Home The Hopf Lemma for the Schrödinger Operator
Article Open Access

The Hopf Lemma for the Schrödinger Operator

  • Augusto C. Ponce ORCID logo EMAIL logo and Nicolas Wilmet ORCID logo
Published/Copyright: March 24, 2020

Abstract

We prove the Hopf boundary point lemma for solutions of the Dirichlet problem involving the Schrödinger operator -Δ+V with a nonnegative potential V which merely belongs to Lloc1(Ω). More precisely, if uW01,2(Ω)L2(Ω;Vdx) satisfies -Δu+Vu=f on Ω for some nonnegative datum fL(Ω), f0, then we show that at every point aΩ where the classical normal derivative u(a)n exists and satisfies the Poisson representation formula, one has u(a)n>0 if and only if the boundary value problem

{ - Δ v + V v = 0 in  Ω , v = ν on  Ω ,

involving the Dirac measure ν=δa has a solution. More generally, we characterize the nonnegative finite Borel measures ν on Ω for which the boundary value problem above has a solution in terms of the set where the Hopf lemma fails.

1 Introduction and Main Results

Let Ω be a bounded connected open subset of N with smooth boundary and let VL(Ω) be a nonnegative function. The weak maximum principle ensures that the distributional solution uC1(Ω¯) of the Dirichlet problem

(1.1) { - Δ u + V u = f in  Ω , u = 0 on  Ω ,

satisfies u0 on Ω whenever fL(Ω) is a nonnegative function; see Lemma 2.2 below. From the minimality of u on Ω, the normal derivative of u with respect to the inward unit normal vector n thus verifies un0 on Ω. When f0, the classical Hopf lemma (see [10, Lemma 6.4.2] or [11, Lemma 3.4]) gives the stronger conclusion

(1.2) u n > 0 on  Ω .

Boundedness of V is an important element to obtain (1.2) as it allows one to construct a positive minorant of u on Ω with positive normal derivative at any given point on Ω. To understand in what respect this assumption on V can be relaxed, we assume henceforth that

V L loc 1 ( Ω )  and  V 0  almost everywhere on  Ω ,

but we restrict ourselves to the class of nonnegative data fL(Ω). In this setting, a solution of (1.1) is a function u that belongs to W01,2(Ω)L2(Ω;Vdx) and satisfies the equation

- Δ u + V u = f in the sense of distributions in  Ω .

Observe that u is the unique minimizer of the energy functional

E ( z ) = 1 2 Ω ( | z | 2 + V z 2 ) d x - Ω f z d x

with zW01,2(Ω)L2(Ω;Vdx).

As the solution of (1.1) need not be C1, nor even continuous, due to some possible singularity from V, we first need to address the pointwise meaning of the normal derivative un. Since u is the difference between a continuous and a bounded superharmonic function, every xΩ is a Lebesgue point and the precise representative of u satisfies the following representation formula in terms of the Green function G of -Δ on Ω:

u ^ ( x ) = Ω G ( x , y ) ( - Δ u ( y ) ) d y for every  x Ω .

Then, from a formal computation, one presumably gets at a point aΩ:

(1.3) u ^ n ( a ) = Ω K ( a , y ) ( - Δ u ( y ) ) d y ,

where K:=Gn denotes the Poisson kernel of -Δ on Ω. This formula can be rigorously justified when VL(Ω), and then ΔuL(Ω), using standard estimates on G.

There is no reason why (1.3) should remain valid in general as we do not assume any particular behavior of V near Ω. We show nevertheless that, for any fixed V, there is a common property which is shared by all nontrivial solutions of (1.1) with nonnegative fL(Ω). To this end, let ζ1 be the solution of (1.1) with constant density f1 and define the set

𝒩 = { a Ω : the classical normal derivative  ζ 1 ^ n  exists at  a  and (1.3) is valid with  u = ζ 1 } .

To simplify the notation, we do not explicit the dependence of 𝒩 on V.

We prove:

Theorem 1.

For every nonnegative function fL(Ω), f0, the solution u of (1.1) involving f has a classical normal derivative at aΩ that satisfies (1.3) if and only if aN.

The set 𝒩 thus provides one with a common ground where a normal derivative exists, independently of the solution of (1.1). We can now address the question of whether the Hopf lemma is valid on 𝒩. We rely on the characterization of the set of points aΩ for which the boundary value problem

(1.4) { - Δ v + V v = 0 in  Ω , v = δ a on  Ω ,

involving the Dirac measure δa has a distributional solution in the sense that the function vL1(Ω) is such that VvL1(Ω;dΩdx) and satisfies

(1.5) ζ n ( a ) = Ω v ( - Δ ζ + V ζ ) d x

for every ζC(Ω¯) with ζ=0 on Ω, where dΩ:Ω+ is the distance to the boundary. When the test function ζ is non-identically zero and satisfies -Δζ+Vζ0 on Ω, it follows from (1.5) and the strong maximum principle for the Schrödinger operator with potential in Lloc1 (see [1, Théorème 9], [5, Theorem 1] or [19]) that ζ(a)n>0. It is therefore reasonable to expect the validity of the Hopf lemma for (1.1) on the set of points aΩ for which the boundary value problem (1.4) has a solution. This motivates the following:

Definition 1.1.

The exceptional boundary set Σ associated to -Δ+V is the set of points aΩ for which the boundary value problem (1.4) with datum δa does not have a distributional solution.

We can now state the Hopf lemma on 𝒩:

Theorem 2.

Let u be the solution of (1.1) for some nonnegative datum fL(Ω), f0. For every aN, we have

u ^ n ( a ) > 0 if and only if a Σ .

In the case where VLq(Ω) for some q>N, one has 𝒩=Ω and Σ=. Hence,

u ^ n ( a ) > 0 for every  a Ω ;

see Corollary 7.2 below. As one can expect, the validity of the Hopf lemma depends on the behavior of V near the boundary. For example, under the assumption that

(1.6) V C d Ω 2 almost everywhere on  Ω

for some constant C0, Ancona established in [2] (see also the Appendix in [20]) a beautiful characterization of the set of points where (1.4) has a solution: aΩΣ if and only if the Poisson kernel of -Δ at a is a supersolution of (1.1). Using his result and the pointwise behavior of K, we can state the following:

Corollary 1.1.

Assume that V satisfies (1.6) and let u be the solution of (1.1) for some nonnegative datum fL(Ω), f0. For every aN, we have

u ^ n ( a ) > 0 if and only if Ω d Ω 2 ( y ) | y - a | N V ( y ) d y < + .

Quadratic blow-up of the potential as in (1.6) is a threshold for the validity of the Hopf lemma. More precisely:

Corollary 1.2.

Assume that V satisfies

V C d Ω 2 almost everywhere on  Ω

for some C>0. Then N=Ω and, for every solution u of (1.1) with fL(Ω), we have

u ^ n = 0 on  Ω .

Corollary 1.2 is a consequence of our Theorem 1 and a result from Díaz [8] which establishes the existence of a bounded nonnegative eigenfunction u for -Δ+V that satisfies a Dirichlet problem of the type (1.1) and such that u^(a)n=0 for every aΩ.

Although we have introduced the exceptional set Σ by dealing with Dirac masses on Ω, the set Σ allows one to characterize all nonnegative finite Borel measures ν on Ω for which the boundary value problem

(1.7) { - Δ v + V v = 0 in  Ω , v = ν on  Ω ,

has a distributional solution. This is the content of our next theorem that extends a previous result by Véron and Yarur [20]:

Theorem 3.

The boundary value problem (1.7) associated to a nonnegative finite Borel measure ν on Ω has a distributional solution if and only if ν(Σ)=0.

The proof of Theorem 3 is inspired by the recent paper of Orsina and the first author [17] concerning the failure of the strong maximum principle for the Schrödinger operator -Δ+V in the case where V is merely a nonnegative Borel measurable function. In this respect, we introduce in Section 2 a notion of pointwise normal derivative for solutions of (1.1) that is defined everywhere on Ω, but possibly depends on the potential V. In Section 3, we present a counterpart for (1.7) of the notion of duality solution introduced by Malusa and Orsina [13]. The exceptional boundary set Σ is then identified in Section 4 with the set of boundary points at which all such normal derivatives vanish. Using the tools developed in Sections 3 and 4, we prove Theorem 3 in Section 5. Theorems 1 and 2 are proved in Sections 6 and 7, respectively.

2 Pointwise Normal Derivative Associated to the Schrödinger Operator

A property that is common to all solutions of (1.1) concerns the existence of a distributional normal derivative as an element in L1(Ω). This is a general feature that relies on the facts that

u W 0 1 , 1 ( Ω ) and Δ u  is a finite Borel measure on  Ω .

Brezis and the first author proved in [7] that in this general setting there exists a function in L1(Ω), which is denoted by un and coincides with the classical normal derivative when u is a C2 function, that satisfies

u n L 1 ( Ω ) | Δ u | ( Ω )

and

(2.1) Ω u ψ d x = - Ω ψ Δ u - Ω u n ψ d σ for every  ψ C ( Ω ¯ ) ,

where σ=N-1Ω is the surface measure on Ω; see [7, Theorem 1.2] or [18, Proposition 7.3]. We recall that n is the inward unit normal vector, which explains the minus sign in front of the second integral in the right-hand side of (2.1). When u0 almost everywhere on Ω, one additionally has

u n 0 almost everywhere on  Ω ;

see [7, Corollary 6.1] or [18, Lemma 12.15]. Since the mapping

{ u W 0 1 , 1 ( Ω ) : Δ u L 1 ( Ω ) } L 1 ( Ω ) , u u n

is linear, such a property yields a handy comparison principle: if v and w both satisfy (1.1), with possibly different potentials V and data f, and if vw almost everywhere on Ω, then

v n w n almost everywhere on  Ω .

More specific to solutions of (1.1), we show that there is a notion of pointwise normal derivative that is adapted to the Schrödinger operator -Δ+V and used in the proofs of Theorems 1, 2 and 3. For this purpose, let (Vk) be a nondecreasing sequence of nonnegative functions in L(Ω) that converges almost everywhere to V on Ω. The construction of this pointwise normal derivative relies on the main result of this section which is the following:

Proposition 2.1.

Let u be the solution of (1.1) associated to fL(Ω) and denote by uk the solution of

(2.2) { - Δ u k + V k u k = f in  Ω , u k = 0 on  Ω .

Then:

  1. u k u and V k u k V u in L 1 ( Ω ) and almost everywhere on Ω,

  2. ( u k n ) is uniformly bounded on Ω ,

  3. ( u k n ) converges pointwise to a function g : Ω such that g = u n almost everywhere on Ω and, for every N < p ,

    g L ( Ω ) C f L p ( Ω )

    with a constant C > 0 depending on p and Ω . Moreover, g0 on Ω whenever f0 almost everywhere on Ω.

Since Vk is bounded, we have ukC1(Ω¯) and in particular the classical normal derivative ukn is well-defined on Ω. To see why this is true, let wC1(Ω¯) be the solution of

{ - Δ w = | f | in  Ω , w = 0 on  Ω .

The weak maximum principle implies that |uk|w almost everywhere on Ω; thus ukL(Ω). Since Vk and f are bounded, we have ΔukL(Ω), hence ukW2,p(Ω) for every 1<p<; see [11, Theorem 9.15 and Lemma 9.17]. Taking any p>N, it follows from the Morrey–Sobolev embedding theorem that ukC1(Ω¯); see [21, Theorem 6.4.4]. In addition, one has the estimate

(2.3) w C 1 ( Ω ¯ ) C Δ w L p ( Ω ) = C f L p ( Ω )

for some constant C>0 depending on p and Ω. Since

| u k n | w n on  Ω ,

one deduces from (2.3) that

(2.4) u k n L ( Ω ) w C 1 ( Ω ¯ ) C f L p ( Ω ) .

Using Proposition 2.1, we then define the pointwise normal derivative of u with respect to -Δ+V as

u ^ n ( a ) := g ( a ) for every  a Ω .

At first sight, this definition could depend on the choice of approximation (Vk) like

V k = min { V , k } ,

but as we shall see later on it does not; see Remark 5.1. As a consequence of assertion (iii) in Proposition 2.1, u^n is a distributional normal derivative of u.

Before proceeding with the proof of Proposition 2.1, we recall standard estimates for solutions of the Dirichlet problem

(2.5) { - Δ u + V u = μ in  Ω , u = 0 on  Ω ,

where μL1(Ω). By a solution of (2.5), we mean a function uW01,1(Ω)L1(Ω;Vdx) that satisfies the equation in the sense of distributions in Ω. For all 1p<NN-1, the solution exists, is unique and belongs to W01,p(Ω) with

(2.6) u W 1 , p ( Ω ) C μ L 1 ( Ω )

for some constant C>0 depending on p and Ω. This can be deduced from elliptic estimates due to Littman, Stampacchia and Weinberger [12, Theorem 5.1] and from the absorption estimate

(2.7) V u L 1 ( Ω ) μ L 1 ( Ω ) .

The latter inequality can be obtained using as test function a suitable approximation of the sign function sgnu; see [4, Proposition 4.B.3] or [18, Proposition 21.5].

The weak maximum principle for (2.5) that is mentioned in the introduction is justified by the following lemma.

Lemma 2.2.

Let u be the solution of (2.5) involving μL1(Ω). If μ0 almost everywhere on Ω, then u0 almost everywhere on Ω.

The proof of Lemma 2.2 relies on a variant of Kato’s inequality: if wL1(Ω), hL1(Ω,dΩdx) and ν(Ω) satisfy

(2.8) - Ω w Δ ζ d x = Ω h ζ d x + Ω ζ n d ν for every  ζ C 0 ( Ω ¯ ) ,

then

(2.9) - Ω w + Δ ζ d x { w 0 } h ζ d x + Ω ζ n d ν + for every  ζ C 0 ( Ω ¯ ) ζ 0  on  Ω ¯ ;

see [14, Lemma 1.5] or [15, Proposition 1.5.9]. Here (Ω) denotes the vector space of finite Borel measures on Ω and

C 0 ( Ω ¯ ) = { ζ C ( Ω ¯ ) : ζ = 0  on  Ω } .

When ν=0, the integral identity (2.8) implicitly encodes the fact that w=0 on Ω in an average sense as test functions need not have compact support in Ω; see [18, Proposition 20.2] and also [9] for related questions. To deduce the weak maximum principle, it now suffices to take w=-u, h=Vu-μ and ν=0, and then (2.9) becomes

- Ω ( - u ) + Δ ζ d x 0 for every  ζ C 0 ( Ω ¯ ) ζ 0  on  Ω ¯ .

One last ingredient involved in the proof of Proposition 2.1 is the following comparison principle:

Lemma 2.3.

Let V1,V2Lloc1(Ω) be two nonnegative functions such that V1V2 almost everywhere on Ω, and let uiL1(Ω)L1(Ω;VidΩdx), with i{1,2}, be two nonnegative functions such that

- Ω ( u 2 - u 1 ) Δ ζ d x + Ω ( V 2 u 2 - V 1 u 1 ) ζ d x = 0 for every  ζ C 0 ( Ω ¯ ) .

Then u2u1 almost everywhere on Ω.

Lemma 2.3 can be deduced using Kato’s inequality as above by taking w=u2-u1.

Proof of Proposition 2.1.

We assume that f is nonnegative; the general case follows by solving the Dirichlet problem with the positive and negative parts of f, and then conclude using the linearity of the equation in (1.1) and uniqueness of solutions. Hence, by the weak maximum principle, u and uk are nonnegative. Since u satisfies

- Δ u + V k u = f - ( V - V k ) u in the sense of distributions in  Ω ,

we have

- Δ ( u k - u ) + V k ( u k - u ) = ( V - V k ) u in the sense of distributions in  Ω .

One deduces from (2.6) applied to uk-u that

u k - u L 1 ( Ω ) C ( V - V k ) u L 1 ( Ω ) .

By Lebesgue’s dominated convergence theorem, the right-hand side of this inequality tends to 0 as k. Hence uku in L1(Ω). Since (uk) is nonincreasing as a consequence of Lemma 2.3, the convergence also holds everywhere on Ω. The triangle inequality and the absorption estimate (2.7) applied to uk-u imply that

V k u k - V u L 1 ( Ω ) V k ( u k - u ) L 1 ( Ω ) + ( V k - V ) u L 1 ( Ω ) 2 ( V - V k ) u L 1 ( Ω ) ,

and then VkukVu in L1(Ω).

By comparison of normal derivatives, the sequence (ukn) is nonincreasing and nonnegative. In particular, it is uniformly bounded on Ω and converges in L1(Ω) and everywhere on Ω to some nonnegative bounded measurable function g:Ω. Let us show that

g = u n almost everywhere on  Ω ,

where un is the distributional normal derivative of u. For this purpose, we recall that each uk satisfies

(2.10) Ω u k ψ d x = Ω f ψ d x - Ω V k u k ψ d x - Ω u k n ψ d σ

for every ψC(Ω¯). By standard interpolation, which in this case follows from an integration by parts, one also has the estimate

u k L 2 ( Ω ) 2 u k L ( Ω ) Δ u k L 1 ( Ω ) ;

see [18, Lemma 5.8]. We claim that the right-hand side of this inequality is bounded. Indeed, as (uk) is nonincreasing, it is bounded from above by u0. On the other hand, we deduce from the triangle inequality and the absorption estimate (2.7) that

Δ u k L 1 ( Ω ) f L 1 ( Ω ) + V k u k L 1 ( Ω ) 2 f L 1 ( Ω ) ,

which validates our claim.

Since (uk) is bounded in L2(Ω;N) and uku in L1(Ω), we have

u k u weakly in  L 2 ( Ω ; N ) .

Taking the limit as k in (2.10), we obtain

Ω u n ψ d σ = Ω g ψ d σ for every  ψ C ( Ω ¯ ) .

Hence un=g almost everywhere on Ω. The estimate

g L ( Ω ) C f L p ( Ω )

follows from (2.4) since 0gukn on Ω. ∎

3 Duality Solution with Measure Data on the Boundary

We investigate the boundary value problem (1.7) involving a finite Borel measure ν on Ω by comparing two notions of solution based on different choices of test functions.

Definition 3.1.

A function vL1(Ω) is a distributional solution of (1.7) with datum ν(Ω) whenever VvL1(Ω;dΩdx) and

Ω v ( - Δ ζ + V ζ ) d x = Ω ζ n d ν for every  ζ C 0 ( Ω ¯ ) .

The boundary value problem for this type of solutions has been studied by Véron and Yarur [20] with nonnegative potentials VLloc(Ω). In particular, the authors prove that nonnegative measures for which (1.7) has a solution cannot charge Σ; see [20, Theorem 4.4]. Their approach is based on the careful study of some capacity associated to the Poisson kernel of -Δ on Ω. In our case, we rely instead on the concept of duality solution in the spirit of the work of Malusa and Orsina [13] that has its roots in the seminal paper of Littman, Stampacchia and Weinberger [12].

Definition 3.2.

A function vL1(Ω) is a duality solution of (1.7) with datum ν(Ω) whenever

Ω v f d x = Ω ζ f ^ n d ν for every  f L ( Ω ) ,

where ζf is the solution of (1.1) with datum f.

Existence of duality solutions is a straightforward consequence of the Riesz representation theorem:

Proposition 3.1.

The boundary value problem (1.7) has a unique duality solution for every datum νM(Ω).

Proof.

Let N<p<. It follows from Proposition 2.1 that, for every fL(Ω),

| Ω ζ f ^ n d ν | ν ( Ω ) ζ f ^ n L ( Ω ) C ν ( Ω ) f L p ( Ω ) ,

where

ν ( Ω ) := | ν | ( Ω ) .

Hence, the linear functional

L ( Ω ) , f Ω ζ f ^ n d ν

is continuous on L(Ω), endowed with the Lp norm. The Riesz representation theorem implies the existence of a unique vLp(Ω) such that

Ω v f d x = Ω ζ f ^ n d ν for every  f L ( Ω ) ,

where p=pp-1 is the conjugate exponent with respect to p. Hence v is the unique duality solution of (1.7) involving ν. ∎

We now prove that distributional solutions are duality solutions:

Proposition 3.2.

If v is a distributional solution of (1.7) with datum νM(Ω), then v is also a duality solution of (1.7) with datum ν.

For the proof of Proposition 3.2, we need a couple of lemmas. We begin with:

Lemma 3.3.

Assume that (1.7) has a distributional solution v with datum νM(Ω) and let vk be the distributional solution of

(3.1) { - Δ v k + V k v k = 0 in  Ω , v k = ν on  Ω .

Then vkv in L1(Ω).

Proof.

First notice that the function v-vk satisfies

Ω ( v - v k ) ( - Δ ζ + V k ζ ) d x = Ω ( V k - V ) v ζ d x for every  ζ C 0 ( Ω ¯ ) .

Kato’s inequality (2.9) applied to v-vk and -(v-vk) with ν=0 implies that

Ω | v - v k | ( - Δ ζ + V k ζ ) d x Ω sgn ( v - v k ) ( V k - V ) v ζ d x ,

and then

Ω | v - v k | ( - Δ ζ + V k ζ ) d x Ω ζ ( V - V k ) | v | d x

for every ζC0(Ω¯), ζ0 on Ω¯. We take as test function the unique solution of

(3.2) { - Δ θ = 1 in  Ω , θ = 0 on  Ω .

Observing that Vkθ0 and θθL(Ω)dΩ on Ω, we obtain the estimate

(3.3) v k - v L 1 ( Ω ) θ L ( Ω ) ( V - V k ) v L 1 ( Ω ; d Ω d x ) .

Since 0VkV and VvL1(Ω;dΩdx), Lebesgue’s dominated convergence theorem implies that

V k v V v in  L 1 ( Ω ; d Ω d x ) .

The lemma then follows by letting k in (3.3). ∎

Lemma 3.4.

Assume that VLq(Ω) for some q>N. Then the boundary value problem (1.7) has a distributional solution for every νM(Ω). In particular, Σ=.

We recall that if vL1(Ω), fL1(Ω;dΩdx) and ν(Ω) satisfy

- Ω v Δ ζ d x = Ω f ζ d x + Ω ζ n d ν for every  ζ C 0 ( Ω ¯ ) ,

which is the weak formulation of v being a solution of

{ - Δ v = f in  Ω , v = ν on  Ω ,

then:

  1. for every 1p<NN-1, we have vLp(Ω) and there exists a constant C>0 depending on p and Ω such that

    (3.4) v L p ( Ω ) C ( f L 1 ( Ω ; d Ω d x ) + ν ( Ω ) ) ,

  2. for every 1p<NN-1 and every ωΩ, we have vLp(ω;N) and there exists a constant C>0 depending on p and ω such that

    (3.5) v L p ( ω ; N ) C ( f L 1 ( Ω ; d Ω d x ) + ν ( Ω ) ) .

We refer the reader to [15, Theorem 1.2.2] for a proof of these assertions.

Proof of Lemma 3.4.

Let (gk) be a sequence of smooth functions on Ω such that gkL1(Ω)ν(Ω) and gkν weak* in (Ω), i.e.,

lim k Ω ϕ g k d x = Ω ϕ d ν for every  ϕ C ( Ω ) .

Such a sequence can be obtained, for example, from a convolution of ν with a sequence of mollifiers. Denote by vk the distributional solution of (1.7) associated to gk. Given ωΩ, we deduce from (3.4) and (3.5) that

v k W 1 , 1 ( ω ) C 1 ( V v k L 1 ( Ω ; d Ω d x ) + g k L 1 ( Ω ) )

for some constant C1>0 depending on ω. Taking a subsequence if necessary, we have

(3.6) v k v almost everywhere on  Ω .

On the other hand, since q<NN-1, we deduce from (3.4) that

v k L q ( Ω ) C 2 ( V v k L 1 ( Ω ; d Ω d x ) + g k L 1 ( Ω ) )

for some constant C2>0 depending on q and Ω. Hence (vk) is bounded in Lq(Ω). This, together with (3.6), implies that

v k v weakly in  L q ( Ω ) .

Recalling that VLq(Ω) and taking the limit as k in the equation

Ω v k ( - Δ ζ + V ζ ) d x = Ω ζ n g k d σ for every  ζ C 0 ( Ω ¯ ) ,

we get the conclusion. ∎

We now turn to the

Proof of Proposition 3.2.

We first assume that V is bounded. In this case, ζfC01(Ω¯) for every fL(Ω). Since ζf need not be smooth enough to be used as test function for v, we approximate ζf in C1(Ω¯) by a sequence (ζfk) in C0(Ω¯), where (fk) is a bounded sequence in L(Ω) such that fkf almost everywhere on Ω. For this purpose, we follow the construction given in [17]: for each k we define the function gk=ρk*g, where g=f-Vζf and (ρk) is a sequence of mollifiers, and we denote by wkC0(Ω¯) the solution of

{ - Δ w k = g k in  Ω , w k = 0 on  Ω .

Observe that wk=ζfk with fk=gk+Vwk. Moreover, estimate (2.3) ensures that for some fixed N<p<,

ζ f k - ζ f C 1 ( Ω ¯ ) C g k - g L p ( Ω ) .

Letting k in this estimate, we have

ζ f k ζ f uniformly on  Ω    and    ζ f k n ζ f n uniformly on  Ω .

Since

Ω v f k d x = Ω v ( - Δ ζ f k + V ζ f k ) d x = Ω ζ f k n d ν ,

taking the limit as k, we obtain

Ω v f d x = Ω ζ f n d ν .

In the general case where VLloc1(Ω), let vk be the solution of

{ - Δ v k + V k v k = 0 in  Ω , v k = ν on  Ω ,

whose existence is ensured by Lemma 3.4. Given fL(Ω), we denote by zkC01(Ω¯) the solution of

{ - Δ z k + V k z k = f in  Ω , z k = 0 on  Ω .

It follows from the first part of the proof that

Ω v k f d x = Ω z k n d ν .

Proposition 2.1 implies that (zkn) is uniformly bounded and converges pointwise to ζf^n on Ω. By Lemma 3.3, we obtain

Ω v f d x = Ω ζ f ^ n d ν .

4 Pointwise Normal Derivative on the Exceptional Set Σ

In this section, we characterize the exceptional boundary set Σ using the pointwise normal derivative with respect to the Schrödinger operator -Δ+V introduced in Section 2. We prove:

Proposition 4.1.

For every aΩ, we have aΣ if and only if

ζ f ^ n ( a ) = 0 for every  f L ( Ω ) .

As a fundamental property that is used in the proofs of Proposition 4.1 and Theorem 3, we first extend Lemma 3.3 to the case where (1.7) need not have a distributional solution.

Proposition 4.2.

Let νM(Ω) be a nonnegative measure and let v be the duality solution of (1.7) associated to ν. We have:

  1. if v k is the distributional solution of ( 3.1 ), then v k v in L 1 ( Ω ) ,

  2. there exists a nonnegative measure λ ( Ω ) such that v is the distributional solution of ( 1.7 ) associated to ν - λ .

We recall the following estimate whose proof is sketched for the convenience of the reader:

Lemma 4.3.

If v is a distributional solution of (1.7) with νM(Ω), then

v L 1 ( Ω ) + V v L 1 ( Ω ; d Ω d x ) C ν ( Ω )

for some constant C>0 depending on Ω.

Proof.

One deduces using Kato’s inequality (2.9) with w=v and w=-v that

Ω | v | ( - Δ ζ + V ζ ) d x ζ n L ( Ω ) ν ( Ω )

for every ζC0(Ω¯), ζ0 on Ω¯. Take as test function the solution θ of (3.2). As a consequence of the classical Hopf lemma, there exists C1>0 such that θC1dΩ on Ω. Therefore, by nonnegativity of V,

v L 1 ( Ω ) + C 1 V v L 1 ( Ω ; d Ω d x ) θ n L ( Ω ) ν ( Ω ) .

Proof of Proposition 4.2.

By a straightforward counterpart of Lemmas 2.2 and 2.3 for distributional solutions of (1.7), the sequence (vk) is nonnegative and nonincreasing. Hence (vk) converges in L1(Ω) to some nonnegative function w. Let fL(Ω) and let uk be the solution of (2.2). Proposition 3.2 implies that

Ω v k f d x = Ω u k n d ν .

By Proposition 2.1, the sequence (ukn) is uniformly bounded and converges pointwise to ζf^n on Ω. Taking the limit as k in the identity above, we deduce from Lebesgue’s dominated convergence theorem that

Ω w f d x = Ω ζ f ^ n d ν for every  f L ( Ω ) .

We have thus proved that w is a duality solution of (1.7) involving ν. By uniqueness of duality solutions, we have v=w.

For every k1, we have

0 V k v k V v 1 almost everywhere on  Ω .

Since v1 is subharmonic, it is locally bounded on Ω; see [21, Theorem 8.1.5]. Then, Lebesgue’s dominated convergence theorem implies that

V k v k V v in  L loc 1 ( Ω ) .

Let θ be the unique solution of (3.2). Since 0<θθL(Ω)dΩ on Ω, by Lemma 4.3 the sequence (Vkvkθ) is bounded in L1(Ω). Therefore, taking a subsequence if necessary, we may assume that there exist nonnegative finite Borel measures μ on Ω and τ on Ω such that, for every ψC(Ω¯),

(4.1) lim k Ω V k v k θ ψ d x = Ω ψ d μ + Ω ψ d τ .

On the other hand, for every φCc(Ω),

lim k Ω V k v k θ φ d x = Ω V v θ φ d x .

Hence μ=Vvθdx. Given ζC0(Ω¯), we define γ=ζθ on Ω. Since θn>0 on Ω, the function γ extends continuously to Ω, and

γ = ζ n 1 θ n on  Ω .

Taking ψ=γ in (4.1), we obtain

lim k Ω V k v k ζ d x = Ω V v ζ d x + Ω ζ n 1 θ n d τ for every  ζ C 0 ( Ω ¯ ) .

The result follows with λ=1θnτ. ∎

Another ingredient involved in the proof of Proposition 4.1 is the inverse maximum principle for distributional solutions of (1.7); see [6, Lemma 1].

Lemma 4.4.

Let νM(Ω) and let hL1(Ω;dΩdx). Assume that vL1(Ω) satisfies

- Ω v Δ ζ d x = Ω h ζ d x + Ω ζ n d ν for every  ζ C 0 ( Ω ¯ ) .

If v0 almost everywhere on Ω, then ν0 on Ω.

Given aΩ, we denote by Pa the duality solution of (1.7) associated to the Dirac measure δa, that is,

(4.2) ζ f ^ n ( a ) = Ω P a f d x for every  f L ( Ω ) .

One deduces from the definition of duality solution using f=χ{Pa<0} as test function that

P a 0 almost everywhere on  Ω .

We apply this simple observation:

Proof of Proposition 4.1.

We first assume that aΣ. By Proposition 4.2, Pa is a distributional solution of (1.7) with datum δa-λ for some nonnegative measure λ(Ω). Since Pa0 almost everywhere on Ω, we deduce from Lemma 4.4 that

δ a λ 0 on  Ω .

Thus, λ=αδa for some 0α1. If we had α1, then Pa1-α would be a distributional solution of (1.4), in contradiction with the assumption that aΣ. Hence, α=1 and

Ω P a ( - Δ ζ + V ζ ) d x = 0 for every  ζ C 0 ( Ω ¯ ) .

Taking ζ=θ, where θ satisfies (3.2), we deduce that

Ω P a d x = 0 .

Since Pa is nonnegative, we have Pa=0 almost everywhere on Ω. The representation formula (4.2) then implies that

ζ f ^ n ( a ) = 0 for every  f L ( Ω ) .

For the converse, one deduces from the assumption on a and (4.2) applied to f1 that

Ω P a d x = 0 .

Hence Pa=0 almost everywhere on Ω, so that Pa cannot be a distributional solution of (1.7) involving δa. Since Pa is the only candidate for such a solution due to Proposition 3.2, we conclude that aΣ. ∎

5 Proof of Theorem 3

() Let v be the duality solution of (1.7) associated to ν. Proposition 4.2 implies the existence of a nonnegative measure λ(Ω) such that v is a distributional solution of (1.7) involving ν-λ. We claim that λ(ΩΣ)=0. By Proposition 3.2, v is also a duality solution of (1.7) with datum ν-λ. Hence

Ω ζ f ^ n d ν = Ω v f d x = Ω ζ f ^ n d ( ν - λ ) for every  f L ( Ω ) ,

which implies that

Ω ζ f ^ n d λ = 0 for every  f L ( Ω ) .

By Proposition 4.1, we have ζ1^n>0 on ΩΣ. Since λ is nonnegative, we conclude that λ(ΩΣ)=0 as claimed. On the other hand, since v0 almost everywhere on Ω, we deduce from Lemma 4.4 that

ν λ 0 on  Ω .

By assumption, ν(Σ)=0. Hence λ(Σ)=0. We thus have

λ ( Ω ) = λ ( Ω Σ ) + λ ( Σ ) = 0 ,

that is, λ=0.

() Let v be the distributional solution of (1.7) associated to ν. Proposition 3.2 implies that v is also a duality solution of (1.7) involving the same datum. By Proposition 4.1, we have

Ω v f d x = Ω ζ f ^ n d ν = Ω ζ f ^ n d ν Ω Σ for every  f L ( Ω ) ,

so that v is also a duality solution of (1.7) with datum νΩΣ. The reverse implication in Theorem 3 implies that (1.7) associated to νΩΣ has a unique distributional solution z. But then, Proposition 3.2 ensures that z is also a duality solution of (1.7) with datum νΩΣ. Since duality solutions are unique, we have v=z almost everywhere on Ω. Thus, v is a distributional solution of (1.7) with both ν and νΩΣ, which implies that ν=νΩΣ, and then ν(Σ)=0.

Remark 5.1.

Using Theorem 3 and Proposition 4.1, we can now explain why u^n does not depend upon the particular choice of approximating sequence (Vk) in Proposition 2.1. Indeed, we have by Proposition 4.1 that

u ^ n ( a ) = 0 for every  a Σ ,

while Σ is independent of (Vk). When aΩΣ, it follows from Theorem 3 that Pa is a distributional solution of (1.4), whose definition does not involve (Vk). Thus, by the representation formula (4.2), u^n is independent of (Vk) also on ΩΣ.

6 Proof of Theorem 1

We deduce Theorem 1 as a consequence of the following proposition:

Proposition 6.1.

Let u be the solution of (1.1) for some nonnegative datum in L(Ω) such that u^ has a classical normal derivative at aΩ which satisfies (1.3). If v is another solution of (1.1) for some nonnegative datum in L(Ω), and if vu almost everywhere on Ω, then v^ also has a classical normal derivative at a which satisfies (1.3).

We recall that whenever v satisfies (1.1) with datum hL(Ω), one has

v ^ ( x ) = Ω G ( x , y ) ( h - V v ) ( x ) d x for every  x Ω .

Using standard estimates on the Green function G, by boundedness of h one shows that

lim t 0 Ω G ( a + t n , y ) t h ( y ) d y = Ω K ( a , y ) h ( y ) d y ,

where n=n(a). The main difficulty in the proof of Proposition 6.1 thus consists in getting that

lim t 0 Ω G ( a + t n , y ) t V v ( y ) d y = Ω K ( a , y ) V v ( y ) d y .

Proof of Proposition 6.1.

Let (εk) be a nonincreasing sequence of positive numbers converging to 0. We define on Ω

g k ( y ) = G ( a + ε k n , y ) ε k .

On the one hand, we have gk(y)K(a,y) for all yΩ. On the other hand, by assumption on u,

Ω g k ( y ) V u ( y ) d y Ω K ( a , y ) V u ( y ) d y

or, equivalently,

g k u L 1 ( Ω ; V d x ) K ( a , ) u L 1 ( Ω ; V d x ) .

We thus have pointwise convergence of (gku) and also convergence of norms. Therefore,

g k u K ( a , ) u in  L 1 ( Ω ; V d x ) ,

which is a special case of the Brezis–Lieb lemma [3]; see [21, Proposition 4.2.6]. Since 0vu, we deduce from Lebesgue’s dominated convergence theorem that gkvK(a,)v in L1(Ω;Vdx). The conclusion is now straightforward. ∎

For the proof of Theorem 1, we also need the following:

Lemma 6.2.

Let u be the solution of (1.1) for some nonnegative datum fL(Ω). Then, for every aΩ, we have

lim sup ε 0 u ^ ( a + ε n ) ε u ^ n ( a ) Ω K ( a , y ) ( f - V u ) ( y ) d y .

Proof.

Let uk satisfy (2.2). By comparison, we have u^uk on Ω. Hence,

lim sup ε 0 u ^ ( a + ε n ) ε u k n ( a ) for every  a Ω .

Then, as k we deduce the first inequality in the statement.

Next, since Δuk is bounded, for every aΩ we have

u k n ( a ) = Ω K ( a , y ) ( - Δ u k ( y ) ) d y = Ω K ( a , y ) ( f - V k u k ) ( y ) d y .

By Proposition 2.1, the sequence (Vkuk) converges almost everywhere to Vu. Fatou’s lemma implies that, for every aΩ,

Ω K ( a , y ) V u ( y ) d y lim inf k Ω K ( a , y ) V k u k ( y ) d y .

Observe that the right-hand side is finite since the sequence (ukn) is bounded. Hence, for every aΩ, we have

u ^ n ( a ) = lim k u k n ( a ) = lim k Ω K ( a , y ) ( f - V k u k ) ( y ) d y Ω K ( a , y ) ( f - V u ) ( y ) d y ,

which implies the second inequality in the statement. ∎

Proof of Theorem 1.

The reverse implication () follows from Proposition 6.1 and the fact that

u ζ f L ( Ω ) = f L ( Ω ) ζ 1 almost everywhere on  Ω .

We now prove the direct implication (). One shows the existence of a constant C>0 such that, for every ε>0, the solution vε of the Dirichlet problem

{ - Δ v ε + V v ε = χ { u C > ε } in  Ω , v ε = 0 on  Ω ,

satisfies vεuε almost everywhere on Ω; this is a consequence of Kato’s inequality, as explained in [16, proof of Proposition 4.1]. Let (εj) be a nonincreasing sequence of positive numbers converging to 0. Since χ{uC>εj}1, by Proposition 6.1 the function vεj^ has a normal derivative at a which satisfies (1.3), which means that

v ε j ^ n ( a ) = Ω K ( a , y ) ( χ { u C > ε j } - V v ε j ) ( y ) d y .

By the strong maximum principle for the Schrödinger operator with potential in Lloc1(Ω), we have u>0 almost everywhere on Ω. Hence

χ { u C > ε j } 1 almost everywhere on  Ω .

The convergence thus holds in L1(Ω), which implies that

v ε j ζ 1 in  L 1 ( Ω ) .

As the sequence (vεj) is nondecreasing, we deduce from Levi’s monotone convergence theorem that

lim j v ε j ^ n ( a ) = Ω K ( a , y ) ( 1 - V ζ 1 ( y ) ) d y .

Since vεj^ζ1^ on Ω, we also have, by classical comparison of limits,

lim j v ε j ^ n ( a ) lim inf j ζ 1 ^ ( a + ε j n ) ε j .

Lemma 6.2 implies that

lim inf k ζ 1 ^ ( a + ε k n ) ε k lim sup k ζ 1 ^ ( a + ε k n ) ε k Ω K ( a , y ) ( 1 - V ζ 1 ( y ) ) d y .

Combining the inequalities above, we deduce that ζ1^(a)n exists and

ζ 1 ^ n ( a ) = lim k v ε k ^ n ( a ) = Ω K ( a , y ) ( 1 - V ζ 1 ( y ) ) d y .

Hence, by definition, a𝒩. ∎

7 Proof of Theorem 2

We first prove a version of the Hopf lemma in terms of the pointwise normal derivative associated to the Schrödinger operator -Δ+V. In this case, the answer does not involve the set 𝒩.

Proposition 7.1.

Let u be the solution of (1.1) for some nonnegative datum fL(Ω), f0. Then, for every aΩ, we have

u ^ n ( a ) > 0 if and only if a Σ .

Proof.

By Proposition 4.1, we only have to prove that u^n>0 on ΩΣ. For this purpose, let aΩΣ. In this case, Pa is both a duality and a distributional solution of (1.7) involving δa. As a distributional solution, it satisfies the strong maximum principle for the Schrödinger operator with potential in Lloc1(Ω). Hence,

(7.1) P a > 0 almost everywhere on  Ω .

As a duality solution, Pa satisfies the representation formula (4.2). Since f is nonzero, we then deduce from this formula and (7.1) that

u ^ n ( a ) > 0 .

We now turn to the

Proof of Theorem 2.

By Theorem 1, the classical normal derivative u^n exists on 𝒩 and satisfies (1.3). A direct application of Lemma 6.2 gives, for every a𝒩,

u ^ n ( a ) u ^ n ( a ) Ω K ( a , y ) ( f - V u ) ( y ) d y .

As the integral in the right-hand side equals u^(a)n, equality holds everywhere and we get

u ^ n = u ^ n on  𝒩 .

The theorem then follows from Proposition 7.1. ∎

We conclude this section with the following particular case of Theorem 2.

Corollary 7.2.

Assume that VLq(Ω) for some q>N. Then, for every solution u of (1.1) involving a nonnegative datum fL(Ω), f0, the normal derivative of u^ exists at every point aΩ and satisfies

u ^ n ( a ) > 0 .

Proof.

We prove that 𝒩=Ω and Σ=. Let u be the solution of (1.1) involving some nonnegative datum fL(Ω). Since uL(Ω), we have ΔuLq(Ω), and then uW2,q(Ω). The Morrey–Sobolev embedding theorem ensures that uC1(Ω¯), which gives 𝒩=Ω. That Σ= follows from Lemma 3.4. We then have the conclusion using Theorem 2. ∎


Dedicated to Laurent Véron, with affection and esteem, on the occasion of his 70th birthday.



Communicated by Julián López-Gómez and Patrizia Pucci


Award Identifier / Grant number: J.0020.18

Funding statement: The first author (A.C.P.) was supported by the Fonds de la Recherche scientifique (F.R.S.–FNRS) under research grant J.0020.18.

Acknowledgements

The authors would like to thank Moshe Marcus for discussions on the definition of the weak normal derivative.

References

[1] A. Ancona, Une propriété d’invariance des ensembles absorbants par perturbation d’un opérateur elliptique, Comm. Partial Differential Equations 4 (1979), no. 4, 321–337. 10.1080/03605307908820097Search in Google Scholar

[2] A. Ancona, Positive solutions of Schrödinger equations and fine regularity of boundary points, Math. Z. 272 (2012), no. 1–2, 405–427. 10.1007/s00209-011-0940-5Search in Google Scholar

[3] H. Brezis and E. Lieb, A relation between pointwise convergence of functions and convergence of functionals, Proc. Amer. Math. Soc. 88 (1983), no. 3, 486–490. 10.1007/978-3-642-55925-9_42Search in Google Scholar

[4] H. Brezis, M. Marcus and A. C. Ponce, Nonlinear elliptic equations with measures revisited, Mathematical Aspects of Nonlinear Dispersive Equations, Ann. of Math. Stud. 163, Princeton University, Princeton (2007), 55–109. 10.1515/9781400827794.55Search in Google Scholar

[5] H. Brezis and A. C. Ponce, Remarks on the strong maximum principle, Differential Integral Equations 16 (2003), no. 1, 1–12. 10.57262/die/1356060694Search in Google Scholar

[6] H. Brezis and A. C. Ponce, Reduced measures on the boundary, J. Funct. Anal. 229 (2005), no. 1, 95–120. 10.1016/j.jfa.2004.12.001Search in Google Scholar

[7] H. Brezis and A. C. Ponce, Kato’s inequality up to the boundary, Commun. Contemp. Math. 10 (2008), no. 6, 1217–1241. 10.1142/S0219199708003241Search in Google Scholar

[8] J. I. Díaz, On the ambiguous treatment of the Schrödinger equation for the infinite potential well and an alternative via singular potentials: the multi-dimensional case, SeMA J. 74 (2017), no. 3, 255–278; erratum, SeMA J. 75 (2018), no. 3, 563–568. Search in Google Scholar

[9] J. I. Díaz and J. M. Rakotoson, On the differentiability of very weak solutions with right-hand side data integrable with respect to the distance to the boundary, J. Funct. Anal. 257 (2009), no. 3, 807–831. 10.1016/j.jfa.2009.03.002Search in Google Scholar

[10] L. C. Evans, Partial Differential Equations, 2nd ed., Grad. Stud. in Math. 19, American Mathematical Society, Providence, 2010. Search in Google Scholar

[11] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Class. Math., Springer, Berlin, 2001. 10.1007/978-3-642-61798-0Search in Google Scholar

[12] W. Littman, G. Stampacchia and H. F. Weinberger, Regular points for elliptic equations with discontinuous coefficients, Ann. Sc. Norm. Super. Pisa Cl. Sci. (3) 17 (1963), 43–77. Search in Google Scholar

[13] A. Malusa and L. Orsina, Existence and regularity results for relaxed Dirichlet problems with measure data, Ann. Mat. Pura Appl. (4) 170 (1996), 57–87. 10.1007/BF01758983Search in Google Scholar

[14] M. Marcus and L. Veron, The boundary trace of positive solutions of semilinear elliptic equations: The supercritical case, J. Math. Pures Appl. (9) 77 (1998), no. 5, 481–524. 10.1016/S0021-7824(98)80028-7Search in Google Scholar

[15] M. Marcus and L. Véron, Nonlinear Second Order Elliptic Equations Involving Measures, De Gruyter Ser. Nonlinear Anal. Appl. 21, De Gruyter, Berlin, 2014. 10.1515/9783110305319Search in Google Scholar

[16] L. Orsina and A. C. Ponce, Hopf potentials for the Schrödinger operator, Anal. PDE 11 (2018), no. 8, 2015–2047. 10.2140/apde.2018.11.2015Search in Google Scholar

[17] L. Orsina and A. C. Ponce, On the nonexistence of Green’s function and failure of the strong maximum principle, J. Math. Pures Appl. (9) 134 (2020), 72–121. 10.1016/j.matpur.2019.06.001Search in Google Scholar

[18] A. C. Ponce, Elliptic PDEs, Measures and Capacities, EMS Tracts Math. 23, European Mathematical Society (EMS), Zürich, 2016. 10.4171/140Search in Google Scholar

[19] N. S. Trudinger, On the positivity of weak supersolutions of nonuniformly elliptic equations, Bull. Austral. Math. Soc. 19 (1978), no. 3, 321–324. 10.1017/S0004972700008868Search in Google Scholar

[20] L. Véron and C. Yarur, Boundary value problems with measures for elliptic equations with singular potentials, J. Funct. Anal. 262 (2012), no. 3, 733–772. 10.1016/j.jfa.2010.12.032Search in Google Scholar

[21] M. Willem, Functional Analysis. Fundamentals and Applications, Cornerstones, Birkhäuser/Springer, New York, 2013. 10.1007/978-1-4614-7004-5Search in Google Scholar

Received: 2020-01-10
Revised: 2020-02-20
Accepted: 2020-02-20
Published Online: 2020-03-24
Published in Print: 2020-05-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 20.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2020-2078/html
Scroll to top button