Equivariant solutions to the optimal partition problem for the prescribed -curvature equation
Abstract.
We study the optimal partition problem for the prescribed constant -curvature equation induced by the higher order conformal operators under the effect of cohomogeneity one actions on Einstein manifolds with positive scalar curvature. This allows us to give a precise description of the solution domains and their boundaries in terms of the orbits of the action. We also prove the existence of least energy symmetric solutions to a weakly coupled elliptic system of prescribed -curvature equations under weaker assumptions and conclude a multiplicity result of sign-changing solutions to the prescribed constant -curvature problem induced by the Paneitz-Branson operator. Moreover, we study the coercivity of -operators on Ricci solitons, compute the -curvature of these manifolds and give a multiplicity result for the sign-changing solutions to the Yamabe problem with prescribed number of nodal domains on the Koiso-Cao Ricci soliton.
1. Introduction
A milestone in Geometric Analysis and Conformal Geometry is the Yamabe problem, which consists in finding a conformal metric on a given Riemannian manifold in order to obtain constant scalar curvature (see, for instance, the book [AubinBook] or the classic survey [LePa], and the references therein). There are many generalizations of this problem. One direction is to consider higher-order operators generalizing the conformal Laplacian. For instance, a fourth-order operator on Riemannian manifolds of dimension four, which is the analogue to the bilaplacian on the Euclidean space, was discovered by S. Paneitz [Paneitz1983] and later generalized to higher dimension by T.P Branson in [Branson1995], and by Branson and B. Ørsted in [BransonOrsted1991]. A systematic construction of conformally invariant operators of higher order was given by C. R. Graham, R. Jenne, J. Mason and G. A. J. Sparling in [GJMS] (-operators for short) and, more recently extended to fractional order conformally invariant operators by S.-Y. A. Chang and M. d. M. González in [ChangGonzalez2011]. This paper deals with optimal partition problems related to equations given by GJMS-operators.
In order to briefly describe these operators, let be a closed (compact without boundary) smooth manifold of dimension and let such that . For any Riemannian metric for , there exists an operator , satisfying the following properties.
-
()
is a differential operator and , where is the Laplace Beltrami operator on .
-
()
is natural, i.e., for every diffeomorphism , , where “” denotes the pullback of a tensor.
-
()
is self-adjoint with respect to the -scalar product.
-
()
is conformally invariant, that is, given any function , if we define the conformal metric , then the following identity holds true:
(1)
There is a natural conformal invariant associated with given by
called the Branson -curvature, after Branson-Ørsted [BransonOrsted1991] and Branson [BransonBook, Branson1995], or simply the -curvature. When , is the conformal Laplacian and is the scalar curvature , while for , is the Paneitz-Branson operator and is the usual -curvature [DjadliHebeyLedoux2000].
When considering conformal metrics with in , equation (1) is written as
(2) |
where is the critical Sobolev exponent of the embedding . Here denotes the Sobolev space of order , which is the closure of in under the norm
(3) |
Taking in (2), one obtains the prescribed -curvature equation
(4) |
For and constant, we recover the Yamabe Problem. Some results about the existence and multiplicity of solutions to the prescribed -curvature problem can be found in [AzaizBoughazi2020, BaScWe, BenaliliBoughazi2016, DjadliHebeyLedoux2000, Ro, Tahri2020].
Given , we will deal with the -partition problem associated with (4) in the presence of symmetries. In order to describe the problem, let be a closed subgroup of Isom under suitable conditions (see hypotheses () and () below), and let be an open and -invariant subset, namely, if , then for every isometry . In what follows, denotes the closure of under the Sobolev norm given by (3).
We consider the symmetric Dirichlet boundary problem
(5) |
where is said to be -invariant if for every and any .
Denote by the least energy among the nontrivial solutions, that is,
In the absence of symmetries, the lack of compactness due to the critical Sobolev exponent nonlinearity in equation (5), implies that this number is not achieved in general [StruweBook, Chapter III]. However, when the domain is smooth, -invariant and the -orbits have positive dimension, this number is attained (see Proposition 2.3 below). The -partition problem consists in finding mutually disjoint and non empty -invariant open subsets such that
(6) |
where
The aim of this paper is to show that this problem has a solution on Einstein manifolds with positive scalar curvature, for every , and with less restrictive hypotheses on the metric, when .
In order to state our main result, we need to impose some conditions over and its isometry group (conditions to below). For the reader convenience, we recall some basic facts about isometric actions (see Chapter 3 and Section 6.3 in [AlexBettiol] for a detailed exposition). For any , the orbit of is the set , and the isotropy subgroup of is defined as . When is a closed subgroup of , the -orbits are closed submanifolds of which are -diffeomorphic to the homogeneous space , meaning the existence of a -invariant diffeomorphism between these two manifolds. An orbit is called principal, if there exists a neighborhood of in such that for each , for some . All points lying in a principal orbit have, up to conjugacy, the same isotropy group. Denote this group by , called the principal isotropy group. We say that induces a cohomogeneity one action on if the principal orbits have codimension one. In presence of a cohomogeneity one action, each principal -orbit is a closed hypersurface in diffeomorphic to , and there are exactly two orbits of bigger codimension, called singular orbits. Denote them by and . are closed submanifolds of codimension , and all points lying in have, up to conjugacy, the same isotropy group.
Denoting these groups by , we have that is -diffeomorphic to .
We can now state the conditions on the group . In what follows, will denote the geodesic distance between and , that is, .
-
is a closed subgroup of inducing a cohomogeneity one action on .
-
for every .
-
The metric on is a -invariant metric of the form:
where is one parameter family of metrics on the principal orbit, for some positive smooth functions defined on the interval , with suitable asymptotic conditions at and . More precisely, the (smooth) metrics are defined on the principal orbit at , for , and are defined around the singular orbits in such a way that , , satisfy appropriate smoothness conditions at the endpoints and of , that ensure that can be extended to the singular orbits:
(7) See [EscWang00, Section 1].
In Section 6, we give concrete examples where these hypotheses hold true.
We are now ready to state our main result, which describes the solution to the -optimal partition problem in terms of the principal orbits of cohomogeneity one actions. The symbol “” will stand for “-diffeomorphic to”.
Theorem 1.1.
Let and let be a closed Riemannian manifold of dimension with scalar curvature . If
-
•
for ; or
-
•
is Einstein with ,
then for any and any subgroup of Isom satisfying (), () and (), the -invariant -partition problem (6) has a solution such that:
-
(1)
is connected for every , , if and .
-
(2)
The sets and are -diffeomorphic to disk bundles at and , respectively. More precisely,
where are normal bundles over the singular orbits . See Section 5 for details.
-
(3)
For each , , and , where denotes the disjoint union of sets.
In Section 6 we give several examples of Einstein manifolds with positive scalar curvature satisfying conditions (), () and (). This result extends Theorem 1.1 in [CSS21] and Theorem 1.2 in [ClappFernandezSaldana2021] to more general manifolds and actions than the sphere with the -action, . In the case of , M. Clapp and A. Pistoia [ClappPistoia21] solved the -invariant -partition problem for actions with higher cohomogeneity and in [ClappPistoiaTavares21], the authors solved the -partition problem without any symmetry assumption. However, neither the structure nor the domains solving the problem are explicit in these works.
Notice that in case , the scalar curvature is -invariant, since we are regarding isometric actions. An interesting application of Theorem 1.1 in this case, is the following result for the Yamabe problem.
Corollary 1.2.
Let be a closed Riemannian manifold of dimension and let be a closed subgroup of satisfying to . If , then for any , the Yamabe problem
admits a -invariant sign changing solution with exactly -nodal domains. Moreover, it has least energy among all such solutions and the nodal set is a disjoint union of principal orbits.
This was proven in the case in [ClappPistoia21] for general manifolds in the presence of symmetries, and in [ClappPistoiaTavares21] in the non-symmetric case; for any , the only manifold where this was known is the round sphere in the presence of symmetries given by isoparametric functions [FdzPetean20, CSS21].
This corollary clearly holds on Einstein manifolds with positive scalar curvature, for which the scalar curvature is constant. Then, a natural setting for extending its applications is Ricci soliton metrics. Recall that a Ricci soliton on a closed smooth manifold is a Riemannian metric satisfying the equation
(8) |
for some constant or , and some smooth function , called the Ricci potential. Here, as usual, denotes the Ricci curvature of , and denotes the Hessian of with respect to . A Ricci soliton is called steady, shrinking or expanding according to , or , respectively.
In these terms, a nontrivial explicit example that fulfills the hypotheses of Corollary 1.2 and that does not reduce to an Einstein metric, is the Koiso-Cao Ricci soliton [Koiso, Cao1]. It is known that this metric is a cohomogeneity one Kähler metric on , with respect to the action of U, the unitary group of dimension . Here, as usual, denotes the connected sum of smooth manifolds and denotes a smooth manifold taken with reverse orientation. It is also known that it is non-Einstein with positive Ricci curvature. The singular orbits of the U-action are both diffeomorphic to , and the principal orbits are diffeomorphic to . The associated fibrations, as a cohomogeneity one manifold, are both given by the Hopf fibration:
Here, for an orientable manifold , we will denote by to refer to the opposite orientation.
See [TOR17, Section 2] for details on the construction of the Koiso-Cao soliton from the Hopf fibration.
In order to establish the existence of a solution to the optimal partition problem (6), we follow the approach introduced in [ContiTerraciniVerzini2002], consisting in the study of a weakly coupled competitive system together with a segregation phenomenon. To this end, we will study the existence of -invariant solutions to the following weakly coupled competitive -curvature system
(9) |
where , and are such that and . We will say that a solution to the system (9) is fully nontrivial, if, for each , is non trivial.
To assure the existence of a fully nontrivial least energy -invariant solution to the system (9) we will only assume that satisfies (), allowing actions with bigger codimension of its principal orbits, and also we will assume that the operator is coercive, meaning the existence of a constant such that
We will say that a sequence of fully nontrivial elements in the Sobolev space
is unbounded if as , for every In this direction, we state our next multiplicity result.
Theorem 1.3.
If the operator is coercive and () holds true, then the system (9) admits an unbounded sequence of -invariant fully nontrivial solutions. One of them has least energy among all fully nontrivial -invariant solutions.
When and , by a well-known argument given in [BenciCerami1991, Proof of Theorem A], the least energy solutions are positive and, hence, they give rise to a -invariant solution to the Yamabe problem. Seeking for this kind of metrics is a classical problem posed by E. Hebey and M. Vaugon in [HebeyVaugon1993]. However, there is a gap in Hebey and Vaugon’s proof, for they used Schoen’s Weyl vanishing conjecture, which turns out to be false in higher dimensions by Brendle’s counterexample in [Brendle2008]. F. Madani solved the equivariant Yamabe problem in [Madani2012] assuming the positive mass theorem to construct good test functions. Here, the positive dimension of the orbits given by hypothesis (), avoids these problems in higher dimensions. For , the Maximum Principle for the operator is not true in general, and the least energy solutions may change sign. In fact, it is not clear whether the components of the solutions to the system (9) are sign-changing or not. In case and , by a recent result by J. Vétois [Vetois2022, Theorem 2.2], we can assure that an infinite number of the solutions to
(10) |
must change sign when considering to be Einstein with positive scalar curvature, as we next state.
Corollary 1.4.
Let , be an Einstein manifold of dimension , with positive scalar curvature and not conformally diffeomorphic to the standard sphere. If is a closed subgroup of satisfying (), then the problem with Paneitz-Branson operator (10) admits an unbounded sequence of -invariant sign-changing solutions.
There are many examples of manifolds admitting an Einstein metric with positive scalar curvature, in order to ensure the coercivity of . In Section 6, we describe explicit examples and some construction that provide large classes of examples. However, it is difficult to find nontrivial examples of non-Einstein manifolds for which the operator is coercive. Towards this direction, the next result for the Paneitz-Branson operator gives a sufficient condition for this to happen.
Proposition 1.5.
For and , if and , then is coercive.
Proof.
It is a direct consequence of a calculation obtained in [GurskyMal03]. It appears as equation (2.18):
∎
Again, it is complicated to compute the -curvature of an arbitrary manifold and the literature lacks examples, different from Einstein manifolds. We explore the possibility of obtaining positive -curvature and coercivity for more general manifolds, such as Ricci solitons. This is difficult to check because almost nothing is known about the Ricci curvature tensor in Ricci solitons. Our next result gives an explicit formula of the -curvature of these metrics.
Theorem 1.6.
The -curvature of a shrinking Ricci soliton , with Ricci potential , is given by:
where
If has radially positive Ricci curvature, i.e., , and
-
•
for , ,
-
•
or for , and ,
then .
Remark 1.1.
The condition of having radially positive Ricci curvature is plausible, due to the fact that if everywhere, then the Ricci soliton is trivial [PetWyl09a, Theorem 1.1]. Moreover, if it is radially flat, i.e., , then it is also trivial [PetWyl09b, Proposition 2]). If is a cohomogeneity one Kähler-Ricci soliton under the action of , then one may average over to obtain a Ricci potential that is a -invariant function on . Then the following identity holds true:
Since a non-trivial shrinking Kähler-Ricci soliton has positive Ricci curvature, then under a cohomogeneity one action, it also has radially positive scalar curvature. Even more, all the known compact non-Einstein Ricci solitons are Kähler[Cao2010]. ∎
Corollary 1.7.
Let be Einstein with or a shrinking Ricci soliton with radially positive Ricci curvature of dimension . If () holds true, then the system (9) admits an unbounded sequence of fully nontrivial -invariant solutions, one of them with least energy among all fully nontrivial -invariant solutions.
Our work is organized as follows. In Section 2, we describe the variational setting in order to prove the existence of least energy -invariant solutions to the problem (5). Next, in Section 3, we give the variational setting to study system (9) and prove Theorem 1.3 and Corollary 1.4. In Section 4, we describe how hypotheses and allow us to reduce the partition problem to a simpler one dimensional problem. In Section 5, we study the segregation phenomenon that gives the description of the domains and -invariant functions that solve the optimal partition problem, showing Theorem 1.1. As an application of Theorem 1.1 with , we prove Corollary 1.2. We give several examples where hypotheses to hold true in Section 6. Finally, in Section LABEL:Section:Q-curvatureRicciSolitons, we compute the -curvature of shrinking Ricci solitons, proving Theorem 1.6.
2. Symmetries and least energy solutions
In this section, we study the existence of least energy solutions to problem (5).
From now on, will denote either or an open, connected -invariant subset of with smooth boundary, will denote the usual norm in , . For , the -th covariant derivative of will be denoted by and we define its norm as the function given by
where we used the Einstein notation convention.
The Sobolev space is the closure of under the norm
where, with some abuse of notation,
Notice that in case , then and ; if , then is a closed subspace of .
Let be the corresponding GJMS-operator in of order . For each , there exists a symmetric -tensor field on , which we will denote by , such that the operator can be written as
where the indices are raised via the musical isomorphism. In particular, for any , integration by parts yields that
See [Ro, Proposition 1] for the details.
As a consequence, the bilinear form can be extended to a continuous symmetric bilinear form on . When is coercive, this bilinear form is actually a well-defined interior product on that induces a norm equivalent to (see [Ro, Proposition 2]). We will denote this interior product and norm by and respectively, and endow with it in what follows. Notice that, by definition,
for every .
The group acts on in the usual way:
Every element induces a linear map
We next show that the norm is invariant under the action of isometries.
Lemma 2.1.
For every and every
In particular, is a linear isometry.
Proof.
Let and . Then is a diffeomorphism and, by the naturality of , property () in the introduction, we obtain for any isometry that
where denotes the pullback metric.
Now, to see that induces a linear isometry, recall that if , then
(11) |
(Cf. [HebeyBook1997, Théorème 4.1.2]). Then, for every ,
Density of in yields the result. ∎
Every also induces a linear map , , given by , which is also an isometry, thanks to (11).
Let be any closed subgroup of Isom such that for any . From now on, suppose that is -invariant, namely, if , then . In this way, for any and every isometry , it follows that
and also induces linear isometries
(12) |
for any .
We define the Sobolev space of -invariant functions as
This is a closed subspace of . In fact, if denotes the space of smooth functions with compact support in which are -invariant, then coincides with the closure of this space under the Sobolev norm . As the dimension of any orbit is strictly less than , the space is infinite dimensional, thanks to the existence of -invariant partitions of the unity (Cf. [Palais1961, Theorem 4.3.1] and also [AlexBettiol, Claim 3.66]).
We will need the following Sobolev embedding result.
Lemma 2.2.
Let be a closed subgroup of , , and let be a -invariant domain. Define
Then
is continuous and compact for every .
Proof.
The case for is just Theorem 2.4 in [IvanovNazarov2007].The case for follows from a bootstrap argument as in Proposition 2.11 in [AubinBook]. ∎
In what follows, we will suppose that satisfies condition (). Under this hypothesis, the existence of least energy -invariant solutions to (5) follows directly from standard variational methods using the Palais’ Principle of Symmetric Criticality [Palais1979] together with Lemma 2.2. For the reader’s convenience, we sketch the proof of a slightly more general result, namely, we show the existence of -invariant solutions to the problem
(13) |
where .
Consider the functional
Given that is a well defined norm equivalent to the standard norm , Sobolev inequality yields that it is a functional for any . From (12), this functional is -invariant, namely it satisfies that
Hence, due to Palais’ Principle of Symmetric Criticality [Palais1979], the critical points of restricted to correspond to the -invariant solutions to the problem (13). The nontrivial ones belong to the set
which is a codimension one Hilbert manifold in . Notice that
Thanks to the Sobolev inequalities [AubinBook, Theorem 2.30], is closed and
We say that satisfies condition in if every sequence such that and in as , has a convergent subsequence.
As satisfies (), then and . Hence, by Lemma 2.2, satisfies condition and Theorem 7.12 in [AmbrosettiMalchiodiBook] yields that is attained. Thus, there exists a least energy -invariant solution to the problem (13).
We summarize this analysis in the following proposition.
Proposition 2.3.
If satisfies () and if is coercive, then, for any the problem (13) admits a least energy -invariant solution.
To our knowledge, this is the first existence result of symmetric least energy solutions to the homogeneous Dirichlet boundary problem (13). Another result for non-homogeneous Dirichlet boundary conditions to problems involving the -operators, in the absence of symmetries, can be found in [BekiriBenalili2018, BekiriBenalili2019, BekiriBenalili2022].
Remark 2.1.
With slight modifications, the same result is true for operators given by a linear combination of Laplacians, i.e., for operators having the form
where is positive, and are constants for if is even, and and are constants for if is odd. This is true because the Principle of Symmetric Criticality can be applied by noticing that
for any and every isometry (see, for instance, [AmannEscherBook, Remark 6.9 (c)]).∎
3. The polyharmonic system
We next study the system (9). Fix and consider the product space endowed with the norm
Let be the functional given by
This is a functional and its partial derivatives are given by
for every and every Hence, solutions to the system (9) correspond to the critical points of
Fix now a closed subgroup of isometries satisfying () and define . This is a closed subspace of . By Lemma 2.1, is -invariant and by the Principle of Symmetric Criticality [Palais1979], the critical points of restricted to are the -invariant solutions to the system (9). Hence, we can restrict ourselves to seek critical points of in . Observe that the fully nontrivial ones belong to the set
It is readily seen that
(14) |
Lemma 3.1.
There exists , independent of , such that if . Thus, is a closed subset of and .
Proof.
Since and as the norm is equivalent to the standard norm in , for any , it follows from the Sobolev inequality the existence of a constant such that
The result follows from this inequality. ∎
A fully nontrivial solution to the system (9) satisfying is called a -invariant least energy solution. To establish the existence of fully nontrivial critical points of , we follow the variational approach introduced in [ClappSzulkin19]. The proof of Theorem 1.3 is, up to minor modifications, the same as in [ClappFernandezSaldana2021, Theorem 1.1], but we sketch it for the reader’s convenience.
Given and , we write
Let , define and
The next result is proved exactly in the same way as [ClappSzulkin19, Proposition 3.1].
Lemma 3.2.
-
Let . If there exists such that , then is unique and satisfies
-
is a nonempty open subset of , and the map given by is continuous.
-
The map given by is a homeomorphism.
-
If is a sequence in and , then .
Define as
According to Lemma 3.2, is a Hilbert manifold, for it is an open subset of the smooth Hilbert submanifold of . When is differentiable at , we write for the norm of in the cotangent space to at , i.e.,
where is the tangent space to at .
Recall that a sequence in is called a -sequence for if and , and is said to satisfy the -condition if every such sequence has a convergent subsequence. Similarly, a -sequence for is a sequence in such that and , and satisfies the -condition if any such sequence has a convergent subsequence. Here denotes the dual space of .
Lemma 3.3.
-
and its derivative is given by
Moreover, there exists such that
-
If is a -sequence for in , then is a -sequence for in .
-
is a critical point of if and only if is a fully nontrivial critical point of .
-
If is a sequence in and , then .
-
if and only if , and .
We omit the proof of this lemma, because the argument is exactly the same as in [ClappSzulkin19, Theorem 3.3].
Lemma 3.4.
For every , satisfies the -condition.
Proof.
First observe that a -sequence for is bounded. Indeed, there exists such that
and the claim follows.
Using this, let be a -sequence for . By Lemma 3.3, the sequence is a -sequence for and it is bounded by the above claim. A standard argument using Lemma 2.2 as in [ClappPistoia2018, Proposition 3.6], shows that contains a convergent subsequence, converging to some . As for every and as is closed by Lemma 3.1, it follows that . Finally, since is a homeomorphism between and , this yields that converges to in a subsequence, and satisfies the -condition ∎
Given a nonempty subset of such that if and only if , the genus of , denoted by , is the smallest integer such that there exists an odd continuous function into the unit sphere in . If no such exists, we define ; finally, we set .
Lemma 3.5.
.
Proof.
Proof of Theorem 1.3.
Lemma 3.3 implies that is positively invariant under the negative pseudogradient flow of , so the usual deformation lemma holds true for , see e.g. [StruweBook, Section II.3] or [WillemBook, Section 5.3]. As satisfies the -condition for every , standard variational arguments show that attains its minimum on at some . By Lemma 3.3 and the Principle of Symmetric Criticality, is a -invariant least energy fully nontrivial solution for the system (9). Moreover, as is even and , arguing as in the proof of Theorem 3.4 (c) in [ClappSzulkin19], it follows that has an unbounded sequence of critical points. Using Lemma 3.3 (iii), and the fact that by (14), the system (9) has an unbounded sequence of fully nontrivial -invariant solutions. ∎
We next apply Theorem 1.3 to the case and a recent result by J. Vétois to prove the multiplicity result stated in Corollary 1.4.
Proof of Corollary 1.4.
Theorem 2.2 in [Vetois2022] states that the positive solutions to the problem (10) must be constant and by the concrete expression of the Paneitz-Branson operator and the -curvature on Einstein manifolds (see, for instance, [DjadliHebeyLedoux2000] and [Gover06] respectively) it is unique. As is Einstein with positive scalar curvature, the operator is coercive [Ro, Proposition 4] and Theorem 1.3 for , yields the existence of an unbounded sequence of -invariant solutions, and the corollary follows. ∎
4. One dimensional reduction
In this section, we will strongly use that the group satisfies properties () and (). Recall that and denote the singular orbits, as it was given in the introduction, and let and , . Since is compact, the geodesic distance between and ,
is attained and the distance function given by
is well defined. This function is a Riemannian submersion and satisfies for any that
where denotes the mean curvature of . See [Petersen06, Chapter 2, Section 4.1], and also [BBP21, Section 2]. For the mean curvature, we explicitly have:
for some matrices and not depending on , and it further satisfies that
(see [GeTang2014], and also [BBP21]). Therefore, for any the following identity holds true:
(15) |
Moreover, notice that is a principal orbit for any , while and . Hence, for every , it follows that
Thus, for any , the function and, conversely, for any there exists a unique such that . In this way, we have a linear isomorphism
(16) |
As in the introduction, will denote the principal isotropy, i.e., the stabilizer of the -action at any point , for some . Such a group is the same at the preimage of any interior point of under . All the regular orbits are diffeomorphic to , and we will fix one, say .
Lemma 4.1.
Assume () and () hold true. Then there exists a metric on , a diffeomorphism and a smooth function such that
-
(1)
for every , .
-
(2)
.
Proof.
The first item only depends on () as follows: For any , consider the unique minimizing horizontal geodesic joining and such that (Cf. [AlexBettiol, Proposition 3.78]). Then, the diffeomorphism is given by
If necessary, we may reparametrize so that .
The second item follows from the fact that the volume form of a metric given as in (), is the volume product. That is, for a local coordinate system in , around an arbitrary point , the set is a basis of the tangent space . Then, around the metric is given by the matrix
where is the matrix corresponding to the metric , . If is the size of , then , and the volume form of is given by
Define and take , which is a metric on . Therefore, is given by
and we conclude the result. ∎
We can adapt Lemma 2.2 in [FdzPetean20] to the context of cohomogeneity one actions. Recall that denotes the metric given to the principal orbits in ().
Lemma 4.2.
For any integrable function ,
In particular,
Proof.
As has Lebesgue measure zero on , by Lemma 4.1 and Fubini’s Theorem, we obtain on the one hand that
On the other hand, using the second expression for the volume element in in 4.1,
where we conclude the integral identity.
For the volume identity, subtracting the above identities we obtain
for any integrable function . We conclude that almost everywhere in . As the volume function and are continuous, we conclude the identity. ∎
Now we study the preimage of measure zero subsets in under the distance function . To this end, denote the Lebesgue measure in by , and by the Lebesgue measure in .
Lemma 4.3.
If satisfies , then .
Proof.
Observe that the critical point set of is exactly the union of the singular orbits of the action, and . Those points correspond to the endpoints of , under . By dimension reasons, . Therefore, it is enough to prove the statement for any proper subset of zero Lebesgue measure. Hence, we may assume that at any point in . Recall that, is a copy of the principal orbit, i.e., it is a submanifold of of dimension , for any . Therefore , for any .
Write . If is countable, then is a countable union of zero measure sets, so has measure zero. As has measure zero in and as in , if is uncountable, the co-area formula yields that
where is the characteristic function of in . Since , then the Lebesgue integral of any measurable function over is zero. In particular, this implies that . ∎
In what follows, we will denote the set of positive and smooth -invariant functions on by and by the set of positive and smooth functions in . Next, we study how the symmetries allow us to reduce the operator into an operator acting on smooth functions defined in the interval . In order to motivate a more general differential operator for which our theory holds true, first observe that if is Einstein with positive scalar curvature , the higher order conformal operator, , can be written as
for some suitable constants (see [Gover06, Juhl13]). On the other hand, in case , when the scalar curvature is positive, the conformal Laplacian is simply
with . Hence, for any and any , we are led to define the operator
(17) |
As for any and any pair of functions we have that
(18) |
then the bilinear form
(19) |
is positive definite and yields a norm in , equivalent to the standard norm in . Note that the term for is simply and but not necessarily constant.
On the other hand, let be such that , , , and define the operator by
For define
where denotes the -fold composition of , and
Proposition 4.4.
For any and any ,
Proof.
From this result, it follows that is a well defined norm in and we define the weighted Sobolev space to be the closure of under this norm.
We have the following direct consequence of the previous result.
Theorem 4.5.
For any , the linear isomorphism given in (16), induces a well defined continuous isometric isomorphism
Proof.
Take and as dense subspaces of and under the norms and , respectively. By Proposition 4.4, given any , , we have that , and the map is a linear and continuous isometric isomorphism. Thus, can be extended, in a unique way, to a linear and continuous isometric isomorphism defined on the whole Sobolev space , as we wanted to prove. ∎
Next we see how the standard norm in is related with the weighted norms
Lemma 4.6.
For each , there exist , with constant, and , depending on , such that
Proof.
As is continuous and positive in , . Then, it is readily seen that, for any , the inequality
holds true, where is a suitable constant depending only on , and .
The proof of the second inequality is exactly the same as in [ClappFernandezSaldana2021, Lemma 2.3].
∎
Corollary 4.7.
For any there exists such that
Proof.
The proof is virtually the same as in [ClappFernandezSaldana2021, Proposition 3], but we include it for the sake of completeness.
First observe that for any , the operator is coercive and, therefore the norm is equivalent to the standard norm in . Next, fix , take and as in Lemma 4.6 and let . Since is equivalent to the standard norm in , the map is continuous. Since the Sobolev embedding is also continuous, for any and such that , there exists such that a.e. in . Applying Lemma 4.3 it follows that a.e. in , where . As has measure zero in , the function given by if , is well defined, is of class on and coincides a.e. with on . ∎
For any , let . We now show that the Dirichlet boundary problem
(21) |
induces a one dimensional Dirichlet boundary problem. We need some preliminary lemmas.
Lemma 4.8.
Let for a smooth function . Then for ,
where denotes the -th derivative of over , is a tensor, varying with , which is a combination of tensor products of the tensors . Moreover, for any ,
Proof.
We will proceed by induction over .
Case . Denote by an arbitrary vector field that is tangent to the level sets of the distance function . By definition of gradient and by the chain rule,
Analogously, computing in the direction of the normal to and using the fact that , we get that
Therefore,
Case . Recall that the Levi-Civita connection induces a covariant derivative for higher-order tensors. Given a tensor of order , the derivative is a tensor of order given by the formula:
We then compute for any vector fields on :
Therefore:
Case . Now, suppose that
where the tensors satisfy the conditions in the Lemma. Take any vector fields on and compute:
(22) | ||||
We substitute the expression for . For the sake of clarity, let us analyze the first summand:
Now,
Note that gives the only term with the factor in the expression for . Observe also that is one of the terms in the definition of ; the others will be obtained from the remaining terms in (22). If is a combination of tensor products of , the same happens with . After a long but straightforward calculation, we have
for some tensors . By the inductive hypothesis,
which proves the lemma. ∎
Proposition 4.9.
Let , for a smooth function . Let , and a fixed integer such that for all , then , for all .
Proof.
We will proceed by induction over . For , we have:
Evaluating at and ,
Now take and suppose that , for all and . Using this and the previous lemma,
Evaluating at and ,
The result follows. ∎
Corollary 4.10.
Let such that and define the operator
where If is a solution to (21), then is a solution to the problem
(23) |
5. Segregation and optimal partitions
In this section, we will suppose that is such that the operator can be written in the form (17) and that satisfies () and (), so that the results of the previous section hold true. Remember that this is possible, for example, if is an Einstein manifold with positive scalar curvature or when in case . See Section 6 for concrete examples.
Recall that for a compact Lie group , a principal -bundle is a fiber bundle , whose structure group is , together with a -action on itself by left translations, and a free right -action on , whose orbits are the fibers of the bundle. Let be another smooth manifold that admits a left action by . The orbit space of the diagonal action on is a smooth manifold denoted by , given as the total space of the fiber bundle
In the literature, the latter is known as the associated bundle to the principal bundle , and is called the twisted space. See [AlexBettiol, Section 3.1] for definitions and a detailed explanation.
Let be a -invariant open subset of with smooth boundary and recall the definitions of the energy functional and the Hilbert manifold given in Section 2. By Proposition 2.3, problem (5) admits a least energy -invariant solution. So the quantity defined in the introduction is attained.
Theorem 1.1 will follow from the next segregation result.
Theorem 5.1.
Suppose satisfies conditions (), () and (), and that can be written as a sum of Laplacians of the form (17). For , fix and for each , , let be such that and as . Let be a least energy fully nontrivial solution to the system (9) with . Then, there exists such that, up to a subsequence,
-
is a solution to the -invariant –optimal partition problem (6) satisfying the following properties:
-
(1)
is smooth and connected for every , , if and ;
-
(2)
-
(3)
For each ,
where denote disk bundles at the singular orbits.
-
(1)
To prove this theorem we will need the following lemma, which is a version of the unique continuation principle that is suitable to our situation.
Lemma 5.2.
Let and let be a -invariant solution to the Dirichlet boundary problem (21) in . If in any subdomain of the form , , then in the whole interval .
Proof.
As is a strong -invariant solution to (21), for some and Corollary 4.10, is a strong solution to 23. Since in and in , then necessarily for any .
Now, if there is no such that , then in and there is nothing to prove. If this is not the case, there must be such that . Without loss of generality, suppose that ; therefore, there must exist such that in and in . As is of class , all its derivatives of lower order are continuous and it follows that for every . By existence and uniqueness of the initial value problem
vanishes identically in a small neighborhood of , contradicting that in . Something similar holds true if for some . Hence in , as we wanted to show. ∎
The following topological lemma will be useful in what follows.
Lemma 5.3.
Let be two topological spaces and be a quotient map. If is connected for every , then is connected for every connected subset .
Proof.
Let be connected and consider . Let be any continuous function. Since is connected for every , then for any , implies that , for maps connected sets into connected sets. Therefore induces a continuous function such that . As is connected, is constant and so is . As was an arbitrary continuous function, it follows that must be connected. ∎
The following result allow us to describe the nodal domains in terms of the orbit structure.
Proposition 5.4.
Proof.
First notice that every connected -invariant open set must be of the form for some . Now take three points and define the sets , and . Note that these sets are -invariant by construction, their boundaries are smooth because is a principal orbit, and since every orbit is connected and is a quotient map, then and are also connected by Lemma 5.3. Hence, by Proposition 2.3, the least energy solution to the problem (5) is attained in each domain and
where this inequality holds true because and every nontrivial function in can be extended by zero to a nontrivial function in . We next prove that the inequality is strict. Suppose, to get a contradiction, and without loss of generality, that and let be a least energy solution to the Dirichlet boundary problem (5) in . Therefore the function given by in and in is a least energy solution to (5) in and by interior regularity [UhlenbeckViaclovsky2000], this function has a class representative. Proposition 5.2 yields that must vanish in , which is a contradiction and the strict inequality follows.
Hence, if is a -invariant solution to the -partition problem (6), then consists exactly of points, say .
Now define as in the statement. As is either a connected principal orbit or a connected singular orbit for each , by Lemma 5.3 these sets are connected and consists in one or two disjoint principal orbits, from which it follows that these sets are smooth. As , for (where and ), then . By definition, if , , and
The equivariant form of the sets and follows from the Tubular Neighborhood Theorem [AlexBettiol, Theorem 3.57]. In fact, they are the associated bundles to the principal -bundle .
Finally as , then and is also a solution to problem (6). ∎
Proof of Theorem 5.1.
With minor modifications, the proof is the same as in [ClappFernandezSaldana2021, Theorem 1.2]. We sketch it for the reader’s convenience.
Fix in (9) for each , and let be a sequence of negative numbers such that and as . To highlight the role of , we write and for the functional and the set associated to the system (9), introduced in Section 3, with replaced by . By Theorem 1.3, for each we can take such that
Let
Then, for all and, therefore,
(24) |
We claim that
(25) |
Indeed, if and , then extending this function by zero outside , we get that and a.e. in , for . Therefore and
As was arbitrary, it follows that
and as was arbitrary, inequality (25) follows.
From (24), it follows that the sequence is bounded in . So, using Lemma 2.2, after passing to a subsequence, we get that weakly in , strongly in , and a.e. in for each . Moreover, as , we have for each ,
Then, Fatou’s lemma yields
Hence, a.e. in . By Lemma 3.1,
and, as strongly in and weakly in , we get
(26) |
Since , there is a unique such that . So . The inequality (26) implies that . Therefore,
It follows that strongly in and , yielding
(27) |
By Corollary 4.7, we can take so that in , . It follows from continuity that the set
is nonempty, open, -invariant and if .
Set
These sets are also nonempty, -invariant and open, and satisfy that if and in . Hence . Since each connected component of is of the form for some , it is smooth and
(Cf. [ClappFernandezSaldana2021, Lemma A.1] and [GrisvardBook, Theorem 1.4.2.2]). Hence, as in , in and satisfies (27), it follows that . As , using the claim (25) we obtain that
(28) |
Also, from this inequality we obtain that for every , for otherwise, we would get that second inequality in (28) is strict, yielding a contradiction. Hence is a (weak) solution to the Dirichlet boundary problem (5) and is a solution to the -invariant -partition problem (6). Proposition 5.4 yields that, actually, satisfy properties (b.1) to (b.3) in Theorem 5.1. ∎
Remark 5.1.
For the case , that is, when is just the conformal Laplacian, we have the following result from which Corollary 1.2 follows immediately.
Corollary 5.5.
Let be a closed Riemannian manifold of dimension and let be a closed subgroup of satisfying to . If the scalar curvature is positive, and if is the solution to the optimal -invariant -partition problem given in Theorem 5.1, then the function
is a -invariant sign-changing solution to the Yamabe problem
having exactly nodal domains and having least energy among all such solutions.
Proof.
Since the maximum principle is valid for the operator , the least energy -invariant fully nontrivial solution to the system (9) given by Theorem 1.3 can be taken to be positive in each of its components [ClappSzulkin19, Theorem 3.4 a)]. In this way, each component of the functions , defined in the proof of Theorem 1.3, is nonnegative, yielding that is also nonnegative for every and . The rest of the proof is, up to minor details, the same as in the proof of item (iii) of Theorem 4.1 in [CSS21]. ∎
6. Examples of cohomogeneity one actions satisfying hypothesis of Theorem 1.1
In this section we will see concrete examples in which Theorem 1.1 can be applied. First, we will discuss cohomogeneity one actions where the metric decomposition () holds true.
Let be a closed subgroup of isometries of inducing a cohomogeneity one action. As before, the principal orbits of the action correspond to the hypersurfaces given by the regular level sets of , and if denote the principal isotropy, all of them are diffeomorphic to (see [AlexBettiol, Proposition 6.41]). Hence, we can fix one of these level hypersurfaces, say . When we have a cohomogeneity one action by isometries, we can describe the decomposition of the metric in terms of a one-parameter family of metrics on as follows. It is known that, given a minimizing horizontal geodesic between the and , a -invariant metric away from the singular orbits can be written as:
for , where is a smooth family of homogeneous metrics on . Let , be the Lie algebras of and , respectively. Let be the orthogonal complement of in , with respect to a bi-invariant metric on . Since is compact, then it admits such a metric. There is a natural identification of with the tangent space to a principal orbit . Namely, for each ,
is a tangent vector at . Then corresponds to a -parameter family of invariant inner products on .
One way to obtain the metric decomposition () is when decomposes into mutually orthogonal Ad-invariant subspaces:
(29) |
such that the metric can be written as:
for some positive smooth functions , on , and satisfying some smoothness conditions at and . Such conditions describe a compactification of by adding two compact submanifolds, corresponding to the endpoints of .
These types of cohomogeneity one metrics are called diagonal metrics. A decomposition in this form can be obtained in several settings:
-
•
If is a simple Lie group (i.e., if is simple), by the Schur’s Lemma [Hall15, Theorem 4.29].
-
•
If is a semi-simple Lie group. It follows from the Weyl’s Theorem on complete reducibility, which ensures that one obtains a decomposition (29) with irreducible factors, with respect to the adjoint representation; and by the Schur’s Lemma. See [Hall15, Theorem 7.8].
-
•
If the Killing form of is negative definite. More generally, if the adjoint representation of is unitary, one also has such a decomposition. See [Hall15, Proposition 4.27].
Recall also a fundamental fact: A cohomogeneity one action can be determined through a group diagram , provided that are spheres (see Section 6.3 from [AlexBettiol]).
We develop this theory for the following well-known example of an isometric action on the round sphere that has been used in several papers to obtain sign-changing solutions to semilinear elliptic problems (see, for instance, [Ding1986, BaScWe, ClappFdz17, FdzPetean20, CSS21, ClappFernandezSaldana2021]).
Example 6.1.
Consider the sphere with its canonical metric, which is an Einstein metric with positive scalar curvature. Let be integers such that . Set
where is the group of linear isometries of . Note that we are regarding as acting on , trivially in one component, and with the transitive action by rotations in the other one. Then we obtain two possible isotropy groups . A similar approach considers as a subgroup of . In those cases, the isotropy is a copy of or , for each corresponding case. Using that the -sphere can be described as the quotient , we obtain the following quotients:
Hence, the group diagram defines a cohomogeneity one action of on with singular orbits and , and with principal orbit . The orbit space is diffeomorphic to . Therefore, this shows that conditions and are fulfilled.
The Lie algebra of is isomorphic to , where denotes the Lie algebra of the skew-symmetric matrices. It is a simple Lie algebra of dimension , except for the case which is semi-simple. In this last case, its decomposition into simple factors is:
As previously, denote by the -invariant complement of in , where is the Lie algebra of . It is canonically identified with the tangent space at , . Then can be decomposed into two factors and given by:
with and . If (or ), then
Therefore, in any case an invariant metric on can be written as:
where is a bi-invariant metric on . We may then take:
Observe that these functions satisfy the smoothness conditions (7). Then condition is also satisfied.
The mean curvature of the principal orbit is:
Then the volume of the principal orbits along is:
where is the -dimensional measure of the sphere , for . This setting was under consideration in [ClappFernandezSaldana2021], where the authors studied the system (9) on and on ∎