• Non ci sono risultati.

Representation formulae of solutions of second order elliptic inequalities

N/A
N/A
Protected

Academic year: 2021

Condividi "Representation formulae of solutions of second order elliptic inequalities"

Copied!
29
0
0

Testo completo

(1)

Representation Formulae of solutions of second order

elliptic inequalities

Lorenzo D’Ambrosio

(1)

, Enzo Mitidieri

(2)

August 21, 2018

Abstract

We prove various representation formulas of solutions of second order differential inequalities on RN. As an outcome we obtain positivity results of the solutions of −Lu ≥ 0, where L is a general second order elliptic operator in divergence form. Results of this type are very useful for proving related Liouville theorems of second order semilinear inequalities.

Contents

1 Introduction 1

2 Main assumptions and examples 2

3 Representation Theorems 6

3.1 Lemmata . . . 12

3.2 A priori estimates and the ring condition . . . 19

3.3 Ring condition (3.2) related to ∇u . . . 20

3.4 Proof of Theorem 6 . . . 22

3.5 Proof of Theorems 14 and 19 . . . 23

3.6 Additional remarks on the ring conditions . . . 23

3.6.1 Ring condition and semilinear problems. . . 23

3.6.2 Other assumptions . . . 25

4 Applications in the Carnot groups setting 26

1

Introduction

The search of representation formulae of solutions of second order elliptic equations or inequalities has been an active research area over the past several years. See for instance the book [5] and the references therein for recent results and applications. Our goal in this paper is to find representation formulae of solution of problems of the type,

−Lu ≥ 0 on RN,

K eywords. Representation formulae, Fundamental solution, sub-Laplacians in Carnot group. 2 000 Mathematics Subject Classification. Primary 35H10, 35C15, 26D10.

(2)

where L is a general second order elliptic operator in divergence form (see section 2 below for the precise assumptions and fundamental properties) possessing a fundamental solution. In general if u ∈ L1loc(RN) is a distributional solution of −Lu ≥ 0, then −Lu := ν ≥ 0 is a nonnegative Radon measure, see [20]. One of the motivations for searching representation formulae, besides their intrinsic interest, relies on the fact that under suitable assumptions, nonnegative solutions of the problem,

− Lu = f (u) in RN, (1.1) where f : R → R+ is a given continuous function, are indeed solutions of the integral

equation

u(x) = l + Z

RN

Γ(x, y)f (u(y))dy,

where l = infRNu(x) and Γ is a fundamental solution of L. See [8, 15] for some results in

this direction.

When looking for Liouville theorems for (1.1), a typical simple statement can be formu-lated as follows. Let L be a differential operator and let f : R → R be a given continuous function. Consider u ∈ Γ(RN), where Γ(RN) is a certain function class that depends on the

differential operator L and on the properties of f.

Liouville theorem. If u is a nonnegative solution of (1.1) then u ≡ constant in RN. General Liouville theorems can be proved for several classes of semilinear/quasilinear equations and inequalities under some kind homogeneity assumptions on L. See [7, 10, 21] and the references therein for further information.

However, for operators L with variable coefficients considered in this paper (see section 2 below), the approach based on the nonlinear capacity method developed in [21] needs to be modified. See3.6.1. In the forthcoming paper [13] we shall prove that the representation formulae proved here are the suitable substitute of the nonlinear capacity method for proving Liouville theorems of (1.1).

Another motivation for searching representation formula of solution of (1.1) are the so called positivity results. These kind of results are related to the following fundamental question:

Positivity problem. When f nonnegative implies that the possible solutions of (1.1) are positive?

Some contributions to this problem have been obtained in [15, 22] for some semilinear inequalities and in [10, 14] in the quasilinear context.

This paper is organised as follows. In Section 2 we state our main assumptions on the operator L and we present some important examples. Section3is devoted to the formulation of the main results, while in subsection 3.1. we prove some necessary lemmata that we will use throughout the paper. In subsection 3.6.1 we comment and prove some results related to the ring condition, see (3.2) below. Finally section 4 is devoted to some applications in the Carnot group setting.

2

Main assumptions and examples

Let A be an N × N matrix with entries aij ∈C1(RN). We shall deal with the operator L of

the form L(u) := div(A∇u) = N X ij=1 ∂ ∂xi  aij(x) ∂u ∂xj  . (2.1)

(3)

H1. The matrix A is semi-positive definite, that is:

(A(x)ξ · ξ) ≥ 0 for any x, ξ ∈ RN. H2. For any x ∈ RN there exists a fundamental solution Γ

x = Γ(x, ·) ∈ L1loc(RN) of −L.

This means that Γ : RN × RN \ {(x, y) | x = y ∈ RN} → R is smooth and

Z

RN

−L(φ)(y)Γ(x, y) dy = φ(x)

for any test function φ ∈ C∞0 (RN). Denote with A0 the transpose matrix of A and with L∗ the adjoint operator of L. We have L∗(u) = div(A0∇u) and −L∗

x) = δx.

H3. limy→xΓ(x, y) = +∞ for any x ∈ RN

H4. lim|y|→∞Γ(x, y) = 0 for any x ∈ RN

H5. Γ(x, y) > 0 for any x 6= y.

H6. For any y ∈ RN, Γ(·, y) is a fundamental solution of −L∗, that is −L(Γ(·, y)) = δy.

H7. For any x ∈ RN and r > 0, we set

Ωr(x) :=y ∈ RN| Γ(x, y) >

1

r ∪ {x}. On the family Ωr(x), we shall assume that

lim inf

→0

|Ω(x)|

 = 0 for any x ∈ R

N. (2.2)

From the above assumptions on Γ, it is easy to see that • For any x ∈ R and r > 0, Ωr(x) is a bounded, open set;

• For any x ∈ R and r > 0, there exists  > 0 such that B(x, ) ⊂ Ωr(x);

• For any x ∈ R, Ωr(x) is a family of sets nondecreasing by inclusion such that Ωr(x) %

RN (as r → +∞).

Notice that if A is symmetric then assumption H6. implies Γ(x, y) = Γ(y, x) and L = L∗. The above hypotheses are fulfilled in several concrete cases.

Example 1. Let L = Pl

i=1X 2

i where Xi are smooth vector fields in RN satisfying

Hör-mander’s condition of hypoellipticity:

rank Lie [X1, . . . , Xl] = N at every x ∈ RN.

Assuming that: Xi∗ = −Xi, where Xi∗ is the formal adjoint of Xi, there exists a symmetric

semi-positive definite matrix A such that L can be wtitten as Lu = div(A∇u).

In this setting hypotheses [H1.] . . . [H7.] and the properties on the family Ω hold. See [9, 23] and the reference therein.

Further special cases is when it is possible to endow RN with a Lie groups law and the

(4)

Example 2 (Carnot Groups). We begin quoting some preliminary results concerning Carnot groups (for more informations and proofs we refer the interested reader to [5, 16, 17]; see also the survey [19]). A Carnot group is a connected, simply connected, nilpotent Lie group G of dimension N ≥ 2 with graded Lie algebra G = V1⊕ · · · ⊕ Vr such that [V1, Vi] = Vi+1

for i = 1 . . . r − 1 and [V1, Vr] = 0. A Carnot group G of dimension N can be identified, up

to an isomorphism, with the structure of homogeneous Carnot group (RN, ◦, δλ) defined as

follows. We identify G with RN endowed with a Lie group law ◦. We consider RN splitted

in r subspaces RN = Rn1×Rn2×· · ·×Rnr with n

1+ n2+ · · · + nr = N and ξ = (ξ(1), . . . , ξ(r))

with ξ(i) ∈ Rni. We shall assume that there exists a family of Lie group automorphisms,

called dilation, δλ with λ > 0 of the form δλ(ξ) = (λξ(1), λ2ξ(2), . . . , λrξ(r)). The Lie algebra

of left-invariant vector fields on (RN, ◦) is G. For i = 1, . . . , n

1 = l let Xi be the unique

vector field in G that coincides with ∂/∂ξ(1)i at the origin. We require that the Lie algebra generated by X1, . . . , Xn1 is the whole G.

With the above hypotheses, we call G=(RN, ◦, δλ) a homogeneous Carnot group. The

canonical sub-Laplacian on G is the second order differential operator L = Pl

i=1Xi2. Let

Y1, . . . , Yl be a basis of span{X1, . . . , Xl}, the second order differential operator ∆G =

Pl

i=1Y 2

i is called a sub-Laplacian on G. We denote by Q =

Pr

i=1ini the homogeneous

dimension of G. In what follows we shall assume that Q ≥ 3.

We shall list some properties and know results about Homogeneous Carnot groups. The Lebesgue measure is the bi-invariant Haar measure. For any measurable set E ⊂ RN, we have |δλ(E)| = λQ|E|. Since Y1, . . . , Yl generate the whole G, any sub-Laplacian ∆G

satisfies the Hörmander’s hypoellipticity condition. Moreover, the vector fields Y1, . . . , Yl

are homogeneous of degree 1 with respect to δλ.

In what follows we fix such vector fields Y1, . . . , Yl and when we shall refer to this setting

we shall use the symbol symbol ∇L to denote the vector field (Y1, . . . , Yl). As particular

case of Example 1 there exists a symmetric semi-positive definite matrix A such that ∆G

can be written as ∆G= div(A∇u).

A nonnegative continuous function N∗ : RN → R+ is called a homogeneous norm on G,

if N∗(ξ) = 0 if and only if ξ = 0 and it is homogeneous of degree 1 with respect to δλ (i.e.

N∗(δλ(ξ)) = λN∗(ξ)). We say that a homogeneous norm is symmetric if N∗(ξ−1) = N∗(ξ).

A homogeneous norm N∗ defines on G a pseudo-distance defined as d(ξ, η) := N∗(ξ−1◦ η). For such function d, there holds only a pseudo-triangular inequality:

d(ξ, η) ≤ Cd(ξ, ζ) + Cd(ζ, η) (ξ, ζ, η ∈ G) (2.3) with C ≥ 1. Hence in general d, is not a distance. For a given homogeneous norm N∗, the symbol BN∗(η, R) denotes the ball BN∗(η, R) := {ξ ∈ RN|N∗(η−1 ◦ ξ) < R}. Then

|BN∗(x, R)| = |BN∗(0, R)| = cN∗RQ.

If N∗ and ˜N are two homogeneous norms, then they are equivalent, that is, there exists a constant C > 0 such that C−1N∗(ξ) ≤ ˜N (ξ) ≤ CN∗(ξ).

Let N∗ be a homogeneous norm, then there exists a constant C > 0 such that C−1|ξ| ≤ N∗(ξ) ≤ C |ξ|1/r, for N∗(ξ) ≤ 1 and |·| stands for the Euclidean norm. An example of symmetric homogeneous norm is the following

NS(ξ) := r X i=1 |ξi| 2r!/i !1/2r! . (2.4)

Notice that if N∗ is a homogeneous norm differentiable a.e., then |∇LN∗| is homogeneous

(5)

In [16] it is proved that for any sub-Laplacian ∆G there exists a homogeneous symmetric

norm N2 on G , often called gauge, such that Γη(ξ) := (N2(η−1 ◦ ξ))2−Q is a fundamental

solution of −∆G at η (see also [5]). A such homogeneous norm (by hypoellipticity of −∆G)

is smooth off of the origin. We shall denote by ψη the quantity ψη := |∇LN2(η−1◦ ·)|.

Since YiN2 is homogeneous of degree 0 with respect to δλ, we have ||ψη||∞ = ||ψ0||∞ by left

invariance of Yi.

With the symbols used at point H7. above we have that Ωr(x) = {y ∈ RN|N2(x−1◦ y) <

rQ−21 } and hence |Ω r(x)| = BN2(r 1 Q−2) = Cr Q Q−2.

In what follows a relevant role will be played by the quantity (∇Γη·A∇Γη)

Γη , which in the

Carnot group setting can be rewritten as (Q − 2)2 ψ2η(ξ)

N2Q(η−1◦ξ) = (Q − 2)

2 ψ20

N2Q(η

−1◦ ξ).

Finally, we remind that in the Carnot group setting a Liouville theorem holds. Namely if L is a sublaplacian then any L-harmonic nonnegative function is constant. See [4, 5]. Example 3. Simple examples of Carnot groups are the usual Euclidean spaces RQ.

More-over, if Q ≤ 3 then G is the ordinary Euclidean space RQ.

The simplest nontrivial example of a Carnot group is the Heisenberg group H1 = R3.

For an integer n ≥ 1, the Heisenberg group Hn is defined as follows. Let ξ = (ξ(1), ξ(2)) =

(x1, . . . , xn, y1, . . . , yn, t) = (x, y, t) ∈ R2n× R. The Heisenberg group Hn is the set R2n+1

endowed with the group law ˆ ξ ◦ ˜ξ := (ˆx + ˜x, ˆy + ˜y, ˆt + ˜t + 2 n X i=1 (˜xiyˆi− ˆxiy˜i)).

For i = 1, . . . , n, consider the vector fields Xi := ∂ ∂xi + 2yi ∂ ∂t, Yi := ∂ ∂yi − 2xi ∂ ∂t,

and the associated Heisenberg gradient as follows ∇H := (X1, . . . , Xn, Y1, . . . , Yn). The

Kohn-Laplacian ∆H is then the operator defined by ∆H :=

Pn

i=1Xi2 + Yi2. The family

of dilation is given by δλ(ξ) := (λx, λy, λ2t). In Hn is defined the homogeneous norm

|ξ|H :=   n X i=1 x2i + yi2 !2 + t2   1/4 .

The homogeneous dimension is Q = 2n+2 and the fundamental solution of the sub-Laplacian −∆H at point η is given by Γη(ξ) = N22−Q(η

−1◦ ξ) = c |η−1◦ ξ|−2n H .

Example 4. As particular case of the Example 1 is when the vector fields satisfying some further conditions. Namely, let {X1, . . . , Xl} be a fixed set of linearly independent smooth

vector fields on Euclidean space RN, satisfying the following properties:

• There exits a family of non-isotropic diagonal maps δλ of the form

δλ(x) = (λσ1x1, . . . , λσNxN) with 1 = σ1 ≤ · · · ≤ σN.

We assume that such that X1, . . . , Xl are δλ-homogeneous of degree 1 with respect to the

family of non-isotropic dilations {δλ}λ≥0.

• X1, . . . , Xl satisfy Hörmander’s rank condition at 0, i.e.,

(6)

• P

jσj > 2.

In this case by a lifting technique in [3] the authors proved that for the operator L = Pl

i=1X 2

i all the hypotheses [H1.] . . . [H7.] are fulfilled. We emphasise that even the

hypothesis that Γ vanishes at infinity is satisfied.

More precisely they show the existence of an homogeneous Carnot group (RN +p, ◦, ˜δ) and

vector fields Z1, . . . , Zl, such that Pli=1Zi2 is a sub-Laplacian on RN +p and the projection

of Zi on RN is Xi (i = 1, . . . , l).

This last remark implies that for the operator L = Pl

i=1Xi2 Liouville theorem holds,

that is any L-harmonic nonnegative function is constant1. Example 5. The Grushin operator

x2+ x2∂y2

acting on R2, satisfies our assumptions [H1.] . . . [H7.]. Indeed this operator is of the type

described in the Example 4. See also the seminal paper [18] and the references in [3]. The Grushin type operator

x2+ x4∂y2

is of the same type, see [3]. Our results apply to this operator too.

For further generalisations of these operators (dealing with several variables, as well as several weights), the explicit fundamental solutions have been computed in [2].

As a final example taken from [3], we remark that our results apply to the following Engel-type operator where the vector fields acting on R3 are given by

X1 = ∂x1; X2 = x1∂x2 + x

2 1∂x3.

3

Representation Theorems

Let L be an operator of type (2.1) satisfying the assumptions H1. . . . H7. of Section 2. In this section we study the solutions of the inequality

− L(u) ≥ 0 on RN. (3.1)

If u ∈ L1

loc(RN) is a distributional solution of (3.1) it follows that −L(u) is in general a

distribution. In our case it is a nonnegative distribution, therefore it is, actually, a nonneg-ative Radon measure. See [20]. Consequently, in general we may assume that −L(u) =: ν is a nonnegative Radon measure.

The following representation theorem of the solutions of (3.1) has its roots in [15]. Theorem 6 (Riesz representation). Let u ∈C2(RN) be such that −L(u) = ν ≥ 0.

A. Let x ∈ RN and l x ∈ R be such that lim inf R→+∞ Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx (u − lx)dy = 0. (3.2) Then u(x) = lx+ Z RN Γ(x, y)ν(y)dy. (3.3) Assume that for any x ∈ RN (3.2) holds and set l(x) := lx. Then

1Indeed if u us a nonnegative function such that Lu = 0, then by lifting ˜Lu =Pl i=1Z

2

iu = 0, and since

˜

(7)

(a) inf u = inf l (finite or not), sup u ≥ sup l (finite or not), and the following alter-native holds,

either u(x) > l(x), ∀x ∈ RN, or u ≡ l and ν ≡ 0. (3.4) (b) If w(·) :=

Z

RN

Γ(·, y)ν(y)dy ∈ L1loc(RN) or equivalently l ∈ L1

loc(RN), then the

func-tion l is a distribufunc-tional solufunc-tion of

−L(h) = 0, on RN.

Moreover if ν 6≡ 0, then there exists a constant C > 0 such that lim inf R→+∞R Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx (u − lx)dy ≥ C > 0, (3.5)

where C depends on ν and it is independent on u nor to the particular structure of the operator L, namely if ν 6≡ 0 on the set Ωs(o) for some s > 0 and o ∈ RN then

C = M Z

Ωs(o)

ν(y)dy, where M is a universal constant.

(c) if u ≥ 0, then w ∈ L1

loc(RN).

(d) Finally, if lx does not depends on x, lx = l ∈ R, then for any x ∈ RN

u(x) = l + Z

RN

Γ(x, y)ν(y)dy. (3.6) B. If u is bounded from below then (3.2) is fulfilled for any x ∈ RN. Hence the claims in

A hold true.

C. If there exists a function h ∈C2(RN) such that L(h) = 0 and u(x) = h(x) +

Z

RN

Γ(x, y)ν(y)dy, then for any x ∈ RN,

h(x) = 1 ln 2lim infR→+∞ Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx u. (3.7)

Remark 7. We notice, see (3.36) below, that condition (3.2) is equivalent to require that

lx= 1 ln 2lim infR→+∞ Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx u (3.8)

exists and it is finite. See also Remark 10below.

The next result deals with an assumption on the gradient of ∇u assuring that (3.2) is satisfied. For additional results of this type see Lemma 29below.

(8)

Theorem 8. Let x ∈ RN and u ∈ C2

(RN) be such that −L(u) = ν ≥ 0. Assume that

|A0∇Γ| ∈ L1 loc(RN). If lim inf r→+∞ Z Ωr(x) (∇Γx· A∇u)dy < +∞,

then there exists lx ∈ R such that (3.2) is satisfied and (3.3) holds. Moreover, we have

lim r→+∞ Z Ωr(x) (∇Γx· A∇u)dy = Z RN Γ(x, y)ν(y)dy.

Remark 9. The fact the integral in (3.3) is well posed (finite) is a consequence of our result. Remark 10. Theorem 6 can be read as that an L-superharmonic function u (with no assumptions on its sign) can be decomposed into a positive superharmonic function (w) and an harmonic function (l). Clearly in general, this fact is not true without further assumptions. Indeed the function u(x) := −|x|2 is superharmonic: −∆u = 2N , but it cannot be written as u = l + w with l harmonic and w nonnegative and superharmonic2. In

this respect, the hypothesis (3.2) plays a crucial role. Roughly speaking, the limit (3.2) or more explicitily (3.8), can be seen as a procedure to extract from a superharmonic function its harmonic component.

Notice that if we assume that the integral in (3.3) is finite everywhere, then this kind of decomposition is immediate. Hence we can claim that,

If the equation

− L(h) = ν ≥ 0 (3.9)

has a solution u ∈C2(RN) satisfying (3.2), then any solution v of (3.9) can be decomposed into an harmonic function and a nonnegative superharmonic function.

Finally a natural question arises: is the ring condition (3.2) necessary to obtain such a decomposition? That is if −Lu = ν ≥ 0 and assume that u can be decomposed as above, i.e. u(x) = h(x) +

Z

RN

Γ(x, y)ν(y)dy. Is the limit in (3.8) finite? A positive answer to this question is given by the claim Cof Theorem 6.

Remark 11. We emphasise that in Theorem6-A we do not assume that u is bounded from below. Indeed this fact is a consequence of the positivity of the fundamental solution and (3.6). In particular if (3.2) holds with l ≥ 0, this gives a first answer to the positivity problem.

We note that a different representation formula under different assumptions (see in par-ticular the assumptions on the symmetry of the matrix A and on the not-totally degeneracy of A) has been proved in [1], under the hypothesis that u is bounded from below. See Theorem 2.4.5 and Corollary 2.4.6 of [1] for details.

Remark 12. The ring condition (3.2) plays a central role in Theorem 6, and the relation (3.5) give us a rate of convergence in the limit in (3.2), namely

Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx (u − lx)dy ≥ C R, ∀x ∈ R N and f or R large. (3.10)

Therefore if the integral, for some x ∈ RN, vanishes too fast as R → +∞, then the function

u is harmonic.

2Indeed if this would be true, since w is nonnegative it must satisfies w(x) ≥ C

Z

(9)

Furthermore the estimate (3.10), at least in this generality is sharp, in the sense that there exists a solution of −L(u) ≥ 0 such that

lim sup R→+∞ R Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx (u − lx)dy ≤ C0 < +∞.

To this end consider L = ∆ the Euclidean Laplacian operator, N > 2, u(z) = (1+|z|2)(2−N )/2 and x = 0 and l = 0. It is clear that u is a strict positive superharmonic function. By computation, taking into account that

AR := Ω2R(0) \ ΩR(0) = {(2R)−1 < (|y|2−N < R−1} = B(2R)1/N −2 \ BR1/N −2, we have Z AR (∇Γ0· A∇Γ0) Γ0 udy = Z AR (N − 2)2 |y|N 1 (1 + |y|2)(N −2)/2dy ≤ cR − N N −2R−1|A R| = cR−1,

proving the sharpness of the rate (3.10)

The proof of Theorems6and8as well as some other sufficient conditions that guarantee the representation formulae (3.3) or (3.6) are contained in Lemma26and Lemma 29below.

Remark 13. In the case L = ∆, if u is a nonnegative superharmonic function then from Theorem6it follows that l is a positive classical harmonic function. Hence l is a nonnegative constant by the classical Liouville theorem. In our general setting, the Liouville theorem may not be true. Actually, the next results claims the equivalence between Riesz representation result, Liouville type theorem and the validity of some ring condition as in (3.2).

Theorem 14. Assume that if u ∈ L1

loc(RN) is a distributional solution of Lu = 0, then u

coincides almost everywhere with a C2(RN) function. The following statements are equivalent.

1. Let u ∈ C2(RN). Then −L(u) = ν ≥ 0 and inf u = m ∈ R if and only if

u(x) = m + Z

RN

Γ(x, y)ν(y)dy ∀x ∈ RN. (3.11)

2. Let u ∈ C2(RN). If u is bounded from below and −L(u) = 0 then u is constant. 3. Let u ∈ C2(RN) be such that −L(u) ≥ 0 and l ∈ R. Then, inf u = l if and only if

lim inf R→+∞ Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx (u − l)dy = 0 ∀x ∈ RN (3.12) if and only if lim R→+∞ Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx |u − l|dy = 0 ∀x ∈ RN (3.13) if and only if lim inf R→+∞ Z ∂ΩR(x) (∇Γx· A∇Γx) |∇Γx| (u − l)dHn−1 = 0 ∀x ∈ RN (3.14)

(10)

if and only if lim R→+∞ Z ∂ΩR(x) (∇Γx· A∇Γx) |∇Γx| |u − l|dHn−1= 0 ∀x ∈ RN (3.15) if and only if l = lim R→+∞ 1 + α R1+α Z ΩR(x) (∇Γx· A∇Γx) Γ2+α x udy α > −1 ∀x ∈ RN. (3.16)

4. Let u ∈ C2(RN). Suppose that u is bounded from below and −L(u) ≥ 0. Set

cx:= lim inf R→+∞ 1 ln 2 Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy ∀x ∈ RN, (3.17) then cx does not depend on x: cx = c ∈ R.

If one of the above statements is true then l = m = c.

Remark 15. Condition 3. of the above Theorem 14, can be restated as

inf u = lim inf

R→+∞ 1 ln 2 Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy ∀x ∈ RN, (3.18) provided the quantities involved are finite. A natural question arises: does the relation (3.18) hold even when the quantities are not finite?

If for some x ∈ RN lim inf R→+∞ 1 ln 2 Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy = −∞,

then from (3.36) one deduces that inf u = −∞. The converse is not true as the following example shows. Let L = ∆ the Laplacian in the Euclidean case, and let u be a non constant harmonic function. Clearly u is superharmonic and inf u = −∞. However, since u is harmonic, we have (see (3.44) below),

lim R→+∞ 1 ln 2 Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy = 1 ln 2 Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy = u(x) ∈ R.

Remark 16. An alternative formulation of Theorem 14 is the following. The classical Liouville theorem holds for the operator L if and only if a representation formula for non negative supersolution holds. These properties are equivalent to the fact that the infimum of a supersolution can be characterised by the ring condition (3.12). Finally all the above statements are equivalent to the fact that the limits of the means involved in the conditions (3.12)-(3.17) on the ring centered at x do not depend on x. Indeed they are constant. Remark 17. The main Theorems 6 and 14 can be formulated even in a more general situation, for instance replacing the whole Euclidean space RN with a domain Ω (open

connected set).

To this end the hypotheses on the operator (and hence on the fundamental solution) must be changed accordingly replacing in all H1. . . . H7., RN with Ω. In addition since H4.

(11)

H4. for any x ∈ RN and for any z ∈ ∂Ω, lim

y→zΓ(x, y) = 0 uniformly, that is

∀ > 0, ∃n > 0 such that ∀y ∈ Ω \ Cn : Γ(x, y) < 

where Cn is the compact set Cn := {z ∈ Ω : dist(z, ∂Ω) ≥ 1/n and |x| ≤ n}.

With these changes Theorems 6 and 14 hold as they are stated replacing RN with Ω. Very simple examples related to the Euclidean Laplacian are for instance, bounded regular domains for which there exists a Green function and half spaces.

Remark 18. In what follows we are going to show a variant of Theorem 14. For a given operator L described in section 2. the classical Liouville theorem may not be true. However, it may be true for functions belonging to some subspace ofC2(RN). For instance for bounded

functions or more generally for functions with a prescribed growth at infinity or even for a summability space. In this case the equivalences of the Theorem 14remain valid.

Indeed we have,

Theorem 19. Let X be a subspace of C2(RN) satisfying3

if u ∈ X and v ∈C2(RN) such that |v| ≤ u, then v ∈ X. If in Theorem 14 we replace the space C2(RN) with C2

(RN) ∩ X, then the equivalences

1. ⇔ 2.. ⇔ 3. ⇔ 4. hold.

See subsections 3.1–3.5 for the proof.

Remark 20. In order to reformulate the ring condition (3.2) in particular cases we shall use the following notation. Fix x ∈ RN and choose a real number D

x > 2. Next we

define gx(y) := Γ

1 2−Dx

x (y). The function gx : RN → R is continuous, positive and smooth

in RN \ {x} and gx(x) = 0. We set h2x := (∇gx · A∇gx). The function gx plays the

the same role that the gauge N2(x−1 ◦ ·) plays in the Carnot group setting. Notice that

y ∈ ΩR(x) ⇔ gx(y) < R1/(Dx−2).

We wish to point out that the choice of Dx > 2 is left free.

Let us to explicitly compute some relevant quantities involved in this paper. From definition of gx we have (∇Γx· A∇Γx) Γx = (2−Dx)2g2−2Dx x (∇gx· A∇gx) g2−Dx x = (2−Dx)2 h2 x gDx x = (Dx−2)2Γ Dx Dx−2h2 x. (3.19)

Hence from Lemma 26(see below) we see that Z Ω2R(x)\ΩR(x) ΓDx−2Dx h2 xdy = 1 (2 − Dx)2 Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx dy = ln 2 (2 − Dx)2 , (3.20) for any R > 0 and x ∈ RN. This justifies why we shall refer to the integral

1 ln 2 Z Ωr(x)\Ωr/2(x) (∇Γx· A∇Γx) Γx udy, 3For instance X = Lp (RN) ∩C2 (RN

) or if G : RN →]0, +∞] is a positive function then X could be the

space of functions with a growth prescribed by the weight G, namely

(12)

as a mean.

With this notation the ring conditions (3.2), see also Lemma 26, can be stated as

lim inf R→+∞ Z R<gx(y)<2R h2 x(y) gDx x (y) (u(y) − lx)dy = 0. (3.21)

For instance in the case L is a sub-Laplacian on a Carnot group, Dx = Q > 2 is

constant and gx(y) = N2(x−1 ◦ y) where N2 is a symmetric homogeneous norm. With

this notation, h2x = ψ2x and it is bounded, Ωr(x) = {y ∈ RN|N2(x, y) < r

1 Q−2} and since |Ωr(x)| = BN(r 1 Q−2) = Cr Q

Q−2, for functions that are bounded from below the condition

(3.21) and hence (3.2), becomes

lim inf R→+∞ Z − R<N2(x−1◦y)<2R ∇LN2(x−1◦ y) 2 (u(y) − l)dy = 0, (3.22) where Z − U := 1 |U | Z U .

See subsection 4 for the analysis in Carnot Groups.

Moreover in the special case in which the Carnot group is the classical Euclidean space with L = ∆, then (3.2) for functions that are bounded from below can be read as follows,

lim inf R→+∞ Z − R<|x−y|<2R (u(y) − l)dy = 0. (3.23)

3.1

Lemmata

In this subsection we prove some lemmata which will be used in proving Theorems 6, 8and

14.

In similar settings in [15,7,8] the authors proved the implication 3. ⇒ 1. of Theorem14

directly using a technique based on the test functions method. Even in our setting this technique can be applied: in this case we have to choose tests functions that depends on the fundamental solution of L. This idea is used in subsections 3.2 and 3.6.1. However in this paper for the proof of the representation formulae we shall use a different pattern. Indeed we will give a proof which is based on the local representation formulae (3.25) and (3.26). This choice is motivated by the fact that even the other implications are based on the local representation formula for a function u.

As a consequence of divergence theorem, we have that Z U Lu = Z ∂U (A∇u · ν) dHN −1, (3.24) for any smooth open set U and any u smooth function on U . In (3.24), ν denotes the exterior normal unitary vector to ∂U and dHN −1 denotes the the (N − 1)-dimensional Hausdorff measure.

We shall use (3.24) with U = Ωr(x). We notice that by Sard’s Lemma these sets are

regular for almost every r > 0.

(13)

Lemma 21. Let u ∈C2(RN), x ∈ RN, R > 0, 0 < δ < 1, α 6= −1, for a.e. r > 0, then u(x) = Z ∂Ωr(x) (∇Γx· A∇Γx) |∇Γx| udHn−1+ Z Ωr(x) (−Lu)(Γx− 1 r)dy (3.25) u(x) = 1 − ln δ Z ΩR(x)\ΩδR(x) (∇Γx· A∇Γx) Γx udy+ (3.26) + Z R δR 1 r Z Ωr(x) (−Lu)(Γx− 1 r)dy dr  , u(x) = α + 1 (1 − δα+1)Rα+1 Z ΩR(x)\ΩδR(x) (∇Γx· A∇Γx) Γα+2 x udy+ (3.27) + Z R δR rα Z Ωr(x) (−Lu)(Γx− 1 r)dy dr  , and δ = 0 is allowed in (3.27) provided α > −1.

Moreover if |A0∇Γ| ∈ L1 loc(RN), then u(x) = Z ∂Ωr(x) (∇Γx· A∇Γx) |∇Γx| udHn−1+ Z Ωr(x) (∇Γx· A∇u)dy. (3.28)

Proof. In what follows the dependence on x will be omitted, hence Ωs stands for Ωs(x)

and Γ stands for Γx. Let 0 <  < r and set U := Ωr\ Ω. Since −L∗Γ = 0 on U, we have

Z U (A0∇Γ · ∇u) = Z ∂U (A0∇Γ · ν)u dHN −1 Z U (L∗Γ)u = Z ∂Ωr (A0∇Γ · ν)u dHN −1 Z ∂Ω (A0∇Γ · ν)u dHN −1. (3.29) On the other hand,

Z U (A0∇Γ · ∇u)dy = Z 1 r<Γ< 1  (∇(Γ −1 r) · A∇u)dy = Z ∂Ωr (Γ − 1 r)(A∇u · ν)dH N −1 Z ∂Ω (Γ −1 r)(A∇u · ν)dH N −1 Z 1 r<Γ< 1  (Γ −1 r)Lu dy = −(1  − 1 r) Z ∂Ω (A∇u · ν)dHN −1− Z 1 r<Γ< 1  (Γ −1 r)Lu dy = −(1  − 1 r) Z Ω Lu dy − Z 1 r<Γ< 1  (Γ − 1 r)Lu dy, where in the last identity we have used (3.24).

Let φ ∈C∞0 (RN) such that φ = 1 on Ω

. By hypotheses we have u(x) = − Z RN ΓxL(uφ) dy = − lim →0 Z Γ≤1 ΓxL(uφ) dy. (3.30)

Therefore integrating by parts, and taking into account that φ(y) = 1 for Γx(y) = 1, we

obtain − Z Γ≤1 ΓxL(uφ)dy = Z Γ≤1 (A0∇Γx· ∇(uφ))dy − 1  Z ∂Ω (A∇(uφ) · ν)dHN −1 = − Z Γ≤1 L∗(Γx)(uφ)dy − Z ∂Ω (A0∇Γx· ν)uφdHN −1− 1  Z ∂Ω (A∇u · ν)dHN −1 = − Z ∂Ω (A0∇Γx· ν)u dHN −1− 1  Z Ω Lu dy, (3.31)

(14)

where in the last identity we have used the fact that L∗(Γx(y)) = 0 for y 6= x and the

relation (3.24). Gluing together the last relations, and observing that the normal ν can be written as ν = −|∇Γ|∇Γ we have u(x) = Z ∂Ωr (A0∇Γx· ∇Γx |∇Γx| )u dHN −1− Z Ωr (Γ −1 r)Lu dy + lim→0( 1 r − 2 ) Z Ω Lu dy. From (2.2) we get (3.25).

The identity (3.28) follows formally from (3.25) by integration by parts. In a more precise way, as before gluing (3.30), (3.29) and (3.31), we have

u(x) = lim →0 Z U (A0∇Γ · ∇u) dy − Z ∂Ωr (A0∇Γ · ν)u dHN −1− 1  Z Ω Lu dy, which yields (3.28).

In order to obtain the missing relations, it is enough to multiply (3.25) by rα, then integrate with respect to variable r between [δR, R] and using the coarea formula to show that Z R δR  rα Z ∂Ωr (∇Γx· A∇Γx) |∇Γx| u  dr = Z R δR Z ∂Ωr (∇Γx· A∇Γx) Γα|∇Γ x| udr = Z ΩR(x)\ΩδR(x) (∇Γx· A∇Γx) Γα+2 x u dy. This completes the proof of (3.26) and (3.27) for δ > 0.

In order to get the proof for δ = 0, we remark that choosing u = 1 in (3.27) we have

1 = α + 1 (1 − δα+1)Rα+1 Z ΩR(x)\ΩδR(x) (∇Γx· A∇Γx) Γα+2 x dy.

By letting δ → 0 we have that (∇Γx·A∇Γx)

Γα+2x ∈ L

1

loc(RN), and hence we can pass to the limit

(as δ → 0) in (3.27) completing the proof. 2 Remark 22. With the notation of Remark 20, by a rescaling (r → rDx−1), the identities

(3.25), (3.26) and (3.27) of Lemma 21, for any x ∈ RN, R > 0, 0 < δ < 1, β 6= 0, and for

a.e. r > 0, read respectively as u(x) = Dx− 2 rDx−1 Z gx=r h2x |∇gx| udHn−1+ Z gx<r (−Lu)(Γx− 1 rDx−2)dy (3.32) u(x) = 1 − ln δ  (Dx− 2) Z δR<gx<R h2x gDx x (y)u(y)dy+ (3.33) + Z R δR 1 r Z gx<r (−Lu)(Γx− 1 rDx−2)dy dr  , u(x) = β (1 − δβ)Rβ  (Dx− 2) Z δR<gx<R h2x gDx−β x udy (3.34) + Z R δR rβ r Z gx<r (−Lu)(Γx− 1 rDx−2)dy dr  , and δ = 0 is allowed in (3.34) provided β > 0.

(15)

Choosing u ≡ 1 in Lemma 21 we have Corollary 23. Let x ∈ RN, α > −1, then

Z ∂Ωr(x) (∇Γx· A∇Γx) |∇Γx| dHn−1 = 1 f or a.e. r > 0; (3.35) 1 ln η Z ΩηR(x)\ΩR(x) (∇Γx· A∇Γx) Γx dy = 1 ∀R > 0, η > 1; (3.36) α + 1 Rα+1 Z ΩR(x) (∇Γx· A∇Γx) Γα+2 x dy = 1 ∀R > 0. (3.37) Lemma 24. Let u ∈C2(RN) be such that −L(u) ≥ 0. For any x ∈ RN, the functions

r 7−→ Nx(r) := Z ∂Ωr(x) (∇Γx· A∇Γx) |∇Γx| udHn−1, (3.38) r 7−→ Mx(r) := Z Ωr(x) (−Lu)(Γx− 1 r)dy, (3.39)

are nonnincreasing and nondecreasing, respectively. Moreover for any x ∈ RN and R > 0, there holds

u(x) ≥ Z ∂Ωr(x) (∇Γx· A∇Γx) |∇Γx| udHn−1, (3.40) u(x) ≥ 1 ln η Z ΩηR(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy, η > 1 (3.41) u(x) ≥ α + 1 (1 − δα+1)Rα+1 Z ΩR(x)\ΩδR(x) (∇Γx· A∇Γx) Γα+2 x udy, R > 0, α > −1, 1 > δ ≥ 0.(3.42) In particular if L(u) = 0 then

u(x) = Z ∂Ωr(x) (∇Γx· A∇Γx) |∇Γx| udHn−1, (3.43) u(x) = 1 ln η Z ΩηR(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy, η > 1. (3.44) Proof. Let s > r. Since the integrand in Mx is nonnegative and Ωr ⊂ Ωs, we have

Mx(s) = Z Ωs(x) (−Lu)(Γx− 1 s)dy ≥ Z Ωr(x) (−Lu)(Γx− 1 s)dy ≥ Mx(r).

On the other hand from (3.25), Mx(r) + Nx(r) = u(x) is constant with respect to r. Thus

Nx is non-increasing.

The remaining relations are an easy consequence of the identities in Lemma 21 and the

fact that −L(u) ≥ 0. 2

Lemma 25. Let ν be a regular measure. Let w ∈ L1loc(RN) defined by w(x) :=

Z

RN

Γ(x, y)ν(y)dy a.e x ∈ RN. Then −L(w) = ν in the distributional sense.

(16)

Proof. To this end it is enough to show that, for any φ ∈C∞0 (RN) we have, Z RN −L∗(φ)(x)w(x) = Z RN φ(x)ν(x). Indeed, multiplying w by −L∗(φ) we have,

Z RN −L∗(φ)(x)w(x)dx = Z RN Z RN −L∗(φ)(x)Γ(x, y)ν(y)dydx = Z RN φ(y)ν(y).

Here we have used the fact that Γ(·, y) is a fundamental solution of −L∗. 2 Lemma 26. Let u ∈ C2(RN) be such that −L(u) ≥ 0. Let x ∈ RN and lx ∈ R. Then the

following implications hold.

(3.47) ⇒ (3.46) ⇒ (3.51) ⇔ (3.50) ⇔ (3.54) ⇓ ⇑ m m (3.49) ⇒ (3.48) ⇒ (3.53) ⇔ (3.52) (3.45) where, for η > 1, lim inf R→+∞ Z ∂ΩR(x) (∇Γx· A∇Γx) |∇Γx| |u − lx| dHn−1 = 0; (3.46) lim R→+∞ Z ∂ΩR(x) (∇Γx· A∇Γx) |∇Γx| |u − lx| dHn−1 = 0; (3.47) lim inf R→+∞ Z ΩηR(x)\ΩR(x) (∇Γx· A∇Γx) Γx |u − lx| dy = 0; (3.48) lim R→+∞ Z ΩηR(x)\ΩR(x) (∇Γx· A∇Γx) Γx |u − lx| dy = 0; (3.49) lim inf R→+∞ Z ∂ΩR(x) (∇Γx· A∇Γx) |∇Γx| udHn−1 = lx; (3.50) lim R→+∞ Z ∂ΩR(x) (∇Γx· A∇Γx) |∇Γx| udHn−1 = lx; (3.51) lim inf R→+∞ 1 ln η Z ΩηR(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy = lx; (3.52) lim R→+∞ 1 ln η Z ΩηR(x)\ΩR(x) (∇Γx· A∇Γx) Γx udy = lx; (3.53) lim R→+∞ 1 + α R1+α Z ΩR(x) (∇Γx· A∇Γx) Γ2+α x udy = lx α > −1. (3.54)

Moreover, if one of the above statements holds, then u(x) = lx+

Z

RN

Γ(x, y)ν(y)dy.

Here ν := −L(u). If (3.52), or equivalently (3.53) or (3.50) or (3.51), holds and u(y) ≥ lx

(17)

Proof. Taking into account Corollary23, the implications (3.47)⇒ (3.46), (3.47)⇒ (3.51)⇒ (3.50), (3.49)⇒ (3.48), (3.49)⇒ (3.53)⇒ (3.52), are obvious.

Without loss of generality we can assume lx = 0. Let us also remind that Nx(r) is non

increasing with respect to r as stated in Lemma 24. We will use the following notation

Kx := (∇Γx· A∇Γx) |∇Γx| , Kx := (∇Γx· A∇Γx) Γx . (3.46)⇒ (3.51). Since lim inf

R(x)→∞

Z

∂ΩR(x)

Kx|u| dσ = 0, there exists a sequence (Rn) such

that lim

n

Z

∂ΩRn(x)

Kx|u| dσ = 0. Therefore lim n

Z

∂ΩRn(x)

Kx|u| dσ = 0, that is Nx(Rn) → 0.

From the monotonicity of Nx we get the claim.

(3.50)⇒ (3.51). It follows from monotonicity of Nx.

(3.48)⇒ (3.46). Arguing by contradiction, we suppose that

lim inf

r

Z

∂Ωr(x)

Kx|u| dσ = c > 0.

This implies that for r large, we have Z

∂Ωr(x)

Kx|u| dσ ≥ c/2.

Using the coarea formula we have Z ΩηR(x)\ΩR(x) Kx|u| dy = Z ηR R 1 r Z ∂Ωr Kx|u| dσ ≥ c 2 Z ηR R dr r = c ln η 2 > 0, which contradicts the assumption (3.48).

(3.47)⇒ (3.49) Let  > 0. By assumption (3.47) there exist r0 > 0 such that for r > r0

we have R∂Ω

r(x)Kx|u| < . Integrating we have

Z ΩηR(x)\ΩR(x) Kx|u| dy = Z ηR R 1 r Z ∂Ωr Kx|u| <  Z ηR R dr r =  ln η. (3.51)⇒ (3.53) Since Nx is nonincreasing we have

ln ηNx(ηR) = Z ηR R Nx(ηR) dr r ≤ Z ηR R Nx(r) dr r = Z ΩηR(x)\ΩR(x) Kxu dy ln ηNx(R) = Z ηR R Nx(R) dr r ≥ Z ηR R Nx(r) dr r = Z ΩηR(x)\ΩR(x) Kxu dy.

Since N (R) → 0 (as R → ∞), these inequalities imply the claim. (3.52)⇒ (3.50) Let

b := lim inf

r→∞ Nx(r).

Arguing as in the proof of the implication (3.51)⇒ (3.53) we have that

lim inf R→∞ 1 ln η Z ΩηR(x)\ΩR(x) Kxu dy = b.

(18)

This implies b = 0.

(3.51)⇒ (3.54) Since Nx is nonincreasing we have,

R1+α 1 + αNx(R) = Z R 0 Nx(R)rαdr ≤ Z R 0 Nx(r)rαdr = Z Ω(R) Kx Γ1+α x u dy. Hence, Nx(R) ≤ 1 + α R1+α Z Ω(R) Kx Γ1+α x u, (3.55) which implies 0 ≤ lim inf 1 + α R1+α Z Ω(R) Kx Γ1+α x u dy. The reverse inequality follows from the simple calculation,

1 + α R1+α Z Ω(R) Kx Γ1+α x udr = 1 + α R1+α Z R 0 rαNx(r)dr = 1 + α R1+α Z s 0 rαNx(r)dr + Z R s rαNx(r)dr  ≤ 1 + α R1+α Z s 0 rαNx(r)dr + 1 + α R1+αNx(s) Z R s rαdr = 1 + α R1+α Z s 0 rαNx(r)dr + Nx(s) R1+α− s1+α R1+α

where 0 < s < R. Indeed, letting R → +∞ we get, 0 ≤ lim sup R 1 + α R1+α Z Ω(R) Kx Γ1+α x u dy ≤ Nx(s),

which implies the claim.

(3.54)⇒ (3.51) Let b := lim infr→∞Nx(r) = limr→∞Nx(r). We note that b = −∞ is

allowed. From (3.55) we get b ≤ 0. Taking the limit for R → +∞ in (3.1) we obtain Nx(s) ≥ 0 which implies the claim.

To conclude the proof assume that (3.51) is satisfied and u(y) ≥ lx for any y ∈ RN.

Therefore |u(y) − lx| = u(y) − lx then (3.47) holds. Thus from the scheme (3.45) it follows

that all implications are verified. 2

Remark 27. The missing implication (3.53) =⇒ (3.48) in general is not true. Indeed, in the Euclidean setting with L = ∆, we choose u(x) = u(x1, . . . , xN) := x1. This function is

harmonic in RN. Hence by (3.26) and using the notation of the above proof, we have u(0) = 0 = 1 ln 2 Z Ω2R(x)\ΩR(x) Kxu dy. However 1 ln 2 Z Ω2R(x)\ΩR(x) Kx|u| dy = CR → +∞.

(19)

3.2

A priori estimates and the ring condition

In this section we do not make any assumption on the sign of the solution u. We begin with an universal estimate which involves a ring condition.

Let q > 1, in what follows by q-admissible function we mean a function φ ∈ C∞0 (R) in the form φ = ϕγ0 with γ > q−12q an integer and ϕ0 ∈C∞0 (R) such that 0 ≤ ϕ0 ≤ 1, ϕ0(t) = 1

if |t| ≤ 1, ϕ0(t) = 0 if |t| ≥ 2.

Theorem 28. Let q > 1. For any φ1 ∈C∞0 (R) q-admissible function, there exists M such

that if u is a weak solution of −L(u) ≥ ν ≥ 0, then we have Z Ω2R(x) φR(y)ν(y)dy ≤ M R2 Z Ω2R(x)\ΩR(x)

|u(y)| φ1/qR (y)(∇Γx(y) · A(y)∇Γx(y)) Γ4

x(y)

dy (3.56)

for any R > 0 and a.e. x ∈ RN, where φ

R(y) := φ1(1x(y)). In particular we have Z Ω2R(x) φR(y)ν(y)dy ≤ 8M R Z Ω2R(x)\ΩR(x)

|u(y)| φ1/qR (y)(∇Γx(y) · A(y)∇Γx(y)) Γx(y)

dy. (3.57)

Here the constant M depends only on φ1: M := sup1<t<2

|φ001(t)+2φ01(t)/t|

φ1/q1 (t) < +∞.

Proof. In what follows we shall omit to write the dependence on x (ΩR= ΩR(x), Γ = Γx).

From our choice of φR it follows that it vanishes outside Ω2R(x) and φR = 1 in ΩR(x).

Using φR in Definition 3.1 we have

Z φRν ≤ − Z uL∗(φR) ≤ Z |u| |div(A0∇φ R)| . By computation we have ∇φR= −φ01( 1 RΓ) ∇Γ RΓ, L ∗ (φR) = div  A0∇Γ RΓφ 0 1( 1 RΓ)  = = −φ01( 1 RΓ) L∗Γ RΓ + (∇Γ(y) · A∇Γ(y))  φ00 1(t) R2Γ4 + 2 φ01(t) RΓ3  t=1/RΓ = 0 +(∇Γ(y) · A∇Γ(y)) R2Γ4  φ001(t) + 2φ 0 1(t) t  t=1/RΓ

where we have used the fact that φ0R= 0 in a neighbourhood of x and L∗(Γ) = 0 away from x. Therefore we have, Z Ω2R(x) φRνdy ≤ Z Ω2R(x)\ΩR(x) |u|(∇Γ · A∇Γ) R2Γ4 φ001(t) + 2φ 0 1(t) t t=1/RΓ dy ≤ M R2 Z Ω2R(x)\ΩR(x) |u|(∇Γ · A∇Γ) Γ4 dy,

(20)

3.3

Ring condition (3.2) related to ∇u

The results of this subsection deal with some conditions on the gradient of u assuring that some form of the ring condition as in Lemma (26) is satisfied.

Lemma 29. Assume that |A0∇Γ| ∈ L1

loc(RN). Let u ∈C 2

(RN) be such that −L(u) ≥ 0 and

let x ∈ RN. Then the following statements are equivalent. 1. There exists lx ∈ R such that (3.50) holds that is

lim inf R→+∞ Z ∂ΩR(x) (∇Γx· A∇Γx) |∇Γx| udHn−1 = lx. 2. lim inf r→+∞ Z Ωr(x) (∇Γx· A∇u)dy < +∞. 3. sup r>0 Z Ωr(x) (∇Γx· A∇u)dy < +∞,

4. Let Dx > 2 and set gx := Γ

1 2−Dx x . Then (2 − Dx) Z +∞ 1 dt tD x Z gx(y)<t (∇gx· A∇u)dy < +∞. 5. Z +∞ 1 dt t2 Z ∂Ωt(x) (∇Γx· A∇u) |∇Γx| < +∞. 6. Let σ := σ(x) > 1. Then Z +∞ 1 dt tσ Z Ωt(x) (∇Γx· A∇u) Γxσ−1 < +∞. Moreover, if one of the above statements holds, then

u(x) = lx+

Z

RN

Γ(x, y)ν(y)dy. Here ν is the Radon measure ν := −L(u) and

lim inf r→+∞ Z Ωr(x) (∇Γx· A∇u)dy = Z RN Γ(x, y)ν(y)dy.

Remark 30. We notice that after combining the above lemma with what we have already proved and Theorem 6, we have now completed the proof of the Theorem 8.

Before proving the lemma we state another result which seems to be of some interest in itself.

Proposition 31. Let x ∈ RN and let u ∈C2

(RN) be such that −L(u) ≥ 0. Then

Z ∂Ωr(x) (∇Γx· A∇u) |∇Γx| ≥ 0, for a.e. r > 0; (3.58) Z Ωr(x)\Ω1(x) (∇Γx· A∇u) Γxγ

(21)

Proof. Let R > 1 and γ ∈ R. Using coarea formula we have Z 1>Γx>R1 (∇Γx· A∇u) Γxγ = Z R 1 Z ∂Ωr(x) (∇Γx· A∇u) Γγ−2x |∇Γx|  dr (3.60) = Z R 1 rγ−2 Z ∂Ωr(x) (∇Γx· A∇u) |∇Γx|  dr. (3.61) In particular if γ = 0, the above identity becomes

Mx(R) − Mx(1) = Z R 1 r−2 Z ∂Ωr(x) (∇Γx· A∇u) |∇Γx|  dr. (3.62) Since M is nondecreasing with respect to R it follows that its derivative is a.e. nonnegative, and (3.58) follows.

Finally, inequality (3.59) follows from (3.58) and (3.60). 2 Proof of Lemma 29 From Lemma 21, and Lemma 24we know that for any r > 0

u(x) = Nx(r) + Mx(r),

where Mx(r) =

Z

Ωr(x)

(∇Γx· A∇u)dy.

Therefore the equivalence 1. ⇔ 2. easily follows.

The monotonicity of Mx(r) gives the equivalence 2. ⇔ 3.

The equivalence 2. ⇔ 5. follows directly from identity (3.62). Since Mx(r) is nondecreasing statement 3., can be rewritten as

lim

r→+∞

Z

Ωr(x)\Ω1(x)

(∇Γx· A∇u)dy < +∞.

Let σ > 1. Observing that

1 sσ−1 = Z +∞ s σ − 1 tσ dt, we obtain Z ΩR(x)\Ω1(x) (∇Γx· A∇u) = Z 1>Γx>R1 (1/Γ)σ−1(A∇Γx· ∇u) (1/Γ)σ−1 = Z 1>Γx>R1 (1/Γ)σ−1(∇Γx· A∇u) Z +∞ 1/Γ σ − 1 tσ dt = = (σ − 1) Z +∞ 1 dt tσ Z 1>Γ>max(1t,R1) (∇Γx· A∇u) Γσ−1 ! = (σ − 1) Z +∞ 1 dt tσ Z Ωt∩ΩR\Ω1 (∇Γx· A∇u) Γσ−1  . Since the integrand of

Z

Ωt∩ΩR\Ω1

(∇Γx· A∇u)

Γσ−1 dy,

is nonnegative we can pass to the limit as R → +∞, and by using Beppo Levi theorem the equivalence 3. ⇔ 6. follows.

Hence the equivalence 3. ⇔ 4. is proved.

(22)

3.4

Proof of Theorem

6

We begin noticing that the proof of the representation (3.3) in Theorem 6 follows from Lemma26.

It remains to prove the others claims of the statement.

Proof. [ Aa] First we prove that inf l ≥ inf u. Without loss of generality we assume that inf u > −∞. From the ring condition 3.2 we have

l(x) = 1 ln 2 lim infR→∞ Z Ω2R (x) \ ΩR(x) (∇Γx· A∇Γx) Γx u dy ≥ ≥ 1 ln 2lim infR→∞ Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx inf u dy = inf u. Hence inf l ≥ inf u.

Next, since u(x) ≥ l(x), let (xn) be a minimising sequence of u. From the inequality

u(xn) ≥ l(xn) ≥ inf u,

the claim follows.

Analogously we obtain the estimate of sup. The alternative (3.4) follows directly from (3.3). Proof. [Ab.] The first claim follows from Lemma25.

In order to prove (3.5), we set v(z) := u(z) − l(x) therefore v is a distributional positive solution of −L(v) = ν. Therefore from the estimate (3.57) (see below Theorem 28), for R large we have Z Ωs(0) ν(y)dy ≤ Z ΩR(x) ν(y)dy ≤ 8M R Z Ω2R(x)\ΩR(x)

v(y)(∇Γx(y) · A(y)∇Γx(y)) Γx(y)

dy, (3.63) and hence the claim follows.

Proof. [Ac.] The function w is finite on the whole space, and it is measurable. Therefore l is measurable too. Since l and w are nonnegative and their sum is the regular function u, it follows that they belong L1

loc(RN).

Proof. [Ad] It follows directly from (3.3).

Proof. [B] Without loss of generality we can assume that u is nonnegative.

From Lemma 24, it follows that the function N (r) is nonincreasing and since the inte-grand is nonnegative, N admits a finite nonnegative limit (as r → +∞), that is (3.51) is fulfilled and by Lemma 26, we get the claim.

Proof. [C] Set Kx:= 1 ln 2 (∇Γx· A∇Γx) Γx , and AR(x) := Ω2R(x) \ ΩR(x). Let w(x) := Z RN

Γ(x, y)ν(y)dy. Since w ∈ C2(RN) is nonnegative and −L(w) = ν from B

of Theorem 6 it follows that for w satisfying (3.8) and

w(x) = h1(x) +

Z

RN

Γ(x, y)ν(y)dy, (3.64) we have h1(x) ≡ 0. On the other hand, h1 is the harmonic function given by

h1(x) = lim inf R→∞

Z

AR(x)

(23)

Hence, lim inf R→∞ Z AR(x) Kxw dy = 0 ∀x ∈ RN.

Therefore, from the decomposition of u, by using (3.44) for the harmonic function h we have Z AR(x) Kxu dy = Z AR(x) Kxh dy + Z AR(x) Kxw dy = h(x) + Z AR(x) Kxw dy.

By taking the limit as R → ∞ we get the claim.

3.5

Proof of Theorems

14

and

19

Theorem 14 can be seen as a particular case of Theorem19 by taking X ≡C2(RN). So we

shall present the proof for functions belonging toC2(RN) ∩ X.

The implications 1. ⇒ 2. and 3. ⇒ 4. are immediate.

2. ⇒ 3. Let u ∈ C2(RN) ∩ X be such that −L(u) = ν ≥ 0. Let l ∈ R and assume that

the ring condition (3.12) holds. We have to prove that inf u = l.

Since the (3.12) holds for any x ∈ R, we are in the position to apply pointA of Theorem

6and from point Aa of the same theorem, we have that inf u = inf lx = l.

To prove the converse implication, assume that inf u = l ∈ R, we have to prove (3.12). Since u is bounded from below, fromB of Theorem6it follows that the ring condition (3.2) holds for any x. Now our claim is to prove such a function l(x) is constant.

FromAc andAbof Theorem 6we deduce that l(x) is a distributional solution of L(h) = 0, and hence by hypothesis it results that l(x) is smooth. From (3.3) we get u ≥ l which implies that l ∈ X. From the Hypothesis 2., we have that l(x) = l is constant, that is the claim.

4. ⇒ 1. Let u ∈ C2(RN) ∩ X. Assume that u can be represented as in (3.11), we have to prove that −L(u) = ν and inf u = m. From Lemma 25 we have that u is a solution of −L(u) = ν. By hypothesis 4., the ring condition is satisfied with a constant function cx = c.

Therefore, the remaining claim follows from Theorem6.

We prove the converse implication. Let u ∈ C2(RN) ∩ X be such that −L(u) = ν ≥ 0 and inf u = m ∈ R: we have to prove that (3.11) holds. Since u is bounded from below, from hypothesis 4. for u the ring condition (3.2) holds with a constant function. The conclusion follows applying Theorem6.

Remark 32. We point out that the proof of Theorem14depends on the validity of Lemma

21. Consequently a version of Theorem 14 for distributional solutions u ∈ L1

loc(RN) holds

provided (3.25) and (3.26) are satisfied in the L1

loc(RN) setting and −L(u) ≥ 0. We do not

know if this is possible in the general framework of section 2.

In the context of homogeneous Carnot group this is possible as we shall see in the following section 4.

3.6

Additional remarks on the ring conditions

In this section we show some cases when the ring conditions are satisfied. 3.6.1 Ring condition and semilinear problems

In this section we investigate the connection between the ring condition and the solution of the inequality

(24)

As in Section 3.6.1, here we do not make any assumption on the sign of the solution u and from the universal esimate in Section3.6.1 we derive that, in some cases the ring conditions (3.2) and (3.48) hold.

For several reasons that will be clarified at the end of this section, we shall deal with a more general inequalities involving a bounded coefficent a (see (3.65) below).

In order to formulate the results, we shall use the same notation of Remark 20, that is, fix x ∈ RN and choose a real number Dx > 2. Next we define gx(y) := Γ

1 2−Dx

x (y). The

function gx : RN → R is continuous, positive and smooth in RN \ {x} and gx(x) = 0. We

set h2

x := (∇gx· A∇gx).

Theorem 33. Let q > 1, a ∈ L∞(RN) and x ∈ RN. There exists a constant C = C(x) > 0

such that if u ∈ C2(RN) is a solution of

− L(au) ≥ h2 x|u|

q

on RN, (3.66)

then for any R > 0 we have, Z ΩR(x) h2x|u|qdy ≤ ||a||∞CR Dx−2q0 Dx−2 , (3.67) where C = C1(q)(Dx− 2) 2

q−1. Moreover the following chain of inequalities hold,

Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx |au| dy ≤ (3.68) c Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx |u| dy ≤ (3.69) c Z Ω2R(x)\ΩR(x) (∇Γx· A∇Γx) Γx |u|qdy 1/q ≤ CR−q−12 1 Dx−2. (3.70)

Therefore, if au is smooth, we have a(x)u(x) ≥

Z

RN

Γ(x, y)hx(y) |u(y)|qdy. (3.71)

In particular, if

hx is bounded on RN \ ΩS(x) for a large S, (3.72)

and for a positive constant c > 0, u solves also

− L(u) ≥ c |u|q on RN, (3.73) then Z ΩR(x) |u|q dy ≤ CRDx−2q 0 Dx−2 , f or R > S. (3.74)

Proof. From Theorem28, with ν = h2x|u|q, we have Z Ω2R(x) h2x|u|qφRdy ≤ 8M R||a||∞ Z Ω2R(x)\ΩR(x) |u| φ1/qR (∇Γ · A∇Γ) Γ dy = cR Z Ω2R\ΩR(x) |u| φ1/qR h2xΓDx−2Dx dy.

(25)

Using Hölder inequality we obtain Z Ω2R h2x|u|qφR ≤ cR Z Ω2R(x)\ΩR h2x|u|qφR 1/qZ Ω2R(x)\ΩR h2xΓDx−2Dx q 01/q 0 dy which implies, Z ΩR(x) h2x|u|q dy ≤ Z Ω2R(x) h2x|u|qφRdy ≤ cRq 0Z Ω2R(x)\ΩR(x) h2xΓDx−2Dx Γ Dx Dx−2(q 0−1) dy ≤ cRDx−2q 0 Dx−2 Z Ω2R(x)\ΩR(x) ΓDx−2Dx h2 xdy ≤ cR Dx−2q0 Dx−2 .

Here we have used the fact that if y ∈ Ω2R(x) \ ΩR(x) then Γ(y) ≤ R1 and (3.20) holds.

The chain of inequalities follow directly from Hölder inequalities, (3.20) and (3.67). A slight modification of the above argument proves the estimate (3.74). 2 An immediate application of above result yields several Liouville type theorems for semi-linear elliptic inequalities as well as qualitative information on the solutions. We refer the interested reader to [13].

Here for sake of brevity, we point out a simple example.

First of all from (3.71) we immediately get the information that if u ∈ C2(RN) is non trivial solution of (3.66) with a ∈ L∞(RN) ∩ C2

(RN) then a(x)u(x) > 0. Therefore if

the weight a vanishes in one point x then (3.66) has no nontrivial solution (without any assumption on the sign).

Looking at “coercive” inequality

L(u) ≥ h2x|u|q on RN,

from (3.71) it follows that either u(x) < 0 or u ≡ 0. Hence we can state the following. Theorem 34. Let q > 1 and u ∈C2(RN) be a solution of the equation

L(u) = |u|q−1u on RN. (3.75) If for some point x ∈ RN there exists D

x > 2 such that the function h2x (as defined at the

beginning of this section and in Remark 20) satisfies (3.72), then u ≡ 0. Proof. By computation we have that u2 satisfies

L(u2) = 2(A∇u · ∇u) + 2uL(u) = 2(A∇u · ∇u) + 2 |u|q+1 ≥ 2 |u|q+1 ≥ ch2 x|u|

q+1

. Therefore, as observed before, u2(x) < 0 or u2 ≡ 0, which concludes the proof. 2.

The above results when L = ∆ in the Euclidean space and for distributional solution has been proved by Brezis in [6]. While for distributional solutions in the Carnot group setting and for sub-Laplacians see Corollary 4.5 of the author’s paper [12]. See also [11] for some quasilinear generalisations.

3.6.2 Other assumptions

In this section we shall present some condition on u that assure the validity of the ring conditions (3.2) and (3.48) with lx = 0.

(26)

We recall that for a given positive measure µ on an open set U and q > 1 the weak-Lq

space is set Lq

w(U, µ) of the functions f ∈ L1loc(U ) such that

sup µ(A)<∞ (µ(A))−1/q0 Z A |f (y)| dµ(y) < +∞, here A is a measurable set contained in U.

Theorem 35. Let x ∈ RN be fixed. In what follows µ

x shall denote the measure h2x(y) dy

or if (3.72) holds µx stands for the Lebesgue measure.

Assume that u ∈ Lqw(RN\ ΩS(x), µx) for S large and q > 1. Then u satisfies (3.48) with

lx = 0.

In addition, if u ∈C2(RN) and −L(u) ≥ 0 with 1 < q < Dx/(Dx− 2) then u ≡ 0.

The threshold Dx/(Dx− 2) is sharp. Indeed in the Euclidean setting there exist positive

superharmonic function in Lq

w(RN) with q ≥ N/(N − 2).

Before proving the theorem, we notice that from Lemma 23 we have that

µx(Ω2R(x) \ ΩR(x)) ≈ R

Dx

Dx−2 f or R > 0.

Proof. Set AR := Ω2R(x) \ ΩR(x). Taking into account that u ∈ Lqw(RN \ ΩS(x), µx) and

the estimate on µx(AR), we have

Z AR (∇Γx· A∇Γx) Γx |u| dy ≈ 1 RDx−2Dx Z AR |u| dµ ≤ c 1 RDx−2Dx µx(AR)1/q 0 ≤ cR−1q Dx Dx−2. (3.76)

Letting R → +∞, we get the claim.

The second part of the theorem follows from Theorem 6. Indeed u can be represented as in (3.3) with lx = 0. If 1 < q < Dx/(Dx− 2), then from (3.76) it follows that,

lim R→+∞R Z AR (∇Γx· A∇Γx) Γx |u| dy = 0. This contradicts (3.5). 2

4

Applications in the Carnot groups setting

In the setting of Carnot groups the previous results can be extended to distributional solu-tions. The main ingredient for this goal is a regularisation procedure by a family of mollifiers that commutes with the sub-Laplacian. See [5].

Therefore all the result in previous section can be restated for distributional solutions and the representation formula hold almost everywhere. For the convenience of the reader we shall rewrite them here. We state the results in terms of the balls BN2 generated by the

gauge N2 instead of Ωr(x).

Lemma 36. Let u ∈ L1loc(RN) be weakly superharmonic and let ν be its Radon measure,

that is −∆Gu = ν in distributional sense. Then, for any β 6= 0, R > 0, 0 < δ < 1, and a.e.

x ∈ RN and a.e. r > 0, the identities (3.32), (3.33), (3.34) in Remark 22 hold by replacing Dx = Q, gx(y) = N2(x−1◦ y), and h2x = φ2x.

The first representation formula for distributional solutions in the Carnot group setting can be stated as follows. This is the analogue of theorem 6.

(27)

Theorem 37. Let u ∈ L1

loc(RN) be such that −∆Gu = ν ≥ 0. Let x ∈ RN and lx ∈ R be

such that lim inf R→+∞ Z BN2(x,2R)\BN2(x,R) ψ2 0 N2Q(x −1◦ y))(u(y) − l x)dy = 0. (4.1)

Then if x is a Lebesgue point for u we have u(x) = lx+

Z

RN

dν(y)

N2Q−2(x−1◦ y). (4.2)

In particular, if (4.1) holds for a.e. x ∈ RN, and l

x = l does not depend on x, then

l = ess infRNu ∈ R and

u(x) = l + Z

RN

Γ(x, y)ν(y)dy. a.e. x ∈ RN. (4.3) Similarly Theorem 14, can be formulated as follows.

Theorem 38. The following statements are equivalent.

1. Let u ∈ L1loc(RN) be such that −∆Gu = ν ≥ 0 and ess infRNu = l ∈ R. Then

u(x) = l + Z RN dν(y) N2Q−2(x−1◦ y) a.e.x ∈ R N. (4.4)

2. Let u ∈ L1loc(RN). If ess infRNu = l ∈ R and −∆Gu = 0 then u ≡ l a.e. in RN.

3. Let u ∈ L1

loc(RN) be such that −∆Gu = ν ≥ 0. Then, ess infRNu = l ∈ R if and only if

lim inf R→+∞ Z − BN2(x,2R)\BN2(x,R) ψ2x(u − l)dy = 0 a.e. x ∈ RN (4.5) if and only if lim R→+∞ Z − BN2(x,2R)\BN2(x,R) ψx2|u − l|dy = 0 a.e. x ∈ RN. (4.6)

Remark 39. In the Carnot group setting a classical Liouville theorem has been proved in [4] for distributional solutions. See also [5]. Therefore the representation formula holds and the infimum of superharmonic function can be characterised by (4.5) and (4.6).

In the same framework a proof of the representation formula based on harmonic analysis argument is contained in the book [5] under the assumption that u is bounded from below. Notice that in [15] the authors prove the first implication of Theorem 37. That is if u ∈ L1

loc(RN) is a distributional solution of −∆Gu = f ≥ 0 and satisfies (4.6) then

u(x) = l + Z

RN

Γ(x, y)f (y)dy,

a.e. on RN. Finally we want to point out that if u ∈ Lqw(RN) for some q > 1 then u satisfies (4.6) with l = 0.

Acknowledgement

We warmly thank Professor Ermanno Lanconelli for useful comments and remarks, and for pointing out to us the reference [1].

(28)

References

[1] B. Abbondanza, Potential Analysis for Hypoelliptic Second Order PDEs with Non-negative Characteristic Form, Tesi di Dottorato, Università di Bologna, (2015), 1–129. [2] W.Bauer, K. Furutani, C. Iwasaki , Fundamental solution of a higher step Grushin

type operator, Adv. Math. 271 (2015), 188–234.

[3] S. Biagi, A. Bonfiglioli, The existence of a global fundamental solution for homo-geneous Hörmander operators via a global lifting method, Proc. London Math. Soc. 114 (2017) 855–889.

[4] A. Bonfiglioli, E. Lanconelli, Liouville-type theorems for real sub-Laplacians, Manuscripta Math. 105 (2001), no. 1, 111–124.

[5] A. Bonfiglioli, E. Lanconelli, F. Uguzzoni, Stratified Lie Groups and Potential Theory for their sub-Laplacians, Springer, 2007.

[6] H. Brezis, Semilinear equations in Rn without condition at infinity, Appl. Math. Op-timization, 12 (1984), 271–282.

[7] G. Caristi, L. D’Ambrosio and E. Mitidieri, Liouville Theorems for Some Non-linear Inequalities, Proc. Steklov Inst. Math., 260, (2008) 90-111.

[8] G. Caristi, L. D’Ambrosio and E. Mitidieri, Liouville theorems and represen-tation formulae for higher order problems and systems and related Liouville theorems, Milan J. Math. 76 (2008), 27–67.

[9] G. Citti, N. Garofalo and E. Lanconelli, Harnack’s inequality for sum of squares of vector fields plus a potential, Amer. J. Math., 115, (1993), 699–734.

[10] L. D’Ambrosio and E. Mitidieri, Nonnegative solutions of some quasilinear elliptic inequalities and applications, SB MATH, 201(6), (2010) 855-871.

[11] L. D’Ambrosio and E. Mitidieri, A priori estimates and reduction principles for quasilinear elliptic problems and applications, Adv. Diff.Eq., 17, 9-10, (2012) 935–1000. [12] L. D’Ambrosio and E. Mitidieri, Uniqueness and comparison principles for semi-linear equations and inequalities in Carnot groups, Adv. Nonsemi-linear Anal. 7(3), (2018) 313–325.

[13] L. D’Ambrosio and E. Mitidieri, Liouville Theorems of some second order elliptic inequalities, Preprint, (2018) 1-40.

[14] L. D’Ambrosio and E. Mitidieri, A priori estimates, positivity results, and nonex-istence theorems for quasilinear degenerate elliptic inequalities Adv. Math., 224 (2010) 967-1020.

[15] L. D’Ambrosio, E. Mitidieri and S. I. Pohozaev, Representation formulae and inequalities for solutions of a class of second order partial differential equations, Trans. Amer. Math. Soc., 358 (2006), 893-910.

[16] G.B. Folland, Subelliptic estimates and function spaces on nilpotent Lie groups, Arkiv gör Mat. 13 (1975), 161–207.

[17] G.B. Folland, E.M. Stein, Hardy spaces on Homogeneous Groups, Math. Notes, vol. 28, Princeton Univ. Press, Princeton, 1982.

[18] P.C. Greiner, A fundamental solution for a nonelliptic partial differential operator, Can. J. Math. 31 (1979), 1107–1120.

(29)

[19] J. Heinonen, Calculus on Carnot Groups, in Fall School in Analysis (Jyväskylä, 1994), vol. 68, Ber. Univ. Jyväskylä Math. Inst. Jyväskylä, 1995, 1–31.

[20] E. H. Lieb and M. Loss, Analysis, Graduate Studies in Mathematics Vol. 14, AMS, 2001.

[21] E. Mitidieri and S.I. Pohozaev, A priori estimates and the absence of solutions of nonlinear partial differential equations and inequalities, Proc. Steklov Inst. Math. 234 (2001), 1–362.

[22] E. Mitidieri and S.I. Pohozaev, Positivity property of solutions of some elliptic inequalities on Rn, Doklady Math. 68 (2003), 339–344.

[23] Sánchez-Calle, Antonio, Fundamental solutions and geometry of the sum of squares of vector fields, Invent. Math. 78 (1984), 143–160.

(1)

Lorenzo D’Ambrosio Dipartimento di Matematica Università degli Studi di Bari via E.Orabona, 4, I-70125 Bari e-mail: dambros@dm.uniba.it

(2)

Enzo Mitidieri

Dipartimento di Matematica e Geoscienze Università degli Studi di Trieste

via A.Valerio, 12/1, I-34127 Trieste e-mail: mitidier@units.it

Riferimenti

Documenti correlati

Therefore, in this work, the attention is focused on Ag 73.2 Cu 17.1 Zr 9.7 (at.%) amorphous/crystalline composite, by studying the effect of cold rolling on the stability of

Inoltre, merita forse qualche attenzione in più, nella direzione della ricerca delle possibili interferenze tra il pisano e il sardo, nell’ambito della morfologia

The aim of the present paper is to provide multiplicity results regarding homoclinic and heteroclinic solutions for equation (.) under particular assumptions on the weight

The existence, uniqueness and other properties of solutions of nonlinear second order Volterra integrodifferential equation in a general Banach space are studied.. The main tools in

We prove stability of the unit ball for the free energy F in the family of nearly-spherical sets, getting as a corollary Theorem 1.1.9 for nearly-spherical sets.. For the sake of

Using the weighted residual formulation we derive a-posteriori estimates for Discontinuous Galerkin approximations of second order elliptic problems in mixed form... In recent

[r]