Home The best approximation of a given function in L 2-norm by Lipschitz functions with gradient constraint
Article Open Access

The best approximation of a given function in L 2-norm by Lipschitz functions with gradient constraint

  • Stefano Buccheri ORCID logo EMAIL logo , Tommaso Leonori and Julio D. Rossi
Published/Copyright: April 24, 2024

Abstract

The starting point of this paper is the study of the asymptotic behavior, as p , of the following minimization problem:

min { 1 p Ω | v | p + 1 2 Ω ( v - f ) 2 , v W 1 , p ( Ω ) } .

We show that the limit problem provides the best approximation, in the L 2 -norm, of the datum f among all Lipschitz functions with Lipschitz constant less or equal than one. Moreover, such an approximation verifies a suitable PDE in the viscosity sense. After the analysis of the model problem above, we consider the asymptotic behavior of a related family of nonvariational equations and, finally, we also deal with some functionals involving the ( N - 1 ) -Hausdorff measure of the jump set of the function.

MSC 2020: 35J92; 35J94; 35J60

1 Introduction

Let us assume that Ω N , N 1 , is a bounded open set with smooth boundary, say C 1 , γ , for some γ ( 0 , 1 ) . The main goal of this work is to study the optimal approximation in the L 2 -norm of a given function f L 2 ( Ω ) by functions in W 1 , ( Ω ) with a constraint on the gradient. Namely, we consider the minimization problem

(1.1) inf 𝕏 ( Ω ) J ( v ) with  J ( v ) = 1 2 Ω ( v - f ) 2 ,

where the set 𝕏 ( Ω ) is given by

𝕏 ( Ω ) = { v W 1 , ( Ω ) : | v | 1  for a.e.  x Ω } .

In order to obtain a minimizer to (1.1), one can argue by direct methods (taking a minimizing sequence), or rather noticing that such a problem appears naturally as Γ-limit, for n , of the following classical energy functional

(1.2) J n ( v ) = 1 p n Ω | v | p n + 1 2 Ω ( v - f ) 2 , v W 1 , p n ( Ω ) ,

where p n is a (non-decreasing) sequence of numbers that diverges to + . Setting u n the minimizer of J n , the heuristic of the limiting process is that the measure of the set { | u n | > 1 } has to go to zero in order to keep the energy bounded (notice that J n ( u n ) J n ( 0 ) = f L 2 ( Ω ) 2 ). Therefore, we expect the sequence { u n } to converge to the unique minimizer u 𝕏 ( Ω ) of J (see Proposition 2.2 below).

An interesting feature of the minimizer of (1.1) that we find here, is that it satisfies a kind of representation formula by cones inside the region { u f } .

Figure 1
Figure 1
Figure 2
Figure 2

Before stating our first result, in order to gain some intuition, let us consider three concrete examples in the special case Ω = ( - 1 , 1 ) .

  1. Let f 1 ( x ) = k χ ( - r , r ) with r ( 0 , 1 ] and k . Observe that if r = 1 , then f 1 𝕏 ( - 1 , 1 ) and so u 1 f 1 is the minimizer. Otherwise, it is not hard to see that

    u 1 = min { ( r + k 2 - | x | ) + , k }

    is the minimizer. In Figure 1 the case r > k 2 is represented.

  2. Let f 2 ( x ) = 2 | x | . Then the minimizer is given by

    u 2 ( x ) = | x | + 1 2 .

    See Figure 2.

  3. Let f 3 ( x ) = | x | . Then it is not hard to see that the minimizer is

    u 3 ( x ) = { | x | + 2 9 for  | x | 4 9 , | x | for  | x | 4 9 .

    See Figure 3.

In the three examples above (see Section 6 for the explicit computations) the behavior of the solution u i , i = 1 , 2 , 3 follows the same idea: any u i tries to be as close as possible to the datum f i as long as f i is smooth and has gradient bounded by one. Otherwise, the best that the solution can do, in order to minimize the L 2 norm of the difference, is to grow as much as it is allowed, that is, as a line with slope + 1 or -1.

Figure 3
Figure 3

More in general we have the following result.

Theorem 1.1.

Let f L 2 ( Ω ) , let { u n } W 1 , p n ( Ω ) be the sequence of minimizers of J n (see (1.2)), and let u be the unique minimizer of (1.1). Then

u n u uniformly in  Ω ¯ .

In addition, the following representations formulas hold true:

(1.3)

u ( x ) = max y A + [ u ( y ) - d ( x , y ) ] for all  x A + ,
u ( x ) = min y A - [ u ( y ) + d ( x , y ) ] for all  x A - ,

where

(1.4) A - = supp ( u - f ) - , A + = supp ( u - f ) + ,

and d ( x , y ) denotes the geodesic distance inside Ω (see Definition (1.22) below).

After considering the problem from a variational viewpoint, we want to address it from a PDE perspective. Since it is not clear a priori what equation the minimizer of (1.1) solves, we go back to the approximating sequence of minimizers of J n (see (1.2)). The minimizers of (1.2) are solutions to

(1.5) { u n ( x ) - Δ p n u n ( x ) = f ( x ) in  Ω , u n ν ( x ) = 0 on  Ω ,

where Δ p v = div ( v | v | p - 2 ) , with p > 1 , is the p-Laplacian and we denote by ν the external unit normal to Ω . As usual, we consider such solutions in the weak sense (see Definition 2.1). Our aim is, now, to pass to the limit in (1.5) to find the PDE solved by the limit function, u.

Typically, the limit as p of solutions to equations related to the p-Laplacian leads to equations that involve the infinity Laplacian, namely,

Δ u = D 2 u u , u = i , j = 1 N u x i x j u x i u x j .

In particular, we refer to the pioneer paper by T. Bhattacharya, E. Di Benedetto and J. Manfredi (see [9]) that first studied such type of problem. We recall that the infinity Laplacian is a second order differential operator (in nondivergence form) that appears in many contexts. For instance, infinity harmonic functions (i.e., solutions to - Δ u = 0 ) appear naturally as limits of p-harmonic functions (i.e., solutions to - Δ p u = - div ( | D u | p - 2 D u ) = 0 ) and they have applications to optimal transport problems, image processing, etc. See, among the others, [4, 9, 24] and references therein. Moreover, the infinity Laplacian plays a fundamental role in the calculus of variations of L functionals, see e.g. [3, 6, 7, 17, 18, 19, 20, 26] and the survey [4]. Notice that such operator is degenerate elliptic (that is non-degenerate only in the direction of the gradient).

Before stating our next result, let us introduce some notations: here and in the rest of the paper for u , f C ( Ω ¯ ) we denote

Ω u + = { x Ω : u ( x ) - f ( x ) > 0 } and Ω u - = { x Ω : u ( x ) - f ( x ) < 0 }

and analogously,

( Ω ) u + = { x Ω : u ( x ) - f ( x ) > 0 } and ( Ω ) u - = { x Ω : u ( x ) - f ( x ) < 0 } .

For the sake of clarity, notice that in Theorem 1.2 f is continuous and the open set Ω u + coincides with the interior of A + .

Now we are ready to write the problem solved by u.

Theorem 1.2.

Assume f C ( Ω ¯ ) , then the unique minimizer u X ( Ω ) of (1.1) is a viscosity solution to

(1.6) { 1 - | u | = 0 in  Ω u + , | u | - 1 = 0 in  Ω u - .

Moreover, it satisfies the following boundary conditions:

(1.7) { 1 - | u | 0 on  ( Ω ) u + , max { 1 - | u | , u ν } 0 on  ( Ω ) u + , 𝑎𝑛𝑑 { | u | - 1 0 on  ( Ω ) u - , min { | u | - 1 , u ν } 0 on  ( Ω ) u - .

Quite surprisingly, in the limit the second order operator is somehow lost: indeed there is no trace of the infinity Laplacian and, in fact, u satisfies a first order equation. We will provide an explanation for the absence of the second order operator in (1.6) in Proposition 1.6 below. Let us stress here that the two equations in (1.6) are not equivalent from the point of view of viscosity solutions and, indeed, they are the proper ones to reflect the information carried out by (1.3): in the region { u - f < 0 } the function u looks like a cone pointing upwards and therefore it solves the eikonal equation; analogously, in { u - f > 0 } the cone points downwards and it solves the eikonal equation with the reverse sign. The same double structure is recognizable in the boundary conditions (1.7).

As far as the boundary conditions are concerned, it is natural to wonder if it would be possible to extend the equation in (1.6) up to Ω ¯ . However the answer is negative, and, to get convinced that (1.7) are the natural conditions to impose on Ω , it is enough to look at the second example above. Indeed, observe that in this case Ω = { ± 1 } { u 2 - f < 0 } , so we have to focus on the second pair of conditions in (1.7). If x 0 = ± 1 , we have that u 2 satisfies (in viscosity sense) | u 2 ( x 0 ) | - 1 0 (when u 2 is touched from below by any smooth function φ, it has to be | φ ( x 0 ) | 1 ). However, when we test the behavior of u 2 touching it from above by a smooth function φ, it is easy to check that it is not necessarily true that | φ ( x 0 ) | 1 ; but if | φ ( x 0 ) | > 1 then ν φ ( x 0 ) < 0 . This is exactly the condition prescribed by (1.7).

It is also worth to mention explicitly that the regions Ω u + and Ω u - (and their boundary counterparts) depend on the solution u itself. A consequence of this fact is that, even if the minimizer of problem (1.1) is unique, problem (1.6)–(1.7) does not posses a unique solution. Indeed, any affine function with modulus of the gradient equal to 1 that lies above or below f solves (1.6)–(1.7). Clearly, this type of solutions are not interesting for our study. However, it is still possible to construct a nontrivial solution to (1.6)–(1.7), that does not coincide with the minimizer of J , via a suitable obstacle problem.

Theorem 1.3.

Assume that f C ( Ω ¯ ) . Then among all z X ( Ω ) solutions to the obstacle problem

(1.8) { z f in  Ω , 1 - | z | 0 in  Ω , 1 - | z | = 0 in  Ω z + ,

with boundary conditions

(1.9) { 1 - | z | 0 on  ( Ω ) z + , max { 1 - | z | , z ν } 0 on  ( Ω ) z + ,

there exists a minimal one u ¯ X ( Ω ) that also solves (1.6)–(1.7).

Observe that u ¯ is not a solution of the minimization problem (1.1) (unless f 𝕏 ( Ω ) ). Indeed, as the intuition suggests, since the minimizer u has to be close to f in L 2 , then u - f has to change sign in Ω (see Figure 4 and Proposition 2.4 for more details). Considering the reverse inequalities in the obstacle problem above, via a similar procedure, it is possible to construct another nonvariational solution u ¯ 𝕏 ( Ω ) , such that u ¯ u u ¯ .

Figure 4
Figure 4

The limit procedure showed above can be, in fact, generalized to a larger class of differential equations that involve lower order terms that have a polynomial growth with respect to the gradient. More precisely, we consider the following problem:

(1.10) { v n ( x ) - Δ p n v n ( x ) = H n ( x , v n ( x ) ) + f ( x ) in  Ω , v n ν ( x ) = 0 on  Ω ,

where the Hamiltonian term H n C ( Ω ¯ × N ) satisfies the following growth conditions:

(1.11) K 1 , K 2 > 0 : | H n ( x , ξ ) | p n K 1 | ξ | p n + K 2 for all  x Ω  and all  ξ N

and

(1.12) H C ( Ω ¯ × N ) : lim n H n ( x , ξ ) p n | ξ | p n - 4 = H ( x , ξ ) uniformly with respect to  x Ω .

Remark 1.4.

Notice that the second assumption in (1.12) implies that

(1.13) lim n H n ( x , ξ ) = 0 for any  | ξ | < 1 , uniformly with respect to  x Ω .

Among all the possible choices of H n that satisfy (1.11), let us mention two important examples:

H n ( x , ξ ) = p n | ξ | p n - 2 ξ B ( x ) with  B ( C ( Ω ¯ ) ) N    and    H n ( x , ξ ) = p n | ξ | p n .

These two types of lower order terms, accounting for the “linear ( p - 1 ) growth” and “natural growth”, are thoroughly studied in literature (for fixed p). Typically, the presence of the Hamiltonian affects the coercivity of the operator and some specific techniques are required to obtain a priori estimates: the first case can be dealt with symmetrization procedures [8], slicing techniques [15] or arguing by contradiction [10]. For the case of natural growth we mention the exponential test function approach developed in [11, 12, 13], again symmetrization in [1], and an argument based on Schauder fixed point Theorem in [29].

Let us notice that, under assumption (1.11), for a smooth f and a fixed n, the existence of a solution v n W 1 , p n ( Ω ) L ( Ω ) to problem (1.10) is a straightforward consequence of the results contained in [13] (see Theorem 3.1 below for more details). We also explicitly point out that the presence of the zeroth lower order term allows us to solve (1.10) without any smallness assumption on the datum f.

The main difficulty when one wants to pass to the limit as p diverges in (1.10) is to obtain a priori estimates for v n that are stable with respect to n. Up to our knowledge, the only related result for quasilinear equations with gradient lower order terms is contained in [16], where a large solution problem is considered.

Our main result concerning problem (1.10) is the following.

Theorem 1.5.

Assume (1.11)–(1.12) and let f C ( Ω ¯ ) . For any n N , let v n W 1 , p n ( Ω ) L ( Ω ) be a solution to (1.10). Then there exists v W 1 , ( Ω ) L ( Ω ) such that

v n v uniformly in  Ω ¯

which solves (in viscosity sense) the equation

(1.14) { max { 1 - | v | , - Δ v - H ( x , v ) } = 0 in  Ω v + , min { | v | - 1 , - Δ v - H ( x , v ) } = 0 in  Ω v - .

together with the boundary conditions

(1.15) { min { max { 1 - | v | , - Δ v - H ( x , v ) } , v ν } 0 on  ( Ω ) v + , max { max { 1 - | v | , - Δ v - H ( x , v ) } , v ν } 0 on  ( Ω ) v + ,

and

(1.16) { min { min { | v | - 1 , - Δ v - H ( x , v ) } , v ν } 0 on  ( Ω ) v - , max { min { | v | - 1 , - Δ v - H ( x , v ) } , v ν } 0 on  ( Ω ) v - .

At the first glance we notice that the structure of (1.14) is more involved then the one in (1.6). First of all v does not belong in general to 𝕏 ( Ω ) since the bound on v depends also on the Hamiltonian (see Section 3). Moreover, as expected, the presence of the Hamiltonian term changes both the limit equation and the boundary conditions.

Let us now go back to the minimizer u 𝕏 ( Ω ) of (1.1) and notice that it also solves equation (1.14) with H 0 . Apparently, this provides some more information about - Δ u that is missing in (1.6). However, we have the following result.

Proposition 1.6.

Assume that f C ( Ω ¯ ) , and that w X ( Ω ) . Then w solves (1.6) if and only if it solves

{ max { 1 - | w | , - Δ w } = 0 in  Ω w + , min { | w | - 1 , - Δ w } = 0 in  Ω w - .

Roughly speaking this is due to the bound | w | 1 and the fact that cones are solutions to the eikonal equation and are also infinity harmonic away from their singular points.

Finally, we include some considerations about a very well-known family of functionals that allow free discontinuities. Unfortunately, we are able to provide only partial results in this case and we left a more complete analysis of the limiting behavior of such functionals for p for future research. Let us consider the functional p : SBV ( Ω ) { + } defined as

(1.17) p ( v ) = 1 p Ω | v | p + 1 r Ω | v - f | r + N - 1 ( S v ) ,

with r 1 . Here SBV ( Ω ) is the space of Special Functions of Bounded Variation, v is the absolutely continuous part of Dv with respect to the Lebesgue measure, S v is the singular set of v and N - 1 is the ( N - 1 ) -dimensional Hausdorff measure (more details and formal definitions can be found in Section 5). We set p ( v ) = + when v L p ( Ω ) .

The functional p ( v ) was introduced in the seminal paper by De Giorgi, Carriero and Leaci ([22]) to provide a weak formulation of the Mumford–Shah image segmentation problem, namely, minimizing

(1.18) F p ( K , v ) = 1 p Ω | v | p + 1 r Ω | v - f | r + N - 1 ( K ) ,

among all K Ω ¯ closed sets, with v W 1 , p ( Ω K ) . An in-depth overview about the history of these functionals, their relation with applications, and the huge impact that their study has had in the field of Calculus of Variations is definitively out of reach for our contribution. We refer the interested reader to the original paper [22], the classical monograph [2], and to the nice review [25]. For our aims, we simply recall that the main difficulty in dealing directly with (1.18) is that the Hausdorff measure is not lower semicontinuous with respect to any reasonable metric the collection of closed set of Ω can be equipped with. Therefore, the strategy of [22] was to find a minimizer u SBV ( Ω ) of p and then show that ( S u ¯ , u ) was also a minimizer to (1.18). The most delicate step in their argument was to prove that N - 1 ( S u ¯ Ω ) = N - 1 ( S u ) and this crucial step was achieved by a lower bound on the density of S u , namely

(1.19) θ > 0 : N - 1 ( S u B ρ ( x ) ) > θ ρ N - 1 for all  x S u ¯  and all  ρ > 0

(see [2, Theorems 7.15, 7.21 and 7.22]).

At this point it is natural to wonder what happens with the limiting minimization problem relative to (1.18) as n . The answer is given by the following proposition.

Proposition 1.7.

Assume f L ( Ω ) and p n + . Then the sequence of functionals I p n : SBV ( Ω ) R { + } defined in (1.18) Γ-converges to I : SBV ( Ω ) R { + } defined as

(1.20) ( v ) = { 1 r Ω | v - f | r + N - 1 ( S v ) if  | v | 1 , + 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒 .

The next step would be to prove that a minimizer of (1.20) is also a minimizer to

(1.21) F ( K , u ) = 1 r Ω | u - f | r + N - 1 ( K ) ,

with K Ω ¯ a closed set, u W 1 , ( Ω K ) and | u | 1 . Unfortunately, we are not able to show such a property since we do not know how to obtain a lower density bound, as in (1.19), for the limit function u. We leave this as an interesting open problem.

Notations and basic definitions

Let us introduce the notion of geodesic distance (see [30], for instance): for x , y Ω we define

(1.22)

Γ = { γ x , y C 1 ( [ 0 , 1 ] ; Ω ) : γ ( 0 ) = x , γ ( 1 ) = y } ,
d ( x , y ) = inf γ Γ { ( γ x , y ) } , where  ( γ x , y ) = length ( γ x , y ) .

Of course, if Ω is convex, then d ( x , y ) | x - y | , but for general domains we just have that d ( x , y ) | x - y | . The geodesic distance is the natural quantity to consider in our setting since the condition

v 𝕏 ( Ω ) = { v W 1 , ( Ω ) : | D v | 1  for a.e.  x Ω }

can be rephrased as

v { v W 1 , ( Ω ) : | v ( x ) - v ( y ) | d ( x , y )  for all  x , y Ω } .

Now, let us introduce some extra notations that will be used in the rest of the paper. Let us recall that (see [14, Proposition 4.17]) the support of a general function f : Ω is defined as

supp ( f ) = ( ω i I ω i ) c , where  I = { ω  open subset of  Ω : f = 0  a.e. in  ω } .

By definition, supp ( f ) is a closed set with respect to the subset topology relative to Ω. For any f L 2 ( Ω ) , and for any continuous function u we define the following sets:

(1.23) A - = supp ( u - f ) - , A + = supp ( u - f ) + .

For future use we need to give the precise definition of a solution to a partial differential equation in the viscosity sense.

Definition 1.8.

Let 𝒪 be a locally compact subset of and let F : 𝒪 × × × 𝒮 ( N ) be a continuous function that satisfies

F ( x , r , ξ , X ) F ( x , s , ξ , Y ) whenever  r s  and  Y X .

We say that u C ( 𝒪 ) is a viscosity sub-solution (super-solution) to

(1.24) F ( x , u , u , D 2 u ) = 0 in  𝒪 ,

and we write

F ( x , u , u , D 2 u ) ( )  0 in  𝒪 ,

if, for any C 2 ( 𝒪 ) function φ such that u - φ has a strict local maximum (minimum) at x 0 𝒪 , it holds true

F ( x 0 , φ ( x 0 ) , φ ( x 0 ) , D 2 φ ( x 0 ) ) ( )  0 .

Finally, u C ( 𝒪 ) is a viscosity solution of (1.24) if it is both sub- and super-solution.

Let us denote by δ ( x ) = dist ( x , Ω ) and let us recall that δ ( x ) W 0 1 , ( Ω ) and that it satisfies (in the viscosity sense) the eikonal equation

| δ | - 1 = 0 in  Ω .

In the sequel we will use the following result.

Proposition 1.9.

Any function in X ( Ω ) satisfies both

(1.25) | u | - 1 0 𝑎𝑛𝑑 - | u | + 1 0 in  Ω

in the viscosity sense.

Proof of Proposition 1.9.

Let us observe that (1.25) is equivalent to say that the following inequalities hold true at any x 0 Ω :

  1. for any φ C 2 ( Ω ) such that there exists r > 0 with u ( x ) - φ ( x ) u ( x 0 ) - φ ( x 0 ) in B r ( x 0 ) Ω , we have that | φ ( x 0 ) | - 1 0 ,

  2. for any φ C 2 ( Ω ) such that there exists r > 0 with u ( x ) - φ ( x ) u ( x 0 ) - φ ( x 0 ) in B r ( x 0 ) Ω , we have that - | φ ( x 0 ) | + 1 0 .

In order to prove the first inequality in (1.25) let x 0 Ω and assume that there exists φ, a C 2 function of a neighborhood of x 0 , such that u - φ has a strict local maximum at x 0 , with u ( x 0 ) = φ ( x 0 ) . Suppose by contradiction that | φ ( x 0 ) | > 1 . Taking the first order Taylor expansion of φ ( x ) at x 0 it follows that

φ ( x ) - φ ( x 0 ) = φ ( x 0 ) ( x - x 0 ) + o ( | x - x 0 | ) as  x x 0 .

Let us consider

x t = - φ ( x 0 ) | φ ( x 0 ) | t + x 0 for  t > 0  small .

We have | x t - x 0 | = t and thus

u ( x t ) - u ( x 0 ) φ ( x t ) - φ ( x 0 ) = φ ( x 0 ) ( x t - x 0 ) + o ( | x t - x 0 | ) = - | φ ( x 0 ) | t ( 1 + o ( 1 ) ) as  t 0 .

This contradicts u 𝕏 ( Ω ) .

The second inequality in (1.25) follows in the same way. ∎

2 The best Lipschitz approximation with gradient constraint

Let p n be an increasing sequence such that p n as n . Let us set, for any f L 2 ( Ω ) , the following functional

(2.1) J n ( v ) = 1 p n Ω | v | p n + 1 2 Ω ( v - f ) 2 , v W 1 , p n ( Ω ) .

For any fixed n, it is not hard to see that there exists a unique (exploiting the convexity of (2.1)) minimizer u n W 1 , p n ( Ω ) to (2.1). Thus, from direct variational arguments, the minimizer, u n , turns out to be a weak solution to

(2.2) { u n ( x ) - Δ p n u n ( x ) = f ( x ) in  Ω , u n ν ( x ) = 0 on  Ω .

Let us recall here such a formulation.

Definition 2.1.

A weak solution u n to (2.2) is a W 1 , p n ( Ω ) function that satisfies

(2.3) Ω u n φ + Ω | u n | p n - 2 u n φ = Ω f φ for all  φ W 1 , p n ( Ω ) .

The aim of this section is to deal with the limiting behavior of these three objects – the minimizer, the functional, and the equation – as p n diverges.

We start with the following proposition that provides a limit for the sequence of minimizers together with a limiting minimization problem associated to (2.1).

Proposition 2.2.

Assume f L 2 ( Ω ) and let { u n } W 1 , p n ( Ω ) be the sequence of minimizers of (2.1). Then there exists u X ( Ω ) such that

u n u uniformly in  Ω ¯ .

Moreover, u is the unique minimizer of the functional

(2.4) J ( v ) = 1 2 Ω ( v - f ) 2 , v 𝕏 ( Ω ) .

Proof.

The minimality of u n gives us that

1 2 Ω ( u n - f ) 2 + 1 p n Ω | u n | p n J n ( v ) J n ( 0 ) = 1 2 Ω | f | 2 ,

where J n ( v ) has been defined in (2.1). This implies that

(2.5) u n L 2 ( Ω ) f L 2 ( Ω ) and u n L p n ( Ω ) ( p n f L 2 ( Ω ) 2 ) 1 p n .

For N < p n , let us recall that Morrey inequality (see [14, Corollary 9.14]) implies that, for any x , y Ω ,

(2.6) | u n ( x ) - u n ( y ) | C M u n L p n ( Ω ) ,

where C M = C M ( N , Ω ) can be chosen independent on p (see in particular [14, formula (28) in the proof of Theorem 9.12]). Therefore, we deduce that

(2.7) | u n ( x ) | | u n ( y ) | + C M u n L p n ( Ω ) for all  x , y Ω ,

and consequently

(2.8) u n L ( Ω ) | Ω | 1 2 u n L 2 ( Ω ) + C ~ M u n L p n ( Ω ) .

Combining (2.5) with (2.8), we conclude that

(2.9) u n L ( Ω ) | Ω | 1 2 f L 2 ( Ω ) + c Ω f L 2 ( Ω ) 2 p n ,

where c Ω = e 1 e C ~ M . On the other hand, using Hölder’s inequality, it follows that, for any q such that N < q < p n ,

u n L q ( Ω ) u n L p n ( Ω ) | Ω | 1 q - 1 p n p n 1 p n f L 2 ( Ω ) 2 p n | Ω | 1 q - 1 p n .

Thanks to the two estimates above, we deduce that the sequence { u n } is bounded in W 1 , q ( Ω ) and then up to a (not relabeled) subsequence, u n u weakly in W 1 , q ( Ω ) and uniformly in Ω. The lower semicontinuity of the norm also implies that for any q 1 we have

u L q ( Ω ) | Ω | 1 q .

Therefore we have that u L q ( Ω ) max { 1 , | Ω | } and we deduce that | u | L ( Ω ) . Passing to the limit as q in the formula above, we conclude that u 𝕏 ( Ω ) .

In order to show that u is, in fact, a minimizer to (2.4), let us notice that, by definition, u n satisfies

1 2 Ω ( u n - f ) 2 1 2 Ω ( u n - f ) 2 + 1 p n Ω | u n | p n 1 2 Ω ( z - f ) 2 + 1 p n Ω | z | p n for all  z 𝕏 ( Ω ) W 1 , p n ( Ω ) .

Since | z | 1 a.e. in Ω, we get

1 2 Ω ( u n - f ) 2 1 2 Ω ( z - f ) 2 + 1 p n | Ω | for all  z 𝕏 ( Ω ) W 1 , p n ( Ω ) .

We can now pass to the limit in the previous inequality (recall that u n u uniformly in Ω), and we deduce that u is a minimizer. The uniqueness follows by the convexity of J . ∎

Remark 2.3.

When f L ( Ω ) , we have a explicit uniform bound for u n ,

| u n | f L ( Ω ) = k ¯ .

Indeed, T k ¯ ( u n ) is a competitor for u n for the energy (2.1) and the uniqueness of the minimizer of J n implies u n T k ¯ ( u n ) .

In other words, the minimizer of (2.4) is the best approximation of f among all the functions in 𝕏 ( Ω ) . The gradient constraint imposes on the minimizer a specific geometric structure. The goal of the following results is to clarify such a structure.

Proposition 2.4.

Assume that f L 2 ( Ω ) and let u X ( Ω ) be the minimizer to (2.4). Then

| A - | | A + | = 0 if and only if f 𝕏 ( Ω ) .

Proof.

We observe that if f 𝕏 ( Ω ) , then u f and hence | A - | = 0 and | A + | = 0 .

Assume now that | A - | | A + | = 0 and by contradiction suppose that f L 2 ( Ω ) 𝕏 ( Ω ) . If both | A - | and | A + | are zero, then u = f , that implies f 𝕏 ( Ω ) , that is a contradiction. Let us deal with the case | A - | > 0 and | A + | = 0 (the case | A + | > 0 and | A - | = 0 is analogous). Then, for any ε > 0 , u + ε 𝕏 ( Ω ) satisfies

J ( u + ε ) = 1 2 Ω ( u + ϵ - f ) 2 = 1 2 Ω ( u - f ) 2 + ε Ω ( u - f ) + 1 2 ε 2 | Ω | = J ( u ) + 1 2 ε 2 | Ω | + ε A - ( u - f ) = J ( u ) + 1 2 ε ( ε | Ω | - 2 A - | u - f | ) ,

and the term in the brackets in the last line is negative for ε small enough. This contradicts the minimality of u. ∎

Remark 2.5.

In fact, the proof shows that

| A - | = 0  and  | A + | = 0 f 𝕏 ( Ω ) .

However, note that the statement of the proposition highlights that it is enough to have | A - | = 0 or | A + | = 0 to obtain that u = f 𝕏 ( Ω ) .

Next we deduce a representation formula for the unique minimizer to (2.4).

Theorem 2.6.

Assume f L 2 ( Ω ) and let u be the unique minimizer to (2.4). Then the following representations formulas hold true:

(2.10) u ( x ) = max y A + [ u ( y ) - d ( x , y ) ] for all  x A +

and

(2.11) u ( x ) = min y A - [ u ( y ) + d ( x , y ) ] for all  x A - .

Proof.

We assume that f 𝕏 ( Ω ) , otherwise u f and there is nothing to prove. Proposition 2.4 assures us that both A + and A - are nontrivial. Then the maximum and the minimum in (2.10) and (2.11) are well defined. We start by proving the second formula above. Let us define

v ( x ) = { min y A - [ u ( y ) + d ( x , y ) ] in  A - , u ( x ) in  Ω A - .

The strategy of the proof is to show that such a v belongs to 𝕏 ( Ω ) and that it is a competitor for u in the minimization of (2.4). Then uniqueness of the minimizer implies that

v u .

The first step is to show that

(2.12) | v ( x ) - v ( z ) | d ( x , z ) for all  x , z A - .

By definition. we have that

(2.13)

v ( x ) = u ( x ¯ ) + d ( x , x ¯ ) = min y A - [ u ( y ) + d ( x , y ) ] ,
v ( z ) = u ( z ¯ ) + d ( z , z ¯ ) = min y A - [ u ( y ) + d ( z , y ) ]

for some x ¯ , z ¯ A - . If x ¯ = z ¯ , we have that

| v ( x ) - v ( z ) | = d ( x , x ¯ ) - d ( z , x ¯ ) d ( x , z ) .

On the other hand, if x ¯ z ¯ , we get

(2.14)

v ( x ) = u ( x ¯ ) + d ( x , x ¯ ) u ( z ¯ ) + d ( x , z ¯ ) ,
v ( z ) = u ( z ¯ ) + d ( z , z ¯ ) u ( x ¯ ) + d ( z , x ¯ ) .

Thus by (2.13)–(2.14), we deduce that

- d ( x , z ) d ( x , x ¯ ) - d ( z , x ¯ ) v ( x ) - v ( z ) d ( x , z ¯ ) - d ( z , z ¯ ) d ( x , z ) ,

and (2.12) follows.

Next we prove that

v u on  A - .

By contradiction, assume that there exists x A - such that, say, u ( x ) > v ( x ) . Thus, there exists x ¯ A - , x x ¯ , such that v ( x ) = u ( x ¯ ) + d ( x , x ¯ ) < u ( x ) that contradicts u 𝕏 ( Ω ) .

We need to prove now that v 𝕏 ( Ω ) . To this end, we show that | v ( x ) - v ( y ) | d ( x , y ) for all x , y Ω . The only relevant case we have to consider is when y Ω A - and x A - . Consider γ Γ (see (1.22) for the definition), let t * = inf { t ( 0 , 1 ) : γ ( t ) A - } , and z = γ ( t * ) . Thus, since z A - , we have u ( z ) = v ( z ) and

| v ( x ) - v ( y ) | | v ( x ) - v ( z ) | + | u ( z ) - u ( y ) | d ( x , z ) + d ( z , y ) ( γ 1 ) + ( γ 2 ) = ( γ ) ,

where γ 1 = γ ( t [ 0 , t * ] ) and γ 2 = γ ( t [ t * , 1 ] ) . Minimizing the above expression with respect to any γ Γ , we deduce that v 𝕏 ( Ω ) .

To conclude the proof, we need to show that v u in A - . If we assume that there exists x ¯ A - such that u ( x ¯ ) > v ( x ¯ ) , then, by the definition of v, we have that

v ( x ¯ ) = u ( y ¯ ) + d ( x ¯ , y ¯ ) for some  y ¯ A - .

It follows that

u ( x ¯ ) - u ( y ¯ ) > v ( x ¯ ) - v ( y ¯ ) = d ( x ¯ , y ¯ ) ,

that contradicts the fact that u 𝕏 ( Ω ) . Next we show that u v in A - . Therefore, let us assume, arguing again by contradiction, that there exists x ¯ A - such that

sup y A - ( v - u ) ( y ) = ( v - u ) ( x ¯ ) = μ > 0 .

We set

D = { x A - : v ( x ) - u ( x ) > μ 2 } , η = ess inf y D ( f - u ) 0

and we define

u ~ = { max { u , v - ε } in  D , u in  Ω D ,

with ε = max { μ - η 2 , μ 2 } . Since both u and v are continuous, it follows that u ~ is continuous in Ω, and moreover u ~ 𝕏 ( Ω ) (since both u and v are). Observe now that u ~ satisfies u ~ < f in A - : indeed, in A - D this follows by the definition of A - , while

u ~ ( x ) - f ( x ) = v ( x ) - ε - f ( x ) = v ( x ) - u ( x ) μ - ε + u ( x ) - f ( x ) - η μ - ε - η < 0 in  D .

Consequently,

(2.15) ( u ( x ) - f ( x ) ) 2 ( u ~ ( x ) - f ( x ) ) 2 in  Ω ,

since the two quantities above coincide in Ω D while

u ( x ) - f ( x ) < max { u ( x ) , v ( x ) - ε } - f ( x ) < 0 in  D .

Then we have that

J ( u ~ ) < J ( u )

and this is in contradiction with the fact that u minimizes J .

Thus, we have that v ( x ) = u ( x ) in Ω and in particular v ( x ) = u ( x ) in A - and, consequently, the representation formula given by (2.11) holds true in A - . Analogously one can prove that also (2.10) is in force. ∎

Proof of Theorem 1.1.

The proof follows immediately putting together the results of Proposition 2.2 and Theorem 2.6. ∎

Remark 2.7.

Let us observe that a similar version of the previous results can be obtained working in W 0 1 , p n ( Ω ) , i.e., considering homogeneous Dirichlet boundary conditions. Actually, this latter case can be recovered as a sort of corollary of the W 1 , p n ( Ω ) . To get convinced of this fact, notice that the limit space for the Dirichlet problem is

𝕏 0 ( Ω ) = { v W 1 , ( Ω ) : | D v | 1  a.e. in  Ω  and  u = 0  on  Ω } .

We consider the two minimization problems

inf 𝕏 0 ( Ω ) 1 2 Ω ( v - f ) 2 and inf 𝕏 ( Ω ) 1 2 Ω ( v - f ~ ) 2 , where  f ~ ( x ) = max { - δ ( x ) , min { f ( x ) , δ ( x ) } } .

We claim that the minimizer u 𝕏 0 ( Ω ) of the first functional coincides with the minimizer u ~ 𝕏 0 ( Ω ) of the second one. Indeed, thanks to the definition of f ~ ( x ) , one easily deduces that u ~ 𝕏 0 ( Ω ) . Moreover, if u and u ~ would differ in some region, one could build a competitor for one of the two, contradicting uniqueness.

3 The problem with gradient lower order term

In this section we address the limit, as n , of the family of problems (1.10).

Our first result provides existence and uniqueness of a solution for problem (1.10), for any fixed p n ( 1 , ) , together with some estimates useful in order to study the asymptotic behavior as n .

Theorem 3.1.

Let us assume (1.11) and f L ( Ω ) . Then, for any fixed n N , there exists a unique solution v n W 1 , p ( Ω ) L ( Ω ) to problem (1.10) in the following weak sense:

(3.1) Ω v n ϕ + Ω | v n | p n - 2 v n ϕ = Ω ( H n ( x , v n ) + f ) ϕ for all  ϕ W 1 , p n ( Ω ) L ( Ω ) .

Moreover, the following estimates hold true:

(3.2) v n L ( Ω ) a = K 2 + f L ( Ω ) 𝑎𝑛𝑑 v n L p n ( Ω ) b n = a 1 p n e ( K 1 + 1 / p n ) a | Ω | 1 p n .

Remark 3.2.

Observe that

lim n + b n = e K 1 ( K 2 + f ) = b > 1 .

Proof.

The existence of a bounded weak solution to (3.1) follows by [11, 12, 13], while the uniqueness is a consequence of [27, Theorem 1.2]. In the following, we shall need the following truncation functions

(3.3) T k ( s ) = max { - k , min { s , k } } and G k ( s ) = s - T k ( s ) .

Set μ n = p n K 1 + 1 and let us consider, for k 0 ,

ϕ = ( e μ n | G k ( v n ) | - 1 ) sign ( v n ) .

Using the definition (3.3), it follows that

ϕ = μ n e μ n | G k ( v n ) | G k ( v n ) = μ n e μ n | G k ( v n ) | v n χ { | v n | > k } .

Therefore, using ϕ as a test function in (3.1) and taking advantage of assumption (1.11), we get that

Ω | v n | ( e μ n | G k ( v n ) | - 1 ) + μ n Ω | G k ( v n ) | p n e μ n | G k ( v n ) | Ω ( | H ( x , v n ) | + | f | ) ( e μ n | G k ( v n ) | - 1 ) p n K 1 Ω | G k ( v n ) | p n e μ n | G k ( v n ) | + ( K 2 + f L ( Ω ) ) Ω ( e μ n | G k ( v n ) | - 1 ) .

Noticing that supo ( ϕ ) = { | v n | > k } , recalling the choice of μ n , and the definition of a given in (3.2), we deduce that

(3.4) Ω | G k ( v n ) | p n + ( k - a ) { | v n | > k } ( e μ n | G k ( v n ) | - 1 ) 0 for all  k 0 .

In order to prove the L bound, assume by contradiction that

meas ( { x Ω : | v n | > a } ) > 0 ,

namely, there exists k ¯ > a such that

meas ( { x Ω : | v n | > k ¯ } ) > 0 .

Dropping the positive energy term in (3.4), it follows that

( k ¯ - a ) { | v n | > k ¯ } ( e μ n | G k ¯ ( v n ) | - 1 ) 0 .

Since, by assumption k ¯ > a and meas ( { | v n | > k ¯ } ) > 0 , the left-hand side above turns out to be strictly positive, that leads to a contradiction.

To obtain the estimate on the gradient, let us now go back to (3.4) and chose k = 0 . We get that

Ω | v n | p n a Ω ( e μ n | v n | - 1 ) a e μ n a | Ω | = b n p n ,

that implies (3.2). ∎

Now we are ready to prove the main result of the section.

Proof of Theorem 1.5.

Thanks to (3.2), we have that

v n L ( Ω ) a and v n L p n ( Ω ) b n .

Using Hölder’s inequality, it follows that for any N < q < p n we have

v n L q ( Ω ) v n L p n ( Ω ) | Ω | 1 q - 1 p n .

Since b n b (see Remark 3.2) and thanks to the L bound, we deduce that the sequence { v n } is bounded in W 1 , q ( Ω ) and, up to a (not relabeled) sequence, we get that v n v weakly in W 1 , q ( Ω ) and uniformly in Ω. The lower semicontinuity of the norm also implies that

v L q ( Ω ) b for all  q > 1 .

Therefore we conclude that v ( x ) satisfies

v W 1 , ( Ω ) a + b .

Let us now focus on the equation solved by v. We start by proving that

(3.5) max { 1 - | v | , - Δ v - H ( x , v ) } 0 in  Ω v + .

Indeed, let us pick any x 0 Ω v + and any φ C 2 ( Ω ) such that v ( x ) - φ ( x ) has a strict local maximum at x 0 with v ( x 0 ) - φ ( x 0 ) = 0 . We need to check that

max { 1 - | φ ( x 0 ) | , - Δ φ ( x 0 ) - H ( x 0 , φ ( x 0 ) ) } 0 .

Recall that v is the (uniform) limit of solutions to (3.1), so that there exist a sequence of real numbers ε n 0 and a sequence of points x n x 0 such that v n ( x ) - φ ( x ) - ε n has a local maximum at x n and v n ( x n ) - φ ( x n ) = ε n .

Since both f and H belong to C 0 ( Ω ¯ ) , then (see [28, Theorem 1]) any v n belongs to C 1 , α ( Ω ¯ ) for some α ( 0 , 1 ) (see also [23]) and thus v n turns out to be also a viscosity solution to (1.10). Hence we have that

(3.6) φ ( x n ) + ε n - f ( x n ) - Δ φ ( x n ) | φ ( x n ) | p n - 2 - ( p n - 2 ) Δ φ ( x n ) | φ ( x n ) | p n - 4 - H n ( x n , φ ( x n ) ) 0 .

Assume at first that 1 - | φ ( x 0 ) | 0 . It follows that | φ ( x n ) | > 0 for n large enough and since x 0 belongs to the (open) set Ω v + ; then φ ( x n ) + ε n - f ( x n ) is strictly positive and we can drop it from the previous inequality. Hence, we divide (3.6) by ( p n - 2 ) | φ ( x n ) | p n - 4 so that

Δ φ ( x n ) | φ ( x n ) | 2 p n - 2 - Δ φ ( x n ) - H n ( x n , φ ( x n ) ) ( p n - 2 ) | φ ( x n ) | p n - 4 0 ,

and we take the limit with respect to n. Since φ C 2 and exploiting (1.12), we deduce that

- Δ φ ( x 0 ) - H ( x 0 , φ ( x 0 ) ) 0 .

Conversely, if | φ ( x 0 ) | < 1 , we pass to the limit directly in (3.6). Indeed, we use once again assumption (1.12) and Remark 1.4 and we deduce that the last three terms in (3.6) vanish. Thus we get φ ( x 0 ) - f ( x 0 ) = v ( x 0 ) - f ( x 0 ) 0 , that is in contradiction with the case under consideration.

Next we prove that

(3.7) max { 1 - | v | , - Δ v - H ( x , v ) } 0 in  { v ( x ) - f ( x ) > 0 } .

Consider any x 0 Ω v + and take any φ C 2 ( Ω ) such that v ( x ) - φ ( x ) has a strict local minimum at x 0 with v ( x 0 ) - φ ( x 0 ) = 0 . We want to prove that for such a φ we have that

max { 1 - | φ ( x 0 ) | , - Δ φ ( x 0 ) - H ( x 0 , φ ( x 0 ) ) } 0 .

If 1 - | φ ( x 0 ) | 0 , the inequality is trivially satisfied, so we deal with the case | φ ( x 0 ) | > 1 . As before, we recall that there exist a sequence of real numbers ε n 0 and a sequence of points x n x 0 such that v n ( x ) - φ ( x ) - ε n has a local minimum at x n and v n ( x n ) - φ ( x n ) = ε n such that

φ ( x n ) + ε n - f ( x n ) - Δ φ ( x n ) | φ ( x n ) | p n - 2 - ( p n - 2 ) Δ φ ( x n ) | φ ( x n ) | p n - 4 - H n ( x n , φ ( x n ) ) 0 .

Dividing the above inequality by ( p n - 2 ) | φ ( x n ) | p n - 4 and taking the limit with respect to n, we conclude that

- Δ φ ( x 0 ) - H ( x 0 , φ ( x 0 ) ) 0 .

Gathering (3.5) and (3.7), we conclude that v is a viscosity solution to

max { 1 - | v | , - Δ v - H ( x , v ) } = 0 in  Ω v + .

The proof that v solves

min { | v | - 1 , - Δ v - H ( x , v ) } = 0 in  Ω v - ,

follows similarly to the previous case.

Let us focus now on the boundary conditions. For the sake of brevity we only prove the first one of the four inequalities (1.15)–(1.16), since the proofs of the other ones follow in the same way.

Assume that there exists φ C 2 ( Ω ¯ ) such that v - φ has a strict local maximum at x 0 , with v ( x 0 ) = φ ( x 0 ) . Since { v n } converges uniformly in Ω ¯ to v, there exist a sequence of real numbers ε n 0 and a sequence of points Ω ¯ x n x 0 such that v n - φ - ε n attains a local maximum at x n and v n ( x n ) = φ ( x n ) + ε n .

Let us assume that (up to a subsequence) { x n } ( Ω ) v + ; we recall that since any v n belongs to C 1 , α ( Ω ¯ ) , we have

v n ( x ) ν = 0 for all  x Ω .

Using that φ ( x ) + ε n touches v n from above at x n , we have that

φ ( x n ) ν v ( x n ) ν = 0 .

Passing to the limit as n , it follows

φ ( x 0 ) ν 0 .

On the contrary, if we assume that { x n } Ω v + , we take advantage to the fact that v n is a viscosity subsolution to (2.2), i.e., inequality (3.6) holds true. Going back to the proof of Theorem 1.5 we observe that | φ ( x 0 ) | < 1 leads to a contradiction and that | φ ( x 0 ) | - 1 0 implies that - Δ φ ( x 0 ) - H ( x , φ ( x 0 ) ) 0 , as desired. ∎

4 The limit equation for the model problem

Let us go back to the limit as p to (1.10) with H n H 0 .

Proof of Theorem 1.2.

Since u 𝕏 ( Ω ) , by Proposition 1.9 we have that

1 - | u | 0 and | u | - 1 0 in  Ω .

To prove the reverse inequalities in Ω u + and Ω u - , let us notice that Proposition 2.2 shows that u n u uniformly in Ω ¯ , where u n are the weak solutions to (2.2). We then use Theorem 1.5 with H 0 to deduce that

{ max { 1 - | u | , - Δ u } 0 in  Ω u + , min { | u | - 1 , - Δ u } 0 in  Ω u - .

Since max { α , β } 0 implies both α 0 and β 0 , we deduce that

1 - | u | 0 in  Ω u + and | u | - 1 0 in  Ω u - .

As far as the boundary conditions are concerned, we use again Theorem 1.5 in the special case H 0 to deduce that

max { 1 - | u | , u ν } 0 on  ( Ω ) u + .

In order to prove that

(4.1) 1 - | u | 0 on  ( Ω ) u + ,

assume by contradiction that there exists φ C 2 that touches u from above at x 0 and that verifies

1 - | φ ( x 0 ) | > 0 .

Take a small r such that B r ( x 0 ) Ω ( Ω ) + and | φ ( y ) | < 1 for all | y - x 0 | r and B r ( x 0 ) . Therefore, set

μ = inf x B r ( x 0 ) Ω ( φ ( x ) - u ( x ) ) > 0 and η = inf x B r ( x 0 ) Ω ( φ ( x ) - f ( x ) ) > 0

and consider, for ε = min { μ , η } , the function

u ~ = { min { u , φ - ε } in  B r ( x 0 ) Ω ¯ , u in  Ω ¯ B r ( x 0 ) .

Thanks to the definition of ε, u ~ is continuous, u ~ > f in A + and u ~ ( x 0 ) < u ( x 0 ) . Moreover, the choice of r also implies that u ~ 𝕏 ( Ω ) . This would imply J ( u ~ ) < J ( u ) , that contradicts the minimality of u. Therefore (4.1) holds true.

To deal with the boundary condition on ( Ω ) - , we can follow exactly the same strategy and hence we omit the details. ∎

Even if the minimizer of (2.4) is unique, in general problem (1.6)–(1.7) possesses more than one solution. In the following we provide a solution to (1.6), that satisfies the same boundary condition, and that differs from the minimizer of J (unless f 𝕏 ( Ω ) ).

Proof of Theorem 1.3.

Let us consider the set

M = { v 𝕏 ( Ω ) : v f } ,

and notice that it is not empty since the constant f L ( Ω ) M , and let us set

α = inf { Ω v : v M } .

By the definition of M, we have that α Ω f and we consider a sequence { v n } M whose integral on Ω converges to α. Observe that for any n , we have that v n f and moreover one can assume, without loss of generality, that v n f L ( Ω ) (otherwise it is sufficient to truncate v n at high f L ( Ω ) ). Hence the sequence { v n } is equibounded. Moreover, since { v n } 𝕏 ( Ω ) , it is also equicontinuous. Thus by the Ascoli–Arzelà Theorem, we have that v n converges uniformly in Ω ¯ to some u ¯ 𝕏 , that satisfies α = Ω u ¯ .

Next, we prove that u ¯ turns out to be a viscosity solution to (1.8). Thanks to Proposition 1.9, we immediately deduce that u ¯ satisfies

1 - | u ¯ | 0 in  Ω .

Hence, to conclude the proof, we take any x 0 Ω u + ¯ ( Ω ) u + and assume by contradiction that there exists φ C 2 that touches u ¯ from above at x 0 and that satisfies

1 - | φ ( x 0 ) | > 0 .

Since φ is smooth, there exists a small r > 0 such that | φ ( y ) | < 1 for all y B r ( x 0 ) Ω ¯ . Therefore, we set

μ = inf x B r ( x 0 ) Ω ( φ ( x ) - u ¯ ( x ) ) > 0 and η = inf x B r ( x 0 ) Ω ( φ ( x ) - f ( x ) ) > 0

and we consider

u ~ = { min { u ¯ , φ - ε } in  B r ( x 0 ) , u ¯ in  Ω B r ( x 0 ) , with  ε = min { μ , η } .

Thus, thanks to the definition of ε, u ~ is continuous, u ~ > f in Ω u ¯ + and u ~ ( x 0 ) < u ¯ ( x 0 ) . Moreover, the choice of r also implies that u ~ 𝕏 ( Ω ) . However, by construction this would imply that Ω u ~ < α , that leads to a contradiction. Hence, we obtain that 1 - | φ ( x 0 ) | 0 .

It remains to verify that, for any x 0 Ω u ¯ + and any φ C 2 that touches u ¯ from below at x 0 , we have

max { 1 - | φ ( x 0 ) | , φ ( x 0 ) ν } 0 .

Assume by contradiction that 1 - | φ ( x 0 ) | < 0 and φ ( x 0 ) ν = - φ ( x 0 ) δ ( x 0 ) < 0 . This latter condition implies that φ ( x 0 ) directs inside Ω; moreover since | φ ( x 0 ) | > 1 , we deduce that there exists y 0 Ω (close enough to x 0 ) such that φ ( y 0 ) - φ ( x 0 ) > | y 0 - x 0 | . Using that φ touches u ¯ from below at x 0 , we have that

u ¯ ( y 0 ) - u ¯ ( y 0 ) φ ( y 0 ) - φ ( y 0 ) > | y 0 - x 0 | ,

that leads to a contradiction.

Finally, the uniqueness of the solution of the obstacle problem follows from the minimality of u ¯ . ∎

Proof of Proposition 1.6.

Let us define

(4.2)

v + ( x ) = max y Ω w + [ f ( y ) - d ( x , y ) ] for all  x Ω w + ,
v - ( x ) = min y Ω w - [ f ( y ) + d ( x , y ) ] for all  x Ω w - .

We prove, at first, that v + solves 1 - | v + | = 0 and that - Δ v + 0 in Ω w + . Thanks to Lemma 2.6 (we recall that v + 𝕏 ( Ω + ) ) we already now that 1 - | v + | 0 . Notice now that for any y Ω w + , the function

x f ( y ) - d ( x , y )

is locally a cone (recall that the geodesic distance coincides locally with the euclidean one). Therefore, it solves the eikonal equation and it is infinity harmonic outside its vertex. Since v + is defined as maximum of such functions, the stability properties of viscosity solutions (see Proposition 4.3 of [21]) imply that 1 - | v + | 0 and - Δ v + 0 in Ω w + .

A similar argument proves that | v - | - 1 = 0 and - Δ v - 0 in Ω w - . Consequently, we found that

max { 1 - | v + | , - Δ v + } = 0 and 1 - | v + | = 0 in  Ω w + ,
min { | v - | - 1 , - Δ v - } = 0 and | v - | - 1 = 0 in  Ω w - .

Let us assume now that w solves

(4.3) { max { 1 - | w | , - Δ w } = 0 in  Ω w + , min { | w | - 1 , - Δ w } = 0 in  Ω w - .

Thanks to the comparison principle (see Theorem 2.1 of [26]) we have that

w v ± in  Ω w ± ,

respectively. Thanks to the properties of v ± , we deduce that w is also a solution to

(4.4) { 1 - | w | = 0 in  Ω w + , | w | - 1 = 0 in  Ω w - .

On the other hand, if w is a solution to (4.4), then again for the comparison principle (see [5, Theorem 5.9]) w v ± in Ω w ± and then it also solves (4.3). ∎

5 A functional allowing jumps

We start this section recalling some basic facts abound the space of special functions of bounded variation SBV ( Ω ) . For more details we address the interested reader to the overview [25] or the classical monograph [2].Let us recall that a function v L 1 ( Ω ) belongs to BV ( Ω ) (bounded variation) if and only if there exists a vector valued Radon measure Du such that

Ω v div ( φ ) = - Ω φ d ( D v ) for all  φ C c ( Ω , N ) .

Thanks to the Radon–Nikodym Theorem, we can uniquely decompose Dv as D v = D a v + D s v with

D a v N and D s v N .

To give a better description of the absolutely continuous part and the singular one, we need to introduce some more concepts. For any x Ω , we say that v BV ( Ω ) admits approximate limit v ¯ ( x ) if

(5.1) lim ρ 0 1 | B ρ ( x ) | B ρ ( x ) | v ( y ) - v ¯ ( x ) | 𝑑 y = 0

and approximate gradient v ( x ) N if

lim ρ 0 1 ρ | B ρ ( x ) | B ρ ( x ) | v ( y ) - v ¯ ( x ) - v ( x ) ( y - x ) | 𝑑 y = 0 .

It can be proved (see [2, Theorem 3.83]) that any v BV ( Ω ) is approximately differentiable a.e. and that v ( L 1 ( Ω ) ) N is the density of the absolutely continuous part of Dv with respect to the Lebesgue measure, namely

D a v = D v N = v N .

In order to describe also the singular part of Dv, let us consider the set C v Ω of all points where (5.1) holds true and denote S u = Ω C v . The set S v is called the set of approximated discontinuity of v and it is ( N - 1 ) -countably rectifiable. Now, we introduce

D j v = D v S u and D c v = D v ( Ω S u ) .

The measure D j v takes into account the jumps of the function v, while D c v is the Cantor part of Dv. We notice that in [2] D j v is defined through the slightly smaller set J v , the set of approximated jumps point of v; since N - 1 ( S v J v ) = 0 , we keep using S v for the sake of simplicity (see also [25]). At this point it is easy to define the space of special functions of bounded variation as

SBV ( Ω ) = { u BV ( Ω ) : D c u 0 } .

Now we are ready to define our functional : SBV ( Ω ) as

(5.2) p ( v ) = 1 p Ω | v | p + 1 r Ω | v - f | r + N - 1 ( S v ) .

Comparing (5.2) with (1.18), we note that the new functional framework allows to pass from a two variables functional to a single variable one, where the discontinuity set K is replaced by the singular (jump) set S u .

We also stress that BV ( Ω ) would not be a good ambient space for (5.2) since the functional is not coercive with respect to the cantor part of gradient measure. To be convinced that SBV ( Ω ) is the right ambient space to settle the minimization of p , it is enough to look at the following result.

Theorem 5.1 ([2, Theorems 4.7 and 4.8]).

Let { u n } SBV ( Ω ) such that

Ω | v n | p + N - 1 ( S v n ) + v n L ( Ω ) C .

Then there exists v SBV ( Ω ) such that, up to a subsequence v n v strongly in L 1 ( Ω ) , v n v weakly in L p ( Ω ) , D v n D v weakly star in the sense of measures and

Ω | v | p lim inf n Ω | v n | p n 𝑎𝑛𝑑 N - 1 ( S v ) lim inf n N - 1 ( S v n ) .

Theorem 5.1 easily implies the existence of a minimizer u p SBV ( Ω ) for (5.2). Indeed it is easy to see that any minimizing sequence v n can be chosen so that v n L ( Ω ) f L ( Ω ) , basically because truncations with a large constant always decrease the energy of the functional.

What is not clear is that if the couple ( S u p , u p ) is a minimizer of (1.18). The main difficulty here is that in general the singular set of a SBV function need not to be closed and can be even dense in Ω. What is possible to prove is that any minimizer u p SBV ( Ω ) of (5.2) satisfies N - 1 ( S ¯ u p S u p ) = 0 and that indeed ( S ¯ u p , u p ) is a minimizer of (1.18). We gather all this information in the following result.

Theorem 5.2 ([2, Theorems 7.15, 7.21 and 7.22]).

Set p ( 1 , ) , r [ 1 , ) and f L ( Ω ) . Then there exists a minimizer u p SBV ( Ω ) to (5.2). Moreover, there exists θ depending on N and p such that

N - 1 ( S u p B ρ ( x ) ) > θ ρ N - 1

for any x S u p ¯ and ρ > 0 . This lower bounds implies that H N - 1 ( S u ¯ S u ) = 0 . Thus the pair given by K = S ¯ u p and u W 1 , p ( Ω S ¯ u p ) and is a minimizer of the classical Mumford–Shah problem (1.18).

Our next step is to prove that any sequence { u n } SBV ( Ω ) of minimizers of (5.2), associated to a sequence p n , converges (up to a subsequence) to a minimizer of

(5.3) ( v ) = 1 r Ω | v - f | r + N - 1 ( S v ) with  v SBV ( Ω ) , | v | 1 .

Theorem 5.3.

Let { u n } SBV ( Ω ) be the sequence of minimizer of the functional (5.2) relative to a diverging sequence of real number p n . Then any accumulation point u SBV ( Ω ) of { u n } (in L 1 ( Ω ) ), is a minimizer of (5.3).

Proof.

Recalling that p n ( u n ) p n ( 0 ) and that u n L ( Ω ) f L ( Ω ) , it results

1 p n Ω | u n | p n + u n L ( Ω ) + N - 1 ( S u n ) C f L ( Ω ) .

Using Hölder inequality, we get

( Ω | u n | q ) 1 q ( p n C f L ( Ω ) ) 1 p n | Ω | 1 q - 1 p n

for any q < p n . Theorem 5.1 and a diagonal argument then implies that there exists u SBV ( Ω ) , with | u | 1 such that u n u strongly in any L q ( Ω ) , u n u weakly in any ( L q ( Ω ) ) N , with q [ 1 , ) , D u n D u weakly star in the sense of measure and

(5.4) ( u ) lim inf n p n ( u n ) .

Now we claim that

( u ) ( z ) for all  z SBV ( Ω ) L ( Ω ) , | z | 1 .

This concludes the proof of the theorem since, by a simple truncation argument, if reached, the infimum of in SBV ( Ω ) L ( Ω ) has to be bounded. By the definition of u n , we have that

(5.5) p n ( u n ) p n ( z ) 1 p n | Ω | + 1 r Ω | z - f | r + N - 1 ( S z ) .

Taking the liminf with respect to n, we obtain the desired result. ∎

Notice that in the proof of Theorem 5.3 we have proved that the functional p n Γ-converges to the functional . For completeness let us recall the definition of Γ-convergence.

Definition 5.4.

Let F n : 𝕏 be a sequence of functionals. We say that F n Γ--converges to F : 𝕏 if

  1. for any x 𝕏 and any x n x , we have that

    lim inf n + F n ( x n ) F ( x ) ,

  2. for any x 𝕏 , there exists x n 𝕏 such that x n x , and

    lim sup n + F n ( x n ) F ( x ) .

In our case (i) has been proved in (5.4) while (ii) is trivially obtained taking the limit on the right-hand side of (5.5).

6 Examples

In this section we collect some explicit examples that clarify the behavior of the minimizers of some of the functional presented in this manuscript.

Examples 6.1.

Suppose that Ω = ( - 1 , 1 ) . We are interested in considering the following minimization problem:

inf 𝕏 ( - 1 , 1 ) J ( v ) with  J ( v ) = 1 2 - 1 1 ( v - f ) 2 .

Case 1: Choose f 1 ( x ) = k χ ( - r , r ) ( x ) with r ( 0 , 1 ] and k R (see Figure 1). We first observe that if r = 1 , then f 1 𝕏 ( - 1 , 1 ) and thus the choice u 1 ( x ) f 1 ( x ) is allowed. Conversely, if r ( 0 , 1 ) , then f 1 𝕏 ( - 1 , 1 ) , and we look for a solution u 1 ( x ) of the form u h ( x ) = min { ( h - | x | ) + , k } , with h to be determined. Indeed, notice that it must be 0 u 1 1 (if not max { 0 , min { u 1 , 1 } } would be a competitor for u 1 ). Moreover, since f 1 is even and the minimizer is unique, also u 1 has to be even. In addition, notice that the minimizer must be a solution to the Eikonal equation or to minus the Eikonal equation, hence u h must be composed by lines with slope one. Thus J ( u h ) can be explicitly computed and it turns out that

J ( u h ) = h - k r ( h - k - x ) 2 𝑑 x + r h ( h - x ) 2 𝑑 x .

Minimizing its value with respect to h, we have that the minimum is achieved at h * = r + k 2 .

The heuristic that motivates this example is that in a neighbor where f is not smooth the solution is a line with maximum slope allowed (i.e., ± 1 ) and far away it tries to be as close as possible to f.

Case 2: Consider f 2 ( x ) = 2 | x | (see Figure 2). In this case the picture is much more clear. Since f 2 𝕏 ( - 1 , 1 ) , the closest function to f 2 in 𝕏 ( - 1 , 1 ) is a line with slope + 1 in ( 0 , 1 ] and with slope -1 in [ - 1 , 0 ] . Since f 2 is even, it follows that u 2 inherits the same property, so that

J ( u h ) = 0 1 ( 2 x - ( x + h ) ) 2 𝑑 x ,

attains its minimum value for h = 1 2 .

Case 3: Consider f 3 ( x ) = | x | (see Figure 3). By symmetry it follows that the solution is even and that it has the form

u 3 ( x ) = { | x | - s + s for  | x | s , | x | for  | x | > s ,

with s > 1 4 . To determine the value of s we have to compute

s J ( u ) = s 0 s ( | x | - | x | + s - s ) 2 𝑑 x = 2 ( 1 - 1 2 s ) 0 s ( | x | - | x | + s - s ) 𝑑 x = 2 ( 1 - 1 2 s ) ( s 2 2 - s 3 2 3 ) .

Therefore we deduce that the minimum is achieved for s = 4 9 .

Example 6.2.

Let us underline the importance of minimizing J in SBV ( Ω ) by showing that for certain f, when it is allowed, the minimizer does jump.

In order to avoid technicalities, we present first full details in dimension 1, being the extension of these ideas to any dimension quite similar. As already observed in the previous example the minimizer of

J ( u ) = 1 2 - R R ( u ( x ) - k χ ( - r , r ) ) 2 𝑑 x with  u 𝕏 ( - R , R )

is

u h ( x ) = min { ( h - | x | ) + , k } ,

with h * = r + k 2 . In addition, we have

J ( u h * ) = 1 12 k 3 .

Observe that, since such an f SBV ( - R , R ) , we can consider

( v ) = 1 2 - R R ( u ( x ) - k χ ( - r , r ) ) 2 𝑑 x + ( S u ) with  u SBV ( - R , R )

with ( S u ) the measure of the singular set of u. Observe that ( S f ) = 2 and since f SBV ( - R , R ) , we have ( f ) = 2 . Consequently, if

2 < 1 12 k 3 ,

we have

( f ) < J ( u h * ) ,

and we conclude that the minimizer indeed has jumps.

As far as the case in dimension N 2 is concerned, we repeat the same argument: in the domain Ω = B R ( 0 ) we set

f ( x ) = k χ B r ( 0 ) ( x )

with r > 0 to be chosen large enough. Thus the minimizer of

J ( u ) = 1 2 B R ( 0 ) ( u ( x ) - k χ B r ( 0 ) ) 2 𝑑 x with  u 𝕏 ( B R ( 0 ) )

has the form

u h ( x ) = min { k , ( h - | x | ) + } .

The minimum value of J ( u h ) is achieved for a suitable value h * > 0 that is not so easy to be computed explicitly and which satisfies

lim r r h * = 1 .

Thus, tedious but not difficult computations lead to

J ( u h * ) k 3 24 ω N r 3 as  r + .

Analogously, it is not hard to see that for the functional

( u ) = 1 2 B R ( 0 ) ( u ( x ) - k χ B r ( 0 ) ) 2 𝑑 x + N - 1 ( S u ) with  u SBV ( B R ( 0 ) )

since f SBV ( B R ( 0 ) ) , we have that

( f ) = N ω N r N - 1 .

Consequently, if r is large enough and k 3 > 24 N , then it holds that

( f ) = N ω N r N - 1 < J ( u 0 ) ,

and the minimizer jumps.


Communicated by Juan Manfredi


Funding source: Austrian Science Fund

Award Identifier / Grant number: 10.55776/F65

Award Identifier / Grant number: 10.55776/P32788

Award Identifier / Grant number: 10.55776/ESP9

Award Identifier / Grant number: CR 40006150

Award Identifier / Grant number: 2023 CUP E53C22001930001

Funding statement: Stefano Buccheri is supported by the Austrian Science Fund (FWF), projects 10.55776/F65, 10.55776/P32788, 10.55776/ESP9, by the Belgian National Fund for Scientific Research (NFSR), project CR 40006150, and by the Istituto Nazionale di Alta Matematica Francesco Severi (INDAM), GRAMPA project 2023 CUP E53C22001930001 “Regolarità per problemi ellittici e parabolici con crescite non standard.

References

[1] A. Alvino, G. Trombetti and P.-L. Lions, Comparison results for elliptic and parabolic equations via Schwarz symmetrization, Ann. Inst. H. Poincaré C Anal. Non Linéaire 7 (1990), no. 2, 37–65. 10.1016/s0294-1449(16)30303-1Search in Google Scholar

[2] L. Ambrosio, N. Fusco and D. Pallara, Functions of Bounded Variation and Free Discontinuity Problems, Oxford Math. Monogr., Oxford University, New York, 2000. 10.1093/oso/9780198502456.001.0001Search in Google Scholar

[3] G. Aronsson, Extension of functions satisfying Lipschitz conditions, Ark. Mat. 6 (1967), 551–561. 10.1007/BF02591928Search in Google Scholar

[4] G. Aronsson, M. G. Crandall and P. Juutinen, A tour of the theory of absolutely minimizing functions, Bull. Amer. Math. Soc. (N. S.) 41 (2004), no. 4, 439–505. 10.1090/S0273-0979-04-01035-3Search in Google Scholar

[5] M. Bardi and I. Capuzzo-Dolcetta, Optimal Control and Viscosity Solutions of Hamilton–Jacobi–Bellman Equations, Mod. Birkhäuser Class., Birkhäuser, Basel, 1997. 10.1007/978-0-8176-4755-1Search in Google Scholar

[6] E. N. Barron, L. C. Evans and R. Jensen, The infinity Laplacian, Aronsson’s equation and their generalizations, Trans. Amer. Math. Soc. 360 (2008), no. 1, 77–101. 10.1090/S0002-9947-07-04338-3Search in Google Scholar

[7] E. N. Barron, R. R. Jensen and C. Y. Wang, The Euler equation and absolute minimizers of L functionals, Arch. Ration. Mech. Anal. 157 (2001), no. 4, 255–283. 10.1007/PL00004239Search in Google Scholar

[8] M. F. Betta, V. Ferone and A. Mercaldo, Regularity for solutions of nonlinear elliptic equations, Bull. Sci. Math. 118 (1994), no. 6, 539–567. Search in Google Scholar

[9] T. Bhattacharya, E. DiBenedetto and J. Manfredi, Limits as p of Δ p u p = f and related extremal problems, Rend. Semin. Mat. Univ. Politec. Torino (1991), 15–68. Search in Google Scholar

[10] L. Boccardo, S. Buccheri and G. R. Cirmi, Calderon–Zygmund–Stampacchia theory for infinite energy solutions of nonlinear elliptic equations with singular drift, NoDEA Nonlinear Differential Equations Appl. 27 (2020), no. 4, Paper No. 38. 10.1007/s00030-020-00641-zSearch in Google Scholar

[11] L. Boccardo, F. Murat and J.-P. Puel, Résultats d’existence pour certains problèmes elliptiques quasilinéaires, Ann. Sc. Norm. Super. Pisa Cl. Sci. (4) 11 (1984), no. 2, 213–235. Search in Google Scholar

[12] L. Boccardo, F. Murat and J.-P. Puel, Quelques propriétés des opérateurs elliptiques quasi linéaires, C. R. Acad. Sci. Paris Sér. I Math. 307 (1988), no. 14, 749–752. Search in Google Scholar

[13] L. Boccardo, F. Murat and J.-P. Puel, L estimate for some nonlinear elliptic partial differential equations and application to an existence result, SIAM J. Math. Anal. 23 (1992), no. 2, 326–333. 10.1137/0523016Search in Google Scholar

[14] H. Brezis, Functional Analysis, Sobolev Spaces and Partial Differential Equations, Universitext, Springer, New York, 2011. 10.1007/978-0-387-70914-7Search in Google Scholar

[15] S. Buccheri, The Bottaro–Marina slice method for distributional solutions to elliptic equations with drift term, AIP Conf. Proc. 2425 (2022), Article ID 210001. 10.1063/5.0082154Search in Google Scholar

[16] S. Buccheri and T. Leonori, Large solutions to quasilinear problems involving the p-Laplacian as p diverges, Calc. Var. Partial Differential Equations 60 (2021), no. 1, Paper No. 30. 10.1007/s00526-020-01883-6Search in Google Scholar

[17] S. Buccheri, T. Leonori and J. D. Rossi, Strong convergence of the gradients for p-Laplacian problems as p , J. Math. Anal. Appl. 495 (2021), no. 1, Article ID 124724. 10.1016/j.jmaa.2020.124724Search in Google Scholar

[18] T. Champion and L. De Pascale, Principles of comparison with distance functions for absolute minimizers, J. Convex Anal. 14 (2007), no. 3, 515–541. Search in Google Scholar

[19] M. G. Crandall, An efficient derivation of the Aronsson equation, Arch. Ration. Mech. Anal. 167 (2003), no. 4, 271–279. 10.1007/s00205-002-0236-3Search in Google Scholar

[20] M. G. Crandall, L. C. Evans and R. F. Gariepy, Optimal Lipschitz extensions and the infinity Laplacian, Calc. Var. Partial Differential Equations 13 (2001), no. 2, 123–139. 10.1007/s005260000065Search in Google Scholar

[21] M. G. Crandall, H. Ishii and P.-L. Lions, User’s guide to viscosity solutions of second order partial differential equations, Bull. Amer. Math. Soc. (N. S.) 27 (1992), no. 1, 1–67. 10.1090/S0273-0979-1992-00266-5Search in Google Scholar

[22] E. De Giorgi, M. Carriero and A. Leaci, Existence theorem for a minimum problem with free discontinuity set, Arch. Ration. Mech. Anal. 108 (1989), no. 3, 195–218. 10.1007/BF01052971Search in Google Scholar

[23] E. DiBenedetto, C 1 + α local regularity of weak solutions of degenerate elliptic equations, Nonlinear Anal. 7 (1983), no. 8, 827–850. 10.1016/0362-546X(83)90061-5Search in Google Scholar

[24] L. C. Evans and W. Gangbo, Differential equations methods for the Monge–Kantorovich mass transfer problem, Mem. Amer. Math. Soc. 137 (1999), no. 653, 1–66. 10.1090/memo/0653Search in Google Scholar

[25] N. Fusco, An overview of the Mumford–Shah problem, Milan J. Math. 71 (2003), 95–119. 10.1007/s00032-003-0016-zSearch in Google Scholar

[26] R. Jensen, Uniqueness of Lipschitz extensions: Minimizing the sup norm of the gradient, Arch. Ration. Mech. Anal. 123 (1993), no. 1, 51–74. 10.1007/BF00386368Search in Google Scholar

[27] T. Leonori, A. Porretta and G. Riey, Comparison principles for p-Laplace equations with lower order terms, Ann. Mat. Pura Appl. (4) 196 (2017), no. 3, 877–903. 10.1007/s10231-016-0600-9Search in Google Scholar

[28] G. M. Lieberman, Boundary regularity for solutions of degenerate elliptic equations, Nonlinear Anal. 12 (1988), no. 11, 1203–1219. 10.1016/0362-546X(88)90053-3Search in Google Scholar

[29] T. Mengesha and N. C. Phuc, Quasilinear Riccati type equations with distributional data in Morrey space framework, J. Differential Equations 260 (2016), no. 6, 5421–5449. 10.1016/j.jde.2015.12.007Search in Google Scholar

[30] J. D. Rossi and N. Saintier, On the first nontrivial eigenvalue of the -Laplacian with Neumann boundary conditions, Houston J. Math. 42 (2016), no. 2, 613–635. Search in Google Scholar

Received: 2023-05-09
Accepted: 2024-03-10
Published Online: 2024-04-24
Published in Print: 2025-01-01

© 2024 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 16.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/acv-2023-0058/html
Scroll to top button