A Serrin-type problem with
partial knowledge of the domain
Abstract.
We present a quantitative estimate for the radially symmetric configuration concerning a Serrin-type overdetermined problem for the torsional rigidity in a bounded domain , when the equation is known on only, for some open subset .
The problem has concrete motivations in optimal heating with malfunctioning, laminar flows and beams with small inhomogeneities.
Key words and phrases:
Serrin’s overdetermined problem, torsional rigidity, integral identities, stability, quantitative estimates2010 Mathematics Subject Classification:
Primary 35N25, 53A10, 35B35; Secondary 35A231. Introduction
In this article we consider a variation of the classical Serrin’s overdetermined problem [Se] in which the equation is only known in a subset of the domain. We will provide quantitative results showing, roughly speaking, that when the part of the domain in which we do not have information is “small”, then the domain is “close” to a ball.
1.1. Statement of the problem and main result
The precise problem that we consider may be stated as follows. Let be a bounded domain – that is a bounded, open, connected set, whose boundary will be denoted by , and let be an open (not necessarily connected) subset of with boundary denoted by . We consider the following problem:
(1.1) |
under the overdetermined condition
(1.2) |
for some . Here and in what follows denotes the outward unit normal of and the derivative of in the direction . Concerning the setting in (1.2), we remark that even without explicitly imposing any regularity assumptions on , [Vo, Theorem 1] guarantees that
(1.3) |
therefore the notation on is well posed in the classical sense, being by standard elliptic regularity theory.
Concerning the regularity of the domain taken into account, to avoid unessential technicalities we will first assume that
(1.5) | is of class |
and that
(1.6) |
We recall, for instance, that domains with boundaries satisfy (1.6), see e.g. [ROV, Lemma A.1].
To state our main result we introduce some notation. For a given domain , we denote by and the -dimensional Lebesgue measure of and the surface measure of , respectively. Our main result aims at considering a convenient point and at obtaining suitable bounds on the “pseudo-distance”
(1.7) |
and on the “asymmetry”
(1.8) |
where denotes the symmetric difference of and the ball of radius centered at . In addition, we provide a “geometric” bound on the set by estimating the difference between the largest ball centered at contained in and the smallest ball centered at that contains .
The precise result that we have here goes as follows:
Theorem 1.1.
Let be a bounded domain satisfying assumptions (1.5) and (1.6). Let satisfy (1.1), (1.2) and (1.4). Assume that on . Set
(1.9) |
Then,
(1.10) |
and
(1.11) |
In (1.10) and (1.11), the quantity denotes a positive constant depending only on the dimension , on the internal radius in (1.6), on the diameter of , and on .
Moreover, if
(1.12) |
then there exist such that
(1.13) |
and
(1.14) |
where is a positive constant depending only on the dimension , on the internal radius in (1.6), on the diameter of and on , and is a constant depending only on .
We point out that Theorem 1.1 collects three different aspects on the problem: indeed, the statement in (1.10) is an “integral” stability result in (see Section 5), the estimate in (1.11) provides a stability “in measure”, and the result in (1.14) deals with a “pointwise” notion of stability.
The exponent in (1.14) can be better precised. Indeed, we can take
Moreover, can be taken arbitrarily close to one, in the sense that for any , we have that (1.14) holds with and depending also on .
Furthermore, for , we can take
We think that it is an interesting open problem to detect whether these choices of exponents , , in (1.14), as well as the exponents appearing in (1.10) and (1.11), are optimal in Theorem 1.1. It would also be very interesting to have explicit examples to check the optimality of the structural assumptions in Theorem 1.1.
We also point out that condition (1.12) is naturally satisfied in many concrete situations (see also Remark 7.2): in particular, for “small” , the point is “close” to the baricenter of , hence condition (1.12) is fulfilled in this case when the baricenter of lies in (as it happens, for instance, for convex sets).
Of course, when , we have that (1.14) reduces to and therefore (1.13) gives that is a ball: in this sense Theorem 1.1 recovers the classical results of [Se, We] for overdetermined problems. The main difference here is that, differently from the existing literature, the equation is supposed to hold possibly only outside a subdomain : as a counterpart, Theorem 1.1 does not prove that the full domain is a ball, but only that it is geometrically close to a ball whenever the subdomain has a small Lebesgue measure.
In the present setting, Theorem 1.1 will be in fact a particular case of more general quantitative results, presented in Section 7, and relying on a number of auxiliary integral identities. For more details, see Theorems 7.1, 7.3, 7.4, and 7.6.
In this spirit, Theorem 1.1 falls within the broad stream of research aiming at obtaining quantitative rigidity results, see e.g. [ABR, BNST, CMV, CV1, Fe, Ma, MP1, MP2, MP3, Pog2] and the references therein. More generally, overdetermined problems have been also considered e.g. in [AB, CV2, EP, FV1, FV2, FV3, FV4, FG, FGK, FGLP, MR1808686, MR2002730, GL, GS, GX, PS, Pog] and in the references therein.
1.2. Comments on the structural assumptions and generalizations
We remark that assumption (1.6) gives a lower bound on the distance not only between and (being ), but also between the boundaries of the different connected components of (if any).
In particular, suitable counterparts of (1.10) and (1.11) can be obtained if (1.5) and (1.6) are replaced by the weaker assumptions
(1.15) | is a John domain |
and
(1.16) | is of finite perimeter. |
Moreover, a counterpart of the pointwise estimate (1.14) can be obtained if (1.5) and (1.6) are dropped and replaced by (1.15), (1.16), and the assumption
(1.17) |
We stress that (1.17) is equivalent to assume a lower bound only for . Indeed, being of class , the set surely satisfies the uniform interior sphere condition on .
In this situation, Theorem 1.1 remains valid, with the following structural modifications:
-
•
The boundary measure of is replaced by its perimeter, namely by the -dimensional Hausdorff measure of its reduced boundary . In turn, on has to be intended as the (measure-theoretic) outer unit normal (see Section 8).
-
•
The constants depend on defined in (1.17) and on the structural constant of the given -John domain;
- •
The definition and details for -John domains can be found in Subsection 8.1. Here, we just stress that the class of John domains is huge: in particular, if (1.6) is satisfied then is surely a -John domain with (see [Pog2, (iii) of Remark 3.12]).
The precise statement that we have in this framework is stated next, and can be deduced from more general results presented in Theorems 8.3, 8.6, 8.9 and 8.10.
Theorem 1.2.
Let be a bounded domain satisfying assumptions (1.15) and (1.16). Let satisfy (1.1), (1.2) and (1.4). Assume that on . Set
Then, the pseudodistance defined in (1.7) and the asymmetry defined in (1.8) satisfy
(1.18) |
(1.19) |
where the constants appearing in (1.18) and (1.19) depend only on , , , , and .
The dependence of the constants on in (1.18) and (1.19) could be replaced with the dependence on the surface measure , as explained in Remark 8.11.
The explicit values of in (1.20) are the following ones. We have that can be taken as close as we wish to , namely one can fix any and take (in this case, the constant in (1.20) will also depend on ). When , one can take .
We notice that these exponents are all smaller (i.e., “worse”) than the ones obtained in Theorem 1.1. Nevertheless, it is possible to get the pointwise estimate (1.20) with the better exponents obtained in (1.14), and by removing the John condition (1.15), provided that we make a different choice of .
The precise statement, that can be deduced from more general results obtained in Theorems 8.15 and 8.16, is the following:
Theorem 1.3.
Let be a bounded domain satisfying assumptions (1.16) and (1.17). Let satisfy (1.1), (1.2) and (1.4). Assume that on . Set
(1.21) |
where denotes the points in which lie at distance strictly less than from , and denotes the points in which lie at distance from .
If , then there exist such that
and
where is a constant depending only on , the internal radius in (1.17), , and . The exponents depend only on .
We stress that the approach used in Theorem 1.3 does not need the assumption in (1.15) that is a John domain and hence the dependence on , present in (1.20), has been dropped.
Interestingly, the values of the exponents in Theorem 1.3 are the same as those in Theorem 1.1 (and therefore they are “better” than the ones obtained in (1.20), though they rely on a different choice of ).
We think that it would be interesting to investigate the possible optimality of these exponents also in the framework provided by Theorem 1.3.
1.3. Organization of the paper
The rest of this paper is organized as follows. Section 2 contains some detailed motivations from shape optimization, fluid dynamics and mechanics which naturally lead to the problem considered in this paper.
In Section 3 we make some notation precise.
In Section 4, we present some integral identities of Rellich-Pohozaev-type for solutions of (1.1). In these computations, one does not need to impose the additional condition in (1.2) from the beginning, and aims at comparing a weighted “deficit” on (measuring “how far from rotational invariant” the solution is) with suitable surface integrands on and . From these identities, the auxiliary information in (1.2) provides a more precise, and simpler information.
In Section 5 we collect useful estimates and we use them to obtain a suitable stability bound on the spherical pseudo-distance defined in (1.7) (see Theorem 5.7). We also put in relation this pseudo-distance with the asymmetry defined in (1.8) (see Lemma 5.1).
2. Models and motivations
2.1. Heating with source malfunctioning
A natural motivation for problem (1.1)-(1.2) comes from the optimal heating theory. In this setting, a region is given which is in direct contact with an external environment having constant (say, zero) temperature. In this setting, at the equilibrium, the temperature on the boundary of is set to be zero. The region is also provided by a fine set of heating devices. All these devices are the same and produce the same heating effect with the exception of those placed in a small subregion which are malfunctioning, see Figure 1.

In this setting, at the equilibrium, the local heat flow, normalized by the flow surface, is constant, say equal to , in , but it may be different from in . In a nutshell, the mathematical description of this situation is given by
(2.1) |
with for every .
A natural question in this setting is to optimize the shape of in order to store in the domain the biggest possible amount of caloric energy. Concretely, given satisfying (2.1), one may try to select in order to maximize (among the domains of fixed measure) the functional
(2.2) |
This variational problem is well posed (see e.g. Theorem 4.5.2 in [HP]) and it is naturally related to the boundary derivative prescription
(2.3) | on , |
thus leading to the problem described in (1.1)-(1.2). We refer to Proposition 6.1.10 in [HP] and Appendix A here for a direct computation relating the shape optimization of (2.2) and the Neumann condition in (2.3).
In this framework, our result111Strictly speaking, (2.1) with a source should be described as an optimal cooling, rather than heating, problem: speaking of heating problem should require to add a minus sign in front of in (2.1). Nevertheless, we preferred to keep the sign convention in accordance with (1.1) and the rest of the paper. in Theorem 1.1 states that, for a small region of malfunctioning of the heat source, the optimal domain is necessarily close to a ball (with a quantitative information on the proximity between and a suitable ball). This result is close to intuition, since jagged domains end up dissipating most of the caloric energy from their boundaries.
2.2. Laminar flows with a small tube of unknown density
Another motivation of the problem in (1.1)-(1.2) comes from laminar flows, as modeled by a Navier-Stokes equation of the type
(2.4) |
where is the vectorial velocity of the fluid, is its density, is its viscosity coefficient, is the pressure, is the gravity acceleration and denotes the outer product (i.e., given two vectors and , is the matrix whose th entry is ) see e.g. [Da]. The incompressibility condition
(2.5) |
leads to
In this way, one obtains from (2.4) that
(2.6) |
One assumes that the flow is “vertical”, namely for some scalar function , thus obtaining that
(2.7) |
We also suppose that the laminar flow occurs in a “vertical tube” of the type , with , and that the density of the fluid only depends on the horizontal position (that is, the fluid maintains the same density along its vertical flow). In this setting, we have that and accordingly, by (2.5) and (2.7),
Hence, we deduce from the third component of (2.6) that
(2.8) |
The case of a “steady state” flow (that is ) with constant pressure (hence ) reduces (2.8) to
(2.9) |
If the fluid has constant (say, up to changing the inertial reference frame, zero) velocity at the boundary of the pipe, equation (2.9) is complemented by the boundary condition
(2.10) |
Also, if the fluid presents a constant tangential stress on the pipe, we have that
(2.11) |
The case described in (1.1)-(1.2) is, in this setting, a byproduct of (2.9), (2.10) and (2.11) in which the density of the fluid (as well as its viscosity and the gravity acceleration) is constant in the region , but it is possibly unknown in . In this framework, our result in Theorem 1.1 says that if a laminar fluid is homogeneous out of a small region and presents a constant tangential stress on the pipe, then necessarily the pipe is close to a right circular cylinder.
2.3. Traction of beams with small inhomogeneity
In the theory of elasticity, one can consider the displacement vector that describes the deformation of some material. Also, it is customary to introduce the stress tensor
(2.12) |
describing the force (per unit area) in the th direction along the infinitesimal surface orthogonal to , being , and , see e.g. formula (129) in [Hj] (here, we are supposing the strain and the stress to be proportional, setting the proportionality constant equal to for the sake of simplicity).
In this framework, the equilibrium configurations are those for which the forces are infinitesimally balanced, that is
(2.13) |
see e.g. formula (189) in [Hj].
More specifically, we focus here on the torsion of a vertical beam (see e.g. [So, pages 100-120]). The fact that the beam is vertical means, in our setting, that
(2.14) |
for a suitable (bounded and smooth) family of domains . We assume that each point of the beam lying on a given horizontal plane performs a horizontal rotation of a small angle , plus a vertical movement that is the same for every height of the bar (we will denote this vertical movement by ). In this setting, the torsion of the bar moves a point to the point
and so, in this approximation, we can write the displacement vector as
From this and (2.12), we can write
(2.15) |
As a result, exploiting the latter equation and (2.13) with ,
(2.16) |
We then fix and consider the -form on given by
(2.17) |
and we deduce from (2.16) that
This gives that there exists a “warping potential” such that as -forms in . Accordingly, by (2.17), we have that
(2.18) |
Combining this with (2.15), one sees that
and therefore
(2.19) |
with these functions evaluated at . As a special case, one can take into account the situation in which the angle depends linearly on the height of the beam, say (this is physically reasonable, for instance, if the beam is constrained at and some torque is applied from the top of the beam). In this setting, (2.19) reduces to
(2.20) |
If we suppose that the beam is built by two different materials, one occupying and the other occupying (where here represents the domain in (2.14) with ), the expression in (2.20) takes two different forms in and . In particular, if we know that the material in is homogeneous we can suppose that the horizontal rotation is uniform there and thus is independent on the point (i.e., ), hence deducing from (2.20) that
(2.21) |
Also, the surface traction, as a force per unit of area, at a boundary point of the beam is defined as the normal component of the vertical stress, that is
being the normal to the beam (see e.g. the third component in formula (174) of [Hj]). Recalling (2.15), we have that
being normal to . Thus, in view of (2.18),
being a unit tangent vector to . Therefore, if the traction vanishes, the tangential derivative of vanishes as well, hence is constant along . Since was introduced as a potential, it is defined up to an additive constant, hence we can rephrase these considerations by saying that if the traction vanishes, then
(2.22) |
We also remark that, if , then
Hence, for constant stress intensity , we deduce from (2.22) that
(2.23) |
We thus observe that problem (1.1)-(1.2) arises naturally from (2.21), (2.22) and (2.23), and, in this framework, Theorem 1.1 says that if a beam is homogeneous out of a small region and presents zero traction and constant stress intensity, then necessarily the horizontal section of the beam is close to a disk.
3. Notation
Unless differently specified, we will denote by , with , a connected, bounded open set, and call its boundary. We will denote indifferently by and the -dimensional Lebesgue measure of and the surface measure of .
Moreover, we will denote by the diameter of , that is
(3.1) |
Also, we shall denote by an open subset of , such that .
For all , we consider the distance function
(3.2) |
As already mentioned in the Introduction, we first consider the case in which, being (1.5) in force and by recalling (1.3), is of class : in this setting denotes the (exterior) unit normal vector field to . Then, we will clarify in Section 8 the notation that we use when the regularity assumption in (1.5) is dropped.
Now, we clarify the notation used for the spaces and we discuss the regularity assumption in (1.4).
As usual, for an open set , and a positive integer, denotes the space of functions possessing continuous derivatives up to order on .
By we denote the space of functions that are restrictions to of functions in . In other words, is the space of the functions in that can be extended to . The same definition has been adopted for instance in [Leoni, Appendix C, pag. 562] and also in many books of Differential Geometry.
We recall that, thanks to Whitney extension theorem (see the original paper [Wh] or [KP-Primer, Theorem 2.3.6]), this definition is equivalent to [KP-Primer, Definition 2.3.5] based on a Taylor-expansion condition.
Moreover, if satisfies property (P) of [Wh-boundaries] – i.e., quasiconvexity, that in particular is surely satisfied if is Lipschitz (see [Br, Sections 2.5.1, 2.5.2]) – then, thanks to the main theorem on page 485 in [Wh-boundaries], our definition of (as well as [KP-Primer, Definition 2.3.5]) is also equivalent to the definition used in many books of PDEs (e.g., [GT]), that is: is the space of functions in whose derivatives up to order have continuous extensions to the closure .
For more general domains, [KP-Primer, Definition 2.3.5] (as well as our definition) is stronger than that adopted in [GT]. For more details on this subject we refer to [KP-Primer, Section 2.3], [KP, Section 5.2], [Kra], [Br], and [Leoni, Appendix C].
In our setting, by taking and in the definition of , we have that the assumption in (1.4) guarantees that can be extended to a function throughout . This will allow us to perform the generalizations described in Section 8 when the regularity assumption on in (1.5) is dropped and replaced just by (1.16). In particular, when integrating by parts, we can still write the derivatives of up to the second order on the reduced boundary .
4. Integral identities
The goal of this section is to develop a series of integral identities which will be conveniently exploited to deduce quantitative bounds on the solution of (1.1)-(1.2) and on its domain of definition. We start by proving a Rellich-Pohozaev-type identity and its consequences. Notice that in the next two statements we are not imposing yet the overdetermined condition in (1.2).
Lemma 4.1 (A Rellich-Pohozaev-type identity).
Suppose that is of class . If satisfies (1.1), then the following identity holds:
(4.1) |
Proof.
By a direct computation, it is easy to verify the following differential identity (valid for every function ):
(4.2) |
We now specialize this identity to a solution of (1.1). For this, we remark that the Dirichlet boundary condition in (1.1) gives that
(4.3) | on . |
We also remark that
(4.4) |
Hence, by integrating (4.2) over , exploiting (1.1), (4.3) and (4.4), and applying the divergence theorem, we see that
(4.5) |
Now we observe that, in ,
where the equation in (1.1) has been used in the last step.
As a consequence,
By integrating this identity over , and using the boundary condition in (1.1), we deduce that
By putting together the last identity and (LABEL:eq:1perPohozaev), we obtain (LABEL:eq:Pohozaevadhoc), as desired. ∎
We now obtain another useful integral identity, which is based on (LABEL:eq:Pohozaevadhoc) and a suitable -function computation.
Theorem 4.2.
Suppose that is of class . If satisfies (1.1), then the following identity holds:
(4.6) |
We point out that the term in the braces in the left-hand side of (4.6) could be written as . Nevertheless, we preferred to use the notation
(4.7) |
to emphasize that this quantity plays the role of a Cauchy-Schwarz deficit. In fact, by Cauchy-Schwarz inequality we have that (4.7) is nonnegative, and equals in if and only if is a ball of radius and in , up to translations (see, e.g., [Pog, Lemma 1.9]).
Proof of Theorem 4.2.
If we set
(4.8) |
a direct computation, valid for any function, informs us that
(4.9) |
where denotes the Hessian matrix of and denotes its Frobenius norm, that is
Then, using (4.9) and the equation in (1.1), we conclude that
(4.10) |
On the other hand, by (1.1) and the Green identity,
(4.11) |
Let us work on the first integral on the right-hand side of (4.11). For this, integrating over the differential identity
and recalling (1.1), we get that
Thus, in light of (4.8),
Consequently, using (LABEL:eq:Pohozaevadhoc),
(4.12) |
Moreover, to deal with the second integral in the right-hand side of (4.11), by recalling (1.1) and (4.8), we have
(4.13) |
We now impose the overdetermined condition (1.2), and we obtain from (4.6) the following integral identity:
Corollary 4.3.
Proof.
We observe that the constant in (1.2) can be determined explicitly in terms of , , and the values of along . Indeed, by using (1.1) and (1.2) together with the divergence theorem, we deduce that
(4.16) |
In particular, we will use that
(4.17) |
On the other hand, by applying again the divergence theorem,
From this and (4.17), we conclude that
Plugging this information into (4.6) we obtain the desired result in (4.15). ∎
5. Some estimates on a spherical pseudo-distance
In this section, we will obtain a suitable bound on the following pseudo-distance
(5.1) |
for a suitable .
We point out that the quantity in (5.1) plays the role of an “integral distance” of from the sphere centered at a point of radius : indeed, when , the quantity in (5.1) vanishes, and, in general, this quantity can be considered an -measure on how far points on are from points on .
We also notice that the pseudo-distance in (5.1) can be put in relation with the following asymmetry:
(5.2) |
where denotes the symmetric difference of and the ball of radius centered at .
In particular, the asymmetry in (5.2) is bounded from above by the pseudo-distance in (5.1), as stated in the following result:
Lemma 5.1.
Let be a bounded domain with Lipschitz boundary , satisfying the uniform interior sphere condition with radius . Then, there exists a positive constant , only depending on , and , such that
Proof.
The desired result follows by applying [Fe, Lemma 11] with
Notice that [Fe, Lemma 11] can be applied with these choices for and because the following two relations hold true: the first is
(5.3) |
where in the last inequality we used that ; the second is
(5.4) |
where denotes the inradius of and in the last inequality we used that, by definition, . ∎
To obtain our bounds for the pseudo-distance introduced in (5.1), we recall the notation in (3.2) and we detect an optimal growth of the solution from the boundary, by adapting an idea from [MP2, Lemma 3.1]:
Lemma 5.2.
Let satisfy (1.1). Assume that on . Then,
(5.5) |
Moreover, if is of class and satisfies the uniform interior sphere condition with radius , that is (1.6), then it holds that
(5.6) |
Proof.
Let and set . We consider
We remark that is the solution of the classical torsion problem in , namely
(5.7) |
By comparison we have that on . In particular,
and (5.5) follows.
We point out that (5.6) follows from (5.5) if . Hence, from now on, we can suppose that
(5.8) |
Let be the closest point in to and call the ball of radius touching at and containing . Up to a translation, we can always suppose that
(5.9) | the center of the ball is the origin. |
Now, we let be the solution of (5.7) in , that is . By comparison, we have that in , and hence, being ,
(5.10) |
Moreover, from (5.9),
We recall now some Hardy-Poincaré-type inequalities that have been proved in [Pog2, Section 3.2] by exploiting the works of Hurri-Syrjänen [Hu, HS]. In what follows, for a set and a function , denotes the mean value of in , that is
(5.11) |
Also, for a function we define by its -norm in , that is
(5.12) |
and
for and . Here and whenever no confusion is possible, we will use the abbreviated notation
that agrees with (3.2) when .
Lemma 5.3.
Let be a bounded domain satisfying the uniform interior sphere condition with radius , and consider three real numbers , and such that either
(5.13) |
or
(5.14) |
Then,
(i) given , there exists a positive constant such that
(5.15) |
for every function which is harmonic in and such that ;
(ii) there exists a positive constant such that
(5.16) |
for every function which is harmonic in .
Lemma 5.3 follows from [MP3, item(i) of Lemma 2.1 and items (i) and (ii) of Remark 2.4].
From Lemma 5.3 we can derive estimates for the derivatives of harmonic functions, as stated in the next result (a proof of this can be found in [MP3, Corollary 2.3]).
Corollary 5.4.
Let be a bounded domain satisfying the uniform interior sphere condition with radius , and let be a harmonic function in . Consider three real numbers , and satisfying either (5.13) or (5.14).
(i) If is a critical point of in , then it holds that
(ii) If
then it holds that
Remark 5.5.
For later use, we mention that Lemma 5.3 and Corollary 5.4 hold true more in general if the assumption of the uniform interior sphere condition is dropped and replaced by the assumption that is a John domain (see [Pog2, Section 3.2] or [MP3, Lemma 2.1 and Corollary 2.3]): in this case explicit estimates of the relevant constants now depending on the John parameter can be found in [MP3, Remark 2.4].
With the aid of Corollary 5.4 we now prove the following lemma, which, together with the forthcoming Theorem 5.7, leads to a stability estimate in terms of the pseudo-distance introduced in (5.1).
Lemma 5.6.
Let be a bounded domain of class satisfying the uniform interior sphere condition with radius , that is (1.6), and let be a harmonic function in . Let satisfy (1.1) and assume that on .
(i) If is a critical point of in , then it holds that
(ii) If
then it holds that
Proof.
We begin with the following differential identity:
(5.21) |
that holds in for any harmonic function in , if is satisfies (1.1).
Next, we integrate (5.21) on and, by the divergence theorem, we get
We use this identity replacing the harmonic function with its derivative , and then we sum up over . In this way, we obtain
(5.22) |
We observe that, by an adaptation of Hopf’s lemma (see [MP1, Theorem 3.10]), the term in the left-hand side of (5.22) can be bounded from below by , namely
(5.23) |
Hence, we obtain from (5.22) that
(5.24) |
Now we suppose that is a critical point of in and we use item (i) in Corollary 5.4, applied here with , and , and we deduce from (5.24) that
From this and (5.6), one obtains the desired estimate in item (i). In a similar way, using item (ii) in Corollary 5.4, one shows item (ii) here, thus completing the proof. ∎
Now, we turn our attention to the harmonic function
(5.25) |
where
(5.26) |
for some choice of and .
We remark that, by a direct computation, it is easy to check that the Cauchy-Schwarz deficit appearing in the left-hand side of (4.15) can be written in terms of as
(5.27) |
Now we specify the choice of the point in (5.26) as follows
(5.28) |
We notice that as tends to the empty set and tends to 0, tends to the baricenter of (however, is not the baricenter of ).
With this choice of we have that
(5.29) |
Indeed, by a direct computation we get that
(5.30) |
and therefore, using Green’s identity and the fact that on ,
thus proving (5.29).
Gathering the previous results, we thus obtain the desired estimate on the pseudo-distance:
Theorem 5.7.
6. Some estimates on
The aim of the present section is to obtain quantitative estimates for the difference .
We remark that, recalling the notation in (3.1),
(6.3) |
Indeed, for every , , we have that
As a result, taking maximizing the distance to , and minimizing the distance to , we obtain that , that is (6.3).
Also, by using to denote the distance of a point to the boundary we define the complementary parallel set as
(6.4) |
Notice that, since satisfies the uniform interior sphere condition of radius , it holds that
Lemma 6.1 below contains an inequality for the oscillation of a harmonic function in terms of its -norm in and of a bound for its gradient in . More precisely, recalling the notation in (5.11) and (5.12), we have:
Lemma 6.1.
Let satisfy the uniform interior sphere condition of radius on , that is (1.17), and suppose that is of class . Let be a harmonic function in of class , and let be an upper bound for the gradient of on .
Then, given , there exist two positive constants and depending only on and such that if
(6.5) |
then we have that
(6.6) |
Lemmata 6.1 and 6.3 and Theorem 6.4 here adapt to the present situation ideas originating from [Pog2] and [MP3] – see also [MP4] for generalizations in other directions of those ideas. Here, we obtain Lemma 6.1 as an immediate consequence of the following new refined estimate, that will be crucial in Subsection 8.2.
Lemma 6.2.
Let satisfy the uniform interior sphere condition of radius on , that is (1.17), and suppose that is of class . Let be a harmonic function in of class , and let be an upper bound for the gradient of on .
Given , we choose for which
(6.7) |
and set
(6.8) |
Then, given , there exist two positive constants and depending only on and such that if
(6.9) |
then we have that
(6.10) |
Proof.
By (6.7), it holds that
(6.11) |
For , we define
By the fundamental theorem of calculus we have that
(6.13) |
Furthermore, since is harmonic in , we can use the mean value property for the balls with radius centered at , thanks to (6.12), and find that
where we used an application of Hölder’s inequality and (6.12) once again.
Therefore, by minimizing the right-hand side of the last inequality, we can conveniently choose
(6.15) |
The computations show that
(6.16) |
∎
Proof of Lemma 6.1.
We now turn our attention to the harmonic function introduced in (5.25), and we modify Lemma 6.1 to link to the -norm of . Since on , we have that
(6.17) |
We also observe that, by definition of , it follows that
(6.18) |
Then, from (6.17) and (6.18) we obtain that
(6.19) |
The next result gives an explicit bound on the difference :
Lemma 6.3.
Let satisfy the uniform interior sphere condition of radius on , that is (1.17), and suppose that is of class . Let satisfy (1.1) and , let be as in (5.26) with , and let be as in (5.25).
Then, there exists a positive constant such that
(6.20) |
The constant depends on , , , , , where
(6.21) |
Proof.
We now consider the constants and defined in (6.16) and we distinguish two cases, according on whether
(6.22) |
or
(6.23) |
If (6.22) holds true, we can apply Lemma 6.1 with and . Thus, by means of (6.19) we deduce that (6.20) holds with
(6.24) |
Theorem 6.4.
Let satisfy the uniform interior sphere condition of radius , that is (1.6), and suppose that is of class . Let satisfy (1.1), , and suppose that on . Let be as in (5.26) with chosen as in (5.28), and assume that belongs to . Let be as in (5.25).
Then, there exists a positive constant such that
(6.25) |
with the following specifications:
-
(i)
;
-
(ii)
is arbitrarily close to , in the sense that, for any sufficiently small, there exists a positive constant such that (6.25) holds with ;
-
(iii)
for .
The constant depends on , , , (as defined in (6.21)), and (the latter, only in the case ).
Proof.
For the sake of clarity, we will always use the letter to denote the constants in all the inequalities appearing in the proof. Their explicit computation will be clear by following the steps of the proof (see the forthcoming Remark 6.5).
(i) Let . By the Sobolev immersion theorem (for instance we apply [Fr, Theorem 9.1] to the function ), we deduce that there exists a positive constant such that
(6.26) |
As noticed in [MP3, Remark 2.9], the immersion constant in (6.26) depends on and only.
Applying (5.16) with , , , and leads to
(6.27) |
Also, since (5.29) holds true, we can apply item (ii) of Corollary 5.4 with , , , , and and obtain that
From this and (6.27), we get that
This inequality, together with (6.26), gives that
(ii) Let . For any sufficiently small, we notice that
satisfy (5.13) in Lemma 5.3. Hence, we can apply the estimate in (5.16) with and , obtaining that
(6.28) |
Furthermore,
satisfy (5.13), and therefore item (ii) of Corollary 5.4, applied again with and , yields that
This and (6.28) give that
Thus, by using Lemma 6.3 with , we obtain that
which is (6.25) with .
(iii) Let . In light of (5.29), we can apply to item (ii) of Corollary 5.4 with , , , and (noticing that they satisfy (5.13)), and obtain that
(6.29) |
Being , we can also apply (5.16) with , , , , and , and get
(6.30) |
Thus, from (6.29) and (6.30) we conclude that
Then, Lemma 6.3, applied with , gives that (6.25) holds with . ∎
Remark 6.5 (On the constant in (6.25)).
The constant in (6.25) can be explicitly computed by following the steps of the proof of Theorem 6.4, and it can be shown to depend only on the parameters mentioned in the statement of Theorem 6.4. Indeed, the parameters and , can be estimated by means of (5.17) and (5.19). Then, we notice that and
(6.31) |
Finally, to remove the dependence on the volume, we use the trivial bound
(6.32) |
Remark 6.6 (Another choice for the point in (6.1)).
Another possible way to choose in (6.1) (different from (5.28)) is
where is any point such that . In fact, with this choice we obtain that and we can thus use item (i) of Corollary 5.4 instead of item (ii).
Also with this choice, in order to obtain an analogue of Theorem 6.4, we should additionally require that , to be sure that the ball is contained in .
7. Stability results and proof of Theorem 1.1
For the sake of clarity, we state here some notation that will be used throughout the rest of this paper.
We denote by the supremum of the diameters of all the connected components of (of course, if is connected, then coincides with the diameter of , and, in this case, according to the notation in (3.1), it holds that ). Then, we have:
Theorem 7.1 (General stability result for ).
Proof.
Remark 7.2.
We notice that with the choice of as in (5.28), (if is small enough) the assumption is satisfied, at least if the baricenter of lies in (in particular, if is convex).
With this preliminary work, we are now in the position of obtaining a quantitative rigidity result bounding the averaged squared pseudodistance of the form
where is as in (5.28) and is that in (1.2). The precise result goes as follows:
Theorem 7.3 (General stability result for a pseudodistance).
Let satisfy (1.1) and (1.2), and suppose that on . Let assumptions (1.5) and (1.6) be verified, and be as in (5.28).
If is a continuous function vanishing at such that (7.1) holds true together with
(7.3) |
then
(7.4) |
where is a constant depending on , , .
Proof.
By means of Lemma 5.1, from Theorem 7.3 we also obtain a quantitative rigidity result for the asymmetry (5.2):
Theorem 7.4 (General stability result for an asymmetry).
Remark 7.5.
The dependence on of the constant in (7.5) can be dropped by exploiting suitable bounds for it. Indeed, on the one hand, putting together (1.2) and (5.23) one obtains the lower bound
(7.6) |
On the other hand, from the expression in (4.16) one can obtain an upper bound for in terms of and , when is small enough. More precisely, formula (4.16) implies that
(7.7) |
thanks to (7.1).
We can now obtain a quantitative symmetry result by assuming a -bound of the solution along :
Theorem 7.6.
Let satisfy (1.1) and (1.2), and suppose that on . Let assumptions (1.5) and (1.6) be verified, and be as in (5.28).
If there exists such that
(7.10) |
then
(7.11) |
where is a positive constant depending on , , , and .
Also, it holds that
(7.12) |
where is a positive constant depending on , , , and .
Moreover, if , we have that
(7.13) |
where are as in Theorem 6.4 and is a positive constant depending on , , , and .
Proof.
We notice that, by (7.10), we have that the assumptions in (7.1) and (7.3) are satisfied with
and so we are in the position of applying Theorems 7.1, 7.3, and 7.4, thus obtaining the desired estimates.
Notice that the constant in (7.13) does not depend on (as defined in (6.21)), differently from that appearing in (7.2). Indeed, we claim that, when ,
(7.14) |
for some positive constant depending on , , , and . We point out that, since , the estimate in (7.14) provides a bound for in terms of , , , and . Hence, we now focus on the proof of (7.14).
For this, we observe that, since attains its maximum on , recalling (1.2) and (7.10), we have that
(7.15) |
As a result, to obtain the desired estimate in (7.14), it remains to find an upper bound for depending on , , , and .
For this, we notice that, by (4.16) and (7.10),
and hence
(7.16) |
On the one hand, by (7.8) we already know that
(7.17) |
On the other hand, by combining the inequality
that holds true since a ball of radius is surely contained in , with the classical isoperimetric inequality , we get that
(7.18) |
Putting together (7.16), (7.17), and (7.18), we find the desired explicit upper bound for :
(7.19) |
In turn, this and (7.15) give that
that is the desired estimate in (7.14).
The estimate in (7.14) proves that, if , (7.13) holds true (with not depending on ). On the other hand, if , (7.13) trivially holds true with , being .
Notice also that (7.12) is a global estimate in which the constant does not depend on , differently from that appearing in (7.5).
Indeed, if , we can remove the dependence of the constant on thanks to the bounds in (7.6) and (7.19).
This proves that, if , (7.12) holds true with not depending on . On the other hand, when , (7.12) trivially holds true with , being
These observations complete the proof of Theorem 7.6. ∎
Another instance in which the assumptions of Theorems 7.1, 7.3, and 7.4, are surely satisfied is the following. Notice that in high dimensions (i.e., when ) the following result allows to blow up on as either the perimeter or the diameter tends to zero.
Theorem 7.7.
Let satisfy (1.1) and (1.2), and suppose that on . Let assumption (1.6) be verified, and be as in (5.28). Let be the union of finitely many disjoint balls of radius and assume that
Then,
are as small as we wish for small .
Furthermore, if in addition , then also is as small as we wish for small .
8. The case of general domains and proofs of Theorems 1.2 and 1.3
As mentioned in the Introduction, analogous stability results can be obtained by weakening the assumptions in (1.5) and (1.6).
Subsection 8.1 is devoted to the case in which is a John domain. We extend the stability estimates for the spherical pseudodistance defined in (5.1) and for the asymmetry defined in (5.2) – i.e., Theorems 7.3, 7.4, and their corresponding consequences in Theorem 7.6 –, when (1.5) and (1.6) are dropped and replaced by the weaker assumptions (1.15), (1.16); that is,
(8.1) | when is a bounded John domain of finite perimeter. |
We also show that the pointwise results of Theorem 7.1 and its corresponding consequences in Theorem 7.6 can be obtained when (1.5) and (1.6) are replaced with the weaker assumptions (1.15) (1.16), (1.17), that is,
(8.2) |
at the cost of getting a worse stability exponent .
All the generalizations presented in Subsection 8.1 are obtained by using the same choice (5.28) for the point . As a particular case of those generalizations, we obtain Theorem 1.2.
In Subsection 8.2 we show how a different choice of the point allows to obtain Theorem 7.1 and its corresponding consequences in Theorem 7.6 in their full power – i.e., with given in Theorem 6.4 – under the weaker assumptions (1.16) and (1.17). We stress that this approach does not need the assumption (1.15) that is a John domain, requested in the generalizations of Subsection 8.1. In fact, the set of assumptions on in Subsection 8.2 is
(8.3) |
To deal with sets of finite perimeter, which is common both in (8.2), (8.3), and (8.1), we recall that, thanks to De Giorgi’s structure theorem (see [Maggi, Theorem 15.9] or [Giusti]), the assumptions in (1.4) and (1.16) (in the sense explained in Section 3) guarantee that the integral identities proved in Section 4 still hold true, provided that one replaces with the reduced boundary and agrees to still use to denote the (measure-theoretic) outer unit normal (see e.g., [Maggi, Chapter 15]).
We observe that when in particular is of class , that is when (1.5) is in force, then and the (measure-theoretic) outer unit normal coincides with the classical notion of outer unit normal (see [Maggi, Remark 15.1]).
Moreover, in the setting of assumption (1.16), the surface measure has to be replaced with the perimeter of . We recall indeed that the perimeter of equals the -dimensional Hausdorff measure of , denoted by (see [Giusti, Chapter 4] or [Maggi, Chapter 15]). Of course, when is of class , as given by (1.5), those notions agree (since in that case we have ).
8.1. Generalizations for John domains and proof of Theorem 1.2
We first deal with the setting in (8.1) and we obtain the generalizations of Theorems 7.3 and 7.4. Then, by further assuming (1.17) (and hence in the setting (8.2)), we establish the generalizations of Theorem 7.1. As a consequence, we thus obtain Theorem 1.2.
The formal framework in which we work is the following. A domain in is a -John domain, with , if each pair of distinct points and in can be joined by a curve such that , , and
(8.4) |
We emphasize that the class of John domains is huge: it contains Lipschitz domains, but also very irregular domains with fractal boundaries such as, e.g., the Koch snowflake. For more details on John’s domains, see [Pog2, Section 3.2] or [Ai, MS, NV], and references therein. Here, we just notice that, if (1.6) is satisfied then is surely a -John domain with (see [Pog2, (iii) of Remark 3.12]).
As already mentioned in Remark 5.5, Lemma 5.3 and Corollary 5.4 still hold true without the assumption of the uniform interior sphere condition, if is a John domain (see also [Pog2, Section 3.2]). In this case, explicit estimates (now depending on the John parameter instead that on ) of the relevant constants of Lemma 5.3 and Corollary 5.4 can be found in [MP3, Remark 2.4]. For the reader’s convenience we report here the only estimate that we need to conclude our reasoning, that is,
(8.5) |
where , and are as in (5.13).
Now we point out the main changes to perform in this situation in order to get Theorem 7.3 and its corresponding consequences in Theorems 7.6 when (1.5) and (1.6) are dropped and replaced just by (1.16).
We notice that assumption (1.5) had been exploited only to allow the use of (5.6) in the proof of Lemma 5.6. However, we notice that (5.5) still holds true without that assumption.
Thus, in order to generalize the stability result of Theorem 7.3, we replace item (ii) of Lemma 5.6 with the following result:
Lemma 8.1.
Let be a bounded -John domain satisfying (1.16). Let satisfy (1.1) and (1.2), and assume that on . Let be a harmonic function in .
If
(8.6) |
then it holds that
Proof.
We remark that the same generalization could be applied – using item (i) in Corollary 5.4 – also to item (i) of Lemma 5.6, that could be useful for other choices of .
In the present subsection we maintain the same choice (5.28) for the point , that is,
(8.8) |
Theorem 8.2.
In the same way in which Theorem 5.7 led to Theorem 7.3, now Theorem 8.2 easily leads to a general stability result for John domains of finite perimeter. In this setting, using Theorem 8.2 in place of Theorem 5.7, one obtains a statement analogous to Theorem 7.3, with replaced with in (7.1) and (7.3), even if assumptions (1.5) and (1.6) are replaced by (1.16). The precise result is the following:
Theorem 8.3 (General stability result for a pseudodistance under relaxed assumptions).
Let a -John domain. Let satisfy (1.1) and (1.2), and suppose that on . Let assumption (1.16) be verified, and be as in (8.8).
If is a continuous function vanishing at such that
(8.10) |
and
(8.11) |
then
(8.12) |
where is a constant depending on , , , and .
Remark 8.4.
Concerning the constant in formula (8.12) of Theorem 8.3, we observe that the dependence from comes from the estimate of in (8.5). We also remark that the volume appearing in (8.5) can be estimated from above in terms of , by means of (6.31) and (6.32).
The dependence of in (8.12) on comes from the fact that the quantity plays a role in the estimates, as it can be seen from formula (LABEL:0895654tfbvnblk) in Theorem 8.2.
We point out that such a dependence can in fact be replaced with the dependence on , by means of a suitable lower bound for . Namely, we claim that
(8.13) |
if is small enough, where is a positive constant depending on . To prove it, we first observe that
(8.14) | for any , , a ball of radius is contained in . |
Indeed, since is a -John domain, we have that there exists a curve such that , and (8.4) holds true. Moreover, by inspection one sees that there exists such that , and so, by (8.4),
which implies that
(8.15) |
Furthermore, by the triangle inequality,
Theorem 8.3 also leads to a stability bound for the asymmetry defined in (5.2) by means of the following generalization of Lemma 5.1 to the case of John domains:
Lemma 8.5.
Let be a bounded -John domain with Lipschitz boundary . Then, there exists a positive constant only depending on , , , such that
Proof.
The desired result follows by applying [Fe, Lemma 11] with
Notice that [Fe, Lemma 11] can be applied with these choices for and . Indeed, formula (5.3) is still satisfied. On the other hand, to deduce formula (5.4) in this setting, we observe that
(8.18) |
where denotes the inradius of . To prove (8.18), one can use (8.14) and choose and in such that .
Then, from (8.18) one deduce that
In light of Lemma 8.5, we also deduce from Theorem 8.3 a stability result for an asymmetry in this setting:
Theorem 8.6 (General stability result for an asymmetry under relaxed assumptions).
Let a -John domain. Let satisfy (1.1) and (1.2), and suppose that on . Let assumption (1.16) be verified, and be as in (8.8).
Then, it holds that
(8.19) |
where is a constant depending on , , , and .
Remark 8.7.
Our next objective is to provide a stability estimate for , as given in formula (7.2), in the more general framework of John domains. This will lead to a general version of Theorem 7.1, which in turn will produce a general version of Theorem 7.6.
Notice that, when assumption (1.5) is replaced by (1.16), we have to use (5.5) instead of (5.6) to bound from below the left-hand side of (4.15). Thus, in this case, the quantity that has to be put in relation with is instead of . For this reason, the exponents in Theorem 8.9, stated next, become worse with respect to those of Theorem 7.1.
More precisely, the counterpart of Theorem 6.4 in this more general setting is the following:
Theorem 8.8.
Let be a bounded -John domain, satisfying (1.17), and suppose that is of class . Let satisfy (1.1), , and suppose that on . Let be as in (5.26) with chosen as in (8.8), and assume that belongs to . Let be as in (5.25).
Then, there exists a positive constant such that
(8.20) |
The constant depends on , , , , (as defined in (6.21)).
Proof.
The desired estimate easily follows by reasoning as in the proof of items (ii) and (iii) of Theorem 6.4. The only difference is that now we apply Poincaré inequalities of item (ii) of Corollary 5.4 with (instead that ).
More precisely, when we apply, with and , item (ii) of Corollary 5.4 with , , and (5.16) with , , . When we apply, with and , item (ii) of Corollary 5.4 with , and (5.16) with , , .
We stress that assumption (1.6) in items (ii) and (iii) of Theorem 6.4 was assumed only to guarantee that the Poincaré inequalities of Lemma 5.3 and Corollary 5.4 could be applied with ; in fact, notice that Lemma 6.3, that was the main ingredient of those proofs, was already stated under the weaker assumption (1.17). Being now a -John domain, in light of Remark 5.5 and (8.5) it is clear that we can still apply those Poincaré inequalities with , even if (1.6) has been dropped. ∎
By applying Theorem 8.8 in the place of Theorem 6.4, we thus derive a counterpart of Theorem 7.1 for John domains. Namely, in this setting, the result in Theorem 7.1 still holds true if assumptions (1.5), (1.6) are replaced with (1.16), (1.17) – with replaced with in (7.1) – but in this situation the exponents are those in (8.20). The precise result is indeed the following one:
Theorem 8.9.
Let be a bounded -John domain. Let satisfy (1.1) and (1.2), and suppose that on . Let assumptions (1.16) and (1.17) be verified.
Assume also that the point chosen in (8.8) belongs to .
Proof.
As can be deduced from the discussion before Theorem 8.8, if (1.5) holds true, then the statement in Theorem 8.9 can be strengthen, since in this case one can obtain as in Theorem 6.4 (we do not enter into these details, since this special statement will not be used in what follows).
The corresponding generalization of Theorem 7.6 easily follows from Theorems 8.3, 8.6, and 8.9, and it can be stated as follows:
Theorem 8.10.
Let be a bounded -John domain. Let satisfy (1.1) and (1.2), and suppose that on . Let assumption (1.16) be verified, and be as in (8.8).
If there exists such that
(8.22) |
then
(8.23) |
where is a positive constant depending on , , , and .
Also, it holds that
(8.24) |
where is a positive constant depending on , , , , and .
Remark 8.11.
As in Theorem 7.6, the constant in (8.25) does not depend on (as defined in (6.21)), differently from that appearing in (8.21).
8.2. A new different choice for and proof of Theorem 1.3
Our next goal is to show that it is possible to obtain Theorem 7.1 and its consequences in Theorem 7.6, with given in Theorem 6.4 and in the more general setting (8.3), provided that we make a different choice of .
The main difficulty in this setting is that no regularity information is available on (not even being a John domain), and therefore we cannot apply Poincaré inequalities on all , making it difficult to establish an appropriate variant of Theorem 6.4.
The key idea to overcome this difficulty is to perform the necessary Poincaré inequalities on a suitable subset of . Such a suitable subset is , where we are using the notation introduced in (1.17) and (6.4). Notice that, by (1.17), it holds that
With this setting, we start with the following estimate:
Lemma 8.12.
Let satisfy (1.17), and suppose that is of class . Let be a harmonic function in of class , and let be an upper bound for the gradient of on .
Then, given , there exist two positive constants and depending only on and such that if
(8.27) |
then we have that
(8.28) |
Proof.
We recall that, thanks to (1.17), inherits the same regularity of . More precisely, we have that:
(8.29) |
This fact relies on the regularity of the distance function. The case has been proved in Appendix of [GT] (see also [KrantzParks-distance, Theorem 3]). The case can be deduced from [KrantzParks-distance, Theorem 2], but we do not need this refinement here.
Moreover, we have that:
Lemma 8.13.
Proof.
In order to use the Poincaré inequality of item (ii) of Corollary 5.4 (with and ), we have to make a new appropriate choice of .
To this aim, a possible choice of is:
(8.30) |
where
(8.31) |
We remark that, with the choice in (8.30), by (5.30) and Green’s identity it follows that
(8.32) |
We are now in position to prove the counterpart of Theorem 6.4 in the present setting.
Theorem 8.14.
Let satisfy (1.17), and suppose that is of class . Let satisfy (1.1), , and suppose that on . Let be as in (5.26) with chosen as in (8.30), and assume that belongs to . Let be as in (5.25).
Then, there exists a positive constant such that
(8.33) |
with the following specifications:
-
(i)
;
-
(ii)
is arbitrarily close to , in the sense that, for any sufficiently small, there exists a positive constant such that (8.33) holds with ;
-
(iii)
for .
The constant depends on , , , , and (the latter, only in the case ).
Proof.
Then, we modify the proof of Theorem 6.4 by using the Poincaré inequalities of Lemma 5.3 and Corollary 5.4 with on (instead of taking ).
By Lemma 8.13 we have that the domain satisfies the uniform interior sphere condition with radius , and hence Lemma 5.3 and Corollary 5.4 can be applied with . Also, in light of (8.32), we have that the choice in (8.30) for guarantees that the Poincaré inequalities of Corollary 5.4 can be applied with and .
From the previous work, we can now obtain the counterpart of Theorem 7.1 under the relaxed assumptions (1.16), (1.17):
Theorem 8.15 (General stability result for under relaxed assumptions).
Proof.
In the notation of (8.31) we have that , and, by recalling (1.3) and (8.29), is of class . Moreover, since satisfies the uniform interior sphere condition with radius , Lemma 5.2 can be applied to in place of . Hence, in this setting, formula (5.6) can be rephrased as
(8.34) |
Moreover, by the maximum principle,
(8.35) |
Hence, since
we deduce from (8.33), (8.34) and (8.35) that
Consequently, the desired result follows from (4.15) (with replaced by ) and (5.27). ∎
Theorem 8.16.
Let satisfy (1.1) and (1.2), and suppose that on . Let assumptions (1.16) and (1.17) be verified, and be as in (8.30).
Assume that there exists such that
and that .
Appendix A Motivation from an optimal heating problem (with possible malfunctioning of the source)
In this section we briefly recall how the simple model from optimal heating described in (2.2) directly produces the overdetermined condition in (2.3). For this, we consider a divergence free vector field . Also, for small , we introduce the diffeomorphism given by
We set and, given a source , we let be the solution of
We consider the energy functional
We also define
In this way, we have that
and
By the Hadamard’s Differentiation Formula (see Theorem 5.2.2 in [HP]), we know that
Since for all , we have that
for all , and so, taking derivatives in ,
(A.1) |
As a consequence,
and therefore
(A.2) |
We also remark that, since on , we have that
(A.3) |
Thus,
In this way, (A.2) can be written as
(A.4) |
We also observe that in , hence
This, (A.1) and (A.3) entail that
Consequently, (A.4) becomes
That is, being stationary for all divergence free vector fields is equivalent to the constancy of , that is (2.3).
Acknowledgements
The authors are supported by the Australian Research Council Discovery Project DP170104880 “N.E.W. Nonlocal Equations at Work” and are members of AustMS and INdAM/GNAMPA.
SD is supported by the DECRA Project DE180100957 “PDEs, free boundaries and applications”.