Home Uniqueness and comparison principles for semilinear equations and inequalities in Carnot groups
Article Open Access

Uniqueness and comparison principles for semilinear equations and inequalities in Carnot groups

  • Lorenzo D’Ambrosio ORCID logo and Enzo Mitidieri EMAIL logo
Published/Copyright: September 15, 2017

Abstract

Variants of the Kato inequality are proved for distributional solutions of semilinear equations and inequalities on Carnot groups. Various applications to uniqueness, comparison of solutions and Liouville theorems are presented.

1 Introduction

It is well known that one of the fundamental tools for studying different questions related to coercive elliptic equations and inequalities on N is the so-called Kato inequality [14].

One of the earlier and main contributions in this direction has been proved by Brezis [3]. As a consequence of a modified Kato inequality he considered, among other things, distributional solutions of elliptic inequalities of the form

(1.1)Δu|u|q-1uon N,

where q>1. The main conclusion of Brezis is that if uLlocq(N) solves (1.1), then

u(x)0a.e. on N.

A number of important results can be deduced from this simple statement (see [3] for details).

Quasilinear versions of the Kato inequality have been studied recently in [8], where general a-priori estimates and Liouville theorems have been proved for weak solutions of coercive quasilinear elliptic equations and inequalities in divergence form; see also [1, 10, 11, 5, 16, 6] for related results.

The goal of this paper is to prove a modified version of the Kato inequality (see (3.1) below) for distributional solutions for a Laplacian operator on a Carnot group; see [2].

It should be noted that a similar Kato inequality has been proved in [8] for weak solutions, i.e., Wloc1,2 solutions. We point out, see Remark 3.5 below, that a Kato inequality for distributional solutions cannot be deduced from the corresponding inequality valid for weak solutions even in the standard Euclidean framework; see [8, Theorem 2.1].

This paper is organized as follows: Section 2 contains some preliminary material on Carnot groups. In Section 3, we prove one of the main results of this paper (see Theorem 3.2) and discuss its relation with the results proved in [8]. In Section 4, we prove some uniqueness results for a general semilinear second-order inequality and give some concrete applications. In Section 5, we shall briefly discuss the ideas pointed out in the preceding section to systems of semilinear inequalities; see [9] for other applications of Kato inequalities to semilinear elliptic systems. Finally, in Section 6 we prove a modified version of Kato complex inequalities in the setting of Carnot groups and present some applications to the so-called reduction principles and to uniqueness of solutions of complex problems; see [6].

2 Preliminaries on Carnot groups

In this section, we recall some preliminary facts concerning Carnot groups (for more information and proofs we refer the interested reader to [2, 12]).

A Carnot group is a connected, simply connected, nilpotent Lie group 𝔾 of dimension N2 with graded Lie algebra 𝒢=V1Vr 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 a homogeneous Carnot group(N,,δλ) defined as follows: We identify 𝔾 with N endowed with a Lie group law . We consider N split into r subspaces N=n1×n2××nr with n1+n2++nr=N and ξ=(ξ(1),,ξ(r)) with ξ(i)ni. 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 (N,) is 𝒢. For i=1,,n1=l, let Xi be the unique vector field in 𝒢 that coincides with /ξi(1) at the origin. We require that the Lie algebra generated by X1,,Xn1 is the whole 𝒢.

With the above hypotheses, we call 𝔾=(N,,δλ) a homogeneous Carnot group. The canonical sub-Laplacian on 𝔾 is the second-order differential operator =i=1lXi2. Now, let Y1,,Yl be a basis of span{X1,,Xl}; the second-order differential operator

ΔG=i=1lYi2

is called a sub-Laplacian on 𝔾. We denote by Q=i=1rini the homogeneous dimension of 𝔾. In the sequel, we assume Q3.

A nonnegative continuous function S:N+ is called a homogeneous norm on 𝔾 in the case that S(ξ)=0 if and only if ξ=0 and it is homogeneous of degree 1 with respect to δλ (i.e., S(δλ(ξ))=λS(ξ)). We say that a homogeneous norm is symmetric if S(ξ-1)=S(ξ).

The Lebesgue measure is the bi-invariant Haar measure. For any measurable set EN, we have |δλ(E)|=λQ|E|. Since Y1,,Yl generate the whole 𝒢, any sub-Laplacian ΔG satisfies the Hörmander hypoellipticity condition. Moreover, the vector fields Y1,,Yl are homogeneous of degree 1 with respect to δλ.

In what follows, we fix the vector fields Y1,,Yl. In this setting, we use the symbol 0 to denote the vector field (Y1,,Yl), and -div0:=0*, where 0* is the formal adjoint of 0. Finally, we set

Wloc1,2:={uLloc2:|0u|Lloc2}.

3 Kato’s inequality for a sub-Laplacian operator on 𝔾

In this section, we shall prove that a modified version of the Kato inequality for distributional solutions holds for a sub-Laplacian operator on a Carnot group 𝔾.

Similar inequalities can be proved for more general classes of linear differential operators. For instance, one can handle second-order operators generated by a system of smooth vector fields in N satisfying the Hörmander condition, and left invariant differential operators on homogeneous groups; see [12]. However, we shall not discuss these kinds of generalizations here.

As usual, we denote by sign, sign+ and u+ the functions defined by

sign(t):={1if t>0,0if t=0,-1if t<0,
sign+(t):={1if t>0,0if t0,
u+:=sign+(u)u.

Throughout this paper, ΩN denotes an open subset.

Definition 3.1.

Let fLloc1(Ω). A distributional solution of the inequality

ΔGufin 𝒟(Ω)

is a function uLloc1(Ω) such that for any nonnegative ϕ𝒞02(Ω) we have that

ΩuΔGϕΩfϕ.

Theorem 3.2 (Kato inequality).

Let u,fLloc1(Ω) be such that

ΔGufin 𝒟(Ω).

Then

(3.1)ΔGu+sign+(u)fin 𝒟(Ω)

and

(3.2)ΔG|u|sign(u)fin 𝒟(Ω).

The proof is a consequence of the following lemma; see [1] for a related result.

Lemma 3.3.

Let γC2(R) be a convex function with bounded first derivative. Let u,fLloc1(Ω) be such that

ΔGufin 𝒟(Ω).

Then γ(u)Lloc1(Ω) and

ΔGγ(u)γ(u)fin 𝒟(Ω).

Proof.

We need to prove that for any nonnegative ϕ𝒞02(Ω) we have

γ(u)ϕL1(Ω),

and that the following inequality holds:

Ωγ(u)ΔGϕΩγ(u)fϕ.

Fix ϕ𝒞02(Ω). Let (mη)η be a family of symmetric mollifiers associated to a fixed homogeneous norm S. Set uη:=u𝔾mη in Ωη:={xΩ:dist(x,Ω)>η}, that is,

uη(x):=Ωu(y)mη(xy-1)𝑑y=Ωu(y-1x)mη(y)𝑑y,xΩη.

For η small enough, it follows that supp(ϕ)Ωη.

Let xΩη. Since mη(x-1) is a nonnegative test function in Ω, by using the Fubini–Tonelli theorem we obtain

Ωuη(x)ΔGϕ(x)𝑑x=ΩΔGϕ(x)(Ωu(y-1x)mη(y)𝑑y)𝑑x
=Ωmη(y)(Ωu(y-1x)ΔGϕ(x)𝑑x)𝑑y
=Ωmη(y)(Ωu(z)(ΔGϕ)(yz)𝑑z)𝑑y
=Ωmη(y)(Ωu(z)ΔG(zϕ(yz))dz)dy
Ωmη(y)(Ωf(z)ϕ(yz)𝑑z)𝑑y
=Ωmη(y)(Ωf(y-1x)ϕ(x)𝑑x)𝑑y
=Ωϕ(x)(Ωf(y-1x)mη(y)𝑑y)𝑑x
=Ωϕ(x)fη(x)𝑑x,

that is,[1]

ΔGuη(x)fη(x)on Ωη.

On the other hand, by the convexity of γ it follows that

ΔGγ(uη)=γ(uη)ΔGuη+γ′′(uη)|Lu|2γ(uη)ΔGuη,

which implies

Ωγ(uη)ΔGϕΩγ(uη)ΔGuηϕΩγ(uη)fηϕ.

The convergence of γ(uη)γ(u) in Lloc1(Ω) is assured by the convergence of uηu in Lloc1(Ω) and the fact that γ is a Lipschitz function (since γ is bounded). By observing that

Ωγ(uη)(x)fη(x)ϕ(x)𝑑x=ΩΩγ(uη)(x)ϕ(x)mη(xy-1)f(y)𝑑y𝑑x
=Ωmη𝔾(γ(uη)ϕ)(y)f(y)𝑑y,

it suffices to prove that

Ωmη𝔾(γ(uη)ϕ)(y)f(y)Ωγ(u)ϕf.

To this end, we first claim that

(3.3)mη𝔾(γ(uη)ϕ)γ(u)ϕin L1(Ω).

Indeed, since γ is continuous and uηu a.e. in Ω (if necessary by passing to a subsequence), it follows that γ(uη)ϕγ(u)ϕ a.e. in Ω. Now, by γ being bounded, an application of the Lebesgue dominated convergence gives γ(uη)ϕγ(u)ϕ in L1(Ω). Moreover,

|Ωmη𝔾(γ(uη)ϕ)(y)-γ(u)ϕ|=|Ωmη𝔾(γ(uη)ϕ)(y)-mη𝔾γ(u)ϕ+mη𝔾γ(u)ϕ-γ(u)ϕ|
|Ωmη𝔾(γ(uη)ϕ)(y)-mη𝔾γ(u)ϕ|+|Ωmη𝔾γ(u)ϕ-γ(u)ϕ|
(Ωmη)|Ω(γ(uη)ϕ)-γ(u)ϕ|+|Ωmη𝔾γ(u)ϕ-γ(u)ϕ|0.

Next, if necessary by passing to a subsequence, we may suppose that the convergence in (3.3) is a.e. on Ω.

Now, since γ(uη)ϕ is uniformly bounded by M:=|γ||ϕ|, we deduce that

|mη𝔾(γ(uη)ϕ)|ΩmηMM.

Noticing that mη𝔾(γ(uη)ϕ) has compact support contained in suptϕ+Bηsuptϕ+B1=:K, it follows that

|mη𝔾(γ(uη)ϕ)|fMfχKLloc1(Ω).

Finally, by the Lebesgue theorem we have

Ωγ(uη)fηϕ=Ωmη𝔾(γ(uη)ϕ)(y)f(y)𝑑yΩγ(u)fϕ𝑑y.

This completes the proof. ∎

Proof of Theorem 3.2.

The idea is first to approximate the function sign+ with a family of convex functions γϵ having bounded derivatives, and then apply Lemma 3.3 above.

Let m𝒞() be nonnegative with supt(m)[-1,1] and m=1. For ϵ>0, set mϵ:=1ϵm(t-ϵϵ) and consider γϵ as the solution of the problem

γϵ′′=mϵwithγϵ(0)=γϵ(0)=0.

Clearly, we have γϵ(t)=γϵ(t)=0 for t0. In addition, γϵ(t)=1 for t>2ϵ and 0γϵ(t)t+, 0γϵ(t)1. This implies the pointwise convergence, as ϵ0, of γϵ(t)t+ and γϵ(t)sign+t. Finally, by Lemma 3.3 we have

Ωγϵ(u)ΔGϕΩγϵ(u)fϕ,

and by the Lebesgue theorem we obtain

Ωu+ΔGϕΩsign+ufϕ.

The proof of (3.2) follows from a similar argument as above, so we shall omit it. ∎

Remark 3.4.

Theorem 3.2 holds if we replace the functions sign+ and u+ respectively with

signh+(t):={1if t>h,0if th,

and uh+:=(u-h)+, where h. To this end, we can argue as in the proof of Theorem 3.2, replacing γϵ(t) by γϵ(t-h).

Remark 3.5.

Theorem 3.2 deals with Lloc1(Ω) solutions of the inequality

ΔGufin Ω,

while [8, Theorem 2.1] allows to consider Wloc1,2(Ω) solutions.

One may try to prove (3.1) by mollifying the solution and then applying [8, Theorem 2.1]. In this case, one would obtain

ΔGuη+sign+(uη)fηin Ω.

Clearly, in order to prove (3.1) we need to know that

sign+(uη)sign+(u)

at least a.e. This is not always possible. Indeed, we can construct a function u (even continuous) such that each mollification uη has sign+(uη)1, while sign+(u)1. We shall prove this when Ω=]0,1[.

Let {qn}n1 be the set of rational numbers contained in ]0,1[. Fix 1>ϵ>0 and set

In:=]qn-ϵ2-n,qn+ϵ2-n[]0,1[,I:=n1In,S:=[0,1]I.

The set I is open and dense in [0,1]. Moreover, 0<|I|ϵ<1, thus |S|>0.

Next, for each n1 let ϕn:[0,1] be a continuous nonnegative function such that ϕn(x)>0 if and only if xIn and ϕn1. Set

u:=n1ϕn2-n.

Since the above series is uniformly convergent, the function u is continuous. Moreover, u(x)>0 if and only if there exists n1 such that ϕn(x)>0. This is obviously equivalent to the fact that xIn. In other words, u vanishes on S and it is positive on I.

Let η>0 and let uη be a mollification of u, that is, umη, where (mη)η is a standard family of mollifiers. We claim that uη(x0)>0 for any x0]0,1[. Indeed, let x0]0,1[. By our choice of {qn}n there exists n1 such that |qn-x0|<η. Hence In]x0-η,x0+η[ and

uη(x0)=u(y)mη(x0-y)𝑑yIn|y-x0|<ηϕn(y)2-nmη(x0-y)𝑑y>0.

4 Applications to uniqueness of solutions

In this section, we consider weakly elliptic linear differential operators of the form

Lu:=div(B(x)u)=divL(Lu),

and the associated uniqueness problem for the semilinear equation

Lu=f(u)+hon Ω.

Notice that since Lu=div(B(x)u), where B is a positive semidefinite matrix, by writing B as B=μTμ and defining divL=div(μT) and L=μ, it follows that

Lu=div(μTμu)=divL(Lu).

This means that a Kato inequality holds for L; see [8].

Definition 4.1.

Let f𝒞() and hLloc1(Ω). A weak solution of

(4.1)Luf(u)+hon Ω,

is a function

uWL,loc1,2(Ω):={uLloc2:|Lu|Lloc2}

with f(u)Lloc1(Ω), such that for any nonnegative ϕ𝒞01(Ω) we have

-ΩLuLϕΩ(f(u)+h)ϕ.

If L=ΔG is a sub-Laplacian on a Carnot group, then a distributional solution of (4.1) is a function uLloc1(Ω) such that f(u)Lloc1(Ω), and for any nonnegative ϕ𝒞02(Ω) we have

ΩuΔGϕΩ(f(u)+h)ϕ.

Theorem 4.2.

Let X be a subspace of Lloc1(Ω) such that if uX, then u+X. Let b:[0,[[0,+[ be a continuous function such that b(0)=0 and the problem

(4.2)Lvb(v)[ΔGub(v)],v0,on Ω,

has no nontrivial weak [distributional] solution belonging to X.

Let hLloc1(Ω) and let fC(R) be such that

f(t)-f(s)b(t-s)for any t>s.

Then the equation

(4.3)Lv=f(v)+h[ΔGv=f(v)+h]on Ω

has at most one weak [distributional] solution belonging to X.

Proof.

Let hLloc1(Ω) and let u,vX be solutions of (4.3). The function u-vX is a weak solution of

L(u-v)=f(u)-f(v)on Ω.

An application of the appropriate Kato inequality (3.1) or [8, Theorem 2.1] yields

L((u-v)+)sign+(u-v)(f(u)-f(v))on Ω,

which in turn implies that the function w:=(u-v)+ is a weak (or distributional) solution of

Lwsign+(u-v)(f(u)-f(v))sign+(u-v)b(u-v)=b(w)on Ω.

In other words, w solves (4.2). Hence w0 a.e. on Ω, that is, uv a.e. on Ω. Inverting the role of u and v, the claim follows. ∎

A concrete application of Theorem 4.2 is contained in the following result.

Theorem 4.3.

Let fC(R) be such that

(4.4)f(t)-f(s)b(t-s)for any t>s,

where b:[0,+[[0,+[ is a continuous function satisfying the following assumptions:

  1. b(0)=0, b(t)>0 for t>0;

  2. it holds that

    (4.5)1+(1tb(s)𝑑s)-12𝑑t<+;
  3. b is convex.

Let hLloc1(RN). Then the problem

ΔGu=f(u)+hin 𝒟(N)

has at most one distributional solution uLloc1(RN). Moreover, if h0, then u0 a.e. on RN.

Proof.

The obvious idea is to apply Theorem 4.2. To this end, it is enough to check that the inequality

(4.6)ΔGvb(v),v0,in 𝒟(N)

has only the trivial solution. Indeed, let us assume that vLloc1(N) is a solution of (4.6). By a mollification argument (as in the proof of Lemma 3.3) we have

ΔG(vη)(b(v))η.

Next, by the convexity of b and the Jensen inequality, it follows that

(4.7)ΔG(vη)b(vη),vη0,on N.

Now vη is smooth and solves (4.7) with the function b nondecreasing (indeed, it satisfies (i) and it is convex) and satisfying (4.5), thus we are in the position to apply [7, Theorem 3.10] (by changing u:=-vη), so we deduce that vη0. Thus, by letting η0 we obtain v0. ∎

Remark 4.4.

When dealing with 𝒞1 solutions, hypothesis (iii) can be relaxed by assuming that b is nonincreasing; see [7].

Corollary 4.5.

Let q>1 and let hLloc1(RN). The problem

ΔGu=|u|q-1u+hin 𝒟(N)

has at most one solution uLlocq(RN). Moreover, if h0, then u0 a.e. on RN.

Remark 4.6.

The above result, as far it is concerned with uniqueness and nonpositivity of the possible solutions, is the analog on Carnot groups of [3, Theorem 2].

Remark 4.7.

All the above results still hold when one replaces the function hLloc1 with a distribution h𝒟.

Theorem 4.3 allows us to generalize Corollary 4.5 to a more general class of nonlinearities, as the following example shows.

Example 4.8.

Let f be defined by

f(t):={tq1if t0,-|t|q2if t<0,

where q1,q2>1. Theorem 4.3 applies to such f. Indeed, for t0 define g(t):=min{tq1,tq2}. The function b that we need is the convexification of cg for a small constant c>0.

We claim that there exists a constant c>0 such that for any t>s we have

f(t)-f(s)cg(t-s).

Assume that q1q2. By the well-known inequality

tp-spcp(t-s)pfor t>s and p>1,

we have the following three cases:

  1. Let t>s>0. Then

    f(t)-f(s)=tq1-sq1cq1(t-s)q1cq1g(t-s).
  2. Let 0>t>s. Then

    f(t)-f(s)=-|t|q2+|s|q2cq2(|s|-|t|)q2cq2g(|s|-|t|)=cq2g(t-s).
  3. Let t>0>-s. The proof of the claim will follow if we prove that

    tq1+sq2cg(t+s)for any t,s>0.

By using the inequality

ap+bp21-p(a+b)pfor a,b>0 and p>1

and distinguishing three different cases, we have the following:

  1. Let s1 and t>0. Then

    tq1+sq2tq1+sq121-q1(t+s)q121-q1g(t+s).
  2. Let 1>t>0 and 1>s>0. Then

    tq1+sq2tq2+sq221-q2(t+s)q221-q2g(t+s).
  3. Let t1 and 1>s>0. Then

    tq1+sq2tq12-q1(t+1)q12-q1(t+s)q12-q1g(t+s).

Next, by choosing

c:=min{cq1,cq2,21-q1,21-q2,2-q1},

we get the claim.

By defining b:=conv(cg), it follows that assumptions (4.4) and Theorem 4.3 (i) and (iii) are fulfilled. Notice that Theorem 4.3 (ii) is satisfied since at infinity the function b behaves like tq1 with q1>1.

We point out that f does not satisfy the Brezis condition f(t)|t|q-1 for any t unless q1=q2. The interested reader may compare this with [3].

5 Some applications to a class of semilinear systems

In this section, as in the previous Section 4, we consider weakly elliptic linear differential operators of the form Lu=divL(Lu). We refer to Definition 4.1 for the appropriate notion of solutions.

Theorem 5.1.

Let X be a subspace of Lloc1(Ω) such that if uX, then u+X. Let b:[0,[[0,+[ be a continuous function such that b(0)=0 and the problem

(5.1)Lvb(v)[ΔGvb(v)],v0 on Ω

has no nontrivial weak [distributional] solutions belonging to X. Let fC(R) be such that

(5.2)f(t)+f(s)b(t+s)for any t>-s.

Let (u,v)X×X be a weak [distributional] solution of the system of inequalities

(5.3){Lvf(u),Luf(v)[{ΔGvf(u),ΔGuf(v)]on Ω.

Then the following assertions hold:

  1. u+v0 a.e. on Ω.

  2. Let C1 and assume that the function f¯(t):=-Cf(-t) satisfies (5.2). Let (u,v)X×X be a weak [distributional] solution of the system

    (5.4){Cf(u)Lvf(u),Cf(v)Luf(v)[{Cf(u)ΔGvf(u),Cf(v)ΔGuf(v)]on Ω.

    Then u=-v a.e. on Ω . Therefore, u satisfies

    (5.5)Cf(u)-Luf(u)[Cf(u)-ΔGuf(u)]  on Ω,

    and the function f must be odd on the range of u, that is, for any tu(Ω) the condition f(t)=-f(-t) holds.

Proof.

Let (u,v)X×X be a solution of (5.3). The function u+vX solves

L(u+v)f(v)+f(u)on Ω.

An application of the Kato inequality yields

L((u+v)+)sign+(u+v)(f(v)+f(u))on Ω,

which in turn implies that the function w:=(u+v)+ is a weak solution of

Lwsign+(u+v)(f(u)+f(v))sign+(u+v)b(u+v)=b(w)on Ω,

that is, w solves (5.1). Hence w0 a.e. on Ω, that is, u+v0 a.e. on Ω. This proves case (i).

(ii) The functions u¯:=-u and v¯:=-v satisfy also the inequalities

Lu¯-Cf(v)=f¯(v¯)andLv¯f¯(u¯).

Since condition (5.2) is satisfied by f¯, from (i) we have u¯+v¯0, that is, u=-v.

From the first inequality in (5.4) it follows that u solves (5.5). Adding (5.5) and the second inequality of (5.4) (and taking into account that v=-u), we obtain

C(f(u)+f(-u))0f(u)+f(-u).

This last chain of inequalities implies that f(u)=-f(-u), completing the proof. ∎

Remark 5.2.

  1. If f is odd and (5.2) holds, then the function f¯ in statement (ii) satisfies condition (5.2) as well.

  2. If f is odd and (5.2) holds, then f is nondecreasing.

  3. If f is odd, then (5.2) is equivalent to (4.4).

A concrete application of Theorem 5.1 is given by the following result.

Theorem 5.3.

Let fC(R) satisfy (5.2), where b:[0,+[[0,+[ is a continuous function such that

  1. b(0)=0, b(t)>0 for t>0;

  2. it holds

    1+(1tb(s)𝑑s)-12𝑑t<+;
  3. b is convex.

Let (u,v) be a distributional solution of the problem

{ΔGvf(u),ΔGuf(v)in 𝒟(N).

Then the conclusions of Theorem 5.1 hold.

Proof.

It is enough to check that the inequality

ΔGwb(w),w0,in 𝒟(N),

has only the trivial solution. This follows from the proof of Theorem 4.3. ∎

Remark 5.4.

Dealing with 𝒞1 solutions, hypothesis (iii) can be weakened, assuming that b is nonincreasing.

Corollary 5.5.

Let q>1. Let (u,v) be a distributional solution of the problem

{ΔGv=|u|q-1u,ΔGu=|v|q-1vin 𝒟(N).

Then u=-v a.e. on RN and

-ΔGu=|u|q-1uin 𝒟(N).

An immediate consequence is the following corollary.

Corollary 5.6.

Let q>1. Let (u,v) be a distributional solution of the problem

{-ΔGv=|u|q-1u,-ΔGu=|v|q-1vin 𝒟(N).

Then u=v a.e. on RN.

The above results improve some theorems obtained in [4].

6 A note on the complex case

In this section, we shall prove a complex version of some results stated in Section 3 and [8] in the framework of Carnot groups. For the Euclidean case, see [14, 13].

Theorem 6.1 (Kato’s inequality: The complex case).

Let u,fLloc1(Ω;C) be such that

ΔGu=fin 𝒟(Ω).

Then

(6.1)ΔG|u|(u¯|u|f)in 𝒟(Ω).

The proof is based on the following lemma.

Lemma 6.2.

Let γC2(R2) be a convex function with bounded first derivatives. Let u,fLloc1(Ω;C) be such that

ΔGu=fin 𝒟(Ω).

Then γ(u)Lloc1(Ω) and

ΔGγ(u)(2γz(u)f),

where γz is the Wirtinger operator defined by

γz(x,y)=12(γx-iγy).

Proof.

We shall use the same notations as in the proof of Lemma 3.3. Without loss of generality, we assume that u and f are smooth (if this is not the case we can use a mollification process as in the proof of Lemma 3.3).

Let u:=s+it. By computation it follows

ΔGγ(u)=γxx|Ls|2+2γxyLsLt+γyy|Lt|2+γxΔGs+γyΔGt.

We claim that

ΔGγ(u)γxΔGs+γyΔGt.

Indeed, taking into account that γ is convex and writing α1e1:=Ls and α2e2:=Lt with unitary vectors ei and real numbers αi, we have

γxx|Ls|2+2γxyLsLt+γyy|Lt|2=γxxα12+ϵ2γxyα1α2+γyyα22+2γxyα1α2[e1e2-ϵ]
(6.2)2γxyα1α2[e1e2-ϵ],

where ϵ{1,-1}. By a suitable choice of ϵ, the right-hand side of inequality (6.2) becomes nonnegative, and we get the claim.

Since

2fγz=(ΔGs+iΔGt)(γx-iγy)=γxΔGs+γyΔGt+i(γxΔGt-γyΔGs),

we complete the proof. ∎

Proof of Theorem 6.1.

Apply Lemma 6.2 to the convex function γ(x,y):=ϵ2+x2+y2 and let ϵ0. We leave the remaining details to the interested reader. ∎

As an application of Theorem 6.1 we have the following result.

Theorem 6.3 (Reduction principle: Complex case).

Let ΩRN be an open set and let f:Ω×RR be a Caratheodory function. Let XLloc1(Ω). Assume that the problem

ΔGvf(x,v),v0,in 𝒟(Ω)X,

has no nontrivial distributional solutions. If uLloc1(Ω;C) is a complex distributional solution of

ΔGu=f(x,|u|)u|u|in 𝒟(Ω)

such that |u|X, then u0 a.e. on Ω.

Proof.

By (6.1) it follows that the function |u| is a nonnegative distributional solution of

ΔG|u|f(x,|u|)in 𝒟(Ω)X.

By assumption it follows that |u|0 a.e. on Ω. ∎

We end this section with easy consequences that follow from the proof of Theorem 4.3.

Theorem 6.4.

Let fC(R) be such that

-f(-t),f(t)b(t)>0for any t>0,

where b:[0,+[[0,+[ is a continuous convex function satisfying (4.5). If uLloc1(RN;C) is a complex distributional solution of

ΔGu=f(x,|u|)u|u|in 𝒟(N),

then u0 a.e. on RN.

Corollary 6.5.

Let q>1 and hLloc1(RN;C). Then the problem

(6.3)ΔGu=|u|q-1u+hin 𝒟(N)

has at most one distributional solution uLlocq(RN;C). Moreover, if there exists θR such that eiθhR, then eiθuR.

Proof.

Let u and v be distributional solutions of (6.3) and set w:=u-v. The function w satisfies

ΔGw=|u|q-1u-|v|q-1v.

Hence, by the Kato inequality (6.1) we have

ΔG|w|((|u|q-1u-|v|q-1v)w¯|w|).

Now, by a well-known inequality (see for example [15]) it follows that

((|u|q-1u-|v|q-1v)(u¯-v¯)|u-v|)=(|u|q-1u-|v|q-1v)(u-v)|u-v|21-q|u-v|q.

Thus the uniqueness follows from the fact that

ΔG|w|21-q|w|qw=0a.e. on N.

The second claim is a consequence of the uniqueness property. Indeed, if θ=0, that is, if h is a real function, since u and u¯ are solutions of (6.3), it follows that u=u¯. This proves the claim for θ=0. If θ0 it suffices to multiply (6.3) by eiθ and apply the uniqueness property. ∎

Funding statement: The authors are supported by the MIUR Bando PRIN 2015 2015KB9WPT_001. The first author is a member of GNAMPA. The second author acknowledges the support from FRA 2015: Equazioni differenziali: teoria qualitativa e computazionale, Università degli Studi di Trieste.

References

[1] A. Baldi, A non-existence problem for degenerate elliptic PDEs, Comm. Partial Differential Equations 25 (2000), no. 7–8, 1371–1398. 10.1080/03605300008821552Search in Google Scholar

[2] A. Bonfiglioli, E. Lanconelli and F. Uguzzoni, Stratified Lie Groups and Potential Theory for Their sub-Laplacians, Springer Monogr. Math., Springer, Berlin, 2007. Search in Google Scholar

[3] H. Brezis, Semilinear equations in 𝐑N without condition at infinity, Appl. Math. Optim. 12 (1984), no. 3, 271–282. 10.1007/BF01449045Search in Google Scholar

[4] L. D’Ambrosio, A new critical curve for a class of quasilinear elliptic systems, Nonlinear Anal. 78 (2013), 62–78. 10.1016/j.na.2012.09.015Search in Google Scholar

[5] L. D’Ambrosio, A. Farina, E. Mitidieri and J. Serrin, Comparison principles, uniqueness and symmetry results of solutions of quasilinear elliptic equations and inequalities, Nonlinear Anal. 90 (2013), 135–158. 10.1016/j.na.2013.06.004Search in Google Scholar

[6] L. D’Ambrosio and E. Mitidieri, A priori estimates, positivity results, and nonexistence theorems for quasilinear degenerate elliptic inequalities, Adv. Math. 224 (2010), no. 3, 967–1020. 10.1016/j.aim.2009.12.017Search in Google Scholar

[7] L. D’Ambrosio and E. Mitidieri, Nonnegative solutions of some quasilinear elliptic inequalities and their applications, Mat. Sb. 201 (2010), no. 6, 75–92. 10.1070/SM2010v201n06ABEH004094Search in Google Scholar

[8] L. D’Ambrosio and E. Mitidieri, A priori estimates and reduction principles for quasilinear elliptic problems and applications, Adv. Differential Equations 17 (2012), no. 9–10, 935–1000. 10.57262/ade/1355702928Search in Google Scholar

[9] L. D’Ambrosio and E. Mitidieri, Liouville theorems for elliptic systems and applications, J. Math. Anal. Appl. 413 (2014), no. 1, 121–138. 10.1016/j.jmaa.2013.11.052Search in Google Scholar

[10] A. Farina and J. Serrin, Entire solutions of completely coercive quasilinear elliptic equations, J. Differential Equations 250 (2011), no. 12, 4367–4408. 10.1016/j.jde.2011.02.007Search in Google Scholar

[11] A. Farina and J. Serrin, Entire solutions of completely coercive quasilinear elliptic equations. II, J. Differential Equations 250 (2011), no. 12, 4409–4436. 10.1016/j.jde.2011.02.016Search in Google Scholar

[12] G. B. Folland and E. M. Stein, Hardy Spaces on Homogeneous Groups, Math. Notes 28, Princeton University Press, Princeton, 1982. 10.1515/9780691222455Search in Google Scholar

[13] B. Helffer, Spectral Theory and Its Applications, Cambridge Stud. Adv. Math. 139, Cambridge University Press, Cambridge, 2013. 10.1017/CBO9781139505727Search in Google Scholar

[14] T. Kato, Schrödinger operators with singular potentials, Israel J. Math. 13 (1972), 135–148. 10.1007/BF02760233Search in Google Scholar

[15] P. Lindqvist, Notes on the p-Laplace equation, Report 102, University of Jyväskylä, Jyväskylä, 2006. Search in Google Scholar

[16] P. Pucci and J. Serrin, A remark on entire solutions of quasilinear elliptic equations, J. Differential Equations 250 (2011), no. 2, 675–689. 10.1016/j.jde.2010.04.018Search in Google Scholar

Received: 2017-07-15
Accepted: 2017-07-30
Published Online: 2017-09-15
Published in Print: 2018-08-01

© 2018 Walter de Gruyter GmbH, Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 2.11.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2017-0164/html
Scroll to top button