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

Equivariant solutions to the optimal partition problem for the prescribed QQ-curvature equation

Juan Carlos Fernández Facultad de Ciencias, Universidad Nacional Autónoma de México (UNAM), Ciudad de México, México. jcfmor@ciencias.unam.mx Oscar Palmas∗∗ Facultad de Ciencias, Universidad Nacional Autónoma de México (UNAM), Ciudad de México, México. oscar.palmas@ciencias.unam.mx  and  Jonatán Torres Orozco∗∗∗ Facultad de Ciencias, Universidad Nacional Autónoma de México (UNAM), Ciudad de México, México. jonatan.tto@gmail.com
Abstract.

We study the optimal partition problem for the prescribed constant QQ-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 QQ-curvature equations under weaker assumptions and conclude a multiplicity result of sign-changing solutions to the prescribed constant QQ-curvature problem induced by the Paneitz-Branson operator. Moreover, we study the coercivity of GJMSGJMS-operators on Ricci solitons, compute the QQ-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.

Partially supported by CONACYT A1-S-10457 grant and by UNAM-DGAPA-PAPIIT IN101322 grant.
∗∗ Supported by UNAM-DGAPA-PAPIIT IN101322 grant.
∗∗∗ Supported by a CONACyT postdoctoral fellowship (Mexico) and partially supported by UNAM-DGAPA-PAPIIT IN101322 grant.

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] (GJMSGJMS-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 MM be a closed (compact without boundary) smooth manifold of dimension NN and let mm\in\mathbb{N} such that 2m<N2m<N. For any Riemannian metric gg for MM, there exists an operator Pg:𝒞(M)𝒞(M)P_{g}:\mathcal{C}^{\infty}(M)\rightarrow\mathcal{C}^{\infty}(M), satisfying the following properties.

  1. (P1P_{1})

    PgP_{g} is a differential operator and Pg=(Δg)m+lower order termsP_{g}=(-\Delta_{g})^{m}+\emph{\text{lower order terms}}, where Δ=divg(g)\Delta=\text{div}_{g}(\nabla_{g}) is the Laplace Beltrami operator on (M,g)(M,g).

  2. (P2P_{2})

    PgP_{g} is natural, i.e., for every diffeomorphism φ:MM\varphi:M\rightarrow M, φPg=Pφgφ\varphi^{\ast}P_{g}=P_{\varphi^{\ast}g}\circ\varphi^{\ast}, where “\!\!\!\phantom{K}{}^{\ast}\;” denotes the pullback of a tensor.

  3. (P3P_{3})

    PgP_{g} is self-adjoint with respect to the L2L^{2}-scalar product.

  4. (P4P_{4})

    PgP_{g} is conformally invariant, that is, given any function ω𝒞(M)\omega\in\mathcal{C}^{\infty}(M), if we define the conformal metric g~=e2ωg\widetilde{g}=e^{2\omega}g, then the following identity holds true:

    (1) Pg~(f)=eN+2m2ωPg(eN2m2ωf), for all f𝒞(M).P_{\widetilde{g}}(f)=e^{-\frac{N+2m}{2}\omega}P_{g}\left(e^{\frac{N-2m}{2}\omega}f\right),\text{ for all }f\in\mathcal{C}^{\infty}(M).

There is a natural conformal invariant associated with PgP_{g} given by

Qg:=2N2mPg(1),Q_{g}:=\frac{2}{N-2m}P_{g}(1),

called the Branson QQ-curvature, after Branson-Ørsted [BransonOrsted1991] and Branson [BransonBook, Branson1995], or simply the QQ-curvature. When m=1m=1, PgP_{g} is the conformal Laplacian and QgQ_{g} is the scalar curvature RgR_{g}, while for m=2m=2, PgP_{g} is the Paneitz-Branson operator and QgQ_{g} is the usual QQ-curvature [DjadliHebeyLedoux2000].

When considering conformal metrics g~=u4/(N2m)g\widetilde{g}=u^{4/(N-2m)}g with u>0u>0 in 𝒞(M)\mathcal{C}^{\infty}(M), equation (1) is written as

(2) Pg~ϕ=u12mPg(uϕ),P_{\widetilde{g}}\phi=u^{1-2_{m}^{\ast}}P_{g}(u\phi),

where 2m:=2NN2m2_{m}^{\ast}:=\frac{2N}{N-2m} is the critical Sobolev exponent of the embedding Hgm(M)Lp(M,g)H_{g}^{m}(M)\hookrightarrow L^{p}(M,g). Here Hgm(M)H_{g}^{m}(M) denotes the Sobolev space of order mm, which is the closure of 𝒞(M)\mathcal{C}^{\infty}(M) in Lg2(M)L^{2}_{g}(M) under the norm

(3) uHm:=(i=1mM|giu|2𝑑Vg)1/2.\|u\|_{H^{m}}:=\left(\sum_{i=1}^{m}\int_{M}|\nabla_{g}^{i}u|^{2}dV_{g}\right)^{1/2}.

Taking ϕ1\phi\equiv 1 in (2), one obtains the prescribed QQ-curvature equation

(4) Pgu=N2m2Qg~u2m1, on M.P_{g}u=\frac{N-2m}{2}Q_{\widetilde{g}}u^{2_{m}^{\ast}-1},\quad\text{ on }M.

For m=1m=1 and Qg~Q_{\widetilde{g}} constant, we recover the Yamabe Problem. Some results about the existence and multiplicity of solutions to the prescribed QQ-curvature problem can be found in [AzaizBoughazi2020, BaScWe, BenaliliBoughazi2016, DjadliHebeyLedoux2000, Ro, Tahri2020].

Given \ell\in\mathbb{N}, we will deal with the \ell-partition problem associated with (4) in the presence of symmetries. In order to describe the problem, let Γ\Gamma be a closed subgroup of Isom(M,g)(M,g) under suitable conditions (see hypotheses (Γ1\Gamma 1) and (Γ2\Gamma 2) below), and let ΩM\Omega\subset M be an open and Γ\Gamma-invariant subset, namely, if xΩx\in\Omega, then γxΩ\gamma x\in\Omega for every isometry γΓ\gamma\in\Gamma. In what follows, H0,gm(Ω)H_{0,g}^{m}(\Omega) denotes the closure of 𝒞c(Ω)\mathcal{C}_{c}^{\infty}(\Omega) under the Sobolev norm given by (3).

We consider the symmetric Dirichlet boundary problem

(5) {Pgu=|u|2m2u, in Ω,iu=0,i=0,1,,2m1, on Ω,u is Γ-invariant,\begin{cases}P_{g}u=|u|^{2_{m}^{*}-2}u,&\text{ in }\Omega,\\ \nabla^{i}u=0,i=0,1,\ldots,2m-1,&\text{ on }\partial\Omega,\\ u\text{ is }\Gamma\text{-invariant},\end{cases}

where u:Ωu:\Omega\rightarrow\mathbb{R} is said to be Γ\Gamma-invariant if u(x)=u(γx)u(x)=u(\gamma x) for every xMx\in M and any γΓ\gamma\in\Gamma.

Denote by cΩΓc_{\Omega}^{\Gamma} the least energy among the nontrivial solutions, that is,

cΩΓ:=inf{mNΩ|u|2mdVg:u0 and u solves (5)}.c^{\Gamma}_{\Omega}:=\inf\left\{\frac{m}{N}\int_{\Omega}|u|^{2^{\ast}_{m}}dV_{g}:u\neq 0\text{ and }u\text{ solves }\eqref{eq:dirichlet}\right\}.

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, Γ\Gamma-invariant and the Γ\Gamma-orbits have positive dimension, this number is attained (see Proposition 2.3 below). The \ell-partition problem consists in finding mutually disjoint and non empty Γ\Gamma-invariant open subsets Ω1,,Ω\Omega_{1},\ldots,\Omega_{\ell} such that

(6) i=1cΩiΓinf(Φ1,,Φ)𝒫Γi=1cΦiΓ\sum_{i=1}^{\ell}c_{\Omega_{i}}^{\Gamma}\leq\inf_{(\Phi_{1},\ldots,\Phi_{\ell})\in\mathcal{P}_{\ell}^{\Gamma}}\sum_{i=1}^{\ell}c_{\Phi_{i}}^{\Gamma}

where

𝒫Γ:={{Ω1,,Ω}:\displaystyle\mathcal{P}_{\ell}^{\Gamma}:=\{\{\Omega_{1},\ldots,\Omega_{\ell}\}: Ωi is a Γ-invariant open subset of M and ΩiΩj= if ij}.\displaystyle\,\Omega_{i}\neq\emptyset\text{ is a }\Gamma\text{-invariant open subset of }M\text{ and }\Omega_{i}\cap\Omega_{j}=\emptyset\text{ if }i\neq j\}.

The aim of this paper is to show that this problem has a solution on Einstein manifolds with positive scalar curvature, for every m2m\geq 2, and with less restrictive hypotheses on the metric, when m=1m=1.

In order to state our main result, we need to impose some conditions over (M,g)(M,g) and its isometry group (conditions (Γ1)(\Gamma 1) to (Γ3)(\Gamma 3) 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 xMx\in M, the Γ\Gamma orbit of xx is the set Γx:={γx:γΓ}\Gamma x:=\{\gamma x\;:\;\gamma\in\Gamma\}, and the isotropy subgroup of xx is defined as Γx:={γΓ:γx=x}\Gamma_{x}:=\{\gamma\in\Gamma\;:\;\gamma x=x\}. When Γ\Gamma is a closed subgroup of Isom(M,g)\text{Isom}(M,g), the Γ\Gamma-orbits are closed submanifolds of MM which are Γ\Gamma-diffeomorphic to the homogeneous space Γ/Γx\Gamma/\Gamma_{x}, meaning the existence of a Γ\Gamma-invariant diffeomorphism between these two manifolds. An orbit Γx\Gamma x is called principal, if there exists a neighborhood VV of xx in MM such that for each yVy\in V, ΓxΓγy\Gamma_{x}\subset\Gamma_{\gamma y} for some γΓ\gamma\in\Gamma. All points lying in a principal orbit have, up to conjugacy, the same isotropy group. Denote this group by KK, called the principal isotropy group. We say that Γ\Gamma induces a cohomogeneity one action on MM if the principal orbits have codimension one. In presence of a cohomogeneity one action, each principal Γ\Gamma-orbit is a closed hypersurface in MM diffeomorphic to Γ/K\Gamma/K, and there are exactly two orbits of bigger codimension, called singular orbits. Denote them by MM_{-} and M+M_{+}. M±M_{\pm} are closed submanifolds of codimension 2\geq 2, and all points lying in M±M_{\pm} have, up to conjugacy, the same isotropy group. Denoting these groups by K±K_{\pm}, we have that M±M_{\pm} is Γ\Gamma-diffeomorphic to Γ/K±\Gamma/K_{\pm}.

We can now state the conditions on the group Γ\Gamma. In what follows, dd will denote the geodesic distance between M+M_{+} and MM_{-}, that is, d:=distg(M,M+)d:=\text{dist}_{g}(M_{-},M_{+}).

  1. (Γ1)(\Gamma 1)

    Γ\Gamma is a closed subgroup of Isom(M,g)\text{Isom}(M,g) inducing a cohomogeneity one action on MM.

  2. (Γ2)(\Gamma 2)

    1dimΓxN11\leq\dim\Gamma x\leq N-1 for every xMx\in M.

  3. (Γ3)(\Gamma 3)

    The metric gg on MM is a Γ\Gamma-invariant metric of the form:

    g=dt2+j=1kfj2(t)gj,g=dt^{2}+\sum_{j=1}^{k}f_{j}^{2}(t)\ g_{j},

    where gt:=j=1kfj2(t)gjg_{t}:=\sum_{j=1}^{k}f_{j}^{2}(t)\ g_{j} is one parameter family of metrics on the principal orbit, for some positive smooth functions defined on the interval I=[0,d]I=[0,d], with suitable asymptotic conditions at 0 and dd. More precisely, the (smooth) metrics gjg_{j} are defined on the principal orbit at tt, for t(0,d)t\in(0,d), and are defined around the singular orbits in such a way that fjf_{j}, j=1,,kj=1,\dots,k, satisfy appropriate smoothness conditions at the endpoints t=0t_{-}=0 and t+=dt_{+}=d of II, that ensure that gg can be extended to the singular orbits:

    (7) f1(t±)=0,f1(t±)=1;fj(t±)>0,fj(t±)=0for 1<jk.f_{1}(t_{\pm})=0,f_{1}^{\prime}(t_{\pm})=1;\ f_{j}(t_{\pm})>0,f_{j}^{\prime}(t_{\pm})=0\ \mbox{for $1<j\leq k$}.

    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 \ell-optimal partition problem in terms of the principal orbits of cohomogeneity one actions. The symbol “\approx” will stand for “Γ\Gamma-diffeomorphic to”.

Theorem 1.1.

Let mm\in\mathbb{N} and let (M,g)(M,g) be a closed Riemannian manifold of dimension N>2mN>2m with scalar curvature RgR_{g}. If

  • Rg>0R_{g}>0 for m=1m=1; or

  • (M,g)(M,g) is Einstein with Rg>0R_{g}>0,

then for any \ell\in\mathbb{N} and any subgroup Γ\Gamma of Isom(M,g)(M,g) satisfying (Γ1\Gamma 1), (Γ2\Gamma 2) and (Γ3\Gamma 3), the Γ\Gamma-invariant \ell-partition problem (6) has a solution (Ω1,,Ω)𝒫ΩΓ(\Omega_{1},\ldots,\Omega_{\ell})\in\mathcal{P}_{\Omega}^{\Gamma} such that:

  1. (1)

    Ωi\Omega_{i} is connected for every i=1,,i=1,\ldots,\ell, Ω¯iΩ¯i+1\overline{\Omega}_{i}\cap\overline{\Omega}_{i+1}\neq\emptyset, ΩiΩj=\Omega_{i}\cap\Omega_{j}=\emptyset if |ij|2|i-j|\geq 2 and Ω1Ω¯=M\overline{\Omega_{1}\cup\ldots\cup\Omega_{\ell}}=M.

  2. (2)

    The sets Ω1\Omega_{1} and Ω\Omega_{\ell} are Γ\Gamma-diffeomorphic to disk bundles at MM_{-} and M+M_{+}, respectively. More precisely,

    Ω1Γ×KD,ΩΓ×K+D+,Ω1ΩΓ/K,\Omega_{1}\approx\Gamma\times_{K-}D_{-},\quad\Omega_{\ell}\approx\Gamma\times_{K+}D_{+},\quad\partial\Omega_{1}\approx\partial\Omega_{\ell}\approx\Gamma/K,

    where Γ×K±D±\Gamma\times_{K_{\pm}}D_{\pm} are normal bundles over the singular orbits M±M_{\pm}. See Section 5 for details.

  3. (3)

    For each i1,i\neq 1,\ell, ΩiΓ/K×(0,1)\Omega_{i}\approx\Gamma/K\times(0,1), Ω¯iΩ¯i+1Γ/K\overline{\Omega}_{i}\cap\overline{\Omega}_{i+1}\approx\Gamma/K and ΩiΓ/KΓ/K\partial\Omega_{i}\approx\Gamma/K\sqcup\Gamma/K, where \sqcup denotes the disjoint union of sets.

In Section 6 we give several examples of Einstein manifolds with positive scalar curvature satisfying conditions (Γ1\Gamma 1), (Γ2\Gamma 2) and (Γ3\Gamma 3). 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 O(n)×O(k)O(n)\times O(k)-action, n+k=N+1n+k=N+1. In the case of m=1m=1, M. Clapp and A. Pistoia [ClappPistoia21] solved the Γ\Gamma-invariant \ell-partition problem for actions with higher cohomogeneity and in [ClappPistoiaTavares21], the authors solved the \ell-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 m=1m=1, the scalar curvature is Γ\Gamma-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 (M,g)(M,g) be a closed Riemannian manifold of dimension N3N\geq 3 and let Γ\Gamma be a closed subgroup of Isom(M,g)\text{Isom}(M,g) satisfying (Γ1)(\Gamma 1) to (Γ3)(\Gamma 3). If Rg>0R_{g}>0, then for any \ell\in\mathbb{N}, the Yamabe problem

Δgu+N24(N1)Rgu=|u|212u,on M,-\Delta_{g}u+\frac{N-2}{4(N-1)}R_{g}u=|u|^{2_{1}^{\ast}-2}u,\quad\text{on }M,

admits a Γ\Gamma-invariant sign changing solution with exactly \ell-nodal domains. Moreover, it has least energy among all such solutions and the nodal set is a disjoint union of 1\ell-1 principal orbits.

This was proven in the case =2\ell=2 in [ClappPistoia21] for general manifolds in the presence of symmetries, and in [ClappPistoiaTavares21] in the non-symmetric case; for any \ell\in\mathbb{N}, 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 MM is a Riemannian metric gg satisfying the equation

(8) Ric(g)+Hessg(f)=μg.Ric(g)+{\rm Hess}_{g}(f)=\mu g.

for some constant μ=1,0,\mu=-1,0, or 11, and some smooth function ff, called the Ricci potential. Here, as usual, Ric(g)Ric(g) denotes the Ricci curvature of gg, and Hessg(f){\rm Hess}_{g}(f) denotes the Hessian of ff with respect to gg. A Ricci soliton is called steady, shrinking or expanding according to μ=0\mu=0, μ>0\mu>0 or μ<0\mu<0, 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 2#2¯\mathbb{CP}^{2}\#\overline{\mathbb{CP}^{2}}, with respect to the action of U(2)(2), the unitary group of dimension 22. Here, as usual, #\# denotes the connected sum of smooth manifolds and M¯\overline{M} denotes a smooth manifold MM taken with reverse orientation. It is also known that it is non-Einstein with positive Ricci curvature. The singular orbits of the U(2)(2)-action are both diffeomorphic to 𝕊2\mathbb{S}^{2}, and the principal orbits are diffeomorphic to 𝕊3\mathbb{S}^{3}. The associated fibrations, as a cohomogeneity one manifold, are both given by the Hopf fibration:

U(1)SU(2)SU(2)/U(1).{\rm U}(1)\to{\rm SU}(2)\to{\rm SU}(2)/{\rm U}(1).

Here, for an orientable manifold MM, we will denote by M¯\overline{M} 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 Γ\Gamma-invariant solutions to the following weakly coupled competitive QQ-curvature system

(9) Pgui=νi|ui|2m2ui+ijηijβij|uj|αij|ui|βij2ui, on M,i,j=1,,,P_{g}u_{i}=\nu_{i}|u_{i}|^{2_{m}^{\ast}-2}u_{i}+\sum_{i\neq j}\eta_{ij}\beta_{ij}|u_{j}|^{\alpha_{ij}}|u_{i}|^{\beta_{ij}-2}u_{i},\quad\text{ on }M,\ \ i,j=1,\ldots,\ell,

where νi>0\nu_{i}>0, ηij=ηji<0\eta_{ij}=\eta_{ji}<0 and αij,βij>1\alpha_{ij},\beta_{ij}>1 are such that αij=βji\alpha_{ij}=\beta_{ji} and αij+βij=2m\alpha_{ij}+\beta_{ij}=2_{m}^{\ast}. We will say that a solution u¯=(u1,,u)\overline{u}=(u_{1},\ldots,u_{\ell}) to the system (9) is fully nontrivial, if, for each i=1,,i=1,\ldots,\ell, uiu_{i} is non trivial.

To assure the existence of a fully nontrivial least energy Γ\Gamma-invariant solution to the system (9) we will only assume that Γ\Gamma satisfies (Γ2\Gamma 2), allowing actions with bigger codimension of its principal orbits, and also we will assume that the operator PgP_{g} is coercive, meaning the existence of a constant C>0C>0 such that

MuPgu𝑑VgCuHm2, for every u𝒞(M).\int_{M}uP_{g}u\;dV_{g}\geq C\|u\|_{H^{m}}^{2},\qquad\text{ for every }u\in\mathcal{C}^{\infty}(M).

We will say that a sequence of fully nontrivial elements u¯k\overline{u}_{k} in the Sobolev space

Hgm(M):=Hgm(M)××Hgm(M) timesH_{g}^{m}(M)^{\ell}:=\underbrace{H_{g}^{m}(M)\times\cdots\times H_{g}^{m}(M)}_{\ell\text{ times}}

is unbounded if uk,iHm,\|u_{k,i}\|_{H^{m}}\rightarrow\infty, as kk\rightarrow\infty, for every i=1,,.i=1,\ldots,\ell. In this direction, we state our next multiplicity result.

Theorem 1.3.

If the operator PgP_{g} is coercive and (Γ2\Gamma 2) holds true, then the system (9) admits an unbounded sequence of Γ\Gamma-invariant fully nontrivial solutions. One of them has least energy among all fully nontrivial Γ\Gamma-invariant solutions.

When =1\ell=1 and m=1m=1, 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 Γ\Gamma-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 (Γ2\Gamma 2), avoids these problems in higher dimensions. For m2m\geq 2, the Maximum Principle for the operator PgP_{g} 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 m=2m=2 and =1\ell=1, by a recent result by J. Vétois [Vetois2022, Theorem 2.2], we can assure that an infinite number of the solutions to

(10) Pgu=|u|224uin M,P_{g}u=|u|^{2_{2}^{\ast}-4}u\qquad\text{in }M,

must change sign when considering (M,g)(M,g) to be Einstein with positive scalar curvature, as we next state.

Corollary 1.4.

Let m=2m=2, (M,g)(M,g) be an Einstein manifold of dimension N>4N>4, with positive scalar curvature and not conformally diffeomorphic to the standard sphere. If Γ\Gamma is a closed subgroup of Isom(M,g)\text{Isom}(M,g) satisfying (Γ2\Gamma 2), then the problem with Paneitz-Branson operator (10) admits an unbounded sequence of Γ\Gamma-invariant sign-changing solutions.

There are many examples of manifolds admitting an Einstein metric gg with positive scalar curvature, in order to ensure the coercivity of PgP_{g}. 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 PgP_{g} 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 m=2m=2 and N6N\geq 6, if Qg>0Q_{g}>0 and Rg>0R_{g}>0, then PgP_{g} is coercive.

Proof.

It is a direct consequence of a calculation obtained in [GurskyMal03]. It appears as equation (2.18):

MϕPgϕ=M[N6N2(Δϕ)2+4|Hess(ϕ)|2N2+(N2)2+42(N1)(N2)Rg|ϕ|2+N42Qgϕ2]𝑑vol(g).\begin{split}&\int_{M}\phi P_{g}\phi\\ &=\int_{M}\left[\frac{N-6}{N-2}(\Delta\phi)^{2}+\frac{4|{\rm Hess}(\phi)|^{2}}{N-2}+\frac{(N-2)^{2}+4}{2(N-1)(N-2)}R_{g}|\nabla\phi|^{2}+\frac{N-4}{2}Q_{g}\phi^{2}\right]d{\rm vol}(g).\end{split}

Again, it is complicated to compute the QQ-curvature of an arbitrary manifold and the literature lacks examples, different from Einstein manifolds. We explore the possibility of obtaining positive QQ-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 QQ-curvature of these metrics.

Theorem 1.6.

The QQ-curvature of a shrinking Ricci soliton (M,g)(M,g), with Ricci potential ff, is given by:

Qg=(2𝐚𝐜)|Ricg|g2+𝐛Rg22𝐚R2𝐚Ricg(f,f).Q_{g}=(2{\mathbf{a}}-{\mathbf{c}})|Ric_{g}|_{g}^{2}+{\mathbf{b}}R_{g}^{2}-2{\mathbf{a}}R-2{\mathbf{a}}\ Ric_{g}(\nabla f,\nabla f).

where

𝐚=12(N1),𝐛=N34N2+16N168(N1)2(N2)2,𝐜=2(N2)2.{\mathbf{a}}=\frac{1}{2(N-1)},\quad{\mathbf{b}}=\frac{N^{3}-4N^{2}+16N-16}{8(N-1)^{2}(N-2)^{2}},\quad{\mathbf{c}}=\frac{2}{(N-2)^{2}}.

If (M,g)(M,g) has radially positive Ricci curvature, i.e., Ricg(f,f)>0Ric_{g}(\nabla f,\nabla f)>0, and

  • for N=4N=4, R2>|Ricg|g2+2Rg+2Ricg(f,f)R^{2}>|Ric_{g}|_{g}^{2}+2R_{g}+2Ric_{g}(\nabla f,\nabla f),

  • or for N>4N>4, (2𝐚𝐜)|Ricg|g2>2𝐚Ricg(f,f)>0(2{\mathbf{a}}-{\mathbf{c}})|Ric_{g}|_{g}^{2}>2{\mathbf{a}}\ Ric_{g}(\nabla f,\nabla f)>0 and Rg>2𝐚/𝐛R_{g}>2{\mathbf{a}}/{\mathbf{b}},

then Qg>0Q_{g}>0.

Remark 1.1.

The condition of having radially positive Ricci curvature is plausible, due to the fact that if Ricg(f,f)0Ric_{g}(\nabla f,\nabla f)\leq 0 everywhere, then the Ricci soliton is trivial [PetWyl09a, Theorem 1.1]. Moreover, if it is radially flat, i.e., Ricg(f,f)0Ric_{g}(\nabla f,\nabla f)\equiv 0, then it is also trivial [PetWyl09b, Proposition 2]). If (M,g)(M,g) is a cohomogeneity one Kähler-Ricci soliton under the action of Γ\Gamma, then one may average over Γ\Gamma to obtain a Ricci potential that is a Γ\Gamma-invariant function on MM. Then the following identity holds true:

Ricg(f,f)=Ricg(ft,ft)=(f)2Ricg(t,t).Ric_{g}(\nabla f,\nabla f)=Ric_{g}\left(f^{\prime}\frac{\partial}{\partial t},f^{\prime}\frac{\partial}{\partial t}\right)=(f^{\prime})^{2}Ric_{g}\left(\frac{\partial}{\partial t},\frac{\partial}{\partial t}\right).

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]. ∎


We have the following immediate consequence of Theorems 1.3 and 1.6, together with Proposition 1.5.

Corollary 1.7.

Let (M,g)(M,g) be Einstein with Rg>0R_{g}>0 or a shrinking Ricci soliton with radially positive Ricci curvature of dimension N>2mN>2m. If (Γ2\Gamma 2) holds true, then the system (9) admits an unbounded sequence of fully nontrivial Γ\Gamma-invariant solutions, one of them with least energy among all fully nontrivial Γ\Gamma-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 Γ\Gamma-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 (Γ1)(\Gamma 1) and (Γ3)(\Gamma 3) 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 Γ\Gamma-invariant functions that solve the optimal partition problem, showing Theorem 1.1. As an application of Theorem 1.1 with m=1m=1, we prove Corollary 1.2. We give several examples where hypotheses (Γ1)(\Gamma 1) to (Γ3)(\Gamma 3) hold true in Section 6. Finally, in Section LABEL:Section:Q-curvatureRicciSolitons, we compute the QQ-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, Ω\Omega will denote either MM or an open, connected Γ\Gamma-invariant subset of MM with smooth boundary, p\|\cdot\|_{p} will denote the usual norm in Lgp(Ω)L_{g}^{p}(\Omega), p1p\geq 1. For u𝒞(M)u\in\mathcal{C}^{\infty}(M), the kk-th covariant derivative of uu will be denoted by ku\nabla^{k}u and we define its norm as the function |ku|g:M|\nabla^{k}u|_{g}:M\rightarrow\mathbb{R} given by

|ku|g2:=α1αkα1αku,|\nabla^{k}u|_{g}^{2}:=\nabla^{\alpha_{1}}\cdots\nabla^{\alpha_{k}}\nabla_{\alpha_{1}}\cdots\nabla_{\alpha_{k}}u,

where we used the Einstein notation convention.

The Sobolev space H0,gm(Ω)H_{0,g}^{m}(\Omega) is the closure of 𝒞c(Ω)\mathcal{C}_{c}^{\infty}(\Omega) under the norm

uHm:=(i=0miu22)1/2=(i=0mΩ|iu|g2𝑑Vg)1/2,\|u\|_{H^{m}}:=\left(\sum_{i=0}^{m}\|\nabla^{i}u\|_{2}^{2}\right)^{1/2}=\left(\sum_{i=0}^{m}\int_{\Omega}|\nabla^{i}u|_{g}^{2}dV_{g}\right)^{1/2},

where, with some abuse of notation, iu2:=|iu|g2\|\nabla^{i}u\|_{2}:=\|\,|\nabla^{i}u|_{g}\|_{2} Notice that in case Ω=M\Omega=M, then 𝒞c(M)=𝒞(M)\mathcal{C}_{c}^{\infty}(M)=\mathcal{C}^{\infty}(M) and H0,gm(M)=Hgm(M)H_{0,g}^{m}(M)=H_{g}^{m}(M); if ΩM\Omega\neq M, then H0,gm(Ω)H_{0,g}^{m}(\Omega) is a closed subspace of Hgm(Ω)H^{m}_{g}(\Omega).

Let PgP_{g} be the corresponding GJMS-operator in (M,g)(M,g) of order mm. For each k{0,1,,m1}k\in\{0,1,\ldots,m-1\}, there exists a symmetric T2k0T^{0}_{2k}-tensor field on MM, which we will denote by A(k)(g)A_{(k)}(g), such that the operator PgP_{g} can be written as

Pg=(Δ)gm+k=0m1(1)kjkj1(A(k)(g)iki1j1jki1ik),P_{g}=(-\Delta)^{m}_{g}+\sum_{k=0}^{m-1}(-1)^{k}\nabla^{j_{k}\cdots j_{1}}\left(A_{(k)}(g)_{i_{k}\cdots i_{1}j_{1}\cdots j_{k}}\nabla^{i_{1}\cdots i_{k}}\right),

where the indices are raised via the musical isomorphism. In particular, for any u,v𝒞c(Ω)u,v\in\mathcal{C}_{c}^{\infty}(\Omega), integration by parts yields that

ΩuPgv𝑑Vg={Ω[Δgm/2uΔm/2v+k=0m1Ak(g)(gku,gkv)]𝑑Vg,m even,Ω[gΔg(m1)/2u,gΔg(m1)/2vg+k=0m1Ak(g)(gku,gkv)]𝑑Vg,m odd.\begin{split}&\int_{\Omega}uP_{g}vdV_{g}\\ &=\begin{cases}\int_{\Omega}\left[\Delta_{g}^{m/2}u\Delta^{m/2}v+\sum_{k=0}^{m-1}A_{k}(g)(\nabla_{g}^{k}u,\nabla_{g}^{k}v)\right]dV_{g},&m\text{ even},\\ \int_{\Omega}\left[\langle\nabla_{g}\Delta_{g}^{(m-1)/2}u,\nabla_{g}\Delta_{g}^{(m-1)/2}v\rangle_{g}+\sum_{k=0}^{m-1}A_{k}(g)(\nabla_{g}^{k}u,\nabla_{g}^{k}v)\right]dV_{g},&m\text{ odd}.\end{cases}\end{split}

See [Ro, Proposition 1] for the details.

As a consequence, the bilinear form (u,v)ΩuPgv𝑑Vg(u,v)\mapsto\int_{\Omega}uP_{g}vdV_{g} can be extended to a continuous symmetric bilinear form on H0,gm(Ω)H^{m}_{0,g}(\Omega). When PgP_{g} is coercive, this bilinear form is actually a well-defined interior product on H0,gm(Ω)H_{0,g}^{m}(\Omega) that induces a norm equivalent to Hm\|\cdot\|_{H^{m}} (see [Ro, Proposition 2]). We will denote this interior product and norm by ,Pg\langle\cdot,\cdot\rangle_{P_{g}} and Pg\|\cdot\|_{P_{g}} respectively, and endow Hgm(Ω)H_{g}^{m}(\Omega) with it in what follows. Notice that, by definition,

u,vPg:=ΩuPgv𝑑Vg,anduPg2=ΩuPgu𝑑Vg,\langle u,v\rangle_{P_{g}}:=\int_{\Omega}uP_{g}vdV_{g},\quad\text{and}\quad\|u\|^{2}_{P_{g}}=\int_{\Omega}uP_{g}udV_{g},

for every u,v𝒞(Ω)u,v\in\mathcal{C}^{\infty}(\Omega).

The group Isom(M,g)\text{Isom}(M,g) acts on Hgm(M)H_{g}^{m}(M) in the usual way:

γu:=uγ1,uHgm(M),γIsom(M,g).\gamma u:=u\circ\gamma^{-1},\qquad u\in H_{g}^{m}(M),\ \gamma\in\text{Isom}(M,g).

Every element γIsom(M,g)\gamma\in\text{Isom}(M,g) induces a linear map

γ:Hgm(M)Hgm(M),uuγ1.\gamma:H_{g}^{m}(M)\rightarrow H_{g}^{m}(M),\quad u\mapsto u\circ\gamma^{-1}.

We next show that the norm is invariant under the action of isometries.

Lemma 2.1.

For every γIsom(M,g)\gamma\in\text{Isom}(M,g) and every u𝒞(M)u\in\mathcal{C}^{\infty}(M)

Pg(uγ)=(Pgu)γ.P_{g}(u\circ\gamma)=(P_{g}u)\circ\gamma.

In particular, γ:Hgm(M)Hgm(M)\gamma:H_{g}^{m}(M)\rightarrow H_{g}^{m}(M) is a linear isometry.

Proof.

Let u𝒞(M)u\in\mathcal{C}^{\infty}(M) and γIsom(M,g)\gamma\in\text{Isom}(M,g). Then γ:MM\gamma:M\rightarrow M is a diffeomorphism and, by the naturality of PgP_{g}, property (P2P_{2}) in the introduction, we obtain for any isometry γIsomg(M,g)\gamma\in\text{Isom}_{g}(M,g) that

Pg(uγ)=(Pgγ)(u)=Pγgγ(u)=(γPg)(u)=Pg(u)γ,P_{g}(u\circ\gamma)=(P_{g}\circ\gamma^{\ast})(u)=P_{\gamma^{\ast}g}\circ\gamma^{\ast}(u)=(\gamma^{\ast}\circ P_{g})(u)=P_{g}(u)\circ\gamma,

where γg\gamma^{\ast}g denotes the pullback metric.

Now, to see that γ\gamma induces a linear isometry, recall that if γIsom(M,g)\gamma\in\text{Isom}(M,g), then

(11) Muγ𝑑Vg=Mu𝑑Vg,\int_{M}u\circ\gamma\;dV_{g}=\int_{M}u\;dV_{g},

(Cf. [HebeyBook1997, Théorème 4.1.2]). Then, for every u𝒞(M)u\in\mathcal{C}^{\infty}(M),

γuPg2=M(uγ1)Pg(uγ1)𝑑Vg=M(uγ1)Pg(u)γ1𝑑Vg=MuPgu𝑑Vg=uPg2.\|\gamma u\|_{P_{g}}^{2}=\int_{M}(u\circ\gamma^{-1})P_{g}(u\circ\gamma^{-1})dV_{g}=\int_{M}(u\circ\gamma^{-1})P_{g}(u)\circ\gamma^{-1}\,dV_{g}=\int_{M}uP_{g}u\,dV_{g}=\|u\|^{2}_{P_{g}}.

Density of 𝒞(M)\mathcal{C}^{\infty}(M) in Hgm(M)H_{g}^{m}(M) yields the result. ∎


Every γIsom(M,g)\gamma\in\text{Isom}(M,g) also induces a linear map γ:Lgp(M)Lgp(M)\gamma:L^{p}_{g}(M)\rightarrow L^{p}_{g}(M), p1p\geq 1, given by uuγ1u\mapsto u\circ\gamma^{-1}, which is also an isometry, thanks to (11).


Let Γ\Gamma be any closed subgroup of Isom(M,g)(M,g) such that dimΓxN1\dim\Gamma x\leq N-1 for any xMx\in M. From now on, suppose that Ω¯\overline{\Omega} is Γ\Gamma-invariant, namely, if xΩ¯x\in\overline{\Omega}, then ΓxΩ¯\Gamma x\subset\overline{\Omega}. In this way, for any u𝒞c(Ω)u\in\mathcal{C}_{c}^{\infty}(\Omega) and every isometry γΓ\gamma\in\Gamma, it follows that

Ωuγ𝑑Vg=Ωu𝑑Vg\int_{\Omega}u\circ\gamma\;dV_{g}=\int_{\Omega}u\;dV_{g}

and γ\gamma also induces linear isometries

(12) γ:H0,gm(Ω)H0,gm(Ω), and γ:Lgp(Ω)Lgp(Ω).\gamma:H_{0,g}^{m}(\Omega)\rightarrow H_{0,g}^{m}(\Omega),\quad\text{ and }\quad\gamma:L_{g}^{p}(\Omega)\rightarrow L_{g}^{p}(\Omega).

for any p1p\geq 1.

We define the Sobolev space of Γ\Gamma-invariant functions as

H0,gm(Ω)Γ:={uH0,gm(Ω):u is Γ-invariant}.H_{0,g}^{m}(\Omega)^{\Gamma}:=\{u\in H_{0,g}^{m}(\Omega)\;:\;u\text{ is }\Gamma\text{-invariant}\}.

This is a closed subspace of H0,gm(Ω)H_{0,g}^{m}(\Omega). In fact, if 𝒞c(Ω)Γ\mathcal{C}_{c}^{\infty}(\Omega)^{\Gamma} denotes the space of smooth functions with compact support in Ω\Omega which are Γ\Gamma-invariant, then H0,gm(Ω)H_{0,g}^{m}(\Omega) coincides with the closure of this space under the Sobolev norm Pg\|\cdot\|_{P_{g}}. As the dimension of any Γ\Gamma orbit is strictly less than NN, the space H0,gm(Ω)ΓH_{0,g}^{m}(\Omega)^{\Gamma} is infinite dimensional, thanks to the existence of Γ\Gamma-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 Γ\Gamma be a closed subgroup of Isom(M,g)\text{Isom}(M,g), κ:=min{dimΓx:xM}\kappa:=\min\{\dim\Gamma x\;:\;x\in M\}, and let Ω\Omega be a Γ\Gamma-invariant domain. Define

2m,Γ:={2(Nκ)(Nκ)2m,Nκ>2m,,Nκ2m.2^{\ast}_{m,\Gamma}:=\begin{cases}\frac{2(N-\kappa)}{(N-\kappa)-2m},&N-\kappa>2m,\\ \infty,&N-\kappa\leq 2m.\end{cases}

Then

H0,gm(Ω)ΓLgr(Ω)H_{0,g}^{m}(\Omega)^{\Gamma}\hookrightarrow L_{g}^{r}(\Omega)

is continuous and compact for every 1r<2m,Γ1\leq r<2^{\ast}_{m,\Gamma}.

Proof.

The case for m=1m=1 is just Theorem 2.4 in [IvanovNazarov2007].The case for m>1m>1 follows from a bootstrap argument as in Proposition 2.11 in [AubinBook]. ∎

In what follows, we will suppose that Γ\Gamma satisfies condition (Γ2\Gamma 2). Under this hypothesis, the existence of least energy Γ\Gamma-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 Γ\Gamma-invariant solutions to the problem

(13) {Pgu=|u|p2u, in Ω,uH0,gm(Ω)Γ,\begin{cases}P_{g}u=|u|^{p-2}u,&\text{ in }\Omega,\\ u\in H_{0,g}^{m}(\Omega)^{\Gamma},\end{cases}

where p(2,2m]p\in(2,2_{m}^{\ast}].

Consider the functional

JΩ:H0,gm(Ω),JΩ(u):=12uPg21pΩ|u|p𝑑Vg.J_{\Omega}:H_{0,g}^{m}(\Omega)\rightarrow\mathbb{R},\qquad J_{\Omega}(u):=\frac{1}{2}\|u\|_{P_{g}}^{2}-\frac{1}{p}\int_{\Omega}|u|^{p}\;dV_{g}.

Given that Pg\|\cdot\|_{P_{g}} is a well defined norm equivalent to the standard norm Hm(Ω)\|\cdot\|_{H^{m}(\Omega)}, Sobolev inequality yields that it is a C1C^{1} functional for any p(2,2m]p\in(2,2_{m}^{\ast}]. From (12), this functional is Γ\Gamma-invariant, namely it satisfies that

JΩ(uγ1)=JΩ(u).J_{\Omega}(u\circ\gamma^{-1})=J_{\Omega}(u).

Hence, due to Palais’ Principle of Symmetric Criticality [Palais1979], the critical points of JΩJ_{\Omega} restricted to H0,gm(Ω)ΓH_{0,g}^{m}(\Omega)^{\Gamma} correspond to the Γ\Gamma-invariant solutions to the problem (13). The nontrivial ones belong to the set

ΩΓ:={uH0,gm(Ω):u0,JΩ(u)u=0},\mathcal{M}_{\Omega}^{\Gamma}:=\{u\in H_{0,g}^{m}(\Omega)\;:\;u\neq 0,J^{\prime}_{\Omega}(u)u=0\},

which is a C1C^{1} codimension one Hilbert manifold in H0,gm(Ω)ΓH_{0,g}^{m}(\Omega)^{\Gamma}. Notice that

JΩ(u)=mNuPg,uΩΓ.J_{\Omega}(u)=\frac{m}{N}\|u\|_{P_{g}},\qquad u\in\mathcal{M}_{\Omega}^{\Gamma}.

Thanks to the Sobolev inequalities [AubinBook, Theorem 2.30], ΩΓ\mathcal{M}_{\Omega}^{\Gamma} is closed and

0<cΩΓ=infuΩΓJΩ(u).0<c_{\Omega}^{\Gamma}=\inf_{u\in\mathcal{M}_{\Omega}^{\Gamma}}J_{\Omega}(u).

We say that JΩJ_{\Omega} satisfies condition (PS)cΓ(PS)_{c}^{\Gamma} in H0,gm(Ω)ΓH_{0,g}^{m}(\Omega)^{\Gamma} if every sequence ukH0,gm(Ω)Γu_{k}\in H_{0,g}^{m}(\Omega)^{\Gamma} such that JΩ(uk)cJ_{\Omega}(u_{k})\rightarrow c and JΩ(uk)0\nabla J_{\Omega}(u_{k})\rightarrow 0 in H0,gm(Ω)ΓH_{0,g}^{m}(\Omega)^{\Gamma} as kk\rightarrow\infty, has a convergent subsequence.

As Γ\Gamma satisfies (Γ2\Gamma 2), then κ1\kappa\geq 1 and p2m<2m,Γp\leq 2_{m}^{\ast}<2_{m,\Gamma}^{\ast}. Hence, by Lemma 2.2, JΩJ_{\Omega} satisfies condition (PS)cΩΓΓ(PS)_{c_{\Omega}^{\Gamma}}^{\Gamma} and Theorem 7.12 in [AmbrosettiMalchiodiBook] yields that cΩΓc_{\Omega}^{\Gamma} is attained. Thus, there exists a least energy Γ\Gamma-invariant solution to the problem (13).

We summarize this analysis in the following proposition.

Proposition 2.3.

If Γ\Gamma satisfies (Γ2\Gamma 2) and if PgP_{g} is coercive, then, for any p(2,2m]p\in(2,2_{m}^{\ast}] the problem (13) admits a least energy Γ\Gamma-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 GJMSGJMS-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

P^g={j=0m/2ai(Δ)j,m even,j=0(m1)/2ai(Δ)j,m odd,\widehat{P}_{g}=\begin{cases}\sum_{j=0}^{m/2}a_{i}(-\Delta)^{j},&m\text{ even},\\ \sum_{j=0}^{(m-1)/2}a_{i}(-\Delta)^{j},&m\text{ odd},\end{cases}

where a0𝒞(M)Γa_{0}\in\mathcal{C}^{\infty}(M)^{\Gamma} is positive, am/2>0a_{m/2}>0 and aj0a_{j}\geq 0 are constants for j=1,,m1j=1,\ldots,m-1 if mm is even, and a(m1)/2>0a_{(m-1)/2}>0 and aj0a_{j}\geq 0 are constants for j=1,,(m3)/2j=1,\ldots,(m-3)/2 if mm is odd. This is true because the Principle of Symmetric Criticality can be applied by noticing that

Δi+1(uγ)=Δi+1(u)γ\Delta^{i+1}(u\circ\gamma)=\Delta^{i+1}(u)\circ\gamma

for any u𝒞c(Ω)Γu\in\mathcal{C}_{c}^{\infty}(\Omega)^{\Gamma} and every isometry γΓ\gamma\in\Gamma (see, for instance, [AmannEscherBook, Remark 6.9 (c)]).∎

3. The polyharmonic system

We next study the system (9). Fix \ell\in\mathbb{N} and consider the product space (Hgm(M))\left(H_{g}^{m}(M)\right)^{\ell} endowed with the norm

u¯:=(u1,,u):=(i=1uiPg2)1/2.\|\overline{u}\|:=\|(u_{1},\ldots,u_{\ell})\|:=\Big{(}\sum_{i=1}^{\ell}\|u_{i}\|_{P_{g}}^{2}\Big{)}^{1/2}.

Let 𝒥:(Hgm(M))\mathcal{J}:\left(H_{g}^{m}(M)\right)^{\ell}\rightarrow\mathbb{R} be the functional given by

𝒥(u¯):=12i=1uiPg212mi=1Mνi|ui|2m12i,j=1jiMηij|uj|αij|ui|βij.\mathcal{J}(\overline{u}):=\frac{1}{2}\sum_{i=1}^{\ell}\|u_{i}\|_{P_{g}}^{2}-\frac{1}{2^{*}_{m}}\sum_{i=1}^{\ell}\int_{M}\nu_{i}|u_{i}|^{2^{*}_{m}}-\frac{1}{2}\sum_{\begin{subarray}{c}i,j=1\\ j\neq i\end{subarray}}^{\ell}\int_{M}\eta_{ij}|u_{j}|^{\alpha_{ij}}|u_{i}|^{\beta_{ij}}.

This is a C1C^{1} functional and its partial derivatives are given by

i𝒥(u¯)vi=ui,viPgMνi|ui|2m2uivij=1jiMηijβij|uj|αij|ui|βij2uivi,\displaystyle\partial_{i}\mathcal{J}(\overline{u})v_{i}=\langle u_{i},v_{i}\rangle_{P_{g}}-\int_{M}\nu_{i}|u_{i}|^{2_{m}^{*}-2}u_{i}v_{i}-\sum_{\begin{subarray}{c}j=1\\ j\neq i\end{subarray}}^{\ell}\int_{M}\eta_{ij}\beta_{ij}|u_{j}|^{\alpha_{ij}}|u_{i}|^{\beta_{ij}-2}u_{i}v_{i},

for every v¯(Hgm(M))\overline{v}\in\left(H_{g}^{m}(M)\right)^{\ell} and every i=1,,.i=1,\ldots,\ell. Hence, solutions to the system (9) correspond to the critical points of 𝒥.\mathcal{J}.

Fix now a closed subgroup Γ\Gamma of isometries satisfying (Γ2\Gamma 2) and define :=(Hgm(M)Γ)\mathcal{H}:=(H^{m}_{g}(M)^{\Gamma})^{\ell}. This is a closed subspace of (Hg1(M))\left(H_{g}^{1}(M)\right)^{\ell}. By Lemma 2.1, 𝒥\mathcal{J} is Γ\Gamma-invariant and by the Principle of Symmetric Criticality [Palais1979], the critical points of 𝒥\mathcal{J} restricted to \mathcal{H} are the Γ\Gamma-invariant solutions to the system (9). Hence, we can restrict ourselves to seek critical points of 𝒥\mathcal{J} in \mathcal{H}. Observe that the fully nontrivial ones belong to the set

𝒩:={u¯:ui0,i𝒥(u¯)ui=0, for each i=1,,}.\mathcal{N}:=\{\overline{u}\in\mathcal{H}\;:\;u_{i}\neq 0,\ \partial_{i}\mathcal{J}(\overline{u})u_{i}=0,\text{ for each }i=1,\ldots,\ell\}.

It is readily seen that

(14) 𝒥(u¯)=mNu¯2,if u¯𝒩.\mathcal{J}(\overline{u})=\frac{m}{N}\|\overline{u}\|^{2},\qquad\text{if \ }\overline{u}\in\mathcal{N}.
Lemma 3.1.

There exists d0>0d_{0}>0, independent of ηij\eta_{ij}, such that mini=1,,uid0\min_{i=1,\ldots,\ell}\|u_{i}\|\geq d_{0} if u¯=(u1,,u)𝒩\overline{u}=(u_{1},\ldots,u_{\ell})\in\mathcal{N}. Thus, 𝒩\mathcal{N} is a closed subset of \mathcal{H} and inf𝒩𝒥>0\inf_{\mathcal{N}}\mathcal{J}>0.

Proof.

Since ηij<0\eta_{ij}<0 and as the norm Pg\|\cdot\|_{P_{g}} is equivalent to the standard norm in Hgm(M)ΓH_{g}^{m}(M)^{\Gamma}, for any u¯𝒩\overline{u}\in\mathcal{N}, it follows from the Sobolev inequality the existence of a constant C>0C>0 such that

uiPg2Mνi|ui|2mCuiPg2m for u¯𝒩,i=1,,.\displaystyle\|u_{i}\|_{P_{g}}^{2}\leq\int_{M}\nu_{i}|u_{i}|^{2_{m}^{*}}\leq C\|u_{i}\|_{P_{g}}^{2_{m}^{*}}\quad\text{ for \ }\overline{u}\in\mathcal{N},\ i=1,\ldots,\ell.

The result follows from this inequality. ∎

A fully nontrivial solution u¯\overline{u} to the system (9) satisfying 𝒥(u¯)=inf𝒩𝒥\mathcal{J}(\overline{u})=\inf_{\mathcal{N}}\mathcal{J} is called a Γ\Gamma-invariant least energy solution. To establish the existence of fully nontrivial critical points of 𝒥\mathcal{J}, 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 u¯=(u1,,u)\overline{u}=(u_{1},\ldots,u_{\ell}) and s¯=(s1,,s)(0,)\overline{s}=(s_{1},\ldots,s_{\ell})\in(0,\infty)^{\ell}, we write

s¯u¯:=(s1u1,,su).\overline{s}\,\overline{u}:=(s_{1}u_{1},\ldots,s_{\ell}u_{\ell}).

Let 𝒮:={uHgm(M)Γ:u=1}\mathcal{S}:=\{u\in H_{g}^{m}(M)^{\Gamma}:\|u\|=1\}, define 𝒯:=𝒮\mathcal{T}:=\mathcal{S}^{\ell} and

𝒰:={u¯𝒯:s¯u¯𝒩 for some s¯(0,)}.\mathcal{U}:=\{\overline{u}\in\mathcal{T}:\overline{s}\,\overline{u}\in\mathcal{N}\text{ \ for some \ }\overline{s}\in(0,\infty)^{\ell}\}.

The next result is proved exactly in the same way as [ClappSzulkin19, Proposition 3.1].

Lemma 3.2.
  • (i)(i)

    Let u¯𝒯\overline{u}\in\mathcal{T}. If there exists s¯u¯(0,)\overline{s}_{\overline{u}}\in(0,\infty)^{\ell} such that s¯u¯u¯𝒩\overline{s}_{\overline{u}}\overline{u}\in\mathcal{N}, then s¯u¯\overline{s}_{\overline{u}} is unique and satisfies

    𝒥(s¯u¯u¯)=maxs¯(0,)𝒥(s¯u¯).\mathcal{J}(\overline{s}_{\overline{u}}\overline{u})=\max_{\overline{s}\in(0,\infty)^{\ell}}\mathcal{J}(\overline{s}\,\overline{u}).
  • (ii)(ii)

    𝒰\mathcal{U} is a nonempty open subset of 𝒯\mathcal{T}, and the map 𝒰(0,)\mathcal{U}\to(0,\infty)^{\ell} given by u¯s¯u¯\overline{u}\mapsto\overline{s}_{\overline{u}} is continuous.

  • (iii)(iii)

    The map ρ:𝒰𝒩\rho:\mathcal{U}\to\mathcal{N} given by u¯s¯u¯u¯\overline{u}\mapsto\overline{s}_{\overline{u}}\overline{u} is a homeomorphism.

  • (iv)(iv)

    If (u¯n)(\overline{u}_{n}) is a sequence in 𝒰\mathcal{U} and u¯nu¯𝒰\overline{u}_{n}\to\overline{u}\in\partial\mathcal{U}, then |s¯u¯n||\overline{s}_{\overline{u}_{n}}|\to\infty.

Define Ψ:𝒰\Psi:\mathcal{U}\to\mathbb{R} as

Ψ(u¯):=𝒥(s¯u¯u¯).\Psi(\overline{u}):=\mathcal{J}(\overline{s}_{\overline{u}}\overline{u}).

According to Lemma 3.2, 𝒰\mathcal{U} is a Hilbert manifold, for it is an open subset of the smooth Hilbert submanifold 𝒯\mathcal{T} of \mathcal{H}. When Ψ\Psi is differentiable at u¯\overline{u}, we write Ψ(u¯)\|\Psi^{\prime}(\overline{u})\|_{*} for the norm of Ψ(u¯)\Psi^{\prime}(\overline{u}) in the cotangent space Tu¯(𝒯)\mathrm{T}_{\overline{u}}^{*}(\mathcal{T}) to 𝒯\mathcal{T} at u¯\overline{u}, i.e.,

Ψ(u¯):=supv¯Tu¯(𝒰)v¯0|Ψ(u¯)v¯|v¯,\|\Psi^{\prime}(\overline{u})\|_{*}:=\sup\limits_{\begin{subarray}{c}\overline{v}\in\mathrm{T}_{\overline{u}}(\mathcal{U})\\ \overline{v}\neq 0\end{subarray}}\frac{|\Psi^{\prime}(\overline{u})\overline{v}|}{\|\overline{v}\|},

where Tu¯(𝒰)\mathrm{T}_{\overline{u}}(\mathcal{U}) is the tangent space to 𝒰\mathcal{U} at u¯\overline{u}.

Recall that a sequence (u¯n)(\overline{u}_{n}) in 𝒰\mathcal{U} is called a (PS)c(PS)_{c}-sequence for Ψ\Psi if Ψ(u¯n)c\Psi(\overline{u}_{n})\to c and Ψ(u¯n)0\|\Psi^{\prime}(\overline{u}_{n})\|_{*}\to 0, and Ψ\Psi is said to satisfy the (PS)c(PS)_{c}-condition if every such sequence has a convergent subsequence. Similarly, a (PS)c(PS)_{c}-sequence for 𝒥\mathcal{J} is a sequence (u¯n)(\overline{u}_{n}) in \mathcal{H} such that 𝒥(u¯n)0\mathcal{J}(\overline{u}_{n})\to 0 and 𝒥(u¯n)0\|\mathcal{J}^{\prime}(\overline{u}_{n})\|_{\mathcal{H}^{\prime}}\to 0, and 𝒥\mathcal{J} satisfies the (PS)c(PS)_{c}-condition if any such sequence has a convergent subsequence. Here \mathcal{H}^{\prime} denotes the dual space of \mathcal{H}.

Lemma 3.3.
  • (i)(i)

    Ψ𝒞1(𝒰)\Psi\in\mathcal{C}^{1}(\mathcal{U}) and its derivative is given by

    Ψ(u¯)v¯=𝒥(s¯u¯u¯)[s¯u¯v¯]for all u¯𝒰 and v¯Tu¯(𝒰).\Psi^{\prime}(\overline{u})\overline{v}=\mathcal{J}^{\prime}(\overline{s}_{\overline{u}}\overline{u})[\overline{s}_{\overline{u}}\overline{v}]\quad\text{for all }\overline{u}\in\mathcal{U}\text{ and }\overline{v}\in\mathrm{T}_{\overline{u}}(\mathcal{U}).

    Moreover, there exists d0>0d_{0}>0 such that

    d0mini{su,i}𝒥(s¯u¯u¯)Ψ(u¯)maxi{su,i}𝒥(s¯u¯u¯)for all u¯𝒰.d_{0}\min_{i}\{s_{u,i}\}\|\mathcal{J}^{\prime}(\overline{s}_{\overline{u}}\overline{u})\|_{\mathcal{H}^{\prime}}\leq\|\Psi^{\prime}(\overline{u})\|_{*}\leq\max_{i}\{s_{u,i}\}\|\mathcal{J}^{\prime}(\overline{s}_{\overline{u}}\overline{u})\|_{\mathcal{H}^{\prime}}\quad\text{for all }\overline{u}\in\mathcal{U}.
  • (ii)(ii)

    If (u¯n)(\overline{u}_{n}) is a (PS)c(PS)_{c}-sequence for Ψ\Psi in 𝒰\mathcal{U}, then (s¯u¯nu¯n)(\overline{s}_{\overline{u}_{n}}\overline{u}_{n}) is a (PS)c(PS)_{c}-sequence for 𝒥\mathcal{J} in \mathcal{H}.

  • (iii)(iii)

    u¯\overline{u} is a critical point of Ψ\Psi if and only if s¯u¯u¯\overline{s}_{\overline{u}}\overline{u} is a fully nontrivial critical point of 𝒥\mathcal{J}.

  • (iv)(iv)

    If (u¯n)(\overline{u}_{n}) is a sequence in 𝒰\mathcal{U} and u¯nu¯𝒰\overline{u}_{n}\to\overline{u}\in\partial\mathcal{U}, then |Ψ(u¯n)||\Psi(\overline{u}_{n})|\to\infty.

  • (v)(v)

    u¯𝒰\overline{u}\in\mathcal{U} if and only if u¯𝒰-\overline{u}\in\mathcal{U}, and Ψ(u¯)=Ψ(u¯)\Psi(\overline{u})=\Psi(-\overline{u}).

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 cc\in\mathbb{R}, Ψ\Psi satisfies the (PS)c(PS)_{c}-condition.

Proof.

First observe that a (PS)c(PS)_{c}-sequence (v¯n)(\overline{v}_{n}) for 𝒥\mathcal{J} is bounded. Indeed, there exists C>0C>0 such that

mNv¯n2=𝒥(v¯n)12m𝒥(v¯n)v¯nC(1+v¯n),\frac{m}{N}\|\overline{v}_{n}\|^{2}=\mathcal{J}(\overline{v}_{n})-\frac{1}{2_{m}^{\ast}}\mathcal{J}^{\prime}(\overline{v}_{n})\overline{v}_{n}\leq C(1+\|\overline{v}_{n}\|),

and the claim follows.

Using this, let (u¯n)𝒰(\overline{u}_{n})\subset\mathcal{U} be a (PS)c(PS)_{c}-sequence for Ψ\Psi. By Lemma 3.3, the sequence v¯n:=ρ(u¯)𝒩\overline{v}_{n}:=\rho(\overline{u})\in\mathcal{N} is a (PS)c(PS)_{c}-sequence for 𝒥\mathcal{J} and it is bounded by the above claim. A standard argument using Lemma 2.2 as in [ClappPistoia2018, Proposition 3.6], shows that (v¯n)(\overline{v}_{n}) contains a convergent subsequence, converging to some v¯\overline{v}\in\mathcal{H}. As v¯n𝒩\overline{v}_{n}\in\mathcal{N} for every nn\in\mathbb{N} and as 𝒩\mathcal{N} is closed by Lemma 3.1, it follows that v¯𝒩\overline{v}\in\mathcal{N}. Finally, since ρ\rho is a homeomorphism between 𝒩\mathcal{N} and 𝒰\mathcal{U}, this yields that u¯n\overline{u}_{n} converges to ρ1(v¯)\rho^{-1}(\overline{v}) in a subsequence, and Ψ\Psi satisfies the (PS)c(PS)_{c}-condition ∎

Given a nonempty subset 𝒵\mathcal{Z} of 𝒯\mathcal{T} such that u¯𝒵\overline{u}\in\mathcal{Z} if and only if u¯𝒵-\overline{u}\in\mathcal{Z}, the genus of 𝒵\mathcal{Z}, denoted by genus(𝒵)\mathrm{genus}(\mathcal{Z}), is the smallest integer k1k\geq 1 such that there exists an odd continuous function 𝒵𝕊k1\mathcal{Z}\rightarrow\mathbb{S}^{k-1} into the unit sphere 𝕊k1\mathbb{S}^{k-1} in k\mathbb{R}^{k}. If no such kk exists, we define genus(𝒵)=\mathrm{genus}(\mathcal{Z})=\infty; finally, we set genus()=0\mathrm{genus}(\emptyset)=0.

Lemma 3.5.

genus(𝒰)=\mathrm{genus}(\mathcal{U})=\infty.

Proof.

Condition (Γ2\Gamma 2) together with the existence of Γ\Gamma-invariant partitions of the unity (see [Palais1961]), one obtains an arbitrarily large number of positive Γ\Gamma-invariant functions in 𝒞(M)\mathcal{C}^{\infty}(M) with mutually disjoint supports. Then, arguing as in [ClappSzulkin19, Lemma 4.5], one shows that genus(𝒰)=\mathrm{genus}(\mathcal{U})=\infty. ∎

Proof of Theorem 1.3.

Lemma 3.3 (iv)(iv) implies that 𝒰\mathcal{U} is positively invariant under the negative pseudogradient flow of Ψ\Psi, so the usual deformation lemma holds true for Ψ\Psi, see e.g. [StruweBook, Section II.3] or [WillemBook, Section 5.3]. As Ψ\Psi satisfies the (PS)c(PS)_{c}-condition for every cc\in\mathbb{R}, standard variational arguments show that Ψ\Psi attains its minimum on 𝒰\mathcal{U} at some u¯\overline{u}. By Lemma 3.3(iii)(iii) and the Principle of Symmetric Criticality, s¯u¯u¯\overline{s}_{\overline{u}}\overline{u} is a Γ\Gamma-invariant least energy fully nontrivial solution for the system (9). Moreover, as Ψ\Psi is even and genus(𝒰)=\mathrm{genus}(\mathcal{U})=\infty, arguing as in the proof of Theorem 3.4 (c) in [ClappSzulkin19], it follows that Ψ\Psi has an unbounded sequence of critical points. Using Lemma 3.3 (iii), and the fact that Ψ(u¯)=𝒥(s¯u¯u¯)=mNs¯u¯u¯2\Psi(\overline{u})=\mathcal{J}(\overline{s}_{\overline{u}}\overline{u})=\frac{m}{N}\|\overline{s}_{\overline{u}}\overline{u}\|^{2} by (14), the system (9) has an unbounded sequence of fully nontrivial Γ\Gamma-invariant solutions. ∎

We next apply Theorem 1.3 to the case =1\ell=1 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 QQ-curvature on Einstein manifolds (see, for instance, [DjadliHebeyLedoux2000] and [Gover06] respectively) it is unique. As (M,g)(M,g) is Einstein with positive scalar curvature, the operator PgP_{g} is coercive [Ro, Proposition 4] and Theorem 1.3 for =1\ell=1, yields the existence of an unbounded sequence of Γ\Gamma-invariant solutions, and the corollary follows. ∎

4. One dimensional reduction

In this section, we will strongly use that the group Γ\Gamma satisfies properties (Γ1\Gamma 1) and (Γ3\Gamma 3). Recall that MM_{-} and M+M_{+} denote the singular orbits, as it was given in the introduction, and let n1=dimMn_{1}=\dim M_{-} and n2=dimM+n_{2}=\dim M_{+}, Nni2N-n_{i}\geq 2. Since MM is compact, the geodesic distance between MM_{-} and M+M_{+},

d:=distg(M,M+),d:=\text{dist}_{g}(M_{-},M_{+}),

is attained and the distance function r:M[0,d]r:M\rightarrow[0,d] given by

r(x):=distg(M,x),r(x):=\text{dist}_{g}(M_{-},x),

is well defined. This function is a Riemannian submersion and satisfies for any xM(M+M)x\in M\setminus(M_{+}\cup M_{-}) that

|r(x)|g2=1andΔgr(x)=h(r(x)),|\nabla r(x)|_{g}^{2}=1\quad\text{and}\quad\Delta_{g}r(x)=h(r(x)),

where h(t)h(t) denotes the mean curvature of r1(t)r^{-1}(t). See [Petersen06, Chapter 2, Section 4.1], and also [BBP21, Section 2]. For the mean curvature, we explicitly have:

h(t):=codim(M)1t+t(trace(A))+trace(B)+o(t2),h(t):=\frac{\text{codim}(M_{-})-1}{t}+t(\text{trace}(A))+\text{trace}(B)+o(t^{2}),

for some matrices AA and BB not depending on tt, and it further satisfies that

limt0th(t)=Nn11andlimtd(td)h(t)=Nn21\lim_{t\rightarrow 0}t\cdot h(t)=N-n_{1}-1\quad\text{and}\quad\lim_{t\rightarrow d}(t-d)h(t)=N-n_{2}-1

(see [GeTang2014], and also [BBP21]). Therefore, for any w𝒞([0,d])w\in\mathcal{C}^{\infty}([0,d]) the following identity holds true:

(15) Δg(wr)={(Nn1)w′′(0),in M,(w′′+hw)r,in M(MM+),(Nn2)w′′(d),in M+.\Delta_{g}(w\circ r)=\begin{cases}(N-n_{1})w^{\prime\prime}(0),&\text{in }M_{-},\\ (w^{\prime\prime}+hw^{\prime})\circ r,&\text{in }M\setminus(M_{-}\cup M_{+}),\\ (N-n_{2})w^{\prime\prime}(d),&\text{in }M_{+}.\end{cases}

Moreover, notice that Mt:=r1(t)M_{t}:=r^{-1}(t) is a principal orbit for any t(0,d)t\in(0,d), while M=r1(0)M_{-}=r^{-1}(0) and M+=r1(d)M_{+}=r^{-1}(d). Hence, for every x,yMx,y\in M, it follows that

r(x)=r(y)x,yMt for some t[0,d]Γx=Γy.r(x)=r(y)\Longleftrightarrow x,y\in M_{t}\text{ for some }t\in[0,d]\Longleftrightarrow\Gamma x=\Gamma y.

Thus, for any w𝒞([0,d])w\in\mathcal{C}^{\infty}([0,d]), the function wr𝒞(M)Γw\circ r\in\mathcal{C}^{\infty}(M)^{\Gamma} and, conversely, for any u𝒞(M)Γu\in\mathcal{C}^{\infty}(M)^{\Gamma} there exists a unique w𝒞([0,d])w\in\mathcal{C}^{\infty}([0,d]) such that u=wru=w\circ r. In this way, we have a linear isomorphism

(16) ι:C(M)ΓC([0,d]),u=wrw.\iota:C^{\infty}(M)^{\Gamma}\rightarrow C^{\infty}([0,d]),\qquad u=w\circ r\mapsto w.

As in the introduction, KK will denote the principal isotropy, i.e., the stabilizer of the Γ\Gamma-action at any point p0r1(t0)p_{0}\in r^{-1}(t_{0}), for some t0(0,d)t_{0}\in(0,d). Such a group is the same at the preimage of any interior point of (0,d)(0,d) under rr. All the regular orbits are diffeomorphic to Γ/K\Gamma/K, and we will fix one, say Md/2:=r1(d/2)M_{d/2}:=r^{-1}(d/2).

Lemma 4.1.

Assume (Γ1\Gamma 1) and (Γ3\Gamma 3) hold true. Then there exists a metric gg_{\ast} on Md/2M_{d/2}, a diffeomorphism φ:(0,d)×Md/2M(MM+)\varphi:(0,d)\times M_{d/2}\rightarrow M\setminus(M_{-}\cup M_{+}) and a smooth function ϕ:[0,d]\phi:[0,d]\rightarrow\mathbb{R} such that

  1. (1)

    for every (x,t)Md/2×(0,d)(x,t)\in M_{d/2}\times(0,d), rφ(x,t)=tr\circ\varphi(x,t)=t.

  2. (2)

    dVg=ϕ(t)dtdVgdV_{g}=\phi(t)dt\wedge dV_{g_{\ast}}.

Proof.

The first item only depends on (Γ1\Gamma 1) as follows: For any xMd/2x\in M_{d/2}, consider the unique minimizing horizontal geodesic c:[0,d]Mc:[0,d]\to M joining MM_{-} and M+M_{+} such that c(d/2)=xc(d/2)=x (Cf. [AlexBettiol, Proposition 3.78]). Then, the diffeomorphism φ\varphi is given by

φ:(0,d)×Md/2M(MM+)(t,x)c(t).\begin{split}\varphi:(0,d)\times M_{d/2}&\to M\setminus(M_{-}\cup M_{+})\\ (t,x)&\mapsto c(t).\end{split}

If necessary, we may reparametrize cc so that rφ(t,x)=tr\circ\varphi(t,x)=t.

The second item follows from the fact that the volume form of a metric given as in (Γ3\Gamma 3), is the volume product. That is, for a local coordinate system (t,x1,,xN1)(t,x^{1},\dots,x^{N-1}) in MM, around an arbitrary point xMx\in M, the set {t,x1,,xN1}\left\{\frac{\partial}{\partial t},\frac{\partial}{\partial x^{1}},\dots,\frac{\partial}{\partial x^{N-1}}\right\} is a basis of the tangent space TxMT_{x}M. Then, around xx the metric gg is given by the matrix

[g]=[10000f12(t)[g1]00000fk2(t)[gk]],[g]=\left[\begin{array}[]{ccccc}1&0&0&\cdots&0\\ 0&f_{1}^{2}(t)[g_{1}]&0&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&f_{k}^{2}(t)[g_{k}]\end{array}\right],

where [gj][g_{j}] is the matrix corresponding to the metric gjg_{j}, j=1,,kj=1,\dots,k. If dj×djd_{j}\times d_{j} is the size of [gj][g_{j}], then dj=N1\sum d_{j}=N-1, and the volume form of gg is given by

dVg\displaystyle dV_{g} =\displaystyle= det([g])dtdx1dxN1\displaystyle\sqrt{{\rm det}([g])}\ dt\wedge dx^{1}\wedge\cdots\wedge dx^{N-1}
=\displaystyle= j=1kfdj(t)det([gj])dtdx1dxN1.\displaystyle\prod_{j=1}^{k}f^{d_{j}}(t)\sqrt{{\rm det}([g_{j}])}\ dt\wedge dx^{1}\wedge\cdots\wedge dx^{N-1}.

Define ϕ(t):=j=1kfdj(t)\phi(t):={\prod_{j=1}^{k}}f^{d_{j}}(t) and take g:=i=1kgig_{\ast}:={\sum_{i=1}^{k}}g_{i}, which is a metric on Md/2M_{d/2}. Therefore, dVgdV_{g_{\ast}} is given by

dVg=j=1kdet([gj])dx1dxN1,dV_{g_{\ast}}=\prod_{j=1}^{k}\sqrt{{\rm det}([g_{j}])}\ dx^{1}\wedge\cdots\wedge dx^{N-1},

and we conclude the result. ∎

We can adapt Lemma 2.2 in [FdzPetean20] to the context of cohomogeneity one actions. Recall that gt=fj2(t)gjg_{t}=\sum f_{j}^{2}(t)g_{j} denotes the metric given to the principal orbits in (Γ3\Gamma 3).

Lemma 4.2.

For any integrable function ψ:[0,d]\psi:[0,d]\rightarrow\mathbb{R},

Mψr𝑑Vg=0dVol(Md/2,gt)ψ(t)𝑑t=Vol(Md/2,g)0dψ(t)ϕ(t)𝑑t.\int_{M}\psi\circ r\;dV_{g}=\int_{0}^{d}\text{Vol}(M_{d/2},g_{t})\psi(t)\,dt=\text{Vol}(M_{d/2},g_{\ast})\int_{0}^{d}\psi(t)\phi(t)\,dt.

In particular,

Vol(Md/2,gt)=Vol(Md/2,g)ϕ(t).\text{Vol}(M_{d/2},g_{t})=\text{Vol}(M_{d/2},g_{\ast})\phi(t).
Proof.

As M+MM_{+}\cup M_{-} has Lebesgue measure zero on MM, by Lemma 4.1 and Fubini’s Theorem, we obtain on the one hand that

Mψr𝑑Vg=M(MM+)ψr𝑑Vg\displaystyle\int_{M}\psi\circ r\;dV_{g}=\int_{M\setminus(M_{-}\cup M_{+})}\psi\circ r\;dV_{g} =Md/2×(0,d)ψ(rφ)(t,x)𝑑tdVgt\displaystyle=\int_{M_{d/2}\times(0,d)}\psi\circ(r\circ\varphi)(t,x)\;dt\wedge dV_{g_{t}}
=0dMd/2ψ(t)𝑑Vgt𝑑t\displaystyle=\int_{0}^{d}\int_{M_{d/2}}\psi(t)\;dV_{g_{t}}\;dt
=0dψ(t)Md/2𝑑Vgt𝑑t\displaystyle=\int_{0}^{d}\psi(t)\int_{M_{d/2}}\;dV_{g_{t}}\,dt
=0dψ(t)Vol(Md/2,gt)𝑑t.\displaystyle=\int_{0}^{d}\psi(t)\text{Vol}(M_{d/2},g_{t})\,dt.

On the other hand, using the second expression for the volume element in Md/2×(0,d)M_{d/2}\times(0,d) in 4.1,

Mψr𝑑Vg=M(MM+)ψr𝑑Vg\displaystyle\int_{M}\psi\circ r\;dV_{g}=\int_{M\setminus(M_{-}\cup M_{+})}\psi\circ r\;dV_{g} =Md/2×(0,d)ψ(rφ)(t,x)ϕ(t)𝑑tdVg\displaystyle=\int_{M_{d/2}\times(0,d)}\psi\circ(r\circ\varphi)(t,x)\;\phi(t)dt\wedge dV_{g_{\ast}}
=0dMd/2ψ(t)ϕ(t)𝑑Vg𝑑t\displaystyle=\int_{0}^{d}\int_{M_{d/2}}\psi(t)\phi(t)\;dV_{g_{\ast}}\;dt
=Vol(Md/2,g)0dψ(t)ϕ(t)𝑑t\displaystyle=\text{Vol}(M_{d/2},g_{\ast})\int_{0}^{d}\psi(t)\phi(t)\,dt

where we conclude the integral identity.

For the volume identity, subtracting the above identities we obtain

0d[Vol(Md/2,g)ϕ(t)Vol(Md/2,gt)]ψ(t)𝑑t=0,\int_{0}^{d}[\text{Vol}(M_{d/2},g_{\ast})\phi(t)-\text{Vol}(M_{d/2},g_{t})]\psi(t)dt=0,

for any integrable function ψ\psi. We conclude that Vol(Md/2,g)ϕ(t)=Vol(Md/2,gt)\text{Vol}(M_{d/2},g_{\ast})\phi(t)=\text{Vol}(M_{d/2},g_{t}) almost everywhere in [0,d][0,d]. As the volume function and ϕ\phi are continuous, we conclude the identity. ∎

Now we study the preimage of measure zero subsets in [0,d][0,d] under the distance function rr. To this end, denote the Lebesgue measure in [0,d][0,d] by λ\lambda, and by λg\lambda_{g} the Lebesgue measure in (M,g)(M,g).

Lemma 4.3.

If E[0,d]E\subseteq[0,d] satisfies λ(E)=0\lambda(E)=0, then λg(r1(E))=0\lambda_{g}(r^{-1}(E))=0.

Proof.

Observe that the critical point set of rr is exactly the union of the singular orbits of the action, MM_{-} and M+M_{+}. Those points correspond to the endpoints of [0,d][0,d], under rr. By dimension reasons, λg(MM+)=0\lambda_{g}(M_{-}\cup M_{+})=0. Therefore, it is enough to prove the statement for any proper subset E(0,d)E\subset(0,d) of zero Lebesgue measure. Hence, we may assume that r0\nabla r\neq 0 at any point in r1(E)r^{-1}(E). Recall that, r1(c)r^{-1}(c) is a copy of the principal orbit, i.e., it is a submanifold of MM of dimension n1n-1, for any cEc\in E. Therefore λg(r1(c))=0\lambda_{g}(r^{-1}(c))=0, for any cEc\in E.

Write A:=r1(E)A:=r^{-1}(E). If EE is countable, then AA is a countable union of zero measure sets, so AA has measure zero. As M+MM_{+}\cup M_{-} has measure zero in MM and as |r|g=1|\nabla r|_{g}=1 in M(M+M)M\smallsetminus(M_{+}\cup M_{-}), if EE is uncountable, the co-area formula yields that

λg(A)\displaystyle\lambda_{g}(A) =\displaystyle= M(MM+)χA𝑑Vg\displaystyle\int_{M\setminus(M_{-}\cup M_{+})}\chi_{A}\ dV_{g}
=\displaystyle= (0,d)[{xA|r(x)=t}χA𝑑Vgt]𝑑t\displaystyle\int_{(0,d)}\left[\int_{\{x\in A|\ r(x)=t\}}\chi_{A}\ dV_{g_{t}}\right]\ dt
=\displaystyle= (0,d)[r1(t)AχA𝑑Vgt]𝑑t\displaystyle\int_{(0,d)}\left[\int_{r^{-1}(t)\cap A}\chi_{A}\ dV_{g_{t}}\right]\ dt
=\displaystyle= E[r1(t)A𝑑Vgt]𝑑t\displaystyle\int_{E}\left[\int_{r^{-1}(t)\cap A}\ dV_{g_{t}}\right]\ dt

where χA\chi_{A} is the characteristic function of AA in MM. Since λ(E)=0\lambda(E)=0, then the Lebesgue integral of any measurable function over EE is zero. In particular, this implies that λg(A)=0\lambda_{g}(A)=0. ∎

In what follows, we will denote the set of positive and smooth Γ\Gamma-invariant functions on MM by 𝒞+(M)Γ\mathcal{C}_{+}^{\infty}(M)^{\Gamma} and by 𝒞+([0,d])\mathcal{C}^{\infty}_{+}([0,d]) the set of positive and smooth functions in [0,d][0,d]. Next, we study how the symmetries allow us to reduce the operator PgP_{g} into an operator acting on smooth functions defined in the interval [0,d][0,d]. In order to motivate a more general differential operator for which our theory holds true, first observe that if (M,g)(M,g) is Einstein with positive scalar curvature μ\mu, the higher order conformal operator, PgP_{g}, can be written as

Pg=i=1m(Δg+ci)P_{g}=\prod_{i=1}^{m}\left(-\Delta_{g}+c_{i}\right)

for some suitable constants ci>0c_{i}>0 (see [Gover06, Juhl13]). On the other hand, in case m=1m=1, when the scalar curvature RgR_{g} is positive, the conformal Laplacian is simply

Pg=Δg+N24(N1)Rg=Δg+N24(N1)Rg(Δg)0,P_{g}=-\Delta_{g}+\frac{N-2}{4(N-1)}R_{g}=-\Delta_{g}+\frac{N-2}{4(N-1)}R_{g}(-\Delta_{g})^{0},

with N24(N1)RgC+(M)Γ\frac{N-2}{4(N-1)}R_{g}\in C^{\infty}_{+}(M)^{\Gamma}. Hence, for any 2m<N2m<N and any a:=(a0,a1,,am)𝒞+(M)Γ×(0,)m\textbf{a}:=(a_{0},a_{1},\ldots,a_{m})\in\mathcal{C}_{+}^{\infty}(M)^{\Gamma}\times(0,\infty)^{m}, we are led to define the operator

(17) Pa:=i=0mai(Δg)i.P_{\textbf{a}}:=\sum_{i=0}^{m}a_{i}(-\Delta_{g})^{i}.

As for any i{0}i\in\mathbb{N}\cup\{0\} and any pair of functions u,v𝒞(M)u,v\in\mathcal{C}^{\infty}(M) we have that

(18) Mv(Δg)iu𝑑Vg={MΔgi/2vΔgi/2u𝑑Vg,i even,MΔg(i1)/2v,Δg(i1)/2ug𝑑Vg,i odd,\int_{M}v(-\Delta_{g})^{i}u\;dV_{g}=\begin{cases}\int_{M}\Delta_{g}^{i/2}v\Delta_{g}^{i/2}u\;dV_{g},&i\text{ even,}\\ \int_{M}\langle\nabla\Delta_{g}^{(i-1)/2}v,\nabla\Delta_{g}^{(i-1)/2}u\rangle_{g}\;dV_{g},&i\text{ odd},\end{cases}

then the bilinear form

(19) (u,v)g,a:=MvPau𝑑Vg=i=0ievenmMaiΔgi/2vΔgi/2u𝑑Vg+i=0ioddmMaiΔg(i1)/2v,Δg(i1)/2ug𝑑Vg,u,v𝒞(M)\begin{split}&(u,v)_{g,\textbf{a}}:=\int_{M}vP_{\textbf{a}}u\;dV_{g}\\ &=\sum_{\begin{subarray}{c}i=0\\ i\ even\end{subarray}}^{m}\int_{M}a_{i}\Delta_{g}^{i/2}v\Delta_{g}^{i/2}u\;dV_{g}+\sum_{\begin{subarray}{c}i=0\\ i\ odd\end{subarray}}^{m}\int_{M}a_{i}\langle\nabla\Delta_{g}^{(i-1)/2}v,\nabla\Delta_{g}^{(i-1)/2}u\rangle_{g}\;dV_{g},\quad u,v\in\mathcal{C}^{\infty}(M)\end{split}

is positive definite and yields a norm g,a\|\cdot\|_{g,\textbf{a}} in Hgm(M)H_{g}^{m}(M), equivalent to the standard norm in Hg1(M)H_{g}^{1}(M). Note that the term for i=0i=0 is simply Ma0uv𝑑Vg,\int_{M}a_{0}uv\;dV_{g}, and a0>0a_{0}>0 but not necessarily constant.

On the other hand, let α0C+([0,d])\alpha_{0}\in C^{\infty}_{+}([0,d]) be such that a0=α0ra_{0}=\alpha_{0}\circ r, β(t):=Vol(Md/2,gt)\beta(t):=\text{Vol}(M_{d/2},g_{t}), t[0,d]t\in[0,d], and define the operator :𝒞(0,d)𝒞(0,d)\mathcal{L}:\mathcal{C}^{\infty}(0,d)\rightarrow\mathcal{C}^{\infty}(0,d) by

:=d2dt2+h(t)ddt.\mathcal{L}:=\frac{d^{2}}{dt^{2}}+h(t)\frac{d}{dt}.

For w𝒞([0,d])w\in\mathcal{C}^{\infty}([0,d]) define

wβ,a:=(i0ievenmai0d|i/2w|2β𝑑t+i=0ioddmai0d|((i1)/2w)|2β𝑑t+0dα0|w|2β𝑑t)1/2\|w\|_{\beta,\textbf{a}}:=\left(\sum_{\begin{subarray}{c}i\neq 0\\ i\ even\end{subarray}}^{m}a_{i}\int_{0}^{d}|\mathcal{L}^{i/2}w|^{2}\beta\ dt+\sum_{\begin{subarray}{c}i=0\\ i\ odd\end{subarray}}^{m}a_{i}\int_{0}^{d}|\left(\mathcal{L}^{(i-1)/2}w\right)^{\prime}|^{2}\beta\ dt+\int_{0}^{d}\alpha_{0}|w|^{2}\beta\,dt\right)^{1/2}

where i\mathcal{L}^{i} denotes the ii-fold composition of \mathcal{L}, 0=Id\mathcal{L}^{0}=Id and

(iw):=ddt[(d2dt2+h(t)ddt)iw].\left(\mathcal{L}^{i}w\right)^{\prime}:=\frac{d}{dt}\left[\left(\frac{d^{2}}{dt^{2}}+h(t)\frac{d}{dt}\right)^{i}w\right].
Proposition 4.4.

For any a𝒞+(M)Γ×(0,)m\textbf{a}\in\mathcal{C}_{+}^{\infty}(M)^{\Gamma}\times(0,\infty)^{m} and any u=wr𝒞(M)Γu=w\circ r\in\mathcal{C}^{\infty}(M)^{\Gamma},

wrg,a=wrβ,a.\|w\circ r\|_{g,\textbf{a}}=\|w\circ r\|_{\beta,\textbf{a}}.
Proof.

Using (15), we obtain that

(20) Δgi(wr)=(iw)r,i{0},\Delta_{g}^{i}(w\circ r)=\left(\mathcal{L}^{i}w\right)\circ r,\qquad i\in\mathbb{N}\cup\{0\},

and therefore

|(Δgiwr)|g2\displaystyle|\nabla(\Delta_{g}^{i}w\circ r)|_{g}^{2} =|((iw)r)|g2\displaystyle=|\nabla\left((\mathcal{L}^{i}w)\circ r\right)|^{2}_{g}
=((iw)r),((iw)r)g\displaystyle=\left\langle\nabla\left((\mathcal{L}^{i}w)\circ r\right),\nabla\left((\mathcal{L}^{i}w)\circ r\right)\right\rangle_{g}
=|((iw))r|2|r|2\displaystyle=\left|\left((\mathcal{L}^{i}w)\right)^{\prime}\circ r\right|^{2}|\nabla r|^{2}
=|(iw)|2r.\displaystyle=\left|\left(\mathcal{L}^{i}w\right)^{\prime}\right|^{2}\circ r.

By Lemma 4.2 this implies that

M|Δgi(wr)|2𝑑Vg=0d|iw|2β(t)𝑑t and M|Δgi(wr)|g2𝑑Vg=0d|(iw)|2β(t)𝑑t,\int_{M}|\Delta_{g}^{i}(w\circ r)|^{2}\;dV_{g}=\int_{0}^{d}|\mathcal{L}^{i}w|^{2}\beta(t)\;dt\quad\text{ and }\quad\int_{M}|\nabla\Delta_{g}^{i}(w\circ r)|_{g}^{2}\;dV_{g}=\int_{0}^{d}\left|\left(\mathcal{L}^{i}w\right)^{\prime}\right|^{2}\beta(t)\;dt,

for every ii\in\mathbb{N}, while for i=0i=0 we get

Ma0|wr|2𝑑Vg=M(α0r)|wr|2𝑑Vg=0dα0|w|2β(t)𝑑t\int_{M}a_{0}|w\circ r|^{2}\;dV_{g}=\int_{M}(\alpha_{0}\circ r)|w\circ r|^{2}\;dV_{g}=\int_{0}^{d}\alpha_{0}|w|^{2}\beta(t)\ dt

In this way, using (18) and (19) we obtain that

wrg,a2=M(wr)Pa(wr)𝑑Vg\displaystyle\|w\circ r\|_{g,\textbf{a}}^{2}=\int_{M}(w\circ r)P_{\textbf{a}}(w\circ r)\;dV_{g}
=i0ievenmaiM|Δgi/2(wr)|2𝑑Vg+i=0ioddmaiM|Δg(i1)/2(wr)|g2𝑑Vg+Ma0|wr|2𝑑Vg\displaystyle=\sum_{\begin{subarray}{c}i\neq 0\\ i\ even\end{subarray}}^{m}a_{i}\int_{M}|\Delta_{g}^{i/2}(w\circ r)|^{2}\;dV_{g}+\sum_{\begin{subarray}{c}i=0\\ i\ odd\end{subarray}}^{m}a_{i}\int_{M}|\nabla\Delta_{g}^{(i-1)/2}(w\circ r)|_{g}^{2}\;dV_{g}+\int_{M}a_{0}|w\circ r|^{2}\;dV_{g}
=i0ievenmai0d|i/2w|2β(t)dt+i=0ioddmai0d(|(i1)/2w)|2β(t)dt+0dα0|w|2β(t)dt\displaystyle=\sum_{\begin{subarray}{c}i\neq 0\\ i\ even\end{subarray}}^{m}a_{i}\int_{0}^{d}|\mathcal{L}^{i/2}w|^{2}\beta(t)\;dt+\sum_{\begin{subarray}{c}i=0\\ i\ odd\end{subarray}}^{m}a_{i}\int_{0}^{d}\left(\left|\mathcal{L}^{(i-1)/2}w\right)^{\prime}\right|^{2}\beta(t)\;dt+\int_{0}^{d}\alpha_{0}|w|^{2}\beta(t)\ dt
=wβ,a,\displaystyle=\|w\|_{\beta,\textbf{a}},

as we wanted to prove. ∎

From this result, it follows that β,a\|\cdot\|_{\beta,\textbf{a}} is a well defined norm in C[0,d]C^{\infty}[0,d] and we define the weighted Sobolev space Hβm(0,d)H_{\beta}^{m}(0,d) to be the closure of C[0,d]C^{\infty}[0,d] under this norm.

We have the following direct consequence of the previous result.

Theorem 4.5.

For any a𝒞+(M)Γ×(0,)m\textbf{a}\in\mathcal{C}_{+}^{\infty}(M)^{\Gamma}\times(0,\infty)^{m}, the linear isomorphism ι\iota given in (16), induces a well defined continuous isometric isomorphism

ι:(Hgm(M)Γ,g,a)(Hβm(0,d),β,a).\iota:\left(H_{g}^{m}(M)^{\Gamma},\|\cdot\|_{g,\textbf{a}}\right)\rightarrow\left(H_{\beta}^{m}(0,d),\|\cdot\|_{\beta,\textbf{a}}\right).
Proof.

Take C(M)ΓC^{\infty}(M)^{\Gamma} and C([0,d])C^{\infty}([0,d]) as dense subspaces of Hgm(M)ΓH_{g}^{m}(M)^{\Gamma} and Hβm(0,d)H_{\beta}^{m}(0,d) under the norms g,a\|\cdot\|_{g,\textbf{a}} and β,a\|\cdot\|_{\beta,\textbf{a}}, respectively. By Proposition 4.4, given any uC(M)Γu\in C^{\infty}(M)^{\Gamma}, u=wru=w\circ r, we have that ug,a=wβ,a=ι(u)β,a\|u\|_{g,\textbf{a}}=\|w\|_{\beta,\textbf{a}}=\|\iota(u)\|_{\beta,\textbf{a}}, and the map ι:C(M)ΓC([0,d])\iota:C^{\infty}(M)^{\Gamma}\rightarrow C^{\infty}([0,d]) is a linear and continuous isometric isomorphism. Thus, ι\iota can be extended, in a unique way, to a linear and continuous isometric isomorphism defined on the whole Sobolev space Hgm(M)ΓH_{g}^{m}(M)^{\Gamma}, as we wanted to prove. ∎

Next we see how the standard norm in Hm(0,d)H^{m}(0,d) is related with the weighted norms β,a.\|\cdot\|_{\beta,\textbf{a}}.

Lemma 4.6.

For each ε>0\varepsilon>0, there exist k=(k0,,km)𝒞+(M)Γ×(0,)m\textbf{k}=(k_{0},\ldots,k_{m})\in\mathcal{C}_{+}^{\infty}(M)^{\Gamma}\times(0,\infty)^{m}, with k0k_{0} constant, and A,B>0A,B>0, depending on ε\varepsilon, such that

BwHm(ε,dε)wβ,kAwHm(ε,dε)B\|w\|_{H^{m}(\varepsilon,d-\varepsilon)}\geq\|w\|_{\beta,\textbf{k}}\geq A\|w\|_{H^{m}(\varepsilon,d-\varepsilon)}
Proof.

As β\beta is continuous and positive in (0,d)(0,d), max[ε,dε]β>0\max_{[\varepsilon,d-\varepsilon]}\beta>0. Then, it is readily seen that, for any a=(α0r,a1,,am)𝒞+(M)Γ×(0,)m\textbf{a}=(\alpha_{0}\circ r,a_{1},\ldots,a_{m})\in\mathcal{C}_{+}^{\infty}(M)^{\Gamma}\times(0,\infty)^{m}, the inequality

wβ,aBwHm(ε,dε)\|w\|_{\beta,\textbf{a}}\leq B\|w\|_{H^{m}(\varepsilon,d-\varepsilon)}

holds true, where BB is a suitable constant depending only on maxiai\max_{i}a_{i}, max[ε,dε]β\max_{[\varepsilon,d-\varepsilon]}\beta and max[ε,dε]α0β>0\max_{[\varepsilon,d-\varepsilon]}\alpha_{0}\beta>0.

The proof of the second inequality is exactly the same as in [ClappFernandezSaldana2021, Lemma 2.3].

Corollary 4.7.

For any uHgm(M)Γu\in H_{g}^{m}(M)^{\Gamma} there exists u~Cm1(M(MM+))Γ\widetilde{u}\in C^{m-1}(M\setminus(M_{-}\cup M_{+}))^{\Gamma} such that

u=u~, a.e. in M.u=\widetilde{u},\quad\text{ a.e. in }M.
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 a𝒞+(M)×(0,)m\textbf{a}\in\mathcal{C}_{+}^{\infty}(M)\times(0,\infty)^{m}, the operator PaP_{\textbf{a}} is coercive and, therefore the norm g,a\|\cdot\|_{g,\textbf{a}} is equivalent to the standard norm in Hgm(M)\|\cdot\|_{H^{m}_{g}(M)}. Next, fix ε>0\varepsilon>0, take k𝒞+(M)Γ×(0,)m\textbf{k}\in\mathcal{C}_{+}^{\infty}(M)^{\Gamma}\times(0,\infty)^{m} and A,B>0A,B>0 as in Lemma 4.6 and let Ωε,dε:=r1(ε,dε)\Omega_{\varepsilon,d-\varepsilon}:=r^{-1}(\varepsilon,d-\varepsilon). Since g,a\|\cdot\|_{g,\textbf{a}} is equivalent to the standard norm in Hgm(M)H_{g}^{m}(M), the map ι:(Hgm(Ωε,dε)Γ,Hgm(M))(Hm(ε,dε),Hm)\iota:(H_{g}^{m}(\Omega_{\varepsilon,d-\varepsilon})^{\Gamma},\|\cdot\|_{H_{g}^{m}(M)})\rightarrow(H^{m}(\varepsilon,d-\varepsilon),\|\cdot\|_{H^{m}}) is continuous. Since the Sobolev embedding Hm(ε,dε)Cm1(ε,dε)H^{m}(\varepsilon,d-\varepsilon)\hookrightarrow C^{m-1}(\varepsilon,d-\varepsilon) is also continuous, for any uHgm(M)Γu\in H_{g}^{m}(M)^{\Gamma} and wHβm(0,d)w\in H_{\beta}^{m}(0,d) such that u=wru=w\circ r, there exists wεCm1[ε,dε]w_{\varepsilon}\in C^{m-1}[\varepsilon,d-\varepsilon] such that w=wεw=w_{\varepsilon} a.e. in [ε,dε][\varepsilon,d-\varepsilon]. Applying Lemma 4.3 it follows that u=uεu=u_{\varepsilon} a.e. in Ωε,dε\Omega_{\varepsilon,d-\varepsilon}, where uε=wεrCm1(Ωε,dε)u_{\varepsilon}=w_{\varepsilon}\circ r\in C^{m-1}(\Omega_{\varepsilon,d-\varepsilon}). As MM+M_{-}\cup M_{+} has measure zero in MM, the function u:M(MM+)u:M\setminus(M_{-}\cup M_{+})\rightarrow\mathbb{R} given by u~(p):=uε(p)\widetilde{u}(p):=u_{\varepsilon}(p) if pΩε,dεp\in\Omega_{\varepsilon,d-\varepsilon}, is well defined, is of class Cm1C^{m-1} on M(MM+)M\setminus(M_{-}\cup M_{+}) and coincides a.e. with uu on MM. ∎

For any a,b(0,d)a,b\in(0,d), let Ωa,b:=r1(a,b)\Omega_{a,b}:=r^{-1}(a,b). We now show that the Dirichlet boundary problem

(21) {Pgu=|u|p2u, in Ωa,b,ku=0,k=0,,2m1, on Ωa,b,\begin{cases}P_{g}u=|u|^{p-2}u,&\text{ in }\Omega_{a,b},\\ \nabla^{k}u=0,k=0,\ldots,2m-1,&\text{ on }\partial\Omega_{a,b},\end{cases}

induces a one dimensional Dirichlet boundary problem. We need some preliminary lemmas.

Lemma 4.8.

Let u=wru=w\circ r for a smooth function w:[0,d]w:[0,d]\to\mathbb{R}. Then for k1k\geq 1,

ku=j=0k1(w(kj)r)Tk,j+1,\nabla^{k}u=\sum_{j=0}^{k-1}(w^{(k-j)}\circ r)\ T^{k,j+1},

where w(i)w^{(i)} denotes the ii-th derivative of ww over \mathbb{R}, Tk,j+1T^{k,j+1} is a kk tensor, varying with kk, which is a combination of tensor products of the tensors r,2r,,j+1r\nabla r,\nabla^{2}r,\dots,\nabla^{j+1}r. Moreover, for any kk,

Tk,1=rr,(k factors).T^{k,1}=\nabla r\otimes\cdots\otimes\nabla r,\qquad(k\text{ factors}).
Proof.

We will proceed by induction over kk.

Case k=1k=1. Denote by XX an arbitrary vector field that is tangent to the level sets MtM_{t} of the distance function rr. By definition of gradient and by the chain rule,

u,Xg\displaystyle\langle\nabla u,X\rangle_{g} =\displaystyle= X(u)\displaystyle X(u)
=\displaystyle= (wr)X(r)=0.\displaystyle(w^{\prime}\circ r)X(r)=0.

Analogously, computing in the direction of the normal to MtM_{t} and using the fact that |r|g2=1|\nabla r|^{2}_{g}=1, we get that

u,rg\displaystyle\langle\nabla u,\nabla r\rangle_{g} =\displaystyle= r(u)\displaystyle\nabla r(u)
=\displaystyle= (wr)r(r)\displaystyle(w^{\prime}\circ r)\nabla r(r)
=\displaystyle= (wr)r,rg\displaystyle(w^{\prime}\circ r)\langle\nabla r,\nabla r\rangle_{g}
=\displaystyle= (wr)|r|g\displaystyle(w^{\prime}\circ r)|\nabla r|_{g}
=\displaystyle= (wr).\displaystyle(w^{\prime}\circ r).

Therefore,

u=(wr)r.\nabla u=(w^{\prime}\circ r)\nabla r.

Case k=2k=2. Recall that the Levi-Civita connection induces a covariant derivative for higher-order tensors. Given a tensor TT of order kk, the derivative T\nabla T is a tensor of order (k+1)(k+1) given by the formula:

T(X1,,Xk,Xk+1)=Xk+1(T(X1,,Xk))T(Xk+1X1,,Xk)T(X1,,Xk+1Xk).\begin{split}\nabla T(X_{1},\dots,X_{k},X_{k+1})=&X_{k+1}(T(X_{1},\dots,X_{k}))-T(\nabla_{X_{k+1}}X_{1},\dots,X_{k})-...\\ &-T(X_{1},\dots,\nabla_{X_{k+1}}X_{k}).\end{split}

We then compute for any vector fields X1,X2X_{1},X_{2} on MM:

2u(X1,X2)\displaystyle\nabla^{2}u(X_{1},X_{2}) =\displaystyle= ((wr)r)(X1,X2)\displaystyle\nabla\left((w^{\prime}\circ r)\nabla r\right)(X_{1},X_{2})
=\displaystyle= X2((wr)r(X1))(wr)r(X2X1)\displaystyle X_{2}\left((w^{\prime}\circ r)\nabla r(X_{1})\right)-(w^{\prime}\circ r)\nabla r\left(\nabla_{X_{2}}X_{1}\right)
=\displaystyle= (wr)[X2(r(X1))r(X2X1)]+X2(wr)r(X1)\displaystyle(w^{\prime}\circ r)\left[X_{2}(\nabla r(X_{1}))-\nabla r\left(\nabla_{X_{2}}X_{1}\right)\right]+X_{2}(w^{\prime}\circ r)\nabla r(X_{1})
=\displaystyle= (wr)2r(X1,X2)+X2(wr)r(X1)\displaystyle(w^{\prime}\circ r)\nabla^{2}r(X_{1},X_{2})+X_{2}(w^{\prime}\circ r)\nabla r(X_{1})
=\displaystyle= (wr)2r(X1,X2)+(w′′r)r(X1)r(X2).\displaystyle(w^{\prime}\circ r)\nabla^{2}r(X_{1},X_{2})+(w^{\prime\prime}\circ r)\nabla r(X_{1})\nabla r(X_{2}).

Therefore:

2u=(wr)2r+(w′′r)rr.\nabla^{2}u=(w^{\prime}\circ r)\nabla^{2}r+(w^{\prime\prime}\circ r)\nabla r\otimes\nabla r.

Case k3k\geq 3. Now, suppose that

k1u=j=0k1(w(k1j)r)Tk1,j+1,\nabla^{k-1}u=\sum_{j=0}^{k-1}(w^{(k-1-j)}\circ r)T^{k-1,j+1},

where the tensors Tk1,j+1rT^{k-1,j+1}r satisfy the conditions in the Lemma. Take any X1,,XkX_{1},\dots,X_{k} vector fields on MM and compute:

(22) ku(X1,,Xk)\displaystyle\nabla^{k}u(X_{1},\dots,X_{k}) =Xk(k1u(X1,,Xk1))k1u(XkX1,,Xk1)\displaystyle=X_{k}(\nabla^{k-1}u\left(X_{1},\dots,X_{k-1}\right))-\nabla^{k-1}u(\nabla_{X_{k}}X_{1},\dots,X_{k-1})
k1u(X1,,XkXk1)\displaystyle\quad-\cdots-\nabla^{k-1}u(X_{1},\dots,\nabla_{X_{k}}X_{k-1})

We substitute the expression for k1u\nabla^{k-1}u. For the sake of clarity, let us analyze the first summand:

Xk(k1u(X1,,Xk1))\displaystyle X_{k}(\nabla^{k-1}u\left(X_{1},\dots,X_{k-1}\right)) =Xk(j=0k1(w(k1j)r)Tk1,j+1(X1,,Xk1))\displaystyle=X_{k}\left(\sum_{j=0}^{k-1}(w^{(k-1-j)}\circ r)T^{k-1,j+1}(X_{1},\dots,X_{k-1})\right)
=j=0k1Xk(w(k1j)r)Tk1,j+1(X1,,Xk1)\displaystyle=\sum_{j=0}^{k-1}X_{k}(w^{(k-1-j)}\circ r)T^{k-1,j+1}(X_{1},\dots,X_{k-1})
+j=0k1(w(k1j)r)Xk(Tk1,j+1(X1,,Xk1)).\displaystyle\quad+\sum_{j=0}^{k-1}(w^{(k-1-j)}\circ r)X_{k}(T^{k-1,j+1}(X_{1},\dots,X_{k-1})).

Now,

Xk(w(k1j)r)=(w(kj)r)Xk(r).X_{k}(w^{(k-1-j)}\circ r)=(w^{(k-j)}\circ r)X_{k}(r).

Note that j=0j=0 gives the only term with the factor (w(k)r)(w^{(k)}\circ r) in the expression for ku\nabla^{k}u. Observe also that Xk(Tk1,j+1(X1,,Xk1))X_{k}(T^{k-1,j+1}(X_{1},\dots,X_{k-1})) is one of the terms in the definition of Tk1,j+1(X1,,Xk)\nabla T^{k-1,j+1}(X_{1},\dots,X_{k}); the others will be obtained from the remaining terms in (22). If Tk1,j+1T^{k-1,j+1} is a combination of tensor products of r,2r,,j+1\nabla r,\nabla^{2}r,\dots,\nabla^{j+1}, the same happens with Tk1,j+1\nabla T^{k-1,j+1}. After a long but straightforward calculation, we have

ku(X1,,Xk)\displaystyle\nabla^{k}u(X_{1},\dots,X_{k}) =(w(k)r)Xk(r)Tk1,1(X1,,Xk1)\displaystyle=(w^{(k)}\circ r)X_{k}(r)T^{k-1,1}(X_{1},\dots,X_{k-1})
+j=0k2(w(k1j)r)Tk,j+1(X1,,Xk),\displaystyle\quad+\sum_{j=0}^{k-2}(w^{(k-1-j)}\circ r)T^{k,j+1}(X_{1},\dots,X_{k}),

for some tensors Tk,j+1T^{k,j+1}. By the inductive hypothesis,

Xk(r)Tk1,1(X1,,Xk1)\displaystyle X_{k}(r)T^{k-1,1}(X_{1},\dots,X_{k-1}) =Xk(r)(rr)(X1,,Xk1)((k1) factors)\displaystyle=X_{k}(r)(\nabla r\otimes\cdots\otimes\nabla r)(X_{1},\dots,X_{k-1})\quad\left((k-1)\text{ factors}\right)
=(rr)(X1,,Xk)(k factors),\displaystyle=(\nabla r\otimes\cdots\otimes\nabla r)(X_{1},\dots,X_{k})\quad(k\text{ factors}),

which proves the lemma. ∎

Proposition 4.9.

Let u=wru=w\circ r, for a smooth function w:[0,d]w\colon[0,d]\to\mathbb{R}. Let xMx\in M, and a fixed integer k1k\geq 1 such that lu(x)=0\nabla^{l}u(x)=0 for all 1lk1\leq l\leq k, then (w(l)r)(x)=0(w^{(l)}\circ r)(x)=0, for all 1lk1\leq l\leq k.

Proof.

We will proceed by induction over kk. For k=1k=1, we have:

u=(wr)r.\nabla u=(w^{\prime}\circ r)\nabla r.

Evaluating at xx and r\nabla r,

0=u(x)(r)=(wr)(x)r(r)(x)=(wr)(x).0=\nabla u(x)(\nabla r)=(w^{\prime}\circ r)(x)\nabla r(\nabla r)(x)=(w^{\prime}\circ r)(x).

Now take k>1k>1 and suppose that lu(x)=0\nabla^{l}u(x)=0, (w(l)r)(x)=0(w^{(l)}\circ r)(x)=0 for all 1l<k1\leq l<k and ku(x)=0\nabla^{k}u(x)=0. Using this and the previous lemma,

ku=j=0k1(w(kj)r)Tk,j+1=(w(k)r)Tk,1=(w(k)r)(rr)\nabla^{k}u=\sum_{j=0}^{k-1}(w^{(k-j)}\circ r)T^{k,j+1}=(w^{(k)}\circ r)T^{k,1}=(w^{(k)}\circ r)(\nabla r\otimes\cdots\otimes\nabla r)

Evaluating at xx and (r,,r)(\nabla r,\dots,\nabla r),

0=ku(x)(r,,r)=(wkr)(x).0=\nabla^{k}u(x)(\nabla r,\dots,\nabla r)=(w^{k}\circ r)(x).

The result follows. ∎

Corollary 4.10.

Let (a0,a1,,am)𝒞+(M)Γ×(0,)m(a_{0},a_{1},\ldots,a_{m})\in\mathcal{C}_{+}^{\infty}(M)^{\Gamma}\times(0,\infty)^{m} such that Pg=i=0mai(Δg)iP_{g}=\sum_{i=0}^{m}a_{i}(-\Delta_{g})^{i} and define the operator

^=α0+i=1mai()i,\widehat{\mathcal{L}}=\alpha_{0}+\sum_{i=1}^{m}a_{i}(-\mathcal{L})^{i},

where a0=α0r.a_{0}=\alpha_{0}\circ r. If u=wrC2m(Ωa,b)u=w\circ r\in C^{2m}(\Omega_{a,b}) is a solution to (21), then ww is a solution to the problem

(23) {^w=|w|p2w, in (a,b),w(k)(a)=w(k)(b)=0,k=0,,2m1.\begin{cases}\widehat{\mathcal{L}}w=|w|^{p-2}w,&\text{ in }(a,b),\\ w^{(k)}(a)=w^{(k)}(b)=0,&k=0,\ldots,2m-1.\end{cases}
Proof.

Let u=wru=w\circ r be a smooth Γ\Gamma-invariant solution to (21). From (20) it follows that

Pgu=i=0mai(Δg)iu=i=0mai(()iw)r=^(w)rP_{g}u=\sum_{i=0}^{m}a_{i}(-\Delta_{g})^{i}u=\sum_{i=0}^{m}a_{i}\left((-\mathcal{L})^{i}w\right)\circ r=\widehat{\mathcal{L}}(w)\circ r

and ww satisfies ^w=|w|p2w\widehat{\mathcal{L}}w=|w|^{p-2}w in (a,b)(a,b). Moreover, as ku(x)=0\nabla^{k}u(x)=0 for every 0km10\leq k\leq m-1 and every xΩa,bx\in\partial\Omega_{a,b}, by Proposition 4.9,

0=w(k)r(x)={w(k)(a),if xr1(a)w(k)(b),if xr1(b),0=w^{(k)}\circ r(x)=\begin{cases}w^{(k)}(a),&\text{if }x\in r^{-1}(a)\\ w^{(k)}(b),&\text{if }x\in r^{-1}(b),\end{cases}

and ww is a (strong) solution to (23). ∎

5. Segregation and optimal partitions

In this section, we will suppose that (M,g)(M,g) is such that the operator PgP_{g} can be written in the form (17) and that Γ\Gamma satisfies (Γ1\Gamma 1) and (Γ2\Gamma 2), so that the results of the previous section hold true. Remember that this is possible, for example, if (M,g)(M,g) is an Einstein manifold with positive scalar curvature or when Rg>0R_{g}>0 in case m=1m=1. See Section 6 for concrete examples.

Recall that for a compact Lie group Γ\Gamma, a principal Γ\Gamma-bundle is a fiber bundle ΓPB\Gamma\to P\to B, whose structure group is Γ\Gamma, together with a Γ\Gamma-action on Γ\Gamma itself by left translations, and a free right Γ\Gamma-action on PP, whose orbits are the fibers of the bundle. Let FF be another smooth manifold that admits a left action by Γ\Gamma. The orbit space of the diagonal action on P×FP\times F is a smooth manifold denoted by P×ΓFP\times_{\Gamma}F, given as the total space of the fiber bundle

FP×ΓFB.F\to P\times_{\Gamma}F\to B.

In the literature, the latter is known as the associated bundle to the principal bundle ΓPB\Gamma\to P\to B, and P×ΓFP\times_{\Gamma}F is called the twisted space. See [AlexBettiol, Section 3.1] for definitions and a detailed explanation.

Let Ω\Omega be a Γ\Gamma-invariant open subset of MM with smooth boundary and recall the definitions of the energy functional JΩJ_{\Omega} and the Hilbert manifold ΩΓ\mathcal{M}_{\Omega}^{\Gamma} given in Section 2. By Proposition 2.3, problem (5) admits a least energy Γ\Gamma-invariant solution. So the quantity cΩΓc_{\Omega}^{\Gamma} defined in the introduction is attained.

Theorem 1.1 will follow from the next segregation result.

Theorem 5.1.

Suppose Γ\Gamma satisfies conditions (Γ1\Gamma 1), (Γ2\Gamma 2) and (Γ3\Gamma 3), and that PgP_{g} can be written as a sum of Laplacians of the form (17). For i=1,,i=1,\ldots,\ell, fix νi=1\nu_{i}=1 and for each iji\neq j, kk\in\mathbb{N}, let ηij,k<0\eta_{ij,k}<0 be such that ηij,k=ηji,k\eta_{ij,k}=\eta_{ji,k} and ηij,k\eta_{ij,k}\to-\infty as kk\to\infty. Let (uk,1,,uk,)(u_{k,1},\ldots,u_{k,\ell}) be a least energy fully nontrivial solution to the system (9) with ηij=ηij,k\eta_{ij}=\eta_{ij,k}. Then, there exists u,1,u,Hgm(M)Γu_{\infty,1},\ldots u_{\infty,\ell}\in H_{g}^{m}(M)^{\Gamma} such that, up to a subsequence,

  • (a)(a)

    uk,iu,iu_{k,i}\to u_{\infty,i} strongly in Hgm(M)H^{m}_{g}(M), u,i𝒞m1(M)u_{\infty,i}\in\mathcal{C}^{m-1}(M), u,i0u_{\infty,i}\neq 0. Let

    Ωi:=int{xM:u,i(x)0}¯ for i=1,,.\displaystyle\Omega_{i}:=\operatorname{int}\overline{\{x\in M:u_{\infty,i}(x)\neq 0\}}\qquad\text{ for \ }i=1,\ldots,\ell.

    Then u,iH0,gm(Ω)Γu_{\infty,i}\in H_{0,g}^{m}(\Omega)^{\Gamma} is a least energy solution of (5) in Ωi\Omega_{i} for each i=1,,i=1,\ldots,\ell.

  • (b)(b)

    {Ω1,,Ω}𝒫Γ\{\Omega_{1},\ldots,\Omega_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma} is a solution to the Γ\Gamma-invariant \ell–optimal partition problem (6) satisfying the following properties:

    1. (1)

      Ωi\Omega_{i} is smooth and connected for every i=1,,i=1,\ldots,\ell, Ω¯iΩ¯i+1\overline{\Omega}_{i}\cap\overline{\Omega}_{i+1}\neq\emptyset, ΩiΩj=\Omega_{i}\cap\Omega_{j}=\emptyset if |ij|2|i-j|\geq 2 and Ω1Ω¯=M\overline{\Omega_{1}\cup\ldots\cup\Omega_{\ell}}=M;

    2. (2)
      Ω1G×KD,ΩG×K+D+,Ω1ΩΓ/K;\Omega_{1}\approx G\times_{K-}D_{-},\quad\Omega_{\ell}\approx G\times_{K+}D_{+},\quad\partial\Omega_{1}\approx\partial\Omega_{\ell}\approx\Gamma/K;
    3. (3)

      For each i1,i\neq 1,\ell,

      ΩiΓ/K×(0,1),Ω¯iΩ¯i+1Γ/K,andΩiΓ/KΓ/K,\Omega_{i}\approx\Gamma/K\times(0,1),\quad\overline{\Omega}_{i}\cap\overline{\Omega}_{i+1}\approx\Gamma/K,\quad\text{and}\quad\partial\Omega_{i}\approx\Gamma/K\sqcup\Gamma/K,

      where G×K±D±G\times_{K\pm}D_{\pm} denote disk bundles at the singular orbits.

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 a,b(0,d)a,b\in(0,d) and let uC2m(Ωa,b)u\in C^{2m}(\Omega_{a,b}) be a Γ\Gamma-invariant solution to the Dirichlet boundary problem (21) in Ωa,b:=r1(a,b)\Omega_{a,b}:=r^{-1}(a,b). If u=0u=0 in any subdomain of the form Ωc,d:=r1(c,d)\Omega_{c,d}:=r^{-1}(c,d), c,d[a,b]c,d\in[a,b], then u=0u=0 in the whole interval [a,b][a,b].

Proof.

As uu is a strong Γ\Gamma-invariant solution to (21), u=wru=w\circ r for some w:[0,d]w:[0,d]\rightarrow\mathbb{R} and Corollary 4.10, wC2m[a,b]w\in C^{2m}[a,b] is a strong solution to 23. Since u=0u=0 in Ωc,d\Omega_{c,d} and r0r\neq 0 in MM=r1(0,d]M\smallsetminus M_{-}=r^{-1}(0,d], then necessarily w(t)=0w(t)=0 for any t[c,d]t\in[c,d].

Now, if there is no t1[a,c][d,b]t_{1}\in[a,c]\cup[d,b] such that w(t)0w(t)\neq 0, then w0w\equiv 0 in [a,b][a,b] and there is nothing to prove. If this is not the case, there must be t0(a,c)(d,b)t_{0}\in(a,c)\cup(d,b) such that w(t0)0w(t_{0})\neq 0. Without loss of generality, suppose that t0(a,c)t_{0}\in(a,c); therefore, there must exist a<t1<t2ca<t_{1}<t_{2}\leq c such that w0w\neq 0 in (t1,t2)(t_{1},t_{2}) and w=0w=0 in [t2,c][t_{2},c]. As ww is of class C2mC^{2m}, all its derivatives of lower order are continuous and it follows that w(k)(t2)=0w^{(k)}(t_{2})=0 for every k=0,1,,2m1k=0,1,\ldots,2m-1. By existence and uniqueness of the initial value problem

^w=|w|p2w,w(k)(t2)=0,k=0,1,,2m1\widehat{\mathcal{L}}w=|w|^{p-2}w,\qquad w^{(k)}(t_{2})=0,k=0,1,\ldots,2m-1

ww vanishes identically in a small neighborhood of t2t_{2}, contradicting that w0w\neq 0 in (t1,t2)(t_{1},t_{2}). Something similar holds true if w(t0)0w(t_{0})\neq 0 for some t0[d,b)t_{0}\in[d,b). Hence w=0w=0 in [a,b][a,b], as we wanted to show. ∎

The following topological lemma will be useful in what follows.

Lemma 5.3.

Let X,YX,Y be two topological spaces and r:XYr:X\rightarrow Y be a quotient map. If r1(y)r^{-1}(y) is connected for every yYy\in Y, then r1(B)r^{-1}(B) is connected for every connected subset BYB\subset Y.

Proof.

Let BYB\subset Y be connected and consider A:=r1(B)A:=r^{-1}(B). Let f:A2:={1,1}f:A\rightarrow\mathbb{Z}_{2}:=\{-1,1\} be any continuous function. Since r1(y)r^{-1}(y) is connected for every yBy\in B, then for any x1,x2Xx_{1},x_{2}\in X, r(x1)=r(x2)r(x_{1})=r(x_{2}) implies that f(x1)=f(x2)f(x_{1})=f(x_{2}), for ff maps connected sets into connected sets. Therefore ff induces a continuous function f^:B2\widehat{f}:B\rightarrow\mathbb{Z}_{2} such that f^r=f\widehat{f}\circ r=f. As BB is connected, f^\widehat{f} is constant and so is ff. As f:A2f:A\rightarrow\mathbb{Z}_{2} was an arbitrary continuous function, it follows that AA must be connected. ∎

The following result allow us to describe the nodal domains in terms of the orbit structure.

Proposition 5.4.

Given a solution {Θ1,,Θ}𝒫Γ\{\Theta_{1},\dots,\Theta_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma} to the optimal Γ\Gamma-invariant \ell-partition problem (6), there exist points a1,,a1(0,d)a_{1},\dots,a_{\ell-1}\in(0,d) such that:

(0,d)i=1r(Θi)={a1,,a1}.(0,d)\setminus\bigcup_{i=1}^{\ell}r(\Theta_{i})=\{a_{1},\dots,a_{\ell-1}\}.

and, up to a relabeling,

Ω1:=Θ1M\displaystyle\Omega_{1}:=\Theta_{1}\cup M_{-} =\displaystyle= r1[0,a1)G×KD\displaystyle r^{-1}[0,a_{1})\approx G\times_{K-}D_{-}
Ωi:=Θi\displaystyle\Omega_{i}:=\Theta_{i} =\displaystyle= r1(ai1,ai)Md/2×(0,1)if i=2,,1\displaystyle r^{-1}(a_{i-1},a_{i})\approx M_{d/2}\times(0,1)\quad\quad\mbox{if $i=2,\dots,\ell-1$}
Ω:=ΘM+\displaystyle\Omega_{\ell}:=\Theta_{\ell}\cup M_{+} =\displaystyle= r1(a1,d]G×K+D+\displaystyle r^{-1}(a_{\ell-1},d]\approx G\times_{K+}D_{+}

Moreover, the sets Ω1,,Ω\Omega_{1},\dots,\Omega_{\ell} satisfy properties (b.1) to (b.3) of Theorem (5.1) and {Ω1,,Ω}\{\Omega_{1},\ldots,\Omega_{\ell}\} is also a solution to the Γ\Gamma-invariant \ell-optimal partition problem (6).

Proof.

First notice that every connected Γ\Gamma-invariant open set must be of the form r1(t,s)r^{-1}(t,s) for some t,s[0,d]t,s\in[0,d]. Now take three points t1,t2,t3(0,d)t_{1},t_{2},t_{3}\in(0,d) and define the sets V1=π1(t1,t2)V_{1}=\pi^{-1}(t_{1},t_{2}), V2=π1(t2,t3)V_{2}=\pi^{-1}(t_{2},t_{3}) and V=π1(t1,t3)V=\pi^{-1}(t_{1},t_{3}). Note that these sets are Γ\Gamma-invariant by construction, their boundaries are smooth because r1(ti)r^{-1}(t_{i}) is a principal orbit, and since every orbit is connected and rr is a quotient map, then V1,V2V_{1},V_{2} and VV 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

cVΓmin{cV1Γ,cV2Γ},c_{V}^{\Gamma}\leq\min\{c_{V_{1}}^{\Gamma},c_{V_{2}}^{\Gamma}\},

where this inequality holds true because ViVV_{i}\subset V and every nontrivial function in H0,gm(Vi)ΓH_{0,g}^{m}(V_{i})^{\Gamma} can be extended by zero to a nontrivial function in H0,gm(Vi)ΓH_{0,g}^{m}(V_{i})^{\Gamma}. We next prove that the inequality is strict. Suppose, to get a contradiction, and without loss of generality, that cVΓ=cV1Γc_{V}^{\Gamma}=c_{V_{1}}^{\Gamma} and let uH0,g(V1)Γu\in H_{0,g}(V_{1})^{\Gamma} be a least energy solution to the Dirichlet boundary problem (5) in V1V_{1}. Therefore the function u^H0,gm(V)Γ\widehat{u}\in H_{0,g}^{m}(V)^{\Gamma} given by u^=u\widehat{u}=u in V1V_{1} and u^=0\widehat{u}=0 in VV1V\smallsetminus V_{1} is a least energy solution to (5) in VV and by interior regularity [UhlenbeckViaclovsky2000], this function has a C2mC^{2m} class representative. Proposition 5.2 yields that u^\widehat{u} must vanish in VV, which is a contradiction and the strict inequality follows.

Hence, if {Θ1,,Θ}\{\Theta_{1},\ldots,\Theta_{\ell}\} is a Γ\Gamma-invariant solution to the \ell-partition problem (6), then (0,d)i=1r(Θi)(0,d)\setminus\cup^{\ell}_{i=1}r(\Theta_{i}) consists exactly of 1\ell-1 points, say a1,,a1a_{1},\dots,a_{\ell-1}.

Now define Ωi\Omega_{i} as in the statement. As r1(t)r^{-1}(t) is either a connected principal orbit or a connected singular orbit for each t[0,d]t\in[0,d], by Lemma 5.3 these sets Ωi\Omega_{i} are connected and Ωi\partial\Omega_{i} consists in one or two disjoint principal orbits, from which it follows that these sets are smooth. As r(Ω¯i)=[ai1,ai]r(\overline{\Omega}_{i})=[a_{i-1},a_{i}], for i=1,i=1,\ldots\ell (where a0:=0a_{0}:=0 and a=da_{\ell}=d), then Ω1Ω¯=M\overline{\Omega_{1}\cup\ldots\cup\Omega_{\ell}}=M. By definition, ΩiΩj=\Omega_{i}\cap\Omega_{j}=\emptyset if |ij|2|i-j|\geq 2, Ω¯iΩ¯i+1=r1(i)Md/2Γ/K\overline{\Omega}_{i}\cap\overline{\Omega}_{i+1}=r^{-1}(i)\approx M_{d/2}\approx\Gamma/K, and

Ωi={r1(ai,ai+1)Md/2×(ai1,ai)Γ/K×(0,1),i=2,,1,r1[0,a1)G×KD,i=1,r11(a1,d]G×K+D+,i=.\Omega_{i}=\begin{cases}r^{-1}(a_{i},a_{i+1})\approx M_{d/2}\times(a_{i-1},a_{i})\approx\Gamma/K\times(0,1),&i=2,\ldots,\ell-1,\\ r^{-1}[0,a_{1})\approx G\times_{K-}D_{-},&i=1,\\ r_{1}^{-1}(a_{\ell-1},d]\approx G\times_{K+}D_{+},&i=\ell.\end{cases}

The equivariant form of the sets Ω1\Omega_{1} and Ω\Omega_{\ell} follows from the Tubular Neighborhood Theorem [AlexBettiol, Theorem 3.57]. In fact, they are the associated bundles to the principal KK-bundle KΓΓ/KK\to\Gamma\to\Gamma/K.

Finally as ΘiΩi\Theta_{i}\subset\Omega_{i}, then cΩiΓcΘiΓc_{\Omega_{i}}^{\Gamma}\leq c_{\Theta_{i}}^{\Gamma} and {Ω1,,Ω}𝒫Γ\{\Omega_{1},\ldots,\Omega_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma} 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 νi=1\nu_{i}=1 in (9) for each i=1,,i=1,\ldots,\ell, and let (ηij,k)k(\eta_{ij,k})_{k\in\mathbb{N}} be a sequence of negative numbers such that ηij,k=ηji,k\eta_{ij,k}=\eta_{ji,k} and ηij,k\eta_{ij,k}\to-\infty as kk\to\infty. To highlight the role of ηij,k\eta_{ij,k}, we write 𝒥k\mathcal{J}_{k} and 𝒩k\mathcal{N}_{k} for the functional and the set associated to the system (9), introduced in Section 3, with ηij\eta_{ij} replaced by ηij,k\eta_{ij,k}. By Theorem 1.3, for each kk\in\mathbb{N} we can take u¯k=(uk,1,,uk,)𝒩k\overline{u}_{k}=(u_{k,1},\ldots,u_{k,\ell})\in\mathcal{N}_{k} such that

ckΓ:=inf𝒩k𝒥k=𝒥k(u¯k)=mNi=1uk,iPg2.c_{k}^{\Gamma}:=\inf_{\mathcal{N}_{k}}\mathcal{J}_{k}=\mathcal{J}_{k}(\overline{u}_{k})=\frac{m}{N}\sum_{i=1}^{\ell}\|u_{k,i}\|_{P_{g}}^{2}.

Let

𝒩0:={(v1,,v):\displaystyle\mathcal{N}_{0}:=\{(v_{1},\ldots,v_{\ell})\in\mathcal{H}:\, vi0,viPg2=M|vi|2m, and vivj=0 a.e. in M if ij}.\displaystyle v_{i}\neq 0,\;\|v_{i}\|_{P_{g}}^{2}=\int_{M}|v_{i}|^{{2^{*}_{m}}},\text{ and }v_{i}v_{j}=0\text{ a.e. in }M\text{ if }i\neq j\}.

Then, 𝒩0𝒩k\mathcal{N}_{0}\subset\mathcal{N}_{k} for all kk\in\mathbb{N} and, therefore,

(24) 0<ckΓc0Γ:=inf{mNi=1viPg2:(v1,,v)𝒩0}<.0<c_{k}^{\Gamma}\leq c_{0}^{\Gamma}:=\inf\left\{\frac{m}{N}\sum_{i=1}^{\ell}\|v_{i}\|_{P_{g}}^{2}:(v_{1},\ldots,v_{\ell})\in\mathcal{N}_{0}\right\}<\infty.

We claim that

(25) c0Γinf{Φ1,,Φ}𝒫Γi=1cΦiΓc_{0}^{\Gamma}\leq\inf_{\{\Phi_{1},\ldots,\Phi_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma}}\sum_{i=1}^{\ell}c_{\Phi_{i}}^{\Gamma}

Indeed, if {Φ1,,Φ}𝒫Γ\{\Phi_{1},\ldots,\Phi_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma} and viΦiΓH0,gm(Φi)Γv_{i}\in\mathcal{M}_{\Phi_{i}}^{\Gamma}\subset H_{0,g}^{m}(\Phi_{i})^{\Gamma}, then extending this function by zero outside Φi\Phi_{i}, we get that viHgm(M)Γv_{i}\in H^{m}_{g}(M)^{\Gamma} and vi,vj=0v_{i},v_{j}=0 a.e. in MM, for ΦiΦj=\Phi_{i}\cap\Phi_{j}=\emptyset. Therefore v¯:=(v1,,v)𝒩0𝒩1\overline{v}:=(v_{1},\ldots,v_{\ell})\in\mathcal{N}_{0}\subset\mathcal{N}_{1} and

c0ΓmNviPg2=𝒥1(v¯)=i=1JΦ(vi).c_{0}^{\Gamma}\leq\frac{m}{N}\|v_{i}\|^{2}_{P_{g}}=\mathcal{J}_{1}(\overline{v})=\sum_{i=1}^{\ell}J_{\Phi}(v_{i}).

As viΦiΓv_{i}\in\mathcal{M}_{\Phi_{i}}^{\Gamma} was arbitrary, it follows that

c0Γi=1cΦiΓ,c_{0}^{\Gamma}\leq\sum_{i=1}^{\ell}c_{\Phi_{i}}^{\Gamma},

and as {Φ1,,Φ}𝒫Γ\{\Phi_{1},\ldots,\Phi_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma} was arbitrary, inequality (25) follows.

From (24), it follows that the sequence (u¯k)(\overline{u}_{k}) is bounded in \mathcal{H}. So, using Lemma 2.2, after passing to a subsequence, we get that uk,iu,iu_{k,i}\rightharpoonup u_{\infty,i} weakly in Hgm(M)ΓH_{g}^{m}(M)^{\Gamma}, uk,iu,iu_{k,i}\to u_{\infty,i} strongly in Lg2m(M)L_{g}^{{2^{*}_{m}}}(M), and uk,iu,iu_{k,i}\to u_{\infty,i} a.e. in MM for each i=1,,i=1,\ldots,\ell. Moreover, as i𝒥k(u¯k)[uk,i]=0\partial_{i}\mathcal{J}_{k}(\overline{u}_{k})[u_{k,i}]=0, we have for each jij\neq i,

0Mβij|uk,j|αij|uk,i|βij1ηij,kM|uk,i|2mCηij,k.\displaystyle 0\leq\int_{M}\beta_{ij}|u_{k,j}|^{\alpha_{ij}}|u_{k,i}|^{\beta_{ij}}\leq\frac{1}{-\eta_{ij,k}}\int_{M}|u_{k,i}|^{{2^{*}_{m}}}\leq\frac{C}{-\eta_{ij,k}}.

Then, Fatou’s lemma yields

0M|u,j|αij|u,i|βijlim infkM|uk,j|αij|uk,i|βij=0.0\leq\int_{M}|u_{\infty,j}|^{\alpha_{ij}}|u_{\infty,i}|^{\beta_{ij}}\leq\liminf_{k\to\infty}\int_{M}|u_{k,j}|^{\alpha_{ij}}|u_{k,i}|^{\beta_{ij}}=0.

Hence, u,ju,i=0u_{\infty,j}u_{\infty,i}=0 a.e. in MM. By Lemma 3.1,

0<d0uk,iPg2M|uk,i|2mfor all k,i=1,,,0<d_{0}\leq\|u_{k,i}\|_{P_{g}}^{2}\leq\int_{M}|u_{k,i}|^{{2^{*}_{m}}}\qquad\text{for all \ }k\in\mathbb{N},\;i=1,\ldots,\ell,

and, as uk,iu,iu_{k,i}\to u_{\infty,i} strongly in L2m(M)L^{{2^{*}_{m}}}(M) and uk,iu,iu_{k,i}\rightharpoonup u_{\infty,i} weakly in Hgm(M)H_{g}^{m}(M), we get

(26) 0<u,iPg2M|u,i|2mfor every i=1,,.0<\|u_{\infty,i}\|_{P_{g}}^{2}\leq\int_{M}|u_{\infty,i}|^{{2^{*}_{m}}}\qquad\text{for every \ }i=1,\ldots,\ell.

Since u,i0u_{\infty,i}\neq 0, there is a unique ti(0,)t_{i}\in(0,\infty) such that tiu,iPg2=M|tiu,i|2m\|t_{i}u_{\infty,i}\|_{P_{g}}^{2}=\int_{M}|t_{i}u_{\infty,i}|^{{2^{*}_{m}}}. So (t1u,1,,tu,)𝒩0(t_{1}u_{\infty,1},\ldots,t_{\ell}u_{\infty,\ell})\in\mathcal{N}_{0}. The inequality (26) implies that ti(0,1]t_{i}\in(0,1]. Therefore,

c0Γ\displaystyle c_{0}^{\Gamma} mNi=1tiu,iPg2mNi=1u,iPg2mNlim infki=1uk,iPg2=lim infkckΓc0Γ.\displaystyle\leq\frac{m}{N}\sum_{i=1}^{\ell}\|t_{i}u_{\infty,i}\|_{P_{g}}^{2}\leq\frac{m}{N}\sum_{i=1}^{\ell}\|u_{\infty,i}\|_{P_{g}}^{2}\leq\frac{m}{N}\liminf_{k\to\infty}\sum_{i=1}^{\ell}\|u_{k,i}\|_{P_{g}}^{2}=\liminf_{k\to\infty}c_{k}^{\Gamma}\leq c_{0}^{\Gamma}.

It follows that uk,iu,iu_{k,i}\to u_{\infty,i} strongly in Hgm(M)ΓH_{g}^{m}(M)^{\Gamma} and ti=1t_{i}=1, yielding

(27) u,iPg2=M|u,i|2m,andmNi=1u,iPg2=c0Γ.\|u_{\infty,i}\|_{P_{g}}^{2}=\int_{M}|u_{\infty,i}|^{{2^{*}_{m}}},\qquad\text{and}\qquad\frac{m}{N}\sum_{i=1}^{\ell}\|u_{\infty,i}\|_{P_{g}}^{2}=c_{0}^{\Gamma}.

By Corollary 4.7, we can take u,iCm1(M(M+M))u_{\infty,i}\in C^{m-1}(M\smallsetminus(M_{+}\cup M_{-})) so that u,iu,j=0u_{\infty,i}u_{\infty,j}=0 in M(MM+)M\smallsetminus(M_{-}\cup M_{+}), iji\neq j. It follows from continuity that the set

Θi:={xM(MM+):u,i0},i=1,,\Theta_{i}:=\{x\in M\smallsetminus(M_{-}\cup M_{+})\;:\;u_{\infty,i}\neq 0\},\ i=1,\ldots,\ell

is nonempty, open, Γ\Gamma-invariant and ΘiΘj=\Theta_{i}\cap\Theta_{j}=\emptyset if iji\neq j.

Set

Ωi=int(Θ¯i),i=1,,.\Omega_{i}=\text{int}(\overline{\Theta}_{i}),\ i=1,\ldots,\ell.

These sets are also nonempty, Γ\Gamma-invariant and open, and satisfy that ΩiΩj=\Omega_{i}\cap\Omega_{j}=\emptyset if iji\neq j and u,i=0u_{\infty,i}=0 in MΩiM\smallsetminus{\Omega_{i}}. Hence {Ω1,,Ω}𝒫Γ\{\Omega_{1},\ldots,\Omega_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma}. Since each connected component of Ωi\partial\Omega_{i} is of the form r1(t)Md/2r^{-1}(t)\approx M_{d/2} for some t(0,d)t\in(0,d), it is smooth and

H0,gm(Ωi)={uHgm(M):u=0 in MΩi}.H_{0,g}^{m}(\Omega_{i})=\{u\in H_{g}^{m}(M)\;:\;u=0\text{ in }M\smallsetminus\Omega_{i}\}.

(Cf. [ClappFernandezSaldana2021, Lemma A.1] and [GrisvardBook, Theorem 1.4.2.2]). Hence, as u,i=0u_{\infty,i}=0 in MΩiM\smallsetminus\Omega_{i}, u,i0u_{\infty,i}\neq 0 in Ωi\Omega_{i} and satisfies (27), it follows that u,iΩiΓH0,gm(Ω)Γu_{\infty,i}\in\mathcal{M}_{\Omega_{i}}^{\Gamma}\subset H_{0,g}^{m}(\Omega)^{\Gamma}. As cΩiJΩi(u,i)c_{\Omega_{i}}\leq J_{\Omega_{i}}(u_{\infty,i}), using the claim (25) we obtain that

(28) inf{Φ1,,Φ}𝒫Γi=1cΦiΓi=1cΩiΓi=1JΩi(u,i)=mNi=1u,i2=c0Γinf(Φ1,,Φ)𝒫Γi=1cΦiΓ.\begin{split}\inf_{\{\Phi_{1},\ldots,\Phi_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma}}&\sum_{i=1}^{\ell}c_{\Phi_{i}}^{\Gamma}\leq\ \sum_{i=1}^{\ell}c_{\Omega_{i}}^{\Gamma}\leq\sum_{i=1}^{\ell}J_{\Omega_{i}}(u_{\infty,i})\\ &=\frac{m}{N}\sum_{i=1}^{\ell}\|u_{\infty,i}\|^{2}=c_{0}^{\Gamma}\leq\inf_{(\Phi_{1},\ldots,\Phi_{\ell})\in\mathcal{P}_{\ell}^{\Gamma}}\;\sum_{i=1}^{\ell}c_{\Phi_{i}}^{\Gamma}.\end{split}

Also, from this inequality we obtain that JΩi(u,i)=cΩiΓJ_{\Omega_{i}}(u_{\infty,i})=c_{\Omega_{i}}^{\Gamma} for every i=1,,i=1,\ldots,\ell, for otherwise, we would get that second inequality in (28) is strict, yielding a contradiction. Hence u,iu_{\infty,i} is a (weak) solution to the Dirichlet boundary problem (5) and {Ω1,,Ω}\{\Omega_{1},\ldots,\Omega_{\ell}\} is a solution to the Γ\Gamma-invariant \ell-partition problem (6). Proposition 5.4 yields that, actually, Ω1,,Ω\Omega_{1},\ldots,\Omega_{\ell} satisfy properties (b.1) to (b.3) in Theorem 5.1. ∎

Remark 5.1.

Changing the exponent p=2mp=2_{m}^{\ast} by any 2p2m2\leq p\leq 2_{m}^{\ast}, the arguments in this section yield a solution {Ω1,,Ω}\{\Omega_{1},\ldots,\Omega_{\ell}\} to the Γ\Gamma-invariant \ell-partition problem associated to the more general Dirichlet boundary problem (13), where the sets Ωi\Omega_{i} satisfy properties (b.1) to (b.3) in Theorem 5.1.∎

For the case m=1m=1, that is, when Pg=Δg+RgP_{g}=-\Delta_{g}+R_{g} is just the conformal Laplacian, we have the following result from which Corollary 1.2 follows immediately.

Corollary 5.5.

Let (M,g)(M,g) be a closed Riemannian manifold of dimension N3N\geq 3 and let Γ\Gamma be a closed subgroup of Isom(M,g)\text{Isom}(M,g) satisfying (Γ1)(\Gamma 1) to (Γ3)(\Gamma 3). If the scalar curvature RgR_{g} is positive, and if {Ω1,,Ω}𝒫Γ\{\Omega_{1},\dots,\Omega_{\ell}\}\in\mathcal{P}_{\ell}^{\Gamma} is the solution to the optimal Γ\Gamma-invariant \ell-partition problem given in Theorem 5.1, then the function

u:=i=1(1)iu,iu_{\ell}:=\sum_{i=1}^{\ell}(-1)^{i}u_{\infty,i}

is a Γ\Gamma-invariant sign-changing solution to the Yamabe problem

Δgu+N24(N1)Rgu=|u|212u,on M-\Delta_{g}u+\frac{N-2}{4(N-1)}R_{g}u=|u|^{2_{1}^{\ast}-2}u,\quad\text{on }M

having exactly \ell nodal domains and having least energy among all such solutions.

Proof.

Since the maximum principle is valid for the operator Pg=Δg+N24(N1)RgP_{g}=-\Delta_{g}+\frac{N-2}{4(N-1)}R_{g}, the least energy Γ\Gamma-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 u¯k\overline{u}_{k}, defined in the proof of Theorem 1.3, is nonnegative, yielding that u,iu_{\infty,i} is also nonnegative for every i=1,,i=1,\ldots,\ell and Ωi={xM(MM+):u,i>0}\Omega_{i}=\{x\in M\smallsetminus(M_{-}\cup M_{+})\;:\;u_{\infty,i}>0\}. 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]. ∎

Remark 5.2.

As it was already noticed in Remark 5.1, Theorem 5.1 holds true for QQ-curvature type equations with power nonlinearities 2p2m2\leq p\leq 2^{\ast}_{m}. Hence, the previous corollary is also true for Yamabe-type problems of the form

Δgu+Ru=|u|p2u,on M,-\Delta_{g}u+Ru=|u|^{p-2}u,\quad\text{on }M,

where 2p212\leq p\leq 2_{1}^{\ast} and R𝒞(M)R\in\mathcal{C}^{\infty}(M) is positive and Γ\Gamma-invariant.∎

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 (Γ3\Gamma 3) holds true.

Let Γ\Gamma be a closed subgroup of isometries of (M,g)(M,g) inducing a cohomogeneity one action. As before, the principal orbits of the action correspond to the hypersurfaces given by the regular level sets of rr, and if KK denote the principal isotropy, all of them are diffeomorphic to Γ/K\Gamma/K (see [AlexBettiol, Proposition 6.41]). Hence, we can fix one of these level hypersurfaces, say Md/2:=r1(d/2)M_{d/2}:=r^{-1}(d/2). When we have a cohomogeneity one action by isometries, we can describe the decomposition of the metric gg in terms of a one-parameter family of metrics on Md/2M_{d/2} as follows. It is known that, given a minimizing horizontal geodesic between the M+M_{+} and MM_{-}, a Γ\Gamma-invariant metric gg away from the singular orbits can be written as:

g=dt2+gt,g=dt^{2}+g_{t},

for t(0,d)t\in(0,d), where gtg_{t} is a smooth family of homogeneous metrics on Γ/K=Md/2\Gamma/K=M_{d/2}. Let 𝔤\mathfrak{g}, 𝔨\mathfrak{k} be the Lie algebras of Γ\Gamma and KK, respectively. Let 𝔪\mathfrak{m} be the orthogonal complement of 𝔨\mathfrak{k} in 𝔤\mathfrak{g}, with respect to a bi-invariant metric BB on 𝔤\mathfrak{g}. Since Γ\Gamma is compact, then it admits such a metric. There is a natural identification of 𝔪\mathfrak{m} with the tangent space to a principal orbit Md/2M_{d/2}. Namely, for each X𝔪X\in\mathfrak{m},

Xp:=ddt|t=0exp(tX)pX_{p}^{*}:=\left.\frac{d}{dt}\right|_{t=0}\exp(tX)\cdot p

is a tangent vector at pMd/2p\in M_{d/2}. Then gtg_{t} corresponds to a 11-parameter family of invariant inner products on 𝔪\mathfrak{m}.

One way to obtain the metric decomposition (Γ3\Gamma 3) is when 𝔪\mathfrak{m} decomposes into kk mutually orthogonal Ad(K)(K)-invariant subspaces:

(29) 𝔪=𝔪1𝔪k\mathfrak{m}=\mathfrak{m}_{1}\oplus\cdots\oplus\mathfrak{m}_{k}

such that the metric gtg_{t} can be written as:

gt=j=1kfj2(t)B|𝔪j,g_{t}=\sum_{j=1}^{k}f_{j}^{2}(t)\ B|_{\mathfrak{m}_{j}},

for some positive smooth functions fjf_{j}, j=1,,kj=1,\dots,k on (0,d)(0,d), and satisfying some smoothness conditions at 0 and dd. Such conditions describe a compactification of Md/2×(0,d)M_{d/2}\times(0,d) by adding two compact submanifolds, corresponding to the endpoints of (0,d)(0,d).

These types of cohomogeneity one metrics are called diagonal metrics. A decomposition in this form can be obtained in several settings:

  • If Γ\Gamma is a simple Lie group (i.e., if 𝔤\mathfrak{g} is simple), by the Schur’s Lemma [Hall15, Theorem 4.29].

  • If Γ\Gamma 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 𝔤\mathfrak{g} is negative definite. More generally, if the adjoint representation of 𝔤\mathfrak{g} is unitary, one also has such a decomposition. See [Hall15, Proposition 4.27].

Denote dj:=dim𝔪jd_{j}:=\dim\mathfrak{m}_{j}, j=1,,kj=1,\dots,k, then the volume form of gg is given by:

dVg=j=1kfjdjdtdV(B|𝔪j),dV_{g}=\prod_{j=1}^{k}f_{j}^{d_{j}}\ dt\ dV_{(B|_{\mathfrak{m}_{j}})},

and we recover the formula in Lemma 4.1.

Recall also a fundamental fact: A cohomogeneity one action can be determined through a group diagram K{K+,K}ΓK\subset\{K_{+},K_{-}\}\subset\Gamma, provided that K±/KK_{\pm}/K 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 (𝕊N,g)(\mathbb{S}^{N},g) with its canonical metric, which is an Einstein metric with positive scalar curvature. Let n1,n22n_{1},n_{2}\geq 2 be integers such that n1+n2=N+1n_{1}+n_{2}=N+1. Set

Γ=O(n1)×O(n2),\displaystyle\Gamma=O(n_{1})\times O(n_{2}), K=O(n11)×O(n21)\displaystyle K=O(n_{1}-1)\times O(n_{2}-1)
K+=O(n11)×O(n2),\displaystyle K_{+}=O(n_{1}-1)\times O(n_{2}), K=O(n1)×O(n21),\displaystyle K_{-}=O(n_{1})\times O(n_{2}-1),

where O(n)O(n) is the group of linear isometries of n\mathbb{R}^{n}. Note that we are regarding Γ\Gamma as acting on 𝕊n11×𝕊n21\mathbb{S}^{n_{1}-1}\times\mathbb{S}^{n_{2}-1}, trivially in one component, and with the transitive action by rotations in the other one. Then we obtain two possible isotropy groups K±K_{\pm}. A similar approach considers KK as a subgroup of K±K_{\pm}. In those cases, the isotropy is a copy of O(n11)O(n_{1}-1) or O(n21)O(n_{2}-1), for each corresponding case. Using that the (n1)(n-1)-sphere can be described as the quotient 𝕊n1O(n)/O(n1)\mathbb{S}^{n-1}\simeq O(n)/O(n-1), we obtain the following quotients:

Γ/K+=𝕊n11,Γ/K=𝕊n21,Γ/K=𝕊n11×𝕊n21,K+/K=𝕊n21,K/K=𝕊n11.\begin{array}[]{cc}\Gamma/K_{+}=\mathbb{S}^{n_{1}-1},&\Gamma/K_{-}=\mathbb{S}^{n_{2}-1},\\ \Gamma/K=\mathbb{S}^{n_{1}-1}\times\mathbb{S}^{n_{2}-1},&K_{+}/K=\mathbb{S}^{n_{2}-1},\quad K_{-}/K=\mathbb{S}^{n_{1}-1}.\end{array}

Hence, the group diagram K{K+,K}ΓK\subset\{K_{+},K_{-}\}\subset\Gamma defines a cohomogeneity one action of Γ=O(n11)×O(n21)\Gamma=O(n_{1}-1)\times O(n_{2}-1) on 𝕊N\mathbb{S}^{N} with singular orbits 𝕊n11\mathbb{S}^{n_{1}-1} and 𝕊n21\mathbb{S}^{n_{2}-1}, and with principal orbit 𝕊n11×𝕊n21\mathbb{S}^{n_{1}-1}\times\mathbb{S}^{n_{2}-1}. The orbit space is diffeomorphic to [0,π][0,\pi]. Therefore, this shows that conditions (Γ1)(\Gamma 1) and (Γ2)(\Gamma 2) are fulfilled.

The Lie algebra 𝔤\mathfrak{g} of Γ\Gamma is isomorphic to 𝔰𝔬(n1)𝔰𝔬(n2)\mathfrak{so}(n_{1})\oplus\mathfrak{so}(n_{2}), where 𝔰𝔬(n)\mathfrak{so}(n) denotes the Lie algebra of the n×nn\times n skew-symmetric matrices. It is a simple Lie algebra of dimension n(n1)/2n(n-1)/2, except for the case 𝔰𝔬(4)\mathfrak{so}(4) which is semi-simple. In this last case, its decomposition into simple factors is:

𝔰𝔬(4)=𝔰𝔬(3)𝔰𝔬(3).\mathfrak{so}(4)=\mathfrak{so}(3)\oplus\mathfrak{so}(3).

As previously, denote by 𝔪\mathfrak{m} the Ad(K){\rm Ad}(K)-invariant complement of 𝔨\mathfrak{k} in 𝔤\mathfrak{g}, where 𝔨\mathfrak{k} is the Lie algebra of KK. It is canonically identified with the tangent space at eKeK, TeKΓ/K𝕊n11×𝕊n21T_{eK}\Gamma/K\simeq\mathbb{S}^{n_{1}-1}\times\mathbb{S}^{n_{2}-1}. Then 𝔪\mathfrak{m} can be decomposed into two factors 𝔭1\mathfrak{p}_{1} and 𝔭2\mathfrak{p}_{2} given by:

𝔭1𝔰𝔬(n1)/𝔰𝔬(n11),and 𝔭2𝔰𝔬(n2)/𝔰𝔬(n21),\mathfrak{p}_{1}\simeq\mathfrak{so}(n_{1})/\mathfrak{so}(n_{1}-1),\quad\mbox{and\quad}\mathfrak{p}_{2}\simeq\mathfrak{so}(n_{2})/\mathfrak{so}(n_{2}-1),

with dim(𝔭1)=n11{\rm dim}(\mathfrak{p}_{1})=n_{1}-1 and dim(𝔭2)=n21{\rm dim}(\mathfrak{p}_{2})=n_{2}-1. If n1=4n_{1}=4 (or n2=4n_{2}=4), then

𝔭1(𝔰𝔬(3)𝔰𝔬(3))/𝔰𝔬(3)𝔰𝔬(3).\mathfrak{p}_{1}\simeq\left(\mathfrak{so}(3)\oplus\mathfrak{so}(3)\right)/\mathfrak{so}(3)\simeq\mathfrak{so}(3).

Therefore, in any case an invariant metric gg on 𝕊N\mathbb{S}^{N} can be written as:

g=dt2+f1(t)2B|𝔭1+f2(t)2B|𝔭2g=dt^{2}+f_{1}(t)^{2}B|_{\mathfrak{p}_{1}}+f_{2}(t)^{2}B|_{\mathfrak{p}_{2}}

where BB is a bi-invariant metric on 𝔤\mathfrak{g}. We may then take:

f1(t)=cos(t/2),f2(t)=sin(t/2).f_{1}(t)=\cos(t/2),\quad f_{2}(t)=\sin(t/2).

Observe that these functions satisfy the smoothness conditions (7). Then condition (Γ3)(\Gamma 3) is also satisfied.

The mean curvature h(t)h(t) of the principal orbit is:

h(t)=(n11)f1(t)f1(t)+(n21)f2(t)f2(t)\displaystyle h(t)=(n_{1}-1)\frac{f_{1}^{\prime}(t)}{f_{1}(t)}+(n_{2}-1)\frac{f_{2}^{\prime}(t)}{f_{2}(t)} =\displaystyle= (n11)2sin(t/2)cos(t/2)+(n21)2cos(t/2)sin(t/2)\displaystyle-\frac{(n_{1}-1)}{2}\frac{\sin(t/2)}{\cos(t/2)}+\frac{(n_{2}-1)}{2}\frac{\cos(t/2)}{\sin(t/2)}
=\displaystyle= 12(n21)cos2(t/2)(n11)sin2(t/2)sin(t/2)cos(t/2)\displaystyle\frac{1}{2}\frac{(n_{2}-1)\cos^{2}(t/2)-(n_{1}-1)\sin^{2}(t/2)}{\sin(t/2)\cos(t/2)}
=\displaystyle= 2(n1+n22)cos(t)sin(t)2(n2n1)sin(t).\displaystyle\frac{2(n_{1}+n_{2}-2)\cos(t)}{\sin(t)}-\frac{2(n_{2}-n_{1})}{\sin(t)}.

Then the volume of the principal orbits along (0,π)(0,\pi) is:

2|𝕊n11||𝕊n21|cosn11(t/2)sinn21(t/2)2|\mathbb{S}^{n_{1}-1}||\mathbb{S}^{n_{2}-1}|\cos^{n_{1}-1}(t/2)\sin^{n_{2}-1}(t/2)

where |𝕊ni1||\mathbb{S}^{n_{i}-1}| is the (ni1)(n_{i}-1)-dimensional measure of the sphere 𝕊ni1\mathbb{S}^{n_{i}-1}, for i=1,2i=1,2. This setting was under consideration in [ClappFernandezSaldana2021], where the authors studied the system (9) on N\mathbb{R}^{N} and on 𝕊N.\mathbb{S}^{N}.