Abstract
We consider the exterior as well as the interior free-boundary Bernoulli problem associated with the infinity-Laplacian under a non-autonomous boundary condition. Recall that the Bernoulli problem involves two domains: one is given, the other is unknown. Concerning the exterior problem we assume that the given domain has a positive reach, and prove an existence and uniqueness result together with an explicit representation of the solution. Concerning the interior problem, we obtain a similar result under the assumption that the complement of the given domain has a positive reach. In particular, for the interior problem we show that uniqueness holds in contrast to the usual problem associated to the Laplace operator.
MSC:
35N25, 35B06, 35R35
1. Introduction
Bernoulli’s exterior problem consists in finding a couple where is a bounded domain ( an open, connected set) in containing a given compact set K, and u is a solution of
where a is a constant. Similarly, Bernoulli’s interior problem consists in finding such that the closure is included in a given, bounded domain , and v satisfies
Both problems have been widely investigated, and several generalizations have been taken into consideration: see, for instance, [,,,,,,,,] and the references therein. In particular, in [,,] the Laplace operator is replaced by the infinity-Laplacian. Roughly speaking, the infinity-Laplacian is the operator , where the subscripts denote differentiation with respect to , and the sum over repeated indices is understood. If we may also write where the subscript denotes differentiation in the direction of . However, the (viscosity) solution of the associated boundary-value problem
where denotes the prescribed boundary values, fails to have second derivatives in general. Furthermore, quoting ([], p. 238): “since the equation is not in divergence form, we cannot expect a notion of weak solution”. In fact, solutions are usually intended in the viscosity sense: a thorough presentation of the concept is found in []. For the present purposes it suffices to recall the following.
Definition 1.
Let Ω be a bounded domain in . A viscosity solution of problem (3) is a function agreeing with φ on and satisfying both of the following conditions at each interior point : for every function φ of class in a neighborhood of such that the difference has a local minimum at , the inequality is satisfied; for every function ψ of class in a neighborhood of such that the difference has a local maximum at , the inequality holds. It is worth noticing that if u does not allow the difference to have a local minimum at for any -function φ (think about ), then condition is trivially satisfied, and a similar remark holds for .
Example 1.
Let and . The function is the unique solution of in satisfying and on (the assertion follows by letting in Lemma 3). In particular, the origin is not a removable singularity as in the case of the Laplacian.
The equation , whose (viscosity) solutions are called infinity-harmonic, appears as the Euler-Lagrange equation of the minimal Lipschitz extension problem (see [,]). The name of infinity-harmonic is due to the fact that the solution of the boundary-value problem (3) can be seen as the limit, as , of the p-harmonic functions coinciding with u along the boundary. As usual, by a p-harmonic function we mean a weak solution of , where is the p-Laplace operator. Such an asymptotic representation also holds for the Bernoulli problem (see []). Nevertheless, when the domain is let to vary, the behavior of the infinity-harmonic functions may differ substantially from the one of the p-harmonic with finite p:
Example 2.
Choose and denote by the weak solution of in the annulus , , , satisfying when , and when . Let us focus on the boundary gradient along the inner boundary . A straightforward computation shows that
where the integral is elementary but takes two different expressions according to or . If the inner radius R tends to zero, the right-hand side tends to (more on this subject is found in ([], Section 3) for the special case ). By contrast, the infinity-harmonic function attaining the same boundary values as satisfies
Now, the right-hand side decreases and tends to 1 as . This difference reflects on the results obtained in the present paper for the interior Bernoulli problem: see Theorem 3 and the comments thereafter.
Concerning Bernoulli’s exterior problem, in the paper [], Manfredi, Petrosyan and Shahgholian proved the result quoted below. Denote by
the (Lipschitz continuous) distance function from a closed, nonempty subset , and let stand for the Minkowski sum
also called the tubular neighborhood of X of radius r.
Theorem 1
(cf. ([], Theorem 3.3)). If the compact, nonempty subset is convex then for every there exists a unique solution of Bernoulli’s problem
The solution is given by and .
In the present paper the result is extended in several directions. First, the convexity assumption on K is relaxed and replaced with the weaker assumption that K is a set with positive reach according to the following definition:
Definition 2.
Let be a closed subset of . Following [] we will use the notation
If for some there exists a unique such that , then we say that y is the projection of x onto X, and we write . A closed, nonempty set is a set with positive reach if there exists such that for all there exists a unique such that . The largest possible value of is called the reach of X and is denoted by (see Figure 1).
Figure 1.
The half-moon satisfies . The point has infinitely many nearest points .
A further extension lies in the fact that the Neumann condition in (4) is replaced here with the non-autonomous condition
where is a prescribed function that is required not to decrease too fast. In particular, is allowed to be a constant, hence condition (7) includes the Neumann condition in (4) as a special case. To be more precise, in Theorem 2 we consider Bernoulli’s exterior problem
where the domain is required to have a differentiable boundary and to satisfy
The boundary gradient of u occurring in (8) is well defined: indeed, since has a differentiable boundary, the infinity-harmonic function attaining constant values at the boundary is differentiable up to (see []). Contrary to what one may expect, if we allow q to be any function of the distance then problem (8) may well admit a solution although is not given by , as the following example shows.
Example 3.
Let , and let Ω be an ellipse in canonical position. Denote by the semi-axes of Ω, with . Clearly, Ω does not have the form : nevertheless, let us construct a function such that Bernoulli’s problem (8) is solvable. Recall that the boundary-value problem
admits a unique solution u (see, for instance, ([], Section 5)). Since problem (10) is invariant under reflections with respect to the coordinate axes, and by uniqueness, the equality holds for every and for every choice of the sign in front of the variables and in the right-hand side. Consequently, we also have for every , and we are allowed to define as follows: for each we first pick such that , then we let
Since the boundary gradient possesses the symmetry property mentioned before, the definition of does not depend on the particular choice of as long as , and therefore the definition is well posed. However, then Bernoulli’s problem (8) with this choice of q is solvable, although for every we have .
We prove that if does not decrease too fast, for instance if the product is strictly increasing, then problem (8) is solvable if and only if for a convenient , and the solution has the form (12):
Theorem 2 (On Bernoulli’s exterior problem).
Let be a compact, connected subset of , , with positive reach , and let be any real-valued function of one real variable.
- (i)
- (ii)
- (iii)
Observe that if Equation (11) has a solution , and the product is strictly increasing, then assertions (i) and (ii) apply. To see that Theorem 2 implies Theorem 1, recall that any compact, convex set is connected and satisfies (see ([], Corollary 4.6) or ([], p. 433)), hence (9) always holds. Furthermore the constant function clearly makes the product strictly increasing. Hence Claim (i) (existence) and Claim (ii) (uniqueness) of Theorem 2 imply the statement in Theorem 1. Theorem 2 also extends ([], Theorem 1.4), where problem (8) is considered in the special case when .
Next we consider the Bernoulli interior problem
where the complement of the given bounded domain is assumed to be a set with positive reach . For instance, cannot be a square in . The unknown domain , instead, is searched for in the class of all domains having a differentiable boundary and containing all points out of reach, i.e., must satisfy the inclusion
where is defined according to (6). For instance, if then and every domain containing the origin satisfies (14).
Theorem 3 (On Bernoulli’s interior problem).
Let be a bounded domain of , , whose complement is a set with positive reach. Define , and let be any real-valued function of one real variable.
- (i)
- (ii)
- (iii)
As before, if Equation (11) has a solution and the product is strictly increasing, then assertions (i) and (ii) apply. A corresponding result for the Laplace operator is illustrated in ([], Theorem 4.1) and in the subsequent ([], Example (1), p. 108). Remarkably, the monotonicity condition required there for the standard Laplacian (namely, non-increasing) excludes the case constant and is opposite to the one in Theorem 3. It is also to be recalled that the usual interior Bernoulli problem (2) lacks uniqueness of the solution. By contrast, if the Laplacian in (2) is replaced with the infinity-Laplacian, or equivalently if (constant) in (13), with , then the assumptions in Claim (i) and Claim (ii) of Theorem 3 are satisfied, and existence and uniqueness follow. These differences between and are related to the different behavior of the radial solutions which was put into evidence in Example 2.
The proofs of both Theorem 2 and Theorem 3 are given in Section 4, using Jensen’s comparison principle ([], Theorem 3.11). The explicit construction of prospective solutions is done in Section 3, and it is based on some fundamental properties of the distance function, which are in their turn recalled in the next section. The method of proof was introduced in [] in connection with the Laplacian, and it is a refinement of the approach in []. Further applications are found in [,,,,,].
2. Basic Properties of the Distance Function
The function measures the distance from the running point to a given nonempty closed subset . The properties of needed to prove Theorem 2 and Theorem 3 are found in [,,] (see also []). Here we collect the main statements under a unified notation, and give precise references to the sources. We start with the notion of proximal normal and proximal smoothness.
Definition 3.
(Cf. ([], Definitions 3.6.3 and 3.6.5), and ([], pp. 119–120)). Let be a proper subset of .
- (i)
- A unit vector is a perpendicular, or a proximal normal, shortly a P-normal, to X at if there exists such that .
- (ii)
- Any vector is also a P-normal at y if the unit vector is a P-normal at y in the sense given above. In this case we say that ζ is realized by an r-ball, where r is as before.
- (iii)
- Finally, the set X is proximally smooth with radius if for every and for every unit P-normal ν (if there exist any) at y we have . Equivalently, X is proximally smooth with radius if every P-normal is realized by an -ball.
From the definition it is clear that if X is proximally smooth with radius then X is also proximally smooth with radius r for every . Proximal smoothness can be considered equivalent to positive reach in the following sense:
Proposition 1.
Let be a closed, proper subset of .
- (i)
- If X is proximally smooth with radius then X is a set with positive reach and .
- (ii)
- If X is a set with positive reach then for every finite the function belongs to the class and X is proximally smooth with radius r.
Proof.
Suppose that X is proximally smooth with radius , and define according to (5). Let us check that every point has a unique projection onto X. Take , define and suppose, contrary to the claim, that there exist such that and for . By the definition of r, the open ball does not intersect X, hence the unit vector is a perpendicular to at for . Since X is proximally smooth with radius by assumption, the ball does not intersect X as well. However B contains together with all boundary points of excepted . In particular, B contains the point . However, since X is closed, we have which shows that B does intersect X: a contradiction. Hence every point must have a unique projection onto X and Claim follows.
Assume that X is a set with positive reach and choose a finite . By Definition 2, every has a unique projection onto X. Since the projection is well defined for all , by Claim of ([], Theorem 4.8) the distance function belongs to the class . By ([], Theorem 4.1 (a),(d)), this is equivalent to say that every P-normal is realized by an r-ball, hence X is proximally smooth with radius r. □
Claim (ii) of Proposition 1 implies that if X is a set with positive reach then for every the set has a boundary (in fact : see ([], Theorem 4.8, Claim (9))). The last assertion fails, in general, when :
Example 4.
The closed, unbounded set satisfies . The corresponding set equals the punctured space and does not have a differentiable boundary. For every the set defined in (6) satisfies , and therefore (see Corollary 1 for a general statement).
We now recall equality (17), which is essential for our purposes.
Lemma 1.
Let be a closed, proper subset of with positive reach. For every finite define and as in (5) , (6) , and take .
- (i)
- The projection is uniquely determined.
- (ii)
- The distance function , which is differentiable at by Proposition 1 , satisfies
- (iii)
- The set is not empty, and the following equality holds:
- (iv)
- The projection is uniquely determined, and the three points , , are aligned.
- (v)
- For every x on the segment whose endpoints are and we have
Proof.
The projection is uniquely defined for all because X is a set with positive reach.
Formula (16) is found, for instance, in ([], Corollary 3.4.5 ) as well as in ([], Theorem 1).
We first use part of Proposition 1 to see that X is proximally smooth with radius r. Then we follow the proof of ([], Theorem 3.6.7): in particular, formula (3.52) in [] corresponds to (17) above.
Choose such that . From (17) and the triangle inequality we get
However by definition (6) we also have , hence the triangle inequality holds with equality, and therefore the projection is uniquely determined and the three points are aligned, as claimed.
Observe that for every x on the segment whose endpoints are and we obviously have , with equality at . Let us check that the equality also holds for . Suppose, by contradiction, that there is such that . By the definition (6) of , the point satisfies . This and the triangle inequality imply
Since both and x lie on the segment whose endpoints are and , we may replace the right-hand side with . Thus, the inequality above becomes
which contradicts (17). Claim (v) follows, and the proof is complete. □
Corollary 1.
Let be a closed, proper subset of with positive reach. For every finite the set given by (6) is also a set with positive reach, and .
Proof.
The set is not empty by Lemma 1 (iii). In view of Definition 2, let us check that every point in has a unique projection onto . This follows from Lemma 1 (iv) provided we show that . We note in passing that the reverse inclusion follows immediately from (17). To prove that we expand and observe that , hence . It remains to verify that . To this aim, observe that for every and we have , hence and the conclusion follows. □
We conclude with a lemma that is needed in the proof of Claim of both Theorem 2 and Theorem 3, to manage the extremal case when .
Lemma 2.
Let be a closed, proper subset of , , with positive reach . If the open set has a differentiable boundary, then the distance function is differentiable at every , and (16) holds.
Proof.
The lemma follows from ([], Corollary 3.4.5 ), as well as from ([], Theorem 1) after having shown that every has a unique projection onto X. To simplify the notation, without loss of generality let and suppose that the inner normal to at 0 is the unit vector . We claim that the projection is uniquely determined, and it is given by . Let be any point on X that realizes . By definition of distance we have
By assumption, in a neighborhood of the boundary is the graph of a differentiable function such that , and the intersection lies above that graph. Letting and we may write
in a neighborhood of , with equality at the origin. Hence the right-hand side (say ) is minimal at and therefore its gradient must vanish. However, then the only possible value for is , and therefore the projection is uniquely determined, as claimed. □
3. Solutions in Parallel Sets
The proofs of Theorem 2 and Theorem 3 are based on a comparison with the particular solutions and that are constructed below.
Lemma 3.
Proof.
The uniqueness of the solution of (19) follows from the comparison principle in ([], Theorem 3.11). The boundary conditions are easily verified. Let us check that the equality
holds in the viscosity sense whenever . Since K is a set with positive reach, by Proposition 1 and by (16) the distance function is differentiable at and its gradient is the unit vector given by . Consequently the function defined in (12) is also differentiable at , and by differentiation we find
Concerning the second derivatives, since may fail to be of class in a neighborhood of we investigate its restriction to the line ℓ passing through and directed by . Define the set by letting and in (6), and notice that by Lemma 1 (iv) the three points , , are aligned, hence the line ℓ passes through all of them. Using Claim (v) of Lemma 1 we may write
Lemma 4.
Proof.
The argument is similar to the proof of Lemma 3. In the present case, for we find
where the unit vector is given by . We may write for every , where ℓ is the line passing through and directed by , and the proof proceeds as before. □
4. Proofs of Theorem 2 and Theorem 3
Proof of Theorem 2.
Claim (i). The boundary of the domain is differentiable for because is a level surface of the function , which is of class by Proposition 1 (ii) and has a nonvanishing gradient by (16). Let be given by (12). From Lemma 3 we know that is the unique solution of the boundary-value problem (19). To prove that the couple is a solution of Bernoulli’s exterior problem (8) it remains to check that the last condition there, namely condition (7), is satisfied for every . Observe that (7) reduces to for , i.e. for x satisfying . However, since is a solution of (11), we have to check that . In the case when , we know that the distance function is differentiable along and therefore (21) holds. If, instead, , then has a differentiable boundary by assumption, and (21) follows from Lemma 2. From (21) we get , as expected.
To prove Claim (ii), suppose that Bernoulli’s exterior problem (8) admits a solution where is a bounded domain satisfying the assumptions. Define
Assume, contrary to the claim, that . Define the parallel sets , , and consider the functions given by (12). Observe that . Since on as well as on , by the comparison principle ([], Theorem 3.11) it follows that on . However, then
Similarly, since along the boundary , we obtain
Let us consider a point , i.e., a point on such that . By (9) we also have , hence the boundary , which is the level set of the continuously differentiable function , is differentiable at and it is tangent to there. Since , and by (23), we deduce
where the last equality comes from (7). Thus, we have . Since Equation (11) has a unique solution by assumption, and the inequality holds for all finite , we deduce . Now we argue at a point . Notice that the function may fail to be differentiable at in case : indeed, although has a differentiable boundary, we have not proven that is a parallel set to K, yet, and therefore Lemma 2 is not applicable. To overcome this difficulty we let and in (17) and obtain for all . Hence writing in place of in (12) we get
Hence the gradient of u, which exists by assumption, must satisfy the estimate
and consequently , contradicting the assumption . Hence we must have , and Claim (ii) is followed by uniqueness (Lemma 3).
To prove Claim (iii) we suppose, by contradiction, that problem (8) is solvable, and show that Equation (11) has a solution , in contrast with the assumption. We follow the same argument as before. In the case when we arrive again at (24) and (25), hence the difference is non-negative at and non-positive at : a contradiction arises because q is continuous. If, instead, , then we may write where denotes the common value of . By uniqueness (Lemma 3), the alleged solution u must coincide with the function in (12). Since has a differentiable boundary, is differentiable along (by Lemma 2) and (21) holds. Hence for , which shows that Equation (11) still has a solution . The proof is complete. □
Proof of Theorem 3.
The argument is similar to the proof of Theorem 2, with minor differences. In particular, in the proof of Claim (i) we use Lemma 4 and (22) to show that the couple is a solution of problem (13). The conclusion also holds in case by Lemma 2 because the two sets and have the same boundary. To prove Claim (ii), denote by a solution of (13) and let
Define and for . Let us check that , or equivalently , where and are defined according to (5),(6). Suppose, by contradiction, that there exists . The segment joining to must intersect the boundary at some point z (possibly ). By Lemma 1 (v) we have , but this contradicts the definition of . Now let us check that , which is equivalent to . Suppose, by contradiction, that there exists . Now the set comes into play. Recall that is a set with positive reach by Corollary 1. By (14), the segment joining to must intersect the boundary at some point , and we have . Using (17), the last inequality leads to . However we also have because , hence we get in contrast with the definition of . In summary, we have
and by comparison we get
Now choose such that for . Assume, by contradiction, that . By (26) we obtain
Funding
The author is partially supported by the research project Integro-differential Equations and Non-Local Problems, funded by Fondazione di Sardegna (2017).
Acknowledgments
The author is a member of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM).
Conflicts of Interest
The author declares no conflict of interest.
References
- Bianchini, C. A Bernoulli problem with non constant gradient boundary constraint. Appl. Anal. 2012, 91, 517–527. [Google Scholar] [CrossRef]
- Bianchini, C.; Salani, P. Concavity properties for elliptic free boundary problems. Nonlinear Anal. Theor. Methods Appl. 2009, 10, 4461–4470. [Google Scholar] [CrossRef]
- Díaz, J.I.; Padial, J.F.; Rakotoson, J.M. On some Bernoulli free boundary type problems for general elliptic operators. Proc. R. Soc. Edinburgh A Math. 2007, 137, 895–911. [Google Scholar] [CrossRef]
- Flucher, M.; Rumpf, M. Bernoulli’s free-boundary problem, qualitative theory and numerical approximation. J. Reine Angew. Math. 1997, 486, 165–204. [Google Scholar]
- Greco, A.; Kawohl, B. On the convexity of some free boundaries. Interfaces Free Bound. 2009, 11, 509–514. [Google Scholar] [CrossRef]
- Hayouni, M.; Henrot, A.; Samouh, N. On the Bernoulli free boundary problem and related shape optimization problems. Interfaces Free Bound. 2001, 3, 1–13. [Google Scholar] [CrossRef]
- Henrot, A.; Shahgholian, H. Existence of classical solutions to a free boundary problem for the p-Laplace operator. I. The exterior convex case. J. Reine Angew. Math. 2000, 521, 85–97. [Google Scholar] [CrossRef]
- Henrot, A.; Shahgholian, H. Existence of classical solutions to a free boundary problem for the p-Laplace operator. II. The interior convex case. Indiana Univ. Math. J. 2000, 49, 311–323. [Google Scholar] [CrossRef]
- Henrot, A.; Shahgholian, H. The one phase free boundary problem for the p-Laplacian with non-constant Bernoulli boundary condition. Trans. Am. Math. Soc. 2002, 354, 2399–2416. [Google Scholar] [CrossRef]
- Crasta, G.; Fragalà, I. Bernoulli free boundary problem for the infinity laplacian. arXiv, 2018; arXiv:1804. [Google Scholar]
- Greco, A. Constrained radial symmetry for the infinity-Laplacian. Nonlinear Anal. Real World Appl. 2017, 37, 239–248. [Google Scholar] [CrossRef]
- Manfredi, J.; Petrosyan, A.; Shahgholian, H. A free boundary problem for ∞-Laplace equation. Calc. Var. Partial Differ. Equ. 2002, 14, 359–384. [Google Scholar] [CrossRef]
- Buttazzo, G.; Kawohl, B. Overdetermined boundary value problems for the ∞-Laplacian. Int. Math. Res. Not. 2011, 2, 237–247. [Google Scholar] [CrossRef]
- Crandall, M.G.; Ishii, H.; Lions, P.-L. User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc. (N.S.) 1992, 27, 1–67. [Google Scholar] [CrossRef]
- Crandall, M.G. A visit with the ∞-Laplace equation. In Calculus of Variations and Non-Linear Partial Differential Equations; Lecture Notes in Mathematics 1927; Dacorogna, B., Marcellini, P., Eds.; Springer: Berlin/Heidelberg, Germany, 2008; pp. 75–122. [Google Scholar]
- Jensen, R. Uniqueness of Lipschitz extensions: minimizing the sup norm of the gradient. Arch. Ration. Mech. Anal. 1993, 123, 51–74. [Google Scholar] [CrossRef]
- Clarke, F.H.; Stern, R.J.; Wolenski, P.R. Proximal smoothness and the lower-C2 property. J. Convex Anal. 1995, 2, 117–144. [Google Scholar]
- Hong, G. Boundary differentiability of infinity harmonic functions. Nonlinear Anal. 2013, 93, 15–20. [Google Scholar] [CrossRef]
- Federer, H. Curvature measures. Trans. Amer. Math. Soc. 1959, 93, 418–491. [Google Scholar] [CrossRef]
- Greco, A. Radial symmetry and uniqueness for an overdetermined problem. Math. Methods Appl. Sci. 2001, 24, 103–115. [Google Scholar] [CrossRef]
- Henrot, A.; Philippin, G.A.; Prébet, H. Overdetermined problems on ring shaped domains. Adv. Math. Sci. Appl. 1999, 9, 737–747. [Google Scholar]
- Babaoglu, C.; Shahgholian, H. Symmetry in multi-phase overdetermined problems. J. Convex Anal. 2011, 18, 1013–1024. [Google Scholar] [CrossRef]
- Greco, A. An overdetermined problem for the infinity-Laplacian around a set of positive reach. Analysis (München) 2018, 38, 155–165. [Google Scholar] [CrossRef]
- Greco, A. Comparison principle and constrained radial symmetry for the subdiffusive p-Laplacian. Publ. Mat. 2014, 58, 485–498. [Google Scholar] [CrossRef]
- Greco, A. Constrained radial symmetry for monotone elliptic quasilinear operators. J. Anal. Math. 2013, 121, 223–234. [Google Scholar] [CrossRef]
- Henrot, A.; Philippin, G.A. Approximate radial symmetry for solutions of a class of boundary value problems in ring-shaped domains. J. Appl. Math. Phys. (ZAMP) 2003, 54, 784–796. [Google Scholar] [CrossRef]
- Cannarsa, P.; Sinestrari, C. Semiconcave Functions, Hamilton-Jacobi Equations, and Optimal Control; Birkhäuser: Basel, Switzerland, 2004. [Google Scholar]
- Crasta, G.; Fragalà, I. On the characterization of some classes of proximally smooth sets. ESAIM Control Optim. Calc. Var. 2016, 22, 710–727. [Google Scholar] [CrossRef]
- Ambrosio, L. Geometric evolution problems, distance function and viscosity solutions. In Calculus of Variations and Partial Differential Equations; Buttazzo, G., Marino, A., Murthy, M.K.V., Eds.; Springer: Berlin/Heidelberg, Germany, 2000; pp. 5–93. [Google Scholar]
© 2019 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).