Startseite Symmetry and asymmetry of minimizers of a class of noncoercive functionals
Artikel Öffentlich zugänglich

Symmetry and asymmetry of minimizers of a class of noncoercive functionals

  • Friedemann Brock , Gisella Croce EMAIL logo , Olivier Guibé und Anna Mercaldo
Veröffentlicht/Copyright: 9. September 2017

Abstract

In this paper we prove symmetry results for minimizers of a noncoercive functional defined on the class of Sobolev functions with zero mean value. We prove that the minimizers are foliated Schwarz symmetric, i.e., they are axially symmetric with respect to an axis passing through the origin and nonincreasing in the polar angle from this axis. In the two-dimensional case, we show a symmetry breaking.

MSC 2010: 49J40; 49K20; 49K30

1 Introduction

Consider the functional

vH01(Ω)12Ω|v|2

subjected to the constraint Ωv2=1, where Ω is the unit ball in the plane. Its critical values are the eigenvalues of the classical fixed membrane problem

(1.1){-Δu=λuin Ω,u=0on Ω.

It is known that the first eigenfunctions are positive and Schwarz symmetric, that is, radial and decreasing in the radial variable. By contrast, the second eigenfunctions are sign-changing; they are not radial, but they are symmetric with respect to the reflection at some line e, and they are decreasing in the angle arccos[x|x|e](0,π). These properties can be seen as a spherical version of the Schwarz symmetry along the foliation of the underlying ball Ω by circles. For this reason, this property is called foliated Schwarz symmetry in the literature.

In the last years much interest has been devoted to the shape of sign changing minimizers of integral functionals, see, for example, [8, 14, 23, 7, 20]. In [14] Girao and Weth studied the symmetry properties of the minimizers of the problem

(1.2)vv2vp,vH1(Ω),Ωv=0

for 2p<2*. In view of the zero average constraint, (1.2) is similar to the problem of finding the second eigenfunctions of problem (1.1). They proved that the minimizers are foliated Schwarz symmetric.

In [14] Girao and Weth pointed out another interesting phenomenon related to the shape of the minimizers of (1.2). If p is close to 2, then any minimizer of the above functional is anti-symmetric with respect to the reflection at the hyperplane {xe=0}. By contrast, the minimizers are not anti-symmetric when N=2 and p is sufficiently large. A similar break of symmetry was already observed in [12, 11, 16, 3, 10, 19] for the minimizers of the functional

vvLp(0,1)vLq(0,1),vW1,p((0,1)),v(0)=v(1),01v=0.

Indeed, it has been shown that any minimizer is an anti-symmetric function if and only if q3p.

In this paper we will prove similar symmetry results for the minimizers of a generalized version of the functional studied by Girao and Weth in [14]. We consider

(1.3)λθ,p(Ω)=inf{Ω|v|2(1+|v|)2θdx,vW1,q(Ω),v0,Ωvdx=0,vLp(Ω)=1},

where Ω is either a ball or an annulus centered in the origin in N, N2, and θ and q satisfy

(1.4)0<2θ<1,
(1.5)q=2N(1-θ)N-2θif N3,
(1.6)2(1-θ)q<2if N=2,
(1.7)1<p<q*if N3,
(1.8)1<p<+if N=2.

Observe that if one defines

Ψ(ξ):=0ξ(1+|t|)-θ𝑑t=sgnξ1-θ[(1+|ξ|)1-θ-1],ξ,

then our functional is the integral of |Ψ(u)|2, that is, (1.3) is equivalent to

λθ,p(Ω)=inf{Ω|Ψ(v)|2dx,vW1,q(Ω),v0,Ωvdx=0,vLp(Ω)=1}.

The main feature of this functional is that it is not coercive on H01(Ω), even if it is well defined on this Sobolev space. The lack of coercivity has unpleasant consequences for the minimizers of

vΩ[|v|2(1+|v|)2θ-G(x,v)]𝑑x

for functions G having various growth assumptions. Indeed, it was shown in [6, 2, 18, 13, 21, 4] that the minimizers are less regular than the minimizers of coercive functionals on H1(Ω).

After recalling the definition of foliated Schwarz symmetry and proving some new sufficient conditions for this symmetry in Section 3, we will prove the foliated Schwarz symmetry of the minimizers for N2. As already pointed out, the same result has been obtained by Girao and Weth in [14] in the ‘coercive’ case, that is, for θ=0. We observed that in their proof, Girao and Weth make use of a well-known regularity result of the solutions of the Euler equation. In our case, we have to prove the analogous regularity result for our noncoercive functional (see Section 4). Actually we are able to prove the foliated Schwarz symmetry of the minimizers of a more general functional, that is, we consider

(1.9)λθ,p(Ω)=inf{Ω|v|2-F(|x|,v)(1+|v|)2θdx,vW1,q(Ω),v0,Ωvdx=0,vLp(Ω)=1},

where we assume that F:+× is a measurable function in r=|x|[0,+) and continuously differentiable in t, which satisfies

(1.10)F(r,0)=0

and the growth conditions

(1.11)|F(r,t)|c(1+|t|)p,c>0,
(1.12)|Ft(r,t)|C1(1+|t|)p-1,C1>0,

for any r[0,+), t. If p(1,2), we add the requirement

(1.13)t(1+|t|)Ft(r,t)-2θ|t|F(r,t)0

for any r[0,+), t.

In the last two sections we will focus on the two-dimensional setting, in the case where Ω is a ball. We will prove that there exists a unique minimizer, which is anti-symmetric, for p=2 and sufficiently small θ. However, the minimizers are not anti-symmetric for p sufficiently large. This shows a symmetry breaking phenomenon, which generalizes the results proved by Girao and Weth in the case θ=0. Note that because of the difficulty given by the lack of coercivity of our functional, our technique is quite different from the one used in [14].

2 Existence of a minimizer

In this section we prove the existence of a minimizer for problem (1.9) by adapting the technique of [12]. We will also make use of an estimate proved in [6] (see also [1, 5]).

Theorem 2.1.

Under assumptions (1.4)–(1.8) and (1.10)–(1.13), there exists a minimizer u which realizes λθ,p(Ω) as defined in (1.9).

Proof.

We first observe that the growth assumption (1.11) on F and the condition uLp(Ω)=1 in the functional imply that λθ,p. For any fixed n, let us define

Hn(v)=Ω|v|2-F(|x|,v)(1+|v|)2θ𝑑x-(λθ,p(Ω)+1n),
H(v)=Ω|v|2-F(|x|,v)(1+|v|)2θ𝑑x-λθ,p(Ω)

for any vW1,q(Ω) such that v0, vLp(Ω)=1 and Ωv=0. By the definition of infimum, for any fixed n, there exists unW1,q(Ω), un0, such that

(2.1)unLp(Ω)=1,Ωun𝑑x=0,Hn(un)<0.

Now, by the growth assumption (1.11) on F, since the functions un have Lp-norm equal to 1, we have

(2.2)Ω|F(|x|,un)|(1+|un|)2θ𝑑xC,

where C is a positive constant which does not depend on n. From now on we will denote by C a positive constant which depends on the data and which can vary from line to line.

Since Hn(un)<0, estimates (2.1) and (2.2) imply that

(2.3)Ω|un|2(1+|un|)2θ𝑑xC.

Now we prove that |un| is bounded in Lq(Ω), that is, for any n,

(2.4)unLq(Ω)C.

We adapt the estimate used in [6, Theorem 2.1], and we distinguish the cases N3 and N=2.

Let N3 with q=2N(1-θ)N-2θ. We begin by applying the Hölder inequality, since q<2. Then we use estimate (2.3) and, since the mean value of un is null, by the Sobolev inequality, we get

Ω|un|qdx(Ω|un|2(1+|un|)2θdx)q2(Ω(1+|un|)2θq2-qdx)1-q2
C(Ω|un|2(1+|un|)2θdx)q2(1+Ω|un|q*dx)1-q2
C(1+Ω|un|qdx)q*q(1-q2),

where we have used the equality 2θq2-q=q*. Since N3, we deduce that q*q(1-q2)<1 and (2.4) is proved.

Let N=2. Similarly to above, using the Hölder inequality, estimate (2.3), the inclusion L2q2-q(Ω)L2qθ2-q(Ω) and the Sobolev inequality, we get

Ω|un|qdx(Ω|un|2(1+|un|)2θdx)q2(Ω(1+|un|)2θq2-qdx)1-q2
(Ω|un|2(1+|un|)2θdx)q2(1+Ω|un|2q2-qdx)(1-q2)θ
C(1+Ω|un|qdx)θ.

Since θ<1, inequality (2.4) follows again.

By the Poincaré–Wirtinger inequality, since the mean value of un is zero, we deduce, by (2.4), that

(2.5)un is bounded in W1,q(Ω),

and therefore there exists a function uW1,q(Ω) such that, as n, up to a subsequence,

(2.6)unuin W1,q(Ω) weakly,
(2.7)unuin Lr(Ω)1r<q*,
(2.8)unua.e. in Ω.

Let N3 with q=2N(1-θ)N-2θ. Note that (2.7), since p<q*, implies that

Ωu𝑑x=0,uLp(Ω)=1.

We claim that

(2.9)H(u)0.

Let

(2.10)Ψ(ξ):=0ξ(1+|t|)-θ𝑑t=sgnξ1-θ[(1+|ξ|)1-θ-1],ξ,

and observe, by (2.3), that

Ω|Ψ(un)|2dx=Ω|un|2(1+|un|)2θdxC.

Moreover, Ψ(un) is bounded in L2(Ω), since 2(1-θ)q, and un is bounded in Lq(Ω) by (2.5). We infer that Ψ(un) is bounded in W1,2(Ω) and, up to a subsequence, Ψ(un) converges weakly in W1,2(Ω) to a limit which is necessarily Ψ(u), by (2.8). Therefore, by the weak semi-continuity of the norm and the inequality in (2.1), as n, up to a subsequence,

Ψ(u)W1,22lim infn+Ψ(un)W1,22limn+ΩF(|x|,un)(1+|un|)2θdx+limn+(λθ,p(Ω)+1n).

To pass to the limit in the first term of the right-hand side, one can use the Lebesgue theorem. Indeed, the pointwise convergence is given by (2.8). The growth assumptions (1.11) on F and (2.7), since p<q*, imply the existence of a function hLp(Ω) such that

|F(|x|,un)|(1+|un|)2θh(x)a.e. in Ω.

Finally, we get

Ω|u|2(1+|u|)2θ𝑑xΩF(|x|,u)(1+|u|)2θ𝑑x+λθ,p(Ω),

that is, (2.9) holds. By the definition of λθ,p(Ω), necessarily we have H(u)=0. We observe that Ψ(u)0, since uLp(Ω)=1.

It remains to conclude the proof for the case N=2. Indeed, when N=2, we have 1<p<+ (see (1.8)) and 2(1-θ)q<2 (see (1.6)), so that the convergences (2.6)–(2.8) do not imply, in general, that uLp(Ω)=1. However, in view of (2.3) and since 2(1-θ)q, we obtain that Ψ(un) is bounded in W1,2(Ω). From the Sobolev embedding theorem, it follows that Ψ(un) is bounded in Lr(Ω) for any 1r<+. Since Ψ(ξ) grows like |ξ|1-θ, with 1-θ>0, we conclude that un is bounded in Lr(Ω) for any 1r<+. We obtain that uLp(Ω)=1, and the arguments developed in the case N=3 allow us to conclude that H(u)=0. ∎

3 Identification of symmetry

In this section we generalize some known symmetry criteria (cf. [9]). We first introduce some notations and definitions. Let Ω be a domain that is radially symmetric with respect to the origin. In other words, Ω is either an annulus, a ball, or the exterior of a ball in N. If u:Ω is a measurable function, we will, for convenience, always extend u onto N by setting u(x)=0 for xNΩ.

Definition 3.1.

Let 0 be the family of open half-spaces H in N such that 0H. For any H0, let σH denote the reflection in H. We write

σHu(x):=u(σHx),xN.

The two-point rearrangement with respect to H is given by

uH(x):={max{u(x),u(σHx)}if xH,min{u(x),u(σHx)}if xH.

The notion of two-point rearrangement was introduced more than fifty years ago as a set transformation in [24], and was applied to variational problems for the first time by Brock and Solynin in [9].

Note that one has u=uH if and only if u(x)u(σHx) for all xH. Similarly, σHu=uH if and only if u(x)u(σHx) for all xH.

We will make use of the following properties of the two-point rearrangement (see [9]).

Lemma 3.2.

Let HH0.

  1. If AC([0,+),), u:Ω is measurable and A(|x|,u)L1(Ω), then A(|x|,uH)L1(Ω) and

    (3.1)ΩA(|x|,u)𝑑x=ΩA(|x|,uH)𝑑x.
  2. If BL() and uW1,p(Ω) for some p[1,+), then

    (3.2)ΩB(u)|u|p𝑑x=ΩB(uH)|uH|p𝑑x.

Proof.

Since |σHx|=|x|, for a.e. xHΩ, we have

A(|x|,u(x))+A(|σHx|,u(σHx))=A(|x|,uH(x))+A(|σHx|,uH(σHx))

and

B(u(x))|u(x)|p+B(u(σHx))|u(σHx)|p=B(uH(x))|uH(x)|p+B(uH(σHx))|uH(σHx)|p.

Now (3.1) and (3.2) follow from this by integration over HΩ. ∎

In order to study the symmetry of minimizers of (1.9), we introduce the notion of foliated Schwarz symmetrization of a function, that is, a function which is axially symmetric with respect to an axis passing through the origin and nonincreasing in the polar angle from this axis.

Definition 3.3.

If u:Ω is measurable, the foliated Schwarz symmetrizationu* of u is defined as the (unique) function satisfying the following properties:

  1. There exists a function w:[0,+)×[0,π), w=w(r,θ), which is nonincreasing in θ, such that

    u*(x)=w(|x|,arccos(x1|x|)),xΩ.
  2. We have

    N-1{x:a<u(x)b,|x|=r}=N-1{x:a<u*(x)b,|x|=r}

    for all a,b, with a<b and r0.

Definition 3.4.

Let PN denote the point (1,0,,0), the ‘north pole’ of the unit sphere 𝒮N-1. We say that u is foliated Schwarz symmetric with respect to PN if u=u*, that is, u depends solely on r and θ (the ‘geographical width’), and is nonincreasing in θ.

We also say that u is foliated Schwarz symmetric with respect to a point P𝒮N-1 if there exists a rotation about the origin ρ such that ρ(PN)=P and u(ρ())=u*().

In other words, a function u:Ω is foliated Schwarz symmetric with respect to P if for every r>0 and c, the restricted superlevel set {x:|x|=r,u(x)c} is equal to {x:|x|=r} or a geodesic ball in the sphere {x:|x|=r} centered at rP. In particular, u is axially symmetric with respect to the axis P. Moreover, a measurable function u:Ω is foliated Schwarz symmetric with respect to P𝒮N-1 if and only if u=uH for all H0, with PH.

The main result of this section is the following, which gives a tool to establish if a measurable function is foliated Schwarz symmetric with respect to some point P.

Theorem 3.5.

Let uLp(Ω) for some p[1,+), and assume that for every HH0, one has either u=uH or σHu=uH. Then u is foliated Schwarz symmetric with respect to some point PSN-1.

Note that the above result has been shown for continuous functions in [23].

Theorem 3.6.

Let uC(RN), and assume that for every HH0 one has either u=uH or σHu=uH. Then u is foliated Schwarz symmetric with respect to some point PSN-1.

The idea in our proof is to use an approximation argument. Let φC0(N), φ0, with Nφ(x)𝑑x=1. Moreover, assume that φ is radial and radially nonincreasing, that is, there exists a nonincreasing function h:[0,+)[0,+) such that φ(x)=h(|x|) for all xN. For any function ε>0, define φε by φε(x):=ε-Nφ(ε-1x), xN. For any uLloc1(N), let uε be the standard mollifier of u given by

uε(x):=(u*φε)(x)Nu(y)φε(x-y)𝑑y,xN.

The following property is crucial. It allows a reduction to C-functions.

Lemma 3.7.

Let uLp(RN) for some p[1,+), and let HH0 be such that u=uH. Then uε=(uε)H for every ε>0.

Proof.

It is easy to see that

|x-y|=|σHx-σHy||σHx-y|=|x-σHy|

whenever x,yH. Since u(y)u(σHy) and since φε is radial and radially nonincreasing, for every xH, we have

uε(x)-uε(σHx)=Nu(y)[φε(x-y)-φε(σHx-y)]𝑑x
=H{u(y)[φε(x-y)-φε(σHx-y)]+u(σHy)[φε(x-σHy)-φε(σHx-σHy)]}𝑑x
=H(u(y)-u(σHy))[φε(x-y)-φε(σHx-y)]𝑑x0.

The lemma is proved. ∎

Corollary 3.8.

Let uLp(RN) for some p[1,+), and let HH0 be such that σHu=uH. Then we have σH(uε)=(uε)H for every ε>0.

We are now able to prove Theorem 3.5.

Proof of Theorem 3.5.

Since for every H0 one has either u=uH or σHu=uH, Lemma 3.7 and Corollary 3.8 apply. Then either uε=(uε)H or σHuε=(uε)H for every ε>0. Since uεC(N), Theorem 3.6 tells us that uε is foliated Schwarz symmetric with respect to some point Pε𝒮N-1, for every ε>0. Since 𝒮N-1 is compact, there exist a sequence of positive numbers {εn} and a point P𝒮N-1 such that uεn is foliated Schwarz symmetric with respect to a point Pn𝒮N-1, and εn0, PnP as n+. Let ρn and ρ be rotations such that ρn(N)=Pn, n, and ρ(N)=P. Writing un:=uεn, we have that

(3.3)un(ρn())=(un)*(),n.

Since unu, it follows that (un)*u* in Lp(N), and since PnP, we also have that un(ρn())u(ρ()) in Lp(N) as n. This, together with (3.3), implies that u(ρ())=u*(). The theorem is proved. ∎

4 Symmetry of minimizers

In this section we study the properties of symmetry of minimizers of (1.9). The main result is the following.

Theorem 4.1.

Assume (1.11)–(1.13) if p(1,2). Then every minimizer of (1.9) is foliated Schwarz symmetric with respect to some point PSN-1.

Remark 4.2.

Condition (1.13) is equivalent to

(4.1)tt(F(r,t)(1+|t|)2θ)0for all (r,t)[0,+)×.

It is satisfied, for instance, if F(r,t)=F(r,-t) and if

(1+t)Ft(r,t)-2θF(r,t)0for t0.

An example is

F(r,t)=-c0|t|α,α2θ,c00.

Observe that F satisfies the growth condition (1.11) with suitable c0 and α such that 2θαp.

Proof.

We divide the proof into four steps. Step 1. Let H0, and let u be a minimizer of (1.9). The Euler equation satisfied by u is

(4.2)-(u(1+|u|)2θ)-θ|u|2sgnu(1+|u|)2θ+1+c+d|u|p-2u=g(|x|,u)in Ω,
uν=0on Ω,

where c,d,

g(r,t):=t(F(r,t)2(1+|t|)2θ)for all (r,t)[0,+)×,

and ν denotes the exterior unit normal to Ω. Setting

I(v):=Ω|v|2-F(|x|,v)(1+|v|)2θ𝑑x,

by Lemma 3.2, we have

uH0,uHW1,q(Ω),ΩuH𝑑x=0,uHLp=1,I(u)=I(uH).

Hence, uH is a minimizer, too, so that it satisfies

(4.3)-(uH(1+|uH|)2θ)-θ|uH|2sgnuH(1+|uH|)2θ+1+c+d|uH|p-2uH=g(|x|,uH)in Ω,
uHν=0on Ω,

where c,d. Step 2. We claim that u,uHW1,q(Ω)W2,2(Ω)C1(Ω¯). Set

Φ(η):=Ψ-1(η),

where Ψ has been defined in (2.10). Let U:=Ψ(u). Note that u=Φ(U), uH=Φ(UH),

Φ(η)=([1+(1-θ)|η|]11-θ-1)sgnη,

and that Φ is locally Lipschitz continuous. Rewriting (4.2) and (4.3) in terms of U and UH, we find

(4.4)-ΔU+dM(U)=N(|x|,U),
(4.5)-ΔUH+dM(UH)=N(|x|,UH)

in Ω, where

M(t):=|Φ(t)|p-2Φ(t)(1+|Φ(t)|)θ,
N(r,t):=(g(r,t)-c)(1+|Φ(t)|)θ

for any r[0,+), t.

Observe that, by the growth conditions (1.11)–(1.12) and the definition of Φ(t), we have

|g(r,t)|cθ(1+|t|)p-1-2θ,
|M(t)|cθ(1+|t|)p-1+θ1-θ,
|N(r,t)|cθ′′(1+|t|)p-1-2θ+θ1-θ.

Now, the growths of M and N allow us to apply classical techniques for Neumann problems (see [17, p. 272] and [22, p. 271]) to state that UH1(Ω) is in fact C1,β(Ω¯), with β(0,1). Therefore, u has the same regularity. Step 3. Integrating (4.2) and (4.3) gives

(4.6)-θΩ|u|2sgnu(1+|u|)2θdx+cΩdx+dΩ|u|p-2udx=Ωg(|x|,u)dx,
(4.7)-θΩ|uH|2sgnuH(1+|uH|)2θdx+cΩdx+dΩ|uH|p-2uHdx=Ωg(|x|,uH)dx.

Further, multiplying (4.2) and (4.3) with u and uH, respectively, and then integrating and using the constraints, yields

(4.8)Ω|u|2[1+(1-θ)|u|](1+|u|)2θ+1𝑑x+d=Ωug(|x|,u)𝑑x,
(4.9)Ω|uH|2[1+(1-θ)|uH|](1+|uH|)2θ+1𝑑x+d=ΩuHg(|x|,uH)𝑑x.

Now (4.6)–(4.9), together with Lemma 3.2, show that necessarily c=c and d=d. Moreover, if p(1,2), then (4.1) holds, and so (4.8) yields d0. Step 4. Note that tM(t) is nondecreasing. Set h:=UH-U, and note that h0 in ΩH. We subtract (4.5) from (4.4) and split into two cases.

  1. Let p2. Then we find that

    -Δh=L(x)hin ΩH,

    where

    L(x):={N(|x|,UH(x))-N(|x|,U(x))-d[M(UH(x))-M(U(x))]h(x)if h(x)>0,0if h(x)=0

    is a bounded function.

  2. Let p(1,2). Then d0, and so

    -ΔhP(x)hin ΩH,

    where

    P(x):={N(|x|,UH(x))-N(|x|,U(x))h(x)if h(x)>0,0if h(x)=0

    is a bounded function.

Thus, in both cases the strong maximum principle tells us that either h(x)0 or h(x)>0 throughout ΩH. This implies that we have either u=uH or σHu=uH in Ω. By Theorem 3.5, we deduce that u is foliated Schwarz symmetric. ∎

5 Anti-symmetry for p=2 in dimension 2

In this section we study symmetry properties of the solutions to (1.9) in the case p=2, Ω=B, where B is a ball in 2, and F0. We will show that for small parameter values θ, there exists a unique minimizer of

vB|v|2(1+|v|)2θ𝑑x,vW1,q(B),v0,Bv𝑑x=0,vL2(B)=1,

which is anti-symmetric. Recall that θ satisfies (1.4) and q satisfies (1.6). With abuse of notations, we will denote the infimum of the above functional by λθ,2(B).

In the following we will use the notations of the proof of Theorem 4.1. In particular, let uθ be a minimizer for λθ,2(B), with corresponding constants c=cθ and d=dθ, see equation (4.6). By (4.8), we have

(5.1)dθ=-B|uθ|2[1+(1-θ)|uθ|)](1+|uθ|)2θ+1𝑑x.

We will also frequently work with the functions

Uθ:=Ψθ(uθ),where Ψθ(ξ)=sgn(ξ)1-θ[(1+|ξ|)1-θ-1]

(see (2.10)) and

Φθ(η)=Ψθ-1(η)=([1+(1-θ)|η|]11-θ-1)sgn(η).

Our calculations will often contain a generic constant C that may vary from line to line, but will be independent of θ.

Furthermore, as a consequence of Theorem 4.1, we will assume that uθ is foliated Schwarz symmetric with respect to the positive x1-half axis, that is,

(5.2)uθ(x1,x2)=uθ(x1,-x2).

The anti-symmetry of uθ then reads as uθ(x1,x2):=-uθ(-x1,x2) if θ is small.

Lemma 5.1.

Under assumptions (1.4), (1.6), the function θλθ,2(B) is decreasing. Moreover, λθ,2(B)λ2(B), where

λ2(B)=inf{B|u|2dx,uH1(B),Budx=0,uL2(B)=1}.

Proof.

Let θ1<θ2, let uθ1 be a minimizer for λθ1,2(B), and let 2(1-θ1)q<2. Then we obtain

λθ1,2(B)=B|u1|2(1+|u1|)2θ1𝑑xB|u1|2(1+|u1|)2θ2𝑑xλθ2,2(B).

Next let u be a minimizer for λ2(B). Then

λ2(B)B|u|2dxB|u|2(1+|u|)2θdxλθ,2(B).

Lemma 5.2.

Under assumptions (1.4), (1.6), let uθ be a minimizer for λθ,2(B). Let also dθ be defined by (5.1). Then limθ0dθ=-λ2(B).

Proof.

First we observe that

(5.3)-dθ=B|uθ|2(1+(1-θ)|uθ|))(1+|uθ|)2θ+1𝑑xB|uθ|2(1+|uθ|)2θ𝑑xλθ,2(B)λ2(B),

by Lemma 5.1. On the other hand,

(5.4)-dθB|uθ|2(1-θ)(1+|uθ|)2θdx=(1-θ)B|Uθ|2dx.

Moreover, it is easy to see that UθL2(B) is uniformly bounded, since uθL2(B)=1 and θ12. Therefore, UθH1(B) is also uniformly bounded. By compactness, as θ0, Uθ converges weakly to some function VH1(B) and strongly in L2(B). By the lower semi-continuity of the norm, from (5.4), we get

(5.5)lim infθ0(-dθ)VL2(B)2.

Further, the a.e. limit of Uθ is the limit of uθ, say u. By the uniqueness of the limit, u=V a.e. in B. We recall that uθL2(B)=1 and Buθ𝑑x=0. Since Ψθ(uθ)H1(B), by the growth of Ψθ, we deduce that uθu in L2(B) and that VL2(B)=1 and BV𝑑x=0. Together with (5.5), this implies that

(5.6)lim infθ0(-dθ)λ2(B).

Now the lemma follows from inequalities (5.3) and (5.6). ∎

Proposition 5.3.

Let uθ be a minimizer for λθ,2(B). Under assumptions (1.4), (1.6), the following hold:

  1. We have uθW1,(B)C, where C does not depend on θ.

  2. Let u be the limit of uθ, as θ0, in W1,r(Ω), for every r(1,+). Then uL2(B)=1 and Bu𝑑x=0.

Proof.

The H1(B)-norm of Uθ is uniformly bounded by Lemma 5.1. By multiplying equation (4.4) by Uθ and integrating on B, one has that the right-hand side of the equality is uniformly bounded, due to Lemma 5.1. The growth of M and Lemma 5.2 imply that UθLp-1+θ1-θ(B)C. This allows us to use the bootstrap argument described in [22, p. 271]. ∎

Let vθ(x1,x2):=-uθ(-x1,x2). Then, from (4.6), we obtain

(5.7)cθ=θ|B|B|uθ|2sgn(uθ)(1+|uθ|)2θ+1𝑑x=-θ|B|B|vθ|2sgn(vθ)(1+|vθ|)2θ+1𝑑x.

Lemma 5.4.

Under assumptions (1.4), (1.6), let uθ be a minimizer for λθ,2(B). Let also cθ be defined by (5.7). Then |cθ|Cθuθ-vθL2(B) for a positive constant C independent on θ.

Proof.

By multiplying equation (4.2) (with p=2 and g=0) by (1+|uθ|)θ, we have

(5.8)-((1+|uθ|)-θuθ)+cθ(1+|uθ|)θ+dθuθ(1+|uθ|)θ=0.

Integrating this gives

cθB(1+|uθ|)θ𝑑x+dθBuθ(1+|uθ|)θ𝑑x=0,

since uθν=0 on B. The first integral in this identity is bounded from below, and |dθ| is bounded by Lemma 5.2. If J denotes Buθ(1+|uθ|)θ𝑑x, we deduce that

(5.9)|cθ|C|J|

for a constant C independent on θ. A change of variables gives

J=12B[uθ(1+|uθ|)θ-vθ(1+|vθ|)θ]𝑑x.

Since Buθ𝑑x=Bvθ𝑑x=0, we get

J=12B(uθ-vθ)[(1+|vθ|)θ-1]𝑑x+12Buθ[(1+|uθ|)θ-(1+|vθ|)θ]𝑑x.

Let J1 denote the first term and J2 the second one in this identity. A short computation shows that

|J1|θ2B|uθ-vθ||vθ|dx,|J2|θ2B|uθ-vθ||uθ|dx.

Since uθ and vθ are uniformly bounded by Proposition 5.3, this gives

|J|Cθuθ-vθL2(B).

The conclusion follows from estimate (5.9). ∎

Now we can prove the main result of the section.

Theorem 5.5.

There exists a number θ0>0, such that every minimizer uθ of λθ,2(B) satisfying (5.2) is unique and anti-symmetric with respect to x1, that is,

uθ(x1,x2):=-uθ(-x1,x2)

for any 0<θ<θ0.

Proof.

We first prove that any minimizer is anti-symmetric. Let Uθ:=Ψθ(uθ) and Vθ:=Ψθ(vθ), where vθ(x1,x2)=-uθ(-x1,x2). Writing equation (5.8) in terms of Uθ, we have

-ΔUθ+cθ(1+|uθ|)θ+dθuθ(1+|uθ|)θ=0.

Similarly,

-ΔVθ-cθ(1+|vθ|)θ+dθvθ(1+|vθ|)θ=0.

Subtract both equations from each other. Assuming that Uθ-Vθ0 along a sequence θ0, we multiply by Uθ-VθUθ-VθL2(B)2 and integrate. Then we obtain

(Uθ-Vθ)L2(B)2Uθ-VθL2(B)2+cθUθ-VθL2(B)B[(1+|uθ|)θ+(1+|vθ|)θ]Uθ-VθUθ-VθL2(B)𝑑x
=-dθB[uθ)(1+|uθ|)θ-vθ(1+|vθ)|)θ]Uθ-VθUθ-VθL2(B)2dx.

The second term of the left-hand side tends to zero, by Lemma 5.4 and since (1+|uθ|)θ+(1+|vθ|)θ is uniformly bounded by Proposition 5.3. To estimate the right-hand side, we first observe that -dθλ2(B), by Lemma 5.2. Moreover, it is not difficult to prove the following estimate:

|uθ(1+|uθ|)θ-vθ(1+|vθ|)θ|(1+θ)[1+(1-θ)|ξθ|]2θ1-θ|Uθ-Vθ|,

where ξθ is between Uθ=Ψθ(uθ) and Vθ=Ψθ(vθ). By Proposition 5.3, we deduce that

(1+θ)[1+(1-θ)|ξθ|]2θ1-θ1as θ0,

uniformly in B.

Now, set

Wθ:=Uθ-VθUθ-VθL2(B).

By the above identity, the norms WθL2(B) are uniformly bounded. Hence, there exists a function W~H1(B) such that, along a subsequence, WθW~ weakly in L2(B) and WθW~ in L2(B), and so W~L2(B)=1. This also implies

W~L2(B)2λ2(B).

Next we claim that |BWθ𝑑x|Cθ. Indeed, since Buθ𝑑x=Bvθ𝑑x=0, we have

|B(Uθ-Vθ)𝑑x|=B(Ψθ(uθ)-uθ-Ψθ(vθ)+vθ)𝑑x
B|vθuθ(Ψθ(t)-1)(uθ-vθ)dt|dx
θB|uθ-vθ|dx.

Now, Φθ is locally Lipschitz continuous, uniformly in θ. By Proposition 5.3, we obtain

|B(Uθ-Vθ)dx|θB|uθ-vθ|dxCθB|Uθ-Vθ|dxCθUθ-VθL2(B),

and the claim follows. This and the fact that WθW~ in L2(B) prove that BW~𝑑x=0. Then, by the definition of λ2(B), we have that

W~L2(B)2λ2(B).

Hence, W~ is a (nonzero) eigenfunction for the Neumann Laplacian in B. By the properties of Uθ and Vθ and by (5.2), one has Wθ(x1,x2)=Wθ(-x1,x2)=Wθ(x1,-x2). Thus, the same symmetry properties hold for W~. But this is in contradiction with the shape of the eigenfunction in a ball, which is given by

Jn(αnk|(x1,x2)|R){cos(nφ),l=1,sin(nφ),l=2(n0),

where we have used the polar coordinates, R is the radius of the ball, and αnk are the positive roots of the derivative of the Bessel function Jn (see, for example, [15]). We thus have proved that any minimizer is anti-symmetric. Note that the anti-symmetry also implies that cθ=0, which can be seen by integrating (4.2).

It remains to prove that the minimizer is unique for small θ. Assume this is not the case. Then there is a sequence θ0 along which there exist two distinct minimizers uθ and uθ. Let the corresponding constants d of (4.2) be denoted by dθ and dθ. Multiplying (4.2) by uθ and integrating by parts gives

dθ=-B|uθ|2(1+|uθ|)2θ𝑑x+θB|uθ|2|uθ|(1+|uθ|)2θ+1𝑑x
(5.10)=-λ2,θ(B)+θB|uθ|2|uθ|(1+|uθ|)2θ+1𝑑x.

Similarly,

(5.11)dθ=-λ2,θ(B)+θB|uθ|2|uθ|(1+|uθ|)2θ+1𝑑x.

We define

gθ(ξ):=|ξ|(1+|ξ|)2θ+1.

Since the functions gθ are locally Lipschitz continuous, uniformly in θ, using Proposition 5.3, we can estimate

||uθ|2|uθ|(1+|uθ|)2θ+1-|uθ|2|uθ|(1+|uθ|)2θ+1|=||uθ|(1+|uθ|)2θ+1(|uθ|2-|uθ|2)+|uθ|2(gθ(uθ)-gθ(uθ))|
(5.12)C(|uθ-uθ|+|uθ-uθ|).

Subtracting (5.11) from (5.10) and taking into account (5.12), we obtain

|dθ-dθ|θB||uθ|2|uθ|(1+|uθ|)2θ+1-|uθ|2|uθ|(1+|uθ|)2θ+1|𝑑x
(5.13)Cθ(uθ-uθL2(B)+(uθ-uθ)L2(B)).

Now we claim that

(5.14)|dθ-dθ|CθUθ-UθL2(B).

As in the proof of Lemma 5.4, by multiplying equation (4.2) (with p=2 and g=0) by (1+|uθ|)θ, we have

-((1+|uθ|)-θuθ)+cθ(1+|uθ|)θ+dθuθ(1+|uθ|)θ=0.

Moreover, the analog of this equality holds true for uθ. In addition, by multiplying this equation by uθ-uθ, we get

B(uθ(1+|uθ|)θ-uθ(1+|uθ|)θ)(uθ-uθ)dx
+dθB[uθ(1+|uθ|)θ-uθ(1+|uθ|)θ](uθ-uθ)𝑑x+(dθ-dθ)Buθ(1+|uθ|)θ(uθ-uθ)𝑑x=0.

Now we evaluate the three integrals on the left-hand side. For a small enough value of θ, we have

B(uθ(1+|uθ|)θ-uθ(1+|uθ|)θ)(uθ-uθ)dx
=B1(1+|uθ|)θ|(uθ-uθ)|2𝑑x+Buθ(uθ-uθ)(1(1+|uθ|)θ-1(1+|uθ|)θ)𝑑x
(5.15)12(uθ-uθ)L2(B)2-C(uθ-uθ)L2(B)uθ-uθL2(B).

Moreover, as in the calculation of J in the previous arguments (see after (5.9)), we get

(5.16)|dθB[uθ(1+|uθ|)θ-uθ(1+|uθ|)θ](uθ-uθ)𝑑x|Cuθ-uθL2(B)2,
Buθ(uθ-uθ)(1(1+|uθ|)θ-1(1+|uθ|)θ)𝑑x
(5.17)Cθuθ-uθL2(B)(uθ-uθL2(B)+(uθ-uθ)L2(B)).

Combining (5.15)–(5.17), from the Young inequality, we get

(uθ-uθ)L2(B)2Cuθ-uθL2(B)2.

Now as in the previous calculation, we get

uθ-uθL2(B)Uθ-UθL2(B),

which gives (5.14).

Next we define

hθ(ξ):=ξ(1+|ξ|)θ-Ψθ(ξ)=ξ(1+|ξ|)θ-sgnξ1-θ[(1+|ξ|)1-θ-1].

It is easy to see that hθ is locally Lipschitz continuous with

|hθ(ξ)|Cθ,|ξ|M,

where the constant C depends only on M (M>0). From this, using Proposition 5.3, we obtain

|hθ(uθ)-hθ(uθ)|Cθ|uθ-uθ|.

Now let Uθ:=Ψθ(uθ) and Uθ:=Ψθ(uθ). Arguing as before, we first observe that

(5.18)|uθ-uθ|C|Ψθ(uθ)-Ψθ(uθ)|=C|Uθ-Uθ|.

By (4.2), we have

(5.19)-ΔUθ+dθuθ(1+|uθ|)θ=0,
(5.20)-ΔUθ+dθuθ(1+|uθ|)θ=0.

Subtracting (5.20) from (5.19), multiplying with (Uθ-Uθ) and integrating by parts gives

0=(Uθ-Uθ)22+dθB(uθ(1+|uθ|)θ-uθ(1+|uθ|)θ)(Uθ-Uθ)𝑑x
(5.21)+(dθ-dθ)Buθ(1+|uθ|)θ(Uθ-Uθ)𝑑x.

Now define Wθ:=(Uθ-Uθ)Uθ-UθL2(B)-1. Then, from (5.21), we obtain

0=WθL2(B)2+dθB(uθ(1+|uθ|)θ-uθ(1+|uθ|)θ)WθUθ-UθL2(B)-1𝑑x
+(dθ-dθ)Buθ(1+|uθ|)θWθUθ-UθL2(B)-1𝑑x.
=WθL2(B)2+dθWθL2(B)2+dθB(hθ(uθ)-hθ(uθ))WθUθ-UθL2(B)-1𝑑x
(5.22)+(dθ-dθ)Buθ(1+|uθ|)θWθUθ-UθL2(B)-1𝑑x.

Since WθL2(B)=1, and in view of estimates (5) and (5.18), we see that the last two terms in (5.22) tend to zero as θ0. Hence, the functions Wθ are uniformly bounded in H1(B). By passing to another subsequence if necessary, we find a function W¯H1(B) such that WθW¯ weakly in H1(B) and WθW¯ in L2(B). Then, passing to the limit in (5.22), since lim infθ0WθL2(B)W¯L2(B), we obtain

(5.23)W¯L2(B)2λ2(B)W¯L2(B)2.

Since also BW¯𝑑x=0 and BW¯2𝑑x=1, we must have equality in (5.23), and W¯ is an anti-symmetric eigenfunction for the Neumann Laplacian in B, that is, W¯=u. In other words, we have

BWθUθ𝑑xBu2𝑑x=1as θ0.

On the other hand, we calculate

BWθUθ𝑑x=BUθ2𝑑x-BUθUθ𝑑xUθ-UθL2(B)
=1-BUθUθ𝑑x2-2BUθUθ𝑑x
=121-BUθUθ𝑑x0as θ0,

which gives a contradiction. The proof of Theorem 5.5 is complete. ∎

6 Symmetry breaking in dimension 2

In this section we continue studying the two-dimensional case, assuming again that F0. We show that for p sufficiently large, the minimizers of λθ,p do not verify the properties of anti-symmetry described in the previous section; therefore a phenomenon of symmetry breaking occurs.

Let us denote by Was1,q(B) the subset of the Sobolev space W1,q(B) of the functions which are anti-symmetric with respect to the plane P{xN+1:xN=0}, that is,

Was1,q(B):={vW1,q(B):u(x,-xN)=-u(x,xN)}.

Let

(v)=B|v|2(1+|v|)2θ𝑑x,vW1,q(B),v0,Bv𝑑x=0,vL2(B)=1.

Recall that θ satisfies (1.4) and q satisfies (1.6). Let

λθ,p(B):=inf{(v),vW1,q(B),v0,Bvdx=0,vLp(B)=1}

and

λasθ,p(B):=inf{(v),vWas1,q(B),v0,Bvdx=0,vLp(B)=1}.

Observe that the existence of a function realizing λasθ,p(B) can be proved as in Theorem 2.1. Let us also recall a well-known result. For any bounded smooth domain Ω2, let

Λasp(Ω)=inf{vL2(Ω)2,vWas1,2(Ω),v0,Ωvdx=0,vLp(Ω)=1}.

In [14] the behavior of Λasp(Ω) is studied and it is proved that

(6.1)Λasp(Ω)0as p.

It is easy to prove the same result for λasθ,p(B).

Proposition 6.1.

We have λasθ,p(B)0 as p.

Proof.

Since B|v|2dx(v), one has

Λasp(B)inf{(v),vWas1,2(B),v0,Bvdx=0,vLp(B)=1}
inf{(v),vWas1,q(B),v0,Bvdx=0,vLp(B)=1}
=λasθ,p(B).

By (6.1), the conclusion follows. ∎

Now we can prove the main result of the section.

Theorem 6.2.

For p sufficiently large, λθ,p(B)<λasθ,p(B). Therefore, the minimizers of F are not anti-symmetric for p sufficiently large.

Proof.

Let vp be an eigenfunction for λasθ,p(B). Then vpLp(B)=1. Let B+={(x1,x2)B:x2>0}, and let u¯p be defined by

u¯p(x)={vp(x),xB+,0,xBB+.

Then

(6.2)B|u¯p|2(1+|u¯p|)2θ𝑑x=λasθ,p(B)2,u¯pLp(B)p=12.

We claim that

(6.3)Bu¯p𝑑x0as p.

By Proposition 6.1, we deduce that

λasθ,p(Ω)=2Ψ(u¯p)L2(B)0as p,

where Ψ has been defined in (2.10). Since u¯p=0 in BB+, we can use the Poincaré–Wirtinger inequality, which implies

Ψ(u¯p)L2(B)0as p.

Therefore, up to subsequence, Ψ(u¯p)0 and up0 a.e. in B. On the other hand, there exists a function hL2(B) such that |Ψ(u¯p)|h a.e. in B. By the definition of Ψ(t), we deduce the existence of a function kL2(1-θ)(B) such that |u¯p|k a.e. in B. Hence, Lebesgue’s theorem applies and we get B|u¯p|dx0. This proves (6.3).

Next we define

u~p:=u¯p-1|B|Bu¯p𝑑xu¯p-1|B|Bu¯p𝑑xLp(B).

Then

λθ,p(B)B|u~p|2(1+|u~p|)2θ𝑑x=1u¯p-1|B|Bu¯p𝑑xLp(B)2B|u¯p|2(1+|u¯p-1|B|Bu¯p𝑑x|u¯p-1|B|Bu¯p𝑑xLp(B))2θ𝑑x.

Let ε>0 be sufficiently small. For a suitable p(ε)>0 and for any p>p(ε), by (6.2), one has

(6.4)λθ,p(B)1[(1/2)1p-ε]2B|u¯p|2[1+|u¯p-1|B|Bu¯p𝑑x|(1/2)1p+ε]2θ𝑑x.

Let us set Mε=1+ε1-ε. We claim that

(6.5)G(u¯p)1+|u¯p|1+|u¯p-1|B|Bu¯p𝑑x|(1/2)1p+εMε.

First of all, it is easy to verifies that for any p>p(ε),

(6.6)1+|u¯p-1|B|Bu¯p𝑑x|1+||u¯p|-1|B||Bu¯p𝑑x||1+||u¯p|-ε|(1-ε)(1+|u¯p|).

Now we distinguish two cases.

  1. If (1/2)1p+ε1, then

    G(u¯p)1+|u¯p|1+|u¯p-1|B|Bu¯p𝑑x|.

    By (6.6), one has G(u¯p)11-ε.

  2. If (1/2)1p+ε>1, then, by (6.6),

    1+|u¯p-1|B|Bu¯p𝑑x|(1/2)1p+ε1+|u¯p-1|B|Bu¯p𝑑x|(1/2)1p+ε1-ε(1/2)1p+ε(1+|u¯p|).

Therefore, (6.5) is proved, that is,

(6.7)11+|u¯p-1|B|Bu¯p𝑑x|(1/2)1p+εMε1+|u¯p|.

Combining estimates (6.4) and (6.7), we get

λθ,p(B)Mε2θ[(1/2)1p-ε]2B|u¯p|2(1+|u¯p|)2θ𝑑x.

It is clear that

Mε2θ[(1/2)1p-ε]2=(1+ε1-ε)2θ1[(1/2)1p-ε]2<2

for p>p(ε). Therefore, for p sufficiently large, one has

λθ,p(B)<2B|u¯p|2(1+|u¯p|)2θ𝑑x=λasθ,p(B),

by (6.2). ∎


Communicated by Bernard Dacorogna


Funding statement: This work started during a visit of A. Mercaldo to Université de Rouen which was financed by the Fédération Normandie Mathématiques. The last author is a member of Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM) which supported visitings of F. Brock to Università degli Studi di Napoli Federico II. All these institutions are gratefully acknowledged.

References

[1] A. Alvino, L. Boccardo, V. Ferone, L. Orsina and G. Trombetti, Existence results for nonlinear elliptic equations with degenerate coercivity, Ann. Mat. Pura Appl. (4) 182 (2003), no. 1, 53–79. 10.1007/s10231-002-0056-ySuche in Google Scholar

[2] D. Arcoya, L. Boccardo and L. Orsina, Existence of critical points for some noncoercive functionals, Ann. Inst. H. Poincaré Anal. Non Linéaire 18 (2001), no. 4, 437–457. 10.1016/s0294-1449(01)00069-5Suche in Google Scholar

[3] M. Belloni and B. Kawohl, A symmetry problem related to Wirtinger’s and Poincaré’s inequality, J. Differential Equations 156 (1999), no. 1, 211–218. 10.1006/jdeq.1998.3603Suche in Google Scholar

[4] L. Boccardo, G. Croce and L. Orsina, W01,1 minima of noncoercive functionals, Atti Accad. Naz. Lincei Rend. Lincei Mat. Appl. 22 (2011), no. 4, 513–523. 10.4171/RLM/612Suche in Google Scholar

[5] L. Boccardo, A. Dall’Aglio and L. Orsina, Existence and regularity results for some elliptic equations with degenerate coercivity, Atti Semin. Mat. Fis. Univ. Modena 46 (1998), 51–81. Suche in Google Scholar

[6] L. Boccardo and L. Orsina, Existence and regularity of minima for integral functionals noncoercive in the energy space, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 25 (1997), no. 1–2, 95–130. Suche in Google Scholar

[7] B. Brandolini, F. Della Pietra, C. Nitsch and C. Trombetti, Symmetry breaking in a constrained Cheeger type isoperimetric inequality, ESAIM Control Optim. Calc. Var. 21 (2015), no. 2, 359–371. 10.1051/cocv/2014016Suche in Google Scholar

[8] B. Brandolini, P. Freitas, C. Nitsch and C. Trombetti, Sharp estimates and saturation phenomena for a nonlocal eigenvalue problem, Adv. Math. 228 (2011), no. 4, 2352–2365. 10.1016/j.aim.2011.07.007Suche in Google Scholar

[9] F. Brock and A. Y. Solynin, An approach to symmetrization via polarization, Trans. Amer. Math. Soc. 352 (2000), no. 4, 1759–1796. 10.1090/S0002-9947-99-02558-1Suche in Google Scholar

[10] A. P. Buslaev, V. A. Kondrat’ev and A. I. Nazarov, On a family of extremal problems and related properties of an integral, Mat. Zametki 64 (1998), no. 6, 830–838. 10.4213/mzm1462Suche in Google Scholar

[11] G. Croce and B. Dacorogna, On a generalized Wirtinger inequality, Discrete Contin. Dyn. Syst. 9 (2003), no. 5, 1329–1341. 10.3934/dcds.2003.9.1329Suche in Google Scholar

[12] B. Dacorogna, W. Gangbo and N. Subía, Sur une généralisation de l’inégalité de Wirtinger, Ann. Inst. H. Poincaré Anal. Non Linéaire 9 (1992), no. 1, 29–50. 10.1016/s0294-1449(16)30249-9Suche in Google Scholar

[13] F. Faraci, A bifurcation theorem for noncoercive integral functionals, Comment. Math. Univ. Carolin. 45 (2004), no. 3, 443–456. Suche in Google Scholar

[14] P. Girão and T. Weth, The shape of extremal functions for Poincaré–Sobolev-type inequalities in a ball, J. Funct. Anal. 237 (2006), no. 1, 194–223. 10.1016/j.jfa.2006.01.001Suche in Google Scholar

[15] A. Henrot, Extremum Problems for Eigenvalues of Elliptic Operators, Front. Math., Birkhäuser, Basel, 2006. 10.1007/3-7643-7706-2Suche in Google Scholar

[16] B. Kawohl, Symmetry results for functions yielding best constants in Sobolev-type inequalities, Discrete Contin. Dyn. Syst. 6 (2000), no. 3, 683–690. 10.3934/dcds.2000.6.683Suche in Google Scholar

[17] J. Mawhin, J. R. Ward, Jr. and M. Willem, Variational methods and semilinear elliptic equations, Arch. Ration. Mech. Anal. 95 (1986), no. 3, 269–277. 10.1007/BF00251362Suche in Google Scholar

[18] A. Mercaldo, Existence and boundedness of minimizers of a class of integral functionals, Boll. Unione Mat. Ital. Sez. B Artic. Ric. Mat. (8) 6 (2003), no. 1, 125–139. Suche in Google Scholar

[19] A. I. Nazarov, On exact constant in the generalized Poincaré inequality. Function theory and applications, J. Math. Sci. (New York) 112 (2002), no. 1, 4029–4047. 10.1023/A:1020006108806Suche in Google Scholar

[20] E. Parini and T. Weth, Existence, unique continuation and symmetry of least energy nodal solutions to sublinear Neumann problems, Math. Z. 280 (2015), no. 3–4, 707–732. 10.1007/s00209-015-1444-5Suche in Google Scholar

[21] A. Porretta, Remarks on existence or loss of minima of infinite energy, Asymptot. Anal. 52 (2007), no. 1–2, 53–94. Suche in Google Scholar

[22] M. Struwe, Variational Methods. Applications to Nonlinear Partial Differential Equations and Hamiltonian Systems, 4th ed., Ergeb. Math. Grenzgeb. (3) 34, Springer, Berlin, 2008. Suche in Google Scholar

[23] T. Weth, Symmetry of solutions to variational problems for nonlinear elliptic equations via reflection methods, Jahresber. Dtsch. Math.-Ver. 112 (2010), no. 3, 119–158. 10.1365/s13291-010-0005-4Suche in Google Scholar

[24] V. Wolontis, Properties of conformal invariants, Amer. J. Math. 74 (1952), 587–606. 10.2307/2372264Suche in Google Scholar

Received: 2017-02-01
Accepted: 2017-08-14
Published Online: 2017-09-09
Published in Print: 2020-01-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 22.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/acv-2017-0005/html
Button zum nach oben scrollen