-1}(y)|}\right)\varepsilon+o(\varepsilon)

with Vρ0(y):=φ(|θρ01(y)|1)V_{\rho_{0}}(y):=\varphi\left(\left|\theta_{\rho_{0}}^{-1}(y)\right|-1\right). The function pp is the so-called shape derivative of uρ0u_{\rho_{0}} with respect to the domain variation from Ωρ0\Omega_{\rho_{0}} to Ωρ0+ερ~\Omega_{\rho_{0}+\varepsilon\tilde{\rho}}.

Next, we reformulate problem (2.1) in terms of ρ~\tilde{\rho} and pp. Using the representation as before and considering the Dirichlet problem (2.1a) for uρ0+ερ~u_{\rho_{0}+\varepsilon\tilde{\rho}} as well as for uρ0u_{\rho_{0}}, and letting ε0\varepsilon\to 0, we get

(2.4) Δyp(y)=0in Ωρ0,p(y)=u(ρ0,y)νρ0(θρ0ρ~)(y)1|Nρ0|on Γρ0.\displaystyle\begin{aligned} \Delta_{y}p(y)&=0&\quad\text{in }\Omega_{\rho_{0}},\\ p(y)&=-\frac{\partial u(\rho_{0},y)}{\partial\nu_{\rho_{0}}}\left(\theta_{*}^{\rho_{0}}\tilde{\rho}\right)(y)\frac{1}{|\nabla N_{\rho_{0}}|}&\quad\text{on }\Gamma_{\rho_{0}}.\end{aligned}

Here, Nρ(x):=|x|1ρ(x|x|)N_{\rho}(x):=|x|-1-\rho(\frac{x}{|x|}) is defined for xθ(𝕊×(1,))x\in\theta(\mathbb{S}\times(-1,\infty)) and we note that Γρ\Gamma_{\rho} is the zero-level set of NρN_{\rho}. Therefore, for the outer unit normal vector field νρ0\nu_{\rho_{0}} at Γρ0\Gamma_{\rho_{0}}, we have νρ0(x)=Nρ0(x)|Nρ0(x)|\nu_{\rho_{0}}(x)=\frac{\nabla N_{\rho_{0}}(x)}{|\nabla N_{\rho_{0}}(x)|} and Nρ0(θρ0(ζ)).N0(ζ)=1\nabla N_{\rho_{0}}\left(\theta_{\rho_{0}}(\zeta)\right).\nabla N_{0}(\zeta)=1 for ζ𝕊\zeta\in\mathbb{S}. We also note that

θρ0(ν0)=(νρ0.ν0)νρ0+τρ0=1|Nρ0|νρ0+τρ0,\theta_{*}^{\rho_{0}}(\nu_{0})=(\nu_{\rho_{0}}.\nu_{0})\nu_{\rho_{0}}+\tau_{\rho_{0}}=\frac{1}{|\nabla N_{\rho_{0}}|}\nu_{\rho_{0}}+\tau_{\rho_{0}},

where τρ0\tau_{\rho_{0}} is a tangent vector field. Here, we used (νρ0.ν0)=1|Nρ0(θρ0(ζ))|(\nu_{\rho_{0}}.\nu_{0})=\frac{1}{\left|\nabla N_{\rho_{0}}\left(\theta_{\rho_{0}}(\zeta)\right)\right|}.

Remark 2.2.

The regularity of uρ0u_{\rho_{0}} and νρ0\nu_{\rho_{0}}, ρ0Uγ,3\rho_{0}\in U_{\gamma,3}, imposes restrictions on the regularity of pp and in general, we can only expect ph2+α(Γρ0)p\in h^{2+\alpha}(\Gamma_{\rho_{0}}). We have νρ0uρ0h2+α(Γρ0)\partial_{\nu_{\rho_{0}}}{u_{\rho_{0}}}\in h^{2+\alpha}(\Gamma_{\rho_{0}}) and for the mean curvature Hρ0h1+α(Γρ0)H_{\rho_{0}}\in h^{1+\alpha}(\Gamma_{\rho_{0}}), but in general, no more. If, however, we are in the setting for ρ0=0\rho_{0}=0, then we have ν0u0C(𝕊)\partial_{\nu_{0}}{u_{0}}\in C^{\infty}(\mathbb{S}), ν0u0=1n-\partial_{\nu_{0}}{u_{0}}=\frac{1}{n} and p=1nρ~p=\frac{1}{n}\tilde{\rho}, thus ph3+α(𝕊)p\in h^{3+\alpha}(\mathbb{S}), the same regularity as ρ~\tilde{\rho}.

Now we calculate the Fréchet-derivative of FF. With the notation as before, let x𝕊x\in\mathbb{S}, y=θρ0(x)Γρ0y=\theta_{\rho_{0}}(x)\in\Gamma_{\rho_{0}} and z=θρ0+ερ~(x)Γρ0+ρz=\theta_{\rho_{0}+\varepsilon\tilde{\rho}}(x)\in\Gamma_{\rho_{0}+\rho}. Note that in this case, Vρ0(y)=1V_{\rho_{0}}(y)=1 and θρ01(y)|θρ01(y)|=x\frac{\theta_{\rho_{0}}^{-1}(y)}{|\theta_{\rho_{0}}^{-1}(y)|}=x.

First note that there exists a tangent vector τ\tau at yΓρ0y\in\Gamma_{\rho_{0}} such that

νρ0+ερ~(z)=νρ0(y)+ετ(y)+o(ε).\nu_{\rho_{0}+\varepsilon\tilde{\rho}}(z)=\nu_{\rho_{0}}(y)+\varepsilon\tau(y)+o(\varepsilon).

Next, we calculate

ziuρ0+ερ~(z)\displaystyle\partial_{z_{i}}u_{\rho_{0}+\varepsilon\tilde{\rho}}(z) =ziu(ρ0+ερ~,θρ0+ερ~(x))\displaystyle=\partial_{z_{i}}u({\rho_{0}+\varepsilon\tilde{\rho}},\theta_{\rho_{0}+\varepsilon\tilde{\rho}}(x))
=yiu(ρ0,y)+εyip(y)+εyiyk(u(ρ0,y))θρ0(ν0ρ~)(y)+o(ε).\displaystyle=\frac{\partial}{\partial y_{i}}u(\rho_{0},y)+\varepsilon\frac{\partial}{\partial y_{i}}p(y)+\varepsilon\frac{\partial}{\partial y_{i}}\frac{\partial}{\partial y_{k}}\left(u(\rho_{0},y)\right)\theta_{*}^{\rho_{0}}\left(\nu_{0}\tilde{\rho}\right)(y)+o(\varepsilon).

This implies

zuρ0+ερ~(z).νρ0+ερ~(z)=ziuρ0+ερ~(z)νρ0+ερ~i(z)\displaystyle\quad\nabla_{z}u_{\rho_{0}+\varepsilon\tilde{\rho}}(z).\nu_{\rho_{0}+\varepsilon\tilde{\rho}}(z)=\partial_{z_{i}}u_{\rho_{0}+\varepsilon\tilde{\rho}}(z)\nu_{\rho_{0}+\varepsilon\tilde{\rho}}^{i}(z)
=[yiu(ρ0,y)+εyip(y)+εyiyk(u(ρ0,y))θρ0(ν0ρ~)(y)][νρ0(y)+ετ(y)]i+o(ε)\displaystyle=\left[\frac{\partial}{\partial y_{i}}u(\rho_{0},y)+\varepsilon\frac{\partial}{\partial y_{i}}p(y)+\varepsilon\frac{\partial}{\partial y_{i}}\frac{\partial}{\partial y_{k}}\left(u(\rho_{0},y)\right)\theta_{*}^{\rho_{0}}\left(\nu_{0}\tilde{\rho}\right)(y)\right]\left[\nu_{\rho_{0}}(y)+\varepsilon\tau(y)\right]^{i}+o(\varepsilon)
=yiu(ρ0,y)νρ0i(y)+εyiu(ρ0,y)τρ0i(y)+εyip(y)νρ0i(y)\displaystyle=\frac{\partial}{\partial y_{i}}u(\rho_{0},y)\nu_{\rho_{0}}^{i}(y)+\varepsilon\frac{\partial}{\partial y_{i}}u(\rho_{0},y)\tau_{\rho_{0}}^{i}(y)+\varepsilon\frac{\partial}{\partial y_{i}}p(y)\nu_{\rho_{0}}^{i}(y)
+εyiyk(u(ρ0,y))θρ0(ρ~)(y)[1|Nρ0|νρ0+τρ0]νρ0i(y)+o(ε)\displaystyle\quad+\varepsilon\frac{\partial}{\partial y_{i}}\frac{\partial}{\partial y_{k}}\left(u(\rho_{0},y)\right)\theta_{*}^{\rho_{0}}\left(\tilde{\rho}\right)(y)\left[\frac{1}{|\nabla N_{\rho_{0}}|}\nu_{\rho_{0}}+\tau_{\rho_{0}}\right]\nu_{\rho_{0}}^{i}(y)+o(\varepsilon)
=νρ0u(ρ0,y)+ενρ0p(y)+ε2νρ02u(ρ0,y)θρ0(ρ~)(y)1|Nρ0|\displaystyle=\frac{\partial}{\partial\nu_{\rho_{0}}}u(\rho_{0},y)+\varepsilon\frac{\partial}{\partial\nu_{\rho_{0}}}p(y)+\varepsilon\frac{\partial^{2}}{\partial\nu_{\rho_{0}}^{2}}u(\rho_{0},y)\theta_{*}^{\rho_{0}}\left(\tilde{\rho}\right)(y)\frac{1}{|\nabla N_{\rho_{0}}|}
+ετρ0νρ0u(ρ0,y)θρ0(ρ~)(y)+o(ε)\displaystyle\quad+\varepsilon\frac{\partial}{\partial\tau_{\rho_{0}}}\frac{\partial}{\partial\nu_{\rho_{0}}}u(\rho_{0},y)\theta_{*}^{\rho_{0}}\left(\tilde{\rho}\right)(y)+o(\varepsilon)
=νρ0u(ρ0,y)+ενρ0p(y)+ε(1Hρ0νρ0u(ρ0,y))θρ0(ρ~)(y)1|Nρ0|\displaystyle=\frac{\partial}{\partial\nu_{\rho_{0}}}u(\rho_{0},y)+\varepsilon\frac{\partial}{\partial\nu_{\rho_{0}}}p(y)+\varepsilon\left(-1-H_{\rho_{0}}\frac{\partial}{\partial\nu_{\rho_{0}}}u(\rho_{0},y)\right)\theta_{*}^{\rho_{0}}\left(\tilde{\rho}\right)(y)\frac{1}{|\nabla N_{\rho_{0}}|}
+ετρ0νρ0u(ρ0,y)θρ0(ρ~)(y)+o(ε),\displaystyle\quad+\varepsilon\frac{\partial}{\partial\tau_{\rho_{0}}}\frac{\partial}{\partial\nu_{\rho_{0}}}u(\rho_{0},y)\theta_{*}^{\rho_{0}}\left(\tilde{\rho}\right)(y)+o(\varepsilon),

where in the last step we used the identity Δuρ0=ΔΓρ0uρ0+νρ02uρ0+Hρ0νρ0uρ0,\Delta u_{\rho_{0}}=\Delta_{\Gamma_{\rho_{0}}}u_{\rho_{0}}+\partial^{2}_{\nu_{\rho_{0}}}u_{\rho_{0}}+H_{\rho_{0}}\partial_{\nu_{\rho_{0}}}u_{\rho_{0}}, with ΔΓρ0\Delta_{\Gamma_{\rho_{0}}} being the Laplace-Beltrami operator and Hρ0H_{\rho_{0}} the mean curvature on Γρ0\Gamma_{\rho_{0}}. Using the Dirichlet boundary condition for pp in (2.4), we arrive at

F(ρ0+ερ~,g)\displaystyle F(\rho_{0}+\varepsilon\tilde{\rho},g) =F(ρ0,g)+εθρ0[νρ0p+Hρ0pθρ0ρ~|Nρ0|+τρ0νρ0u(ρ0,y)θρ0ρ~]+o(ε).\displaystyle=F(\rho_{0},g)+\varepsilon\theta_{\rho_{0}}^{*}\left[\frac{\partial}{\partial\nu_{\rho_{0}}}p+H_{\rho_{0}}p-\frac{\theta_{*}^{\rho_{0}}\tilde{\rho}}{|\nabla N_{\rho_{0}}|}+\frac{\partial}{\partial\tau_{\rho_{0}}}\frac{\partial}{\partial\nu_{\rho_{0}}}u(\rho_{0},y)\theta_{*}^{\rho_{0}}\tilde{\rho}\right]+o(\varepsilon).

We see that the term

θρ0[νρ0p+Hρ0pθρ0ρ~|Nρ0|+τρ0νρ0u(ρ0,y)θρ0ρ~]\theta_{\rho_{0}}^{*}\left[\frac{\partial}{\partial\nu_{\rho_{0}}}p+H_{\rho_{0}}p-\frac{\theta_{*}^{\rho_{0}}\tilde{\rho}}{|\nabla N_{\rho_{0}}|}+\frac{\partial}{\partial\tau_{\rho_{0}}}\frac{\partial}{\partial\nu_{\rho_{0}}}u(\rho_{0},y)\theta_{*}^{\rho_{0}}\tilde{\rho}\right]

is well-defined and lies in h1+α(𝕊)h^{1+\alpha}(\mathbb{S}) even when only assuming ρ~h2+α(𝕊)\tilde{\rho}\in h^{2+\alpha}(\mathbb{S}). Note that to verify this, one also needs to take into account the impact of the regularity assumption of ρ~\tilde{\rho} on the regularity of the solution pp of (2.4) as mentioned in Remark 2.5. This gives us the following lemma.

Lemma 2.3.

The mapping FF as defined in (2.2) satisfies

(2.5) FC(Uγ,2×h1+α(𝕊),h1+α(𝕊))C1(Uγ,3×h1+α(𝕊),h1+α(𝕊)).F\in C(U_{\gamma,2}\times h^{1+\alpha}(\mathbb{S}),h^{1+\alpha}(\mathbb{S}))\cap C^{1}(U_{\gamma,3}\times h^{1+\alpha}(\mathbb{S}),h^{1+\alpha}(\mathbb{S})).

Furthermore, we have the following.

  1. 1.

    The ρ\rho-Fréchet-derivative of FF at (ρ0,g)Uγ,3×h1+α(𝕊)(\rho_{0},g)\in U_{\gamma,3}\times h^{1+\alpha}(\mathbb{S}) is

    ρF(ρ0,g)[ρ~]=θρ[pνρ0+Hρ0pθρ0ρ~|Nρ0|+τρ0(uρ0νρ0)θρ0ρ~]h1+α(𝕊),\displaystyle\begin{aligned} &\partial_{\rho}F(\rho_{0},g)\left[\tilde{\rho}\right]\\ &=\theta_{\rho}^{*}\left[\frac{\partial p}{\partial\nu_{\rho_{0}}}+H_{\rho_{0}}p-\frac{\theta_{*}^{\rho_{0}}\tilde{\rho}}{|\nabla N_{\rho_{0}}|}+\frac{\partial}{\partial\tau_{\rho_{0}}}\left(\frac{\partial u_{\rho_{0}}}{\partial\nu_{\rho_{0}}}\right)\theta_{*}^{\rho_{0}}\tilde{\rho}\right]\in h^{1+\alpha}(\mathbb{S}),\end{aligned}

    where pp is a unique solution to (2.4), i.e.

    Δp=0in Ωρ0,p=uρ0νρ0θρ0ρ~1|Nρ0|on Γρ0.\displaystyle\begin{aligned} \Delta p&=0&\quad\text{in }\Omega_{\rho_{0}},\\ p&=-\frac{\partial u_{\rho_{0}}}{\partial\nu_{\rho_{0}}}\theta_{*}^{\rho_{0}}\tilde{\rho}\frac{1}{|\nabla N_{\rho_{0}}|}&\quad\text{on }\Gamma_{\rho_{0}}.\end{aligned}
  2. 2.

    We have

    ρF(0,0)\displaystyle\partial_{\rho}F(0,0) (h3+α(𝕊),h2+α(𝕊)) and\displaystyle\in\mathcal{L}(h^{3+\alpha}(\mathbb{S}),h^{2+\alpha}(\mathbb{S}))\text{ and }
    ρF(ρ,g)\displaystyle\partial_{\rho}F(\rho,g) (h3+α(𝕊),h1+α(𝕊)).\displaystyle\in\mathcal{L}(h^{3+\alpha}(\mathbb{S}),h^{1+\alpha}(\mathbb{S})).
  3. 3.

    The linear operator ρF(ρ,g)\partial_{\rho}F(\rho,g) has the extension

    ρF(ρ,g)(h2+α(𝕊),h1+α(𝕊)).\partial_{\rho}F(\rho,g)\in\mathcal{L}(h^{2+\alpha}(\mathbb{S}),h^{1+\alpha}(\mathbb{S})).

In view of (2.3), one may verify that for m=2,3m=2,3 and gh2+α(𝕊)g\in h^{2+\alpha}(\mathbb{S}), one has

FC(Uγ,m×h2+α(𝕊),hm1+α(𝕊))C1(Uγ,m+1×h2+α(𝕊),hm1+α(𝕊)).F\in C(U_{\gamma,m}\times h^{2+\alpha}(\mathbb{S}),h^{m-1+\alpha}(\mathbb{S}))\cap C^{1}(U_{\gamma,m+1}\times h^{2+\alpha}(\mathbb{S}),h^{m-1+\alpha}(\mathbb{S})).
Remark 2.4.

We have the following characterisation of bijectivity of the ρ\rho-Fréchet-deriva-tive of FF at a point (ρ,g)(\rho,g) with F(ρ,g)=0F(\rho,g)=0, i.e. when Ωρ\Omega_{\rho} is a solution to (2.1) for gh1+α(𝕊)g\in h^{1+\alpha}(\mathbb{S}):

The extended operator ρF(ρ,g)\partial_{\rho}F(\rho,g), with

ρFC(Uγ,3×h1+α(𝕊),(h2+α(𝕊),h1+α(𝕊))),\partial_{\rho}F\in C\left(U_{\gamma,3}\times{h^{1+\alpha}(\mathbb{S})},\mathcal{L}\left(h^{2+\alpha}(\mathbb{S}),h^{1+\alpha}(\mathbb{S})\right)\right),

has the bounded inverse

ρF(ρ,g)1(h1+α(𝕊),h2+α(𝕊))\partial_{\rho}F(\rho,g)^{-1}\in\mathcal{L}\left(h^{1+\alpha}(\mathbb{S}),h^{2+\alpha}(\mathbb{S})\right)

if and only if the boundary problem

(2.6) Δp=0 in Ωρ(Hρ11n+θρg)p+pνρ=φ on Γρ.\displaystyle\begin{aligned} -\Delta p&=0&\text{ in }\Omega_{\rho}\\ \left(H_{\rho}-\frac{1}{\frac{1}{n}+\theta_{*}^{\rho}g}\right)p+\frac{\partial p}{\partial\nu_{\rho}}&=-\varphi&\text{ on }\Gamma_{\rho}.\end{aligned}

is uniquely solvable for any φh1+α(Γρ)\varphi\in h^{1+\alpha}(\Gamma_{\rho}). Unique solvability of (2.6) is given provided that (Hρ11n+θρg)>0,\left(H_{\rho}-\frac{1}{\frac{1}{n}+\theta_{*}^{\rho}g}\right)>0, see [11, Thm. 6.31], in which case we would have ph2+α(Ω¯ρ)Cφh1+α(Γρ)\left\|p\right\|_{h^{2+\alpha}(\overline{\Omega}_{\rho})}\leq C\left\|\varphi\right\|_{h^{1+\alpha}(\Gamma_{\rho})}. This does not hold in the current setting. Thus, we have to examine the bijectivity in a different manner.

3 Degeneracy of ρF\partial_{\rho}F

3.1 Non-Bijectivity of the partial derivative of FF

We examine the ρ\rho-derivative of FF at (ρ,g)=(0,0)(\rho,g)=(0,0), as we merely require the existence of an inverse of ρF(0,0)\partial_{\rho}F(0,0) to use the modified implicit function theorem, Theorem 4.2. We have

ρF(0,0)[ρ~]=1nρ~+1n𝒩ρ~,for ρ~h3+α(𝕊).\partial_{\rho}F(0,0)[\tilde{\rho}]=-\frac{1}{n}\tilde{\rho}+\frac{1}{n}\mathscr{N}\tilde{\rho},\quad\text{for }\tilde{\rho}\in h^{3+\alpha}(\mathbb{S}).

Here, 𝒩\mathscr{N} denotes the Dirichlet-to-Neumann operator on the sphere 𝕊\mathbb{S}.

Definition 3.1.

Let φhl+α(𝕊)\varphi\in h^{l+\alpha}(\mathbb{S}), l2l\geq 2 arbitrary. The Dirichlet-to-Neumann operator on the sphere 𝒩:hl+α(𝕊)hl1+α(𝕊)\mathscr{N}\colon h^{l+\alpha}(\mathbb{S})\to h^{l-1+\alpha}(\mathbb{S}) is defined as

𝒩φ:=νu, with u the unique solution of {Δu=0in 𝔹,u=φon 𝕊.\mathscr{N}\varphi:=\partial_{\nu}u,\text{ with }u\text{ the unique solution of }\begin{cases}-\Delta u&=0\quad\text{in }\mathbb{B},\\ \hfil u&=\varphi\quad\text{on }\mathbb{S}.\end{cases}

We see that

𝒩(hl+α(𝕊),hl1+α(𝕊)),l.\mathscr{N}\in\mathscr{L}\left(h^{l+\alpha}(\mathbb{S}),h^{l-1+\alpha}(\mathbb{S})\right),\,l\in\mathbb{N}.

It suffices to test bijectivity of ρF(0,0)\partial_{\rho}F(0,0) for ρ~Hk\tilde{\rho}\in H_{k}, k{0}k\in\mathbb{N}\cup\{0\}, where

Hk=span{hk,j|j=1,,dkn},dkn<,H_{k}=\text{span}\left\{h_{k,j}\,\big{|}\,j=1,\ldots,d_{k}^{n}\right\},\quad d_{k}^{n}<\infty,

is the set of harmonic homogeneous polynomials on the unit sphere of degree kk. Indeed, the hk,jh_{k,j}, k{0},j=1,,dknk\in\mathbb{N}\cup\{0\},\,j=1,\ldots,d_{k}^{n}, form an orthonormal basis of L2(𝕊)L^{2}(\mathbb{S}), see e.g. [10, Thm. 2.53]. One may show that

:=span{hk,j|k{0},j=1,,dkn}\mathscr{H}:=\text{span}\left\{h_{k,j}\,\big{|}\,k\in\mathbb{N}\cup\{0\},\,j=1,\ldots,d_{k}^{n}\right\}

is dense in hl+α(𝕊)h^{l+\alpha}(\mathbb{S}), which is why it is sufficient to consider ρ~\tilde{\rho}\in\mathscr{H}.

Therefore, let ρ~=hk,j\tilde{\rho}=h_{k,j} be a harmonic homogeneous polynomial on the unit sphere of order k{0}k\in\mathbb{N}\cup\{0\}, with j{1,,dkn}j\in\{1,\ldots,d_{k}^{n}\}. We get

ρF(0,0)[hk,j]=1nhk,j+knhk,j=k1nhk,j{=0if k=1,0else.\partial_{\rho}F(0,0)[h_{k,j}]=-\frac{1}{n}h_{k,j}+\frac{k}{n}h_{k,j}=\frac{k-1}{n}h_{k,j}\begin{cases}=0&\text{if }k=1,\\ \neq 0&\text{else}.\end{cases}

This shows that ρF(0,0)\partial_{\rho}F(0,0) is not bijective and that its kernel is

(3.1) ker(ρF(0,0))=span{h1,j|j=1,,n},\text{ker}\left(\partial_{\rho}F(0,0)\right)=\text{span}\left\{h_{1,j}\,\big{|}\,j=1,\ldots,n\right\},

with dim(ker(ρF(0,0)))=n<\dim(\text{ker}(\partial_{\rho}F(0,0)))=n<\infty. Furthermore, we see that the range of ρF(0,0)\partial_{\rho}F(0,0) is

range(ρF(0,0))=span{hk,j|k2{0},j=1,,dkn}¯h2+α(𝕊).\text{range}\left(\partial_{\rho}F(0,0)\right)=\overline{\text{span}\left\{h_{k,j}\,\big{|}\,k\in\mathbb{N}_{\geq 2}\cup\{0\},j=1,\ldots,d_{k}^{n}\right\}}^{\left\|\cdot\right\|_{h^{2+\alpha}(\mathbb{S})}}.
Notation 3.2.

In view of the calculations to come, we define

Xl=span{hk,j|k2{0},j=1,,dkn}¯hl+α(𝕊), for l.X_{l}=\overline{\text{span}\left\{h_{k,j}\,\big{|}\,k\in\mathbb{N}_{\geq 2}\cup\{0\},j=1,\ldots,d_{k}^{n}\right\}}^{\left\|\cdot\right\|_{h^{l+\alpha}(\mathbb{S})}},\text{ for }l\in\mathbb{N}.

Note that this is equivalent to defining

Xl:={ρhl+α(𝕊)|ρ,h1,jL2(𝕊)=0,j=1,,n},X_{l}:=\left\{\rho\in h^{l+\alpha}(\mathbb{S})\,\big{|}\,\langle{\rho},h_{1,j}\rangle_{L^{2}(\mathbb{S})}=0,\,j=1,\ldots,n\right\},

as in (1.8). To confirm this, also note that h1,j(x)=ωn12xjh_{1,j}(x)=\omega_{n}^{-\frac{1}{2}}x_{j}, j=1,,nj=1,\ldots,n. Since XlX_{l} is a finite-codimensional subspace of hl+α(𝕊)h^{l+\alpha}(\mathbb{S}), we have XlXl=hl+α(𝕊)X_{l}\oplus X_{l}^{\bot}=h^{l+\alpha}(\mathbb{S}), where XlX_{l}^{\bot} denotes the orthogonal complement of XlX_{l}. Because dim(Xl)=n<\dim(X_{l}^{\bot})=n<\infty for all ll\in\mathbb{N}, we do not need to differentiate between those spaces depending on ll, as we have XlnX_{l}^{\bot}\cong\mathbb{R}^{n}. Therefore, we may define

K:=span{h1,j|j=1,,n}K:=\text{span}\left\{h_{1,j}\,\big{|}\,j=1,\ldots,n\right\}

and get XlK=hl+α(𝕊)X_{l}\oplus K=h^{l+\alpha}(\mathbb{S}). Finally, we denote

𝒰l:={ρUγ,l|ρ,h1,jL2(𝕊)=0,j=1,,n},\mathcal{U}_{l}:=\left\{\rho\in U_{\gamma,l}\,\big{|}\,\langle{\rho},{h_{1,j}}\rangle_{L^{2}(\mathbb{S})}=0,\,j=1,\ldots,n\right\},

as well as 𝒰l\mathcal{U}_{l}^{\bot} for the subset of KK such that Uγ,l=𝒰l𝒰lU_{\gamma,l}=\mathcal{U}_{l}\oplus\mathcal{U}_{l}^{\bot}.

Remark 3.3.

That the kernel of ρF(0,0)\partial_{\rho}F(0,0) is non-trivial is not surprising, in fact it is an obvious property resulting from the translational invariance of (2.1) with g=0g=0. The overdetermined problem is in this case solvable for any translated sphere 𝔹+c:=B1(c)\mathbb{B}+c:=B_{1}(c), with cnc\in\mathbb{R}^{n}, and with solution uc(x)=12n(1|xc|2)u_{c}(x)=\frac{1}{2n}\left(1-|x-c|^{2}\right).

To show the connection to the kernel of ρF(0,0)\partial_{\rho}F(0,0), we find ρUγ,3\rho\in U_{\gamma,3} such that Γρ=(𝔹+c)\Gamma_{\rho}=\partial\left(\mathbb{B}+c\right). With the ansatz x+ρ(x)x=y+c,x,y𝕊,x+\rho(x)x=y+c,\ x,y\in\mathbb{S}, we arrive at ρ(x)=ρ(c,x)=x.c1±1|c|2+(x.c)2\rho(x)=\rho(c,x)=x.c-1\pm\sqrt{1-|c|^{2}+(x.c)^{2}} for any |c|1|c|\leq 1. For c=tejc=te_{j}, we then arrive at

ddtρ(tej,x)|t=0=[xj+121t2+t2xj2(2t+2txj2)]t=0=xj.\frac{d}{dt}\rho(te_{j},x)\big{|}_{t=0}=\left[x_{j}+\frac{1}{2\sqrt{1-t^{2}+t^{2}x_{j}^{2}}}\left(-2t+2tx_{j}^{2}\right)\right]_{t=0}=x_{j}.

Now, because of the translational invariance of (2.1), we have F(ρ(c,x),0)=F(0,0)=0F(\rho(c,x),0)=F(0,0)=0, which implies

ρF(0,0)[ddt(ρ(tej,))|t=0]=0\partial_{\rho}F(0,0)\left[\frac{d}{dt}\left(\rho(te_{j},\cdot)\right)\big{|}_{t=0}\right]=0

and we arrive at ρF(0,0)[xj]=0\partial\rho F(0,0)[x_{j}]=0 for j{1,,n}j\in\{1,\ldots,n\}. This coincides with (3.1).

3.2 Re-formulation to eliminate degeneracy

To eliminate the problem of non-bijectivity, i.e. the degeneracy of the problem (2.6), we need to eliminate the translation invariance in the original problem (2.1). Therefore, we replace FF as defined in (2.2) by a mapping

GC(𝒰2×𝒰2×X1×K,X1×n×K),G\in C(\mathcal{U}_{2}\times\mathcal{U}_{2}^{\bot}\times X_{1}\times K,X_{1}\times\mathbb{R}^{n}\times K),
G(ρ1,ρ2,g1,g2):=(G0G1GnGn+1)(ρ1,ρ2,g1,g2):=(P2F(ρ1+ρ2,g1+g2)Ωρ1+ρ2x1dxΩρ1+ρ2xndx(IdP2)F(ρ1+ρ2,g1+g2)).\displaystyle\begin{aligned} G&(\rho_{1},\rho_{2},g_{1},g_{2}):=\begin{pmatrix}G_{0}\\ G_{1}\\ \vdots\\ G_{n}\\ G_{n+1}\end{pmatrix}(\rho_{1},\rho_{2},g_{1},g_{2}):=\begin{pmatrix}P_{2}F(\rho_{1}+\rho_{2},g_{1}+g_{2})\\ \int_{\Omega_{\rho_{1}+\rho_{2}}}x_{1}\,\mathrm{d}x\\ \vdots\\ \int_{\Omega_{\rho_{1}+\rho_{2}}}x_{n}\,\mathrm{d}x\\ (\text{Id}-P_{2})F(\rho_{1}+\rho_{2},g_{1}+g_{2})\end{pmatrix}.\end{aligned}

Pl:hl+α(𝕊)XlP_{l}\colon h^{l+\alpha}(\mathbb{S})\to X_{l} denotes the projection onto XlX_{l}. If it is clear which ll\in\mathbb{N} is to be used, we write PP for PlP_{l}.

Note that for m=2,3m=2,3, and g1X2g_{1}\in X_{2}, GG is also well-defined and we have

GC(𝒰m×𝒰m×X2×K,Xm1×n×K).\displaystyle G\in C\left(\mathcal{U}_{m}\times\mathcal{U}_{m}^{\bot}\times X_{2}\times K,X_{m-1}\times\mathbb{R}^{n}\times K\right).

By the condition Ωρxidx=0\int_{\Omega_{\rho}}x_{i}\,\mathrm{d}x=0 for i=1,,ni=1,\ldots,n, we achieve that the center of mass of Ωρ\Omega_{\rho} is in the origin and thus eliminate the possibility of translations, and thus, admissible ρ\rho will be in the set

={ρUγ,3|Ωρxidx=0,i=1,,n}.\mathcal{M}=\left\{\rho\in U_{\gamma,3}\,\big{|}\,\int_{\Omega_{\rho}}x_{i}\,\mathrm{d}x=0,\,i=1,\ldots,n\right\}.

As a direct consequence, we have

Lemma 3.4.

Ωρ\Omega_{\rho} with barycenter zero admits a solution to (2.1) for given gh1+α(𝕊)g\in h^{1+\alpha}(\mathbb{S}) if and only if G(ρ1,ρ2,g1,g2)=0G(\rho_{1},\rho_{2},g_{1},g_{2})=0, with ρ=ρ1+ρ2\rho=\rho_{1}+\rho_{2} and g=g1+g2g=g_{1}+g_{2}.

3.3 Bijectivity of the partial derivative of GG

The mapping GG has the following regularity properties.

Lemma 3.5.

GG is Fréchet-differentiable as a map from 𝒰3×𝒰3×K\mathcal{U}_{3}\times\mathcal{U}_{3}^{\bot}\times K to X1×n×KX_{1}\times\mathbb{R}^{n}\times K. We have

ρ1,ρ2,g2G(ρ1,ρ2,g1,g2)(X3×K×K,X1×n×K),\partial_{\rho_{1},\rho_{2},g_{2}}G(\rho_{1},\rho_{2},g_{1},g_{2})\in\mathscr{L}(X_{3}\times K\times K,X_{1}\times\mathbb{R}^{n}\times K),

for (ρ1+ρ2,g1+g2)Uγ,3×h1+α(𝕊)(\rho_{1}+\rho_{2},g_{1}+g_{2})\in U_{\gamma,3}\times h^{1+\alpha}(\mathbb{S}), which can be extended to

ρ1,ρ2,g2G(ρ1,ρ2,g1,g2)(X2×K×K,X1×n×K).\partial_{\rho_{1},\rho_{2},g_{2}}G(\rho_{1},\rho_{2},g_{1},g_{2})\in\mathscr{L}(X_{2}\times K\times K,X_{1}\times\mathbb{R}^{n}\times K).

Furthermore,

G\displaystyle G\in\ C(𝒰2×𝒰2×X1×K,X1×n×K)\displaystyle C(\mathcal{U}_{2}\times\mathcal{U}_{2}^{\bot}\times X_{1}\times K,X_{1}\times\mathbb{R}^{n}\times K)
C1(𝒰3×𝒰3×X1×K,X1×n×K).\displaystyle\cap\ C^{1}(\mathcal{U}_{3}\times\mathcal{U}_{3}^{\bot}\times X_{1}\times K,X_{1}\times\mathbb{R}^{n}\times K).

In view of the application of the modified function theorem introduced in Section 4, we also need the following observation concerning the regularity of GG. For m=2,3m=2,3, we have

(3.2) G\displaystyle G\in C(𝒰m×𝒰m×X2×K,Xm1×n×K)\displaystyle C\left(\mathcal{U}_{m}\times\mathcal{U}_{m}^{\bot}\times X_{2}\times K,X_{m-1}\times\mathbb{R}^{n}\times K\right)
C1(𝒰m+1×𝒰m+1×X2×K,Xm1×n×K).\displaystyle\cap\ C^{1}\left(\mathcal{U}_{m+1}\times\mathcal{U}_{m+1}^{\bot}\times X_{2}\times K,X_{m-1}\times\mathbb{R}^{n}\times K\right).

as well as for the extension of the partial derivative

(3.3) ρ1,ρ2,g2G(ρ1,ρ2,g1,g2)(Xm×K×K,Xm1×n×K),\partial_{\rho_{1},\rho_{2},g_{2}}G(\rho_{1},\rho_{2},g_{1},g_{2})\in\mathscr{L}(X_{m}\times{K\times K},X_{m-1}\times\mathbb{R}^{n}\times K),

where (ρ1,ρ2,g1,g2)Xm+1×K×X2×K(\rho_{1},\rho_{2},g_{1},g_{2})\in X_{m+1}\times K\times X_{2}\times K.

Proof.

We have for i=1,2i=1,2 and ρ~1X3h3+α(𝕊)\tilde{\rho}_{1}\in X_{3}\subset h^{3+\alpha}(\mathbb{S}), ρ~2Kh3+α(𝕊)\tilde{\rho}_{2}\in K\subset h^{3+\alpha}(\mathbb{S})

ρiG0(ρ1,ρ2,g1,g2)[ρ~i]\displaystyle\partial_{\rho_{i}}G_{0}(\rho_{1},\rho_{2},g_{1},g_{2})[\tilde{\rho}_{i}] =ρiPF(ρ1+ρ2,g1+g2)[ρ~i]\displaystyle=\partial_{\rho_{i}}PF(\rho_{1}+\rho_{2},g_{1}+g_{2})[\tilde{\rho}_{i}] X1,\displaystyle\in X_{1},
ρiGn+1(ρ1,ρ2,g1,g2)[ρ~i]\displaystyle\partial_{\rho_{i}}G_{n+1}(\rho_{1},\rho_{2},g_{1},g_{2})[\tilde{\rho}_{i}] =ρi(IdP)F(ρ1+ρ2,g1+g2)[ρ~i]\displaystyle=\partial_{\rho_{i}}(\text{Id}-P)F(\rho_{1}+\rho_{2},g_{1}+g_{2})[\tilde{\rho}_{i}] K\displaystyle\in K

and further for j=1,,nj=1,\ldots,n

ρiGj(ρ1,ρ2,g1,g2)[ρ~i]=Γρ1+ρ21|Nρ|θρρ~iσjdσ,for j=1,,n.\partial_{\rho_{i}}G_{j}(\rho_{1},\rho_{2},g_{1},g_{2})[\tilde{\rho}_{i}]=\int_{\Gamma_{\rho_{1}+\rho_{2}}}\frac{1}{|\nabla N_{\rho}|}\theta_{*}^{\rho}\tilde{\rho}_{i}\cdot\sigma_{j}\,\mathrm{d}\sigma,\quad\text{for }j=1,\ldots,n.

These expressions are still well-defined and of the same regularity for ρ~h2+α(𝕊)\tilde{\rho}\in h^{2+\alpha}(\mathbb{S}), implying the existence of an extension of ρ1,ρ2G(ρ1,ρ2,g1,g2)\partial_{\rho_{1},\rho_{2}}G(\rho_{1},\rho_{2},g_{1},g_{2}) onto X2×KX_{2}\times K and thus

ρ1,ρ2G(ρ1,ρ2,g1,g2)(X3×K,X1×n×K)(X2×K,X1×n×K).\partial_{\rho_{1},\rho_{2}}G(\rho_{1},\rho_{2},g_{1},g_{2})\in\mathscr{L}(X_{3}\times K,X_{1}\times\mathbb{R}^{n}\times K)\cap\mathscr{L}(X_{2}\times K,X_{1}\times\mathbb{R}^{n}\times K).

For the g2g_{2}-partial derivative and g2~K\tilde{g_{2}}\in K, we get

g2G0(ρ1,ρ2,g1,g2)[g2~]=g2PF(ρ1+ρ2,g1+g2)[g2~],\displaystyle\partial_{g_{2}}G_{0}(\rho_{1},\rho_{2},g_{1},g_{2})[\tilde{g_{2}}]=\partial_{g_{2}}PF(\rho_{1}+\rho_{2},g_{1}+g_{2})[\tilde{g_{2}}],
g2Gj(ρ1,ρ2,g1,g2)[g2~]=0and\displaystyle\partial_{g_{2}}G_{j}(\rho_{1},\rho_{2},g_{1},g_{2})[\tilde{g_{2}}]=0\quad\text{and}
g2Gn+1(ρ1,ρ2,g1,g2)[g2~]=g2(IdP)F(ρ1+ρ2,g1+g2)[g2~].\displaystyle\partial_{g_{2}}G_{n+1}(\rho_{1},\rho_{2},g_{1},g_{2})[\tilde{g_{2}}]=\partial_{g_{2}}(\text{Id}-P)F(\rho_{1}+\rho_{2},g_{1}+g_{2})[\tilde{g_{2}}].

This implies g2G(ρ1,ρ2,g1,g2)(K,X1×n×K)\partial_{g_{2}}G(\rho_{1},\rho_{2},g_{1},g_{2})\in\mathscr{L}(K,X_{1}\times\mathbb{R}^{n}\times K). ∎

At zero, the partial derivative ρ1,ρ2,g2G\partial_{\rho_{1},\rho_{2},g_{2}}G is bijective. We abbreviate (0,0,0,0)(0,0,0,0) by (0)(0).

Lemma 3.6.

GG is Fréchet-differentiable at (ρ1,ρ2,g1,g2)=(0)(\rho_{1},\rho_{2},g_{1},g_{2})=(0), and we have

ρ1,ρ2,g2G(0)(X3×K×K,X2×n×K)\displaystyle\partial_{\rho_{1},\rho_{2},g_{2}}G(0)\in\mathscr{L}(X_{3}\times K\times K,X_{2}\times\mathbb{R}^{n}\times K)

with

(3.4) ρ1,ρ2,g2G(0)[ρ~1,ρ~2,g~2]=(1n(𝒩Id)ρ~1ωn1/2ρ~2g~2).\displaystyle\begin{aligned} \partial_{\rho_{1},\rho_{2},g_{2}}G(0)[\tilde{\rho}_{1},\tilde{\rho}_{2},\tilde{g}_{2}]&=\begin{pmatrix}\frac{1}{n}\left(\mathscr{N}-\text{Id}\right)\tilde{\rho}_{1}\\ \omega_{n}^{1/2}\tilde{\rho}_{2}\\ \tilde{g}_{2}\end{pmatrix}.\end{aligned}

Indeed, for arbitrary mm\in\mathbb{N}, we have

ρ1,ρ2,g2G(0)(Xm+1×K×K,Xm×n×K).\partial_{\rho_{1},\rho_{2},g_{2}}G(0)\in\mathscr{L}(X_{m+1}\times K\times K,X_{m}\times\mathbb{R}^{n}\times K).

Further, ρ1,ρ2,g2G(0)\partial_{\rho_{1},\rho_{2},g_{2}}G(0) is invertible with

ρ1,ρ2,g2G(0)1(Xm×n×K,Xm+1×K×K)\partial_{\rho_{1},\rho_{2},g_{2}}G(0)^{-1}\in\mathscr{L}(X_{m}\times\mathbb{R}^{n}\times K,X_{m+1}\times K\times K)

for m,m,\in\mathbb{N}, and

(3.5) ρ1,ρ2,g2G(0)1[ϕ,α1,,αn,ψ]=(n((𝒩1)|X2)1ϕωn1/2j=1nαjh1,jψ)Xm+1×K×K\partial_{\rho_{1},\rho_{2},g_{2}}G(0)^{-1}[\phi,\alpha_{1},\ldots,\alpha_{n},\psi]=\begin{pmatrix}n\left(\left(\mathscr{N}-1\right)\big{|}_{X_{2}}\right)^{-1}\phi\\ \omega_{n}^{-1/2}\sum_{j=1}^{n}\alpha_{j}h_{1,j}\\ \psi\end{pmatrix}\in X_{m+1}\times K\times K

where ϕXm\phi\in X_{m}, αj\alpha_{j}\in\mathbb{R} for j=1,,nj=1,\ldots,n and ψK\psi\in K.

Proof.

Let mm\in\mathbb{N}. Let (ρ~1,ρ~2,g~2)Xm+1×K×K(\tilde{\rho}_{1},\tilde{\rho}_{2},\tilde{g}_{2})\in X_{m+1}\times K\times K and j=1,,nj=1,\ldots,n. We calculate

ρ1G0(0)[ρ~1]=ρ1PF(0)[ρ~1]=1n(𝒩Id)ρ~1,\displaystyle\partial_{\rho_{1}}G_{0}(0)[\tilde{\rho}_{1}]=\partial_{\rho_{1}}PF(0)[\tilde{\rho}_{1}]=\frac{1}{n}\left(\mathscr{N}-\text{Id}\right)\tilde{\rho}_{1},
ρ2G0(0)[ρ~2]=ρ2PF(0)[ρ~2]=0,\displaystyle\partial_{\rho_{2}}G_{0}(0)[\tilde{\rho}_{2}]=\partial_{\rho_{2}}PF(0)[\tilde{\rho}_{2}]=0,
ρ1Gj(0)[ρ~1]=𝕊ρ~1σjdσ=ωn1/2𝕊ρ~1h1,jdσ=0,\displaystyle\partial_{\rho_{1}}G_{j}(0)[\tilde{\rho}_{1}]=\int_{\mathbb{S}}\tilde{\rho}_{1}\sigma_{j}\,\mathrm{d}\sigma=\omega_{n}^{1/2}\int_{\mathbb{S}}\tilde{\rho}_{1}h_{1,j}\,\mathrm{d}\sigma=0,
ρ2Gj(0)[ρ~2]=𝕊ρ~2σjdσ=ωn1/2𝕊ρ~2h1,jdσ=ωn1/2(ρ~2)j,\displaystyle\partial_{\rho_{2}}G_{j}(0)[\tilde{\rho}_{2}]=\int_{\mathbb{S}}\tilde{\rho}_{2}\sigma_{j}\,\mathrm{d}\sigma=\omega_{n}^{1/2}\int_{\mathbb{S}}\tilde{\rho}_{2}h_{1,j}\,\mathrm{d}\sigma=\omega_{n}^{1/2}\left(\tilde{\rho}_{2}\right)_{j},
ρiGn+1(0)[ρ~i]=ρi(IdP)F(0)[ρ~i]=0,for i=1,2.\displaystyle\partial_{\rho_{i}}G_{n+1}(0)[\tilde{\rho}_{i}]=\partial_{\rho_{i}}(\text{Id}-P)F(0)[\tilde{\rho}_{i}]=0,\quad\text{for }i=1,2.

𝒩\mathscr{N} again denotes the Dirichlet-to-Neumann operator. Recall that the linear operator (𝒩Id)\left(\mathscr{N}-\text{Id}\right) is bijective as an operator in (Xm+1,Xm)\mathscr{L}(X_{m+1},X_{m}). For the g2g_{2}-partial derivative, we find

g2G0(0)[g2~]=g2PF(0)[g2~]=0,\displaystyle\partial_{g_{2}}G_{0}(0)[\tilde{g_{2}}]=\partial_{g_{2}}PF(0)[\tilde{g_{2}}]=0,
g2Gj(0)[g2~]=0and\displaystyle\partial_{g_{2}}G_{j}(0)[\tilde{g_{2}}]=0\quad\text{and}
g2Gn+1(0)[g2~]=g2(IdP)F(0)[g2~]=g~2.\displaystyle\partial_{g_{2}}G_{n+1}(0)[\tilde{g_{2}}]=\partial_{g_{2}}(\text{Id}-P)F(0)[\tilde{g_{2}}]=\tilde{g}_{2}.

This implies (3.4). We directly arrive at (3.5) and also at the regularity properties of ρ1,ρ2,g2G(0)1\partial_{\rho_{1},\rho_{2},g_{2}}G(0)^{-1}. ∎

4 A modified implicit function theorem

To arrive at the existence, uniqueness and at a stability result for Problem 2.1, we introduce a modified version of the implicit function theorem, Theorem 4.2. Because of the regularity issues stated in Remark 2.5, we are not able to apply the classical implicit function theorem. In preparation, we need

Theorem 4.1.

Assume the following.

  1. (I)

    Let 𝒳0,𝒳1,𝒴,𝒵0,𝒵1\mathcal{X}_{0},\mathcal{X}_{1},\mathcal{Y},\mathcal{Z}_{0},\mathcal{Z}_{1} be Banach spaces with 𝒳1𝒳0\mathcal{X}_{1}\hookrightarrow\mathcal{X}_{0} and 𝒵1𝒵0\mathcal{Z}_{1}\hookrightarrow\mathcal{Z}_{0}. Let D1D0D_{1}\subset D_{0} be open sets such that (0,0)Dj𝒳j×𝒴(0,0)\in D_{j}\subset\mathcal{X}_{j}\times\mathcal{Y} for j=0,1j=0,1.

  2. (II)

    Let FC1(D1,𝒵0)C(D0,𝒵0)F\in C^{1}(D_{1},\mathcal{Z}_{0})\cap C(D_{0},\mathcal{Z}_{0}) with F(0,0)=0F(0,0)=0 and xFC(D1,(𝒳0,𝒵0))\partial_{x}F\in C(D_{1},\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0})), which is to be understood such that for (x,y)D1(x,y)\in D_{1}, the partial deriative xF(x,y)(𝒳1,𝒵0)\partial_{x}F(x,y)\in\mathscr{L}(\mathcal{X}_{1},\mathcal{Z}_{0}) can be extended to xF¯(x,y)(𝒳0,𝒵0)\overline{\partial_{x}F}(x,y)\in\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0}) and xF¯C(D1,(𝒳0,𝒵0))\overline{\partial_{x}F}\in C(D_{1},\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0})).

  3. (III)

    We have F:D1𝒵1F\colon D_{1}\to\mathcal{Z}_{1} and FF is Fréchet-differentiable at (0,0)(0,0), hence xF(0,0)(𝒳1,𝒵1)\partial_{x}F(0,0)\in\mathscr{L}(\mathcal{X}_{1},\mathcal{Z}_{1}) and yF(0,0)(𝒴,𝒵1)\partial_{y}F(0,0)\in\mathscr{L}(\mathcal{Y},\mathcal{Z}_{1}).

  4. (IV)

    The inverse xF(0,0)1(𝒵1,𝒳1)(𝒵0,𝒳0)\partial_{x}F(0,0)^{-1}\in\mathscr{L}(\mathcal{Z}_{1},\mathcal{X}_{1})\cap\mathscr{L}(\mathcal{Z}_{0},\mathcal{X}_{0}) exists.

Then there exist neighbourhoods of zero 0U0𝒳00\in U_{0}\subset\mathcal{X}_{0}, 0U1𝒳10\in U_{1}\subset\mathcal{X}_{1} and 0V𝒴0\in V\subset\mathcal{Y}, as well as a function u:VU0u\colon V\to U_{0} such that

  1. (i)

    F(u(y),y)=0F(u(y),y)=0 for all yVy\in V, u(0)=0u(0)=0, and

  2. (ii)

    for x1,x2U1x_{1},x_{2}\in U_{1}, yVy\in V such that F(xi,y)=0F(x_{i},y)=0 for i=1,2i=1,2, we have x1=x2x_{1}=x_{2}.

Proof.

Let ε,δ>0\varepsilon,\delta>0 – we will redefine both later – and define

U1:={x𝒳1|x𝒳1ε},\displaystyle U_{1}:=\left\{x\in\mathcal{X}_{1}\,\big{|}\,\left\|x\right\|_{\mathcal{X}_{1}}\leq\varepsilon\right\},
U0:={x𝒳0|x𝒳0Cε},\displaystyle U_{0}:=\left\{x\in\mathcal{X}_{0}\,\big{|}\,\left\|x\right\|_{\mathcal{X}_{0}}\leq C\varepsilon\right\},
V:={y𝒴|y𝒴<δ},\displaystyle V:=\left\{y\in\mathcal{Y}\,\big{|}\,\left\|y\right\|_{\mathcal{Y}}<\delta\right\},

with C>0C>0 a constant satisfying x𝒳0Cx𝒳1\left\|x\right\|_{\mathcal{X}_{0}}\leq C\left\|x\right\|_{\mathcal{X}_{1}}, thus U1U0U_{1}\subset U_{0}.

Step 1: Show that for all yVy\in V, the function

Φy(x):=xxF(0,0)1F(x,y)=xF(0,0)1(xF(0,0)xF(x,y))\Phi_{y}(x):=x-\partial_{x}F(0,0)^{-1}F(x,y)=\partial_{x}F(0,0)^{-1}\left(\partial_{x}F(0,0)x-F(x,y)\right)

is a contraction mapping from (U1,𝒳0)\left(U_{1},\left\|\cdot\right\|_{\mathcal{X}_{0}}\right) to itself, provided that ε,δ>0\varepsilon,\delta>0 are sufficiently small.

As the fundamental theorem of calculus holds on Banach spaces as well, we have for FC1(D1,𝒵0)F\in C^{1}(D_{1},\mathcal{Z}_{0}) and for all x1,x2D1x_{1},x_{2}\in D_{1}

(4.1) F(x1,y)F(x2,y)=01xF(x2+t(x1x2),y)(x1x2)dt.F(x_{1},y)-F(x_{2},y)=\int_{0}^{1}\partial_{x}F(x_{2}+t(x_{1}-x_{2}),y)(x_{1}-x_{2})\,\mathrm{d}t.

Note that xF(x2+t(x1x2),y)(𝒳1,𝒵0)\partial_{x}F(x_{2}+t(x_{1}-x_{2}),y)\in\mathscr{L}(\mathcal{X}_{1},\mathcal{Z}_{0}) with extension in (𝒳0,𝒵0)\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0}).

Now let x1,x2U1x_{1},x_{2}\in U_{1}, yVy\in V. Then (xj,y)D1(x_{j},y)\in D_{1} for j=1,2j=1,2, and we use (4.1) to arrive at

Φy(x1)Φy(x2)=xF(0,0)1[01(xF(0,0)xF(x2+t(x1x2),y))dt(x1x2)].\Phi_{y}(x_{1})-\Phi_{y}(x_{2})=\partial_{x}F(0,0)^{-1}\left[\int_{0}^{1}\left(\partial_{x}F(0,0)-\partial_{x}F(x_{2}+t(x_{1}-x_{2}),y)\right)\,\mathrm{d}t(x_{1}-x_{2})\right].

By choosing ε,δ>0\varepsilon,\delta>0 smaller, if necessary, we get

(4.2) Φy(x1)Φy(x2)𝒳0xF(0,0)1(𝒵0,𝒳0)x1x2𝒳0sup0t1xF(0,0)xF(x2+t(x1x2),y)(𝒳0,𝒵0)12x1x2𝒳0,\displaystyle\begin{aligned} \left\|\Phi_{y}(x_{1})-\Phi_{y}(x_{2})\right\|_{\mathcal{X}_{0}}&\leq\left\|\partial_{x}F(0,0)^{-1}\right\|_{\mathscr{L}(\mathcal{Z}_{0},\mathcal{X}_{0})}\left\|x_{1}-x_{2}\right\|_{\mathcal{X}_{0}}\\ &\quad\cdot\sup_{0\leq t\leq 1}\left\|\partial_{x}F(0,0)-\partial_{x}F(x_{2}+t(x_{1}-x_{2}),y)\right\|_{\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0})}\\ &\leq\frac{1}{2}\left\|x_{1}-x_{2}\right\|_{\mathcal{X}_{0}},\end{aligned}

where the second inequality holds because of the condition xFC(D1,(𝒳0,𝒵0))\partial_{x}F\in C(D_{1},\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0})) in assumption (II), which for sufficiently small ε,δ>0\varepsilon,\delta>0 implies

sup0t1xF(0,0)xF(x2+t(x1x2),y)(𝒳0,𝒵0)<<1.\sup_{0\leq t\leq 1}\left\|\partial_{x}F(0,0)-\partial_{x}F(x_{2}+t(x_{1}-x_{2}),y)\right\|_{\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0})}<<1.

Next, we show that Φy(x)U1\Phi_{y}(x)\in U_{1} for (x,y)D1(x,y)\in D_{1}. We estimate Φy(x)𝒳1\left\|\Phi_{y}(x)\right\|_{\mathcal{X}_{1}} for (x,y)D1(x,y)\in D_{1}: Choosing δ=δ(ε)>0\delta=\delta(\varepsilon)>0 smaller, if necessary, we obtain

Φy(x)𝒳1\displaystyle\left\|\Phi_{y}(x)\right\|_{\mathcal{X}_{1}} xF(0,0)1(𝒵1,𝒳1)xF(0,0)xF(x,y)𝒵1\displaystyle\leq\left\|\partial_{x}F(0,0)^{-1}\right\|_{\mathscr{L}(\mathcal{Z}_{1},\mathcal{X}_{1})}\left\|\partial_{x}F(0,0)x-F(x,y)\right\|_{\mathcal{Z}_{1}}
xF(0,0)1(𝒵1,𝒳1)yF(0,0)(𝒴,𝒵1)δ+o(ε+δ(ε))\displaystyle\leq\left\|\partial_{x}F(0,0)^{-1}\right\|_{\mathscr{L}(\mathcal{Z}_{1},\mathcal{X}_{1})}\left\|\partial_{y}F(0,0)\right\|_{\mathscr{L}(\mathcal{Y},\mathcal{Z}_{1})}\delta+o(\varepsilon+\delta(\varepsilon))
ε.\displaystyle\leq\varepsilon.

Step 2: Construct a mapping u:U0Vu\colon U_{0}\to V.

Let yVy\in V arbitrary but fixed. The inductively defined sequence (xj)j0\left(x_{j}\right)_{j\in\mathbb{N}_{0}} with x0:=0x_{0}:=0, xj+1:=Φy(xj)U1U0x_{j+1}:=\Phi_{y}(x_{j})\in U_{1}\subset U_{0} for jj\in\mathbb{N} is a Cauchy sequence and thus converges in 𝒳0\left\|\cdot\right\|_{\mathcal{X}_{0}} to some xU0x_{\infty}\in U_{0}. Because FC(D0,𝒵0)F\in C(D_{0},\mathcal{Z}_{0}), this implies

F(x,y)𝒵0=limjF(xj,y)𝒵0limjxF(0,0)(𝒳0,𝒵0)xjxj+1𝒳0=0,\left\|F(x_{\infty},y)\right\|_{\mathcal{Z}_{0}}=\lim_{j\to\infty}\left\|F(x_{j},y)\right\|_{\mathcal{Z}_{0}}\leq\lim_{j\to\infty}\left\|\partial_{x}F(0,0)\right\|_{\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0})}\left\|x_{j}-x_{j+1}\right\|_{\mathcal{X}_{0}}=0,

where we used F(xj,y)=xF(0,0)(xjxj+1)F(x_{j},y)=\partial_{x}F(0,0)(x_{j}-x_{j+1}) for j0j\in\mathbb{N}_{0} by definition of Φy(x)\Phi_{y}(x). We set u(y):=xU0u(y):=x_{\infty}\in U_{0} for yVy\in V.

Step 3: Show (ii) of the theorem.

If x1,x2U1x_{1},x_{2}\in U_{1} and yVy\in V with F(xj,y)=0F(x_{j},y)=0 for j=1,2j=1,2, then

x1x2𝒳0=Φy(x1)Φy(x2)𝒳0(4.2)12x1x2𝒳0,\left\|x_{1}-x_{2}\right\|_{\mathcal{X}_{0}}=\left\|\Phi_{y}(x_{1})-\Phi_{y}(x_{2})\right\|_{\mathcal{X}_{0}}\overset{\eqref{eq:contraction}}{\leq}\frac{1}{2}\left\|x_{1}-x_{2}\right\|_{\mathcal{X}_{0}},

and therefore x1=x2x_{1}=x_{2}. ∎

Theorem 4.2.

Assume the following.

  1. (I)

    Consider Banach spaces 𝒳2𝒳1𝒳0\mathcal{X}_{2}\hookrightarrow\mathcal{X}_{1}\hookrightarrow\mathcal{X}_{0}, 𝒵2𝒵1𝒵0\mathcal{Z}_{2}\hookrightarrow\mathcal{Z}_{1}\hookrightarrow\mathcal{Z}_{0} and 𝒴\mathcal{Y}. Let D2D1D0D_{2}\subset D_{1}\subset D_{0} be open sets such that (0,0)Dj𝒳j×𝒴(0,0)\in D_{j}\subset\mathcal{X}_{j}\times\mathcal{Y} for j=0,1,2j=0,1,2.

  2. (II)

    For j=1,2j=1,2, let FC1(Dj,𝒵j1)C(Dj1,𝒵j1)F\in C^{1}(D_{j},\mathcal{Z}_{j-1})\cap C(D_{j-1},\mathcal{Z}_{j-1}) with F(0,0)=0F(0,0)=0 and further xFC(Dj,(𝒳j1,𝒵j1))\partial_{x}F\in C(D_{j},\mathscr{L}(\mathcal{X}_{j-1},\mathcal{Z}_{j-1})). This is to be understood such that for (x,y)Dj(x,y)\in D_{j}, the partial deriative xF(x,y)(𝒳j,𝒵j1)\partial_{x}F(x,y)\in\mathscr{L}(\mathcal{X}_{j},\mathcal{Z}_{j-1}) can be extended to xF¯(x,y)(𝒳j1,𝒵j1)\overline{\partial_{x}F}(x,y)\in\mathscr{L}(\mathcal{X}_{j-1},\mathcal{Z}_{j-1}) and xF¯C(Dj,(𝒳j1,𝒵j1))\overline{\partial_{x}F}\in C(D_{j},\mathscr{L}(\mathcal{X}_{j-1},\mathcal{Z}_{j-1})).

  3. (III)

    For j=1,2j=1,2, the mapping F:Dj𝒵jF\colon D_{j}\to\mathcal{Z}_{j} is Fréchet-differentiable at (0,0)(0,0).

  4. (IV)

    For j=0,1,2j=0,1,2, the inverse xF(0,0)1(𝒵j,𝒳j)\partial_{x}F(0,0)^{-1}\in\mathscr{L}(\mathcal{Z}_{j},\mathcal{X}_{j}) exists.

Then there exist neighbourhoods of zero 0U0𝒳00\in U_{0}\subset\mathcal{X}_{0}, 0U1𝒳10\in U_{1}\subset\mathcal{X}_{1} and 0V𝒴0\in V\subset\mathcal{Y} such that there is a function u:VU1u\colon V\to U_{1} satisfying

  1. (i)

    F(u(y),y)=0F(u(y),y)=0 for all yVy\in V, u(0)=0u(0)=0,

  2. (ii)

    if xU1x\in U_{1}, yVy\in V such that F(x,y)=0F(x,y)=0, then x=u(y)x=u(y), and

  3. (iii)

    it holds uC1(V,𝒳0)u\in C^{1}(V,\mathcal{X}_{0}), and

    u(y)=xF(u(y),y)1yF(u(y),y),u^{\prime}(y)=-\partial_{x}F(u(y),y)^{-1}\partial_{y}F(u(y),y),

    with xF(u(y),y)1(𝒵0,𝒳0)\partial_{x}F(u(y),y)^{-1}\in\mathscr{L}(\mathcal{Z}_{0},\mathcal{X}_{0}) and yF(u(y),y)(𝒴,𝒵0)\partial_{y}F(u(y),y)\in\mathscr{L}(\mathcal{Y},\mathcal{Z}_{0}).

Proof.

Step 1: Existence and uniqueness of uu.

Applying Theorem 4.1 twice, we arrive at the existence of neighbourhoods 0V𝒴0\in V\subset\mathcal{Y}, 0Uj𝒳j0\in U_{j}\subset\mathcal{X}_{j}, j=0,1j=0,1, and at the existence of a mapping u:VU1u\colon V\to U_{1} such that

  • \circ

    F(u(y),y)=0F(u(y),y)=0 for yVy\in V, u(0)=0u(0)=0, and

  • \circ

    for x1,x2U1x_{1},x_{2}\in U_{1}, yVy\in V with F(xi,y)=0F(x_{i},y)=0, i=1,2i=1,2, we have x1=x2x_{1}=x_{2}.

Thus, for x1U1x_{1}\in U_{1} and yVy\in V such that F(x,y)=0F(x,y)=0, we have x=u(y)x=u(y). This shows (i) and (ii)(ii) of Theorem 4.2.

Step 2: Show Lipschitz-continuity of uu, i.e. uC0,1(V,U0)u\in C^{0,1}(V,U_{0}).

Consider y1,y2Vy_{1},y_{2}\in V. Then with uu as above, i.e. u(yi)U1u(y_{i})\in U_{1} for i=1,2i=1,2, we have

u(y1)u(y2)𝒳0\displaystyle\left\|u(y_{1})-u(y_{2})\right\|_{\mathcal{X}_{0}} =Φy1(u(y1))Φy2(u(y2))𝒳0\displaystyle=\left\|\Phi_{y_{1}}(u(y_{1}))-\Phi_{y_{2}}(u(y_{2}))\right\|_{\mathcal{X}_{0}}
Φy1(u(y1))Φy1(u(y2))X0+Φy1(u(y2))Φy2(u(y2))𝒳0\displaystyle\leq\left\|\Phi_{y_{1}}(u(y_{1}))-\Phi_{y_{1}}(u(y_{2}))\right\|_{X_{0}}+\left\|\Phi_{y_{1}}(u(y_{2}))-\Phi_{y_{2}}(u(y_{2}))\right\|_{\mathcal{X}_{0}}
(4.2)12u(y1u(y2))𝒳0\displaystyle\overset{\eqref{eq:contraction}}{\leq}\frac{1}{2}\left\|u(y_{1}-u(y_{2}))\right\|_{\mathcal{X}_{0}}
+xF(0,0)1(𝒵0,𝒳0)F(u(y2),y1)F(u(y2),y2)𝒵0.\displaystyle\quad+\left\|\partial_{x}F(0,0)^{-1}\right\|_{\mathscr{L}(\mathcal{Z}_{0},\mathcal{X}_{0})}\left\|F(u(y_{2}),y_{1})-F(u(y_{2}),y_{2})\right\|_{\mathcal{Z}_{0}}.

This implies

u(y1)u(y2)𝒳0\displaystyle\left\|u(y_{1})-u(y_{2})\right\|_{\mathcal{X}_{0}} 2xF(0,0)1(𝒵0,𝒳0)\displaystyle\leq 2\left\|\partial_{x}F(0,0)^{-1}\right\|_{\mathscr{L}(\mathcal{Z}_{0},\mathcal{X}_{0})}
sup0t1yF(u(y2),y2+t(y1y2))(𝒴,𝒵0)y1y2𝒴\displaystyle\quad\cdot\sup_{0\leq t\leq 1}\left\|\partial_{y}F(u(y_{2}),y_{2}+t(y_{1}-y_{2}))\right\|_{\mathscr{L}(\mathcal{Y},\mathcal{Z}_{0})}\left\|y_{1}-y_{2}\right\|_{\mathcal{Y}}
Ly1y2𝒴,\displaystyle\leq L\left\|y_{1}-y_{2}\right\|_{\mathcal{Y}},

because the sup\sup-term is uniformly bounded in VV for sufficiently small ε,δ>0\varepsilon,\delta>0 defining the sets as in the proof of Theorem 4.1, since yF\partial_{y}F is continuous around (0,0)(0,0).

Step 3: Show uC1(V,𝒳0)u\in C^{1}(V,\mathcal{X}_{0}).

We have for y,hVy,h\in V s.th. y+hVy+h\in V

0\displaystyle 0 =F(u(y+h),y+h)F(u(y),y)\displaystyle=F(u(y+h),y+h)-F(u(y),y)
=F(u(y+h),y+h)F(u(y+h),y)+F(u(y+h),y)F(u(y),y)\displaystyle=F(u(y+h),y+h)-F(u(y+h),y)+F(u(y+h),y)-F(u(y),y)
=yF(u(y+h),y)[h]+oZ0(h𝒴)+F(u(y+h),y)F(u(y),y)\displaystyle=\partial_{y}F(u(y+h),y)[h]+o_{Z_{0}}(\left\|h\right\|_{\mathcal{Y}})+F(u(y+h),y)-F(u(y),y)
=yF(u(y),y)[h]+(yF(u(y+h),y)+yF(u(y),y))[h]+oZ0(h𝒴)\displaystyle=\partial_{y}F(u(y),y)[h]+\left(\partial_{y}F(u(y+h),y)+\partial_{y}F(u(y),y)\right)[h]+o_{Z_{0}}(\left\|h\right\|_{\mathcal{Y}})
+F(u(y+h),y)F(u(y),y)\displaystyle\quad+F(u(y+h),y)-F(u(y),y)
=yF(u(y),y)[h]+oZ0(h𝒴)\displaystyle=\partial_{y}F(u(y),y)[h]+o_{Z_{0}}(\left\|h\right\|_{\mathcal{Y}})
+01xF(u(y)+t(u(y+h)u(y)),y)dt[u(y+h)u(y)],\displaystyle\quad+\int_{0}^{1}\partial_{x}F(u(y)+t(u(y+h)-u(y)),y)\,\mathrm{d}t[u(y+h)-u(y)],

and further

01xF(u(y)+t(u(y+h)u(y)),y)dt[u(y+h)u(y)]\displaystyle\int_{0}^{1}\partial_{x}F(u(y)+t(u(y+h)-u(y)),y)\,\mathrm{d}t[u(y+h)-u(y)]
=xF(u(y),y)[u(y+h)u(y)]+o𝒵0(h𝒴).\displaystyle=\partial_{x}F(u(y),y)[u(y+h)-u(y)]+o_{\mathcal{Z}_{0}}(\left\|h\right\|_{\mathcal{Y}}).

Therefore,

u(y+h)u(y)=xF(u(y),y)1yF(u(y),y)[h]+o𝒳0(h𝒴),u(y+h)-u(y)=\partial_{x}F(u(y),y)^{-1}\partial_{y}F(u(y),y)[h]+o_{\mathcal{X}_{0}}(\left\|h\right\|_{\mathcal{Y}}),

yielding uC1(V,𝒳0)u\in C^{1}(V,\mathcal{X}_{0}) and (iii).

Note that xF(u(y),y)(𝒳0,𝒵0)\partial_{x}F(u(y),y)\in\mathscr{L}(\mathcal{X}_{0},\mathcal{Z}_{0}) is invertible for yVy\in V, since xF(0,0)\partial_{x}F(0,0) is invertible and u(y)𝒳1+y𝒴<<1\left\|u(y)\right\|_{\mathcal{X}_{1}}+\left\|y\right\|_{\mathcal{Y}}<<1. ∎

5 Proof of Theorem 1.1

With the tool of the modified implicit function theorem, Theorem 4.2, at hand, we are now able to prove Theorem 1.1, that is, the existence and uniqueness of admissible sets Ωρ\Omega_{\rho} with barycenter zero that solve the perturbed overdetermined problem (2.1), as well as a stability estimate.

Remark 5.1.

Considering the somewhat unintuitive partial derivative ρ1,ρ2,g2\partial_{\rho_{1},\rho_{2},g_{2}} in Lemma 3.6 was necessary to arrive at bijectivity and to be able to apply Theorem 4.2. The partial derivative ρG(0)\partial_{\rho}G(0) is not bijective.

In addition to that, keeping in mind the nature of the problem discussed in Section 3, the set Ωρ\Omega_{\rho} will only depend on the perturbations that do not induce a mere translation of the problem. ρ\rho depending on g1g_{1} (instead of gg) is a consequence of that setting.

Proof of Theorem 1.1.

We confirm the requirements for Theorem 4.2. For (I), we set

𝒳j\displaystyle\mathcal{X}_{j} =hj+2+α(𝕊)×X2=hj+2+α(𝕊)×K and\displaystyle=h^{j+2+\alpha}(\mathbb{S})\times X_{2}^{\bot}=h^{j+2+\alpha}(\mathbb{S})\times K\text{ and }
𝒵j\displaystyle\mathcal{Z}_{j} =Xj+1×n×X2=Xj+1×n×K\displaystyle=X_{j+1}\times\mathbb{R}^{n}\times X_{2}^{\bot}=X_{j+1}\times\mathbb{R}^{n}\times K

for j=0,1,2j=0,1,2, 𝒴=X2\mathcal{Y}=X_{2}, and DjD_{j} accordingly. By Lemma 3.5, (3.2) and (3.3), (II) is satisfied. Lemma 3.6 implies (III) and (IV).

Thus, Theorem 4.2 implies the existence of a neighbourhood 0VX20\in V\subset X_{2} as well as neighbourhoods 0Ujhj+α(𝕊)×X2=hj+α(𝕊)×K0\in U_{j}\subset h^{j+\alpha}(\mathbb{S})\times X_{2}^{\bot}=h^{j+\alpha}(\mathbb{S})\times K, j=2,3j=2,3, such that there is a function (ρ,g2):VU3(\rho,g_{2})\colon V\to U_{3} with G(ρ(g1),g1+g2(g1))=0G(\rho(g_{1}),g_{1}+g_{2}(g_{1}))=0 for all g1Vg_{1}\in V and (ρ,g2)(0)=0(\rho,g_{2})(0)=0.

Furthermore, (ρ,g2)(\rho,g_{2}) is unique in U3U_{3} and we have (ρ,g2)C1(V,U2)(\rho,g_{2})\in C^{1}(V,U_{2}). Differentiating G(ρ(g1),g1+g2(g1))=0G(\rho(g_{1}),g_{1}+g_{2}(g_{1}))=0 with respect to g1g_{1} and evaluating it at g1=0g_{1}=0 in direction g~1\tilde{g}_{1}, we get

0\displaystyle 0 =Dg1G(ρ1(g1),ρ2(g1),g1+g2(g1))|(0)[g~1]\displaystyle=D_{g_{1}}G\left(\rho_{1}(g_{1}),\rho_{2}(g_{1}),g_{1}+g_{2}(g_{1})\right)\big{|}_{(0)}[\tilde{g}_{1}]
=Dρ1,ρ2,g2G(0)[g1ρ1(0)[g~1],g1ρ2(0)[g~1],g1g2(0)[g~1]]+g1G(0)[g~1]\displaystyle=D_{\rho_{1},\rho_{2},g_{2}}G(0)\left[\partial_{g_{1}}\rho_{1}(0)[\tilde{g}_{1}],\partial_{g_{1}}\rho_{2}(0)[\tilde{g}_{1}],\partial_{g_{1}}g_{2}(0)[\tilde{g}_{1}]\right]+\partial_{g_{1}}G(0)[\tilde{g}_{1}]
=(ρ1F(0)[g1ρ1(0)[g~1]]𝕊σ1g1ρ2(0)[g~1]dσ𝕊σng1ρ2(0)[g~1]dσg1g2(0)[g~1])+(g~100).\displaystyle=\begin{pmatrix}\partial_{\rho_{1}}F(0)[\partial_{g_{1}}\rho_{1}(0)[\tilde{g}_{1}]]\\ \int_{\mathbb{S}}\sigma_{1}\partial_{g_{1}}\rho_{2}(0)[\tilde{g}_{1}]\,\mathrm{d}\sigma\\ \vdots\\ \int_{\mathbb{S}}\sigma_{n}\partial_{g_{1}}\rho_{2}(0)[\tilde{g}_{1}]\,\mathrm{d}\sigma\\ \partial_{g_{1}}g_{2}(0)[\tilde{g}_{1}]\end{pmatrix}+\begin{pmatrix}\tilde{g}_{1}\\ 0\\ 0\end{pmatrix}.

This yields

g1ρ1(0)[g~1]=ρ1F(0)1[g~1],g1ρ2(0)[g~1]=0, andg1g2(0)[g~1]=0,\displaystyle\begin{aligned} \partial_{g_{1}}\rho_{1}(0)[\tilde{g}_{1}]&=\partial_{\rho_{1}}F(0)^{-1}[\tilde{g}_{1}],\\ \partial_{g_{1}}\rho_{2}(0)[\tilde{g}_{1}]&=0,\text{ and}\\ \partial_{g_{1}}g_{2}(0)[\tilde{g}_{1}]&=0,\end{aligned}

where the last equation results from g1ρ2(0)[g~1]X2=K\partial_{g_{1}}\rho_{2}(0)[\tilde{g}_{1}]\in X_{2}^{\bot}=K, and we arrive at the stability estimates in (1.9). ∎

References

  • Aftalion et al. [1999] Amandine Aftalion, Jérôme Busca, and Wolfgang Reichel. Approximate radial symmetry for overdetermined boundary value problems. Adv. Differential Equations, 4(6):907–932, 1999.
  • Alexandrov [1962] Aleksandr D. Alexandrov. A characteristic property of spheres. Ann. Mat. Pura Appl., 58(1):303–315, 1962.
  • Baldi and Haus [2017] Pietro Baldi and Emanuele Haus. A Nash–Moser–Hörmander implicit function theorem with applications to control and Cauchy problems for PDEs. J. Funct. Anal., 273(12):3875–3900, 2017.
  • Bianchini et al. [2014] Chiara Bianchini, Antoine Henrot, and Paolo Salani. An overdetermined problem with non-constant boundary condition. Interfaces and Free Bound., 16(2):215–242, 2014.
  • Brandolini et al. [2008a] Barbara Brandolini, Carlo Nitsch, Paolo Salani, and Cristina Trombetti. Serrin-type overdetermined problems: an alternative proof. Arch. Ration. Mech. Anal., 190(2):267–280, 2008a.
  • Brandolini et al. [2008b] Barbara Brandolini, Carlo Nitsch, Paolo Salani, and Cristina Trombetti. On the stability of the Serrin problem. J. Diff. Equations, 245(6):1566–1583, 2008b.
  • Brock and Henrot [2002] Friedemann Brock and Antoine Henrot. A symmetry result for an overdetermined elliptic problem using continuous rearrangement and domain derivative. Rend. Circ. Mat. Palermo, 51(3):375–390, 2002.
  • Ciraolo et al. [2016] Giulio Ciraolo, Rolando Magnanini, and Vincenzo Vespri. Hölder stability for Serrin’s overdetermined problem. Ann. Mat. Pura Appl., 195(4):1333–1345, 2016.
  • Feldman [2018] William M. Feldman. Stability of Serrin’s problem and dynamic stability of a model for contact angle motion. SIAM J. Math. Anal., 50(3):3303–3326, 2018.
  • Folland [1995] Gerald B. Folland. Introduction to partial differential equations. Princeton University Press, 1995.
  • Gilbarg and Trudinger [2001] D. Gilbarg and N.S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. reprint of the 1998 edition.
  • Henrot and Pierre [2018] Antoine Henrot and Michel Pierre. Shape variation and optimization. European Mathematical Society, 2018.
  • Lunardi [2012] Alessandra Lunardi. Analytic semigroups and optimal regularity in parabolic problems. Springer Science & Business Media, 2012.
  • Magnanini and Poggesi [2020] Rolando Magnanini and Giorgio Poggesi. Nearly optimal stability for Serrin’s problem and the Soap Bubble theorem. Calc. Var. Partial Differential Equations, 59(1):35, 2020.
  • Moser [1961] Jürgen Moser. A new technique for the construction of solutions of nonlinear differential equations. Proc. Nat. Acad. Sci. USA, 47(11):1824, 1961.
  • Nash [1956] John Nash. The imbedding problem for Riemannian manifolds. Ann. of Math., pages 20–63, 1956.
  • Pólya [1948] George Pólya. Torsional rigidity, principal frequency, electrostatic capacity and symmetrization. Q. Appl. Math., 6(3):267–277, 1948.
  • Serrin [1971] James Serrin. A symmetry problem in potential theory. Arch. Ration. Mech. Anal., 43(4):304–318, 1971.
  • Weinberger [1971] Hans F. Weinberger. Remark on the preceding paper of Serrin. Arch. Ration. Mech. Anal., 43:319–320, 1971.