• No results found

A mean value formula for the variational p-Laplacian

N/A
N/A
Protected

Academic year: 2022

Share "A mean value formula for the variational p-Laplacian"

Copied!
33
0
0

Loading.... (view fulltext now)

Full text

(1)

 2021 The Author(s)c

https://doi.org/10.1007/s00030-021-00688-6

Nonlinear Differential Equations and Applications NoDEA

A mean value formula for the variational p-Laplacian

F´ elix del Teso and Erik Lindgren

Abstract. We prove a new asymptotic mean value formula for the p- Laplace operator,

Δ

p

u = div(|∇u|

p−2

∇u), 1 < p < ∞

valid in the viscosity sense. In the plane, and for a certain range of p, the mean value formula holds in the pointwise sense. We also study the exis- tence, uniqueness and convergence of the related dynamic programming principle.

Mathematics Subject Classification. 35J60, 35J70, 35J75, 35J92, 35D40, 35B05, 49L20.

Keywords. p-Laplacian, Mean value property, Viscosity solutions, Dynamic programming principle.

Contents 1. Introduction

2. Main results 3. Related results

4. Comments on our results Comments on Theorem 2.3

Comments on the definition of viscosity solution and the proof Theorem 2.5

Comments on the limit p → 1 More general datum

Plan of the paper

5. Notation and prerequisites

6. The mean value formula for C

2

-functions 7. Viscosity solutions

8. The pointwise property in the plane

9. Study of the dynamic programming principle

0123456789().: V,-vol

(2)

9.1. Existence and uniqueness: the proof of Theorem 2.5 (i) 9.2. Convergence: The proof of Theorem 2.5 (ii)

Acknowledgements

Appendix A: Auxiliary inequalities References

1. Introduction

It is well known that a function is harmonic if and only if it is satisfies

Br

(u(x + y) − u(x)) dy = 0,

for all r small enough. This can be relaxed: a function is harmonic at a point x if and only if

Br

(u(x + y) − u(x)) dy = o(r

2

), as r → 0.

In this paper, we study a new

1

asymptotic mean value property for p-harmonic functions, i.e., solutions of the equation

Δ

p

u = 0.

Here p ∈ (1, ∞) and Δ

p

is the p-Laplace operator

Δ

p

u = div(|∇u|

p−2

∇u), (1.1)

the first variation of the functional u →

ˆ

Ω

|∇u|

p

dx.

Our result implies in particular that a function is p-harmonic at a point x if and only if it is satisfies

Br

|u(x + y) − u(x)|

p−2

(u(x + y) − u(x))dy = o(r

p

), as r → 0.

The major strength and novelty of our mean value formula is that it recovers the variational p-Laplace operator (1.1) in contrast to the other known mean value formulas that recover the normalized p-Laplacian,

Δ

Np

u = 1

p |∇u|

2−p

Δ

p

u.

In particular, it allows us to deal with non-homogeneous problems of the form

−Δ

p

u = f with f = 0, which was not possible with previous approaches.

The drawback is that it cannot be written in the form u(x) = A

r

[u](x) + o(r

p

)

1It is new for 1 < p < 2. For p≥ 2, it has also been found in [6].

(3)

for some monotone operator A

r

. However, the mean value formula is still monotonically increasing in u and monotonically decreasing in u(x), which is decisive in the context of viscosity solutions.

2. Main results

Our main results concern the asymptotic behavior as r → 0 of the quantities I

rp

[φ](x) = 1

C

d,p

r

p ∂Br

|φ(x + y) − φ(x)|

p−2

(φ(x + y) − φ(x)) dσ(y) and

M

pr

[φ](x) = 1

D

d,p

r

p Br

|φ(x + y) − φ(x)|

p−2

(φ(x + y) − φ(x)) dy, where

C

d,p

= 1

2

∂B1

|y

1

|

p

dσ(y), D

d,p

= dC

d,p

p + d and d is the dimension.

2

Our first result, that will be proved in Sect. 6, provides the mean value formula for C

2

functions. It reads:

Theorem 2.1. Let p ∈ (1, ∞), x ∈ R

d

and φ ∈ C

2

(B

R

(x)) for some R > 0. If p ∈ (1, 2) assume also that |∇φ(x)| = 0. Then, we have

I

rp

[φ](x) = Δ

p

φ(x) + o

r

(1) and M

pr

[φ](x) = Δ

p

φ(x) + o

r

(1), as r → 0.

The second of our results relates the mean value property in the viscosity sense to the p-Laplace equation.

Theorem 2.2. Let Ω ⊂ R

d

be bounded and open, p ∈ (1, ∞) and f be a contin- uous function. Then u is a viscosity solution of

−I

rp

[u] = f + o

r

(1), as r → 0, in Ω if and only if it is a viscosity solution of

−Δ

p

u = f in Ω.

We refer to Sect. 7 for the proof of the above result, and to Sect. 5 for the definition of viscosity solutions.

We wish to point out that for p ≥ 2, the above results have been proved independently in [6], see Proposition 2.10 and Theorem 2.12 therein.

2Cd,p can be expressed in terms of the so-called β-functions. We thank ´Angel Arroyo and an anonymous referee for pointing this out.

(4)

Our third result states that in the plane, and for a certain range of p, functions that satisfy the (homogeneous) mean value property in a pointwise sense are the same as the p-harmonic functions. Let p

0

be the root of

2

 −p + 

(36(p − 1) + (p − 2)

2



 − p + 

16(p − 1) + (p − 2)

2



2

= 1 p − 1 that lies in the interval (1, 2). We have p

0

≈ 1.117.

Theorem 2.3. Let Ω ⊂ R

2

be bounded and open, and p ∈ (p

0

, ∞). Then a continuous function u satisfies

−I

rp

[u] = o

r

(1), as r → 0,

in the pointwise sense in Ω if and only if it is a viscosity solution of

−Δ

p

u = 0 in Ω.

We refer to Sect. 8 for the proof of this theorem and to page 4 for a heuristic explanation on the technical limitation p > p

0

.

Remark 2.4. Theorem 2.2 and Theorem 2.3 remain true if one replaces I

rp

by M

pr

.

The fourth and the last of our main results concerns the associated dy- namic programming principle. Consider the following boundary value problem

 −M

pr

[U

r

](x) = f (x), x ∈ Ω

U

r

(x) = G(x), x ∈ ∂Ω

r

:= {x ∈ Ω

c

: dist(x, Ω) ≤ r}, (2.1) where G is a continuous extension of g from ∂Ω to ∂Ω

r

.

Theorem 2.5. Let Ω ⊂ R

d

be a bounded, open and C

2

domain, p ∈ (1, ∞), f be a continuous function in Ω and g a continuous function on ∂Ω. Then

(i) there is a unique classical solution U

r

of (2.1),

(ii) U

r

→ u as r → 0 uniformly in Ω, where u is the viscosity solution of

 −Δ

p

u(x) = f (x), x ∈ Ω

u(x) = g(x), x ∈ ∂Ω. (2.2)

Remark 2.6. We have stated all our results in the context of viscosity solutions.

Since weak and viscosity solutions are equivalent (cf. [12] and [13]), the same results hold true for weak solutions.

3. Related results

Recently, there has been a surge of interest around mean value properties of equations involving the p-Laplacian. In [20], it is proved that a function is p-harmonic if and only if

u(x) = A

r

[u](x) + o(r

2

),

(5)

as r → 0. Here A

r

is the monotone operator A

r

[u](x) = p − 2

2(p + d)

 max

Br(x)

u + min

Br(x)

u

+ 2 + d

p + d

Br(x)

u(y)dy

. This was first proved to be valid in the viscosity sense. In [19], this was proved to hold in the pointwise sense, in the plane and for 1 < p < ˆ p ≈ 9.52. Shortly after, this was extended to all p ∈ (1, ∞), in [3]. Linked to a mean value for- mula, there is a corresponding dynamic programming principle (DPP), which is the solution U

r

of the problem U

r

= A

r

[U

r

] subject to the corresponding boundary conditions. The typical result is to show that U

r

→ u where u is a viscosity solution of the boundary value problem associated to the p-Laplacian.

The above mentioned results are based on the following identity for the so-called normalized p-Laplacian

Δ

Np

u := 1

p Δu + p − 2

p |∇u|

−2

Δ

u (3.1)

and the now well-known mean value formulas for the Laplacian and ∞-Lapla- cian. More precisely, for a smooth function φ,

A

r

[φ](x) − φ(x)

C

d,p

r

2

− Δ

Np

φ(x) = o

r

(1),

for some constant C

d,p

> 0. In the last years, several other mean value for- mulas for the normalized p-Laplacian have been found, and the corresponding program (equivalence of solutions in the viscosity and classical sense and study of the associated dynamic programming principle) has been developed. See for instance [2,7,9,14, 16–18], and [22].

We also want to mention [8] and [10], where two other nonlinear mean value formulas are studied, with some similarities with ours.

It is noteworthy to mention that our results are also related to asymptotic mean value formulas for nonlocal operators involving for instance fractional or non-local versions of the p-Laplacian. See [1] and [6]. In particular, in [6], a mean value formula and equivalence of viscosity solutions have been obtained in the case p ≥ 2.

4. Comments on our results

Comments on Theorem 2.3

The curious reader might wonder why we are not able to prove the pointwise validity for the mean value formula for the full range p ∈ (1, ∞), as in [3].

To make a long story short this has to do with the fact that the mean value

formula considered in [3] has quadratic scaling. It is therefore enough with

an error term of order strictly larger than 2. The mean value formula in the

present paper however, has scaling p/(p −1), which makes it necessary with an

error term of order strictly larger than p/(p − 1). When p < 2, this certainly

comes with some difficulties that for the moment forces us to assume the larger

(6)

lower bound p > p

0

. However, we still believe that such a result holds in the full range p ∈ (1, ∞).

Comments on the definition of viscosity solution and the proof Theorem 2.5 The operator Δ

p

φ(x) is singular in the range p ∈ (1, 2) when ∇φ(x) = 0. This fact forces us to choose a modified version of viscosity solution (see Definition 5.1. As expected, when p ≥ 2 or ∇φ(x) = 0, this definition is equivalent to the usual one (cf. [4]).

This definition of viscosity solution adds some extra technicalities in the proofs of this manuscript. In particular in the proof of convergence of Theorem 2.5. Here we follow the classical program developed in [5] and adapted to the context of homogeneous problems involving the p-Laplacian.

Comments on the limit p → 1

Formally, when p = 1 the mean value formula becomes

Br

sign(u(x + y) − u(x))dy = o(r), as r → 0, or

1 r

 |{y ∈ B

r

: u(x + y) > u(x) }|

|B

r

| |{y ∈ B

r

: u(x + y) < u(x) }|

|B

r

|

= o

r

(1), which could relate to 1-harmonic functions. We plan to study this possibility in the future.

More general datum

It would also be interesting to study problems where f = f (x, u, ∇u(x)) has the right monotonicity assumptions as described in [21]. Theorem 2.2 follows in a straightforward way. However, the convergence of dynamic programming principles like in Theorem 2.5 would require a more delicate study, both in terms of existence and properties of the r-scheme, and the study of convergence based on the Barles-Souganidis approach.

Plan of the paper

The plan of the paper is as follows. In Sect. 5, we introduce some notation

and the notions of viscosity solutions. This is followed by Sect. 6, where we

prove the mean value formula for C

2

functions. This result is then used in

Sect. 7, where we prove the mean value formula for viscosity solutions. In

Sect. 8, we prove that in dimension d = 2, and for a certain range of p,

functions that satisfy the (homogeneous) mean value property in a pointwise

sense are the same as the p-harmonic functions. In Sect. 9, we study existence,

uniqueness and convergence for the dynamical programming principle. Finally,

in the Appendix, we prove and state some auxiliary inequalities.

(7)

5. Notation and prerequisites

Throughout this paper, d will denote the dimension and we will for p ∈ (1, ∞) use the notation

J

p

(t) = |t|

p−2

t.

We now define viscosity solutions of the related equations and mean value properties. We adopt the definition of solutions from [12].

Definition 5.1. (Viscosity solutions of the equation) Suppose that f is contin- uous function in Ω. We say that a lower (resp. upper) semicontinuous function u in Ω is a viscosity supersolution (resp. subsolution) of the equation

−Δ

p

u = f

in Ω if the following holds: whenever x

0

∈ Ω and ϕ ∈ C

2

(B

R

(x

0

)) for some R > 0 are such that |∇ϕ(x)| = 0 for x ∈ B

R

(x

0

) \{x

0

},

ϕ(x

0

) = u(x

0

) and ϕ(x) ≤ u(x)

(resp.ϕ(x) ≥ u(x)) for all x ∈ B

R

(x

0

) ∩ Ω, then we have

ρ→0

lim sup

Bρ(x0)\{x0}

( −Δ

p

ϕ(x)) ≥ f(x

0

) (resp. lim

ρ→0

inf

Bρ(x0)\{x0}

( −Δ

p

ϕ(x)) ≤ f(x

0

)). (5.1) A viscosity solution is a continuous function being both a viscosity supersolu- tion and a viscosity subsolution.

Remark 5.1. We consider condition (5.1) to avoid problems with the definition of −Δ

p

φ(x

0

) when ∇ϕ(x

0

) = 0 and p ∈ (1, 2). However, when either p ≥ 2 or

∇ϕ(x

0

) = 0, (5.1) can be replaced by the standard one, i.e.,

− Δ

p

ϕ(x

0

) ≥ f(x

0

) (resp. − Δ

p

ϕ(x

0

) ≤ f(x

0

)). (5.2) Definition 5.2. (The mean value property in the viscosity sense) Suppose that f is continuous function in Ω. We say that a lower (resp. upper) semicontinuous function u in Ω is a viscosity supersolution (resp. subsolution) of the equation

−I

rp

[u] = f + o

r

(1)

in Ω if the following holds: whenever x

0

∈ Ω and ϕ ∈ C

2

(B

R

(x

0

)) for some R > 0 are such that |∇ϕ(x)| = 0 for x ∈ B

R

(x

0

)\{x

0

},

ϕ(x

0

) = u(x

0

) and ϕ(x) ≤ u(x)

(resp.ϕ(x) ≥ u(x)) for all x ∈ B

R

(x

0

) ∩ Ω, then we have

ρ→0

lim sup

Bρ(x0)\{x0}

(−I

rp

[ϕ](x)) ≥ f(x

0

) + o

r

(1) (resp. lim

ρ→0

inf

Bρ(x0)\{x0}

(−I

rp

[ϕ](x)) ≤ f(x

0

) + o

r

(1)).

A viscosity solution is a continuous function being both a viscosity supersolu-

tion and a viscosity subsolution.

(8)

Remark 5.2. The above definition can also be considered with M

pr

instead of I

rp

.

Finally, we define the concept of viscosity solution for the the boundary value problem (2.2).

Definition 5.3. (Viscosity solutions of the boundary value problem) Suppose that f is continuous function in Ω, and that g is a continuous function in ∂Ω.

We say that a lower (resp. upper) semicontinuous function u in Ω is a viscosity supersolution (resp. subsolution) of (2.2) if

(a) u is a viscosity supersolution (resp. subsolution) of −Δ

p

u = f in Ω (as in Definition 5.1);

(b) u(x) ≥ g(x) (resp. u(x) ≤ g(x)) for x ∈ ∂Ω.

A viscosity solution of (2.2) is a continuous function in Ω being both a viscosity supersolution and a viscosity subsolution.

6. The mean value formula for C

2

-functions

In this section we prove the mean value formulas for C

2

-functions as presented in Theorem 2.1. The proof is split into two different cases: p > 2 and p < 2.

The case p = 2 is well known so we leave that out. We restate the results for convenience.

Theorem 6.1. Let p ∈ (2, ∞) and φ ∈ C

2

(B

R

(x)) for some R > 0. Then 1

C

d,p

r

p ∂Br

|φ(x + y) − φ(x)|

p−2

(φ(x + y) − φ(x)) dσ(y) = Δ

p

φ(x) + o

r

(1), where C

d,p

=

12

∂B1

|y

1

|

p

dσ(y).

Proof. Since φ ∈ C

2

near x, we have that φ(x + y) − φ(x) = y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y + o( |y|

2

).

Using Lemma A.1 for ε = 0 and with a = y · ∇φ(x) +

12

y

T

D

2

φ(x)y and b = o( |y|

2

) we get

J

p

(φ(x + y) − φ(x)) = J

p



y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y + o( |y|

2

)



= J

p



y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y



+ o( |y|

p

).

Therefore, A

r

:=

∂Br

J

p

(φ(x + y) − φ(x)) dσ(y)

=

∂Br

J

p



y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y



dσ(y) + o(r

p

).

(6.1)

(9)

Now we use Lemma A.1 for some ε ∈ (0, p − 2) with a = y · ∇φ(x) and b =

12

y

T

D

2

φ(x)y and obtain

J

p

(y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y) = |y · ∇φ(x)|

p−2

y ·

∇φ(x) + (p − 1)|y · ∇φ(x)|

p−2

1

2 y

T

D

2

φ(x)y + O(|y|

p−2−ε

)O(y

2(1+ε)

)

 

o(|y|p)

.

Since the first term is odd and we are integrating over a sphere in (6.1), we get

A

r

= 1 2 (p − 1)

∂Br

|y · ∇φ(x)|

p−2

y

T

D

2

φ(x)y dσ(y) + o(r

p

).

Without loss of generality, assume that, ∇φ(x) = ce

1

for some c ≥ 0. Note that this assumption implies that |∇φ(x)| = c and Δ

φ(x) = c

2

D

11

φ(x). The symmetry of the integral and the term y

T

D

2

φ(x)y imply that

A

r

= 1

2 c

p−2

(p − 1)

∂Br

|y

1

|

p−2

d



i=1

y

2i

D

ii

φ(x)

dσ(y) + o(r

p

)

= 1

2 c

p−2

(p − 1)

d



i=1

D

ii

φ(x)

∂Br

|y

1

|

p−2

y

i2

dσ(y)

+ o(r

p

).

Note that if d ≥ 2, for all i = 1, integration by parts implies C

d,p

r

p

= 1

2

∂Br

|y

1

|

p

dσ(y) = 1 2 (p − 1)

∂Br

|y

1

|

p−2

y

i2

dσ(y).

Thus,

A

r

= C

d,p

r

p

c

p−2

(p − 1)D

11

φ(x) +



d i=2

D

ii

φ(x)

+ o(r

p

)

= C

d,p

r

p

c

p−2

d



i=1

D

ii

φ(x) + (p − 2)D

11

φ(x)

+ o(r

p

)

= C

d,p

r

p



|∇φ(x)|

p−2

Δφ(x) + (p − 2)|∇φ(x)|

p−4

Δ

φ(x) 

+ o(r

p

).

Now, from identity (3.1) we get I

rp

[φ](x) = 1

C

d,p

r

p

A

r

= |∇φ(x)|

p−2

Δφ(x) + (p − 2)|∇φ(x)|

p−4

Δ

φ(x) + o

r

(1)

= Δ

p

φ(x) + o

r

(1),

which concludes the proof. 

We now proceed to the case p < 2, which is slightly more involved.

(10)

Theorem 6.2. Let p ∈ (1, 2) and φ ∈ C

2

(B

R

(x)) for some R > 0. Assume also that |∇φ(x)| = 0. Then

1

C

d,p

r

p ∂Br

|φ(x + y) − φ(x)|

p−2

(φ(x + y) − φ(x)) dσ(y) = Δ

p

φ(x) + o

r

(1), where C

d,p

=

12

∂B1

|y

1

|

p

dσ(y).

Proof. We keep the notation A

r

of (6.1). Without loss of generality, we assume that ∇φ(x) = ce

1

for some c > 0. We split the proof into several parts.

Part 1: First we prove an estimate that will be used several times along the proof. Let α ∈ (0, 1) and ρ ≥ 0 small enough. Then

∂Br

 

 z

|z| · ∇φ(x) + ρ

 z

|z|

T

D

2

φ(x)

 z

|z|

  

−α

dσ(z) ≤ C

1

(6.2) for some C

1

= C

1

(α, d) ≥ 0. To prove (6.2), we first note that its left hand side is equal to

C

2

c

−α

ˆ

∂B1

|ze

1

+ ρc

−1

z

T

D

2

φ(x)z|

−α

dσ(z)

for some constant C

2

= C

2

(d) > 0. Estimate (6.2) follows from applying Lemma A.3 with L(ω, ω) = ρc

−1

ω

T

D

2

φ(x)ω choosing ρ small enough such that (A.1) holds.

Part 2: In this part, we prove A

r

=

∂Br

J

p

(φ(x + y) − φ(x)) dσ(y)

=

∂Br

J

p

(y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y) dσ(y) + o(r

p

).

By Taylor expansion, A

r

=

∂Br

J

p

(y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y + o( |y|

2

)) dσ(y).

Lemma A.2 with a = y · ∇φ(x) +

12

y

T

D

2

φ(x)y and b = o(|y|

2

) implies

 J

p

(y · ∇φ(x) + 1 2 y

T

D

2

φ(x)y + o(|y|

2

)) − J

p

(y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y) 

≤ C



|y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y| + |o(|y|

2

)|

p−2

o(|y|

2

)

≤ C 

y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y 



p−2

o( |y|

2

)

= C 

ˆy· ∇φ(x) + 1

2 |y|ˆy

T

D

2

φ(x)ˆ y 



p−2

o( |y|

p

),

(11)

where ˆ y := y/ |y|. Thus,

 A

r

∂Br

J

p

(y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y) dσ(y) 



≤ o(r

p

)

∂Br

 ˆy· ∇φ(x) + 1

2 r ˆ y

T

D

2

φ(x)ˆ y 



p−2

dσ(y)

= o(r

p

)

where the last identity follows from applying (6.2) with ρ = r (choosing r small enough).

Part 3: This part amounts to proving that

|B

r,γ

| := 

 

∂Br∩{|ˆy·e1|≤γ}

J

p

(y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y) dσ(y)

 

 ≤ C

γ

r

p

, where C

γ

→ 0 as γ → 0. First we note that with our notation we have

J

p

(y · ∇φ(x) + y

T

D

2

φ(x)y) = J

p



cy · e

1

+ 1

2 y

T

D

2

φ(x)y



= (c|y|)

p−1

J

p

 ˆ

y · e

1

+ 1

2 c

−1

|y|ˆy

T

D

2

φ(x)ˆ y

 . (6.3) Lemma A.2 with a = ˆ y · e

1

and b =

12

c

−1

|y|ˆy

T

D

2

φ(x)ˆ y implies

 (c|y|)

p−1

J

p

y · e

1

+ 1

2 c

−1

|y|ˆy

T

D

2

φ(x)ˆ y) − (c|y|)

p−1

J

p

y · e

1

)  

≤ C(c|y|)

p−1



|ˆy · e

1

| + 1

2 c

−1

|y||ˆy

T

D

2

φ(x)ˆ y|

p−2

1

2 c

−1

|y||ˆy

T

D

2

φ(x)ˆ y|

≤ C|y|

p

|ˆy · e

1

|

p−2

. By antisymmetry

∂Br∩{|ˆy·e1|≤γ}

(c|y|)

p−1

J

p

y · e

1

) dσ(y) = 0.

This, (6.2) with α = (p − 2)(1 + δ) > −1 and ρ = 0, and H¨older’s inequality imply

|B

r,γ

| ≤ Cr

p

∂Br

|ˆy · e

1

|

p−2

χ

{|ˆy·e1|≤γ}

dσ(y)

≤ Cr

p



∂Br

|ˆy · e

1

|

α

dσ(y)

1+δ1



∂Br

χ

{|ˆy·e1|≤γ}

dσ(y)

1+δδ

≤ CC

γ

r

p

, where

C

γ

=



∂Br

χ

{|ˆy·e1|≤γ}

dσ(y)

1+δδ

γ→0

−→ 0.

(12)

Part 4: We will now prove that for fixed γ > 0, D

r,γ

:=

∂Br∩{|ˆy·e1|>γ}

J

p



y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y

 dσ(y)

= 1

2

∂Br∩{|ˆy·e1|>γ}

(p − 1)|y · ∇φ(x)|

p−2

y

T

D

2

φ(x)y dσ(y) + o(r

p

).

(6.4) Here it is crucial that the integrals are restricted to the set {|ˆy · e

1

| > γ}.

We observe that outside ξ = 0 the function ξ → J

p

(ξ) is smooth. In particular, for a = 0 and b such that |b| < |a|/2, we have the following estimate

|J

p

(a + b) − J

p

(a) − J

p

(a)b| ≤ C(|a + b|

p−2−δ

+ |a|

p−2−δ

)|b|

1+δ

for any δ ∈ (0, p−1) ⊂ (0, 1). For any y such that |ˆy·e

1

| > γ, the above estimate with a = ˆ y · e

1

and b =

12

c

−1

|y|ˆy

T

D

2

φ(x)ˆ y (since a = 0 and b < γ/2 < |a|/2 by choosing r = |y| small enough), together with ( 6.3) imply

J

p

(y · ∇φ(x) + 1

2 y

T

D

2

φ(x)y) = (c |y|)

p−1

|ˆy

· e

1

|

p−2

y ˆ · e

1

+ (p − 1)|y · ∇φ(x)|

p−2

1

2 y

T

D

2

φ(x)y + R(y), where R(y) is bounded by

C

3

|y|

p+δ

  ˆy · e

1

+ 1

2 c

−1

|y|ˆy

T

D

2

φ(x)ˆ y 



p−2−δ

+ |ˆy · e

1

|

p−2−δ

× 

 1

2 y ˆ

T

D

2

φ(x)ˆ y 



1+δ

≤ C

4

|y|

p+δ

  ˆy · e

1

+ 1

2 c

−1

|y|ˆy

T

D

2

φ(x)ˆ y 



p−2−δ

+ |ˆy · e

1

|

p−2−δ

. For some constants C

3

, C

4

≥ 0 and r small enough (depending on γ). Moreover, by antisymmetry,

∂Br∩{|ˆy·e1|>γ}

(c |y|)

p−1

|ˆy

1

|

p−2

y ˆ

1

dσ(y) = 0.

We apply (6.2) with α = −p + 2 + δ ∈ (0, 1) two times, first with ρ = r and later with ρ = 0 to get

∂Br∩{|ˆy·e1|>γ}

|R(y)| dσ(y) ≤

∂Br

|R(y)| dσ(y) ≤ O(r

p+δ

) = o(r

p

)

where the bound is uniform for fixed γ. This implies (6.4).

Part 5: From parts 2 and 3 we have lim sup

r→0

|A

r

− D

r,γ

|

r

p

≤ lim sup

r→0

|B

r,γ

|

r

p

≤ C

γ γ→0

−→ 0,

(13)

Moreover, by part 4

r→0

lim D

r,γ

r

p

= 1 2 lim

r→0

r

−p

∂Br∩{|ˆy·e1|>γ}

(p − 1)|y · ∇φ(x)|

p−2

y

T

D

2

φ(x)y dσ(y)

= 1 2 (p − 1)

∂B1∩{|z·e1|>γ}

|z · ∇φ(x)|

p−2

z

T

D

2

φ(x)z dσ(z).

Since the last term is independent of r and converges to 1

2 (p − 1)

∂B1

|z · ∇φ(x)|

p−2

z

T

D

2

φ(x)z dσ(z) = C

d,p

Δ

p

φ(x),

as γ → 0, where the last equality follows from the proof of Theorem 6.1, the

result follows. 

As an immediate corollary, we obtain that also the mean over balls have the same asymptotic limit.

Corollary 6.3. Let p ∈ (1, ∞) and φ ∈ C

2

(B

R

(x)). If p < 2, assume also that

|∇φ(x)| = 0. Then 1

D

d,p

r

p Br

|φ(x + y) − φ(x)|

p−2

(φ(x + y) − φ(x)) dy = Δ

p

φ(x) + o

r

(1), where D

d,p

=

dCp+dd,p

.

7. Viscosity solutions

Now we prove that satisfying the asymptotic mean value property in the vis- cosity sense is equivalent to being a viscosity solution of the corresponding PDE.

Proof of Theorem 2.2. We only prove that the notion of supersolutions are equivalent. The case of a subsolution can be treated similarly. Suppose first that u is a viscosity supersolution of −Δ

p

u = f in Ω. Take x

0

∈ Ω and ϕ ∈ C

2

(B

R

(x

0

)) for some R > 0 such that |∇ϕ(x)| = 0 when x = x

0

,

ϕ(x

0

) = u(x

0

) and ϕ(x) ≤ u(x) for all x ∈ Ω.

Since u is viscosity supersolution of −Δ

p

u = f we have that for given ε > 0 there is x ∈ B

ρ

(x

0

)\{x

0

} with ρ = ρ(ε) such that

−Δ

p

ϕ(x) ≥ f(x

0

) − ε.

By Theorem 2.1

Δ

p

ϕ(x) = I

rp

[ϕ](x) + o

r

(1).

Therefore,

−I

rp

[ϕ](x) ≥ f(x

0

) + o

r

(1) − ε.

Since ε was arbitrary, this proves the mean value supersolution property. Now suppose instead that u is a viscosity supersolution of

−I

rp

[u] = f + o

r

(1)

(14)

in Ω. Take again x

0

∈ Ω and ϕ ∈ C

2

(B

R

(x

0

)) for some R > 0 such that

|∇ϕ(x)| = 0 when x = x

0

,

ϕ(x

0

) = u(x

0

) and ϕ(x) ≤ u(x) for all x ∈ Ω.

By the definition of a supersolution, for given ε > 0 there is x ∈ B

ρ

(x

0

)\{x

0

} with ρ = ρ(ε) such that

−I

rp

[ϕ](x) ≥ f(x

0

) + o

r

(1) − ε.

Again by Theorem 2.1

Δ

p

ϕ(x) = I

rp

[ϕ](x) + o

r

(1), which implies

−Δ

p

ϕ(x) ≥ f(x

0

) + o

r

(1) − ε.

Passing r → 0 implies −Δ

p

ϕ(x) ≥ f(x

0

) − ε. Again, since ε was arbitrary, the

proof is complete. 

8. The pointwise property in the plane

Now we are ready to prove that the mean value property is satisfied in a pointwise sense in the aforementioned range of p.

Proof of Theorem 2.3. Assume that u satisfies

− I

rp

[u] = o

r

(1) (8.1)

in the pointwise sense in Ω. Then it is obviously also a viscosity solution. By Theorem 2.2 it is also a viscosity solution of −Δ

p

u = 0 which proves the first implication.

Assume now instead that u is a viscosity solution of −Δ

p

u = 0 and let x

0

∈ Ω. If |∇u(x

0

)| = 0, then u is real analytic near x

0

and the mean value formula holds trivially at x

0

by Theorem 2.1. If |∇u(x

0

)| = 0 we need different arguments depending on p.

Case p ≥ 2: The case p = 2 is well-known and we do not comment on it.

If p > 2, Theorem 1 in [11] implies that u ∈ C

1,α

for some 1 > α > 1/(p − 1).

Then

|u(x

0

+ y) − u(x

0

)| ≤ C|y|

1+α

, which implies that

|I

rp

[u](x

0

) | ≤ Cr

−p

r

(p−1)(1+α)

= o

r

(1), (8.2) which ends the proof in this case.

Case p

0

< p < 2: First we use that on page 146 in [19] it is proved that for some integer n ≥ 1 we have that

|D

2

u| = O  r

ηn−1



in B

r

(x

0

) where

1 η

n

:= 1

2

⎝−p +

 4

 1 + 1

n

2

(p − 1) + (p − 2)

2

⎠ .

(15)

In particular, when n ≥ 3 we have that 1/η

n

< p − 1 which implies |D

2

u | = o 

r

p−11 −1



in B

r

(x

0

). By Taylor expansion we thus get,

|u(x

0

+ y) − u(x

0

)|

p−1

≤ ( D

2

u

L(Br(x0))

r

2

)

p−1

= o(r

p−11 +1

)

p−1

= o(r

p

) which in turn implies −I

rp

[u](x

0

) = o

r

(1) as in (8.2).

We still need to check the cases n = 1 and n = 2. We do it in several steps.

Step 1. For this we need a refined expansion around a critical point x

0

(and assume u(x

0

) = 0 for simplicity) taken from pages 147–148 in [19]. It reads

u(x) = A(x) + O (r

γ

) for all x ∈ B

r

(x

0

), (8.3) with

γ = 1 + λ

(n)n+2

 λ

(n)n+1



2

and λ

(n)k

= 1 2

 −np + 

4k

2

(p − 1) + n

2

(p − 2)

2

 (8.4) and where the function A(x) is defined by (see pages 3864–3865 in [3])

A = A ◦ A

−1

. Here A and A are defined in complex variables by

A(re

) = r

β

e

−inθ



e

i(n+1)θ

+ εe

−i(n+1)θ



|A(re

)| = r

β

m(θ), m(θ) = 

1 + ε

2

+ 2ε cos(2(n + 1)θ), and

A(re

) = Cr

α

cos((n + 1)θ).

In the above, C, α, β and ε are constants depending on n, but their values will not be important in what follow, except the fact that |ε| < (2n + 1)

−1

, see equation (2.4) on page 3861 in [3]. Note that by (8.3), we necessarily have

A(x

0

) = |∇A(x

0

)| = 0.

Step 2. We prove now that A satisfies the mean value property, i.e.

I

rp

[A](x

0

) = 0.

We define,

B 

R

= A

−1

(B

R

) =



re

: r

β

< R m(θ)

 ,

where the equality follows from the fact that |A(re

) | = r

β

m(θ). We also compute the jacobian of A and find

J (re

) = |DA|(re

) = βr

2(β−1)

(1 − (2n + 1)ε

2

− 2nε cos(2(n + 1)θ)) > 0, where we used that |ε| < (2n + 1)

−1

. By a change of variables

ˆ

BR

|A(re

) |

p−2

A(re

)dA = ˆ

B˜R

|A(re

) |

p−2

A(re

)J (re

)rdrdθ

= C

p−1

β ˆ

0

ˆ

r(θ)

0

r

α(p−1)+2β−1

| cos((n + 1)θ)|

p−2

cos((n + 1)θ)j(θ)drdθ

(16)

where r(θ) =

 R

m(θ)

β1

, j(θ) = 1 − (2n + 1)ε

2

− 2nε cos(2(n + 1)θ).

Hence, we see that we are left with an integral of the form ˆ

0

f (cos(2(n + 1)θ)) | cos((n + 1)θ)|

p−2

cos((n + 1)θ)dθ.

By change of variables we can reduce this to computing ˆ

0

f (cos(2θ))| cos(θ)|

p−2

cos(θ)dθ = 0, by symmetry. Therefore,

ˆ

BR

|A(re

)|

p−2

A(re

)dA = 0 and A satisfies the mean value property.

Step 3. Now we go back to u. Using (8.3), we have together with Lemma A.2

|J

p

(u(x

0

+ y) − u(x

0

)) − J

p

( A(x

0

+ y) − A(x

0

)) | = O 

r

(p−2)γ

 O 

r

γ



= O 

r

(p−1)γ

 ,

with γ given in (8.4). By Step 2, A satisfies the mean value property at x

0

and thus

|I

rp

[u](x

0

) | ≤ Cr

−p+(p−1)γ

.

The proof will be finished if we verify that γ > p/(p − 1), that is, λ

(n)n+2

 λ

(n)(n+1)



2

> 1

p − 1 . (8.5)

First we verify (8.5) when n = 1. In this case λ

(1)k

= 1

2

 − p + 

4k

2

(p − 1) + (p − 2)

2

 so that (8.5) becomes

2

 −p + 

36(p − 1) + (p − 2)

2



 − p + 

16(p − 1) + (p − 2)

2



2

> 1 p − 1 . This inequality is exactly true when p ∈ (p

0

, 2).

If n = 2 then λ

(2)k

= 1

2

 −2p + 

4k

2

(p − 1) + 4(p − 2)

2

 and (8.5) becomes

2

 −2p + 

64(p − 1) + 4(p − 2)

2



 − 2p + 

36(p − 1) + 4(p − 2)

2



2

> 1

p − 1 .

(17)

This inequality turns out to be true for p > 1.06 and therefore it is true for

p > p

0

. 

9. Study of the dynamic programming principle

Recall the notation M

pr

[φ](x) = 1

D

d,p

r

p Br

|φ(x + y) − φ(x)|

p−2

(φ(x + y) − φ(x)) dy.

Given an open domain Ω and r > 0, we will in this section denote by

∂Ω

r

= {x ∈ Ω

c

: dist(x, Ω) ≤ r}

and Ω

r

= Ω ∪ ∂Ω

r

.

We want to study solutions of the (extended) boundary value problem

 −M

pr

[U

r

](x) = f (x) x ∈ Ω

U

r

(x) = G(x) x ∈ ∂Ω

r

:= {x ∈ Ω

c

: dist(x, Ω) ≤ r}, (9.1) where f ∈ C(Ω) and G ∈ C(∂Ω

r

) (a continuous extension of g ∈ C(∂Ω)).

These will be our running assumptions in this section.

9.1. Existence and uniqueness: the proof of Theorem 2.5 (i)

For convenience, we will write M

p

instead of M

pr

when the subindex r plays no role.

We first prove a comparison principle which immediately implies unique- ness and then we prove the existence.

Proposition 9.1. Let p ∈ (1, ∞) and U, V ∈ L

r

) be such that

 −M

p

[V ](x) ≥ f(x) x ∈ Ω,

V (x) ≥ G(x) x ∈ ∂Ω

r

, and

 −M

p

[U ](x) ≤ f(x) x ∈ Ω, U (x) ≤ G(x) x ∈ ∂Ω

r

. Then U ≤ V in Ω

r

.

Proof. Assume by contradiction that U (x) > V (x) for some x ∈ Ω. It has to be in the interior of Ω since by definition U ≤ G ≤ V in ∂Ω

r

.

Let M > 0 and x

0

∈ Ω be such that M = U (x

0

) − V (x

0

) = sup

x∈Ω

{U(x) − V (x)}.

Define ˜ U = U − M. Then ˜ U (x

0

) = V (x

0

), ˜ U ≤ V in Ω, ˜ U < V in ∂Ω

r

, and

 −M

p

[ ˜ U ](x) ≤ f(x) x ∈ Ω, U (x) ˜ ≤ G(x) − M x ∈ ∂Ω

r

. By the monotonicity of J

p

J

p

(V (x

0

+ y) − V (x

0

)) − J

p

( ˜ U (x

0

+ y) − ˜ U (x

0

))

≥ J

p

( ˜ U (x

0

+ y) − V (x

0

)) − J

p

( ˜ U (x

0

+ y) − ˜ U (x

0

))

= J

p

( ˜ U (x

0

+ y) − ˜ U (x

0

)) − J

p

( ˜ U (x

0

+ y) − ˜ U (x

0

)) = 0.

(18)

From the equations satisfied by U and V we have 0 ≥ M

p

[V ](x

0

) − M

p

[U ](x

0

)

= 1

D

d,p

r

p B

r

J

p

(V (x

0

+ y) − V (x

0

)) − J

p

( ˜ U (x

0

+ y) − ˜ U (x

0

)) dy.

Hence, the average of the non-negative integrand is non-positive. This means that

J

p

(V (x

0

+ y) − V (x

0

)) = J

p

( ˜ U (x

0

+ y) − ˜ U (x

0

)).

By the strict monotonicity of J

p

this implies

V (x

0

+ y) − V (x

0

) = ˜ U (x

0

+ y) − ˜ U (x

0

),

that is, V (x

0

+ y) = ˜ U (x

0

+ y) for all y ∈ B

r

. This means that ˜ U (x) = V (x) for all x ∈ B

r

(x

0

). Repeating this process in the contact points of ˜ U and V and iterating, we will eventually arrive at the conclusion that ˜ U (x) = V (x) for some x ∈ ∂Ω

r

. This contradicts the fact ˜ U < V in ∂Ω

r

.  In order to prove the existence and to study the limit as r → 0, we will first derive uniform bounds (in r) for the solution of (9.1).

Proposition 9.2. (L

-bound) Let p ∈ (1, ∞), let R > 0 and U

r

be the solution of (9.1) corresponding to some r ≤ R. Then

U

r

≤ A

with A > 0 depending on p, Ω, f, g and R (but not on r).

Proof. Consider the function h(x) = |x|

p−1p

. Then h ∈ C

(R

d

\B

1

(0)) and Δ

p

h(x) = d

 p

p − 1

p−1

for all x = 0.

Let C, D ∈ R and z ∈ R

d

to be chosen later and define ψ(x) = C − D|x − z|

p−1p

1

d

 p − 1 p

p−1

. Then

Δ

p

ψ(x) = −D for all x = z and ψ ∈ C

(R

d

\B

1

(z)).

Now take z such that

B

1

(z) ∩ Ω

R

= ∅.

Then ψ ∈ C

R

). By Corollary 6.3, for all x ∈ Ω we have

−M

p

[ψ](x) = −Δ

p

ψ(x) + o

r

(1) = D + o

r

(1) ≥ D − ˜ D

where ˜ D > 0 depends only on R but not on r. Then choose D = ˜ D + f

to get

−M

p

[ψ](x) ≥ D − ˜ D = f

for all x ∈ Ω.

(19)

Finally, we choose C such that ψ(x) ≥ G

for all x ∈ ∂Ω

R

. Thus

 −M

p

[ψ](x) ≥ f

x ∈ Ω ψ(x) ≥ G

x ∈ ∂Ω

r

, for all r ≤ R. Then, by comparison (Proposition 9.1)

U (x) ≤ ψ(x) ≤ ψ

.

Note that this bound depends on R but not on r. A similar argument with

−ψ as barrier shows that U(x) ≥ − ψ

and thus, U

≤ ψ

,

which concludes the proof. 

The aim is now to prove the existence of a solution of (9.1). Before doing that, we need some auxiliary results. Define

L[ψ, φ](x) := 1

D

d,p

r

p Br

J

p

(φ(x + y) − ψ(x)) dy.

Lemma 9.3. Let r > 0 and φ ∈ L

r

).

(a) Then there exists a unique ψ ∈ L

(Ω) such that

−L[ψ, φ](x) = f(x) for all x ∈ Ω.

(b) Let ψ

1

and ψ

2

be such that

−L[ψ

1

, φ](x) ≤ f(x) and − L[ψ

2

, φ](x) ≥ f(x) for all x ∈ Ω, then ψ

1

≤ ψ

2

in Ω.

Proof. We start by proving the comparison principle. This will imply unique- ness. Assume that ψ

1

(x) > ψ

2

(x) for some x ∈ Ω. Then

0 = (−f(x) + f(x))r

p

D

d,p

Br

J

p

(φ(x + y) − ψ

2

(x)) − J

p

(φ(x + y) − ψ

1

(x)) dy

>

Br

J

p

(φ(x + y) − ψ

2

(x)) − J

p

(φ(x + y) − ψ

2

(x)) dy = 0 which is a contradiction. To prove existence we start by defining

ψ

I

(x) = sup

Ωr

φ + J

p−1

(D

d,p

r

p

f (x)) . Since

sup

Ωr

φ − φ(x + y) ≥ 0, we have

−L[ψ

I

, φ](x) = 1

D

d,p

r

p Br

J

p



φ(x + y) − sup

Ωr

φ − J

p−1

(D

d,p

r

p

f (x))

dy

1

D

d,p

r

p Br

J

p



J

p−1

(D

d,p

r

p

f (x)) 

dy = f (x).

References

Related documents

1 Introduction to the magnetic Schrödinger operator 1 1.1 A non-relativistic quantum theory. 5 1.3 Magnetic Schrödinger operators. 12 2.2 Spectral inequalities. 19.. 3 Overview of

Industrial Emissions Directive, supplemented by horizontal legislation (e.g., Framework Directives on Waste and Water, Emissions Trading System, etc) and guidance on operating

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

Från den teoretiska modellen vet vi att när det finns två budgivare på marknaden, och marknadsandelen för månadens vara ökar, så leder detta till lägre

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

Parallellmarknader innebär dock inte en drivkraft för en grön omställning Ökad andel direktförsäljning räddar många lokala producenter och kan tyckas utgöra en drivkraft

I dag uppgår denna del av befolkningen till knappt 4 200 personer och år 2030 beräknas det finnas drygt 4 800 personer i Gällivare kommun som är 65 år eller äldre i