Yamabe problem in the presence of singular Riemannian Foliations
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 foliation2020 Mathematics Subject Classification:
Primary 58J05, 53C12; Secondary 35J61, 35B06, 35J20, 35R01, 53C21, 57R301. Introduction
The Yamabe problem, stated by Yamabe in [52], asks if for a given closed Riemannian manifold there exists a conformal Riemannian metric , for some smooth positive function , such that the scalar curvature of 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 , 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 which are sign changing and satisfy the Yamabe equation. For , Ding [22] established the existence of infinitely many sign-changing solutions for the round -sphere, using the fact that there is a (linear) action by isometries. Moreover, Ammann and Humbert proved in [4] that in dimension at least 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 be a closed Riemannian manifold of dimension and consider the following Yamabe type problem:
(Y) |
where is the Laplace-Beltrami operator, , with . We will assume that the operator is coercive in the Sobolev space , meaning, that there exists such that
(1) |
for every .
Let be a singular Riemannian foliation on with closed leaves and nontrivial, meaning that is different to the foliation that consists of only one leaf, , and to the foliation (see Section 2.1 for definitions). We will further assume that all the leaves of have dimension greater than or equal to one. For that, we define the number
(2) |
We say that a function is -invariant if is constant on each leaf of .
Theorem A.
þ Let be an -dimensional Riemannian manifold, , together with a nontrivial closed singular Riemannian foliation such that . Assume that is coercive in , that and are -invariant functions, with and . Then (Y) admits an infinite number of -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.
Remark 1.2.
þ Observe that when is a positive function then the operator is coercive, since we can take
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 be a nontrivial singular Riemannian foliation with closed leaves on a compact Riemannian manifold . Assume that , is constant and that is -invariant. Then there exists an infinite number of sign changing -invariant solutions to (Y). In particular, this is true for the Yamabe problem:
(3) |
provided that the scalar curvature of , is -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 , 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 is constant, then for any nontrivial singular Riemannian foliation on the round sphere with , there exist an infinite number of sign changing -invariant solutions to the Yamabe problem:
(4) |
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 -invariant solution with minimal energy among any other -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 equals the scalar curvature of it is not clear if the energy minimizing solutions from þB attain the Yamabe constant.
Recently in [14, 15], given a compact Lie group acting by isometries with positive dimensional orbits, the authors give -invariant solutions to the Yamabe problem which have minimal energy among any other -invariant sign-changing solutions with a fixed number of nodal domains, by showing the existence of regular optimal -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 -invariant function for a singular Riemannian foliation 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 and on a fixed Riemannian manifold , it is difficult to know if there is a relation between the -invariant and the -invariant sign-changing solutions to (Y) given by our method. We say that if for any leaf there exists a leaf such that . Thus we have the following question:
Question.
Consider a fixed Riemannian manifold and two singular Riemannian foliations and with respect to such that . Does there exist a solution to (Y) which is -invariant but not -invariant?
As a particular example, note that for a fixed Riemannian manifold and a given group action by a compact Lie group via isometries on , and a fixed closed subgroup , then the -orbits are contained in the -orbits. That is we have . 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 -invariant Sobolev functions is isometric to the Sobolev space , where is the pushfoward measure of the Riemannian volume via quotient the map . Thus from our proof it is clear that if is isometric to then we have a negative answer to the question posted above. That is, the -invariant sign-changing solutions given by þA are the same as the -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 is an infinite dimensional subspace of , 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, . 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 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 (þF), 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, will be a closed (compact and without boundary) Riemannian manifold of dimension . By a singular Riemannian foliation of we mean a decomposition of into connected injectively immersed submanifolds, called leaves, which may have different dimensions, satisfying:
-
(i)
is a transnormal system, i.e., every geodesic orthogonal to one leaf remains orthogonal to all the leaves that it intersects.
-
(ii)
is smooth, that is, for every leaf and ,
where is the module of smooth vector fields on which are everywhere tangent to the leaves of .
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 .
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 -dimensional manifold with a smooth foliation by -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 is called a regular Riemannian foliation. The dimension of , denoted by , is the dimension of the regular leaves and the codimension of is equal to .
2.2. Variational Setting
Let be a closed Riemannian manifold of dimension , , and let be a singular Riemannian foliation on . In what follows, will denote the Riemannian metric in , while will denote the gradient with respect to of a smooth function . The Sobolev space is the closure of under the norm, , induced by the interior product
(5) |
By density, this bilinear form extends to a continuous symmetric bilinear form on Even if, formally, 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 .
Fix two functions with and suppose that the operator is coercive (see (1)). Since is coercive, the bilinear product
(6) |
is definite positive. Thus, it is an inner product and consequently induces a norm in , which furthermore is equivalent to the norm given that is compact, is continuous and is coercive.
Since is positive and continuous and is compact, we also have a norm in the space,
(7) |
equivalent to the standard norm of . Indeed, denoting by the maximum of , and by its minimum, we have that , and thus
The energy functional given by
is a well defined functional for all , and its derivative is given by
(8) |
Moreover, is a critical point of if and only if is a solution to (Y). The nontrivial critical points of lie in the Nehari manifold
(9) |
which is a closed Hilbert submanifold of of codimension one and radially diffeomorphic to , the unit sphere in with respect to the norm . Concretely, for each , there exists a unique such that , and it is given explicitly by
Hence, we can define a projection given by
(10) |
which induces the radial diffeomorphism with inverse . In addition, if , then
(11) |
(see [51, Lemma 4.1]).
For any singular Riemannian foliation, define the space
and let be the closure of under the norm . Hence is a closed subspace of and if and only if and is -a.e. constant on the leaves of .
For with defined as in (2) we show in þD that is an infinite dimensional Hilbert space endowed with the restriction of the interior product . The critical points of that are contained in will be called -invariant critical points of . Hence, the nontrivial -invariant critical points of lie in the restricted Nehari manifold
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) |
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 restricted to the space (þG), and a variational principle (þF), which we state below.
First we notice that is an infinite dimensional space, see Section 3 for the proof. Recall that .
Theorem D.
þ Let be an -dimensional closed Riemannian manifold together with a nontrivial singular Riemannian foliation with . Then, for any , there exist nontrivial nonnegative functions with pairwise disjoint supports. In particular, it follows that is infinite dimensional.
Given , we say that a sequence in is a -sequence for if and in , where the last space is the dual of . We say that satisfies the -condition in if every -sequence for has a strongly convergent subsequence in . 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 and to be non foliated, the functional satisfies the -condition for every .
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 be a nontrivial finite dimensional subspace of . If satisfies the condition in for every , then has at least one positive critical point and pairs of sign-changing critical points in such that and for .
The Principle of Symmetric Criticality needed to prove our main theorem is the following.
Theorem G (Principle of Symmetric Criticality for ).
þ Under the hypotheses of Theorem A, if is a critical point of the functional restricted to , then is a critical point of in the space .
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.
Proof of Theorem A.
In order to find nontrivial foliated critical points of , by Theorem G, it suffices to consider critical points of the restriction of to the space . To do so, we proceed as follows. By Theorem D, since we are assuming that the singular Riemannian foliation is nontrivial, for any given , we may choose nontrivial functions with disjoint supports. Let be the linear subspace of spanned by . Since and have disjoint supports for the set is orthogonal in . Hence,
Now, as , Theorem E yields that satisfies in for every . Therefore, we can apply Theorem F and get at least one positive and sign-changing -invariant critical points for the restriction of to . Theorem G yields that these are also critical points of in and, hence nontrivial -invariant solutions to problem (Y). Given that 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 -invariant partitions of unity and the Slice Theorem. Furthermore, we prove that 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 be a singular Riemannian foliation with closed leaves on a closed Riemannian manifold . Given , denote by the leaf containing , by the dimension of and by the natural projection. The following lemma describes the foliation in a small neighborhood of .
Lemma 3.1 (Proposition 2.17 in [44], Figure 1).
þ For any , there exists a coordinate system such that
-
(1)
, , where and are open and bounded subsets with smooth boundary;
-
(2)
for any , .

Consider fixed. Take such that is a diffeomorphism. Denote by the closed ball of radius in the normal space to the leaf at . Consider . For denote by the connected component of containing . Since is smaller than the injectivity radius at , then if and only if for a unique . For such we set . The following theorem states that the partition of by is a singular Riemannian foliation.
Theorem 3.2 (Infinitesimal Foliation, Proposition 6.5 in [41]).
þ Let be a singular Riemannian foliation on the compact manifold and fix . Take smaller than the injectivity radius of at . Then equipped with the Euclidean metric is a singular Riemannian foliation.
Moreover, does not depend on the radius chosen.
Lemma 3.3 ([44]).
þ Let and be such that and are smaller than the injectivity radius of at . Then the map given by is a foliated diffeomorphism between and
The previous result implies that for small enough is independent of , and thus we set . We call the infinitesimal foliation at .
Observe that the image under of different leaves of may be contained in the same leaf of . 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 the leaf of that contains , the following statements hold:
-
a)
Given a piece-wise smooth curve starting at there exists a map which is a foliated isometry between and .
-
b)
The set
is a group under the composition operation.
-
c)
Consider , and a piece-wise curve starting at and ending at . Then
-
d)
Given two piece-wise smooth curves and starting at and homotopic relative to its end points, the composition of their corresponding foliated isometries, , is homotopic to the identity map.
The group given in þ3.4 is called the holonomy group of the leaf at . For any other due to and the groups and are conjugates, and we denote their conjugacy class as . The regular leaves of for whom are called principal leaves. We denote by the set of points contained in the principal leaves. We remark that this is an open and dense subset of .
In order to prove that is infinite dimensional, we define the leaf space of to be the quotient space induced by the partition equipped with the quotient topology and denote it by . We denote the quotient map as . Given a subset , we denote its image by .
Proof of þD.
Let be a singular Riemannian foliation with closed leaves on the complete manifold . Then the leaf space is a complete length metric space. Thus for any open cover of we have a subordinate partition of unit such that . We can refine the cover so that for any , at least open neighborhoods of the cover are disjoint. Taking and setting we get the desired continuous foliated functions with disjoint support.
This approach can be improved to get smooth foliated functions. We can restrict to and , so that . We endow with the induced metric by the inclusion into . Then is a regular Riemannian foliation. Moreover the leaf space is a manifold. By Theorem [30] there is a Riemannian metric on such that the map is actually a Riemannian submersion.
Then by restricting an open cover of to , we get a smooth partition of unity subordinated to the cover . Setting gives us a smooth partition of unity subordinated to the open cover of ; observe that by construction these functions are constant along the leaves. We can refine this open cover so that for any given , there are at least functions which have disjoint support, contained in . Setting , we extend these functions trivially on to get the desired smooth functions , 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 be a singular Riemannian foliation with closed leaves. Given a leaf and , denote by the tubular neighborhood of radius of . Then there exist small enough and a -principal bundle , such that is foliated diffeomorphic to
Here denotes the quotient with respect to the product action of on , i.e. .
Given the leaf of consists of all the classes such that for some .
Lemma 3.6.
þ Let be a closed foliation on a compact manifold. Then for any there exists an open cover of such that each is a tubular neighborhood of radius at most and as in þ3.5 such that there exists a finite subcover and a smooth partition of unity adapted to the subcover which is -invariant; i.e. for each if then .
Proof.
For each consider such that þ3.5 holds for the tubular neighborhood of of radius . If then we replace by , and note that þ3.5 still holds for this value. The collection is an open cover of and since is compact there exist , , such that .
For each , let be the foliated diffeomorphism given by þ3.5. We will define a -foliated function and then define functions
so that defined as
will be the desired partition of unity. Observe that is -invariant since it is a composition of a foliated diffeomorphism with an -invariant function, and thus is an -invariant function.
For each , to define , first consider a smooth nonnegative decreasing function such that for , for . Let be given by . Observe that over the spheres in centered at , is constant. Since the leaves of are contained in such spheres we conclude that is -foliated.
Set , , and let be the total space of the principal -bundle from þ3.5. Define
Now note that for any we have . Indeed, this follows from the fact that as can be seen below,
Thus we have a well defined map
and we will check using local trivializations that it is also smooth.
Consider the projection map of the principal -bundle, and denote by the projection of the associated bundle. Fix , and consider a sufficiently small open neighborhood of , such that there exists a trivialization of . Observe that the map given by
where denotes the identity element, is a trivialization of the associated bundle . Then we have that is given over a local trivialization as
Therefore can be written as the composition of smooth maps and so it is also smooth.
We check now that is -foliated. Consider , and take . Then there exists some such that . Then
This concludes the construction of 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 we can define a singular Riemannian foliation where the foliation consists of the set of preimages under of the image of . We give more details below.
Recall that a surjective differentiable map between smooth manifolds is a submersion if at any point , the differential map is a surjective linear map. This implies that the dimension of is greater than or equal to the dimension of . From now on assume that the dimension of is strictly greater than the dimension of . Assume that has a Riemannian metric , and for any denote by the subspace of tangent at to the fiber of through , and let to be the subbunddle over with fiber equal to . We say that is a Riemannian submersion if for any the Lie derivative of in the direction of satisfies , that is, the metric is invariant in the directions tangent to the fibers. Setting to be the -orthogonal complement of in , this implies that for any the inner product given as
is a Riemannian metric on .
Homogeneous foliations: Another large family of examples stem from a compact Lie group acting by isometries on a given Riemannian manifold . Then, for the partition consisting of the set of orbits of the action of on , , we have that 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 foliations on round spheres 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 of dimension , equipped with a positive definite inner-product the Clifford algebra is the quotient of the tensor algebra by the ideal generated by , where is the unit element in . The vector space embeds naturally into . A representation of a Clifford algebra is an algebra homomorphism . A Clifford system is the restriction of to , and we denote by the image . Given a Clifford system we can find an inner product on , such that for every , the matrix is a symmetric matrix. We endow the space of all symmetric matrices with the inner product . With these choices of inner products on and the map is an isometry onto its image.
Remark 3.7.
Given a Clifford system , the dimension is even (see [43]).
Consider the map that takes to the unique element which for all satisfies
The image of this map is contained in the unit disk of , denoted by (see [43, Proposition 2.4]).
Proposition 3.8 (Proposition 2.6 in [43]).
þ The set is a singular Riemannian foliation on the unit sphere with the round metric.
We refer to the foliation 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 , for the map defined as , the preimages of the composition induce a singular Riemannian foliation of codimension on the round sphere . These are the singular Riemannian foliations described in [26].
In general, given two singular Riemannian foliations and on a given Riemannian manifold we say that if for any leaf , there exists a leaf , such that . With this definition observe that .
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 a closed Riemannian manifold and consider the closed disk with the Euclidean metric, equipped with a singular Riemannian foliation given by the cone of a non-homogeneous RFKM-foliation on the sphere . Now consider the foliated product . Next we glue two copies of along the boundary via the identity map to obtain a new smooth foliated closed manifold whose orbit space is , the double of . For any of the two singular leaves given by , in each copy of , the infinitesimal foliation corresponds to the RFKM-foliation . Thus 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 any closed Riemmannian manifold, and for a smooth manifold which is homeomorphic but not diffeomorphic to the -torus, i.e. an exotic torus. These manifolds exist due to [34]. Now we consider with any product metric. This produces a Riemannian foliation 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 -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 be a singular Riemannian foliation with closed leaves. As before, denote by the closure of the smooth foliated real valued functions under and by the orthogonal complement of with respect to the -inner product. The aim of this section is to prove Theorem G. The proof relies on establishing that for any and , and using the explicit formula of the derivative of the energy functional .
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 , given a a function we define , the Lipschitz constant of , as
Denote by
We now let be the distance function on induced by the Riemannian metric and consider the subset
Denote by the closure of in , and by the orthogonal complement of with respect to the -inner product.
Recall that the leaf space of is the space equipped with the quotient topology induced by the projection . For complete and with closed leaves, this map also endows with a metric and a measure , which is the pushforward of the Riemannian volume form of under the quotient map . Moreover is a complete separable metric space with a non-trivial locally finite Borel measure ([37, p. 119]); that is is a metric measure space [32, p. 65]. So that is dense in (see for example [32, Theorem 4.2.4]).
Given a function which is constant along the leaves, we define a function as . Moreover, given a function , we define a function which is constant along the leaves of by . For a function constant along the leaves we have .
Using the fact that is the pushforward measure, we have well defined functions and . Indeed, take . Then it holds that
On the other hand, for we have that
Lemma 4.1.
þ For a singular Riemannian foliation with closed leaves, it holds true that .
Proof.
Given , since is dense in , there exists a sequence of Lipschitz bounded functions converging to in . Applying a change of variable we get that in . To prove this, fix . Then there exists , such that for we have:
Thus, in . Finally note that by definition is the composition of Lipschitz functions, being a -Lipschitz function. Hence, . Thus, we conclude that as desired. ∎
Lemma 4.2.
þ For a singular Riemannian foliation with closed leaves, it holds true that .
Proof.
We first show that consists only of the equivalent class of the zero function. Take . Since there exists a sequence of Lipschitz functions which are constant along the leaves, , converging to with respect to the -norm. Then is a Cauchy sequence in ; fix , there exists such that for all it holds:
Since is complete there is a limit of . As in the proof of the previous lemma, converges to in . Thus by uniqueness of the limit, is equal to -a.e., i.e. is constant along the leaves of up to zero measure. This implies that the class of in is an element in and since by hypothesis, we conclude that corresponds to the class of the zero function in (recall that ). Thus .
Finally, since consists only of the equivalence class of the zero function, by the first paragraph we conclude that . ∎
Lemma 4.3.
Let be a singular Riemannian foliation with closed leaves. If and then .
We are ready to prove the Principle of Symmetric Criticality, that is, that critical points of restricted to are critical points of .
Proof of Theorem G.
Write any as with and . Since is a critical point of restricted to the space , we have that . Hence,
where in the last part we used the expression (8) for the derivative of . By Lemma 4.3 both terms above equal zero. To see this is true for the first term, by hypothesis we have and by Lemma 4.3 we have that . This and being an -invariant function implies that . For the second term, since is -invariant we have and, we know that . 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 satisfies the -condition for every .
For any , consider the Sobolev space , which is the closure of with respect to the norm
As before, let be the closure of under the norm above, so that it is a Banach space. Observe that is just .
Recall that is the smallest dimension of the leaves of . We now state the Sobolev embedding theorem needed for the proof of Theorem E.
Theorem H.
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 , to get a chart of the form . As is compact, we can cover it by open sets of the form . Then one passes the information of any function in to functions defined on sets , where as in þ3.1, reducing the dimension of to the dimension of , which is . This allows us to apply the subcritical Sobolev embedding theorem for the sets . To prove the compactness of the map, we also apply þ3.5 and þ3.6 in order to have a covering of 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 . Fix , and let be a chart around given by þ3.1. For any function , let . By the properties of the coordinate chart, for any and any , we have that , so we have a well defined function given by . Denote by the gradient of , for then we denote by the gradient of , and by the gradient of .
Lemma 5.1.
þ Consider a fixed , and as above. For any open subsets and and any there exists a constant , where , such that
holds over .
Proof.
Since we have . Hence, and thus .
Writing the metric in coordinates as , at each the matrix is definite positive and symmetric, hence it is diagonalizable and with positive eigenvalues . Then, at each we have that
Since the functions are continuous, and then
If is such that , substituting everything in the previous inequality, and using that we conclude. ∎
Lemma 5.2.
þ For , and any as in the previous lemma, there exists a constant , such that, for any and any function
and
In particular, there exists such that for any ,
(13) |
Proof.
For the first inequality, the function is continuous and attains its maximum and its minimum, say and .
First we have,
Second we have
We are now ready to prove the Sobolev embedding theorem.
Proof of Theorem H.
For any , take a chart around , , as in þ3.1. Take a further open subset around as in Lemma 5.2. As is compact, there exist a finite number of points such that the charts cover . Set and note that for some open sets.
To prove the continuity of the inclusion map, take and consider the function as defined in this section, i.e. , where is an arbitrary element in . Then Lemma 5.2 implies that . Using (13) twice together with the Sobolev embedding theorems for [1, Theorem 5.4], we have the existence of a constant depending on such that
if either and , or and .
We now prove that for every index , either or and holds. If (C1) holds, then by the definition of we get for every . If (C2) holds, fix . Then either or . Assume that . Then the fact and the definition of yields the inequality .
Therefore, by taking for all ,
Hence the embedding is continuous.
Now we prove the compactness of the embedding. Without loss of generality assume that each open set is a trivialization of the disk bundle given by þ3.5 centered at some . We can identify each element of the finite open cover of , with . By þ3.6 there exists a smooth foliated partition of unity subordinated to . For each index , due to the identification we may assume that is defined over .
Let be a bounded sequence in and define , for arbitrary . Observe is well defined and is compactly supported in and that the sequence is bounded in . Then, if either or if and , we have that for every . Hence, the embedding is compact by the Rellich-Kondrachov theorem [1, Theorem 6.5] and for each , there is a subsequence of , which we denote in the same way, which is a Cauchy sequence in . Inequalities (13) and the fact that are bounded yield that has a Cauchy subsequence in and, hence, this subsequence converges in this space. ∎
We now show that satisfies the -condition for every .
Proof of Theorem E.
The proof is standard and consists in showing that any -sequence for , , is bounded in , so there exists such that up to a subsequence, converges weakly to in . The next step is to apply Theorem H to show that , which implies that strongly in . We provide more details below.
Let be a -sequence for in , i.e. and in . We first uniformly bound in terms of and .
From the definition of we get
Therefore, we have
On the other hand, since in , there exists a positive constant such that for large the following holds
Since , there exists a positive constant such that for large the following holds
Therefore for large we obtain that
This, and the fact that the norm is equivalent to the standard norm of , imply that is bounded in .
Since is bounded in , there exists such that up to a subsequence, also denoted by , it converges weakly to in . As is weakly closed in , is -invariant. To show that strongly in , it suffices to show that .
As weakly in , we have that
It follows that
and so, as 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 is bounded, then is also bounded, say by a positive constant , and thus we get,
(14) |
For the term to the right, we proceed as follows. Observe that by þH when , then for any the map is compact. In the case when due to the fact that we get that , and thus for the map is also compact by þH.
Hence, up to a subsequence, we have that converges weakly to in . Using this, Hölder’s inequality for , Sobolev inequality for the embedding , the fact that is bounded in and , we get that
(15) |
where denotes some positive constant.
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, denotes the usual norm. Recall that are foliated functions with , and that as defined in (7) is a norm in equivalent to the standard norm of .
For any , where was defined in (1), we have a well defined interior product
which induces a norm, , that is equivalent to the standard norm of .
Recall that we are studying the functional , given by
restricted to the space , and where is the norm induced by the inner product defined in (6). For , this functional is of class and its gradient with respect to the interior product defines a function . We next show that actually . To do so, we decompose the gradient of at as follows.
First, note that with respect to , , is the vector in which satisfies:
for any (see 8). Then, we need to apply the following result that we prove in Section 6.1.
Lemma 6.1.
þ Let and a nonnegative constant. Then the non-homogeneous problem
admits a unique solution in . Moreover, if is -invariant, then the solution lies in .
Thus, for each , by applying þ6.1 to equal to and , which are elements of , there exist unique solutions and to the non-homogeneous linear problems
Note that these solutions are uniquely determined by the relations
(16) |
for every . It follows that
If , then also and are -invariant. Thus, Lemma 6.1 yields that the functions and are -invariant, and so we conclude that for , the function is -invariant as we claimed.
Recall that the restricted Nehari manifold was defined to be
Observe that if , then also , where . The nontrivial -invariant sign-changing critical points of , and thus, sign changing solutions to (Y) in virtue of Theorem G, must belong to the set
This set is nonempty. Indeed, Theorem D gives the existence of at least two foliated smooth functions with disjoint supports. Then recalling the definition of the projection onto the Nehari manifold given in (10), we can define the function , which is an element in , given that and .
Let be the convex cone of nonnegative functions. Then the set of functions in which do not change sign is given as .
As the -gradient of , , is of class , it is locally Lipschitz and thus for each , the Cauchy problem:
has a unique solution defined for all , where is the maximal existence time of the solution. With this at hand, the negative gradient flow of , is simply the map where and is as above. A subset of is said to be strictly positively invariant under if
For any set , we define the set
and the entrance time , given by
If is strictly positively invariant under , then is open in and is continuous.
In what follows, for any subset of and any , will denote the set
To find sign-changing critical points of we use the relative genus between symmetric subsets of . But before that we need the following preliminary result that we prove in Subsection 6.1.
Lemma 6.2.
þ þ There exists such that for every
-
(a)
, , and
-
(b)
and are strictly positively invariant under the flow of the negative gradient of with respect to .
Fix as in the previous lemma. For , set
where .
The next result says that, under suitable conditions, is a neighborhood retract.
Lemma 6.3.
þ If has no sign-changing critical point with , then is strictly positively invariant under and the map
is odd and continuous, and satisfies for every .
Proof.
Note that by definition is odd. To show that is strictly positively invariant under , by þ6.2, it suffices to consider . By definition of the flow, given any , we have that
(17) |
Thus we conclude that for every . So, if , it follows that for every . Next, suppose that . By hypothesis, cannot be a sign changing critical point for in , hence we have two possible cases: either is a critical point that does not change sign, or .
The first case reduces to the first paragraph of the proof, that is we have , and hence there is nothing to prove. Now, suppose that the second case holds true. Then, there exist real numbers such that for every . Thus, for small enough, and (17) yields that . Thus, is strictly positively invariant under .
It is now easy to check that has the desired properties. ∎
Remark 6.4.
þ Observe that is always strictly positively invariant under the flow , for there are no nontrivial critical points of satisfying . In fact, if is a critical point of satisfying , then ; this implies that and .
Remark 6.5.
þ Notice that a critical point of changes sign if and only if it lies in the complement of in . Indeed, if lies in , then and it must change sign. Conversely, if is a sign changing critical point of , as we pointed before, it belongs to and þ6.2 says that this set has empty intersection with ; moreover, as it is a nontrivial critical point, we have that and . Thus, .
A subset of will be called symmetric if for every .
Definition 6.6.
Let and be symmetric subsets of . The genus of relative to , denoted by , is the smallest number such that can be covered by open symmetric subsets of with the following two properties:
-
(i)
and there exists an odd continuous map such that for .
-
(ii)
There exist odd continuous maps for every .
If no such cover exists, we define .
As in [13, Section 3] in order to obtain a variational principle for sign changing solutions, we need a refined version of the condition. Given , we say that satisfies the condition relative to in , if every sequence in such that
has a strongly convergent subsequence in . When , we recover the condition given in Section 2.3. Also notice that if satisfies the condition in , then it satisfies the condition relative to any subset of
Lemma 6.7.
þ Fix with given as in Lemma 6.2. For , define
Assume that satisfies relative to in . Then, the following statements hold true:
-
(a)
has a sign-changing critical point with .
-
(b)
If , then has infinitely many sign-changing critical points with .
Consequently, for , if satisfies relative to in for every , then has at least distinct pairs of sign-changing critical points: , , in with for all .
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 by contradiction, using the fact that is strictly positively invariant under the flow (see þ6.4). More precisely, we assume that there does not exist a sign-changing critical point with . We claim that the condition relative to implies the existence of and such that
(18) |
To see this, suppose in order to get a contradiction, the existence of a sequence in such that and . Then is a sequence, and since satisfies the condition relative to there exists such that, up to a subsequence which we also denote by , it holds that strongly in . As and are continuous, it follows that and . By hypothesis, cannot change sign. Hence is a critical point of lying in , which is open. Therefore, there exists such that for every , which is a contradiction and the claim follows.
As we pointed out in þ6.5, every sign changing critical point lies in , and (18) implies that there are no sign changing critical points of in . Therefore is strictly positively invariant under for every by þ6.3. As , identity (18) yields that flows under to and . In this way, þ6.3 implies that is odd, continuous and for every , given that . As is a symmetric neighborhood retract by þ6.4 and þ6.3, from the monotonicity of the genus [13, Lemma 3.4] it follows that which is a contradiction.
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 . Let such that . Since we get
Thus, is bounded in . As and , as in the proof of Theorem E, there exists such that converges weakly to in and strongly in . Then
implying that . Therefore, there exists such that , where is the projection onto . As , identity (11) yields that . Therefore, using basic properties of weak convergence and that in , we obtain that
Again, since strongly in , we conclude from this inequalities that exists and is equal to . Hence strongly in , and since is closed in then it follows that , and . As is a natural restriction for the functional (see [51, Chapter 4]), is a nontrivial -invariant critical point for in attaining .
Now we see that does not change sign. Suppose, in order to get a contradiction, that this is not true. Then, and . As and as is a critical point for , we have that
concluding that . Hence
which is a contradiction since . Thus does not change sign. If , we can take , since it is also a critical point for and it is positive.
We proceed to prove the existence of the sign-changing critical points: Let . By Lemma 6.7, we only need to show that
Let be open symmetric subsets of covering with and let and , , be odd continuous maps such that for all . Since is an absolute retract, we may assume that is the restriction of an odd continuous map . Let be the connected component of the complement of the Nehari manifold in which contains the origin, and set . Then, is a bounded open symmetric neighborhood of in .
Let . Then, are symmetric and open in , and they cover . Further, by Lemma 6.2,
The set consists of two connected components, see for example [10, Lemmas 2.5 and 2.6]. Therefore there exists an odd continuous map . Let be the restriction of the map if and the restriction of if . Take a partition of the unity subordinated to the cover consisting of even functions, and let be the canonical basis of . Then, the map given by
is odd and continuous, and satisfies for every . The Borsuk-Ulam theorem allow us to conclude that as claimed. ∎
6.1. Proof of auxiliary results
Proof of Lemma 6.1.
Fix , and define as
Observe that is linear and bounded because and since for any it holds that . By the Frèchet-Riesz representation theorem, there exists a unique such that the following equality holds for any ,
This implies that is a weak solution in to the equation
Now we show that if is -invariant then is -invariant. Consider the operator given by
Observe that for any we have that
Note that for any ,
Thus we see that is a critical value of by construction.
Recall that is a closed linear subspace of . We note that the orthogonal decomposition with respect to the standard inner product is also an orthogonal decomposition with respect to . Indeed if and , so that , then we have
Since we conclude that the previous expression equals zero applying Lemma 4.3.
Write , with and . We claim that is a critical point of . If the claim holds, then for any we have that , and since we know that for any , we conclude that , that is, as desired.
Now we prove the claim. For ,
This implies that is a critical point of . Now consider a fixed . Then,
Since , þ4.2 implies that is a critical point of . By linearity, it follows that is a critical point of . ∎
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 .
For every , Sobolev’s inequality yields a positive constant such that
(19) |
where in the last part we used the fact that for every with , and .
Now we assume that . We bound by noticing that implies that,
Then by definition of , see (12), we get
Taking , it follows that for all . Similarly, for all . Thus, , for every . ∎
To prove part of Lemma 6.2 we will use the following result.
Proposition 6.8 (Theorem 5.2 in [20]).
þ Let be a real vector space with a norm inducing a distance , open and closed convex with non empty interior with and such that the distance from any point in to is achieved by some point. Let be a locally Lipschitz function that satisfies
Then any continuous such that and in satisfies for all .
Proof of Lemma 6.2 part .
By symmetry, we will only prove this part for . The proof consists in the following two steps:
Step 1: For small enough, we apply þ6.8 taking , , and which is closed and convex, and , to obtain that
Step 2: We then apply Mazur’s separation theorem [38, Section 2.2.19] to prove that
Proof of Step 1: Recall that the gradient of with respect to is . In order to apply Proposition 6.8, for , we have to show that the limit as of the following expression equals zero:
This will be achieved by showing that . In particular, we show that for some and any ,
Since is given by a norm, by the triangle inequality it is enough to bound and separately.
We first note that and if . This follows from the fact that is compact, , and and the Maximum Principle. Indeed consider and denote by the derivative of at . Since is compact, then attains it minimum at some point . If we assume that , then since we also know that and , we have , which is a contradiction. The same reasoning holds for .
As in [12, Lemma 5.1], the following inequality holds true
(20) |
For let be such that dist Then, the previous paragraph, linearity of and inequality (20) yield
(21) |
Now we bound . For any function we defined as and . Then, since for we have , by the definition of , using that and identity (16), we get
To bound the integral above, recall that and . Thus, we can apply Hölder’s inequality, with , to the functions and , and then apply (19) and Sobolev’s inequality, to get
Combining the inequalities above, we conclude that for all with it holds
(22) |
This inequality is also true if , for, in this case and . So, inequality (22) holds true for every
Fix and let be such that . Then, for any by adding inequalities (21) and (22) we obtain
for all . Therefore, if .
Proof of Step 2: Observe that we have not ruled out . We now argue as in the proof of Lemma 2 in [17].
By contradiction, assume that there exists such that for some . By Mazur’s separation theorem, there exists a continuous linear functional and such that and for any . It follows that
Since by hypothesis , then by Step 1 above . Thus we obtain that
and consequently . Hence, there exists such that for . Thus, for . This contradicts Step 1.
Letting 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 , 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.