This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Yamabe problem in the presence of singular Riemannian Foliations

Diego Corro Mathematics Institute of the National Autonomous University of Mexico (UNAM), Oaxaca, Mexico. Mathematisches Institut, Universität zu Köln, Köln, Deutschland. [email protected] Juan Carlos Fernandez∗∗ Facultad de Ciencias, Universidad Nacional Autónoma de México (UNAM), Ciudad de México, México. [email protected]  and  Raquel Perales Investigadora por México at the Mathematics Institute of the National Autonomous University of Mexico (UNAM), Oaxaca, Mexico. [email protected]
Abstract.

Using variational methods together with symmetries given by singular Riemannian foliations with positive dimensional leaves, we prove the existence of an infinite number of sign-changing solutions to Yamabe type problems, which are constant along the leaves of the foliation, and one positive solution of minimal energy among any other solution with these symmetries. In particular, we find sign-changing solutions to the Yamabe problem on the round sphere with new qualitative behavior when compared to previous results, that is, these solutions are constant along the leaves of a singular Riemannian foliation which is not induced neither by a group action nor by an isoparametric function. To prove the existence of these solutions, we prove a Sobolev embedding theorem for general singular Riemannian foliations, and a Principle of Symmetric Criticality for the associated energy functional to a Yamabe type problem.

Key words and phrases:
Yamabe problem, singular Riemannian foliation
2020 Mathematics Subject Classification:
Primary 58J05, 53C12; Secondary 35J61, 35B06, 35J20, 35R01, 53C21, 57R30
Supported by DGAPA-Fellowship associated to the Mathematics Institute of UNAM, campus Oaxaca, and by DFG-Eigenestelle Fellowship CO 2359/1-1
∗∗Partially supported by Professor Christina Sormani’s NSF Reaserch Grant DMS-1612049

1. Introduction

The Yamabe problem, stated by Yamabe in [52], asks if for a given closed Riemannian manifold (M,g)(M,g) there exists a conformal Riemannian metric g^=fg\hat{g}=fg, for some smooth positive function f:Mf\colon M\to\mathbb{R}, such that the scalar curvature of g^\hat{g} is constant. This problem can be written in terms of a PDE, and has been completely solved in the positive by the combined work of Yamabe [52], Trudinger [47], Aubin [5], and Schoen [45]. Solutions to this problem are not necessarily unique, although for manifolds not conformally equivalent to the round sphere with dimension less than or equal to 2424, the set of solutions is compact (see [35]). In contrast, for higher dimensions the set of solutions is not compact (see [8, 9]). For an extensive exposition of the Yamabe problem, the interested reader may consult [6, 36].

In the past years, sign-changing solutions to the Yamabe equation have been studied. These are functions uu which are sign changing and satisfy the Yamabe equation. For k,n2k,n\geq 2, Ding [22] established the existence of infinitely many sign-changing solutions for the round (k+n)(k+n)-sphere, using the fact that there is a (linear) O(k)×O(n)\mathrm{O}(k)\times\mathrm{O}(n) action by isometries. Moreover, Ammann and Humbert proved in [4] that in dimension at least 1111 for a closed Riemannian manifold with positive Yamabe invariant and not locally conformally flat, there exists a minimal energy sign-changing solution to the Yamabe equation.

The study of the existence of sign-changing solutions has been recently carried out by several authors using very different techniques [4, 13, 12, 21, 25, 33, 48]. One of the approaches considered has been to find equivariant solutions with respect to a given compact Lie group action by isometries with positive dimensional orbits. The approach of finding equivariant solutions to the Yamabe equation has also led to an equivariant solution to the Kazdan-Warner problem in [11].

Recently, singular Riemannian foliations have been considered as a notion of symmetry for Riemannian manifolds in the context of manifolds with nonnegative sectional curvature, since they are a natural extension to the concepts of group actions and Riemannian submersions (see for example [2, 19, 28, 29]). Moreover, several results which hold for group actions that depend only on the geometry that is transverse to the orbits can be extended to the setting of singular Riemannian foliations.

In light of this, we study the existence of sign-changing solutions for a family of elliptic partial differential equations related to the Yamabe problem in the presence of singular Riemannian foliations.

Namely, let (M,g)(M,g) be a closed Riemannian manifold of dimension m3m\geq 3 and consider the following Yamabe type problem:

(Y) Δgu+bu=c|u|p2uon M,-\Delta_{g}u+bu=c|u|^{p-2}u\quad\text{on }\ M,

where Δg=divggradg\Delta_{g}=\text{div}_{g}\text{grad}_{g} is the Laplace-Beltrami operator, p>2p>2, b,c𝒞(M)b,c\in\mathcal{C}^{\infty}(M) with c>0c>0. We will assume that the operator Δg+b-\Delta_{g}+b is coercive in the Sobolev space Hg1(M)H_{g}^{1}(M), meaning, that there exists μ>0\mu>0 such that

(1) Mgu,gvg+buvdVgμMgu,gvg+uvdVg\int_{M}\langle\nabla_{g}u,\nabla_{g}v\rangle_{g}+b\,uv\;dV_{g}\geq\mu\int_{M}\langle\nabla_{g}u,\nabla_{g}v\rangle_{g}+uv\;dV_{g}

for every u,vHg1(M)u,v\in H_{g}^{1}(M).

Let \mathcal{F} be a singular Riemannian foliation on MM with closed leaves and nontrivial, meaning that \mathcal{F} is different to the foliation that consists of only one leaf, {M}\{M\}, and to the foliation {{p}pM}\{\{p\}\mid p\in M\} (see Section 2.1 for definitions). We will further assume that all the leaves of \mathcal{F} have dimension greater than or equal to one. For that, we define the number

(2) κ:=min{dimLL}.\kappa_{\mathcal{F}}:=\min\{\dim L\mid L\in\mathcal{F}\}.

We say that a function u:Mu\colon M\to\mathbb{R} is \mathcal{F}-invariant if uu is constant on each leaf LL of \mathcal{F}.

Theorem A.

þ Let (M,g,)(M,g,\mathcal{F}) be an mm-dimensional Riemannian manifold, m3m\geq 3, together with a nontrivial closed singular Riemannian foliation such that κ1\kappa_{\mathcal{F}}\geq 1. Assume that Δg+b-\Delta_{g}+b is coercive in Hg1(M)H_{g}^{1}(M), that bb and cc are \mathcal{F}-invariant functions, with c>0c>0 and 2<p2m:=2mm22<p\leq 2_{m}^{\ast}:=\frac{2m}{m-2}. Then (Y) admits an infinite number of \mathcal{F}-invariant solutions, one of them is positive and has least energy among any other such solutions. That is, it attains (12) given below, and the rest are sign-changing solutions.

Remark 1.1.

We point out that if (Y) admits a constant function as solution, then this is a foliated solution and hence is obtained applying þA. In case (Y) does not admit a constant function as solution, then the solution given by þA is a solution with non trivial geometry.

Remark 1.2.

þ Observe that when bb is a positive function then the operator Δg+b-\Delta_{g}+b is coercive, since we can take

μ=min{1,inf{b(x)xM}}.\mu=\min\{1,\inf\{b(x)\mid x\in M\}\}.

As an application of \threfTheorem Main and þ1.2 we give new sign changing solutions to the Yamabe problem and a positive minimal energy solution on any compact manifold of positive scalar curvature that admits a singular foliation:

Corollary B.

þ Let \mathcal{F} be a nontrivial singular Riemannian foliation with closed leaves on a compact Riemannian manifold (M,g)(M,g). Assume that 1κ1\leqslant\kappa_{\mathcal{F}}, c>0c>0 is constant and that b>0b>0 is \mathcal{F}-invariant. Then there exists an infinite number of sign changing \mathcal{F}-invariant solutions to (Y). In particular, this is true for the Yamabe problem:

(3) Δgu+Sgu=c|u|4m2uon M,-\Delta_{g}u+S_{g}u=c|u|^{\frac{4}{m-2}}u\quad\text{on }\ M,

provided that the scalar curvature of (M,g)(M,g), b=Sg>0b=S_{g}>0 is \mathcal{F}-invariant.

Observe that for a singular Riemannian foliation with positive dimensional leaves with respect to a Riemannian metric of constant scalar curvature all the conditions in þB are satisfied. Due to the fact that the Yamabe problem has been extensively studied for the unit sphere in n\mathbb{R}^{n}, and the extensive examples of singular Riemannian foliations for this manifold (see Section 3.2) we restate the previous corollary specifically for the case of the unit sphere:

Corollary C.

þ If c>0c>0 is constant, then for any nontrivial singular Riemannian foliation \mathcal{F} on the round sphere (𝕊m,g)(\mathbb{S}^{m},g) with 1κ1\leq\kappa_{\mathcal{F}}, there exist an infinite number of sign changing \mathcal{F}-invariant solutions to the Yamabe problem:

(4) Δgu+m(m1)4u=c|u|4m2uon 𝕊m.-\Delta_{g}u+\frac{m(m-1)}{4}u=c|u|^{\frac{4}{m-2}}u\quad\text{on }\ \mathbb{S}^{m}.

þA and Corollaries B and C directly expand the results obtained in [12, 25, 24]. In [12] solutions to (Y) were found for foliations arising from closed subgroups of isometries acting on a given Riemannian manifold, meanwhile [25, 24] provided solutions for foliations arising from isoparametric functions (codimension one foliations). We point out that by the work of Radeschi [43] and of Farrell and Wu [23], the notion of a singular Riemannian foliation is more general than the one of a group action and a Riemannian submersion; that is, there are singular Riemannian foliations of arbitrary dimension which cannot arise from a group action nor from a Riemannian submersion (see Subsection 3.2 for a description of Radeschi’s examples). Hence þA gives a strict generalization of the current literature in finding sign-changing solutions to problem (Y).

We stress out that þA gives an \mathcal{F}-invariant solution with minimal energy among any other \mathcal{F}-invariant solution. In the absence of symmetries, even if the Yamabe invariant is always attained, problem (Y) may not have a ground state. That is, there may not be a solution with minimal energy, as it was shown for instance in [12, Theorem 1.5]. For the case when bb equals the scalar curvature of MM it is not clear if the energy minimizing solutions from þB attain the Yamabe constant.

Recently in [14, 15], given a compact Lie group GG acting by isometries with positive dimensional orbits, the authors give GG-invariant solutions to the Yamabe problem which have minimal energy among any other GG-invariant sign-changing solutions with a fixed number of nodal domains, by showing the existence of regular optimal GG-invariant partitions for an arbitrary number of components. This is a more controlled way of finding sign-changing solutions to the Yamabe equation, and highlights the noncompactness of the set of sign-changing solutions. We point out that for singular Riemannian foliations given by group actions it is not clear if there exists a relation between the solutions in [14, 15] and the ones given by þA.

To prove þA we state and prove a Rellich–Kondrachov embedding theorem for the subspace of foliated Sobolev functions in þH below, and a Principle of Symmetric criticality for the energy functional associated to problem (Y) in þG. This is a generalization of Palais’ Principle of Symmetric criticality [42], and also generalizes the work of Henry [33] in codimension one foliations. We point out that þH has been proven independently by Alexandrino and Cavenaghi [3]. Nonetheless there are examples of foliations for which þG still holds (see Section 3.2) but not the one in [3]. We point out that a general statement of the Palais’ Principle of Symmetric criticality for an arbitrary foliated functional is probably false in general. Nonetheless þG can be applied to the classical Yamabe equation in the context of the foliated Kazdan-Warner problem, that is, finding Riemannian metrics with scalar curvature equal to a prescribed \mathcal{F}-invariant function for a singular Riemannian foliation (M,)(M,\mathcal{F}) as in [11, Section 2]. This problem has also been considered for regular Riemannian foliations on closed manifolds in [49].

When we have two singular Riemannian foliations 1\mathcal{F}_{1} and 2\mathcal{F}_{2} on a fixed Riemannian manifold (M,g)(M,g), it is difficult to know if there is a relation between the 1\mathcal{F}_{1}-invariant and the 2\mathcal{F}_{2}-invariant sign-changing solutions to (Y) given by our method. We say that 12\mathcal{F}_{1}\subset\mathcal{F}_{2} if for any leaf L11L_{1}\in\mathcal{F}_{1} there exists a leaf L22L_{2}\in\mathcal{F}_{2} such that L1L2L_{1}\subset L_{2}. Thus we have the following question:

Question.

Consider a fixed Riemannian manifold (M,g)(M,g) and two singular Riemannian foliations 1\mathcal{F}_{1} and 2\mathcal{F}_{2} with respect to gg such that 12\mathcal{F}_{1}\subset\mathcal{F}_{2}. Does there exist a solution to (Y) which is 1\mathcal{F}_{1}-invariant but not 2\mathcal{F}_{2}-invariant?

As a particular example, note that for a fixed Riemannian manifold (M,g)(M,g) and a given group action by a compact Lie group GG via isometries on (M,g)(M,g), and a fixed closed subgroup HGH\subset G, then the HH-orbits are contained in the GG-orbits. That is we have H={H(p)pM}G={G(p)pM}\mathcal{F}_{H}=\{H(p)\mid p\in M\}\subset\mathcal{F}_{G}=\{G(p)\mid p\in M\}. Even in this homogeneous setting it is not clear how to answer the previous question. Although due to the work of Galaz-Garcia, Kell, Mondino, Sosa [27, Theorem 5.12] the subspace of GG-invariant Sobolev functions Hg1(M)GHg1(M)H^{1}_{g}(M)^{G}\subset H^{1}_{g}(M) is isometric to the Sobolev space H1(M/G,μG)H^{1}(M/G,\mu^{\ast}_{G}), where μG\mu^{\ast}_{G} is the pushfoward measure of the Riemannian volume VgV_{g} via quotient the map MM/GM\to M/G. Thus from our proof it is clear that if H1(M/H,μH)H^{1}(M/H,\mu^{\ast}_{H}) is isometric to H1(M/G,μG)H^{1}(M/G,\mu^{\ast}_{G}) then we have a negative answer to the question posted above. That is, the HH-invariant sign-changing solutions given by þA are the same as the GG-invariant solutions given by the theorem.

Our work is organized as follows. In Section 2 we give the proof of þA together with all the necessary concepts to write the proof. In Section 3 we present some preliminaries of singular Riemannian foliations, give a proof of þD, which says that Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} is an infinite dimensional subspace of Hg1(M)H_{g}^{1}(M), and finalize presenting examples of singular Riemannian foliations which are not coming from group actions or induced by isoparametric functions (i.e. are codimension one foliations). In Section 4 we prove þG, which is a kind of Principle of Symmetric Criticality for the energy functional associated to our problem, JJ. In Section 5 we give a proof of þH which is a foliated version of the Sobolev and Rellich–Kondrachov embedding theorems. We also prove þE which states that JJ satisfies some Palais-Smale condition. We end with Section 6, where, employing a variational principle for sign-changing solutions, we show the existence of arbitrarily large number of critical points for the energy functional in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}}F), which due to þG are the sign-changing solutions to (Y).

Acknowledgments

We thank Monica Clapp for useful conversations about the question posted in the introduction. We also thank Jimmy Petean for useful comments.

2. Proof of Theorem A

In this section we briefly state some definitions and results needed to prove our main theorem and we conclude the section by giving its proof.

2.1. Some words about singular Riemannian foliations

Here we define some basic things of singular Riemannian foliations, and provide more details in Section 3.


In what follows, (M,g)(M,g) will be a closed (compact and without boundary) Riemannian manifold of dimension m3m\geq 3. By a singular Riemannian foliation \mathcal{F} of (M,g)(M,g) we mean a decomposition of MM into connected injectively immersed submanifolds, called leaves, which may have different dimensions, satisfying:

  1. (i)

    \mathcal{F} is a transnormal system, i.e., every geodesic orthogonal to one leaf remains orthogonal to all the leaves that it intersects.

  2. (ii)

    \mathcal{F} is smooth, that is, for every leaf LL\in\mathcal{F} and pLp\in L,

    TpL=span{XpX𝒳},T_{p}L=\mathrm{span}\{X_{p}\mid X\in\mathcal{X}_{\mathcal{F}}\},

    where 𝒳\mathcal{X}_{\mathcal{F}} is the module of smooth vector fields on MM which are everywhere tangent to the leaves of \mathcal{F}.

Remark 2.1.

Observe that for a compact leaf, an injective immersion is the same as an embedding. Thus for foliations whose leaves are all compact, we consider the leaves embedded in MM.

Remark 2.2.

It is conjectured that (i) implies (ii), that is, a transnormal system is a smooth foliation. For some discussion on this conjecture see [50, Final Remarks (c)].

Remark 2.3.

Condition (i) is independent of (ii), since there are singular foliations satisfying (ii) but not (i). For example, in [46] a 55-dimensional manifold with a smooth foliation by 11-dimensional circles, that does not have finite holonomy is given. By [40, Theorem 2.6] any foliation with compact leaves, all of the same dimension, which admits a Riemannian metric satisfying (i) has finite holonomy. Thus, the smooth foliation given in [46] does not admit a Riemannian metric satisfying (i).

The leaves of maximal dimension are called regular and the other ones singular. If all the leaves are regular, then \mathcal{F} is called a regular Riemannian foliation. The dimension of \mathcal{F}, denoted by dim\dim\mathcal{F}, is the dimension of the regular leaves and the codimension of \mathcal{F} is equal to dimMdim\dim M-\dim\mathcal{F}.

2.2. Variational Setting

Let (M,g)(M,g) be a closed Riemannian manifold of dimension m3m\geq 3, p(2,2)p\in(2,2^{\ast}), and let \mathcal{F} be a singular Riemannian foliation on MM. In what follows, ,g\langle\cdot,\cdot\rangle_{g} will denote the Riemannian metric in MM, while gu\nabla_{g}u will denote the gradient with respect to gg of a smooth function u:Mu\colon M\to\mathbb{R}. The Sobolev space Hg1(M)H_{g}^{1}(M) is the closure of 𝒞(M)\mathcal{C}^{\infty}(M) under the norm, Hg1(M)\|\cdot\|_{H^{1}_{g}(M)}, induced by the interior product

(5) u,vHg1(M):=Mgu,gvg+uvdVg,u,v𝒞(M).\langle u,v\rangle_{H^{1}_{g}(M)}:=\int_{M}\langle\nabla_{g}u,\nabla_{g}v\rangle_{g}+uv\;dV_{g},\quad u,v\in\mathcal{C}^{\infty}(M).

By density, this bilinear form extends to a continuous symmetric bilinear form on Hg1(M)×Hg1(M).H_{g}^{1}(M)\times H_{g}^{1}(M). Even if, formally, gu,gvg\langle\nabla_{g}u,\nabla_{g}v\rangle_{g} is not defined for an arbitrary weak gradient, we will understand (5) as a limit and conserve the notation on the right hand side for every u,vHg1(M)u,v\in H_{g}^{1}(M).

Fix two functions b,cC(M)b,c\in C^{\infty}(M) with c>0c>0 and suppose that the operator Δ+b-\Delta+b is coercive (see (1)). Since Δ+b-\Delta+b is coercive, the bilinear product

(6) u,vb:=Mgu,gvg+buvdVg,u,vHg1(M)\langle u,v\rangle_{b}:=\int_{M}\langle\nabla_{g}u,\nabla_{g}v\rangle_{g}+b\,uv\ dV_{g},\qquad u,v\in H_{g}^{1}(M)

is definite positive. Thus, it is an inner product and consequently induces a norm in Hg1(M)H_{g}^{1}(M), which furthermore is equivalent to the norm Hg1(M)\|\cdot\|_{H^{1}_{g}(M)} given that MM is compact, bb is continuous and Δ+b-\Delta+b is coercive.

Since cc is positive and continuous and MM is compact, we also have a norm in the Lgp(M)L^{p}_{g}(M) space,

(7) |u|c,p:=(Mc|u|p𝑑Vg)1/puLgp(M),|u|_{c,p}:=\left(\int_{M}c|u|^{p}\;dV_{g}\right)^{1/p}\quad u\in L^{p}_{g}(M),

equivalent to the standard norm of Lgp(M)L^{p}_{g}(M). Indeed, denoting by KK the maximum of cc, and by kk its minimum, we have that Kk>0K\geqslant k>0, and thus

k1/puLgp(M)|u|c,pK1/puLgp(M).k^{1/p}\|u\|_{L^{p}_{g}(M)}\leqslant|u|_{c,p}\leqslant K^{1/p}\|u\|_{L^{p}_{g}(M)}.

The energy functional J:Hg1(M)J\colon H^{1}_{g}(M)\to\mathbb{R} given by

J(u):=12ub21p|u|c,ppJ(u):=\frac{1}{2}\|u\|_{b}^{2}-\frac{1}{p}|u|_{c,p}^{p}

is a well defined C2C^{2} functional for all p(2,2m]p\in(2,2^{\ast}_{m}], and its derivative is given by

(8) J(u)v=u,vbMc|u|p2uv𝑑Vg,u,vHg1(M).J^{\prime}(u)v=\langle u,v\rangle_{b}-\int_{M}c|u|^{p-2}uv\,dV_{g},\qquad u,v\in H^{1}_{g}(M).

Moreover, uHg1(M)u\in H^{1}_{g}(M) is a critical point of JJ if and only if uu is a solution to (Y). The nontrivial critical points of JJ lie in the Nehari manifold

(9) 𝒩g:={uHg1(M)u0,J(u)u=0}={uHg1(M)u0,ub2=|u|c,pp},\begin{split}\mathcal{N}_{g}&:=\{u\in H_{g}^{1}(M)\mid u\neq 0,J^{\prime}(u)u=0\}\\ &=\{u\in H_{g}^{1}(M)\mid u\neq 0,\|u\|_{b}^{2}=|u|_{c,p}^{p}\},\end{split}

which is a closed C1C^{1} Hilbert submanifold of Hg1(M)H^{1}_{g}(M) of codimension one and radially diffeomorphic to 𝕊Hg1(M)\mathbb{S}_{H_{g}^{1}(M)}, the unit sphere in Hg1(M)H^{1}_{g}(M) with respect to the norm b\|\cdot\|_{b}. Concretely, for each uHg1(M){0}u\in H_{g}^{1}(M)\smallsetminus\{0\}, there exists a unique tu(0,)t_{u}\in(0,\infty) such that tuu𝒩gt_{u}u\in\mathcal{N}_{g}, and it is given explicitly by

tu:=(ub2|u|c,pp)1/(p2).t_{u}:=\left(\frac{\|u\|_{b}^{2}}{|u|_{c,p}^{p}}\right)^{1/(p-2)}.

Hence, we can define a projection σ:Hg1(M){0}𝒩g\sigma\colon H_{g}^{1}(M)\smallsetminus\{0\}\to\mathcal{N}_{g} given by

(10) σ(u):=tuu=(ub2|u|c,pp)1/(p2)u,\sigma(u):=t_{u}u=\left(\frac{\|u\|_{b}^{2}}{|u|_{c,p}^{p}}\right)^{1/(p-2)}u,

which induces the radial diffeomorphism σ:𝕊Hg1(M)𝒩g\sigma\colon\mathbb{S}_{H_{g}^{1}(M)}\to\mathcal{N}_{g} with inverse uuubu\mapsto\frac{u}{\|u\|_{b}}. In addition, if u𝒩gu\in\mathcal{N}_{g}, then

(11) J(u)=max{J(tu)t0},J(u)=\max\{J(tu)\mid t\geqslant 0\},

(see [51, Lemma 4.1]).

For any singular Riemannian foliation, define the space

𝒞(M):={u𝒞(M)u is constant on L for any L}\mathcal{C}^{\infty}(M)^{\mathcal{F}}:=\{u\in\mathcal{C}^{\infty}(M)\mid u\text{ is constant on }L\text{ for any }L\in\mathcal{F}\}

and let Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} be the closure of 𝒞(M)\mathcal{C}^{\infty}(M)^{\mathcal{F}} under the norm Hg1(M)\|\cdot\|_{H^{1}_{g}(M)}. Hence Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} is a closed subspace of Hg1(M)H_{g}^{1}(M) and uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} if and only if uHg1(M)u\in H_{g}^{1}(M) and is dVgdV_{g}-a.e. constant on the leaves of \mathcal{F}.

For 1κ<m1\leqslant\kappa_{\mathcal{F}}<m with κ\kappa_{\mathcal{F}} defined as in (2) we show in þD that Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} is an infinite dimensional Hilbert space endowed with the restriction of the interior product ,Hg1(M)\langle\cdot,\cdot\rangle_{H^{1}_{g}(M)}. The critical points of JJ that are contained in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} will be called \mathcal{F}-invariant critical points of JJ. Hence, the nontrivial \mathcal{F}-invariant critical points of JJ lie in the restricted Nehari manifold

𝒩g:=𝒩gHg1(M),\mathcal{N}_{g}^{\mathcal{F}}:=\mathcal{N}_{g}\cap H_{g}^{1}(M)^{\mathcal{F}},

which is nonempty by virtue of Theorem D below and the definition of the projection (10). The least energy of the functions contained in the restricted Nehari manifold is given by

(12) τg:=inf{J(u)u𝒩g},\tau_{g}^{\mathcal{F}}:=\inf\{J(u)\mid u\in\mathcal{N}_{g}^{\mathcal{F}}\},

which is a positive number (see [51, Theorem 4.2]). We recall that the positive solution obtained in þA will attain this number, and thus it will be a solution of least energy.

2.3. Main results to prove Theorem A and proof of Theorem A

The main ingredients for the proof of þA are þD, a compactness result (þE), a Principle of Symmetric Criticality for the functional JJ restricted to the space Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}G), and a variational principle (þF), which we state below.


First we notice that Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} is an infinite dimensional space, see Section 3 for the proof. Recall that κ:=min{dimLL}\kappa_{\mathcal{F}}:=\min\{\dim L\mid L\in\mathcal{F}\}.

Theorem D.

þ Let (M,g,)(M,g,\mathcal{F}) be an mm-dimensional closed Riemannian manifold together with a nontrivial singular Riemannian foliation with 1κ1\leq\kappa_{\mathcal{F}}. Then, for any kk\in\mathbb{N}, there exist nontrivial nonnegative functions u1,,uk𝒞(M)u_{1},\ldots,u_{k}\in\mathcal{C}^{\infty}(M)^{\mathcal{F}} with pairwise disjoint supports. In particular, it follows that Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} is infinite dimensional.

Given τ\tau\in\mathbb{R}, we say that a sequence (un)n(u_{n})_{n\in\mathbb{N}} in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} is a (PS)τ(PS)^{\mathcal{F}}_{\tau}-sequence for JJ if J(un)τJ(u_{n})\to\tau and J(un)0J^{\prime}(u_{n})\to 0 in (Hg1(M))(H_{g}^{1}(M)^{\mathcal{F}})^{\ast}, where the last space is the dual of Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. We say that JJ satisfies the (PS)τ(PS)^{\mathcal{F}}_{\tau}-condition in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} if every (PS)τ(PS)^{\mathcal{F}}_{\tau}-sequence for JJ has a strongly convergent subsequence in Hg1(M)H^{1}_{g}(M). With these definitions, the statement of the compactness results is as follows.

Theorem E (Compactness).

þ Under the hypotheses of þA, except that we also allow bb and cc to be non foliated, the functional JJ satisfies the (PS)τ(PS)^{\mathcal{F}}_{\tau}-condition for every τ\tau\in\mathbb{R}.

We prove this result in Section 5. This is a generalization of a classic result by E. Hebey and M. Vaugon [31] and a consequence of a Sobolev embedding theorem for singular Riemannian foliations (þH). We believe this result is interesting in its own right, since previous versions of it have been applied to restore compactness in many variational problems involving symmetries given by compact groups actions via isometries (see for instance, [12, 7, 16]).

With this at hand, in Section 6 we will be able to generalize Clapp-Pacella’s variational principle in [13] (see also [12, 17]).

Theorem F (Variational Principle).

þ Under the hypotheses of Theorem A, let WW be a nontrivial finite dimensional subspace of Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. If JJ satisfies the (PS)τ(PS)^{\mathcal{F}}_{\tau} condition in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} for every τsupWJ\tau\leq\sup_{W}J, then JJ has at least one positive critical point u1u_{1} and k1:=dimW1k-1:=\dim W-1 pairs of sign-changing critical points ±u2,,±uk\pm u_{2},\ldots,\pm u_{k} in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} such that J(u1)=τgJ(u_{1})=\tau_{g}^{\mathcal{F}} and J(ui)supWJJ(u_{i})\leq\sup_{W}J for i=2,,ki=2,\ldots,k.

The Principle of Symmetric Criticality needed to prove our main theorem is the following.

Theorem G (Principle of Symmetric Criticality for JJ).

þ Under the hypotheses of Theorem A, if uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} is a critical point of the functional JJ restricted to Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, then uu is a critical point of JJ in the space Hg1(M)H_{g}^{1}(M).

The proof of this result appears in Section 4. In [33], a similar argument is used for the case of foliations given by isoparametric functions.


We now prove þA in an analogous fashion to the first Main Result in [12].

Proof of Theorem A.

In order to find nontrivial foliated critical points of JJ, by Theorem G, it suffices to consider critical points of the restriction of JJ to the space Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. To do so, we proceed as follows. By Theorem D, since we are assuming that the singular Riemannian foliation is nontrivial, for any given kk\in\mathbb{N}, we may choose kk nontrivial functions ωi𝒞(M)\omega_{i}\in\mathcal{C}^{\infty}(M)^{\mathcal{F}} i=1,,ki=1,\ldots,k with disjoint supports. Let WW be the linear subspace of Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} spanned by {ω1,,ωk}\{\omega_{1},\ldots,\omega_{k}\}. Since ωi\omega_{i} and ωj\omega_{j} have disjoint supports for ij,i\neq j, the set {ω1,,ωk}\{\omega_{1},\dots,\omega_{k}\} is orthogonal in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. Hence, dimW=k.\dim W=k.

Now, as κ1\kappa_{\mathcal{F}}\geq 1, Theorem E yields that JJ satisfies (PS)τ(PS)^{\mathcal{F}}_{\tau} in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} for every τ\tau\in\mathbb{R}. Therefore, we can apply Theorem F and get at least one positive and (k1)(k-1) sign-changing \mathcal{F}-invariant critical points for the restriction of JJ to Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. Theorem G yields that these are also critical points of JJ in Hg1(M)H_{g}^{1}(M) and, hence nontrivial \mathcal{F}-invariant solutions to problem (Y). Given that kk\in\mathbb{N} is arbitrary, we conclude that there are infinitely many sign-changing solutions. ∎

3. Singular Riemannian Foliations

In the first part of this section we provide the material needed in Section 5 to show a Sobolev embedding theorem (Theorem H), such as the existence of smooth and \mathcal{F}-invariant partitions of unity and the Slice Theorem. Furthermore, we prove that Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} is infinite dimensional (Theorem D). In the second part we give examples of singular Riemannian foliations. In particular, we present examples for which previous methods to find solutions to (Y) such as those [12, 25, 24, 3] do not apply.

3.1. Singular Riemannian Foliations Revisited


Let \mathcal{F} be a singular Riemannian foliation with closed leaves on a closed Riemannian manifold (Mm,g)(M^{m},g). Given qMq\in M, denote by LqL_{q} the leaf containing qq, by kqk_{q} the dimension of LqL_{q} and by πq:kq×mkqmkq\pi_{q}\colon\mathbb{R}^{k_{q}}\times\mathbb{R}^{m-k_{q}}\to\mathbb{R}^{m-k_{q}} the natural projection. The following lemma describes the foliation in a small neighborhood of qq.

Lemma 3.1 (Proposition 2.17 in [44], Figure 1).

þ For any qMq\in M, there exists a coordinate system (W,φ)(W,\varphi) such that

  1. (1)

    qWq\in W, φ(W)=U×V\varphi(W)=U\times V, where UkqU\subset\mathbb{R}^{k_{q}} and VmkqV\subset\mathbb{R}^{m-k_{q}} are open and bounded subsets with smooth boundary;

  2. (2)

    for any qWq^{\prime}\in W, U×πq(φ(q))φ(LqW)U\times\pi_{q}(\varphi(q^{\prime}))\subset\varphi(L_{q^{\prime}}\cap W).

Refer to caption
Figure 1. Local trivialization of a foliation

Consider pMp\in M fixed. Take ε>0\varepsilon>0 such that expp:Bε(0)TpMM\operatorname{exp}_{p}\colon B_{\varepsilon}(0)\subset T_{p}M\to M is a diffeomorphism. Denote by 𝔻p(ε)νp(M,Lp)\mathbb{D}^{\perp}_{p}(\varepsilon)\subset\nu_{p}(M,L_{p}) the closed ball of radius ε\varepsilon in the normal space to the leaf at pp. Consider Sp(ε)=expp(𝔻p(ε))S_{p}(\varepsilon)=\operatorname{exp}_{p}(\mathbb{D}^{\perp}_{p}(\varepsilon)). For qSp(ε)q\in S_{p}(\varepsilon) denote by L~q\tilde{L}_{q} the connected component of LqSp(ε)L_{q}\cap S_{p}(\varepsilon) containing qq. Since ε\varepsilon is smaller than the injectivity radius at pp, then qSp(ε)q\in S_{p}(\varepsilon) if and only if q=expp(v)q=\operatorname{exp}_{p}(v) for a unique v𝔻p(ε)v\in\mathbb{D}^{\perp}_{p}(\varepsilon). For such qq we set v=expp1(L~q)\mathcal{L}_{v}=\operatorname{exp}_{p}^{-1}(\tilde{L}_{q}). The following theorem states that the partition of 𝔻p(ε)\mathbb{D}^{\perp}_{p}(\varepsilon) by p(ε)={vv𝔻p(ε)}\mathcal{F}_{p}^{\perp}(\varepsilon)=\{\mathcal{L}_{v}\mid v\in\mathbb{D}^{\perp}_{p}(\varepsilon)\} is a singular Riemannian foliation.

Theorem 3.2 (Infinitesimal Foliation, Proposition 6.5 in [41]).

þ Let (M,)(M,\mathcal{F}) be a singular Riemannian foliation on the compact manifold MM and fix pMp\in M. Take ε>0\varepsilon>0 smaller than the injectivity radius of MM at pp. Then (𝔻p(ε),p(ε))(\mathbb{D}_{p}^{\perp}(\varepsilon),\mathcal{F}_{p}^{\perp}(\varepsilon)) equipped with the Euclidean metric is a singular Riemannian foliation.

Moreover, (𝔻p(ε),p(ε))(\mathbb{D}_{p}^{\perp}(\varepsilon),\mathcal{F}_{p}^{\perp}(\varepsilon)) does not depend on the radius chosen.

Lemma 3.3 ([44]).

þ Let ε>0\varepsilon>0 and λ>0\lambda>0 be such that ε\varepsilon and λε\lambda\varepsilon are smaller than the injectivity radius of MM at pp. Then the map hλ:𝔻p(ε)𝔻p(λε)h_{\lambda}\colon\mathbb{D}^{\perp}_{p}(\varepsilon)\to\mathbb{D}^{\perp}_{p}(\lambda\varepsilon) given by hλ(v)=λvh_{\lambda}(v)=\lambda v is a foliated diffeomorphism between (𝔻p(ε),p(ε))(\mathbb{D}_{p}^{\perp}(\varepsilon),\mathcal{F}_{p}^{\perp}(\varepsilon)) and (𝔻p(λε),p(λε))(\mathbb{D}_{p}^{\perp}(\lambda\varepsilon),\mathcal{F}_{p}^{\perp}(\lambda\varepsilon))

The previous result implies that for ε>0\varepsilon>0 small enough p(ε)\mathcal{F}_{p}^{\perp}(\varepsilon) is independent of ε\varepsilon, and thus we set p=p(ε)\mathcal{F}_{p}^{\perp}=\mathcal{F}_{p}^{\perp}(\varepsilon). We call p\mathcal{F}_{p}^{\perp} the infinitesimal foliation at pp.

Observe that the image under expp\operatorname{exp}_{p} of different leaves of p\mathcal{F}_{p}^{\perp} may be contained in the same leaf of \mathcal{F}. To account for this, we define the holonomy group of a leaf.

Lemma 3.4 (See [18, 39, 44]).

þ Under the hypotheses of Theorem 3.2 and denoting by LL the leaf of \mathcal{F} that contains pp, the following statements hold:

  1. a)

    Given a piece-wise smooth curve γ:[0,1]L\gamma\colon[0,1]\to L starting at pp there exists a map G(γ)(t):𝔻p(ε)𝔻γ(t)(ε)G(\gamma)(t)\colon\mathbb{D}^{\perp}_{p}(\varepsilon)\to\mathbb{D}^{\perp}_{\gamma(t)}(\varepsilon) which is a foliated isometry between (𝔻p(ε),p)(\mathbb{D}^{\perp}_{p}(\varepsilon),\mathcal{F}_{p}^{\perp}) and (𝔻γ(t),γ(t))(\mathbb{D}^{\perp}_{\gamma(t)},\mathcal{F}^{\perp}_{\gamma(t)}).

  2. b)

    The set

    Hol(L,p)={G(γ)(1)γ is a closed loop in L centered at p}\mathrm{Hol}(L,p)=\{G(\gamma)(1)\mid\gamma\mbox{ is a closed loop in }L\mbox{ centered at }p\}

    is a group under the composition operation.

  3. c)

    Consider pqLp\neq q\in L, and α:[0,1]L\alpha\colon[0,1]\to L a piece-wise curve starting at pp and ending at qq. Then

    Hol(L,p)=G(α)(1)1Hol(L,q)G(α)(1).\mathrm{Hol}(L,p)=G(\alpha)(1)^{-1}\mathrm{Hol}(L,q)G(\alpha)(1).
  4. d)

    Given two piece-wise smooth curves γ1\gamma_{1} and γ2\gamma_{2} starting at pp and homotopic relative to its end points, the composition of their corresponding foliated isometries, G(γ2)(1)1G(γ1)(1)G(\gamma_{2})(1)^{-1}\circ G(\gamma_{1})(1), is homotopic to the identity map.

The group Hol(L,p)\mathrm{Hol}(L,p) given in þ3.4 is called the holonomy group of the leaf LL at pp. For any other qLq\in L due to c)c) and d)d) the groups Hol(L,p)\mathrm{Hol}(L,p) and Hol(L,q)\mathrm{Hol}(L,q) are conjugates, and we denote their conjugacy class as Hol(L)\mathrm{Hol}(L). The regular leaves LL of \mathcal{F} for whom Hol(L)=[{Id}]\mathrm{Hol}(L)=[\{Id\}] are called principal leaves. We denote by MprinMM_{\mathrm{prin}}\subset M the set of points contained in the principal leaves. We remark that this is an open and dense subset of MM.

In order to prove that Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} is infinite dimensional, we define the leaf space of (M,g,)(M,g,\mathcal{F}) to be the quotient space induced by the partition \mathcal{F} equipped with the quotient topology and denote it by M=M/M^{\ast}=M/\mathcal{F}. We denote the quotient map as π:MM/\pi\colon M\to M/\mathcal{F}. Given a subset AMA\subset M, we denote its image π(A)\pi(A) by AA^{\ast}.

Proof of þD.

Let (M,)(M,\mathcal{F}) be a singular Riemannian foliation with closed leaves on the complete manifold MM. Then the leaf space M/M/\mathcal{F} is a complete length metric space. Thus for any open cover {U}\{U^{\ast}\} of M/M/\mathcal{F} we have a subordinate partition of unit ϕi:M/\phi_{i}^{\ast}\colon M/\mathcal{F}\to\mathbb{R} such that supp(ϕi)Ui\operatorname{supp}(\phi_{i}^{\ast})\subset U_{i}^{\ast}. We can refine the cover {Ui}\{U_{i}^{\ast}\} so that for any kk\in\mathbb{N}, at least kk open neighborhoods of the cover are disjoint. Taking π:MM/\pi\colon M\to M/\mathcal{F} and setting uj=ϕjπu_{j}=\phi_{j}^{\ast}\circ\pi we get the kk desired continuous foliated functions with disjoint support.

This approach can be improved to get smooth foliated functions. We can restrict π:MM/\pi\colon M\to M/\mathcal{F} to MprinMM_{\mathrm{prin}}\subset M and Mprin=Mprin/M/M_{\mathrm{prin}}^{\ast}=M_{\mathrm{prin}}/\mathcal{F}\subset M/\mathcal{F}, so that π:MprinMprin\pi\colon M_{\mathrm{prin}}\to M_{\mathrm{prin}}^{\ast}. We endow MprinM_{\mathrm{prin}} with g|Mpring|_{M_{\mathrm{prin}}} the induced metric by the inclusion into MM. Then (Mprin,g|Mprin,|Mprin)(M_{\mathrm{prin}},g|_{M_{\mathrm{prin}}},\mathcal{F}|_{M_{\mathrm{prin}}}) is a regular Riemannian foliation. Moreover the leaf space MprinM^{\ast}_{\mathrm{prin}} is a manifold. By Theorem [30] there is a Riemannian metric gpring^{\ast}_{\mathrm{prin}} on MprinM^{\ast}_{\mathrm{prin}} such that the map π:(Mprin,g|Mprin)(Mprin,gprin)\pi\colon(M_{\mathrm{prin}},g|_{M_{\mathrm{prin}}})\to(M^{\ast}_{\mathrm{prin}},g^{\ast}_{\mathrm{prin}}) is actually a Riemannian submersion.

Then by restricting an open cover {Ui}\{U^{\ast}_{i}\} of M/M/\mathcal{F} to MprinM_{\mathrm{prin}}^{\ast}, we get a smooth partition of unity {ϕi:Mprin}\{\phi_{i}^{\ast}\colon M_{\mathrm{prin}}^{\ast}\to\mathbb{R}\} subordinated to the cover {UiMprin}\{U_{i}^{\ast}\cap M_{\mathrm{prin}}^{\ast}\}. Setting ϕi=ϕiπ\phi_{i}=\phi_{i}^{\ast}\circ\pi gives us a smooth partition of unity subordinated to the open cover {π1(UiMprin)}\{\pi^{-1}(U_{i}^{\ast}\cap M_{\mathrm{prin}}^{\ast})\} of MprinM_{\mathrm{prin}}; observe that by construction these functions are constant along the leaves. We can refine this open cover so that for any given kk\in\mathbb{N}, there are at least kk functions ϕ1,,ϕk\phi_{1},\ldots,\phi_{k} which have disjoint support, contained in MprinM_{\mathrm{prin}}. Setting Ui=π1(UiMprin)U_{i}=\pi^{-1}(U_{i}^{\ast}\cap M_{\mathrm{prin}}^{\ast}), we extend these functions trivially on M(U1Uk)M\smallsetminus(U_{1}\cup\ldots\cup U_{k}) to get the desired smooth functions uiu_{i}, constant along the leaves with disjoint support. ∎

The following theorem states how a foliated tubular neighborhood looks like:

Theorem 3.5 (Slice Theorem in [39]).

þ Let (M,)(M,\mathcal{F}) be a singular Riemannian foliation with closed leaves. Given a leaf LL\in\mathcal{F} and pLp\in L, denote by Tubε(L)\mathrm{Tub}^{\varepsilon}(L) the tubular neighborhood of radius ε\varepsilon of LL. Then there exist ε>0\varepsilon>0 small enough and a Hol(L,p)\mathrm{Hol}(L,p)-principal bundle PLpP\to L_{p}, such that (Tubε(L),Tubε(L))(\mathrm{Tub}^{\varepsilon}(L),\mathcal{F}\cap\mathrm{Tub}^{\varepsilon}(L)) is foliated diffeomorphic to

(P×Hol(L,p)𝔻p(ε),P×Hol(L,p)p).(P\times_{\mathrm{Hol}(L,p)}\mathbb{D}^{\perp}_{p}(\varepsilon),P\times_{\mathrm{Hol}(L,p)}\mathcal{F}_{p}^{\perp}).

Here P×Hol(L,p)𝔻(ε)P\times_{\mathrm{Hol}(L,p)}\mathbb{D}^{\perp}(\varepsilon) denotes the quotient (P×𝔻(ε))/Hol(L,p)(P\times\mathbb{D}^{\perp}(\varepsilon))/\mathrm{Hol}(L,p) with respect to the product action of Hol(L,p)\mathrm{Hol}(L,p) on P×𝔻(ε)P\times\mathbb{D}^{\perp}(\varepsilon), i.e. h(p,v)=(ph,h1v)h(p,v)=(ph,h^{-1}v).

Given [x,v]P×Hol(L)𝔻p(ε)[x,v]\in P\times_{\mathrm{Hol}(L)}\mathbb{D}^{\perp}_{p}(\varepsilon) the leaf L[x,v]L_{[x,v]} of P×Hol(L)pP\times_{\mathrm{Hol}(L)}\mathcal{F}_{p}^{\perp} consists of all the classes [y,w]P×Hol(L)𝔻p(ε)[y,w]\in P\times_{\mathrm{Hol}(L)}\mathbb{D}^{\perp}_{p}(\varepsilon) such that (yh,h1w)P×v(yh,h^{-1}w)\in P\times\mathcal{L}_{v} for some hHol(L)h\in\mathrm{Hol}(L).

Lemma 3.6.

þ Let (M,)(M,\mathcal{F}) be a closed foliation on a compact manifold. Then for any ε>0\varepsilon>0 there exists an open cover {Zi}iΛ\{Z_{i}\}_{i\in\Lambda} of MM such that each ZiZ_{i} is a tubular neighborhood of radius at most ε\varepsilon and as in þ3.5 such that there exists a finite subcover {Zi}i=1N\{Z_{i}\}_{i=1}^{N} and a smooth partition of unity {ϕi}i=1N\{\phi_{i}\}^{N}_{i=1} adapted to the subcover which is \mathcal{F}-invariant; i.e. for each pMp\in M if qLpq\in L_{p} then ϕi(q)=ϕi(p)\phi_{i}(q)=\phi_{i}(p).

Proof.

For each pMp\in M consider ε(p)>0\varepsilon(p)>0 such that þ3.5 holds for the tubular neighborhood of LpL_{p} of radius ε(p)\varepsilon(p). If ε(p)>ε\varepsilon(p)>\varepsilon then we replace ε(p)\varepsilon(p) by ε\varepsilon, and note that þ3.5 still holds for this value. The collection {Tubε(p)(Lp)pM}\{\mathrm{Tub}^{\varepsilon(p)}(L_{p})\mid p\in M\} is an open cover of MM and since MM is compact there exist Zi=Tubε(pi)(Lpi)Z_{i}=\mathrm{Tub}^{\varepsilon(p_{i})}(L_{p_{i}}), i=1,,Ni=1,\ldots,N, such that M=i=1NZiM=\cup_{i=1}^{N}Z_{i}.

For each 1iN1\leqslant i\leqslant N, let ψi:ZiPi×Holi𝔻i\psi_{i}\colon Z_{i}\to P_{i}\times_{\mathrm{Hol}_{i}}\mathbb{D}^{\perp}_{i} be the foliated diffeomorphism given by þ3.5. We will define a (Pi×Holii)(P_{i}\times_{\mathrm{Hol}_{i}}\mathcal{F}_{i})-foliated function ϕ¯i:Pi×Holi𝔻i\bar{\phi}_{i}\colon P_{i}\times_{\mathrm{Hol}_{i}}\mathbb{D}^{\perp}_{i}\to\mathbb{R} and then define functions fi:Mf_{i}\colon M\to\mathbb{R}

fi(q)={ϕ¯iψi(q)qZi0qZi,f_{i}(q)=\begin{cases}\bar{\phi}_{i}\circ\psi_{i}(q)&q\in Z_{i}\\ 0&q\not\in Z_{i},\end{cases}

so that ϕi:M\phi_{i}\colon M\to\mathbb{R} defined as

ϕi(q)=fi(q)j=1Nfj(q)qM,\phi_{i}(q)=\frac{f_{i}(q)}{\sum_{j=1}^{N}f_{j}(q)}\quad q\in M,

will be the desired partition of unity. Observe that fif_{i} is \mathcal{F}-invariant since it is a composition of a foliated diffeomorphism with an \mathcal{F}-invariant function, and thus ϕi\phi_{i} is an \mathcal{F}-invariant function.

For each 1iN1\leqslant i\leqslant N, to define ϕ¯i\bar{\phi}_{i}, first consider a smooth nonnegative decreasing function φ~i:\tilde{\varphi}_{i}\colon\mathbb{R}\to\mathbb{R} such that φ~i(t)=1\tilde{\varphi}_{i}(t)=1 for t0t\leqslant 0, φ~i(t)=0\tilde{\varphi}_{i}(t)=0 for tε(pi)t\geqslant\varepsilon(p_{i}). Let φ¯i:𝔻pi(ε(pi))\overline{\varphi}_{i}\colon\mathbb{D}^{\perp}_{p_{i}}(\varepsilon(p_{i}))\to\mathbb{R} be given by φ¯i(v)=φ~i(v)\overline{\varphi}_{i}(v)=\tilde{\varphi}_{i}(\|v\|). Observe that over the spheres in 𝔻pi(ε(pi))\mathbb{D}^{\perp}_{p_{i}}(\varepsilon(p_{i})) centered at 0, φ¯i\overline{\varphi}_{i} is constant. Since the leaves of pi\mathcal{F}_{p_{i}}^{\perp} are contained in such spheres we conclude that φ¯i\overline{\varphi}_{i} is pi\mathcal{F}_{p_{i}}^{\perp}-foliated.

Set Holi=Hol(Lpi,pi)\mathrm{Hol}_{i}=\mathrm{Hol}(L_{p_{i}},p_{i}), 𝔻i=𝔻pi(ε(pi))\mathbb{D}^{\perp}_{i}=\mathbb{D}^{\perp}_{p_{i}}(\varepsilon(p_{i})), i=pi\mathcal{F}_{i}=\mathcal{F}_{p_{i}}^{\perp} and let PiP_{i} be the total space of the principal Holi\mathrm{Hol}_{i}-bundle from þ3.5. Define

ϕ~i:Pi×𝔻iasϕ~i(x,v)=φ¯i(v).\tilde{\phi}_{i}\colon P_{i}\times\mathbb{D}_{i}^{\perp}\to\mathbb{R}\quad\textrm{as}\quad\tilde{\phi}_{i}(x,v)=\bar{\varphi}_{i}(v).

Now note that for any hHolih\in\mathrm{Hol}_{i} we have ϕ~i(h(x,v))=ϕ~i(x,v)\tilde{\phi}_{i}(h(x,v))=\tilde{\phi}_{i}(x,v). Indeed, this follows from the fact that h1v=v\|h^{-1}v\|=\|v\| as can be seen below,

ϕ~i(h(x,v))=\displaystyle\tilde{\phi}_{i}(h(x,v))= ϕ~i(xh,h1v)=φ¯i(h1v)=φ~i(h1v)\displaystyle\tilde{\phi}_{i}(xh,h^{-1}v)=\bar{\varphi}_{i}(h^{-1}v)=\tilde{\varphi}_{i}(\|h^{-1}v\|)
=\displaystyle= φ~i(v)=φ¯i(v)=ϕ~i(x,v).\displaystyle\tilde{\varphi}_{i}(\|v\|)=\bar{\varphi}_{i}(v)=\tilde{\phi}_{i}(x,v).

Thus we have a well defined map

ϕ¯i:Pi×Holi𝔻iasϕi¯[p,v]=φi¯(v),\bar{\phi}_{i}\colon P_{i}\times_{\mathrm{Hol}_{i}}\mathbb{D}^{\perp}_{i}\to\mathbb{R}\quad{as}\quad\bar{\phi_{i}}[p,v]=\bar{\varphi_{i}}(v),

and we will check using local trivializations that it is also smooth.

Consider the projection map πi:PiLpi\pi_{i}\colon P_{i}\to L_{p_{i}} of the principal Holi\mathrm{Hol}_{i}-bundle, and denote by π¯i:Pi×Holi𝔻iLpi\bar{\pi}_{i}\colon P_{i}\times_{\mathrm{Hol}_{i}}\mathbb{D}^{\perp}_{i}\to L_{p_{i}} the projection of the associated bundle. Fix qLpiq\in L_{p_{i}}, and consider UαiLpiU_{\alpha i}\subset L_{p_{i}} a sufficiently small open neighborhood of qq, such that there exists a trivialization ταi:Uαi×Holiπi1(Uαi)Pi\tau_{\alpha i}\colon U_{\alpha i}\times\mathrm{Hol}_{i}\to\pi_{i}^{-1}(U_{\alpha i})\subset P_{i} of πi\pi_{i}. Observe that the map τ¯αi:Uαi×𝔻iπ¯i1(Uαi)Pi×Holi𝔻i\bar{\tau}_{\alpha i}\colon U_{\alpha i}\times\mathbb{D}^{\perp}_{i}\to\bar{\pi}_{i}^{-1}(U_{\alpha i})\subset P_{i}\times_{\mathrm{Hol}_{i}}\mathbb{D}^{\perp}_{i} given by

τ¯αi(p,v)=[ταi(p,Id),v],\bar{\tau}_{\alpha i}(p,v)=[\tau_{\alpha i}(p,\mathrm{Id}),v],

where IdHoli\mathrm{Id}\in\mathrm{Hol}_{i} denotes the identity element, is a trivialization of the associated bundle π¯i\bar{\pi}_{i}. Then we have that ϕ¯\bar{\phi} is given over a local trivialization as

ϕ¯iτ¯αi(p,v)=φ¯i(v).\bar{\phi}_{i}\circ\bar{\tau}_{\alpha i}(p,v)=\bar{\varphi}_{i}(v).

Therefore ϕ¯i:Pi×Holi𝔻i\bar{\phi}_{i}\colon P_{i}\times_{\mathrm{Hol}_{i}}\mathbb{D}^{\perp}_{i}\to\mathbb{R} can be written as the composition of smooth maps and so it is also smooth.

We check now that ϕ¯i\bar{\phi}_{i} is (Pi×Holii)(P_{i}\times_{\mathrm{Hol}_{i}}\mathcal{F}_{i})-foliated. Consider [x,v]Pi×Holi𝔻i[x,v]\in P_{i}\times_{\mathrm{Hol}_{i}}\mathbb{D}^{\perp}_{i}, and take [y,w]L[x,v][y,w]\in L_{[x,v]}. Then there exists some hHolih\in\mathrm{Hol}_{i} such that hwvhw\in\mathcal{L}_{v}. Then

ϕ¯i[y,w]\displaystyle\bar{\phi}_{i}[y,w] =ϕ~i(h(y,w))=ϕ~i(hy,hw)\displaystyle=\tilde{\phi}_{i}(h(y,w))=\tilde{\phi}_{i}(hy,hw)
=φ¯i(hw)=φ¯i(v)\displaystyle=\bar{\varphi}_{i}(hw)=\bar{\varphi}_{i}(v)
=ϕ~i(x,v)=ϕ¯i[x,v].\displaystyle=\tilde{\phi}_{i}(x,v)=\bar{\phi}_{i}[x,v].

This concludes the construction of ϕ¯i\bar{\phi}_{i} and thus the proof of the lemma. ∎

3.2. Examples of singular Riemannian foliations

Here we present in detail some examples of singular Riemannian foliations already mentioned in the introduction.

Riemmannian submersions: Given a Riemannian submersion π:MB\pi\colon M\to B we can define a singular Riemannian foliation where the foliation consists of the set of preimages under π\pi of the image of π\pi. We give more details below.

Recall that a surjective differentiable map π:MB\pi\colon M\to B between smooth manifolds is a submersion if at any point pMp\in M, the differential map Dpπ:TpMTpBD_{p}\pi\colon T_{p}M\to T_{p}B is a surjective linear map. This implies that the dimension of MM is greater than or equal to the dimension of BB. From now on assume that the dimension of MM is strictly greater than the dimension of BB. Assume that MM has a Riemannian metric gg, and for any pMp\in M denote by V(p)V(p) the subspace of TpMT_{p}M tangent at pp to the fiber Lp=π1(π(p))L_{p}=\pi^{-1}(\pi(p)) of π\pi through pp, and let VV to be the subbunddle over MM with fiber equal to V(p)V(p). We say that π\pi is a Riemannian submersion if for any XVX\in V the Lie derivative of gg in the direction of XX satisfies Xg=0\mathcal{L}_{X}g=0, that is, the metric gg is invariant in the directions tangent to the fibers. Setting H(p)H(p) to be the gg-orthogonal complement of V(p)V(p) in TpMT_{p}M, this implies that for any U,VH(p)U,V\in H(p) the inner product given as

hπ(p)(πU,πV)=gp(U,V)h_{\pi(p)}(\pi_{\ast}U,\pi_{\ast}V)=g_{p}(U,V)

is a Riemannian metric on BB.

Homogeneous foliations: Another large family of examples stem from a compact Lie group GG acting by isometries on a given Riemannian manifold (M,g)(M,g). Then, for the partition consisting of the set of orbits of the action of GG on MM, \mathcal{F}, we have that (M,g,)(M,g,\mathcal{F}) is a singular Riemannian foliation which is known as a homogeneous foliation.

RFKM-foliations: By the work of Ferus, Karcher, and Münzner in [26] there is an infinity of non-homogeneous closed codimension 11 foliations on round spheres 𝕊kk+1\mathbb{S}^{k}\subset\mathbb{R}^{k+1} given by Clifford systems. This construction was generalized by Radeschi in [43] and we present some details here. We recall that solutions to (Y) such as [12, 25, 24] do not apply to the foliations in [43] while our main theorems do apply.

Given a real vector space VV of dimension mm, equipped with a positive definite inner-product ,\langle\,,\rangle the Clifford algebra C(V)C\ell(V) is the quotient of the tensor algebra T(V)T(V) by the ideal generated by xx+yy2x,y1x\otimes x+y\otimes y-2\langle x,y\rangle 1, where 11 is the unit element in T(V)T(V). The vector space VV embeds naturally into C(V)C\ell(V). A representation of a Clifford algebra C(V)C\ell(V) is an algebra homomorphism ρ:C(V)End(n)\rho\colon C\ell(V)\to\mathrm{End}(\mathbb{R}^{n}). A Clifford system is the restriction of ρ\rho to VC(V)V\subset C\ell(V), and we denote by ρ\mathbb{R}_{\rho} the image ρ(V)\rho(V). Given a Clifford system we can find an inner product on n\mathbb{R}^{n}, such that for every vVv\in V, the matrix ρ(v)\rho(v) is a symmetric matrix. We endow the space of all symmetric matrices Sym2(n)\mathrm{Sym}^{2}(\mathbb{R}^{n}) with the inner product A,B=(1/n)tr(AB)\langle A,B\rangle=(1/n)\mathrm{tr}(AB). With these choices of inner products on n\mathbb{R}^{n} and Sym2(n)\mathrm{Sym}^{2}(\mathbb{R}^{n}) the map ρ:VρSym2(n)\rho\colon V\to\mathbb{R}_{\rho}\subset\mathrm{Sym}^{2}(\mathbb{R}^{n}) is an isometry onto its image.

Remark 3.7.

Given a Clifford system ρ:C(V)End(n)\rho\colon C\ell(V)\to\mathrm{End}(\mathbb{R}^{n}), the dimension nn is even (see [43]).

Consider the map πρ:𝕊n1ρ\pi_{\rho}\colon\mathbb{S}^{n-1}\to\mathbb{R}_{\rho} that takes x𝕊n1x\in\mathbb{S}^{n-1} to the unique element πρ(x)ρ\pi_{\rho}(x)\in\mathbb{R}_{\rho} which for all PρP\in\mathbb{R}_{\rho} satisfies

πρ(x),P=P(x),x.\langle\pi_{\rho}(x),P\rangle=\langle P(x),x\rangle.

The image of this map is contained in the unit disk of ρ\mathbb{R}_{\rho}, denoted by 𝔻ρ\mathbb{D}_{\rho} (see [43, Proposition 2.4]).

Proposition 3.8 (Proposition 2.6 in [43]).

þ The set ρ={πρ1(P)P𝔻ρ}\mathcal{F}_{\rho}=\{\pi_{\rho}^{-1}(P)\mid P\in\mathbb{D}_{\rho}\} is a singular Riemannian foliation on the unit sphere 𝕊n1\mathbb{S}^{n-1} with the round metric.

We refer to the foliation ρ\mathcal{F}_{\rho} in þ3.8 as an RFKM-foliation. As proven in [43, Section 5], only a finite number of such foliations are homogeneous.

Given a Clifford system ρ:VC(V)ρ\rho\colon V\subset C\ell(V)\to\mathbb{R}_{\rho}, for the map f:ρf\colon\mathbb{R}_{\rho}\to\mathbb{R} defined as f(P)=12P2f(P)=1-2\|P\|^{2}, the preimages of the composition fπρ:𝕊n1f\circ\pi_{\rho}\colon\mathbb{S}^{n-1}\to\mathbb{R} induce a singular Riemannian foliation of codimension 11 on the round sphere 𝕊n1\mathbb{S}^{n-1}. These are the singular Riemannian foliations FKM\mathcal{F}_{\mathrm{FKM}} described in [26].

In general, given two singular Riemannian foliations 1\mathcal{F}_{1} and 2\mathcal{F}_{2} on a given Riemannian manifold (M,g)(M,g) we say that 12\mathcal{F}_{1}\subset\mathcal{F}_{2} if for any leaf L1L\in\mathcal{F}_{1}, there exists a leaf L2L^{\prime}\in\mathcal{F}_{2}, such that LLL\subset L^{\prime}. With this definition observe that ρFKM\mathcal{F}_{\rho}\subset\mathcal{F}_{\mathrm{FKM}}.

Non-orbit like foliations

With RFKM-foliations we are able to give examples of closed manifolds with singular Riemannian foliations for which the infinitesimal foliation is not given by a group action. Namely, fix (L,gL)(L,g_{L}) a closed Riemannian manifold and consider the closed disk (𝔻n,g0,0)(\mathbb{D}^{n},g_{0},\mathcal{F}_{0}) with the Euclidean metric, equipped with a singular Riemannian foliation given by the cone of a non-homogeneous RFKM-foliation 0\mathcal{F}_{0} on the sphere 𝕊n1\mathbb{S}^{n-1}. Now consider the foliated product (M1,g1,1)=(L×𝔻n,gLg0,{L×vv0})(M_{1},g_{1},\mathcal{F}_{1})=(L\times\mathbb{D}^{n},g_{L}\oplus g_{0},\{L\times\mathcal{L}_{v}\mid\mathcal{L}_{v}\in\mathcal{F}_{0}\}). Next we glue two copies of M1M_{1} along the boundary via the identity map to obtain a new smooth foliated closed manifold (M,g,)(M,g,\mathcal{F}) whose orbit space is 𝕊n\mathbb{S}^{n}, the double of 𝔻n\mathbb{D}^{n}. For any of the two singular leaves given by L×{0}L\times\{0\}, in each copy of M1M_{1}, the infinitesimal foliation corresponds to the RFKM-foliation (𝕊n1,g0,0)(\mathbb{S}^{n-1},g_{0},\mathcal{F}_{0}). Thus \mathcal{F} cannot be an orbit-like foliation. This implies that the results in [3] do not apply to this foliation, but þG, and in turn þA do apply to this foliation.

Other non-homogeneous singular Riemannian foliations

It is easy to construct closed non simply-connected manifolds with a singular Riemannian foliation which is not given by a group action. For example as suggested in [28], we consider NN any closed Riemmannian manifold, and for n5n\geq 5 a smooth manifold 𝒯n\mathcal{T}^{n} which is homeomorphic but not diffeomorphic to the nn-torus, i.e. an exotic torus. These manifolds exist due to [34]. Now we consider M=N×𝒯nM=N\times\mathcal{T}^{n} with any product metric. This produces a Riemannian foliation ={{p}×𝒯npN}\mathcal{F}=\{\{p\}\times\mathcal{T}^{n}\mid p\in N\} with leaves of the same dimension, whose leaves are not homogeneous spaces and therefore, this foliation cannot be given by a group action. Example 3.6 in [28] gives a way to construct other non-homogeneous examples. As a particular case of these examples, we can see that the Klein bottle has a foliation by circles, which is not given by a circle action. Moreover, in [23] Farrell and Wu gave examples of Riemannian foliations with all leaves of the same dimension. The leaves of one of theses foliations are the fibers of a fiber bundle over 44-dimensional manifolds, and moreover these fibers are exotic tori. Again we point out that the previous results in the literature for finding sign-changing solutions to (Y) do not apply for these foliations, but þA does.

4. A Principle of Symmetric Criticality for singular Riemannian foliations

Let (M,g,)(M,g,\mathcal{F}) be a singular Riemannian foliation with closed leaves. As before, denote by Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} the closure of the smooth foliated real valued functions 𝒞(M)\mathcal{C}^{\infty}(M)^{\mathcal{F}} under Hg1(M)\|\cdot\|_{H^{1}_{g}(M)} and by HH^{\perp}_{\mathcal{F}} the orthogonal complement of Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} with respect to the Hg1(M)H^{1}_{g}(M)-inner product. The aim of this section is to prove Theorem G. The proof relies on establishing that f,uLg2(M)=0\langle f,u\rangle_{L^{2}_{g}(M)}=0 for any fHf\in H^{\perp}_{\mathcal{F}} and uHg1(M)u\in H^{1}_{g}(M)^{\mathcal{F}}, and using the explicit formula of the derivative of the energy functional JJ.

To our knowledge, Theorem G fills in a part in [33, Proof of Lemma 4.1] where a particular case of Theorem G for foliations given by isoparametric functions was used without proof to obtain certain solutions to the subcritical Yamabe equation.


For a metric space (X,d)(X,d), given a a function f:Xf\colon X\to\mathbb{R} we define Lip(f)\mathrm{Lip}(f), the Lipschitz constant of ff, as

Lip(f)=sup{|f(x)f(y)|d(x,y)xyX}.\mathrm{Lip}(f)=\sup\left\{\frac{|f(x)-f(y)|}{d(x,y)}\mid x\neq y\in X\right\}.

Denote by

LIP(X)={f:XLip(f)<}.\mathrm{LIP}(X)=\Big{\{}f\colon X\to\mathbb{R}\mid\mathrm{Lip}(f)<\infty\Big{\}}.

We now let dMd_{M} be the distance function on MM induced by the Riemannian metric gg and consider the subset

Lip={fLIP(M)f is constant along the leaves of }.\mathrm{Lip}_{\mathcal{F}}=\{f\in\mathrm{LIP}(M)\mid f\mbox{ is constant along the leaves of $\mathcal{F}$}\}.

Denote by L2L^{2}_{\mathcal{F}} the closure of Lip\mathrm{Lip}_{\mathcal{F}} in Lg2(M)L_{g}^{2}(M), and by (L2)(L^{2}_{\mathcal{F}})^{\perp} the orthogonal complement of L2L^{2}_{\mathcal{F}} with respect to the Lg2(M)L^{2}_{g}(M)-inner product.

Recall that the leaf space of (M,g,)(M,g,\mathcal{F}) is the space M=M/M^{\ast}=M/\mathcal{F} equipped with the quotient topology induced by the projection π:MM\pi\colon M\to M^{\ast}. For MM complete and \mathcal{F} with closed leaves, this map also endows MM^{\ast} with a metric dd^{\ast} and a measure μ=πdVg\mu^{\ast}=\pi_{\ast}dV_{g}, which is the pushforward of the Riemannian volume form dVgdV_{g} of (M,g)(M,g) under the quotient map π\pi. Moreover (M,d,μ)(M^{\ast},d^{\ast},\mu^{\ast}) is a complete separable metric space with a non-trivial locally finite Borel measure ([37, p. 119]); that is (M,d,μ)(M^{\ast},d^{\ast},\mu^{\ast}) is a metric measure space [32, p. 65]. So that LIP(M)\mathrm{LIP}(M^{\ast}) is dense in L2(M,μ)L^{2}(M^{\ast},\mu^{\ast}) (see for example [32, Theorem 4.2.4]).

Given a function f:Mf\colon M\to\mathbb{R} which is constant along the leaves, we define a function f:Mf^{\ast}\colon M^{\ast}\to\mathbb{R} as f(π(x))=f(x)f^{\ast}(\pi(x))=f(x). Moreover, given a function h:Mh\colon M^{\ast}\to\mathbb{R}, we define a function h^:M\hat{h}\colon M\to\mathbb{R} which is constant along the leaves of \mathcal{F} by h^(x)=h(π(x))\hat{h}(x)=h(\pi(x)). For a function f:Mf\colon M\to\mathbb{R} constant along the leaves we have f=f^f=\widehat{f}^{\ast}.

Using the fact that μ\mu^{\ast} is the pushforward measure, we have well defined functions :L2L2(M,μ)\ast\colon L_{\mathcal{F}}^{2}\to L^{2}(M^{\ast},\mu^{\ast}) and ^:L2(M,μ)L2\ \widehat{}\,\colon L^{2}(M^{\ast},\mu^{\ast})\to L_{\mathcal{F}}^{2}. Indeed, take fL2f\in L^{2}_{\mathcal{F}}. Then it holds that

M(f)2μ=M(fπ)2𝑑Vg=Mf2𝑑Vg<.\int_{M^{\ast}}(f^{\ast})^{2}\,\mu^{\ast}=\int_{M}(f^{\ast}\circ\pi)^{2}\,dV_{g}=\int_{M}f^{2}\,dV_{g}<\infty.

On the other hand, for hL2(M,μ)h\in L^{2}(M^{\ast},\mu^{\ast}) we have that

Mh^2𝑑Vg=M(hπ)2𝑑Vg=Mh2μ<.\int_{M}\hat{h}^{2}\,dV_{g}=\int_{M}(h\circ\pi)^{2}\,dV_{g}=\int_{M^{\ast}}h^{2}\,\mu^{\ast}<\infty.
Lemma 4.1.

þ For (M,g,)(M,g,\mathcal{F}) a singular Riemannian foliation with closed leaves, it holds true that Hg1(M)L2H^{1}_{g}(M)^{\mathcal{F}}\subset L^{2}_{\mathcal{F}}.

Proof.

Given uHg1(M)u\in H^{1}_{g}(M)^{\mathcal{F}}, since LIP(M)\mathrm{LIP}(M^{\ast}) is dense in L2(M,μ)L^{2}(M^{\ast},\mu^{\ast}), there exists a sequence {hn}L2(M,μ)\{h_{n}\}\subset L^{2}(M^{\ast},\mu^{\ast}) of Lipschitz bounded functions converging to uu^{\ast} in L2(M,μ)L^{2}(M^{\ast},\mu^{\ast}). Applying a change of variable we get that h^nu^=u\hat{h}_{n}\to\widehat{u}^{\ast}=u in Lg2(M)L_{g}^{2}(M). To prove this, fix ε>0\varepsilon>0. Then there exists NN\in\mathbb{N}, such that for nNn\geqslant N we have:

M|h^nu|2𝑑Vg\displaystyle\int_{M}|\hat{h}_{n}-u|^{2}\,dV_{g} =M|h^nu^|2𝑑Vg=M|hnπuπ|2𝑑Vg\displaystyle=\int_{M}|\hat{h}_{n}-\widehat{u^{\ast}}|^{2}\,dV_{g}=\int_{M}|h_{n}\circ\pi-u^{\ast}\circ\pi|^{2}\,dV_{g}
=M|hnu|2π𝑑Vg=M|hnu|2μ\displaystyle=\int_{M^{\ast}}|h_{n}-u^{\ast}|^{2}\,\pi_{\ast}dV_{g}=\int_{M^{\ast}}|h_{n}-u^{\ast}|^{2}\,\mu^{\ast}
<ε.\displaystyle<\varepsilon.

Thus, h^nu\hat{h}_{n}\to u in Lg2(M)L_{g}^{2}(M). Finally note that by definition h^n=hnπ\hat{h}_{n}=h_{n}\circ\pi is the composition of Lipschitz functions, π\pi being a 11-Lipschitz function. Hence, h^nLIP\hat{h}_{n}\in\mathrm{LIP}_{\mathcal{F}}. Thus, we conclude that uL2u\in L^{2}_{\mathcal{F}} as desired. ∎

Lemma 4.2.

þ For (M,g,)(M,g,\mathcal{F}) a singular Riemannian foliation with closed leaves, it holds true that H(L2)H_{\mathcal{F}}^{\perp}\subset(L^{2}_{\mathcal{F}})^{\perp}.

Proof.

We first show that HL2L2H^{\perp}_{\mathcal{F}}\cap L^{2}_{\mathcal{F}}\subset L^{2}_{\mathcal{F}} consists only of the equivalent class of the zero function. Take uHL2L2u\in H^{\perp}_{\mathcal{F}}\cap L^{2}_{\mathcal{F}}\subset L^{2}_{\mathcal{F}}. Since uL2u\in L^{2}_{\mathcal{F}} there exists a sequence of Lipschitz functions which are constant along the leaves, {fn}Lip\{f_{n}\}\subset\mathrm{Lip}_{\mathcal{F}}, converging to uu with respect to the Lg2(M)L^{2}_{g}(M)-norm. Then {fn}\{f_{n}^{\ast}\} is a Cauchy sequence in L2(M,μ)L^{2}(M^{\ast},\mu^{\ast}); fix ε>0\varepsilon>0, there exists NN\in\mathbb{N} such that for all m,nNm,n\geqslant N it holds:

ε\displaystyle\varepsilon >M|fnfm|2𝑑Vg=M|fn^fm^|2𝑑Vg\displaystyle>\int_{M}|f_{n}-f_{m}|^{2}\,dV_{g}=\int_{M}|\widehat{f_{n}^{\ast}}-\widehat{f_{m}^{\ast}}|^{2}\,dV_{g}
=M|fnπfmπ|2𝑑Vg=M|fnfm|2π𝑑Vg\displaystyle=\int_{M}|f_{n}^{\ast}\circ\pi-f_{m}^{\ast}\circ\pi|^{2}\,dV_{g}=\int_{M^{\ast}}|f_{n}^{\ast}-f_{m}^{\ast}|^{2}\,\pi_{\ast}dV_{g}
=M|fnfm|2μ.\displaystyle=\int_{M^{\ast}}|f_{n}^{\ast}-f_{m}^{\ast}|^{2}\,\mu^{\ast}.

Since L2(M,μ)L^{2}(M^{\ast},\mu^{\ast}) is complete there is a limit hL2(M,μ)h\in L^{2}(M^{\ast},\mu^{\ast}) of {fn}\{f_{n}^{\ast}\}. As in the proof of the previous lemma, fn^=fn\widehat{f_{n}^{\ast}}=f_{n} converges to h^\hat{h} in Lg2(M)L_{g}^{2}(M). Thus by uniqueness of the limit, uu is equal to h^\hat{h} μ\mu-a.e., i.e. uu is constant along the leaves of \mathcal{F} up to zero measure. This implies that the class of uu in Hg1(M)H^{1}_{g}(M) is an element in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} and since uHu\in H^{\perp}_{\mathcal{F}} by hypothesis, we conclude that uu corresponds to the class of the zero function in Hg1(M)H^{1}_{g}(M) (recall that Hg1(M)=Hg1(M)HH^{1}_{g}(M)=H^{1}_{g}(M)^{\mathcal{F}}\oplus H^{\perp}_{\mathcal{F}}). Thus HL2={0}H^{\perp}_{\mathcal{F}}\cap L^{2}_{\mathcal{F}}=\{0\}.

Finally, since L2(L2)L^{2}_{\mathcal{F}}\cap(L^{2}_{\mathcal{F}})^{\perp} consists only of the equivalence class of the zero function, by the first paragraph we conclude that H(L2)H_{\mathcal{F}}^{\perp}\subset(L^{2}_{\mathcal{F}})^{\perp}. ∎

Lemma 4.3.

Let (M,g,)(M,g,\mathcal{F}) be a singular Riemannian foliation with closed leaves. If fHf\in H^{\perp}_{\mathcal{F}} and uHg1(M)u\in H^{1}_{g}(M)^{\mathcal{F}} then f,uLg2(M)=0\langle f,u\rangle_{L^{2}_{g}(M)}=0.

Proof.

By Lemma 4.1 and Lemma 4.2, we have that f(L2)f\in(L^{2}_{\mathcal{F}})^{\perp} and uL2u\in L^{2}_{\mathcal{F}}. The lemma then follows. ∎

We are ready to prove the Principle of Symmetric Criticality, that is, that critical points of JJ restricted to Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} are critical points of JJ.

Proof of Theorem G.

Write any vHg1(M)v\in H^{1}_{g}(M) as v=v1+v2v=v_{1}+v_{2} with v1Hg1(M)v_{1}\in H_{g}^{1}(M)^{\mathcal{F}} and v2Hv_{2}\in H_{\mathcal{F}}^{\perp}. Since uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} is a critical point of JJ restricted to the space Hg1(M)H^{1}_{g}(M)^{\mathcal{F}}, we have that J(u)v1=0J^{\prime}(u)v_{1}=0. Hence,

J(u)v\displaystyle J^{\prime}(u)v =J(u)v1+J(u)v2=J(u)v2\displaystyle=J^{\prime}(u)v_{1}+J^{\prime}(u)v_{2}=J^{\prime}(u)v_{2}
=u,v2bMc|u|p2uv2𝑑Vg,\displaystyle=\langle u,v_{2}\rangle_{b}-\int_{M}c|u|^{p-2}uv_{2}dV_{g},

where in the last part we used the expression (8) for the derivative of JJ. By Lemma 4.3 both terms above equal zero. To see this is true for the first term, by hypothesis we have u,v2Hg1(M)=0\langle u,v_{2}\rangle_{H^{1}_{g}(M)}=0 and by Lemma 4.3 we have that Muv2𝑑Vg=0\int_{M}uv_{2}\ dV_{g}=0. This and bb being an \mathcal{F}-invariant function implies that u,v2b=0\langle u,v_{2}\rangle_{b}=0. For the second term, since cc is \mathcal{F}-invariant we have c|u|p2uL2c|u|^{p-2}u\in L_{\mathcal{F}}^{2} and, we know that v2H(L2)v_{2}\in H_{\mathcal{F}}^{\perp}\subset(L^{2}_{\mathcal{F}})^{\perp}. Thus, we can apply again Lemma 4.3 to conclude the claim. ∎

5. Compactness

In this section we first prove þH which is a Sobolev embedding theorem for singular Riemannian foliations that generalizes the classical result by Hebey and Vaugon in [31] and Lemma 6.1 in [33]. Then we apply þH to show þE, which says that JJ satisfies the (PS)τ(PS)^{\mathcal{F}}_{\tau}-condition for every τ\tau\in\mathbb{R}.


For any s1s\geq 1, consider the Sobolev space Hg1,s(M)H_{g}^{1,s}(M), which is the closure of 𝒞(M)\mathcal{C}^{\infty}(M) with respect to the norm

uHg1,s(M):=(M|gu|gs+|u|sdVg)1/s.\|u\|_{H_{g}^{1,s}(M)}:=\left(\int_{M}|\nabla_{g}u|^{s}_{g}+|u|^{s}dV_{g}\right)^{1/s}.

As before, let Hg1,s(M)H_{g}^{1,s}(M)^{\mathcal{F}} be the closure of 𝒞(M)\mathcal{C}^{\infty}(M)^{\mathcal{F}} under the norm above, so that it is a Banach space. Observe that Hg1(M)H^{1}_{g}(M) is just Hg1,2(M)H^{1,2}_{g}(M).

Recall that κ\kappa_{\mathcal{F}} is the smallest dimension of the leaves of \mathcal{F}. We now state the Sobolev embedding theorem needed for the proof of Theorem E.

Theorem H.

þ Let \mathcal{F} be a singular Riemannian foliation on the mm-dimensional Riemannain manifold (M,g)(M,g) and let s1s\geq 1. If any of the following conditions hold,

(C1) smκandp1s\geq m-\kappa_{\mathcal{F}}\quad\text{and}\quad p\geq 1

or

(C2) s<mκandp[1,s(mκ)mκs],s<m-\kappa_{\mathcal{F}}\quad\text{and}\quad p\in\Big{[}1,\frac{s(m-\kappa_{\mathcal{F}})}{m-\kappa_{\mathcal{F}}-s}\Big{]},

then the inclusion map Hg1,s(M)Lgp(M)H_{g}^{1,s}(M)^{\mathcal{F}}\hookrightarrow L^{p}_{g}(M) is continuous. For either p1p\geqslant 1 in (C1) or 1p<s(mκ)mκs1\leqslant p<\frac{s(m-\kappa_{\mathcal{F}})}{m-\kappa_{\mathcal{F}}-s} in (C2), the map is also compact.

The proof of this is virtually the same as the proof of Lemma 6.1 in [33] or the proof of the Main Lemma in [31]. But we present it below for the convenience of the reader. To show the continuity of the inclusion, the idea is to apply þ3.1 to any qMq\in M, to get a chart of the form (Wq,φq)(W_{q},\varphi_{q}). As MM is compact, we can cover it by open sets of the form WqW_{q}. Then one passes the information of any function in Hg1,s(M)H_{g}^{1,s}(M)^{\mathcal{F}} to functions defined on sets VqV_{q}, where φq(Wq)=Uq×Vq\varphi_{q}(W_{q})=U_{q}\times V_{q} as in þ3.1, reducing the dimension of MM to the dimension of VqV_{q}, which is mkqm-k_{q}. This allows us to apply the subcritical Sobolev embedding theorem for the sets VqV_{q}. To prove the compactness of the map, we also apply þ3.5 and þ3.6 in order to have a covering of MM consisting of tubular neighborhoods and finalize the proof by applying the Rellich-Kondrachov compactness theorem.


We start by fixing some notation and studying the behavior of functions in Hg1,s(M)H_{g}^{1,s}(M)^{\mathcal{F}}. Fix qMq\in M, and let φ:WU×V\varphi\colon W\to U\times V be a chart around qq given by þ3.1. For any function uHg1,s(M)u\in H_{g}^{1,s}(M)^{\mathcal{F}}, let u^:=uφ1:U×V\hat{u}:=u\circ\varphi^{-1}\colon U\times V\to\mathbb{R}. By the properties of the coordinate chart, for any x,xUx,x^{\prime}\in U and any yVy\in V, we have that u^(x,y)=u^(x,y)\hat{u}(x,y)=\hat{u}(x^{\prime},y), so we have a well defined function uV:Vu_{V}\colon V\to\mathbb{R} given by uV(y):=u^(x,y)u_{V}(y):=\hat{u}(x,y). Denote by gu\nabla_{g}u the gradient of uu, for z=(x,y)U×Vz=(x,y)\in U\times V then we denote by u^=(xu^,yu^)\nabla\hat{u}=(\nabla_{x}\hat{u},\nabla_{y}\hat{u}) the gradient of u^\hat{u}, and by yuV\nabla_{y}u_{V} the gradient of uVu_{V}.

Lemma 5.1.

þ Consider a fixed qMq\in M, and WW as above. For any open subsets UUU^{\prime}\subset\subset U and VVV^{\prime}\subset\subset V and any u𝒞(M)u\in\mathcal{C}^{\infty}(M)^{\mathcal{F}} there exists a constant C=C(W)C=C(W^{\prime}), where W=φ1(U×V)W^{\prime}=\varphi^{-1}(U^{\prime}\times V^{\prime}), such that

C1|yuV|2|gu|g2C|yuV|2C^{-1}|\nabla_{y}u_{V}|^{2}\leq|\nabla_{g}u|_{g}^{2}\leq C|\nabla_{y}u_{V}|^{2}

holds over WW^{\prime}.

Proof.

Since u𝒞(M)u\in\mathcal{C}^{\infty}(M)^{\mathcal{F}} we have u^xi(x,y)=0\frac{\partial\hat{u}}{\partial x^{i}}(x,y)=0. Hence, u^=(xu^,yu^)=(0,yuV)\nabla\hat{u}=(\nabla_{x}\hat{u},\nabla_{y}\hat{u})=(0,\nabla_{y}u_{V}) and thus |gu|g2=|u^|2=|yuV|2|\nabla_{g}u|^{2}_{g}=|\nabla\hat{u}|^{2}=|\nabla_{y}u_{V}|^{2}.

Writing the metric gg in coordinates as g=i,j=1mgijdxidxjg=\sum_{i,j=1}^{m}g_{ij}dx^{i}\otimes dx^{j}, at each qWq^{\prime}\in W the matrix (gij(q))(g_{ij}(q^{\prime})) is definite positive and symmetric, hence it is diagonalizable and with positive eigenvalues λi(q)\lambda_{i}(q^{\prime}). Then, at each qq^{\prime} we have that

min{λi(q)}i=1m(u^yi(q))2\displaystyle\min\{\lambda_{i}(q^{\prime})\}\sum_{i=1}^{m}\left(\frac{\partial\hat{u}}{\partial y^{i}}(q^{\prime})\right)^{2} i,j=1mgij(q)(u^yi(q))(u^yj(q))\displaystyle\leq\sum_{i,j=1}^{m}g_{ij}(q^{\prime})\left(\frac{\partial\hat{u}}{\partial y^{i}}(q^{\prime})\right)\left(\frac{\partial\hat{u}}{\partial y^{j}}(q^{\prime})\right)
max{λi(q)}i=1m(u^yi(q))2.\displaystyle\leq\max\{\lambda_{i}(q^{\prime})\}\sum_{i=1}^{m}\left(\frac{\partial\hat{u}}{\partial y^{i}}(q^{\prime})\right)^{2}.

Since the functions λi\lambda_{i} are continuous, and WWW^{\prime}\subset\subset W then

A:=minqW¯min{λi(q)},B:=maxqW¯max{λi(q)}>0.A:=\min_{q^{\prime}\in\overline{W^{\prime}}}\min\{\lambda_{i}(q^{\prime})\},\quad B:=\max_{q^{\prime}\in\overline{W^{\prime}}}\max\{\lambda_{i}(q^{\prime})\}>0.

If C>0C>0 is such that C1<AB<CC^{-1}<A\leq B<C, substituting everything in the previous inequality, and using that |u^|2=|yuV|2|\nabla\hat{u}|^{2}=|\nabla_{y}u_{V}|^{2} we conclude. ∎

Lemma 5.2.

þ For qMq\in M, WW and any WW^{\prime} as in the previous lemma, there exists a constant C=C(W)C=C(W^{\prime}), such that, for any s1s\geq 1 and any function u𝒞(M)u\in\mathcal{C}^{\infty}(M)^{\mathcal{F}}

C1V|uV|s𝑑yW|u|s𝑑VgCV|uV|s𝑑yC^{-1}\int_{V^{\prime}}|u_{V}|^{s}dy\leq\int_{W^{\prime}}|u|^{s}dV_{g}\leq C\int_{V^{\prime}}|u_{V}|^{s}dy

and

C1V|yuV|s𝑑yW|gu|gs𝑑VgCV|yuV|s𝑑y.C^{-1}\int_{V^{\prime}}|\nabla_{y}u_{V}|^{s}dy\leq\int_{W^{\prime}}|\nabla_{g}u|_{g}^{s}dV_{g}\leq C\int_{V^{\prime}}|\nabla_{y}u_{V}|^{s}dy.

In particular, there exists C=C(W)>0C=C(W^{\prime})>0 such that for any uHg1,s(M)u\in H^{1,s}_{g}(M)^{\mathcal{F}},

(13) C1|u|Lgs(W)|uV|Ls(V)C|u|Lgs(W), and C1|gu|Lgs(W)|uV|Ls(V)C|gu|Lgs(W).\displaystyle\begin{split}&C^{-1}|u|_{L_{g}^{s}(W^{\prime})}\leq|u_{V}|_{L^{s}(V^{\prime})}\leq C|u|_{L_{g}^{s}(W^{\prime})},\\ \mbox{ and }&\\ &C^{-1}|\nabla_{g}u|_{L_{g}^{s}(W^{\prime})}\leq|\nabla u_{V}|_{L^{s}(V^{\prime})}\leq C|\nabla_{g}u|_{L_{g}^{s}(W^{\prime})}.\end{split}
Proof.

For the first inequality, the function |g|:W¯\sqrt{|g|}\colon\overline{W^{\prime}}\to\mathbb{R} is continuous and attains its maximum and its minimum, say 0<A:=min|g|0<A:=\min\sqrt{|g|} and 0<B=max|g|0<B=\max\sqrt{|g|}.

First we have,

W|u|s𝑑Vg\displaystyle\int_{W^{\prime}}|u|^{s}dV_{g} =U×V|g||u^|s𝑑x𝑑yBU×V|uV|s𝑑x𝑑y\displaystyle=\int_{U^{\prime}\times V^{\prime}}\sqrt{|g|}|\hat{u}|^{s}dxdy\leq B\int_{U^{\prime}\times V^{\prime}}|u_{V}|^{s}dxdy
=BU𝑑xV|uV|s𝑑y=C1V|uV|s𝑑y.\displaystyle=B\int_{U^{\prime}}dx\int_{V^{\prime}}|u_{V}|^{s}dy=C_{1}\int_{V^{\prime}}|u_{V}|^{s}dy.

Second we have

V|uV|s𝑑y\displaystyle\int_{V^{\prime}}|u_{V}|^{s}dy =A1VA|uV|s𝑑y=C2U×VA|u|s𝑑x𝑑y\displaystyle=A^{-1}\int_{V^{\prime}}A|u_{V}|^{s}dy=C_{2}\int_{U^{\prime}\times V^{\prime}}A|u|^{s}dxdy
C2U×V|g||u|s𝑑x𝑑y=C2W|u|s𝑑Vg.\displaystyle\leq C_{2}\int_{U^{\prime}\times V^{\prime}}\sqrt{|g|}|u|^{s}dxdy=C_{2}\int_{W^{\prime}}|u|^{s}dV_{g}.

To obtain the gradient estimate, we use þ5.1:

W|gu|gs𝑑Vg\displaystyle\int_{W^{\prime}}|\nabla_{g}u|_{g}^{s}dV_{g} =U×V|g||gu|gs𝑑x𝑑yB2U×V|yuV|s𝑑x𝑑y\displaystyle=\int_{U^{\prime}\times V^{\prime}}\sqrt{|g|}|\nabla_{g}u|_{g}^{s}dxdy\leq B_{2}\int_{U^{\prime}\times V^{\prime}}|\nabla_{y}u_{V}|^{s}dxdy
=B2U𝑑xV|yuV|s𝑑y=C3V|yuV|s𝑑y.\displaystyle=B_{2}\int_{U^{\prime}}dx\int_{V^{\prime}}|\nabla_{y}u_{V}|^{s}dy=C_{3}\int_{V^{\prime}}|\nabla_{y}u_{V}|^{s}dy.

And again:

V|yuV|s𝑑y=A1VA|yuV|s𝑑x𝑑y=C4U×VA|yuV|s𝑑x𝑑y\displaystyle\int_{V^{\prime}}|\nabla_{y}u_{V}|^{s}dy=A^{-1}\int_{V^{\prime}}A|\nabla_{y}u_{V}|^{s}dxdy=C_{4}\int_{U^{\prime}\times V^{\prime}}A|\nabla_{y}u_{V}|^{s}dxdy
C4U×V|g||yuV|s𝑑x𝑑yC5U×V|g||gu|gs𝑑x𝑑y\displaystyle\leq C_{4}\int_{U^{\prime}\times V^{\prime}}\sqrt{|g|}|\nabla_{y}u_{V}|^{s}dxdy\leq C_{5}\int_{U^{\prime}\times V^{\prime}}\sqrt{|g|}|\nabla_{g}u|_{g}^{s}dxdy
=C5W|gu|gs𝑑Vg.\displaystyle=C_{5}\int_{W^{\prime}}|\nabla_{g}u|_{g}^{s}dV_{g}.

Taking CC to be any real constant such that

C1min{C11/s,C31/s} and max{C21/s,C51/s}CC^{-1}\leq\min\{C_{1}^{-1/s},C_{3}^{-1/s}\}\mbox{ and }\max\{C_{2}^{1/s},C_{5}^{1/s}\}\leq C

we obtain the inequalities (13) for any u𝒞(M)u\in\mathcal{C}^{\infty}(M)^{\mathcal{F}}. Since 𝒞(M)\mathcal{C}^{\infty}(M)^{\mathcal{F}} is dense in Hg1,s(M)H^{1,s}_{g}(M)^{\mathcal{F}} and in Lgs(M)L_{g}^{s}(M)^{\mathcal{F}}, we conclude that (13) holds. ∎

We are now ready to prove the Sobolev embedding theorem.

Proof of Theorem H.

For any qMq\in M, take a chart around qq, (Wq,φq)(W_{q},\varphi_{q}), as in þ3.1. Take a further open subset WqWqW^{\prime}_{q}\subset\subset W_{q} around qq as in Lemma 5.2. As MM is compact, there exist a finite number of points q1,,qMq_{1},\ldots,q_{\ell}\in M such that the charts (Wi,φi):=(Wqi,φqi|Wqi)(W_{i},\varphi_{i}):=(W^{\prime}_{q_{i}},\varphi_{q_{i}}|_{W^{\prime}_{q_{i}}}) cover MM. Set ki:=kqik_{i}:=k_{q_{i}} and note that φi(Wi)=Ui×Vi\varphi_{i}(W_{i})=U_{i}\times V_{i} for some open sets.

To prove the continuity of the inclusion map, take uHg1,s(M)u\in H_{g}^{1,s}(M)^{\mathcal{F}} and consider the function ui=uVi:Viu_{i}=u_{V_{i}}:V_{i}\to\mathbb{R} as defined in this section, i.e. uVi(y):=uφi1(x,y)u_{V_{i}}(y):=u\circ\varphi_{i}^{-1}(x,y), where xx is an arbitrary element in UiU_{i}. Then Lemma 5.2 implies that uiH1,s(Vi)u_{i}\in H^{1,s}(V_{i}). Using (13) twice together with the Sobolev embedding theorems for VimkiV_{i}\subset\mathbb{R}^{m-k_{i}} [1, Theorem 5.4], we have the existence of a constant CiC_{i} depending on WiW_{i} such that

|u|Lgpi(Wi)\displaystyle|u|_{L^{p_{i}}_{g}(W_{i})} Ci|ui|Lpi(Vi)\displaystyle\leq C_{i}|u_{i}|_{L^{p_{i}}(V_{i})}
Ci(|yui|Ls(Vi)+|ui|Ls(Vi))\displaystyle\leq C_{i}\left(|\nabla_{y}u_{i}|_{L^{s}(V_{i})}+|u_{i}|_{L^{s}(V_{i})}\right)
Ci(|gu|Lgs(Wi)+|u|Lgs(Wi))\displaystyle\leq C_{i}\left(|\nabla_{g}u|_{L^{s}_{g}(W_{i})}+|u|_{L^{s}_{g}(W_{i})}\right)

if either smkis\geq m-k_{i} and pi1p_{i}\geq 1, or s<mkis<m-k_{i} and 1pis(mki)/(mkis)1\leq p_{i}\leq s(m-k_{i})/(m-k_{i}-s).

We now prove that for every index i=1,,i=1,\ldots,\ell, either smkis\geq m-k_{i} or s<mkis<m-k_{i} and ps(mki)/(mkis)p\leq s(m-k_{i})/(m-k_{i}-s) holds. If (C1) holds, then by the definition of κ\kappa_{\mathcal{F}} we get smkis\geq m-k_{i} for every i=1,,i=1,\ldots,\ell. If (C2) holds, fix ii. Then either smkis\geq m-k_{i} or s<mkis<m-k_{i}. Assume that s<mkis<m-k_{i}. Then the fact s0s\geqslant 0 and the definition of κ\kappa_{\mathcal{F}} yields the inequality s(mκ)/(mκs)s(mki)/(mkis)s(m-\kappa_{\mathcal{F}})/(m-\kappa_{\mathcal{F}}-s)\leq s(m-k_{i})/(m-k_{i}-s).

Therefore, by taking CCiC\geqslant C_{i} for all ii,

|u|Lgp(M)\displaystyle|u|_{L_{g}^{p}(M)} i=1|u|Lgp(Wi)Ci=1|gu|Lgs(Wi)+|u|Lgs(Wi)\displaystyle\leq\sum_{i=1}^{\ell}|u|_{L_{g}^{p}(W_{i})}\leq C\sum_{i=1}^{\ell}|\nabla_{g}u|_{L_{g}^{s}(W_{i})}+|u|_{L_{g}^{s}(W_{i})}
Ci=1uHg1,s(M)CuHg1,s(M).\displaystyle\leq C\sum_{i=1}^{\ell}\|u\|_{H_{g}^{1,s}(M)}\leq C\ell\|u\|_{H_{g}^{1,s}(M)}.

Hence the embedding Hg1,s(M)Lgp(M)H_{g}^{1,s}(M)^{\mathcal{F}}\hookrightarrow L^{p}_{g}(M) is continuous.

Now we prove the compactness of the embedding. Without loss of generality assume that each open set Wi=Ui×ViW_{i}=U_{i}\times V_{i} is a trivialization of the disk bundle Pi×Hol(Li)𝔻iLiP_{i}\times_{\mathrm{Hol}(L_{i})}\mathbb{D}_{i}^{\perp}\to L_{i} given by þ3.5 centered at some piMp_{i}\in M. We can identify each element of the finite open cover 𝒜={Tubε(pi)(Lpi)}i=1N\mathcal{A}=\{\mathrm{Tub}^{\varepsilon(p_{i})}(L_{p_{i}})\}_{i=1}^{N} of MM, with Pi×Hol(Li)𝔻iP_{i}\times_{\mathrm{Hol}(L_{i})}\mathbb{D}_{i}^{\perp}. By þ3.6 there exists a smooth foliated partition of unity {ϕi}\{\phi_{i}\} subordinated to 𝒜\mathcal{A}. For each index ii, due to the identification we may assume that ϕi\phi_{i} is defined over Pi×Hol(Li)𝔻iP_{i}\times_{\mathrm{Hol}(L_{i})}\mathbb{D}_{i}^{\perp}.

Let (uj)j(u_{j})_{j\in\mathbb{N}} be a bounded sequence in Hg1,s(M)H^{1,s}_{g}(M)^{\mathcal{F}} and define uji(y):=(ϕiφi1)(ujφi1)(x,y)u_{ji}(y):=(\phi_{i}\circ\varphi_{i}^{-1})(u_{j}\circ\varphi_{i}^{-1})(x,y), for arbitrary xUix\in U_{i}. Observe ujiu_{ji} is well defined and is compactly supported in ViV_{i} and that the sequence (uji)j(u_{ji})_{j\in\mathbb{N}} is bounded in H1,s(Vi)H^{1,s}(V_{i}). Then, if either smκs\geq m-\kappa_{\mathcal{F}} or if s<mκs<m-\kappa_{\mathcal{F}} and 1p<s(mκ)mκs1\leq p<\frac{s(m-\kappa_{\mathcal{F}})}{m-\kappa_{\mathcal{F}}-s}, we have that 1p>1s1mki\frac{1}{p}>\frac{1}{s}-\frac{1}{m-k_{i}} for every i=1,i=1,\ldots\ell. Hence, the embedding H01,s(Vi)Lp(Vi)H_{0}^{1,s}(V_{i})\hookrightarrow L^{p}(V_{i}) is compact by the Rellich-Kondrachov theorem [1, Theorem 6.5] and for each i=1,,i=1,\ldots,\ell, there is a subsequence of (uji)j(u_{ji})_{j\in\mathbb{N}}, which we denote in the same way, which is a Cauchy sequence in Lp(Vi)L^{p}(V_{i}). Inequalities (13) and the fact that {ϕi}i=1\{\phi_{i}\}_{i=1}^{\ell} are bounded yield that (uj)j(u_{j})_{j\in\mathbb{N}} has a Cauchy subsequence in Lgp(M)L^{p}_{g}(M) and, hence, this subsequence converges in this space. ∎

We now show that JJ satisfies the (PS)τ(PS)^{\mathcal{F}}_{\tau}-condition for every τ\tau\in\mathbb{R}.

Proof of Theorem E.

The proof is standard and consists in showing that any (PS)τ(PS)^{\mathcal{F}}_{\tau}-sequence for JJ, (vn)nHg1(M)(v_{n})_{n\in\mathbb{N}}\subset H_{g}^{1}(M)^{\mathcal{F}}, is bounded in Hg1(M)H_{g}^{1}(M), so there exists vHg1(M)v\in H_{g}^{1}(M) such that up to a subsequence, (vn)n(v_{n})_{n\in\mathbb{N}} converges weakly to vv in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. The next step is to apply Theorem H to show that vnHg1(M)=vnHg1(M)vHg1(M)\|v_{n}\|_{H^{1}_{g}(M)}=\|v_{n}\|_{H^{1}_{g}(M)}\to\|v\|_{H^{1}_{g}(M)}, which implies that vnvv_{n}\to v strongly in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. We provide more details below.

Let (vn)nHg1(M)(v_{n})_{n\in\mathbb{N}}\subset H_{g}^{1}(M)^{\mathcal{F}} be a (PS)τ(PS)^{\mathcal{F}}_{\tau}-sequence for JJ in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, i.e. J(vn)τJ(v_{n})\to\tau and J(vn)0J^{\prime}(v_{n})\to 0 in (Hg1(M))(H_{g}^{1}(M)^{\mathcal{F}})^{\ast}. We first uniformly bound vnb\|v_{n}\|_{b} in terms of J(vn)J(v_{n}) and J(vn)vnJ^{\prime}(v_{n})v_{n}.

From the definition of JJ we get

J(vn)vn=vnb2|vn|c,pp.J^{\prime}(v_{n})v_{n}=\|v_{n}\|^{2}_{b}-|v_{n}|^{p}_{c,p}.

Therefore, we have

J(vn)1pJ(vn)vn\displaystyle J(v_{n})-\tfrac{1}{p}J^{\prime}(v_{n})v_{n} =12vnb21p|vn|c,pp1pvnb2+1p|vn|c,pp\displaystyle=\tfrac{1}{2}\|v_{n}\|^{2}_{b}-\tfrac{1}{p}|v_{n}|^{p}_{c,p}-\tfrac{1}{p}\|v_{n}\|^{2}_{b}+\tfrac{1}{p}|v_{n}|^{p}_{c,p}
=p22pvnb2.\displaystyle=\tfrac{p-2}{2p}\|v_{n}\|^{2}_{b}.

On the other hand, since J(vn)0J^{\prime}(v_{n})\to 0 in (Hg1(M))(H_{g}^{1}(M)^{\mathcal{F}})^{\ast}, there exists a positive constant C1C_{1} such that for large nn\in\mathbb{N} the following holds

|J(vn)vn|J(vn)Hg1(M)vnbC1vnb.|J^{\prime}(v_{n})v_{n}|\leq\|J^{\prime}(v_{n})\|_{H_{g}^{1}(M)^{\ast}}\|v_{n}\|_{b}\leq C_{1}\|v_{n}\|_{b}.

Since J(vn)τJ(v_{n})\to\tau, there exists a positive constant C2C_{2} such that for large nn\in\mathbb{N} the following holds

|J(vn)|C2.|J(v_{n})|\leq C_{2}.

Therefore for large nn\in\mathbb{N} we obtain that

p22pvnb2\displaystyle\tfrac{p-2}{2p}\|v_{n}\|^{2}_{b} =J(vn)1p(J)(vn)vn|J(vn)1p(J)(vn)vn|\displaystyle=J(v_{n})-\tfrac{1}{p}(J)^{\prime}(v_{n})v_{n}\leqslant\big{|}J(v_{n})-\tfrac{1}{p}(J)^{\prime}(v_{n})v_{n}\big{|}
|J(vn)|+1p|(J)(vn)vn|C2+C1pvnb.\displaystyle\leqslant|J(v_{n})|+\tfrac{1}{p}|(J)^{\prime}(v_{n})v_{n}|\leqslant C_{2}+\tfrac{C_{1}}{p}\|v_{n}\|_{b}.

This, and the fact that the norm b\|\cdot\|_{b} is equivalent to the standard norm of Hg1(M)H^{1}_{g}(M), imply that (vn)n(v_{n})_{n\in\mathbb{N}} is bounded in Hg1(M)H^{1}_{g}(M).

Since (vn)n(v_{n})_{n\in\mathbb{N}} is bounded in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, there exists vHg1(M)v\in H_{g}^{1}(M) such that up to a subsequence, also denoted by (vn)n(v_{n})_{n\in\mathbb{N}}, it converges weakly to vv in Hg1(M)H_{g}^{1}(M). As Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} is weakly closed in Hg1(M)H_{g}^{1}(M), vv is \mathcal{F}-invariant. To show that vnvv_{n}\to v strongly in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, it suffices to show that vnHg1(M)vHg1(M)\|v_{n}\|_{H^{1}_{g}(M)}\to\|v\|_{H^{1}_{g}(M)}.

As vnvv_{n}\rightharpoonup v weakly in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, we have that

vn,vHg1(M)=vHg1(M)2+o(1).\langle v_{n},v\rangle_{H^{1}_{g}(M)}=\|v\|^{2}_{H^{1}_{g}(M)}+o(1).

It follows that

J(vn)(vnv)=\displaystyle J^{\prime}(v_{n})(v_{n}-v)= vn,vnvHg1(M)Mc|vn|p2vn(vnv)𝑑Vg\displaystyle\langle v_{n},v_{n}-v\rangle_{H^{1}_{g}(M)}-\int_{M}c\,|v_{n}|^{p-2}v_{n}(v_{n}-v)dV_{g}
=\displaystyle= vnHg1(M)2vHg1(M)2+o(1)\displaystyle\|v_{n}\|^{2}_{H^{1}_{g}(M)}-\|v\|^{2}_{H^{1}_{g}(M)}+o(1)
Mc|vn|p2vn(vnv)𝑑Vg,\displaystyle\qquad-\int_{M}c\,|v_{n}|^{p-2}v_{n}(v_{n}-v)dV_{g},

and so, vnHg1(M)vHg1(M)\|v_{n}\|_{H^{1}_{g}(M)}\to\|v\|_{H^{1}_{g}(M)} as nn\to\infty if both quantities to the left and to the right of the previous expression converge to zero.

We first deal with the term to the left. Since (vn)n(v_{n})_{n\in\mathbb{N}} is bounded, then (vnv)n(v_{n}-v)_{n\in\mathbb{N}} is also bounded, say by a positive constant C3C_{3}, and thus we get,

(14) |J(vn)(vnv)|J(vn)(Hg1(M))vnvHg1(M)C3J(vn)(Hg1(M))0.\displaystyle\begin{split}\left|J^{\prime}(v_{n})(v_{n}-v)\right|&\leq\|J^{\prime}(v_{n})\|_{(H_{g}^{1}(M)^{\mathcal{F}})^{\ast}}\|v_{n}-v\|_{H_{g}^{1}(M)}\\ &\leq C_{3}\|J^{\prime}(v_{n})\|_{(H_{g}^{1}(M)^{\mathcal{F}})^{\ast}}\rightarrow 0.\end{split}

For the term to the right, we proceed as follows. Observe that by þH when 2mκ2\geqslant m-\kappa_{\mathcal{F}}, then for any 2<p<2m2<p<2^{\ast}_{m} the map Hg1(M)Lgp(M)H^{1}_{g}(M)^{\mathcal{F}}\to L^{p}_{g}(M) is compact. In the case when mκ>2m-\kappa_{\mathcal{F}}>2 due to the fact that κ1\kappa_{\mathcal{F}}\geqslant 1 we get that 2m<2(mκ)/(mκ2)2^{\ast}_{m}<2(m-\kappa_{\mathcal{F}})/(m-\kappa_{\mathcal{F}}-2), and thus for 1p<2(mκ)/(mκ2)1\leqslant p<2(m-\kappa_{\mathcal{F}})/(m-\kappa_{\mathcal{F}}-2) the map Hg1(M)Lgp(M)H^{1}_{g}(M)^{\mathcal{F}}\to L^{p}_{g}(M) is also compact by þH.

Hence, up to a subsequence, we have that (vn)n(v_{n})_{n\in\mathbb{N}} converges weakly to vv in Lgp(M)L_{g}^{p}(M). Using this, Hölder’s inequality for p1p+1p=1\tfrac{p-1}{p}+\tfrac{1}{p}=1, Sobolev inequality for the embedding Hg1(M)Lgp(M)H_{g}^{1}(M)\hookrightarrow L_{g}^{p}(M), the fact that (vn)n(v_{n})_{n\in\mathbb{N}} is bounded in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} and ccLg(M)c\leq\|c\|_{L^{\infty}_{g}(M)}, we get that

(15) |Mc|vn|p2vn(vnv)dVg|Mc|vn|p1|vnv|dVgcLg(M)|vn|Lgp(M)p1|vnv|Lgp(M)cLg(M)|vnv|Lgp(M)(C1p1vnHg1(M)p1)C|vnv|Lgp(M)0,\displaystyle\begin{split}\Big{|}\int_{M}c\,|v_{n}&|^{p-2}v_{n}(v_{n}-v)dV_{g}\Big{|}\leq\int_{M}c\,|v_{n}|^{p-1}|v_{n}-v|dV_{g}\\ &\leq\|c\|_{L^{\infty}_{g}(M)}|v_{n}|_{L^{p}_{g}(M)}^{p-1}|v_{n}-v|_{L^{p}_{g}(M)}\\ &\leq\|c\|_{L^{\infty}_{g}(M)}|v_{n}-v|_{L^{p}_{g}(M)}\left(C_{1}^{p-1}\|v_{n}\|^{p-1}_{H^{1}_{g}(M)}\right)\\ &\leq C|v_{n}-v|_{L^{p}_{g}(M)}\to 0,\end{split}

where CC denotes some positive constant.

Finally, from the expression we have for J(vn)(vnv)J^{\prime}(v_{n})(v_{n}-v), (14) and (15), it follows that vnHg1(M)vHg1(M)\|v_{n}\|_{H^{1}_{g}(M)}\to\|v\|_{H^{1}_{g}(M)} as nn\to\infty, as we wanted to prove. ∎

6. Variational Principle

This section is devoted to the proof of Theorem F. We adapt the proof of [12, Theorem 2.3] (see also [13, 17]) to our context using Lemma 6.1 below.


In what follows, ||Lg(M)|\cdot|_{L^{\infty}_{g}(M)} denotes the usual LL^{\infty} norm. Recall that c,b𝒞(M)c,b\in\mathcal{C}^{\infty}(M) are foliated functions with c>0c>0, and that |u|c,p|u|_{c,p} as defined in (7) is a norm in Lgp(M)L^{p}_{g}(M) equivalent to the standard norm of Lgp(M)L^{p}_{g}(M).

For any θ>max{1,μ,|b|}\theta>\max\{1,\mu,|b|_{\infty}\}, where μ\mu was defined in (1), we have a well defined interior product

u,vθ:=Mgu,gvg+θuvdVg,u,vHg1(M),\langle u,v\rangle_{\theta}:=\int_{M}\langle\nabla_{g}u,\nabla_{g}v\rangle_{g}+\theta\,uv\;dV_{g},\quad u,v\in H_{g}^{1}(M),

which induces a norm, θ\|\cdot\|_{\theta}, that is equivalent to the standard norm of Hg1(M)H_{g}^{1}(M).

Recall that we are studying the functional J:Hg1(M)J\colon H^{1}_{g}(M)\to\mathbb{R}, given by

J(u)=12ub21p|u|c,pp,J(u)=\tfrac{1}{2}\|u\|^{2}_{b}-\tfrac{1}{p}|u|_{c,p}^{p},

restricted to the space Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, and where b\|\cdot\|_{b} is the norm induced by the inner product defined in (6). For p(2,2m]p\in(2,2^{\ast}_{m}], this functional is of class C2C^{2} and its gradient with respect to the interior product ,θ\langle\cdot,\cdot\rangle_{\theta} defines a C1C^{1} function J:Hg1(M)Hg1(M)\nabla J\colon H_{g}^{1}(M)\to H_{g}^{1}(M). We next show that actually J:Hg1(M)Hg1(M)\nabla J\colon H_{g}^{1}(M)^{\mathcal{F}}\to H_{g}^{1}(M)^{\mathcal{F}}. To do so, we decompose the gradient of JJ at uHg1(M)u\in H^{1}_{g}(M) as follows.

First, note that with respect to ,θ\langle\cdot,\cdot\rangle_{\theta}, J(u)\nabla J(u), is the vector in Hg1(M)H^{1}_{g}(M) which satisfies:

J(u),vθ\displaystyle\langle\nabla J(u),v\rangle_{\theta} =J(u)v\displaystyle=J^{\prime}(u)v
=u,vθM(θb)uv𝑑VgMc|u|p2uv𝑑Vg.\displaystyle=\langle u,v\rangle_{\theta}-\int_{M}(\theta-b)uvdV_{g}-\int_{M}c|u|^{p-2}uvdV_{g}.

for any vHg1(M)v\in H^{1}_{g}(M) (see 8). Then, we need to apply the following result that we prove in Section 6.1.

Lemma 6.1.

þ Let fLg2(M)f\in L_{g}^{2}(M) and θ\theta a nonnegative constant. Then the non-homogeneous problem

Δv+θv=f,on M,\displaystyle-\Delta v+\theta v=f,\quad\text{on }M,

admits a unique solution in Hg1(M)H^{1}_{g}(M). Moreover, if ff is \mathcal{F}-invariant, then the solution lies in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}.

Thus, for each uHg1(M)u\in H_{g}^{1}(M), by applying þ6.1 to ff equal to (θb)u(\theta-b)u and c|u|p2uc|u|^{p-2}u, which are elements of Lg2(M)L_{g}^{2}(M), there exist unique solutions LuLu and GuGu to the non-homogeneous linear problems

Δg(Lu)+θLu\displaystyle-\Delta_{g}(Lu)+\theta Lu =(θb)u,on M\displaystyle=(\theta-b)u,\quad\text{on }M
Δg(Gu)+θGu\displaystyle-\Delta_{g}(Gu)+\theta Gu =c|u|p2u,on M.\displaystyle=c\left|u\right|^{p-2}u,\quad\text{on }M.

Note that these solutions are uniquely determined by the relations

(16) Lu,vθ=M(θb)uv𝑑Vg,Gu,vθ=Mc|u|p2uv𝑑Vg,\displaystyle\begin{split}\langle Lu,v\rangle_{\theta}&=\int_{M}(\theta-b)uv\;dV_{g},\quad\\ \langle Gu,v\rangle_{\theta}&=\int_{M}c|u|^{p-2}uv\;dV_{g},\end{split}

for every vHg1(M)v\in H_{g}^{1}(M). It follows that

J(u)=uLuGu, for every uHg1(M).\nabla J(u)=u-Lu-Gu,\quad\text{ for every }u\in H^{1}_{g}(M).

If uHg1(M)u\in H^{1}_{g}(M)^{\mathcal{F}}, then also (θb)u(\theta-b)u and c|u|p2c|u|^{p-2} are \mathcal{F}-invariant. Thus, Lemma 6.1 yields that the functions LuLu and GuGu are \mathcal{F}-invariant, and so we conclude that for uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}}, the function J(u)\nabla J(u) is \mathcal{F}-invariant as we claimed.

Recall that the restricted Nehari manifold was defined to be

𝒩g={uHg1(M)u0,ub2=|u|c,pp}.\mathcal{N}^{\mathcal{F}}_{g}=\{u\in H_{g}^{1}(M)^{\mathcal{F}}\mid u\neq 0,\ \|u\|_{b}^{2}=|u|_{c,p}^{p}\}.

Observe that if uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}}, then also u±Hg1(M)u^{\pm}\in H_{g}^{1}(M)^{\mathcal{F}}, where u+:=max{0,u},u^{+}:=\max\{0,u\}, u:=min{0,u}u^{-}:=\min\{0,u\}. The nontrivial \mathcal{F}-invariant sign-changing critical points of JJ, and thus, sign changing solutions to (Y) in virtue of Theorem G, must belong to the set

g:={u𝒩gu+,u𝒩g}.\mathcal{E}_{g}^{\mathcal{F}}:=\{u\in\mathcal{N}_{g}^{\mathcal{F}}\mid u^{+},u^{-}\in\mathcal{N}_{g}^{\mathcal{F}}\}.

This set is nonempty. Indeed, Theorem D gives the existence of at least two foliated smooth functions u1,u20u_{1},u_{2}\geq 0 with disjoint supports. Then recalling the definition of the projection onto the Nehari manifold given in (10), we can define the function u:=σ(u1)σ(u2)u:=\sigma(u_{1})-\sigma(u_{2}), which is an element in g\mathcal{E}_{g}^{\mathcal{F}}, given that u+=σ(u1)𝒩gu^{+}=\sigma(u_{1})\in\mathcal{N}_{g}^{\mathcal{F}} and u=σ(u2)𝒩gu^{-}=-\sigma(u_{2})\in\mathcal{N}_{g}^{\mathcal{F}}.

Let 𝒫:={uHg1(M)u0}\mathcal{P}^{\mathcal{F}}:=\{u\in H^{1}_{g}(M)^{\mathcal{F}}\mid u\geq 0\} be the convex cone of nonnegative functions. Then the set of functions in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} which do not change sign is given as 𝒫𝒫\mathcal{P}^{\mathcal{F}}\cup-\mathcal{P}^{\mathcal{F}}.

As the ,θ\langle\cdot,\cdot\rangle_{\theta}-gradient of JJ, J:Hg1(M)Hg1(M)-\nabla J\colon H_{g}^{1}(M)^{\mathcal{F}}\to H_{g}^{1}(M)^{\mathcal{F}}, is of class C1C^{1}, it is locally Lipschitz and thus for each uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}}, the Cauchy problem:

{tψ(t,u)=J(ψ(t,u))ψ(0,u)=u,\begin{cases}\frac{\partial}{\partial t}\psi(t,u)=-\nabla J(\psi(t,u))\\ \psi(0,u)=u,\end{cases}

has a unique solution defined for all 0t<T(u)0\leq t<T(u), where T(u)(0,)T(u)\in(0,\infty) is the maximal existence time of the solution. With this at hand, the negative gradient flow of JJ, is simply the map ψ:𝒢Hg1(M)\psi\colon\mathcal{G}\to H^{1}_{g}(M)^{\mathcal{F}} where 𝒢:={(t,u)uHg1(M),\mathcal{G}:=\{(t,u)\mid u\in H^{1}_{g}(M)^{\mathcal{F}}, 0t<T(u)}0\leq t<T(u)\} and ψ\psi is as above. A subset 𝒟\mathcal{D} of Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} is said to be strictly positively invariant under ψ\psi if

ψ(t,u)int𝒟 for every u𝒟 and t(0,T(u)).\psi(t,u)\in\text{int}\mathcal{D}\text{\qquad for every }u\in\mathcal{D}\text{ and }t\in(0,T(u)).

For any set 𝒟Hg1(M)\mathcal{D}\subset H_{g}^{1}(M)^{\mathcal{F}}, we define the set

𝒜(𝒟):={uHg1(M)ψ(t,u)𝒟 for some t(0,T(u))},\mathcal{A}(\mathcal{D}):=\{u\in H_{g}^{1}(M)^{\mathcal{F}}\mid\psi(t,u)\in\mathcal{D}\text{ for some }t\in(0,T(u))\},

and the entrance time e𝒟:𝒜(𝒟)e_{\mathcal{D}}\colon\mathcal{A}(\mathcal{D})\to\mathbb{R}, given by

e𝒟(u):=inf{t0ψ(t,u)𝒟}.e_{\mathcal{D}}(u):=\inf\{t\geq 0\mid\psi(t,u)\in\mathcal{D}\}.

If 𝒟\mathcal{D} is strictly positively invariant under ψ\psi, then 𝒜(𝒟)\mathcal{A}(\mathcal{D}) is open in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} and e𝒟(u)e_{\mathcal{D}}(u) is continuous.

In what follows, for any subset 𝒟\mathcal{D} of Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} and any ε>0\varepsilon>0, B¯ε(𝒟)\overline{B}_{\varepsilon}(\mathcal{D}) will denote the set

B¯ε(𝒟):={uHg1(M)distθ(u,𝒟):=infv𝒟uvθε}.\overline{B}_{\varepsilon}(\mathcal{D}):=\{u\in H^{1}_{g}(M)^{\mathcal{F}}\mid\text{dist}_{\theta}(u,\mathcal{D}):=\inf_{v\in\mathcal{D}}\left\|u-v\right\|_{\theta}\leq\varepsilon\}.

To find sign-changing critical points of JJ we use the relative genus between symmetric subsets of Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. But before that we need the following preliminary result that we prove in Subsection 6.1.

Lemma 6.2.

þ þ There exists α0>0\alpha_{0}>0 such that for every α(0,α0),\alpha\in(0,\alpha_{0}),

  1. (a)

    B¯α(𝒫)g=\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\cap\mathcal{E}_{g}^{\mathcal{F}}=\emptyset, B¯α(𝒫)g=\overline{B}_{\alpha}(-\mathcal{P}^{\mathcal{F}})\cap\mathcal{E}_{g}^{\mathcal{F}}=\emptyset, and

  2. (b)

    B¯α(𝒫)\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}) and B¯α(𝒫)\overline{B}_{\alpha}(-\mathcal{P}^{\mathcal{F}}) are strictly positively invariant under the flow of the negative gradient of JJ with respect to ,θ\langle\cdot,\cdot\rangle_{\theta}.

Fix α\alpha as in the previous lemma. For dd\in\mathbb{R}, set

𝒟d:=B¯α(𝒫)B¯α(𝒫)Jd,\mathcal{D}_{d}^{\mathcal{F}}:=\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\cup\overline{B}_{\alpha}(\mathcal{-P}^{\mathcal{F}})\cup J^{d},

where Jd:={uHg1(M)J(u)d}J^{d}:=\{u\in H^{1}_{g}(M)^{\mathcal{F}}\mid J(u)\leq d\}.

The next result says that, under suitable conditions, 𝒟d\mathcal{D}_{d}^{\mathcal{F}} is a neighborhood retract.

Lemma 6.3.

þ If JJ has no sign-changing critical point uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} with J(u)=dJ(u)=d, then 𝒟d\mathcal{D}_{d}^{\mathcal{F}} is strictly positively invariant under ψ\psi and the map

ρd:𝒜(𝒟d)𝒟d,ρd(u):=ψ(e𝒟d(u),u),\rho_{d}\colon\mathcal{A}(\mathcal{D}_{d}^{\mathcal{F}})\to\mathcal{D}_{d}^{\mathcal{F}},\quad\rho_{d}(u):=\psi(e_{\mathcal{D}_{d}^{\mathcal{F}}}(u),u),

is odd and continuous, and satisfies ρd(u)=u\rho_{d}(u)=u for every u𝒟du\in\mathcal{D}_{d}^{\mathcal{F}}.

Proof.

Note that by definition ρd\rho_{d} is odd. To show that 𝒟d\mathcal{D}_{d}^{\mathcal{F}} is strictly positively invariant under ψ\psi, by þ6.2, it suffices to consider uJdu\in J^{d}. By definition of the flow, given any uJdu\in J^{d}, we have that

(17) ddtJψ(t,u)=J(ψ(t,u))tψ(t,u)=J(ψ(t,u))(J(ψ(t,u)))=J(ψ(t,u))θ20,t[0,T(u)).\begin{split}\frac{d}{dt}J\circ\psi(t,u)&=J^{\prime}(\psi(t,u))\frac{\partial}{\partial t}\psi(t,u)\\ &=J^{\prime}(\psi(t,u))(-\nabla J(\psi(t,u)))\\ &=-\|\nabla J(\psi(t,u))\|_{\theta}^{2}\leq 0,\quad t\in[0,T(u)).\end{split}

Thus we conclude that ψ(t,u)Jd\psi(t,u)\in J^{d} for every t[0,T(u))t\in[0,T(u)). So, if J(u)<dJ(u)<d, it follows that ψ(t,u)intJd\psi(t,u)\in\text{int}J^{d} for every t[0,T(u))t\in[0,T(u)). Next, suppose that J(u)=dJ(u)=d. By hypothesis, uu cannot be a sign changing critical point for JJ in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, hence we have two possible cases: either uu is a critical point that does not change sign, or J(u)0J^{\prime}(u)\neq 0.

The first case reduces to the first paragraph of the proof, that is we have uB¯α(𝒫)B¯α(𝒫)u\in\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\cup\overline{B}_{\alpha}(-\mathcal{P}^{\mathcal{F}}), and hence there is nothing to prove. Now, suppose that the second case holds true. Then, there exist real numbers β,ε>0\beta,\varepsilon>0 such that J(v)θ>β\|\nabla J(v)\|_{\theta}>\beta for every v[dε,d+ε]v\in[d-\varepsilon,d+\varepsilon]. Thus, J(ψ(t,u))θβ>0\|\nabla J(\psi(t,u))\|_{\theta}\geq\beta>0 for tt small enough, and (17) yields that ψ(t,u)intJd\psi(t,u)\in\text{int}J^{d}. Thus, 𝒟d\mathcal{D}_{d}^{\mathcal{F}} is strictly positively invariant under ψ\psi.

It is now easy to check that ρd\rho_{d} has the desired properties. ∎

Remark 6.4.

þ Observe that 𝒟0\mathcal{D}_{0}^{\mathcal{F}} is always strictly positively invariant under the flow ψ\psi, for there are no nontrivial critical points of JJ satisfying J(u)=0J(u)=0. In fact, if uu is a critical point of JJ satisfying J(u)=0J(u)=0, then 0=J(u)u=ub|u|c,p0=J^{\prime}(u)u=\|u\|_{b}-|u|_{c,p}; this implies that J(u)=p22pub=0J(u)=\frac{p-2}{2p}\|u\|_{b}=0 and u=0u=0.

Remark 6.5.

þ Notice that a critical point of JJ changes sign if and only if it lies in the complement of 𝒟0\mathcal{D}_{0}^{\mathcal{F}} in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}}. Indeed, if uu lies in Hg1(M)𝒟0H^{1}_{g}(M)^{\mathcal{F}}\smallsetminus\mathcal{D}_{0}^{\mathcal{F}}, then u𝒫(𝒫)u\notin\mathcal{P}^{\mathcal{F}}\cup(-\mathcal{P}^{\mathcal{F}}) and it must change sign. Conversely, if uu is a sign changing critical point of JJ, as we pointed before, it belongs to g\mathcal{E}_{g}^{\mathcal{F}} and þ6.2 says that this set has empty intersection with B¯α(𝒫)B¯α(𝒫)\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\cup\overline{B}_{\alpha}(-\mathcal{P}^{\mathcal{F}}); moreover, as it is a nontrivial critical point, we have that J(u)=p22pub>0J(u)=\frac{p-2}{2p}\|u\|_{b}>0 and uJ0u\notin J^{0}. Thus, u𝒟0u\notin\mathcal{D}_{0}^{\mathcal{F}}.

A subset 𝒴\mathcal{Y} of Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} will be called symmetric if u𝒴-u\in\mathcal{Y} for every u𝒴u\in\mathcal{Y}.

Definition 6.6.

Let 𝒟\mathcal{D} and 𝒴\mathcal{Y} be symmetric subsets of Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. The genus of 𝒴\mathcal{Y} relative to 𝒟\mathcal{D}, denoted by 𝔤(𝒴,𝒟)\mathfrak{g}(\mathcal{Y},\mathcal{D}), is the smallest number nn such that 𝒴\mathcal{Y} can be covered by n+1n+1 open symmetric subsets 𝒰0,𝒰1,,𝒰n\;\mathcal{U}_{0},\mathcal{U}_{1},\ldots,\mathcal{U}_{n} of Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} with the following two properties:

  • (i)

    𝒴𝒟𝒰0\mathcal{Y}\cap\mathcal{D}\subset\mathcal{U}_{0} and there exists an odd continuous map ϑ0:𝒰0𝒟\vartheta_{0}\colon\mathcal{U}_{0}\to\mathcal{D} such that ϑ0(u)=u\vartheta_{0}(u)=u for u𝒴𝒟u\in\mathcal{Y}\cap\mathcal{D}.

  • (ii)

    There exist odd continuous maps ϑj:𝒰j{1,1}\vartheta_{j}\colon\mathcal{U}_{j}\to\{1,-1\} for every j=1,,nj=1,\ldots,n.

If no such cover exists, we define 𝔤(𝒴,𝒟):=\mathfrak{g}(\mathcal{Y},\mathcal{D}):=\infty.

As in [13, Section 3] in order to obtain a variational principle for sign changing solutions, we need a refined version of the (PS)τ(PS)^{\mathcal{F}}_{\tau} condition. Given 𝒟Hg1(M)\mathcal{D}\subseteq H_{g}^{1}(M)^{\mathcal{F}}, we say that JJ satisfies the (PS)τ(PS)_{\tau}^{\mathcal{F}} condition relative to 𝒟\mathcal{D} in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, if every sequence (un)n(u_{n})_{n\in\mathbb{N}} in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} such that

un𝒟,J(un)τ,J(un)0 in (Hg1(M))u_{n}\notin\mathcal{D},\quad J(u_{n})\to\tau,\quad J^{\prime}(u_{n})\to 0\text{ in }(H_{g}^{1}(M)^{\mathcal{F}})^{\ast}

has a strongly convergent subsequence in Hg1(M)H^{1}_{g}(M). When 𝒟=\mathcal{D}=\emptyset, we recover the (PS)τ(PS)_{\tau}^{\mathcal{F}} condition given in Section 2.3. Also notice that if JJ satisfies the (PS)τ(PS)_{\tau}^{\mathcal{F}} condition in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, then it satisfies the condition relative to any subset of Hg1(M).H_{g}^{1}(M)^{\mathcal{F}}.

Lemma 6.7.

þ Fix α(0,α0)\alpha\in(0,\alpha_{0}) with α0\alpha_{0} given as in Lemma 6.2. For jj\in\mathbb{N}, define

τj:=inf{τ𝔤(𝒟τ,𝒟0)j}.\tau_{j}:=\inf\{\tau\in\mathbb{R}\mid\mathfrak{g}(\mathcal{D}_{\tau}^{\mathcal{F}},\mathcal{D}_{0}^{\mathcal{F}})\geq j\}.

Assume that JJ satisfies (PS)τj(PS)_{\tau_{j}}^{\mathcal{F}} relative to 𝒟0\mathcal{D}_{0}^{\mathcal{F}} in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}}. Then, the following statements hold true:

  • (a)

    JJ has a sign-changing critical point uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} with J(u)=τjJ(u)=\tau_{j}.

  • (b)

    If τj=τj+1\tau_{j}=\tau_{j+1}, then JJ has infinitely many sign-changing critical points uHg1(M)u\in H^{1}_{g}(M)^{\mathcal{F}} with J(u)=τjJ(u)=\tau_{j}.

Consequently, for dd\in\mathbb{R}, if JJ satisfies (PS)τ(PS)_{\tau}^{\mathcal{F}} relative to 𝒟0\mathcal{D}_{0}^{\mathcal{F}} in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} for every τd\tau\leq d, then JJ has at least 𝔤(𝒟d,𝒟0)\mathfrak{g}(\mathcal{D}_{d}^{\mathcal{F}},\mathcal{D}_{0}^{\mathcal{F}}) distinct pairs of sign-changing critical points: ±u1,,±uk\pm u_{1},...,\pm u_{k}, k𝔤(𝒟d,𝒟0)k\geq\mathfrak{g}(\mathcal{D}_{d}^{\mathcal{F}},\mathcal{D}_{0}^{\mathcal{F}}), in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} with J(±ui)dJ(\pm u_{i})\leq d for all i=1,,ki=1,...,k.

Proof.

The proof is exactly the same as that of Proposition 3.6 in [13], but we include it for the sake of completeness.

We prove part (a)(a) by contradiction, using the fact that 𝒟0\mathcal{D}_{0}^{\mathcal{F}} is strictly positively invariant under the flow ψ\psi (see þ6.4). More precisely, we assume that there does not exist a sign-changing critical point uHg1(M)u\in H^{1}_{g}(M)^{\mathcal{F}} with J(u)=τjJ(u)=\tau_{j}. We claim that the (PS)τj(PS)_{\tau_{j}}^{\mathcal{F}} condition relative to 𝒟0\mathcal{D}_{0}^{\mathcal{F}} implies the existence of β>0\beta>0 and ε(0,τj)\varepsilon\in(0,\tau_{j}) such that

(18) J(u)θβ>0,for every uJ1[τjε,τj+ε]𝒟0.\|\nabla J(u)\|_{\theta}\geq\beta>0,\quad\text{for every }u\in J^{-1}[\tau_{j}-\varepsilon,\tau_{j}+\varepsilon]\setminus\mathcal{D}_{0}^{\mathcal{F}}.

To see this, suppose in order to get a contradiction, the existence of a sequence (un)n(u_{n})_{n\in\mathbb{N}} in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} such that J(un)τjJ(u_{n})\to\tau_{j} and J(un)θ0\|\nabla J(u_{n})\|_{\theta}\to 0. Then (un)n(u_{n})_{n\in\mathbb{N}} is a (PS)τj(PS)_{\tau_{j}} sequence, and since JJ satisfies the (PS)τj(PS)_{\tau_{j}} condition relative to 𝒟0\mathcal{D}_{0}^{\mathcal{F}} there exists uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} such that, up to a subsequence which we also denote by unu_{n}, it holds that unuu_{n}\to u strongly in Hg1(M)H_{g}^{1}(M). As JJ and JJ^{\prime} are continuous, it follows that J(u)=τjJ(u)=\tau_{j} and J(u)=0J^{\prime}(u)=0. By hypothesis, uu cannot change sign. Hence uu is a critical point of JJ lying in 𝒫(𝒫)int𝒟0\mathcal{P}^{\mathcal{F}}\cup(-\mathcal{P}^{\mathcal{F}})\subset\text{int}\mathcal{D}_{0}^{\mathcal{F}}, which is open. Therefore, there exists n0n_{0} such that un𝒟0u_{n}\in\mathcal{D}_{0}^{\mathcal{F}} for every nn0n\geq n_{0}, which is a contradiction and the claim follows.

As we pointed out in þ6.5, every sign changing critical point lies in Hg1(M)𝒟0H_{g}^{1}(M)^{\mathcal{F}}\smallsetminus\mathcal{D}_{0}^{\mathcal{F}}, and (18) implies that there are no sign changing critical points of JJ in J1[τjε,τj+ε]J^{-1}[\tau_{j}-\varepsilon,\tau_{j}+\varepsilon]. Therefore 𝒟d\mathcal{D}_{d}^{\mathcal{F}} is strictly positively invariant under ψ\psi for every d[τjε,τj+ε]d\in[\tau_{j}-\varepsilon,\tau_{j}+\varepsilon] by þ6.3. As Jθ>0\|\nabla J\|_{\theta}>0, identity (18) yields that 𝒟τj+ε\mathcal{D}_{\tau_{j}+\varepsilon}^{\mathcal{F}} flows under ψ\psi to 𝒟τjε\mathcal{D}_{\tau_{j}-\varepsilon}^{\mathcal{F}} and 𝒟τj+ε𝒜(𝒟τjε)\mathcal{D}_{\tau_{j}+\varepsilon}^{\mathcal{F}}\subset\mathcal{A}(\mathcal{D}_{\tau_{j}-\varepsilon}^{\mathcal{F}}). In this way, þ6.3 implies that ρτjε:𝒟τi+ε𝒟τiε\rho_{\tau_{j}-\varepsilon}\colon\mathcal{D}_{\tau_{i}+\varepsilon}^{\mathcal{F}}\to\mathcal{D}_{\tau_{i}-\varepsilon}^{\mathcal{F}} is odd, continuous and ρτjε(u)=u\rho_{\tau_{j}-\varepsilon}(u)=u for every u𝒟0u\in\mathcal{D}_{0}^{\mathcal{F}}, given that 𝒟0𝒟τiε\mathcal{D}_{0}^{\mathcal{F}}\subset\mathcal{D}_{\tau_{i}-\varepsilon}^{\mathcal{F}}. As 𝒟0\mathcal{D}_{0}^{\mathcal{F}} is a symmetric neighborhood retract by þ6.4 and þ6.3, from the monotonicity of the genus 𝔤\mathfrak{g} [13, Lemma 3.4] it follows that j𝔤(𝒟τj+ε,𝒟0)𝔤(𝒟τjε,𝒟0)<jj\leq\mathfrak{g}(\mathcal{D}_{\tau_{j}+\varepsilon}^{\mathcal{F}},\mathcal{D}_{0}^{\mathcal{F}})\leq\mathfrak{g}(\mathcal{D}_{\tau_{j}-\varepsilon}^{\mathcal{F}},\mathcal{D}_{0}^{\mathcal{F}})<j which is a contradiction.

The proof of part (b)(b) is exactly the same as the proof of Proposition 3.6 (b) in [13]. The argument uses again the fact that 𝒟0\mathcal{D}_{0}^{\mathcal{F}} is strictly positively invariant under the flow ψ\psi, the contradicting hypothesis, and, the monotonicity and subadditive properties of the genus (see [13, Lemma 3.4]) to get j+1𝔤(𝒟τj+ε,𝒟0)𝔤(𝒟τjε,𝒟0)+1<j+1j+1\leq\mathfrak{g}(\mathcal{D}_{\tau_{j}+\varepsilon}^{\mathcal{F}},\mathcal{D}_{0}^{\mathcal{F}})\leq\mathfrak{g}(\mathcal{D}_{\tau_{j}-\varepsilon}^{\mathcal{F}},\mathcal{D}_{0}^{\mathcal{F}})+1<j+1. ∎

We now prove þF. We start by showing the existence of a positive solution of minimal energy using the ideas developed in Section 5. Then we show the existence of the sign-changing solutions following the proof of Theorem 3.7 in [13].

Proof of Theorem F.

We first show the existence of a positive critical point attaining τg=inf{J(u)u𝒩g}\tau_{g}^{\mathcal{F}}=\inf\{J(u)\mid u\in\mathcal{N}_{g}^{\mathcal{F}}\}. Let un𝒩gu_{n}\in\mathcal{N}_{g}^{\mathcal{F}} such that J(un)τgJ(u_{n})\rightarrow\tau_{g}^{\mathcal{F}}. Since un𝒩gu_{n}\in\mathcal{N}_{g}^{\mathcal{F}} we get

J(un)=12unb21p|un|c,pp=p22p|un|c,pp.J(u_{n})=\frac{1}{2}\|u_{n}\|_{b}^{2}-\frac{1}{p}|u_{n}|_{c,p}^{p}=\frac{p-2}{2p}|u_{n}|_{c,p}^{p}.

Thus, |un|c,pp|u_{n}|_{c,p}^{p} is bounded in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}. As 2<p2m2<p\leq 2_{m}^{\ast} and κ1\kappa_{\mathcal{F}}\geq 1, as in the proof of Theorem E, there exists uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} such that unu_{n} converges weakly to uu in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} and strongly in Lgp(M)L_{g}^{p}(M). Then

|u|c,pp=limn|un|c,pp=2pp2τg>0,|u|_{c,p}^{p}=\lim_{n\rightarrow\infty}|u_{n}|_{c,p}^{p}=\frac{2p}{p-2}\tau_{g}^{\mathcal{F}}>0,

implying that u0u\neq 0. Therefore, there exists tu>0t_{u}>0 such that σ(u)=tuu𝒩g\sigma(u)=t_{u}u\in\mathcal{N}_{g}^{\mathcal{F}}, where σ\sigma is the projection onto 𝒩g\mathcal{N}_{g}. As un𝒩gu_{n}\in\mathcal{N}_{g}^{\mathcal{F}}, identity (11) yields that J(tuun)J(un)J(t_{u}u_{n})\leq J(u_{n}). Therefore, using basic properties of weak convergence and that unuu_{n}\rightarrow u in Lgp(M)L_{g}^{p}(M), we obtain that

τgJ(tuu)=12tuub21p|tuu|c,pplim infn(12tuunb2)lim infn(1p|tuun|c,pp)=lim infnJ(tuun)limnJ(un)=τg.\begin{split}\tau_{g}^{\mathcal{F}}&\leq J(t_{u}u)=\frac{1}{2}\|t_{u}u\|_{b}^{2}-\frac{1}{p}|t_{u}u|_{c,p}^{p}\\ &\leq\liminf_{n\rightarrow\infty}\left(\frac{1}{2}\|t_{u}u_{n}\|_{b}^{2}\right)-\liminf_{n\rightarrow\infty}\left(\frac{1}{p}|t_{u}u_{n}|_{c,p}^{p}\right)\\ &=\liminf_{n\rightarrow\infty}J(t_{u}u_{n})\leq\lim_{n\rightarrow\infty}J(u_{n})=\tau_{g}^{\mathcal{F}}.\end{split}

Again, since unuu_{n}\rightarrow u strongly in Lgp(M)L^{p}_{g}(M), we conclude from this inequalities that limnunb2\lim_{n\rightarrow\infty}\|u_{n}\|^{2}_{b} exists and is equal to ub2\|u\|^{2}_{b}. Hence unuu_{n}\rightarrow u strongly in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}}, and since 𝒩g\mathcal{N}_{g}^{\mathcal{F}} is closed in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} then it follows that u𝒩gu\in\mathcal{N}_{g}^{\mathcal{F}}, and J(u)=τgJ(u)=\tau_{g}^{\mathcal{F}}. As 𝒩g\mathcal{N}_{g}^{\mathcal{F}} is a natural restriction for the functional JJ (see [51, Chapter 4]), uu is a nontrivial \mathcal{F}-invariant critical point for JJ in Hg1(M)H_{g}^{1}(M)^{\mathcal{F}} attaining τg\tau_{g}^{\mathcal{F}}.

Now we see that uu does not change sign. Suppose, in order to get a contradiction, that this is not true. Then, u+0u^{+}\neq 0 and u0u^{-}\neq 0. As u±Hg1(M)u^{\pm}\in H_{g}^{1}(M) and as uu is a critical point for JJ, we have that

0=J(u)u±=M[gu,g(u±)g+bu(u±)]Mc|u|p2u(u±)=u±b2|u±|c,pp,\begin{split}0&=J^{\prime}(u)u^{\pm}=\int_{M}[\langle\nabla_{g}u,\nabla_{g}(u^{\pm})\rangle_{g}+bu(u^{\pm})]-\int_{M}c|u|^{p-2}u(u^{\pm})\\ &=\|u^{\pm}\|_{b}^{2}-|u^{\pm}|_{c,p}^{p},\end{split}

concluding that u±𝒩gu^{\pm}\in\mathcal{N}_{g}^{\mathcal{F}}. Hence

τg=J(u)=J(u+)+J(u)2infv𝒩gJ(v)=2τg,\tau_{g}^{\mathcal{F}}=J(u)=J(u^{+})+J(u^{-})\geq 2\inf_{v\in\mathcal{N}_{g}^{\mathcal{F}}}J(v)=2\tau_{g}^{\mathcal{F}},

which is a contradiction since τg>0\tau_{g}^{\mathcal{F}}>0. Thus uu does not change sign. If u0u\leq 0, we can take u-u, since it is also a critical point for JJ and it is positive.

We proceed to prove the existence of the sign-changing critical points: Let d:=supWJd:=\sup_{W}J. By Lemma 6.7, we only need to show that

n:=𝔤(𝒟d,𝒟0)dim(W)1.n:=\mathfrak{g}\left(\mathcal{D}_{d}^{\mathcal{F}},\mathcal{D}_{0}^{\mathcal{F}}\right)\geq\dim(W)-1.

Let 𝒰0,𝒰1,,𝒰n\mathcal{U}_{0},\mathcal{U}_{1},\ldots,\mathcal{U}_{n} be open symmetric subsets of Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} covering 𝒟d\mathcal{D}_{d}^{\mathcal{F}} with 𝒟0𝒰0\mathcal{D}_{0}^{\mathcal{F}}\subset\mathcal{U}_{0} and let ϑ0:𝒰0𝒟0\vartheta_{0}\colon\mathcal{U}_{0}\to\mathcal{D}_{0}^{\mathcal{F}} and ϑj:𝒰j{1,1}\vartheta_{j}\colon\mathcal{U}_{j}\to\{1,-1\}, j=1,,nj=1,\ldots,n, be odd continuous maps such that ϑ0(u)=u\vartheta_{0}(u)=u for all u𝒟0u\in\mathcal{D}_{0}^{\mathcal{F}}. Since Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} is an absolute retract, we may assume that ϑ0\vartheta_{0} is the restriction of an odd continuous map ϑ~0:Hg1(M)Hg1(M)\widetilde{\vartheta}_{0}\colon H^{1}_{g}(M)^{\mathcal{F}}\to H^{1}_{g}(M)^{\mathcal{F}}. Let \mathcal{B} be the connected component of the complement of the Nehari manifold 𝒩g\mathcal{N}_{g}^{\mathcal{F}} in Hg1(M)H^{1}_{g}(M)^{\mathcal{F}} which contains the origin, and set 𝒪:={uWϑ~0(u)}\mathcal{O}:=\{u\in W\mid\widetilde{\vartheta}_{0}(u)\in\mathcal{B}\}. Then, 𝒪\mathcal{O} is a bounded open symmetric neighborhood of 0 in WW.

Let 𝒱j:=𝒰j𝒪\mathcal{V}_{j}:=\mathcal{U}_{j}\cap\partial\mathcal{O}. Then, 𝒱0,𝒱1,,𝒱n\mathcal{V}_{0},\mathcal{V}_{1},\ldots,\mathcal{V}_{n} are symmetric and open in 𝒪\partial\mathcal{O}, and they cover 𝒪\partial\mathcal{O}. Further, by Lemma 6.2,

ϑ0(𝒱0)𝒟0𝒩g𝒩gg.\vartheta_{0}(\mathcal{V}_{0})\subset\mathcal{D}_{0}^{\mathcal{F}}\cap\mathcal{N}_{g}^{\mathcal{F}}\subset\mathcal{N}_{g}^{\mathcal{F}}\smallsetminus\mathcal{E}_{g}^{\mathcal{F}}.

The set 𝒩gg\mathcal{N}_{g}^{\mathcal{F}}\smallsetminus\mathcal{E}_{g}^{\mathcal{F}} consists of two connected components, see for example [10, Lemmas 2.5 and 2.6]. Therefore there exists an odd continuous map η:𝒩gg{1,1}\eta\colon\mathcal{N}_{g}^{\mathcal{F}}\smallsetminus\mathcal{E}_{g}^{\mathcal{F}}\to\{1,-1\}. Let ηj:𝒱j{1,1}\eta_{j}\colon\mathcal{V}_{j}\to\{1,-1\} be the restriction of the map ηϑ0\eta\circ\vartheta_{0} if j=0,j=0, and the restriction of ϑj\vartheta_{j} if j=1,,nj=1,\ldots,n. Take a partition of the unity {πj:𝒪[0,1]j=0,1,,n}\{\pi_{j}\colon\partial\mathcal{O}\to[0,1]\mid j=0,1,\ldots,n\} subordinated to the cover {𝒱0,𝒱1,,𝒱n}\{\mathcal{V}_{0},\mathcal{V}_{1},\ldots,\mathcal{V}_{n}\} consisting of even functions, and let {e1,,en+1}\{e_{1},\ldots,e_{n+1}\} be the canonical basis of n+1\mathbb{R}^{n+1}. Then, the map Ψ:𝒪n+1\Psi\colon\partial\mathcal{O}\to\mathbb{R}^{n+1} given by

Ψ(u):=j=0nηj(u)πj(u)ej+1\Psi(u):=\sum_{j=0}^{n}\eta_{j}(u)\pi_{j}(u)e_{j+1}

is odd and continuous, and satisfies Ψ(u)0\Psi(u)\neq 0 for every u𝒪u\in\partial\mathcal{O}. The Borsuk-Ulam theorem allow us to conclude that dim(W)n+1,\dim(W)\leq n+1, as claimed. ∎

6.1. Proof of auxiliary results

Proof of Lemma 6.1.

Fix fLg2(M)f\in L^{2}_{g}(M), and define If:Hg1(M)I_{f}\colon H^{1}_{g}(M)\to\mathbb{R} as

If(u)=Mfu𝑑Vg.I_{f}(u)=\int_{M}fudV_{g}.

Observe that IfI_{f} is linear and bounded because fL2(M)f\in L^{2}(M) and since for any uHg1(M)u\in H^{1}_{g}(M) it holds that uLg2(M)u\in L_{g}^{2}(M). By the Frèchet-Riesz representation theorem, there exists a unique vfHg1(M)v_{f}\in H^{1}_{g}(M) such that the following equality holds for any uHg1(M)u\in H^{1}_{g}(M),

vf,uθ=If(u).\langle v_{f},u\rangle_{\theta}=I_{f}(u).

This implies that vfv_{f} is a weak solution in MM to the equation

Δgv+θv=f.-\Delta_{g}v+\theta v=f.

Now we show that if ff is \mathcal{F}-invariant then vfv_{f} is \mathcal{F}-invariant. Consider the operator Jθ:Hg1(M)J_{\theta}\colon H^{1}_{g}(M)\to\mathbb{R} given by

Jθ(u)=12uθIf(u).J_{\theta}(u)=\tfrac{1}{2}\|u\|_{\theta}-I_{f}(u).

Observe that for any u,wHg1(M)u,w\in H^{1}_{g}(M) we have that

Jθ(w)(u)=w,uθIf(w)(u).J^{\prime}_{\theta}(w)(u)=\langle w,u\rangle_{\theta}-I^{\prime}_{f}(w)(u).

Note that for any u,wHg1(M)u,w\in H^{1}_{g}(M),

If(w)(u)=If(u).I^{\prime}_{f}(w)(u)=I_{f}(u).

Thus we see that vfv_{f} is a critical value of JθJ_{\theta} by construction.

Recall that H=Hg1(M)H=H^{1}_{g}(M)^{\mathcal{F}} is a closed linear subspace of Hg1(M)H^{1}_{g}(M). We note that the orthogonal decomposition Hg1(M)=HHH^{1}_{g}(M)=H\oplus H^{\perp} with respect to the standard inner product is also an orthogonal decomposition with respect to ,θ\langle\,,\,\rangle_{\theta}. Indeed if uHu\in H and vHv\in H^{\perp}, so that u,vHg1(M)=0\langle u,v\rangle_{H^{1}_{g}(M)}=0, then we have

u,vθ=M(θ1)uv𝑑Vg.\langle u,v\rangle_{\theta}=\int_{M}(\theta-1)uvdV_{g}.

Since θ>1\theta>1 we conclude that the previous expression equals zero applying Lemma 4.3.

Write vf=vf+vfv_{f}=v_{f}^{\top}+v_{f}^{\perp}, with vfHv_{f}^{\top}\in H and vfHv_{f}^{\perp}\in H^{\perp}. We claim that vfv^{\top}_{f} is a critical point of JθJ_{\theta}. If the claim holds, then for any uHg1(M)u\in H^{1}_{g}(M) we have that If(u)=vf,uθI_{f}(u)=\langle v^{\top}_{f},u\rangle_{\theta}, and since we know that If(u)=vf,uθI_{f}(u)=\langle v_{f},u\rangle_{\theta} for any uHg1(M)u\in H^{1}_{g}(M), we conclude that vf=vfv_{f}=v_{f}^{\top}, that is, vfHg1(M)v_{f}\in H^{1}_{g}(M)^{\mathcal{F}} as desired.

Now we prove the claim. For uHu\in H,

Jθ(vf)(u)=\displaystyle J^{\prime}_{\theta}(v^{\top}_{f})(u)= vf,uθIf(vf)(u)\displaystyle\langle v_{f}^{\top},u\rangle_{\theta}-I^{\prime}_{f}(v^{\top}_{f})(u)
=\displaystyle= vf,uθIf(vf)(u)=J(vf)(u)=0.\displaystyle\langle v_{f},u\rangle_{\theta}-I^{\prime}_{f}(v_{f})(u)=J^{\prime}(v_{f})(u)=0.

This implies that vfv^{\top}_{f} is a critical point of Jθ|Hg1(M)J_{\theta}|_{H^{1}_{g}(M)^{\mathcal{F}}}. Now consider a fixed wHw\in H^{\perp}. Then,

Jθ(vf)(w)=vf,wθIf(vf)(w)=If(w)=Mfw𝑑Vg.J^{\prime}_{\theta}(v^{\top}_{f})(w)=\langle v^{\top}_{f},w\rangle_{\theta}-I^{\prime}_{f}(v^{\top}_{f})(w)=I_{f}(w)=\int_{M}fwdV_{g}.

Since fL2(M)f\in L^{2}(M)^{\mathcal{F}}, þ4.2 implies that vfv^{\top}_{f} is a critical point of Jθ|HJ_{\theta}|_{H^{\perp}}. By linearity, it follows that vfv^{\top}_{f} is a critical point of JθJ_{\theta}. ∎

We proceed to prove Lemma 6.2. The proof, up to minor modifications, is the same as in [12, Lemma 5.2], but we sketch it for the convenience of the reader.

Proof of Lemma 6.2 part (a)(a).

For every uHg1(M)u\in H^{1}_{g}(M)^{\mathcal{F}}, Sobolev’s inequality yields a positive constant CC such that

(19) distθ(u,𝒫)=minv𝒫uvθC1minv𝒫|uv|c,pC1|u|c,p,\text{dist}_{\theta}(u,\mathcal{P}^{\mathcal{F}})=\min_{v\in\mathcal{P}^{\mathcal{F}}}\left\|u-v\right\|_{\theta}\geq C^{-1}\min_{v\in\mathcal{P}^{\mathcal{F}}}\left|u-v\right|_{c,p}\geq C^{-1}|u^{-}|_{c,p},

where in the last part we used the fact that |uv||u|\left|u-v\right|\geq\left|u^{-}\right| for every u,v:Mu,v\colon M\to\mathbb{R} with v0v\geq 0, and u=min{0,u}u^{-}=-\min\{0,u\}.

Now we assume that ugu\in\mathcal{E}_{g}^{\mathcal{F}}. We bound |u|c,p|u^{-}|_{c,p} by noticing that u𝒩gu^{-}\in\mathcal{N}_{g}^{\mathcal{F}} implies that,

J(u)=12ub21p|u|c,pp=p22p|u|c,pp.J(u^{-})=\tfrac{1}{2}\|u^{-}\|^{2}_{b}-\tfrac{1}{p}|u^{-}|_{c,p}^{p}=\tfrac{p-2}{2p}|u^{-}|^{p}_{c,p}.

Then by definition of τg\tau_{g}^{\mathcal{F}}, see (12), we get

|u|c,pp=2pp2J(u)2pp2τg>0.|u^{-}|_{c,p}^{p}=\tfrac{2p}{p-2}J(u^{-})\geq\tfrac{2p}{p-2}\tau_{g}^{\mathcal{F}}>0.

Taking α1:=C1(2pp2τg)1p>0\alpha_{1}:=C^{-1}\left(\tfrac{2p}{p-2}\tau_{g}^{\mathcal{F}}\right)^{\tfrac{1}{p}}>0, it follows that distθ(u,𝒫)α1\mathrm{dist}_{\theta}(u,\mathcal{P}^{\mathcal{F}})\geq\alpha_{1} for all ugu\in\mathcal{E}^{\mathcal{F}}_{g}. Similarly, distθ(u,𝒫)α1\mathrm{dist}_{\theta}(u,-\mathcal{P}^{\mathcal{F}})\geq\alpha_{1} for all ugu\in\mathcal{E}^{\mathcal{F}}_{g}. Thus, B¯α(𝒫)g,B¯α(𝒫)g=\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\cap\mathcal{E}_{g}^{\mathcal{F}},\,\overline{B}_{\alpha}(-\mathcal{P}^{\mathcal{F}})\cap\mathcal{E}_{g}^{\mathcal{F}}=\emptyset, for every α(0,α1)\alpha\in(0,\alpha_{1}). ∎

To prove part (b)(b) of Lemma 6.2 we will use the following result.

Proposition 6.8 (Theorem 5.2 in [20]).

þ Let XX be a real vector space with a norm inducing a distance ρ\rho, ΩX\Omega\subset X open and DXD\subset X closed convex with non empty interior with DΩD\cap\Omega\neq\emptyset and such that the distance from any point in XX to DD is achieved by some point. Let f:(0,a)×ΩXf\colon(0,a)\times\Omega\to X be a locally Lipschitz function that satisfies

limλ0λ1ρ(u+λf(t,u),D)=0foruΩDandt(0,a).\lim_{\lambda\to 0}\lambda^{-1}\rho(u+\lambda f(t,u),D)=0\quad\text{for}\quad u\in\Omega\cap\partial D\quad\text{and}\quad t\in(0,a).

Then any continuous x:[0,b)Ωx\colon[0,b)\to\Omega such that x(0)Dx(0)\in D and x(t)=f(t,x(t))x^{\prime}(t)=f(t,x(t)) in (0,b)(0,b) satisfies x(t)Dx(t)\in D for all t[0,b)t\in[0,b).

Proof of Lemma 6.2 part (b)(b).

By symmetry, we will only prove this part for 𝒫)\mathcal{P}^{\mathcal{F}}). The proof consists in the following two steps:

Step 1: For α>0\alpha>0 small enough, we apply þ6.8 taking X=Ω=Hg1(M)X=\Omega=H_{g}^{1}(M)^{\mathcal{F}}, ρ=distθ\rho=\text{dist}_{\theta}, and D=B¯α(𝒫)D=\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}) which is closed and convex, and f(t,u)=J(u)f(t,u)=-\nabla J(u), to obtain that

ψ(t,u)B¯α(𝒫) for all t(0,T(u)) if uB¯α(𝒫).\psi(t,u)\in\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\text{ \ for all \ }t\in(0,T(u))\text{ \ if }u\in\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}).

Step 2: We then apply Mazur’s separation theorem [38, Section 2.2.19] to prove that

ψ(t,u)Bα(𝒫) for all t(0,T(u)) if uB¯α(𝒫).\psi(t,u)\in B_{\alpha}(\mathcal{P}^{\mathcal{F}})\text{ \ for all \ }t\in(0,T(u))\text{ \ if }u\in\partial\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}).

Proof of Step 1: Recall that the gradient of JJ with respect to ,θ\langle\cdot,\cdot\rangle_{\theta} is J(u)=uLuGu\nabla J(u)=u-Lu-Gu. In order to apply Proposition 6.8, for uB¯α(𝒫)u\in\partial\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}), we have to show that the limit as λ0\lambda\searrow 0 of the following expression equals zero:

λ1distθ(u+λ(J(u)),B¯α(𝒫))=λ1distθ((1λ)u+λ(Lu+Gu),B¯α(𝒫))λ1[(1λ)distθ(u,B¯α(𝒫))+λdistθ(Lu+Gu,B¯α(𝒫))]=distθ(Lu+Gu,B¯α(𝒫)).\begin{split}&\lambda^{-1}\text{dist}_{\theta}\left(u+\lambda(-\nabla J(u)),\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\right)\\ &=\lambda^{-1}\text{dist}_{\theta}\left((1-\lambda)u+\lambda(Lu+Gu),\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\right)\\ &\leq\lambda^{-1}\left[(1-\lambda)\text{dist}_{\theta}\left(u,\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\right)+\lambda\text{dist}_{\theta}\left(Lu+Gu,\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\right)\right]\\ &=\text{dist}_{\theta}\left(Lu+Gu,\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}})\right).\end{split}

This will be achieved by showing that Lu+GuBα(𝒫)Lu+Gu\in\,B_{\alpha}(\mathcal{P}^{\mathcal{F}}). In particular, we show that for some α2>0\alpha_{2}>0 and any α(0,α2)\alpha\in(0,\alpha_{2}),

distθ(Lu+Gu,𝒫)<distθ(u,𝒫) uB¯α(𝒫).\text{dist}_{\theta}(Lu+Gu,\mathcal{P}^{\mathcal{F}})<\text{dist}_{\theta}(u,\mathcal{P}^{\mathcal{F}})\text{\qquad}\forall u\in\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}).

Since distθ\text{dist}_{\theta} is given by a norm, by the triangle inequality it is enough to bound distθ(Lu,𝒫)\text{dist}_{\theta}(Lu,\mathcal{P}^{\mathcal{F}}) and distθ(Gu,𝒫)\text{dist}_{\theta}(Gu,\mathcal{P}^{\mathcal{F}}) separately.

We first note that Lv𝒫Lv\in\mathcal{P}^{\mathcal{F}} and Gv𝒫Gv\in\mathcal{P}^{\mathcal{F}} if v𝒫v\in\mathcal{P}^{\mathcal{F}}. This follows from the fact that MM is compact, θ>0\theta>0, θb0\theta-b\geq 0 and c>0c>0 and the Maximum Principle. Indeed consider Lv:MLv\colon M\to\mathbb{R} and denote by D(Lv)(p)D(Lv)(p) the derivative of LvLv at pMp\in M. Since MM is compact, then LvLv attains it minimum at some point pMp\in M. If we assume that Lv(p)<0Lv(p)<0, then since we also know that D(Lv)(p)=0D(Lv)(p)=0 and Δ(Lv)(p)0-\Delta(Lv)(p)\leqslant 0, we have (θb)v(p)=Δ(Lv)(p)+θLv(p)<0(\theta-b)v(p)=-\Delta(Lv)(p)+\theta Lv(p)<0, which is a contradiction. The same reasoning holds for GvGv.

As in [12, Lemma 5.1], the following inequality holds true

(20) Luθθμθ+μuθ.\|Lu\|_{\theta}\leq\frac{\theta-\mu}{\theta+\mu}\|u\|_{\theta}.

For uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} let v𝒫v\in\mathcal{P}^{\mathcal{F}} be such that dist(u,𝒫)θ=uvθ.{}_{\theta}(u,\mathcal{P}^{\mathcal{F}})=\left\|u-v\right\|_{\theta}. Then, the previous paragraph, linearity of LL and inequality (20) yield

(21) distθ(Lu,𝒫)LuLvθθμθ+μuvθ=θμθ+μdistθ(u,𝒫).\displaystyle\begin{split}\text{dist}_{\theta}(Lu,\mathcal{P}^{\mathcal{F}})&\leq\left\|Lu-Lv\right\|_{\theta}\\ &\leq\tfrac{\theta-\mu}{\theta+\mu}\left\|u-v\right\|_{\theta}\\ &=\tfrac{\theta-\mu}{\theta+\mu}\,\text{dist}_{\theta}(u,\mathcal{P}^{\mathcal{F}}).\end{split}

Now we bound distθ(Gu,𝒫)\text{dist}_{\theta}(Gu,\mathcal{P}^{\mathcal{F}}). For any function v:Mv\colon M\to\mathbb{R} we defined v±v^{\pm} as v+=max{0,v}v^{+}=\max\{0,v\} and v=min{0,v}v^{-}=\min\{0,v\}. Then, since for uHg1(M)u\in H^{1}_{g}(M)^{\mathcal{F}} we have (Gu)+𝒫(Gu)^{+}\in\mathcal{P}^{\mathcal{F}}, by the definition of distθ\text{dist}_{\theta}, using that G(u)+,G(u)θ=0\left\langle G(u)^{+},G(u)^{-}\right\rangle_{\theta}=0 and identity (16), we get

distθ(Gu,𝒫)G(u)θ\displaystyle\text{dist}_{\theta}(Gu,\mathcal{P}^{\mathcal{F}})\left\|G(u)^{-}\right\|_{\theta}\leqslant G(u)G(u)+θG(u)θ\displaystyle\|G(u)-G(u)^{+}\|_{\theta}\,\|G(u)^{-}\|_{\theta}
=\displaystyle= G(u)θ2\displaystyle\left\|G(u)^{-}\right\|_{\theta}^{2}
=\displaystyle= G(u),G(u)θ\displaystyle\left\langle G(u),G(u)^{-}\right\rangle_{\theta}
=\displaystyle= Mc|u|p2uG(u)𝑑Vg.\displaystyle\int_{M}c\left|u\right|^{p-2}uG(u)^{-}dV_{g}.

To bound the integral above, recall that c>0c>0 and G(u)0G(u)^{-}\leq 0. Thus, we can apply Hölder’s inequality, with p1p+1p=1\frac{p-1}{p}+\frac{1}{p}=1, to the functions c(p1)/p|u|p1c^{(p-1)/p}|u^{-}|^{p-1} and c1/p|G(u)|c^{1/p}|G(u)^{-}|, and then apply (19) and Sobolev’s inequality, to get

Mc|u|p2uG(u)𝑑Vg\displaystyle\int_{M}c\left|u\right|^{p-2}uG(u)^{-}dV_{g}
={u>0}c|u|p2uG(u)𝑑Vg+{u<0}c|u|p2uG(u)𝑑Vg\displaystyle=\int_{\{u>0\}}c\left|u\right|^{p-2}uG(u)^{-}dV_{g}+\int_{\{u<0\}}c\left|u\right|^{p-2}uG(u)^{-}dV_{g}
{u<0}c|u|p2uG(u)𝑑Vg\displaystyle\leq\int_{\{u<0\}}c\left|u\right|^{p-2}uG(u)^{-}dV_{g}
=Mc|u|p2uG(u)𝑑Vg\displaystyle=\int_{M}c|u^{-}|^{p-2}u^{-}G(u)^{-}\,dV_{g}
|u|c,pp1|G(u)|c,p\displaystyle\leq|u^{-}|_{c,p}^{p-1}|G(u)^{-}|_{c,p}
Cpdistθ(u,𝒫)p1G(u)θ.\displaystyle\leq C^{p}\,\text{dist}_{\theta}(u,\mathcal{P}^{\mathcal{F}})^{p-1}\|G(u)^{-}\|_{\theta}.

Combining the inequalities above, we conclude that for all uHg1(M)u\in H_{g}^{1}(M)^{\mathcal{F}} with Gu0\|Gu^{-}\|\neq 0 it holds

(22) distθ(Gu,𝒫)Cpdistθ(u,𝒫)p1.\text{dist}_{\theta}(Gu,\mathcal{P}^{\mathcal{F}})\leq C^{p}\,\text{dist}_{\theta}(u,\mathcal{P}^{\mathcal{F}})^{p-1}.

This inequality is also true if Gu=0\|Gu^{-}\|=0, for, in this case Gu=Gu+Gu=Gu^{+} and distθ(Gu,𝒫)=0\text{dist}_{\theta}(Gu,\mathcal{P}^{\mathcal{F}})=0. So, inequality (22) holds true for every uHg1(M).u\in H_{g}^{1}(M)^{\mathcal{F}}.

Fix ν(θμθ+μ,1)\nu\in\left(\tfrac{\theta-\mu}{\theta+\mu},1\right) and let α2>0\alpha_{2}>0 be such that Cpα2p2ν(θμθ+μ)C^{p}\alpha_{2}^{p-2}\leq\nu-\left(\tfrac{\theta-\mu}{\theta+\mu}\right). Then, for any α(0,α2),\alpha\in(0,\alpha_{2}), by adding inequalities (21) and (22) we obtain

distθ(Lu+Gu,𝒫)νdistθ(u,𝒫)<distθ(u,𝒫),\text{dist}_{\theta}(Lu+Gu,\mathcal{P}^{\mathcal{F}})\leq\nu\,\text{dist}_{\theta}(u,\mathcal{P}^{\mathcal{F}})<\text{dist}_{\theta}(u,\mathcal{P}^{\mathcal{F}}),

for all uB¯α(𝒫)u\in\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}). Therefore, Lu+GuBα(𝒫)Lu+Gu\in\,B_{\alpha}(\mathcal{P}^{\mathcal{F}}) if uB¯α(𝒫)u\in\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}).

Proof of Step 2: Observe that we have not ruled out ψ(u,t)Bα(𝒫)\psi(u,t)\in\partial B_{\alpha}(\mathcal{P}^{\mathcal{F}}). We now argue as in the proof of Lemma 2 in [17].

By contradiction, assume that there exists uB¯α(𝒫)u\in\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}) such that ψ(u,t)Bα(𝒫)\psi(u,t)\in\partial B_{\alpha}(\mathcal{P}^{\mathcal{F}}) for some t(0,T(u))t\in(0,T(u)). By Mazur’s separation theorem, there exists a continuous linear functional R(Hg1(M))R\in(H^{1}_{g}(M)^{\mathcal{F}})^{*} and β>0\beta>0 such that R(ψ(u,t))βR(\psi(u,t))-\beta and R(v)>βR(v)>\beta for any vBα(𝒫)v\in B_{\alpha}(\mathcal{P}^{\mathcal{F}}). It follows that

s|s=tR(ψ(u,s))=\displaystyle\tfrac{\partial}{\partial s}\big{|}_{s=t}R(\psi(u,s))= R(Jψ(u,t))\displaystyle R(-\nabla J\psi(u,t))
=R((L+G)ψ(u,t))R(ψ(u,t)).\displaystyle=R((L+G)\psi(u,t))-R(\psi(u,t)).

Since by hypothesis ψ(u,t)Bα(𝒫)B¯α(𝒫)\psi(u,t)\in\partial B_{\alpha}(\mathcal{P}^{\mathcal{F}})\subset\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}), then by Step 1 above (L+G)(ψ(u,t))Bα(𝒫)(L+G)(\psi(u,t))\in B_{\alpha}(\mathcal{P}^{\mathcal{F}}). Thus we obtain that

R((L+G)ψ(u,t))R(ψ(u,t))=R((L+G)ψ(u,t))β>0,R((L+G)\psi(u,t))-R(\psi(u,t))=R((L+G)\psi(u,t))-\beta>0,

and consequently s|s=tR(ψ(u,s))>0\tfrac{\partial}{\partial s}\big{|}_{s=t}R(\psi(u,s))>0. Hence, there exists ε>0\varepsilon>0 such that R(ψ(u,s))<βR(\psi(u,s))<\beta for s(tε,t)s\in(t-\varepsilon,t). Thus, ψ(u,s)B¯α(𝒫)\psi(u,s)\not\in\overline{B}_{\alpha}(\mathcal{P}^{\mathcal{F}}) for s(tε,t)s\in(t-\varepsilon,t). This contradicts Step 1.

Letting α0:=min{α1,α2}\alpha_{0}:=\min\{\alpha_{1},\alpha_{2}\} we conclude the proof of the lemma. ∎

References

  • [1] R. A. Adams, Sobolev Spaces, vol. 65 of Pure and applied mathematics series, Academic Press, 1975.
  • [2] M. M. Alexandrino and R. G. Bettiol, Lie groups and geometric aspects of isometric actions, Springer, Cham, 2015.
  • [3] M. M. Alexandrino and L. F. Cavenaghi, Singular Riemannian Foliations and the prescribing scalar curvature problem, arXiv:2111.13257 [math.DG], (2021).
  • [4] B. Ammann and E. Humbert, The second Yamabe invariant, J. Funct. Anal., 235 (2006), pp. 377–412.
  • [5] T. Aubin, Équations différentielles non linéaires et problème de Yamabe concernant la courbure scalaire, J. Math. Pures Appl. (9), 55 (1976), pp. 269–296.
  • [6]  , Some nonlinear problems in Riemannian geometry, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 1998.
  • [7] T. Bartsch, M. Schneider, and T. Weth, Multiple solutions of a critical polyharmonic equation, J. Reine Angew. Math., 571 (2004), pp. 131–143.
  • [8] S. Brendle, Blow-up phenomena for the Yamabe equation, J. Amer. Math. Soc., 21 (2008), pp. 951–979.
  • [9] S. Brendle and F. C. Marques, Blow-up phenomena for the Yamabe equation. II, J. Differential Geom., 81 (2009), pp. 225–250.
  • [10] A. Castro, J. Cossio, and J. M. Neuberger, A sign-changing solution for a superlinear dirichlet problem, Rocky Mountain J. Math., 27 (1997), pp. 1041–1053.
  • [11] L. F. Cavenaghi, J. M. do Ó, and L. D. Sperança, The symmetric Kazdan-Warner problem and applications, arXiv:2106.14709 [math.DG], (2021).
  • [12] M. Clapp and J. C. Fernández, Multiplicity of nodal solutions to the yamabe problem, Calc. Var. Partial Differential Equations, 56 (2017), pp. Paper No. 145, 22.
  • [13] M. Clapp and F. Pacella, Multiple solutions to the pure critical exponent problem in domains with a hole of arbitrary size, Math. Z., 259 (2008), pp. 575–589.
  • [14] M. Clapp and A. Pistoia, Yamabe systems and optimal partitions on manifolds with symmetries, Electronic Research Archive, 29 (2021), pp. 4327–433.
  • [15] M. Clapp, A. Saldaña, and A. Szulkin, Phase separation, optimal partitions, and nodal solutions to the Yamabe equation on the sphere, Int. Math. Res. Not. IMRN, (2021), pp. 3633–3652.
  • [16] M. Clapp and S. Tiwari, Multiple solutions to a pure supercritical problem for the p-laplacian, Calc. Var. Partial Differential Equations, 55 (2016), pp. Paper No. 7, 23.
  • [17] M. Clapp and T. Weth, Multiple solutions for the brezis-nirenberg problem, Adv. Differential Equations, 10 (2005), pp. 463–480.
  • [18] D. Corro, Manifolds with aspherical singular Riemannian foliations, PhD thesis, Karlsruhe Institute of Technology, 2018. https://publikationen.bibliothek.kit.edu/1000085363.
  • [19] D. Corro and A. Moreno, Core reduction for singular riemannian foliations in positive curvature, arXiv:2011.05303 [math.DG], (2020).
  • [20] K. Deimling, Ordinary differential equations in Banach spaces, vol. 596 of Lecture Notes in Mathematics, Springer-Verlag, Berlin-New York, 1977.
  • [21] M. del Pino, M. Musso, F. Pacard, and A. Pistoia, Large energy entire solutions for the Yamabe equation, J. Differential Equations, 251 (2011), pp. 2568–2597.
  • [22] W. Y. Ding, On a conformally invariant elliptic equation on 𝐑n{\bf R}^{n}, Comm. Math. Phys., 107 (1986), pp. 331–335.
  • [23] F. T. Farrell and X. Wu, Riemannian foliation with exotic tori as leaves, Bull. Lond. Math. Soc., 51 (2019), pp. 745–750.
  • [24] J. C. Fernández, O. Palmas, and J. Petean, Supercritical elliptic problems on the round sphere and nodal solutions to the Yamabe problem in projective spaces, Discrete Contin. Dyn. Syst., 40 (2020), pp. 2495–2514.
  • [25] J. C. Fernández and J. Petean, Low energy nodal solutions to the Yamabe equation, J. Differential Equations, 268 (2020), pp. 6576–6597.
  • [26] D. Ferus, H. Karcher, and H. F. Münzner, Cliffordalgebren und neue isoparametrische Hyperflächen, Math. Z., 177 (1981), pp. 479–502.
  • [27] F. Galaz-García, M. Kell, A. Mondino, and G. Sosa, On quotients of spaces with Ricci curvature bounded below, J. Funct. Anal., 275 (2018), pp. 1368–1446.
  • [28] F. Galaz-Garcia and M. Radeschi, Singular Riemannian foliations and applications to positive and non-negative curvature, J. Topol., 8 (2015), pp. 603–620.
  • [29] J. Ge and M. Radeschi, Differentiable classification of 4-manifolds with singular Riemannian foliations, Math. Ann., 363 (2015), pp. 525–548.
  • [30] D. Gromoll and G. Walschap, Metric foliations and curvature, vol. 268 of Progress in Mathematics, Birkhäuser Verlag, Basel, 2009.
  • [31] E. Hebey and M. Vaugon, Sobolev spaces in the presence of symmetries, J. Math. Pures Appl., 76 (1997), pp. 859–881.
  • [32] J. Heinonen, P. Koskela, N. Shanmugalingam, and J. T. Tyson, Sobolev spaces on metric measure spaces, vol. 27 of New Mathematical Monographs, Cambridge University Press, Cambridge, 2015. An approach based on upper gradients.
  • [33] G. Henry, Isoparametric functions and nodal solutions of the yamabe equation, Ann. Glob. Anal. Geom., 56 (2019), pp. 203––219.
  • [34] W. C. Hsiang and J. L. Shaneson, Fake tori, the annulus conjecture, and the conjectures of Kirby, Proc. Nat. Acad. Sci. U.S.A., 62 (1969), pp. 687–691.
  • [35] M. A. Khuri, F. C. Marques, and R. M. Schoen, A compactness theorem for the Yamabe problem, J. Differential Geom., 81 (2009), pp. 143–196.
  • [36] J. M. Lee and T. H. Parker, The Yamabe problem, Bull. Amer. Math. Soc. (N.S.), 17 (1987), pp. 37–91.
  • [37] A. Lytchak and G. Thorbergsson, Curvature explosion in quotients and applications, J. Differential Geom., 85 (2010), pp. 117–139.
  • [38] R. E. Megginson, An introduction to Banach space theory, vol. 183 of Graduate texts in mathematics, Springer-Verlag New York, 1998.
  • [39] R. A. E. Mendes and M. Radeschi, A slice theorem for singular Riemannian foliations, with applications, Trans. Amer. Math. Soc., 371 (2019), pp. 4931–4949.
  • [40] I. Moerdijk and J. Mrčun, Introduction to foliations and Lie groupoids, vol. 91 of Cambridge Studies in Advanced Mathematics, Cambridge University Press, Cambridge, 2003.
  • [41] P. Molino, Riemannian foliations, vol. 73 of Progress in Mathematics, Birkhäuser Boston, Inc., Boston, MA, 1988.
  • [42] R. S. Palais, The principle of symmetric criticality, Comm. Math. Phys., 69 (1979), pp. 19–30.
  • [43] M. Radeschi, Clifford algebras and new singular Riemannian foliations in spheres, Geom. Funct. Anal., 24 (2014), pp. 1660–1682.
  • [44]  , Lecture notes on singular Riemannian foliations, 2017. URL: https://static1.squarespace.com/static/5994498937c5815907f7eb12/t/5998477717bffc656afd46e0/1503151996268/SRF+Lecture+Notes.pdf. Last visited on 26 April 2021.
  • [45] R. Schoen, Conformal deformation of a Riemannian metric to constant scalar curvature, J. Differential Geom., 20 (1984), pp. 479–495.
  • [46] D. Sullivan, A counterexample to the periodic orbit conjecture, Inst. Hautes Études Sci. Publ. Math., (1976), pp. 5–14.
  • [47] N. S. Trudinger, Remarks concerning the conformal deformation of Riemannian structures on compact manifolds, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3), 22 (1968), pp. 265–274.
  • [48] J. Vétois, Multiple solutions for nonlinear elliptic equations on compact riemannian manifolds, Internat. J. Math., 18 (2007), pp. 1071–1111.
  • [49] G. Wang and Y. Zhang, A conformal integral invariant on Riemannian foliations, Proc. Amer. Math. Soc., 141 (2013), pp. 1405–1414.
  • [50] B. Wilking, A duality theorem for Riemannian foliations in nonnegative sectional curvature, Geom. Funct. Anal., 17 (2007), pp. 1297–1320.
  • [51] M. Willem, Minimax theorems, vol. 24 of Progress in Nonlinear Differential Equations and their Applications, Birkhäuser, 1996.
  • [52] H. Yamabe, On a deformation of Riemannian structures on compact manifolds, Osaka Math. J., 12 (1960), pp. 21–37.